content
stringlengths 1
15.9M
|
---|
\section{Introduction}
\label{intro}
The nature of the dark matter in the haloes of galaxies is an
outstanding problem in astrophysics. Over the last several
decades there has been great debate about whether this matter
is baryonic or must be exotic. Many astronomers believed
that a stellar or substellar solution to this problem might be
the most simple and therefore most plausible explanation.
However, recent analysis of various data sets has shown
that faint stars and brown dwarfs probably constitute no more than
a few percent of the mass of our Galaxy (Bahcall, Flynn, Gould,
and Kirhakos \cite{bfgk}); Graff and Freese \cite{gf96a};
Graff and Freese \cite{gf96b}; Mera, Chabrier, and Schaeffer \cite{mcs};
Flynn, Gould, and Bahcall \cite{fgb};
Freese, Fields, and Graff \cite{freese}). Hence the only surviving
stellar candidates of known populations are stellar remnants.
In this paper we consider severe constraints on white dwarf
stellar remnants. The situation for neutron stars is probably
even more restrictive. If indeed stellar candidates are ruled
out, one may be forced to more exotic nonbaryonic halo
dark matter.
We have been particularly motivated to consider white
dwarfs as Halo dark matter by recent results from
microlensing experiments (Alcock et al. \cite{macho:2yr};
Renault \cite{ren}), which have reported evidence
for Massive Compact Halo Objects (Machos) in the Halo of our
Galaxy. White dwarfs have been identified as plausible Macho
candidates because of the best-fit Macho mass of ($0.1-1$) $M_\odot$.
While some of our results are presented in the context of
a possible Macho interpretation, our chemical abundance
results constrain a white dwarf population in the Halo
regardless of what the Machos are.
In a previous paper (Fields, Freese, and Graff
\cite{ffg}), we discussed the baryonic mass budget
implied by a Galactic Halo interpretation of the LMC Macho events.
We found that a simple extrapolation of the Galactic
population (out to 50 kpc)
of Machos to cosmic scales gives a cosmic density
$\rho_{\rm Macho} = (1-5) \times 10^9 f_{\rm gal}
\, h \, $$M_\odot$$ \, {\rm Mpc}^{-3}$,
which in terms of the critical density corresponds to
\begin{equation}
\label{omegam}
\Omega_{\rm Macho}=(0.0036-0.017) h^{-1} f_{\rm gal} \, .
\end{equation}
Here the factor $f_{\rm gal} \geq 0.17$
is the fraction of galaxies that contain
Machos, as we argued in Fields, Freese, and Graff \cite{ffg},
and $h$ is the Hubble constant in units of 100 km s$^{-1}$ Mpc$^{-1}$.
This estimate applies regardless of the nature of the Machos,
and shows that Machos (if indeed they are in the Galactic Halo) are
a significant fraction of all baryons. Similar results
have been obtained by Steigman \& Tkachev \pcite{st}.
If one assumes--as we will hereafter--that the Machos are white dwarfs,
then stronger constraints result.
In particular, since white dwarfs are stellar remnants,
their formation necessarily requires both the formation of
progenitor stars, and ejection of the bulk of the progenitor mass
when the white dwarf is formed.
The simple requirement that the formation of white dwarfs is accompanied by
the release of at least as much mass in the form of hot gas
ejecta has profound consequences
which constrain white dwarfs as Machos.
For example, including progenitors in the Macho mass budget
increases the cosmological density of material needed to make Machos.
If Machos are white dwarfs resulting from a single
burst of star formation (without reprocessing of ejecta gas), then their
main sequence progenitors would have been
at least twice more massive:
$\Omega_\star \geq (0.007 - 0.034) h^{-1} f_{\rm gal}$.
Accounting for ejecta mass also has implications
on the scale of our Galaxy. The gaseous ejecta produced
along with the Galaxy's Machos would have had a
mass larger than what is measured in the known stellar
and gaseous components of the Galaxy. Thus, mass budget
considerations demand that most of the ejecta
left the Galaxy, which in turn requires
some kind of Galactic wind to remove it.
The ejecta produced by the white dwarf progenitors
lead to constraints not only due to
their mass, but also due to their composition.
The latter is the focus of this paper:
chemical abundance constraints on white dwarfs as Halo dark matter.
The ejecta contain the products of
nucleosynthesis--enrichment of some elements,
depletion of others--which become signatures
of white dwarf formation. We will show
that current models for low-mass stellar nucleosynthesis
predict a degree of processing which is so severe
that it rules out white dwarf Machos.
The most powerful constraints on white dwarfs as halo dark matter
come from carbon and nitrogen. However, the amount of these
produced is also dependent on the stellar model. Hence
we also consider the less powerful but unavoidable constraints from the
light element abundances, deuterium and helium.
We find that \he4 can be kept within observational
limits only for the lowest possible Macho density
$\Omega_{\rm Macho}$ compatible with Eq. 1, together
with high cosmic baryon density, and
Macho progenitor initial mass function (IMF) peaked at 2$M_\odot$ (so that there
are very few progenitor stars heavier than 3$M_\odot$).
The carbon and nitrogen yields
from white dwarf progenitors depend on the IMF of the stars
and on the amount of Hot Bottom Burning,
and are uncertain for zero metallicity stars. Still, best
estimates for these yields are in excess of observations of these elements
in our Galaxy (as first discussed for the case of carbon
by Gibson and Mould (1997)). Hence a galactic wind would
be required to eject these elements from the Galaxy
along with the excess mass. We show that such a wind could be driven
by Type Ia supernovae, which are produced by the same
white dwarfs in binary orbits with other stars. To produce
a successful wind, we find that at least
0.5\% (by mass) of stars must
to explode as supernovae. Such a scenario is reasonable,
since a comparable fraction of stars become supernovae in the Disk
of the Galaxy, if the star formation rate is
$\sim 1 \hbox{$M_{\odot}$}/{\rm yr}$ and the Type Ia rate is
$\sim 10^{-2}/{\rm yr}$ (Tutukov, Yungelson, \& Iben \cite{tyi}).
However, gas cooling may be
rapid enough to keep the bulk of the ejecta from being
evaporated. Furthermore, even if the C and N are ejected
from the Galaxy, they are still constrained by extragalactic
observations. Measurements of C and N in damped Lyman systems
and the Ly$\alpha$ forest are in excess of what would
be produced by a white dwarf Halo.
In addition, the Type Ia supernovae overproduce iron.
In Section \ref{IMF} we discuss white dwarf properties; we discuss
the initial mass function of the progenitor stars and the
relation between the masses of progenitor stars and the resultant
white dwarfs.
In Section \ref{chemev}, we present our
chemical evolution models which calculate the
effect of white dwarf production on
D, He, C, and O.
In Section \ref{sect:DHe}, we compare the expected
chemical abundances arising from white dwarf production
with observed D and He abundances
in various systems, and derive constraints on $\Omega_{\rm WD}$;
in section \ref{sect:CN} we derive constraints from
C and N, which in fact is more restrictive.
In Section \ref{wind},
we discuss the requirements for a Galactic
wind to remove chemical debris from the Galaxy.
We finish with a discussion in Section \ref{conclude}.
\section{White Dwarf Properties: IMF and Initial/Final Mass Relation:}
\label{IMF}
{\it Initial Mass Function:}
The progenitor stars of any white dwarf halo had to arise from
an initial mass function (IMF) that is strikingly different from any
observationally inferred IMF: a white dwarf progenitor IMF must have
very few stars less massive than $\sim 1$ $M_\odot$,
many intermediate mass stars, and few high mass
stars with mass greater than $\sim 8$$M_\odot$.
Adams and Laughlin (1996)
argued that the initial masses of halo white dwarf progenitors have to
be between 1 and 8 M$_\odot$. The lower limit on the range of
initial masses comes from the fact that stars with mass $< 1 M_\odot$
would still be on the main sequence.
The upper bound arises from
the fact that progenitor stars
heavier than $\sim 8 M_\odot$ explode as Type II supernovae, and leave behind
neutron stars rather than white dwarfs.
We can allow the IMF to have a small contribution
to higher masses so that there are some Type II supernovae and corresponding
remnant neutron stars, but not so many as to overproduce heavy elements.
Because low mass main sequence halo stars are intrinsically scarce
(Bahcall {\it et al. } \cite{bfgk}; Graff \& Freese 1996a,b),
an IMF of the usual Salpeter (1955)
type $dN/dm \propto m^{-2.35}$
is not appropriate, as it would imply a gross overabundance of low mass
stars in the Halo. Adams \& Laughlin (1996) propose a log-normal
mass function motivated by Adams \& Fatuzzo's (1996) theory of the IMF:
\begin{equation}
\label{lognormal}
\ln {dN \over dm}(\ln m) = A - {1 \over 2 \avg{\sigma}^2}
\Bigl\{ \ln \bigl[ m / m_C \bigr] \Bigr\}^2 \, .
\end{equation}
The parameter $A$ sets the overall normalization. The mass scale
$m_C$ (which determines the center of the distribution) and the
effective width $\avg{\sigma}$ of the distribution are set by the
star-forming conditions which gave rise to the present day population of
remnants. Possible values of the parameters are $m_C=2.3 M_{\odot}$
and $\avg{\sigma}=0.44$, which imply warm, uniform star-forming conditions
for the progenitor population. These parameters saturate the twin constraints
required by the low-mass and high-mass tails of the IMF, as discussed
by Adams \& Laughlin (1996), i.e., this IMF is as wide as possible.
Stars in the mass range 2-4 $M_\odot$
will produce different abundances of He, C, and N than an IMF with most
of the stars in the mass range 4-8 $M_\odot$.
Thus we will also examine the effect of narrowly peaked
IMFs chosen to highlight
the different nucleosynthesis within
the $1-8 \hbox{$M_{\odot}$}$ mass range.
{\it Initial/Final Mass Relation:}
The relation between the mass of a progenitor star and
the mass of its resultant white dwarf relies on an
(imperfect) understanding of mass loss from red giants. We use
the results of
Van den Hoek \& Groenewegen \pcite{vdhg}; these
are consistent with the results of Iben \& Tutukov \pcite{it}.
At the progenitor mass limits of interest, we have white dwarf masses
$m_{\rm WD}(1 $$M_\odot$$) = 0.55$ $M_\odot$, and
$m_{\rm WD}(8 $$M_\odot$$) = 1.2$ $M_\odot$.
\section{Chemical Evolution Calculations}
\label{chemev}
It is our goal to compare light element abundances
produced by white dwarf progenitors with the measurements
of the these abundances. In this section we describe our approach to
evolution calculations to estimate the element abundances
arising from MACHO progenitors.
First, in Section \ref{sect:analytic}, we describe
two different extreme approximations to bracket the
possible abundances that can arise. This analytic approach is
also useful in that it provides insight. Then, in
Section \ref{sect:numeric},
we discuss the numerical calculations. Below, in Sections
4 and 5, we will apply these calculations to
D and He, and then C and N. There we will present
the results of our calculations and compare them
with observations of these elements.
Chemical evolution calculates the
history of gas as it is processed into stars,
which ultimately die, leaving remnants and ejecting
processed material. Specifically, one calculates
the time development of the gas and comoving remnant
densities $\rho_{\rm gas}$ and $\rho_{\rm Macho}$,
as well as the gas density $\rho_{{\rm gas},i}$
in each isotope $i$. The abundances $i$ are expressed
in terms of mass fractions $X_i = \rho_{{\rm gas},i}/\rho_{\rm gas}$.
All of these components
change according to star formation and the resulting
star death.
As initial conditions for all models,
we take the baryons to be in gaseous form
with density $\rho_{\rm B}$. We take the primordial composition
of elements to be the big bang nucleosynthesis abundances
appropriate for the chosen $\rho_{\rm B}$,
$X_i^0 = \rho_{{\rm gas},i}^0/\rho_{\rm gas}^0 = \rho_{{\rm gas},i}^0/\rho_{\rm B}$.
Here superscript $0$ refers to primordial abundances.
{\it Homogeneity:}
In both analytic and numerical calculations, we assume that at high
redshifts the gas exists in a single ``homogeneous" chemical phase;
i.e., concentrations of various element
abundances are independent of spatial position.
A corollary of this
assumption is that outflow from stars is instantly and evenly mixed with
the primordial gas.
This approximation allows us to use the average co-moving density of a
chemical species as a useful parameter. We will refer to $\rho_{\rm B}$ as the
total co-moving baryon density, $\rho_g$ as the co-moving gas density,
$\rho_{\rm WD}$ as the comoving white dwarf density,
$\rho_{\rm H}$ as the comoving hydrogen density, etc.
This picture thus amounts to a universal ``post-processing''
of baryons that occurs after primordial nucleosynthesis.
In reality some regions are likely to have abundances enhanced over the
homogeneous levels, while other regions are likely to have abundances
closer to primordial. For example, the numerical simulations
of Cen and Ostriker (\cite{ceno}) suggest that the universe
is far from being chemically
homogeneous: high density regions tend to have
higher metallicity than low density regions.
If mixing is less efficient than we have assumed, the element
abundances inside dense star forming galaxies due to progenitors
of white dwarf Machos
would be higher than our predictions, while the abundances
outside these regions would be lower.
Lack of homogeneity makes the
measured galactic abundances harder to match and
thus more constraining. In the simulations
of Cen and Ostriker, the Ly$\alpha$ forest has a metallicity
roughly equal to the mean metallicity of the universe.
Thus, these forest lines are representative of the mean metallicity
results we calculate in our homogeneous models, and we
will use these lines below to compare theory with observation.
We do note, however, that a galactic wind
which drives material out of galaxies is likely to exist and
might be stronger than the one used in the Cen and Ostriker
simulations; such a wind drives the system towards homogeneity.
One can treat our results
as constraints on the efficiency with which the
enriched material is segregated from sites of subsequent
star formation.
\subsection{Abundances obtained with two Analytic Approximations}
\label{sect:analytic}
In this section we present analytic results of chemical abundances
obtained with two extreme approximations.
We consider two limits relating the star
formation time-scale $t_{\rm SFR}$ to the lifetime of a typical
star $t_*$ in our strongly peaked IMF.
In the limit where $t_{\rm SFR} \ll t_*$, or the {\it star burst}
limit, all the Machos are formed in a short time. Their ejecta mix into
the IGM, but are not incorporated into any second generation of Machos.
The opposite case where $t_{\rm SFR} \gg t_*$ is
the {\it instantaneous recycling} limit.
Here several generations of stars are
created, and the ejecta from stars of one generation
are mixed into the next generations of
stars. Within this limit, we can use the instantaneous recycling
approximation of chemical evolution which ignores the lifetime of
stars. Note that a very efficient wind, which removes ejecta
into the IGM as soon as they are produced, would make the
recycling case look more like a burst; in this case the ejecta from
a star are not mixed into the next generation of stars.
These two limits bracket any possible star formation scenario.
\subsubsection{Burst Model:}
We take the baryons in the universe at any time to consist of
three components, with comoving densities:
\begin{equation}
\rho_{\rm B} = \rho_{\rm gas} + \rho_{\rm star} + \rho_{\rm WD} \, ,
\end{equation}
where subscripts ``star" and ``Macho" refer to stars and
remnant white dwarfs respectively.
Initially all the baryons are in gaseous form with different
primordial abundances of various species.
During the star burst, a fraction $f_{\rm pro}$ of the
gas goes into stars, reducing $\rho_{\rm gas}$
from its initial density $\rho_{\rm B}$ by
an amount $f_{\rm pro} \rho_{\rm B}$. Once the stars die,
a fraction $R$
of the progenitor
mass is returned as processed gas.
Given a white dwarf progenitor IMF
$\xi_*(m) = dN_*/dm$, the gas return fraction is
\begin{equation}
\label{gasfraction}
R = {\int_{1M_{\odot}}^\infty dm \, m_{\rm ej}(m) \, \xi_*(m)
\over \int_0^\infty dm \, m \, \xi_*(m)} \, ,
\end{equation}
where $m$ is the mass of the progenitor,
which upon its death produces a remnant of mass $m_{\rm rem}$
and ejecta of mass $m_{\rm ej} = m - m_{\rm rem}$.
Thus, the density of ejected, processed gas is
$R f_{\rm pro} \rho_{\rm B}$;
there is no further processing of the ejecta.
A portion of the progenitor stars is left in the form of white dwarf Machos.
These objects will have a cosmic density
$\rho_{\rm WD} = f_{\rm pro} (1-R) \rho_{\rm B}$.
Thus a ``white dwarf Macho fraction''
\begin{equation}
f_{\rm WD} \equiv \rho_{\rm WD}/\rho_{\rm B} = f_{\rm pro} (1-R)
\end{equation}
of the baryons is turned into white dwarfs.
Note that in the burst scenario,
$f_{\rm WD} \le (1-R) < 1$.
In terms of the Macho fraction, the gas
density
after the burst is just
$\rho_{\rm gas} = \rho_{\rm B} - \rho_{\rm Macho} = [1-f_{\rm pro} (1-R)] \rho_{\rm B}$
by baryon conservation,
and the gas
fraction is $\mu = 1-f_{\rm WD} = 1 - f_{\rm pro}(1-R)$.
Hence, after the burst of star formation and the
evolution of the stars to stellar remnants has ended,
we are left with only gas and white dwarfs on the right hand
side of eqn. (3), with gas fraction $\mu$ and white dwarf
fraction $f_M$.
{\it Gas Composition:}
The initial gas density in each isotope $i$
is given by $\rho_{{\rm gas},i}^0 = X_i^0 \rho_{\rm B}$
where $X_i^0$ is the primordial abundance.
As a result of star formation and the subsequent evolution
of the stars, the composition of the gas has changed to:
$\rho_{{\rm gas},i} = \rho_{{\rm gas},i}^0 - f_{\rm pro} X_i^0 \rho_{\rm B}
+ \rho_i^{\rm eject}$.
The production of stars has lowered $\rho_{{\rm gas},i}$ by an amount
$f_{\rm pro} X_i^0 \rho_{\rm B}$. The ejecta of these stars once they die
has further changed it by $\rho_{{\rm gas},i}^{\rm eject}$. The
details of this latter quantity depend on the element.
In the process of stellar evolution, some gas is turned into helium
and some primordial deuterium is destroyed. In the remainder
of this section we describe our analysis of specific element
abundances in the burst model.
{\it Deuterium:}
All deuterium that passes through a star is destroyed.
Thus, $\rho_{{\rm gas},D}^{\rm eject}
= 0$, and
the post-Macho D density is
just that in unprocessed material:
$\rho_{{\rm gas},D} = (1 - f_{\rm pro}) X_D^0 \rho_{\rm B}$.
Thus the deuterium mass
fraction $X_{\rm D}$ after the burst is
\begin{equation}
\label{Dburst}
X_{\rm D} = \frac{1-f_{\rm WD}/(1-R)}{1-f_{\rm WD}} \
X_{\rm D}^0 \, .
\end{equation}
{\it Helium:}
As our notation we use $Y \equiv X_{\he4}$ to be the abundance
of \he4; we take the initial abundance to be $Y^0$. Some
of this helium is removed from the Galaxy by Machos, while
additional helium is added by the stellar evolution of the white dwarf progenitors.
In the case of helium, the ejecta are enriched:
$\rho_{{\rm gas},He}^{\rm eject}
= ( Y^0 R + \yld{He}) f_{\rm pro} \rho_{\rm B}$,
where the first term is the fraction of the primordial helium that is returned
as processed gas after the stars die and the second term is
the He production during stellar evolution. The helium yield in the second term,
\begin{equation}
\label{eq:yld}
\yld{He} =
{\int_{1M_{\odot}}^\infty dm \, (m_{{\rm ej,He}} - Y^0 m_{\rm ej}) \,
\xi_*(m) \over \int_0^\infty dm \, m \, \xi_*(m)} \, ,
\end{equation}
measures the He production, over and above the initial abundance
$Y^0$. Here $m_{\rm ej,He}$ is the mass of He ejected,
and $m_{\rm ej}$ is the total mass ejected.
For the Adams and Laughlin IMF (eq.\ \ref{lognormal}), and the Halo
metallicity stellar yields of Van Den Hoek \& Groenewegen (1997),
$\yld{He}=0.02$.
Since the helium yield is a roughly constant function of mass,
$\yld{He}$ is roughly independent of IMF for a range of white dwarf
IMFs.
The final, post-Macho He abundance is
thus $Y = (Y^0 \rho_{\rm B} - f_{\rm pro} Y^0 \rho_{\rm B} +
\rho_{{\rm gas,He}}^{\rm eject})/\rho_{\rm gas}$, which simplifies to
\begin{equation}
\label{y}
\Delta Y = \frac{ \yld{He} }{1-R} \ \frac{f_{\rm WD}}{1-f_{\rm WD}}
\end{equation}
{\it Carbon and Nitrogen:}
These elements have no primordial component, but are made by
stars. Thus the production of C and N is formally
similar to that of He (eq.\ \ref{y}),
with the exception that the
lack of a primordial component means that
$X_{\rm C}^0 = X_{\rm N}^0 = 0$.
Thus we have, after the burst,
\begin{eqnarray}
\label{cn}
X_{\rm C} & = & \frac{ \yld{C} }{1-R} \ \frac{f_{\rm WD}}{1-f_{\rm WD}} \\
X_{\rm N} & = & \frac{ \yld{N} }{1-R} \ \frac{f_{\rm WD}}{1-f_{\rm WD}} \, ,
\end{eqnarray}
where $\yld{C}$ and $\yld{N}$ are
defined in a way analogous to eq.\ \pref{eq:yld}.
\subsubsection{Instantaneous Recycling Approximation}
Within the instantaneous recycling approximation (IRA), we have
the well known results (e.g., Tinsley \cite{tins})
\begin{eqnarray}
\label{recyc}
X_{\rm D} &=& ( 1-f_{\rm WD} )^{R/(1-R)} \ X_{\rm D}^0 \\
\Delta Y &=& \frac{\yld{He}}{1-R} \ln \frac{1}{1-f_{\rm WD}} \\
X_{\rm C} &=& \frac{\yld{C}}{1-R} \ln \frac{1}{1-f_{\rm WD}} \\
X_{\rm N} &=& \frac{\yld{N}}{1-R} \ln \frac{1}{1-f_{\rm WD}} \, .
\end{eqnarray}
Note that
our ${\cal Y}_i \rightarrow (1-R) {\cal Y}_{{\rm Tins},i}$
in Tinsley's notation.
In this approximation there is no restriction on
$f_{\rm WD}$, unlike the burst case (see below eqn. (5)).
Note also that as in the burst case,
the ratios $\Delta$He:C:N are constant.
The burst and recycling solutions
agree to first order in $f_{\rm WD}$, but disagree at higher orders.
In particular, for a fixed $f_{\rm WD}$,
the burst model always gives a larger $\Delta Y$ and
a smaller $X_D/X_D^0$ than the instantaneous
recycling approximation does.
\subsection{Numerical Models}
\label{sect:numeric}
The chemical evolution model used here
is based on a code described in detail
elsewhere (Fields \& Olive \cite{fo98}).
The model allows for finite stellar ages
prior to the stellar death and the concomitant
remnant and ejecta production.
Thus the model assumes neither
instantaneous recycling nor the burst approximation,
which are equivalent to zero and infinite stellar lifetimes
respectively, relative to the timescale for star formation.
The star formation rate is chosen as an exponential
$\psi \propto e^{-t/\tau}$ with an $e$-folding time
$\tau = 0.1$ Gyr. We have investigated other $e$-folding times
up to $\tau = 1$ Gyr
and find that the results are insensitive to details
of the star formation rate.
The initial mass function will vary as indicated.
The model results are only as reliable as the nucleosynthesis
yields used.
For stars of $1-8 \hbox{$M_{\odot}$}$ we use the
yields of Van den Hoek \& Groenewegen (1997),
which allow for metallicity-dependence (but the lowest
calculated metallicity is $Z=0.001$, i.e., 1/20 solar).
For higher mass stars we use the yields of Woosley \& Weaver \pcite{ww},
though the IMFs we examine put very little mass into these stars.
For the initial D and He abundances of our calculations,
we have adopted the results of big bang
nucleosynthesis calculations, which
relate these quantities directly to $\rho_{\rm B}$ and the number of light
neutrino species $N_\nu$. We shall assume that $N_\nu=3$.
As we will illustrate below, we find that our numerical calculations
yield results very similar to those of the burst approximation.
The reason for this similarity is that many of the stars are
in the low mass range, so that they have long lifetimes
compared to reasonable star formation rates.
By the time they die, they can no longer contribute to
recycling in other stars.
\section{Deuterium and Helium}
\label{sect:DHe}
A large white dwarf component in the Galactic Halo
may lead to possible overproduction of helium and depletion
of deuterium. The results of our calculations for these
two elements are presented in this section, and compared with
observations. We will find that these elements can be kept
within observational limits only for $\Omega_{\rm WD} \leq 0.003$
and for a white dwarf progenitor initial mass function
sharply peaked at low mass (2$M_\odot$).
The problem of helium overproduction
has previously been investigated by Ryu, Olive,
and Silk (1990). In their work, they took the Galaxy to be
a closed box, in which there is no infall of unprocessed
gas to the Galaxy from the intergalactic medium (IGM),
and no outflow of processed gas from the
Galaxy into the IGM. They concluded that, in this closed box model,
the Halo could contain only a few white dwarfs, or else
the Galaxy would have no hydrogen left; all the hydrogen would have been turned
into helium. We will generalize their work here: we will move
beyond the closed box model and
consider the possibility that the processed gas is able to leave
the Galaxy via a galactic wind. The details of such a wind
will be discussed in a later section.
As we will see in Section 5, the overproduction of C and N
provide by far the severest chemical abundance constraint on a white dwarf
population in the Halo. However, this statement assumes that
we understand the dredge-up of C and N from the core
of the low-metallicity white dwarf progenitors (Chabrier
\cite{chabriernew}). Hence, in this section
we consider D and He, whose yields are far less uncertain.
Of all of the elements considered here,
the evolution of D is the cleanest: D is always destroyed
by stars and is not produced in significant amounts
by any astrophysical process other than the big bang
\pcite{els}.
Although He is produced by stars, as are C and N,
He production is farther out from the core of the star
so that the He yields are thus less uncertain than those of C and N.
On the other hand, Fields \& Olive \pcite{fo98}
found that published He yields have trouble with the $Y-Z$ slope in
dwarf galaxies. However, the difficulty was that
the model predictions {\em underestimate}
the slope compared to the observations, suggesting
that in fact the He yields themselves may be an underestimate.
In this sense, therefore, the constraints on He production
are conservative.
\subsection{Observational Constraints}
With the assumption of homogeneous abundances,
D and He are universally altered from their
primordial values. In this view, then,
the apparently ``primordial'' abundances
of D and He used to constrain BBN
are really ``pregalactic'' abundances
which have already had some
processing from their initial values.
We want to quote
D and He abundances in different environments and use
these as constraints on processing by white dwarf progenitors.
{\it Deuterium:}
The best available Galactic measurement
of deuterium is the abundance in the present day
local ISM.
Linsky \pcite{linsky} find
${\rm D/H}=(1.5\pm 0.1) \times 10^{-5}$.
The present day value has been
depleted by an unknown amount from
the original low metallicity value by galactic
disk stars, and thus provides a very conservative
lower limit on the D abundance and thus on pre-Galactic processing.
A stronger limit arises from
measurements of D in
quasar absorption line systems.
At present, different groups report different D/H values.
The strongest claims include
``high'' D/H $\simeq (8-25) \times 10^{-5}$
(Webb et al. \cite{webb};
Tytler et al.\ \cite{hityt})
measured in a system at
$z = 0.701$;
and ``low'' D/H $= (3-5) \times 10^{-5}$
(Burles \& Tytler \cite{bt98a}; Burles \& Tytler \cite{bt98b})
measured in two systems at $z > 3$.
These measurements are difficult and subject to systematic
errors (principally affecting H, rather than D).
It is thus unclear which (if either) of these values best
represents the primordial abundance.
Thus we will allowing a very generous range:
\begin{equation}
\label{eq:D}
{\rm D/H}_p = (3-25) \times 10^{-5} \, .
\end{equation}
{\it Helium:}
A best estimate of pre-galactic (i.e., normally
``primordial'') helium comes from extragalactic HII regions,
the lowest metallicity cases of which are in
blue compact dwarf galaxies.
The data are summarized in, e.g., Fields \& Olive \pcite{fo98}.
The large number of measurements now lead to a small statistical
error, so that {\em systematic} errors are now the limiting factor.
Again, we will take generous limits, adding the systematic error
linearly with the statistical errors (both at $1\sigma$):
\begin{equation}
\label{eq:He}
Y_p = 0.231 - 0.245
\end{equation}
\subsection{Model Results and Constraints}
\label{sect:DY-results}
The results of our calculation
depend on several parameters: the IMF of the white dwarf population,
the total density of white dwarfs
$\rho_{\rm WD}$, the Hubble constant,
and the total baryon density $\rho_{\rm B}$.
In general, the departure
from the big bang nucleosynthesis initial conditions increases as
$f_{\rm WD} = \rho_{\rm WD} / \rho_{\rm B}$ increases, i.e., as
white dwarfs become a larger fraction of the
baryons. We can see this in the analytical results.
As the white dwarf fraction increases
in Eqs. \ref{y} and eq. \ref{recyc},
helium and CNO enrichment increases, and
more deuterium is depleted.
We present results for four different sets of parameter choices here.
In the first model, we take $\Omega_{\rm WD} h=0.0036$, the lowest value
allowed by a simple extrapolation of
the Galactic Macho results to a cosmic scale in Eq. (1)
(Fields, Freese, \& Graff \cite{ffg}). In this model
we take the white dwarf IMF of
Adams and Laughlin (eq. \ref{lognormal}).
Figure \ref{fig:std} summarizes the nucleosynthetic processing
in two panels.
In Figure \ref{fig:std}a, we show
the values of $Y$ and D/H
which result from our calculations
(for various values of $\rho_{\rm B}$, and with $h=0.7$).
Shown are the
full numerical model, as well as the burst
and instantaneous recycling models.
Also shown are the initial values from big
bang nucleosynthesis and the (very generous)
range of
primordial values from eqs.\ \pref{eq:D} and \pref{eq:He}.
Note that the numerical model falls between the burst and
IRA, as expected. It is interesting to see that the full model
falls very close to the burst case. Thus we can conclude that
the burst model well-approximates the full results;
also, as the burst model gives stronger constraints,
the IRA results are in fact the most generous (and thus the most
conservative) bounds.
Since the previous model is obviously not consistent with measurements, we
also present, in Figure \ref{fig:min_consis}, a threshold model with results
barely consistent with measurements
of deuterium and helium. For this
model, we have kept the log-normal IMF suggested by Adam and Laughlin, but
with different parameters: our IMF is centered at $M_c=2$$M_\odot$ instead of
$2.3$$M_\odot$, and is narrower, with an effective width $\sigma=0.05$ instead
of 0.44. This IMF contains far fewer stars with initial mass $M>5$$M_\odot$,
and so produces less helium enriched gas, represented by the fact that $R$
drops slightly from 0.69 to 0.66.
We also drop $\Omega_{\rm WD} h$ down to 0.002,
somewhat below the lower bound of what is
suggested by the simple extrapolation
in eq. \ref{omegam} for $f_{gal}=1$.
This model is most constrained by the upper limit of the
He data. The allowed range in $\Omega_{\rm B}$ is $0.01-0.03$
(for $h=0.7$). Note that
to prevent over-production of helium, Machos are
a relatively modest $\sim 10\% $ of Baryons.
Figures \ref{fig:min_imf2} and \ref{fig:min_imf4}
represent the {\em minimum} cosmic processing
required if Machos are contained only in spiral Galaxies of
luminosities similar to the Milky Way:
$\Omega_{\rm WD} h = 6.1 \times 10^{-4}$
(Fields, Freese, \& Graff \cite{ffg}).
Figure \ref{fig:min_imf2} uses an IMF peaked at
$2 \hbox{$M_{\odot}$}$, designed
to minimize the effect on deuterium and helium abundances.
Figure \ref{fig:min_imf2}{\bf (a)} shows
that the effect on D and He is small and permissible
(but see the following section for discussion of
C and N production in this model).
Figure \ref{fig:min_imf4} uses the same
$\Omega_{\rm WD}$, but adopts an IMF peaked at
$4 \hbox{$M_{\odot}$}$. Note the increased D and He processing now
becomes unallowably large. Thus we are driven to a low
initial progenitor mass by the helium and deuterium abundances alone.
Note that white dwarf progenitors would lead to a raised floor in the
\he4 abundance. From Eq. (\ref{y}), one can see that, to obtain
the primordial helium abundance from the measured values,
one should really subtract the contribution due to white dwarf
progenitors. This would complicate the usual big bang nucleosynthesis
comparison of observed pregalactic abundances with
the primordial yields.
\section{Carbon and Nitrogen}
\label{sect:CN}
We illustrate here the difficulties of
reconciling the carbon and nitrogen
production with the abundance of white dwarfs in the Halo
suggested by the microlensing experiments.
\subsection{Production of C and N}
\label{sect:CNprod}
White dwarf progenitors are expected to produce prodigious amounts
of C and N. Here we discuss the relative production of these
two elements.
The relative amounts of C and N produced in
the asymptotic giant branch (AGB) phase are determined
by a process known as Hot Bottom Burning (hereafter HBB). During
HBB, the temperature at the bottom of a star's convective
envelope is sufficiently high for nucleosynthesis to take place
(Sackmann et al \cite{sack}, Scalo {\it et al. } \cite{scalo2}, Lattanzio
\cite{lattanzio}).
One of the main effects of HBB is to take the \c12 which is
dredged to the surface and process it into \n14 via the CN cycle.
Significant destruction of \c12 together with production
of \c13 and \n14 requires temperatures of at least
80 $\times 10^6$K. For low mass AGB stars ($m < 4$$M_\odot$), the effect of
HBB is negligible due to the low
temperature at the bottom of their envelopes. For high mass AGB
stars (m $ >$ 4$M_\odot$), the effect of HBB depends on the amount of matter
exposed to the high temperatures at the bottom of their envelopes,
the net result being the conversion of carbon and oxygen to nitrogen
(Boothroyd {\it et al. } \cite{booth}).
Yields of H, He, are not affected by HBB;
moreover, the total CNO yields also remain the same.
Since the CNO production is dominated by C and N,
this means that the sum C+N is independent of Hot Bottom Burning.
Thus, the main effect of Hot Bottom Burning is to
determine the degree to which C is processed into N,
but the sum remains the same.
With Hot Bottom Burning,
progenitor stars less massive than about 4 $M_\odot$\ produce significant
amounts of carbon and negligible nitrogen,
while heavier stars produce significant amounts of nitrogen and
negligible carbon.
Van den Hoek \& Groenewegen (1997) find that a star of mass
2.5$M_\odot$ and metallicity $Z = 0.001$
will produce $1.76 $ $M_\odot$ of ejecta of which $0.012 $ $M_\odot$ is new
carbon,
for an ejected mass fraction of $7 \times 10^{-3}$.
In comparison, the solar system composition
has a carbon mass fraction of $3.0\times 10^{-3}$.
In other words, the ejecta of a
typical intermediate mass star have more than twice the
solar enrichment of carbon.
If a substantial fraction of all
baryons pass through $1-4 \hbox{$M_{\odot}$}$ stars, the carbon abundance in this
model will be near solar.
These stars also produce $2.2 \times 10^{-4}$$M_\odot$ of N,
leading to an ejected mass fraction
$1.25 \times 10^{-4} \simeq X_{\rm N,\odot}/8$,
a much lower enrichment. On the other hand,
a 5$M_\odot$\ progenitor at the same metallicity
produces $X_{\rm C} = 7.2 \times 10^{-4} = 0.24X_{\rm C,\odot}$ and
$X_{\rm N} = 8.2 \times 10^{-3} = 7.4 X_{\rm N,\odot}$.
Hence, with Hot Bottom Burning, a white dwarf
IMF with stars in the mass range 1-4 $M_\odot$ produces a twice-solar
enrichment of
carbon, whereas a white dwarf IMF with stars in the mass range 4-8 $M_\odot$
produces seven times solar enrichment of nitrogen. An
IMF with stars in both regimes,
such as the Adams and Laughlin IMF in Eq. (2), produces both elements.
For comparison, van den Hoek and Groenewegen (1997) considered
the case of no HBB. Then stellar yields of carbon are seen to
dominate the total CNO-yields over the entire mass range, with
C production at the level of solar enrichment.
Models with HBB are favored as they are in excellent
agreement with observations, e.g. for AGB stars in the
Magellenic Clouds (Plez {\it et al. } \cite{plez}, Smith {\it et al. } \cite{smith}).
In the next section we will present results from our models
without Hot Bottom Burning; however, the presence of HBB
would not change our results as it merely trades a C overproduction
problem for a N overproduction problem.
A possible loophole to C and N overproduction
stems from the primordial, zero-metallicity composition
that the Macho progenitors would have.
Stellar carbon and nitrogen yields for zero
metallicity stars are quite uncertain,
and have not been systematically calculated
for the $1-8 \hbox{$M_{\odot}$}$ mass range of interest to us here.
Thus we use the yields of Van den Hoek \& Groenewegen (1997),
at the lowest metallicity, $Z=0.001 = Z_\odot/20$, and
as an approximation of the true $Z=0$ yields.
However, it is possible (although not likely) that carbon never
leaves the white dwarf progenitors, so that carbon overproduction is
not a problem (Chabrier \cite{chabriernew}). Carbon is produced
exclusively in the stellar core. In order to be ejected, carbon must
convect to the outer layers in the ``dredge up'' process. Since
convection is less efficient in a zero metallicity star, it is
possible that no carbon would be ejected in a primordial star. In
that case, it would be impossible to place limits on the density of
white dwarfs using carbon abundances.
On the other hand,
the 1$M_\odot$ model of Fujimoto et al.\ \pcite{fkih}
suggests that C and N are in fact highly enriched
due to strong mixing.
Indeed, there is
evidence (Norris, Ryan, \& Beers \cite{nrb})
for very strong C enrichment in some Halo giants,
suggesting a mixing effect.
The basic result of typical models with HBB is then that
a white dwarf IMF with stars in the mass range 1-4 $M_\odot$ produces a twice-solar
enrichment of
carbon, whereas a white dwarf IMF with stars in the mass range 4-8 $M_\odot$
produces seven times solar enrichment of nitrogen. An
IMF with stars in both regimes,
such as the Adams and Laughlin IMF in Eq. (2), produces both elements.
Without HBB, a solar enrichment of C is produced by all WD progenitor stars.
\subsection{Model Results}
In the figures, in panels b), we show
CNO abundances from the same four models discussed
previously for deuterium and helium. The CNO abundances are presented
relative to solar via the usual notation of the form
\begin{equation}
[{\rm C/H}] = \log_{10} \frac{{\rm C/H}}{({\rm C/H})_\odot} \, .
\end{equation}
For example, in this notation $[{\rm C/H}]=0$ represents
a solar abundance of C, while $[{\rm C/H}]=-1$ is 1/10 solar,
etc. Our C and N abundances were obtained without
including Hot Bottom Burning, which would exchange a C overproduction
problem for a N overproduction problem.
The effect of HBB would be to increase N at the expense of C,
keeping the sum C+N constant.
In Figure 1, we have $\Omega_{\rm WD} h = 0.0036$, the lowest
value allowed by Eq. (1). We take $h=0.7$ and the Adams-Laughlin
IMF in Eq. (2). We see that, even after dilution with the primordial
baryons,
the C and N abundances are still both greater than 1/10 solar
(e.g. [C/H] $>$ -0.8) over the entire range of $\Omega_B$.
Lower values of $\Omega_B$ correspond to higher C abundances
because there are fewer primordial baryons to dilute the C emerging
from the white dwarf progenitors.
In Figure 2, we have $\Omega_{\rm WD} h = 0.002$, $h=0.7$, and an IMF
peaked at 2$M_\odot$ as described previously. In Figures 3 and 4,
we have $\Omega_{\rm WD} h = 0.00061$, the minimum amount of WD
required to explain the microlensing results
if only Galaxies similar to ours produce WD Machos.
Figure 3 uses an IMF peaked at 2$M_\odot$ while Figure 4 uses an
IMF peaked at 4$M_\odot$. In all cases there is substantial
C and N production: in particular, the resultant C abundance is above
1/10 solar.
In the next section, we will show that, with or without HBB,
C and N exceed by at least 2 orders of magnitude the levels
seen in halo stars in our own Galaxy as well as by an order
of magnitude those in quasar absorbers.
\subsection{Observational Constraints}
White dwarf progenitors produce a huge amount of C and/or N.
With the assumption of homogeneity, the C and N produced
would give rise to a universal ``floor", i.e.,
an apparent Pop III component which might even be
mistaken as primordial. If the abundances are not homogeneous,
then the observations of C and N in various sites
can be used to obtain the required segregation of these
elements to keep them out of certain regions. In addition,
if one argues that C and N are underrepresented in some region,
then they must be enhanced elsewhere.
The overproduction of carbon and nitrogen can be a serious problem,
as emphasized by Gibson \& Mould \pcite{gm}.
They noted that white dwarf progenitors are expected to be the
main source of carbon. Thus the production of a white
dwarf population would be accompanied by a
copious production of carbon, without a corresponding
enrichment of oxygen, which is made predominantly by
Type II supernovae. The expected signature of
white dwarf production would be
anomalously high ratios of C/O and N/O,
i.e., ${\rm C/O} \mathrel{\mathpalette\fun >} 3 ({\rm C/O})_\odot$
and ${\rm N/O} \mathrel{\mathpalette\fun >} 3 ({\rm N/O})_\odot$.
However, metal-poor stars in our galactic halo have
C/O and N/O that are about 1/3 solar,
i.e., {\em below} and not above levels in Population I disk stars.
Thus Gibson \& Mould \pcite{gm} concluded that the gas
which formed these stars cannot have been polluted by the ejecta of a
large population of white dwarfs.
In using Galactic Halo star abundance
ratios as constraints,
the Gibson \& Mould \pcite{gm}
analysis assumes that 1) the Halo stars form at the same time
as the white dwarf progenitors, and 2) the Galaxy's Macho progenitor
ejecta would remain {\em in situ}.
It is possible that the observed low C spheroid
stars formed {\it before} the white
dwarf progenitors, in which case they would not be affected
by the metals produced later on by the white dwarf progenitors.
The authors
note that galactic winds could intervene but argue
these to be unlikely. However, they did not consider
the effect of Type Ia supernovae, which may in fact
be a natural engine to drive such winds
(though at the price of iron production;
see \S\ref{wind}).
Thus, in order to be generous to the
white dwarf scheme, we will examine C and N
production in terms of the absolute abundances produced,
and use these as constraints on the degree of
efficiency of the winds.
If the spheroid stars do not predate the white dwarf progenitors, then,
in our own Galaxy, the metal-poor Halo stars provide a strong constraint:
in these stars, neither C nor N has a
detectable ``floor" that would indicate a pre-Galactic component.
However, there is no evidence for such a floor,
which would appear as a constant C and/or N abundance as,
e.g., Fe decreases.
C has been observed with abundances at least as low as
$10^{-3} {\rm C/H}_\odot$;
and, N has been observed with abundances as low as
$10^{-3} {\rm N/H}_\odot$.
Thus if the production of these elements is of order solar,
as we have seen in the previous section,
the segregation between white dwarf progenitor ejecta and these Halo
stars must be very effective.
Mixing must be prevented with a
$\sim 99\%$ efficiency.
A way to achieve this segregation is with a Galactic wind,
which can remove C and N from the Galaxy.
If the C and N are expelled from the Galaxy,
the abundances of these
elements are constrained by measurements
in the intergalactic medium.
Carbon abundances in intermediate redshift
Ly$\alpha$\ forest lines have been measured to be
quite low.
Carbon is indeed present, but only at the
$\sim 10^{-2}$ solar level,
(Songaila \& Cowie \cite{sc}) in the Ly$\alpha$\ forest at $z \sim 3$
with column densities $N \ge 3 \times 10^{15} \, {\rm cm}^{-2}$.
Ly$\alpha$ forest abundances have also been recently measured
at low redshifts with HST (Shull {\it et al. } \cite{shull}) to be
less than $3 \times 10^{-2}$ solar.
The forest lines sample the neutral intergalactic medium.
With HBB, white dwarf progenitors in the mass range ($1-4$)$M_\odot$
$\,$ typically produce solar abundances of carbon; without HBB,
all white dwarf progenitors do so.
If we assume that the nucleosynthesis products of
the white dwarf progenitors do not avoid the neutral medium,
then these observations
offer strong constraints on scenarios for
ubiquitous white dwarf formation.
In order to maintain carbon abundances as low as $10^{-2}$ solar, only about
$10^{-2}$ of all baryons can have passed through the intermediate mass
stars that were the predecessors of Machos. Such a fraction can barely
be accommodated by the results in our previous paper (Fields, Freese,
and Graff \cite{ffg})
for the remnant density predicted from our extrapolation
of the Macho group results, and would be in conflict with
$\Omega_\star$ in the case of a single burst of star formation.
Note that, while the Halo star limit is not absolutely robust,
in that it could be avoided if the Halo stars predate the Machos,
the Ly$\alpha$ constraint cannot be avoided. Hence, below,
in obtaining numbers, we use the Ly$\alpha$ constraint.
Furthermore, in
an ensemble average of systems
within the redshift interval $2.2 \le z \le 3.6$,
with lower column densities
($10^{13.5} \, {\rm cm}^{-2} \le N \le 10^{14} \, {\rm cm}^{-2}$),
the mean C/H drops to $\sim 10^{-3.5}$ solar
(Lu, Sargent, Barlow, \& Rauch \cite{lsbr}).
One can immediately infer that, however carbon is produced
at high redshift, the sources do not enrich all material
uniformly. Any carbon that {\em had} been produced more
uniformly prior to these observations (i.e., at still
higher redshift) cannot
have been made above the $10^{-3.5}$ solar level.
These damped Ly$\alpha$ systems are thought to be possible
precursors to today's galaxies.
While measurements of nitrogen abundance have not been
made in the Ly$\alpha$ forest, there are measurements
in damped Ly$\alpha$ systems. The value of N/H in these systems
is measured to be typically $< 10^{-2}$ of solar,
and in one case at $z_{\rm DLA}=0.28443$ reported to
be as low as ${\rm N/H} = 10^{-3.79\pm0.08} {\rm N/H}_\odot$
(Lu et al \cite{lsbr}). In contrast, with HBB, white dwarf
progenitors in the mass range (4-8)$M_\odot$ produce seven times
the solar abundance of nitrogen.
In order to reconcile measurements of C and N
in damped Lyman systems with the much higher abundances
predicted by white dwarf progenitors, one would have to argue
that these elements are ejected from the damped Ly$\alpha$ systems, which
may be protogalaxies. Again a wind may be operative here. However,
the segregation requirements are even stronger,
particularly if N/H of $10^{-4}$ solar is to be taken seriously.
{\it Comparison with Model Results:}
We can compare these observations with our model results to
obtain more quantitative constraints when specific parameter
choices are made. Again, our models have no HBB included.
First let us assume that the abundances
we obtained in the figures apply homogeneously throughout the universe.
We will compare our results to the Ly$\alpha$ carbon measurements
of 10$^{-2}$ and the Halo measurements of $10^{-3}$.
Then in order to obtain agreement of
the C and N abundances we find in our Model 1 (see Fig. 1)
with the Ly$\alpha$ observations described above (which are a factor of 30
below the predicted values),
we must reduce the white dwarf densities by a factor of 30.
Hence we require $\Omega_{\rm WD} h \leq 0.0036/30 = 1 \times 10^{-4}$.
Alternatively, we require an actual abundance distribution
that is quite heterogeneous: those regions in which the observations
are made must be underprocessed. This implies departure from
the mean of a factor of at least 30, i.e., there must be segregation
efficiency of $1-1/30=97\%$.
The other figures confirm the results of Figure 1.
While the parameter choices of Figures 2 and 3 give acceptably low
D and He reprocessing, the C and N abundances are again
10-100 times what is observed. In Fig. 2 and 3, agreement with
Ly$\alpha$ forest requires $\Omega_{\rm WD} h \leq 1 \times 10^{-4}$.
Figure 4, with
an IMF peaked at 4$M_\odot$, overproduces all four elements.
This last model is the least restrictive when comparing
with the Ly$\alpha$ measurements, $\Omega_{\rm WD} h \leq
2 \times 10^{-4}$. Note that if C and N remain
inside the Galaxy and Halo stars do not
predate the white dwarf progenitors, then all these limits would
be an order of magnitude more powerful; the abundances must match
the measured C values of 10$^{-3}$ solar of the Halo stars.
Our results are mildly dependent on the redshift when C and N are expelled
into the IGM. If the C and N are not expelled until low redshifts, then
they would not be seen in intermediate redshift $(z=2-3)$ absorbers. Our
limits at low redshifts will be $\sim 3$ times
less restrictive since the observatonal limits are less restrictive.
However, removing the C and N from the Galaxy requires
supernovae. Since
large numbers of SN Type Ia are not seen out to
$z \sim 1$ (Hardin {\it et al. } \cite{hardin}), one must ensure that the
supernovae have mostly gone off by $z \sim 1$.
Thus the stronger bounds quoted previously in the session
apply unless the supernovae that ejected the material
take place precisely at $z \sim (1-2)$.
Hence the low measurements of C and N in the damped Ly$\alpha$
systems are hard to reconcile with the higher
predictions of C and N from white dwarf progenitors.
Thus, C and N indeed prove to be very restrictive;
in {\em all} models the mean cosmic production is
unacceptably large if it is homogeneously distributed.
As mentioned above, however, the abundances could well
be inhomogeneous due to galactic winds,
which would blow the C, N, and other
products of the white dwarf progenitors out of galaxies.
The D, He, C, and N measurements could be avoided as constraints only if
there is not much mixing, e.g. of hot outflowing gas and cool
infalling gas; with mixing, the material
essentially reenters the galaxies with a universal proportion.
In summary, low mass stellar progenitors produce a solar enrichment
of carbon; high mass stellar progenitors produce either
a solar abundance of carbon (without HBB) or a ten times
solar enrichment of nitrogen (with HBB).
Both elements are in conflict with measurements
inside our Galaxy and {\it must} be ejected from the Galaxy
if white dwarfs are to survive as Macho candidates.
Even outside our Galaxy, these abundances are hard to reconcile
with measurements of the Ly$\alpha$ systems.
We do wish to repeat the caveat, however, that the C and N
yields from low metallicity stars are still uncertain.
We close this section by
pointing out that extragalactic HII regions cannot contain
a substantial number of white dwarf Machos. These regions are
observed to have N and C increasing as the oxygen abundance
increases. White dwarf progenitors,
on the other hand, produce C and/or N without producing
O enrichment. One would have to argue that extragalactic
HII regions missed out in white dwarf formation.
\section{Galactic Wind}
\label{wind}
We have seen that the progenitors of a substantial white dwarf Halo
population would have produced a significant amount of pollution,
in conflict with observations.
In general one could avoid these constraints
by arguing for strong segregation between the hot gas emerging
from the progenitors and the cold gas where the element abundances
are measured. Then one views the incompatibility of the
predicted abundances with the observations as a measure of the
required efficiency of segregation of the hot ejecta from
the rest of the universe.
A possible means of removing excess abundances from the Galaxy is a Galactic
wind. As discussed in the Introduction, such
a wind is required to remove the excess gaseous baryonic material
left over from the Macho progenitors; this excess material
has more mass than the Disk and Spheroid combined,
is extremely polluted (with carbon, nitrogen, etc.)
and must be ejected from the Galaxy.
Indeed, as pointed out
by Fields, Mathews, \& Schramm \pcite{fms},
such a wind may be a virtue, as hot gas containing
metals is ubiquitous in the universe, seen in
galaxy clusters and groups, and present as an ionized
intergalactic medium that dominates the observed
neutral Ly$\alpha$\ forest. Thus, it seems mandatory
that many galaxies do manage to shed hot, processed material.
Here a galactic wind could remove helium, carbon and nitrogen
from the star forming regions and mix it throughout the universe.
Such a wind could be produced by supernova explosions providing
the energy source. The white dwarf IMF must therefore include
the stars responsible for the supernovae.
Possibilities include Type II supernovae from neutron stars
arising from massive progenitor stars; in this case the IMF
must contain some stars heavier than 8 $M_\odot$. The disadvantage
of such a scenario is that these heavy stars evolve {\it{more}}
quickly than the lighter stars that give rise to the white dwarfs;
i.e., the supernovae explosions would naturally take place
before the white dwarf progenitors have produced their
polluting materials. Then it would be hard to see how the
excess carbon and nitrogen could be ejected from the Galaxy.
We therefore propose the alternate possibility of Type Ia supernovae.
Here the same white dwarfs that are Macho candidates would
also be responsible for the supernova explosions. These
white dwarfs are in binary systems. Smecker \& Wyse (\cite{wyse})
have shown a problem with a binary system of two merging white dwarfs
as being responsible for the supernova explosions: too few
such explosions are seen in haloes today to allow us to
have enough of these earlier on to provide the required wind.
However, a scenario in which the white dwarf has a red giant companion can
be quite successful. The red giant
loses mass onto the white dwarf. When the white dwarf
mass approaches the Chandrasekhar mass, then there is a supernova
explosion. The timing is just right, since the supernova and
accompanying galactic wind takes place when low mass stars become
red giants. Thus the explosion and wind take place after the
white dwarf progenitors pollute the Galaxy with excess element
abundances, so that the wind is able to eject any excess helium, carbon and/or
nitrogen from the galaxy.
Here we now show that about 0.5\% (by mass) of the stars
must explode as Type Ia supernovae in order
to provide sufficient energy to produce the required
Galactic wind. Such a number is very reasonable, as it
is comparable to the number of Type Ia supernovae per white dwarf in the
disk of Galaxy.
Consider a protogalaxy with a baryonic mass
$M_B$, total mass $\hbox{$M_{\rm tot}$} = M_B + M_{\rm DM} \sim 10^{12} \hbox{$M_{\odot}$}$,
and size $R \sim 100 \, {\rm kpc}$.
The escape velocity is thus
\begin{equation}
\label{eq:vesc}
\hbox{$v_{\rm esc}$}^2 = 2 \ \frac{G \hbox{$M_{\rm tot}$}}{R} \sim (300 \, {\rm km} \, {\rm s}^{-1})^2
\end{equation}
For a supernova wind to be effective in evaporating gas
from the protogalaxy, it must heat the gas
to a temperature $T_{\rm gas}$ such that the wind condition
\begin{equation}
\label{eq:evap}
\frac{3}{2} kT_{\rm gas} = \frac{1}{2} m_p v_{\rm gas}^2
> \frac{1}{2}m_p {\hbox{$v_{\rm esc}$}^2}
\end{equation}
is satisfied,
or $kT_{\rm gas} \mathrel{\mathpalette\fun >} 0.3$ keV for the $\hbox{$v_{\rm esc}$}$ value in
eq.\ \pref{eq:vesc}.
This condition sets a lower limit to the number (and fraction) of
supernovae needed, as follows.
We envision a scenario wherein some baryons (i.e., gas) become
stars and ultimately their remnants and refuse,
while other gas remains
unprocessed. We thus write
\begin{equation}
M_B = M_\star + M_{\rm unpro} \ \ ,
\end{equation}
and we will denote the ``processed fraction''
$f_\star = M_\star/M_B$.
Furthermore, we note that some of the white dwarfs
will occur in binaries and will lead to Type Ia supernovae.
Consequently, some (most) of the stars
will meet their demise as white dwarfs and planetary nebulae (PN), while
some will die as supernovae:
$M_\star = M_{\rm PN} + M_{\rm SN}$.
We thus
denote the ``supernova fraction''
$f_{\rm SN} = M_{\rm SN}/M_\star$;
our goal here is to constrain $f_{\rm SN}$.
To get the constraint, we assume that the
three gas components--unprocessed, planetary nebulae, and
supernova ejecta--are mixed, and come to some temperature
$T_{\rm gas}$. Since the unprocessed and planetary nebula
components are much cooler than the supernova ejecta,
we can, to good approximation, put their temperatures to zero.
In this case, the temperature of the mixed gas is just
given by energy conservation:
\begin{equation}
\frac{3}{2} N_{\rm gas} \, kT_{\rm gas} = E_{\rm SN} N_{\rm SN}
\end{equation}
where $N_{\rm gas} = M_{\rm gas}/m_p$ is the number of gas molecules,
$N_{SN}$ is the number of supernovae that have gone off.
Also, $E_{\rm SN} \sim 10^{51} \, {\rm erg}$ is the mechanical
energy of the supernova, which is ultimately thermalized.
Furthermore, since $N_{\rm SN} = M_{\rm SN}/\avg{m_{\rm SN}}$, we have
\begin{equation}
\frac{3}{2} M_B \, kT_{\rm gas} = m_p \hbox{$\varepsilon_{\rm SN}$} M_{\rm SN}
\end{equation}
where $\hbox{$\varepsilon_{\rm SN}$} \equiv E_{\rm SN}/\avg{m_{\rm SN}}$ is the
specific energy per supernova. For Type Ia supernovae,
$\hbox{$\varepsilon_{\rm SN}$} \sim 10^{51} \, {\rm erg} / 5 \hbox{$M_{\odot}$}
= (3000 \, {\rm km} \, {\rm s}^{-1})^2$.
Collecting, then, we have
\begin{equation}
\frac{M_{\rm SN}}{M_B}
= \frac{3}{2} \frac{kT_{\rm gas}}{m_p \hbox{$\varepsilon_{\rm SN}$}}
\end{equation}
and since $M_{\rm SN}/M_B = f_{\rm SN} M_\star/M_B = f_{\rm SN} f_\star$,
we have
\begin{equation}
f_{\rm SN} f_\star = \frac{3}{2} \frac{kT_{\rm gas}}{m_p \hbox{$\varepsilon_{\rm SN}$}}
\end{equation}
Thus the condition of eq.,\ (\ref{eq:evap}) gives
\begin{eqnarray}
f_{\rm SN} f_\star & > & \frac{1}{2} \frac{\hbox{$v_{\rm esc}$}^2}{\hbox{$\varepsilon_{\rm SN}$}} \\
\Rightarrow f_{\rm SN}
& > & \frac{1}{2} \frac{\hbox{$v_{\rm esc}$}^2}{\hbox{$\varepsilon_{\rm SN}$}} f_\star^{-1} \\
& \sim & 5 \times 10^{-3} \ f_\star^{-1}
\end{eqnarray}
Thus we see that we need at least about 0.5\% (by mass) of the stars to
explode as Type Ia supernovae; more, if the processed fraction $f_\star$
is significantly lower than unity.
Thus far, we have only accounted for gas heating due to
the Type Ia supernovae, ignoring any cooling processes.
However, cooling processes will operate; for the temperatures
of interest, the dominant cooling mechanism is bremsstrahlung.
We can estimate the importance of cooling by computing the
cooling rate, $\tau_{\rm cool} = E/\dot{E}$, where $E \sim kT \sim 0.3$ keV
is the energy per gas particle, and $\dot{E}$ is the cooling rate
per particle. The cooling rate is $\dot{E} = \Lambda n$,
with $\Lambda \simeq 10^{-23} \, {\rm erg}\, {\rm cm}^{3} \, {\rm s}^{-1}$,
and $n$ the gas density. Assuming a constant density, we have
$n = {M_{\rm gas} \over {4\pi \over 3} R^3}$, where $M_{\rm gas}$ and $R$
are the mass and radius respectively of the WD gaseous ejecta. Thus
\begin{equation}
\tau_{\rm cool} = 0.2 \ {\rm Gyr} \
\left( \frac{M_{\rm gas}}{10^{11} \hbox{$M_{\odot}$}} \right)^{-1} \
\left( \frac{R}{50 \, {\rm kpc}} \right)^{3}
\end{equation}
for the fiducial gas mass and radii indicated.
We see that the cooling timescale is shorter
than longest stellar lifetime considered, $\tau(2\hbox{$M_{\odot}$}) = 1$ Gyr.
Thus cooling can be effective if the Type Ia supernova burst is
not rapid or the WD progenitors have masses $\mathrel{\mathpalette\fun <} 3 \hbox{$M_{\odot}$}$.
Furthermore, the cooling will be all the more effective
if the gas is inhomogeneous, as denser regions will cool much faster.
On the other hand, the cooling is very sensitive to the assumed
total radius $R$ of the WD gaseous ejecta.
Hence, cooling cannot rule out such a wind, but
it does demand that the wind be driven out on timescales
more rapid than $\sim 0.2$ Gyr.
Thus, if the cooling is indeed inefficient,
it is quite reasonable to use some of the white dwarf Macho
candidates as Type Ia supernovae to remove excess carbon
and nitrogen from the Galaxy.
However, SN Ia make prodigious amounts of iron, about
$m_{\rm ej}({\rm Fe}) \sim 1 \hbox{$M_{\odot}$}$ per event,
i.e., a large fraction of the mass going into Ia's
becomes iron
(Canal, R., Isern, J., \& Ruiz-Lapuente \cite{r-l}).
Thus we will expect a mass fraction of iron
of order
\begin{equation}
X({\rm Fe}) \sim M_{\rm SN}/M_{\rm B} = f_{\star} f_{\rm SN}
\sim 5 \times 10^{-3} \sim 4 \, X({\rm Fe})_\odot
\end{equation}
i.e., a very large enrichment.
Thus, while the SN Ia's can remove the gas
from the galaxies, they add their own contamination
which must be kept segregated from the
observable neutral material at a high precision.
(And the iron makes things all the worse as it also
adds to the cooling of the hot gas.)
\section{Conclusions and Discussion}
\label{conclude}
In conclusion, we have found that the chemical abundance constraints
on white dwarfs as candidate Machos are formidable.
The D and \he4 production by the progenitors of white
dwarfs can be in agreement with observation for low
$\Omega_{\rm WD}$ and an IMF sharply peaked at low masses
$\sim 2$$M_\odot$. Unless carbon is never dredged up from
the stellar core (as has been suggested by Chabrier \cite{chabriernew}),
overproduction of carbon and/or nitrogen is problematic.
The relative amounts of these elements that is produced
depends on Hot Bottom Burning, but both elements are produced
at the level of at least solar enrichment. Such enrichment
is in excess of what is observed in our Galaxy and must
be removed. A Galactic wind may have been driven by Type Ia supernovae,
which emerged from some of the same white dwarfs that are
the Machos. However, Ly$\alpha$ measurements in the IGM
are extremely restrictive and imply that these elements
must somehow be kept out of damped Ly$\alpha$ systems.
In addition these Type Ia supernovae overproduce iron
(Canal, R., Isern, J., \& Ruiz-Lapuente \cite{r-l}).
In sum, there is no evidence in Galactic halo stars, in external
galaxies, or in quasar absorbers for the patterns of chemical
pollution that should be formed along with a massive population
of white dwarfs. While this debris does carry the seeds of its
own removal in the form of Type Ia supernovae, the required
galactic winds must be effective in all protogalaxies, must arise
at redshifts $1 < z < 2$, and the debris must remain hot and segregated
from cooler neutral matter. Given these requirements,
we conclude that white dwarfs are very unlikely Macho candidates
unless they are formed in an unknown and unconventional manner.
With the failure of known stellar candidates as significant sources
of dark matter, one may be driven to exotic candidates. These
include Supersymmetric particles, axions, massive neutrinos,
primordial black holes (Carr \cite{carr}; Jedamzik \cite{jedam})
and mirror matter Machos (Mohapatra \cite{mohap}).
\bigskip
We thank Elisabeth Vangioni-Flam,
Grant Mathews, Scott Burles, Joe Silk, Julien Devriendt, Michel Cass\'e,
Jim Truran, Nick Suntzeff, Sean Scully, and Dave Spergel for helpful
discussions. We especially wish to thank
Dave Schramm, without whom none of us would be working
in the field of cosmology.
We are grateful for the
hospitality of the Aspen Center for Physics,
where part of this work was done.
DG acknowledges the financial support
of the French Ministry of Foreign Affairs' Bourse Chateaubriand
and the Physics and Astronomy Departments at Ohio State University.
KF acknowledges support from the DOE at the
University of Michigan.
The work of BDF was
supported in part by
DoE grant DE-FG02-94ER-40823.
|
\section{Introduction}
In recent years, atom optics has been rapidly emerging as a new and exciting
subfield of atomic physics. The objective of atom optics is
to manipulate atomic beams in a way similar to conventional optics by
exploiting the wave properties of the atoms. Supported by advances in
laser technology and microstructure fabrication a number of significant
accomplishments have been realized in the laboratory with the demonstration
of, e.g., mirrors, lenses, and diffraction gratings for atomic beams
\cite{AdaSigMly94}.
A natural way of extending these studies consists in exploring the
possibilities of holographic imaging with atoms, the conventional optics
analogue of which has been well-known for several decades
\cite{Gab48,Gab49}.
Optical holography can be described as the three-dimensional
reconstruction of the optical image of an arbitrarily shaped object.
Typically, this is done in a two-step process where first the
information about the object is stored in a hologram. This hologram is
created by recording, e.g., with the help of a photographic film, the
interference pattern between scattered light originating from the illuminated
object and a (plane-wave) reference beam. The second step is the
reconstruction, which is performed by shining a reading
beam similar to the reference beam onto the hologram. The diffraction
of the reading beam from the recorded pattern yields a virtual
as well as a real optical image of the original object.
Drawing on this concept, the characteristic property of atomic
holography is that at least the final reading step is performed
with an atomic beam. In this way, an atom-optical image of the object is
created which in certain situations can be thought of as some sort
of material replica of the original. There are several reasons why the
realization of atom holography is of interest: From a basic point of view,
it significantly extends the already well-established range of analogies
between light and matter waves. But more importantly perhaps, it may also
have useful practical applications from atom lithography to the
manufacturing of microstructures, or quantum microfabrication.
One of the prerequisites for an actual implementation of atomic
holography is the availability of a reading beam of sufficient
monochromaticity and coherence. Given the rapid advances in atom optics
and especially in the realization of atom lasers, this requirement can be
expected to be met in the near future. Another important question concerns the
potentially detrimental influence of gravitional effects. One of the greatest
challenges, however, is the manufacturing of the actual hologram where
the information to be reconstructed is stored. Several schemes can be
considered. One possibility is to diffract the atoms from a mechanical mask.
The first successful realizations of such an approach have recently been
reported in Ref.\ \cite{MorYasKis96}. In these experiments the hologram was
manufactured as a binary mask written onto a thin silicon nitride
membrane. Such a hologram has the advantage of being
permanent, however, as the mask only allows for complete or
vanishing (binary) transmission of the beam at a given point one loses a
significant amount of information about the optical image. Another
interesting proposal was recently made in Ref.\ \cite{Sor97}. In
this setup the atomic beam is diffracted from the inhomogeneous light
field created by the superposition of object and reference beam. These
beams thus directly form the hologram.
The purpose of the present paper is to investigate the perspectives of
an alternative approach, namely the manufacturing of the hologram directly
within a Bose-Einstein condensate (BEC). Atomic Bose condensates have recently
been realized experimentally \cite{AndEnsMat95,DavMewJofAndKet95} and are
now available almost routinely in several laboratories. The motivation for
the present study is twofold. First, the possibility of such an ``all-atomic''
scheme is interesting in itself and deserves further examination.
Furthermore, it illustrates the wide potential applicability of condensates
in atoms optics as a tool to influence the trajectories of atoms from
external sources.
Our approach is based on two main ideas. The holographic information
is encoded into the condensate in the form of density modulations by using
writing and reference laser beams that form an optical potential for the
condensate atoms. As we show later on, the density modulations follow
closely the optical beam interference pattern if the condensate is of
sufficiently high density, i.e., if a Thomas-Fermi description is applicable
\cite{DalGioPit98}. All-atomic reading is then accomplished in a way
reminiscent of the Raman-Nath regime of diffraction between an atomic beam
and a light field. \cite{AdaSigMly94} Specifically, the reading beam atoms,
that have a suitably chosen velocity, interact with the condensate atoms
via $s$-wave scattering and acquire a spatially dependent phase shift
reflecting the density modulations of the condensate. In the further
spatial propagation of the atoms, this phase shift gives rise to the formation
of the atom-optical image.
The proposed method is hence fundamentally different from a recent suggestion
to arbitrarily shape the center-of-mass wave function of an atom
(``wave-front engineering''). \ \cite{OlsDekHer98} Instead of pursuing an
holographic approach, this latter method makes use of a sequence of suitably
shaped laser pulses to obtain the desired wave front. In fact, our proposal
is more closely related to Ref. \cite{ChiFor98}, which also suggests using
of Bose-Einstein condensates to control particle deflection. Finally, we note
that the present work is also related to the discussion of
the analogy between matter-wave mixing phonomena in ultracold atomic samples
and conventional nonlinear optics, including in particular matter-wave
phase conjugation \cite{GolPlaMey95,GolPlaMey96,GolMey99} and four-wave mixing
\cite{LawPuBig98,TriBanJul98,Phi99}.
The paper is organized as follows. After briefly
recollecting the principles of optical holography Sec.\ II gives a
general discussion of our approach to atomic holography introduced
above. As an illustration in Sec.\ III the atom-optical imaging of a simple
object is worked out in detail. Summary and conclusions are given in
Sec.\ IV.
\section{Holographic imaging with atomic beams}
\subsection{Principles of optical holography}
The principles of optical holography in their most basic form are shown
in Fig.\ 1a; detailed expositions can be found, e.g., in
Ref.\ \cite{VelRey67}. An object is illuminated with a laser
wave and the resulting field $E_o({\bf r})$ is brought to interference
with the reference beam $E_r({\bf r})$. The ensuing superposition field
is recorded on a suitable medium, e.g., a photographic plate, in such
a way that the optical transmission of the medium becomes proportional
to the total field intensity
\begin{equation}
I=|E_r+E_o|^2=|E_o|^2+ |E_r|^2+E_r^\star E_o +E_rE_o^\star.
\label{inten}
\end{equation}
The wave front of a reading beam $E_{rd}({\bf r})$ impinging upon this
hologram is thus proportional to $I({\bf r})E_{rd}({\bf r})$
after it has penetrated the medium. In this expression,
the terms of interest are
$E_oE_r^\star E_{rd}+E_o^\star E_rE_{rd}.$
They contain the original object wavefront and its conjugate, and can be
used to construct a virtual and a real image of the object.
In optics several techniques have been developed which allow to
separately view each of these terms, such as side-band Fresnel holography,
Fraunhofer holography, Fourier transform holography, etc. \cite{VelRey67}.
For the present discussion of atom-optical holography we make use of
\centerline{\psfig{figure=holo_fig1a.ps,width=8.6cm,clip=}}
\centerline{\psfig{figure=holo_fig1b.ps,width=8.6cm,clip=}}
\begin{figure}
\caption{Set-ups for Fresnel side-mode
holography: (a) all-optical realization; (b) optical/matter-wave realization.
(Note the $-z$-axis in the last sketch.)}
\label{fig1}
\end{figure}
ideas from side-band Fresnel holography, which does not rely on lenses and
thus allows for a simple extension to matter waves.
\subsection{Atomic Bose-Einstein condensates as recording media}
The idea of storing information onto atomic condensates is based on the
observation that the density distribution of a condensate in the Thomas-Fermi
limit closely reflects the behavior of the confining potential. This yields
the possibility of accurate external control.
The Gross-Pitaevskii equation which governs the evolution of the
macroscopic wave function $\Phi({\bf r},t)$ describing the state of
an atomic condensate with $N$ atoms is given by \cite{DalGioPit98}
\begin{equation}
i\hbar \dot\Phi = \frac{{\bf p}^2}{2M}\Phi+V({\bf r})\Phi+
g|\Phi|^2\Phi,
\label{GP}
\end{equation}
where ${\bf p}$ denotes the atomic center-of-mass
momentum, $M$ the atomic mass, and $V({\bf r})$ the external potential.
The strength of atomic two-body interactions is determined by
$g=4\pi\hbar^2 a/M$ with $a$ being the $s$-wave scattering length.
The normalization condition for the condensate wave function reads
\begin{equation}
\int d^3{\bf r}\,|\Phi({\bf r})|^2=N.
\label{norm}
\end{equation}
The steady state of a condensate can thus be described with a
time-independent wave function $\phi({\bf r})$ which is defined by
$$\Phi({\bf r},t)=e^{-i\mu t/\hbar}\phi({\bf r}),$$
where $\mu$ is the chemical potential.
In the Thomas-Fermi limit, where the effect of kinetic energy is much weaker
than the mean-field potential, the contribution of the term ${\bf p}^2/2M$
can be neglected and the condensate density becomes
\begin{equation}
|\phi({\bf r})|^2=[\mu-V({\bf r})]/g,
\label{TFGS}
\end{equation} where $\mu$ is determined by the
normalization condition Eq.\ (\ref{norm}). From this expression we see
immediately that the form of the external potential is replicated
in the density profile of the atomic condensate.
Consider then replacing the photographic plate in Fig.1b by a pancake-shaped
BEC as the recording medium. In case the writing of the information into
the condensate is achieved by optical beams, which are assumed to be far
detuned from atomic resonance, they create an optical potential proportional
to $I/\delta$ where $I$ is given by Eq.\ (\ref{inten}) and
$\delta=\omega-\omega_L$ is the detuning between the atomic resonance
$\omega$ and the laser frequency $\omega_L$.
The total potential $V({\bf r})$ acting on the condensate
is then the sum of the trap potential, taken to be slowly varying,
and this optical potential. From Eq.\ (\ref{TFGS}), it then follows in
full analogy with optical holography that all terms in Eq.\ (\ref{inten})
are stored in the density distribution of the condensate ground
state. However, this atomic-condensate recording is in some ways more akin
to ``real-time'' holography, since the optical fields should be continuously
present in order to maintain the density modulations in the condensate.
\footnote{Note that the use of optical fields is not essential to the present
discussion: other interactions susceptible of imposing a spatially dependent
potential $V({\bf r})$ on the condensate can also be considered.}
\subsection{Reading from atomic condensates}
As already mentioned in the introduction, we consider an all-atomic reading
scheme, which has the fundamental advantage of allowing one to reconstruct
a {\em material} ``replica'' of the stored object. Specifically, the
reading beam is a monoenergetic atomic beam of velocity ${\bf v}_{rd}$
impinging at some angle onto the condensate. We assume that the
internal state of these incoming atoms is such that they are only weakly
perturbed by the writing and trap potentials, so that their dominant
interaction is scattering by the atoms in the condensate. It is important at
this point to emphasize that the atoms in the reading beam {\em need not}
be of the same species as the condensate atoms. In principle they
could be of just about any element or even molecule.
We consider specifically reading beam velocities such that the interaction
between the incoming atoms and the condensate can be described in terms
of $s$-wave scattering. This condition is fulfilled provided that \cite{LifPit80}
\begin{equation}
a_{rc}m_{rc}v_{rd}/\hbar \ll 1,
\label{scat_cond}
\end{equation}
where $a_{rc}$ denotes the $s$-wave scattering length for collisions
between reading and condensate atoms and $m_{rc}$ is their relative mass.
For simplicity, we further assume that the density of the reading beam is
low enough that collisions between atoms in that beam can be neglected.
Under these conditions the time evolution of the reading atoms' wave function
$\varphi({\bf r},t)$ in the mean field of the condensate is determined by
the equation
\begin{equation}
i\hbar \dot{\varphi}({\bf r},t)=\left[
\frac{{\bf p}^2}{2M_{rd}}+g_{rd}|\phi({\bf r})|^2\right]
\varphi({\bf r},t)
\label{Schr}
\end{equation}
where $M_{rd}$ is the mass of the atoms in the reading beam,
$g_{rd}=2\pi \hbar^2 a_{rc}/m_{rc}$. This equation
assumes that to a good degree of approximation, the condensate stays in
its ground state during the whole reading process.
Over the course of time, the condensate gradually loses atoms due to scattering
by the incoming atoms and other processes, but it is assumed that its
density distribution remains given by Eq.\ (\ref{TFGS}) with $\mu$
slowly varying due to the change in the number of atoms $N$, so that
$|\phi({\bf r})|^2$ has to be changed adiabatically in Eq.\ (\ref{Schr}).
Under these circumstances the shape of the holographic image will gradually change
and eventually distort when $|\phi({\bf r})|^2$ deviates too
much from the Thomas-Fermi expression. However, the time scale for this
process, the lifetime of the condensate, can be long in comparison to the
time necessary to form the image, the flight time of the reading atoms.
The reading and reconstruction of the condensate information into a
material ``replica'' is easily achieved if the condensate is sufficiently
thin that its density distribution can be regarded as effectively
two-dimensional, and its interaction with the reading atoms is short enough that
the Raman-Nath (or thin hologram) approximation can be invoked. The condensate
then acts as a phase grating for the reading beam, whose wavefront after
penetrating the condensate is given by
\begin{equation}
\varphi({\bf r},z_{c+},\tau)=\exp[-i g_{rd}|\phi({\bf r})|^2\tau /\hbar]
\varphi({\bf r},z_{c-},0).
\label{read}
\end{equation}
Here $\tau$ denotes the time it takes the probe atoms to pass through
the condensate of length $l_z$, $z_{c-}$ and $z_{c+}$ are the $z$-coordinates
just before and past the condensate respectively.
The situation described by Eq.\ (\ref{read}) is reminiscent
of phase holography in optics. In case
\begin{equation}
g_{rd} \max[|\phi({\bf r})|^2] \tau/\hbar\ll 1
\label{RNcond}
\end{equation}
we obtain $$\varphi({\bf r},z_{c+},\tau)\simeq (1-
i g_{rd}|\phi({\bf r})|^2\tau/\hbar)\varphi({\bf r},z_{c-},0),$$
i.e., the holographic information stored in the condensate is
indeed transferred to the reading beam. The subsequent free space
propagation allows to separate the different terms contained in
$|\phi({\bf r})|^2$ and to reconstruct the atom-optical replica of the stored
object.
\section{Example: Atom-optical imaging of a small aperture}
In this section we illustrate the principle of atom holography in the case of
imaging of a simple object. This example allows one to investigate
more closely under which conditions and to which degree the general scheme
of Section II can be realized in practice.
\subsection{Optical potentials}
The geometry we are considering is shown in Fig.\ 1b. The aperture and
the condensate are parallel to each other, their centers being located at
the points $(0,0,0)$ and $(0,0,z_c)$, respectively.
The aperture is illuminated by a plane optical wave $E_{inc}$
of amplitude ${\cal E}_0$ and wave vector $k_L$ propagating
along the $z-$direction. The emerging electric field is the well-known
Kirchhoff's solution to the associated diffraction problem \cite{VelRey67}
\begin{eqnarray}
E_o({\bf r},z_c)&=&-\frac{1}{2\pi}\int_{obj}d^2{\bf r}_0\,E_{inc}({\bf r}_0,0)
\left(ik_L-\frac{1}{R}\right)\times\nonumber\\
&& \cos \theta \frac{\exp(ik_LR)}{R},
\label{Kirh}
\end{eqnarray}
where ${\bf r}_0$ is a two-dimensional vector in the object plane and
$R=\sqrt{|{\bf r}-{\bf r}_0|^2+z_c^2}$ the distance from an object
point $({\bf r}_0,z=0)$ to a point $({\bf r},z_c)$ on the thin condensate
acting as a recording medium. Finally, $\cos\theta=z_c/R$, and the
integration is performed over the object boundaries.
The diffracted electric field acquires a simpler form in the Fresnel regime
(i.e., the paraxial or parabolic approximation) where $z_c\gg \lambda_L
=2\pi/k_L$, $|{\bf r}-{\bf r}_0|$ so that $\cos \theta\sim 1$ and
$R\sim z_c+|{\bf r}-{\bf r}_0|^2/z_c$. The object field at the location of
the condensate can then be approximated as
\begin{eqnarray}
E_o({\bf r},z_c)&\propto&e^{ik_Lz_c}e^{i\pi {\bf r}^2/\lambda_L z_c}
\nonumber\\
&&{\cal F}_2[{\cal E}_o O({\bf r}_0)e^{i\pi{\bf r}_0^2/\lambda z_c}
]\left |_{{\bbox \rho}={\bf r}/\lambda_L z_c},\right .
\label{fresn}
\end{eqnarray}
where $O({\bf r}_0)$ defines the shape of the object and
${\cal F}_2[]_{\bbox \rho}$ denotes the two-dimensional Fourier
transform with respect to ${\bbox \rho}$.
In order to construct the optical potential $V({\bf r})$ that imprints the
information onto the condensate, the object field $E_o({\bf r},z_c)$ is
interfered as in conventional holography with a plane wave reference
beam of amplitude ${\cal E}_r$ and wave vector ${\bf k}_r=k_L(\sin\beta,
0,\cos\beta)=(k_\perp,0,k_z)$, see Fig.\ 1b. The intensity of the
superposition at the location of the condensate is thus
\begin{eqnarray}
I({\bf r},z_c)&=& |E_o({\bf r},z_c)|^2+|{\cal E}_r|^2
\nonumber \\
&+&[{\cal F}_2[\dots] {\cal E}_r^\star e^{i(k_L-k_z)z_c
-i k_\perp x }e^{i\pi {\bf r}^2/\lambda_L z_c} +c.c.]
\label{int}
\end{eqnarray}
To make things clear, let us backtrack for a moment and imagine that
instead of a matter-wave hologram, we create an optical hologram from the
intensity distribution (\ref{int}). When illuminating that hologram with the
reading beam $E_{rd}={\cal E}_{rd}\exp{(ik_z z-i k_\perp x) }$ ,
we see the emergence of three wavefronts: the background wave
$E_{rd}(|E_o|^2+|E_r|^2)$ travelling along the direction $(-k_{\perp},0,k_z)$;
the object wave $E_{rd}E_o E_r^\star$; and the conjugate beam
$E_{rd}E_o^\star E_r$ which constitutes a converging wave front travelling
in the $z$-direction. Upon propagating a distance $z_c$ after the plane of
the hologram, the quadratic phase in the conjugate beam is undone and a
real image is created. Indeed, by applying again Kirchhoff's solution in
the Fresnel approximation of Eq. (\ref{fresn}) one obtains
\begin{eqnarray}
E_{im}({\bf r},2z_c)&\rightarrow& E_{rd}E_0^\star E_r
\propto e^{ik_Lz_c}e^{i\pi {\bf r}^2/\lambda_L z_c}
\nonumber\\
&\times&{\cal F}_2[E_{rd}E_0^\star E_r
e^{i\pi{\bf r}_0^2/\lambda z_c}
]\left |_{{\bbox \rho}={\bf r}/\lambda_L z_c},\right .
\label{image}
\end{eqnarray}
which is precisely proportional to $O({\bf r})$. The object wavefront,
on the other hand, corresponds to a virtual image. We now study the
conditions under which this same procedure can be applied to atom holography.
\subsection{The writing process}
We assume for concreteness that the condensate consists of sodium atoms, so
that laser fields with wavelengths of about $\lambda_L\sim 10^{-6}$m can
be used to create the optical potentials \cite{StaAndChi98}. It is assumed
to be trapped in a square well potential, as this provides a homogeneous
density of a condensate and thus avoids distortions of the holographic image.
However, one could also work with the approximately constant density
distribution near the center of a harmonic trap that is very wide in the
transverse directions.
To be specific, we investigate the imaging of a
rectangular aperture of width $D=10 \lambda_L$ located at a distance
$z\sim 1000 \lambda_L$ from the condensate. The emerging diffraction
pattern has an angular width of $\theta_d=\lambda_L/D\sim 0.1$, so that
the condensate must have an extension of at least $l_x\sim
100\lambda_L\sim 10^{-4}$m in the $x$-direction. Its extension $l_y$ along
the $y$-axis, as well as the width of the aperture, are
both assumed large enough that diffraction effects are negligible in that
direction, reducing the problem to an effective two-dimensional geometry.
This allows us to express the condensate wave function as
$\phi({\bf r})=\psi(x) /\sqrt{l_yl_z}$ where $l_z$ is the condensate
thickness, which is assumed to be very small as we recall from sec. II C.
Indeed, an upper limit to $l_z$ is provided by the condition that the
density distribution has to be effectively two-dimensional.
The periodicity of the intensity distribution (\ref{int})
along the $z$-direction is determined by the angle between reference and
writing beams; quantitatively, one obtains the requirement
$$
l_z\ll 2\pi/k_L(1-\cos\beta).
$$
In our numerical example we choose $\beta= 30^{\circ}$ and $l_z=10^{-6}$m
so that this condition is well satisfied. From the normalization condition
for a condensate in a square well potential one immediately finds that the
chemical potential is given by
\begin{equation}
\mu=\frac{Ng}{l_xl_yl_z}
\label{mu}
\end{equation}
and thus from Eq.\ (\ref{TFGS})
\begin{equation}
|\psi(x)|^2=\frac{N}{l_x}
\label{dens}
\end{equation}
The strength of the optical fields is determined from the requirement that
the recorded density profile of the condensate is mainly determined by the
light field intensity, and not by the trap ground state profile (pedestal).
This means that the field intensity must yield modulations of the optical
potential deeper than the trapping potential, i.e.
\begin{equation}
\max|\hbar\Omega^2({\bf r},z_c)/\delta|\gg \mu ,
\end{equation}
where the Rabi frequency $\Omega^2({\bf r},z_c)\propto I({\bf r},z_c)$.
In addition, since the atomic density of Eq. (\ref{TFGS}),
\begin{equation}
|\psi({\bf r},z_c)|^2=\frac{1}{g}{\left[\mu-V_{trap}({\bf r})-
\frac{\hbar\Omega^2({\bf r},z_c)}{\delta}\right]}
\end{equation}
where $V_{trap}$ is the trap potential, is non-negative, one should choose
a negative detuning $\delta$ so that $|\psi({\bf r},z_c)|^2$ reaches the
approximate value
\begin{equation}
|\psi({\bf r},z_c)|^2 \simeq \left|\frac{\hbar\Omega^2({\bf r},z_c)}
{g\delta}\right|.
\label{TFGS1}
\end{equation}
For the Thomas-Fermi approximation and thus Eq.\ (\ref{TFGS}) to be valid,
the condensate healing length $\xi$ needs to be much smaller than the
characteristic length scale $\lambda$ over which the external potential
varies, i.e.,
$$
\xi=(8\pi a n)^{-1/2} \ll \lambda
$$
where $n$ is the condensate density. In our case, $\lambda$ is of course
of the order of the optical writing beam wavelength, $\lambda \simeq
\lambda_L$. The healing length determines the length scale of a density
variation whose quantum pressure (kinetic energy contribution) is of the
order of the interaction energy \cite{DalGioPit98}. This leads to the
requirement
\begin{equation}
8\pi N a l_x/l_y l_z \gg (l_x/\lambda_c)^2
\end{equation}
which, with all other values previously fixed, translates into $N/l_y
\gg 2\times 10^{9}$. In our numerical example we work with the value
$N/l_y=2\times 10^{11}$ which correspond to $N=2\times
10^7$ for $l_y=10^{-4}$m.
We proceed by first determining numerically the ground state of the
condensate subject to the writing optical potentials, with the goal of
justifying the approximate density profile (\ref{TFGS1}). This is achieved
by solving the Hartree-Fock wave function evolution governed by the
Gross-Pitaevskii equation in imaginary time. For the parameter values listed,
we find a very good agreement between the shape of the density modulation
of the condensate ground state, see Fig.\ \ref{fig2}, and the optical
intensity distribution, see Fig.\ \ref{fig3}.
Further numerical simulations show that this density profile
can actually be prepared by adiabatically turning on the writing and reference
beams: Starting from the condensate in the trap ground state, the optical
fields are switched on slowly enough that no density oscillations become
significantly excited. In our specific example, we consider first the
square-well ground state $\psi(x)=\sqrt{N/l_x}$
and turn on the optical potential Eq.\ (\ref{int}) with a switch-on function
$[1-\exp(-t/t_c)]^2$, where $t_c\simeq 0.01$s is short in comparison to
typical condensate lifetimes. The resulting condensate density profile is
given in Fig.\ \ref{fig4}. This illustrates that in this way the condensate
wave function is transferred without difficulty from the trap ground state
to the new ground state in presence of the optical
potential, see Fig.\ \ref{fig3}.
\centerline{\psfig{figure=holo_fig2.ps,width=8.6cm,clip=}}
\begin{figure}
\caption{Condensate ground state density [m$^{-1}$] in a square-well potential
with the optical fields on. The object size is $ 10\lambda_L$, and the distance
from the object to the condensate is $z_c=1000\cdot \lambda_L$. The sodium
condensate parameters are $N=2\cdot 10^{7}$, $l_{x}=3\cdot 10^{-4}$m,
$l_y=10^{-4}$m, $l_z=10^{-6}$m, $V_{trap}(|x|>l_{x})=\mu$. The
optical field parameters are $|\hbar\Omega^2({\bf r},z_c)/\delta|
=100\cdot \mu$, ${\cal E}_r=0.1{\cal E}_o$, $\beta=30^0$}
\label{fig2}
\end{figure}
\centerline{\psfig{figure=holo_fig3.ps,width=8.6cm,clip=}}
\begin{figure}
\caption{Optical hologram for the same optical fields as used for
writing on the atomic condensate in Fig.\ \ref{fig2}.}
\label{fig3}
\end{figure}
\subsection{The reading process}
The reading and construction of a replica of the store object is achieved by
an off-axis atomic beam impinging upon the condensate from the side opposite
to the writing optical beams. The velocity of this beam needs to be
carefully selected, as it must be confined between a lower and an upper
bound resulting from the thin hologram and $s$-wave scattering approximations,
respectively.
As we have seen, a lower bound in atomic
velocities $v_{rd}$ in that beam is determined by the condition of validity of
Raman-Nath, or thin hologram, approximation. The physical meaning of this
condition is that the longitudinal
\centerline{\psfig{figure=holo_fig4.ps,width=8.6cm,clip=}}
\begin{figure}
\caption{Steady-state atomic density profile [m$^{-1}$] for the system of
Fig.\ \ref{fig2}, but achieved by adiabatically turning on the optical
potential in a characteristic time $t_c=0.01 sec$}
\label{fig4}
\end{figure}
\noindent
contribution to the kinetic energy of the probe atoms must be large
compared to their interaction energy with the condensate atoms
\begin{equation}
\frac{p_z^2}{2 M_{rd}}\gg g_{rd}max|\phi({\bf r})|^2
\label{ineq2}
\end{equation}
and the typical transverse deflection inside the condensate must remain
small compared to the length scale of the density fluctuations,
\begin{equation}
\frac{p_x \tau}{M_{rd}}\ll \lambda_L.
\label{ineq3}
\end{equation}
Under these circumstances, the kinetic energy term in
Eq.\ (\ref{Schr}) can be dropped. Eq.\ (\ref{ineq2}), together with the
requirement of small phase variations in the reading beam upon propagation
through the condensate (see Eq.\ (\ref{RNcond})), gives the lower bound for
$v_{rd}$.
In addition, Eq.\ (\ref{ineq3}) together with the requirement that
the $s$-wave scattering approximation is valid, see Eq. (\ref{scat_cond}),
determines an upper bound for $v_{rd}$.
Our numerical simulations are for $v_{rd}\sim 10^{-1}$ m/sec, which is a
reasonable value for the experiments with ultra cold atomic beams and lies
within these lower and upper bounds.
Similarly to the optical case, the atomic wave function acquires a quadratic
phase upon free propagation, a result of the fact that in the paraxial
approximation, the dispersion relations of optical and matter waves are
both quadratic and essentially the same. Indeed,
\begin{eqnarray}
& &\varphi(x,z_{c+}+\Delta z,\tau+\Delta t)=
\nonumber\\ & &
\int d\xi e^{i x \xi} e^{ -i\frac {\hbar \xi^2} {2 M_{rd}} \Delta t }
\int dx' e^{-i x \xi} \varphi(x',z_{c+},\tau) \nonumber \\
&&= e^{ i\frac {M_{rd}} {2 \hbar \Delta t} x^2 }
\int dx' \varphi(x',z_{c+},\tau)
e^ {i\frac{M_{rd}}{2 \hbar \Delta t}x'^2}
e^{-i\frac{M_{rd}}{\hbar \Delta t}x x'} \nonumber \\
&&=e^{i\frac{M_{rd}}{2 \hbar \Delta t}x^2} {\cal F}_2 [\varphi(x',z_{c+},\tau)
e^{i\frac{M_{rd}}{2 \hbar \Delta t}x'^2}]\left
|_{\xi=\frac{M_{rd}}{\hbar \Delta t}x}\right .
\label{atom_rd}
\end{eqnarray}
Here $\phi(x',z_{c+},\tau)$ is defined according with Eq.\ (\ref{read})
and $\Delta z = v_{rd}\Delta t$. The free propagation time $\Delta t$ is
determined from the same requirement as in the optical case, namely that
the quadratic phase $\exp[-i\pi x^2/\lambda_L z_c]$ in the recorded atomic
density (\ref{TFGS1}) be exactly compensated. This gives
\begin{equation}
\Delta t = \frac{M_{rd}}{2\hbar} \left (\frac{\lambda_L z_c}{\pi} \right ) .
\end{equation}
We finally observe that in order for the image to be formed on-axis,
the reading beam needs to propagate off-axis at an angle $\beta_A$ such
that it compensates the angle of an optical reference beam, i. e.
\begin{equation}
\sin\beta_A=k_L\sin\beta/k_A .
\end{equation}
This is a very small angle since $k_A\equiv v_{rd}M_{rd}/\hbar \gg k_L$.
Fig.\ \ref{fig5} shows the numerically computed atomic density profile in the
plane $z_c+\Delta z$ for this choice of parameters. This demonstrates
explicitly that the original rectangular aperture is indeed reconstructed
on axis. In contrast, the virtual image, for which the quadratic phase
stored in the condensate is not compensated, propagates off-axis, and so does
the background contribution. \\
\centerline{\psfig{figure=holo_fig5.ps,width=8.6cm,clip=}}
\begin{figure}
\caption{Reconstructing a replica of the original object from the atomic
hologram. The reading beam consists of a monochromatic beam of sodium atoms
moving at an angle $\beta_A$ from the $z$-axis at a velocity
$v_{rd}=0.1$ m/sec. Shown is the atomic density profile at a distance
$\Delta z$ from the condensate such that the quadratic phase shift of the
conjugate image is precisely canceled. The insert compares the reconstructed
and original objects. The off-axis feature for positive $x$ corresponds to
the real object, for which the quadratic phase is still present. The
large off-axis feature at negative $x$ is background.}
\label{fig5}
\end{figure}
\noindent
\section{Summary and Conclusions}
In this paper, we have theoretically discussed an atom holography scheme
where the hologram is stored in a Bose-Einstein condensate. An
important feature of the proposed scheme is that the reading beam needs not
consist of the same element as the condensate atoms. It merely needs to be a slow
monochromatic atomic or molecular beam that interacts with the condensate
atoms via $s$-wave scattering, or any other interaction leading to a cubic
nonlinearity in the nonlinear Schr\"odinger equation. Matter-wave holography,
once experimentally realized, is certain to open up the way to numerous
potential applications, in particular in microfabrication. One should note
that the slow atoms discussed in the present paper have large de Broglie
wavelengths, which severely limit the spatial resolution of the material
replica that can be reconstructed. However, the wavelength of matter waves
can easily be shortened, for example by gravitation. It is readily
conceivable that this can be used to achieve miniaturized structures with
nanometer-scale features. This, and other aspects of atom holography, will
be the subject of future studies.
\acknowledgements
The authors aknowledge numerious discussions and valuable suggestions from
M. G. Moore. This work is supported in part by the U.S.\ Office of Naval Research
under Contract No.\ 14-91-J1205, by the National Science Foundation under
Grant No.\ PHY98-01099, by the U.S.\ Army Research Office, and by the
Joint Services Optics Program.
|
\section{Introduction}
In astronomy the problem of correcting images for imperfect telescope
optics, atmospheric turbulence and other effects that adversely
influence the image quality is very well known. There exist therefore
many different ways to improve on images or imaging techniques in
order to obtain more detailed spatial information.
First of all there are hardware based solutions such as adaptive optics
and interferometry. It is of course always desirable to make use
of such techniques whenever possible. However, given a recorded
image with a known point spread function (PSF) and an estimate of
the point by point error it is still possible to improve on the
spatial resolution by means of software~: numerical image deconvolution.
A number of algorithms already exist that attempt to do this
image reconstruction. Best known in the field of interferometry at radio
wavelengths are probably the CLEAN method (H\"ogbom, 1974~;
Schwarz, 1978~; Wakker \& Schwarz, 1988) and the maximum entropy
method (MEM) (cf. Narayan \& Nityananda, 1986). For the deconvolution
of optical images the Richardson-Lucy (RL) method is well known
(cf. Richardson, 1972~; Lucy, 1974, 1992, 1994). Quite recently
a new method was presented by Magain, Courbin, \& Sohy (MCS, 1998).
A characteristic of all of these methods is that images are reconstructed
by placing flux (MEM and RL), or building blocks such as point sources
(MCS, CLEAN, two channel RL) in the field of view and minimizing
the difference between the image of this model (after convolution with
the PSF) and the actual image. Since this is an inverse
problem the solution is generally not unique and it can be quite sensitive
to errors in the data. Thus in the minimization there is some need for
regularization which is usually a smoothness constraint.
The various method differ in which building blocks are used and in the
form of regularization applied.
It is however somewhat unsatisfactory to proceed in this manner since
astronomical images are not generally easily described by just
point sources or objects of a certain shape or size, and may
not conform to the smoothness constraint applied. Forcing an
algorithm to nevertheless build the image up within such constraints
may well introduce an undesirable bias. For this reason it is useful to
consider an alternative that does not assume anything about properties
of the image in the deconvolution. In fact the method presented here
does not even use the image actually recorded until its very last step.
The method of subtractive optimally localized averages (SOLA) was
originally developed for application in the field of helioseismology
(cf. Pijpers \& Thompson, 1992, 1994) to determine
internal solar structure and rotation. Since then it has also been
successfully applied to the reverberation mapping of the broad line
region of active galactic nuclei (AGN) (Pijpers \& Wanders, 1994).
Instead of operating on the image, the SOLA method uses the PSF with
which the image was recorded. With this PSF and a user-supplied desired
PSF a linear transformation is constructed between any recorded
image for which that PSF applies and its deconvolved counterpart. The
resolution that can be attained in
this way is only limited by the sampling of the recording device
(the pixel size of the CCD) and by the level of the flux errors in the
recorded image. Since the transformation is linear, it is quite
straightforward to impose photometric accuracy. Astrometric
accuracy at the pixel scale is similarly guaranteed since there is
no `positioning of sources' in the image by the algorithm. Sub-pixel
accuracy, claimed for some deconvolution methods, implies subdividing each
pixel into subpixels. It requires knowledge of the PSF at very high
accuracy and very small errors in the data in order to deconvolve
down to a sub-pixel scale. If such information is available the
SOLA method can easily accommodate sub-pixel scale deconvolution,
without substantial modifications. In what follows however, it is assumed
that a single pixel is the smallest scale required.
In section 2 the SOLA method is presented. In section 3 the method
is applied to an example image to demonstrate the workings of
SOLA. In section 4 the method is applied to some astronomical images.
Some conclusions are presented in section 5.
\section{The SOLA method}
\subsection{arbitrary PSFs}
The strategy of the SOLA method in general is to find a set of linear
coefficients $c$ which, when combined with the data, produce a weighted
average of the unknown convolved function under the integral sign, where
the weighting function is sharply peaked. In the application at hand this
means finding the linear transformation between an image recorded at
a given resolution and an image appropriate to a different (better)
resolution.
The relation between a recorded image $D$ and the actual distribution of
flux over the field of view $I$ is~:
\eqnam\blurone{blurone}
$$
D(x,y)\ =\ \int\ {\rm d}x'{\rm d}y'\ K(x',y'~; x, y) I(x',y')
\eqno{\rm(\the\eqnumber)}\global\advance\eqnumber by 1
$$
where $K$ is the PSF. If one assumes that the PSF is constant over the
field of view then~:
\eqnam\psfcon{psfcon}
$$
K(x',y'~; x, y) \equiv K (x-x', y-y')
\eqno{\rm(\the\eqnumber)}\global\advance\eqnumber by 1
$$
Of course generally $D$ is not known as a continuous function of
$(x,y)$, but instead it is sampled discretely as for instance
an image recorded on a CCD. Thus one has as available data the recorded
pixel-by-pixel values of flux $D(x_i, y_j)$. These measured fluxes
will usually be corrupted by noise and thus the discretized version
of equation \blurone) is~:
\eqnam\blurdis{blurdis}
$$
D_{ij}\ =\ \int\ {\rm d}x'{\rm d}y'\ K_{ij}(x',y') I(x',y') + n_{ij}
\eqno{\rm(\the\eqnumber)}\global\advance\eqnumber by 1
$$
where now $K_{ij}$ refers to the PSF appropriate for
the pixel at $(x_i,y_j)$ and $D_{ij}$ is the flux value recorded in
that pixel. In the vocabulary usual for the SOLA method the $K_{ij}$
are referred to as integration kernels.
In the SOLA technique a set of linear coefficients ${c_l}$ is sought
which, when combined with the data, produces a value for the flux
$R$ in any given pixel that would correspond to an image recorded
with a much narrower PSF. Writing this out explicitly and using \blurdis)
yields~:
$$
\eqalign{
R &\equiv \sum c_l D_l = \cr
&\int\ {\rm d}x'{\rm d}y'\ \left\{\sum c_l K_{l}(x',y')
\right\} I(x',y') + \sum c_l n_{l} \cr}
\eqno{\rm(\the\eqnumber)}\global\advance\eqnumber by 1
$$
in which the double subscript $ij$ has been replaced by a single one
$l$ for convenience. Thus one would construct the ${c_l}$ such that
the averaging kernel ${\cal K}$ defined by~:
$$
{\cal K}(x',y') \equiv \sum c_l K_{l}(x',y')
\eqno{\rm(\the\eqnumber)}\global\advance\eqnumber by 1
$$
is as sharply peaked as possible. If one does this for all locations on
the CCD the collected values $R_m$ are then the fluxes
corresponding to the image at this (better) resolution with
a (improved) ``point spread function'' ${\cal K}$. The so-called propagated
error, the error in the flux $R$ is~:
\eqnam\fullerr{fullerr}
$$
\sigma_R^2 \equiv \sum\sum c_l c_m N_{lm}
\eqno{\rm(\the\eqnumber)}\global\advance\eqnumber by 1$$
Here the $N_{lm}$ is the error variance-covariance matrix of the
recorded CCD images where both $l$ and $m$ run over all $(i,j)$ combinations
of the pixel coordinates. If the errors are uncorrelated between pixels
then \fullerr) reduces to~:
\eqnam\simperr{simperr}
$$
\sigma_R^2 = \sum c_l^2 \sigma_{l}^2
\eqno{\rm(\the\eqnumber)}\global\advance\eqnumber by 1$$
which is trivially computed once the coefficients $c_l$ are known.
Ideally one would
wish to construct an image corresponding to an infinitely narrow
PSF~: a Dirac delta function. In practice this
cannot be achieved with a finite amount of recorded data. As
has already been pointed out by Magain {et al.} (1998) one must
in the deconvolved image still satisfy the sampling theorem.
A further restriction arises because of the noise term in equation
\blurdis). As is well known in helioseismology the linear combination
of data corresponding to a very highly resolved measurement usually
bears with it a very large propagated error. In order to obtain
a flux value for each pixel in the deconvolved image that does
not have an excessively large error estimate associated with it, one
needs to remain modest in the resolution sought for in the deconvolved
image.
Finding the optimal set of coefficients taking these limitations
into account can be expressed mathematically in the following minimization
problem. One needs to minimize for the coefficients $c_l$ the
following~:
\eqnam\minimize{minimize}
$$
\int\ {\rm d}x{\rm d}y\ \left[ {\cal K} - {\cal T}\right]^2 + \mu \sum
\sum c_{l} c_{m} N_{lm}
\eqno{\rm(\the\eqnumber)}\global\advance\eqnumber by 1
$$
Here $\mu$ is a free parameter which is used
to adjust the relative weight given to minimizing the errors in the
deconvolved image and to producing a more sharply peaked kernel
${\cal K}$. The higher the value of $\mu$ the lower this error but
the less successful one will be in producing a narrow PSF.
In SOLA one is free to choose the function ${\cal T}$. A common choice
in SOLA applications is a Gaussian~:
$$
{\cal T} = {1\over f \Delta^2} \exp \left[ - \left({(x-x_0)^2+(y-y_0)^2
\over\Delta^2}\right)\right]
\eqno{\rm(\the\eqnumber)}\global\advance\eqnumber by 1
$$
Here $(x_0, y_0)$ is the location for which one wishes to know the flux
at the resolution corresponding to the width $\Delta$. $f$ is a
normalization factor chosen such that~:
$$
\int\ {\rm d}x{\rm d}y\ {\cal T} \equiv 1
\eqno{\rm(\the\eqnumber)}\global\advance\eqnumber by 1
$$
although any set of locations $(x_0, y_0)$ can be chosen, a natural
choice in the application at hand is to take all original pixel locations
$(x_i, y_j)$. If one wishes to deconvolve to sub-pixel scales this
can be done by an appropriate choice of the $(x_0, y_0)$ and $\Delta$.
In terms of an algorithm the problem of minimizing the function \minimize)
leads to a set of linear equations~:
\eqnam\lineqs{lineqs}
$$
A_{lm} c_{l} = b_{m}
\eqno{\rm(\the\eqnumber)}\global\advance\eqnumber by 1
$$
The elements of the matrix $A$ are given by~:
\eqnam\matrix{matrix}
$$
A_{lm} \equiv \int {\rm d}x{\rm d}y\ K_l (x,y) K_m (x,y) + \mu N_{lm}
\eqno{\rm(\the\eqnumber)}\global\advance\eqnumber by 1
$$
The elements of the vector $b$ are given by~:
\eqnam\vector{vector}
$$
b_{m} \equiv \int {\rm d}x{\rm d}y\ {\cal T} (x,y) K_m (x,y)
\eqno{\rm(\the\eqnumber)}\global\advance\eqnumber by 1
$$
Writing out the dependencies on the free parameters explicitly,
determining the coefficients $c_l$ results from a straightforward
matrix inversion~:
\eqnam\matinv{matinv}
$$
c_l (x_0, y_0~; \Delta, \mu) = A^{-1}_{lm} (\mu) b_m (x_0, y_0~; \Delta)
\eqno{\rm(\the\eqnumber)}\global\advance\eqnumber by 1
$$
It is clear that for each point $(x_0, y_0)$ there is a separate set
of coefficients $c_l$ which will depend on the resolution width $\Delta$
required and on the error weighting $\mu$.
Note that it is not necessary to invert a matrix for every location
$(x_0, y_0)$, which would certainly be prohibitive if one wishes to
calculate the entire deconvolved image. For a given error weighting
$\mu$ one needs to invert $A$ only once. Only the elements of the
vector $b$ need be recomputed for different locations or different
resolutions.
In order to ensure that at every point in the reconstructed image the
summed weight of all measurements is equal and thus a true (weighted)
average it is necessary to additionally impose the condition~:
\eqnam\unitav{unitav}
$$
\sum c_l \equiv 1
\eqno{\rm(\the\eqnumber)}\global\advance\eqnumber by 1
$$
It is this condition that imposes photometric accuracy on the reconstructed
image. Using the method of Lagrange multipliers this condition is easily
incorporated into the matrix equation \lineqs) by augmenting the
matrix $A$ with a row and column of $1$'s, and a corner element
equal to $0$. The vector $b$ gains one extra element equal to $1$
as well. The details of this procedure can be found in Pijpers \& Thompson
(1992, 1994).
\subsection{translationally invariant PSFs}
Although the method described above can work in principle with general
PSFs $K$, the matrix inversion becomes intractable
very quickly as the number of pixels increases. For an image of
$M \times M$ pixels the number of elements in the matrix A is
$M^2 \times M^2$. The matrix $A$ is symmetric but even so a naive
matrix inversion routine would require a number of operations
scaling as $M^6$.
However the entire procedure for obtaining the transformation coefficients
for all locations of the CCD can be speeded up considerably if one
accepts some restrictions for the properties of the PSF $K$ and of the
expected errors $N_{lm}$. The first restriction is to assume that the PSF is
constant over the field of view, that is to say that equation \psfcon)
is valid.
When condition \psfcon) is met one can easily demonstrate that
in equation \matrix) the integrals of the cross products of the PSFs
are a convolution~:
\eqnam\matconvlv{matconvlv}
$$
\eqalign{
\int &{\rm d}x{\rm d}y\ K_l (x,y) K_m (x,y) \equiv \cr
&\int {\rm d}x{\rm d}y\ K(x-x_{i_l}, y-y_{j_l}) K(x-x_{i_m}, y-y_{j_m}) \cr
= &\int {\rm d}x'{\rm d}y'\ K(x', y') K'(\Delta x_{i_m\, i_l}-x',
\Delta y_{j_m\, j_l}-y') \cr}
\eqno{\rm(\the\eqnumber)}\global\advance\eqnumber by 1$$
in which~:
$$
\eqalign{
K'(x-x', y-y') &\equiv K(x'-x, y'-y) \cr
\Delta x_{i_m\, i_l} &\equiv x_{i_m}-x_{i_l}\cr
\Delta y_{j_m\, j_l} &\equiv y_{j_m}-y_{j_l}\cr}
\eqno{\rm(\the\eqnumber)}\global\advance\eqnumber by 1$$
Evaluating all the $M^4$ elements of the matrix $A$ is much simplified
by doing this two-dimensional convolution as a multiplication in the
Fourier domain. This calculation is then dominated by the FFT calculation
which requires ${\cal O}( M^2 \log (M) )$ operations. Similarly the
vectors $b$ in \vector) can be evaluated for all locations $(x_i, y_j)$
with a single two-dimensional convolution of $K$ and $T$, again dominated
by the FFT.
If the CCD pixels are assumed to be equally spaced the matrix $A$ for
$\mu = 0$ can be constructed in such a way that it becomes of a special
type known as symmetric block circulant with circulant blocks (BCCB),
for which very fast inversion algorithms exist.
Circulant matrices have the property that every row is identical to the
previous row, but shifted to the right by one element. The shifting is
`wrapped around' so that the first element on each row is equal to the
last element of the previous row. Thus the main diagonal elements are
all equal and on every diagonal parallel to the main diagonal of the
matrix all elements are equal as well. A BCCB matrix is a matrix that
can be partitioned into blocks in such a way that each row of blocks is
repeated by shifting (and wrapping around) by one block in the subsequent
row of blocks and each individual block is circulant.
It can be shown that circulant matrices can be multiplied and inverted
using Fourier transforms, and by extension BCCB matrices
can be multiplied and inverted using two-dimensional Fourier
transforms. The detailed steps of the algorithm are worked out in the appendix.
\fignam{\testimgs}{testimgs}
\beginfigure*{1}
\epsfysize=16.8cm
\epsfbox{fig1.ps}
\caption{{\bf Figure \testimgs.}
{The $128\times 128$ pixels image used in testing the algorithm.
Top left panel~: the original
image. Top right panel~: the original convolved with the target PSF~:
a Gaussian with $\Delta = 1.5$ pixels. Bottom left panel~: the
original convolved with a PSF which is the sum of a Gaussian with
$\Delta = 10$ pixels and an $0.1\%$ contribution from a Gaussian with
$\Delta = 1$ pixel.
Bottom right panel~: the image after SOLA deconvolution of the bottom
left image. In all images the grey-scale is linear. In the
bottom left image noise is added before deconvolution. In the
bottom right image the noise propagated in the deconvolution has
an expectation value of $\sim 0.5$ in arbitrary flux units and the S/N
ratio for the brightest pixel is $\sim 1000$.
}}
\endfigure
\global\advance\fignumber by 1
The restriction on the matrix $N_{lm}$ is that is must also be a
symmetric BCCB matrix for the fast inversion algorithm to
work. It is evident that fully optimal results can only be obtained if
the full $N^2 \times N^2$ covariance matrix of the errors is used.
However, the error correlation function for the pixels is expected to
behave similarly to the point spread function in the sense that it is large
(in absolute value) for small pixel separations and small for large pixel
separations, independently of where on the CCD the pixel is located. It
is therefore likely that the error covariance matrix will already be
BCCB or be very nearly so. Since its role in the minimization of \minimize)
is to regularize the inversion it is in practice not essential that the
exact variance-covariance matrix be used. Experience in using SOLA in
other fields has shown that the results of linear inversions are robust
to inaccuracies in the error matrix, as long as those are not orders of
magnitude large~:
if for instance substantial amounts of data (fluxes in pixels) are to
be given small weight in the resulting linear combination, because of
large errors associated with them, this can give rise to large departures
from BCCB behaviour of the error covariance matrix. This would then cause
problems for the fast version of the SOLA method presented here.
Thus if $N_{lm}$ is not circulant it should in
most cases be sufficient to use a BCCB matrix that is close to the
original~: one could think of using a modified matrix $\overline{N}_{lm}$ on
the diagonals of which are the average values over those diagonals of
the true $N_{lm}$.
Of course once the coefficients have been determined, when calculating the
propagated errors one should use equation \fullerr) with the proper
variance-covariance matrix $N_{lm}$.
As is shown in the appendix the matrix corresponding to the collection
of all vectors of coefficients $c$, which results from the multiplication
of $A$ with the matrix corresponding to the collection of all vectors
$b$ (one for every pixel), is also a BCCB matrix.
The process of combining these coefficients $c_{l}$ with the recorded
fluxes on the CCD to form the improved image is~:
$$
R_{ij} = \sum\limits_{k}\sum\limits_{l} C_{kl} D_{i+k-1\, j+l-1}
\eqno{\rm(\the\eqnumber)}\global\advance\eqnumber by 1
$$
Since the matrix $C$ is a BCCB matrix and therefore its transpose $C^T$ is
as well, the following holds~:
\eqnam\imaconv{imaconv}
$$
\eqalign{
R_{ij} &= \sum\limits_{k,\, l} C_{kl} D_{i+k-1\, j+l-1}
= \sum\limits_{k,\, l} C^T_{lk} D_{i+k-1\, j+l-1}\cr
&= \sum\limits_{k,\, l} C^T_{2-k\, 2-l} D_{i+k-1\, j+l-1}\cr
&= \sum\limits_{k',\, l} C^T_{k'\, l'} D_{i+1-k'\, j+1-l'}\cr
}\eqno{\rm(\the\eqnumber)}\global\advance\eqnumber by 1
$$
From the final equality in \imaconv) it is clear that the process of combining
the matrix of coefficients with the image is a convolution, and
hence can also be done using FFTs.
From the above it is clear that limiting the algorithm to the case of
a PSF that is constant over the CCD implies a profound reduction of the
computing time. If the PSFs $K$ satisfy the condition \psfcon) the
vectors $b$ collected together for all $(x_0, y_0) = (x_i, y_j)$
form a BCCB matrix, and therefore the matrix inversion of $A$ {\bf and}
its subsequent multiplication with {\bf all} vectors $b$, shown in
equation \matinv), can be done in ${\cal O}( M^2 \log (M) )$ operations.
The entire deconvolved image is thus produced in ${\cal O}( M^2 \log (M) )$
operations. This acceleration of the algorithm over the version described
in the previous section is so substantial that even when
the PSF is not constant over the CCD it is worthwhile subdividing the
image into subsections in which the PSF can be closely approximated by
a single function $K$. The error introduced in this way can be estimated
in a way similar to what is done in the application of SOLA to the
reverberation mapping of AGN (Pijpers \& Wanders, 1994), and should
generally be much smaller than the propagated error from equation \fullerr).
If such a subdivision is undesirable, there is the possibility of
reverting to the more general algorithm of section 2.1. For a single
peaked PSF the matrix $A$ should have a banded structure to which
fast sparse matrix solvers can be applied. In this case one could use
the inverse of the matrix $A$ for an approximated PSF that is
translationally invariant as a pre-conditioner to speed up the matrix
inversion for the case of the true PSF.
\section{application to a test image}
\subsection{constructing a narrow PSF}
\fignam{\diffbw}{diffbw}
\global\advance\fignumber by 1
\beginfigure{2}
\epsfysize=8.5cm
\epsfbox{fig2.ps}
\caption{{\bf Figure \diffbw.}
{The difference between the image that is SOLA deconvolved
and the original image convolved to the target PSF.
The gray scale is adjusted so that the full scale is $0.01\times$ the scale
in the right-hand side images of figure \testimgs{}, which corresponds
to $10\sigma$ of the noise in the deconvolved image.
}}
\endfigure
\fignam{\AvTarker}{AvTarker}
\global\advance\fignumber by 1
\beginfigure{3}
\epsfysize=5.5cm
\epsfbox{fig3.ps}
\caption{{\bf Figure \AvTarker.}
{A slice through the peak of the target PSF specified
in the SOLA algorithm (solid line) and the averaging kernel ${\cal K}$
(dashed) constructed from the linear combination of the pixel PSFs.
}}
\endfigure
\noindent
In the first instance it is useful to test the algorithm on a
test image for which the result and the errors are known. To this end an
artificial image of a cluster of stars is convolved with two different
PSFs. One PSF is a sum of two Gaussians~;
one with a width $\Delta = 10$ pixels in which $99.9\%$ of the flux is
collected, and a second one with a width $\Delta = 1$ pixels
which collects the other $0.1\%$ of the flux. Poisson distributed noise
is added to every pixel and this `dirty' image serves as the image to
be deconvolved.
The other PSF is a Gaussian with a width $\Delta = 1.5$ pixels which
is also the target chosen for the SOLA algorithm. Thus the deconvolved image
can be compared directly with the image obtained from direct convolution
of the original with the narrow target PSF. The results are shown in
figure \testimgs{}. In order to get an optimal reproduction of the target
PSF, the error weighting parameter $\mu$ is chosen to be equal to $0$.
\fignam{\DelLam}{DelLam}
\beginfigure{4}
\epsfysize=5.5cm
\epsfbox{fig4.ps}
\caption{{\bf Figure \DelLam.}
{The error magnification $\Lambda$ as a function of the width
$\Delta$ of the target PSF specified. The PSF of the blurred image
is as described in the text in all cases, the error weighting is
$\mu = 0$.
}}
\endfigure
\global\advance\fignumber by 1
It is clear that the bottom and top right panels are very similar
and thus the image appears to be recovered quite well.
To illustrate this further the two images can be subtracted.
Figure \diffbw{} shows the SOLA deconvolved image minus the image convolved
with the target PSF, with an adjusted gray scale to bring out the
differences, which in the central portion of the image are all $< 1 \%$.
Although there is no strong evidence for it in this image, the deconvolution
can suffer from edge effects because part of the original image can
`leak away' in the convolution with the broad PSF. When deconvolving,
the region outside the image is assumed to be empty and so a spurious
negative signature is then introduced in the image. The magnitude of such
edge effects must clearly depend both on the image and on the PSF of
the `dirty' image, since they are determined by the information that has been
lost at the edges of the CCD. Of course it is desirable to demonstrate
this method on a more realistic suite of images than just the simple
one used here, which is work currently in progress.
\fignam{\difmag}{difmag}
\beginfigure{5}
\epsfxsize=8.5cm
\epsfbox{fig5.ps}
\caption{{\bf Figure \difmag.}
{The difference on an arbitrary magnitude scale between the magnitude of
the stars in the deconvolved image $m_{\rm decon}$ and the magnitude
$m_{\rm aim}$ of their counterparts in the reference image constructed by
convolving the original with the target PSF.
}}
\endfigure
\global\advance\fignumber by 1
\fignam{\difpos}{difpos}
\beginfigure{6}
\epsfxsize=8.5cm
\epsfbox{fig6.ps}
\caption{{\bf Figure \difpos.}
{The position difference in pixel units between the stars in the deconvolved
image $m_{\rm decon}$ and their counterparts in the image constructed
by convolving the original with the target PSF. Open circles : all
stars with magnitudes between $15.9$ and $16.9$, crosses : stars with
magnitudes between $16.9$ and $17.9$, open squares : stars with magnitudes
between $17.9$ and $18.9$, open triangles : all stars with magnitude
greater than 18.9.
}}
\endfigure
\global\advance\fignumber by 1
The averaging kernel that is constructed cannot in general match perfectly
the target form, even in the absence of errors. In general any function
can be completely reconstructed only out of a {\it complete set} of base
functions. Since function space is infinite dimensional this would
require an infinite number of base functions. In this test image there
are no more available than the $128^2$ PSFs corresponding to each of the
pixels and so there can never be a perfect matching of ${\cal K}$
with ${\cal T}$. In figure \AvTarker{} a section through the maximum of
both ${\cal T}$ and ${\cal K}$ is shown. It is clear that at the $10\ {\rm ppm}$
level, the constructed averaging kernel starts getting wider than the target.
If the ratio of the widths of the target form and actual PSF is
even smaller than for this image, alternating negative and positive side
lobes can show up in the averaging kernel which cause ringing. The
amplitude of the sidelobes, and the width $\Delta$ below which ringing
starts occurring, will in general depend on the weighting of the errors
$\mu$, as has been demonstrated from the application of SOLA to
helioseismology (Pijpers \& Thompson, 1994).
As it stands the SOLA algorithm does not impose positivity on the image.
One could attempt to use a positivity constraint to extrapolate the image
beyond the recorded edges in such a way that it eradicates any negative
fluxes in the image, which would in principle also remove associated positive
artifacts around the edges. However, in the presence of errors this might be
somewhat hazardous. Furthermore, in the presence of errors any
edge effects might well disappear into the noise.
If one assumes that
the covariance of errors between pixels is equal to $0$ and the flux
error in each pixel is equal to $\sigma^2$, or $N \equiv \sigma^2 I$, then
it is particularly simple to calculate the flux error for each pixel
in the deconvolved image from equation \simperr) since it is
$$
\sigma_R^2\ =\ \sigma^2 \sum c_l^2\ \equiv\ \Lambda^2 \sigma^2
\eqno{\rm(\the\eqnumber)}\global\advance\eqnumber by 1
$$
This factor $\Lambda$ is usually referred to as the error magnification
and is equal for all pixels in the reconstructed image. In general $\Lambda$
increases as the ratio of $\Delta_{\rm target}/\Delta_{\rm PSF}$
decreases. For the deconvolved image of figure \testimgs{} the error
magnification is $\sim 321$.
The magnitude of the error magnification for this simple example
illustrates that the true limitation of deconvolving images may
in practice not lie in the sampling theorem, but instead in the
S/N of the recorded image. For example for a point source the
peak flux in its central pixel will increase in the deconvolution by a factor
which is roughly ${\rm FWHM}_{\rm PSF}/{\rm FWHM}_{\rm target}$. The noise
will increase by a factor $\Lambda$ and so the signal-to-noise
ratio for point sources will scale roughly as~:
$$
\left({S\over N}\right)_{\rm decon} \approx {1\over\Lambda}
\left({ {\rm FWHM}_{\rm PSF} \over {\rm FWHM}_{\rm target}}\right)
\left({S\over N}\right)_{\rm dirty}
\eqno{\rm(\the\eqnumber)}\global\advance\eqnumber by 1$$
Thus in the example shown the signal-to-noise ratio for point sources
is degraded by a factor of roughly $\sim 48$ between the dirty image
and the deconvolved image. This clearly requires a very high a
signal-to-noise ratio in the dirty image, which means that in practice
as dramatic a resolution enhancement as attempted here will not
usually be possible. A more modest resolution enhancement of
around a factor of 2 should in most cases be possible however,
as can be deduced from figure \DelLam{}.
In order to show the relation between resolution and error
magnification in figure \DelLam{} is shown the value of the error
magnification as a function of the width $\Delta$ specified for the
target function.
If only the broad component had been present the error magnification
would have been unity for a target $\Delta = 10$. Effectively
because of the narrow component which captures a mere $0.1\%$ of the flux
the image can be deconvolved to a PSF with $\Delta = 8$ pixels
without significant penalty in the magnification of the errors.
The CPU time used to construct the matrix $A$, all the vectors $b$,
to invert $A$ and multiply with all $b$ and finally to combine the
coefficients with the image takes $\sim 0.5\ {\rm min}$ on an SGI
workstation for this $128\times 128$ image.
\subsection{photometric and astrometric accuracy}
Since there is no placement of flux or point sources, the algorithm should
automatically be astrometrically accurate. Photometric accuracy is ensured
by explicitly constraining the linear coefficients to sum to unity, i.e.
imposing constraint \unitav). In order to demonstrate both these properties
for this test image a standard photometric package DAOPHOT was used to
do aperture photometry on the deconvolved image, and on the image obtained by
convolving the original with the target PSF. The deconvolved image of
course has noise propagated from the dirty image. The other image is
kept noise-free to properly serve as a reference. The errors are calculated
by DAOPHOT and are consistent with what is expected from the noise in
the deconvolved image. In figure \difmag{} is shown the difference
between the magnitudes of the stars in the two images as a function of
the stellar magnitudes in the reference image. The difference is clearly
consistent with zero over the entire range of 5 magnitudes, and does not
show any trend.
DAOPHOT calculates the error bar assuming that the error in different
pixels is uncorrelated. For the SOLA deconvolved image this is not
the case. In the deconvolved image the error correlation function
falls below $0.01$ in absolute value only at inter-pixel distances
larger than $\sim 12$ for this test case. Furthermore, because fluxes
are combined with positive and negative coefficients, the error
is not distributed as for a Poisson process. If this is taken
into account properly the error bars in figure \difmag{} should be
decreased and are then compatible with the actual scatter of the points.
\fignam{\sectst}{sectst}
\beginfigure{7}
\epsfxsize=6.8cm
\epsfbox{fig7.ps}
\caption{{\bf Figure \sectst.}
{Images of the galaxy UGC 5041. The top image is convolved with the same
broad PSF used on the image at the bottom left in figure \testimgs{}.
The middle image is the deconvolved image with a PSF with a FWHM of $2.5$. The
bottom image is the difference between the original image convolved
with the target PSF, and the deconvolved image, no noise has been added.
The gray scale of the bottom panel extends between $\pm 0.1\%$
of the gray scale of the middle image.
}}
\endfigure
\global\advance\fignumber by 1
\fignam{\omcdet}{omcdet}
\beginfigure*{8}
\epsfxsize=18.0cm
\epsfbox{fig8.ps}
\caption{{\bf Figure \omcdet.}
{Image of a detail of the Orion Molecular Cloud obtained using adaptive
optics in IR lines of shocked molecular hydrogen, north is at the top
of the image, east is to the left. The image size is
$256\times 256$ pixels at $50 {\rm mas/pixel}$. The top left image is
as obtained with the Adonis instrument on ESO's $3.6{\rm m}$ telescope.
The PSF for this image is shown on a $4\times$ enlarged scale as the
$32\times 32$ pixel image in the top left-hand corner of the top
right-hand panel. The bottom left- and
right-hand panels are the deconvolved image using different gray-scales.
The two bottom images in the top right-hand panel show $32\times 32$
pixels images of the PSF for the deconvolved image, using the same
gray scales as the corresponding images, and the same spatial scale
as the PSF for the original. For all images
the dynamic range between lightest and darkest colour is a factor of
$\sim 8$ in flux level.
}}
\endfigure
\global\advance\fignumber by 1
DAOPHOT also determines the positions of point sources in the image
and therefore those positions can be used to determine astrometric
accuracy. In figure \difpos{} are shown the difference in units of a
pixel between the DAOPHOT determined positions in the two images.
Here the stars have been grouped into 4 magnitude bins, each
bin 1 magnitude in range, and starting from the brightest star with
magnitude $15.9$. The right-hand panel is a blow-up of the central
portion of the left-hand panel. Figure 4 shows that there is a trend in
that the fainter stars show a greater scatter in position, the largest
position difference being of the order of $\sim 0.2$ pixels. In the
right-hand panel it can be seen that for stars brighter than magnitude $19$
the difference in positions is smaller than 0.03 pixels. These
uncertainties are entirely consistent with the accuracy with which
DAOPHOT can determine stellar positions.
From figures \difmag{} and \difpos{} it is clear that if any errors
in photometry or in position are introduced by the deconvolution
process, they are much smaller than the errors due to the random noise.
\section{application to astronomical images}
\subsection{UGC 5041}
To give a somewhat more interesting example, a high resolution HST image
of a galaxy is blurred and then deconvolved to demonstrate that the
method also works on an image which contains a combination of extended
structure and point sources.
The galaxy UGC~5041 is an Sc type galaxy at a redshift of 0.027 (Haynes
{et al.}., 1997). It has been part of various surveys for use in studies
of clustering and in establishing distance scales for the Tully-Fisher
distance method. In figure \sectst{} is shown a $512 \times 512$ image
of this galaxy obtained in March 1997 using the WFPC2 (WF3) instrument
on board the Hubble Space Telescope (HST) with the F814W filter which
corresponds to the I-band. The resolution of the original image has
a FWHM of $1.4$ pixels and is convolved with the same broad PSF used
in the bottom left panel of figure \testimgs{}. It is then deconvolved
to the same resolution as the right-hand images of figure \testimgs{}
and the result is shown in the middle panel of figure \sectst{}.
The difference image between this deconvolved image and a reference
is shown in the bottom panel of figure \sectst{}, where the gray scale
is enhanced to demonstrate that in the absence of noise the differences
between the deconvolved image and the reference image are less than
$0.1\%$ of the peak flux apart from edge effects.
\subsection{Orion Molecular Cloud}
Observations in IR lines of shocked ${\rm H}_2$ of the SE part of the
Orion Molecular Cloud complex (OMC1) have been performed at the ESO 3.6m
telescope taking advantage of the high spatial resolution given by
adaptive optics (Adonis at 50 marcsec/pixel) combined with the high
spectral resolution given by a Fabry-Perot (R=1000) (Vannier {et al.}, 1998).
The image is deconvolved using a target PSF with FWHM $3$ pixels
and an error weighting $\mu = 0$. The resulting error magnification
factor is $\Lambda = 6.2$. The dynamic range in the deconvolved image
is $\sim 30$ as opposed to $\sim 8$ in the original. For this
reason the deconvolved image is shown twice~: bottom left in figure \omcdet{} is
shown a linear gray scale extending from an estimated noise level to
$8$ times that, bottom right is shown a linear gray scale extending
from the peak level in the deconvolved image which is $\sim 30$ times
the estimated noise level, to $1/8$ of that. The PSF of the deconvolved
image, i.e. the constructed averaging kernel ${\cal K}$, is shown
using the same two gray scales in the bottom part of the
top right-hand panel of figure \omcdet{}.
Although some of the `graininess' in the bottom panels must be due to
the increased noise compared to the top left-hand image, fine structure
can clearly be seen in the bottom panels of figure \omcdet{}. There
is also some evidence of edge effects at the top and right of the image.
From the image of the PSF of the deconvolved image it is also clear that,
as expected, the Gaussian target is not reproduced perfectly over the
entire dynamic range of the deconvolved image. Comparing the PSF images
with the same dynamic range from the peak down (top left and bottom right
in the top right-hand panel of figure 8) it is clear that
the PSF is indeed much narrower for the deconvolved image.
\section{Conclusions}
In this paper the SOLA inversion method, well known in helioseismology, is
applied to the reconstruction of astronomical images. It is demonstrated
how a linear transformation is constructed between any image recorded
with a known PSF and its deconvolved counterpart with a different
(narrower) PSF. The method itself uses {\bf only} the PSF and no assumptions
are made concerning what is contained within the image(s) to be deconvolved.
It is furthermore shown that in the case of
translationally invariant PSFs, a fast algorithm, using ${\cal O} (N\log N)$
operations where $N$ is the total number of pixels in the image, can be
constructed, which allows deconvolution of even $1024 \times 1024$ images
within half an hour on medium-sized workstations.
\section*{Acknowledgments}
Steve Holland is thanked for a number of helpful discussions.
D. Rouan, J.-L. Lemaire, D. Field, and L. Vannier are thanked for making
available their data prior to publication.
The observations of UGC~5041 were made with the NASA/ESA
Hubble Space Telescope, obtained from the data archive at the Space
Telescope Science Institute. STScI is operated by the Association of
Universities for Research in Astronomy, Inc. under NASA contract
NAS 5-26555.
The Theoretical Astrophysics Center is a collaboration
between Copenhagen University and Aarhus University and is funded by
Danmarks Grundforskningsfonden.
\section*{References}
\ref
Haynes M.P, Giovanelli R., Herter T., Vogt N.P., Freudling W., Maia M.A.G.,
Salzer J.J., Wegner G., 1997,
{Astron. J.}
113, 1197
\ref
H\"ogbom J.A., 1974,
{A\&{}A Supp.}
15, 417
\ref
Lucy L.B., 1974,
{Astron. J.}
79, 745
\ref
Lucy L.B., 1992,
{Astron. J.}
104, 1260
\ref
Lucy L.B., 1994,
{A\&{}A}
289, 983
\ref
Magain P., Courbin F., Sohy S., 1998,
{ApJ}
494, 472
\ref
Narayan R., Nityananda R., 1986,
{Ann. Rev. A\&{}A}
24, 127
\ref
Pijpers F.P., Thompson M.J., 1992,
{A\&{}A}
262, L33
\ref
Pijpers F.P., Thompson M.J., 1994,
{A\&{}A}
281, 231
\ref
Pijpers F.P., Wanders I., 1994,
{MNRAS}
271, 183
\ref
Press W.H., Teukolsky S.A., Vetterling W.T., Flannery B.P., 1992,
Numerical Recipes, the art of scientific computing 2$^{nd}$ Ed.,
CUP, Cambridge, 70
\ref
Richardson W.H., 1972,
{J. Opt. Soc. Am.}
62, 55
\ref
Schwarz U.J., 1978,
{A\&{}A}
65, 345
\ref
Vannier L., Lemaire J.-L., Field D., Rouan D., Pijpers F.P., Pineau des
For\^e{}ts G., Gerin M., Falgarone E., 1998,
{ESO/OSA Meeting on Astronomy with Adaptive Optics. Present results and
future programs, Sonthofen Germany 7-11 sept 98},
in press
\ref
Wakker B.P., Schwarz U.J., 1988,
{A\&{}A}
200, 312
|
\section{Introduction}
One of the most remarkable and
intriguing discoveries in hadron physics in the last years was the
observation of hard interactions (namely, partonic activity) in
diffractive events by the UA8 Collaboration \cite{ua8} at the CERN
Collider. This discovery was inspired by a seminal paper by Ingelman
and Schlein \cite{ingelman} in which such a possibility was foreseen as
a consequence of the pomeron having an internal structure and a
quark/gluon content.
Experiments performed at HERA by ZEUS \cite{zeus_diff} and H1
\cite{h1_diff} collaborations provided further elements to this view
through the obtainment of deep inelastic electron-proton scattering
(DIS) events accompanied by large rapidity gaps adjacent to the proton
beam direction. The presence of rapidity gaps in such events is
interpreted as indicative that the internal structure of a colourless
object carrying the vaccum quantum numbers, namely the pomeron, is
being probed \cite{zeus_diff,h1_diff}.
Further experimental evidences of hard diffraction have been reported by
CDF and D0 collaborations in terms of diffractive production of W's and
dijets \cite{cdf,d0}.
Taking together, these evidences seem to indicate that the
Ingelman-Schlein (IS) picture \cite{ingelman} is right at least {\em
qualitatively}. However, a serious problem arises when one checks this
model quantitatively. For instance, the model predictions
systematically overestimate the diffractive production rates of jets
and W's \cite{alvero}. There are reasons to believe that the
discrepancy between predictions and data comes from the so-called {\em
pomeron flux factor} \cite{dino}. In fact, a recent analysis
\cite{Covolan} has shown that the pomeron structure function extracted
from HERA data by using the IS approach is strongly dependent on which
expression is used for the flux factor.
Recently, new data from H1 Collaboration~\cite{H1} on final hadronic
states in diffractive deep inelastic process (DDIS) of the type
\begin{equation}
ep \rightarrow e'XY, \label{I}
\end{equation}
where $X$ and $Y$ are hadronic systems, have been presented. The system
$X$ and $Y$ in (\ref{I}) are separated by the largest rapidity gap. $Y$
is the closest to the proton beem ($M_Y < 1.6$ GeV) and squared
momentum transfer at the $pY$ vertex, $t$, is limited to $|t| < 1 \
\mbox{\rm GeV}^2$, while $\langle Q^2 \rangle$ is $21 \div 27 \ \mbox{\rm
GeV}^2$~\cite{H1}. Both invariant masses,
$M_X$ and $M_Y$, are small compared to $W$, the centre-of-mass energy
of the $\gamma^* p$ system.
In particular, charged hadron multiplicity have been studied
in~\cite{H1} as a function of $M_X$ in the centre--of--mass system of
$X$. The data obtained have been compared with the calculations of
JETSET $e^+e^-$ (which is known to reproduce well the $e^+e^-$ data)
and with the data on hadron multiplicities in deep inelastic scattering
(DIS)~\cite{EMC}.
The interesting observations presented in~\cite{H1} are the following:
i) for $M_X > 10$ GeV $\langle n \rangle^{DDIS}(M_X, Q^2)$ is larger
than charged hadron multiplicity in DIS,
$\langle n \rangle^{DIS}(W, Q^2)$, at comparable values
of $W = \langle M_X \rangle$; ii) $\langle n \rangle^{DDIS}(M_X, Q^2)$
is also larger than charged hadron multiplicity in $e^+e^-$ annihilation,
$\langle n \rangle^{e^+e^-}(s)$, taken at $\sqrt{s} = \langle M_X \rangle$.
In the present paper we demonstrate that perturbative QCD is able to
explain (qualitatively and quantitavely) the more rapid growth of
hadron multiplicity in DDIS. So, no mechanisms besides gluon/quark jet
emission with subsequent jet fragmentation into hadrons are needed.
It is shown that the results on hadron multiplicity are very sensitive
to the quark-gluon structure of the pomeron. An advantage of the method developed here is that it does not depend on the pomeron flux factor.
The paper is organized as follows. In Section II, we present the QCD
formalism for the description of final hadron states in ordinary DIS.
In Section III, we extend this formalism to diffractive DIS and
apply it to describe the H1 data. Our main conclusions are summarized
in Section IV.
\section{Hadron Multiplicities in Hard Processes in
Perturbative QCD}
For the time being, we cannot describe a transition
of quark and gluons into hadrons in the framework of QCD. Nevertheless,
perturbative QCD enables one to calculate an energy dependence of
characteristics of final hadrons produced in hard process
($e^+e^-$ annihilation into hadrons, DIS, Drell--Yan process etc.).
Mean hadron multiplicity, $\langle n \rangle$, is one of the main
features of final hadronic states. Hadron multiplicity in $e^+e^-$
annihilation has
been studied in a number of papers. The result looks like (see, for
instance, \cite{Webber})
\begin{equation}
\langle n \rangle^{e^+e^-}(s) = a (\ln s)^b \exp \left( c \sqrt{\ln s}
\right), \label{2.1}
\end{equation}
where $\sqrt{s}$ is a total c.m.s. energy of colliding leptons. As one
can see, $\langle n \rangle^{e^+e^-}(s)$ rises more rapidly than $\ln s$,
although more slowly than any power of $s$. Expression~(\ref{2.1})
describes well the available data.
Hadron multiplicity in DIS was calculated first in the framework of
perturbative QCD in Refs.~\cite{Kisselev/Bassetto} (see
also~\cite{Kisselev1}). It was shown that average multiplicity in
lepton scattering off parton
\begin{equation}
eq(g) \rightarrow e'X, \label{II}
\end{equation}
$\langle n \rangle^{DIS}_{q/g}(W,Q^2)$, is related to $\langle n
\rangle^{e^+e^-}(s)$, the average
multiplicity in $e^+e^-$ annihilation, taken at $\sqrt{s} = W$ (up to
small NLO corrections which descrease in $W$ and
$Q^2$)~\cite{Kisselev/Bassetto}.
In a case when quark distribution dominates (say, at $x \simeq 1$), the
relation between $\langle n \rangle^{DIS}$ and $\langle n
\rangle^{e^+e^-}$ looks like
(up to small NLO corrections which decrease in $W$ and
$Q^2$)~\cite{Kisselev/Bassetto}
\begin{equation}
\hat{\langle n \rangle}^{DIS}_q(W, Q^2) = \langle n \rangle^{e^+e^-}(W).
\label{2.3}
\end{equation}
If we consider small $x$, we have to account for the gluon distribution
and the result is of the following form~\cite{Kisselev2}
\begin{eqnarray}
\hat{\langle n \rangle}^{DIS}_g(W, Q^2) = \langle n \rangle^{e^+e^-}(W)
\left[\,1 +
\frac{C_A}{2C_F} \varepsilon \left( 1 - \frac{3}{2} \,\varepsilon
\right) \right], \label{2.4}
\end{eqnarray}
where $C_A = N_c$, $C_F = (N_c^2-1)/2N_c$ and $N_c$ is a number of
colours.
The quantity $\varepsilon$ is defined via gluon distribution
(see~\cite{Kisselev2} for details)
\begin{equation}
\varepsilon = \sqrt{\frac{\alpha_s(W^2)}{2 \pi
C_A}}\frac{\partial}{\partial \xi} \ln D^g(\xi, x), \label{2.5}
\end{equation}
$\xi$ being the QCD--evolution parameter $\xi(W^2)=\int^{W^2}
(dk^2/k^2) (\alpha_s(k^2)/2\pi)$.
Let us notice that $\varepsilon$~(\ref{2.5}) does not depend on a type
of a target, and $\varepsilon$ is completely defined by the evolution
of $D^g$ in $\xi$.
Starting from the well--known expression for $D^g_g$ at small
$x$~\cite{Dokshitzer},
\begin{equation}
D^g_g(\xi, x) = \frac{1}{\ln (1/x)} v \mbox{\rm I}_1(2v) \exp (-d \xi),
\label{2.7}
\end{equation}
where $v=\sqrt{4N_c \xi \ln (1/x)}$, $d=(1/6) \left( 11N_c + 2N_f/N_c^2
\right)$, $N_f$ is a number of flavours, we get
\begin{equation}
\varepsilon = \varepsilon (W^2, x) = \sqrt{\frac{\alpha_s(W^2)}{\pi}}
\left[\sqrt{\frac{\ln (1/x)}{\xi (W^2)}} - \frac{d}{\sqrt{2N_c}}
\right]. \label{2.8}
\end{equation}
At $\ln (1/x) \gg 1$ (that is omitting the second term in RHS of
~(\ref{2.8}))
we come to the asymptotical expression for $\varepsilon$ from
Ref.~\cite{Kisselev2}.
Using (\ref{2.3}) and (\ref{2.4}), we get in a general case
\begin{eqnarray}
\hat{\langle n \rangle}^{DIS} (W, Q^2) = \langle n \rangle^{e^+e^-}(W)
\left[ \Delta + \left( 1 +
\frac{C_A}{2C_F} \varepsilon (W^2, x) \left( 1 - \frac{3}{2}
\,\varepsilon (W^2, x) \right) \right) (1 - \Delta) \right],
\label{2.9}
\end{eqnarray}
where $\Delta/(1 - \Delta)$ defines the quark/gluon ratio inside the
nucleon.
As was shown in \cite{Kisselev2}, $\varepsilon$ (\ref{2.8}) is a
leading correction to $\hat{\langle n \rangle}^{DIS}$ which rises with the
decreasing of $x$.
All mentioned above is related to the multiple production of the
hadrons in DIS of lepton off the parton~(\ref{II}). It is a subprocess
of the process of lepton deep inelastic scattering off the nucleon
\begin{equation}
ep \rightarrow e'X. \label{III}
\end{equation}
According to Refs.~\cite{Kisselev1,Petrov,Kisselev3}, the corresponding
formula for mean hadron multiplicity in DIS looks like
\begin{equation}
\langle n \rangle^{DIS}(W, Q^2) = \hat{\langle n \rangle}^{DIS}(W_{eff},
Q^2) + n_0(M_0^2).
\label{2.11}
\end{equation}
The quantitiy $\hat{\langle n \rangle}^{DIS}$ has been defined above (see
(\ref{2.9})). It depends on the effective energy, $W_{eff}$, available
for hadron production in the parton subprocess~(\ref{II}):
\begin{equation}
W_{eff}^2 = k W^2. \label{2.13}
\end{equation}
The average multiplicity of nucleon remnant fragments, $n_0$, is a
function of its invariant mass $M_0^2 = (1 - z)(Q_0^2/z + M^2)$, where
$M$ is a nucleon mass.
The efficiency factor $k$ in (\ref{2.13}) was estimated in
\cite{Kisselev1} to be $\langle k \rangle \simeq 0.15 \div 0.20$ at present
energies. That is why $\langle n \rangle^{DIS}(W, Q^2)$ is less than
$\langle n \rangle^{e^+e^-}(W)$ and
the growth of $\langle n \rangle^{DIS}(W, Q^2)$ in $W$ at fixed $Q^2$
is delayed in
comparison with that of $\langle n \rangle^{e^+e^-}(W)$.
\section{Average Hadron Multiplicity in Diffractive DIS}
Hadronic system $X$ in diffractive DIS (\ref{I}) is produced as a
result of the virtual photon--pomeron interaction:
\begin{equation}
e{\rm I\! P} \rightarrow e'X. \label{IV}
\end{equation}
The kinematical variables usually used to describe DDIS (in addition to
$W$ and $Q^2$) are
\begin{equation}
x_{{\rm I\! P}} \simeq \frac{M_X^2 + Q^2}{W^2 + Q^2} \quad \mbox{\rm
and} \quad
\beta \simeq \frac{Q^2}{M_X^2 + Q^2}. \label{3.02}
\end{equation}
We start from the fact that pomeron has quark--gluon structure. It
means that hadron production in process~(\ref{IV}) is similar to that
in parton subprocess of DIS (\ref{II}).
Recently a factorization theorem for DDIS has been
proven~\cite{Collins}. Structure functions of DDIS coincide with
structure functions of DIS and have the same dependence on a
factorization scale. As a result, distribution functions of quark and
gluons inside the pomeron, $D_{{\rm I\! P}}^{q(g)}$, must obey standard
DGLAP evolution equations at high $Q^2$~\cite{Altarelli}.
So, we can conclude that a hadron multiplicity of system $X$ in DDIS,
$\langle n \rangle^{DDIS}$, is given by expression (\ref{2.9}), in which $W$
is replaced by $M_X$, while $x$ is replaced by $\beta$. Namely, we get
\begin{eqnarray}
\langle n \rangle^{DDIS}(M_X, Q^2) &=& \langle n \rangle^{e^+e^-}(M_X) \times
\nonumber \\
&\times& \left[\,\Delta_{{\rm I\! P}} +
\left(1 + \frac{C_A}{2C_F} \varepsilon (M_X^2, \beta) \left( 1 -
\frac{3}{2} \,\varepsilon (M_X^2, \beta) \right) \right) (1 -
\Delta_{{\rm I\! P}}) \right]. \label{3.1}
\end{eqnarray}
Here $\varepsilon (M_X^2, \beta)$ is given by formula (\ref{2.8}) and
$\Delta_{{\rm I\! P}}/(1 - \Delta_{{\rm I\! P}})$ defines quark/gluon
ratio inside the pomeron.
As both $D^q_{{\rm I\! P}}$ and $D^g_{{\rm I\! P}}$ obey DGLAP
equations, the quantity $\varepsilon (M_X^2, \beta)$ and,
consequently, hadronic multiplicity in DDIS are sensitive to the
quark--gluon structure of the pomeron at starting scale $Q_0$.
As folllows from (\ref{2.11}) and (\ref{3.1}),
\begin{equation}
\langle n \rangle^{DDIS}(M_X, Q^2) > \langle n \rangle^{DIS}(M_X, Q^2)
\label{3.4}
\end{equation}
because $\langle n \rangle^{DDIS}(M_X, Q^2)$ is defined by $\langle n
\rangle^{e^+e^-}(M_X)$, while $\langle n \rangle^{DIS}(M_X, Q^2)$ is
expressed via $\langle n \rangle^{e^+e^-}(M_X^{eff})$, the quantity
$M_X^{eff}$ being much smaller than $M_X$ by factor $k$ (see
(\ref{2.13})).
For $M_X < 29$ GeV we have $M_X^{eff} < 5 \div 6$ GeV. It is known
that in the region $W < 5 \div 6$ GeV function $\langle n
\rangle^{e^+e^-}(W)$ rises
logarithmically ($\sim \ln W$) that is slower than (\ref{2.1}). This
results in more rapid growth of $\langle n \rangle^{DDIS}(M_X, Q^2)$
in $M_X$ in comparison with $\langle n \rangle^{DIS}(M_X, Q^2)$.
We conclude from formula (\ref{3.1}) that $\langle n
\rangle^{DDIS}(M_X, Q^2)$
should exceed $\langle n \rangle^{e^+e^-}(M_X)$. Moreover, the ratio
$\langle n \rangle^{DDIS}(M_X, Q^2)/\langle n \rangle^{e^+e^-}(M_X)$
has to grow in $M_X$ at fixed $Q^2$
(that corresponds to the rise in $1/\beta$ at fixed $Q^2$). All said
above is in good qualitative agreement with the data from H1
Collaboration~\cite{H1}.
In Fig.~1 we show the result of the fits of the H1 data by using our
formula~(\ref{3.1}) (solid curve). In order to obtain this result, we
proceeded as follows. For the quantity $\Delta_{{\rm I\! P}}$ which
enters expression~(\ref{3.1}) we used
\begin{equation}
\Delta_{{\rm I\! P}}(\beta, Q^2) = \frac{D^{q}_{{\rm I\! P}}(\beta,
Q^2)}{{D^{g}_{{\rm I\! P}}(\beta, Q^2)}+{D^{q}_{{\rm I\! P}}(\beta,
Q^2)}}, \label{3.5}
\end{equation}
where ${D^{g}_{{\rm I\! P}}}$ and ${D^{q}_{{\rm I\! P}}}$ are
respectively the gluon and the singlet quark distributions inside the
pomeron that, as mentioned above, obey DGLAP evolution equations
\cite{Altarelli}. For the distributions at the initial scale
$Q_0^2=4\ GeV^2$ the following forms were employed:
\begin{eqnarray}
D^{q}_{{\rm I\! P}}(\beta, Q_0^2) &=& a_1\ \beta (1-\beta), \nonumber \\
D^{g}_{{\rm I\! P}}(\beta, Q_0^2) &=& b_1\ {\beta}^{b_2} (1-\beta)^{b_3}.
\label{gluon}
\end{eqnarray}
Thus, for the quark distribution we fixed an initial hard profile
leaving free the normalization parameter, while for the the gluon
distribution we have left all parameters free without imposing any sum
rule.
In order to perform the fit, other elements are needed.
In Eq.~(\ref{3.1}), for $\langle n \rangle^{e^+e^-}(M_X)$ we have used
the parametrization
\begin{equation}
\langle n \rangle^{e^+e^-}(M_X)= 2.392 + 0.024\ln(\frac{M^2_X}{M^2_0})
+ 0.913\ln^2(\frac{M^2_X}{M^2_0})
\end{equation}
with $M_0=1\ GeV$, taken from \cite{mirian}. Besides that, the
experimental data shown in Fig.1 are given in terms of average values
of $M_X, \beta,$ and $Q^2$ that do not obey strictly the kinematical
relation (\ref{3.02}). In order to be faithful to data, we employed in
our fit the parametrization
\begin{equation}
\langle \beta \rangle = \frac{a}{(1 + b\ \langle M_X \rangle)^{c}},
\end{equation}
with $a = 1.63, b = 0.165\ GeV^{-1}$, and $c=2.202$.
The dotted (dashed) line in Fig.~1 corresponds to a pure quark (gluon)
content of the pomeron.
The solid line, which gives the fit of the H1 data, is obtained
with parameters $a_1=2.400,\ b_1=3.600,\ b_2=5.279,$ and $b_3=0.204$
in Eqs.~(\ref{gluon}), which are evolved to the respective $Q^2$-value
corresponding to each experimental point.
The initial distributions and their evolution with $Q^2$ are shown in
Fig.~2. As can be seen, a super-hard profile (peaked at
$\beta \sim 1$) was obtained for the gluon distribution at the initial
scale, in qualitative agreeement with H1 analysis \cite{h1_novo}.
The $Q^2$--evolution of the normalized fractions of the pomeron
momentum carried by quarks and gluons is presented in Fig.~3. It is
shown that quarks are slightly predominant over gluons at the initial
scale, but that this relation reverses as $Q^2$ evolves.
In this case, our results do not follow those obtained by H1
Collaboration since their analysis \cite{H1} favour a fit in which
predominate gluons carrying $\geq 80\%$ of pomeron momentum at the
initial scale. It must be said, however, that the results of Fig.3 are
consistent with limits established in other experiments (see, for
instance, \cite{cdf}).
\section{Conclusion}
We have presented in this paper a description of the hadron
multiplicity obtained in diffractive DIS and recently reported by the
H1 Collaboration \cite{H1}. This description was derived from a
formalism previously developed for ordinary DIS within the framework
of the perturbative QCD.
The formula obtained enables us to explain the observed excess of
hadron multiplicity in diffractive DIS, $\langle n
\rangle^{DDIS}(M_X, Q^2)$,
relative to those in DIS and $e^+e^-$ annihilation taken at $W = \langle
M_X \rangle$ and $\sqrt{s} =
\langle M_X \rangle$, correspondingly. The more rapid growth of
$\langle n \rangle^{DDIS}(M_X, Q^2)$
is also understood.
It was shown (Fig.1) that neither a pure quark nor a pure gluon
content of the pomeron satisfy the data, but that a mixing of both
components in approximately equal shares is able to provide a good
description of the experimental results.
The pomeron structure function that comes out from the present
analysis consists of a hard distribution for the quark singlet and
a super-hard distribution for gluons at the initial scale of
evolution, in agreement with what has been reported lately in the
literature \cite{h1_novo}.
This result is remarkably good if one considers that it was obtained
from a very small data set (only 7 data points), covering a limited
$\beta$ range ($0.03 < \beta < 0.43$) for an equally restricted
$Q^2$-range ($22 < Q^2 < 27\ GeV^2$).
Concluding, the most important result presented here is the theoretical
framework summarized by Eq.(\ref{3.1}), that has made it possible to
explain anomalous H1 data on hadron multiplicity in DDIS. It also
enabled us to extract information about the pomeron structure function
from such a limited data set, without the usual complications and
ambiguities with flux factor normalization and $x_{{\rm I\! P}}$-dependence.
This conclusion points out to the need of more and more precise DDIS
multiplicity data, taken at extended kinematical ranges. Such a
possibility would improve a lot the analytical capacity of the scheme
presented here.
\section*{Acknowledgements}
We would like to thank the Brazilian governmental agencies CNPq and
FAPESP for financial support.
|
\section{Introduction}
The triangular-lattice Heisenberg antiferromagnet (TLHA) has played a
fundamental role in the understanding of frustrated quantum spin systems. In
particular, the existence of a nonmagnetic ground state in the $S=1/2$
system has been strongly debated in the literature, although there is now a
widespread conviction that it displays the classical 120$^{\circ }$ spiral
order.\cite{orden,Colo1} In the last years, this particular problem was
linked to more general questions concerning spin-liquid states and their
possible connections to high-$T_c$ superconductivity. In this context, the
most studied model has been the square-lattice Heisenberg antiferromagnet
with first- and second-neighbor interactions, the so-called $J_1%
\negthinspace -\negthinspace J_2$ model.\cite{12} The introduction of
frustrating second-neighbor interactions leads, at intermediate values of
the couplings, to the existence of a disordered spin-liquid state which
intervenes between the N\'eel and ''striped '' (ferromagnetic in one
direction, antiferromagnetic in the other) magnetic orders. This situation
is by now fairly well established by a variety of numerical and analytical
methods.\cite{12,letter} On the other hand, an extension of the TLHA
including second-neighbor interactions, the $J_1\negthinspace -\negthinspace %
J_2$ TLHA, has also been
considered.\cite{TLHA12}$^{-}$\cite{Leche1} This extension is a
natural development after the thorough investigation of the model on the
square lattice, but the existence or not of an intermediate spin-liquid
phase in this case is not so clear, with most works favoring the
non-existence of such a state.\cite{Chu}$^{-}$\cite{Leche1} This is
somehow paradoxical,
since the TLHA was considered for many years the best candidate to have a
spin-liquid ground state just on the basis of the frustrating lattice
topology. One would expect that the introduction of additional frustration
through the $J_{2\text{ }}$interaction should contribute to melt the already
weak order of the nearest-neighbor model.
The study of the $J_1\negthinspace -\negthinspace J_2$ TLHA was mostly
driven by purely theoretical motivations. On the other hand, recent
experimental results have produced a surge of
interest\cite{McK}$^{-}$\cite{series} in
the ground-state properties of the nearest-neighbor TLHA with spatially
anisotropic couplings: $J_1$ along two bond directions and $J_1^{\prime }$
in the third one. It has been argued\cite{McK} that this model should
describe the magnetic phases of the quasi-two-dimensional organic
superconductors ${\it \kappa }-$(BEDT-TTF)$_2$X. For these layered molecular
crystals the relevant values of $J_1^{\prime }/J_1$ are around $0.3-1.0$,
with the ratio varying with the anion X and with uniaxial stress along the
layer diagonal.
We have previously performed a study of both the $J_1\negthinspace -%
\negthinspace J_2$ and $J_1\negthinspace -\negthinspace J_1^{\prime }$ TLHA
using the rotational invariant Schwinger-boson approach (SBA) in a
mean-field approximation.\cite{GC} At that time, our motivations for
studying the latter model was the natural interpolation that it provides
between the nearest-neighbor TLHA ($J_1^{\prime }=J_1$) and the
square-lattice antiferromagnet ($J_1^{\prime }=0$). We obtained a good
agreement between the ground-state energy predicted in our approach and
exact numerical values on finite lattices. Furthermore, no indication of a
disordered state was found for the values of $J_2/J_1$ and $J_1^{\prime
}/J_1 $ of interest. In this work we extend the calculations in \cite{GC} to
include one-loop corrections to the mean-field picture. We have shown\cite
{letter,fluctri} that these corrections bring the saddle-point results in
line with exact diagonalization values on finite clusters, which lends
support to the SBA predictions for the thermodynamic limit. In particular,
for the square-lattice $J_1\negthinspace -\negthinspace J_2$ model we found
that there is a quantum nonmagnetic phase for $0.53\negthinspace \lesssim
\negthinspace J_2/J_1\negthinspace \lesssim \negthinspace 0.64$. This result
was obtained by considering the spin stiffness on large lattices and
extrapolating to the thermodynamic limit, a procedure that avoids the
infinite-lattice infrared divergencies associated to Bose condensation. We
show here that similar behaviors are predicted for both the $J_1%
\negthinspace -\negthinspace J_2$ and $J_1\negthinspace -\negthinspace %
J_1^{\prime }$ TLHA.
\section{The calculational method}
We consider a general Hamiltonian of the form
\begin{equation}
\label{H}H=\frac 12\sum_{{\bf r,r^{\prime }}}J({\bf r-r^{\prime }})\ {\vec S}%
_{{\bf r}}\cdot {\vec S}_{{\bf r^{\prime }}},
\end{equation}
where ${\bf r,r^{\prime }}$\ indicate sites on a triangular lattice. As
usual, we write spin operators in terms of Schwinger bosons: \cite{A} ${\vec
S_i}\negthinspace =\negthinspace {\frac 12}{\bf a}_i^{\dagger }.{\bf \vec
\sigma }.{\bf a}_i$, where ${\bf a_i^{\dagger }}\negthinspace
=\negthinspace (a_{i\uparrow }^{\dagger },a_{i\downarrow }^{\dagger })$ is a
bosonic spinor, ${\bf \vec \sigma }$ is the vector of Pauli matrices, and
there is a boson-number restriction $\sum_\sigma a_{i\sigma }^{\dagger
}a_{i\sigma }\negthinspace =\negthinspace 2S$ on each site. With this
representation the spin-spin interaction can be written as ${\vec S_i}.{\vec
S_j}\negthinspace =:\negthinspace
B_{ij}^{\dagger }B_{ij}\negthinspace :-A_{ij}^{\dagger }A_{ij},$ where we
defined the $SU(2)-$invariant order parameters $A_{ij}=\frac 12\sum_\sigma
\sigma a_{i\sigma }a_{j{\bar \sigma }}$ and $B_{ij}^{\dagger }=\frac
12\sum_\sigma a_{i\sigma }^{\dagger }a_{j\sigma }\ ({\bar \sigma }=-\sigma
,\ \sigma =\pm 1)$, and the notation $:\negthinspace O\negthinspace :$
indicates the normal order of operator $O$. Thus, we can formulate the
partition function for the Hamiltonian (\ref{H}) as a functional integral
over boson coherent states, which allows its evaluation by a saddle-point
expansion. Since the theory presents a local $U(1)$ symmetry, we use
collective coordinate methods ---as developed in the context of relativistic
lattice gauge theories\cite{P}--- to handle the infinitely-many zero modes
associated to the symmetry breaking in the saddle point. These modes without
restoring forces, which correspond to local gauge transformations, are
eliminated by the exact integration of the collective coordinates. Such a
program can be carried out by enforcing in the functional integral measure
the so-called background gauge condition, or ``natural'' gauge, \cite{P} by
means of the Fadeev-Popoff trick. In this way we restrict the integrations
in the partition function to field fluctuations that are orthogonal to the
collective coordinates. At $T=0$, after carrying out the integrations on
these genuine fluctuations, we obtain the one-loop correction to the
ground-state energy {\it per site},
\begin{equation}
\label{E1}E_1=-\frac 1{2\pi }\int_{-\infty }^\infty d\omega \sum_{{\bf k}%
}\ln \left( \frac{\Delta _{{\rm FP}}({\bf k},\omega )}{|\omega |\sqrt{\det
{\cal A}_{\perp }^{(2)}({\bf k},\omega )}}\right) .
\end{equation}
Here the Fadeev-Popoff determinant $\Delta _{{\rm FP}}({{\bf k},\omega })%
\negthinspace =\negthinspace \left| {\vec \varphi }_0^L({{\bf k},\omega }).{%
\vec \varphi }_0^R({{\bf k},\omega })\right| ,$where ${\vec \varphi }_0^L({%
{\bf k},\omega })$ is the {\it left} zero mode of the dynamical matrix $%
{\cal A}^{(2)}$ in the ${\bf k}\negthinspace -\negthinspace \omega $
subspace, and ${\cal A}_{\perp }^{(2)}$ is the projection of this matrix in
the subspace orthogonal to the collective coordinates. The dynamical matrix $%
{\cal A}^{(2)}$ is the Hessian of the effective action as a function of the
decoupling (Hubbard-Stratonovich) fields (see \cite{letter} for details).
In the ordered phases of the model the Bose condensate breaks the global $%
SU(2)$ symmetry and its density gives the local magnetization.\cite{A} The
associated physical Goldstone modes at ${\bf k}\negthinspace =\negthinspace
0,\pm {\bf Q}$\ (${\bf Q}$ is the magnetic wavevector) lead to serious
infrared divergencies of intermediate quantities, which have to be cured by
standard renormalization prescriptions. In order to avoid these problems we
have computed physical quantities (which are free of divergencies) on large
by finite lattices, and finally extrapolated these values to the
thermodynamic limit. We considered clusters with the spatial symmetries of
the infinite triangular lattice, corresponding to translation vectors ${\bf T%
}_1=(n+m){\bf e}_1+m{\bf e}_2$, ${\bf T}_2=n{\bf e}_1+(n+m){\bf e}_2.$\cite
{Bernu,Leche2}
Here ${\bf e}_1={\bf (}a,0{\bf )\ }$and ${\bf e}_2=(-%
{\bf \frac 12}a{\bf ,\frac{\sqrt{3}}2}a{\bf )\ }$are the basic triangular
lattice vectors.${\bf \ }$To fit to the cluster shapes the expected (spiral)
magnetic orders, and also to allow the calculation of the stiffness, we
impose arbitrary boundary conditions ${\vec S}_{{\bf r+T}_i}={\cal R}_{
\widehat{n}}(\Phi _i){\vec S}_{{\bf r}}$ ($i=1,2$) , where ${\cal R}_{
\widehat{n}}(\Phi _i)$ is the matrix that rotates an angle $\Phi _i$ around
some axis $\widehat{n}$ (notice that we are using boldface (arrows) for
vectors in real (spin) space). It is convenient to perform local rotations ${%
\vec S}_{{\bf r}}\rightarrow {\cal R}_{\widehat{n}}(\theta _{{\bf r}}){\vec S%
}_{{\bf r}}$ of angle $\theta _{{\bf r}}=\Delta {\bf Q\cdot r}${\bf ,} so
that with the choice $\Delta $${\bf Q\cdot T}_i=\Phi _i$ the boundary
conditions become the standard periodic ones ${\vec S}_{{\bf r+T}_i}={\vec S}%
_{{\bf r}}$. Then, we define the ($T=0$) stiffness tensor $\rho _{\widehat{n}%
}$ by
\begin{equation}
\label{RHO}\rho _{\widehat{n}}^{ij}=\left. \frac{\partial ^2E_{{\rm GS}}(%
{\bf Q+}\Delta {\bf Q)}}{\partial \theta _i\partial \theta _j}\right|
_{\Delta {\bf Q=}0},
\end{equation}
where $E_{{\rm GS}}$ is the ground-state energy {\it per spin} and $\theta
_i=\Delta {\bf Q\cdot e}_i$ ($i=1,2$) are the twist angles along the basis
vectors ${\bf e}_i$. In the next two sections we will apply this formalism
to the models under consideration.
\section{The $J_1\negthinspace -\negthinspace J_2$ TLHA}
The $J_1\negthinspace -\negthinspace J_2$ TLHA is given by (\ref{H}) with $J(%
{\bf r-r^{\prime }})=J_1$($J_2$) for ${\bf r}$ and ${\bf r^{\prime }}$%
nearest (next-nearest) neighbor sites, and $0$ otherwise. For classical spin
vectors, when $\alpha \equiv J_2/J_1<1/8$ the lowest-energy configuration in
this model is the commensurate spiral with magnetic wavevector ${\bf Q}=(
\frac{4\pi }{3a},0)$; for $1/8<\alpha <1$ there is a degeneracy between the
two-sublattice and four-sublattice N\'eel orders. Quantum fluctuations lift
this degeneracy and select, through an ''order from disorder '' phenomenon,
the two-sublattice collinear ground-state with magnetic wavevector ${\bf Q}%
=(\frac \pi a,-\frac \pi {\sqrt{3}a})$. This scenario was first proposed
using spin-wave theory\cite{Chu} and later confirmed by a study of the
thermodynamic-limit collapse to the ground state of low lying levels.\cite
{Leche1} The correction (\ref{E1}) to the saddle-point value $E_0$ leads to
the ground-state energy $E_{{\rm GS}}\negthinspace =\negthinspace E_0+E_1$
shown in Fig. 1. This figure contains the result for the infinite lattice
and also for a finite cluster of 12 sites, which allows a comparison with
exact results obtained by numerical diagonalization.\cite{SW} We see that
the addition of the Gaussian correction $E_1$ improves the saddle-point
value $E_0$, particularly for the 120$^{\circ }$ spiral phase. Moreover, the
departure of the fluctuation-corrected results from the exact values in the
range $0.1\lesssim\negthinspace \alpha \negthinspace \lesssim 0.2$ hints to
the possible existence of a disordered phase in this region. In the
thermodynamic limit, at saddle-point order the theory predicts a first order
transition between the two magnetic ground states at some intermediate
frustration $\alpha \simeq 0.16$, with no intervening disordered phase.\cite
{GC} After the inclusion of the Gaussian fluctuations there is a window $0.12%
\lesssim\negthinspace \alpha \negthinspace \lesssim 0.19$ where the
stiffness vanishes (see below) and the magnetic order is melted by the
combined action of quantum fluctuations and frustration.This result should
be compared with the linear spin wave results of \cite{Iva}, which predict a
quantum nonmagnetic phase in the range $(0.1,0.14)$. Notice, however, that
within spin-wave theory this window closes when corrections to the
leading-order calculations are included.\cite{Chu} We stress that the same
happens in this approach for the $J_1\negthinspace -\negthinspace J_2$ model
on the square lattice, where however other methods confirm the existence of
a nonclassical phase.
As stated above, the existence or not of magnetic long-range order in the
thermodynamic limit was investigated by considering the spin-stiffness
tensor (\ref{RHO}). \cite{stiff} For spins lying on the $xy$-plane, it is
necessary to consider both the parallel stiffness $\rho _{\Vert }\equiv \rho
_{\widehat{z}}$ under a twist around $\widehat{z}$ and the perpendicular
stiffness $\rho _{\bot }\equiv \overline{\rho }_{xy}$ for twists around an
arbitrary versor $\widehat{n}$ on this plane. However, on clusters with
periodic boundary conditions our approach is rotational invariant and we
only have access to $\overline{\rho }=\frac 13(\rho _{\Vert }+2\rho _{\bot
}) $. On the other hand, we found that the large-$S$ Schwinger-boson
prediction for this quantity is exactly 4/3 smaller than the corresponding
classical result. As discussed in \cite{fluctri}, to solve these problems we
considered clusters that fit the magnetic orders in the $xy$-plane with
appropriate twisted boundary conditions.\cite{Bernu,Leche2} In this case,
since the rotational invariance is explicitly broken by the boundary
conditions, one is able to compute the parallel stiffness $\rho _{\Vert }$
and, moreover, the large-$S$ predictions have the correct behavior (no {\it %
ad hoc} factors are required to correct this quantity).\cite{fluctri}
Finally, using the values for $\overline{\rho }$ obtained from clusters with
periodic boundary conditions it is possible to determine $\rho _{\bot
}\equiv \frac{3.}2$ $\overline{\rho }-\frac 12\rho _{\Vert }$. The
corresponding results are presented in Fig. 2, both at saddle-point and
one-loop orders for comparison. Notice that in the 120$^{\circ }$ spiral
phase $\rho _{\Vert }$ softens first than $\rho _{\bot }$, while the
behavior is the opposite for the colinear state. Since the indirect
calculation of $\rho _{\bot }$ requires the use of {\it ad hoc} factors to
renormalize $\overline{\rho }$, the determination of the upper limit for the
disordered phase might be unreliable. However, based on our previous
experience with the SBA, we believe the existence of an intermediate
nonmagnetic phase can be trusted.
\section{The $J_1\negthinspace-\negthinspace J_1^{\prime }$ TLHA}
For the $J_1\negthinspace -\negthinspace J_1^{\prime }$ TLHA, in (\ref{H})
we take $J({\bf r-r^{\prime }})=J_1$ for ${\bf r-r^{\prime }=e}_{i\text{ }}$(%
$i=1,2$) and $J({\bf r-r^{\prime }})=J_1^{\prime }$ for ${\bf r-r^{\prime }=e%
}_3\equiv {\bf e}_1+{\bf e}_2.$ As mentioned above, this model is
interesting in view of its connections to the spin degrees of freedom in the
insulating phase of some layered organic superconductors.\cite{McK} Very
recent works have investigated its ground-state phase diagram and other
properties, using spin-wave theory\cite{Colo2} and series expansions.\cite
{series} For classical vectors, with $\eta \equiv J_1^{\prime }/J_1<1/2$ the
lowest-energy configuration is the two-sublattice N\'eel order discussed in
the previous section, with the (frustrated) ferromagnetic arrangement of the
spins along the weakly-coupled ${\bf e}_3$ direction. For $\eta >1/2$ the
preferred spin configuration becomes an incommensurate spiral, with the
angle $Q{\rm _{Class}}$ between neighboring spins along the ${\bf e}_1,{\bf e%
}_2$ directions given by $Q{\rm _{Class}}=\arccos (-\frac 1{2\eta }).$
The inclusion of the quantum nature of spins by means of the SBA produces
the results for the ground-state energy shown in Fig. 3. In this figure we
give the mean-field values and the fluctuation-corrected results, and
compare them with the series expansion predictions from \cite{series}. We
see that the fluctuations produce regions in which the saddle-point
solutions become unstable. On the other hand, for those values of $\eta $
where the magnetic phases considered are stable there is a very good
agreement between both methods. It is also of interest to compare the
quantum renormalization of the spiral pitch; in Fig. 4 we plot the classical
result for $Q$ given above, and the corresponding angle that minimize the
quantum ground-state energy. Again, in the region where the magnetic phases
are stable the results nicely agree with those coming from series
expansions. Furthermore, we have checked at mean-field order that the angle
that minimizes the total energy corresponds also to the minimum of the
quasiparticle dispersion relation, a fact that was used in \cite{series} to
determine $Q.$
To establish the regions without magnetic order we studied again the
spin-stiffness behavior. Like for classical vectors, the stiffness tensor (%
\ref{RHO}) is diagonal in the (perpendicular) directions ${\bf e}_3$ and $%
{\bf e}_{2}-{\bf e}_1$; the corresponding stiffness along these directions
are plotted in Fig. 5. We found a qualitatively different behavior between
the saddle-point and one-loop results; most notably, in the N\'eel phase the
Gaussian fluctuations soften first the classically stronger stiffness in the
${\bf e}_3$ direction. At mean-field order our calculations indicate a
continuous transition between the two magnetic phases at $\eta \simeq 0.621$%
, and the absence of a magnetic saddle-point solution beyond $\eta \simeq
2.20$. The corrected results show the melting of the N\'eel phase at $\eta
\simeq 0.61$, still above the classical point $\eta {\rm _{Class}}=1/2$, and
predict a disordered quantum phase in the range $0.61\lesssim \eta \lesssim %
0.96$. Furthermore, the incommensurate phase becomes stable only in a
reduced parameter region $0.96\lesssim \eta \lesssim 1.10.$ In this case the
instability appears as a negative eigenvalue of the projected dynamical
matrix ${\cal A}_{\perp }^{(2)}$ in (\ref{E1}). These results are in fair
agreement with the series-expansion predictions from \cite{series}, which
indicate that the N\'eel order disappears at $\eta \sim 0.65-0.7$, the
disorder region extends from this value up to $\eta \sim 0.9$, and there is
an incommensurate phase for $\eta \gtrsim 0.9$ with no clear ending point.
\section{Conclusions}
In conclusion, we have considered, within the SBA, the Gaussian-fluctuation
corrections to the spin stiffness of the $J_1\negthinspace-\negthinspace J_2$
and $J_1\negthinspace-\negthinspace J_1^{\prime }$ TLHA. For the $J_1%
\negthinspace-\negthinspace J_2$ model we found that the order-parameter
fluctuations weaken the stiffness, which is reduced by increasing the
frustration until it vanishes leaving a small window $0.12\lesssim \alpha
\lesssim 0.19$ where the system has no long-range magnetic order. Like in
previous studies, we found that the consideration of clusters which require
twisted boundary conditions to fit the magnetic orders avoids the use of
{\it ad hoc} factors to correct the Schwinger-boson predictions. This fact
points to a subtle interplay between rotational invariance and the
relaxation of local constraints in this approach.
In the case of the $J_1\negthinspace-\negthinspace J_1^{\prime }$ TLHA our
results indicate a rich phase diagram, with at least two magnetic (N\'eel
and incommensurate spiral) phases and two disordered quantum states in the
parameter region of interest. One of the most notable aspects of our
calculations is the quantum renormalization of the magnetic wavevector in
the spiral phase, which agrees remarkably with the series expansion
prediction. We also found that the N\'eel order extends beyond its classical
stability point up to a value $\eta \simeq 0.61,$ where it seems to melt
continuously into a purely quantum phase. On the contrary, the spiral order
is favored in a reduced parameter range, with our results indicating first
order transitions from this phase to the nonmagnetic states. These first
order transitions appear in our calculations as local instabilities against
the order-parameter fluctuations.
Finally, there still remains the difficult task of characterizing the
disordered quantum phases of the models under consideration. Some
attempts in this
direction have already been done, for both the $J_1\negthinspace -%
\negthinspace J_2$ \cite{TLHA12} and $J_1\negthinspace -\negthinspace %
J_1^{\prime }$ \cite{series} TLHA, but there is not clear understanding of
these phases yet. They are usually considered to be of the
resonant-valence-bond type, and are in general described starting from a
regular strong-bond (dimer) covering of the lattice.\cite{series} These
investigations can be performed within the formalism developed here, since
the SBA does not rely on having magnetic order in the system like, for
instance, spin-wave theory. However, these studies would require the
extension of the present calculations to larger magnetic unit cells and the
computation of physical quantities able to characterize the new phases, a
question that is far from trivial.
\acknowledgements
{We are grateful to Adolfo E. Trumper for useful discussions and for
calling our attention to Ref. 12.}
|
\section{Introduction}
Let $S$ be a compact connected orientable surface. The mapping class group
${\cal M}_S$ of the surface $S$ is the group of isotopy classes of
orientation preserving diffeomorphisms $S\to S$. The extended mapping
class group ${\cal M}_S^*$ of $S$ is the group of isotopy classes of all
diffeomorphisms $S\to S$. Note that the isotopy classes of orientation
reversing diffeomorphisms are also included in ${\cal M}_S^*$,
and hence ${\cal M}_S$ is a subgroup of ${\cal M}_S^*$ of index two.
Recall that a group $G$ is called residually finite if for each
$x\neq 1$ in $G$ there exists a homomorphism $f$ from $G$ onto
some finite group such that $f(x)$ is nontrivial. Equivalently,
there is some finite index normal subgroup of $G$ that does not
contain $x$. $G$ is called hopfian if every surjective
endomorphism of $G$ is an automorphism. It is well known that
finitely generated residually finite groups are hopfian \cite{ls}.
$G$ is called cohopfian if every injective endomorphism of $G$ is
an automorphism.
The mapping class group of an orientable surface is finitely
generated \cite{l,b} and residually finite \cite{g,i1}. Hence it
is hopfian. N.V. Ivanov and J.D. McCarthy \cite{im} proved that
${\cal M}_S$ is also cohopfian. Author and J.D. McCarthy \cite{km}
proved that if $\phi :{\cal M}_S\to{\cal M}_S$ is a homomorphism
such that $\phi({\cal M}_S)$ is a normal subgroup and ${\cal
M}_S/\phi({\cal M}_S)$ is abelian, then $\phi$ is an automorphism.
In this paper, we prove further that if $\phi$ is an endomorphism
of the mapping class group ${\cal M}_S$ onto a finite index
subgroup, then $\phi$ is in fact an automorphism, with a few
exceptions. The proof of this result relies on a result of R.
Hirshon, which states that if $\phi$ is an endomorphism of a
finitely generated residually finite group $G$ such that $\phi(G)$
is of finite index in $G$, then $\phi$ restricted to $\phi^n(G)$
is an injection for some $n$.
D.T. Wise \cite{w} gave an example of a finitely generated
residually finite group $G$ and an endomorphism $\Phi$ of $G$ such
that the restriction of $\Phi$ to $\Phi^n(G)$ is not injective for
any $n$, answering a question of R. Hirshon in negative. It might
be interesting to consider the same question for mapping class
groups of surfaces.
\section{Endomorphisms of mapping class groups}
Let $S$ be a compact connected oriented surface of genus $g$ with
$b$ boundary components. For any simple closed curve $a$ on $S$,
there is a well known diffeomorphism, called a right Dehn twist,
supported in a regular neighborhood of $a$. We denote by $t_a$ the
isotopy class of a right Dehn twist about $a$, also called a Dehn
twist. Note that $ft_af^{-1}=t_{f(a)}$ for any orientation
preserving mapping class $f$.
The pure mapping class group ${\cal PM}_S$ is the subgroup of
${\cal M}_S$ consisting of those orientation preserving mapping
classes which preserve each boundary component.
For a group $G$ and a subgroup $H$ of it, we denote by $C_G(H)$
the centralizer of $H$ in $G$. The center of $G$ is denoted by $C(G)$.
\begin{thm}
Let $G$ be a finitely generated residually finite group, and let
$\phi$ be an endomorphism of $G$ onto a finite index subgroup.
Then there exists an $n$ such that the restriction of $\phi$ to
$\phi^n(G)$ is an injection.
\label{thm1}
\end{thm}
\begin{thm}
Let $S$ be a compact connected orientable surface of genus $g$ with $b$
boundary components. Suppose, in addition, that if $g=0$ then $b\geq 5$, if
$g=1$ then $b\geq 3$, and if $g=2$ then $b\geq 1$. Then any
isomorphism between two finite index subgroups of the extended mapping
class group ${\cal M}_S^*$ is the restriction of an inner automorphism
of ${\cal M}_S^*$.
\label{thm2}
\end{thm}
Theorem \ref{thm1} was proved by R. Hirshon (\cite{h}), and
Theorem \ref{thm2} was proved by N.V. Ivanov \cite{i2} for
surfaces of genus at least two and by the author \cite{k} for the
remaining cases. Since the mapping class group ${\cal M}_S$ is
normal in ${\cal M}_S^*$, we deduce the following theorem.
\begin{thm}
Let $S$ be a compact connected orientable surface of genus $g$ with $b$
boundary components. Suppose, in addition, that if $g=0$ then $b\geq 5$, if
$g=1$ then $b\geq 3$, and if $g=2$ then $b\geq 1$. Then any
isomorphism between two finite index subgroups of the mapping
class group ${\cal M}_S$ is the restriction of an automorphism
of ${\cal M}_S$.
\end{thm}
\begin{lemma}
Let $S$ be a closed orientable surface of genus two and let $\Gamma$ be a
finite index subgroup of ${\cal M}_S$. Then the center $C(\Gamma)$ of
$\Gamma$ is equal to $\Gamma\cap\langle\sigma\rangle$, where $\sigma$ is
the hyperelliptic involution.
\label{lemma}
\end{lemma}
{\it Proof:} Since the subgroup $\langle\sigma\rangle=
\{1,\sigma\}$ is the center
of ${\cal M}_S$, its intersection with $\Gamma$ is contained in the
center of $\Gamma$.
Now let $f\in C(\Gamma)$ and let $N$ be the index of $\Gamma$ in
${\cal M}_S$. Since $t_a^N\in\Gamma$ for all simple closed curves
$a$, we have $t_{f(a)}^N=ft_a^Nf^{-1}=t_a^N$. It follows that
$f(a)=a$ (cf. \cite{im}). Hence, $ft_af^{-1}=t_{f(a)}=t_a$. Since
${\cal M}_S$ is generated by Dehn twists, $f\in C({\cal M}_S)
=\langle\sigma\rangle$. $\Box$
\bigskip
We are now ready to state and prove the main result of
this paper.
\begin{thm}
Let $S$ be a compact connected orientable surface of genus $g$
with $b$ boundary components. Suppose, in addition, that if $g=0$
then $b\neq 2,3,4$, and if $g=1$ then $b\neq 2$. If $\phi$ is an
endomorphism of ${\cal M}_S$ such that $\phi({\cal M}_S)$ is of
finite index in ${\cal M}_S$, then $\phi$ is an automorphism.
\label{mthm}
\end{thm}
{\it Proof:} If $S$ is a (closed) sphere or a disk, then ${\cal
M}_S$ is trivial. Clearly, the conclusion of the theorem holds.
Suppose first that $S$ is a torus with $b\leq 1$ boundary
component. It is well known that ${\cal M}_S$ is isomorphic to
$SL_2({\bf Z})$. The commutator subgroup of $SL_2({\bf Z})$ is a nonabelian free group
of rank $2$ and its index in $SL_2({\bf Z})$ is $12$. Let us denote it by
$F_2$. $\phi(F_2)$ is contained in $F_2$ as a finite index
subgroup. If this index is $k$, $\phi(F_2)$ is a free group of
rank $k+1$. Since there is no homomorphism from $F_2$ onto a free
group of rank $\geq 3$, it follows that $k=1$. That is,
$\phi(F_2)=F_2$. In particular, $\phi(SL_2({\bf Z}))$ contains $F_2$. The
fact that $\phi$ is an automorphism in this case was proved in
\cite{km}.
Suppose now that $S$ is not one of the surface above and not a
closed a surface of genus $2$. Let us orient $S$ arbitrarily.
Since ${\cal M}_S$ is finitely generated and residually finite,
there exists an $n$ such that the restriction of $\phi$ to
$\phi^n({\cal M}_S)$ is an isomorphism onto $\phi^{n+1}({\cal
M}_S)$. Note that the subgroups $\phi^n({\cal M}_S)$ and
$\phi^{n+1}({\cal M}_S)$ are of finite index in ${\cal M}_S$.
Hence, there is an automorphisms $\alpha$ of ${\cal M}_S$ such
that the restrictions of $\alpha$ and $\phi$ to $\phi^n({\cal
M}_S)$ coincide.
Let $N$ be the index of $\phi^n({\cal M}_S)$ in ${\cal M}_S$.
For any simple closed curve $a$ on $S$, $t_a^N$ is contained in
$\phi^n({\cal M}_S)$. Hence,
$\alpha(t_a^N)=\phi(t_a^N)$. Let $f\in{\cal M}_S$ be any element. Then
\begin{eqnarray*}
t_{f(a)}^N&=&\alpha^{-1}(\phi(t_{f(a)}^N))\\
&=&\alpha^{-1}(\phi(ft_a^Nf^{-1}))\\
&=&\alpha^{-1}(\phi(f))\alpha^{-1}(\phi(t_a^N))\alpha^{-1}(\phi(f^{-1}))\\
&=&\alpha^{-1}(\phi(f))t_a^N\alpha^{-1}(\phi(f))^{-1}\\
&=&t_{\alpha^{-1}(\phi(f))(a)}^N.
\end{eqnarray*}
Hence, $\alpha^{-1}(\phi(f))(a)=f(a)$ for all $a$ (cf. \cite{im}).
It follows that $f^{-1}\alpha^{-1}(\phi(f))$ commutes with all
Dehn twists. Since ${\cal PM}_S$ is generated by Dehn twists, it
is in $C_{{\cal M}_S}({\cal PM}_S)$, the centralizer of ${\cal
PM}_S$ in ${\cal M}_S$. But $C_{{\cal M}_S}({\cal PM}_S)$ is
trivial \cite{im}. Hence, $\alpha^{-1}(\phi(f))=f$. Therefore,
$\phi=\alpha$. In particular, $\phi$ is an automorphism.
Suppose finally that $S$ is a closed surface of genus two. Let $R$
be a sphere with six holes. Then ${\cal M}_R$ is isomorphic to the
quotient of ${\cal M}_S$ with its center $\langle\sigma\rangle$,
where $\sigma$ is the hyperelliptic involution (cf. \cite{bh}) .
Let us identify ${\cal M}_R$ and ${\cal M}_s/ \langle
\sigma\rangle$, and let $\pi: {\cal M}_S\to {\cal M}_R$ be the
quotient map. Since $\phi(\sigma)$ is in the center of $\phi({\cal
M}_S)$, either $\phi(\sigma)=\sigma$ or $\phi(\sigma)=1$ by the
lemma above.
If $\phi(\sigma)=\sigma$, then $\phi$ induces an endomorphism
$\Phi$ of ${\cal M}_R$, such that $\pi\phi=\Phi\pi$. Then we have
a diagram in which all squares are commutative:
\[
\begin{array}{ccccccccc}
1& \longrightarrow &\langle\sigma\rangle& \longrightarrow&{\cal M}_S&
\stackrel{\pi}\longrightarrow&{\cal M}_R&\longrightarrow &1\\
& &\Big\downarrow {\rm I}& &\Big\downarrow \phi& &\Big\downarrow\Phi& & \\
1& \longrightarrow &\langle\sigma\rangle& \longrightarrow&{\cal M}_S&
\stackrel{\pi}\longrightarrow &{\cal M}_R & \longrightarrow &1
\end{array}
\]
where ${\rm I}$ is the identity homomorphism. Since the image
$\Phi({\cal M}_R)$ of $\Phi$ is of finite index, $\Phi$ is an
automorphism by the first part. By $5$-lemma, $\phi$ is an
automorphism.
If $\phi(\sigma)=1$, then $\phi$ induces a homomorphism
$\overline{\phi}:{\cal M}_R\longrightarrow {\cal M}_S$ such that
$\overline{\phi}\pi=\phi$.
The image of the endomorphism
$\Phi=\pi\overline{\phi}$ of ${\cal M}_R$ has finite index. Since
$R$ is a sphere with six holes, $\Phi$ is an automorphism by the
first part. Then, $\phi$ is an automorphism, and hence $\sigma=1$.
This contradiction finishes the proof of our theorem. $\Box$
\bigskip
\noindent {\bf Remark}: If $S$ is a sphere with two holes, then
${\cal M}_S$ is a group of order two, and if $S$ is a sphere with
three holes, then ${\cal M}_S$ is isomorphic to the symmetric
group on three letters. Hence, in these cases the trivial
homomorphism is an endomorphism onto a finite index subgroup which
is not automorphism. We do not know if the conclusion of Theorem
\ref{mthm} holds if a sphere with four holes and a torus with two
holes.
\bigskip
|
\section{Introduction}
\label{intro}
Self-organized criticality (SOC) is observed in various dynamical systems
that develop
algebraic correlations. The systems drive themselves into critical states
independent of initial states or tuning of parameters. The complex
behaviour is characterized by the absence of characteristic time or
length scales. It is possible
to define a set of critical scaling exponents that characterize the large scale
behaviour and describe the universal 1/f-power spectra (for an overview
see Ref.\ 2 in \cite{Vespa}).
Sandpile automata \cite{BTW} are among the simplest models that develope SOC
behaviour in the avalanche propagation.
The model is defined as follows: on each site of a lattice, a variable $z_i$
is defined (number of grains of sand). At each time step, the variable at a
randomly chosen
site is increased $z_i \rightarrow z_i+1$. Above a critical value $z_c$,
the site topples $z_i \rightarrow z_i-z_c$ and the nearest neighbors $j$ of
site $i$ are increased $z_j \rightarrow z_j+1$. Thereby, $z_c$ grains of sand
have been transported from site $i$ to the neighbors $j$.
A toppling can induce nearest neighbor topplings and can create avalanches
of arbitrary sizes $s$.
For finite systems, one typically uses
open boundary conditions (periodic boundary conditions can lead to never
ending avalanches), i.e.\ the sand is just dropping from the table
if it reaches the border. This introduces the concept of
finite size scaling (FFS) to compensate cutoff effects.
Under the assumption that FFS is valid the avalanches
are described by three variables: the number of toppling $s$, the area
$a$ covered by an avalanche and the avalanche duration $T$. The probability
distributions of these variables are algebraic:
\begin{equation}
P(x)=x^{-\tau_x}G(\frac{x}{x_c})
\end{equation}
with $x\in\{s,a,T\}$. $G$ is a cutoff function and the cutoff diverges
$x_c \sim L^{\beta_x}$ with the system size $L \rightarrow \infty$.
The universality class is determined by the set of exponents $\{\tau_x,
\beta_x\}$. Very recently, the validity of FFS for sandpile models
was questioned \cite{Drossel} but this discussion is far beyond the
scope of this article. We refer to \cite{Drossel} and references therein
and take FFS as a working hypothesis hereafter.
The sandpile automata can be isotropic \cite{BTW}, i.e the sand is
distributed equally between nearest neighbors, or anisotropic
\cite{Manna}. The question whether both belong to the same universality
class is still under discussion. Whereas real-space normalization group
\cite{Pietro}, field theoretical \cite{Dickman} and late numerical approaches
\cite{Vespa} suggest that both belong to the same universality class,
Ben-Hur and Biham \cite{Biham} question these results.
In what follows we implement the sandpile model on the 8-fold quasiperiodic
rhombus-tiling, also called Am\-mann-Been\-ker tiling. Due to the specific
geometry of qua\-si\-pe\-rio\-dic tilings it can
become either anisotropic with complete non-deterministic toppling rules,
or isotropic with deterministic topplings, or a mixture of both.
\section{Non-periodic tilings and SOC}
\label{sec:1}
Non-periodic tilings \cite{Gruenbaum} are used to describe the geometry
of quasicrystals \cite{Shechtman}. They appear in two distinguished classes.
The ideal quasiperiodic tilings are produced by a cut-and-pro\-ject-scheme
from a higher dimensional (minimal) embedding lattice \cite{Kramer,BJKS}.
Their vertices are given by so-called model sets \cite{Moody} and they
show pure Bragg-peak diffraction spectra with non-crystallographic
symmetries (5,7,8,...-fold). Typically, the tilings are built by two or more
prototiles. Fig.\ \ref{fig1} shows the 8-fold Am\-mann-Been\-ker tiling
consisting of two prototiles, a square and a $\pi/4$ rhombus.
The second class of non-periodic tilings are the so-called
random tilings. They are constructed by the same prototiles
as the ideal ones but this time these prototiles fit together stochastically
in such a way
that they fill the entire space, face to face without gaps and overlaps
\cite{Elser,Henley}. They also can be embedded into a higher dimensional
lattice but their diffraction properties change. They can keep the same
symmetry of the ideal counterparts but this time, their diffraction
spectrum consists of Bragg-peaks and diffuse scattering (in 3D) or
algebraic peaks (in 2D).
Beneath the long-range correlations which govern the diffraction properties,
the main difference of non-periodic tilings compared to crystallographic
tilings is in their local neighborhood. Whereas we have a fixed translational
invariant neighborhood for crystals, we have a finite atlas of
different vertex configurations without translational invariance for
quasicrystals. For the example of the Am\-mann-Been\-ker tiling, we find
6 different vertices in the ideal and 16 different configurations in the
random version. The number of nearest neighbors ranges from 3 to 8
\cite{BJ}.
Concerning SOC-models, one now could expect two major influences.
Both, long-range correlations and the diversity of the local neighborhoods
of the tilings, could develop a difference in the avalanche distribution.
It is not necessary to test a lot of different quasiperiodic tilings
because the generic features of quasiperiodicity are the same.
Symmetry or the type of the tiling plays a minor r{\^o}le.
Additionally, the local configurations allow different implementations
of e.g.\ the sandpile model. For simplicity we will focus on three
potentially different versions (there are more possible but typically
they fall in the same categories) on the 8-fold Am\-mann-Been\-ker tiling:
\begin{itemize}
\item[(I)]
anisotropic and non-deterministic: The critical height $z_c$ and the number
of toppling grains $n_i=z_c>0$ are equal for all vertices $i$ of the tiling.
At every toppling step one has to make a random choice to distribute the
$n_i$ grains among the $N_i$ neighbors $(3 \leq N_i \leq 8)$ of the
toppling vertex $i$.
\item[(II)]
partly isotropic and deterministic, partly not:
$z_c$ is equal for all vertices $i$ but greater or equal to the minimal number
of neighbors $N_{\min}=3$ and less than the maximal number of neighbors
$N_{\max}=8$. The number of topplings at site $i$ is $n_i=\min(z_c,N_i)$
where $N_i$ again is the number of neighbors of vertex $i$. This way, the
toppling rules are locally isotropic and deterministic for vertices with
$N_i \leq z_c$ but anisotropic and non-deterministic for all the others because
of the random choice one has to make in order to distribute the $n_i=z_c$
grains among the $N_i>z_c$ neighbors.
\item[(III)]
isotropic and deterministic: $z_c$ is greater or equal than the
maximal number of neighbors $N_{\max}=8$, but the number of toppling grains
of vertex $i$ is always $n_i=N_i$. This way, every bond will transport the
same amount of grain in both directions and the number of topplings is given
locally by the number of neighbors.
\end{itemize}
\begin{figure}
\resizebox{0.45\textwidth}{!}{%
\includegraphics{figure1.ps}
}
\caption{
The ideal Am\-mann-Been\-ker tiling and a critical state at the end of an
avalanche of type {\bf (I)}. The symbols on top of the vertices code the
number of grains of sand $z_i$ and $z_c=5$.}
\label{fig1}
\end{figure}
After a suitable thermalization, we perform $2*10^6$ ($L=16, 32, 64$)
and $5*10^5$ ($L=128,256$) avalanches for the
different scenarios. Each avalanche is seperated by an appropriate
thermalization time to prevent cross-correlations. The system size
($L \times L)$ is given by the length $L$ which varies from 16 to
256 whereas the absolute scale is given by the bond length of the tiling
that is equal to one. For each setting, the avalanche distributions
are stored. According to the finite size scaling assumption, the data
for the different length scales should collaps to a single curve that
is determined by the critical exponents $\tau_x, \beta_x$. Fig.\ \ref{fig3}
shows the data collaps of the probability distributions of the number of
topplings $s$ for version {\bf (I)}, Fig.\ \ref{fig4} for version {\bf (II)},
and Fig.\ \ref{fig5} for the isotropic deterministic version {\bf (III)}.
Table \ref{tab1} gives the extracted numerical values for the exponents.
The errors given in the brackets are the statictical errors of the fitting
procedure.
\begin{figure}
\resizebox{0.38\textwidth}{!}{%
\includegraphics{figure2.ps}
}
\caption{
Comparsion of anisotropic avalanche distributions of an ideal and
a random tiling version of the Am\-mann-Been\-ker tiling at L=128.}
\label{fig2}
\end{figure}
\begin{figure}
\resizebox{0.4\textwidth}{!}{%
\includegraphics{figure3.ps}
}
\caption{
Data Collaps of the probability distribution of the number of topplings $s$
for the anisotropic, non-deterministic version {\bf (I)} on the
ideal Am\-mann-Been\-ker tiling. System lengths are L=16 ($\circ$),
32 ($\Box$), 64 ($\diamond$), 128 ($\triangle$), and 256 (+).}
\label{fig3}
\end{figure}
\begin{figure}
\resizebox{0.4\textwidth}{!}{%
\includegraphics{figure4.ps}
}
\caption{
Data Collaps of the probability distribution of the number of topplings $s$
for version {\bf (II)} on the ideal Am\-mann-Been\-ker tiling.
System lengths and symbols as in Fig.\ 3.}
\label{fig4}
\end{figure}
\begin{figure}
\resizebox{0.4\textwidth}{!}{%
\includegraphics{figure5.ps}
}
\caption{
Data Collaps of the probability distribution of the number of topplings $s$
for the isotropic, deterministic version {\bf (III)} on the
ideal Am\-mann-Been\-ker tiling.
System lengths and symbols as in Fig.\ 3.}
\label{fig5}
\end{figure}
Within the limits set by the small statistic, all three version
belong to the same universality class: the crtitical exponents and also
the scaling functions seem to be the same. If one compares the
distributions of the ideal quasiperiodic tilings and the corresponding
random tiling, there is no apparent difference (see Fig.\ \ref{fig2}).
Looking at the crystalline counterpart, the 2D square lattice, the state of
the art values for the exponents are $\tau_s=1.27(1)$ and $\beta_s=2.73(2)$
\cite{Vespa}. These values are determined by massive parallelized
computation. From the difference to Table \ref{tab1} one cannot conclude
that there exists a different universality class for quasiperiodic graphs.
The values of Table \ref{tab1} are more comparable to older estimates for
the square lattice based on similar statistics \cite{Grassberger}.
The distributions for the area $a$ and duration $T$ are not shown because
they behave in the same way.
\begin{table}
\caption{The citical exponents $\tau_s, \beta_s$ of the three tested versions
(Statistical errors are given in brackets).}
\label{tab1}
\begin{tabular}{llll}
\hline\noalign{\smallskip}
& Version {\bf (I)} & Version {\bf (II)} & Version {\bf (III)} \\
\noalign{\smallskip}\hline\noalign{\smallskip}
$\tau_s$ & 1.20(3) & 1.19(3) & 1.19(3) \\
$\beta_s$ & 2.5(2) & 2.5(2) & 2.5(2) \\
\noalign{\smallskip}\hline
\end{tabular}
\end{table}
\section{Conclusion}
We have performed numerical estimates of the critical exponents of
the 2D sandpile model on the 8-fold quasiperiodic Am\-mann-Been\-ker tiling.
Thereby, three different version (anisotropic-non-deterministic,
isotropic-deterministic and a mixed one) were implemented on the ideal
and on the random Am\-mann-Been\-ker tiling. The measured exponents are
comparable
to early numerical results of the 2D square lattice and might suggest that all
tested version belong to the same universality class as the 2D square lattice.
The problem however remains to substantiate this, e.g.\ by better statistics.
This can only be achieved by a massive use of a parallel computing
which is not at hand at the moment. But the situation seems to be similair
to the investigation of Ising-models \cite{micha} which led to the conlcusion
that aperiodic and periodic versions belong to the same universality class.
The author wants to thank S.\ Maslov and M.\ Baake for discussion and
helpful comments.
|
\section{Introduction}
\label{INTRODUCTION}
The braid groups $B_n$ appear in many different guises.
We recall here the definition we will be using
and some of the main properties we will need.
For other equivalent definitions
see \cite{birman:blam}.
Let $D$ denote a disc
and let $q_1,\dots,q_n$ be $n$ distinct points
in the interior of $D$.
For concreteness,
take $D$ to be the disc in the complex plane
centered at the origin and having radius $n+1$,
and take $q_1,\dots,q_n$ to be the points $1,\dots,n$.
Let $D_n$ denote the punctured disc $D\setminus \{q_1,\dots,q_n\}$,
with basepoint $p_0$ on $\partial D$, say $p_0 = -(n+1)i$.
\begin{defn}
The braid group $B_n$
is the group of all equivalence classes of
orientation preserving homeomorphisms $h\co D_n\rightarrow D_n$
which fix $\partial D$ pointwise,
where two such homeomorphisms are equivalent
if they are homotopic rel $\partial D$.
\end{defn}
It can be shown that
$B_n$ is generated by $\sigma_1,\dots,\sigma_{n-1}$,
where $\sigma_i$ exchanges punctures $q_i$ and $q_{i+1}$
by means of a clockwise twist.
Let $x_1,\dots,x_n$ be free generators of $\pi_1(D_n,p_0)$,
where $x_i$ passes counterclockwise around $q_i$.
Consider the map $\epsilon\co \pi_1(D_n) \rightarrow {\bf{Z}}$
which takes a word in $x_1,\dots,x_n$
to the sum of its exponents.
Let $\tilde{D}_n$ be the corresponding covering space.
The group of covering transformations of $\tilde{D}_n$ is ${\bf{Z}}$,
which we write as a multiplicative group generated by $t$.
Let $\Lambda$ denote the ring ${\bf{Z}}[t,t^{-1}]$.
The homology group $H_1(\tilde{D}_n)$
can be considered as a $\Lambda$--module,
in which case it becomes a free module of rank $n-1$.
Let $\psi$ be an autohomeomorphism of $D_n$
representing an element of $B_n$.
This can be lifted to a map
$\tilde{\psi}\co \tilde{D}_n\rightarrow\tilde{D}_n$
which fixes the fiber over $p_0$ pointwise.
This in turn induces a $\Lambda$--module automorphism
$\tilde{\psi}_*$ of $H_1(\tilde{D}_n)$.
The {\em (reduced) Burau representation} is the map
\[
\psi \mapsto \tilde{\psi}_*.
\]
This is an $(n-1)$--dimensional representation of $B_n$ over $\Lambda$.
The main result of this paper is the following.
\begin{thm}
\label{burau5}
The Burau representation is not faithful for $n = 5$.
\end{thm}
The idea is to use the fact that
the Dehn twists about two simple closed curves
commute if and only if those simple closed curves can be
freely homotoped off each other.
Our construction will use two simple closed curves
which cannot be freely homotoped off each other
but in some sense ``fool'' the Burau representation
into thinking that they can.
To make this precise, we first make the following definition.
\begin{defn}
Suppose $\alpha$ and $\beta$ are two arcs in $D_n$.
Let $\tilde{\alpha}$ and $\tilde{\beta}$ be lifts
of $\alpha$ and $\beta$ respectively
to $\tilde{D}_n$.
We define
\[
\int_\beta \alpha= \sum_{k\in{\bf{Z}}} (t^k\tilde{\alpha},\tilde{\beta}) t^k,
\]
where $(t^k\tilde{\alpha},\tilde{\beta})$
denotes the algebraic intersection number
of the two arcs in $\tilde{D}_n$.
Note that this is only defined
up to multiplication by a power of $t$,
depending on the choice of lifts $\tilde{\alpha}$ and $\tilde{\beta}$.
This will not pose a problem because we will only be interested in
whether or not $\int_\beta \alpha$ is zero.
\end{defn}
\begin{thm}
\label{lp}
For $n\geq 3$, the Burau representation of $B_n$ is not faithful
if and only if there exist embedded arcs $\alpha$ and $\beta$ on $D_n$
such that $\alpha$ goes from $q_1$ to $q_2$,
$\beta$ goes from $p_0$ to $q_3$ or from $q_3$ to $q_4$,
$\alpha$ cannot be homotoped off $\beta$ rel endpoints,
and $\int_\beta \alpha = 0$.
\end{thm}
The special case in which $\beta$ goes from $p_0$ to $q_3$
follows easily from \cite[Theorem 1.5]{long-paton:burau}.
This special case is all we will need to prove Theorem~\ref{burau5}.
Nevertheless,
we will give a direct proof of Theorem~\ref{lp}
in Section~\ref{LP}.
In Section~\ref{BURAU5} we give a pair of curves
on the $5$--punctured disc
which satisfy the requirements of Theorem~\ref{lp},
thus proving Theorem~\ref{burau5}.
Throughout this paper,
elements of the braid group act on the left.
If $\psi_1$ and $\psi_2$ are elements of the braid group $B_n$
then we denote their commutator by:
\[
[\psi_1,\psi_2] = \psi_1^{-1} \psi_2^{-1} \psi_1 \psi_2.
\]
\section{Proof of Theorem \ref{lp}}
\label{LP}
It will be useful to keep the following lemma in mind.
It can be found in \cite[Proposition 3.10]{flp:tdt}.
\begin{lem}
\label{no_digons}
Suppose $\alpha$ and $\beta$ are simple closed curves on a surface
which intersect transversely at finitely many points.
Then $\alpha$ and $\beta$ can be freely homotoped
to simple closed curves which intersect at fewer points
if and only if there exists a ``digon'',
that is, an embedded disc
whose boundary consists of one subarc of $\alpha$
and one subarc of $\beta$.
\end{lem}
First we prove the ``only if'' direction of Theorem~\ref{lp}.
Let $n\geq 3$ be such that
for any embedded arcs $\alpha$ from $q_1$ to $q_2$
and $\beta$ from $p_0$ to $q_3$ in $D_n$
satisfying $\int_\beta \alpha = 0$
we have that $\alpha$ can be homotoped off $\beta$ rel endpoints.
Let $\psi\co D_n \rightarrow D_n$
lie in the kernel of the Burau representation.
We must show that $\psi$ is homotopic to the identity map.
Let $\alpha$ be the straight arc from $q_1$ to $q_2$
and let $\beta$ be the straight arc from $p_0$ to $q_3$.
Then $\int_\beta \psi(\alpha) = 0$.
Thus $\psi(\alpha)$ can be homotoped off $\beta$.
By applying this same argument to an appropriate conjugate of $\psi$
we see that $\psi(\alpha)$ can be homotoped off
the straight arc from $p_0$ to $q_j$ for any $j=3,\dots,n$.
It follows that we can homotope $\psi$ so as to fix $\alpha$.
Similarly, we can homotope $\psi$ so as to fix
every straight arc from $q_j$ to $q_{j+1}$ for $j=1,\dots,n-1$.
The only braids with this property are powers of
$\Delta$, the Dehn twist about a simple closed curve
parallel to $\partial D$.
But $\Delta$ acts as multiplication by $t^n$ on $H_1(\tilde{D}_n)$.
Thus the only power of $\Delta$ which lies in the kernel of
the Burau representation is the identity.
We now prove the converse for the case in which
$\beta$ is an embedded arc from $q_3$ to $q_4$ in $D_n$.
Let $\alpha$ be an embedded arc from $q_1$ to $q_2$
such that $\alpha$ cannot be homotoped off $\beta$ rel endpoints
but $\int_\beta \alpha = 0$.
Let $\tau_\alpha\co D_n\rightarrow D_n$ be a ``half Dehn twist''
about the boundary of a regular neighborhood of $\alpha$.
This is the homeomorphism which exchanges punctures $q_1$ and $q_2$
and whose square is a full Dehn twist
about the boundary of a regular neighborhood of $\alpha$.
Similarly, let $\tau_\beta$ be a half Dehn twist
about the boundary of a regular neighborhood of $\beta$.
We will show that the commutator of $\tau_\alpha$ and $\tau_\beta$
is a non-trivial element of the kernel of the Burau representation.
Let $\epsilon$ be an embedded arc in $D_n$ which crosses $\alpha$ once.
\begin{figure}[ht!]\small
\centering
\begin{picture}(270,45)
\thinlines
\put(40,0){\vector(0,1){45}}
\put(43,45){$\epsilon$}
\thicklines
\put(0,20){\line(1,0){80}}
\put(60,20){\vector(1,0){0}}
\put(60,22){$\alpha$}
\put(0,20){\circle*{4}}
\put(80,20){\circle*{4}}
\thinlines
\put(110,20){\vector(1,0){10}}
\put(110,18){\line(0,1){4}}
\put(210,0){\oval(20,20)[tl]}
\put(210,10){\line(1,0){30}}
\put(240,20){\oval(20,20)[r]}
\put(240,30){\line(-1,0){30}}
\put(210,20){\oval(20,20)[tl]}
\put(190,20){\oval(20,20)[br]}
\put(190,10){\line(-1,0){30}}
\put(160,20){\oval(20,20)[l]}
\put(160,30){\line(1,0){30}}
\put(190,40){\oval(20,20)[br]}
\put(200,40){\vector(0,1){5}}
\thicklines
\put(160,20){\line(1,0){80}}
\put(180,20){\vector(-1,0){0}}
\put(160,20){\circle*{4}}
\put(240,20){\circle*{4}}
\end{picture}
\caption{The action of $\tau_\alpha$}
\label{fig_twist}
\end{figure}
Figure~\ref{fig_twist} shows $\epsilon$ and
its image under the action of $\tau_\alpha$.
Thus the effect of $\tau_\alpha$
on $\epsilon$ is, up to homotopy rel endpoints,
to insert the ``figure-eight'' $\alpha'$
shown in Figure~\ref{fig_eight}.
\begin{figure}[ht!]\small
\centering
\begin{picture}(90,25)
\thinlines
\put(10,10){\oval(20,20)[l]}
\put(10,20){\vector(1,0){10}}
\put(20,22){$\alpha'$}
\put(20,20){\line(1,0){10}}
\qbezier(30,20)(40,20)(50,10)
\qbezier(50,10)(60,0)(70,0)
\put(70,0){\line(1,0){20}}
\put(90,10){\oval(20,20)[r]}
\put(90,20){\vector(-1,0){10}}
\put(80,20){\line(-1,0){10}}
\qbezier(70,20)(60,20)(50,10)
\qbezier(50,10)(40,0)(30,0)
\put(30,0){\line(-1,0){20}}
\thicklines
\put(10,10){\line(1,0){80}}
\put(10,10){\circle*{4}}
\put(90,10){\circle*{4}}
\end{picture}
\caption{The ``figure eight'' $\alpha'$}
\label{fig_eight}
\end{figure}
Now let $\tilde{\epsilon}$ be a lift of $\epsilon$
to the covering space $\tilde{D}_n$.
Note that $\alpha'$ lifts to a closed curve in $\tilde{D}_n$.
Thus the effect of $\tilde{\tau}_\alpha$
on $\tilde{\epsilon}$ is, up to homotopy rel endpoints,
to insert a lift of~$\alpha'$.
Let $\tilde{\gamma}$ be a closed arc in $\tilde{D}_n$.
The effect of $\tilde{\tau}_\alpha$ on $\tilde{\gamma}$
is to insert some lifts of $\alpha'$.
If we consider $\tilde{\gamma}$ and $\tilde{\alpha}'$
as representing elements of $H_1(\tilde{D}_n)$ then
\[
(\tilde{\tau}_\alpha)_*(\tilde{\gamma})
= \tilde{\gamma} + P(t)\tilde{\alpha}',
\]
where $P(t) \in \Lambda$.
Similarly,
\[
(\tilde{\tau}_\beta)_*(\tilde{\gamma})
= \tilde{\gamma} + Q(t)\tilde{\beta}',
\]
where $Q(t) \in \Lambda$
and $\beta'$ is a figure eight
defined similarly to $\alpha'$.
Any lift of $\alpha$, and hence of $\alpha'$,
has algebraic intersection number zero
with any lift of $\beta$.
It follows that
\[
(\tilde{\tau}_\beta)_*(\tilde{\alpha}') = \tilde{\alpha}'.
\]
Thus
\[
(\tilde{\tau}_\beta \tilde{\tau}_\alpha)_*(\tilde{\gamma})
= (\tilde{\gamma} + Q(t)\tilde{\beta}') + P(t)\tilde{\alpha}'.
\]
Similarly
\[
(\tilde{\tau}_\alpha \tilde{\tau}_\beta)_*(\tilde{\gamma})
= (\tilde{\gamma} + P(t)\tilde{\alpha}') + Q(t)\tilde{\beta}'.
\]
Thus
$(\tilde{\tau}_\alpha)_*$ and $(\tilde{\tau}_\beta)_*$ commute,
so the commutator $[\tau_\alpha, \tau_\beta]$
lies in the kernel of the Burau representation.
It remains to show that $[\tau_\alpha, \tau_\beta]$
is not homotopic to the identity map.
Let $\gamma$ be the boundary of a regular neighborhood of $\alpha$.
Using Lemma~\ref{no_digons} and the fact that
$\gamma$ cannot be freely homotoped off $\beta$,
it is not hard to check that $\tau_\beta(\gamma)$
cannot be freely homotoped off $\alpha$.
A similar check then shows that $\tau_\alpha \tau_\beta(\gamma)$
cannot be freely homotoped off $\tau_\beta(\gamma)$.
Thus $\tau_\alpha \tau_\beta(\gamma)$ is not freely homotopic to
$\tau_\beta(\gamma)$.
But $\tau_\beta(\gamma) = \tau_\beta \tau_\alpha(\gamma)$.
Thus $\tau_\alpha \tau_\beta$ is not homotopic to
$\tau_\beta \tau_\alpha$,
so $[\tau_\alpha, \tau_\beta]$ is not homotopic to the identity map.
The case in which $\beta$ goes from $p_0$ to $q_3$
can be proved by a similar argument.
Instead of a half Dehn twist about
the boundary of a regular neighborhood of $\beta$
we use a full Dehn twist
about the boundary of a regular neighborhood of
$\beta \cup \partial D$.
Instead of a figure eight curve $\beta'$
we obtain a slightly more complicated curve
which is a commutator of $\partial D$
and the boundary of a regular neighborhood of $\beta$.
\section{Proof of Theorem \ref{burau5}}
\label{BURAU5}
Let $\alpha$ and $\beta$ be the embedded arcs on $D_5$
as shown in Figure~\ref{fig_b5}.
\begin{figure}[ht!]\small
\centering
\unitlength0.7pt
\begin{picture}(340,240)(4,0)
\thinlines
\put(324,0){\vector(0,1){66}}
\put(326,66){$\beta$}
\put(324,66){\line(0,1){66}}
\put(252,132){\oval(24,48)[t]}
\put(246,132){\oval(60,72)[t]}
\put(246,132){\oval(84,96)[t]}
\put(246,132){\oval(108,120)[t]}
\put(252,132){\oval(144,144)[t]}
\put(228,132){\oval(24,48)[b]}
\put(234,132){\oval(60,72)[b]}
\put(234,132){\oval(84,96)[b]}
\put(234,132){\oval(108,120)[b]}
\put(234,132){\oval(132,144)[b]}
\put(96,132){\oval(24,48)[t]}
\put(102,132){\oval(60,72)[t]}
\put(102,132){\oval(84,96)[t]}
\put(102,132){\oval(108,120)[t]}
\put(102,132){\oval(132,144)[t]}
\put(96,132){\oval(120,144)[b]}
\put(54,132){\oval(12,36)[b]}
\put(108,132){\oval(48,48)[b]}
\put(108,132){\oval(72,72)[b]}
\put(108,120){\line(0,1){12}}
\thicklines
\put(18,144){\line(1,0){78}}
\put(18,136){\line(1,0){312}}
\put(18,128){\line(1,0){312}}
\put(228,120){\line(1,0){78}}
\put(18,150){\oval(12,12)[bl]}
\put(18,150){\oval(28,28)[bl]}
\put(18,114){\oval(28,28)[tl]}
\put(330,142){\oval(12,12)[br]}
\put(330,142){\oval(28,28)[br]}
\put(306,114){\oval(12,12)[tr]}
\put(4,150){\line(0,1){72}}
\put(12,150){\line(0,1){72}}
\put(18,222){\oval(12,12)[tl]}
\put(18,222){\oval(28,28)[tl]}
\put(18,228){\line(1,0){312}}
\put(18,236){\line(1,0){312}}
\put(330,222){\oval(12,12)[tr]}
\put(330,222){\oval(28,28)[tr]}
\put(336,142){\line(0,1){80}}
\put(344,142){\line(0,1){80}}
\put(4,42){\line(0,1){72}}
\put(18,42){\oval(28,28)[bl]}
\put(18,28){\line(1,0){140}}
\put(298,28){\vector(-1,0){140}}
\put(155,33){$\alpha$}
\put(298,42){\oval(28,28)[br]}
\put(312,42){\line(0,1){72}}
\put(54,120){\circle*{4}}
\put(96,144){\circle*{4}}
\put(108,120){\circle*{4}}
\put(228,120){\circle*{4}}
\put(252,144){\circle*{4}}
\end{picture}
\caption{Arcs on the $5$--punctured disc}
\label{fig_b5}
\end{figure}
These cannot be homotoped off each other rel endpoints,
as can be seen by applying Lemma \ref{no_digons}
to boundaries of regular neighborhoods of $\alpha$ and $\beta\cup\partial D$.
It remains to show that $\int_{\beta}\alpha = 0$.
Let $\tilde{\alpha}$ and $\tilde{\beta}$
be lifts of $\alpha$ and $\beta$ to $\tilde{D}_5$.
Each point $p$ at which $\beta$ crosses $\alpha$
contributes a monomial $\pm t^k$ to $\int_\beta \alpha$.
The exponent $k$ is such that
$\tilde{\beta}$ and $t^k \tilde{\alpha}$ cross at a lift of $p$,
and the sign of the monomial is the sign of that crossing.
We choose our lifts and sign conventions such that
the first point at which $\beta$ crosses $\alpha$
is assigned the monomial $+t^0$.
In Figure \ref{fig_b5},
the sign of the monomial at a crossing $p$ will be
positive if $\beta$ is directed upwards at $p$
and negative if $\beta$ is directed downwards at $p$.
The exponents of the monomials can be computed
using the following remark:
\begin{rem}
Let $p_1,p_2 \in \alpha\cap\beta$
and let $k_1$ and $k_2$ be the exponents of the monomials
at $p_1$ and $p_2$ respectively.
Let $\alpha'$ and $\beta'$ be the arcs from $p_1$ to $p_2$
along $\alpha$ and $\beta$ respectively
and suppose that $\alpha' \cap \beta' = \{p_1,p_2\}$.
Let $k$ be such that
$\alpha' \cup \beta'$ bounds a $k$--punctured disc.
Then $|k_2 - k_1| = k$.
If $\beta'$ is directed counterclockwise around the $k$--punctured disc
then $k_2 \geq k_1$,
otherwise $k_2 \le k_1$.
\end{rem}
One can now progress along $\beta$,
using the above remark to calculate the exponent at each crossing
from the exponent at the previous crossing.
Reading the crossings from left to right, top to bottom,
we obtain the following:
\begin{eqnarray*}
\int_{\alpha} \beta
&=& - t^{-3} - t^0 + t^1 + t^{-1} + t^{-3} \\
& &\mbox{}-t^{-1}-t^2+t^3 + t^1 +t^{-1}-t^{-2}- t^0 - t^2 + t^1 +t^{-2} \\
& &\mbox{}-t^{-1}+t^0-t^1 + t^2 - t^3 + t^2 - t^1 + t^0 - t^{-1} + t^{-2} \\
& &\mbox{}- t^1 - t^4+t^5 + t^3 + t^1 - t^0 - t^2 - t^4 + t^3 + t^0 \\
& &\mbox{}- t^1 + t^2-t^3 + t^4 - t^5 + t^4 - t^3 + t^2 - t^1 + t^0 \\
& &\mbox{}- t^2 + t^1-t^0 + t^{-1} - t^{-2} \\
&=& 0. \\
\end{eqnarray*}
Thus $\alpha$ and $\beta$ satisfy the requirements of Theorem~\ref{lp},
and we conclude that the Burau representation is not faithful for $n=5$.
The proof of Theorem~\ref{lp}
gives an explicit non-trivial element of the kernel,
namely the commutator of a half Dehn twist
about the boundary of a regular neighborhood of $\alpha$
and a full Dehn twist about the boundary of a regular neighborhood
of $\beta\cup\partial D$.
The following element of $B_5$ sends $\alpha$ to
a straight arc from $q_4$ to $q_5$:
\[
\psi_1 =
\sigma_3^{-1} \sigma_2 \sigma_1^2 \sigma_2 \sigma_4^3 \sigma_3 \sigma_2.
\]
The following element of $B_5$
sends $\beta$ to a straight arc from $p_0$ to $q_5$:
\[
\psi_2 =
\sigma_4^{-1} \sigma_3 \sigma_2 \sigma_1^{-2} \sigma_2
\sigma_1^2 \sigma_2^2 \sigma_1 \sigma_4^5.
\]
Thus the required kernel element is:
\[
[
\psi_1^{-1}
\sigma_4
\psi_1,
\psi_2^{-1}
\sigma_4 \sigma_3 \sigma_2 \sigma_1^2 \sigma_2 \sigma_3 \sigma_4
\psi_2
].
\]
This is a word of length $120$ in the generators.
The arcs in Figure~\ref{fig_b5}
were found using a computer search,
although they are simple enough to check by hand.
A similar computer search for the case $n=4$
has shown that any pair of arcs on $D_4$
satisfying the requirements of Theorem~\ref{lp}
must intersect each other at least $500$ times.
We conclude with an example of a non-trivial braid in
the kernel of the Burau representation for $n=6$
which is as simple such a braid as one could reasonably hope
to obtain from Theorem~\ref{lp}.
\begin{figure}[ht!]\small
\centering
\unitlength0.7pt
\begin{picture}(270,100)
\thinlines
\put(30,45){\oval(60,60)[t]}
\put(135,45){\oval(90,90)[t]}
\put(135,90){\vector(1,0){0}}
\put(135,95){$\beta$}
\put(240,45){\oval(60,60)[t]}
\put(45,45){\oval(90,90)[b]}
\put(60,45){\line(0,-1){15}}
\put(225,45){\oval(90,90)[b]}
\put(210,45){\line(0,-1){15}}
\thicklines
\put(30,45){\vector(1,0){105}}
\put(134,48){$\alpha$}
\put(135,45){\line(1,0){105}}
\put(30,45){\circle*{4}}
\put(60,30){\circle*{4}}
\put(120,60){\circle*{4}}
\put(150,60){\circle*{4}}
\put(210,30){\circle*{4}}
\put(240,45){\circle*{4}}
\end{picture}
\caption{Arcs on the $6$--punctured disc}
\label{fig_b6}
\end{figure}
The curves in Figure~\ref{fig_b6}
give us the braid
\[
[\psi_1^{-1}\sigma_3\psi_1, \psi_2^{-1}\sigma_3\psi_2],
\]
where
\[
\psi_1 = \sigma_4 \sigma_5^{-1} \sigma_2^{-1} \sigma_1,
\]
and
\[
\psi_2 = \sigma_4^{-1} \sigma_5^2 \sigma_2 \sigma_1^{-2}.
\]
This is a word of length $44$ in the generators.
|
\section{Introduction}
Ly$\alpha$\ absorption due to groups or clusters of galaxies has only been
detected relatively recently (Lanzetta et al.\ 1996, Ortiz-Gil et al.\
1997, Tripp et al.\ 1998, Shull et al.\ 1998). This is because it had
been difficult to identify suitable combinations of background QSOs
and foreground clusters close enough to the QSO line of sight to
produce Ly$\alpha$\ absorption in the QSO spectrum. The search
for suitable groups was poorly motivated since Ly$\alpha$\ absorption would
not be expected due to the high temperature of the intracluster
medium. Furthermore the majority of Lyman-alpha absorption line data
in the literature are at redshifts z$>1.5$ (when the line is shifted into
the optical band), where most clusters are still on the process of
formation.
Recent observations appear to produce contradictory results:
Morris et al.\ (1993) and Bowen, Blades, \& Pettini (1996) failed to
identify absorption due to clusters of galaxies, in contrast to the
results of Lanzetta, Webb \& Barcons (1996), Ortiz-Gil et al.\ (1997),
Tripp et al.\ (1998), and Shull et al.\ (1998).
However, other recent work suggests that galaxies may partially retain their
gaseous envelopes in a cluster environment (see Cayatte et al.\ 1990 for
an example in the Virgo cluster). Zabludoff \& Mulchaey (1998) find
that galaxies in poor groups ($\approx 20 - 50$ members) lie in a
common halo which contains most of the mass of the group. Blitz et al.\
(1998) suggest that the High Velocity Clouds (HVCs) detected in the
Local Group might be the counterparts of Lyman Limit systems, as they
find similar column densities and internal velocity dispersions and
subsolar metallicites. They also suggest that lower column density
HVCs may correspond to Ly$\alpha$\ clouds. In their models the HVCs trace
the distribution of dark matter in and around the group, following its
filamentary/sheet-like structure.
Lanzetta et al.\ (1996) report the identification of a group of galaxies
toward QSO 1545$+$2101 responsible for a broad absorption feature
present in an HST \sl Faint Object Spectrograph \rm (FOS) spectrum of
this object. The spectral resolution of these data was insufficient to
show whether it was the group as a whole or individual galaxies within
it which were responsible for the observed Ly$\alpha$\ absorption.
Individual galaxies would give rise to discrete absorption components
associated with particular galaxies. Also, galaxies at a smaller
impact parameter should produce higher column densities.
In this paper we present new higher resolution observations of the QSO
1545$+$2101 using the \sl Goddard High Resolution Spectrograph \rm on
HST.
Throughout this paper we have adopted a deceleration parameter
$q_0=0.5$ and a dimensionless Hubble constant $h=H_0/100 (~{\rm km \ s}^{-1}
\mbox{Mpc}^{-1})$.
\section{HST/GHRS observations of the QSO 1545$+$2101}
\label{data}
A high resolution spectrum of the QSO 1545$+$2101 was obtained using
the GHRS spectrograph with the G160M grating on August 23 1996
(Fig. \ref{fig1}). The observations were obtained during a series of
14 exposures each of 300s duration for a total exposure time of 4200s.
The individual exposures were reduced using standard pipeline
techniques and were registered to a common origin and combined using
our own reduction programs. The final spectrum was fitted with a
smooth continuum using an iterative spline fitting technique. The
spectral resolution of the final spectrum was measured to be
FWHM=$0.07$ \AA\ (or FWHM=$ 14 ~{\rm km \ s}^{-1}$), and the continuum
signal-to-noise ratio was measured to be $S/N \approx 12$ per
resolution element. Previous imaging and spectroscopic observations of
the field surrounding QSO 1545$+$2101 are described in Lanzetta et al.\
(1996).
\section{Detection and analysis of the absorption systems in the
field of QSO 1545$+$2101}
We detected all absorption lines above a significance level of $3
\sigma$ in the spectrum of QSO 1545$+$2101. The parameters
characterizing each absorption line were estimated using
multi-component Voigt profile fitting (three different and
independently developed software packages were used, the three of them
providing the same results). The values obtained are shown in Table
\ref{tab1}. The $3 \sigma$ confidence level detection limit was found
to be $0.03$~\AA\ .
Two absorption lines are found at $z=0.2504707 \pm 0.0000030$ and
$z=0.2522505 \pm 0.0000016$, with a velocity separation of $\sim 427\,
~{\rm km \ s}^{-1}$. A group of at least eight lines is also detected, the redshift
centroid of this group being $<\!\!\! z \!\!\!>= 0.2648 \pm 0.0002$ with a velocity
dispersion of $163\pm 57 ~{\rm km \ s}^{-1}$.
Galactic heavy element lines in the spectrum (C {\sc iv}
$\lambda\lambda$1548,1550 and Si {\sc ii}$\lambda$ 1526) were
identified and removed them from subsequent analyses (Fig.\ref{fig1}).
\section{Group of galaxies in the field of QSO 1545+2101}
The spectroscopic galaxy sample was selected on the basis of galaxy
brightness and proximity to the QSO line-of-sight. Although this
sample cannot be considered complete the galaxies were selected
randomly from an essentially flux-limited sample. The sample of
galaxies used in this study is given in Table \ref{tab2}. Figure
\ref{fig0a} is an image of the field. The galaxies in the group which
were observed spectroscopically are indicated in that figure. The
error in the galaxies redshift is estimated to be $ \sim 120\, ~{\rm km \ s}^{-1}$
(Lanzetta et al.\ 1995).
We detected a group of seven galaxies with redshift centroid at
$<\!\!\! z \!\!\!>=0.2645 \pm 0.0004$ and velocity dispersion of
$239\pm 90 ~{\rm km \ s}^{-1} $. Their individual impact parameters range from $7.2$
up to $456.4 \ h^{-1}$ kpc.
We searched in the archives for X-ray data on this field as
detecting X-ray emission from the group might help to
characterize it better. Unfortunately the group is so close to
the QSO (which is an X-ray source itself) that only the {\it ROSAT}
High Resolution Imager data would be of any use, and there are no such
data in the archive. QSO 1545$+$2101 was observed both with the {\it
Einstein} Imaging Proportional Counter and with the {\it ROSAT}
Position Sensitive Proportional Counter, but the extended
Point-Spread-Function of both instruments resulted in QSO emission
severely contaminating the region where the galaxy group might
emit X-rays.
In any case, from our image of this field and the fact that the
velocity dispersion that we have measured for this group is quite
small, it is clear that it is a loose association of galaxies rather
than a galaxy cluster. The maximum value of impact parameter in our
sample ($456.4 \ h^{-1}$ kpc) is also the typical physical size for
poor groups of galaxies (Zabludoff \& Mulchaey 1998). In fact, this
group is most probably the one hosting the QSO itself, as
$z_{em}=0.264 \pm 0.0003$ for this object (Marziani et al.\ 1996). One
might therefore expect this group to have only between 10-20 members
(Bahcall et al.\ 1997).
\section{The relationship between the absorption systems and the galaxies}
\label{relation}
The cluster of absorption lines present in the spectrum of QSO
1545+2101 might arise in different environments. They might be intrinsic
absorbers, arising either in the QSO region itself or in its near
environment. The similar redshifts of the QSO and the group of
absorbers in these data may point to one of these hypotheses as the
right one. In addition, QSO 1545+2101 is a radio-loud object and
there have been suggestions that intrinsic absorption or absorption
arising in the QSO host galaxy would be stronger in radio-loud QSOs
than in radio-quiet ones (Foltz et al.\ 1988; Mathur, Wilkes \& Aldcroft 1997).
But there are also some arguments against an intrinsic origin for
these lines, as discussed by Lanzetta et al.\
(1996). Associated absorbers of the kind described above produce
rather strong absorption lines, while the lines detected in QSO
1545+2101 are relatively weak. Moreover, no corresponding metal
absorption lines have been detected. With the
new data from GHRS we have more evidence against the
associated nature of the absorption systems: the agreement between
the galaxies' and absorbers' redshift centroids and between their
corresponding velocity dispersions suggests that both galaxies and
absorbers groups share---at least---the same physical location.
These characteristics lead us to reject the associated hypothesis
and to consider another possible scenario:
absorption arising in cosmologically intervening objects,
within the same group of galaxies that hosts the QSO.
To further explore this third hypothesis a demonstration of
the non-random coincidence between the galaxies and absorbers positions
in velocity space is necessary. Identifying each galaxy with a single
absorption line would also be very interesting. In what follows we
use statistical methods to address both questions.
\subsection{A cluster of absorbers arising in a group of galaxies}
\label{random}
The group of galaxies detected toward the QSO 1545$+$2101
has a mean redshift of $<z_g>=0.2645 \pm 0.0004$, compatible with the
mean redshift centroid value of the group of absorption lines,
$<z_a>=0.2648 \pm 0.0002$ (to the red of the Galactic Si {\sc ii} line
in Fig. \ref{fig1}). The absorber and galaxy velocity dispersions are
similar: $239\pm 90 ~{\rm km \ s}^{-1}$ for the group of galaxies and $163\pm 57
~{\rm km \ s}^{-1}$ for the group of absorption lines (the error in the galaxy
redshifts is $\sim 120 ~{\rm km \ s}^{-1}$ ). This strongly suggests a
connection between the absorbers and galaxies.
There is also a galaxy in this field whose redshift is $z=0.2510$.
Two Ly$\alpha$\ absorption lines are detected near this redshift at
$z=0.2504707 \pm 0.0000030$ and $z=0.2522505\pm 0.0000020$ (to the
blue of the Galactic Si {\sc ii} line in Fig. \ref{fig1}). The
galaxy-absorber velocity differences are $ \Delta v = 127\pm 120 ~{\rm km \ s}^{-1}$
and $\Delta v = 300\pm 120\, ~{\rm km \ s}^{-1}$ respectively. This implies that
this galaxy, whose impact parameter is $\rho=306.4 \ h^{-1}$ kpc,
could be responsible for one of the absorption lines as both $\Delta
v$ values are compatible with the velocity dispersions that one finds
typically in a galactic halo ($\approx 200 ~{\rm km \ s}^{-1}$). The velocity
difference between the two absorption systems ($427 ~{\rm km \ s}^{-1}$), is perhaps
too large for the same galaxy to be responsible for both of them. It
may also be that we have not observed the actual galaxy giving rise to
either of these absorption lines, since our galaxy sample is not
complete.
\subsubsection{Statistical analysis}
Two statistical tests were carried out to investigate the relationship
between the group of galaxies and absorbers. In the first, we
computed the two-point cross-correlation function ($\xi_{ag}$) between
the absorbers and the galaxies. This function was normalized by
computing the $\xi_{ag}$ expected if there were no relation between
absorbers and galaxies (derived using galaxy redshifts which are
randomly distributed over a redshift range around the real group of
absorption lines). Errors were computed using a bootstrap method,
simulating 1000 samples of 7 galaxy redshifts, each set randomly
selected from the real set, deriving error estimates from the
distribution of 1000 values of $\xi_{ag}$. The final result is
shown in Fig. \ref{fig3} (panel $a$).
In the second statistical test, we applied a similar statistical
method, not to the actual set of absorbers but to the individual pixel
intensities in the spectrum. This way we overcome any potential
problems introduced as a consequence of incorrect determination of the
true velocity structure in the profile fitting process, or the
presence of weak lines falling below the detection threshold. Recent
analyses by Liske, Webb \& Carswell (1999) show that the study of
pixel intensities is more sensitive to clustering than the usual
line--fitting techniques. The test is as follows: we evaluated for
each pixel $i$ the value of the function $g_i=1-f_i$, where $f_i$ is
the intensity of pixel $i$ normalized to the continuum. Clearly,
$g_i$ has larger values in pixels belonging to absorption lines. For
pixels corresponding to metal lines (including galactic lines) we
assigned a value of $g_i=0$. Then, for each galaxy we compute the
function $\zeta_v= \sum_{n=1}^{N} \sum_j g_{nj}$, where $g_{nj}$
refers to all the pixels $j$ located at a distance $v$ in velocity
space from galaxy $n$, $N$ being the total number of galaxies in the
sample. The velocity distances $v$ considered range from $0 ~{\rm km \ s}^{-1}$ up
to the one spanned by the whole spectrum. This function $\zeta_v$ is
analogous in some sense to the two-point cross-correlation function
computed before: high values of $\zeta_v$ at low velocity distances
reflect the tendency of the pixels corresponding to absorption lines
to lie close to the galaxies. The same function $\zeta_v$
corresponding, again, to a randomly selected sample of galaxies chosen
from a uniformly distributed group was computed in a way analogous to
the previous one and the result, after normalizing to this random
case, is shown in Fig. \ref{fig3} (panel $b$). The error bars
in $\zeta_v$ were computed using a bootstrap method as before. The
result obtained is similar to the one obtained by
computing the two-point cross-correlation function: most of the pixels
belonging to absorption lines (i.e., with larger values of $g$) lie
less than $200 ~{\rm km \ s}^{-1}$ away from the galaxies.
\subsection{Are absorption lines related to galaxies on a case-by-case basis?}
As there is a clear complex of discrete Ly$\alpha$\ lines in the spectrum of
QSO 1545+2101, we explored the possibility of a one-to-one match
between galaxies and absorbers.
A Gaussian model was assumed for the distribution of galaxies in the
group. The null hypothesis is that the galaxies are randomly drawn
from a Gaussian whose parameters are derived from the real data. The
two-point correlation functions $\xi_{ag}$ and the function $\zeta_v$
were computed in the same way as in \S \ref{random} (see Fig.
\ref{fig3}, panels $c$ and $d$). No evidence of a one-to-one
correspondence between galaxies and absorbers is found.
We can estimate the maximum statistical velocity dispersion between
individual absorbers and their galaxies of origin that would permit
the detection of a one--to--one correspondence, assuming an intrinsic
one--to--one correspondence exists. A Kolmogorov-Smirnov Monte Carlo
test showed that if the average velocity dispersion between the
galaxies and the corresponding absorbers were $\lesssim 100 ~{\rm km \ s}^{-1}$ then
the null hypothesis would be rejected ($>2 \sigma$) in $90$\% of the
cases. As a typical galactic velocity dispersion
is $\sim 200 ~{\rm km \ s}^{-1}$, this condition is not likely to be satisfied in
practice. Another possibility is to improve the statistics. From Monte Carlo
simulations we estimate that about 100 groups similar to the ones
studied here are needed for a $2 \sigma$ detection of this one--to--one
association, for a velocity dispersion between the absorber and the
galaxy of about $200 ~{\rm km \ s}^{-1}$.
\section{Discussion and conclusions}
\label{discussion}
We have detected a clump of absorption lines along the line-of-sight
towards the QSO 1545$+$2101. A group of galaxies has also been
detected, with impact parameters $\rho$ of the individual galaxies to
the QSO line--of--sight of less than $\sim 460 h^{-1}$ kpc. The group
is probably the one hosting the QSO, so we can expect it to have about
10-20 members (Bahchall et al.\ 1997).
Several scenarios might give rise to the absorption. Due to the close
redshift values of the QSO and the group of absorbers one could easily
think that they arise either in the QSO itself or in the corresponding
host galaxy. We consider that there are compelling arguments
supporting the intervening system hypothesis (see Lanzetta et al.\ 1996)
and contradicting the associated hypothesis. We now summarise those
arguments.
The velocity spanned by the group of Ly$\alpha$\ absorption lines is
consistent with the velocity dispersion of the group of galaxies. This
implies that the Ly$\alpha$\ absorbers arising in that group occupy the same
region of space as the galaxies themselves. Moreover, the spectrum of
QSO 1545$+$2101 reveals a group of \it discrete \rm Ly$\alpha$\ absorption lines
at $z \approx 0.26$. Multi-component Voigt profile fitting provides a
statistically good fit to the data, indicating that the absorption
lines arise in overdense gas regions rather than in some smoothly
distributed intragroup medium.
The average Doppler dispersion parameter of the absorption lines, {\it
b}, is measured to be $19\pm 4 ~{\rm km \ s}^{-1}$ with a dispersion of $10\pm 4
~{\rm km \ s}^{-1}$. This value is in agreement with the values measured in the low
redshift Ly$\alpha$\ forest. Therefore there is no evidence from this case
for any physical difference, in terms of the {\it b}
parameter, between Ly$\alpha$\ clouds lying within or outside of groups.
Two statistical analyses show that the distribution of galaxies with
respect to the absorbers is not random, but that is not possible to
confirm a one--to--one match due to the proximity in velocity space of
the galaxies in the groups and the uncertainty on their redshifts. A
Kolmogorov-Smirnov test showed that a small galaxy-absorber velocity
dispersion (less than $\sim 100 ~{\rm km \ s}^{-1}$) would be required to establish
a one--to--one match. As this is well below the typical values
corresponding to a galaxy potential well, another approach is
required, such as having a large enough sample of clusters or groups
of galaxies related to clusters of Ly$\alpha$\ absorption lines. All the
facts above support the idea of a physical connection between the
group of galaxies and the group of Ly$\alpha$\ absorbers.
Another piece of circumstantial evidence pointing to a one--to--one
relationship between absorbers and galaxies. This concerns the number
of each type of object detected. As mentioned before, the galaxy
group towards QSO 1545$+$2101 contains approximately 10-20 members.
According to Lanzetta et al.\ (1995), only a subset of them will be
close enough to the QSO line of sight to produce observable Ly$\alpha$\
absorption ($\rho \lesssim 160$ kpc for a covering factor of $\approx
1$). We observe eight individual components in the absorption
profile, consistent with the expectations from such a na\"{\i}ve
model. There is no obvious reason why such an agreement would be found
for some other quite different model (be it HVCs, filaments or any
other structure). We note that the absorption lines may break up into
further components at higher spectral resolution although these may
then be substructure within individual galaxies.
\vskip 1truecm
\acknowledgments
A.O.-G., K.M.L. and A.F.-S. were supported by grant NAG-53261; grants
AR-0580-30194A, GO-0594-80194, GO-0594-90194A and GO-0661-20195A from
STScI; and grant AST-9624216 from NSF. A.O.-G. acknowledges support from
a UNSW Honorary Fellowship. A.F.-S. was also supported by
an ARC grant. X.B. was partially supported by the DGES under project
PB95-0122.
\clearpage
|
\section{Introduction}
The importance of dust extinction in the Galaxy has been recognized
since early in this century when star-counting surveys revealed
absorption of optical light by `dark clouds' (Barnard 1919). It is
fortunate that extinction correlates relatively well with reddening in
the Galaxy, because it is difficult enough to accurately measure
either the distance to a typical astronomical source or its intrinsic
luminosity -- let alone both. But knowing the intrinsic color
$(B-V)_i$ (using an unobscured line of sight) along with the observed
$B-V$ color and a reddening-extinction law $A_V =
R_V[(B-V)-(B-V)_i] \equiv R_VE(B-V)$, one can correctly
determine the distance of an object from its distance modulus, or vice
versa. The value of $R_V$ varies markedly within the Milky Way ($ 3
\la R_V \la 6$; Mathis 1990) and among different galaxies ($1.5 \la
R_V \la 7.2$; Falco et al. 1999), but a value of $R_V \simeq 3.2$ is
useful for many estimates of Galactic extinction.
Radiation from extragalactic objects is subject to extinction by dust
both inside and outside of the Galaxy. However, while extragalactic
dust has received attention, our knowledge of its amount and properties is
rather limited, because methods useful for estimating dust density in
galaxies have proved less effective when extended to intergalactic
space. For example, a number of groups have attempted to measure
extinction by dust in clusters using background-object counting, and
several claims of intracluster dust (Bogart \& Wagoner 1973; Boyle,
Fong \& Shanks 1988; Romani \& Maoz 1992) and extragalactic dust cloud
detections (Wsozlek et al. 1988 and references therein) have been
made, but even now these claims remain controversial (see Maoz 1995).
The strongest proposed limits on a diffuse distribution of
intergalactic dust with a Galactic reddening law have come from
studies of the redshift evolution of the mean quasar spectral index
(e.g. Wright 1981; Wright \& Malkan 1988; Cheng, Gaskell \& Koratkar
1991). These studies limit uniform dust of constant comoving density
to have $A_V(z=1) \la 0.05$ mag (from Wright \& Malkan 1988), and are
most sensitive to dust at $z > 1.$
While our knowledge of it is poor, intergalactic dust could have great
cosmological importance, as it could affect results concerning the
cosmic microwave (CMB) and cosmic infrared (CIB) backgrounds, galaxy
and quasar numbers at high $z$, galaxy evolution, large-scale
structure, etc. This paper discusses intergalactic dust chiefly in
the context of its importance in measurements of the cosmological
deceleration parameter -- a subject discussed numerous times, first by
Eigenson (1949) and most recently in Aguirre 1999~(A99).
Conditions in the diffuse intergalactic medium (IGM) strongly disfavor
dust formation, so whatever intergalactic dust exists is probably
either the remnant of an early Population III epoch, or is formed in
galaxies and removed by some mechanism. (The remaining possibility,
that a substantial dust-forming population of extragalactic stars
exists, is not considered here.) Previous investigations of
intergalactic dust have almost invariably assumed that it has
properties similar to that of Galactic dust; but this assumption is
not well justified. Even among galaxies, $R_V$ varies by a factor of
four, and intergalactic dust may have creation, destruction, and
selection mechanisms quite different from dust in galaxies. As argued
in A99, radiation pressure ejects grains with high opacity and a broad
opacity curve more efficiently than other grain types. In
\S\ref{sec-dustdest} I discuss results suggesting that small grains
are preferentially destroyed by sputtering, both in the halos of
galaxies (during the ejection process) and perhaps in the IGM.
The large grains, while having significant mass, give very small
$E(B-V)$ reddening, and actually have higher visual opacity (per unit
mass) than dust which includes small grains. Intergalactic dust of
this type would not have been detected by quasar reddening surveys.
Current data, described in \S\ref{sec-metals}, suggests that the
universe has been enriched to $\ga 1/10$ solar metallicity before
$z\sim 0.5$. Section~\ref{sec-snae} shows that if a significant
fraction of this metal is locked in dust that is distributed fairly
uniformly, the dust extinction to $z\sim 0.5$ can explain the observed
progressive dimming of type Ia supernovae (Riess et al. 1998;
Perlmutter et al. 1999 [P99]) without cosmic acceleration.
The last section outlines ways in which the influence of the type of
intergalactic dust described here can be tested, probably most cleanly
by future supernova observations.
\section{Cosmic Metallicity}
Although still very incomplete, our understanding of the evolution of
cosmic metallicity has improved dramatically during the last few
years. Recent surveys (see e.g. Madau 1999 for a review) of galaxies
in the Hubble Deep Field show that the comoving star formation rate
(SFR) rises with redshift by a factor of ten to $z\sim 1-1.5$, past
which it either declines or levels off. Using the observed star
formation rate and an assumed ratio of metal formation to star
formation of 1/42 (Madau et al. 1996), the total cosmic metallicity
(neglecting any population III contribution) may be estimated by
integrating over time the metal formation rate. Figure~\ref{fig-sfr}
shows the result of this integration starting at $z=10$, using curves
from Madau (1999) and Steidel et al. (1999). These demonstrate that
whether or not the SFR declines for $z \ga 1.5$, the current metal
density is $\Omega_Z \sim (1.5-2)h_{65}^{-2}\times 10^{-4}.$ The
estimates shown for $\Omega_Z(z)$ agree with other results in the
literature using similar methods: Pettini (1999) estimates
$\Omega_Z(z\simeq 2.5) \sim 6\times 10^{-5}$; Cen \& Ostriker (1999b)
calculate $\Omega_Z(z\simeq2.5) \sim 2\times 10^{-5}$ and
$\Omega_Z(z\simeq0.5)\simeq 1.1\times10^{-4}$.
Very interestingly, these estimates coincide with the `fossil
evidence' presented by Renzini (1997; 1998). He argues that clusters
are essentially closed systems which contain all of the metal produced
by their stellar populations. Stars of approximately solar
metallicity comprise only a fraction $\Upsilon_{cl} \approx
0.09h_{65}^{3/2}$ of the total cluster gas mass, yet the remaining
intracluster gas has $\approx 1/3$ solar metallicity. The associated
metal production per unit of stellar mass can be written $M_Z \approx
M_*[1+3.15h_{65}^{-3/2}]Z_\odot$ (Renzini 1997). Unless stars in
clusters produce metals much more efficiently than those in field
galaxies, this figure should apply to the $\Omega_* \approx 0.004$
(Fukugita, Hogan \& Peebles 1996) of stars in the universe, giving
$\Omega_{Z} \approx 3.1 \times 10^{-4}$. Moreover, if the cosmic gas
density is $\Omega_{gas} \approx 0.05$, then the star formation
efficiency of clusters well represents that of the universe,
$\Upsilon_{IGM} \equiv {\Omega_* \over \Omega_{gas}} \approx
\Upsilon_{cl}$, further indicating that clusters are a fair sample.
Two obvious ratios link these numbers to estimates of intergalactic
dust density: the fraction $F_I$ of all metal residing in the IGM, and
the fraction $d_m$ of the intergalactic metal locked in dust. At
high $z$, the primary data bearing upon these ratios are observations
of metal lines in quasar absorption spectra. Pettini et al. (1997)
have measured the depletions of Cr and Zn in damped Lyman-$\alpha$
systems which Pei, Fall \& Hauser (1998) interpret as giving a dust/metal ratio
of $\approx 0.5$ for all $z \la 3$, similar to that of known galaxies.
This suggests that even at high $z$, source regions for the cosmic
metallicity have $d_m \approx 0.5$. Whether this applies to metal as
it reaches the IGM depends on the degree to which the ejection process
destroys dust, or selectively ejects either gaseous metal or dust (to
be discussed in~\S\ref{sec-dustrem}).
Observations of lower column density Ly-$\alpha$ absorbers give
information about the gaseous metals in relatively cool regions of the
IGM. Under some (important) assumptions about ionization corrections,
these line strengths indicates that areas with column density $N(H I)
\approx 10^{14-15}\,{\rm cm^{-2}}$ have $\sim 0.3\%$ solar
metallicity, giving $\Omega_{Z}^{ly}(z\sim3)\sim3\times 10^{-6}$ in
gaseous metal if this enrichment is uniform. Direct integration
(effectively treating $N(H I) \la 10^{14}\,{\rm cm^{-2}}$ regions as
pristine) gives $\Omega_Z^{ly}(z\sim3) \ga 3\times 10^{-7}$ (Songaila
1997). In light of the higher numbers derived from the SFR, the
assumption that most of the cosmic metallicity resides in the
Ly-$\alpha$ forest therefore leads to a `missing metals' problem, as
noted by Pettini (1999) and Renzini (1999), who argue that this
discrepancy may indicate that most metals are located in hot halos
around galaxies, in proto-clusters, or in a phase of the IGM hotter
than that which the Ly-$\alpha$ forest studies probe. Cen \& Ostriker
(1999a) reach the same conclusion on separate grounds, arguing on the
basis of both simple physical arguments and their numerical
simulations that the bulk of universal baryons at low $z$ should be
hot.
The idea that the IGM is fairly metal-rich gains more support from
arguments that hold at low redshift. If enriched to solar metallicity,
the current mass of stars and gas in known galaxies is sufficient to
contain the metals expected at $z\sim 2.5$ (if most galaxies had
already formed by then), but this does not hold for later epochs: as
also argued by Renzini (1998), the cosmic metallicity at $z=0$ derived
from cluster enrichment or by integration of the SFR cannot be
stored in currently observed galaxies, which can hold at most
\begin{equation}
\Omega_Z = 8\times 10^{-5}\left({Z_{gal}\over
0.02}\right)\left({\Omega_{gal}\over 0.004}\right)
\end{equation}
The remaining $\approx 50-75 \%$ must exist in the intergalactic gas,
in extended halos, or hidden in undetected galaxies (which seems
unlikely). This is exactly what one would expect on the basis of the
metal distribution in clusters: Renzini (1997) argues that while
intracluster gas contains about three times as much metal as the
cluster galaxies, there seems to be no compelling reason to expect
that metal ejection is much more efficient in clusters than in the
field\footnote{Even if ram-pressure stripping and/or mergers made
clusters more efficient, this would predict that field galaxies
would be about four times as metal-rich as cluster galaxies,
contrary to observations. Also, clusters with higher velocity
dispersion should have higher metallicity; Renzini finds no evidence
for this.}, so most metals created from field galaxies should also
lie outside them.
To avoid the conclusion that the universe has substantial
intergalactic metallicity, it therefore seems that one would have to
argue both that cluster galaxies eject metals more efficiently
than field galaxies and have a different IMF, and also accept
that estimates of the SFR and/or the density of metal in galaxies are
incorrect by at least a factor of two.
In summary, the density of intergalactic dust can be estimated at
$z\sim 0.5$ as
\begin{equation}
\Omega_{dust}^{igm}(z\sim0.5) = F_I\times d_m \times \chi \times
10^{-4},
\label{eq-metdens1}
\end{equation}
with likely values of $1.5 \la \chi \la 3$ and $d_m \simeq 0.5$. The
argument that stars and gas in known galaxies cannot contain these
metals gives $0.5 \la F_I \la 0.75$; the larger number also
corresponds to the value derived by assuming that field galaxies eject
metal as cluster galaxies do. The resulting estimate does not
take into account grain destruction.
\label{sec-metals}
\section{Removal of dust and metal from galaxies}
The conclusion that the IGM is fairly metal rich implies that metal
can be efficiently removed from the galaxies in which it forms.
Several ways in which galaxies can eject dust and metallic gas have
been studied in the context of clusters (for exactly the same reason),
but it rather unclear which mechanism is dominant. The density of
intracluster gas is high enough that ram-pressure stripping of
galactic gas may be efficient (e.g. Fukumoto \& Ikeuchi 1996; Gunn \&
Gott 1972) in clusters, but it would be much less effective for a
galaxy in the general IGM. The removal of gas by supernova-driven
winds has been widely discussed, and detailed simulations have been
performed investigating this effect in large starburst galaxies
(Suchkov et al. 1994) and in dwarves (Mac Low \& Ferrara 1998). It
is widely thought that galactic winds also remove gas from ellipticals
(e.g. David, Forman \& Jones 1990). Gnedin (1998) has claimed that
mergers provide the dominant metal removal mechanism, at least for $z
\ga 4.$ Finally, dust removal (without metallic gas) by radiation
pressure can be fairly efficient even for present-day spirals;
starburst galaxies with higher luminosities would be correspondingly
more effective. In this section I will concentrate on radiation
pressure as the dominant dust expulsion mechanism because the effects
on the dust have been investigated most carefully in that scenario;
but I will discuss the other mechanisms briefly.
The dynamical evolution of a dust grain in a spiral galaxy is governed
primarily by the radiation pressure, gravity, the viscous gas drag and the
magnetic Lorentz force. Starting with Chiao \& Wickramasinghe (1972), several
groups have studied these forces acting on grains in model galaxies
with some assumptions about the mass, gas, and luminosity
distributions.
Confirming the results of Chiao \& Wickramasinghe, Ferrara et al. (1990) find
that graphite grains of most sizes can escape most spiral galaxies,
and that silicate grains are marginally confined (though silicate grains
with $a\ga 0.05{\,\mu{\rm m}}$ can escape high luminosity spirals.)
Calculating the grain dynamics for two specific galactic models (of
the Milky way and NGC 3198) in more detail, Ferrara et al. (1991) find
that in the Milky Way, silicate grains of radius $a = 0.1{\,\mu{\rm m}}$ have
typical speeds of $\sim 200-600\,{\rm km\,s^{-1}}$ and reach halo
radii $\sim 100$ kpc in $\sim 150-500$ Myr. Graphite grains of $a =
0.05{\,\mu{\rm m}}$ can move approximately twice as fast. Another
investigation, by Shustov \& Vibe (1995), gives similar results. They
find that grains of size $0.07{\,\mu{\rm m}} \le a \le 0.2{\,\mu{\rm m}}$ are ejected most
effectively. Silicate (graphite) grains of $0.1{\,\mu{\rm m}}$ starting 1~kpc
above the galactic disk attain speeds of $1000\,{\rm km\,s^{-1}}$
($2000\,{\rm km\,s^{-1}}$) and reach 100 kpc in 100 (40) Myr. Shustov
\& Vibe stress that only dust starting somewhat above the disk can
escape, but this does not imply that dust ejection is inefficient: in
their model, the dust expulsion rate is up to $0.4M_\odot\,{\rm
yr^{-1}}$, which in a Hubble time exceeds the entire metal content
of the Galaxy and gives $\Omega_{dust}^{igm}$ of order $10^{-4}$, even
assuming a constant SFR. Most recently, Davies et al. (1998) have
performed numerical calculations of dust removal by radiation
pressure, taking into account the opacity of the disk. Disk opacity
reduces the radiation pressure at high galactic latitude, so Davies et
al. find dust expulsion less efficient than the previous
studies.\footnote{Of course a higher assumed intrinsic luminosity of
the galaxy would cancel this effect; Davies et al. only investigate
one mass-to-light ratio.} Nevertheless, Davies et al. predict the
removal of (at least) $0.1{\,\mu{\rm m}}$ graphite grains from their model
galaxy if there is fairly low disk opacity. Smaller grains are
expelled much less efficiently. Notably, Davies et al. still
estimate that up to $90\%$ of the dust formed in spirals may be
ejected, and even calculate an estimate of $\sim 1$ mag of
intergalactic extinction across a Hubble distance.
All studies of radiation pressure driven dust removal have noted the
importance of magnetic fields on grain dynamics but have (effectively)
neglected them in their calculations, with the following
justifications given:
\begin{itemize}
\item{Magnetic field lines are only sometimes
parallel to the disk, and only on large scales (all studies).}
\item{Grains are charged only part of the time (Ferrara et al. 1991; Davies
et al. 1998).}
\item{Radiation pressure can enhance Parker (1972) instabilities
that can lead to open field configurations with lines perpendicular to
the disk. (Chiao \& Wickramasinghe 1972; Ferrara et al. 1991).}
\end{itemize}
An additional justification is empirical: dust is actually observed
outside the disks of galaxies (Howk \& Savage 1999; Alton, Davies \&
Bianchi 1999; Davies et al. 1998 and references therein; Ferrara et
al. 1999 and references therein). However the dust escapes the disk,
its presence proves that while magnetic fields are potentially
important and currently impossible to model in detail, they cannot be
perfectly effective at dust confinement. On the other hand, this does
not prove that dust can fully decouple from the gas.
While dust expelled by radiation pressure could not carry a
significant gas mass with it, other metal expulsion mechanisms
probably remove dust along with gas. Alton et. al. (1999) have
presented observations of dust outflows in three nearby starburst
galaxies, concluding that the `superwinds' driving these outflows can
impart near-escape velocity on the dust, and that up to $10\%$ of the
dust mass of these galaxies could be lost in the observed outflows
alone. These results lend observational support to the numerical
simulations of Suchkov et al. (1994) which predicted such outflows,
and further demonstrate that dust can escape along with metallic gas.
Lehnert \& Heckman (1996) have estimated the efficiency of metal
removal by winds in starburst galaxies using a large sample, and find
that galaxies could enrich the IGM to $\Omega_Z^{igm} \sim 5 \times
10^{-5}$. This figure assumes a constant SFR and the authors estimate
that the enrichment is likely to be ten times higher with a more
realistic SFR.
Supernova-driven winds might also eject
dust from ellipticals. Vereshchagin, Smirnov \& Tutukov (1989)
estimate that the ratio of galactic wind force to gravitational force
on a grain of radius $a$ is
$$F_W/F_G = {3\alpha \over 16\pi G}{V_W/a},$$
where $V_W$ is the wind
velocity and $\alpha$ is the specific mass ejection rate for the wind
in the galaxy. For $\alpha \sim 10^{-19}\,{\rm s^{-1}}$ (applicable
for the present epoch; Vereshchagin et al. 1989; David et al. 1990)
and relatively slow winds ($V_W \sim 10-60\,{\rm km\, s^{-1}}$;
Vereshchagin et al. 1989), only very small grains escape; but for
starburst ellipticals, David et al. find rates of at least $5\times
(10^{-18}-10^{-17})\,{\rm s^{-1}}$ for the first $10^8\,{\rm yr}$ of
starburst activity, implying that the winds dominate gravity in grain
dynamics for grains up to $\sim 0.05-2.5{\,\mu{\rm m}}$, even for winds too
slow to escape the galaxy themselves. Essentially the same argument
would apply to grains subject to winds in spirals, so whether dust
lies in relatively cool outflows or is exposed to the wind itself, it
is difficult to see how it could avoid being driven out with the
metallic gas.
Gnedin (1998) has performed high resolution
cosmological simulations that suggest that mergers are the dominant
metal removal mechanism, at least at high ($z \ga 4$) redshift. This
mechanism would eject gas with roughly the same dust/gas ratio as the
source galaxy.
This brief survey of metal ejection mechanisms suggests that it is
difficult to efficiently remove metallic gas from galaxies without
also removing dust (although the converse of this would not be true if
radiation pressure is the dominant ejection mechanism). Rough
estimates of the ejection efficiencies show that metal ejection rates
sufficient to account for the enrichment of the IGM estimated in
\S\ref{sec-metals} are reasonable. While dust probably accompanies
gas as it leaves galaxies (or leaves by itself), studies of clusters
show that intracluster gas is not dust-rich. The next section
addresses the probable cause of this disagreement: grain destruction
during the ejection process and in the IGM.
\label{sec-dustrem}
\section{Destruction of small grains}
A key result of the investigations by both Ferrara et al. (1991) and
Shustov \& Vibe (1995) is that grain destruction due to sputtering by
hot halo gas is relatively insignificant for grains of $a \sim
0.1{\,\mu{\rm m}}$ but very effective for grains with $a\sim 0.01{\,\mu{\rm m}}.$ While
$0.1{\,\mu{\rm m}}$ grains lose only $\sim 0.005{\,\mu{\rm m}}$ in radius, the small
grains are completely destroyed on a timescale of $\sim 500$ Myr. The
sharp difference arises because for sputtering at fixed gas
temperature and grain velocity the destruction timescale $(1/a)(da/dt)
\propto 1/a$ (Draine \& Salpeter 1979a), and because small grains are
more affected by gas drag yet less propelled by radiation pressure,
hence move more slowly through the halo.\footnote{Size effects can be
even stronger; Draine \& Salpeter (1979b) find that the most
efficient dust destruction, sputtering in the `inter-cloud medium',
is $\sim 500\,\times$ more effective in $0.01{\,\mu{\rm m}}$ grains than for
$a=0.1{\,\mu{\rm m}}.$} Shustov \& Vibe conclude from their calculations that
the grains escaping intact from galaxies will have sizes $0.03{\,\mu{\rm m}} \la
a \la 0.2{\,\mu{\rm m}}$ for graphite particles and $0.07{\,\mu{\rm m}} \la a \la 0.2{\,\mu{\rm m}}$
for silicate particles.\footnote{These numbers are rather approximate
because the authors computed results only for six grain radii.}
These conclusions depend on assumptions about the density,
temperature, and extent of the galactic halos, but the fact that both
groups obtain similar results suggests that the minimal grain size
surviving expulsion is probably of order $a_{min} \sim 0.05{\,\mu{\rm m}}.$
The efficiency of dust destruction in other metal removal processes
has not been calculated in detail and is difficult to estimate. Dust
driven out by winds would be vulnerable to sputtering by the halo gas
as well as by the faster moving wind, though it will be somewhat
shielded if embedded in cool clumps of gas. To estimate the effect of
the wind, let us assume a mass loss rate $\dot M$ (in solar masses/yr)
due to a wind leaving the galaxy radially with velocity $V_W$. The
effect of this wind would be similar to the effect of a hot gas of
temperature $T_W \equiv m_p V_W^2/2k$ and (proton) number density $n_p
\sim 3\dot M / 16\pi R^2 V_W m_p$. For $125\,{\rm km\,s^{-1}} \la v
\la 4000\,{\rm km\,s^{-1}}$ and $R \ga 10$ kpc, this gives $10^6\,{\rm
K} \la T \la 10^9$ K and
\begin{equation}
n_p \la 1.9 \times 10^{-4}\,\dot M
\left({V_W \over 125\,{\rm
km\,s^{-1}}}\right)^{-1}\,{\rm cm^{-3}}.
\end{equation}
Using Draine \& Salpeter's (1979a) sputtering rate for graphite
in this temperature range,
this corresponds to a lifetime of
$$
\tau_W \ga (7-16) \times 10^7\,\dot M^{-1} \left({V_W
\over 125\,{\rm km\,s^{-1}}}\right)\left({a\over0.01{\,\mu{\rm m}}}\right)\,{\rm yr}.
$$
This is comparable to the ejection timescale, so this sputtering
could be important but is unlikely to completely destroy the dust.
Grain-grain collisions provide another important dust destruction
mechanism in galaxies and might be important in the early stages of
the ejection process. For example, grain-grain collisions in
supernova shocks can efficiently shatter large grains into smaller
ones (e.g. Jones, Tielens \& Hollenbach 1996), so if supernova blowout
removes dust, there is a danger that shocks from the same supernovae
might shatter the large grains before the dust is expelled. Shocks
may also play an important role in mergers. On the other hand, it is
unclear whether the dust observed in the ISM is representative of
dust which has just formed, or already been shock-processed, or some
steady-state between the two. Specifically, there is evidence for the
formation of large grains in novae (Shore et al. 1994), and possibly
in supernovae (see Wooden 1997 and Pun et al. 1995), and grains are
presumably larger in molecular clouds where high values of $R_V$ are
measured. It may be, then, that pre-shock grains tend to be somewhat
larger, and the MRN distribution is more characteristic of grains
after significant shattering has occurred. The assumption of this
paper is that dust leaving its progenitor galaxy will have an grain
size distribution characteristic of dust in the ISM. In the absence
of significant shattering, this is probably conservative, since a
significant fraction of dust is contained in dense clouds with high
$R_V$ (e.g. Kim, Martin \& Hendry 1994).
\label{sec-dustdest}
\subsection{Dust Destruction in the IGM}
Rather little is known about the destruction of dust in the IGM.
Schmidt (1974) estimates that soft cosmic rays would provide the most
efficient destruction, but cannot determine whether or not the
destruction time would exceed the Hubble time; moreover, Draine \&
Salpeter (1979b) find that cosmic rays are unimportant dust destroyers
in the Galaxy (where they should be at least as effective as in the
IGM). The hot gas component of the IGM, however, could sputter grains
effectively, even at low density. Using again Draine \& Salpeter's
(1979a) estimate\footnote{More recent calculations by Tielens et al.
give similar sputtering rates for carbon at $T \ga 10^7{\rm\,K}$, while their rates
are somewhat higher for silicates and somewhat lower in both
materials for $10^6{\,\rm K} < T < 10^7\,$K.} , the lifetime can be
written
\begin{equation}
\tau \approx (3.5-9)\times10^9\,\left({a\over0.01{\,\mu{\rm m}}}\right)
h_{65}^{-2}\Omega_{gas}^{-1}
\delta^{-1}(1+z)^{-3}\,{\rm yr},
\end{equation}
where $\Omega_{gas}$ signifies the hot gas density in critical units and
$\delta$ is a clumping factor. The Hubble time (for $\Omega=1$) is
$H^{-1}(z) = 1.6 \times 10^{10}h_{65}^{-1}(1+z)^{-3/2}\,{\rm yr}$,
suggesting the efficient destruction of grains for which
\begin{equation}
Q \equiv 0.14\left({a\over0.01{\,\mu{\rm m}}}\right)^{-1}\left({\Omega_{gas}\over
0.05}\right)h_{65}\delta(1+z)^{3/2} \gg 1.
\end{equation}
The clumping factor (i.e. the overdensity felt by a `typical' grain)
is quite uncertain, but the simulations of Cen \& Ostriker (1999b),
which numerically track the distribution of metallicity, indicate that
at $z\sim 0.5$, the mean universal metallicity approaches the
metallicity of $\delta \sim 100$ regions. Regions of much higher
overdensity do not have much higher metallicity and hence cannot
contain most of the metals -- for example, $\delta \sim 1000$ regions
have only about twice the metallicity, so dense `subregions' can
contain only about 20\% of the metals in $\delta \sim 100$ regions.
If a `typical' grain experiences $\delta \sim 100$, then $Q \sim 26$
for $0.01 {\,\mu{\rm m}}$ grains and $Q \sim 2.6$ for $0.1{\,\mu{\rm m}} $ grains. This is
suggestive (but {only} suggestive) that sputtering by hot
intergalactic gas might provide yet another mechanism by which grains
of $a \la 0.1{\,\mu{\rm m}}$ might be selectively destroyed.
Finally, note that the low mean dust density in the IGM and in
extended galaxy halos would strongly suppress the grain-grain
collisions thought to shatter large grains into small ones in the
galaxy\footnote{Equation~\ref{eq-galopt} gives the dust optical depth
through the halo of a galaxy. The `optical depth' to an emerging
grain would be of similar magnitude, so grain-grain collisions are
probably unimportant unless high-$z$ galaxies are all heavily
obscured by dust in their halos.}; since dust formation is also
inefficient in the IGM there is probably no source of {new} small
grains outside of galaxies.
The efficiency of dust destruction depends in a rather complicated way
on the environment; moreover the type and details of the dominant
mechanism of metal ejection for galaxies are uncertain. Thus the
arguments of this section are intended merely to make plausible the
chief {assumption} of this paper, which is that grains of size $a \la
0.05-0.1 {\,\mu{\rm m}}$ are removed (either by destruction or by failure to
escape their progenitor galaxies) from the grain-size distribution
characterizing dust outside of galaxies, whereas larger grains are
not.
\section{Density of Surviving Intergalactic Dust}
The estimate of the density of intergalactic dust in section
\S\ref{sec-metals} did not take into account dust destruction or the
preferential expulsion of dust. Lets us assume that a mass fraction
$(1-f_{esc})$ of dust is destroyed as it leaves the disk and/or
traverses the halo, and that a further fraction $(1-f_{igm})$ is
destroyed in the IGM after the dust escapes the halos but before $z
\sim 0.5$. There are three general scenarios indicated by the dust
ejection and destruction mechanisms outline above:
\begin{enumerate}
\item{Dust and gas leave together, with the galactic dust/metals ratio
of $\approx 0.5$. A fraction $f_{esc}f_{igm}$ of this survives,
so that $d_m \approx 0.5 f_{esc}f_{igm}$. This scenario predicts a
high enrichment of the IGM near galaxies and perhaps a substantial
density of metal in galactic halos.}
\item{Dust, driven by radiation pressure, decouples from the gas
either in the disk or in the inner halo, but is partially
destroyed. The gaseous metal (both destroyed dust and metal which
escapes the disk but then decouples from dust) could (a) remain in
the halo or could (b) return to the disk to form more dust,
repeating the process. In the former case galactic halos may be
highly enriched and $d_m \approx 0.5 f_{esc}f_{igm}$; in the
latter (unlikely) case galaxies should be very deficient in metals
which are easily incorporated into dust, and $d_m \sim f_{igm}$ is
possible.}
\item{Gaseous metals leave disks but dust remains ($d_m \ll 1$).
While unlikely, this would lead to a highly enriched IGM and/or
halo gas component but little intergalactic obscuration (like the
case $f_{esc}f_{igm} \ll 1$). Disks would be heavily enriched
with elements that {do} form dust.}
\end{enumerate}
Assuming that $0.5 \la F_I \la 0.75$, equation~\ref{eq-metdens1} gives
\begin{equation}
\Omega_{dust}^{igm}(z\sim0.5)
\sim (4-11) \times10^{-5}f_{esc}f_{igm}
\label{eq-metdens2}
\end{equation}
for scenarios 1 and 2a. The dust density would be higher by a factor
of up to $\sim 2/f_{esc}$ for scenario 2b, and very small for 3.
Since small grains are preferentially destroyed, but probably cannot
be created in halos and in the IGM, $f_{esc}f_{igm}$ effectively
determines the minimal grain size $a_{min}$.\footnote{Sputtering will
also change the upper grain-size cutoff, in effect shifting the
whole distribution toward smaller radii. The neglect of this effect
may be justified by the excess of large grains over the MRN
prediction indicated by estimates of the actual grain-size
distribution (see \S\ref{sec-othermodels}), and because sputtering
may be more selective in destroying small grains than the $1/a$
behavior would imply.} The next section discusses the dependence of
dust properties on $a_{min}$, and gives corresponding values of
$f_{esc}f_{igm}$.
\label{sec-dustdens}
\section{Properties of the dust}
The absence of $a \la 0.1{\,\mu{\rm m}}$ grains would have important
implications for the properties of intergalactic dust. The most
commonly used model for Galactic dust is the two component Draine \&
Lee (1984; DL) model consisting of silicate and graphite spheres with
a distribution in radius (as proposed by Mathis, Rumpl and Nordsieck
1977; MRN) of $N(a)da \propto a^{-3.5},\ 0.005{\,\mu{\rm m}} \le a \le
0.25{\,\mu{\rm m}}$. After synthesizing dielectric functions for both graphite
and `astronomical silicate' and assuming a silicate/graphite mass
ratio $\sim 1$, DL demonstrated that the resulting model fits both the
observed opacity and polarization over a wide wavelength range
($0.1{\,\mu{\rm m}} \la \lambda \la 1000{\,\mu{\rm m}}$), most notably fitting the
observed features at $0.2175{\,\mu{\rm m}}$ and $10{\,\mu{\rm m}}$. This paper employs
the DL model not because it is most likely to be correct, but because
there is little agreement as to what the correct grain model might be. The
DL model is widely used and familiar, and hopefully (but by no means
certainly) captures the essential features of the dust. Other models
are discussed briefly below.
An interesting aspect of the MRN distribution is that while the
geometrical cross section ($\propto a^2$) is dominated by small-radius
grains, the mass ($\propto a^3$) is dominated by grains of large
radii. Thus, removing the very small grains can affect the opacity
curve dramatically, without radically changing the total dust mass.
Figure~\ref{fig-opacs} shows the
extinction curve for silicate and graphite with the MRN size
distribution over $a_{min} \le a \le a_{max}$ for $a_{min} = 0.005,
a_{max}=0.1$ and $a_{min} = 0.1, a_{max}=0.25$. These curves use
publicly available extinction data calculated using the method of Laor
\& Draine (1993). With the very small grains gone, the graphite
absorption curve becomes quite flat out to $\lambda \sim 1{\,\mu{\rm m}}.$
Figures~\ref{fig-props} and~\ref{fig-reds} show the extinction,
reddening and mass fraction (relative to the full MRN distribution) of
dust distributions with various value of $a_{min}$. Curves are given
both for (rest-frame) $E(B-V)/V$ reddening concentrated at one
redshift, and for a cosmological dust distribution (as described
in~\S\ref{sec-snae}). These show that even in the (more reddening)
integrated extinction, for $a_{min} \ga 0.06{\,\mu{\rm m}}$, graphite grains
give very little $(B-V)/V$ reddening. Silicate grains do not become
grey for $a_{min} \la 0.2$, but the combined silicate+graphite
reddening falls by 50\% for $a_{min} \ga 0.09{\,\mu{\rm m}}$. Moreover, this
large change in the reddening behavior of the dust does not require a
large change in the mass: these `grey' dust distributions contain
$40-55\%$ of the mass of the MRN distribution.
\label{sec-dustmod}
\subsection{Other Dust Models}
The above conclusions, based on the assumption that dust is
characterized by the DL model, may not hold for other
dust models. Mathis \& Whiffen (1989) have proposed that galactic
grains are composites of very small ($a \la 0.005{\,\mu{\rm m}}$) silicate,
graphite and amorphous carbon particles. These composite grains have
a filling factor of $\sim 0.2-1$ and corresponding maximal size $\sim
0.9 - 0.23{\,\mu{\rm m}}$. Sputtering would be effective at destroying all
sizes of low filling-factor composite grains, since both gas drag
(slowing the grains) and sputtering would be much more effective than
in comparably sizes solid spheres. Also, sputtering might tend to
`cleave' large, filamentary grains into smaller ones. Large
filling-factor particles in this model would be much like the Draine
\& Lee model, although the optical properties of the composite
materials would differ from those of pure graphite or silicate.
Several core-mantle grain models have also been proposed; see e.g.,
Duley, Jones \& Williams (1989) and Li \& Greenberg (1997). The
latter model assumes a three-component model: large silicate
core-organic refractory mantle dust, very small carbonaceous grains,
and polyaromatic hydrocarbons (PAHs). The latter two components would
presumably be destroyed as dust leaves the galaxy, leaving the large
core/mantle grains. Li \& Greenberg take the size distribution of
these grains as Gaussian, strongly dominated by $\sim 0.1{\,\mu{\rm m}}$ grains,
with parameters chosen to fit the observed extinction curve. Such a
distribution would be insignificantly affected by removal of the small
or large-size portions, so intergalactic dust would have properties
exemplified by the large core/mantle grains (these grains will redden
less than the full three-component model, but only slightly). On the
other hand, it seems that there are good reasons to expect a power-law
grain size distribution (Biermann \& Harwit 1979; Mathis \& Whiffen
1989). It would be interesting to investigate whether the model of Li
\& Greenberg could accommodate a power law distribution (as they
assume for the PAHs and the very small grains). The Duley, Jones \&
Williams (1989) model assumes a bimodal grain-size distribution: small
silicate core/graphite coated grains provide UV extinction and the
$0.2175{\,\mu{\rm m}}$ bump, whereas an MRN distribution of cylindrical silicate
grains provides extinction in the IR, with grey extinction in UV and
optical. In this model, intergalactic dust (composed of the large
silicate grains) would be significantly more grey than galactic dust,
assuming that it can escape.
Fractal grains (e.g. Wright 1987) and needles (e.g. A99 and references
therein) provide another possible dust component. Needles and
platelets have been observed in captured dust (Bradley, Brownlee \&
Veblen 1983), and might explain `very cold' dust in the ISM (Reach et
al. 1995). As argued in A99 these grains redden very little
(especially if graphitic) and absorb with high efficiency, hence would
be preferentially ejected by radiation pressure. Along the same
lines, DL grains must be at least somewhat elongated in order to
correctly predict polarization. Elongated grains are somewhat more
grey than spherical grains of the same mass, giving some additional
support to the general assumption that there is a significant grey
sub-component to interstellar dust.
While a different grain model might predict a different effect of
destroying small grains, it is also true that any viable grain model
must be capable of accommodating values of $R_V \ga 6$, since such
values are in fact observed. Large grains seem to be a necessary
component of grain models which match the observed extinction laws
(Kim et al. 1994; Zubko, Krelowski \& Wegner 1996, 1998), and a
greater fraction of these large grains in some regions is probably
responsible the high observed values of $R_V$ in those regions. It
is, then, unlikely that the destruction of very small grains will make
any model {more} reddening, so the assumption of the DL model seems at
least qualitatively safe. The MRN grain-size distribution is probably
safe for the same reasons, and very likely even conservative, in the
sense that inversions of dust opacity curves into grain-size
distributions tend to lead to more large grains than MRN would predict
(Kim et al. 1994), and the grain-size distribution gleaned from
observations by the Ulysses and Galileo satellites (Frisch et al.
1999) shows many large grains up to $1{\,\mu{\rm m}}$ or more in radius.
\label{sec-othermodels}
\section{The supernova results}
Dust of the DL model with $a_{min} \sim 0.1$ would correspond to a
dust survival fraction $f_{esc}f_{igm} \sim 0.4$. Using
equation~\ref{eq-metdens2}, this gives $\Omega_{dust}^{igm} \approx
(1.5- 4.5) \times 10^{-5}.$ This amount of dust would be quite
important cosmologically. Measurements of the redshift-magnitude
relation of type Ia supernovae (P99; Riess et al. 1998) show
statistically significant progressive dimming of supernovae which
has been interpreted as evidence for acceleration in the cosmic
expansion. This section discusses grey intergalactic dust (as
specified in \S\ref{sec-dustmod}) in the context of the these
observations. Only the DL grain model is considered here.
Both supernova groups find that after calibration using a low-$z$
sample, the supernova at $z\sim0.5$ have magnitudes indicative of
acceleration in the cosmic expansion. The best fit (for a flat
cosmology with cosmological constant) of P99 is $\Omega=0.28,
\Omega_\Lambda=0.72$; the results of R99 are similar. The necessary
extinction to account for these results in an $\Omega=0.2$ open
universe (see Fig.~\ref{fig-depth}) is $A_V(z=0.5) \approx 0.15-0.2$
mag.\footnote{ $A_V \approx 0.2$~mag accounts for all of the effect,
but $A_V \approx 0.15$~mag puts an $\Omega=0.2$ open universe within
the stated $1-\sigma$ contour.} An $\Omega=1$ universe requires
$A_V(z=0.5) \approx 0.4$ mag.
Riess et al. argue that grey extinction would cause too much
dispersion in the supernova magnitudes to be compatible with their
observations if the dust is confined to spiral galaxies. Perlmutter
et al. derive from their data an intrinsic dispersion $\Delta$ at
$z\sim 0.5$ almost identical to that at $z\sim 0.05$:
$$\Delta(z\sim0.05) = 0.154\pm0.04,\ \ \ \Delta(z\sim0.5) =
0.157\pm0.025.$$
This suggests that the processes dominating the
intrinsic dispersion do not change significantly in magnitude from low
to high redshift. However, note that -- assuming that the errors as
well as the dispersions add in quadrature -- the amount of {
additional} dispersion $\Delta_{add}$ at high $z$ formally allowed
within the stated errors of is $\Delta_{add} \la 0.13$ mag. This does
not include any systematic errors in the estimation of the intrinsic
dispersion.
P99 also investigate the mean color difference between
the high- and low-$z$ samples, finding
$\langle E(B-V)\rangle_{z\sim 0.05} = 0.033\pm0.014$ and $\langle
E(B-V)\rangle_{z\sim 0.5} = 0.035\pm0.022.$
Again, this suggests that
a systematic effect (in color) is not large, but nevertheless the
errors allow a color difference of
$$\langle E(B-V)\rangle_{z\sim 0.5} - \langle E(B-V)\rangle_{z\sim
0.05} \la 0.03\,{\rm mag}.$$
In addition, this comparison is subject
to a systematic uncertainty of $\approx 0.03$ mag resulting from the
conversion of (observed) $R$ and $I$ magnitudes into rest-frame $B$
and $V$ magnitudes.
To place tighter constraints on systematic reddening, Perlmutter et
al. construct an artificially blue subsample of the high-$z$ points
which is unlikely to be redder (in the mean) than the low-$z$ sample.
The change in fitting that this elimination produces then gives an
indication of systematic extinction by reddening dust.
Perlmutter et al. use this method to place a strong constraint of
$\delta A_V \la 0.025$ mag on the effect of any extinction which (a)
exists at high $z$ but not at low $z$, (b) dominates the dispersion of
both the color and extinction, (c) has a reddening-extinction relation
$R_V$ up to twice that of the Galaxy and (d) occurs in a flat
universe. Assumption (b), unstated in P99, is crucial but seems
unfounded. The limit on a systematic increase in dispersion indicates
that systematic extinction must be a sub-dominant component of the
total computed `intrinsic' dispersion in magnitude; this holds also
for dispersion in color. In this case, removing the reddest
supernovae will not preferentially remove more obscured supernovae,
even if dust accounts for the whole effect at high $z$. Thus the
stated (more stringent) limits on systematic reddening do not apply,
{as long as} the dispersion in brightness and/or color is dominated by
factors other than extinction. Furthermore,
if the assumption of flatness is dropped, the elimination of the seven
reddest supernovae actually changes the fit considerably, in the
direction of an open universe (P98, Table 3, Figure 5c). The shift
corresponds to $\delta A_V \approx 0.07$ mag at z=0.5.
The required extinction and the limits on reddening and dispersion can
now be compared to that expected from intergalactic dust.
Figure~\ref{fig-reds} shows the reddening $R_V$ for graphite and
silicate dust of the DL model, assuming an MRN distribution over
$a_{min} \le a \le 0.25{\,\mu{\rm m}}.$ The extinction in Figure~\ref{fig-props}
is integrated to $z=0.5$ for an $\Omega=0.2$ universe, assuming a
constant comoving dust density of $\Omega_{dust}^{igm}=10^{-5}$ in
each component. The results show that, for example, a distribution
with $a_{min} = 0.1{\,\mu{\rm m}}$, $\Omega_{dust}^{igm}(z=0.5) \approx 4 \times
10^{-5}$ (total) and equal mass density of silicate and graphite
grains (approximately the ratio derived by Draine \& Lee) provides
sufficient extinction to account for the type Ia supernova results.
The induced reddening is $0.025$ mag, comparable to the allowed
reddening due to {either} random {or} systematic errors. Most
of this reddening is provided by the silicate grains, so the 1:1 ratio
is conservative; the real ratio should be biased toward the less
efficiently destroyed (Draine \& Salpeter 1979a) and possibly more
efficiently ejected (assuming radiation pressure expulsion) graphite
grains. The large grains contain $\sim 40\%$ of the full MRN
distribution. Larger values of $a_{min}$ provide less reddening but
values of $a_{min} \ga 0.15$ probably contain too little mass to be
viable in explaining the supernova data. Graphite grains alone (if
silicate grains were preferentially destroyed) with $a_{min} \ga 0.06$
(giving $f_{esc}f_{igm} \la 0.6$) and $\Omega_{dust}^{igm} \sim 3
\times 10^{-5}$ would produce similar effects.
The amount of dispersion induced by the dust is very important but can
be estimated only roughly. Assuming that the dust is uniformly
distributed in randomly placed spheres of radius $R$ with number
density $n$, the dispersion $\Delta$ is given approximately by
$\Delta/A_V(z=0.5) \approx N^{-1/2}$, where $N$ is the number of
spheres intersected by a typical path, and can be written $N \simeq
n\pi R^2D$, where $D\approx 2400h_{65}^{-1}\,$Mpc is the distance to
$z=0.5$ in an $\Omega=0.2$ universe.
Now consider galaxies with $n = 0.008h_{65}^3(1+z)^3\,{\rm
Mpc^{-3}}$ (Lin et al. 1996). For $N^{1/2} \ga 1$ this implies $R
\ga 70h_{65}^{-1}[(1+z)/1.5]^{-3/2}\,{\rm kpc}$. Escape velocities from
spirals are $\ga 250\,{\rm km\,s^{-1}}$, so the dispersion in
integrated optical depth due to dust ejected by radiation pressure or
winds and traveling away from the disk for time $\tau_{esc}$ is
\begin{equation}
\Delta \sim A_V h_{65}^{-1}\left({\tau_{esc} \over 270\,{\rm
Myr}}\right)^{-1}\left({v_{dust}\over 250\,{\rm
km\,s^{-1}}}\right)^{-1}.
\end{equation}
Figure~\ref{fig-depth} shows the
(small) difference in optical depth between the total dust
distribution and the dust which has existed for $> 200$ Myr and $> 1$
Gyr, demonstrating that the dispersion induced by grey dust created
but not yet sufficiently dispersed would be small. Large-scale correlations
between galaxies are unlikely to be important in this analysis. The
dispersion in density on 8 Mpc scales is $\sim 1$, but $\sim 300$ such
domains lie between here and $z\sim 0.5$, leading to $\Delta \sim
0.2/\sqrt{300} = 0.01$.
The best way to estimate dispersion (currently underway) would
probably be to measure the optical depth through random lines of sight
piercing a high-resolution cosmology simulation which either tracks
metals (Cen \& Ostriker 1999b; Gnedin 1998) or allows some
prescription relating gas density to dust density in the IGM. Short
of this, I note that Cen \& Ostriker's simulations show that $\delta
\sim 100$ is characteristic of the bulk of the metal-rich gas. This
corresponds to $R \ga 670h_{65}^{-1}(1+z)^{-1}$ kpc for the spheres
considered above. Fairly uniform regions of this size and number
density would give little dispersion.
If $R \la 70$~kpc, most extinction takes place in a small number of
clumps, in particular in the halo of the supernova host galaxy. For
dust of density $\Omega_{dust} = \chi\times 10^{-5}$ uniformly
distributed in radius $R=\xi \times 100\,{\rm kpc}$ halos of galaxies
with $n = 0.008h_{65}^3(1+z)^3\,{\rm Mpc^{-3}}$, the extinction to the
galaxy center is
\begin{eqnarray}
A_V &=& {1.086\kappa_V\Omega_{dust}\rho_c\over {4\over 3}\pi R^2n} \\ \nonumber
&=& 0.04 \left({\kappa_V \over 5 \times 10^{4}\,{\rm cm^2\,g^{-1}}}\right)
h_{65}^{-1}\chi\xi^{-2}, \\ \nonumber
\label{eq-galopt}
\end{eqnarray}
which gives the required extinction for $\xi \approx 1$ and $\chi
\approx 4.$ As long as $\xi \ga 0.1$ there will not be large
dispersion due to different galactic radii in the supernovae, but
inhomogeneities in the dust distribution in the halo could be
important. The supernova results could could be explained by such
halos with a smaller $\Omega_{dust}$, but this would require that the
halos are somewhat larger at low $z$, and that all halos at $z\sim
0.5$ have similar column density through them. This scenario does not
seem to be as natural an explanation as a more uniform dust
distribution, but it remains a possibility.
Leaving aside these uncertainties, the essential result of my
calculation is that a truncated-MRN distribution of Draine \& Lee dust
with $a_{min} \ga 0.1$, uniformly distributed and with
$\Omega_{dust}^{igm}(z=0.5) \simeq 4\times 10^{-5}$ can account for the
supernova dimming in an $\Omega=0.2$ universe without excessive
reddening. Whether the induced dispersion in extinction is too large
depends crucially on the (uncertain) distance to which ejected dust
can escape from the galaxies in which it forms, and on the clumpiness
of the resulting distribution.
\label{sec-snae}
\section{Cosmic backgrounds}
An intergalactic dust distribution of the magnitude required to
explain the supernova results would have other cosmological
implications. For instance, any intergalactic dust component will
absorb energy from the optical/UV background and re-emit the energy in
the FIR/microwave. Calculation of the evolution of the cosmic mean
density field shows (Aguirre \& Haiman, in preparation) that the dust
considered in this paper will not lead to measurable CMB spectral
distortions,\footnote{Intergalactic dust in the calculations of Loeb
\& Haiman (1997) and Ferrara et al. (1999) has low temperature, but
both papers assume a UV/optical background lower than recent
detections indicate.} but instead adds (significantly) to the CIB.
The FIRAS distortion limits can, however, limit dust types with high
FIR emissivity, since these have lower equilibrium temperatures.
Note also that while the SFR estimates correct for dust extinction
(e.g. Madau, Pozetti \& Dickinson 1998; Pettini 1999 and references
therein), these corrections would not account for intergalactic dust.
Intergalactic absorption of $\sim 0.1 - 1$ mag at $z \sim 0.5-5$ (see
fig.~\ref{fig-depth}) would imply an SFR -- and hence metal density --
a factor of $\sim 2$ higher than that given in~\S\ref{sec-metals}.
\section{Testing for intergalactic dust}
Future observations of supernovae can investigate the importance of
intergalactic dust in two ways. First, accurate high-$z$ observations
in rest-frame $R$-band or longer wavelengths should reveal the dust
with properties of the model developed here. As shown in
fig.~\ref{fig-reds}, the $E(B-R)/B$ reddening to $z=0.5$ is $\approx
0.25$ for both pure graphite dust (with $a_{min} = 0.06{\,\mu{\rm m}}$) or for
silicate+graphite (with $a_{min}=0.1{\,\mu{\rm m}}$). Therefore high-$z$
supernova should have $E(B-R)$ values 0.05 mag higher than the low-$z$
sample. The effect in $B-I$ would be even stronger. Non-spherical
grains (of which the needles of A99 are an example) could provide grey
opacity into $R$-band and beyond, but are constrained limits on
far-infrared/microwave emission (Aguirre \& Haiman, in preparation).
Figure~\ref{fig-depth} includes the deviation of the fit of R99
from the $\Omega=0.2$ model (dashed line). The difference
between this line and the dust optical depth then indicates the
deviation at high $z$ of the dust model from the cosmological constant
($\Omega_\Lambda = 0.72$) model; at $z=1.5$ the difference is $\simeq
0.2-0.3$ mag. This possible difference should be testable once a
significant number of $z > 1$ supernovae are measured. A large number
of supernovae is important both because the dispersion is comparable
to the effect being measured, and because it is crucial to be sure
that the $z>1$ sample is statistically {complete} -- a worry that
is less significant at $z\sim0.5$ since the claimed effect is of
dimming rather than brightening.
An intergalactic dust distribution would also dim other distant
objects. Dust has been proposed several times as an explanation for
the dropoff in quasar number counts for $z\ga 3$ (e.g. Ostriker \&
Heisler 1984; Wright 1986). While this point is still controversial,
observation of the dropoff in radio-selected quasars (Shaver et al.
1996) shows this explanation to be unlikely. The models of this paper
would not predict the dropoff (unless substantial population III dust
exists also) since $A_V(z) \la 1$ for all $z$. Galaxy counts, in
contrast to quasar studies, show an excess in $B$-band counts at
high $z$ as compared to some evolution models (see Shimasaku \&
Fukugita 1998 for a summary). For $B$-band observations probing
$z\sim 0.5\pm0.3$, this `excess' amounts to a deviation of up to $\sim
1$ mag from the predicted curve. The dust extinction of $\la 0.4$ mag
for $z \la 1$ would enlarge this discrepancy, but not greatly. This
holds for $I$ and $K$-band surveys as well. Because of uncertainties
in the models at high $z$, Shimasaku \& Fukugita are reluctant to draw
conclusions about cosmological parameters from the galaxy counts. But
with rapid progress in the field this may soon become a useful test of
the dust model.
Several claims of dust detection in clusters have been made, but
remain controversial. The model of A99 predicted that cluster dust
should be more grey than galactic dust; the arguments of this paper
reinforce this prediction. The intracluster gas destroys dust
efficiently; since sputtering destroys small grains more efficiently
than large grains, it is likely that whatever grains survive in
clusters are large, and thus supply grey opacity (this may help
explain the controversy surrounding the existence on intracluster
dust). Determination of the extinction curve of dust in clusters (if
dust indeed exists) would be an important test of the key idea of the
intergalactic dust model, although grey dust in clusters would not
necessarily imply that dust in the diffuse IGM is also grey.
Dust confined to galactic halos might also be detectable by its IR
emission or by studies of objects seen through the halo. Zaritsky
(1994) has presented preliminary evidence for halo dust in NGC 2835
and NGC 3521 at $\sim 60\,$kpc using $B-I$ reddening; if confirmed
this would be an important dust component, as the inferred mass is
large.
\section{Conclusions}
As discussed in \S\ref{sec-metals}, fairly strong arguments suggest
that the universe currently has $\Omega_Z \approx (1.5-3) \times
10^{-4}$ in metals. This metal cannot all be contained in the stars and
gas in known galaxies, so unless metals are well hidden in unobserved
galaxies, a metal density $\Omega_Z^{igm} \approx (0.75-2.25) \times 10^{-4}$
should exist in the more uniform IGM or in extended halos. Moreover,
most of this metal was probably in place by $z\sim0.5$.
Some fraction $d_m$ of this metal must be in dust. There are several
good reasons to expect that the grain size distribution of this dust should
be different than for dust in galaxies: (1) Ejection by radiation
pressure favors higher opacity, less reddening grains (A99). Within
the DL model, this means large grains are preferentially ejected. (2)
As dust leaves galaxies (where it is almost certainly created), small
grains are preferentially destroyed by sputtering. This preferential
destruction occurs also in the IGM, but the efficiency is rather
uncertain. (3) Small grains are generally assumed to form from the
shattering of larger grains. This shattering will not occur in the
IGM due to the low densities, nor can small grains grow from the
vapor.
The principal assumption of this paper (supported by what detailed
calculations are available) is that very small dust grains leaving the
galaxy -- comprising a fraction $(1-f)$ of the total dust density --
are removed from the grain-size distribution, while large grains are
not. Dust as modeled by DL, with grains of radius ($\la 0.1{\,\mu{\rm m}}$)
removed ($f \approx 0.4$), reddens very little yet has a visual
opacity $\kappa \approx 5\times 10^4\,{\rm cm^2\,g^{-1}}$. Uniformly
distributed dust of constant comoving density in an $\Omega=0.2$ open
universe provides an extinction to $z=0.5$ of
\begin{equation}
A_V \simeq 0.15h_{65}\left({\Omega_{dust}^{igm} \over 4\times 10^{-5}}\right)
\left({\kappa\over 5\times 10^4\,{\rm
cm^2\,g^{-1}}}\right),
\end{equation}
hence it can
account for the dimming of type Ia supernovae at $z\sim 0.5$ in a way
fully consistent with observations. The fact that the expected dust
density of $0.5 f \Omega_Z^{igm} \simeq (0.25-2.25)\times 10^{-4} \times
(0.5) \times (0.4) = (1.5-4.5) \times 10^{-5}$ is so close to
the density required is very interesting.
According to these estimates (see also~\S\ref{sec-metals} and \S\ref{sec-dustdens}) it is
possible, but rather unlikely, that there is sufficient dust
($\Omega_{dust}^{igm}(z=0.5) \sim 9 \times 10^{-5}$) to allow
compatibility with a closed, matter-dominated universe.
The arguments of this paper show that there may be an intergalactic
dust component which is important yet has evaded earlier attempts at
its discovery; from this point of view the supernova observations are
telling us not about cosmic acceleration, but about cosmic opacity.
If the arguments of this paper are substantially correct, an
intergalactic dust distribution of {some} magnitude is inevitable, so
it is worth pointing out the chief caveats (which maintain rough
consistency with other observations) under which the obscuration would
not be sufficient to to account for the supernova dimming:
\begin{itemize}
\item{Both the SFR argument and the cluster enrichment argument may
predict cosmic metallicity several times the true value. In this
case most metal could be locked in galaxies. Taking
this view, however, probably also requires one to assume that cluster
galaxies both have a different IMF and eject metals more
efficiently than field galaxies do.}
\item{A large population of unobserved but compact objects such a low
surface brightness galaxies could contain a large fraction of the
cosmic metallicity. This would, however, also fail to explain the
cluster observations.}
\item{The Draine \& Lee model may not accurately characterize dust, so
the effect of small grain destruction on the opacity curve may be
significantly different than assumed here. But the alternative
grain model would then probably have to exclude the grey
sub-component which most current grain models include, yet still
explain the values $R_V \ga 6$ observed in some regions.}
\item{Dust of all sizes may be efficiently destroyed in the diffuse
IGM or as it leaves galaxies. While total grain destruction does
not seem to be quantitatively supported in the DL model, it is
certainly plausible, and would probably be the case if grains are
very fluffy, high filling-factor composites.}
\end{itemize}
It is unlikely that the issue can be cleanly decided on the basis of
the above points or by the current supernova data (i.e. limits on
dispersion or reddening); on the other hand, future supernova results
can rule decisively. Statistically robust deviation of the
magnitude-redshift curve of $z>1$ supernovae from the dust prediction
would argue strongly for the interpretation of the dimming as cosmic
acceleration and for the relative unimportance of grey intergalactic
dust. Clear evidence of, for example, systematic (rest frame) $B-R$
reddening would argue strongly for dust.
If future observations show that intergalactic dust is indeed
important, observations of type Ia supernovae will be seen to have not
only determined the deceleration parameter (once dust is accounted
for), but to have discovered a component of the universe that will
have important implications for many future observations at high
redshift.
\acknowledgements
I thank David Layzer, George Field, Andrea Ferrara, Zoltan Haiman,
Chris Kochanek, Pat Thaddeus, Bob Kirshner, Eliot Quataert and Saurabh
Jha for helpful discussions. Communications from Ari Laor, Anthony
Jones, Tom Yorke, and an an anonymous referee were helpful and
appreciated. I thank also Bruce Draine for making publicly available
the dust data used, and I am grateful to Bill Press for financial
support. This work was supported in part by the National Science
Foundation grant no. PHY-9507695.
\newpage
|
\section{Introduction}
String theory in the previous decade raised some expectations about the
nature of geometry at very small distance scales. Because strings have a
finite intrinsic length scale $l_s$, it may not be possible to observe
distances smaller than $l_s$. Thus if one uses only strings as probes of
short distance structure, the conventional ideas of general relativity
break down at lengths of the order of $l_s$. This is exemplified through
string modified uncertainty relations \cite{ven} which yield an absolute
lower bound on the measurability of lengths in the spacetime. The spacetime
coordinates thus become smeared out and at short distances the notion of a
``point'' becomes meaningless.
Another piece of evidence is the $T$-duality symmetry of strings compactified
on a circle $S^1$ of radius $R$ \cite{duality}. This is a quantum symmetry
which maps the string theory onto one with target space the circle of dual
radius $\tilde R=l_s^2/R$, and at the same time interchanges the momenta of the
strings with their winding numbers around the $S^1$ in the spectrum of the
quantum string theory. Because of this symmetry, the moduli space of string
theories on $S^1$ is parametrized by radii $R\geq l_s$, and very small circles
are unobservable since the corresponding string theory can be mapped onto a
completely equivalent one living on a very large $S^1$. This leads to the
notion of quantum geometry, defined to be the appropriate modification of
classical
general relativity implied by string theory.
The uncertainty relations tell us that spacetime at very small distances
should be thought of as a quantum object. The appropriate mathematical
arena for the study of such ``pointless'' geometry is the theory of
algebras started by von Neumann, which in more modern times has developed
into noncommutative geometry \cite{connes}. Noncommutative geometry
presents an alternative, algebraic approach to the study of Riemannian
geometry and its generalizations, such as those hinted by string theory. In
this paper we shall discuss the applications of noncommutative geometry
towards a systematic development of the notion of quantum geometry.
Despite the implications on short distance structure presented by string
theory, many questions have been answered using only classical geometry. This
has occured in part because of the extremely rich mathematical structures
embedded into string compactifications which enables one to develop to a great
extent effective field theories on moduli spaces. However, the effective field
theories hide the true internal symmetries of string theory, and to study the
internal Kaluza-Klein spaces
the notion of spacetime geometry needs a drastic modification. This has become
especially clear over the last few years when it has been realized that the
low energy effective field theory for D-branes in string theory has
configuration space which is described in terms of non-commuting, matrix valued
spacetime coordinate fields \cite{wittenp}. This has led to, among other
things, the Matrix Theory conjecture \cite{bfss}, which proposes a light-cone
frame description of M-theory in terms of the Hamiltonian dynamics of
D0-branes. This rich structure indicates that some sort of generalization of
geometry is needed to describe the internal degrees of freedom implied by
D-branes and M-theory, and indeed it has been shown recently that
noncommutative geometry is the natural setting in which to study toroidal
compactifications of Matrix Theory \cite{cds}.
In this paper we shall review the formulation of quantum geometry through
the techniques of noncommutative geometry, based mostly on the approach
developed in \cite{fg}--\cite{size} that constructs a ``space'' in which
string duality is naturally realized as a true geometric symmetry. Starting
from a brief review of the ideas from noncommutative geometry that we shall
need, we shall describe the construction of the Fr\"ohlich-Gaw\c{e}dzki
geometry \cite{fg} which is based upon the algebraic properties of vertex
operator algebras. We will then describe how string duality naturally leads
to the quantum geometry of classical spacetimes within this framework, and
how the dualities manifest themselves as internal gauge symmetries of the
noncommutative geometry \cite{lscmp,lsplb,lscsf,ss}. This latter property
is formalized by a remarkable connection between string geometry and the
noncommutative torus \cite{lls}, which also allows us to relate this
worldsheet approach to the target space descriptions using Matrix Theory
compactifications.
\section{Spectral Triples in Noncommutative Geometry}
Noncommutative geometry is the study of geometric spaces (and their
generalizations) using algebras of fields defined on them. In this article we
shall discuss how to describe stringy spacetime as a noncommutative geometry,
and how the symmetries of the theory (such as $T$-duality and spacetime
diffeomorphisms) are realized as gauge transformations. The starting point is
to discuss an algebraic framework for ordinary {\it commutative} geometry.
Usually a compact Riemannian manifold $M$ is characterized as a topological
space on which
locally it is possible to introduce points $x\in M$ characterized by a finite
number of real numbers $x^i\in\reals$. Distances in $M$ are determined by the
metric of the space,
\begin{equation}
ds^2=g_{ij}(x)~dx^i\,dx^j
\label{ds2}\end{equation}
via the formula for geodesic length
\begin{equation}
d(x,y)=\inf_{\gamma_{x,y}}\,\mbox{$\int_{\gamma_{x,y}}$}\,ds
\label{dxy}\end{equation}
where $\gamma_{x,y}$ is a path from the point $x$ to the point $y$ in $M$.
There is a dual description of the topology and differentiable structure of a
smooth manifold which is provided by the $*$-algebra
$\alg=C^\infty(M,\complexs)$ of smooth complex-valued functions
$f:M\to\complexs$ (This algebra can be thought of as the algebra ``generated''
by the points of $M$). The completion of this algebra is the {\it commutative}
$C^*$-algebra $C^0(M,\complexs)$ of continuous complex-valued functions on $M$
with the $L^\infty$-norm
\begin{equation}
\|f\|_\infty=\sup_{x\in M}|f(x)|
\label{Linfnorm}\end{equation}
The algebra $C^0(M,\complexs)$ encodes all of the information about the
topology of the
space through the continuity criterion. Thus, in general, given a topological
space one may naturally associate to it an abelian $C^*$-algebra. That the
converse is also true is known as the Gel'fand-Naimark theorem \cite{fd}.
Namely, there is an isomorphism between the category of {\it Hausdorff}
topological spaces $M$ and the category of {\it commutative} $C^*$-algebras
$\alg$. The Gel'fand-Naimark functor is constructed by using the fact that
given an abelian $C^*$-algebra $\alg$, it is possible to reconstruct a
topological space $M$ as the structure space of characters of the algebra, i.e.
the $*$-linear multiplicative functionals $\chi:\alg\to\complexs$. Points $x\in
M$ are then obtained via the identification
\begin{equation}
\chi_x(f)=f(x)~~~~~~,~~~~~~\forall f\in\alg
\label{chardef}\end{equation}
and the topology is obtained in an unambiguous way from the notion of pointwise
convergence (in the usual topology of $\complexs$). Note that for a commutative
algebra a character is the same thing as an irreducible representation of
$\alg$.
What this all means is that the study of the properties of topological
spaces can be substituted by a purely algebraic description in terms of
abelian $C^*$-algebras. A {\em noncommutative space} is then obtained by
replacing $C^0(M,\complexs)$ by some non-abelian $C^*$-algebra. In that
case, not all irreducible representations of the algebra are
one-dimensional and the identification of ``points'' becomes ambiguous. But
as we shall see, the purely algebraic approach of noncommutative geometry
is particularly well-suited to describe the intrinsic symmetries of string
theory. Note that the $*$-algebra $C^\infty(M,\complexs)$ also encodes
all of the information about the differentiable structure of a manifold $M$
through the smoothness criterion. In what follows it will suffice to have a
description in terms of only a dense subalgebra of a given $C^*$-algebra.
The metric aspect and other geometrical properties are introduced into this
framework by using the fact that {\it any} $C^*$-algebra can be represented
faithfully and unitarily as a subalgebra of the algebra ${\cal B}({\cal
H})$ of bounded operators acting on some separable Hilbert space $\cal H$
\cite{fd}. In the following often we will not distinguish between the
abstract algebra $\alg$ and its representation $\pi(\alg)$ (the norm on
$\alg$ is thus always understood as the operator norm and the
$*$-involution as Hermitian conjugation). An ``infinitesimal length
element'' is introduced by the relation
\begin{equation}
ds^{-1}=D
\label{dsD}\end{equation}
where $D$ is a (not necessarily bounded) operator on $\cal H$ which is called a
generalized Dirac operator \cite{connes} and which satisfies the following
properties:
\begin{itemize}
\item{$D=D^\dagger$ (this ensures positivity $ds^2\geq0$)}
\item{$[D,f]\in{\cal B}({\cal H})~~\forall f\in\alg$}
\item{$D$ has compact resolvent}
\end{itemize}
A metric space structure is then given by the Connes distance function on the
structure space of $\alg$,
\begin{equation}
d(x,y)=\sup_{f\in\alg\,:\,\|[D,f]\|\leq1}|\chi_x(f)-\chi_y(f)|
\label{ncdxy}\end{equation}
The generalized Dirac operator also enables one to introduce concepts from
ordinary differential geometry. Using $D$ one may define a representation of
abstract differential one-forms as
\begin{equation}
\pi_D(f\,dg)=f\,[D,g]
\label{piDdef}\end{equation}
and similarly one can introduce a representation of higher degree forms (In
that case a quotient space must be considered in order to eliminate the
so-called junk forms) \cite{connes}. Gauge theories are also readily
generalized to this setting. Generally, within this algebraic framework a
vector bundle
corresponds to a finitely-generated projective module over the algebra $\alg$
(the corresponding space of smooth sections when $\alg=C^\infty(M,\complexs)$).
However, in the following we shall consider, for simplicity, only the case of
trivial bundles. A gauge potential $A$ is then a one-form of the type
\eqn{piDdef} and a connection is provided via the Dirac operator through the
definition of a gauge covariant derivative
\begin{equation}
D_A=D+A
\label{DAdef}\end{equation}
The gauge group is defined as the group of unitary elements of the algebra,
\begin{equation}
{\cal U}(\alg)=\left\{u\in\alg~|~u^\dagger u=uu^\dagger=\id\right\}
\label{unitary}\end{equation}
and gauge transformations are the inner automorphisms of the algebra, i.e.
the maps $g_u:\alg\to\alg$ which act as conjugation by a unitary element
$u\in{\cal U}(\alg)$,
\begin{equation}
g_u(f)=ufu^\dagger
\label{gu}\end{equation}
The usual ingredients of a gauge theory on a manifold may then be summarized as
the following algebraic elements:
\begin{itemize}
\item{Connections: $A=\sum_nf_n\,[D,g_n]$, $f_n,g_n\in\alg$}
\item{Curvature: $F=[D,A]+A^2$}
\item{Bosonic Action: ${\int\!\!\!\!\!\!-}~F^2$ (where ${\int\!\!\!\!\!\!-}$ is a regularized trace, e.g.
the Dixmier trace \cite{connes})}
\item{Fermionic Action: ${\int\!\!\!\!\!\!-}~\overline{\psi}\,D_A\psi$}
\end{itemize}
The set of three ingredients $(\alg,{\cal H},D)$, i.e. a $*$-algebra $\alg$ of
bounded operators acting on a separable Hilbert space $\cal H$ and a
generalized Dirac operator $D$ on $\cal H$, is called a spectral triple (or a
Dirac K-cycle). The
spectral triple encodes all the topological and geometrical information about a
Riemannian manifold. But notice that its construction is made with no reference
to any underlying space, and that it also applies to generic (not necessarily
commutative) algebras. We shall now describe a number of examples.
\subsection*{Spin Manifolds}
If $M$ is a compact spin manifold, we take
\begin{eqnarray}
\alg&=&C^\infty(M,\complexs)\nonumber\\{\cal
H}&=&L^2(M,S)\nonumber\\D&=&i\,\gamma^i\,\nabla_i
\label{spinman}\end{eqnarray}
where $L^2(M,S)$ is the Hilbert space of square integrable spinors on $M$ and
the algebra $\alg$ acts diagonally on $\cal H$ by pointwise multiplication.
Here $\gamma^i$ are the usual Dirac matrices generating the Clifford algebra
$\{\gamma^i,\gamma^j\}=2g^{ij}$ of $M$ and $\nabla=d+\Gamma$ is the usual
covariant derivative constructed from the spin connection of $M$. The invariant
line element of $M$ is now represented as the free massless fermion propagator
on $M$ and the distance function \eqn{ncdxy} is
\begin{equation}
d(x,y)=\sup_{|\nabla f|\leq1}|f(x)-f(y)|
\label{dxyspin}\end{equation}
Note that the distance formula \eqn{dxyspin}, which is defined in terms of
complex-valued functions on $M$, is dual to the geodesic distance formula
\eqn{dxy}, which is defined in terms of arcs connecting $x$ to $y$ in $M$.
Notice also how the Riemannian geometry of $M$ is naturally encoded within the
definition of the Dirac operator.
The action functionals described above corresponding to a gauge theory
constructed from this spectral triple yield the usual one for electrodynamics
on the manifold $M$, with gauge group the unimodular loop group
$C^\infty(M,S^1)$ of $U(1)$ gauge transformations on $M$. Note that the
spectral triple \eqn{spinman} is that which naturally arises from quantizing
the free geodesic motion of a test particle on $M$. Then $\alg$ is the algebra
of observables, $\cal H$ is the Hilbert space of physical states, and the
Hamiltonian $H=-D^2$ is the
Laplace-Beltrami operator of $M$. Thus ordinary (commutative) spaces can be
thought of as those probed by quantum mechanical test particles. It is in this
way that noncommutative geometry may be thought of as ``quantum geometry''.
\subsection*{Morita Equivalence}
The natural extension of the previous example is
\begin{eqnarray}
\alg&=&C^\infty(M,\complexs)\otimes M_N(\complexs)\nonumber\\{\cal
H}&=&L^2(M,S)^{\oplus N}\nonumber\\D&=&i\,\gamma^i\,\nabla_i\otimes I_N
\label{spinmanN}\end{eqnarray}
where $M_N(\complexs)$ is the algebra of $N\times N$ complex-valued matrices
and $I_N$ is the $N\times N$ identity matrix. The gauge group is now the group
of $U(N)$ gauge transformations $C^\infty(M,U(N))$ on $M$ and the spinors of
$\cal H$ transform in the vector representation of $U(N)$. Since the algebra
$\alg$ of matrix-valued functions on the manifold $M$ is noncommutative, the
Gel'fand-Naimark theorem does not apply to this case and it is not possible to
formally identify ``points'' of a space. In fact, there is an $N$ dimensional
sphere of pure states at each point (corresponding to the various unitary
equivalent representations) which can be thought of as an internal Kaluza-Klein
isospin
space with points connected by $U(N)$ gauge transformations. Nevertheless, it
is clear that the configuration space of the quantum theory corresponding to
\eqn{spinmanN} is still the manifold $M$. The choices of abelian subalgebras of
$M_N(\complexs)$ create $N$ copies of the same manifold connected by gauge
transformation.
The apparent paradox is resolved by noticing that the algebra $M_N(\complexs)$
has only one non-trivial irreducible representation as a $C^*$-algebra. Thus
the spectral triples \eqn{spinman} and \eqn{spinmanN} both determine the same
space. This phenomenon is captured formally by saying that the algebras
$C^\infty(M,\complexs)$ and $C^\infty(M,\complexs)\otimes M_N(\complexs)$ are
Morita equivalent. A $C^*$-algebra $\cal B$ is Morita equivalent to a
$C^*$-algebra $\alg$ if it is isomorphic to the algebra ${\rm End}_\alg^0({\cal
E})$ of compact endomorphisms of some $\alg$-module $\cal E$. This means that
the two algebras become isomorphic upon tensoring them with the algebra of
compact operators. Morita equivalent $C^*$-algebras have equivalent
representation theories. They therefore differ only in the structure of their
internal spaces (i.e. their gauge symmetries). The action functionals
corresponding to \eqn{spinmanN} yield the usual massless $U(N)$ gauge theory on
the manifold $M$.
\subsection*{The Two-sheeted Spacetime}
Let us now consider the spectral triple \cite{standard}
\begin{eqnarray}
\alg&=&C^\infty(M,\complexs)\otimes\zeds_2\nonumber\\{\cal
H}&=&L^2(M,S)\otimes\zeds_2\nonumber\\D&=&\pmatrix{i\,\gamma^i\,\nabla_i&m\cr
m^*&i\,\gamma^i\,\nabla_i\cr}
\label{2sheet}\end{eqnarray}
where $m$ is a fermion mass. The bosonic action functional corresponding to
\eqn{2sheet} gives not only the usual Yang-Mills term, but also the Higgs
potential with its biquadratic form. This (commutative) space therefore gives a
geometrical origin for the Higgs mechanism associated with the spontaneous
breaking of the $U(1)\times U(1)$ gauge symmetry of $\alg$ down to $U(1)$
corresponding to the diagonal projection of $\alg$ onto
$C^\infty(M,\complexs)$.
\subsection*{The Standard Model}
The previous example can be generalized to the noncommutative geometry
\cite{standard}
\begin{eqnarray}
\alg&=&C^\infty(M,\complexs)\otimes[\complexs\oplus\quater\oplus
M_3(\complexs)]\nonumber\\{\cal
H}&=&L^2(M,S)\otimes[\complexs
\oplus\complexs^2\oplus\complexs^3]
\nonumber\\D&=&\pmatrix{i\,
\gamma^i\,\nabla_i\otimes I_3&{\cal M}
\cr{\cal M}^\dagger&i\,\gamma^i\,\nabla_i\otimes I_3\cr}
\label{standard}\end{eqnarray}
where $\quater$ is the algebra of quaternions and $\cal M$ is the $3\times3$
fermion mass matrix. The unimodular group of the algebra $\alg$ is the familiar
gauge group $C^\infty(M,U(1)\times SU(2)\times SU(3))$ of the standard model
and $\cal H$ represents the physical Hilbert space of six generation fermions.
The action functional consists of the Yang-Mills action and the Higgs term. The
spectral triple \eqn{standard} thereby shows how electroweak and chromodynamic
degrees of freedom are induced by the geometry involving a discrete internal
Kaluza-Klein space.
\subsection*{The Noncommutative Torus}
The spectral triples we have considered thus far all have the property that
they contain an underlying ordinary geometry, i.e. their algebras have the form
$\alg=C^\infty(M,\complexs)\otimes\alg_F$ where $\alg_F$ is a finite
dimensional algebra. We now turn to a genuine example of a noncommutative
geometry, in which it is impossible to think of ``points'', that will turn out
to play a prominant role in the string theory applications.
Consider a $d$ dimensional torus $T_d=\reals^d/2\pi\Gamma$, where $\Gamma$ is a
Euclidean lattice of rank $d$ with bilinear form $g_{ij}$. Let $\omega^{ij}$ be
a real-valued antisymmetric matrix. We define an algebra $\alg^{(\omega)}$ with
generators $U_i$, $i=1,\dots,d$, and relations
\begin{eqnarray}
U_iU_i^\dagger&=&U_i^\dagger U_i~=~\id\nonumber\\U_iU_j&=&{\rm e}^{2\pi
i\omega^{ij}}\,U_jU_i
\label{nctorusrels}\end{eqnarray}
A generic ``smooth'' element $f$ of the completion $\alg_\infty^{(\omega)}$ of
$\alg^{(\omega)}$ is a linear combination of monomials in the $U_i$,
\begin{equation}
f=\sum_{p\in\Gamma^*}f_p\,U_1^{p_1}U_2^{p_2}\cdots U_d^{p_d}
\label{smoothelts}\end{equation}
where $\Gamma^*$ is the dual lattice to $\Gamma$ and $f_p$ are elements of the
Schwartz space ${\cal S}(\Gamma^*)$ of sequences of rapid decrease. The
abstract algebra $\alg^{(\omega)}_\infty$ can be given a concrete
representation as a quantum deformation of the algebra
$C^\infty(T_d,\complexs)$ of functions on the torus. The product of two
functions $f,g\in C^\infty(T_d,\complexs)$ is now given by
\begin{equation}
(f\star_\omega g)(x)=\exp\left(i\pi\omega^{ij}\,\mbox{$\frac\partial{\partial
x^i}\,\frac\partial{\partial x'^j}$}\right)f(x)g(x')\Bigm|_{x'=x}
\label{defproduct}\end{equation}
which is just the usual rule for multiplying Weyl ordered symbols of quantum
mechanical operators. With the product \eqn{defproduct} the generators in
\eqn{nctorusrels} are just the usual basic plane waves
\begin{equation}
U_i={\rm e}^{ix^i}
\label{planewaves}\end{equation}
and the expansion \eqn{smoothelts} can be thought of as a generalized Fourier
series expansion.
It follows that $\alg_\infty^{(0)}\cong C^\infty(T_d,\complexs)$. For
$\omega^{ij}\neq0$, the algebra $\alg_\infty^{(\omega)}$ represents the
quotient of the ordinary torus by the orbit of a free particle in it whose
velocity vector forms an angle $\omega^{ij}$ with respect to cycles $i$ and $j$
of $T_d$. When the $\omega^{ij}$ are all rational numbers, the algebra
$\alg_\infty^{(\omega)}$ is therefore Morita equivalent to
$C^\infty(T_d,\complexs)$. When the $\omega^{ij}$ are irrational, the orbits
are dense and ergodic and the resulting quotient space is not a conventional
Hausdorff manifold. In any of the cases, an invariant integration may be
introduced via the unique trace ${\int\!\!\!\!\!\!-}:\alg_\infty^{(\omega)}\to\complexs$
which
is given by the classical average
\begin{equation}
{\int\!\!\!\!\!\!-}\,f=\int_{T_d}\prod_{i=1}^d\frac{dx^i}{2\pi}~f(x)=f_0
\label{trace}\end{equation}
and a Dirac operator may be introduced via the natural set of linear
derivations $\Delta_i:\alg_\infty^{(\omega)}\to\alg_\infty^{(\omega)}$,
$i=1,\dots,d$, defined by the logarithmic derivatives
\begin{equation}
\Delta_i(U_j)=\delta_{ij}\,U_j
\label{linderiv}\end{equation}
One aspect of this example which will be particularly important for us is the
set of Morita equivalence classes of noncommutative tori \cite{riefsch}. From
\eqn{nctorusrels} it follows that the noncommutative tori with deformation
parameters $\omega^{ij}$ and $\omega^{ij}+\lambda^{ij}$, with $\lambda^{ij}$
any antisymmetric integer valued matrix, are the same. This symmetry is part of
a larger group $O(d,d;\zeds)$ which parametrizes the Morita equivalence
classes. It acts naturally on the deformation matrix $\omega$ as (upon
picking a suitable basis of $\reals^{d,d}$)
\begin{eqnarray}
&\omega\to\omega^*=(A\omega+B)(C\omega+D)^{-1}~~~~~~,\nonumber\\&{\rm
with}~~\pmatrix{A&B\cr C&D\cr}\in O(d,d;\zeds)
\label{moritaom}\end{eqnarray}
where $A,B,C,D$ are $d\times d$ integer valued matrices which satisfy the
relations
\begin{eqnarray}
A^\top C+C^\top A&=&0~=~B^\top D+D^\top B\nonumber\\A^\top D+C^\top B&=&I_d
\label{ABCDrels}\end{eqnarray}
Later on we will see
that for toroidally compactified string theory, Morita equivalence of the
corresponding noncommutative tori is precisely the same notion as target space
duality on the associated toroidal string background.
\section{Noncommutative String\\ Spacetimes}
We now seek some sort of algebra which describes the ``noncommutative
coordinates'' of spacetime as seen by strings. The structure should be such
that at very large distance scales (much larger than $l_s$), where the
strings effectively become point particles which are well described by
ordinary quantum field theory, we recover a usual (commutative) spacetime
manifold $M$ as described in section 2. The elegant proposal of Fr\"ohlich
and Gaw\c{e}dzki \cite{fg}, which as we will see naturally incorporates
duality as a gauge symmetry of the corresponding spectral triple, is to
take $\alg$ to be the vertex operator algebra of the underlying worldsheet
conformal field theory of the string theory. Vertex operators describe
interactions of strings and they operate on the string Hilbert space as
insertions on the worldsheet corresponding to the emission or absorption of
string states. They therefore form the appropriate noncommutative algebra
which describes the quantum geometry of the ``space'' of interacting
strings. In this section we shall construct the spectral triple associated
with a toroidally compactified bosonic string theory. The theory of vertex
operator algebras \cite{fgr,voa} has a
distinguished place in mathematics (having connections with the theory of
modular functions, the Monster sporadic group, etc.). We shall start by
giving a very brief general overview of the definition and properties of a
vertex operator algebra.
\subsection*{Vertex Operator Algebras}
Any conformal field theory naturally has associated to it two chiral algebras
${\cal E}^\pm$ which form the operator product algebras of the
(anti-)holomorphic fields of the theory. They contain two mutually commuting
representations of the infinite dimensional Virasoro algebra (generating the
conformal invariance of the theory)
\begin{equation}
\left[L_k^\pm,L_m^\pm\right]=(k-m)L_{k+m}^\pm+\mbox{$\frac
c{12}$}\,(k^3-k)\delta_{k+m,0}
\label{virasoro}\end{equation}
where $c$ is the central charge of the string theory. The algebras ${\cal
E}^\pm$ act densely on the Hilbert spaces ${\cal H}^\pm$, of left and right
handed string states respectively. The vertex operator algebra is
constructed using the operator-state correspondence of local quantum field
theory. Namely, to each state $\psi^{(\pm)}\in{\cal H}^\pm$ there
corresponds a chiral vertex operator $V_\pm(\psi^{(\pm)};z_\pm)$, where
$(z_+,z_-)$ are local coordinates on a Riemann surface (here we assume that the
worldsheet is the Riemann sphere). For a bosonic string theory, the vertex
operators can be expanded as Laurent series
\begin{equation}
V_\pm(\psi^{(\pm)};z_\pm)=\sum_{n\in\zed}\psi_n^{(\pm)}\,z_\pm^{-n-1}
\label{VLaurent}\end{equation}
The vertex operator algebra $\alg$ is now characterized by some algebraic
relations which can be summarized as follows. The commutation relations are
determined by the braiding relations
\begin{eqnarray}
&
&\left[V_\pm(\psi_I^{(\pm)};z_\pm)\,
V_\pm(\psi_J^{(\pm)};w_\pm)\right]_{\gamma^\pm_{z,w}}\nonumber\\
& &=\sum_{K,L}\left(R^\pm\right)_{IJ}^{KL}~
V_\pm(\psi_K^{(\pm)};w_\pm)\,V_\pm(\psi_L^{(\pm)};z_\pm)\nonumber\\& &
\label{braid}\end{eqnarray}
where $R^\pm$ are called braiding matrices and $[\,\cdot\,]_{\gamma_{z,w}^\pm}$
is the operator which exchanges the two points $z_\pm\leftrightarrow w_\pm$
along a path $\gamma_{z,w}^\pm$ with (anti-)clockwise orientation on the
worldsheet. The products in \eqn{braid} are well defined operators on ${\cal
H}^\pm$ provided that $|z_\pm|<|w_\pm|$. Furthermore, chiral vertex operators
may be composed together using the fusion equations
\begin{eqnarray}
&
&V_\pm(\psi_I^{(\pm)};z_\pm)\,V_\pm(\psi_J^{(\pm)};w_\pm)=\sum_{K,L}\left(F^\pm\right)_{IJ}^{KL}\nonumber\\& &\times\,V_\pm
\left(V_\pm(\psi_K^{(\pm)};z_\pm-w_\pm)\,\psi_L^{(\pm)};w_\pm\right)
\label{fusion}\end{eqnarray}
where $F^\pm$ are called fusion matrices. The braiding and fusion relations
completely characterize the chiral algebras ${\cal E}^\pm$. They can be
combined into a single relation known as the Jacobi identity of the vertex
operator algebra \cite{voa}, which can be thought of as a combination of the
classical Jacobi identity for Lie algebras and the Cauchy residue formula for
meromorphic functions.
The full, left-right symmetric vertex operator algebra is now obtained via the
sewing transformations
\begin{eqnarray}
& &V(\psi_I;z_+,z_-)\nonumber\\& &=\sum_{K,L}D_I^{KL}~V_+(\psi_K^{(+)};z_+)\otimes
V_-(\psi_L^{(-)};z_-)\nonumber\\& &
\label{sewing}\end{eqnarray}
where $D_I^{KL}$ are complex valued sewing coefficients and
$\psi_I=\psi_I^{(+)}\otimes\psi_I^{(-)}$. The local fields \eqn{sewing} act
as operator valued distributions on the Hilbert space ${\cal
H}=\complexs^{\{D\}}\otimes{\cal H}^+\otimes{\cal H}^-$, where
$\complexs^{\{D\}}$ is the finite dimensional multiplicity space which
labels the various left-right sewings. The operator-state correspondence is
represented by the relation $V(\psi;0,0)|{\rm vac}\rangle=\psi\in{\cal H}$,
where $|{\rm vac}\rangle$ is the vacuum state of $\cal H$ (which we assume
is unique). Locality is the constraint that two operators of the type
\eqn{sewing} commute whenever their worldsheet arguments are space-like
separated. From this constraint it is possible to actually combine the
braiding and fusion relations into a reduced set of relations known as the
operator product expansion of two local conformal fields \eqn{sewing}. For
this, we grade the Hilbert space $\cal H$ by conformal dimensions
$\Delta_I^\pm$ which are the highest weights of the representations of the
Virasoro algebra in ${\cal E}^\pm$. The corresponding highest weight
vectors $\psi_I$ for ${\cal E}^+\otimes{\cal E}^-$ are called primary
states and are defined by
\begin{equation}
L_0^\pm\psi_I=\Delta_I^\pm\psi_I~~~~~~,~~~~~~L_k^\pm\psi_I=0~~\forall k>0
\label{highestwt}\end{equation}
The associated vertex operators are called primary fields and they satisfy the
differential equations
\begin{eqnarray}
& &\left[L_k^\pm,V(\psi_I;z_+,z_-)\right]\nonumber\\&
&=\left(z_\pm^{k+1}\frac\partial{\partial z_\pm}+(k+1)\Delta_I^\pm
z_\pm^k\right)V(\psi_I;z_+,z_-)\nonumber\\& &
\label{primfield}\end{eqnarray}
With a suitable normalization of the two-point functions of primary fields of
fixed conformal dimension, one may derive the operator product expansion
\begin{eqnarray}
& &V(\psi_I;z_+,z_-)\,V(\psi_J;w_+,w_-)\nonumber\\&
&=\sum_KC_{IJK}~(z_+-w_+)^{\Delta_K^+-\Delta_I^+-\Delta_J^+}\nonumber\\&
&~~~~\times(z_--w_-)^{\Delta_K^--\Delta_I^--\Delta_J^-}~V(\psi_K;z_+,z_-)\nonumber\\&
&
\label{ope}\end{eqnarray}
where the sum runs over a complete set of primary fields (equivalently
orthonormal primary states $\psi_K\in{\cal H}$), and $C_{IJK}$ are the constant
operator product expansion coefficients which are functions of the braiding,
fusion and sewing coefficients introduced above. The relation \eqn{ope}
completely characterizes the vertex operator algebra, which as we see is a
rather complicated unital $*$-algebra.
\subsection*{Lattice Vertex Operator Algebras}
We shall now specialize the above discussion to the case of closed bosonic
strings propagating in a $d$-dimensional toroidal target space
$T_d=\reals^d/2\pi\Gamma$. The classical string embedding fields in such a
target space are determined as the mod $2\pi\Gamma$ periodic solutions of the
two dimensional wave equation, which are given by the chiral multivalued
Fubini-Veneziano fields (in units with $l_s=1$)
\begin{equation}
X_\pm^i(z_\pm)=x_\pm^i+ig^{ij}p_j^\pm\log
z_\pm+\sum_{k\neq0}\frac1{ik}\,\alpha_k^{(\pm)i}\,z_\pm^{-k}
\label{fvfields}\end{equation}
where $g^{ij}$ is the matrix inverse of the metric $g_{ij}$ of $T_d$,
$(\alpha_k^{(\pm)i})^*=\alpha_{-k}^{(\pm)i}$ and $z_\pm={\rm e}^{-i(\tau\pm\sigma)}$
with $(\tau,\sigma)$ local coordinates on the cylinder $\reals\times S^1$. The
left-right momenta are given by
\begin{equation}
p_i^\pm=\mbox{$\frac1{\sqrt2}$}\left(p_i\pm d_{ij}^\pm w^j\right)
\label{lrmom}\end{equation}
and the background matrices are
\begin{equation}
d_{ij}^\pm=g_{ij}\pm\beta_{ij}
\label{background}\end{equation}
with $\beta_{ij}$ the antisymmetric constant torsion form of the target space.
The zero modes $x^i=\frac1{\sqrt2}(x_+^i+x_-^i)\in T_d$ represent the position
of the center of mass of the string while $p_i\in\Gamma^*$ are the
corresponding momenta. The $\alpha$'s represent the vibrational modes of the
string and
$w^i\in\Gamma$ are the winding numbers which represent the number of times
that the string wraps around the cycles of the torus. The set of momenta
$(p^+,p^-)$ live in the even, self-dual Lorentzian lattice
$\Gamma^*\oplus\Gamma$.
Canonical quantization of this theory identifies the non-vanishing quantum
commutators
\begin{eqnarray}
\left[p_i^\pm,x_\pm^j\right]&=&
i\,\delta_i^j\nonumber\\\left[\alpha_k^{(\pm)i},\alpha_m^{(\pm)j}\right]
&=&k\,g^{ij}\,\delta_{k+m,0}
\label{cancomms}\end{eqnarray}
The quantum fields \eqn{fvfields} therefore act on the Hilbert space
\begin{equation}
{\cal H}=L^2(T_d\times T_d^*,S)\otimes{\cal F}^+\otimes{\cal F}^-
\label{hilbert}\end{equation}
where $T_d^*=\reals^d/2\pi\Gamma^*$ is the dual torus to $T_d$ and $S\to
T_d\times T_d^*$ is the spin bundle over the double torus constructed from the
self-dual lattice $\Gamma^*\oplus\Gamma$. The $L^2$ space in \eqn{hilbert} is
constructed from the zero mode operators and is spanned by the plane wave
states
\begin{equation}
|q^+,q^-\rangle={\rm e}^{-i(q_i-\beta_{ij}v^j)x^i-iv^ix^*_i}
\label{L2states}\end{equation}
where $x^*_i=\frac1{\sqrt2}g_{ij}(x_+^j+x_-^j)\in T_d^*$. The ${\cal
F}^\pm$ are bosonic Fock spaces generated by the oscillatory modes
$\alpha_k^{(\pm)i}$, respectively, and the unique vacuum state of \eqn{hilbert}
is $|{\rm vac}\rangle=|0,0\rangle\otimes|0\rangle_+\otimes|0\rangle_-$.
A vertex operator algebra for the toroidal compactification may now be
constructed using the operator-state correspondence, as described above. The
Hilbert space \eqn{hilbert} is spanned by states of the form
\begin{eqnarray}
\psi&=&|q^+,q^-\rangle\otimes
\prod_{a=1}^Nr_i^{(a)+}\,\alpha_{-n_a}^{(+)i}
|0\rangle_+\nonumber\\& &\otimes\,\prod_{b=1}^Mr_j^{(b)-}\,\alpha_{-m_b}^{(-)j}
|0\rangle_-
\label{spanhilbert}\end{eqnarray}
where $(q^+,q^-),\,(r^{(a)+},r^{(a)-})\in\Gamma^*\oplus\Gamma$ and $n_a,m_b>0$.
To \eqn{spanhilbert} we associate the vertex operator
\begin{eqnarray}
& &V(\psi;z_+,z_-)\nonumber\\& &=\NO
i\,V_{q^+q^-}(z_+,z_-)\,
\prod_{a=1}^N\frac{r_i^{(a)+}}{(n_a-1)!}\,
\frac{d^{n_a}X_+^i(z_+)}{dz_+^{n_a}}\nonumber\\&
&~~~~\times\prod_{b=1}^M\frac{r_j^{(b)-}}{(m_b-1)!}\,
\frac{d^{m_b}X_-^j(z_-)}{dz_-^{m_b}}\NO
\label{vertex}\end{eqnarray}
where $\NO\,\cdot\,\NO$ denotes the usual Wick normal ordering of operators,
and
\begin{eqnarray}
& &V_{q^+q^-}(z_+,z_-)\nonumber\\&
&=(-1)^{q_iw^i}\,\NO{\rm e}^{-iq_i^+X_+^i(z_+)-iq_i^-X_-^i(z_-)}\NO\nonumber\\& &
\label{tachyon}\end{eqnarray}
are the basic, left-right symmetric tachyon vertex operators which generate a
unital $*$-algebra $\alg$ which acts diagonally on \eqn{hilbert}. The
operator-valued cocycle phases in \eqn{tachyon} are inserted to give the vertex
operators the correct locality relations. The algebraic properties of $\alg$
are described by the operator product expansion, which for the tachyon
generators reads
\begin{eqnarray}
& &V_{q^+q^-}(z_+,z_-)\,V_{r^+r^-}(w_+,w_-)\nonumber\\& &={\rm e}^{-\pi i\langle
q,r\rangle}~V_{r^+r^-}(w_+,w_-)\,V_{q^+q^-}(z_+,z_-)\nonumber\\& &
\label{opelattice}\end{eqnarray}
where we have assumed that $\pm\,{\rm arg}\,z_\pm>\pm\,{\rm arg}\,w_\pm$, and
\begin{equation}
\langle q,r\rangle=q_i^+g^{ij}r_j^+-q_i^-g^{ij}r_j^-
\label{naraininner}\end{equation}
is the bilinear form on $\Gamma^*\oplus\Gamma$. We shall discuss this
commutation relation in more detail later on. Note that the tachyon operators
create the states $|q^+,q^-\rangle\in L^2(T_d\times T_d^*,S)$ in which all
vibrational modes of the string are absent, i.e. they correspond to low energy
states of the spacetime, which as we will see lead to ordinary (commutative)
manifolds. On the other hand, the lowest symmetric stringy excitation of
commutative spacetime is the state
$|q^+,q^-\rangle\otimes\alpha_{-1}^{(+)i}|0\rangle_+\otimes
\alpha_{-1}^{(-)j}|0\rangle_-$ which is created by the graviton operator
\begin{eqnarray}
& &V_{q^+q^-}^{ij}(z_+,z_-)\nonumber\\& &=\NO
i\,V_{q^+q^-}(z_+,z_-)\,\frac{dX_+^i(z_+)}{dz_+}\,
\frac{dX_-^j(z_-)}{dz_-}\NO\nonumber\\
\label{graviton}\end{eqnarray}
and represents the Fourier modes of the background matrices $d_{ij}^\pm$.
\subsection*{Spectral Triples for Toroidal Compactifications}
The final object we require to complete the string theory spectral triple
is an appropriate Dirac operator. In addition to the basic conformal
symmetry of the model (generated by reparametrizations of the worldsheet
coordinates), there is the target space reparametrization symmetry
$X_\pm^i(z_\pm)\to X_\pm^i(z_\pm)+\delta X_\pm^i(z_\pm)$, with $\delta
X_\pm^i(z_\pm)$ arbitrary periodic functions on $T_d$, which is generated
by the conserved currents
\begin{equation}
\delta_\pm^i(z_\pm)=-i\,z_\pm\,\frac{dX_\pm^i(z_\pm)}{dz_\pm}
\label{kacmoody}\end{equation}
The operators \eqn{kacmoody} generate a $u(1)_+^d\oplus u(1)_-^d$ Kac-Moody
algebra whose highest weight states are the tachyon vectors described above.
The spin bundle $S\to T_d\times T_d^*$ inherits a natural chirality grading
from the corresponding Clifford module over the double torus, which yields two
sets of Dirac matrices $\gamma_i^\pm$ that generate the Clifford algebra
$\{\gamma_i^\pm,\gamma_j^\pm\}=\pm2g_{ij}$ (with all other anticommutators
vanishing). This grading naturally splits the Hilbert space \eqn{hilbert} as
${\cal H}={\cal H}^+\oplus{\cal H}^-$ into the $\pm1$ eigenspaces of the
corresponding chirality operator.
The chiral structure of the theory now enables us to introduce {\it two}
independent Dirac operators for the noncommutative geometry by \cite{lscmp}
\begin{equation}
D^\pm(z_\pm)=\gamma_i^\pm\,\delta_\pm^i(z_\pm)=\sum_{k\in\zed}
\gamma_i^\pm\,\alpha_k^{(\pm)i}\,z_\pm^{-k}
\label{diracchiral}\end{equation}
where $\alpha_0^{(\pm)i}=g^{ij}p_j^\pm$. The choices \eqn{diracchiral} are not
as ad-hoc as they may first seem. First of all, the operator $\delta_\pm$
generates reparametrizations of the target space, so that the first equality in
\eqn{diracchiral} is just the usual relationship between the Dirac operator and
a covariant ``derivative''. Secondly, it is possible to show that the
energy-momentum tensor of the conformal field theory is related to the Dirac
operator by
\begin{equation}
T^\pm(z_\pm)=\sum_{k\in\zed}L_k^\pm\,z_\pm^{-k-2}=-\NO D^\pm(z_\pm)^2\NO
\label{enmom}\end{equation}
so that the square of the Dirac operator is related to the Hamiltonian in the
same way as in the case of a spin-manifold. It can therefore be thought as
generating the appropriate ``Laplace-Beltrami operator'' for the Riemannian
geometry. Finally, it can also be shown that \eqn{diracchiral} are the Ramond
sector fermionic zero modes of the $N=1$ worldsheet supercharges which generate
supersymmetry transformations of the appropriate supersymmetrization of the
underlying worldsheet conformal sigma model. The quantized fermionic zero modes
of these supercharges, when acting on states of vanishing total spacetime
momentum, generate the deRham complex of the manifold $T_d$ -- the two
operators can be identified with the exterior derivative and co-derivative,
while harmonic forms correspond to supersymmetric (physical) states. This is
just the classic Witten complex associated with two-dimensional $N=1$
supersymmetric sigma models \cite{witten}. Thus the Dirac operators
\eqn{diracchiral} are indeed appropriate to the quantum geometry associated
with the string theory. The existence of two independent Dirac operators means
that there are two spectral triples that may be constructed. We shall see that
this feature severely restricts the effective spacetime geometry and is
ultimately responsible for the occurence of duality symmetries in the quantum
geometry.
\section{Target Space Duality}
As we will now show, the existence of two natural Dirac operators for the
noncommutative geometry of the previous section is not an ambiguous property
and is directly tied with the notion of duality. The main feature is that there
are several isometries of the spectral triple that relate the chirally
symmetric and antisymmetric Dirac operators $D,\overline{D}=D^+\pm D^-$. An
isometry in the present context is a unitary operator $T:{\cal H}\to{\cal H}$
which is an automorphism of the vertex operator algebra $\alg$, i.e. $T\,\alg
\,T^{-1}=\alg$ (and in particular it preserves the operator-state
correspondence), and which relates the two Dirac operators via
\begin{equation}
D\,T=T\,\overline{D}
\label{Diso}\end{equation}
This implies that, at the level of their spectral triples, the two spacetimes
associated with $D$ and $\overline{D}$ are the same,
\begin{equation}
(\alg\,,\,{\cal H}\,,\,D)~\cong~(\alg\,,\,{\cal H}\,,\,\overline{D})
\label{speciso}\end{equation}
Since a change of Dirac operator in noncommutative geometry corresponds to a
change in metric on the ``manifold'', the isomorphism \eqn{speciso} is simply
the statement of general covariance of the noncommutative string spacetime.
A spacetime duality transformation is defined as an isomorphism of this
type which identifies subspaces of the two spectral triples in
\eqn{speciso} representing (classically) distinct ordinary spacetimes. The
idea is represented symbolically by the diagram:
\begin{equation}{\begin{array}{crc}
(\alg,{\cal H},D)&{\buildrel T\over\longrightarrow}&(\alg,{\cal
H},\overline{D})\\{\scriptstyle\Pi_0}\downarrow&~~&
\downarrow{\scriptstyle\overline{\Pi}_0}\\(\alg_0,{\cal H}_0,D_0)&~~&
(\overline{\alg}_0,\overline{\cal H}_0,\overline{D}_0)\end{array}}
\label{dualitydiag}\end{equation}
In this diagram, the top line represents the isomorphisms between the full
spectral triples, while $\Pi_0:{\cal H}\to{\cal H}_0$ and
$\overline{\Pi}_0:{\cal H}\to\overline{\cal H}_0$ are orthogonal projections
onto subspaces which represent classical spacetime geometries, e.g.
$(\alg_0,{\cal
H}_0,D_0)=(C^\infty(T_d,\complexs),L^2(T_d,S),ig^{ij}\gamma_i\partial_j)$
represents the ordinary torus $T_d$ with its flat metric $g_{ij}$. From the
point of view of ordinary geometry, there is no reason for the two classical
spacetimes in the bottom line of \eqn{dualitydiag} to be the same. However,
their embeddings into the full spectral triple representing the noncommutative
string spacetime defines an equivalence relation under the action of the
unitary isomorphism $T$ which identifies them at the level of the quantum
geometry, i.e. the mappings in \eqn{dualitydiag} do not commute with the
projection operators, and hence distinct classical spacetimes are identified.
This is a very natural and powerful way to characterize string geometry in
terms of different projections of the same spectral triple.
First, we shall establish that there indeed does exist a natural projection of
the noncommutative string spacetime onto a classical, low-energy sector.
Consider the subspace
\begin{equation}
\overline{\cal H}_0=\ker D\cong\bigotimes_{i=1}^d\left(\overline{\cal
H}_0^{(+)i}\oplus\overline{\cal H}^{(-)i}_0\right)
\label{barH0}\end{equation}
The states in $\overline{\cal H}_0$ are projected onto the vacuum sectors of
the Fock spaces, i.e. their oscillatory parts vanish, as anticipated for a
low-energy regime of the string theory. The space $\overline{\cal H}_0^{(+)i}$
consists of those states which have vanishing momentum $p_i=0$ and whose spinor
components carry the chiral action of the spin group defined by the action of
the Dirac matrices as $g^{jk}d_{ki}^+\gamma_j^+=g^{jk}d_{ki}^-\gamma_j^-$ (upon
choosing appropriate boundary conditions for the spinors with respect to a
homology basis of $T_d\times T_d^*$) in the $i$-th direction of the target
space. On the other hand, $\overline{\cal H}^{(-)i}_0$ consists of those states
which have zero winding number $w^i=0$ and which transform under the antichiral
spinor representation defined by $\gamma_i^+=-\gamma_i^-$.
The $2^d$ subspaces in the decomposition \eqn{barH0} are all naturally
isomorphic to one another \cite{lscmp} under ``partial'' $T$-duality
transformations $\overline{\cal H}^{(+)i}\leftrightarrow\overline{\cal
H}^{(-)i}$ for each $i$. This simply means that this decomposition is
independent of the choice of spin structure on the torus (or equivalently of
the choice of spinor boundary conditions along the elements of a homology
basis), as it should be, and it is a remarkable fact that this independence is
in itself an internal duality symmetry of the string theory. It suffices then
to consider only the completely antichiral subspace
\begin{equation}
\overline{\cal H}_0^{(-)}=\overline{\cal H}_0^{(-)1}\otimes\overline{\cal
H}_0^{(-)2}\otimes\cdots\otimes\overline{\cal H}_0^{(-)d}
\label{antichiralsubsp}\end{equation}
which consists of those states of $\cal H$ which have the quantum
representations $\gamma_i^+=-\gamma_i^-\equiv\gamma_i$ and $w^i=0$ for all
$i=1,\dots,d$. This is precisely what one expects for the low-energy
(particle-like) regime of the string theory, whereby the chiral structure
disappears and there are no non-local string modes that wind around the
compactified directions of the target space. In fact, since
$|p,p\rangle={\rm e}^{-ip_ix^i}$, it follows that $\overline{\cal H}_0^{(-)}\cong
L^2(T_d,S^-)$, where $S^-\to T_d$ is the projection of the spin bundle $S\to
T_d\times T_d^*$ onto its antichiral representation. It is also straightforward
to see that the restriction of the antichiral Dirac operator to the subspace
\eqn{antichiralsubsp} is
\begin{equation}
\overline{D}\,\overline{\Pi}_0^{(-)}=
i\,g^{ij}\,\gamma_i\,\mbox{$\frac\partial{\partial x^j}$}
\label{barDproj}\end{equation}
where $\overline{\Pi}_0^{(-)}:{\cal H}\to\overline{\cal H}_0^{(-)}$ is the
corresponding orthogonal projection. Finally, using the operator-state
correspondence, the corresponding algebra is taken to be the projection of the
commutant of the chiral Dirac operator in $\alg$ (the ``restriction'' of $\alg$
to \eqn{antichiralsubsp}),
\begin{equation}
\overline{\alg}_0^{(-)}=\overline{\Pi}_0^{(-)}\left({\rm
End}_D\alg\right)\overline{\Pi}_0^{(-)}
\label{baralg0def}\end{equation}
where ${\rm End}_D\alg=\{V\in\alg\,|\,[D,V]=0\}$ determines the largest
subalgebra of $\alg$ which acts densely on \eqn{barH0}. The elements of
\eqn{baralg0def} are those vertex operators which have no oscillatory modes and
which create string states of identical left and right chiral momentum, i.e.
$V_{qq}\sim{\rm e}^{-iq_ix^i}$, from which it follows that
$\overline{\alg}_0^{(-)}\cong C^\infty(T_d,\complexs)$. The construction
described above thereby produces a subspace of the spectral triple $(\alg,{\cal
H},\overline{D})$,
\begin{eqnarray}
& &\left(\overline{\alg}_0^{(-)}\,,\,\overline{\cal
H}_0^{(-)}\,,\,\overline{D}\,\overline{\Pi}_0^{(-)}\right)\nonumber\\&
&~~~~~~\cong\Bigl(C^\infty(T_d,\complexs)\,,\,L^2(T_d,S^-)\,,\,i\,
g^{ij}\,\gamma_i\,\mbox{$\frac\partial{\partial x^j}$}\Bigr)\nonumber\\& &
\label{barlowenergy}\end{eqnarray}
The spectral triple \eqn{barlowenergy} represents the ordinary Riemannian
geometry of the torus $T_d$ with metric $g_{ij}$, and thus the Dirac operator
$D$ naturally defines the appropriate low-energy projection of the full
noncommutative string spacetime via its zero-mode eigenspace $\ker
D\subset{\cal H}$ (or its dual version ${\rm End}_D\alg\subset\alg$).
The target space duality transformation \eqn{dualitydiag} follows from the
observation that one could have chosen $\overline{D}$ instead of $D$ in the
definition \eqn{barH0} and carried out an analogous low energy projection.
Doing so, we define
\begin{equation}
{\cal H}_0=\ker\overline{D}\cong\bigotimes_{i=1}^d\left({\cal
H}_0^{(+)i}\oplus{\cal H}^{(-)i}_0\right)
\label{H0}\end{equation}
where ${\cal H}_0^{(+)i}$ consists of states with $\gamma_i^+=\gamma_i^-$ and
$w^i=0$, while the states of ${\cal H}_0^{(-)i}$ carry the quantum
representation $g^{jk}d_{ki}^+\gamma_j^+=-g^{jk}d_{ki}^-\gamma_j^-$ and
$p_i=0$. Taking again the canonical choice
\begin{equation}
{\cal H}_0^{(-)}={\cal H}_0^{(-)1}\otimes{\cal
H}_0^{(-)2}\otimes\cdots\otimes{\cal H}_0^{(-)d}
\label{chiralsubsp}\end{equation}
in which
$g^{jk}d_{kl}^+g^{li}\gamma_j^+=-g^{jk}d_{kl}^-g^{li}
\gamma_j^-\equiv\gamma_*^i$ and $p_i=0$ for all $i=1,\dots,d$, we find the
isomorphism ${\cal H}_0^{(-)}\cong L^2(T_d^*,S^-)$ under the identification
$|d^+w,-d^-w\rangle\leftrightarrow{\rm e}^{-iw^ix_i^*}$.
The projected chiral Dirac operator is
\begin{equation}
D\,\Pi_0^{(-)}=i\,\gamma_*^i\,\mbox{$\frac\partial{\partial x^*_i}$}
\label{Dproj}\end{equation}
where $\Pi_0^{(-)}:{\cal H}\to{\cal H}_0^{(-)}$. Note that the Dirac matrices
$\gamma_*^i$ generate the Clifford algebra with the dual metric
\begin{equation}
\tilde g^{ij}=(d^+)^{ik}\,g_{kl}\,(d^-)^{lj}
\label{dualmetric}\end{equation}
on $T_d^*$ which defines an inner product on the dual lattice $\Gamma^*$ (equal
to the matrix inverse of $g_{ij}$ when $\beta=0$). The low-energy subalgebra of
$\alg$ is now
\begin{equation}
\alg_0^{(-)}=\Pi_0^{(-)}\left({\rm End}_{\overline{D}}\alg\right)\Pi_0^{(-)}
\label{alg0def}\end{equation}
which consists of vertex operators of the form
$V_{d^+v,-d^-v}\sim{\rm e}^{-iv^ix_i^*}$, so that $\alg_0^{(-)}\cong
C^\infty(T_d^*,\complexs)$. In this way we arrive at another low-energy
commutative subspace
\begin{eqnarray}
& &\left(\alg_0^{(-)}\,,\,{\cal H}_0^{(-)}\,,\,D\,\Pi_0^{(-)}\right)\nonumber\\&
&~~~~~~\cong\Bigl(C^\infty(T_d^*,\complexs)\,,\,L^2(T_d^*,S^-)\,,\,i\,
\gamma_*^i\,\mbox{$\frac\partial{\partial x^*_i}$}\Bigr)\nonumber\\& &
\label{lowenergy}\end{eqnarray}
which represents the dual $d$-torus $T_d^*$ with metric $\tilde g^{ij}$.
{}From the point of view of classical general relativity, the two spacetimes
\eqn{barlowenergy} and \eqn{lowenergy} are inequivalent. However, consider the
unitary transformation $T=T_S\otimes T_X:{\cal H}\to{\cal H}$ which acts as the
gauge transformation
\begin{eqnarray}
T_X&=&{\rm e}^{i{\cal G}_X}\in{\cal U}(\alg)\nonumber\\{\cal
G}_X&=&\mbox{$\frac\pi{2i}$}\left(J_+^+J_+^--J_-^+J_-^-\right)
\nonumber\\J_\pm^a(z_a)&=&\NO{\rm e}^{\pm ik_iX_a^i(z_a)}\NO
\label{Tdualtransf}\end{eqnarray}
where $a=\pm$ and $k_i$ is a Killing vector of $T_d$. It acts on the spectral
data of the full noncommutative geometry as \cite{lscmp}
\begin{eqnarray}
T_X|p^+,p^-\rangle&=&(-1)^{p_iw^i}\nonumber\\&
&\times\,|(d^+)^{-1}p^+,-(d^-)^{-1}p^-\rangle\nonumber\\&
&\label{TXstates}\\T_X\,\alpha_n^{(\pm)i}\,T_X^{-1}&
=&\pm\,g_{jk}\,(d^\mp)^{ij}\,\alpha_n^{(\pm)k}\label{TXosc}
\\T_S\,\gamma_i^\pm\,T_S^{-1}&=&g^{jk}\,d_{ij}^\mp\,
\gamma_k^\pm\label{TSgamma}\\T_X\,V_{q^+q^-}\,T_X^{-1}
&=&(-1)^{q_iw^i}\,V_{q^+(d^+)^{-1},-q^
-(d^-)^{-1}}\nonumber\\& &
\label{TXV}\end{eqnarray}
Note that \eqn{TSgamma} implies that the metric $g_{ij}$ is mapped to its dual
\eqn{dualmetric}. It is straightforward to see that $T$ interchanges the two
Dirac operators as prescribed in \eqn{Diso}, and therefore gives the required
isomorphism of the corresponding spectral triples. In this way, the distinct
low energy classical spacetimes \eqn{barlowenergy} and \eqn{lowenergy} are
identified, leading to the celebrated $T$-duality transformation of toroidally
compactified string theory \cite{duality}, in which $d^\pm\to(d^\pm)^{-1}$,
$g\to\tilde g$, and $p_i\leftrightarrow w^i$. We see therefore that in the
framework of noncommutative geometry, the $T$-duality transformation of a
toroidal target space is just the low energy projection of a gauge
transformation on the noncommutative string spacetime.
The $T$-duality mapping actually determines only a $\zeds_2$ subgroup of a
larger discrete duality group of the string theory. The remaining generators
come from the web of transformations between the other various subspaces of
\eqn{barH0} and \eqn{H0}. These are all constructed explicitly in \cite{lscmp}.
For example, the mapping between the antichiral subspace \eqn{antichiralsubsp}
and the corresponding chiral subspace ${\cal H}_0^{(+)}$ determines the
worldsheet parity symmetry of the string theory, which acts on the worldsheet
by
interchanging the chiral structures and on the background data as
$\beta\to-\beta,d^\pm\to d^\mp$. Similarly, the mapping between
\eqn{antichiralsubsp} and the subspace of $\ker\overline{D}$ in which all
tensor components have chiral conditions except for the $i$-th one leads to the
factorized duality transformation of $T_d$, which can be thought of as the
$R\to1/R$ circle duality along the $i$-th cycle of $T_d$. In the present
framework, this transformation is also accompanied by a worldsheet parity map
in all of the other $d-1$ directions. When $d$ is even, the factorized
dualities lead to the phenomenon of mirror symmetry (and hence of spacetime
topology change) which exchanges the complex and K\"ahler structures of the
torus. In addition there other dualities which act trivially on the spectral
triples, but nonetheless do lead to non-trivial quantum dynamics and are
therefore considered as symmetries of the quantum geometry. These are the
changes of basis of the compactification lattice $\Gamma$, and the torsion
cohomology shift $\beta_{ij}\to\beta_{ij}+\lambda_{ij}$, with $\lambda_{ij}$ an
antisymmetric integer-valued matrix, which can be absorbed into a redefinition
of the momenta
$p_i\to p_i-\lambda_{ij}w^j$. Altogether these transformations generate the
discrete duality group which is the semi-direct product
\begin{equation}
G_d=O(d,d;\zeds)\,{\supset\!\!\!\!\!\!\!\times~}\zeds_2
\label{dualitygp}\end{equation}
of the isometry group ${\rm Aut}(\Gamma^*\oplus\Gamma)=O(d,d;\zeds)$ by the
natural action of the worldsheet parity group $\zeds_2$. The duality group
\eqn{dualitygp} labels a large set of low energy theories and connects various
inequivalent classical spacetimes. Its action also leaves the spectra of the
Dirac operators $D$ and $\overline{D}$ invariant.
\section{Gauge Symmetries}
In the previous section we saw that the duality group \eqn{dualitygp} of
toroidally compactified string theory arises in a very natural way through the
automorphisms of the noncommutative string spacetime. With the exception of
worldsheet parity, the transformations we described above were all realized as
inner automorphisms and so represented gauge symmetries of the noncommutative
geometry \cite{lscmp}. On the other hand, worldsheet parity exchanges the left
and right chiral algebras ${\cal E}^\pm\to{\cal E}^\mp$ and therefore
represents an outer automorphism of the vertex operator algebra (since no
element of $\alg$ can accomplish this chirality reversal). In this final
section we shall study in more depth the automorphism group of the vertex
operator algebra, which represents the homeomorphisms of the noncommutative
space, and focus on the consequences of the realization of duality
transformations as gauge symmetries. Here we shall set $\beta_{ij}=0$.
\subsection*{Automorphisms of Lattice Vertex Operator Algebras}
Generally, the inner automorphisms of a $*$-algebra $\alg$ form a normal
subgroup ${\rm Inn}(\alg)$ of the automorphism group ${\rm Aut}(\alg)$. The
remaining symmetries ${\rm Out}(\alg)={\rm Aut}(\alg)/{\rm Inn}(\alg)$ form the
outer automorphisms of the algebra such that the automorphism group is the
semi-direct product
\begin{equation}
{\rm Aut}(\alg)={\rm Inn}(\alg)\,{\supset\!\!\!\!\!\!\!\times~}{\rm Out}(\alg)
\label{autgp}\end{equation}
of ${\rm Inn}(\alg)$ by the natural action of ${\rm Out}(\alg)$. The inner
automorphisms represent internal fluctuations of the noncommutative geometry
corresponding to rotations \eqn{gu} of the elements of $\alg$. Consider the
example of a spin manifold discussed in section 2. In this case the inner
automorphisms of $\alg=C^\infty(M,\complexs)$ are all trivial (there are no
internal symmetries for a commutative algebra), while ${\rm Out}(\alg)\cong{\rm
Diff}(M)$ (given a diffeomorphism $\phi\in{\rm Diff}(M)$, one may construct the
natural automorphism $g_\phi(f)=f\circ\phi^{-1}~~\forall f\in\alg$). The outer
automorphisms in this case generate the general covariance of the space. On the
other hand, for the case $\alg=C^\infty(M,\complexs)\otimes M_N(\complexs)$,
while the outer automorphisms still generate general coordinate transformations
of the manifold $M$, the inner automorphisms generate the group of $U(N)$ gauge
transformations on $M$.
In these two examples there is a clear distinction between (internal) gauge
symmetries and outer automorphisms, again because there is an underlying
commutative algebra which is simply augmented by a discrete internal space.
This is not the case of the string theory spectral triple, which represents a
genuine noncommutative space. As a dramatic example, we recall that the
conserved currents \eqn{kacmoody} are the generators of target space
reparametrizations, so that a general coordinate transformation $X\to\xi(X)$ of
the spacetime may be represented by the gauge transformation \cite{lscmp}
\begin{eqnarray}
T_\xi&=&{\rm e}^{i{\cal G}_\xi}\in{\cal U}(\alg)\nonumber\\{\cal
G}_\xi&=&\xi_i(X)\left(\delta_+^i+\delta_-^i\right)
\label{gencoordtransfs}\end{eqnarray}
However, these inner automorphisms are mapped onto outer automorphisms of the
commutative algebra $\overline{\alg}_0^{(-)}\cong C^\infty(T_d,\complexs)$,
representing the diffeomorphisms of the torus, under the low energy projection
onto the ordinary geometry of $T_d$. This is a very nontrivial fact, as it
implies that general covariance is manifested as a gauge symmetry of the full
noncommutative string spacetime, and in this sense general relativity is in
fact really a gauge theory. This is of course what one would like of a theory
that unifies all of the fundamental interactions, in that it puts gauge and
gravitational interactions on the same footing. Note that in this sense the
worldsheet parity transformations are implemented as ``diffeomorphisms'' of the
noncommutative string spacetime.
The realization of dualities as gauge symmetries yields a natural explanation
of the long standing puzzle as to the origin of string duality as part of some
mysterious gauge group \cite{duality}. In the present case, the conserved
currents \eqn{kacmoody} along with the chiral tachyon vertex operators generate
the affine Lie group ${\widehat{\rm Inn}}^{(0)}(\alg)$ of primary fields of
weight 1 \cite{lscsf}. The corresponding inner automorphisms give internal
perturbations which preserve the grading by conformal dimension and thereby
yield isomorphic conformal field theories. Generically, this group is just the
$U(1)_+^d\times U(1)_-^d$ Kac-Moody group, but there are points in the moduli
space of toroidal compactifications at which this group is enhanced to a
non-abelian gauge symmetry, such as an affine $SU(2)_+^d\times SU(2)_-^d$
Kac-Moody group. A natural subgroup of the automorphism group of a lattice
vertex
operator algebra is therefore given by \cite{lscsf}
\begin{equation}
{\rm Aut}^{(0)}(\alg)={\widehat{\rm Inn}}^{(0)}(\alg)\,{\supset\!\!\!\!\!\!\!\times~}{\rm Out}(\alg)
\label{aut0alg}\end{equation}
where the group of outer automorphisms is
\begin{equation}
{\rm Out}(\alg)\cong O(d,d;\zeds)\,{\supset\!\!\!\!\!\!\!\times~} O(2;\reals)
\label{outeralg}\end{equation}
where $O(d,d;\zeds)$ represents those automorphisms which preserve the bilinear
form of the Lorentzian lattice $\Gamma^*\oplus\Gamma$, while $O(2;\reals)$ is a
worldsheet symmetry group which is an extension of worldsheet parity that acts
by rotating the two chiral sectors of the theory among each other. The full
automorphism group is very large and is not known in explicit form. It
is related to some exotic mathematical constructions, such as the Monster
sporadic group \cite{voa}. But the qualitative form of \eqn{aut0alg} and the
occurence of duality as a gauge symmetry is a generic property of all duality
symmetries. For instance, all of these structures can be shown to hold for
asymmetric duality groups, such as those associated with heterotic strings
\cite{ss}, and also for non-perturbative $SL(2;\zeds)$ $S$-duality symmetries
\cite{lsplb}.
\subsection*{Duality Equivalence Classes}
In the previous section we saw the effects of quantum geometry at the level of
classical spectral triples. The full-blown noncommutative geometry comes into
play at the level of the automorphisms which implement these quantum
symmetries. For instance, the diffeomorphism gauge symmetries
\eqn{gencoordtransfs} are associated with the graviton operators \eqn{graviton}
which represent the lowest oscillatory excitation of the classical spacetime.
The other set of generators of ${\widehat{\rm Inn}}^{(0)}(\alg)$, i.e. the
tachyon vertex operators, turn out to also give a noncommutative perturbation
of classical spacetime that we shall now proceed to describe.
The operator product of two tachyon operators may be written as
\begin{eqnarray}
& &V_{q^+q^-}(z_+,z_-)\,V_{r^+r^-}(w_+,w_-)\nonumber\\&
&=(z_+-w_+)^{q_i^+g^{ij}r_j^+}\,(z_--w_-)^{q_i^-g^{ij}r_j^-}\nonumber\\&
&~~~~~~\times\NO V_{q^+q^-}(z_+,z_-)\,V_{r^+r^-}(w_+,w_-)\NO\nonumber\\& &
\label{tachyonope}\end{eqnarray}
The expression \eqn{tachyonope} is in general singular at coinciding points of
the operators, and to make proper sense of it the vertex operators must be
suitably regularized. Doing so (see \cite{lls} for details), by interchanging
the order of the two operators the relation
\eqn{tachyonope} induces a local cocycle relation which, up to a permutation of
the coordinate directions of $T_d\times T_d^*$, is independent of the
worldsheet coordinates $z_\pm$ and $w_\pm$. This local cocycle relation may be
easily recognized as the defining relation of the noncommutative torus
\eqn{nctorusrels} with the generators $U_i$ identified with the tachyon
operators $V_{e^i_+e^i_-}$, where $e_\pm^i$ is the canonical basis of the
Lorentzian lattice
$\Gamma^*\oplus\Gamma$. The deformation parameter is related to the metric of
$T_d$ by \cite{lls}
\begin{equation}
\omega^{ij}={\rm sgn}(j-i)\,g^{ij}~~~~~~,~~~~~~i\neq j
\label{voadefpar}\end{equation}
More precisely, the tachyon algebra generated by the operators $V_{q^+q^-}$
defines a unitary equivalence class of self-dual $\zeds_2$-twisted projective
modules of rank $2d$ over the double noncommutative torus
$T_\omega^{(+)d}\times T_\omega^{(-)d}$ (one copy for each chiral sector of the
vertex operator algebra along with a $\zeds_2$ twist coming from the cocycle
factors in \eqn{tachyon}). Notice, however, that according to \eqn{opelattice}
the product of two projected tachyon operators (made via the operator-state
correspondence by mapping onto $L^2(T_d\times T_d^*,S)$) is commutative and
just represents the algebra of functions on $T_d\times T_d^*$. In the case at
hand, two given operators are first multiplied together in the full algebra,
and then the result is projected onto the tachyon sector of the noncommutative
geometry. Namely, given $V_0=\Pi_0\,V\,\Pi_0$ and $W_0=\Pi_0\,W\,\Pi_0$, where
$V,W\in\alg$ and $\Pi_0:{\cal H}\to L^2(T_d\times T_d^*,S)$, the deformed
product is given by
\begin{equation}
V_0\star_\omega W_0=\Pi_0\,(VW)\,\Pi_0
\label{voadefprod}\end{equation}
Thus the algebraic properties of the vertex operator algebra give a very
natural way to deform the algebra of functions on $T_d\times T_d^*$ which takes
into account of the internal oscillatory modes of the strings.
We see therefore that the intermediate regime separating the classical torus
from the highly noncommutative string spacetime with its oscillators turned on
is the tachyon algebra which is a module for the noncommutative torus. Note
that for very large distance scales, i.e. $g_{ij}\to\infty$, the deformation
parameter \eqn{voadefpar} vanishes and the tachyon algebra coincides with the
algebra of functions on the double torus. This is just the statement that at
very large distance scales we recover a classical spacetime. On the other hand,
at very small distance scales $g_{ij}\to0$ and thus, within the framework of
toroidally compactified string theory, spacetime at very short distances is a
noncommutative torus.
Having identified this intermediate regime allows us to explore features of the
noncommutative geometry without having to deal with all of the complicated
structures that make up the full vertex operator algebra. Moreover, it enables
us to put the notion of string duality into a more familiar mathematical
parlance. Namely, the deformation parameter \eqn{voadefpar} is the natural
induced antisymmetric bilinear form on the lattice $\Gamma$, and so it
transforms under the action of the target space duality group $O(d,d;\zeds)$ as
in \eqn{moritaom} \cite{duality,lscmp}. Each such orbit is implemented in the
full algebra $\alg$ by a gauge transformation $T$ as described above. Under the
equivalence relation generated by the isomorphism of full spectral triples, the
modules for the noncommutative tori with deformation parameters $\omega$ and
$\omega^*$ are identified. In other words, target space dualities are naturally
realized as the Morita equivalences among noncommutative tori. This remarkable
feature of string geometry can be thought of as lending a physical
interpretation to the mathematical notion of Morita equivalence in
noncommutative geometry. It is also in accord with the connections that exist
between the noncommutative torus and Matrix Theory \cite{cds}, in which the
deformation parameter $\omega$ is given by the light-like component of the
three-form tensor field coming from the corresponding reduction of
11-dimensional supergravity. In that case Morita equivalence relates dual
quantum theories which have the same BPS spectra \cite{schwarz}.
\subsection*{Universal Gauge Symmetry}
Let us now take the Lorentzian lattice to be
$\Gamma^*\oplus\Gamma=\zeds\oplus\omega\zeds$ and assume that the
deformation parameter is a rational number
\begin{equation}
\omega=M/N
\label{omegarat}\end{equation}
where $M$ and $N$ are relatively prime positive integers.
The center of $\alg^{(\omega)}$
\begin{equation}
{\cal Z}(\alg^{(\omega)})={\rm
span}_\complex\left\{U_1^{mN}U_2^{nN}~\Bigm|~m,n\in\zeds\right\}
\label{centeraomega}\end{equation}
is infinite dimensional, and the quotient
$\hat\alg^{(\omega)}=\alg^{(\omega)}/{\cal I}_\omega$ by the ideal ${\cal
I}_\omega$ generated by ${\cal Z}(\alg^{(\omega)})-\{\id\}$ is isomorphic, as a
unital $*$-algebra, to $M_N(\complexs)$. The clock algebra $U_1U_2={\rm e}^{2\pi
i\omega}U_2U_1$ can be represented by the $N\times N$ clock and shift matrices
\begin{eqnarray}
U_1&=&\pmatrix{1&0&0&\dots&0\cr0&q&0&\dots&0\cr0&0&q^2&\dots&0\cr\vdots&\vdots&
\vdots&\ddots&\vdots\cr0&0&0&\dots&q^{N-1}\cr}\nonumber\\
U_2&=&\pmatrix{0&1&0&\dots&0\cr0&0&1&\dots&0\cr\vdots&
\vdots&\vdots&\ddots&\vdots\cr0&0&0&\dots&1\cr1&0&0&\dots&0\cr}
\label{UVmatrix}\end{eqnarray}
where $q={\rm e}^{2\pi iM/N}$.
The physical relevance of this special case is many-fold. First of all, any
``smooth'' element of the $su(N)$ Lie algebra can be expanded in terms of
products of the generators \eqn{UVmatrix}, such that in the limit
$N\to\infty,\omega\to0$ it becomes the Poisson-Lie algebra of smooth
functions on the 2-torus with respect to the usual Poisson bracket
\cite{bfss,fgr}. This is similar to what occured for the vertex operator
algebra, whereby some of the gauge symmetries of the noncommutative
geometry projected onto outer automorphisms of the classical spacetime. The
large-$N$ limit of $SU(N)$ is relevant to the Matrix Theory description of
M-theory, in which $N$ is proportional to the longitudinal momentum and the
limit $N\to\infty$ describes dynamics in the infinite momentum frame
\cite{bfss}. The gauge symmetry here is then represented by the infinite
unitary group $SU(\infty)$ which is an example of a universal gauge group
\cite{lscsf,Rajeev}. The canonical universal gauge groups of vertex
operator algebras, representing the algebras overlying all of the dynamical
symmetries of string theory, may thus be related, through the connection
with the noncommutative torus, to $SU(\infty)$.
More ambitiously, one can imagine that the noncommutative 2-torus with
rational deformation parameter \eqn{omegarat} is some sort of finite
dimensional approximation to the vertex operator algebra. In the correlated
limit $N,M\to\infty$ with $\omega$ a fixed irrational number, we recover the
genuine noncommutativetorus (the center \eqn{centeraomega} is then trivial)
which is related to a corresponding vertex operator algebra as described
above with an explicit representation of the universal gauge symmetry. The
noncommutative geometry associated with rank 2 lattice vertex operator
algebras may in this way be relevant to the true noncommutative geometric
structure encompassing the large-$N$ limit of Matrix Theory. However, in
order for this to be true, one needs to realize the noncommutative string
spacetime as some limit of a tower of finite dimensional matrix geometries.
The potential relevance to Matrix Theory compactifications comes from the
observation that the Morita equivalence of the noncommutative 2-torus can be
naturally interpreted as the zero area limit of the usual string theoretical
$SL(2;\zeds)$ duality in the presence of a background Neveu-Schwarz two-form
field $\beta$ \cite{dh}. In supergravity, this is not a geometrical symmetry in
the usual sense, but in the framework of noncommutative geometry it is. In the
present case, the field $\beta$ is given explicitly by \eqn{voadefpar} and in
the zero area limit we obtain a highly noncommutative structure. This is in
precise agreement with the idea that toroidal compactifications at very small
distance scales are equivalent to noncommutative tori.
\subsection*{Noncommutative Gauge Theory}
The mapping onto the noncommutative torus can also be used to give a dynamical
model for the duality symmetries of the noncommutative string spacetime,
through the construction of an explicitly duality symmetric noncommutative
gauge theory \cite{lls}. The action functional of the gauge theory defined from
the twisted module over the noncommutative torus is obtained using the
invariant integration \eqn{trace} and the derivations \eqn{linderiv}.
The interaction vertices are constructed from the commutator which is
represented by the Moyal bracket
\begin{equation}
[f,g]=\{f,g\}_\omega=f\star_\omega g-g\star_\omega f
\label{moyal}\end{equation}
where the deformed product of two functions $f,g\in C^\infty(T_d\times
T_d^*,\complexs)$ is given by
\begin{eqnarray}
& &(f\star_\omega g)(x,x^*)\nonumber\\&
&=\exp\left[i\pi\omega^{ij}\left(\mbox{$\frac\partial{\partial
x^i}\,\frac\partial{\partial
x'^j}$}-g_{ik}g_{jl}\,\mbox{$\frac\partial{\partial
x_k^*}\,\frac\partial{\partial x^{\prime*}_l}$}\right)\right]\nonumber\\&
&~~~~~~\times\,f(x,x^*)\,g(x',x^{\prime*})
\Bigm|_{(x',x^{\prime*})=(x,x^*)}\nonumber\\& &
\label{defprodncgauge}\end{eqnarray}
Using the projections of the two Dirac operators $D$ and $\overline{D}$
introduced earlier onto the tachyon algebra, one may naturally write down
bosonic and fermionic Lagrangians in the usual spirit of noncommutative
geometry \cite{lls},
\begin{eqnarray}
&&{\cal L}=\left(F+{ }^\star F\right)_{ij}\left(F+{ }^\star
F\right)^{ij}\nonumber\\ &&-i\overline{\psi_*}\gamma^i\left(\partial_i+
i{\buildrel\leftrightarrow\over
{A_i}}\right)\,\psi-i\overline{\psi}\gamma^*_i\left(\partial_*^i+i
{\buildrel\leftrightarrow\over{A_*^i}}\right)\,\psi_*\nonumber\\
\label{dualsymmaction}\end{eqnarray}
The bosonic part yields a symmetrized
Yang-Mills type functional with the above interaction vertices, constructed
from two vector potentials $A(x,x^*)$ and $A_*(x,x^*)$. The fermionic part
contains fermion fields minimally coupled to the gauge potentials, again
determined by a non-local interaction vertex coming from the representation
of the noncommutative tachyon algebra on the Hilbert space $L^2(T_d\times
T_d^*,S)$, i.e.
\begin{eqnarray}
(V_{q^+q^-}\,f)_{r^+r^-}&=&(-1)^{q_iv^i}~{\rm e}^{2\pi
i(q_i^+g^{ij}r_j^++q_i^-g^{ij}r_j^-)}\nonumber\\&
&\times\,f_{q^++r^+,q^-+r^-}
\label{tachyonaction}\end{eqnarray}
for $f\in{\cal S}(\Gamma^*\oplus\Gamma)$.
The noncommutative gauge theory so constructed in possesses a number of
symmetries. The fact that the action is manifestly gauge invariant leads
immediately to its explicit invariance under target space duality
transformations (represented by the actions of inner automorphisms) and
also its general covariance. It is also invariant under the outer
automorphism group
\eqn{outeralg} representing ``diffeomorphisms'' of the noncommutative geometry.
This includes the Morita equivalence acting on the doublet of vector
potentials by the vector representation of $O(d,d;\zeds)$, and also the
$O(2;\reals)$ chiral rotations of the doublet.
Like the usual
formulations of duality symmetric action functionals, this model possesses
an $O(2;\reals)$ doublet of vector potentials, but the present gauge theory
is generally covariant without the need of introducing auxiliary fields.
The price to pay for this is the nonlocality of the gauge interactions, but
from the point of view of quantum field theoretic renormalizability the
present model is in fact completely well-defined. The gauge theory can be
considered as a dynamical model which measures how much duality symmetry is
present in the target space and hence how far away the stringy perturbation
is from ordinary classical spacetime \cite{lls}. The action functional can
thereby be regarded as an effective measure of distance scales in the
spacetime.
The nonlocality which arises in this field theory can also be shown to be
induced by D1-branes wrapping a highly oblique 2-torus \cite{dh}, whose
worldvolume field theory is thereby described by 2+1 dimensional noncommutative
gauge theory. By $T$-duality, this gauge theory also describes D0-branes on a
very small 2-torus with non-zero Neveu-Schwarz two-form field, which in the
large $N$ limit becomes the Matrix Theory description of M-theory with a
non-zero background three-form field and a light-like compactification circle.
It would be interesting to pursue further the interrelationships that exist
between the two noncommutative geometric approaches, one based on (worldsheet)
vertex operator algebras and the other on (target space) Matrix Theory
compactifications, to string theory and M-theory. Their very natural
relationships to the noncommutative torus hints that the connection may not be
very far away.
|
\section{Introduction}
North (1995) provides an excellent history of the discovery that distant
galaxies had a distant-dependent redshift and of the various theories that
were proposed to explain this redshift.
Although there was strong initial support for tired-light models the lack
of a viable physical explanation and the apparent success of the expansion
cosmology has meant that there has been little consideration of tired-light
models in the last forty years.
However in a short note Grote Reber (1982) argues that Hubble himself was never a
promoter of the expanding universe model and personally thought that a
tired-light model was simpler and less irrational.
LaViolette (1986) has compared the generic tired-light model with the big-bang
model on four kinds of cosmological tests.
He concluded that non-evolving Euclidean tired-light model is a better fit
for the cosmological
tests of angular diameter verses redshift, magnitude verses redshift, number density of
galaxies verses magnitude and number density of radios sources verses flux density. He also
provides references to earlier theories and to his own model for a tired-light mechanism.
The strongest theoretical arguments against a tired light model are that it requires new physics
and that any scattering mechanism that gives rise to an energy loss will also produce an angular
scattering that is not observed.
In the tired-light model considered here the new physics for the redshift is minimal
and the angular scattering is insignificant.
In previous papers (Crawford 1987a, 1991) it was argued that there is a gravitational
interaction such that photons and particles lose energy as they pass through a gas.
The energy loss for photons' results in a redshift that could produce the Hubble redshift.
The argument is that photons can be treated as discrete entities with
a finite extent that are subject to the "focusing" theorem of curved spacetime.
That is the cross-section of a bundle of geodesics (in a space with positive curvature)
will decrease in area with distance.
This is just the analogue of the convergence of lines of longitude.
The hypothesis (Crawford 1987a) is that this focusing produces an interaction that leads to the
loss of energy to very low energy secondary photons and an effective redshift of the primary
photon.
Because the interaction is effectively with the mass that produces the curved spacetime and
that this mass will have a very much larger inertia than the particle the angular
scattering of the photon will be negligible.
If the Hubble redshift is explained by a non-expansion mechanism there is still
the problem that a static cosmological solution to the equations of general relativity
is unstable so that a viable cosmology requires some way to provide stability.
This problem was investigated (Crawford 1993) without curvature pressure in a
static cosmology within a Newtonian context.
That work is superseded by the present paper which shows that, using general
relativity, the use of curvature pressure can provide a static and stable
cosmology.
The solution is the new concept of curvature pressure which is based on the observation
that in plasmas electromagnetic forces completely dominate
the particle motions so that they do not travel along geodesics.
The curvature pressure is the reaction back on the material that generates curved
spacetime from the non-geodesic motion of its component particles.
Where curved spacetime is due to a plasma the reaction is seen as a (curvature) pressure
within the plasma that depends on its density and temperature and acts to prevent compression.
The curvature pressure is investigated in a static cosmological model and for plasmas that
occur in the center of the sun and around compact objects. It is shown that the effect of
curvature pressure will decrease the central solar temperature by an amount that may be
sufficient to explain the observed deficiency of solar neutrinos.
Since curvature pressure acts to oppose contraction and since it increases with temperature
it is unlikely that black holes could form from hot plasmas. However it remains possible to
form black holes from cold material. More significantly curvature pressure is very important
in accretion disks around compact objects and may provide the engine that drives
astrophysical jets.
Since the big-bang cosmological model in all its ramifications is so well entrenched,
to be taken seriously, any alternative model must at least be able to explain
the major cosmological observations.
It is argued that using the Friedmann equations the introduction of curvature pressure leads
to a static and stable cosmological model.
One of the predictions of this model is that there is a background X-ray radiation and an
analysis of the background observations done in a previous paper are used to determine the
average density of the universe.
Because of its essential importance to this static cosmology and because the earlier results
did not include the effects of curvature pressure the hypothesis of a gravitational
interaction is revisited.
The result is a prediction of the Hubble constant
$H=60.2\,\mbox{km}\cdot \mbox{s}^{-1}\cdot \mbox{Mpc}^{-1}$.
and the existence of the microwave background radiation with a temperature of 3.0\,K.
It is shown how the observations that lead to the occurrence of dark matter in the big-bang
cosmology are readily explained without dark matter.
Next previous work on the luminosity function of quasars and the angular sizes of radio
sources is discussed to show that the observations can be fitted without evolution.
The classic Tolman surface brightness test is discussed with respect to recent
observations from Pahre et al (1996).
The theme of evolution, or lack of it, is continued with examination of observations on
quasar absorption lines, a microwave background temperature at high redshift,
type 1a supernovae light curves and the Butcher-Oemler effect.
Finally the topics of nuclear abundance, entropy, and Olber's paradox are briefly covered.
\section{Theoretical background}
A theme that is common to the development of both curvature pressure and the gravitational
interaction is that in four-dimensional-space the effects of centripetal acceleration are
essentially the same as they are in three-dimensional-space.
Mathematically a smooth three-dimensional curved space can be locally approximated as a
four-dimensional Euclidean space.
Provided the volume is small enough and the curvature is smooth enough
higher order spaces can be neglected.
The hypothesis made here is that this four-dimensional space has a physical reality.
Note that is not to be confused with four-dimensional spacetime; here we have
five-dimensional spacetime.
Consider two meridians of longitude at the equator with a perpendicular separation of $h$,
then as we move along the longitudes this separation obeys the differential equation
$h''=-h/r^2$ where the primes denote differentiation with
respect to the path length and $r$ is the radius of the earth.
In addition the particle has a centripetal acceleration of $v^2/r$ where $r$ can be
determined from the behavior of $h''$.
In four-dimensional-space the longitudes become a geodesic bundle and the separation
becomes a cross-sectional area, $A$, where $A''=-A/r^2$.
Again the particle has a centripetal acceleration of $v^2/r$ where now $r$ is the radius
of the hyper-sphere. Although the particle as we know it is confined to three dimensions
there is a centripetal acceleration due to curvature in the fourth dimension that could
have significant effects.
Another fundamental topic considered is the nature of gravitational force.
It is critical to the development of curvature pressure that gravitation produces
accelerations and not forces.
The gravitational interaction theory explicitly requires that photons and particles are
described by localized wave packets.
The wave equations that describe their motion in flat spacetime are carried over to
curved spacetime in which the rays coincide with geodesics.
In particular with the focusing theorem (Misner, Thorne \& Wheeler 1973) there is an
actual focusing of the wave packet in that its cross-sectional area decreases as the
particle (photon) travels along its trajectory.
In this and previous papers (Crawford 1987a, 1991) it is argued that the result is a
gravitational interaction in which the particle loses energy.
\section{The theoretical model for curvature pressure}
In a plasma there are strong, long-range electromagnetic forces that completely dominate
accelerations due to gravitational curvature.
The result is that, especially for electrons, the
particles do not travel along geodesics.
If we stand on the surface of the earth our natural geodesic is one of free fall but
the contact forces of the ground balance the gravitational acceleration with the
consequence that there is a reaction force back on the ground.
The result of stopping our geodesic motion is to produce a force that compresses the ground.
The major hypothesis of this paper is that there is a similar reaction force in
four-dimensional spacetime.
This force acts back on the plasma (that produces the curved spacetime) because
its particles do not follow geodesics.
Thus the plasma appears in two roles.
The first produces the curved spacetime and in the second the failure of its particles to
follow geodesics causes a reaction back on itself acting in the first role.
It is long-range electromagnetic forces that are important, not particle collisions.
For example in a gas without long-range forces and assuming that the time spent during
collisions is negligible the particles will still travel along geodesics between collisions.
Given that there are long range forces that dominate the particle trajectories there is a
reaction force that appears as a pressure, the curvature pressure.
For the cosmological model consider the plasma to occupy the surface of a
four-dimensional hypersphere.
It is easier to imagine if one of the normal dimensions is suppressed then it will appear as
the two-dimensional surface of a three-dimensional sphere.
The nature of this pressure can then be understood by analyzing this reduced model with
Newtonian physics in three-dimensional space.
In this case the curvature pressure acts within the two-dimensional surface and is another
way of describing the effects of the centripetal accelerations of the particles.
By symmetry the gravitational attraction on one particle due to the rest is equivalent to
having the total mass at the center of the sphere.
To start let the shell contain identical particles all with the same velocity,
and let this sphere have a radius r, then the radial acceleration of a particle with
velocity v is $v^2/r$.
At equilibrium the radial accelerations are balanced by the mutual gravitational attraction.
Now for a small change in radius, $dr$, without any change in the particle velocities
and going from one equilibrium position to another we can equate the work done by the
curvature pressure to the work done by the force required to overcome the centripetal
acceleration to get
\begin{equation}
\label{e1}
p_cdA=-{{Mv^2} \over {r}}dr,
\end{equation}
where M is the total mass, but for a two-dimensional area $dA/dr=2A/r$ therefore
$p_c=-Mv^2/2Ar=-\rho v^2/2$ where $\rho $ is the surface density.
Thus the effects of the centripetal accelerations can be represented as a negative
pressure acting within the shell.
The next step is to generalize this result to many types of particles where each
type of particle has a distribution of velocities.
The particles are constrained to stay in the shell by a dimensional constraint that is
not a force.
The experiments of E\"{o}tv\"{o}s and others (Roll et al. 1964 and Braginski and Panov 1971)
show that the Newtonian passive gravitational mass is identical to the inertial mass to
about one part in $10^{12}$.
The logical conclusion is that Newtonian gravitation produces an acceleration and not a force.
The mass is only introduced for consistency with Newton's second law of motion.
The concept of gravitation as an acceleration and not a force is even stronger in general
relativity.
Here the geodesics are the same for all particles independent of their mass and gravitational
motion does not use the concept of force.
Clearly for a single type of particle the averaging over velocities is straightforward so
that the curvature pressure is $p_c=-\rho \overline {v^2}/2$.
The averaging over particles with different masses is more ambiguous.
Traditionally we would weight the squared velocities by their masses; that is we
compute the average energy.
However since the constraint that holds the particles within the two-dimensional shell is not
due to forces and since gravitation produces accelerations and not forces the appropriate
average is over their accelerations.
The result for our simple Newtonian model is
\begin{equation}
\label{e2}
p_c=-\case{1}{2}\rho \sum\limits_i {\overline {v_i^2}},
\end{equation}
where the density is defined as $\rho =\sum\limits_i {n_i}m_i$ and $n_i$ is the number
density of the i'th type of particle.
This simple Newtonian model gives a guide to what the curvature pressure would be for a
more general model in a homogeneous isotropic three-dimensional gas that forms the surface
of a four-dimensional hyper-sphere.
The dimensional change requires that we replace $dA/dr$ by $dV/dr=V/3r$, and then
including the relativistic corrections (a factor of $\gamma ^2$) needed to transform the
accelerations from the particle's reference system to a common system where the average
velocity is zero, we get
\begin{eqnarray}
\label{e3}
p_c& = &-{{\rho} \over {3}}\sum\limits_i {n_i}\overline {\gamma _i^2v_i^2}\nonumber\\ \nonumber
& =& -{{\rho c^2} \over {3}}\sum\limits_i {n_i\left( {\overline {\gamma _i^2}-1} \right)}\nonumber\\
& =& -{{\rho c^2} \over {3}}{\left( \overline {\gamma^2}-1 \right)},
\end{eqnarray}
where the Lorentz factor $\gamma ^2=1/\sqrt {1-v^2/c^2}$.
Note that although the equation for curvature pressure does not explicitly include the
spacetime curvature the derivation requires that it is not zero.
Because this equation was only obtained by a plausibility argument we hypothesize that
the curvature pressure in the cosmological model is given by
equation (\ref{e3}).
Since the particles may have relativistic velocities, and assuming thermodynamic equilibrium,
the ($\overline {\gamma ^2}-1$) factor can be evaluated using the J\"{u}ttner distribution.
For a gas at temperature $T$ and particles with mass $m$ de Groot, Leeuwen \& van Weert (1980) show that
\begin{equation}
\label{e4}
\gamma ^2(\alpha )=3\alpha K_3(1/\alpha)/K_2(1/\alpha)+1,
\end{equation}
where $\alpha = kT/mc^2$ and $K_n(1/\alpha)$ are the modified Bessel functions of the
second kind (Abramowitz and Stegun 1968).
For small $\alpha $ this has the approximation
\begin{equation}
\label{e5}
\gamma ^2(\alpha )=1+3\alpha +{\case{15}{2}}\alpha ^2+{\case{45}{8}}\alpha ^3+\ldots .
\end{equation}
Note for a Maxwellian distribution the first three terms are exact so that the extra
terms are corrections required for the J\"{u}ttner distribution.
For non-relativistic velocities equation (\ref{e5}) can be used and equation (\ref{e3})
becomes
\begin{equation}
\label{e6}
p_c=-\frac{1}{N}\sum\limits_{i=1}^N {\left( {{{n_i} \over {m_i}}}\right)}\overline mkT,
\end{equation}
where $n_i$ is the number density for the i'th type of particle and
$\overline m=\sum\limits_{i=1}^N {n_im_i}/n$ is the mean particle mass.
Except for the inverse mass weighting and the sign this is identical to the expression
for the thermodynamic pressure.
\section{Solar interior and local plasma concentrations}
The equation for curvature pressure derived above for the cosmological model cannot be
used in other situations with different metrics.
The key to understanding the application of curvature pressure in other metrics such as
the Schwartzschild metric used for stellar interiors is to consider the case where the
overall curvature is small and superposition may be assumed.
Since the free fall acceleration of a particle is independent of its mass there is no
curvature pressure associated with external gravitational fields provided they have scale
lengths much greater than the typical ion separation.
Any curvature pressure is due to local curvature of the metric produced by the local density.
This arises because although the electrons and ions have in general different centripetal
accelerations these are completely dominated by accelerations due to the electromagnetic forces.
Let the gravitational potential be $\Phi $, then the self-gravitational energy density
is $\rho\Phi $.
Now it was argued above that the curvature pressure is proportional to the energy density
(it has the same units) but with an averaging over accelerations rather than forces that
results in replacing $\rho$ by $\left( {\overline {\gamma ^2}-1} \right)\rho $.
Consequently we take the curvature pressure in a plasma due to its own density as
\begin{equation}
\label{e7}
p_c=\case{1}{3}\left( {\overline {\gamma ^2}-1} \right)\rho \Phi
\end{equation}
Note that the derivation is essentially one based on
dimensional analysis and therefore the numerical factor of $1/3$ may need modification.
It was used in part for consistency with the cosmological curvature pressure and in part
because it makes the application of equation (\ref{e7}) to a low temperature gas with a
single type of particle have the simple expression $p_c=p_T\Phi /c^2$ where $p_T$
is the thermodynamic pressure.
From potential theory we get for the curvature pressure of a plasma at the point $r_0$
the expression
\begin{equation}
\label{e8}
p_c\left( {r_0} \right)=
\case{1}{3}G\rho \left( {r_0} \right)\left( {\overline {\gamma
^2\left( {r_0} \right)}-1} \right)
\int \frac{\rho \left( {r-r_0} \right)}{\left| {r-r_0} \right|} dV.
\end{equation}
Equation (\ref{e8}) can be simplified for non-relativistic velocities by using the
approximation (equation \ref{e5}) to get
\begin{equation}
\label{e9}
p_c={{G\rho \left( {r_0} \right)kT} \over {c^2}}\left( {\sum\limits_{i=1}^N
{{{n_i} \over {nm_i}}}} \right)
\int \frac{\rho \left( {r-r_0} \right)}{\left| {r-r_0} \right|} dV
\end{equation}
where n is the total number density.
The curvature pressure adds to the thermodynamic pressure (and radiation pressure) to
support the solar atmosphere against its own gravitational attraction.
That is for the same gravitational attraction the required thermodynamic pressure,
and hence the temperature, will be reduced by curvature pressure.
Applying equation (\ref{e9}) to the sun and using pressures, temperatures, and abundance
ratios given by Bahcall (1989), it was found that the curvature pressure at the center
of the sun is $2.8\times 10^{14}\,\mbox{Pa}$ compared to the thermodynamic pressure of
$2.3\times 10^{16}\,\mbox{Pa}$.
Since the temperature is directly proportional to the thermodynamic pressure this implies
that the temperature at the center of the sun is reduced by 1.2\%.
Bahcall (1989 p151) shows that the $^8$B neutrino flux is very sensitive to the temperatures
at the center of the sun with a flux rate that is proportional to the eighteenth power of
the temperature.
Thus this temperature change would decrease the neutrino flux to 80\% of that from the
standard model.
Although the observed ratio of $2.55/9.5=27\% $ (Bahcall 1997) is much smaller the effect
of the pressure curvature is clearly significant and large enough to warrant a more
sophisticated computation.
\section{Cosmology with curvature pressure}
The main application of curvature pressure is to a cosmological model for a homogeneous
and isotropic distribution of a fully ionized gas.
Based on the theory of general relativity and using the Robertson-Walker metric the
Friedmann equations (Weinberg 1972) are
\begin{eqnarray*}
\label{e10}
-\ddot R & =& {{4\pi G} \over {c^2}}\left( {\rho c^2+3p}\right)R\\
R\ddot R+2\dot R^2 & = & {{4\pi G} \over {c^2}}\left( {\rho c^2-p}
\right)R^2 -2kc^2,
\end{eqnarray*}
where R is the radius $\rho $ is the proper density, $p$ is the pressure, $G$ is the
Newtonian gravitational constant, and $c$ is the velocity of light.
The constant $k$ is one for a closed universe, minus one for an open universe and zero
for a universe with zero curvature.
Working to order $m_e/m_H$ the thermodynamic pressure can be neglected but not the
curvature pressure.
The equations including the curvature pressure (equation \ref{e3}) are
\begin{eqnarray*}
\label{e12}
-\ddot R & = & 4\pi G\rho R\{ 1-\left( {\overline {\gamma ^2}-1} \right)\} \\
R\ddot R+2\dot R^2 & =& 4\pi G\rho R^2\{ 1+\case{1}{3}\left( {\overline {\gamma^2}-1}
\right)\} \\ & & - 2kc^2,
\end{eqnarray*}
where $\overline {\gamma ^2}$ is the average over all velocities and particle types.
Clearly
$\ddot R$ is zero if $\overline {\gamma ^2}=2$ and equation (\ref{e4}) can be solved
for a hydrogen plasma to get $\alpha _e=kT_0/m_ec^2=0.335$ or $T_0=1.99\times 10^9$K.
Thus with thermal equilibrium the second derivative of $R$ is zero if the plasma has
this temperature.
Thus the requirement for stability leads to a prediction of the plasma temperature.
This temperature is based on a model in which the plasma is homogeneous, but the occurrence
of galaxies and clusters of galaxies show that it is far from homogeneous.
In order to investigate the effects of inhomogeneity consider a simple and quite arbitrary
model where the plasma is clumped with the probability of a clump having the density n is
given by the exponential distribution $\exp \left( {-n/n_0} \right)/n_0$, where $n_0$
is the average density.
Assuming pressure equilibrium so that $T_e=T_0n_0/n$ then for $\overline {\gamma ^2}=2$
we find that the average temperature $T=1.1\times 10^9$K thus showing that the effect of
inhomogeneity could reduce the observed temperature
by a factor of order two.
Since the right hand side of the second Friedmann equation is positive then the curvature
constant $k$ must be greater or equal to zero.
The only useful static solution requires that $k=1$ and with $\dot R=\ddot R=0$ the
result for the radius of the universe is given by
\begin{equation}
\label{e14}
{{1} \over {R_0^2}}={{8\pi G\rho _0} \over {3c^2}}.
\end{equation}
Thus the model is a static cosmology with positive curvature.
Although the geometry is the
same as the original Einstein static model this cosmology differs in that it does not
require a cosmological constant.
Furthermore it is stable. Consider a perturbation, $\Delta R$, about the equilibrium
position then the perturbation equation is
\begin{equation}
\label{e15}
\Delta \ddot R={{c^2} \over {8\pi R_0}}\left( {{{d\overline {\gamma
^2}} \over {dR}}} \right)\Delta R,
\end{equation}
and since for any realistic equation of state the average velocity (temperature) will
decrease as R increases the right hand side is negative showing that the result of a
perturbation is an oscillation about the equilibrium value.
Thus this model does not suffer from the deficiency that the static Einstein model has of
gross instability.
Since the volume of the three-dimensional surface of the hyper-sphere is $2\pi ^2R_0^3$
the radius of the universe can be written in terms of the total mass of the universe,
$M_0$, as
\begin{equation}
\label{e16}
R_0={{4GM_0} \over {3\pi c^2}},
\end{equation}
which differs by a factor of $2/3$ from that (Crawford 1993) which was derived from a
purely Newtonian model.
For interest the values with a density of $2.05m_H \mbox{ m}^{-3}$ (see below) are
$R_0=2.17\times 10^{26\,}\mbox{m}\ =7.04\,\mbox{Gpc}$, and
$M_0=6.90\times 10^{53}\,\mbox{kg}\ =3.47\times 10^{23}\,\mbox{M}_{\mbox{sun}}$.
\section{Background X-ray radiation }
If this cosmological model is correct there should be a very hot plasma between the
galaxies and in particular between galactic clusters.
This plasma should produce a diffuse background X-ray radiation and indeed such
radiation is observed.
Attempts to explain the X-rays by bremsstrahlung radiation within the standard
model have not been very successful (Fabian \& Barcons 1992), mainly because it must
have originated at earlier epochs when the density was considerably larger
than present.
The hard X-rays could come from discrete sources but if it did there are problems
with the spectral smoothness and strong evolution is required to achieve the
observed flux density (Fabian \& Barcons 1992).
However there is an excellent fit to the data in a static cosmology (Crawford 1987b, 1993)
for X-ray energies between 5\,KeV and 200\,KeV.
Using universal abundances (Allen 1976) the analysis showed a temperature of
$1.11\times 10^9$\,K and a density of $2.05m_H \mbox{ m}^{-3}$.
Comparison of this temperature with that predicted by the homogeneous model of
$171\,\mbox{KeV}$ shows that it is nearly a factor of two too small.
A possible explanation comes from the observation that the universe is not homogeneous.
Although there is fortuitous agreement with the simple inhomogeneous model described above
this can only be interpreted as showing that the observations are consistent with an
inhomogeneous model.
One of the main arguments against the explanation that the background X-ray radiation
comes from a hot inter-cluster plasma is that this plasma would distort the cosmic
microwave background radiation by the Sunyaev-Zel'dovich effect.
This distortion is usually expressed by the dimensionless parameter $y$.
Mather et al (1994) have measured the spectrum of the cosmic microwave background
radiation and conclude that $\left| y \right|<2.5\times 10^{-5}$.
In the big-bang cosmology most of the distortion occurs
at earlier epochs where the predicted density and the temperature of the plasma are
much higher than current values.
However for any static model we can use a constant density of
$2.05m_H \mbox{ m}^{-3}$ in the equation (Peebles 1993)
\begin{equation}
\label{e17}
y={{kT_e\sigma _Tn_er} \over {m_ec^2}},
\end{equation}
where $\sigma _T$ is the Thomson cross-section and $r$ is the path length since the
formation of the radiation.
For a hydrogen plasma we get $y=2.6\times 10^{-29}r$.
The microwave background radiation (see below) it is being
continuously replenished by energy losses from the hot electrons and the typical path
length for the energy lost by electrons to equal the energy of a photon at the peak of
the spectrum is
$3.5\times 10^{18}\,$m which results in $y=9.1\times 10^{-11}$ well within the observed limits.
\section{The Hubble constant}
One of the major requirements of any cosmological model is the necessity to explain the
relationship found by Hubble that the redshift of extra-galactic objects depends on
their distance.
In earlier papers (Crawford 1987a, 1991) the author suggested that there is an interaction
of photons with curved spacetime that produces an energy loss that can explain the
Hubble redshift relationship.
Because the earlier work did not include the effects of curvature pressure and because
this interaction is central to the description of a viable static cosmology a brief
updated description is given here.
The principle is that a photon can be considered as a localized wave traveling along a
geodesic bundle.
Because of the `focusing theorem' (Misner et al 1973) the cross-sectional area of this
bundle will decrease with time, and in applying this theorem to a photon it was argued
that this will cause a change in the photon's properties.
In particular angular momentum will decrease because it is proportional to a spatial
integral over the cross-sectional area.
The change in angular momentum can only be sustained for a time consistent with the
Heisenberg uncertainty principle.
The conclusion is that eventually there will the emission of two (in order to conserve
the total angular momentum) very low energy photons.
The second part of the argument is that the rate at which this energy loss occurs is
proportional to the rate of change of area of the geodesic bundle.
This rate of change of area in the absence of shear and vorticity is given by the
equation (Raychaudhuri 1955),
\begin{equation}
\label{e18}
{{1} \over {A}}{{d^2A} \over {ds^2}}=-R_{\alpha \beta }U^\alpha U^\beta ,
\end{equation}
where $R_{\alpha \beta }$ is the Ricci tensor, $U^\alpha $ is the four-velocity and,
s is a suitable affine parameter.
At any point the trajectory of the geodesic bundle is tangential to the surface of a
four-dimensional hyper-sphere with radius $r$.
Then since the centripetal acceleration is $c^2/r$ where $r$ is defined by
\begin{equation}
\label{e19}
{{1} \over {r^2}}={{1} \over {A}}{{d^2A} \over {ds^2}}
\end{equation}
we can define $\varepsilon $, the fractional rate of energy loss by
\begin{equation}
\label{e20}
\varepsilon =c^2\sqrt {{{1} \over {A}}{{d^2A} \over {ds^2}}}.
\end{equation}
This relationship for $\varepsilon $ is a function only of Riemann geometry and does
not depend on any particular gravitational theory.
However, Einstein's general relativity gives a particularly elegant evaluation.
Direct application of the field equations with the stress-energy-momentum tensor
$T_{\alpha \beta }$ gives
\begin{equation}
\label{e21}
\varepsilon =\sqrt {{{8\pi G} \over {c^2}}\left( {T_{\alpha \beta }U^\alpha
U^\beta -\case{1}{2}Tg_{\alpha \beta }U^\alpha U^\beta } \right)},
\end{equation}
where $T$ is the contraction of $T_{\alpha \beta }$ and $U^\alpha$ is the four-velocity.
Then for a gas with density $\rho $ where the pressures are negligible the energy loss
rate is (Crawford 1987a)
\begin{equation}
\label{e22}
\varepsilon c =-{{1} \over E}{{dE} \over {dt}}=\sqrt {{{8\pi G\left( {\rho
c^2+p} \right)} \over {c^2}}},
\end{equation}
where x is measured along the photon's trajectory.
This equation can be integrated to obtain
\begin{equation}
\label{e22a}
E=E_0\exp (-\varepsilon \mathop x\limits_{}).
\end{equation}
If $\rho =n\,m_H$ and with
(using equation \ref{e3})
\begin{equation}
\label{e23}
p\approx p_c=-{{\rho c^2} \over {3}}\left( {\overline {\gamma ^2}-1} \right)=-{{1} \over
{3}}\rho c^2,
\end{equation}
then $\varepsilon =4.54\times 10^{-27}\sqrt{n}\,\mbox{m}^{-1}$ and the predicted
Hubble's constant is
\begin{equation}
\label{e24}
H=c\varepsilon =42.0\,\sqrt{n} \,\mbox{km}\cdot \mbox{s}^{-1}\cdot \mbox{Mpc}^{-1}.
\end{equation}
With the value $n=2.05m_H \mbox{ m}^{-3}$ we get
$H=60.2\,\mbox{km}\cdot \mbox{s}^{-1}\cdot \mbox{Mpc}^{-1}$.
Note for non-cosmological applications where the curvature pressure is negligible the
results are
$\varepsilon =5.57\times 10^{-27}\sqrt{n} \,\mbox{m}^{-1}$ or
\begin{equation}
\label{e24a}
\varepsilon c = 51.5\sqrt{n} \,\mbox{km}\cdot \mbox{s}^{-1}\cdot \mbox{Mpc}^{-1}.
\end{equation}
Required later is the product of Hubble's constant with the radius of the universe which is
$RH=\sqrt{2}\,c$.
This is identical to that derived earlier (Crawford 1993) for a Newtonian cosmology.
The principle of the focusing theorem can be illustrated by considering a very long
cylinder of gas and Newtonian gravitation.
At the edge of the cylinder of radius $r$ the acceleration to-wards the center of the
cylinder is $\ddot r=2\pi G\rho r$ where the dots denote differentiation with
respect to time.
Hence for the area A we get $\ddot A=4\pi G\rho A$.
Except for the numerical constant this is the same as that for general relativity showing
that it is the local density that determines focusing.
The difference of a factor of one half is because the model only includes space curvature
and not spacetime curvature.
In both cases distant masses have no effect.
In particular there is no focusing and hence no energy loss in the exterior Schwartzschild
field of a spherical mass distribution such as the sun.
Since the excitation of the photon is slowly built up along its trajectory before the
emission of two low energy photons any other interaction that occurs with a path length
shorter than that between the emission of secondaries will clearly diminish their production.
That is the excitation can be dissipated without any extra energy loss.
The average distance between emission of secondaries is (Crawford 1987a;
using Heisenberg's uncertainty principle)
$\Delta x=\sqrt {\lambda _0/4\pi \varepsilon }$ where $h$ Plank's constant,
$\varepsilon $ is the fractional rate of energy loss per unit distance defined above and
$\lambda _0$ is the wavelength of the primary photon.
For the cosmological plasma the secondaries would have a typical frequency of 0.02 Hz for
a 21 cm primary and about 11 Hz for an optical photon which may be compared with the
plasma frequency of 13 Hz. Thus in most cases the secondaries will not propagate
but will be directly absorbed by the plasma.
The classic experiment of Pound and Snyder (1965) is an example of how the hypothesis
of a gravitational interaction may be tested.
They used the Mossbauer effect to measure the energy of 14.4\,KeV (${}^{57}$Co) gamma rays
after they had passed up or down a 22.5\,m path in helium.
Their result for the gravitational redshift was in excellent agreement with the predicted
fractional change in energy of $2.5\times 10^{-15}$.
The gravitational interaction theory predicts a fractional change in
energy due to the gravitational interaction based on the density of helium in the tube of
$1.25\times 10^{-12}$ which is considerably larger.
Since their measurement was for the difference between upward and downward paths any
effects independent of direction will cancel.
However for these conditions although the typical path length between the emission of
secondaries of 11m is less than the length of the apparatus it is still much longer
than the mean free path for coherent forward scattering that is the quantum description
of refractive index.
In this scattering the photon is absorbed by many electrons and after a short time delay
(half a period) a new photon with the same energy and momentum is emitted.
For these high energy gamma rays the binding energy of the electrons can be ignored and
the mean free path for coherent forward scattering is given by the Ewald and Oseen
extinction length (Jackson 1975) of $X=1/\left( {\lambda r_0n_e} \right)$ where
$\lambda $ is the wavelength and $r_0$ is the classical electron radius.
In this case $X=0.15\,$m that is much less than the 11\,m required for secondary emission
and therefore the gravitational interaction energy loss will be minimal.
The major difficulty with a laboratory test is
in devising an experiment where $\Delta x$ is less than the size of the apparatus and also
less than the mean free path of any other interaction.
Nevertheless if there are any residual effects they may be detectable in such an experiment
with a horizontal run using gases of different types and densities.
This inhibition of the gravitational interaction can occur in astrophysical situations.
Consider the propagation of radiation through the Galaxy where there is a fully ionized
plasma with density $\rho =n\,m_H$, then the critical density is defined by when the
Ewald and Oseen extinction length is equal to the distance between emission of secondary photons.
If the density is greater than this critical density then the inhibition by refractive
index impairs the gravitational interaction and there is a greatly reduced redshift.
The critical density (for a hydrogen plasma) is
$n_e=426.5/\lambda ^2\,\mbox{m}^{-3}$.
For 21\,cm radiation the critical density is $n_e=9,700\,\mbox{m}^{-3}$ and since
most inter-stellar densities are much larger than this we do not expect 21\,cm radiation
within the Galaxy to show redshifts due to the gravitational interaction.
However if the gas is very clumpy we could still get uninhibited redshifts from the low
density components.
Thus all redshifts of 21\,cm radiation within the Galaxy may be primarily due to doppler shifts.
However optical radiation in the Galaxy should show the redshift due to the gravitational
interaction. This inhibition could be verified if a neutral hydrogen cloud could be
clearly identified with an object having optical line emission.
It has been argued (Zel'dovich 1963) that tired light cosmologies (such as this) should
show a smearing out of the images of distant sources.
The argument is that if the energy loss is caused by an interaction with inter-galactic matter,
it is accompanied by a transfer of momentum with a corresponding change in direction.
That is the photon is subject to multiple scattering and hence photons from the same source
will eventually have slightly different directions and its image will be smeared.
For the gravitational interaction the interaction is not with some particle with
commensurate mass but with the mass of the gas averaged over a suitable volume.
Since the effective mass is so large the scattering angles will be negligible.
Furthermore in low density gas the photon loses energy to two secondary photons and to
conserve spin and momentum these will be on average be emitted symmetrically so that
there is no scattering of the primary photon. Thus this model overcomes the scattering
objection to tired-light explanations for redshifts.
\section{The microwave background radiation}
Because of their wave nature electrons and other particles will be subject to the focusing
theorem in a way similar to photons.
In Crawford (1991) it was argued that particles such as electrons are subject to a similar
centripetal acceleration that produces a fractional energy loss rate of
$\varepsilon _e$, and for a gas with density $\rho $ and pressure $p$ it is
\begin{equation}
\label{e25}
\varepsilon _e=\sqrt {{{8\pi G} \over {c^2}}\left[ {(\gamma ^2 - \case{1}{2})\rho
c^2+\left( {\gamma ^2 + \case{1}{2} } \right)p} \right]}\ ,
\end{equation}
where $\gamma $ is the usual velocity factor.
Hence the rate of energy loss as a function of distance is
\begin{equation}
\label{e26}
{{dP^0} \over {dx}}=\sqrt {{{8\pi G} \over {c^4}}\left[ {(\gamma^2-
\case{1}{2})\rho c^{^2}+\left( {\gamma ^2+\case{1}{2}} \right)p} \right]}\,\beta^2P^0,
\end{equation}
where $\beta =v/c$ is the particle's velocity relative to the medium and $P^0$ is the energy component of its momentum four-vector.
As it moves along its trajectory the particle will be excited by the focusing of its wave packet.
For charged particles conservation of spin prevents them from removing their excitation by direct emission of low energy photons.
However if there are photons present it may interact by stimulated
emission and thereby lose energy to secondary photons.
The dominant photon field in inter-galactic space is that
associated with the microwave background radiation.
The model proposed is that the electrons lose energy by stimulated emission to the
background radiation so that the local black body spectrum is conserved.
Since the conservation of energy, momentum and spin prevents a free electron from
radiating it can only lose its energy of excitation by interacting with another particle
or in this case the radiation field.
The hypothesis is that it continuously gains energy until an interaction with a photon
stimulates the emission of a new photon. Thus the energy spectrum of the emitted photons
will match that of the existing photons. Thus give a local black body spectrum the emitted
radiation will also have the same black body spectrum. This does not explain how the
black body radiation originally arose but if there is any way in which photons can interact
to alter their energy spectrum then the equilibrium spectrum is that for a black body.
Concurrently because of the gravitational interaction the photons are losing energy that is absorbed by the plasma.
Note that most of the secondary photons have frequencies below the plasma frequency.
Although this means that they cannot propagate it does not prevent direct absorption of
their energy.
After all for frequencies below the plasma frequency the electrons can have bulk motion
and absorb energy from an oscillating field.
Given an equilibrium condition in which the energy lost by the electrons is equal to the
energy lost by the photons we can equate the two energy loss rates and get an expression
for the temperature of the microwave background radiation (Crawford 1991) of
\begin{equation}
\label{e27}
T_M^4={{n_em_ec^3} \over {4\sigma }}\overline{ {\left( {(\gamma ^2-
\case{1}{2})+\left( {\gamma ^2+\case{1}{2}} \right){{p} \over {\rho c^{^2}}}}
\right)\beta ^3\gamma } },
\end{equation}
where $n_e$ is the electron number density, $m_e$ is the electron mass,
$\sigma $ is the Stefan-Boltzmann constant and, an average s done over all electron velocities.
For an electron temperature of $1.11\times 10^9$\,K the bracketed term has the value
of 0.555 with zero pressure or 0.412 with the gravitational curvature pressure.
With an electron density of 1.78\,m$^{-3}$ corresponding to a mass density of
$2.05m_H \mbox{ m}^{-3}$ and with the curvature pressure included the predicted temperature
is 3.0\,K.
Given the deficiencies of the model (mainly its assumption of homogeneity) this is in good
agreement with the observed value of 2.726\,K (Mather et al 1994).
It is interesting that the predicted temperature only depends on the average density and a
function of electron velocities that
is of order one.
\section{Dark matter}
In the standard big-bang cosmology there are three major arguments (Trimble 1987) for the
existence of dark matter, that is matter that has gravitational importance but is not seen at any wavelength.
The first argument is based on theoretical considerations of closure and reasonable cosmological models within the big-bang paradigm.
The second is from the application of the virial theorem to clusters of galaxies and the third is that galactic rotation curves show high velocities at large radii.
The first of these is purely an artifact of the big-bang cosmological model; it is not based on observation and therefore it is not relevant to this cosmology.
The second and third are based on observations and will be discussed at some length.
In the standard big-bang model all the galaxies in a cluster are gravitationally bound and do not partake in the universal expansion.
If they are gravitationally bound then assuming that their differential (peculiar) redshifts are due to differential velocities we can use the virial theorem to estimate the total (gravitational) mass in the cluster.
Typically this gravitational mass is one to several orders of magnitude larger than the mass derived from the luminosities of the galaxies: hence the need for dark matter.
Observations of X-rays from galactic clusters show that there is a large mass of gas in the
space between the galaxies.
Although the mass of this inter-cluster gas is small compared to the mass of the presumed
dark matter it is large enough to give significant redshifts due to the gravitational interaction.
Thus the current model ascribes most of the differential redshifts to gravitational interactions
in the inter-cluster gas.
This model has been quantitatively investigated by Crawford (1991) for the Coma cluster.
The method used was to take the observed differential redshift for each galaxy and by
integrating equation (\ref{e24a}) through the known inter-galactic gas the differential
line-of-sight distance to the galaxy was computed.
The gas density distribution that was used is that given by Gorenstein, Huchra \&
de Lapparent (1979).
The result is that galaxies with lower redshifts than that for the center of the cluster
would be nearer and those with higher redshifts would be further away.
The model assumed that the inter-cluster gas was spherically distributed and the test was
in how well the distribution of Z coordinates compared with those for the X and Y coordinates
that were in the plane of the sky.
Furthermore it was assumed that genuine velocities were negligible compared to the effective
velocities of the differential redshifts.
The median distances for each coordinate were X=0.19\,Mpc, Y=0.17\,Mpc and Z= 0.28\,Mpc.
Given that the Coma cluster has non-spherical structure and that the model is very simple
the agreement of the median Z distance with those for X and Y is good.
Again it should be emphasized that there were no free parameters; the Z distances depend only on the gas distribution, the measured differential redshift, and equation (\ref{e24a}).
If this result can be taken as representative of clusters then there is no need for dark
matter to explain cluster `dynamics'.
The large differential redshifts are mainly due to gravitational interactions in the
inter-galactic gas.
One of the difficulties with the big-bang cosmology is that it is so vague in its
predictions that it is very difficult to refute it with observational evidence.
However the redshifts from a cluster of galaxies can provide a critical test.
Since celestial dynamics is time reversible a galaxy at any point in the cluster is
equally likely to have a line-of sight velocity towards us as away from us.
Then if accurate measurements of magnitude, size or some other variable can be used to
get differential distances there should (in the big-bang cosmology) be no correlation
between differential redshift and distance within the cluster.
Whereas in the static cosmology proposed here there should be a strong correlation with
the more distant galaxies having a higher differential redshift.
Clearly this is a difficult experiment since for the Coma cluster it requires
measurements of differential distances to an accuracy of about 1\,Mpc at a distance of 100\,Mpc.
The third argument for dark matter comes from galactic rotation curves.
What is observed is that velocity plotted as a function of distance along the major
axis shows the expected rapid rise from the center but instead of reaching a maximum and
then declining in an approximately Keplerian manner it tends to stay near its maximum value.
The standard explanation is that there is a halo of dark matter that extends well beyond
the galaxy and that has a larger mass than the visible galaxy.
For this static cosmology a partial explanation is that most of the redshift is due to
gravitational interaction in a halo but one that is commensurate in size with the galaxy.
Consider a spiral galaxy that is inclined to the line of sight and that has a halo
with a Gaussian density distribution (chosen purely for analytic convenience).
Then if the redshift origin is taken to be at the centre of the galaxy light
from points further away will travel through more
halo gas and therefore will be redshifted and nearer points will be blue shifted.
Let the halo density distribution as a function of radius be
$\rho = \rho_0 Exp{-(r/r_0)^2}$
and let $x$ be the distance measured from the galaxy centre along line through the galaxy
that lies in the same plane as the line of sight and the normal to the galaxy.
Then the observed redshift (in velocity units) is
\begin{equation}
\label{e27a}
v - v_0 = \pi(4G\rho_0)^{1/2} r_0 \exp(-(\frac{x \sin(i)}{\sqrt{2}r_0})^2)
erf(\frac{x \cos(i)}{\sqrt{2} r_0})
\end{equation}
where $i$ is the inclination angle.
Since the error function is asymmetric and the exponential function dominates
at large distances the relative velocity shows a rapid increase to a broad maximum
and then a slow decrease back to zero.
For most galaxies it is likely that the maxima well extended
beyond the physical limits of the galaxy and so that only part of the decrease
may be observed.
Clearly the precise shape of the curve and its numerical value will depend on the
precise nature of the density distribution.
Nevertheless the value of the maximum velocity
for this curve will give a good indication of the effect.
For an inclination angle of $\hbox{$^\circ$}{45}$ the maximum is when $x \simeq 0.8r_0$
and it has the value $1.4\times 10^{-2} r_0 \sqrt{n_0}$ where $r_0$ is in kpc
and $n_0$ is the density in H atoms per $m^3$.
For the values $r_0 = 10$kpc and $n_0 = 10^6$ the maximum redshift
in velocity units is 140 km$s^{-1}$ which is within the range of observed values.
The difficulty with this model is that it predicts that the maximum spectral shifts
should occur along the line of sight whereas most observations show that the
maximum velocity gradient is along the major axis.
Although it is possible to devise density distributions that can explain
particular rotation curves there is no universal model that can explain all
rotation curves.
Nevertheless the fact that it predicts the magnitude and shape of a typical
galactic redshift curve must carry some weight.
If the size and shape of the halo is a simple function of the galaxies luminosity
this model can partly explain the Tully-Fisher correlation for spiral galaxies
(Rowan-Robinson 1984) and the Faber-Jackson correlation (Faber and Jackson 1976)
for elliptical galaxies.
They observed that the absolute magnitude of galaxies is correlated with
the width of the 21-cm emission line of neutral hydrogen.
If the line width is primarily due to gravitational interactions in the galactic
halo then its width $W_0$ is
$W_0 = A\pi(4G\rho_0)^{1/2} r_0$ where $A$ is a constant of order unity that
depends on the actual density distribution.
To proceed further requires knowledge of how the halo properties depend on
luminosity.
These two cases illustrate an important aspect of redshifts in this cosmology.
Although the
redshift is on average an excellent measure of distance any particular redshift is only a
measure of the gas in its line of sight.
Any lumpiness in the inter-cluster gas will produce apparent structure in redshifts that
could be falsely interpreted as structure in galaxy distributions.
That is, the apparent "walls", "holes", and other structures may be due to intervening
higher density or lower density clouds.
For example the model predicts an apparent hole behind clusters of galaxies because of
gravitational interactions in intra-cluster gas.
The velocity width of the hole would be of the same magnitude as the velocity dispersion
in the cluster.
For the Coma cluster the velocity width of this hole would vary from about
4100\,kms$^{-1}$ near the center of the cluster to about 1200\,kms$^{-1}$ near the edge.
\section{No evolution}
The most important observational difference between this cosmology and the big-bang
cosmology is that it obeys the perfect cosmological principle: it is homogeneous both
in space and time.
Consequently any unequivocal evidence of evolution would be fatal to its viability.
In contrast the big-bang theory demands evolution.
However it has the difficulty that the theory only provides broad guides as to what
that evolution should be and there is little communality between the evolution required
for different observations.
Nevertheless there is an entrenched view that evolution is observed in the characteristics
of many objects.
Two notable examples are the luminosity distribution of quasars and the angular-size
relationship for radio galaxies.
It will be shown that the observations for both of these phenomena are fully compatible
with a static cosmological model.
\section{Quasar luminosity distribution}
Because of their high redshifts quasars are excellent objects for probing the distant universe.
Since this cosmological model is static neither the density distribution nor the luminosity
distribution of any object should be a function of distance.
Consider the density distribution
$n\left(z\right)$ where z is the usual redshift parameter
$z=(\lambda _{\mbox{observed}}/\lambda _{\mbox{emitted}}-1$) then
\begin{equation}
\label{e28}
z=\exp \left( {Hr/c} \right)-1,
\end{equation}
where $r$ is the distance.
Since the range of $r$ is $0\le r\le \pi R$ the maximum value of z is 84.0 and its value
at the `equator' is 8.2.
Given that the geometry is that for a
three-dimensional hyper-spherical surface with radius R in a four-dimensional space the
volume out to a distance $r$ is
\begin{equation}
\label{e29}
V\left( r \right)=2\pi R^2\left( {r-{{R} \over {2}}\sin \left( {{{2r} \over {R}}}
\right)} \right)
\end{equation}
and the density distribution as a function of redshift for an object with a uniform
density of $n_0$
\begin{eqnarray}
\label{e30}
n\left( z \right)dz&=&n_0{{dV} \over {dr}}{{dr} \over {dz}}dz\nonumber \\
&=&{{4\pi R^2cn_0\sin ^2\left( {c\ln \left( {1+z} \right)/RH} \right)} \over {H\left( {1+z}
\right)}}dz.
\end{eqnarray}
From equations (\ref{e14}) and (\ref{e24}) we find that $HR=\sqrt 2\,c$ and equation
(\ref{e30}) becomes
\begin{equation}
\label{e31}
n\left( z \right)dz={{4\pi R^3n_0\sin ^2\left( {\ln \left( {1+z} \right)/\sqrt 2}
\right)} \over {\sqrt 2\left( {1+z} \right)}}dz,
\end{equation}
which has a maximum when $z=2.861$.
Now the difficulty of using equation (\ref{e31}) with observations is that most quasar
observations have severe selection effects.
Boyle et al (1990) measured the spectra of 1400 objects of which 351 were identified
as quasars with redshifts z $<$ 2.2.
The advantage of their observations is that their selection effects were well defined.
A full analysis is given in Crawford (1995b).
Let a source have a luminosity $L\left( \nu \right)$(Whz$^{-1}$) at the emission
frequency $\nu$.
Then if the energy is conserved the observed flux density
$S\left( \nu \right)$ (Wm$^{-2}$Hz$^{-1}$) at a distance $r$ is the luminosity
divided by the area which is
\begin{equation}
\label{e32}
S\left( \nu \right)={{L\left( \nu \right)} \over {4\pi R^2\sin ^2\left( {r/R}
\right)}}.
\end{equation}
However because of the gravitational interaction there is an energy loss such that the
received frequency $\nu _0$ is related to the emitted frequency $\nu _e$ by
\begin{equation}
\label{e33}
\nu _0=\nu _e\exp \left( {-Hr/c} \right)=\nu _e/\left( {1+z} \right).
\end{equation}
This loss in energy means that the observed flux density is decreased by a factor of $1+z$.
But there is an additional bandwidth factor that exactly balances the energy loss
factor.
In addition allowance must be made for K-correction (Rowan-Robinson 1985)
that relates
the observed spectrum to the emitted spectrum.
Since it is usual to include the bandwidth factor in the K-correction the apparent magnitude is
\begin{eqnarray*}
\label{e34}
m & =& -\case{5}{2}\log \left( {S\left( {\nu _0} \right)} \right)\nonumber \\
& =& -\case{5}{2}\log \left( {L\left( {\nu _0} \right)} \right)+
\case{5}{2}\log \left( {4\pi R^2} \right)
\nonumber \\
&+&
5\log \left( {\sin \left( {{{c\ln (1+z)}
\over {HR}}} \right)} \right) \\
&+&\case{5}{2}\log \left( {1+z} \right)
+K\left( z \right),
\end{eqnarray*}
where $K\left( z \right)$ is the K-correction and from above $RH=\sqrt{2}c$.
The result of the analysis was that the observations were fitted by a (differential)
luminosity function that had a Gaussian shape with a standard deviation of 1.52 magnitudes
and a maximum at $M=-22.2\,\mbox{mag}$ (blue).
The only caveat was that there appeared to be a deficiency of weak nearby quasars in the sample.
Since all cosmological models are locally Euclidean this must be a selection effect.
The fact that the absolute magnitude distribution had a well-defined peak and this was achieved
without requiring any evolution is strong support for the static model.
\section{Angular size of radio sources}
For the geometry of the hyper-sphere the observed angular size $\theta $ for an object with
a redshift of z and projected linear size of D is
$\theta =D/\left( {R\sin \left( {r/R} \right)} \right)$, and in terms of redshift it is
\begin{eqnarray*}
\label{e35}
\theta &=&{D \over {R\sin \left( {c\ln \left( {1+z} \right)/RH} \right)}}\\
&=&{D \over {R\sin \left( {\ln \left( {1+z} \right)/\sqrt 2} \right)}}.
\end{eqnarray*}
The angular size decreases with z until $z=8.2$ where there is a broad minimum and then
it increases again.
This model was used (Crawford 1995a) to analyze 540 double radio sources (all
Faranoff-Riley type II) listed by Nilsson et al (1993).
The result was an excellent fit to the radio-source size measurements, much better than
the big-bang model with a free choice of its acceleration parameter.
\section{Surface brightness of galaxies}
In an expanding universe, bolometric surface brightness will decrease with redshift as
$(1+z)^{-4}$ while, in a nonexpanding cosmology with tired-light it will decrease as
$(1+z)^{-1}$.
Because it is independent of the geometry of space it is a powerful test for discriminating
between the two cosmologies. The problem is that the measurement of surface brightness
is very difficult. Recently Pahre et al (1996) have reported measurements of surface
brightness for elliptic galaxies in the clusters Abell 2380 ($z=0.23$) and Abell 851 ($z=0.41$)
and have compared them with the nearby Coma cluster.
Their final results are surface brightness measurements for an average elliptic galaxy in
the K, B and $R_c$ spectral bands for the Coma and Abell 851 clusters and only the K band
for cluster Abell 2380.
Although they claim that the results are inconsistent with a tired light cosmology
their claim is only true for the K band data.
The B and $R_c$ band data are consistent with a tired light cosmology.
In addition the K-corrections for the K band data seem to have been computed using
evolving galaxy models whereas for a valid comparison it should be done for a
non-evolving galaxy.
Although the difference is small the K corrections are commensurate with the
discrepancy with the tired light model.
It would be more convincing if the K band observations for Abell 851 could be done
with a filter redshifted by a factor of 1.41.
\section{Other evidence for evolution}
There are however more direct observations of evolution that will be discussed.
They are the
distribution of absorption lines in quasar spectra, the measurement of the
microwave background temperature at high redshift, the time dilation of the type I
supernova light curves at large distances and the Butcher-Oemler effect.
For this static cosmology consider a uniform distribution of objects with number density N and cross-sectional area A then their distribution in redshift along
a line of sight is (here $\gamma $ is the exponent and not the Lorentz velocity parameter)
\begin{equation}
\label{e36}
{{dN} \over {dz}}={{NAc} \over {H}}\left( {1+z} \right)^\gamma \ .
\end{equation}
with $\gamma =-1$.
If the absorption lines seen in the spectra of quasars are due to absorption by
the Lyman-$\alpha$ line of hydrogen in intervening clouds of gas and with a
uniform distribution of clouds their predicted redshift distribution should
have $\gamma =-1$.
However observations (Hunstead et al 1988; Morris et al 1991; Williger et al 1994;
Storrie-Lombardie et al 1997; Barlow \& Tytler 1998) show exponents that range from 0.8 to 4.6.
Although there is poor agreement amongst the observations clearly they are all
in disagreement with this model.
The recent observations of Barlow and Tytler (1998) are of interest in that for the
Lyman-$\alpha$ lines they get $\gamma \simeq 1$ but for C IV $\lambda1548.20$ absorption lines
they find that the result from Steidel (1990) of $\gamma = -2.35\pm 0.77$ is inconsistent
with their low z data point and that equation \ref{e36} has a very poor fit.
Observations of absorption lines have complications due to lack of resolution causing
lines to be merged and that only a limited range in z (from Lyman-$\alpha$ to
Lyman-$\beta$) is available from each quasar.
However the major change required in the interpretation of the results for the static
cosmology is in the explanation for the broad absorption lines.
Traditionally these have been ascribed to Doppler
broadening from bulk motions in the clouds but it is also possible that they are due to
energy loss by the gravitational interaction.
For example using equation (\ref{e24a}) the `velocity' width of a cloud of diameter
$10^4$\,pc and density $10\,m_Hm^{-3}$ is $16\,\mbox{km}\cdot \mbox{s}^{-1}$ which is
typical of the observed line widths.
For a typical column density of
$ N_{\mbox{H1}}=10^{15}\,\mbox{cm}^{-2}$ this
cloud would have a ratio of H$_1$ to ionized hydrogen of
$3\times 10^{-5}$.
A further consequence is that because of the clouds the observed redshift
is not a valid measure of the true distance.
For example suppose the quasar is located in a galactic cluster where we would
expect a high local concentration of clouds then its redshift would be increased
over that expected for the cluster by the extra energy loss in the clouds.
The conclusion is that until the nature if the absorption lines are better
understood and analyzed in the context of this theory the evidence for
evolution is not convincing.
Another observation that could refute this theory is if the cosmic microwave background
radiation has a higher temperature at large distances.
Ge et al (1997) measured the absorption from the ground and excited states
of C1 (with a redshift of 1.9731) in the quasar QSO 0013-004.
They measure the strengths of the J=0 and J=1 fine structure levels and derived an
excitation temperature of $11.6\pm 1.0$\,K which after corrections gives a temperature
for the surrounding radiation of $7.9\pm 1.0$\,K that is in good agreement with the
redshifted temperature of 8.1\,K.
On face value this is clear evidence for evolution.
But not only are the measurements difficult they are based on a model for line widths
that does not include broadening due to the gravitational interaction.
Until this is done and the results are confirmed for other quasars and by other
observers a static cosmology is not refuted.
Programs that search for supernovae in high redshift galaxies with large telescopes are now
finding many examples and more importantly some are being detected before they reach their
maximum intensity.
Leibundgut et al (1996), Goldhaber et al (1996), and Riess et al (1997) have reported on
type 1a supernovae in which they believe that they can identify the type of supernova from
its spectral response and by comparing the supernova light curves with reference templates
they measure a time dilation that corresponds to that expected for their redshift in a
big-bang cosmology.
In addition there is a significant correlation between the rate of decay of the light curve
and the intrinsic luminosity (Riess, Press and Kirshner 1996) in that brighter supernovae
have a slower decline.
Hence there may be a bias due to selection effects and the cosmological model used
to get the absolute luminosity that is needed to correct for the correlation.
However because of this correlation and of uncertainties in matching the exact type
of supernova and because of the
occurrence of individual inhomogeneities many more observations are needed before these results
are well established.
The Butcher-Oemler (1978) effect is the observation that galaxies at redshifts $z \geq 0.3$
the galaxies in rich clusters tend to be bluer than is typical of nearby clusters.
Couch et al (1998) have found significant differences in their study of three rich clusters
at a redshift of $z=0.31$.
However Andreon and Ettori (1999) looked at a larger sample of x-ray selected clusters and
found no evidence for the Butcher-Oemler effect.
Their argument is that the effects that are observed are due to selection criteria rather
than differences in look back time.
As well as lack of unambiguous observations the effect (when present) is only seen at redshifts
up to $z \sim 1$ which is only relevant to the local inhomogeneity.
The conclusion is that the Lyman-$\alpha$ forest observations and the cosmic background radiation
temperature observations need to be re-evaluated within the static cosmological model in
order to see if they show evolution and refute the model.
The supernovae results are essentially unchanged in the static model and if they hold up they
make a strong case for evolution that would refute any static model.
The Butcher-Oemler effect observations are still not strong enough to make a good argument
against a static homogeneous universe.
\section{Nuclear abundance}
In this cosmology the universe is dominated by a very high temperature plasma.
Galaxies
condense from this plasma, evolve and die.
Eventually all of their matter is returned to the plasma.
Although nuclear synthesis in stars and supernova can produce the heavy elements it cannot
produce the very light elements.
In big-bang cosmology these are produced early in the expansion when there were high
temperatures and a large number of free neutrons.
This mechanism is not available in a static cosmology.
Nevertheless the temperature of the plasma ($2\times 10^9\,\mbox{K}$) is high enough
to sustain nuclear reactions.
The end result is an abundance distribution dominated by hydrogen and with smaller
quantities of helium and other light elements.
The problem is that the density is so low that the reaction rates may be too small achieve
equilibrium within the recycling time.
Naturally much further work is needed to quantify this hypothesis.
\section{Entropy}
Nearly every textbook on elementary physics quotes a proof based on the second law of
thermodynamics to show that the entropy of the universe is increasing but this is in direct
conflict with the perfect cosmological principle where total entropy is constant.
The conflict can be resolved if it is noted that the formal proof of the second law of
thermodynamics requires consideration of an isolated system and the changes that occur
with reversible and irreversible heat flows between it and its surroundings.
Now there is no doubt that irreversible heat flows occur and lead to an overall increase in
entropy.
However the formal proof is flawed in that with gravitational fields one cannot have an
isolated system.
There is no way to shield gravity.
Furthermore in their delightful book Fang \& Li (1989) argue that a self-gravitating system
has negative thermal capacity and that such systems cannot be in thermal equilibrium.
The crux of their argument is that if energy is added to a self-gravitating system, such as the
solar system, then the velocities and hence the `temperature' of the bodies decrease.
What happens is that from the virial theorem the potential energy (with a zero value for a
fully dispersed system) is equal to minus twice the kinetic energy and the total energy is
the sum of the potential and kinetic energies which is therefore equal to minus the kinetic
energy.
Thus we must be very careful in applying simple thermodynamic laws to gravitational systems.
Now consider the gravitational interaction where photons lose energy to the background
plasma.
Since this process does not depend on temperature it is not a flow of heat energy rather it
is work.
Nevertheless we can compute the entropy loss from the radiation field, considered as a heat
reservoir, as $-W/T_r$ where W is the energy lost, and similarly the entropy gained by the
plasma as $W/T_e$.
Then since $T_e>>T_R$ there is a net entropy loss.
Thus this gravitational interaction not only produces the Hubble redshift but it also acts
to decrease the entropy of the universe thereby balancing the entropy gained in irreversible
processes such as the complementary interaction where electrons lose energy to the radiation
field.
\section{Olber's paradox}
An essential requirement of any cosmology is to be able to explain Olber's paradox (or more
correctly de Chesaux's paradox; Harrison 1981) as to why the sky is dark at night.
For the big-bang cosmology although the paradox is partly explained by the universal redshift
the major reason is that the universe has a finite lifetime.
For this static cosmology the explanation is entirely due to the redshift.
The further we look to distant objects the more the light is redshifted until it is shifted
outside our spectral window.
Thus in effect we only see light from a finite region.
Note that the energy lost by the photons is returned to the inter-galactic plasma as part of
a cyclic process.
\section{Conclusion}
The introduction of curvature pressure has wide ranging astrophysical applications. It is
possible that it may resolve the solar neutrino problem but this must await a full analysis
using the standard solar model.
Although the theory does not prevent the formation of a black hole from cold matter it does
have an important effect on the high temperature accretion rings and curvature pressure
may provide the engine that produces astrophysical jets.
The greatest strength of this model is that it shows how a stable and static cosmology
may exist within the framework of general relativity without a cosmological constant.
The model with a homogeneous plasma depends only on one parameter, the average density
which from X-ray observations is taken to be $2.05m_H \mbox{ m}^{-3}$.
It then predicts that the plasma has a temperature of $2\times 10^9\,K$ and that the
universe has a radius given by equation (\ref{e14}).
It has been shown that for a simple inhomogeneous density distribution the predicted
temperature could easily be much lower and it could be in agreement with the temperature
observed for the X-ray background radiation.
Inclusion of the gravitational interactions permits the prediction of a Hubble constant
of $H=60.2\,\mbox{km}\cdot \mbox{s}^{-1}\cdot \mbox{Mpc}^{-1}$ and a microwave
background radiation with a temperature of $3.0\,K$.
Dark matter does not exist but arises from assuming that non-cosmological redshifts are
genuine velocities and then using the virial theorem.
In this static model most of the non-cosmological velocities are due to gravitational
interactions in intervening clouds.
Analysis of the observations for quasar luminosities and the angular size of radio sources
shows that they can be fully explained in a static cosmology without requiring any evolution.
The implication is that many other observations that require evolution in the big-bang
cosmology need to be re-examined within the static cosmology before evolution can be confirmed.
The strong evolution shown in the distribution of absorption lines (the Lyman-$\alpha$ forest)
is a problem for the static model.
However because of the gravitational interaction that can cause line broadening and the
possibility that some of the redshift may come from the clouds that produce the absorption
lines the results cannot at this stage be taken as a refutation of the static model.
Although the observations of a redshifted background microwave temperature and the evidence
of time dilation in the decay curves of type 1a supernovae appear to show direct evolution
it is too early to be certain.
These observations need better statistics and should be analyzed within this static model
before their apparent evolution is convincing.
The model includes a qualitative model for the generation of the light elements in the
high temperature inter-galactic plasma.
It was also argued that the effects of gravitational interaction of the microwave background
radiation that transfers energy to the high temperature plasma decreases entropy so that overall
total entropy of the universe is constant.
Finally the sky is dark at night because the light from distant stars is redshifted out of our
spectral window.
An important characteristic of this static cosmology is that it is easily refuted: any
unequivocal evidence for evolution would disprove the model.
Apart from evolution the most discriminating test that chooses between it and the big-bang
cosmology would be to compare the differential velocities of galaxies in a cluster with
their distance.
Whereas the big-bang model requires that there is no correlation this static cosmology requires
that the more distant galaxies will have larger redshifts.
\section{Acknowledgments}
I wish the thank the referee for many critical comments of the original draft
and for suggesting several recent references.
This work is supported by the Science Foundation for Physics within the University of Sydney,
and use has made of NASA's Astrophysics Data System Abstract Service.
\section{References}
\setlength{\leftmargin=1em}
\setlength{\itemindent=-\leftmargin}
\reference{1} Allen, C. W. 1976, Astrophysical Quantities, 3rd ed. (Athlone; London)
\reference{1a} Andreon, S., \& Ettori, S., 1999, \refZjnl{Astrophys. J.}, {\bf 516}, 647.
\reference{2} Bahcall, J. 1989, Neutrino Astrophysics, (Cambridge University Press; Cambridge).
\reference{3} Bahcall, J. 1997, Proc. of the 18th Texas Symposium on Relativistic Astrophysics,
eds, A. Olinto, J. Frieman, \& D. Schramm (World Scientific; Singapore).
\reference{4} Barlow, T. A. \& Tytler, D. 1998, \aj, {\bf 115}, 1725.
\reference{4a} Braginsky, V. B. \& Panov, V. I., 1971, {\it Zh. Eksp. \& Teor. Fiz.},
{\bf 61}, 873.
\reference{4a} Boyle, B. J., Fong, R., Shanks, T., \& Peterson, B. A., 1987, \refZjnl{Mon. Not. R. Astron. Soc.},
{\bf 227}, 717.
\reference{5} Butcher, H., \& Oemler, A., 1978, \refZjnl{Astrophys. J.}, {\bf 219}, 18.
\reference{8} Crawford, D. F. 1987a, Australian J. Phys., {\bf 40}, 449.
\reference{9} Crawford, D. F. 1987b, Australian J. Phys., {\bf 40}, 459.
\reference{10} Crawford, D. F. 1991, \refZjnl{Astrophys. J.}, {\bf 377}, 1.
\reference{11} Crawford, D. F. 1993, \refZjnl{Astrophys. J.}, {\bf 410}, 488 .
\reference{12} Crawford, D. F. 1995a, \refZjnl{Astrophys. J.}, {\bf 440}, 466.
\reference{13} Crawford, D. F. 1995b, \refZjnl{Astrophys. J.}, {\bf 441}, 488.
\reference{13a} Faber, S. M. \& Jackson, R. E., 1976, \refZjnl{Astrophys. J.}, {\bf 204}, 668.
\reference{14} Fabian, A. C. \& Barcons, X. 1992, \araa, {\bf 30}, 429.
\reference{15} Fang, Li-Zhi, \& Li, Shu Xian 1989, Creation of the Universe,
(World Scientific; Singapore).
\reference{16} Harrison, E. R. 1981, Cosmology, (Cambridge University Press; Cambridge).
\reference{17} Hunstead, R. W., Murdoch, H. S., Pettini, M., \& Blades, J. C. 1988, \refZjnl{Astrophys. J.},
{\bf 329}, 527.
\reference{18} Ge, J., Bechtold, J., \& Black, J. H. 1997, \refZjnl{Astrophys. J.}, {\bf 474}, 67.
\reference{19} Goldhaber, G., Deustua, S., Gabi, S>, Groom, D., Hook, I., Kim, A.,
Kim, M., Lee, J., Pain, R., Pennypacker, C>, Perlmuter, S., Small, I., Goobar, A.,
Ellis, R., McMahon, R., Boyle, B., Bunclark, P., Carter, D., Glazebrook, K., Irwin, M.,
Newberg, H., Filippenko, A. V., Matheson, T., Dopita, M., Mould, J., \& Couch, W,
1996, Thermonuclear Supernovae (NATO ASI), eds.,
R. Canal, P. Ruiz-Lapuente \& J. Isern (NATO ASI Ser. C, 486),(Kluwer: Dordrecht).
\reference{21} Gorenstein, P., Huchra, J. P., \& de Lapparent, V. 1979, in IAU Symposium 124,
\reference{22} de Groot, S. R. ,Leeuwen, W. A.,\& van Weert, C. G. 1980,
Relativistic Kinetic Theory (North-Holland; Amsterdam).
\reference{23} Jackson, J. D. 1975, Classical Electrodynamics, (John Wiley; New York).
\reference{23a} LaViolette, P. A., 1986, \refZjnl{Astrophys. J.}, {\bf 301}, 544.
\reference{25} Leibundgut, B.,et al., R. 1996, \refZjnl{Astrophys. J. Lett.}, {\bf 466}, L21.
\reference{26} Mather, J. C., Cheng, E. S., Cottingham, D. A., Eplee Jr, R. E.,
Fixsen, D. J., Hewagama, T., Isaacman, R. B., Jensen, K. A., Meyer, S. S.,
Noerdlinger, P. D., Read, S. M., Rosen, L. P., Shafer, R. A., Wright, E. L.,
Bennett, C. I., Boggess, N. W., Hauser, M. G., Kelsall, T., Moseley Jr, S. H.,
Silverberg, R. F., Smoot, G. F., Weiss, R., \& Wilkson, D. T., 1994, \refZjnl{Astrophys. J.}, {\bf 420}, 439.
\reference{27} Misner, C. W., Thorne, K. S., \& Wheeler, J. A. 1973, Gravitation,
(Freeman; San Francisco).
\reference{28} Nilsson, K., Valtonen, M. J., Kotilainen, J., \& Jaak\-kola, T. 1993, \refZjnl{Astrophys. J.},
{\bf 413}, 453.
\reference{28a} North, J. D., "The Norton History of Astronomy and Cosmology",
1995, (W. W. Norton \& Co, N.Y.).
\reference{28a} Peebles, P. J., 1993, 'Principles of Physical Cosmology', (Dover: New York).
\reference{29} Pound, R. V., \& Snider, J. L. 1965, \prb {\bf 140}, 788.
\reference{30} Raychaudhuri, A. K. 1955, Phys. Rev., {\bf 98}, 1123.
\reference{31} Reber, G., 1982, Proc ASA {\bf 4}, 482.
\reference{32a} Riess, A. G., s, W. RH., Kirshner, R. P., 1996, \refZjnl{Astrophys. J.}, {\bf 473}, 88.
\reference{32} Riess, A. G., et al., R. C. 1997, \aj, {\bf 114}, 722.
\reference{33} Roll, P. G., Krotov, R., \& Dicke, R. H. 1964, Annals of Physics, {\bf 26}, 442.
\reference{33a} Rowan-Robinson, M, 1984, "The Cosmological Distance Ladder",
(W. H. Freeman).
\reference{33b} Steidel, C. C., 1990, \apjs, {\bf 72}, 1.
\reference{34} Storrie-Lombardi, L. J., McMahon, R. G., Irwin, M. J., \& Hazard, C. 1997,
ESO Workshop on QSO Absorption Lines, \refZjnl{Astrophys. J.}, {\bf 468}, 121.
\reference{35} Trimble, V. 1987, \araa, {\bf 25}, 425.
\reference{36} Weinberg, S. 1972, Gravitation and Cosmology (Wiley; New York).
\reference{37} Williger, G. M., Baldwin, J. A., Carswell, R. F., Cooke, A. J., Hazard, C.,
Irwin, M. J., McMahon, R. G., \& Storrie-Lombardi, L. J. 1994, \refZjnl{Astrophys. J.}, {\bf 428}, 574.
\end{document}
|
\section*{{\sc Preliminaries}}
\hspace{0.4cm}
{\bf Assumptions.}
A long time ago, the Great Programmer wrote a program that
runs all possible universes on His Big Computer.
``Possible'' means ``computable'':
(1) Each universe evolves on a discrete time scale.
(2) Any universe's state at a given time is
describable by a finite number of bits.
One of the many universes is ours, despite some who evolved
in it and claim it is incomputable.
{\bf Computable universes.}
Let $TM$ denote an arbitrary universal Turing machine
with unidirectional output tape. $TM$'s
input and output symbols are ``0'', ``1'', and ``,'' (comma).
$TM$'s possible input programs can be ordered alphabetically:
``'' (empty program),
``0'',
``1'',
``,'',
``00'',
``01'',
``0,'',
``10'',
``11'',
``1,'',
``,0'',
``,1'',
``,,'',
``000'', etc.
Let $A_k$ denote $TM$'s $k$-th program in this list.
Its output will be a finite or infinite string over the alphabet
$\{$ ``0'',``1'',``,''$\}$. This
sequence of bitstrings separated by commas
will be interpreted as
the evolution $E_k$
of universe $U_k$.
If $E_k$ includes at least one comma,
then let $U_k^l$
denote the $l$-th (possibly empty) bitstring
before the $l$-th comma.
$U_k^l$ represents $U_k$'s state at the $l$-th time step of $E_k$
($k, l \in \{1, 2, \ldots, \}$).
$E_k$ is represented
by the sequence $U_k^1, U_k^2, \ldots$
where $U_k^1$ corresponds to $U_k$'s big bang.
Different algorithms may compute the same universe.
Some universes are finite (those whose programs
cease producing outputs at some point), others are not.
I don't know about ours.
{\bf TM not important.}
The choice of the Turing machine is not important.
This is due to the compiler theorem:
for each universal Turing machine $C$ there exists a constant prefix $\mu_C$
$\in \{$ ``0'',``1'',``,''$\}^*$
such that for all possible programs $p$, $C$'s output in response to
program $\mu_C p$ is identical to
$TM$'s output in response to $p$.
The prefix $\mu_C$ is the compiler that compiles programs for $TM$ into
equivalent programs for $C$.
{\bf Computing all universes.}
One way of sequentially computing all computable universes is
dove-tailing. $A_1$ is run for one instruction every second step, $A_2$ is
run for one instruction every second of the remaining steps, and so on.
Similar methods exist for computing many universes in parallel.
Each time step of each universe that is computable by at least one
finite algorithm will eventually be computed.
{\bf Time.}
The Great Programmer does not worry about computation time.
Nobody presses Him.
Creatures which evolve in any of the universes don't have to worry
either. They run on local time and have no idea of how many instructions
it takes the Big Computer to compute one of their time steps, or how many
instructions it spends on all the other creatures in parallel universes.
\section*{{\sc Regular and Irregular Universes }}
\hspace{0.4cm}
{\bf Finite histories.}
Let $\mid x \mid$ denote the number of symbols in string $x$.
Let the partial history $S_k^{i,j}$ denote the substring between
the $i$-th and the $j$-th symbol of $E_k$, $j > i$.
$S_k^{i,j}$ is regular (or compressible, or non-random)
if the shortest program that computes $S_k^{i,j}$
(and nothing else)
{\em and halts} consists of less than $ \mid S_k^{i,j} \mid $ symbols.
Otherwise $S_k^{i,j}$ is irregular (incompressible, random).
{\bf Infinite histories.}
Similarly,
if some universe's evolution is infinite,
then it is compressible if it can be computed by a finite algorithm.
{\bf Most universes are irregular.}
The evolutions of almost all universes are incompressible.
There are $3^n$ strings of size $n$, but less than
$(1/3)^c * 3^{n} << 3^n$ algorithms
consisting of less than $n-c$ symbols
($c$ is a positive integer).
And for the infinite case, we observe:
the number of infinite symbol strings is incountable.
Only a negligible fraction (namely countably many of them)
can be computed by finite programs.
{\bf The few regular universes.}
There are a few compressible universes which can be computed by very
short algorithms, though. For instance, suppose that some $U_k$ evolves according
to physical laws that tell us how to compute next states from previous states.
All we need to compute $U_k$'s evolution is $U_k^1$ and the algorithm that computes
$U_k^{i+1}$ from $U_k^i$ ($i \in \{1, 2, \ldots, \}$).
{\bf Noise?}
Apparently, we live in one of the few highly regular universes.
Each electron appears to behave the same way.
Dropped breads of butter regularly hit the floor, not the ceiling.
There appear to be deviations from regularity, however,
embodied by what we call noise.
Although certain macroscopic properties (such as pressure in a
gas container) are predictable by physicists,
microscopic properties (such as precise particle positions) seem
subject to noisy fluctuations. Noise represents additional information
absent in the original physical laws. Uniform noise is
incompressible --- there is no short algorithm that computes it and
nothing else.
{\bf Noise does not necessarily prevent compressibility.}
Laws currently used by physicists to model our own universe model
noise. Based on Schr\"{o}dinger's equation,
they are only conditional probability
distributions on possible next states, given previous states.
The evolution of Schr\"{o}dinger's wave function (WF) itself can be computed by
a very compact algorithm (given the quantizability assumptions in the first
paragraph of this paper) --- WF is just a short formula.
Whenever WF collapses in a particular way, however, the
resulting actual state represents additional information (noise) not conveyed
by the algorithm describing the initial state (big bang) and WF.
Still, since the noise
obviously is non-uniform (due to the nature of the physical laws and
WF), our universe's evolution so far is greatly compressible. How?
Well, there is a comparatively short algorithm that
simply codes probable next states by few bits,
and unlikely next states by many bits,
as suggested by standard information
theory \cite{Shannon:48}.
{\bf More regularity than we think?}
The longer the shortest program computing a given universe,
the more random it is.
To certain observers, certain universes appear partly
random although they aren't.
There may be at least two reasons for this:
{\bf 1. Shortest algorithm cannot be found.}
It can be shown that
there is no algorithm that can generate the shortest
program for computing arbitrary given data on a given computer
\cite{Kolmogorov:65,Solomonoff:64,Chaitin:87}.
In particular, our physicists cannot expect to find the
most compact description of our universe.
{\bf 2. Additional problems of the Heisenberg type.}
Heisenberg tells us that we cannot even observe the precise, current state of a
single electron, let alone our universe. In our particular
universe, our actions seem to influence our measurements in a fundamentally
unpredictable way.
This does not mean that there is no predictable
underlying computational process (whose precise results we cannot access).
In fact, rules that hold for observers who are
part of a given universe and evolved according to its
laws need not hold outside of it. There is no reason to believe that
the Great Programmer cannot dump a universe and examine its precise state at any
given time, just because the creatures that evolved in it cannot because their
measurements modify their world.
{\bf How much true randomness?}
Is there ``true'' randomness in our universe,
in addition to the simple physical laws?
True randomness essentially means
that there is no short algorithm computing ``the precise collapse of the
wave function'', and what is perceived as noise by today's physicists.
In fact, if our universe was infinite, and there was true randomness,
then it could not be computed by a finite algorithm that computes nothing
else. Our fundamental inability to perceive
our universe's state does {\em not} imply its true randomness, though.
For instance, there may be a very short algorithm computing the positions
of electrons lightyears apart in a way that seems like noise to
us but actually is highly regular.
\section*{{\sc All Universes are Cheaper Than Just One}}
In general, computing all evolutions of all universes is
much cheaper in terms of information requirements than computing just
one particular, arbitrarily chosen evolution. Why?
Because the Great Programmer's algorithm that systematically
enumerates and runs all universes
(with all imaginable types of physical laws, wave functions, noise etc.)
is {\em very} short (although it takes time).
On the other hand,
computing just one particular universe's evolution (with, say, one particular
instance of noise), without computing the others,
tends to be very expensive,
because almost all individual universes are
incompressible, as has been shown above.
More is less!
{\bf Many worlds.}
Suppose there is true (incompressible) noise
in state transitions of our particular world evolution.
The noise conveys additional information
besides the one for initial state and physical laws.
But from the Great Programmer's point of view, almost no extra
information (nor, equivalently, a random generator) is required.
Instead of computing just one of the many possible evolutions of
a probabilistic universe with fixed laws but random noise of a
certain (e.g., Gaussian) type, the Great Programmer's
simple program computes them all.
An automatic by-product of the Great Programmer's set-up is
the well-known ``many worlds hypothesis'', \copyright Everett III.
According to it, whenever our
universe's quantum mechanics allows for alternative next paths,
all are taken and the world splits into separate universes.
From the Great Programmer's view, however,
there are no real splits --- there are just a bunch of
different algorithms which yield identical results for some time, until
they start computing different outputs corresponding to different noise
in different universes.
From an esthetical point of view that favors simple explanations
of everything, a set-up in which all possible
universes are computed instead of just ours
is more attractive. It is simpler.
\section*{{\sc Are we Run by a Short Algorithm? }}
Since our universes' history so far is regular,
it by itself {\em could} have been computed by a
relatively short algorithm.
Essentially, this algorithm embodies the physical
laws plus the information about the historical noise.
But there are many algorithms whose output sequences start with our
universe's history so far. Most of them are very long.
How likely is it now that our universe is indeed run by a short algorithm?
To attempt an answer, we need a prior probability on the possible
algorithms. The obvious candidate is the ``universal prior''.
{\bf Universal prior.}
Define $P_U(s)$, the {\em a priori probability} of a finite symbol
string $s$ (such as the one representing our universe's history so far),
as the probability of guessing a halting program
that computes $s$ on a universal Turing machine
$U$. Here, the way of guessing is defined
by the following procedure:
initially, the input tape consists of a single square.
Whenever the scanning head of the program tape shifts
to the right, do: (1) Append a new square.
(2) With probability $\frac{1}{3}$ fill it with a ``0'';
with probability $\frac{1}{3}$ fill it with a ``1'';
with probability $\frac{1}{3}$ fill it with a ``,''.
Programs are ``self-delimiting''
\cite{Levin:74,Chaitin:87}
--- once $U$ halts due
to computations based on
the randomly chosen symbols (the program) on its input tape,
there won't be any additional program symbols.
We obtain
\[
P_U(s) = \sum_{p: U~computes~s~from~p~and~halts} (\frac{1}{3})^{\mid p \mid}.
\]
Clearly,
the sum of all probabilities of all halting programs cannot exceed 1
(no halting program can be the prefix of another one).
But certain programs may lead to
non-halting computations.
Under different universal priors (based on different universal
machines), probabilities of a given string
differ by no more than a constant factor independent of the
string size, due to the compiler
theorem (the constant factor corresponds to the probability of
guessing a compiler).
This justifies the name ``{\em universal} prior,''
also known as Solomonoff-Levin distribution.
{\bf Dominance of shortest programs.}
It can be shown (the proof is non-trivial) that
the probability of
guessing any of the programs computing some string
and the probability of
guessing one of its shortest programs
are essentially equal (they differ by no more than a constant factor
depending on the particular Turing machine).
The probability of a string is dominated by the probabilities of its
shortest programs. This is known as the ``coding theorem'' \cite{Levin:74}.
Similar coding theorems exist for the case of non-halting programs
which cease requesting additional input symbols at a certain point.
Now back to our question: are we run by a relatively compact algorithm?
So far our universe {\em could} have been run by one ---
its history {\em could} have been much noisier and thus much less
compressible. Hence universal prior and coding theorems suggest that the
algorithm is indeed short. If it is, then there will be less than
maximal randomness in our future, and more than vanishing predictability.
We may hope that our universe will remain regular, as opposed to
drifting off into irregularity.
\section*{{\sc Life in a Given Universe}}
\hspace{0.4cm}
{\bf Recognizing life.}
What is life? The answer depends on the observer.
For instance, certain substrings of $E_k$ may be interpretable
as the life of a living thing $L_k$ in $U_k$.
Different observers will have different views, though. What's life to
one observer will be noise to another. In particular, if the observer
is not like the Great Programmer but
also inhabits $U_k$, then its own life may
be representable by a similar substring.
Assuming that recognition implies relating observations to previous knowledge,
both $L_k$'s and the observer's life will have to
share mutual algorithmic information \cite{Chaitin:87}:
there will be a comparatively
short algorithm computing $L_k$'s from the observer's life, and vice versa.
Of course, creatures living in a given universe don't
have to have any idea of the symbol strings by which they
are represented.
{\bf Possible limitations of the Great Programmer.}
He does not need not be very smart. For instance, in some of His
universes phenomena will appear that humans would call life.
The Great Programmer won't have to be able to recognize them.
{\bf The Great Programmer reappears.}
Several of the Great Programmer's universes
will feature another Great Programmer who programs another Big Computer
to run all possible universes. Obviously, there are infinite chains of
Great Programmers. If our own universe allowed for enough storage,
enough time, and fault-free computing, then you could be one of them.
\section*{{\sc Generalization and Learning}}
\hspace{0.4cm}
{\bf In general, generalization is impossible.}
Given the history of a particular universe up to a given
time, there are infinitely many possible continuations.
Most of these continuations have nothing to do with the previous history.
To see this, suppose we have observed
a partial history $S_k^{i,j}$ (the substring between
the $i$-th and the $j$-th symbol of $E_k$).
Now we want to generalize from previous
experience to predict
$S_k^{j+1,l}$, $l > j$.
To do this, we need an algorithm that computes $S_k^{j+1,l}$ from $S_k^{i,j}$
($S_k^{i,j}$ may be stored on a separate, additional
input tape for an appropriate
universal Turing machine).
There are $3^{l - j}$ possible futures.
But for $c < l - j$, there are less than $(1/3)^c * 3^{l - j}$ algorithms
with less than $l - j - c$ bits computing such a future,
given $S_k^{i,j}$.
Hence in most cases
the shortest algorithm computing the future, given the past,
won't be much shorter than the
shortest algorithm computing the future from nothing.
Both will have about the size of the entire future.
In other words, the mutual algorithmic information between history
and future will be zero.
As a consequence, in most universes (those that can be computed by
long algorithms only), successful generalization from previous experience is
not possible. Neither is inductive transfer.
This simple insight is related to
results in \cite{Wolpert:96}.
{\bf Learning.}
Given the above, since learning means to use previous experience
to improve future performance, learning is possible only in
the few regular universes
(no learning without compressibility).
On the other hand, regularity by itself is not sufficient
to allow for learning. For instance, there is a highly compressible and
regular universe represented
by ``,,,,,,,...''. It is too simple to allow for processes we would
be willing to call learning.
In what follows, I will assume that a regular universe
is complex enough to allow for
identifying certain permanent data structures of a general learner
to be described below. For convenience,
I will abstract from bitstring models, and instead talk about
environments, rewards, stacks etc. Of course, all these abstract
concepts are representable as bitstrings.
{\bf Scenario.}
In general, the learner's life is limited.
To it, time will be important (not to the Great Programmer though).
Suppose its life in environment $\cal E$
lasts from time 0 to unknown time $T$.
In between it repeats the following cycle
over and over again ($\cal A$ denotes a set of possible actions):
select and execute $a \in \cal A$ with probability $P( a \mid \cal E )$,
where the modifiable policy $P$ is a variable,
conditional probability distribution on the possible actions,
given current $\cal E$.
Action $a$ will consume time and may change $\cal E$ and $P$.
Actions that modify $P$ are called primitive learning algorithms (PLAs).
$P$ influences the way $P$ is modified (``self-modification'').
{\em Policy modification processes} (PMPs)
are action subsequences that include PLAs.
The $i$-th PMP in system life is denoted {\em PMP}$_i$,
starts at time $s_i > 0$, ends at $e_i < T$, $e_i > s_i$,
and computes a sequence of $P$-modifications denoted $M_i$.
Both $s_i$ and $e_i$ are computed dynamically
by special instructions in $\cal A$
executed according to $P$ itself: $P$ says when to start and end PMPs.
Occasionally $\cal E$ provides
real-valued reward. The cumulative reward obtained in between
time 0 and time $t > 0$ is denoted by $R(t)$ (where $R(0) = 0$).
At each PMP-start $s_i$ the learner's goal is to
use experience to generate $P$-modifications to
accelerate future reward intake.
Assuming that reward acceleration is possible at all,
given $E$ and $\cal A$, how can the learner achieve it?
I will describe a rather general way of doing so.
{\bf The success-story criterion.}
Each PMP-start time $s_i$ will trigger an evaluation
of the system's performance so far.
Since $s_i$ is computed according to $P$,
$P$ incorporates information
about when to evaluate itself.
Evaluations may cause
policy modifications to be undone (by restoring
the previous policy --- in practical implementations,
this requires to store previous values of modified
policy components on a stack).
At a given PMP-start $t$ in the learner's life,
let $V(t)$ denot the set of those previous $s_i$ whose
corresponding $M_i$ have not
been undone yet. If $V(t)$ is not empty,
then let $v_i ~~ (i \in \{1, 2, \ldots, \mid V(t) \mid \}$
denote the $i$-th such time, ordered according to size.
The success-story criterion
SSC is satisfied if either $V(t)$ is empty (trivial case) or if
\[
\frac{R(t)}{t}
<
\frac{R(t) - R(v_1)}{t - v_1}
<
\frac{R(t) - R(v_2)}{t - v_2}
<
\ldots
<
\frac{R(t) - R(v_{ \mid V(t) \mid }) }{t - v_{\mid V(t) \mid}}.
\]
SSC essentially says that
each surviving $P$-modification corresponds
to a long term reward acceleration.
Once SSC is satisfied, the learner
continues to act and learn
until the next PMP-start.
Since there may be arbitrary reward delays in response
to certain action subsequences, it is important that
$\cal A$ indeed includes actions for
delaying performance evaluations --- the learner will have
to learn when to trigger evaluations.
Since the success of a policy modification recursively depends on the
success of later modifications for which it is setting the stage,
the framework provides a basis for ``learning how to learn''.
Unlike with previous learning paradigms, the entire life is considered
for performance evaluations. Experiments in
\cite{Schmidhuber:97bias,Schmidhuber:97ssa}
show the paradigm's practical feasibility.
For instance, in \cite{Schmidhuber:97bias}
$\cal A$ includes an extension of Levin search \cite{Levin:84}
for generating the PMPs.
\section*{{\sc Philosophy}}
\hspace{0.4cm}
{\bf Life after death.}
Members of certain religious sects expect resurrection of the dead in a paradise
where lions and lambs cuddle each other. There is a possible continuation
of our world where they will be right. In other possible continuations, however,
lambs will attack lions.
According to the computability-oriented view adopted in this paper,
life after death is a technological problem, not a religious one.
All that is necessary for some human's resurrection is to record his
defining parameters (such as brain connectivity and synapse properties etc.),
and then dump them into a large computing device computing an appropriate
virtual paradise. Similar things have been suggested by
various science fiction authors.
At the moment of this writing, neither appropriate
recording devices nor computers of sufficient size exist. There is
no fundamental reason, however, to believe that they won't exist in
the future.
{\bf Body and soul.}
More than 2000 years of European philosophy dealt
with the distinction between body and soul.
The Great Programmer does not care.
The processes that correspond to our brain firing patterns and the
sound waves they provoke during discussions about body and soul
correspond to computable substrings of our universe's evolution.
Bitstrings representing such talk may evolve in many universes.
For instance, sound wave patterns representing notions
such as body and soul and ``consciousness'' may be useful in
everyday language of certain inhabitants of those universes.
From the view of the Great Programmer, though, such
bitstring subpatterns may be entirely irrelevant.
There is no need for Him to load them with ``meaning''.
{\bf Talking about the incomputable.}
Although we live in a computable universe, we occasionally chat
about incomputable things, such as the halting probability of a
universal Turing machine (which is closely related to G\"{o}del's
incompleteness theorem). And we sometimes discuss inconsistent
worlds in which, say, time travel is possible. Talk about such
worlds, however, does not violate the consistency of the processes
underlying it.
{\bf Conclusion.}
By stepping back and adopting the Great Programmer's point of view,
classic problems of philosophy go away.
\section*{{\sc Acknowledgments}}
Thanks to Christof Schmidhuber for interesting discussions on
wave functions, string theory, and possible universes.
\bibliographystyle{plain}
|
\section{Introduction}
Recently a new set of ideas was put forward, which was called ``the
holographic principle'' \cite{thooft,lenny1}.
According to this set of ideas,
under certain conditions all
the information about
a physical system is coded on its boundary, implying that the entropy
of a system cannot exceed its boundary area in Planck units.
This principle was motivated by the well-known result
in black
hole theory: the total entropy $S_m$ of matter inside of a black hole
cannot be greater than the Bekenstein-Hawking entropy, which is
equal to a quarter of the area of the event horizon in the Planck
units, $S_m \leq S_{BH} = {A \over 4}$ \cite{bekenstein}. One can
interpret
this result as a statement that all the information about the interior of
a black hole is stored on its horizon.
The main aim
of the holographic principle is to extend this
statement to a broader class of situations. This principle, in its
most radical form, would imply that our world
is two-dimensional in a certain sense, because all the
information about physical processes in
the world is stored at its surface.
This conjecture is very interesting, and physical implications of
its most radical version could be quite significant. There has been a lot
of activity related to the use of the holographic principle in quantum
gravity, string theory and M-theory.
For example, there is a
conjecture that the knowledge of a supersymmetric
Yang-Mills theory at the boundary of an Anti-de-Sitter space
may be sufficient
to restore
the information about
supergravity/string theory in the bulk \cite{malda}.
However, if one tries to apply the holographic principle to
cosmology, one
immediately recognizes several problems. For example, a
closed universe has finite size, but it does not have any boundary.
What is the
meaning of the holographic principle in such a case?
If the universe is infinite (open or flat), then it does not have
boundaries either.
In these cases,
one may try to compare the entropy inside of a box
of size $R$ with its area,
and then take the limit as $R \to \infty$. But in this limit the
entropy is
always larger than the area \cite{lenny3}.
Another possibility is to compare the area of a domain of the size of
the particle
horizon (the causally connected part of the universe) with the entropy
of matter
inside this domain.
But this is also problematic. The entropy produced during reheating after
inflation is proportional to the total volume of inflationary
universe. During
inflation, the volume inside the particle horizon grows as
$e^{3Ht}$, whereas
the area of the horizon grows as $e^{2Ht}$. Clearly, the entropy
becomes much
greater than the area of the horizon if the duration of
inflation is sufficiently large. This means that an inflationary
universe is not two-dimensional; information stored at its
``surface'' is not rich enough to describe physical processes in its
interior. In fact, one of the main advantages of inflation is the
possibility to study each domain of size $H^{-1}$ as an independent part
of the universe, due to the no-hair theorem for de Sitter space.
This makes the
events at the boundaries of an inflationary universe irrelevant for
the description of local physics \cite{book}.
Thus, the most radical version of the holographic principle seems to
be at
odds with inflationary
cosmology.
One may try to formulate a weaker form of this
principle, which may still be quite useful. For example, Fischler and
Susskind proposed to put constraints only on the part of the entropy
which passed through the backward light cone \cite{lenny3}. This
formulation
does not confront inflationary cosmology because it eliminates from the
consideration most of the entropy produced inside the light cone
during the
post-inflationary reheating of the universe. They further concentrated on
investigation of those situations where
cosmological evolution is adiabatic. From the point of view of
inflationary
cosmology, this means that they considered the evolution of the
universe after
reheating. The largest domain in which all of the entropy crossed
the boundary
when the evolution is adiabatic is bounded by the light cone
emitted {\it
after} inflation and reheating. In what follows we will loosely call this
light cone of size $O(H^{-1})$ ``particle horizon,'' even though the true
particle horizon, describing the light cone emitted at the beginning of
inflation, is exponentially large.
Fischler and
Susskind argued that in the
case of adiabatic evolution the total entropy of matter within
the particle horizon must
be smaller than the area of the horizon, $S \lesssim A $
\cite{lenny3}.
This conjecture is
rather nontrivial. Indeed, the origin
of the Bekenstein-Hawking constraint on the entropy of a black hole
is the existence of the {\it event} horizon, which serves as a natural
boundary for all processes inside a black hole. But there is no event
horizon in a non-inflationary universe, and the idea to replace it by
the {\it particle} horizon requires some justification. Also, the
Bekenstein-Hawking constraint on the entropy is valid even if the
processes
inside a black hole are non-adiabatic. Thus it would be desirable to
investigate this proposal and find a way to apply it to the
situations when the
processes can be non-adiabatic.
Remarkably, Fischler and
Susskind have shown that their conjecture is valid for a
flat universe
with all possible equations of state satisfying the condition
$0\leq p \leq \rho$.
This result suggests that there may be some deep
reasons for the validity of holography. However, they also
noticed that
their version of the holographic principle is violated in a closed
universe.
One may consider this observation
either as an indication that closed universes are
impossible or as a warning,
showing that the holographic
principle may require additional
justification and/or reformulation. Indeed, this
principle is not a rigid scheme but a theory
in the making. It may be quite successful in many respects, but one
should not be surprised to see some parts of its formulation change.
For example, Bak and Rey suggested to replace the particle
horizon by an
apparent horizon in the formulation of the holographic principle,
claiming that their proposal does not suffer from any problems in
the closed universe case \cite{brtwo}.
There were many attempts to apply various formulations of the holographic
principle to various cosmological models, but the existing literature on
cosmic holography is somewhat controversial. The entropy of the
observed
component of matter (such as photons) is well below $10^{90}$
\cite{book}. Meanwhile the constraint $S \lesssim A$
applied to our part of the universe implies that $S < 10^{120}$
\cite{lenny3}, which does not look particularly restrictive.
Holography could be quite
important if it were able to rule out
some types of cosmological models,
but this possibility depends on the
formulation and the range of validity of
the holographic principle. One may try
to use holography to solve the cosmological constant problem
\cite{banks,cohen}, but the progress
in this direction was very limited.
Recently it was claimed that holography puts strong constraints on
inflationary theory \cite{infl}, but the authors
of Ref. \cite{el}
argued that
this is not the case. Holographic
considerations were
used in investigation of the
pre-big bang theory \cite{br,bmp,gv},
and on the basis of this investigation
it was claimed that this theory solves
the entropy
problem in the pre-big bang theory\cite{gv}, which is at odds
with the results of \cite{klb}.
The main goal of this paper is to
examine the basic assumptions of cosmic
holography and check which of them may require
modifications. We will try to
find out whether holography indeed puts
constraints on various cosmological
models. We will show, in particular,
that the original formulation of
the holographic principle should be reconsidered
more generally, and not only when applied
to closed universes. The holographic entropy bound proposed in
\cite{lenny3}, as well as the
formulation proposed in \cite{brtwo}, is
violated at
late stages of evolution of
open, flat and closed universes
containing usual matter and a small
amount of negative vacuum energy density. At the
beginning of their evolution, such universes
cannot be distinguished from the universe with a positive or
vanishing vacuum energy density. Thus there is no obvious reason to
consider
such universes unphysical and rule them out.
However, when the density of
matter becomes diluted by expansion, a universe with a negative
vacuum energy collapses, and the condition
$S \lesssim A $ becomes violated long before the
universe reaches the Planck density.
The investigation of universes with a negative cosmological constant
gives an additional reason to look for a
reformulation of the cosmological
holographic principle.
Our approach will be most closely related to the approach outlined by
Easther and Lowe \cite{el}, and by Veneziano \cite{gv}. They argued
that the entropy of the interior of a domain of
size $H^{-1}$ cannot be greater than
the entropy of a black hole of a
similar radius. We will
extend their discussion and propose a
justification for the entropy bound obtained in Ref.
\cite{lenny3} for the case of an expanding noninflationary (or
post-inflationary) universe.
We will argue, in agreement with \cite{el,gv}, that in those
cases where the
holographic bound of Ref.
\cite{lenny3} is valid, it is equivalent to the Bekenstein-Hawking
bound, which does not require any assumptions about adiabatic evolution.
This bound alone cannot resolve the entropy problem for
the pre-big bang cosmology and does not lead to any constraints on
inflation.
\section{Cosmology and holography}
\subsection{Flat universe with $p = \gamma \rho$}
Let us begin with a brief review of \cite{lenny3}. We will restrict
our attention to the case when gravitational dynamics is given by
the Einstein's equations, and the evolution is adiabatic. First we
will consider flat homogeneous
and isotropic FRW universes, whose metric is
\begin{equation}
ds^2 = - dt^2 + a^2(t) \left({dr^2} + r^2 d \Omega \right) \ .
\label{metric}
\end{equation}
We will use the units $8\pi{G_N}=1$. For
simplicity we will consider matter with the energy-momentum tensor
$T_{\mu\nu}$ = diag$(\rho,p,p,p)$. The independent equations of
motion are
\begin{equation}
H^2 = \rho/3 \ ,~~~~~~~~ \dot \rho + 3H (\rho + p) =0 \ ,
\label{eoms}
\end{equation}
where $H = \dot a/a$ is the Hubble parameter, $\rho$
and $p$ are the energy density and pressure, and the overdot denotes
the time derivative. We will assume that $\rho > 0$, $p = \gamma
\rho$, and
that
the energy-momentum tensor satisfies the dominant energy condition
$|\gamma| \leq 1$. This will generalize the results of \cite{lenny3}
obtained
for $0\leq \gamma \leq 1$,
and is in fact the correct sufficient condition
for the validity of the holographic bounds in flat
and open
FRW universes.
The solutions of (\ref{eoms}) for $\gamma > -1$ can be written
as
\begin{equation} a(t) = t^{\frac{2}{3(\gamma+1)}}\ .
\label{solscale}
\end{equation}
Here we took by definition $a = 1$ at
the Planck time $t = 1$. Density decreases as $ \rho =
\frac{\rho_0}{a^{3(\gamma+1)}}$, where $\rho_0 = {4\over 3 (\gamma +
1)^2}$ is the
density at $t = 1$.
(For $\gamma =-1$ one has the usual de
Sitter solution.) The particle horizon is defined by the distance
covered by the light cone emitted at the singularity $t = 0$:
\begin{equation}
\label{horizon} L_H(t) = a(t) \int^t_{0} \frac{dt'}{a(t')} = a(t)
r_H(t) \ ,
\label{parthor}
\end{equation}
where $r_H$ is the comoving size of the
horizon defined by the condition ${dt\over a } = dr_H$. Suppose
first that
$\gamma > -1/3$. Then the comoving
horizon is
\begin{equation} r_H = L_H/a = {3(\gamma+1)\over 3\gamma +1}~
t^{3\gamma+1\over 3(\gamma+1)} \ ,
\label{comhorizon}\end{equation}
and
\begin{equation} L_H = {3(\gamma+1)\over 3\gamma +1} t = {2\over
3\gamma
+1} H^{-1} \ .
\label{comhorizon3}\end{equation}
At the Planck time $t = 1$ one has $ L_H = {3(\gamma+1)\over3\gamma
+1}$
which generically is $O(1)$. The volume of space within the
distance $L_H$ from any point was also $O(1)$. The entropy density at
that time could not be greater than $O(1)$, so one may say that
initially $\left({S\over A}\right)_0=\sigma \lesssim 1$. Later the
total entropy inside the horizon grows as
$\sigma L_H^3/a^3$, whereas the
total area $A$ of the particle horizon grows as $L_H^2$.
Therefore
\begin{equation} {S\over A} \sim \sigma {L_H\over a^3} =
\sigma {r_H\over a^2} \ .
\label{comhorizon2}\end{equation}
This yields
\begin{equation}
{S\over A} \sim \sigma t^{\gamma-1\over
\gamma+1}\ .
\label{comhorizon2a}
\end{equation}
Thus the ratio ${S\over
A}$ does not increase in time for $1\ge \gamma > -1/3$,
so if the holographic
constraint ${S\over A}\lesssim 1$ was satisfied at the Planck time,
later on it will be satisfied even
better \cite{lenny3}.
A similar result can be obtained for $-1 \leq \gamma \leq -1/3$.
However,
investigation of this case involves several subtle points. First of
all, in
this case the integral in Eq. (\ref{parthor}) diverges at small $t$.
This is
not a real problem though. It is resolved if one defines the
particle horizon
as an integral not from $t = 0$, but from the Planck time $t = 1$.
A more serious issue is the assumption of adiabatic expansion of the
universe. If one makes this assumption, then one can show that the
holographic
bound is satisfied for all $\gamma$ in the interval $-1\le \gamma
\le 1$, which
generalizes the result obtained in \cite{lenny3}. However, the
universe with
$1+\gamma \ll 2/3$
(i.e. with $\gamma \approx -1$) is inflationary.
The density
of matter after inflation becomes negligibly small, so it must be
created again
in the process of reheating of the universe. This process is strongly
nonadiabatic.
As we already mentioned in the Introduction, in inflationary
cosmology the
bounds of Ref. \cite{lenny3} refer to the post-inflationary particle
horizon,
which means that the integration in Eq. (\ref{parthor}) should begin
not at $t =
0$ or at $t = 1$ but after reheating of the universe. One can easily
verify
that
the bounds
obtained in \cite{lenny3} are valid in this case as well.
\subsection{Closed universe}
The metric of a closed FRW universe is
\begin{equation}
\label{closed} ds^2 = - dt^2 + a^2(t) (d\chi^2 + \sin^2\chi d\Omega) \ ,
\end{equation}
where the spatial part represents a $3$-sphere, with $\chi$
being the azimuthal angle and $d\Omega$ the line element on the polar
$2$-spheres. The lightcones are still bounded by the particle
horizon. However, due to the curvature of the $3$-sphere, the light
rays must now travel along the azimuthal direction in order to
maximize the sphere of causal contact. The comoving horizon is the
extent of the azimuthal angle traveled by light between times $0$ and
$t$:
\begin{equation}
\label{areasph2}\chi_H = {L_H\over a} = \int^t_0 {dt'\over a(t')} \ .
\end{equation}
The boundary area of the
causal sphere is then given by
\begin{equation}
\label{areasph} A \sim 4\pi a^2(t) \sin^2 \chi_H \ .
\end{equation}
The volume inside of this sphere is
\begin{equation}
\label{volsph} V = \int^{\chi_H}_0 d\chi \sin^2 \chi \, d\Omega = \pi
(2\chi_H - \sin 2\chi_H ) \ .
\end{equation}
Assuming a constant comoving
entropy density $\sigma$, we find
\begin{equation} \label{ratio}
\frac{S}{A} = \sigma \frac{2\chi_H - \sin 2\chi_H }{4 a^2(t)
\sin^2 \chi_H } \ .
\end{equation}
Here we have explicitly retained the
contribution from the comoving entropy density $\sigma$, which was
ignored in \cite{lenny3}.
Consider for simplicity a cold dark matter dominated universe, with $p \ll
\rho$. In this case $a = a_{\rm max}
\sin^2(\chi_H/2)$. The moment $\chi_H = \pi$ corresponds to
the maximal
expansion, $a = a_{\rm max}$. But at that time the
light cone
emitted from the ``North pole'' of the universe converges at the ``South
pole,'' the area of the horizon (\ref{areasph2}) vanishes, and the
holographic
bound on the ratio $S/A$ becomes violated \cite{lenny3}. Note that
in all other
respects the point $\chi_H = \pi$ is regular, so one cannot argue,
for example,
that the violation of the holographic bound is a result of violent quantum
fluctuations of the light cone.
\subsection{Open, closed and flat universes with $\Lambda < 0$}
Let us return to the discussion of
the flat universe case and look at Eq. (\ref{comhorizon2})
again. The size of the comoving horizon $r_H$ can only
grow. Despite this growth, the holographic bound is satisfied for
$\rho > 0$, $p > -\rho$,
because the value of $a^2$ grows faster than $r_H$ in this
regime. But this bound can be violated if $a^2$ grows more slowly
than $r_H$, and it will definitely be
violated in all cases where a flat space can collapse.
Usually, cosmologists believe that closed universes collapse,
whereas open or
flat universes expand forever. But the situation is not quite
so simple.
If there is
a sufficiently large positive cosmological constant,
then even a
closed universe will
never collapse.
On the other hand, if the cosmological
constant is negative, then, even
if it is extremely small, eventually it
becomes dominant, and the universe collapses, independently of
whether it is
closed, open or flat. In all of these cases the holographic
principle,
as formulated in \cite{lenny3}, will be violated.
For simplicity, we will consider a flat universe ($k = 0$) with a negative
vacuum energy density $-\lambda < 0$, so that $\rho = p/\gamma
-\lambda$.
We will assume that $\lambda \ll 1$ in Planckian units. For
example, in our
universe $\lambda$ cannot be greater than $10^{-122}$. In an expanding
universe $\rho = \frac{\rho_0}{a^{3(\gamma+1)}} - \lambda$, and the
Friedmann
equation
\begin{equation}
3H^2 = \frac{\rho_0}{a^{3(\gamma+1)}} - \lambda
\label{friedcc}\end{equation} can be rewritten
as
\begin{equation} \dot a = \pm {1\over \sqrt 3}
\sqrt{\frac{\rho_0}{a^{3\gamma+1}} -
\lambda a^2} \ .
\label{adotcc}
\end{equation}
Because of the presence of
the cosmological term, in general we cannot write the integrals in a
simple form. However, the exact form of the solutions is not necessary
for our purpose here.
First of all, we see that $\dot a$ vanishes at $\lambda a^{3(\gamma
+ 1) }=
\rho_0$, after which $\dot a$ becomes negative and the universe
collapses. This
happens within a finite time after the beginning of
the expansion. From the
definition
of the particle horizon and (\ref{adotcc}), one can find the
value of $L_H$ at the turning point:
\begin{equation}
L_H(turning) =
\frac{B(\frac{\gamma}{2(\gamma+1)},\frac12)}{3(\gamma+1)\sqrt{\lambda}}
\ ,
\end{equation}
where $B(p,q)$ is the Euler beta function. Putting these
formulas together, we see that at the turning point
\begin{equation}
\frac{S}{A} \sim \sigma \lambda^{\frac{1-\gamma}{2(1+\gamma)}}
\end{equation}
up to
factors of order unity. For $1\ge \gamma > -1$,
the power of
$\lambda$ is
positive and so the ratio $S/A$
is very small at the turning point. Now, we can consider what happens
near the final stages of collapse, where the energy density reaches
the Planckian scales. By symmetry, $L_H \sim 2 \frac{a_0}{a(turning)}
L_H(turning)
\sim \lambda^{-(3\gamma+1)/[6(\gamma+1)]}$ at this time, whereas
$\sigma/a^3 \sim 1$. Hence, Eq.
(\ref{comhorizon2a}) yields
$S/A\sim \lambda^{-(3\gamma+1)/[6(\gamma+1)]}
\gg 1$ when $\gamma>-1/3$.
Therefore, we see that the ratio $S/A$
reaches unity
at some time after the turning point,
and that the holographic bound becomes violated thereafter,
but still well in the
classical phase, when the universe is still very large. Indeed, we
can estimate the density of matter at that time to be $\rho \sim
\lambda^{\frac{\gamma+1}{2}}\ll 1$.
A universe
where the only energy density is in form of
a negative cosmological constant is called the anti de
Sitter
space (AdS).
In string theory, AdS spaces typically emerge after
compactifying string or M
theory on an internal, compact, Einstein
space of positive constant curvature.
Many interesting
applications
of the holographic principle have been elaborated for the
pure AdS space. It is therefore
quite interesting that in the cosmological context an AdS background
containing
matter describes a
collapsing Friedmann universe with a
negative vacuum energy, in which the
cosmological holographic principle is violated.
\subsection{AdS spaces with matter and an alternative
formulation of cosmic
holography}
In order to cure the problems of the original formulation of the
cosmological
holographic principle, Bak and Rey proposed a different formulation
\cite{brtwo}. They suggested to consider the so-called apparent
horizon instead
of the particle horizon and claimed that in this case the
holographic bound
holds even in a closed universe. We will not present here a detailed
discussion
of their proposal. Instead we will
consider here their holographic
bound in
the three-dimensional
spatially flat universe (d = 3), see Eq. (16) of
\cite{brtwo}:
\begin{equation}
{ 4 \sigma \over 3 a^{2}(t)\dot{a}(t) } \quad \le \quad 1 .
\label{flatcondition}
\end{equation}
This condition is violated when the universe approaches the turning
point at
$\lambda a^{3(\gamma + 1) }= \rho_0$, when one has $\dot a = 0$.
This violation
occurs even much earlier than in the original formulation of the
cosmological
holographic principle of Ref. \cite{lenny3}.
One can propose two possible interpretations of these results. First
of all,
one may argue that closed universes are impossible, and
that the universes with a
negative cosmological constant are also impossible. We do not see
how one could
justify such a statement. After all, the main reason why the holographic
constraint was violated in both cases studied above was related to the
possibility of gravitational collapse. It would be very
odd to
expect
that the holographic principle which was motivated by the study of
black holes
should imply that gravitational collapse cannot occur.
Another possibility is that the formulations of the cosmic holography
proposed in \cite{lenny3,brtwo} should be somewhat modified in the
cases when
the universe may experience collapse. It would also be
interesting to
understand the
reasons why the holographic inequalities
were correct in the flat universe
case. We will discuss this issue in the next section.
\section{Black holes as big as a universe}
The simplest way to understand the holographic bound on the entropy of the
observable part of the universe is related to the theory of black
holes. In
what follows we will develop further an
argument given by Easther
and Lowe
\cite{el}, and by Veneziano \cite{gv}.
The simplest cosmological models are based on the assumption that
our universe
is homogeneous. But how do we know that it is indeed homogeneous if
the only
part of the universe that we can see\footnote{If one takes into account
inflation, then particle horizon is exponentially large. Still we
can see
(by means of electromagnetic radiation) only a small part of the
universe of
size $\sim H^{-1} \sim t$. It is important that this scale, rather
than the
particle horizon, determines the largest size of a black hole which can be
formed in an expanding universe.} has size $H^{-1}$? We cannot
exclude the
possibility that if we wait for another 10 billion years, we will
see that we
live near
the center of an expanding but isolated gravitational system of size
$O(H^{-1})$ in an asymptotically flat space. Then we can apply the
Bekenstein
bound to the entropy of this system,
$S \lesssim ER$,
where $E \sim \rho
R^3$ is the
total energy and $R \sim H^{-1} $ is the size of this system, with
$H^2 \sim
\rho$, in Planck units. This gives $S \lesssim H^{-2}$, which
coincides with
the holographic bound.
Of course, the idea that our part of the universe is a small
isolated island of
size $H^{-1}$ is weird, but we do not really advocate this view here.
Rather, we
simply say that since we cannot tell whether the universe is
homogeneous, or it
is an island of a size somewhat greater than $ H^{-1} $, the bound $S
\lesssim
H^{-2}$ must hold for a usual homogeneous universe as well.
One can look at this constraint from a different perspective. It is
well known
that if our universe is locally overdense on a scale of horizon with
${\delta\rho\over \rho} = O(1)$, the overdense part will collapse
and form a
black hole of a size $H^{-1}$ \cite{carrhawk}. Then the entropy of
this part
of the universe will satisfy the black hole bound $S \lesssim
H^{-2}$. Again,
there is no indication that ${\delta\rho\over \rho} = O(1)$ on
the horizon
scale, but since we cannot exclude this possibility on a scale
somewhat greater
than the present value of $H^{-1}$, the bound should apply to the
homogeneous
universe as well.
Instead of debating the homogeneity of our universe, one can imagine
adding a
sufficient amount of cold dark matter to a part of our universe of
size $R$.
This would not change its entropy, but it would lead to black hole
formation.
Then one can find an upper bound on the entropy of a black hole of
size $R$: $S
\lesssim R^2$.
If one takes
$R \sim H^{-1}$, one again finds that $S \lesssim H^{-2}$.
The bound $S
\lesssim R^2$ implies that the density of entropy satisfies the
constraint $s
= {S/ R^3} < 1/R$. Thus one could expect that it is possible to get a more
stringent constraint on the density of entropy by considering black
holes of
size greater than $H^{-1}$.
However, according to Carr and Hawking \cite{carrhawk}, black
holes formed
in a flat universe cannot have size greater than $O(H^{-1})$. This
constraint
has a dynamical origin, and is not related to the size of the
particle horizon.
Usually the difference between $H^{-1}$ and the particle horizon is
not very
large, but during inflation this difference is very significant: $H^{-1}$
remains nearly constant, whereas the particle horizon grows exponentially.
If an inflationary domain is homogeneous
on a scale $O(H^{-1})$, then it is going to expand
exponentially,
independently of any inhomogeneities on a larger scale. Such a
domain is not
going to
collapse and form a black hole until inflation ends and we wait
long enough
to see the
boundaries
of the domain. But this will not happen for an exponentially long
time.
Nevertheless
the holographic constraints on the entropy can be derived for the
processes
after inflation, just as in the case considered above.
These
constraints
will be related to the size of the largest black hole which can be
formed during
the expansion of the post-inflationary universe, $R \sim H^{-1}$,
rather than
to the
exponentially large size of the particle horizon in an inflationary
universe.
As a result, the holographic bounds do not lead to the constraints on the
duration of inflation, inflationary density perturbations, and other
parameters
of inflationary theory discussed in \cite{infl}.
If the universe is non-inflationary and closed, or if it has a
negative
cosmological constant,
then, prior to the point of maximal expansion, the holographic
constraints on
the entropy within the
regions of size $H^{-1} \sim t$ coincide with the constraints for the
flat universe
case. Once the universe begins to collapse, the constraints cannot
be further
improved because the typical time of formation of a black hole of
size $O(t)$
at that stage
will be of the same order of magnitude
as the lifetime of the universe. But this
fact does
not imply the impossibility of collapsing universes.
Note that in our consideration we did not make any assumptions about the
adiabatic evolution of the universe. Thus, the cosmological holographic
constraints on entropy are as general as their black hole counterparts. In
fact, we believe that these two constraints
have the same origin.
\section{Holography vs. Inflation}
As we already mentioned, all holographic constraints discussed
in this
paper apply only to the post-inflationary universe. Inflationary
cosmology in its spirit is somewhat opposite to holography. The
possibility of
solving the horizon,
homogeneity,
isotropy, and flatness problems is related to the superluminal
stretching of
the universe, which
erases all memory about the boundary conditions. The speed of
rolling of the
inflaton scalar field approaches an asymptotic value which does not
depend on
its initial speed. The gradients of the fields and the density of
particles
which existed prior to inflation (if there were any) become exponentially
small. All particles (and all entropy) which exist now in the
universe have
been created after
inflation in the process of reheating. This process occurs locally, so the
properties of particles as well as their entropy do not depend on the
initial conditions in the universe.
In order to investigate this issue in a more detailed way, let us
consider the
simplest version of inflationary cosmology where the universe during
inflation
expands only $10^{30}$ times (the minimal amount which is necessary for
inflation to work). We will also assume for simplicity that
inflation occurs at
the GUT scale, so that $H \sim 10^{-6}$ and the temperature after
reheating is
$T \sim 10^{-3}$ in the Planck units.
In such a case the size of the particle horizon after inflation will
be $L_H
\sim H^{-1} \times 10^{30} \sim 10^{36}$, the area $A \sim L_H^2 \sim
10^{72}$, and
the entropy $S \sim T^3 L_H^3 \sim 10^{99}$, which clearly violates
the bound
$S < A$.
This means that the information stored at the surface of an
inflationary domain cannot describe dynamics
in its interior.
In practice, it is extremely difficult to invent inflationary
theories where
the universe grows only by a factor of $10^{30}$ because typically in such
models ${\delta\rho\over \rho} = O(1)$ at the scale of the horizon. In the
simplest versions of chaotic inflation the universe grows more than
$10^{1000000}$ times during inflation. The situation becomes especially
dramatic in those versions of inflationary cosmology which lead to
the process
of eternal self-reproduction of inflationary domains. In such models the
universe is not an expanding
ball of a huge size, but a growing fractal consisting of many
exponentially
large balls. In the process of eternal self-reproduction of the
universe all
memory about the boundary conditions and initial conditions becomes
completely
erased \cite{book}.
Of course, one can use the version of the holographic principle
describing the post-inflationary evolution of the universe, as
discussed in the
previous sections. However, in realistic inflationary models the energy
density at the end of inflation falls more than 15
orders of magnitude below the Planck density, and the most
interesting part of dynamics of the universe where quantum gravity
could play
a significant role is already over.
There is another interesting aspect of relations between inflation and
holography.
The holographic bound on the present
entropy of the universe is $S \lesssim H^{-2}$. One has $H^{-1}
\sim 10^{60}$
in the Planck units. This gives the constraint
\begin{equation}\label{bound}
S \lesssim H^{-2} \sim 10^{120} \ .
\end{equation}
Meanwhile, the entropy of matter in the observable part of the
universe is
smaller than $10^{90}$.
If one thinks about cosmology in terms of the
information which can be stored on the horizon (or, to be more
accurate, on a
surface of a sphere of size $H^{-1}$), one can be encouraged by the
fact that
the holographic bound is satisfied with a wide safety margin, $S/A
\lesssim
10^{-30}$. On
the other hand, if, as we have argued, the
information stored on
the sphere of size $H^{-1}$ is not related to the initial conditions
at the
beginning of
inflation, then its importance is somewhat limited. In such a case
the only
information about the universe that we gained is the bound $S \lesssim
10^{120}$, which is $30$ orders of magnitude less precise than the
observational constraint on the entropy. But what is the origin of
these $30$
orders of magnitude?
Let us look back in time and assume that there was no inflation and the
evolution of the universe was adiabatic. Our part of the universe today
has size $\sim 10^{28}$ cm.
At the Planck time its size $l$ would be $10^{28}$ cm
multiplied by $ {T_p\over T_0}$, where $T_0$ is the present value of the
temperature of the universe, and $T_p \sim 1$ is the Planck
temperature. (Note
that the scale of the universe is inversely proportional to $T$ during
adiabatic expansion.) One therefore
finds $l \sim 10^{-3}$ cm, which is
$10^{30}$ times
greater than the Planck length. That is exactly the reason why we need the
universe to inflate by the factor of $10^{30}$. (The true number
depends on the
value of reheating temperature after inflation.)
If the universe did not inflate at all, it would be very
holographic. A typical
homogeneous part of the universe soon after the big bang would have Planck
size, it would contain just one or two particles, and the constraint
$S < A$
would be saturated. But we would be unable to live there.
Let us assume, for the sake of the
argument, that inflation starts and
ends at the
Planck density, and it has Planckian temperature after reheating. If the
universe during this period inflated by more than $10^{30}$ times,
then our
part of the universe after inflation would have the size $10^{-3}$
cm, i.e.
$10^{30}$ in Planck units, just as we estimated above. Its entropy
would be
$10^{90}$. Then the universe expands by $ {T_p\over T_0} \sim 10^{30}$
times, and the area of our domain becomes $10^{120}$. This makes it
clear that
the factor of $10^{30}$ which characterizes the discrepancy between the
holographically natural value of entropy $10^{120}$ and the observed value
$10^{90}$ is the same factor which appears in the formulations of
the entropy
problem and flatness problem \cite{gv}.
Thus, in the final analysis, the reason why one has $S \lesssim
10^{-30} A$ in
our universe is related to inflation.
Without inflation one
would have $S
\sim A$, and a typical locally
homogeneous patch of the universe would collapse within the Planck
time.
The safety
margin of 30 orders of magnitude
created by inflation makes the universe very large and long-living, but
simultaneously
prevents the holographic constraint on entropy from
being very informative.
A nontrivial relation between the holographic constraint and
inflation does not
mean that one can identify the entropy problem (existence of a huge
entropy $S \sim 10^{90}$ in our part of the universe) and the holography
problem (discrepancy between the holography bound $10^{120}$ and the
true value
of entropy $10^{90}$). For example, in one of the recent versions of
the pre-big
bang scenario the stage of the pre-big bang inflation begins from a
state which
can be identified with a black hole with a large area of the black hole
horizon \cite{damour}.
In this case, the initial entropy of the gravitational
configuration by definition satisfies the Bekenstein-Hawking bound, which
coincides
with the holographic bound.
If one assumes that the entropy of
matter inside
the black hole {\it saturates} the Bekenstein-Hawking bound (this is
just an
assumption which does not follow from the black hole theory),
then the holography problem will be resolved \cite{gv}.
However, one should still
determine the
origin of the enormously large black
hole entropy in this scenario, which
constitutes the entropy problem \cite{klb}.
\section{Conclusions}
The idea that all information about physical processes in the world can be
stored
on its surface is very powerful. It has many interesting implications in
investigation of the nonperturbative properties of M-theory.
However, it is
rather difficult to merge this idea with cosmology. The universe may
not have
any boundary at all, or it may expand so fast that boundary effects
become
irrelevant for the description of the local dynamics.
In this paper we have shown that some of the formulations of the
holographic
principle should be modified not only in application to a closed
universe, but
also for open, closed and flat universes with a negative
cosmological constant.
We believe that the cosmological holographic constraints on entropy,
in those
cases where they are valid, can be understood using the Bekenstein-Hawking
bound on the entropy of black holes. These constraints are rather
nontrivial,
but if applied
to our part of the universe they are much weaker
than the
observational constraints, as well as the constraints which follow
from the
theory of creation of matter after inflation. We believe that these
constraints
do not permit one
to rule out the universes which may experience
gravitational
collapse, and they do not impose any additional constraints on
inflationary
cosmology.
The constraints on entropy represent only one aspect of the holographic
principle.
A stronger form which has been advocated
requires the existence of a theory
living on the boundary surface which would
describe physical processes in the enclosed volume.
Validity of this conjecture in the cosmological context has
not been
demonstrated, and in fact one may argue that there exists a general
obstacle on
the way towards the realization of this idea. In the theory of black
holes, the role
of the holographic surface is played by the black hole horizon. Its
area, and
correspondingly the number of degrees of freedom living on the
horizon, remains constant if one neglects quantum gravity effects.
Thus it is not
unreasonable
to assume that there exists a unitary quantum theory associated with
the black hole horizon.
However,
in an expanding universe the number of degrees of freedom
associated with the cosmological horizon, or with apparent horizon,
or with a
horizon of a
would-be black
hole which provides holographic constraints on entropy, rapidly
changes in
time. For example, in a closed universe the initial area
of the horizon is
vanishingly small, then it grows until it reaches the
maximum, and subsequently it disappears. Thus the
number of degrees of freedom associated
with such a surface strongly depends on time even when the
evolution of the
universe is adiabatic and the total number of degrees of freedom in
the bulk is
conserved \cite{RB}.
Therefore one may wonder whether the holographic theory existing on such a
surface will violate unitarity. In addition, the disappearance of
degrees of
freedom after the moment of the maximal expansion implies that the
entropy
measured at the holographic surface will increase during the universe
expansion, but then it will decrease during its contraction, and
eventually it
will vanish. This means that the second law of
thermodynamics may be violated in the holographic theory.
The situation with causality in such a theory is not clear as well.
Indeed,
information about the new degrees of freedom which are going to appear or
disappear on the
holographic surface is stored not on this surface but in the bulk. This
information does not
propagate
along the surface, rather it crosses the surface when new particles
enter the
apparent horizon. But this suggests that the creation of the new
degrees of
freedom in the holographic theory will not look like an effect
caused by the
earlier existing conditions at the surface.
It remains to be seen whether one can overcome all of these
problems and
make the
holographic principle a useful part of the modern cosmological
theory including
inflationary theory.
We should note, however, that quantum cosmology is extremely
complicated and
counterintuitive in many respects. It is still a challenging task to unify
M-theory and inflationary cosmology. Any progress in this direction
would be
very important. One may expect that the ideas borne out by
the investigation of
quantum dynamics of black holes and enriched by the study of
supergravity and
string theory will play the key role in the development of a
nonperturbative
approach to quantum cosmology.
\
We wish to thank R. Bousso, W. Fischler and L. Susskind for valuable
discussions. This work has been supported in part by NSF grant
PHY-9870115.
|
\section{INTRODUCTION}
\end{center}
Due to vanishing of Weyl tensor in $(2+1)$ dimensions, the Riemann curvature
tensor can be identified with zero in matter-void regions
of spacetime. Consequently, spacetime is flat in local vacua.
Addition of cosmological constant term in the absence of matter leads to
solutions with constant curvature, where the sign of cosmological constant
determines the sign of scalar curvature \cite{Jackiw}.
In the last few years many interesting problems have been investigated in $(2+1)$
dimensional gravity, such as the spacetime metric of multi-point particle with
or without spin \cite{Deser,Clement}. The Einstein-Maxwell equations in $(2+1)$ dimensions
have already been treated \cite{Kogan,Reznik}.
Indeed, in many physical situations
in $(3+1)$ dimensions there is no structure along one of spatial dimension like
an infinite cosmic string, where the theory becomes $(2+1)$ dimensional.
There are also some interesting works concerning the quantum mechanics of
a point mass particle in the presence of a very heavy particle in $(2+1)$
dimensions, both relativistically and nonrelativistically \cite{Dese,Sousa,Krzysztof}.
We introduce the Einstein-Maxwell action in the presence of
matter together with cosmological term in $(2+1)$ spacetime dimension.
Then, we choose the solutions that correspond to a spacetime with a spatial part of
locally constant curvature surface with deficit of angle at location of a very heavy
point mass, and magnetic field of a magnetic monopole.
For positive cosmological
term we have Minkowskian and Euclidean spacetimes with spatial part of
locally constant curvature. For vanishing cosmological term
we have a spacetime with locally flat spatial part.
Over these spacetimes, in absence of angular deficit, the quantum Hamiltonian
associated with a free test particle leads to solvable systems with
degeneracy group $GL(2,c)$, where their eigen-states can be obtained
algebraically, too.
We will show that quantum solvable models will be obtained
by restricting the Casimir of $SO(4,c)$ group to the Casimir of $SL(2,c)$ group.
These models possess simultaneously the degeneracy group $SL(2,c)$ and shape
invariance symmetry, where both symmetries are kind of realization of
para-supersymmetry of arbitrary order.
\begin{center}
\section{SOME SPECIAL EXACT SOLUTIONS OF EINSTEIN EQUATIONS}
\end{center}
In $(2+1)$ dimensions the Einstein-Hilbert action of gravity coupled to matter
and electromagnetic field, together with the cosmological term can be written as
\begin{equation}
S=\int d^3x \sqrt{-g} \{ \frac{1}{4 \pi G}(R+2\Lambda) +
\frac{1}{4} F_{\mu\nu}F^{\mu\nu}-{\cal L}_M \},
\end{equation}
where $\cal L_{M}$ is the matter Lagrangian. We have rescaled G by a
factor of $4$.
Variation of the action (2.1) leads to Einstein-Maxwell equations in
$(2+1)$ dimensions
\renewcommand{\theequation}{\thesection.\arabic{equation}{a}}
\begin{equation}
R_{\mu\nu}-\frac{1}{2}g_{\mu\nu}R=2\pi GT_{\mu\nu}^{eff}
\end{equation}
\vspace{-7mm}
\renewcommand{\theequation}{\thesection.\arabic{equation}{b}}
\setcounter{equation}{1}
\begin{equation}
\partial_\mu(\sqrt{-g}F^{\mu\nu})=0
\end{equation}
with
$T_{\mu\nu}^{eff}=T_{\mu\nu}^{(M)}+T_{\mu\nu}^{(EM)}+\frac{\Lambda}{2\pi G} g_{\mu\nu}$.
The energy-momentum tensor of matter $T_{\mu\nu}^{(M)}$ and electromagnetic
$T_{\mu \nu}^{(EM)}$ are respectively
\renewcommand{\theequation}{\thesection.\arabic{equation}{a}}
\begin{equation}
T_{\mu\nu}^{(M)}=\frac{2}{\sqrt{-g}}\frac{\delta (\sqrt{-g}{\cal L}_{M})}{\delta g^{\mu\nu}},
\end{equation}
\vspace{-7mm}
\renewcommand{\theequation}{\thesection.\arabic{equation}{b}}
\setcounter{equation}{2}
\begin{equation}
T_{\mu\nu}^{(EM)}=-g^{\alpha \beta}F_{\mu \alpha} F_{\nu \beta}+\frac{1}{4}
g_{\mu\nu}F_{\lambda \sigma} F^{\lambda \sigma}.
\end{equation}
We assume that our spacetime is described by the axial symmetric static metric, that is \hspace{10mm}
$(\partial_{t}g_{\mu\nu}=0, g_{_{0i}}=0)$ \cite{Jack}:
\renewcommand{\theequation}{\thesection.\arabic{equation}}
\begin{equation}
ds^2_{(3)}=N^2({\bf r})dt^2-\rho({\bf r})(dr^2+r^2d\phi^2),
\end{equation}
where $0\leq r<\infty$ and $0\leq\phi<2\pi$ are the usual polar coordinates.
The non-zero components of electromagnetic field tensor are
$$
F_{0r}=E(r),\hspace{6mm} F_{r\phi}=B(r) \hspace{6mm} and \hspace{6mm} F_{0\phi}=0
$$
with E and B as electric and magnetic fields, respectively.
After, changing the coordinates as
$x^1=r \cos \phi $ and $x^2=r \sin \phi $,
the electromagnetic energy-momentum tensor (2.3b)
takes the following form
\begin{equation}
T_{\alpha \beta }^{(EM)}=\left(\begin{array}{ccc}
\frac{1}{2 \rho}(\frac{N^2}{\rho r^2}B^2+E^2) & \frac{x^2}
{\rho r^2}EB & \frac{-x^1}{\rho r^2}EB \\
\frac{x^2}{\rho r^2}EB & \frac{1}{2\rho r^2}B^2+\frac{(x^2)^2-(x^1)
^2}{2 r^2 N^2} E^2 & \frac{-x^1 x^2}{N^2r^2}E^2 \\
\frac{-x^1}{\rho r^2}EB & \frac{-x^1x^2}{N^2r^2}E^2 & \frac{1}{2\rho r^2}B^2-
\frac{(x^2)^2-(x^1)^2}{2r^2N^2}E^2 \end{array}\right).
\end{equation}
In this article we consider the matter as a point mass located at the
origin $(x^1=x^2=0)$ with the only nonvanishing component of $T_{\mu \nu}^{(M)}$ as
$T_0^{(M) 0}=\frac{1}{\sqrt{^{(2)}g}} M \delta (x^1) \delta (x^2)$,
where $^{(2)}g$ is the determinant of the spatial part of the metric.
We remind that the Ricci scalar for metric (2.4) is
\begin{equation}
R=\frac{2}{\rho N}\nabla^{2}N+\frac{1}{\rho^{2}}\nabla^{2} \rho -
\frac{1}{\rho^{3}}(\vec{\nabla} \rho)^{2},
\end{equation}
where $\nabla^{2}$ and $\vec{\nabla}$ are the usual Euclidean 2-dimensional
Laplacian and gradient operators, respectively.
For a spatially conformal metric with
$\rho({\bf r})=\rho_{_{_{0}}}e^{-2GM \ln r+\chi}$,
the singular term on the right hand side of Eqs. (2.2a) disappears.
Thus, the equations can be written in the following singularity free form
\begin{eqnarray}
&& EB=0 \nonumber \\
\nonumber \\
&& \nabla^{2} \chi +2 \Lambda \rho_{_{_{0}}}r^{-2GM} e^{\chi}+\frac{2\pi G}{\rho_{_{_{0}}}}r^{2GM-2}
e^{-\chi}B^{2}+\frac{2\pi G}{N^{2}}E^{2}=0 \nonumber \\
\nonumber \\
&& \partial_{1}\partial_{2}N-\frac{1}{2}(\partial_{1}N\partial_{2}\chi
+\partial_{1}\chi \partial_{2}N)+\frac{GM}{r^{2}}(x^{2}\partial_{1}N
+x^{1}\partial_{2}N)-2\pi G\frac{x^{1}x^{2}}{Nr^{2}}E^{2}=0 \nonumber \\
\nonumber \\
&& \partial_{2}^{2}N+\Lambda \rho_{_{_{0}}}Nr^{-2GM} e^{\chi}+\frac{1}{2}(
\partial_{1}\chi \partial_{1}N-\partial_{2} \chi \partial_{2}N)-GM
r^{-2}(x^{1}\partial_{1}N-x^{2}\partial_{2}N) \nonumber \\
&&\hspace{45mm} -\frac{\pi G}{\rho_{_{_{0}}}}Nr^{2GM-2}e^{-\chi}B^{2}-\pi Gr^{-2}\frac{(x^{2})^{2}-(x^{1})^{2}}{N}E^{2}=0 \nonumber \\
\nonumber \\
&&\partial_{1}^{2}N+\Lambda \rho_{_{_{0}}}Nr^{-2GM}e^{\chi}+\frac{1}{2}(
\partial_{2} \chi \partial_{2}N-\partial_{1} \chi \partial_{1}N)-GM
r^{-2}(x^{2}\partial_{2}N-x^{1}\partial_{1}N) \nonumber \\
&&\hspace{45mm}-\frac{\pi G}{\rho_{_{_{0}}}}Nr^{2GM-2}e^{-\chi}B^{2}-\pi Gr^{-2}\frac{(x^{1})^{2}-(x^{2})^{2}}{N}E^{2}=0.
\end{eqnarray}
Letting $E=0$ in Eqs. (2.7) and using the following
ansatz for the magnetic field $B$
$$
B^{2}=h \frac{\Lambda \rho_{_{_{0}}}^2}{\pi G} r^{2-4GM}e^{2\chi}
$$
where $h$ is a constant parameter. Then, with a change of variable
$u=r^{1-GM}$, the Eqs. (2.7) can be written as
\renewcommand{\theequation}{\thesection.\arabic{equation}{a}}
\begin{equation}
\frac{1}{u}\frac{d}{du}(u\frac{d}{du})\chi +
\frac{2(1+h)\Lambda \rho_{_{_{0}}}}{(1-GM)^{2}}e^{\chi}=0
\end{equation}
\vspace{-3mm}
\renewcommand{\theequation}{\thesection.\arabic{equation}{b}}
\setcounter{equation}{7}
\begin{equation}
\frac{1}{u}\frac{dN}{du}=\frac{c}{1-GM}e^{\chi}
\end{equation}
\vspace{-3mm}
\setcounter{equation}{7}
\renewcommand{\theequation}{\thesection.\arabic{equation}{c}}
\begin{equation}
\frac{1}{u}\frac{d}{du}(u\frac{d}{du})N+
\frac{2(1-h)\Lambda \rho_{_{_{0}}}}{(1-GM)^{2}}Ne^{\chi}=0,
\end{equation}
where $c$ is a constant of integration. Choosing the following ansatz as a solution for
the Eq. (2.8a)
\renewcommand{\theequation}{\thesection.\arabic{equation}}
\begin{equation}
e^{\chi}=\frac{1}{(1+\frac{(1+h)\Lambda \rho_{_{_{0}}}}{4(1-GM)^{2}}u^{2})^2} ,
\end{equation}
we make use of (2.9) to solve the Eq. (2.8b) as
$$
N=c\frac{\frac{-2(1-GM)}{(1+h)\Lambda \rho_{_{_{0}}}}}{1+\frac{(1+h)\Lambda \rho_{_{_{0}}}}{4(1-GM)^{2}}u^{2}}+d,
$$
where $d$ is another constant of integration.
Next, having inserted the results just obtained for $\chi$ and $N$ in
Eq. (2.8c), we get the following equations for constants $d$ and $c$
\begin{eqnarray}
&&\frac{(1-h^{2})\Lambda \rho_{_{_{0}}}}{1-GM}d-(1+h)c=0 \nonumber \\
&&\frac{(1-h)\Lambda \rho_{_{_{0}}}}{1-GM}d-\frac{1-3h}{1+h}c=0.
\end{eqnarray}
The nontrivial solutions of Eqs. (2.10) exist only for $h=0$ and $1$.
If $h=0$, the magnetic field vanishes and Maxwell Eqs. (2.2b)
are satisfied. Then, (2.4)
becomes the metric of spacetime in the presence of a point mass located at the origin,
together with the cosmological term which has already been studied
in Refs. \cite{Jackiw,Reznik}.
But, for $h=1$ we have $N=1$ and
\begin{equation}
B^{2}=\frac{\Lambda \rho_{_{_{0}}}^2}{\pi G} \frac{r^{2-4GM}}{(1+\frac{\Lambda \rho_{_{_{0}}}}{
2(1-GM)^{2}}r^{2-2GM})^{4}},
\end{equation}
together with the following metric of spacetime
\begin{equation}
ds^{2}_{(3)}=dt^{2}-\frac{\rho_{_{_{0}}}r^{-2GM}}{(1+\frac{\Lambda \rho_{_{_{0}}}}{2(1-GM)^2}
r^{2-2GM})^2}(dr^{2}+r^2d\phi^{2}).
\end{equation}
It turns out that the magnetic field (2.11) and the metric (2.12) satisfy
Einstein-Maxwell Eqs. (2.2).
Using the formula (2.6) we calculate the scalar curvature of metric (2.12)
$$
R=-4\Lambda-4\pi GMr^{2GM}(1+\frac{\Lambda \rho_{_{_{0}}}}{2(1-GM)^2}r^{2-2GM})^2\delta^2(\bf r).
$$
Hence, except for a delta singularity at the origin, the spacetime has a
constant curvature. We will discuss the
solutions of Einstein equations in $(2+1)$-dimensional spacetime
corresponding to a point mass located at origin in the presence of cosmological
constant and magnetic field in the next section.
In the zero limit of cosmological constant, the magnetic field
(2.11) vanishes and the metric (2.12) changes to
the metric of a point mass located at origin
with angular deficit \cite{Deser,Jac}.
Note that, we have considered here only the most simple
solution of Liouville equation (2.8a), while one can take some less trivial solutions
using the Backlund transformations.
\begin{center}
\section{EMBEDDING OF DEGENERACY AND SHAPE INVARIANCE OF QUANTUM STATES
IN $SO(4,c)$ GROUP}
\end{center}
\setcounter{equation}{0}
We investigate the solutions (2.12) of Einstein equation
in $(2+1)$ dimensions with nonnegative cosmological constant $\Lambda$, point mass $M$
and magnetic field (2.11) as sources of energy-momentum tensor.
Now, we introduce
the parameter $\gamma$ which only takes the values $0$, $1$ and
$i=\sqrt{-1}$ and redefine the cosmological constant $\Lambda$ as
$\Lambda=\gamma^2 \lambda$.
In this article, $\lambda$ and $\rho_{_{_{0}}}$ are arbitrary positive (negative) nonvanishing constants
for $\gamma=0$ and $1$ ($i$). With change of variables
\begin{eqnarray}
&&\frac{r^{1-GM}}{1-GM}=\sqrt{\frac{2}{\lambda \rho_{_{_{0}}}}} \frac{\tan \frac{\gamma \theta}{2}}{\gamma} \nonumber \\
&&\Phi:=(1-GM) \phi,
\end{eqnarray}
the metric (2.12) takes the form
\begin{equation}
ds^{2}_{(3)}=dt^{2}-\frac{1}{2 \lambda}(d \theta^{2}+\frac{\sin^{2}\gamma \theta}
{\gamma^{2}}d\Phi^{2}).
\end{equation}
It is obvious that for $\gamma=i$ the spacetime is described
by an Euclidean metric, while for $\gamma=0$ and $1$ the spacetime is described
by Minkowskian metrics.
Using the general coordinate transformation, the magnetic field can be written
as (3.3) in terms of the new coordinates $\theta$ and $\Phi$
\begin{equation}
{\cal B}=q\gamma \sin \gamma \theta,
\end{equation}
where
\begin{eqnarray}
\nonumber
q=\left \{\begin{array}{ll}
\frac{1}{2\sqrt{\pi G\lambda}} & \mbox{if $\gamma=0$ , $1$} \\
\frac{-1}{2\sqrt{\pi G|\lambda|}} & \mbox{if $\gamma=i.$}\end{array} \right.
\end{eqnarray}
The magnetic potential one-form $A$ corresponding to magnetic field (3.3) is
$A=q(1-\cos \gamma \theta)(\frac{i\gamma}{\sin \gamma \theta}d\theta +d\Phi)$.
Obviously, in terms of the new coordinates $\theta$ and $\Phi$, the choice of $\gamma
=0$ leads to a Minkowskian metric with flat spatial part and
angular deficit in its metric. Here, $\theta$ is its
radial coordinate and the magnetic field $\cal B$ is zero.
The choice of $\gamma=1$ leads to Minkowskian metric
of $(2+1)$-dimensional spacetime with local constant curvature
together with angular deficit and magnetic field $\frac{1}{2\sqrt{\pi G\lambda}} \sin \theta$.
Finally, for $\gamma=i$, we get Euclidean metric of
$(2+1)$-dimensional spacetime with local constant curvature and
deficit of angle and magnetic
field $\frac{1}{2\sqrt{\pi G|\lambda|}}\sinh \theta$.
In the presence of the magnetic field (3.3), the quantum states of
a point mass with mass ${\cal M}$, which is
negligible compared to the mass $M$ of a point source located at the origin,
i.e. ${\cal M} \ll M$,
can be described in terms of the energy spectrum of the
Hamiltonian
$H=\frac{-1}{2{\cal M}}D_{i}^{A}D^{A i}$
with $D_{i}^{A}=\nabla_{i}-iA_{i}$ as covariant derivative.
Using the metric (3.2) and the given connection one-form,
the Hamiltonian can be written as
\begin{equation}
H= \frac{-\lambda}{{\cal M}}\{\frac{\partial^{2}}{\partial \theta^{2}} +
\gamma (\frac{1-2q}{\tan \gamma \theta}+\frac{2q}{\sin \gamma \theta})
\frac{\partial}{\partial \theta}+\frac{\gamma^{2}}{\sin^{2} \gamma \theta}
\frac{\partial^{2}}{\partial \Phi^{2}} +2iq \gamma^{2} (\frac{1}{\tan^{2}
\gamma \theta \cos \gamma \theta}-\frac{1}{\sin^{2} \gamma \theta}) \frac{
\partial}{\partial \Phi}+q \gamma^{2} \},
\end{equation}
with the spectrum given by
\begin{equation}
E(n)=\frac{\lambda \gamma^{2}}{{\cal M}}[n(n+1)-q(2n+1)],
\end{equation}
where in absence of angular deficit $n$ is a non-negative integer \cite{Jafar}.
Writting the eigenstates of the Hamitonian (3.4) in the following form
$$
\Psi_{n,m}(\theta, \Phi)=(e^{-i\Phi}\frac{\tan \frac{\gamma \theta}{2}}{\sin \gamma \theta})^{-q}
e^{i\frac{m}{1-GM}\Phi}(1-\cos \gamma \theta)^{\frac{1}{2}|\frac{m}{1-GM}+q|}
(1+\cos \gamma \theta)^{\frac{1}{2}|\frac{m}{1-GM}-q|}F(\cos \gamma \theta)
$$
and choosing the change of variable $z=\frac{1-\cos \gamma \theta}{2}$,
the eigenvalue equation turns in to hypergeometric differential equation
\begin{eqnarray}
z(1-z)\frac{d^2}{dz^2}F+[1+|\frac{m}{1-GM}+q|-(|\frac{m}{1-GM}-q|+
|\frac{m}{1-GM}+q|+2)z]\frac{d}{dz}F \nonumber \\
+[n-q-\frac{1}{2}(|\frac{m}{1-GM}-q|+|\frac{m}{1-GM}+q|)]
[n-q+\frac{1}{2}(|\frac{m}{1-GM}-q|+|\frac{m}{1-GM}+q|)+1]F=0. \nonumber
\end{eqnarray}
The wave equations for relativistic and non-relativistic quantum states are
\renewcommand{\theequation}{\thesection.\arabic{equation}{a}}
\begin{equation}
H\Psi_{nonrel}(\theta,\Phi)=E(n)\Psi_{nonrel}(\theta,\Phi)
\end{equation}
\vspace{-3mm}
\renewcommand{\theequation}{\thesection.\arabic{equation}{b}}
\setcounter{equation}{5}
\begin{equation}
H\Psi_{rel}(\theta,\Phi)=\frac{E_{rel}^{2}-{\cal M}^{2}}{2{\cal M}}
\Psi_{rel}(\theta,\Phi).
\end{equation}
For $\gamma=0$ we define $n \gamma=k$, which for $n$ very large $k$ is an
arbitrarily finite constant, and we get $E(n) \rightarrow \frac{k^{2}}{2{\cal M}}$.
Then, with assumption $E_{rel}>{\cal M}$,
solutions of Eqs. (3.6) are
\renewcommand{\theequation}{\thesection.\arabic{equation}{a}}
\begin{equation}
\Psi_{_{nonrel}}(\theta,\Phi)=e^{i\frac{m}{1-GM}\Phi}J_{\frac{|m|}{1-GM}}(k\theta)
\end{equation}
\vspace{-3mm}
\renewcommand{\theequation}{\thesection.\arabic{equation}{b}}
\setcounter{equation}{6}
\begin{equation}
\Psi_{_{rel}}(\theta,\Phi)=e^{i\frac{m}{1-GM}\Phi}J_{\frac{|m|}{1-GM}}(\frac{1}{2\lambda}\sqrt{E_{rel}^{2}-{\cal M}^{2}} \theta).
\end{equation}
It is straitforward to see that by defining $\gamma=i\delta$ and taking the zero
limit of $\delta$ as $n\delta=k$ together with $E_{rel}<{\cal M}$, we get the
solutions of Eqs. (3.6) in terms of
modified Bessel functions
\renewcommand{\theequation}{\thesection.\arabic{equation}{a}}
\begin{equation}
\Psi_{_{nonrel}}(\theta,\Phi)=e^{i\frac{m}{1-GM}\Phi}K_{\frac{|m|}{1-GM}}(k\theta)
\end{equation}
\vspace{-3mm}
\renewcommand{\theequation}{\thesection.\arabic{equation}{b}}
\setcounter{equation}{7}
\begin{equation}
\Psi_{_{rel}}(\theta,\Phi)=e^{i\frac{m}{1-GM}\Phi}K_{\frac{|m|}{1-GM}}(\frac{1}{2\lambda}\sqrt{{\cal M}^{2}-E_{rel}^{2}} \theta).
\end{equation}
Therefore, in the presence of a heavy particle $M$ which is located at origin
we get scattering and bound states of a point particle with mass ${\cal M}$,
similar to Eqs. (3.7) and (3.8).
These results are in agreement with reference \cite{Krzysztof}.
So far the domain of $\Phi$ was $[0,2 \pi (1-GM)]$, but, from now on, for simplicity, we ignore
the presence of this angular deficit in the rest of the article.
In general, in order to obtain the eigen-spectrum algebraically, we introduce
generators of $gl(2,c)$ Lie algebra as
\renewcommand{\theequation}{\thesection.\arabic{equation}}
\begin{eqnarray}
\nonumber
&&[L_{+}, L_{-}]=2\gamma^{2}L_{3}- 2q\gamma^{2}I \\
\nonumber
&&[L_{3}, L_{\pm}]=\pm L_{\pm} \\
&&[{\bf L}, I]=0.
\end{eqnarray}
Note that the algebra (3.9) becomes $iso(2) \oplus u(1)$ algebra for $\gamma=0$,
$u(2)$ Lie algebra for $\gamma=1$, and $u(1,1)$
Lie algebra for $\gamma=i$.
The raising $L_{_{+}}$ and lowering $L_{_{-}}$ operators
have the following coordinate representations \cite{Jafar}
\begin{eqnarray}
L_{+}&=&e^{i\phi}(\frac{\partial}{\partial \theta}+\frac{i \gamma}{
\tan \gamma \theta}\frac{\partial}{\partial \phi})
\hspace{54mm} L_{3}=\frac{1}{i}\frac{\partial}{\partial \phi} \nonumber \\
L_{-}&=&e^{-i\phi}(-\frac{\partial}{\partial \theta}+\frac{i\gamma}{
\tan \gamma \theta}\frac{\partial}{\partial \phi}+2q\frac{\gamma}{\tan
\gamma \theta}-2q\frac{\gamma}{\sin \gamma \theta})
\hspace{10mm} I=1.
\end{eqnarray}
It is easy to show that the Hamiltonian (3.4) without angular deficit is
the Casimir operator of $gl(2,c)$ Lie algebra, that is
\begin{eqnarray}
H=\frac{2\lambda}{M}[\frac{1}{4}L_{_{+}}L_{_{-}}+\frac{1}{4}
L_{_{-}}L_{_{+}}+\frac{1}{2}\gamma^{2}L_{_{3}}^{2}-q\gamma^{2}L_{_{3}}].
\end{eqnarray}
For $\gamma \neq 0$, its highest weight can be obtained as follows
$$
\Psi_{n,n}(\theta,\phi)=e^{in\phi}(\frac{1}{\gamma}
\sin \frac{\gamma \theta}{2})^{n}(\cos \frac{\gamma \theta}{2})^{n},
$$
where $n$ is a nonnegative integer of integration constant. The other states can be obtained by applying the lowering operator
$L_{-}$ over $\Psi_{n,n}(\theta,\phi)$ repeatedly. That is,
\begin{eqnarray}
\Psi_{_{nonrel}}(\theta,\phi)&=&\Psi_{n,m}(\theta,\phi)=(L_{-})^{n-m}\Psi_{n,n}(\theta,\phi) \nonumber \\
&=&(2q-2n)_{n-m}e^{im \phi}(\frac{\sin \frac{\gamma \theta}{2}}{\gamma})
^{m} (\cos \frac{\gamma \theta}{2})^{2n-m} F(m-n, 2q-n, 2q-2n,
\frac{1}{\cos^{2}\frac{\gamma \theta}{2}}). \nonumber
\end{eqnarray}
Hence, using the usual
properties of hypergeometric function \cite{Vilenkin},
up to a constant coefficient $(-1)^{n-2q}(2q-2n)_{_{n-m}}$,
in the limit $n \rightarrow \infty$ and $\gamma \rightarrow 0$
such that $n\gamma=k=$finite, the solution of algebraic method can
be written as
\begin{eqnarray}
\nonumber
\Psi_{_{m}}(\theta,\phi)&=&\lim_{n \rightarrow \infty}k^{-m}e^{im\phi}
\cos^{4q}(\frac{k\theta}{2n})n^{m}\tan^{m}(\frac{k\theta}{2n})
F(2q-n,n+1,m+1,\sin^{2}(\frac{k\theta}{2n})) \\
&=& k^{-m}e^{im\phi}\Gamma(m+1)J_{m}(k\theta), \nonumber
\end{eqnarray}
which is the same as special case $M=0$ in Eq. (3.7a). It is obvious that with
the definition $\gamma=i \delta$, we will again get the solution (3.8a)
without angular deficit in the zero limit of $\delta$.
For the left and right
invariant generators of $SL(2,c)$ group manifold with $sl(2,c)$ Lie algebra
\begin{eqnarray}
\nonumber
&&J_{_{\pm}}^{(L)}=e^{\pm i\phi}
(\pm \frac{\partial}{\partial \theta}+i\frac{\gamma}{\tan \gamma \theta}
\frac{\partial}{\partial \phi}-i\frac{\gamma}{\sin \gamma \theta}
\frac{\partial}{\partial \psi}) \\
&&J_{_{3}}^{(L)}=-i\frac{\partial}{\partial \phi},
\end{eqnarray}
\begin{eqnarray}
\nonumber
&&J_{_{\pm}}^{(R)}=e^{\pm i\psi}
(\pm \frac{\partial}{\partial \theta}-i\frac{\gamma}{\sin \gamma \theta}
\frac{\partial}{\partial \phi}+i\frac{\gamma}{\tan \gamma \theta}
\frac{\partial}{\partial \psi}) \\
&&J_{_{3}}^{(R)}=-i\frac{\partial}{\partial \psi},
\end{eqnarray}
where $\theta$, $\phi$ and $\psi$ are complex variables at present,
the Casimir operators are equal with each other, i.e.
$H_{_{sl(2,c)}}=H^{(L)}=H^{(R)}$, such that
\begin{eqnarray}
H^{(L,R)}=
\frac{1}{4}J_{_{+}}^{(L,R)}J_{_{-}}^{(L,R)}+\frac{1}{4}J_{_{-}}^{(L,R)}
J_{_{+}}^{(L,R)}+\frac{1}{2}\gamma^{2}J_{_{3}}^{{(L,R)}^2}.
\end{eqnarray}
Since $so(4,c)=sl(2,c) \oplus sl(2,c)$, therefore the Casimir operator of
$so(4,c)$, can be given in terms of the $sl(2,c)$ Casimir operators as
$H_{_{so(4,c)}}=H^{(L)}+H^{(R)}$,
but note that, choosing the coordinates of left and right generators the same
will be equivalent to restricting the space of $so(4,c)$ Lie algebra
to one of the subspaces $sl(2,c)$.
Now we restrict ourselves to real values of $\theta$, $\phi$ and $\psi$
variables such that $0 \leq \phi < 2\pi$ and $0 \leq \psi < 4\pi$
for all values of $\gamma$, while we choose $0 \leq \theta < \pi$ for $\gamma=1$
and $0 \leq \theta < \infty$ for other cases.
Due to this restriction, the $sl(2,c)$ algebra reduces to its different real
forms as follows:
in the case of $\gamma=1$ the $sl(2,c)$ Lie algebra reduces
to $su(2)$, while for $\gamma=i$ it reduces to $su(1,1)$ Lie algebra and
finally for $\gamma=0$ it reduces to $iso(2)$ Lie algebra \cite{Gilmore}. We should remind that
for $\gamma=1$, $\gamma=i$ and $\gamma=0$ the direct sum of left and right invariant
generators becomes $so(4)=su(2) \oplus su(2)$, $so(2,2)=su(1,1) \oplus su(1,1)$
and $iso(2) \oplus iso(2)$ Lie algebra respectively, which are different real forms of $so(4,c)$.
If we define $l:=n-q$, the eigenvalue equation for
Casimir operator $H_{_{sl(2,c)}}$ will be \cite{Vilenkin}
\begin{equation}
H_{_{sl(2,c)}} \Psi_{_{l,m,q}}(\theta,\phi,\psi)=\frac{1}{2} \gamma^{2}
l(l+1) \Psi_{_{l,m,q}}(\theta,\phi,\psi)
\end{equation}
with
$\Psi_{_{l,m,q}}(\theta,\phi,\psi)=e^{im \phi-iq \psi} P_{_{m,-q}}^{l}(\cos
\gamma \theta).$
After transfer the $\psi$ dependence factor $e^{-iq \psi}$ to the left
hand side of equation (3.15), the reduced Casimir operator with degeneracy group $SL(2,c)$, is
\begin{eqnarray}
H_{_{sl(2,c)}}(q)=\frac{1}{4}J_{_{+}}^{(L)}(q)J_{_{-}}^{(L)}(q)+\frac{1}{4}
J_{_{-}}^{(L)}(q)J_{_{+}}^{(L)}(q)+\frac{1}{2}\gamma^{2}J_{_{3}}^{{(L)}^2}(q),
\end{eqnarray}
where
\begin{eqnarray}
&& J_{_{\pm}}^{(L)}(q)=e^{\pm i\phi}(\pm \frac{\partial}{\partial \theta}+\frac{i
\gamma}{\tan \gamma \theta} \frac{\partial}{\partial \phi}- \frac{q\gamma}{
\sin \gamma \theta}) \nonumber \\
&& J_{_{3}}^{(L)}(q)=-i\frac{\partial}{\partial \phi}. \nonumber
\end{eqnarray}
The above reduced $sl(2,c)$ generators are related to (3.10)
generators of $gl(2,c)$ through a similarity transformation together with
a change of basis,
such that, one can obtain the eigenstates of $gl(2,c)$ Casimir operator.
Thus, the eigenvalue equation (3.15) reduces to
\begin{equation}
H_{_{sl(2,c)}}(q)\Psi_{_{l,m,q}}(\theta,\phi)=\frac{1}{2}\gamma^{2}l(l+1)
\Psi_{_{l,m,q}}(\theta,\phi),
\end{equation}
with $\Psi_{_{l,m,q}}(\theta,\phi)=e^{im\phi}P_{_{m,-q}}^{l}(\cos\gamma\theta)$.
From there we can introduce the eigenfunction
\begin{eqnarray}
\Psi_{n,m}(\theta,\phi)=
(2\frac{\tan \frac{\gamma \theta}{2}}{\sin \gamma \theta}e^{-i\phi})^{-q}
e^{im\phi}P_{_{m,-q}}^{l}(\cos\gamma\theta), \nonumber
\end{eqnarray}
for Hamiltonian (3.11) with the same eigenvalue (3.15).
Then, we see that the Dirac's quantization of magnetic charge
follows very naturally from the $SL(2,c)$ representation,
that is integrity of $q$.
Also, the $so(4,c)$ Lie algebra is the
dynamical group of the quantum dynamics of a point mass particle in the
presence of a constant magnetic
field over two dimensional surface with global constant curvature with $gl(2,c)$
sub-algebra as its degeneracy group.
Transferring the function $e^{-iq\psi}$ to the left of the Casimir operator
$H_{_{sl(2,c)}}$
in the eigenvalue equation (3.15),
the Casimir operator $H_{_{sl(2,c)}}(q)$
can be written as
\renewcommand{\theequation}{\thesection.\arabic{equation}{a}}
\begin{equation}
H_{_{sl(2,c)}}(q)=\frac{1}{2}J_{_{+}}^{(R)}(q+1)J_{_{-}}^{(R)}(q)+\frac{1}{
2}q(q+1)\gamma^{2}
\end{equation}
\vspace{-7mm}
\renewcommand{\theequation}{\thesection.\arabic{equation}{b}}
\setcounter{equation}{17}
\begin{equation}
=\frac{1}{2}J_{_{-}}^{(R)}(q-1)J_{_{+}}^{(R)}(q)+\frac{1}{2}q(q-1)\gamma^{2}
\end{equation}
with
\renewcommand{\theequation}{\thesection.\arabic{equation}}
\begin{eqnarray}
J_{_{\pm}}^{(R)}(q)=\pm \frac{\partial}{\partial \theta}-i\frac{
\gamma}{\sin \gamma \theta} \frac{\partial}{\partial \phi}+\frac{q\gamma}{
\tan \gamma \theta}. \nonumber
\end{eqnarray}
Using the Eqs. (3.18) in the eigenvalue equation (3.17)
we obtain
\begin{eqnarray}
&&J_{_{+}}^{(R)}(q)J_{_{-}}^{(R)}(q-1)\Psi_{_{l,m,q-1}}(\theta,\phi)=E_{_{q}}\Psi_{_{l,m,q-1}}(\theta,\phi) \nonumber \\
&&J_{_{-}}^{(R)}(q-1)J_{_{+}}^{(R)}(q)\Psi_{_{l,m,q}}(\theta,\phi)=E_{_{q}}\Psi_{_{l,m,q}}(\theta,\phi), \nonumber
\end{eqnarray}
where $E_{_{q}}$ is defined as
\begin{eqnarray}
E_{_{q}}=\gamma^2[l(l+1)-q(q-1)]. \nonumber
\end{eqnarray}
The restriction of $so(4,c)$ Lie algebra to $sl(2,c)$, together with
$H_{_{sl(2,c)}}$ reduction, lead to eigenvalue equation (3.17) which can
be obtained from Hamiltonian (3.11) via similarity transformation. The
Hamiltonian $H_{_{sl(2,c)}}(q)$, on the one hand, is quadratic Casimir operator of $sl(2,c)$
Lie algebra with $J_{_{\pm}}^{(L)}(q)$ and $J_{_{3}}^{(L)}(q)$ generators
which, as a result, possesses $gl(2,c)$ degeneracy symmetry. On the other hand,
writting the Hamiltonian in terms of $J_{_{\pm}}^{(R)}(q)$ will possess the shape invariance
symmetry. Both properties are related to the realization of
para-supersymmetry of arbitrary order \cite{Jafar,Jaf}.
\vspace{7mm}
\begin{center}
{\bf {\large ACKNOWLEDGEMENT}}
\end{center}
We wish to thank Dr. S.K.A. Seyed Yagoobi for carefully reading the
article and for his constructive comments.
|
\section{Notation}
Using the notation of \cite{M},
we will consider the power
$\left\{ p_\lambda[X]
\right\}_\lambda$, Schur $\left\{ s_\lambda[X]
\right\}_\lambda$, monomial $\left\{ m_\lambda[X] \right\}_\lambda$, homogeneous $\left\{ h_\lambda[X]
\right\}_\lambda$, elementary $\left\{ e_\lambda[X] \right\}_\lambda$ and forgotten $\left\{ f_\lambda[X]
\right\}_\lambda$ bases for the symmetric functions. We will often appeal to \cite{M}
for proofs of identities relating these bases.
These bases are all indexed by partitions, non-increasing sequences of non-negative
integers. The $i^{th}$ entry of the partition will be denoted by $\lambda_i$.
The length of a partition $\lambda$ is the largest $i$ such that $\lambda_i$ is non-zero
and will be denoted by $l(\lambda)$.
The size of the partition will be denoted by $|\lambda|$ and is equal to the sum over all
the entries of $\lambda$. The symbol $n_i(\lambda)$ will be used to represent the number
of parts of size $i$ in the partition $\lambda$. The conjugate partition will be
denoted by $\lambda'$ and is the partition such that $\lambda_i' =$ the number of $j$ such that
$\lambda_j$ is greater than or equal to $i$.
There is a standard inner product on symmetric functions $\left< p_\lambda, p_\mu \right> =
z_\lambda \delta_{\lambda\mu}$ (where $\delta_{xy} = 1$ if $x=y$ and $0$ otherwise and $z_\lambda =
\prod_{i \geq 1} i^{n_i(\lambda)} n_i(\lambda)!$.
We will use a few non-standard operations on partitions that will require
some notation. The first is adding a column (or a sequence of columns) to
a partition. Let $\lambda + a^k$ denote the partition $(\lambda_1 +a, \lambda_2 +a, \ldots,
\lambda_k + a)$. We will assume that this partition is undefined when $l(\lambda)>k$.
Use the notation $\lambda - (\mu)$ to denote the partition formed by removing
the parts that are equal to $\mu$ from the partition $\lambda$. This of course assumes
that there is a sequence $I = \{ i_1, i_2, \ldots i_{l(\mu)} \}
\subset \left\{1,2,\ldots l(\lambda) \right\}$ such that
$\mu = (\lambda_{i_1}, \lambda_{i_2}, \ldots, \lambda_{i_{l(\mu)}})$. If this sequence does not exist
then $\lambda - (\mu)$ is again undefined.
The last operation will be inserting
parts into a partition and will be represented by $\lambda + (\mu)$. This will be
the partition formed by ordering the sequence $(\lambda_1, \lambda_2, \ldots, \lambda_{l(\lambda)},
\mu_1, \mu_2, \\ \ldots, \mu_{l(\mu)})$ into a partition.
We will say that two bases for the symmetric functions
$\{a_\lambda\}_\lambda$ and $\{b_\lambda\}_\lambda$ are
dual if they have the property that $\left< a_\lambda, b_\mu \right> = \delta_{\lambda\mu}$.
By definition, the power symmetric functions are dual to the basis
$\{ p_\lambda/z_\lambda \}_\lambda$. The monomial and homogeneous
symmetric functions are dual. The forgotten and the
elementary symmetric functions are dual. The Schur symmetric functions
are self dual ( $\left< s_\lambda, s_\mu \right> = \delta_{\lambda\mu}$).
There exists an involution, $\omega$,
on symmetric functions that relates the elementary and
homogeneous bases by $\omega h_\mu = e_\mu$ and the monomial and forgotten bases
by $\omega m_\mu = f_\mu$. It also has the property that $\omega s_\lambda = s_{\lambda'}$.
Denote the operation of 'skewing' by a symmetric function $f$ by $f^{\perp}$.
It is defined as being the operation dual to multiplication by the symmetric
function $f$ in the sense that $\left< f^\perp P, Q \right> = \left< P, f Q \right>$.
Its action on an arbitrary symmetric function $P$ may be calculated by the formula
$f^{\perp} P = \sum_\lambda \left< P, f a_\lambda \right> b_\lambda$ for any dual bases $\{a_\lambda\}_\lambda$
and $\{b_\lambda\}_\lambda$.
Using results and notation in \cite{M} (p.92-3 example (I.5.25)),
by setting $\Delta f = \sum_\mu (a_\mu^\perp f) \otimes b_\mu^\perp$
where $\{ a_\mu \}_\mu$ and $\{ b_\mu \}_\mu$ are any dual bases. It follows that
if $\Delta f = \sum_i c_i \otimes d_i$ then $f^\perp(PQ) = \sum_i c_i^\perp(P) d_i^\perp(Q)$.
Using this definition, $\Delta h_k = \sum_i h_i \otimes h_{k-i}$ and
$\Delta e_k = \sum_i e_i \otimes e_{k-i}$ and $\Delta p_k = p_k \otimes 1 + 1 \otimes p_k$.
By the phrase 'vertex operators' we are referring to linear symmetric function operators that
add a row or a column to the partitions indexing a particular family of symmetric functions.
Formulas of this type for symmetric functions are sometimes called Rodrigues formulas. In this
article we look at those symmetric function operators which lie in the linear span of
$\left\{ f_i g_i^\perp \right\}_i$ where $f_i$ and $g_i$ are symmetric functions
to find expressions for vertex operators for each basis.
The vertex operators for the elementary
and forgotten symmetric function basis are related to the operators for the homogeneous
and monomial (resp.) symmetric functions by conjugating by the operator
$\omega$.
The existence of such operators for the Schur (\cite{M} p.96-7, \cite{Ze} p.69),
and (row only)
Hall-Littlewood (\cite{M} p.237-8, \cite{J}) symmetric functions are known.
For the multiplicative
bases, it is clear that there exists operators that add a row to the symmetric
functions of this form since $e_k e_\lambda = e_{\lambda+(k)}$, $h_k h_\lambda = h_{\lambda+(k)}$
and $p_k p_\lambda = p_{\lambda+(k)}$, but adding a
column is not an obvious operation.
In general, formulas for adding a row or a column can be useful in proving a
combinatorial interpretation for a symmetric function or deriving new formulas
or properties. The author's interest in this particular question comes from trying to find
vertex operators for the Macdonald symmetric functions. The Macdonald
vertex operator must specialize to the vertex operators for other symmetric functions
and so understanding these operators is an important first step.
\section{The power vertex operator}
This is the warm up case for the other $5$ bases. The commutation
relation between
$p_k^\perp$ and $p_j$ has a nice expression: $p_k^\perp p_j = p_j p_k^\perp
+ k \delta_{kj}$. This can be used to show the slightly more general relation $p_\lambda^\perp p_k =
p_k p_\lambda^\perp + k n_k(\lambda) p_{\lambda-(k)}^\perp$ (where it is assumed that $p_{\lambda-(k)}^\perp = 0$
if $\lambda$ does not contain a part of size $k$).
The vertex operator is given by the following theorem
\begin{thm} For $k\geq 0$ define the following linear operator
$$CP_{a^k} = \sum_{\ontop{\lambda}{l(\lambda) \leq k}} p_a^{k-l(\lambda)} \prod_{i=1}^{l(\lambda)}
\left( p_{\lambda_i+a} - p_{\lambda_i} p_a \right) p_\lambda^\perp/z_\lambda$$
where the sum is over all partitions $\lambda$ with less than or equal to $k$ parts (if
$k=0$ then $CP_{a^0}=1$).
$CP_{a^k}$ has the property that $CP_{a^k} p_\mu = p_{\mu+a^k}$ for all $\mu$ such
that $l(\mu)<k$.
\end{thm}
\vskip .3in
\noindent
{\bf Proof:}
The proof is by induction on the number of parts of $\mu$.
Clearly this operator has the property that $CP_{a^k} 1 = p_a^k$ since $p_\lambda^\perp$
kills $1$ for $|\lambda|>0$. From the
commutation relation of $p_\lambda^\perp$ and $p_k$ we derive that
$$CP_{a^k} p_j = (p_{j+a} - p_j p_a) CP_{a^{k-1}} + p_{j} CP_{a^k}$$
The proof by induction follows from this relation.
\hskip .1in $\Box$
\vskip .3in
The formula $CP_{a^k}$ was chosen so that it has two properties: it adds a column to the
power symmetric functions, and it has a relatively simple expression when written in this
notation. The action of this operator on $p_\lambda$ when $l(\lambda)>k$ is not specified by
these conditions, but it is determined.
If one wishes to give an expression for an operator that has the same
action on $p_\lambda$ for $l(\lambda) \leq k$ and the action on $p_\lambda$ for $l(\lambda)>k$ is something
else (say for instance $0$), this is possible by adding in terms of the form $p_\mu p_\lambda^\perp$
where $l(\lambda)>k$ to $CP_{a^k}$.
\section{Homogeneous and elementary vertex operators}
Note that
$\left< h_i^\perp m_\lambda, h_\mu \right>
= \left< m_\lambda, h_i h_\mu \right> = \delta_{\mu, \lambda - (i)}$. Therefore
\begin{equation}
h_i^\perp (m_\lambda) = m_{\lambda -(i)} \label{hponm}
\end{equation} and $h_i^\perp (m_\lambda) = 0$ if $\lambda$ does not have a part of size $i$.
We then use these results to prove the following lemma
\begin{lemma} \label{hkmlacomm}
$$h_k^\perp m_\lambda = \sum_{i \geq 0}
m_{\lambda - (i)} h_{k-i}^\perp$$
$$m_\lambda^\perp h_k = \sum_{i \geq 0}
h_{k-i} m_{\lambda - (i)}^\perp$$
where we will assume the convention $m_{\lambda -(i)} = 0$ whenever $\lambda -(i)$ is undefined.
\end{lemma}
\vskip .3in
\noindent
{\bf Proof:}
The first identity follows from the remarks made in the previous section
that say that
$h_k^\perp (m_\lambda P) = \sum_i h_i^\perp (m_\lambda) h_{k-i}^\perp (P)$ (where $P$ is any
arbitrary symmetric function).
The second identity is a restatement of the first since
\begin{equation*}
\left< m_\lambda^\perp h_k P, Q \right> = \left< P, h_k^\perp m_\lambda Q \right>
= \sum_{i \geq 0} \left< P, m_{\lambda - (i)} h_{k-i}^\perp Q \right>
= \sum_{i \geq 0}
\left< h_{k-i} m_{\lambda - (i)}^\perp P, Q \right>
\end{equation*}
\hskip .1in $\Box$
\vskip .3in
Define $CH_{1^k}$ to be the operator $CH_{1^k} = \sum_\lambda (-1)^{|\lambda|} e_{\lambda+1^k} m_\lambda^{\perp}$,
and the operator $CE_{1^k}$ to be $CE_{1^k} = \sum_\lambda (-1)^{|\lambda|} h_{\lambda+1^k} f_\lambda^{\perp}$
where we will assume that $h_{\lambda+1^k} = 0$ if $l(\lambda)>k$ so that the sums
in these equations are over all partitions with parts
smaller than or equal to $k$.
\vskip .2in
The vertex operator property that we prove for the homogeneous and elementary symmetric
functions is
\begin{thm}
If $l(\lambda) \leq k$, then $CH_{1^k} h_\lambda = h_{\lambda+1^k}$ and $CE_{1^k} e_\lambda = e_{\lambda+1^k}$ .
\end{thm}
\vskip .3in
\noindent
{\bf Proof:}
The proof is a matter of showing that for $k>0$ operator $CH_{1^k}$ and $h_n$
(considered as an operator that consists of multiplication by $h_n$) has
the commutation relation $CH_{1^k} h_n = h_{n+1} CH_{1^{k-1}}$.
$$CH_{1^k} h_n = \sum_\lambda (-1)^{|\lambda|} e_{\lambda+1^k} m_\lambda^{\perp} h_n$$
The sum here is over $\lambda$ with the number of parts less than or equal to $k$.
Apply Lemma \ref{hkmlacomm} and rearrange the terms in the sum.
\begin{align*}
CH_{1^k} h_n &= \sum_\lambda (-1)^{|\lambda|} e_{\lambda+1^k} \sum_{i \geq 0}
h_{n-i} m_{\lambda - (i)}^\perp \\
&= \sum_\lambda \sum_{i \geq 0}
(-1)^{|\lambda|-i} (-1)^i e_{\lambda+1^k} h_{n-i} m_{\lambda - (i)}^\perp\\
&= \sum_\lambda \sum_{i \geq 0}
(-1)^{|\lambda|-i} (-1)^i h_{n-i} e_{i+1} e_{\lambda- (i) +1^{k-1}} m_{\lambda - (i)}^\perp
\end{align*}
In the last equation, there is an assumption that $e_{\lambda- (i) +1^{k-1}}=0$
if $\lambda - (i)$ is undefined.
As long as $k>1$, making the substitution $\lambda \rightarrow \lambda+(i)$ yields the
equation:
$$=\left( \sum_{i =1}^n
(-1)^i h_{n-i} e_{i+1} \right) \left( \sum_\lambda (-1)^{|\lambda|}
e_{\lambda +1^{k-1}} m_{\lambda}^\perp \right)$$
Now the sum is over $\lambda$ with less than or equal to $k-1$ parts. This is equal to
$=h_{n+1} CH_{1^{k-1}}$
using the well known relation $\sum_{r=0}^n (-1)^r e_r h_{n-r} = 0$ for $n\geq 0$.
If $k=1$ then
\begin{align*}
CH_{1^1} h_n &= \sum_{i =1}^n (-1)^i h_{n-i} e_{i+1} + \sum_{i \geq 1} (-1)^{i} e_{i+1} h_n
m_{(i)}^{\perp} \\ &= h_{n+1} + \sum_{i \geq 1} (-1)^{i} e_{i+1} h_n m_{(i)}^{\perp}
\end{align*}
which is the 'correct' answer only if $CH_{1^1} h_n$ are acting on $1$. Notice also that
$CH_{1^k}$ acting on $1$, yields $h_1^k$ since only one term is not $0$.
The corresponding result for the $CE_{1^k}$ operator follows by noting that
$CE_{1^k} = \omega CH_{1^k} \omega$.
\hskip .1in $\Box$\vskip .3in
The action of $CH_{1^k}$ on $h_\lambda$ when $l(\lambda) > k$ is not known.
The sum in the formula for $CH_{1^k}$ is only over partitions $\lambda$ such
that $l(\lambda) \leq k$ and by adding terms of the same form but with $l(\lambda)>k$
it is possible to modify the formula so that the action on the
$h_\lambda$ when $l(\lambda)>k$ is $0$, but the formula will not be as simple.
It would be interesting to know the action of these vertex operators on other bases
besides the one that it adds a row and column to. For instance, actions of $e_k$, $h_k$, and
$p_k$ are known on the Schur basis, but what is the action
of an operator that adds a column to the homogeneous, elementary, or power basis when
it acts on the Schur basis?
Note the following two formulas that relate $CH_{1^k}$ and $CE_{1^k}$.
\begin{equation} \label{HErel}
CH_{1^k} = \sum_{\ontop{\lambda}{l(\lambda) \leq k}} (-1)^{|\lambda|} CE_{1^k} (e_\lambda) m_\lambda^\perp
\end{equation}
\begin{equation} \label{EHrel}
CE_{1^k} = \sum_{\ontop{\lambda}{l(\lambda) \leq k}} (-1)^{|\lambda|} CH_{1^k} (h_\lambda) f_\lambda^\perp
\end{equation}
This is the first instance when a pair of operators share a relation like this,
and it will occur with pairs of the other operators that appear in this article.
These relations fall under the category of 'eerie coincidences' (by this I mean that
there is probably some explanation for these relations but they are very unexpected
and I don't know what that explanation might be).
\section{Monomial and forgotten vertex operators}
The vertex operators for the monomial and forgotten symmetric functions requires
a few identities.
\begin{lemma} \label{expan}
Let $r_{\mu} = (-1)^{|\mu| - l(\mu)} \frac{l(\mu)!}{n_1(\mu)! n_2(\mu)! \cdots}$
then for $|\mu| \geq 0$,
$e_k = \sum_{\mu \vdash k} r_{\mu} h_\mu$
\end{lemma}
\noindent
{\bf Proof: }
\cite{M} example I.2.20 p.33 \hskip .1in $\Box$
\vskip .3in
\begin{lemma} \label{ecid}
$\sum_{j \geq 0} (-1)^j r_{\mu - (j)} = 0$ where is is assumed that
if $\mu-(j)$ does not exist then $r_{\mu-(j)}=0$.
\end{lemma}
\noindent
{\bf Proof:}
$\sum_{j=0}^k (-1)^j e_{k-j} h_j = 0$. Now expand $e_{k-j}$ in terms of the
homogeneous basis using the last lemma and equate coefficients of $h_\mu$
on both sides of the equation.
\hskip .1in $\Box$
\vskip .3in
\begin{lemma} \label{ekmlacomm}
$e_k^\perp m_\lambda = \sum_\mu r_{\mu} m_{\lambda - (\mu)} e_{k-|\mu|}^\perp$
where $m_{\lambda - (\mu)} = 0$ if $\lambda - (\mu)$ is undefined.
\end{lemma}
\noindent
{\bf Proof:}
By \cite{M} (example I.5.25, p.92-3), $e_k^\perp m_\lambda =
\sum_{i \geq 0} e_i^\perp (m_\lambda) e_{k-i}^\perp$.
The expansion of the $e_i^\perp$ in terms of $h_\mu^\perp$ is given in the last lemma and so
we have that
\begin{equation*}
e_k^\perp m_\lambda = \sum_{i \geq 0} \left(\sum_{\mu \vdash i} r_{\mu} h_\mu^\perp (m_\lambda)
\right) e_{k-i}^\perp
= \sum_{i \geq 0} \left(\sum_{\mu \vdash i} r_{\mu} m_{\lambda-(\mu)}
\right) e_{k-i}^\perp
\end{equation*}
by the equation (\ref{hponm}) and this is equivalent to the statement of the lemma.\hskip .1in $\Box$
\vskip .3in
\begin{lemma} \label{mkmla}
$m_{(k)} m_\lambda = \sum_{i \geq 0} (1+n_{k+i}(\lambda)) m_{\lambda -(i) + (k+i)}$
where it is assumed that $m_{\lambda -(i) + (k+i)} = 0$ if
$\lambda - (i)$ is undefined.
\end{lemma}
\noindent
{\bf Proof:} For paritions $\mu$ of $|\lambda|+k$, one has that
$$m_{(k)} m_\lambda {\Big|}_{m_\mu} = h_\mu^\perp (m_{(k)} m_\lambda)$$
We note that for all $n \geq 0$ that
$h_n^\perp m_{(k)} = m_{(k)} h_n^\perp + h_{n-k}^\perp$. Apply this to
the expression for the coefficient of $m_{\mu}$
$$h_\mu^\perp (m_{(k)} m_\lambda) = \sum_{j=1}^{l(\mu)} h_{\mu - (\mu_j) + (\mu_j-k)}^\perp( \mu_\lambda)$$
This implies that for the coeffient to be non-zero that $\mu$ must be equal to $\lambda$ with
a part (say of size $i$) pulled away and a part of size $k+i$ added in. The coefficient
will be the number of times that $k+i$ appears in the partition $\mu$ (one more time
than it appears in the partition $\lambda$).
\hskip .1in $\Box$
\vskip .3in
The first vertex operator that is presented here
for the monomial symmetric functions adds a
row but it also multiplies by a coefficient
(a property that is unwanted in the final result), but this operator
provides an easy method for obtaining an operator that does not have
this coefficient.
\begin{prop}
Let $RM_a^{(1)} = \sum_{i\geq 0} (-1)^i m_{(a+i)} e_{i}^\perp$
then
$RM_a^{(1)} m_\lambda = (1 + n_a(\lambda)) m_{\lambda + (a)}$
\end{prop}
\vskip .3in
\noindent
{\bf Proof:}
$$RM_a^{(1)} m_\lambda
= \sum_{i\geq 0} (-1)^i m_{(a+i)} e_{i}^\perp m_\lambda$$
Apply Lemma \ref{ekmlacomm} to get
$$= \sum_{i\geq 0} (-1)^i m_{(a+i)}
\sum_{\mu \vdash i} r_{\mu} m_{\lambda - (\mu)}
$$
The sum over $i$ and $\mu$ may be combined to form
one sum over all partitions $\mu$.
$$= \sum_{\mu} (-1)^{|\mu|} r_{\mu}
m_{(a+|\mu|)} m_{\lambda - (\mu)}
$$
Now multiplying by a monomial symmetric function with
one part has an expansion given in Lemma \ref{mkmla}.
$$= \sum_{\mu} \sum_{j \geq 0}
(-1)^{|\mu|} r_{\mu}
(1+n_{a+|\mu|+j}(\lambda - (\mu))) m_{\lambda - (\mu) - (j) + (j+a+|\mu|)}
$$
The terms indexed by the same monomial symmetric function may be
grouped together by letting $\nu = \mu + (j)$.
$$=\sum_\nu \sum_{j\geq 0} (-1)^{|\nu| - j} r_{\nu-(j)}
(1+n_{a+|\nu|}(\lambda - ((\nu)-(j)))) m_{\lambda - (\nu) + (a+|\nu|)}$$
$$=\sum_\nu (1+n_{a+|\nu|}(\lambda)) m_{\lambda - (\nu) + (a+|\nu|)}
\sum_{j\geq 0} (-1)^{|\nu| - j} r_{\nu-(j)}$$
But $\sum_{j\geq 0} (-1)^{|\nu| - j} r_{\nu-(j)} = 0$
if $|\nu|>0$ by Lemma \ref{ecid}. There is one term left.
$$=(1+n_{a}(\lambda)) m_{\lambda + (a)}$$
\hskip .1in $\Box$
\vskip .3in
An expression for an operator that adds a row without a coefficient
can be written in terms of this operator.
\begin{thm}\label{Mavertex}
For $a>0$
define $RM_a = \sum_{k \geq 0} (-1)^k \frac{\left( RM_a^{(1)} \right)^{k+1} }{(k+1)!}
\left(h_a^k\right)^\perp$
then $RM_a m_\lambda = m_{\lambda+(a)}$.
\end{thm}
\vskip .3in
\noindent
{\bf Proof:}
Apply the previous proposition to this formula and reduce using the following steps.
\begin{align*}
RM_a m_\lambda &= \sum_{i \geq 0} (-1)^i
\frac{\left( RM_a^{(1)} \right)^{i+1} }{(i+1)!}
\left(h_a^i\right)^\perp m_\lambda \\
&= \sum_{i \geq 0} (-1)^i \frac{(n_{a}(\lambda)+1)(n_{a}(\lambda)+2) \cdots (n_{a}(\lambda)+i+1)}{(i+1)!}
m_{\lambda - (a^i) +(a^{i+1})} \\
&= \sum_{i = 0}^{n_{a}(\lambda)} (-1)^i \left( \begin{array}{c} n_{a}(\lambda)+1 \\ i+1 \end{array} \right)
m_{\lambda + (a)} \\
&= m_{\lambda + (a)}
\end{align*}
The last equality is true because for $a>0$,
$\sum_{i = 1}^{a} (-1)^{i-1} \left( \begin{array}{c} a \\ i \end{array} \right) = 1$.
\hskip .1in $\Box$
\vskip .3in
Notice that the action of the $RM_a$ operators on the monomial basis implies
that $RM_a RM_b = RM_b RM_a$. This property is difficult to derive just from the
definition of the operator.
This expression for the operator $RM_a$ is a little unsatisfying
since the computation of $\left( RM_a^{(1)} \right)^i$ can be simplified. The
following operator shows how $RM_a$ can be reduced to closer resemble $CH_{1^k}$.
To add more than one row at a time to a monomial symmetric function, the formula resembles the
vertex operator that adds a column to the homogeneous basis.
\begin{prop} \label{makop}
For $a \geq 0$ and $k \geq 0$, we have that
$$RM_a^{(k)} = \frac{\left( RM_a^{(1)} \right)^k}{k!}
= \sum_\lambda (-1)^{|\lambda|} m_{\lambda + a^k} e_\lambda^\perp$$
with the understanding that $m_{\lambda + a^k} = 0$ if ${\lambda + a^k}$ is undefined and
$n_0(\lambda) = k - l(\lambda)$.
It follows that
$RM_a^{(k)} m_\lambda = \left( \begin{array}{c} n_a(\lambda) + k \\ k \end{array} \right) m_{\lambda + (a^k)}$.
\end{prop}
\vskip .3in
\noindent
{\bf Proof:}
By induction on $k$, we will show that $RM_a RM_a^{(k)} = (k+1) RM_a^{(k+1)}$. It follows
that $(RM_a^{(1)})^k = k! RM_a^{(k)}$. Since $(RM_a^{(1)})^k m_\lambda =
(n_a(\lambda) + 1) (n_a(\lambda) + 2) \cdots (n_a(\lambda) + k) m_{\lambda + (a^k)}$ then
$RM_a^{(k)} m_\lambda = \frac{(n_a(\lambda) + 1) (n_a(\lambda) + 2) \cdots (n_a(\lambda) + k)}{k!} m_{\lambda + (a^k)}$.
$$RM_a RM_{a}^{(k)} =
\sum_{j \geq 0} (-1)^j m_{(a+j)} e_j^\perp \sum_\lambda (-1)^{|\lambda|} m_{ \lambda + a^k} e_\lambda^\perp$$
Commute the action of $e_j^\perp$ and $m_{ \lambda + a^k}$ using Lemma \ref{ekmlacomm}.
\begin{align*}
&=\sum_{j\geq 0} (-1)^j m_{(a+j)} \sum_\lambda (-1)^{|\lambda|} \sum_\mu r_{\mu} m_{\lambda + a^k - (\mu)}
e_{j-|\mu|}^\perp e_\lambda^\perp \\
&=\sum_{j \geq 0} \sum_\lambda \sum_\mu (-1)^{|\lambda|+j} r_{\mu} m_{(a+j)} m_{\lambda + a^k - (\mu)}
e_{j-|\mu|}^\perp e_\lambda^\perp
\end{align*}
The formula for multiplying a monomial symmetric function with one part is given in
Lemma \ref{mkmla}.
$$=\sum_{j\geq 0} \sum_\lambda \sum_\mu \sum_{l \geq 0}
(-1)^{|\lambda|+j} r_{\mu} n_{a+j+l}(\lambda + a^k - (\mu) - (l) + (a+l+j))
m_{\lambda + a^k - (\mu) - (l) + (a+l+j)} e_{j - |\mu|}^\perp e_\lambda^\perp$$
The next step is to change the sum over $\mu$ so that it includes the part of
size $l$, this is equivalent to making the replacement $\mu \rightarrow \mu - (l)$.
$$=\sum_{j \geq 0} \sum_{\lambda} \sum_{\mu} \sum_{l \geq 0}
(-1)^{|\lambda|+j} r_{\mu-(l)} n_{a+j+l}(\lambda + a^k - (\mu) + (a+l+j))
m_{\lambda + a^k - (\mu) + (a+l+j)} e_{j+l - |\mu|}^\perp e_\lambda^\perp$$
Let $i=j+l$, then the sum over $j$ can be converted to a sum over $i$.
$$=\sum_{\lambda} \sum_{\mu} \sum_{l \geq 0} \sum_{i \geq l}
(-1)^{|\lambda|+i-l} r_{\mu-(l)} n_{a+i}(\lambda + a^k - (\mu) + (a+i))
m_{\lambda + a^k - (\mu) + (a+i)} e_{i - |\mu|}^\perp e_\lambda^\perp$$
The next step is to interchange the sum over $i$ and the sum over $l$. Since
$l \geq 0$ and $i \geq l$ then $i \geq 0$ and $0 \leq l \leq i$.
$$=\sum_{\lambda} \sum_{\mu} \sum_{i \geq 0} \sum_{l = 0}^{i} (-1)^l r_{\mu-(l)}
(-1)^{|\lambda|+i} n_{a+i}(\lambda + a^k - (\mu) + (a+i))
m_{\lambda + a^k - (\mu) + (a+i)} e_{i - |\mu|}^\perp e_\lambda^\perp$$
Notice that since $e_{i - |\mu|}^\perp$ is zero for all $i<|\mu|$, then
all terms are zero unless $i \geq |\mu|$.
$$=\sum_{\lambda} \sum_{\mu} \sum_{i \geq |\mu|} \sum_{l = 0}^{i} (-1)^l r_{\mu-(l)}
(-1)^{|\lambda|+i} n_{a+i}(\lambda + a^k - (\mu) + (a+i))
m_{\lambda + a^k - (\mu) + (a+i)} e_{i - |\mu|}^\perp e_\lambda^\perp$$
The sum over $l$ is equal to $0$ as long as $|\mu|>0$ using Lemma \ref{ecid}.
A substitution of $i \rightarrow i+|\mu|$ can be made so that $i$ runs over all
integers greater than or equal to $0$.
$$=\sum_{\lambda} \sum_{i \geq 0} (-1)^{|\lambda|+i} n_{a+i}(\lambda + a^k + (a+i))
m_{\lambda + a^k + (a+i)} e_{i}^\perp e_\lambda^\perp$$
Let the sum over $\lambda$ include the part of size $i$, then $\lambda = \lambda + (i)$.
$$=\sum_{\lambda} (-1)^{|\lambda|} \sum_{i \geq 0} n_{a+i}(\lambda + a^{k+1})
m_{\lambda + a^{k+1}} e_\lambda^\perp$$
The sum over $i$ is now independent of $\lambda$ since $\sum_{i \geq 0} n_{a+i}(\lambda + a^{k+1})$
will always be $k+1$.
$$=(k+1) \sum_{\lambda} (-1)^{|\lambda|} m_{\lambda + a^{k+1}} e_\lambda^\perp = (k+1) RM_a^{(k+1)}$$\hskip .1in $\Box$
\vskip .3in
It follows that the formula for $RM_a^{(k)}$ can be substituted into
Theorem \ref{Mavertex} and this provides a more reduced form of the
first formula given for $RM_a$.
\begin{cor}
$RM_a = \sum_{k \geq 0} \sum_\lambda (-1)^{|\lambda|+k} m_{\lambda + a^{k+1}} e_\lambda^\perp (h_a^k)^\perp$
\end{cor}
\vskip .3in
Since the forgotten basis is related to the monomial basis by an application of the
involution $\omega$, then the formulas for the symmetric function operator that
adds a row to the forgotten symmetric functions follows immediately.
\begin{cor}
$RF_a = \sum_{k \geq 0} \sum_\lambda (-1)^{|\lambda|+k} f_{\lambda + a^{k+1}} h_\lambda^\perp (e_a^k)^\perp$
has the property that
$$RF_a f_\lambda = f_{\lambda + (a)}$$
\end{cor}
\vskip .3in
There exists an operator ${\mathcal T}_{-X}$ of the same form as the operators
that exist already in this paper that has the property that ${\mathcal T}_{-X} P[X] =0$,
if $P[X]$ is a homogeneous symmetric function of degree greater than 0 and
${\mathcal T}_{-X} 1 = 1$. This means that the operator applied to an arbitrary symmetric
function has the property that it picks out the constant term of the symmetric function.
\begin{prop}
Define the operator
$${\mathcal T}_{-X}= \sum_\lambda (-1)^{|\lambda|} s_\lambda s_{\lambda'}^\perp$$
Then for any dual bases $\{ a_\mu \}_\mu$ and $\{ b_\mu \}_\mu$ (that is,
$\left< a_\mu, b_\lambda \right> = \delta_{\lambda \mu}$), this is equivalent to
$${\mathcal T}_{-X}= \sum_\lambda (-1)^{|\lambda|} \omega(a_\lambda) b_\lambda^\perp$$
This operator has the property that ${\mathcal T}_{-X} s_\lambda = 0$ for $|\lambda| > 0$
and ${\mathcal T}_{-X} 1 = 1$.
\end{prop}
\vskip .3in
\noindent
{\bf Proof:}
$${\mathcal T}_{-X} s_\mu = \sum_\lambda (-1)^{|\lambda|} s_\lambda s_{\lambda'}^\perp s_\mu$$
This is exactly the same expression as formula (\cite{M}, p.90, (I.5.23.1))
with the $x$ variables substituted for the $y$. This expression is $0$
unless $s_\mu = 1$.
It requires very little to show that
this operator can be given an expression in terms of any dual
basis.
\begin{align*}
{\mathcal T}_{-X} &= \sum_\lambda (-1)^{|\lambda|} \omega(s_\lambda) s_{\lambda}^\perp \\
&= \sum_\lambda (-1)^{|\lambda|} \sum_{\mu\vdash |\lambda|} \left< \omega(s_\lambda), \omega(b_\mu) \right> \omega(a_\mu) s_{\lambda}^\perp \\
&= \sum_\mu \sum_{\lambda \vdash |\mu|} (-1)^{|\mu|} \omega(a_\mu) \left< s_\lambda, b_\mu \right> s_{\lambda}^\perp \\
&= \sum_\mu (-1)^{|\mu|} \omega(a_\mu) b_\mu^\perp
\end{align*}\hskip .1in $\Box$
\vskip .3in
Note that ${\mathcal T}_{-X}$ is actually a special case of a plethystic operator
${\mathcal T}_{Z} P[X] = P[X+Z]$.
Fix a basis of the symmetric functions, $\{ a_\mu \}_\mu$, we may talk about the
symmetric function linear operator that sends $a_\mu$ to the expression $d_\mu$
(where $\{ d_\mu \}_\mu$ is any family of symmetric function expressions).
We can say that this operator lies in the linear span of the operators
$s_\lambda s_\mu^\perp$ and an expression can be given fairly easily.
\begin{cor} (The everything operator)
Let $\{ a_\mu \}_\mu$ be a basis of the symmetric functions and $\{ b_\mu \}_\mu$ be its dual
basis. Then an operator that sends $a_\mu$ to the expression $d_\mu$ is
given by
$$E_{\{ a_\mu \}}^{\{ d_\mu \}} = \sum_\mu d_\mu {\mathcal T}_{-X} b_\mu^\perp$$
In other words we have that $E_{\{ a_\mu \}}^{\{ d_\mu \}}$ acts linearly,
and on the basis $a_\mu$ it has the action $E_{\{ a_\mu \}}^{\{ d_\mu \}} a_\mu = d_\mu$.
\end{cor}
\vskip .3in
\noindent
{\bf Proof:}
Note that when $b_\mu^\perp$ acts on a homogeneous polynomial, the result
is a homogeneous polynomial of degree $|\mu|$ less. Therefore
if $|\mu| > |\lambda|$, then $b_\mu^\perp a_\lambda = 0$. If $|\mu| < |\lambda|$ then
${\mathcal T}_{-X} b_\mu^\perp a_\lambda = 0$ since ${\mathcal T}_{-X}$ kills all non-constant
terms. When $|\mu| = |\lambda|$, we have that $b_\mu^\perp a_\mu = \delta_{\lambda\mu}$
and therefore, ${\mathcal T}_{-X} b_\mu^\perp a_\lambda = \delta_{\lambda\mu}$. This
also implies that $$\sum_\mu d_\mu {\mathcal T}_{-X} b_\mu^\perp a_\lambda = \sum_\mu d_\mu \delta_{\lambda\mu} =
d_\lambda$$\hskip .1in $\Box$
\vskip .3in
This operator looks too general to be of much use, but using known
symmetric function identities it is possible to reduce and derive
expressions for other operators. For instance, the
symmetric function operator that adds a column to the monomial
symmetric functions is a special case of this.
\begin{thm} \label{colum}
Let $CM_{a^k} = \sum_\lambda (-1)^{|\lambda|} \left( \begin{array}{c} n_a(\lambda) + k \\
k \end{array} \right) m_{\lambda + (a^k)} e_\lambda^\perp$, then
$CM_{a^k} m_\lambda = m_{\lambda + a^k}$ with the convention that
$m_{\lambda + a^k} = 0$ if $\lambda + a^k$ is undefined.
\end{thm}
\vskip .3in
\noindent
{\bf Proof:}
We will reduce an expression for $E_{\{ m_\lambda \}}^{\{ m_{\lambda + a^k} \}}$ to one
for $CM_{a^k}$.
$$E_{\{ m_\lambda \}}^{\{ m_{\lambda + a^k} \}} =
\sum_\lambda m_{\lambda + a^k} \sum_\mu (-1)^{|\mu|} m_\mu e_\mu^\perp h_\lambda^\perp$$
Let $r_{\lambda \mu}$ be the coefficient of $e_\mu$ in $h_\lambda$ (by an application of
the involution $\omega$ it is also the coefficient of $h_\mu$ in $e_\lambda$).
Then the expression is equivalent to
$$=\sum_\lambda m_{\lambda + a^k} \sum_\mu (-1)^{|\mu|} m_\mu
e_\mu^\perp \sum_{\gamma \vdash |\lambda|} r_{\lambda \gamma} e_\gamma^\perp$$
Rearranging the sums this may be rewritten as
$$=\sum_\lambda \sum_\mu \sum_{\gamma} (-1)^{|\mu|} m_{\lambda + a^k}
r_{\lambda \gamma} m_\mu e_\mu^\perp e_\gamma^\perp$$
It is possible to group all the terms that skew by the same elementary symmetric function
by making the substitution $\mu \rightarrow \mu - (\gamma)$ since the sum over
$\mu$ and $\gamma$ are over partitions.
$$=\sum_\lambda \sum_\mu \sum_{\gamma} (-1)^{|\mu|-|\gamma|} m_{\lambda + a^k}
r_{\lambda \gamma} m_{\mu-(\gamma)} e_\mu^\perp$$
Note that $m_{\mu-(\gamma)} = h_\gamma^\perp (m_\mu)$ and $\sum_\gamma (-1)^{|\gamma|} r_{\lambda
\gamma} h_\gamma^\perp = (-1)^{|\lambda|} e_\lambda^\perp$.
$$=\sum_\mu \sum_\lambda (-1)^{|\mu|} m_{\lambda + a^k}
(-1)^{|\lambda|} e_\lambda^\perp (m_{\mu}) e_\mu^\perp$$
Notice that the first part of this expression is exactly the operator $RM_a^{(k)}$ acting
exclusively on $m_\mu$. We may then apply Proposition \ref{makop} and note that the
expression reduces to the sum stated in the theorem.
\hskip .1in $\Box$
\vskip .3in
The symmetric function operator that adds a column (or a group of columns) to
the forgotten symmetric functions can be found by conjugating the $CM_{a^k}$ operator
by the involution $\omega$ to derive the following corollary.
\begin{cor}
Let $CF_{a^k} = \sum_\lambda (-1)^{|\lambda|} \left( \begin{array}{c} n_a(\lambda) + k \\
k \end{array} \right) f_{\lambda + (a^k)} h_\lambda^\perp$, then
$CF_{a^k} f_\lambda = f_{\lambda + a^k}$ with the convention that
$f_{\lambda + a^k} = 0$ if $\lambda + a^k$ is undefined.
\end{cor}
The operator that adds a sequence of rows to the monomial symmetric functions and
the operator that adds a sequence of columns are related by a pair of formulas
similar to in the case of formulas (\ref{HErel}) and (\ref{EHrel}). Notice
that proposition \ref{makop} and theorem \ref{colum} say that
\begin{equation} \label{CMrel}
CM_{a^k} = \sum_\lambda (-1)^{|\lambda|} RM_a^{(k)}( m_{\lambda}) e_\lambda^\perp
\end{equation}
\begin{equation} \label{MCrel}
RM_a^{(k)} = \sum_\lambda (-1)^{|\lambda|} CM_{a^k}(m_{\lambda}) e_\lambda^\perp
\end{equation}
This is 'eerie coincidence' number two. The relation between these two operators is
very similar to the relation between $CH_{1^k}$ and $CE_{1^k}$ but not exactly the same.
Once again this is unexpected and unexplained.
\section{Schur vertex operators}
A symmetric function operator that adds a row to the Schur functions is given in \cite{M}
(p.95-6 I.5.29.d) that is of the same flavor as the other vertex operators presented here.
\begin{thm} (Bernstein) \label{bern}
Let $RS_a = \sum_{i\geq 0} (-1)^i h_{a+i} e_i^\perp$, then $RS_a s_\lambda = s_{\lambda+(a)}$
if $a \geq \lambda_1$. In addition, $RS_a RS_b = - RS_{b-1} RS_{a+1}$.
\end{thm}
\noindent {\bf Proof:}
Repeated applications of this operator yeilds
expressions of the Jacobi-Trudi sort.
Use the relation $RS_a h_k = h_k RS_a - h_{k-1} RS_{a+1}$ (which follows
from \cite{M} example (I.5.29.b.5) and (I.5.29.d)),
$RS_a( 1 )= h_a$ and follow the proof of \cite{M} (I.3.(3.4'') p.43) which does not
actually require that the indexing sequence be a partition.
It follows then that
$$RS_{s_1} RS_{s_2} \cdots RS_{s_n} (1) = det \left| h_{s_j-j+i} \right|_{1\leq i,j \leq n}$$
\hskip .1in $\Box$
\vskip .3in
Conjugating this operator by $\omega$ produces an operator that adds
a column to a Schur symmetric function. We will show in this section that
an nice expression exists for a formula for an operator that adds a column
to a Schur function, but with the property that the result is $0$ if
the partition is longer than the column being added.
It follows from the commutation relation of the $RS_a$, that there is
a combinatorial method for calculating the action of $RS_a$ on a Schur function
when $m < \lambda_1$. Let $ht_k(\mu)$ be the integer $i$ such that
$\mu \rfloor_k = (\mu_2 - 1, \mu_3 -1, \ldots, \mu_i-1,
\mu_1 + i - k, \mu_{i+1}, \ldots, \mu_{l(\mu)})$ is chosen
to be a partition. This amounts to removing the first $k$ cells from the border of
$\mu$. If it is not possible to find such an $i$ such that
$\mu \rfloor_k$ is a partition then say that $\mu \rfloor_k$ is undefined.
\begin{cor}
Let $\nu = \lambda + (m+k)$ where $k \geq \lambda_1-m$ ($\nu$ is $\lambda$ resting on a sufficiently
long first row).
$$RS_a s_{\lambda} = (-1)^{ht_k(\nu)-1} s_{\nu \rfloor_k}$$
where it is assumed that $s_{\nu \rfloor_k} = 0$ if $\nu \rfloor_k$ does not exist.
\end{cor}
\vskip .3in
The proof of this corollary is not difficult, just a matter of showing
that the commutation relation of $RS_a RS_b$ agrees with this definition
of $\nu \rfloor_k$ and that the vanishing condition exists because $RS_a RS_{a+1} = 0$.
This definition and corollary are useful in showing that an expression for $(RS_a)^k$
can be reduced to a form that is very similar to the other vertex operators presented here.
\begin{lemma} \label{Smak}
$$(RS_a)^k = \sum_\lambda (-1)^{|\lambda|} s_{\lambda + a^k} s_{\lambda'}^\perp$$
with the convention that $s_{\lambda+a^k}$ is $0$ if $\lambda+a^k$ is undefined.
\end{lemma}
\vskip .3in
\noindent
{\bf Proof:}
By induction on $k$. The statement agrees with Theorem \ref{bern} for $k=1$.
$$RS_a (RS_a)^k
= \sum_{i \geq 0} (-1)^i h_{a+i} e_i^\perp \sum_{\lambda} (-1)^{|\lambda|} s_{\lambda+a^k} s_{\lambda'}^\perp$$
$e_i^\perp$ can be commuted with the Schur function to produce
$$= \sum_{i \geq 0} (-1)^i h_{a+i} \sum_{\lambda} (-1)^{|\lambda|}
\sum_{j = 0}^i e_j^\perp( s_{\lambda+a^k}) e_{i-j}^\perp s_{\lambda'}^\perp$$
Interchange the order of all of the sums.
$$= \sum_{\lambda} \sum_{j \geq 0} \sum_{i \geq j} (-1)^{|\lambda|+i} h_{a+i}
e_j^\perp( s_{\lambda+a^k}) e_{i-j}^\perp s_{\lambda'}^\perp$$
Make the substitution that $i \rightarrow i+j$, making the sum over all $i \geq 0$ and
expand the product $e_{i}^\perp s_{\lambda'}^\perp$. The notation that
$\gamma \slash \lambda' \in {\mathcal V}_i$ means that $\gamma$ differs from $\lambda'$ by
a vertical $i$ strip ($\lambda'_j \geq \gamma_j \geq \lambda'_j+1$ and $|\gamma| = |\lambda|+i$).
$$= \sum_{\lambda} \sum_{j \geq 0} \sum_{i \geq 0} (-1)^{|\lambda|+i+j} h_{a+i+j}
e_j^\perp( s_{\lambda+a^k}) \sum_{\gamma \slash \lambda' \in {\mathcal V}_i} s_{\gamma}^\perp$$
Make the substitution $\gamma \rightarrow \gamma'$ so that the sum is over
all partitions $\gamma$ that differ from $\lambda$ by a horizontal $i$ strip
and rearrange the sums.
$$= \sum_{\lambda} \sum_{i \geq 0} \sum_{\gamma \slash \lambda \in {\mathcal H}_i} (-1)^{|\lambda|+i}
\sum_{j \geq 0} (-1)^j h_{a+i+j} e_j^\perp( s_{\lambda+a^k}) s_{\gamma'}^\perp$$
Now it is only necessary to notice that the sum over $j$ is actually an application of
the Schur vertex operator acting exclusively on the symmetric function $s_{\lambda+a^k}$.
Switch the order of the sums over the partitions and expression becomes
$$= \sum_{\gamma } (-1)^{|\gamma|}
\sum_{i \geq 0}
\sum_{\lambda: \gamma \slash \lambda \in {\mathcal H}_i}
RS_{a+i}( s_{\lambda+a^k}) s_{\gamma'}^\perp$$
There is a sign reversing involution on these terms so that only one term in the
sum over $i$ and $\lambda$ survives, namely, $s_{\gamma + a^{k+1}}$.
If $i=\gamma_1$ then $RS_{a+\gamma_1}(s_{\gamma-(\gamma_1)+a^k}) = s_{\gamma + a^{k+1}}$.
Take any partition $\lambda$ in this sum such that $\gamma \slash \lambda$ is a horizontal strip
of length less than $\gamma_1$.
If $RS_{a+i}(s_{\lambda+a^k}) = 0$, then this term does not contribute to the sum.
If $RS_{a+i}(s_{\lambda+a^k}) = s_{\nu + a^k}$ then $\nu = \lambda+(i+n) \rfloor_n$, where $n=\gamma_1-i$.
There is a combinatorial statement that can be made about partitions that satisfy this
condition, this is a lemma stated in \cite{Z} (Lemma 3.15, p.34).
\begin{lemma}
There exists an involution $I_\gamma^n$ on partitions $\mu$ such
that $\mu \slash \gamma$ is a horizontal $n$ strip, $\mu \rfloor_n$ exists and
$\gamma \neq \mu \rfloor_n$ with the property that $ht_n(I_\gamma^n(\mu)) =
ht_n(\mu) \pm 1$ and $\mu \rfloor_n = I_\gamma^n(\mu) \rfloor_n$.
\end{lemma}
This is exactly the situation here. Set $\mu = \lambda + (i+n)$ then
$\mu \slash \gamma$ is a horizontal strip of size $|\mu| - |\gamma| = |\lambda|+i+n - |\gamma|
= n$. The result then is that all terms cancel EXCEPT for the terms
such that $\gamma = \lambda+(i+n) \rfloor_n$ or $i=\gamma_1$ and
$RS_{a+i}(s_{\lambda+a^k}) = s_{\gamma + a^{k+1}}$.
The sum therefore reduces to
$$=\sum_\gamma (-1)^{|\gamma|}
s_{\gamma + a^{k+1}} s_{\gamma'}^\perp$$
\hskip .1in $\Box$
\vskip .3in
With this expression for the Schur function vertex operator, it is possible to reduce
the expression for the 'everything operator' that adds a column to the Schur functions
but is zero when the length of the indexing partition is larger than the height of
the column being added.
\begin{thm}
For $a,k \geq 0$, let $CS_{a^k} = \sum_\lambda (-1)^{|\lambda|} (RS_a)^k(s_\lambda) s_{\lambda'}^\perp$.
This operator has the property that $CS_{a^k} s_\lambda = s_{\lambda + a^k}$ if $l(\lambda) \leq k$ and
$CS_{a^k} s_\lambda = 0$ for $l(\lambda) > 0$.
\end{thm}
\vskip .3in
\noindent
{\bf Proof:}
Take the expression for the everything operator that adds $a$ columns of height $k$
using the convention that $s_{\lambda+a^k}$ is zero whenever $\lambda+a^k$ is undefined.
$$E_{s_\lambda}^{s_{\lambda+a^k}} = \sum_\lambda s_{\lambda+a^k} \sum_\mu (-1)^{|\mu|} s_\mu s_{\mu'}^\perp
s_{\lambda'}^\perp$$
The coefficients of the expansion of $s_{\mu} s_\lambda$ in terms of Schur functions
are well studied and
there exists formulas and combinatorial interpretations for their calculation. The only
properties that we require here is that the coefficients in the
the expression $s_\mu s_\lambda = \sum_\nu c_{\lambda\mu}^\nu s_\nu$ have the property that
$c_{\lambda\mu}^\nu = c_{\lambda'\mu'}^{\nu'}$ and $s_\lambda^\perp s_\nu = \sum_\nu c_{\mu\lambda}^\nu s_\mu$.
$$= \sum_\lambda s_{\lambda+a^k} \sum_\mu (-1)^{|\mu|} s_\mu \sum_\nu c_{\lambda\mu'}^{\nu'} s_{\nu'}^\perp$$
Next, we rearrange the sums and make the substitution $c_{\lambda\mu'}^{\nu'} = c_{\lambda'\mu}^{\nu}$.
$$= \sum_\lambda s_{\lambda+a^k} \sum_\nu (-1)^{|\nu| - |\lambda|} \sum_\mu c_{\lambda'\mu}^{\nu} s_\mu
s_{\nu'}^\perp$$
Therefore the sum over $\mu$ is just an application of $s_{\lambda'}^\perp$ on $(s_{\nu})$ and the
sums can be rearranged.
$$= \sum_\nu (-1)^{|\nu|} \sum_\lambda (-1)^{|\lambda|} s_{\lambda+a^k} s_{\lambda'}^\perp (s_{\nu})
s_{\nu'}^\perp$$
The sum over $\lambda$ is now exactly an application of Lemma \ref{Smak}.
$$= \sum_\nu (-1)^{|\nu|} (RS_a)^k(s_{\nu}) s_{\nu'}^\perp$$
This is the expression given in the statement of the theorem.
\hskip .1in $\Box$
\vskip .3in
The last of the 'eerie coincidences' of this article is that the $CS_{a^k}$ and $(RS_a)^k$
are related by a pair of formulas similar to the case of formulas (\ref{HErel}),
(\ref{EHrel}) and (\ref{CMrel}), (\ref{MCrel}).
\begin{equation}
CS_{a^k} = \sum_\lambda (-1)^{|\lambda|} (RS_a)^k(s_{\lambda}) s_{\lambda'}^\perp
\end{equation}
\begin{equation}
(RS_a)^k = \sum_\lambda (-1)^{|\lambda|} CS_{a^k}(s_{\lambda}) s_{\lambda'}^\perp
\end{equation}
\section{An application: the tableaux of bounded height}
One observation about the operator $CS_{a^k}$ that could have an interesting application is that
$CS_{0^k} s_\lambda = 0$ if $l(\lambda) > k$ and $CS_{0^k} s_\lambda = s_\lambda$ if $l(\lambda) \leq k$.
Knowing this and the commutation relation between $RS_a$ and $h_k$ allows us to calculate
the number of pairs of standard tableaux of the same shape of bounded height \cite{B}
$\sum_{\ontop{\lambda \vdash n}{l(\lambda)\leq
k}} f_\lambda^2$ (where $f_\lambda$ is the number of standard tableaux of shape $\lambda$).
\begin{prop}
Let $CP(n,k)$ be the collection of sequences of non-negative integers of
length $k$ such that the sum is $n$.
$$\sum_{\ontop{\lambda \vdash n}{l(\lambda)\leq
k}} f_\lambda^2 = \sum_{s \in CP(n,k)} \left( \begin{array}{c} n \\ s \end{array} \right)
\frac{\prod_{i<j} (s_j+j - (s_i+i)) }{\prod_{i=1}^k (s_i + i -1)!} n!$$
\end{prop}
The formula follows by applying $CS_{0^k}$ to the symmetric function $h_1^n$ to arrive
at a formula for the symmetric function $\sum_{\ontop{\lambda \vdash n}{l(\lambda)\leq k+1}}
f_\lambda s_\lambda$.
\begin{lemma}
$$ CS_{0^k}(h_1^n)
= \sum_{\ontop{\lambda \vdash n}{l(\lambda) \leq k}} f_\lambda s_\lambda
= \sum_{s \in CP(n,k)} \left( \begin{array}{c} n\\ s\end{array} \right)
det \left| h_{s_j-j+i} \right|_{1 \leq i,j \leq k}$$
\end{lemma}
\vskip .3in
\noindent
{\bf Proof:}
Use the relation $RS_a h_k = h_k RS_a - h_{k-1} RS_{a+1}$,
$RS_a 1 = h_a$ and induction to calculate that
\begin{equation} \label{rsform}
RS_0^k (h_1^n) = \sum_{l=0}^n \sum_{s \in CP(n-l,k)} (-1)^{n-l} h_1^l \left(
\begin{array}{c} n \\ l,s \end{array} \right) det \left| h_{s_j-j+i} \right|_{1 \leq i,j \leq k}
\end{equation}
Using the relation that $s_\lambda^\perp (h_1^n) = \left( \begin{array}{c} n \\ |\lambda| \end{array}
\right) f_\lambda h_1^{n-|\lambda|}$ we have that
\begin{align*}
CS_{0^k}(h_1^n) &= \sum_\lambda (-1)^{|\lambda|} (RS_0)^k(s_\lambda) s_{\lambda'}^\perp (h_1^n) \\
&= \sum_\lambda (-1)^{|\lambda|} (RS_0)^k(s_\lambda) \left( \begin{array}{c} n \\ |\lambda| \end{array} \right)
f_\lambda h_1^{n-|\lambda|} \\
&= \sum_{i=0}^n \sum_{\lambda \vdash i} (-1)^{i} \left( \begin{array}{c} n \\ i \end{array}
\right) (RS_0)^k(f_\lambda s_\lambda) h_1^{n-i} \\
&= \sum_{i=0}^n (-1)^{i} \left( \begin{array}{c} n \\ i \end{array}
\right) (RS_0)^k(h_1^i) h_1^{n-i}
\end{align*}
Now using (\ref{rsform}) we can reduce this further to
$$=\sum_{m=0}^n \sum_{l=0}^m \sum_{s \in CP(m-l,k)} (-1)^{l}
\left( \begin{array}{c} n \\ m \end{array} \right) \left(
\begin{array}{c} m \\ l,s \end{array} \right) h_1^{n+l-m}
det \left| h_{s_j-j+i} \right|_{1 \leq i,j \leq k}$$
Now switch the sums indexed by $l$ and $m$ and then make the replacement $m \rightarrow m+l$
$$=\sum_{l=0}^n \sum_{m=0}^{n-l} \sum_{s \in CP(m,k)} (-1)^{l}
\left( \begin{array}{c} n \\ m+l \end{array} \right) \left(
\begin{array}{c} m+l \\ l,s \end{array} \right)
h_1^{n-m} det \left| h_{s_j-j+i} \right|_{1 \leq i,j \leq k}$$
Now switch the sums back and rearrange the binomial coefficients
$$=\sum_{m=0}^n \sum_{l=0}^{n-m} \sum_{s \in CP(m,k)} (-1)^{l}
\left( \begin{array}{c} n \\ n-m,s \end{array} \right)
\left( \begin{array}{c} n-m \\ l \end{array} \right)
h_1^{n-m} det \left| h_{s_j-j+i} \right|_{1
\leq i,j \leq k}$$
Now the sum $\sum_{l=0}^{n-m} (-1)^{l} \left( \begin{array}{c} n-m \\ l \end{array} \right)$
will always be zero unless $n-m=0$ and if $n=m$ then it is $1$
and so the entire sum collapses to
$$=\sum_{s \in CP(n,k)} \left( \begin{array}{c} n \\ s \end{array} \right)
det \left| h_{s_j-j+i} \right|_{1 \leq i,j \leq k}$$ \hskip .1in $\Box$
\vskip .3in
\noindent{\bf Proof:} (of proposition)
The proposition follows from this lemma with a little manipulation. There is a linear
and multiplicative homomorphism that sends the symmetric functions to the space of
polynomials in one variable due to Gessel
defined by $ \theta( h_n ) = x^n/n!$. This homomorphism has the property
that $\theta( s_\lambda ) = f_\lambda x^{|\lambda|}/|\lambda|!$. The image of the formula
in the lemma is then
$$\theta( CS_{0^k}(h_1^n) ) = \theta( \sum_{\ontop{\lambda \vdash n}{l(\lambda) \leq k}} f_\lambda s_\lambda )
= \sum_{\ontop{\lambda \vdash n}{l(\lambda) \leq k}} f_\lambda^2 \frac{x^n}{n!}$$
Therefore if we set $(a)_0 = 1$ and $(a)_i = a (a-1) \cdots (a-i+1)$ then we have
(by making a slight transformation that reverses order of the sequence first...$j \rightarrow
n+1-j$, $i \rightarrow n+1-i$ and $s_i \rightarrow s_{n+1-i}$) that
\begin{equation*}
\sum_{\ontop{\lambda \vdash n}{l(\lambda)\leq
k}} f_\lambda^2 = \sum_{s \in CP(n,k)} \left( \begin{array}{c} n\\ s\end{array} \right)
det \left| \frac{(s_j+j-1)_{i-1}}{(s_j+j-1)!} \right|_{1 \leq i,j \leq k} n!
\end{equation*}
\begin{equation*}
\sum_{\ontop{\lambda \vdash n}{l(\lambda)\leq
k}} f_\lambda^2 = \sum_{s \in CP(n,k)} \left( \begin{array}{c} n\\ s\end{array} \right) \prod_{i=1}^k
\frac{1}{(s_j+j-1)!}
det \left| {(s_j+j-1)_{i-1}} \right|_{1 \leq i,j \leq k} n!
\end{equation*}
The determinant is a specialization of the Vandermonde determinant in the variables
$s_j + j -1$ so the formula reduces to the expression stated in the proposition. \hskip .1in $\Box$
We note that in the case that $k=1$ this sum reduces to $1$ and in the case that $k=2$
we have that
\begin{equation*}
\sum_{\ontop{\lambda \vdash n}{l(\lambda)\leq
2}} f_\lambda^2 = \sum_{j=0}^n \left( \begin{array}{c} n\\ j\end{array} \right)
\frac{n-2j+1}{(j)!(n-j+1)!} n!
= \sum_{j=0}^n \left( \begin{array}{c} n\\ j\end{array} \right)^2
\frac{n-2j+1}{ n-j+1 }
\end{equation*}
And this is an expression for the Catalan numbers.
It would be interesting to see if these expressions and equations
could be $q$ or $q,t$ anlogued.
|
\section{\bf Introduction}
In the standard construction of a first-order equilibrium phase transition via
the Gibbs criteria from a quark-gluon plasma (QGP)
to a hot and dense hadron gas (HG), the appearance of a discontinuity in the
entropy per baryon ratio (s/n) makes the phase transition at fixed
temperature T and fixed chemical potential $\mu$ irreversible.
Recently several papers have been addressed to the
question of conserving the entropy per baryon s/n across the phase boundary.
Leonidov et al. [1] have proposed a bag model equation of state (EOS) for the QGP
consisting of massless, free gas of quarks and gluons using a ($\mu$, T)
dependent bag constant $B(\mu, T)$ in an isentropic equilibrium phase transition
from a QGP to the HG at a constant T and $\mu$. Later Patra and Singh [2] have
extended this idea to remedy some anomalous behaviour of such a bag constant
$B(\mu,T)$ through the inclusion of
perturbative QCD corrections in the EOS for QGP. They have also explored the consequences
of such a bag constant
on the deconfining phase transition in the relativistic heavy-ion collisions as well as in the early Universe case [3].\\
The above mentioned analysis refers only to the stationary systems.
But in the context of modern experiments on ultrarelativistic heavy-ion collisions,
the dynamical evolution of the system within the framework of hydrodynamical
models has to be incorporated also. Recently Chernavskaya [4] have suggested
a double phase transition model via
an intermediate phase containing massive constituent quarks and pions.
He claimed that only the continuity
condition in s/n ratio does not provide equilibrium character of first-order
transition in dynamically evolving systems
as for all equilibrium processes, but enthalpy
of the system as well.
Subsequently he has also criticized the work of Ref. 1 and 2 on
the ground of twin constraints arising from Gibbs - Duham equilibrium relation [5]
and enthalpy conservation [6] for an evolving system. The aim of the present
note is to logically counter both these criticisms below.\vspace{0.22 in}\\
\section{\bf Gibbs - Duham Relation}
Ref.[4] suggests that if $\epsilon$ is the energy density and p the pressure
then form-invariance of the relation
\begin{eqnarray}
\epsilon + p - \mu n = s T
\end{eqnarray}
imposes the condition
\begin{eqnarray}
\mu ~~ \frac{\partial B}{\partial \mu} = - T ~~ \frac{\partial B}{\partial T}
\end{eqnarray}
We comment that eq.(2) is untenable due to two reason. Firstly, it
would imply that B, instead of being a function of two independent variables
$\mu $ and T, will depend only on the single variable $\mu $/T as can be
verified by direct differentiation. Secondly, eq.(2) would make it impossible
to apply the iterative analytical procedure [1,2] of solving the basic
partial differential equation based on s/n in the extreme regions
$\mu \rightarrow $ 0, T $\rightarrow \infty $ as well as $\mu \rightarrow
\infty $, T $\rightarrow$ 0. This would imply a conflict with the QCD
sum rule results [7]. \vspace{.2in}\\
\section{\bf Enthalpy Condition}
Ref.[4] mentions that, if $\omega$ = $\epsilon + P $ is the enthalpy
density, then for an evolving hydrodynamic system containing the mixed phase of the
QGP and hadronic gas, the conservation of enthalpy per baryon $\omega$/n
gives an additional constraint. We comment that this constraint is
redundant, i.e., does not give a new information. This is so because even if the
system is evolving with time, we can sit in the local comoving frame where the
relation (1) is expected to hold. Then
\begin{eqnarray}
\frac{\omega}{n} = \frac{(\epsilon + P)}{n} = T ~~ \frac{s}{n} + \mu
\end{eqnarray}
implying that, for any given T and $\mu$, the conservations of $\omega$/n
and s/n are equivalent.\\
\pagebreak
|
\section{Introduction}
GRS 1915+105 is a transient X-ray source that has been extremely active during
the six years since it was discovered in 1992 with the
WATCH instrument on {\it Granat} (\cite{ct92}).
It is located behind the Sagittarius arm of the Milky Way at an estimated
distance of 12.5 kpc (\cite{mr94}), where extinction from interstellar
dust limits optical/IR studies
to wavelengths greater than 1 $\mu$m (\cite{mir94}).
No measurement has yet been made of a binary mass function or orbital
period.
GRS 1915+105 is one of several galactic X-ray sources
observed to produce superluminal radio jets (\cite{mr94}, \cite{fen99}).
One of these sources, GRO J1655-40 (\cite{zha94}), has been observed
optically to be a binary system containing normal F star and a 7 M$_\odot$
compact
object presumed to be a black hole (\cite{ob97}). Since the spectral
properties of GRS 1915+105 are similar to those of GRO J1655-40
(\cite{gro98}, \cite{rem98}), and the luminosity of GRS 1915+105 in outburst
is $5 \times 10^{39}$ ergs ${\rm s}^{-1}$ (25 times the Eddington
luminosity of a neutron star; \cite{gmr97}), it is
thought that GRS 1915+105 also
contains a black hole.
The X-ray spectrum of GRS 1915+105 is typical of a black hole
candidate, and spectral models require at least two emission components.
The energy spectrum
below 10 keV is dominated by emission which appears to be thermal in
origin. This is usually modeled with a multi-temperature blackbody
representing an optically thick, geometrically thin accretion disk
(\cite{mit84}).
The spectrum above 10 keV can be modeled with a power law function,
and is thought to originate from inverse-Compton scattering (\cite{st80}).
In the case of GRS 1915+105, this power law component
is sometimes seen at energies up to 600 keV (\cite{gro98}).
The physical origin and spatial distributions of the Comptonizing
electrons is unknown. It was thought that the electrons were part
of an optically thin corona above the plane of the accretion disk, but
recent numerical simulations indicate that a self-consistent planar corona
cannot produce the spectra seen in the low state of black hole binaries
such as Cygnus X-1 (\cite{dov98}).
Various recent models have suggested that the optically thick
flow may give way to an optically thin flow close to
the black hole in a manner which would produce relativistic electrons
(\cite{chen95}; \cite{ct95}; \cite{ll98}; \cite{tz98}; \cite{dov98}).
These Comptonizing electrons could either be extremely hot,
or they could be part of a bulk flow of matter
streaming toward the black hole. It is also possible that the
relativistic electrons are contained in a bulk outflow or a jet.
The X-ray variability of GRS 1915+105 is spectacular
(see Figure \ref{lcurves}).
Observations with BATSE on the {\it Compton Gamma Ray Observatory}
(\cite{har94}) and with SIGMA on
{\it Granat} (\cite{fig94}) have revealed that GRS 1915+105 is highly
variable in the hard X-rays. When the
{\it Rossi X-Ray Timing Explorer (RXTE)} began observations in
1996, variations of as much as 3 Crab were observed on time scales
from seconds to days (\cite{gmr97}).
The variations of the X-ray intensity and spectrum on time scales of hundreds
of seconds have invited several interpretations. Most models involve a
thermal-viscous instability in an accretion disk (\cite{bel97a}, 1997b)
and some take into account the dissipation of accretion energy in
a hot corona (\cite{tcs97}).
Moreover, some of these cycles of variability in X-rays
have been strongly linked to non-thermal flares in the
infrared (\cite{sam96}, \cite{ef97}, \cite{eik98}) and in the radio
(\cite{mir98}, \cite{fp98}).
These studies have produced the first observational evidence that directly
links the formation of jets to instabilities in the accretion disk.
Quasi-periodic oscillations (QPOs) seen in power density spectra (PDS) of GRS
1915+105 are another area of interesting research
(\cite{mrg97}; \cite{cts97}). One QPO with a centroid frequency of 67 Hz
appears occasionally, and is likely
caused by one of several effects due to general relativity in the inner
accretion disk. Common lower frequency QPOs (0.001 -- 10 Hz)
are broadened in frequency by a random walk in phase, and
exhibit phase lags of a few percent at $\sim 10$ keV relative to
2 keV. The QPO amplitude increases with photon energy, indicating
that these QPOs are associated with the hard X-ray power
law component. Further studies by Markwardt, Swank, \& Taam (1999)
and Trudolyubov, Churazov, \& Gilfanov (1999) have
shown that the {\it frequency}
of a spectrally hard QPO between 0.5-10 Hz is positively correlated with the
thermal flux from the disk. Thus this QPO appears to be linked {\it both}
to the accretion disk and the population of Compton scattering electrons.
QPOs are therefore a promising means of probing the relationship between the
hard and soft components in the spectra of accreting black holes.
In this paper we examine the relationship between the properties of the
0.5--10 Hz QPO and X-ray spectral parameters from
27 observations, which sample both steady states and
repetitive patterns of variability. We find that
the frequency and fractional normalization of the QPO are best correlated
with the temperature of the
inner accretion disk.
We then show that the source has two distinct tracks of spectral evolution,
which can be distinguished by the presence or absence of the intermediate
frequency (0.5--10 Hz) QPOs. (Note that these QPOs are to be distinguished
from the 67 Hz QPO and from the occasional QPOs seen at lower frequency
(0.05--0.2 Hz)).
We find a relationship between the radius and temperature of the
inner disk which spans these two spectral states, and examine the relationship
between the photon
index of and flux from the power law and the parameters of the inner disk.
We conclude that the 0.5--10 Hz QPO is crucial in understanding the origin
of the power law component and the variable X-ray emission from GRS 1915+105.
\section{Data Selection and Analysis}
There have been over 300 observations of GRS 1915$+$105
by the Proportional Counter Array (PCA) (\cite{jah96}) and the High-Energy
X-ray Timing Experiment (HEXTE) (\cite{roth98}) on the {\it Rossi X-ray Timing
Explorer (RXTE)}.
In this paper, we report on our studies of selected observations (see
Table \ref{obs}). These observations represent 15\% of the time spent
observing GRS 1915+105 with the PCA and HEXTE through December 1998.
Although these observations are not representative of the amount
of time GRS 1915+105 spends in any one state, they cover all types
of variability seen to date.
\placetable{obs}
Figure \ref{lcurves} displays
X-ray light curves, PCA hardness ratios (HRs), and ``dynamic power spectra''
(dynamic PDS) which are demonstrative of the types of variability in
the observations listed in Table \ref{obs}. The light curves represent the
count rate in the PCA band (2--30 KeV) per proportional
counting units (PCU). The PCA HRs are the count rate at 13-25 keV relative
to the rate at 2-13 keV. The dynamic PDS are power density spectra computed
every sixteen seconds and rebinned at 0.25 Hz, plotted with time on the
horizontal axis, frequency on the vertical axis, and the linear Leahy power
density represented by the grey-scale from white (0) to
black ($>$ 50).
Figure \ref{lcurves}a characterizes the observations on 1996 May 5,
1997 July 20, 1997 November 17, 1998 February 3, and 1998 February 14.
These observations have steady count rates and do not exhibit the narrow
QPO between 0.5--10 Hz, although low frequency or broad
QPOs are present in some of these observations (see Table \ref{obs}).
We find that HR $<$ 0.05 for this set of observations, so we refer
to this first set of observations as ``soft-steady''.
Figure \ref{lcurves}b is representative of the observations during three
long time intervals: 1996 July 11
through 1996 August 3; 1996 November 28 through 1997 March 26;
and 1997 October 9 through 25. All of these observations have a steady count
rate and a strong 0.5--10 Hz QPO. Their spectra are harder
(0.08 $<$ HR $<$ 0.15) than the soft-steady group, and Morgan,
Remillard, \& Greiner (1997)\markcite{mrg97} label these as ``low-hard
states''.
Chen, Taam, \& Swank (1997)\markcite{cts97} discuss the hardness ratio,
count rate, and PDS of the observations in the first time interval, and
Trudolyubov, Churazov \& Gilfanov (1999) discuss energy spectra
and PDS of the observations from the second time interval.
The radio flux at 8.3 GHz (or 15.2 GHz for the first interval; see
Table \ref{obs}) distinguishes two subsets in these
intervals: the values listed in Table \ref{obs} are greater than 35 mJy
during the {\it RXTE} observations we used from the
first and third intervals, but less than
15 mJy during the observations from second interval
(there is a radio flare to 110 mJy on 1996 December 6, but there
was no coincident {\it RXTE} observation on that date).
Therefore we refer to the {\it RXTE} observations we used from these
intervals as the ``radio-loud hard-steady'' states and ``the radio-quiet
hard-steady'' states respectively. The radio-loud observations are
given particular attention later in this paper, as the parameters we
derive from spectral fits to these observations are difficult to interpret.
The remainder of the observations exhibit a wide range of variability.
Figure \ref{lcurves}c is taken from the observation on 1996 October 7.
The dips in this observation are spectrally hard and contain a 0.5--10 Hz
QPO, while the brighter portions are soft and void of this QPO. Theoretical
models of thermal-viscous instabilities in an optically thick accretion
disk have been used by Belloni et al. (1997a) \markcite{bel97a} to
explain this series of dips; based upon spectral analyses they conclude that
the inner disk empties and re-fills over the course of each dip.
The light curve in Figure \ref{lcurves}d is from the observation on 1997 May
26. The time-series exhibits a QPO
throughout the low portion of the light curve in addition to the
large quasi-periodic ``ringing'' flares every $\sim$ 120 s.
Taam, Chen, \& Swank (1997) \markcite{tcs97} have used numerical simulations
of an unstable accretion disk which dissipates energy into a hot, optically
thin corona to explain many features of these rapid bursts.
The observations on 1997 August 14, 1997 September 9, and
1997 October 30 are similar to the light curve in Figure \ref{lcurves}e.
The longer dips are hard, and contain a QPO; the other features lack the
0.5--10 Hz QPO, and are generally soft. These observations are distinct
from those represented by Figure \ref{lcurves}c in that they display a
large ``spike'' at the end of the long dips, which is
followed by a spectrally soft dip.
Observations of this type exhibit infrared and radio
flares which follow the X-ray dip-spike cycle (see references in introduction).
The hard X-ray dips can also
be explained by the inner disk emptying and re-filling. The observation on
1997 September 9 has been analyzed by Markwardt, Swank, \& Taam (1999)
\markcite{mst98} as well, in a manner similar
to this paper.
The light curve in Figure \ref{lcurves}f is
from the observation on 1997 September 16. The dip/flare
cycles have properties which are the reverse
of cases 1e and 1c, in that the lowest dips in the observation
have soft spectra and lack a 0.5--10 Hz QPO, while the brighter portions
are spectrally harder and contain a QPO. Spectral analyses indicate that
these patterns are not consistent with the disk instability model of Belloni
et al. (1997b).
High-luminosity soft states are presented in Figure \ref{lcurves}g (from
1997 August 13) and Figure \ref{lcurves}h (representative of 1997 August 19
and 1997 December 22). None of these observations contain a narrow 0.5-10
Hz QPO. The observation
in Figure \ref{lcurves}g is presented in Remillard et al. (1998) in
a discussion of the similarities between GRS 1915$+$105 and GRO J1655-40.
Figure \ref{lcurves}i illustrates a moderately soft and
bright interval from the observation on (1997 September 18). During the
course of the observation, the count rate increases
and the spectrum softens on time scales longer than the {\it RXTE} orbit
for which data is displayed.
This observation exhibits the 0.5--10 Hz QPO, which is particularly
strong at lower count
rates and higher HR. This observation also presented some
problems with spectral fits, which we will address later
in this paper.
Finally, the observation in Figure
\ref{lcurves}j displays another type of X-ray ringing (1997 May 18). It
exhibits a series of soft flares ($\sim$ 100 seconds long) that
recur with increasingly longer time intervals and lower amplitudes, until
the series terminates in a long (1000 s), hard minimum in the
count rate. Thereafter, the cycle begins again. The 0.5--10 Hz
QPO is present throughout the observation.
In order to create energy spectra and PDS with good statistics while avoiding
the intrinsic changes due to the chronic
variability of GRS 1915+105, we separate the observations into three
categories that we analyze differently. If the standard deviation in the
PCA count rate over the full energy band width (effectively 2--30 keV) in 1 s
bins is less than 15\% of
the mean count rate during every 96 min {\it RXTE} orbit of an observation,
we collect a single energy
spectrum and PDS for each orbit.
The resulting interval of on-source exposure time outside of the
South Atlantic
Anomaly is generally $\sim 3000$ s. If the count rate changes slowly
from orbit to orbit, but during a single orbit the variability is less than
15\%, we collect energy spectra and PDS every 512 s. If the count rate in
each orbit varies by more than 15\%, we create energy spectra and PDS every
32 s to track the changes in the light curve. Finally, if the count rate varies
by more than 15\% during 32 s, the time segment is removed from our analysis.
Subsequent to these data selections, each spectrum and PDS is analyzed
identically.
\subsection{Energy Spectra}
We perform fits to the energy spectra from both
the PCA (Table \ref{fits}) and HEXTE (Table \ref{hexte}). We use
the standard background subtraction procedures for each instrument,
and apply these to 128 channel spectra from the Standard 2 mode
of the PCA and 64 channel spectra from the Archive mode of HEXTE.
All spectra are first fit in the PCA band alone.
We then fit the combined PCA/HEXTE spectra for the steady observations
only, because the count rate from GRS 1915+105 in the HEXTE band is too
low to analyze with good statistics
on 32 s time scales. We also are unable to analyze HEXTE spectra from
the steady observations
on 1996 May 05, July 11, July 26, or August 03, because the HEXTE clusters
were not rocking between source and background positions, and accurate
background estimates are not available.
This leaves 12 observations for which we provide the results of combined
PCA/HEXTE spectral fits (Table \ref{hexte}).
In order to investigate systematic errors in the PCA and HEXTE response
matrices (from 1997 October 2 and 1997 March 20 respectively), we have
analyzed spectra from the Crab Nebula before modeling more complex spectra
from GRS 1915+105. We fit the Crab spectrum with a model consisting of
a power law with photo-electric absorption. There is sufficient curvature
in the Crab spectrum to prevent an adequate fit over the complete bandwidth
of the PCA and HEXTE. However, the power law fit to each instrument is
good, and we use the results to identify persistent
local features in the residuals of an individual detector unit.
To lessen the statistical weight of such features, we add 1\% systematic
errors to the
PHA bins in both the PCA and HEXTE. In spectra from the PCA, there
are larger systematic deviations in the residuals below 2.5 keV and above
25 keV,
which a leads us to limit the energy range for PCA analysis to
2.5--25 keV. Moreover, the spectral fits from PCUs 2 and 3 have
systematically larger residuals between 5 and 7 keV than do fits from
the other PCUs, which leads us to use only data from PCUs 0, 1, and 4.
We use both HEXTE clusters, and we find systematic deviation in HEXTE fits
to the crab spectrum below 15
keV, which leads us to limit our analysis of HEXTE spectra to energies
greater than 15 keV. The upper limit to fits to HEXTE spectra is determined
by the energy above which the source is no longer detectable, which
occurs between 30--170 keV depending on the observation.
We have applied many models to our spectral analysis of GRS 1915+105, and we
find that 22 out of 27 of the PCA spectra are best fit
by the standard disk black body and power law component model.
We also needed a Gaussian emission line (with a fixed FWHM of 1.0
keV) to measure iron emission between 6 and 7 keV.
When fitting all of our spectra, the column density is
fixed to $N_H = 6.0\times10^{22}$ cm$^{-2}$ of H, which was chosen by
allowing the column density to float in several spectral fits and taking
the average of the resulting values. Finally, we add a fixed multiplicative
constant normalization on each PCU and each HEXTE cluster (when applicable)
to account for differences in the effective areas of the detectors.
This is the same model that has been used for GRS 1915+105 spectra by
Belloni et al. (1997) and Taam, Chen, and Swank (1997).
There are several systematic issues that cause us to use caution when
interpreting
the absolute values of the parameters quoted throughout this paper.
The multi-temperature disk model does not take into account
electron scattering (\cite{ebi93}, \cite{st95}) and
general relativistic effects at the inner disk (\cite{zcc97}), which modify
the emergent spectrum. It is necessary to correct the observed
model parameters for these effects in order to obtain estimates of
the physical parameters of the disk. The accuracy of such corrections is
largely uncertain. Moreover, the black hole mass
is not known, and the temperature and radius of the inner disk appear
wildly variable in GRS 1915+105. In light of these problems we refrain
from applying any of these corrections until
the discussion in Section 4.
The free parameters in our model for the disk emission
are the normalization on the disk black body component ($N_{bb}$) and the
color temperature of the disk at the inner radius
($T_{col}$). The characteristic radius of the
inner disk ($R_{col} = D_{10 kpc}\sqrt{N_{bb}/cos \theta}$ km) is derived
from the normalization of the disk black body ($N_{bb}$), assuming a
distance of 12.5 kpc ($D_{10 kpc}=1.25$) and
an inclination angle ($\theta$) equal to that of the radio jets,
$70^{\circ}$ (\cite{mr94}). The total flux from the disk
is then $F_{bb} = 1.08\times 10^{-11} N_{bb} \sigma T_{col}^4$
ergs$^{-1}$cm$^{-2}$s, where $\sigma$ is the Stephan-Boltzmann constant.
Our model parameters for the power law component are the
photon index of the power law component ($\Gamma$) and
the flux at 1 keV from the hard component ($N_{\Gamma}$).
The flux from the power law ($F_{pl}$) is calculated by integrating
$1.60 \times 10^{-9} N_{\Gamma} E^{-\Gamma+1}$ from 1 to 25 keV, where $E$
is energy in keV.
Finally, our model provides the centroid ($E_{gauss}$) and
normalization ($N_{gauss}$) of the Gaussian.
\placetable{fits}
The parameters, fluxes, and reduced chi-squared values for the PCA spectral
fits are presented in Table \ref{fits}. In 21 of the 27 cases the
standard spectral model provides the best fit with a reduced
chi-squared $< 2.0$. The observation of 1997 September 18 on average
yields a higher value ($\chi^2_{\nu} = 3.18$), but we find no
better alternative model. The remaining
five spectra (from the soft-steady observations, identified
by the use of $E_c$ in Table \ref{fits}) are not
consistent with our standard model. The reduced chi-squared values for these
observations were initially in
the range of 3 to 35, and we were forced to adopt an alternative model.
After exploring many options,
and found that the hard ($> 10$ keV) emission in these observations falls
off as an exponential in energy rather than a power law.
Although several more complicated physical models (e.g. the thermal
Comptonization model of \cite{tit93})
also fit these spectra well, we are unable to constrain
the extra parameters used in these spectral models with our data.
For these five cases we therefore replace the power law
with an exponential photon spectrum, $N_{\Gamma}\exp(-E/E_c)$,
where $E$ is the photon energy, $E_c$ is the cut-off
energy of the exponential, and $N_{\Gamma}$ is flux at 1keV.
The flux from the exponential ($F_{exp}$) is calculated
by integrating over the energy range 1 to
25 keV. The final reduced chi-squared values for all five of these
remaining observations are in the range of 0.5--2.6, and the corresponding
disk parameters (e.g. $T_{col}$, see Table \ref{fits}) are very similar to the
range of results for observations in which the hard component is a power law.
We must further note that fits with a disk black body and a power law in 6
of the observations (viz. the 5 observations
in the radio-loud hard-steady state, plus many segments of 1997 September 18)
yield very small
values for the color radius ($<$ 10 km) and high values for the effective
temperature of the inner disk (3-5 keV).
Similar episodes have also been observed in GRS 1124-68 (Nova Muscae;
\cite{ebi94}), GRO J1655-40 (\cite{sob99a}), and XTE J1550-524
(\cite{sob99b}) in which the disk spectrum appears very hot
with a normalization that implies a small characteristic radius.
These observations are discussed further in Section 4.
\placetable{hexte}
The results for the combined PCA/HEXTE fits are given in Table \ref{hexte}.
The high-energy upper limits for these fits ($E_{max}$) are given in column 2.
As noted above, joint PCA/HEXTE fits are statistically meaningful
only for long exposures during the steady states in GRS 1915+105.
For the soft-steady and
radio-loud, hard-steady observations, the results from the PCA/HEXTE fits
are consistent with those derived from the PCA alone. However, for the
radio-quiet, hard-steady observations, we find
that the addition of HEXTE data to our analysis requires that we add to our
standard model either a high-energy cut-off to the power law (see Table
\ref{hexte}) or a
component representing reflection from un-ionized matter (not shown) in order
to obtain reduced chi-squared values near 1. The reflection model adds a
single free parameter, the relative amount of hard flux which is reflected.
Assuming that the inclination of the disk is $70^{\circ}$ to the line
of sight and that metals in the disk are of Solar abundance, we sometimes
find values of the reflection parameter $>$ 1.0, which is not physically
possible. Consequently, we have chosen to use the cut-off power law in
characterizing the hard-steady spectra.
The addition of a cut-off to the power law introduces a systematic decrease of
$\Gamma \sim 0.2$ for the power law, but leaves the disk spectral parameters
more or less unchanged.
Our results with a cut-off power law are consistent with those of
Trudolyubov, Churazov, \& Gilfanov (1999), who noticed a cut-off
in the power law became detectable when GRS 1915+105 was at low
luminosity throughout
1996 November to 1997 March. The presence of a cut-off in the
power law would supplement the results of Grove et al. (1998),
who demonstrated that GRS 1915+105 was one of a class of black hole
candidates which exhibit power law spectra with slope 2.7 which can extend
to 600 keV without a cut-off. If our modification of the model
to include a cut-off energy is correct, then Table \ref{hexte} would
imply that two types of cut-off power law (with $E_c \sim 3.5$ and 100 keV)
are exhibited by GRS 1915+105. Clearly, more observations in the
100--600 keV range are needed in order to determine whether the
power law in GRS 1915+105 evolves as a function of time.
\subsection{Power Density Spectra}
To create power density spectra we use data which effectively covers
2--30 keV with a time resolution of 122 $\mu$s.
For data segments longer than 32 s, the light curve is divided into
256 s intervals, and a PDS is created for each interval. The PDS
are then averaged for each segment, weighted by the total
counts, and the results are logarithmically rebinned and subtracted
for dead-time-corrected Poisson noise (\cite{mrg97}; \cite{zha96}).
The shape of the PDS are as diverse as those in Morgan,
Remillard, \& Greiner (1997) and Chen, Swank, \& Taam (1997).
When the energy spectra are analyzed in 32 s intervals, PDS are created
for the identical intervals. These spectra are linearly rebinned into
0.25 Hz bins, but otherwise treated as above.
We search for a QPO peak in the PDS by fitting frequency intervals between
0.5--12 Hz with a Lorentzian profile on top of a power-law background
continuum. Only features with a $Q >$ 3 are considered as candidate QPOs.
In addition we varied this range when low-frequency noise obviously
dominated a portion of the PDS; see Figure \ref{lcurves}d, for instance.
To compensate for systematic difficulties in fitting the QPO profiles,
we estimate the significance of the QPO to be the ratio of its amplitude
to the average statistical uncertainty over the QPO width, divided by the
square root of the reduced-chi squared value from the fit. We estimate the
statistical errors on the parameters of the QPO to be the values of the
covariance
matrix from the least chi-squares minimization routine. The average
uncertainty is $\pm 0.04$ Hz in frequency, $\pm 0.04$ Hz in width,
and $\pm 0.02$ in amplitude (which is expressed as an RMS deviation
divided by the mean count rate).
In order to investigate changes in the energy spectrum which are correlated
with the properties of the 0.5--10 Hz QPO, we separate our database of
PCA spectral and QPO
parameters into three groups based on the strength of the observed QPOs:
(1) definite QPOs, for which the best candidate QPO between
0.5--12 Hz has a significance greater than 6.0; (2) no QPO, for which the
best QPO candidate either has a significance less than 2.0
or a significance less than 3 with an amplitude less than 2\%;
and (3) ambiguous cases not selected with the previous criteria.
The third category is ignored in this paper, as the results are inconclusive or
of poor quality.
The values for the significances of the QPOs used to define these selection
criteria are chosen in order to minimize the number
of false assignments among the large number of trial fits.
The joint condition on the amplitude and significance of candidate features in
category 2 is chosen because many of the PDS
contain broad QPOs ($Q <$ 3), knees, and curvature in the background
continuum that affect the distribution of the amplitudes of marginally
significant features.
Out of the 250 ks of exposure time denoted by Table \ref{obs},
54\% yield QPO detections, 15\% show no QPO, and 31\% are dropped
from further study (including those times when the standard deviation in the
count rate is above 15\%, so that no QPO search or spectral
analysis is performed). The amplitude of the definite
QPOs (category 1) is $9 \pm 1$\% for the PDS covering an {\it RXTE} orbit,
and $7 \pm 3$\% for the PDS from 32 s and 512 s intervals. The amplitude
of the best candidate feature in the no QPO case (category 2) is
$0.6 \pm 0.3$\% for PDS
covering and orbit, and $0.8 \pm 0.8$\% for the PDS from 32 s and
512 s intervals. The distributions of the ``with QPO'' and ``without QPO''
groups are therefore statistically well separated.
We must also note that because the 32 s exposures have
limited statistics, we expect that there will be some errors in assigning
these data to each of the three groups. The
dynamic PDS in Figure 1 demonstrates that the 0.5--10 Hz QPO which dominates
the PDS in this frequency range is generally persistent, and varies slowly
in frequency so long as the energy spectrum varies slowly.
However, in rare instances the QPO disappears in a single 32 s time
interval among a series of intervals that otherwise exhibit a persistent
QPO. In addition, visual inspections of QPO fits reveal occasions when
a QPO is found in low frequency ($<$ 3 Hz) noise during a single time
interval when there is no evidence in the dynamic PDS of a persistent QPO.
Of the $\sim 1000$ points plotted in the figures below, we estimate that
50 points may have been mis-categorized in one of these two manners.
Finally, we must emphasize that we have restricted our interest to
features with $Q > 3$ and to the frequency range of 0.5--10 Hz, which
implicitly presumes that occasional QPOs at both low frequencies
($\sim 0.1 Hz$) and high frequencies (e.g. 67 Hz) are different X-ray
oscillation phenomena from the QPOs we examine in this paper.
\section{Results}
In Figures \ref{qpofreq}--\ref{alphat} we present the results of our
analysis of PCA energy spectral and 0.5--10 Hz QPO parameters.
The size of the symbol in each plot
corresponds to the length of the time interval from which the data point was
taken: the large squares correspond to data points taken from
entire {\it RXTE} orbits, the medium-sized asterisks correspond to
data points from 512 s intervals,
and the small points correspond to data from 32 s intervals.
Most of our conclusions use data in the aggregate, but for the reader who
wishes to examine how each type of variability (as explained in Section 2
and illustrated in Figure \ref{lcurves}) relates to the figures, we plot
each type of variability with a separate color. In our scheme, the most red
symbols indicate variability types with the softest spectra, and the blue
those with the hardest spectra. The key for
the large symbols is as follows: the blue squares are radio-quiet hard-steady
observations, the green squares are radio-loud hard-steady observations,
and the red squares are soft-steady observations (which were fit with
an exponential rather than a power law).
Only one observation (1997 September 18; Figure \ref{lcurves}i) was
analyzed at 512 s intervals; this observation is indicated by medium-sized
green asterisks.
The color code for the 32 s points is as follows: red points correspond to
the observations similar to Figures \ref{lcurves}g and h; orange points
to Figure \ref{lcurves}f; yellow-green to Figure \ref{lcurves}c;
green points to Figure \ref{lcurves}d;
blue to Figure \ref{lcurves}e; and purple to Figure \ref{lcurves}j.
Note that few red symbols appear in Figures \ref{qpofreq}--\ref{qpoq}, as
there is no persistent 0.5--10 Hz QPO evident in these soft observations.
\subsection{The Relationship Between QPO and Spectral Parameters}
\placefigure{qpofreq}
We first use our database of spectral parameters and corresponding
0.5--10 Hz QPO parameters (for the definite QPOs) to explore the
correlation between QPO frequency and
flux from the disk black body over a larger range of light curves
than were studied by Markwardt, Swank, \& Taam (1999)
and Trudolyubov, Churazov, \& Gilfanov (1999). The results
are shown in Figure \ref{qpofreq}. Above 5 Hz, the QPO
frequency increases slightly with disk flux, and at frequencies between
0.5--5 Hz
the QPOs generally occur at a nearly constant disk flux between
0.2--0.5$\times 10^{-8}$ ergs/cm$^2$s. However, four of the variable
observations (1996 October 7, 1997 August 14, 1997 September 9, and 1997
October 30) seem to lie off this track at frequencies below 4 Hz. These
exceptional data points are from dips during observations which exhibit
infrared and radio flares (yellow-green points; see Figure \ref{lcurves}e)
and observations when spectral analysis indicate the inner disk
has been evacuated (yellow-green points; see Figure \ref{lcurves}c).
The QPO frequency is correlated with the power law flux during some
individual observations, even at lower fluxes. However, each
observation traces its own track in Figure \ref{qpofreq}b, and it is difficult
to make a generalization of the relationship between power law flux
and QPO frequency that is valid for the entire set of observations.
\placefigure{freqrt}
To further investigate the relationship between the 0.5--10 Hz QPO and the
disk flux (which is simply proportional to $R_{col}^2 T_{col}^4$) we
plot the QPO frequency versus the temperature and radius of the
inner disk in Figures \ref{freqrt}a and b respectively. It is apparent that
the QPO frequency generally increases with the disk temperature, and
decreases with disk radius.
The observations with both high inner disk
temperature and small radii (green squares and green asterisks) are
exceptions to this trend; these points are off the
temperature scale with $T_{col} > 3$ keV in Figure \ref{freqrt}a, and
approach $R_{col} = 0$ in Figure \ref{freqrt}b.
We have compared the relationship between QPO frequency and photon index as
well (not shown), finding no apparent correlation between the
two parameters.
For the majority of the observations (excepting the green squares and asterisks
with high $T_{col}$), the inner disk temperature is
clearly the parameter that is most useful in predicting QPO frequency.
We next consider the width and the integrated amplitude of the 0.5--10 Hz
QPO as they relate to the spectral parameters.
The integrated fractional amplitude is
$$ A_{QPO} = \sqrt{{\pi W H} \over {2I}},$$
where $W$ is the width of the Lorentzian,
$H$ is the maximum value of the Lorentzian in Leahy normalized units,
and $I$ is the mean count rate.
The coherence parameter, $Q$ is a measure of the relative width of the QPO:
$$Q = \nu/W,$$
where $\nu$ is the centroid frequency of the QPO.
\placefigure{qpoamp}
\placefigure{amprt}
Figure \ref{qpoamp}a shows that there is significant scatter in the
correlations between QPO amplitude and disk flux, but that the QPOs
with the highest amplitude occur at low values of disk flux.
There is no significant correlation between QPO amplitude and power law
flux in Figure \ref{qpoamp}b.
Figure \ref{amprt}a demonstrates that the amplitude of the QPO increases
with decreasing $T_{col}$. However, there is no such correlation between
amplitude and $R_{col}$ (Figure \ref{amprt}b).
\placefigure{qpoq}
Finally, the coherence parameter of the QPO does not appear to be
correlated with either component of the flux, and its average value is $Q
\sim$10 (Figure \ref{qpoq}). Similarly, there is no correlation between either
$T_{col}$ or $R_{col}$ (not shown), in direct contrast to the
variations in the frequency and amplitude of the QPO.
We note however that features in the power
spectrum that could be interpreted as very broad QPOs ($Q < 3$) have been
excluded from consideration here.
\subsection{Comparison of Spectra with and without QPOs}
\placefigure{bbpl}
We next turn our attention to the systematic changes that may occur when
the 0.5--10 Hz QPO appears on or off.
Figure \ref{bbpl} demonstrates that the correlation between disk black body
flux and power law flux are strikingly
different when the QPO is present as opposed to when it is not.
When the QPO is present, the power law flux is much more variable than the
disk flux. For the most part, the changes in the total flux track vertically
in this diagram with a relatively constant disk black body flux
(0.3--1.3 $\times 10^{-8}$ ergs/cm$^2$s) and varying power
law flux (0.5--13 $\times 10^{-8}$ ergs/cm$^2$s).
The points at the minimum power law flux correspond to
the lowest count rates during hard dips, such as those in
Figure \ref{lcurves}c (yellow-green points),
The vertical tracks are traced by the changing power law flux during the
entry into and exit from hard dips, such as those in Figures \ref{lcurves}e and
j (blue and purple points).
There are also several horizontal branches in Figure
\ref{bbpl}a, which represent the change in disk flux that
occurs during the bright, hard emission such as in Figure \ref{lcurves}f
(orange points). On the other hand, no steady observation
(green and blue squares; Figure \ref{lcurves}a) that contains a QPO is seen
with a disk black body flux greater than $0.5 \times 10^{-8}$ ergs/cm$^2$s.
This indicates that when the 0.5--10 Hz QPO is present, changes in
luminosity are basically confined to changes in the power law component.
Substantial increases in the disk flux only occur when the disk structure
is cycling through unstable configurations.
When the QPO is absent, the black body component is much more variable
than the power law.
In most cases, the flux in the power law component remains between
2--5 $\times 10^{-8}$
ergs/cm$^2$s (less than half of the maximum), while the flux from the disk
is seen to vary between 0.7--6 $\times 10^{-8}$ ergs/cm$^2$s.
The horizontal track in Figure
\ref{bbpl}b corresponds to the soft emission that follows the hard dips
during the observations similar to those in Figures \ref{lcurves}c and e
(the small yellow-green and blue points respectively).
However, the soft dips of 1997 September 16
(orange points; Figure \ref{lcurves}f), the variable high luminosity soft
states (red points; Figures \ref{lcurves}g and h),
and the soft-steady observations (red squares; Figure \ref{lcurves}b)
also lie on this same track. Our interpretation of Figure \ref{bbpl}b
is that the absence of the 0.5--10 Hz QPO corresponds to an accretion mode
in which changes in the luminosity are primarily seen in the thermal
component of the spectrum.
However, one may wonder whether we do not detect a 0.5--10 Hz QPO in the
accretion mode when flux from the disk is most variable simply because of
the decrease in QPO amplitude with disk black body flux suggested by
Figure \ref{qpoamp} (see also \cite{tcg99}). This is not an issue
at disk black body fluxes less than $2\times10^{-8}$ ergs s$^{-1}$ cm$^{-2}$,
since the QPOs which we find there have much higher amplitudes that
the upper limits for the ``no QPO'' group. At higher fluxes however, our
upper limits ($\sim$ 0.6\%) are only a factor of a
few smaller than the faintest QPOs which we can detect with certainty.
Nonetheless, several indications lead us to believe that the 0.5--10 Hz
QPO truly is absent along the whole horizontal branch in Figure \ref{bbpl}b.
We have visually inspected both dynamic PDS (as in Figure \ref{lcurves})
and individual PDS integrated for longer times (as in Morgan, Remillard,
\& Greiner 1997) for all of our observations, and it is clear that the
power spectrum differs dramatically during the two accretion modes.
Furthermore, if we plot all of our data regardless of QPO properties
(not shown), the two accretion modes which we report are still evident---
searching
for the 0.5--10 Hz QPO only serves to choose one branch or the other other.
Since the branch without the QPO is continuous from low disk black body fluxes
(where the QPO can clearly be detected) to high fluxes, we feel it is
reasonable to believe that the 0.5--10 Hz QPO is genuinely absent from
the accretion mode in which we do not detect this QPO.
\placefigure{rt}
Having found a fundamental difference in the spectrum of GRS 1915+105
when the 0.5--10 Hz QPO is present and absent, we now examine how these
changes manifest themselves in the soft X-ray component of the
spectrum. Figure \ref{rt} demonstrates
that the inner color radius and temperature of the disk are well
correlated, even when comparing wildly different observations. The correlation
is even more compelling if one ignores those points from segments with small
inner disk radii and large temperatures, many of which lie off beyond the
extent of the $x$-axis at temperatures greater than 3 keV (green asterisks
and green diamonds; compare Table \ref{obs}).
Plotted on
top of the data in Figure \ref{rt} is a line corresponding to the function
$R_{col} = 39 T_{col}^{-2} + 22$, which represents a least-chi squares
fit of the dependence of radius on temperature (the fit excludes the
observations with abnormally small $R_{col}$ and large $T_{col}$).
We believe that the form of this relationship is more instructive than
the parameter values themselves, as these parameters may be systematically
dependent on the methods used to model the energy spectrum (e.g. the
choice of an absorption column) at low $T_{col}$, and because some of the
spectral evolution may be due to extrinsic changes, such as modifications
in the spectrum due to electron scattering. The
$T_{col}^{-2}$ dependence of $R_{col}$ indicates a constant disk flux at
low color temperatures, which is seen clearly in Figure \ref{bbpl}a.
Notice also that the color radius is observed to be
relatively constant when the QPO is absent (Figure \ref{bbpl}b),
although even then there is significant scatter in the data around our
empirical correlation.
\placefigure{plt}
\placefigure{alphat}
Finally, we examine the relationship between the inner
disk and the power law component of the spectrum. Figure \ref{plt}
demonstrates that when the 0.5--10 Hz QPO is present, the flux from the power
law component generally increases from $0.5-15 \times 10^{-8}$ ergs/cm$^2$s
as the color temperature of the inner disk increases from 0.7--1.9 keV.
There are, however, exceptions to this trend.
The horizontal branch at $\sim 3 \times 10^{-8}$ ergs/cm$^2$s of power law
flux is composed of the ringing
observation on 97 May 26 (Figure \ref{lcurves}d) and the ringing portion of
the observation on 97 May 18 (Figure \ref{lcurves}j). There is a second
horizontal branch at $\sim 9 \times 10^{-8}$ ergs/cm$^2$s of power law
flux which is corresponds to the observations with unusually
small inner disk radii and large temperatures (green squares and green
asterisks), and extends the scale of the plot at high temperatures.
When the QPO is absent, the power law flux is generally weaker, and
there is little correlation
between the power law flux and color temperature.
Figure \ref{alphat} investigates the relationship between power law index
and disk color temperature. When the QPO is present
the photon index of the
power law component increases from 2.1--3.0 as the color temperature of
the inner disk increases from 0.7--2.0 keV.
The observations with unusually small inner disk radii lie off the plot
at $T_{col} > 3.0$ keV and $2.6< \Gamma < 3.1$.
When the QPO is absent,
there is no correlation between the temperature of the inner disk and the
photon index.
There are related
correlations between the inner radius of the disk and the flux and
photon index of the power law,
as would be expected given the correlation between radius and temperature,
but these correlations (not shown) are weaker. We therefore
conclude that when the 0.5--10 Hz QPO is present, the temperature of the
inner disk can be used to roughly predict the photon index and the flux from
the power law.
\section{Discussion}
Several of the results of this paper demonstrate the significance of the
0.5--10 Hz QPO in GRS 1915+105.
First, we have shown that the frequency of the QPO is best
correlated with the temperature of the inner
accretion disk (Figure \ref{freqrt}). This indicates that the time scale of
this QPO is set by conditions in the optically thick accretion disk.
Second, we have discovered that when the 0.5--10 Hz QPO is present,
changes in the
luminosity of GRS 1915+105 occur mainly in the power law component,
and that when the QPO is absent changes in luminosity occur in
the thermal emission from the inner accretion disk (Figure \ref{bbpl}).
Chen, Swank, \& Taam (1997) and Markwardt, Swank, \& Taam (1998) have similarly
noted that the 0.5--10 Hz QPO is characteristic of hard spectral states.
Moreover, when the QPO is absent during the steady observations
(where good statistics are available) we find that the hard X-ray component
must be modeled with an exponential that produces
negligible flux above $\sim$50 keV, suggesting that the mechanism generating
the Comptonizing electrons is inhibited or that the electrons have
been quenched by Compton cooling.
These results further establish the link between the 0.5--10 Hz
QPO and the hard X-ray power law component in GRS 1915+105, as Morgan,
Remillard, \& Greiner (1997) \markcite{mrg97}
have already demonstrated that the fractional amplitudes of four QPOs
increase with photon energy, extending well past the
thermal component of the spectrum.
Two general classes of models for the formation of QPOs are relevant in
seeking an understanding of the intimate relationship between the 0.5--10 Hz
QPO and both
components of the X-ray spectrum of GRS 1915+105. First, numerous
authors have proposed that oscillations in the inner accretion disk
could lead to QPOs (e.g. \cite{ct94}, \cite{act95}),
and it has recently been noted that if these oscillations
generate ``seed'' photons for Comptonization,
the QPO may be predominately exhibited in the hard portion
of the energy spectrum (\cite{st98}). However, the 0.5--10 Hz QPO in
GRS 1915+105 disappears when the disk appears stable (in terms of the color
radius, as in Figure \ref{rt}b) and luminous (Figure \ref{bbpl}), and there
is no clear reason why the hypothesized
disk oscillations should cease at these times.
The second class of models postulates that QPOs may form at
a geometric boundary between the
optically thick disk and the population of Comptonizing electrons, allowing
for simultaneous modulations of both the hard and soft components of
the spectrum. Models which involve a shock front
between the optically thick and thin regions of the accretion
flow can provide such effects. Oscillations in the height and width of the
shock could result in QPOs by modulating either the number
of ``seed'' photons or the populations of Comptonizing electrons
(\cite{kht97}). However, the time scales of such oscillations are on the order
of $\sim$100 Hz, and better serve as a explanation of the 67 Hz feature seen
in GRS 1915+105 (\cite{tlm98}). Nonetheless, numerical simulations by
Molteni, Sponholz, \& Chakrabarti (1996) \markcite{msc96}
indicate that oscillations in the radial position of
a shock could generate a $\sim$1--10 Hz QPO. These oscillations
are based on a resonance between the cooling time of the shock and
in the free fall time of matter into the black hole, so the time scale is set
by the radius at which the shock forms. The radius is
in turn is set by the accretion rate in the disk, so the correlation between
$T_{col}$ ($\propto ({\dot M}/R^3)^{1/4}$) and QPO frequency in Figure
\ref{freqrt} may support some aspects of this model.
In order to most simply explain the fact that the power law is only
active when the 0.5--10 Hz QPO is present,
one may hypothesize that the same mechanism
generates both the QPO and the relativistic electrons responsible for the hard
X-rays in GRS 1915+105. On the other hand, a 9--30 Hz QPO with similar
spectral characteristics in GRO J1655-40 {\it decreases}
in frequency with X-ray luminosity and disk temperature
(\cite{rem98}, \cite{sob99a}), precisely the
opposite of the QPO is GRS 1915+105 (Figures \ref{qpofreq} and \ref{freqrt}).
If some
QPOs are indeed part of the mechanism generating Comptonizing electrons,
it is not clear why such a discrepancy would exist between two apparently
similar sources.
The results contained in Figure \ref{rt} contribute to further reasons for
caution in using the disk color radius to
estimate the radius of the last stable orbit in accreting black hole
systems (\cite{zcc97}).
We have demonstrated that the disk color radius approached a stable value
only when the 0.5--10 Hz QPO is absent. If we omit the points at $T_{col}
< 1.2$ keV as likely statistical ``leaks'' which are mis-categorized as
lacking a QPO, we
find a minimum stable color radius, $R_{col,min} = 44 \pm 8 (1\sigma)$
km (the uncertainty here is standard deviation of the values for $R_{col,min}$
used to compute the average,
which by coincidence is equal to our average uncertainty on a single
measurement of $R_{col}$). Note that even after this careful selection
of data, there is still a 20\% scatter in color radius values, which should be
considered in addition to the uncertainty on the particular correction factors
from which a physical radius is estimated from the values of $R_{col}$.
If we now use Equation 3 in Zhang, Cui, \& Chen (1997)\markcite{zcc97}
along with their correction factors to convert our estimate of
$R_{col,min}= 44 \pm 8$ to an estimate of the
last stable orbit, we find the compact object in GRS 1915+105 has a
mass $M \sim 14 \pm 4 M_{\odot}$ km for a
Schwarschild black hole, $M \sim 65 \pm 20 M_{\odot}$ for a prograde Kerr
black hole, and $M \sim 9 \pm 3 M_{\odot}$ for a retrograde Kerr black hole.
However, we note that significant uncertainty in this estimate of the mass
is introduced not only through the unknown spin period, but also through
systematic diffuculties in determining the disk parameters (e.g.
\cite{sob99a}).
The radio-loud hard-steady observations with $R_{col} < 10$ km deserve
further consideration, as they provide exceptions to
many of the global trends in Figures \ref{qpofreq}--\ref{alphat}.
In addition to the exceptional values of $R_{col}$ and $T_{col}$ in
Table \ref{fits}, these observations exhibit
particularly strong iron emission in the energy spectrum ($N_{gauss} >
8.5\times 10^{-2}$). Moreover, these radio-loud hard-steady observations are
distinct from the radio-quiet observations, in that joint PCA/HEXTE fits reveal
no necessity for either a reflection component or a cut-off in the power law
(Table \ref{hexte}). The time intervals when the radio-loud hard-steady
observations occurred (1996 July--August and 1997 October) have also been
singled out as exhibiting optically thick radio emission, and the intervals
are referred to as the ``plateau'' radio state by (\cite{fen99}; see also
references therein). It is evident that these
time intervals deserve additional attention, not only to explore the
limits of the the multi-temperature disk model in the X-ray band, but also
to elucidate the relationship between emission from GRS 1915+105 in the X-ray
and radio wavelengths.
Finally, our accretion modes may be compared with the ``states'' which Belloni
et al. (1997b) via Figure \ref{bbpl} used to characterize the variable
emission from GRS 1915+105. The ``quiescent'' state of Belloni et al., which
corresponds to hard dips such as those in Figure \ref{lcurves}c, is
associated with the accretion mode that exhibits the
0.5--10 Hz QPO (Figure \ref{bbpl}a; see also the discussion in Section
3.2).
The ``flaring'' state of Belloni et al. describes the soft, bright emission
that follows the hard dips (Figure \ref{bbpl}a), and is associated with
the accretion mode without the QPO. We also find that the fluxes taken
from steady observations (large squares in \ref{bbpl})
when the emission from GRS 1915+105 was stable for entire orbits
($> 3000$ s) lie on very similar tracks as the fluxes derived from 32 s
intervals (small points) when the source was highly variable.
This suggests an alternative perspective for understanding the cyclic
variations in GRS 1915+105. The variations may represent transitions
between two quasi-steady accretion modes, signified by the presence or
absence of the 0.5--10 Hz QPO. This concept may be useful
toward understanding the ``reverse'' dip cycles represented in Figure
\ref{lcurves}f (orange points in Figure \ref{bbpl}), which are not understood
in the context of the CV-like disk instability model (\cite{bel97a}).
The physics underlying these accretion modes is
unknown, but the transitions may require an explanation beyond the traditional
thermal disk instability.
\section{Conclusions}
We have analyzed a set of 27 PCA observations of GRS 1915+105 which are
a representative sample of the spectral shapes and variability patterns
from this source. We modeled the energy spectrum with a disk black
body, a Gaussian, a constant interstellar absorption, and either a power
law or exponential spectrum. We also searched the PDS for a 0.5--10
Hz QPO. Finally, we compared the parameters of this QPO with the spectral
parameters, and separated our database into two groups based upon whether
or not they contained this QPO.
We extended the results of Markwardt, Swank, \& Taam (1999) and
Trudolyubov, Churazov, \& Gilfanov (1999)
by demonstrating that the 0.5--10 Hz
QPO frequency increases as the color temperature of the inner disk
increases from 0.7 keV--1.5 keV and as the color radius
decreases from 120--20 km (Figure \ref{freqrt}).
On average, the fractional RMS amplitude of the QPO
decreases from 25\% to less than 3\% as the temperature of the inner
disk increases (Figure \ref{amprt}), while the coherence of the QPO does not
change systematically with any of the spectral parameters (Figure \ref{qpoq}).
We have also demonstrated that the 0.5--10 Hz
QPO serves as a marker for two distinct
accretion modes (Figure \ref{bbpl}). When the QPO is present,
accretion energy is channeled primarily into the
power law component of the spectrum and the power law flux and spectral index
roughly increase as the temperature of the disk increases. When the QPO is
absent, accretion energy is primarily expressed in the disk black body
component, while the power law flux and photon index are largely uncorrelated
with the disk color temperature (Figures \ref{bbpl}, \ref{plt}, and
\ref{alphat}).
The color radius and
temperature of the inner accretion disk are related by
$R_{col} \propto T_{col}^{-2} + const$, and that when the QPO is absent, the
color radius remains near a minimum value of $R_{col,min}= 44 \pm 8
(1 \sigma)$ km.
Assuming this minimum
value of the color radius is indicative of the innermost stable orbit,
rough estimates of
the mass of the black hole in GRS 1915+105 range between $\sim 6-80$
M/M$_{\odot}$, depending on the spin.
Eleven of the 27 observations require particular attention in modeling their
spectra. Six observations (with a hard spectrum and 0.5--10 Hz QPO) are
characterized by relatively high radio flux, high color temperatures, and
color inner disk radii less than 5 km. The reduced chi-squared values for
the spectral fits are good, but there is no physical interpretation of
the dramatic decrease in the normalization of the thermal component.
In five other observations that have soft spectra and lack the
narrow 0.5--10 Hz QPO, the hard component must be modeled with
an exponential function.
Finally, we find that much of the emission from GRS 1915+105 can be reduced to
two accretion modes, distinguished by the presence or absence of
a 0.5-10 Hz QPO.
\acknowledgements
We thank Hale Bradt, Al Levine, Wei Cui, and Elizabeth Waltman for careful
readings of drafts of this paper, and Craig Markwardt, Jean Swank,
Ronald Taam for discussions about the content of this paper. This work
was supported by NASA contract NAS 5-30612.
|
\section{Introduction}
The description of the masses of the baryon octet in $SU(3)$ chiral
perturbation theory is not in a satisfactory state. Using heavy baryon
chiral perturbation theory (HBChPT) Borasoy and Meissner \cite{bm} have
carried calculations
to order $Q^4$, where $Q$ denotes a small momentum or a meson mass.
The contributions of order $Q^2$, $Q^3$ and $Q^4$ are of roughly similar
magnitude with alternating sign, thus casting doubt on the convergence of
the expansion. Donoghue and Holstein \cite{dh} have suggested that
the loop integrals should be regularized by introducing form factors.
While this can be used to make the loop contributions tiny, it has the
disadvantage of being model dependent and it is difficult to see how to apply
it consistently in general situations.
An alternative approach has recently been suggested
by Becher and Leutwyler \cite{bl} following the work of Ref. \cite{et}.
It employs chiral perturbation theory in manifestly Lorentz invariant form
with the loop integrals evaluated in the so-called infrared regularization
(IR) scheme. Our purpose here
is to examine this scheme in the context of the baryon masses.
In Sec. II we recall the necessary aspects of chiral perturbation theory
and the infrared regularization scheme. Our results are given in Sec. III,
first through third order and then including an estimate of fourth order.
Finally our conclusions are presented in Sec. IV.
\section{Theory}
The lowest order $SU(3)$ chiral Lagrangian \cite{kr} is
\begin{equation}
{\cal L}^1={\rm Tr}\left\{i\bar{B}(\partial\hspace{-2mm}/ B+[\Gamma\hspace{-2.25mm}/\,,B]-M_0
\bar{B}B+{\textstyle{\frac{1}{2}}} D\bar{B}\gamma_5\{u\hspace{-2.25mm}/\,,B\}+{\textstyle{\frac{1}{2}}} F\bar{B}\gamma_5[u\hspace{-2.25mm}/\,,B]
\right\}\;. \label{l1}
\end{equation}
Here the baryon octet is $B=2^{-\frac{1}{2}}\sum_{a=1}^8\lambda^aB^a$
in terms of the $SU(3)$ matrices $\lambda^a$ and the quantities
involving the meson fields $\phi^a$ are
\begin{eqnarray}
U&=&u^2=\exp\left(\frac{i}{f}\sum_{a=1}^8\lambda^a\phi^a\right)
\nonumber\\
\Gamma_\mu&=&{\textstyle{\frac{1}{2}}}(u^\dagger\partial_\mu u+u\partial_\mu u^\dagger)
\ ;\ u_\mu=i(u^\dagger\partial_\mu u-u\partial_\mu u^\dagger)\;.
\end{eqnarray}
The second order Lagrangian is
\begin{equation}
{\cal L}^2=b_0{\rm Tr}(\bar{B}B){\rm Tr}\chi_++b_d{\rm Tr}
(\bar{B}\{\chi_+,B\})+b_f{\rm Tr}(\bar{B}[\chi_+,B])+\ldots\;,
\label{eq:l2}
\end{equation}
and in $\chi_+=2B_0(u^\dagger{\cal M}u^\dagger+u{\cal M}u)$. In an
obvious notation the quark mass matrix is
${\cal M}={\rm diag}(\hat{m},\hat{m},m_s)$ and this can be related to the
meson masses to leading order
\begin{equation}
m_\pi^2=2\hat{m}B_0\quad;\quad m_K^2=(\hat{m}+m_s)B_0\quad;\quad
m_\eta^2={\textstyle\frac{1}{3}}(4m_K^2-m_\pi^2)\;.
\end{equation}
For present purposes we can use the above Gell-Mann-Okubo relation
for the $\eta$ mass; we take $m_\pi=0.139$ GeV and $m_K=0.494$ GeV
yielding $m_\eta=0.565$ GeV.
The order $Q^2$ contribution to the baryon masses take the familiar form
\begin{eqnarray}
M_N(2)&=&-2b_0(2m_K^2+m_\pi^2)-4m_K^2b_d
+4(m_K^2-m_\pi^2)b_f\nonumber\\
M_\Lambda(2)&=&-2b_0(2m_K^2+m_\pi^2)-{\textstyle\frac{4}{3}}
(4m_K^2-m_\pi^2)b_d\nonumber\\
M_\Sigma(2)&=&-2b_0(2m_K^2+m_\pi^2)-4m_\pi^2b_d\nonumber\\
M_\Xi(2)&=&-2b_0(2m_K^2+m_\pi^2)-4m_K^2b_d
+4(m_\pi^2-m_K^2)b_f\;.
\end{eqnarray}
The order $Q^3$ contribution arises from the loop diagram pictured in Fig.
1 and the guts of this is the integral
\begin{equation}
H=-i\mu^{4-d}\int\frac{d^d\ell}{(2\pi)^d}\frac{1}{(\ell^2-m_M^2+i\epsilon)
(2\ell\cdot P+\ell^2+P^2-M_{B'}^2+i\epsilon)}\;, \label{heq}
\end{equation}
in dimension $d$, with $P$ denoting the four-momentum of the external baryon
leg, $B$. In \cite{et} it was argued that ``hard" part of the integral,
dominated by poles at momenta of the order of the baryon mass, should be
absorbed in the coefficients of the effective Lagrangian. The ``soft"
part of the integral on the other hand needs to be calculated explicitly
and can be obtained by expanding out $\ell^2$ from the baryon propagator
and interchanging the order of integration and summation (the latter
of course changes the value of the integral).
In dimensional regularization the net effect is that $\ell^2$ is
replaced by $m_M^2$ so that (\ref{heq}) is replaced by
\begin{equation}
I=-i\mu^{4-d}\int\frac{d^d\ell}{(2\pi)^d}\frac{1}{(\ell^2-m_M^2+i\epsilon)
(2\ell\cdot P+m_M^2+P^2-M_{B'}^2+i\epsilon)}\:. \label{ieq}
\end{equation}
A similar result has been obtained by Becher and Leutwyler \cite{bl}.
They combine the denominators
in Eq. (\ref{heq}) using an integration over a Feynman parameter $z$.
The ``soft" part is then defined by extending the integration range from
0--1 to 0--$\infty$ and this includes the infrared singular contribution of
leading order $Q^{d-3}$ in the chiral limit $m_M\rightarrow0$ for
$P^2\simeq M_{B'}^2$. The net
effect of these maneuvers is to replace the denominator $(ab)^{-1}$
in $H$ by the denominator $[a(b-a)]^{-1}$ to give $I$ in (\ref{ieq}). These
authors refer to this as infrared regularization. In heavy baryon theory
the second denominator in (\ref{ieq}) is expanded out thus leaving
$(2\ell\cdot P+i\epsilon)^{-1}$ in leading order. Becher and Leutwyler
have argued against making this expansion since it breaks down in certain
regions of parameter space. The present procedure can be viewed as a
summation of the heavy baryon insertions in the baryon propagator to all
orders. The leading order term preserves the Weinberg power counting,
but in addition higher order contributions are included.
The integral that needs to be evaluated for the loop diagram of Fig. 1 is
\begin{equation}
H=-i\mu^{4-d}\int\frac{d^d\ell}{(2\pi)^d}
\frac{\gamma_5\ell\hspace{-1.9mm}/\, (P\hspace{-2.25mm}/\, +\ell\hspace{-1.9mm}/\, +M_{B'})\gamma_5\ell\hspace{-1.9mm}/\,}{(\ell^2-m_M^2+i\epsilon)
(2\ell\cdot P+\ell^2+P^2-M_{B'}^2+i\epsilon)}\:. \label{hfull}
\end{equation}
We apply the so-called infrared regularization on shell, $P^2=M_B^2$.
The resulting integrals contain ultraviolet divergences and these are
removed in the standard $\overline{MS}$ scheme; this requires polynomial
counterterms of arbitrarily high order in $Q$ which we do not need to
specify. Then Eq. (\ref{hfull}) becomes
\begin{eqnarray}
I(M_B,M_{B'}, m_M)&=&\frac{(M_B^2-M_{B'}^2)m_M^2}{32\pi^2M_B}
\ln\frac{m_M^2}{\mu^2}\nonumber\\
&&+\frac{(M_B+M_{B'})}{16\pi^2M_B}[(M_B-M_{B'})z-m_M^2]J(z,m_M)\;,
\label{eqi}
\end{eqnarray}
where $z=(M_B^2-M_{B'}^2+m_M^2)/(2M_B)$ and
\begin{equation}
J(z,m)=\left\{\matrix{z-z\ln\frac{m^2}{\mu^2}-2\sqrt{m^2-z^2}
\cos^{-1}\left(-\frac{z}{m}\right)&\quad|z|<m\cr
z-z\ln\frac{m^2}{\mu^2}-\sqrt{z^2-m^2}\ln\frac{z+\sqrt{z^2-m^2}}
{z-\sqrt{z^2-m^2}}&\quad|z|>m\;.} \right. \label{jeq}
\end{equation}
This is equivalent to the expression given in Ref. \cite{bl} and is of
leading order $Q^3$. It is natural
to identify the renormalization scale $\mu$ in this equation with the
natural scale in the problem, namely the baryon mass in the chiral
limit, $M_0$. The total third order contribution is then
$M_B(3)=\sum_{B'M}\alpha_{BB'M}I(M_B,M_{B'}, m_M)$, where the coefficients
$\alpha$ are easily evaluated using Eq. (\ref{l1}) and are listed in the
Appendix.
To evaluate the baryon masses in fourth order
${\cal L}^2$ requires additional terms besides those displayed in Eq.
(\ref{eq:l2}). They contribute to the loop integrals, but
the value of the coefficients is {\it a priori} unknown. Further,
the finite parts of ${\cal L}^4$ play a role at tree level, again
introducing unknown coefficients. Borasoy and Meissner \cite{bm}
estimated these low energy constants. They concluded that, while their effect
is not negligible, the dominant fourth order contribution arises from
the meson loop diagram pictured in Fig. 2 evaluated using the three terms
of ${\cal L}^2$ given in Eq. (\ref{eq:l2}). Therefore we will consider just
this contribution in order to get an estimate of the magnitude of fourth
order. Since only loops containing baryons require special treatment in
order to keep track of the chiral order, the meson tadpole diagram of Fig. 2
is evaluated in standard fashion irrespective of whether the IR or HBChPT
schemes are used. For nucleons the result, once divergences have been
absorbed in the counterterms, is
\begin{eqnarray}
M_N(4)&=&\biggl\{3(2b_0+b_d+
b_f)m_\pi^4\ln\frac{m_\pi^2}{\mu^2}+
2(4b_0+3b_d-b_f)m_K^4\ln\frac{m_K^2}{\mu^2}+
2b_0m_\eta^4\ln\frac{m_\eta^2}{\mu^2}\nonumber\\
&&\qquad+(-b_d+{\textstyle{\frac{5}{3}}} b_f)m_\pi^2m_\eta^2\ln\frac{m_\eta^2}{\mu^2}+
{\textstyle{\frac{8}{3}}}(b_d-b_f)m_K^2m_\eta^2\ln\frac{m_\eta^2}{\mu^2}\biggr\}/(4\pi f^2)\;,
\end{eqnarray}
and, as we have remarked, we take $\mu=M_0$.
The contributions for the remaining baryons can be read off from Ref. \cite{bm}.
\section{Results}
\subsection{Through Third Order}
In order to get a first comparison of (\ref{eqi}) with the corresponding
heavy baryon result we simply use the physical masses of the
particles and pick a reasonable value of the renormalization scale, $\mu=1$
GeV. The decay constant $f$ is taken to be the average of the kaon
and pion decay constants \cite{pdg}, $f=0.103$ GeV. For the parameters $D$
and $F$ we take the ratio from Close and Roberts \cite{cr} and their sum
is the axial coupling constant \cite{pdg}, giving $D=0.804$ and $F=0.463$.
The result is labelled IREX in Table 1 and can be compared with the heavy
baryon case, labelled HB, for which Eq. (\ref{eqi}) is simply
replaced by $m_M^3/(8\pi)$.
The dominant effect comes from the ratio $z/m$ in Eq. (\ref{jeq}) which
is set to zero in the heavy baryon approach, but can be as large as 0.7
in magnitude here. Clearly it is largest when $B\neq B'$ which is always
the case
for the kaons, since they carry strangeness, and they dominate numerically.
In comparison to the heavy baryon case the net result is a rather small
increase for the $N$ and $\Sigma$ and a fairly substantial, $\sim30$\%,
reduction for the $\Lambda$ and $\Xi$.
Obviously this is not a consistent procedure and since $m_M^2$ is
of ${\cal O}(Q^2)$ we should treat the baryon masses to the same order.
Thus we should evaluate
$M_B(3)=\sum_{B'M}\alpha_{BB'M}I(M_0+M_B(2),M_0+M_{B'}(2), m_M)$.
The total baryon mass through third order is then
$M_B^{\rm tot}=M_0+M_B(2)+M_B(3)$.
Further, as we have remarked, we wish to choose the renormalization scale
to be the baryon mass in the chiral
limit, {\it i.e.} $\mu=M_0$. In order to disentangle the constants $M_0$ and
$b_0$ we need a further piece of information for which we select
the $\pi N$ sigma term,
$\sigma_{\pi N}(0)=\hat{m}\partial M_N^{\rm tot}/\partial\hat{m}$.
The actual value of the sigma term is not precisely known, but for
present purposes we will take the currently
accepted figure of 45 MeV from Gasser {\it et al.} \cite{gls}. This
allows us to determine the unknown parameters $b_0,\:b_d,\:b_f$
and $M_0$ by performing
a least squares fit to the baryon masses and $\sigma_{\pi N}(0)$.
The strange quark contribution to the nucleon mass
$S\equiv m_s\langle N|\bar{s}s|N\rangle=m_s\partial
M_N^{\rm tot}/\partial\hat{m_s}$
can then be calculated. We obtain $S=360$ MeV. While the
magnitude of $S$ is poorly known, this appears a little large so we
constrained $S$ to be 200 MeV in the
final fit shown under the rubric IR in Tables 1 and 2; this degraded the
accuracy of the fit to the baryon masses by about 10 MeV. Note that
a positive value for $S$ is favored here, whereas the
heavy baryon case \cite{bkm} gives a negative value of similar magnitude.
This would suggest larger $KN$ sigma terms here
for which there is some weak experimental support \cite{gen}.
(To obtain the parameters listed in the heavy baryon column of Table 2 the
masses and $\sigma_{\pi N}(0)$ were
fitted with a resulting $S=-195$ MeV.)
Table 1 gives some indication of the
sensitivity to the baryon masses employed. The net loop results evaluated
in the IR scheme are now in all cases smaller than in the HB case with
the reduction ranging from 9\% for the $\Sigma$ to 63\% for the $\Lambda$
and the bulk of the effect arises from the
kaon loop contribution. Note that, referring to Eq. (\ref{jeq}), the
magnitude of the ratio $z/m$ varies over a wide range, from 0.04 to almost
2. Thus the assumption of the heavy baryon scheme that this ratio is small,
so that a power series expansion can be made, is questionable.
Table 2 shows that the baryon masses are fit in the IR
scheme to an accuracy $\sim50$ MeV. More relevant is the ratio of the third
and second order contributions. In the HB case this ranges from
1.11 for the $N$ to 0.70 for the $\Sigma$ and $\Xi$, whereas in the IR
case it ranges from 0.54 for the $\Sigma$ to 0.30 for the $\Lambda$.
The convergence of the expansion is clearly better in the infrared
regularization scheme, due to the smaller loop integrals.
The values of $\sigma_{\pi N}(0)$
and $S$ used in the fit are uncertain. If $\sigma_{\pi N}(0)$ is
increased to 55 MeV or $S$ is increased to 300 MeV the fit to the masses
is improved slightly and the ratio of third to second order becomes
slightly smaller. Conversely for a reduction of $\sigma_{\pi N}(0)$ to
35 MeV or a reduction of $S$ to 100 MeV the results go in the
opposite direction. The changes are reasonably small so that the tabulated
case is representative of our results.
\subsection{Fourth Order Estimate}
Here we estimate fourth order by considering the diagram of Fig. 2.\
If this is calculated
using the parameters previously determined, the value of the diagram in
the IR scheme is 0.17, 0.21, 0.24 and 0.27 GeV for the $N$,
$\Lambda$, $\Sigma$ and $\Xi$, respectively. For the $N$ and the $\Lambda$
these contributions are similar in magnitude to third order, while for
the remaining two baryons they are about half of third order. In the HB
approach the renormalization scale enters for the first time in fourth order
and, taking $\mu=1$ GeV as in Ref. \cite{bm}, we find 0.26, 0.38, 0.40
and 0.51 GeV for the $N$, $\Lambda$, $\Sigma$ and $\Xi$, respectively. The
difference in the magnitudes in the two schemes largely reflects the
different values chosen for the renormalization scale.
Since the
contributions are sizeable, the parameters should be fitted as before
with the fourth order contribution included. The baryon
masses thus become $M_B^{\rm tot}=M_0+M_B(2)+M_B(3)+M_B(4)$.
The results are given in Table 3. In the IR
approach the fitting produces two minima of similar depth. We reject the
solution with a rather low mass $M_0$ of 0.463 GeV and a negative value of
$b_d$ and display the solution which appears to evolve
more naturally from the third order parameterization. Here the value of
$M_0$, 0.653 GeV, is about 10\% smaller than in Table 2 which causes
the fourth order contributions to be reduced by about 37\%; the sensitivity
is obviously due to the fact that $M_0$ is becoming comparable to the eta
and kaon masses. The second and third order contributions display only
modest changes from Table 2, while the totals give a little better fit to
the masses here with an average deviation of 40 MeV. In the
HB case the results displayed correspond to a renormalization scale
$\mu=1$ GeV; they show qualitatively the same trends as the complete fourth
order calculation of Borasoy and Meissner \cite{bm}. The alternative of
choosing the renormalization scale self-consistently results in
$\mu=M_0=0.868$ GeV and the fourth order values of Table 3 are decreased
by 17\%, while second order shows a small 5\% increase. Either way the
strong cancellation, and in some cases overcancellation, between second
and third order remains. In the IR case however the magnitude of third
order is about half that of second order, and our fourth order estimate
is smaller on average by roughly a further factor of a half.
It is also interesting to examine the order-by-order contributions to
the sigma term
\begin{eqnarray}
\sigma_{\pi N}^{IR}&=&0.074-0.036+0.008=0.045\ {\rm GeV}\nonumber\\
\sigma_{\pi N}^{HB}&=&0.059-0.032+0.018=0.045\ {\rm GeV}\;.
\end{eqnarray}
Our values for $\sigma_{\pi N}^{HB}$ are quite close to those reported by
Borasoy and Meissner \cite{bm}. The convergence of the series for
$\sigma_{\pi N}$ is similar in the two schemes, with the IR approach weakly
favored. For completeness we also give breakdown of the strange quark
contribution to the nucleon mass, while noting that $S$ was fitted in the
IR calculation but left free in the HB case,
\begin{eqnarray}
S^{IR}&=&0.254-0.059+0.005=0.200\ {\rm GeV}\nonumber\\
S^{HB}&=&0.093-0.346+0.230=-0.023\ {\rm GeV}\;.
\end{eqnarray}
While the behavior of the IR series looks much better, we caution that
the very small value of the fourth order contribution is somewhat
fortuitous since it is very sensitive to the value of the scale $\mu$.
\section{Conclusions}
In conclusion we have examined the numerical implications of a new form
of $SU(3)$ chiral perturbation
theory where the loop integrals are evaluated using the so-called infrared
regularization scheme. We have examined the chiral series through third
order and made an estimate of fourth order by evaluating the dominant
contribution. The most important feature is that the magnitude of the third
order loop integral contribution to the baryon octet masses is smaller
in the infrared scheme than in HBChPT. This means that the strong
cancellation, in some
cases overcancellation, of second order that is characteristic of HBChPT
no longer occurs. Thus the convergence of the chiral series appears to be
better when infrared regularization is used with successive terms
decreasing in magnitude by about a factor of a half. Given that the
ratio of the kaon and eta masses to the baryon masses is of this order,
this is probably the most that one could expect.
\section*{Acknowledgements}
We thank the referee for a useful suggestion.
This work was supported in part by the US Department
of Energy under grant DE-FG02-87ER40328.
\section*{Appendix}
We list here the non-zero coefficients $\alpha$ needed in the
evaluation of Fig. 1.
\begin{eqnarray}
&&\alpha_{NN\pi}=\alpha_{\Xi\Sigma K}={\textstyle{\frac{3}{2}}}\alpha_{\Sigma\Xi K}
=-\frac{3(D+F)^2}{4f^2}\nonumber\\
&&\alpha_{N\Sigma K}=\alpha_{\Xi\Xi\pi}={\textstyle{\frac{3}{2}}}\alpha_{\Sigma NK}
=-\frac{3(D-F)^2}{4f^2}\nonumber\\
&&\alpha_{NN\eta}=\alpha_{\Xi\Lambda K}={\textstyle{\frac{1}{2}}}\alpha_{\Lambda\Xi K}=
-\frac{(D-3F)^2}{12f^2}\nonumber\\
&&\alpha_{N\Lambda K}=\alpha_{\Xi\Xi\eta}={\textstyle{\frac{1}{2}}}\alpha_{\Lambda NK}
=-\frac{(D+3F)^2}{12f^2}\nonumber\\
&&\alpha_{\Lambda\Lambda\eta}=\alpha_{\Sigma\Sigma\eta}=
\alpha_{\Sigma\Lambda\pi}={\textstyle{\frac{1}{3}}}\alpha_{\Lambda\Sigma\pi}=
-\frac{D^2}{3f^2}\nonumber\\
&&\alpha_{\Sigma\Sigma\pi}=-\frac{2F^2}{f^2}\;.
\end{eqnarray}
|
\section{Introduction}
It is often noted that bright galaxies were assembled at about the same time that QSOs flare (\eg\ Rees 1997), suggesting a causal connection. Blandford (this volume) reviews both the evidence for short lifetimes for QSOs and some recent models for their formation, but leaves open the question of whether they formed before or after the first galaxies. I suggest that, since QSOs are believed to reside in the centers of galaxies (\eg\ Bahcall \etal\ 1997), it is natural to suppose that they formed there. There may be a loose proportionality between the mass of the BH and that of the bulge in which it resides (Magorrian \etal\ 1998). Thus an attractive model would offer answers to at least the following questions:
\begin{itemize}
\item Why should QSOs flare during an early stage of galaxy formation?
\item Why are the centers of galaxies the preferred sites for QSOs?
\item What interrupts the fuel supply to limit QSO lifetimes?
\item Why is the mass of the central BH related to that of the host galaxy?
\end{itemize}
Here I outline a model that offers dynamical answers to all these questions. The main ideas are: (1) most large galaxies developed a bar at an early stage of their formation, (2) the central engine is created from gas driven to the center by the bar and (3) changes to the galaxy potential, caused by mass inflow itself, shut off the fuel supply to the central engine when the mass concentration reaches a small fraction of the galaxy mass. Furthermore, this central mass is sufficient to weaken or even destroy the bar.
\section{Bars in young galaxies}
I adopt the conventional picture that a galaxy disk forms as gas cools and settles into rotational balance in a dark matter halo. As I argue elsewhere (Sellwood, this volume), the DM halo has a large, low-density core. Unless the cooling gas has very low angular momentum, the disk it forms will be extensive and have a gently rising rotation curve at first. Under these conditions, a global bar instability will become unavoidable as the mass of the disk rises (\eg\ Binney \& Tremaine 1987, \S6.3). Thus almost every galaxy with a dominant disk today would have become barred early in its lifetime.
\section{Gas inflow driven by bars}
Many authors (\eg\ Shlosman \etal\ 1990) have suggested that bar-driven gas inflow could fuel QSO activity. The inflow occurs because large-scale shocks develop in the gas flow pattern within the bar which Prendergast (1962) identified with the dust lanes seen along the leading sides of bars in galaxies today. The gas loses both energy and angular momentum in these shocks, the latter because the gas is asymmetrically distributed about the bar major axis. Thus gas is driven towards the center.
Of the many simulations of gas flows in barred galaxy-like potentials, those by Athanassoula (1992) perhaps illustrate most clearly the difference in flow patterns caused by a central mass. A relatively shallow density profile allows gas to flow right into the center (her Figure 4), whereas a mass concentration causes the gas flow to stall some distance from the center (her Figure 2). The different flow pattern in the second case results from the presence of a generalized inner Lindblad resonance of the bar; outside this resonance the flow pattern is generally aligned along the bar, but it switches to perpendicular alignment inside this radius. The flow pattern inside this resonance ring does not contain shocks, and the gas cannot be driven by the bar any closer to the center. This phenomenon is also seen in nearby barred galaxies: nuclear rings occur at radii of a few hundred parsecs in many barred galaxies (Buta \& Crocker 1993) where gas is often observed to pile up (Helfer \& Blitz 1995; Rubin \etal\ 1997).
\section{Quasar activity}
The bar which forms early in the life of a galaxy lacks a central mass concentration and does not possess an ILR. The abundant gas at this epoch will therefore be driven close to the center by the bar. But as the mass in the center rises to a percent or two of the then galaxy mass, an ILR will develop shutting off the dynamically driven flow of gas into the very inner regions. Thus the amount of gas that can reach the central $\sim 50$ pc is naturally limited by the large-scale dynamics.
It is hard to predict the precise fate of the gas as it accumulates in the center, but an attractive guess is that some fraction of it makes a collapsed object while the rest forms stars.
As bars form on the dynamical time-scale of the inner galaxy, and gas inflow time is not much longer, we expect a central engine to be created soon after a galaxy begins to be assembled. The ILR valve will close shortly thereafter, depriving the central collapsed object of further fuel and limiting its mass to a small fraction of the galactic mass.
\section{Bar destruction}
The majority of galaxies are not strongly barred today, so the above picture requires that most bars be destroyed. Simulators have been reporting for years that stellar bars seem to be robust, long-lived objects (\eg\ Miller \& Smith 1979; Sparke \& Sellwood 1986), but it is now known that bars can be destroyed by growing a central object. The mass and concentration needed for a central object to destroy a bar is not known at all precisely; Norman \etal\ (1996) showed that a dense object containing 5\% of the disk plus bulge mass caused the bar to dissolve on a dynamical time, but Friedli's (1994) work suggests that masses of 1-2\% could lead to slower bar decay (see also Hozumi \& Hernquist 1998). The central masses needed seem too high to be just the collapsed object, but all the gas and stars within a radius $\sim 50$ pc should be included.
Sellwood \& Moore (1999) report simulations that mimic this process. They grow a central mass ``by hand'' after a bar develops and limit its mass to 1.5\% of the initial galaxy mass, to mimic the effect of the formation of an ILR. They find that the bar is weakened at this stage, though not yet totally destroyed.
They go on to mimic later infall of fresh material which causes strong spiral patterns to develop in the disk. In some cases, the spiral patterns are vigorous enough to destroy the already weakened bar, but in other cases, infall of the material with the appropriate angular momentum can re-excite the bar.
\section{Conclusions}
I have argued that every bright galaxy should have developed a bar early in its lifetime. The bar drives gas into the center which creates (in an unspecified manner) a central engine for QSO activity. Once the collapsed object, and its surrounding gas and star cluster, reaches a mass of 1-2\% of the (luminous) galaxy mass at that time, an inner Lindblad resonance forms which shuts off the dynamically driven gas supply to the central engine. The dense center also weakens the bar, which may either be destroyed or re-excited, depending on the angular momentum distribution of later infalling material.
This model leads two significant predictions: (1) Halo dominated galaxies, such as LSB or low-luminosity galaxies, which never suffered a bar instability, should not contain massive BHs. A good example of the latter kind is M33, for which Kormendy \& McClure (1993) have placed a very low upper limit of $10^4\;$\Msun\ on the mass of any central BH. (2) The fraction of barred galaxies should be {\it lower\/} in the early universe. This is because the first barred phase should be very short and occurs when the QSO is bright making the bar hard to see. By the time the QSO fades, any residual bar will be short and weak, and the later development of large-scale bars in galaxies is a more gradual process. Some observational support for this prediction is now available (Abraham \etal\ 1998).
The model proposed here does not exclude the possibility that QSO activity would be re-ignited during galaxy mergers. Indeed, the further growth of the BHs during/after a merger will lead to brighter QSOs than those expected in the early stages because the central engines will be more massive.
\acknowledgments
This work was supported by NSF grant AST 96/17088 and NASA LTSA grant NAG 5-6037.
|
\section{Introduction}
Yang-Mills theories \cite{mills} are non-Abelian gauge theories
which have found a central role in particle physics in describing
both the electroweak and strong interactions. The non-Abelian
nature of Yang-Mills theories make the field equations
non-linear, and therefore much more difficult to handle
compared to Abelian gauge theories such as pure electromagnetism.
For example, at the classical level (and also approximately at the
quantum level if the quantum corrections are not too large --
see the Introduction of Ref. \cite{jackson}) one can use
superposition for Abelian gauge theories, while even at the classical
level, superposition is not valid for Yang-Mills theories.
This non-linear nature of the Yang-Mills field
equations makes finding solutions difficult.
There are some well known and interesting solutions
to the Yang-Mills field equations, such as the t 'Hooft-Polyakov
monopole \cite{thooft}, the Julia-Zee dyon \cite{zee}, the
BPS dyon \cite{bogo} \cite{prasad}, and the instanton \cite{poly}, but
there is no systematic way of arriving at solutions to the Yang-Mills
field equations.
General relativity can also in some sense be considered a non-Abelian
gauge theory \cite{uti} \cite{carmeli}, and a mathematical connection
between the two theories can be made \cite{lun2} \cite{lunj}. Using this
connection between the two theories one can ask if the solutions
to the field equations of one theory could provide a
starting point to look for solutions in the other theory. This is
in fact possible, and one can find a host of solutions in this
manner. In this paper we will give a review of the various solutions
found in this way and discuss some of their properties. All of the
solutions discovered in this way have the apparent weak point
that they have an infinite field energy, {\it i.e.},
there are singularities in the fields of the solutions
which make the field energy infinite. This is to be contrasted with
the finite field energy solutions of Refs. \cite{thooft}
\cite{zee} \cite{bogo} \cite{prasad}. However, aside from the
mathematical interest in studying all types of solutions that occur
in such non-linear field theories, we present some ideas concerning
the possible physical uses of such singular solutions. One speculation
is that some of these solutions may be connected with the confinement
phenomenon of the strong interaction. Just as the various black hole
solutions (Schwarzschild or Kerr black holes) exhibit a type of
confinement for any particle which crosses the event horizon, so too
the Yang-Mills analogs of these solutions may exhibit a
confining behaviour.
The outline of the paper is as follows : First we will discuss
the spherically symmetric solutions of the SU(2) Yang-Mills
equations coupled to a scalar field (these are usually called the
Yang-Mills-Higgs equations). Second we will discuss solutions
for gauge groups other than SU(2).
Finally we will examine the behaviour of a test particle which
is placed in the potential of the Schwarzschild analog solution.
We will see that under certain conditions this analog
solution can confine the test particle, and that this system
has a half-integer angular momentum even though all the fields
involved are of integer angular momentum.
\section{SU(2) Yang-Mills Field Equations for Spherically Symmetric
Field Configurations}
The system studied in this section is an SU(2) gauge theory
coupled to a scalar field, $\phi ^a$, in the triplet representation.
The scalar field is taken to have no mass or self interaction.
The Lagrangian for this system is
\begin{equation}
\label{lagrange}
{\cal L} = -{ 1\over 4} G_{\mu \nu} ^a G^{\mu \nu} _a +
{1 \over 2} (D_{\mu} \phi _a ) (D^{\mu} \phi ^a )
\end{equation}
where $G_{\mu \nu} ^a$ is the field strength tensor of the SU(2)
gauge fields, which is defined in terms of the gauge fields $W_{\mu} ^a$,
as
\begin{equation}
\label{fst}
G_{\mu \nu} ^a = \partial _{\mu} W_{\nu} ^a - \partial _{\nu} W_{\mu} ^a
+ g \epsilon _{abc} W_{\mu} ^b W_{\nu} ^c
\end{equation}
and $D_{\mu}$ is the covariant derivative of the scalar field which is
given by
\begin{equation}
\label{codev}
D_{\mu} \phi ^a = \partial _{\mu} \phi ^a + g \epsilon _{abc} W_{\mu} ^b
\phi ^c
\end{equation}
The general equations of motion for this system are
\begin{eqnarray}
\label{eqnmom}
\partial ^{\nu} G _{\mu \nu } ^a &=& g \epsilon _{abc} \left[
G_{\mu \nu b} W^{\nu} _c - (D_{\mu} \phi _b) \phi _c \right]
\nonumber \\
\partial ^{\mu} D_{\mu} \phi ^a &=& g \epsilon _{abc} (D_{\mu} \phi _b)
W_c ^{\mu}
\end{eqnarray}
these field equations can be simplified through the use of a generalized
Wu-Yang ansatz \cite{yang} which was used by Witten \cite{wit}
to study multi-instanton solutions
\begin{eqnarray}
\label{wuyang}
W_i ^a &=& \epsilon _{aij} {r^j \over g r^2} [1 - K(r)] +
\left( {r_i r_a \over r^2} - \delta _{ia} \right)
{G(r) \over g r} \nonumber \\
W_0 ^a &=& {r^a \over g r^2} J(r) \nonumber \\
\phi ^a &=& {r^a \over g r^2} H(r)
\end{eqnarray}
$K(r)$, $G(r)$, $J(r)$, and $H(r)$ are the ansatz functions to be
determined by the equations of motion. Inserting these
expressions into the field equations in Eq. (\ref{eqnmom}) we
find the following set of coupled, non-linear equations,
\begin{eqnarray}
\label{difeq}
r^2 K'' &=& K (K^2 + G^2 + H^2 - J^2 - 1) \nonumber \\
r^2 G'' &=& G (K^2 + G^2 + H^2 - J^2 -1) \nonumber \\
r^2 J'' &=& 2J (K^2 + G^2) \nonumber \\
r^2 H'' &=& 2 H (K^2 + G^2)
\end{eqnarray}
where the primes denote differentiation with respect to $r$. The
most well known solutions to these equations are those
discovered by Prasad and Sommerfield \cite{prasad} and independently by
Bogomolnyi \cite{bogo}. They are
\begin{eqnarray}
\label{soln}
K(r) &=& cos (\theta) C r \; csch(Cr) \; \; \; \; \; \; \; \; \; \; \; \;
G(r) = sin (\theta) C r \; csch(Cr) \nonumber \\
J(r) &=& sinh(\gamma ) [1 - C r \; coth(C r)] \; \; \; \; \; \; \;
H(r) = cosh (\gamma ) [ 1 - C r \; coth(C r)]
\end{eqnarray}
where $C$, $\theta$ and $\gamma$ are
arbitrary constants. One of the nice
properties of this solution is that it has finite field
energy. In terms of the ansatz functions the energy
density of the fields is
\begin{eqnarray}
\label{energy}
T^{00} &=& {1 \over g^2 r^2 } \Bigg( {K'}^2 + {G'}^2 +
{(K^2 + G^2 - 1)^2 \over 2 r^2} + {J^2 (K^2 + G^2) \over r^2}
+ {(rJ' - J)^2 \over 2 r^2} \nonumber \\
&+& {H^2 (K^2 + G^2) \over r^2}
+ {(r H' - H)^2 \over 2 r^2} \Bigg)
\end{eqnarray}
For the solution in Eq. (\ref{soln}) this gives a non-singular
energy density, which when integrated over all space yields
a finite field energy of $E = 4 \pi C cosh^2(\gamma) / g^2$.
This finite energy property of the BPS solution is one of the
main reasons for the interest in this classical
solution. We now examine the general relativistic analog solutions
of the Yang-Mills equations.
\subsection{Solutions with Spherical Singularities}
To find the general relativistic analog solutions to the Yang-Mills
field equations we begin by examining the Schwarzschild solution
of general relativity.
The Schwarzschild solution in Schwarzschild coordinates has two
non-trivial components to the metric tensor : $g_{00}$ and $g_{rr}$.
The non-trivial spatial element has the form $g_{rr} = {K r \over
1 - K r}$ and $g_{00} = - 1/ g_{rr}$ where $K = 1 / 2GM$.
Trying this form of $g_{rr}$ in Eq. (\ref{difeq})
one immediately finds the following solution
\begin{eqnarray}
\label{schsol}
K(r) &=& {\mp \cos \theta C r \over 1 \pm C r} \; \; \; \;
G(r) = {\mp \sin \theta C r \over 1 \pm C r} \nonumber \\
J(r) &=& {\sinh \gamma \over 1 \pm C r} \; \; \; \; \; \;
H(r) = {\cosh \gamma \over 1 \pm C r}
\end{eqnarray}
where $C, \gamma$ and $\theta$ are arbitrary constants. The solution
with the minus sign in the denominator (which we call the
Schwarzschild-like solution) develops a singularity in the
gauge fields ($W_{\mu} ^a$) and scalar fields ($\phi ^a$) on a
spherical surface of radius $r = r_0 = 1/C$. Both the
Schwarzschild-like solution and the solution with the plus sign
in the denominator
develop singularities in the fields at $r=0$. These field
singularities lead to the field energy of these solutions being
infinite, as can be seen by inserting the ansatz functions from
Eq. (\ref{schsol}) into Eq. (\ref{energy}) and trying to
integrate over all space. The investigation
of such infinite energy solutions to the Yang-Mills equations
has been discussed by several authors \cite{rosen} \cite{swank} \cite{proto}
\cite{maha} \cite{lunev} \cite{sing1}, and the earliest discussion
actually pre-dates \cite{rosen} the study of the finite energy solutions
such as the t 'Hooft-Polyakov monopole or the BPS dyon.
Although the infinite field energy of these solutions could
be seen as a drawback as compared to the finite energy solutions,
there are other classical field theory solutions which nevertheless
have a physical importance. The prime example is
the Coulomb solution in electromagnetism which
has a field singularity at $r=0$ that is similar to the $r=0$
singularities of the solutions in Eq. (\ref{schsol}).
The Schwarzschild-like solution, with its singular spherical
surface, has been speculated to have some connection with the
confinement phenomenon for quarks \cite{swank} \cite{maha}
\cite{lunev} \cite{yoshida} \cite{pav}. By studying the motion of a test
particle which moves in the potentials given by the minus
sign solution in Eq. (\ref{schsol}) one finds that the spherical
singularity in the fields represents a barrier which can trap
the test particle inside the sphere. This is similar in spirit
to bag models of hadron structure where one looks at test
particles moving in some confining potential (such as an
infinite spherical well). Also it is interesting that
this Schwarzschild-like solution was arrived at from the
general relativistic solution for a non-rotating black hole,
which exhibits its own type of ``confinement'' : any particle
which passes within the event horizon becomes permanently trapped.
One should be cautious about pushing this analogy too far, since
the nature of the spherical singularity in general relativity and Yang-Mills
theory are different. The singularity at the event horizon of
the general relativistic Schwarzschild solution is not a physical
singularity, but a coordinate singularity as can be seen by
writing the Schwarzschild solution in Kruskal coordinates, where
the only singularity is the one at $r=0$. Both singularities in
the Yang-Mills analog of the Schwarzschild solution are true
singularities in the fields.
The existence of singular solutions for certain field theories is
not new ({\it e.g.} the singularities in the Coulomb
solution of electromagnetism, the Wu-Yang monopole solution
\cite{yang}, or the meron solutions \cite{dea}). Even
the appearance of a singularity in the gauge fields
on a spherical surface, such as occurs in the Schwarzschild-like
solution of Eq. (\ref{schsol}), which may at first appear unique,
can be found in other infinite energy solutions. These other
solutions possess an infinite set of concentric
spherical surfaces on which the fields develop a singularity. This
could be taken as evidence that such spherical surfaces with
singularities are not uncommon features in classical solutions to
the Yang-Mills field equations. The first of these solutions can be
obtained by exchanging the hyperbolic functions of the BPS
solution in Eq. (\ref{soln}) with their trigonometric counterparts
\begin{eqnarray}
\label{soln1}
K(r) &=& cos( \theta ) C r \; csc(Cr) \; \; \; \; \; \; \; \; \; \; \;
\; G(r) = sin ( \theta ) C r \; csc(Cr) \nonumber \\
J(r) &=& sinh(\gamma ) [1 - C r \; cot(C r)] \; \; \; \; \; \; \;
H(r) = cosh (\gamma ) [ 1 - C r \; cot(C r)]
\end{eqnarray}
This solution was briefly discussed by Hsu and Mac \cite{hsu} in
their derivation of the BPS solution ({\it i.e.} Hsu and
Mac start with a solution like that in Eq. (\ref{soln1}) and
apply the transformation $C \rightarrow i C$ to arrive at the
BPS solution). This solution exhibits a series of concentric
spherical surfaces on which the gauge and scalar fields become
singular. These singularities are located on the spherical surfaces
$C r = n \pi $ where $n =1,2,3,4 ...$. Inserting the ansatz
functions of Eq. (\ref{soln1}) in Eq. (\ref{energy}) we find
that the energy density of this solution is
\begin{equation}
\label{ensoln1}
T^{00} = {2 cosh ^2 (\gamma) \over r^2 g^2} \left[ C^2 csc ^2 (C r)
\Big( 1 - C r \; cot (C r)\Big) ^2 + {\Big( C^2 r^2
csc ^2 (C r) - 1 \Big) ^2 \over 2 r^2} \right]
\end{equation}
The energy density becomes singular on the same spherical surfaces
as the gauge and scalar fields. These spherical shells, on which
the energy density becomes infinite, cause the total field
energy of this solution to diverge.
To obtain the next solution we simply try the complementary
trigonometric functions for the solution in Eq. (\ref{soln1}).
Doing this shows that the following is also a solution \cite{sing3}
to the Eq.(\ref{difeq})
\begin{eqnarray}
\label{soln2}
K(r) &=& cos( \theta ) C r \; sec(Cr) \; \; \; \; \; \; \; \; \; \; \; \;
G(r) = sin( \theta ) C r \; sec(Cr) \nonumber \\
J(r) &=& sinh(\gamma ) [1 + C r \; tan(C r)] \; \; \; \; \; \; \;
H(r) = cosh (\gamma ) [ 1 + C r \; tan(C r)]
\end{eqnarray}
It should be noted that due to the linear $C r$ term in each solution,
one can not obtain the solution in Eq. (\ref{soln2}) from
the other trigonometric solution in Eq. (\ref{soln1}) by
simply letting $C r \rightarrow C r - \pi /2$. Although
these two trigonometric solutions are in this sense
distinct ({\it i.e.} they are not simply related by the
transformation $C r \rightarrow C r - \pi / 2$) they
are physically similar since most of the comments concerning
the solution in Eq. (\ref{soln1}) apply here as well.
Most obviously the ansatz functions, and therefore the gauge
and scalar fields, become singular when $C r = n \pi / 2$
where $n = 1, 3, 5, 7, ... \;$ and at $r=0$.
Thus this solution exhibits a series
of concentric spherical surfaces on which its fields become
singular as well as a point singularity at the origin.
These singularities also show up in the energy density
of this solution as they did for the solution in Eq. (\ref{soln1}).
The point singularity at $r=0$ and the spherical
singular surfaces of the solutions in Eqs. (\ref{soln1})
(\ref{soln2}) are similar to that of the
solutions from Eq. (\ref{schsol}). However, the solutions in Eq.
(\ref{schsol}) only possessed one spherical surface
on which the fields and energy density diverged. One conjectured
use for the Schwarzschild-like solution is as a
possible explanation of the confinement mechanism. When the
Schwarzschild-like solution of Ref. \cite{sing1} is treated as a
background potential in which a test particle is placed it is
found that the spherical singularity can act as an impenetrable
barrier which traps the test particle either in the interior
or the exterior of the sphere \cite{yoshida}, giving a classical
type of confinement. Similar results have been found
for other singular solutions \cite{swank} \cite{maha} \cite{lunev}.
In addition Ref. \cite{swank} points out that such a classical
type of confinement is only possible with infinite energy solutions.
Treating the trigonometric solutions as a background potential would also trap
test particles between any two of the concentric spherical singularities.
These trigonometric solutions could possibly be used to solve
the field equations in some limited range of $r$, and then
it could be patched to one of the other solutions which
would solve the field equations for the remaining range of $r$.
This is similar to what is sometimes done in general relativity
where one tries to patch an exterior solution
with some interior solution.
Finally one can obtain a third solution to Eq. (\ref{difeq}) by
applying the transformation $C \rightarrow i C$ to the
solution \cite{sing3} in Eq. (\ref{soln2}). This yields
\begin{eqnarray}
\label{soln3}
K(r) &=& {\bf i} cos(\theta ) C r \; sech(Cr) \; \; \; \; \; \; \; \; \;
\; \; \; G(r) = {\bf i} sin(\theta ) C r \; sech(Cr) \nonumber \\
J(r) &=& sinh(\gamma ) [1 - C r \; tanh(C r)] \; \; \; \; \; \; \;
H(r) = cosh (\gamma ) [ 1 - C r \; tanh(C r)]
\end{eqnarray}
Since the ansatz functions $K(r)$ and $G(r)$ are imaginary, the
space components of the gauge fields will be complex. Despite this
all the physical quantities associated with this complex solution,
such as energy density, are real. Inserting the ansatz functions of Eq.
(\ref{soln3}) into Eq. (\ref{energy}) we find that the field
energy density is
\begin{equation}
\label{ensoln2}
T^{00} = {2 cosh ^2 (\gamma) \over r^2 g^2} \left[ - C^2 sech ^2 (C r)
\Big( 1 - C r \; tanh(C r) \Big) ^2 + {(C^2 r^2 sech ^2 (C r) + 1)^2
\over 2 r^2} \right]
\end{equation}
This energy density is real, but the total field energy
is infinite due to the singularity
at $r = 0$. Thus the above solution is more like a Wu-Yang
monopole \cite{yang} or a charged point particle, as opposed
to a finite energy BPS dyon.
\subsection{SU(2) Solutions With Increasing Potentials}
In addition to the preceding infinite energy solutions which have
gauge and scalar fields that become singular on some spherical surface,
there are other types of infinite energy, general relativistic
analog solutions. In general relativity if one allows
for a nonzero cosmological constant,
$\Lambda$, then the time-time component of the metric tensor for
the Schwarzschild solution becomes \cite{ohanian}
\begin{equation}
\label{cossch}
g_{00} = 1 - {2 G M \over r} - {\Lambda r^2 \over 3}
\end{equation}
The Newtonian potential for this solution is
\begin{equation}
\label{newsol}
\Phi = {( g_{00} - 1) \over 2} = {-G M \over r} - {\Lambda r^2 \over 6}
\end{equation}
Using Eq. (\ref{newsol}) as a starting point one finds the following
simple solution \cite{sing2} to Eq. (\ref{difeq})
\begin{equation}
\label{linear}
K(r) = cos \theta \; \; \; \; \; G(r) = sin \theta \; \; \; \; \;
J(r) = H(r) = { B \over r} + A r^2
\end{equation}
where $a, B$ and $\theta$ are arbitrary constants. If one sets
$A = 0$ then it can be seen the Schwarzschild-like solutions
of Eq. (\ref{schsol}) and those above in Eq. (\ref{linear}) are
of a similar form in the limit $C \rightarrow \infty$
and $e^{\gamma} / C \rightarrow 2 B$. Inserting the ansatz
functions of Eq. (\ref{linear}) into the gauge and scalar fields
of Eq. (\ref{wuyang}) one finds that the time component of the gauge
field ($W_0 ^a$) and the scalar field ($\phi ^a$) behave like
$A r + B/ r^2$. The space part of the gauge fields
($W_i ^a$) have a $1/r$ dependence. This classical
solution exhibits a linear confining potential similar to those
used in some phenomenological studies of hadronic spectra \cite{eich}.
In addition lattice gauge theory arguments \cite{wilson} seem to indicate
that the confining potential between quarks should be linear. Classical
solutions similar to those in Eq. (\ref{linear}) were also discussed
in Ref. \cite{sivers} in connection with the confinement problem.
This solution also has an infinite field energy. Inserting the
ansatz functions of Eq. (\ref{linear}) into the energy density
of Eq. (\ref{energy}) and integrating to get the total field
energy one finds
\begin{eqnarray}
\label{enlin}
E &=& \int T^{00} d^3 x = { 4 \pi \over g^2} \int _{r_a} ^{r_b} T^{00}
r^2 dr \nonumber \\
&=& {4 \pi A^2 \over g^2} (r_b ^3 - r_a ^3) - {8 \pi B^2 \over g^2}
\left( {1 \over r_b^3} - {1 \over r_a ^3} \right)
\end{eqnarray}
where we have introduced an upper ($r_b$) and lower ($r_a$)
cutoff in the radial coordinate. If one lets $r_b \rightarrow
\infty$ then the field energy becomes infinite due to the linear
part of the gauge and scalar fields, while if one lets $r_a
\rightarrow 0$ then the field energy becomes infinite due to
the singularity at $r=0$. Compared to the solutions in Eq.
(\ref{schsol}), which had infinite field energy from local
singularities (either at $r=0$ or $r= 1/C$), the solution
in Eq. (\ref{linear}) can have a infinite field energy from
the point singularity at $r=0$ and/or the linearly increasing gauge
and scalar fields as $r \rightarrow \infty$. Again, although this
classical solution has some undesirable characteristics, it also
exhibits features which are found in some phenomenological
studies of hadronic bound states.
\subsection{SU(3) Solutions}
Up to this point we have discussed classical solutions to
the Yang-Mills field equations for SU(2) fields. Since QCD
involves the SU(3) gauge group it is natural to ask if
there are any SU(3) or even SU(N) generalizations of the
above solutions. One possibility is to embed the above SU(2)
solutions into an SU(N) gauge theory \cite{sing4}. Recently
\cite{dzhunu} a Schwarzschild-like classical solution was found
which is not a simple embedding of the previous SU(2) solutions
into an SU(N) gauge theory, but is a true SU(3) solution. To
arrive at the SU(3) solution one makes the
following generalization \cite{dzhunu} of the Wu-Wu ansatz
\cite{wuwu} \cite{pagel} \cite{volkov}
\begin{eqnarray}
\label{wuwu}
W _0 &=& {- {\bf i} \phi (r) \over g r^2} \left( \lambda ^7 x
-\lambda ^5 y + \lambda ^2 z \right) + {1 \over 2} \lambda ^a
\left( \lambda ^a _{ij} + \lambda ^a _{ji} \right) {x^i x^j \over
g r^3} w(r) \nonumber \\
W^a _i &=& \left( \lambda ^a _{ij} - \lambda ^a _{ji} \right)
{{\bf i} x^j \over g r^2} (1 - f(r)) + \lambda ^a _{jk} \left(
\epsilon _{ilj} x^k + \epsilon_{ilk} x^j \right) {x^l \over g r^3}
v(r)
\end{eqnarray}
where $\lambda ^a$ are the Gell-Mann matrices. Using this ansatz
in the Yang-Mills field equations yields the following set
of coupled differential equations for the functions $f(r) , v(r) ,
\phi (r)$ and $w(r)$
\begin{eqnarray}
\label{difeqsu3}
r^2 f'' &=& f^3 - f + 7 f v^2 + 2vw\phi - f (w^2 + \phi ^2)
\nonumber \\
r^2 v'' &=& v^3 - v + 7v f^2 + 2 f w \phi - v(w^2 + \phi ^2 )
\nonumber \\
r^2 w'' &=& 6 w (f^2 + v^2) - 12 f v \phi \nonumber \\
r^2 \phi '' &=& 2 \phi (f^2 + v^2) - 4 f v w
\end{eqnarray}
where the primes denote differentiation with respect to $r$.
The nonlinear, coupled differential equations in Eq. (\ref{difeqsu3})
are the SU(3) equivalents of the equations in Eq. (\ref{difeq}).
In Ref. \cite{dzhunu} several simplifying assumptions were made
in order to make the problem more tractable. First, taking
$w = \phi = 0$, reduces Eq. (\ref{difeqsu3}) to
\begin{eqnarray}
\label{su3bag}
r^2 f'' &=& f (f^2 - 1 + 7 v^2) \nonumber \\
r^2 v'' &=& v(v^2 - 1 + 7 f^2)
\end{eqnarray}
Then further simplifying by letting $f^2 = v^2 = q^2 / 8$ one
finds that Eq. (\ref{su3bag}) reduces to the Wu-Yang \cite{yang}
equation for $q(r)$
\begin{equation}
\label{su3bag1}
r^2 q'' = q( q^2 - 1)
\end{equation}
This equation has been shown \cite{rosen} \cite{lunev} \cite{dzhunu}
to have a solution which is singular at some radius $r = r_1$. In
other words near $r = r_1$ the solution will be of the form
\begin{equation}
\label{dzhsol}
q(r) \approx {A \over r_1 - r}
\end{equation}
where $A$ and $r_1$ are constant. Thus, even with the scalar field
absent one can find solutions to the pure gauge field theory
equations which will tend to trap test particles behind a spherical
barrier in much the same way as the Schwarzschild-like solution
of Eq. (\ref{schsol}). It is also possible to find closed form
solutions to a special case of the system in Eq. (\ref{difeqsu3})
Taking $v=w=0$ then the equations of Eq. (\ref{difeqsu3}) become
\begin{eqnarray}
\label{su3bag2}
r^2 f'' &=& f(f^2 - \phi ^2 -1) \nonumber \\
r^2 \phi '' &=& 2 \phi f^2
\end{eqnarray}
which has the following simple closed form solution
\begin{eqnarray}
\label{mysu3}
f(r) &=& \mp {C r \over 1 \pm C r} \nonumber \\
\phi (r) &=& \pm {i \over 1 \pm C r}
\end{eqnarray}
Other, similar solutions can be found by making different
simplifying assumptions such as $f=w=0$.
Thus solutions with singular fields on a spherical surface are not
unique to SU(2) gauge theories, but can also be found for SU(3)
\cite{dzhunu} and in general for SU(N) \cite{sing4}. The interesting
aspect of the solutions given by Dzhunushaliev in Ref. \cite{dzhunu}
is that these solutions are true SU(3) solutions rather than
embeddings of the SU(2) solution into the SU(N) gauge group as
in Ref. \cite{sing4}. Also the SU(3) solutions presented here
are pure gauge field solutions, as opposed to the general SU(2)
solutions for the system given in Eq. (\ref{lagrange}), which involves
scalar fields. In some sense the role of the scalar field
of the SU(2) system is taken up by the time component of the gauge
field in the SU(3) system. This can be seen by comparing the system
of equations of Eq. (\ref{difeq}) with the system of equations of
Eq. (\ref{difeqsu3}) : the equations for $f(r), v(r)$ are
similar to those for $K(r), G(r)$ while the equations for
$w(r), \phi (r)$ are similar to those for $J(r), H(r)$.
\section{Electromagnetic Properties of the SU(2) Solutions}
All of the SU(2) solutions to the Yang-Mills field equations
have interesting ``electromagnetic'' features.
To investigate these properties we will use 't Hooft's
definition of a generalized, gauge invariant, U(1) field
strength tensor \cite{thooft}
\begin{equation}
\label{emfst2}
F_{\mu \nu} = \partial _{\mu} (\hat{ \phi} ^a W^a _{\nu}) -
\partial _{\nu} (\hat{ \phi} ^a W^a _{\mu}) - {1 \over g}
\epsilon _{abc} \hat{\phi } ^a ( \partial _{\mu} \hat{\phi } ^b )
( \partial _{\nu} \hat{\phi } ^c )
\end{equation}
where $\hat{\phi } ^a = \phi ^a (\phi ^b \phi ^b)^{-1/2}$. This
generalized U(1) field strength tensor reduces to the usual
expression for the field strength tensor if one performs a
gauge transformation to the Abelian gauge where the scalar
field only points in one direction in isospin space ({\it i.e.}
$\phi ^a = \delta ^{3a} v$) \cite{arafune}. If one associates
this U(1) field with the photon of electromagnetism then the
solutions in Eqs. (\ref{schsol}) (\ref{soln1}) (\ref{soln2})
(\ref{soln3}) (\ref{linear}) carry magnetic and/or electric charges.
In general the electric and magnetic fields associated with these
solutions are
\begin{eqnarray}
\label{ebschw}
E_i &=& F_{i0} = {r_i \over g r} {d \over dr} \Bigg( {J(r)
\over r} \Bigg) \nonumber \\
B_i &=& {1 \over 2} \epsilon _{ijk} F_{jk} = -{r_i \over
g r^3}
\end{eqnarray}
The magnetic field of all the solutions is that of a point monopole
of strength $-4 \pi / g$. The reason for this will be discussed
at the end of this section.
The electric field of the Schwarzschild-like solutions of Eq.
(\ref{schsol}) is easily found by inserting the ansatz function
$J(r)$ from Eq. (\ref{schsol}) into Eq. (\ref{ebschw}). Doing this
gives
\begin{equation}
\label{eb}
E_i = {- r_i sinh \gamma (1 \pm 2 C r) \over g r^3 (1 \pm C r) ^2}
\end{equation}
As $r \rightarrow \infty$ this electric field goes as $1 / r^3$
which indicates that the net electric charge of this solution
is zero, although there appears to be some kind of dipole charge
distribution.
The electric fields of both of the trigonometric solutions presented
in Eqs. (\ref{soln1}) (\ref{soln2}) are similar to each other in
that they indicate that these solutions carry an infinite electric
charge. Taking the ansatz function $J(r)$ from the trigonometric
solution given in Eq. (\ref{soln1}) and inserting it into Eq.
(\ref{ebschw}) yields the following electric field
\begin{equation}
\label{eb2}
E_i = { - sinh (\gamma) r_i \over g r^3}
\big( 1 - C^2 r^2 csc ^2 (C r) \big)
\end{equation}
The electric field does not fall off for large $r$, but
behaves like $r_i csc ^2 (C r) / r$. This electric field
also becomes singular on the spherical surfaces defined by
$C r = n \pi$ where $n = 1,2,3,4 ...$. The trigonometric
solution of Eq. (\ref{soln2}) exhibits the same type of electric
field except that it becomes
singular on the spherical shells given by $C r = n \pi /2$
(with $n = odd$) and at $r=0$.
The electric charge of this solution is also infinite
since the electric field from Eq. (\ref{eb2}) does not
fall off as $r \rightarrow \infty$. For the special case
where $\gamma = 0$ , one finds that the solution carries
no electric charge, but only a magnetic charge. Even in this
case the energy density becomes singular on the
concentric spherical surfaces and at the origin.
Both the BPS solution and the solutions from Eqs. (\ref{soln1})
(\ref{soln2}) have the same finite magnetic strength of
$-4 \pi / g$. Although this solution is a
dyon in the sense that it carries both magnetic and electric charge
it is probably not correct to view it as a particle-like
solution, since the electric field does not fall off, thus
implying that these solutions have an infinite, spread out
electric charge.
The electric field associated with the
complex solution in Eq. (\ref{soln3}) can be found in the same way
as for the other solutions. Inserting the ansatz function $J(r)$
from Eq. (\ref{soln3}) into Eq. (\ref{ebschw}) yields
\begin{equation}
\label{eb3}
E_i = { - sinh (\gamma) r_i \over g r^3}
\Big( C^2 r^2 sech ^2 (C r) + 1 \Big)
\end{equation}
As with all the other solutions, the complex solution carries a magnetic
charge of strength $-4 \pi / g$. In addition, by examining the behaviour
of the electric field in Eq. (\ref{eb3}) as $r \rightarrow \infty$
one finds that this complex solution carries an electric charge of
$-4 \pi sinh (\gamma )/ g$, which is the same as that carried by the BPS
solution. One interesting feature
of the solution from Eq. (\ref{soln3}) is that
even though the space components of the gauge fields are complex
all the physical quantities ({\it e.g.} field energy, magnetic
charge, electric charge) calculated from it are real. Also, unlike the
solutions of Eqs. (\ref{soln1}) (\ref{soln2}), this complex
solution can be viewed as a point-like dyon since it has a localized
electric charge. The main difference between this solution and
the BPS solution is the infinite field energy of the complex
solution due to the field singularity at $r=0$.
While many of the physical characteristics of these various
solutions are substantially different in each case, the
magnetic charge of all the solutions is the same.
This comes about since the magnetic charge of each solution is
a topological charge which carries the same value for each field
configuration. A topological current, $k _{\mu}$, can be
defined as \cite{arafune}
\begin{equation}
\label{current}
k_{\mu} = {1 \over 8 \pi} \epsilon_{\mu \nu \alpha \beta}
\epsilon _{abc} \partial ^{\nu} {\hat \phi} ^a \partial ^{\alpha}
{\hat \phi} ^b \partial ^{\beta} {\hat \phi} ^c
\end{equation}
The topological charge of this field configuration is then
\begin{eqnarray}
\label{tc}
q &=& \int k_0 d^3 x = {1 \over 8 \pi} \int
(\epsilon_{ijk} \epsilon_{abc}
\partial ^{i} {\hat \phi} ^a \partial ^{j} {\hat \phi} ^b
\partial ^{k} {\hat \phi} ^c ) d^3 x \nonumber \\
&=& {1 \over 8 \pi} \int \epsilon_{ijk} \epsilon_{abc}
\partial ^{i} ( {\hat \phi} ^a \partial ^{j} {\hat \phi} ^b
\partial ^{k} {\hat \phi} ^c ) d^3 x
\end{eqnarray}
For all the solutions one finds that ${\hat \phi } ^a = r^a / r$,
which is the same regardless of the ansatz function $H(r)$. In all
cases we find that the topological charge is $q =1$. In the next
section when we examine the motion of a test particle in the background
field of the Schwarzschild-like solution we will find that there
is a field angular momentum due to the interaction of the test
particle with the field configuration of the Schwarzschild-like
solution. This field angular momentum can be seen to arise from the
interaction of the topological magnetic charge with the charge
of the test particle, in much the same way that the configuration of
a normal magnetic charge and an electric charge lead to a field
angular momentum \cite{thomson} \cite{saha} \cite{jackson}.
\section{Motion of Tests Particles in Schwarzschild-like Potential}
We would now like to study the motion of a test particle in the
background potential of the Schwarzschild-like solution of Eq.
(\ref{schsol}). We will make several assumptions in doing this.
First we will take our test particle to be a scalar particle
as in Ref. \cite{lunev} \cite{yoshida}. One reason for making this
choice is to illustrate the spin from isospin \cite{rebbi} effect
that occurs with these solutions. As discussed in the preceding section
all of these solutions carry a magnetic charge. Many researchers have
remarked on the fact \cite{thomson} \cite{saha} that the composite
system of a magnetic charge and electric charge carry an angular
momentum due to the configuration of electric and magnetic fields.
Even when the magnetic charge is topological in character, as is the
case with 't Hooft-Polyakov monopoles, one finds \cite{rebbi} a similar
effect whereby the composite system of a topological magnetic charge
and a particle with the charge of the gauge group, will carry an
angular momentum in the gauge fields. This has the interesting
consequence that if one wants to construct fermionic objects from
the singular solutions one should use scalar particle which are
in the fundamental representation of the gauge group -- SU(2) for the
solutions considered here. (Fermionic test particles in the adjoint
representation would also give a net fermionic bound state \cite{pav}).
Our second assumption is that the test particle will be coupled to the
scalar field part of the solution of Eq. (\ref{schsol}) via the
following substitution $m^2 \rightarrow (m + \lambda \sigma ^a \phi ^a
/ 2)^2$ where $\lambda$ is an arbitrary coupling constant. Our final
assumption is to set $\theta = 0$ in Eq. (\ref{schsol}) in order to
not have to take the ansatz function $G(r)$ into account. In this way
the scalar particle $\Phi ^A$ moving in the background field of the
Schwarzshild-like solution given in Eq. (\ref{schsol}) becomes
\begin{eqnarray}
\label{kg1}
& &\left( \partial _0 - {i g \over 2} \sigma ^a W_0 ^a \right)
\left( \partial _0 - {i g \over 2} \sigma ^a W_0 ^a \right)^A _B
\Phi ^B (x ,t) - \left( \partial _i - {i g \over 2} \sigma ^a W^a _i \right)
\left( \partial _i - {i g \over 2} \sigma ^a W^a _i \right) ^A _B
\Phi ^B (x , t) \nonumber \\
&=& - \left( m + {\lambda \over 2} \sigma ^a \phi^a
\right) ^2 \Phi ^A (x , t)
\end{eqnarray}
Where $\sigma ^a$ are the standard Pauli matrices, and $A, B$ are SU(2)
group indices which can take on the values $1$ or $2$.
Taking $\Phi ^A (x,t) = \Phi ^A (x) e^{-i E t}$, using $(\sigma ^a v^a) ^2 =
v^a v^a$ and expanding we find that Eq. (\ref{kg1}) becomes
\begin{eqnarray}
\label{kg2}
& &\left[- E^2 -g \sigma ^a W_0 ^a E - {g^2 \over 4} (W_0^a)^2
- \nabla ^2 + i g \sigma ^a W_i^a \partial _i + {g^2 \over 4}
(W_i ^a)^2 \right] ^A _B \Phi ^B (x) \nonumber \\
= &-& \left[m^2 + \lambda m \sigma ^a \phi ^a
+ {\lambda ^2 \over 4} (\phi ^a)^2 \right] ^A _B \Phi^B (x)
\end{eqnarray}
Inserting the ansatz form of the gauge and scalar fields from Eq.
(\ref{wuyang}) into Eq. (\ref{kg2}) then yields
\begin{eqnarray}
\label{kg3}
& & - \left[\nabla ^2 - {(1-K(r) ) \over r^2} \sigma ^a l^a -
{(1-K(r))^2 \over 2 r^2} + {\sigma ^a r^a \over r^2} E J(r)
+ {J(r) ^2 \over 4 r^2} \right] ^A _B \Phi ^B (x) \nonumber \\
= & & \left( E^2 - m^2 - {\lambda m \over g r^2} \sigma ^a r ^a
H(r) - {\lambda ^2 \over 4 g^2 r^2} H(r) ^2 \right) ^A _B \Phi ^B (x)
\end{eqnarray}
where we have used $i g \sigma ^a W_i ^a \partial _i = -i (1 - K(r))
\sigma ^a \epsilon _{aji} r^j \partial ^i / r^2 = (1-K(r)) \sigma ^a
l^a / r^2$ ($l^a$ is the standard orbital angular momentum
operator), and $(W^a _i)^2 = \epsilon_{aij} \epsilon_{aik} r^j r^k
(1-K(r))^2 / g^2 r^4 = 2 (1-K(r))^2 / g^2 r^2$. Since $\sigma ^a
l^a$ does not commute with $\sigma ^a r^a$ Eq. (\ref{kg3}) is
difficult to handle. By taking advantage of the free parameter
$\gamma$ which occurs in the ansatz functions $J(r), H(r)$ one can chose
$\gamma$ such that $E sinh \gamma = \lambda m cosh \gamma / g$. With
this choice the two $\sigma ^a r^a$ terms in Eq. (\ref{kg3})
cancel one another. In order to handle the $\sigma ^a l^a$ term
it is necessary to define the total angular momentum operator as
\begin{equation}
\label{jtot}
J^a = l^a + S^a = l^a + {1 \over 2} \sigma ^a
\end{equation}
Thus the total angular momentum comes not only from the orbital
angular momentum, but has a contribution that looks like a spin
angular momentum. The $\sigma$ matrices in the last term of
Eq. (\ref{jtot}) are, however, connected with the isospin of the
system rather than with spin. This is just the spin from isospin
effect \cite{rebbi}, and is connected with the fact that the
Schwarzschild-like solution of Eq. (\ref{schsol}) carries a topological
magnetic charge. Thus even though our system involves only integer
spin fields ({\it i.e.} $W_{\mu} ^a , \phi ^a , \Phi ^A$) the
combined system is a spin 1/2 object. Using Eq. (\ref{jtot})
the $\sigma ^a l^a$ term can be expanded in the usual way as $\sigma ^a l^a
= 2 S^a l^a = J_{op} ^2 -l _{op} ^2 -S _{op} ^2$ except
now $S _{op}$ is the isospin operator rather than the spin operator.
Finally, we make the simplifying assumption that $\lambda = g$ so that
the $J(r)^2$ and $H(r) ^2$ terms may be more easily combined. At this
point there seems to be nothing special in this choice, but we
will see that according to the arguments of Ref. \cite{lunev} and
\cite{exner}, the barrier at $r = 1/C$ will only absolutely confine
a particle if $\lambda \ge g$ while for $\lambda < g$ there will
be some probability for the test particle to tunnel through the
barrier. Combining all the preceding assumptions we find
\begin{eqnarray}
\label{kg4}
& & - \left[ \nabla ^2 - {(1-K(r)) \over r^2}(J_{op}^2 -l_{op}^2-S_{op}^2) -
{(1-K(r))^2 \over 2 r^2}+ {1 \over 4 r^2}(J(r)^2 - H(r)^2)
\right] ^A _B \Phi ^B (x) \nonumber \\
= & & \left( E^2 - m^2 \right) ^A _B \Phi ^B (x)
\end{eqnarray}
In order to separate out the radial equation from Eq. (\ref{kg4})
we take
\begin{equation}
\label{sep}
\Phi ^A (x) = {1 \over r} f_{Jl} (r) Y^A _{JlM} (\theta , \phi )
\end{equation}
where the $Y^A _{JlM}$ are the standard spinor
spherical harmonics that one gets from adding an orbital angular
momenta $l^a$ to a spin $1/2$. Here spin is replaced by isospin, but
the math, and the spinor spherical harmonics, are exactly the same.
Now inserting Eq. (\ref{sep}) into Eq. (\ref{kg4}) yields
\begin{eqnarray}
\label{kg5}
& &-\left[ {d^2 \over dr^2} - {D \over r^2} -{F(1-K(r)) \over r^2}
- {(1-K(r))^2 \over 2 r^2} + {1 \over 4 r^2}
(J (r)^2 - H (r)^2) \right] f_{Jl} (r) \nonumber \\
& & = (E^2 - m^2) f_{Jl} (r)
\end{eqnarray}
where we have defined the constants $D= l(l+1)$ , $F = J(J+1)-l(l+1)-3/4$.
Then setting $x = C r$ and inserting the ansatz functions
$K(r), J(r), H(r)$ from Eq. (\ref{schsol}) into Eq. (\ref{kg5})
turns the problem into an effective one-dimensional Schr{\"o}dinger
equation
\begin{equation}
\label{kg6}
\left[- {d^2 \over dx^2} + { D \over x^2} + {F (1-2x) \over x^2
(1-x)} + {(1-2x)^2 \over 2 x^2 (1-x)^2} + {1 \over 4 x^2 (1-x)^2}
\right] f_{Jl} (x) = {(E^2 - m^2) \over C^2} f_{Jl} (x)
\end{equation}
where all the non-derivative terms on the left hand side are the
effective potential. The key feature of this effective
potential are the singularities at $x=0$ and $x=1$. Now as
$x \rightarrow 1$ the leading term in the effective potential goes like
\begin{equation}
\label{effv}
V_{eff} (x) = {D \over x^2} + {F (1-2x) \over x^2 (1-x)}
+{(1-2x)^2 \over 2 x^2 (1-x)^2} +{1 \over 4 x^2 (1-x)^2}
\rightarrow {3 \over 4 (1-x)^2}
\end{equation}
It was argued in Refs. \cite{exner} \cite{lunev} that such
a singularity would only present a true barrier to the test
particle ({\it i.e.} the probability of the test particle tunneling
through the barrier would be zero) if the coefficient in Eq.
(\ref{effv}) were greater than or equal to $3/4$. Thus the effective
potential of Eq. (\ref{kg6}) just confines the test particle to remain
in the range $0\le x \le 1$. The fact that the effective potential
is just able to confine the test particle stems from our
choice of $\lambda = g$ for the coupling of the scalar potential
$\phi ^a$ to the test particle $\Phi ^A$. If we had taken $\lambda
< g$ then the coefficient in the limiting form of the effective
potential from Eq. (\ref{effv}) would have been less than $3/4$
and the test particle would no longer be confined ({\it e.g.} if
one took $\lambda =0$ it is straightforward, starting from Eq.
(\ref{kg3}), to show that one gets a coefficient of $1/2$). Conversely,
when $\lambda > g$ then the coefficient in Eq. (\ref{effv}) becomes
greater than $3/4$ and the test particle becomes confined. This has
the interesting implication that the scalar potential plays an
important role in this confinement mechanism. Although, generally
confinement is thought to be just the result of the gauge interaction,
there are phenomenological studies \cite{goebel} \cite{tekuda}
\cite{ram} which indicate
that an effective scalar potential is involved in the confinement
mechanism.
To get more detailed in the solution of Eq. (\ref{kg6}) one must
pick particular values of $J$ and $l$ (which determine the constants
$D$ and $F$ in Eq. (\ref{kg6})), and solve for the eigenfunctions,
$f_{Jl} (x)$ and eigenvalues $(E^2 -m^2) /C^2$. In general this must be done
numerically \cite{pav} \cite{yoshida}, however, the key features of the
effective one-dimensional potential of Eq. (\ref{effv}) ({\it i.e.}
the singularities in the potential at $x=0$ and $x= 1$) make this
potential similar to the P{\"o}schl-Teller potential \cite{flugge}.
\begin{equation}
\label{postel}
V(x) = {1\over 2} V_0 \left[ {\alpha (\alpha -1) \over sin ^2 (\pi x /
2)} + {\beta (\beta -1) \over cos ^2 (\pi x / 2)} \right]
\end{equation}
where $\alpha , \beta , V_0$ are constants. By choosing $\alpha , \beta$
and $V_0$ correctly the P{\"o}schl-Teller potential can be made similar to
the effective potential from Eq. (\ref{effv}). Then the known
eigenfunctions and eigenvalues of the P{\"o}schl-Teller potential
should give a good approximation to the eigenvalues and eigenfunctions
of the potential from Eq. (\ref{effv}). The eigenfunctions for the
P{\"o}schl-Teller potential are \cite{flugge}
\begin{equation}
\label{ef}
f_n (x) = K sin ^{\alpha} (\pi x /2) cos ^{\beta} (\pi x / 2)
\; \; \; _2 F _1 \left( -n , \alpha + \beta +n , \alpha +{1 \over 2} ;
sin ^2 (\pi x / 2) \right)
\end{equation}
where $K$ is a constant fixed by normalization, $n$ is the radial
quantum number which takes on values of $n=0, 1, 2, 3 ...$,
and $_2 F _1 (a,b,c ;x)$ is the hypergeometric function. The
eigenenergies for the P{\"o}schl-Teller potential are \cite{flugge}
\begin{equation}
\label{ee}
E_n = {1 \over 2} V_0 (\alpha + \beta + 2n)^2
\end{equation}
From the shape of both the P{\"o}schl-Teller potential and the
effective potential in Eq. (\ref{effv}) this is exactly the kind of
dependence one would expect for the energy eigenvalues. For small
energies ({\it i.e.} $\alpha + \beta > 2n$) both potentials behave like
a harmonic oscillator potential and so one would expect that the
leading term in $E_n$ should go like $2 V_0 (\alpha + \beta) n \propto
n$. For large energies ({\it i.e.} $2n > \alpha + \beta$) both
potentials behave like infinite spherical wells, and so one
would expect that the leading term in $E_n$ should go like
$2 V_0 n^2 \propto n^2$.
As a simple example we will consider the $l=0$ case for the
potential in Eq. (\ref{effv}). For $l=0$ we find $J= 1/2$, $D=0$
$F=0$ and the potential in Eq. (\ref{effv}) becomes
\begin{equation}
\label{effv1}
V_{eff} (x) = {3 - 8x + 8x^2 \over 4 x^2 (1-x)^2}
\end{equation}
This potential approaches $3/ (4 (1-x)^2)$ as $x \rightarrow 1$ so
the test particle is just confined to the range $0<x<1$. In this
range $V_{eff} (x)$ of Eq. (\ref{effv1}) reaches its minimum value
of $4$ at $x = 1/2$, and the potential is symmetric about this point.
In order for the P{\"o}schl-Teller potential
to also be symmetric about $x=1/2$, and to also take
a value of $4$ at this point we can choose $V_0 = 1$ and
$\alpha = \beta = 2$. Now inserting these into Eq. (\ref{ee}) and
remembering that our eigenvalue from Eq. (\ref{kg6}) is $(E^2 -m^2)
/C^2$ we find that the approximate energy of the bound states for
this case with $l=0$ is
\begin{equation}
\label{ee1}
E^2 _n = m^2 + C^2 (2 + n)^2
\end{equation}
Note that this energy depends on the arbitrary constant $C$, which
sets the radius of the confining sphere ($r= 1/C$). As $C$ increases
the radius of the spherical shell decreases and from Eq. (\ref{ee1})
the energy of the state increases as would be expected.
Although in this $l=0$ case it was particularly easy to determine
$V_0, \alpha , \beta$, the form of the bound state energy given
by Eq. (\ref{ee1}) will be similar even when $l \ne 0$.
\section{Discussion and Conclusions}
In this article we have presented a variety of solutions to
the field equations of Yang-Mills theory. Although finding exact
solutions to non-linear field theories is in general difficult,
many of the present solutions were found by using the mathematical
connection which exists between Yang-Mills theory and general
relativity. Since general relativity has been studied for a longer
time than Yang-Mills theory there exists a body of known solutions
which can serve as guides for finding solutions to the Yang-Mills
or Yang-Mills-Higgs field equations. The Schwarzschild solution
of general relativity, both without and with a cosmological term,
gave rise to the solution with a spherical singularity in
Eq. (\ref{schsol}) and the linearly increasing solution of
Eq. (\ref{linear}). Although both of these solutions suffered
from the apparent drawback of having an infinite field energy,
they also exhibited some possible connection with the confinement
phenomenon. The linear solution of Eq. (\ref{linear}) is of the
form of phenomenological potentials \cite{eich} that are often
used in studies of heavy quark bound states. In addition lattice
gauge theory arguments \cite{wilson} favour a linear type of
confining potential. The Schwarzschild-like solution of Eq.
(\ref{schsol}) has some similarities to bag models for
quark bound states. Spherical singularities, similar to those
of the Schwarzschild-like solution, were also found to occur in
several other solutions as given in Eqs. (\ref{soln1}) (\ref{soln2}).
Actually, the solutions given in Eqs. (\ref{soln1}) (\ref{soln2})
possessed an infinite set of concentric spheres on which the gauge
and scalar fields became infinite. Thus, such spherically singular
surfaces may not be uncommon features of Yang-Mills field
theories. The SU(2) Schwarzschild-like solution can easily be generalized
to SU(N) \cite{sing4} by simply embedding the SU(2)
solutions into an SU(N) gauge theory. It has also recently been found
\cite{dzhunu} that true SU(3) solutions, which are not simply
embeddings of the SU(2) solutions, can be given.
In the previous section the behaviour of a scalar test
particle placed inside the background potential presented by
the Schwarzschild-like solution was examined.
In order for the Schwarzschild-like potential to confine the test
particle, $\Phi ^A$, that it was necessary to couple, $\Phi ^A$,
to the scalar part of the Schwarzschild-like solution, $\phi ^a$,
via the coupling $m^2 \rightarrow (m + \lambda \sigma ^a \phi ^a /2)^2 $,
where $\lambda$ is the strength of the coupling between $\Phi ^A$
and $\phi ^a$. Even with this coupling it was found that confinement
occurred when $\lambda \ge g$, while for $\lambda < g$ there would be
some finite probability for $\Phi ^A$ to tunnel out of the spherical
well. Although normally it is thought that the confinement
phenomenon is the result of only gauge interactions, there has
been some work \cite{goebel} \cite {tekuda} \cite{ram} which indicates
that an effective scalar interaction may be needed to completely
explain confinement. Another interesting aspect of the bound
state system studied in the previous section is that the total
system was a fermion even though only integer spin fields were involved.
The spin 1/2 nature of the bound state system resulted from the fact
that the isospin 1/2 of the test particle $\Phi ^A$ was converted
into spin 1/2 when it was placed inside the Schwarzschild-like
solution. Another way of arriving at this result is to note that almost
all of the solutions presented here could be shown to carry a
topological magnetic charge. Thus, in the same way that a standard magnetic
charge - electric charge system carries a field angular momentum
of $1/2$ in their combined electromagnetic fields, so too the
combined charges of the Schwarzscild-like solution and $\Phi ^A$
carried a field angular momentum of $1/2$ in their combined
non-Abelian fields. If a realistic model of hadronic bound
states can be constructed from these classical field theory solutions,
then the fact that the net angular momentum of these states does not
come entirely from the constituent particles, may offer a possible
explanation of the EMC effect \cite{ashman}, which shows that a large
part of the net spin of the proton does not come from the valence
quarks.
In addition to the Schwarzschild-like solutions presented
here it is also possible to take more complex solutions
from general relativity to find other Yang-Mills solutions.
In Ref. \cite{sing6} the general relativistic Kerr solution
was used to construct a new Yang-Mills solution. Although the
final form of the Yang-Mills Kerr-like solution was not as
simple as the Schwarzschild-like solutions, it did share the
common feature of having confining surfaces on which the
fields became singular. Finally it is also possible to use
this method for finding solutions to non-linear field
equations in reverse : starting from known solutions to
the Yang-Mills equations one can obtain solutions to the
general relativistic field equations \cite{sing7}.
\section{Dedication} This article is dedicated to the memory
of Professor Fyodor Lunev.
|
\section{ New mechanism of frustration near orbital degeneracy }
\label{sec:orbitals}
Quite generally, strongly correlated electron systems involve orbitally
degenerate states,\cite{Ima98} such as $3d$ ($4d$) states in transition metal
compounds, and $4f$ ($5f$) states in rare-earth compounds. Yet, the orbital
degrees of freedom are ignored in most situations and the common approach
is to consider a single correlated orbital per atom
which leads to spin degeneracy alone. Indeed, most of the current studies of
strongly correlated electrons deal with models of nondegenerate orbitals. The
problems discussed recently include mechanisms of ferromagnetism in the
Hubbard model,\cite{ferro} hole propagation and quasiparticles in the $t-J$
model,\cite{Dag94} and magnetic states of the Kondo lattice.\cite{Man97}
Of course, in many actually existing compounds the orbital degeneracy is
removed by the crystal field, and a single-orbital approach is valid {\it per
se\/}. Also, from a fundamental point of view it is often possible to argue
that orbital degeneracy is qualitatively irrelevant, and that a single-orbital
approach can capture the generic mechanisms operative in the presence of
strong correlations.
However, neither of these arguments applies for a class of insulating
strongly correlated transition metal compounds, where the crystal field
leaves the $3d$ orbitals explicitly degenerate and thus the type of occupied
orbitals is not known {\it a priori\/}, while the magnetic interaction
between the spins of neighboring transition metal ions depends on which
orbitals are occupied. In this particular class of Mott-Hubbard insulators
(MHI) the orbital degrees of
freedom acquire a separate existence in much the same way as the spins do.
Thereby, the degeneracy of $t_{2g}$ orbitals is of less importance, as the
magnetic superexchange and the coupling to the lattice are rather weak.
A more interesting situation occurs when $e_g$ orbitals are partly occupied,
which results in stronger magnetic interactions, and strong
Jahn-Teller (JT) effect. Typical examples of such ions are:
Cu$^{2+}$ ($d^9$ configuration, one hole in $e_g$-orbitals), low-spin
Ni$^{3+}$ ($d^7$ configuration, one electron in $e_g$-orbitals), as well as
Mn$^{3+}$ and Cr$^{2+}$ ions (high-spin $d^4$ configuration, one $e_g$
electron). The simplest model, relevant for $d^9$ transition metal ions,
which is also the subject of the present paper, was introduced
by Kugel and Khomskii more than two decades ago,\cite{Kug73} but its
mean-field (MF) phase diagram was analyzed only recently.\cite{Fei97}
It describes magnetic superexchange interactions between spins $S=1/2$,
and the accompanying orbital superexchange interactions.
One might argue that the (classical) orbital degeneracy is not easy to
realize in such systems, as the electron-phonon coupling will lead to the
conventional collective JT instability. In fact, it can be shown that the JT
instability is enhanced by the orbital pattern once this has been established
as the result of effective interactions:\cite{Kug73,Kho97,crete} the lattice
has to react to the symmetry lowering in the orbital sector, which can only
increase the stability of a given magnetic state. So the lattice follows
rather than induces the orbital order, and therefore, as was pointed out in
the early work by Kugel and Khomskii,\cite{Kug73,Kug82} in the orbitally
degenerate MHI one has to consider in first instance the purely electronic
problem. This is supported by the results of recent band structure
calculations using the local density approximation (LDA) with the electron
interactions treated in Hartree-Fock approximation, the so-called LDA+U
method, which permits both orbitals and spins to polarize while keeping the
accurate treatment of the electron-lattice coupling of LDA intact.
These calculations reproduce the observed orbital ordering in
KCuF$_3$ (Ref. \onlinecite{Lie95}) and in LaMnO$_3$ (Ref. \onlinecite{Ani97}),
even when the lattice distortions are suppressed, while allowing the lattice
to relax only yields an energy gain which is minute in comparison with the
energies involved in the orbital ordering.
Effects of orbital degeneracy are expected as soon as crystal-field
splittings become small. Such situations are frequently encountered in
rare-earth systems, where they lead to the so-called singlet-triplet models
discussed in the seventies,\cite{Hsi72} while in the $3d$ oxides only a small
number of so-called Kugel-Khomskii (KK) systems \cite{Kug82} have been
recognized that actually exhibit orbital effects.\cite{Kho97}
As pointed out by Kugel and Khomskii,\cite{Kug73} in such situations the
superexchange interactions have a more complex form than in spin-only models
and one expects that also in some other Mott-Hubbard (or charge-transfer)
insulators new magnetic phases might arise due to the competition of various
magnetic and orbital interactions. Some examples of such a competition of
magnetic interactions are encountered in the heavy fermion
systems,\cite{Man97,Cox87} and in the manganites where the phase diagrams
show a particular frustration of magnetic
interactions.\cite{Miz95,Ish96,Shi97,Fei99}
Even more interesting behavior is expected for the doped systems, as the
competition between the magnetic, orbital, and kinetic energy is then
described by $t-J$ Hamiltonians of a novel type, which exhibit qualitatively
different excitation spectra due to the underlying orbital
degeneracy.\cite{Zaa93}
A few examples of such models have already been discussed in the literature,
such as the triplet $t-J$ model,\cite{Zaa92} the low-spin defects in a $S=1$
background,\cite{Dag96} or a new $t-J$ model for the manganites.\cite{Mul96}
Whether such models are realistic enough is not yet clear, as for example in
the manganites there are experimental\cite{Oki97} and theoretical \cite{Mil95}
indications that the double-exchange model which includes only the spin
degrees of freedom is insufficient to understand the transport properties
under doping. Recent work \cite{Shi97,Fei99,Tak98,Bri99} strongly suggests
that an extension of the $t-J$ and double-exchange models which include fully
the orbital physics should be studied instead.
In this paper we shall consider only the insulating situation, where one can
integrate out the $d-d$ excitations and derive an effective low-energy
Hamiltonian. This approach is justified by the large on-site Coulomb
interaction $U$, being the largest energy scale in MHI.
A low-energy Hilbert space splits off, spanned by {\it spin and orbital\/}
configuration space, with superexchangelike couplings between both spin
and orbital local degrees of freedom. The orbital sector carries a discrete
symmetry and the net outcome is that the clock-like orbital degrees of
freedom get coupled into the $SU(2)$ spin problem. The resulting low-energy
Hamiltonian is called a {\it spin-orbital model\/}. Here we focus
on the simplest situation with two nearly degenerate partially filled $e_g$
orbitals, and completely filled $t_{2g}$ orbitals, as encountered in KCuF$_3$
and related systems.\cite{Kug82} These are JT-distorted cubic crystals,
three-dimensional (3D) analogues of the cuprate superconductors.\cite{Web88}
In the high-$T_c$ cuprates, orbital degeneracy would occur if the Cu-O bonds
which involve apical oxygens were squeezed such
as to recover the cubic symmetry of the perovskite lattice. Of course, such
a degeneracy of $e_g$ orbitals is far from being realized in the
actual high-$T_c$ materials, and in their parent compounds.\cite{Kho91,Gra92}
If only one correlated orbital is present, the system
may be described by the effective single-band Hubbard model (typically with
more extended hopping), as in the cuprate superconductors.\cite{Fei96}
In this simplest case the effective model at half-filling is the Heisenberg
model with antiferromagnetic (AF) superexchange.\cite{And59} This changes
when more than one $3d$ orbital is partly occupied. For example, we show in
Sec. \ref{sec:model} that virtual excitations involving $d^8$ local triplet
states become possible in the case of degenerate $e_g$ orbitals, and this
leads to additional ferromagnetic (FM) interactions. The origin of these new
interactions was first discussed by Kugel and Khomskii\cite{Kug73} and by
Cyrot and Lyon-Caen\cite{Cyr75} who pointed out that the strongest
superexchange constant results from the excitation to the lowest energy
triplet state in the degenerate Hubbard model.
The model proposed by Kugel and Khomskii explains qualitatively the observed
magnetic ordering in KCuF$_3$ as being due to an orbital ordering which gives
planes of perpendicularly oriented orbitals, and the magnetic coupling becomes
then FM according to the Goodenough-Kanamori rules.\cite{Goo63} As mentioned
above, such a state was indeed found in the band structure calculations of
Liechtenstein {\it et al.}\cite{Lie95} using the LDA+U method. An analogous
orbital order is responsible for ferromagnetism in the planar FM insulator
K$_2$CuF$_4$.\cite{Kug82} In the colossal magnetoresistance parent compound
LaMnO$_3$, where the $e_g$ orbitals contain one electron instead of one hole,
a similar orbital ordering occurs,\cite{Kho97,Ish96} although the situation
there is more complex due to the presence of $t_{2g}$ spins, so that the
resulting superexchange is not between spins $S=1/2$ but between total spins
$S=2$.\cite{Fei99} Another example of degenerate orbitals is found in
V$_2$O$_3$, with the orbital ordering studied by Castellani, Natoli and
Ranninger in a series of papers.\cite{Cas78} In fact, their prediction that
the transition into the AF insulator is accompanied by the onset of orbital
ordering was experimentally verified only recently.\cite{Bao97} However,
this case is still open, as recent electronic structure calculations suggest
that doubly degenerate orbitals are occupied by two electrons in the
high-spin state and the orbital degree of freedom plays no role.\cite{Ezh99}
In any of the above situations the orbital ordering breaks the translational
symmetry and represents an analogon of spin antiferromagnetism in orbital
space. So, {\em classically\/} orbital ordering is expected to occur quite
generally whenever one encounters $e_g$ orbitals containing either one hole
or one electron, with important consequences for the magnetism. This
immediately raises a number of questions about what happens in the
{\em quantum regime\/}. Will orbital long-range order (LRO) be robust or
will it give way to an {\em orbital liquid\/}, as proposed by Ishihara,
Yamanaka and Nagaosa (Ref. \onlinecite{Nag97})? In either case, what are the
consequences of the enlarged phase space and the associated additional
channels for quantum fluctuations for the magnetism: can magnetic LRO
survive or will it be replaced by a {\em spin liquid\/}?
Quantum disordered phases are of great current interest. Spin disorder is
well known to occur in one-dimensional (1D) and quasi 1D quantum spin
systems, and the best example is the 1D Heisenberg model, where the famous
exact solution found by Bethe many years ago \cite{Bet31} showed that the
quantum fluctuations prevent true AF LRO, giving instead a slow decay of
spin correlations. A similar situation is encountered in spin ladders with
an even number of legs, which have a spin gap and purely short-range
magnetic order.\cite{Whi94,Ric96} This is one of the realizations of a
spin-liquid ground state due to purely short-range spin correlations. In
the limit of a two-dimensional (2D) Heisenberg model the spin disorder is
replaced by a ground state with AF LRO.
It is well known that frustrated magnetic interactions may lead to spin
disordered states in two dimensions. However, in order to achieve this, i.e.,
to prevent 2D macroscopic spin systems from behaving classically and to make
quantum mechanics take over instead, the frustration of the interactions
must be sufficiently severe. This shows that global SU(2) by itself is not
symmetric enough to defeat classical order in $D>1$ and one has to change
the magnetic interactions in such a way that they lead to sufficiently
strong quantum fluctuations.
So far, this strategy has been shown to lead to spin disorder in (quasi) 2D
systems in three different situations:
(i) Frustrating a 2D square lattice by adding longer-range AF interactions,
as in $J_1-J_2$ and $J_1-J_2-J_3$ models, gives a high degeneracy of
the classical sector, and a disordered state is found for particular
values of the magnetic interactions.\cite{Cha88,Chu91} This mechanism
involves fine-tuning of parameters and therefore such systems are
hard to realize in nature.
(ii) In the bilayer Heisenberg model two planes are coupled by interlayer AF
superexchange $J_{\perp}$ which generates zero-dimensional fluctuations.
This leads to a crossover to the disordered ground state of an
incompressible spin liquid above a certain critical value of
$J_{\perp}$.\cite{Mil94,Chu95} Also this mechanism is hard to
realize experimentally.
(iii) In contrast, a spin disordered state can be obtained in nature by
reducing the number of magnetic bonds in a 2D square lattice. The
model of CaV$_4$O$_9$ studied by Taniguchi {\it et al.}\cite{Tan95} is
a 1/5 depleted square lattice, which gives a plaquette resonating
valence bond (PRVB) ground state for realistic interactions, and
a spin gap which agrees with experimental observations.\cite{Ued96}
A common feature of these systems is a crossover between different magnetic
ground states, either between two different patterns of LRO, as in case (i),
or simply between the ordered and disordered states, which results in all
three situations in a tendency towards the formation of spin singlets on
the bonds with the strongest AF superexchange. One may further note that in
these spin-only models very specific patterns of magnetic interactions are
required already in two dimensions to prevent the system to order classically,
while up to now it has proven impossible to realize a spin liquid in three
dimensions.
In the present paper we address two fundamental questions for the Heisenberg
antiferromagnet (HAF) extended to include the orbital degrees of freedom in
orbitally degenerate MHI:
(i) Which {\em classical\/} states with magnetic LRO do exist in the
neighborhood of orbital degeneracy?
(ii) Are those forms of classical order
always stable against {\em quantum\/} fluctuations?
We will show that the orbitally degenerate MHI represent a class of
systems in which spin disorder occurs due to frustration of {\em spin and
orbital\/} superexchange couplings.
This frustration mechanism is different from that operative in pure spin
systems, and suppresses the magnetic LRO in the ground state {\em even in
three dimensions\/}.
As explained above, the low-energy behavior of such systems is described by
a spin-orbital model. We will show that within the framework of such a
spin-orbital model the occurrence of spin disorder may be regarded
as resulting from a competition between various classical ordered phases,
each one with a simultaneous symmetry breaking
in spin and orbital space. The qualitatively new aspect is that the magnetic
interactions follow the orbital pattern, and thus these systems tend to
"self-tune" to (critical) points of high classical degeneneracy.
We show explicitly that in the vicinity of such a multicritical point
classical order is highly unstable with respect to quantum fluctuations.
As a result, a qualitatively new quantum spin liquid with strong orbital
correlations is expected. We believe that a 3D state of this type is realized
in LiNiO$_2$.
The paper is organized as follows. The spin-orbital model for $d^9$
transition metal ions, such as Cu$^{2+}$ in KCuF$_3$, is derived in Sec.
\ref{sec:model} using the correct multiplet structure of Cu$^{3+}$ excited
configurations. We solve this model first in the MF approximation and
present the resulting classical phases and the accompanying orbital
orderings in Sec. \ref{sec:mfa}. The elementary excitations obtained within
an extension of the linear spin-wave theory (LSW) are presented in Sec.
\ref{sec:magnons}, where we demonstrate that two transverse modes are
strongly coupled to each other. This leads to soft modes next to the
classical transition lines, and to the collapse of LRO due to diverging
quantum corrections, as shown in Sec. \ref{sec:rpa}. We summarize the
results and present our conclusions in Sec. \ref{sec:spinliquid}.
\section{ The spin-orbital model }
\label{sec:model}
Our aim is to construct the effective low-energy Hamiltonian for a 3D
perovskite-like lattice. The original charge-transfer multiband model, as
considered for instance for the cuprates, includes the hybridization elements
between the $3d$ orbitals of transition metal ions and the $2p$ orbitals
of oxygen ions.\cite{Fei96}
If the Coulomb elements at the $3d$ orbitals and the charge-transfer energy
between the $3d$ and $2p$ orbitals are large, this model can be transformed
into an effective spin-fermion model. For example, this transformation
performed for the three-band model gives an effective Hamiltonian with
localized spins at the Cu sites which interact by superexchange
interactions, while the doped carriers interact with them by a Kondo-like
coupling.\cite{Zaa88} In the limit of undoped compounds, one is thus left
with a model which describes interacting transition metal ions.
The simplest form of (superexchange) interaction, namely a spin model, is
obtained for the case of nondegenerate $d$ orbitals, whereas orbital
degeneracy gives a spin-orbital model acting in a larger Hilbert space
defined by both spin and orbital degrees of freedom at each transition
metal site. Having in mind the strongly correlated late transition
metal oxides, we consider specifically the case of one hole per unit cell in
the $3d^9$ configuration, characterized in the absence of JT-distortion by
two degenerate $e_g$ orbitals: $x^2-y^2\sim |x\rangle$ and
$(3z^2-r^2)/\sqrt{3}\sim |z\rangle$. The derivation is, however, more
general and applies as well to the low-spin $d^7$ configuration; in the case
of the early transition metal oxides the $d^1$ case would involve the
$t_{2g}$ orbitals instead.
The holes in the undoped compound which corresponds to the $d^9$
configuration of transition metal ions, as in La$_2$CuO$_4$ or KCuF$_3$, are
fairly localized.\cite{notecov} Hence, we take as a starting point the
following Hamiltonian which describes $d$-holes on transition metal ions,
\begin{equation}
\label{hband}
H_{e_g} = H_{kin} + H_{int} + H_z,
\end{equation}
and consider the kinetic energy, $H_{kin}$, and the electron-electron
interactions, $H_{int}$, within the subspace of the $e_g$ orbitals (the
$t_{2g}$ orbitals are filled by electrons, do not couple to $e_g$ orbitals
due to the hoppings via oxygens, and hence can be neglected). The last term
$H_z$ describes the crystal-field splitting of the $e_g$ orbitals.
Due to the shape of the two $e_g$ orbitals $|x\rangle$ and $|z\rangle$,
their $d-p$ hybridization in the three cubic directions is unequal, and
is different between them, so that the effective hopping elements are
direction dependent and different for $|x\rangle$ and $|z\rangle$.
The only nonvanishing hopping in the $c$-direction connects two $|z\rangle$
orbitals, while the elements in the $(a,b)$ planes fulfill the Slater-Koster
relations,\cite{Sla54} as presented before by two of us.\cite{Zaa93} Taking
the hopping $t$ along the $c$-axis as a unit, the kinetic energy is given by,
\begin{eqnarray}
\label{hkin}
H_{kin}&=&{t\over 4}\sum_{\langle ij\rangle{\parallel}}
\left[ 3 d^{\dagger}_{ix\sigma}d^{}_{ix\sigma}
+ (-1)^{{\vec \delta}\cdot {\vec y}}\sqrt{3}
(d^{\dagger}_{iz\sigma}d^{}_{ix\sigma}\right. \nonumber \\
&+&\left. H.c.)+d^{\dagger}_{iz\sigma}d^{}_{iz\sigma}\right]
+t\sqrt{\beta} \sum_{\langle ij\rangle{\perp}}
d^{\dagger}_{iz\sigma}d^{}_{iz\sigma},
\end{eqnarray}
where $\langle ij\rangle{\parallel}$ and $\langle ij\rangle{\perp}$ stand
for the bonds between nearest neighbors within the $(a,b)$-planes, and
along the
$c$-axis, respectively, and $\beta=1$ in a cubic system. The $x-z$ hopping
in the $(a,b)$ planes depends on the phases of the $x^2-y^2$ orbitals along
$a$- and $b$-axis, respectively, included in the factors
$(-1)^{{\vec \delta}\cdot {\vec y}}$ in Eq. (\ref{hkin}).
The electron-electron interactions are described by the on-site terms,
\begin{eqnarray}
\label{hint}
&H_{int}&=(U+\case{1}{2}J_H)
\sum_{i\alpha}n_{i\alpha \uparrow}n_{i\alpha\downarrow}
+ (U-J_H)\sum_{i\sigma}n_{ix\sigma}n_{iz\sigma}
\nonumber \\
&+&(U-\case{1}{2}J_H)\sum_{i\sigma}n_{ix\sigma}n_{iz\bar{\sigma}}
-\case{1}{2}J_H \sum_{i\sigma}d^{\dagger}_{ix \sigma}d^{}_{ix\bar{\sigma}}
d^{\dagger}_{iz\bar{\sigma}}d^{}_{iz \sigma}
\nonumber \\
&+&\case{1}{2}J_H\sum_{i}
( d^{\dagger}_{ix \uparrow}d^{\dagger}_{ix\downarrow}
d^{ }_{iz\downarrow}d^{ }_{iz \uparrow}
+ d^{\dagger}_{iz \uparrow}d^{\dagger}_{iz\downarrow}
d^{ }_{ix\downarrow}d^{ }_{ix \uparrow} ),
\end{eqnarray}
with $U$ and $J_H$ standing for the Coulomb and Hund's rule exchange
interaction,\cite{noteuj} respectively, and $\alpha=x,z$. For convenience,
we used the simplified notation $\bar{\sigma}=-\sigma$. This Hamiltonian
describes correctly the multiplet structure of $d^8$ (and $d^2$)
ions,\cite{Gri71} and is rotationally invariant in the orbital
space.\cite{Ole83} The wave functions have been assumed to be real which
gives the same element $J_H/2$ for the exchange interaction and for the
pair hopping term between the $e_g$ orbitals, $|x\rangle$ and $|z\rangle$.
In fact, we adopted here the most natural units for the elements of the
Coulomb interaction, with the energy of the central $|^1E\rangle$ doublet
being equal to $U$. By definition this energy does not depend on the Hund's
exchange element $J_H$, as we show below, and is thus the measure of the
average excitation energy in the
$d^9_id^9_j\rightarrow d^{10}_id^8_j$ transition.
The interaction element $J_H$ stands for
the singlet-triplet splitting in the $d^8$ spectrum (Fig. \ref{virtual}) and
is just twice as big as the exchange element $K_{xz}$ used usually in quantum
chemistry.\cite{Gra92} The typical energies for the Coulomb and exchange
elements can be found using constrained-occupation local-density functional
theory.\cite{Hyb89} Unfortunately, such calculations have been performed
only for a few compounds so far. For La$_2$CuO$_4$, a parent compound of
superconducting cuprates, one finds $U=7.77$ eV and $J_H=2.38$ eV;\cite{Gra92}
other estimations of $U$ based on the experimental data report values $6<U<8$
eV for cuprates and nickelates.\cite{Zaa90} This results in the ratio
$J_H/U\simeq 0.3$ which we take as a representative value for the strongly
correlated late transition metal oxides. The values of intersite hopping $t$,
being an effective parameter, are more difficult to estimate. As a
representative value for La$_2$CuO$_4$ one might take $t\approx 0.65$ eV,
which results in the superexchange interaction between the $|x\rangle$
orbitals in $(a,b)$ planes, $J_{(a,b)}=(9/4)t^2/U\simeq 0.13$ eV,\cite{Esk93}
in good agreement with the experimental value.\cite{expJ} Similar values of
the effective $t$ are expected also in the other transition metal oxides, and
thus we can safely assume that at the filling of one hole per ion the
ionic Hamiltonian (\ref{hband}) describes an insulating state, and
that the effective magnetic interactions can be derived in the strongly
correlated regime of $t\ll U$.
The last term in Eq. (\ref{hband}) stands for the crystal field which lifts
the degeneracy of the two $e_g$ orbitals and breaks the symmetry in
the orbital space,
\begin{equation}
\label{hz}
H_{z}=\sum_{i\sigma}(\varepsilon_xn_{ix\sigma}+\varepsilon_zn_{iz\sigma}),
\end{equation}
if $\varepsilon_x\neq \varepsilon_z$. It acts as a magnetic field in the
orbital space, and together with the parameter $\beta$ in
$H_{kin}$ (\ref{hkin}) quantifies the deviation in the electronic structure
from the ideal cubic local point group.
In the atomic limit, i.e., at $t=0$ and $E_z=0$, one has orbital degeneracy
next to spin degeneracy. This gives four basis states per site, as each hole
may occupy either orbital, $|x\rangle$ or $|z\rangle$, and either spin state,
$\sigma=\uparrow$ or $\sigma=\downarrow$. The system of $N$ $d^9$ ions thus
has a large degeneracy $4^N$, which is, however, removed by the effective
interactions between each pair of nearest neighbor ions $\{i,j\}$ which
originate from virtual transitions to the excited states,
$d^9_id^9_j\rightleftharpoons d^{10}_id^8_j$, due to hole hopping. Hence,
we derive the effective spin-orbital model following Kugel and
Khomskii,\cite{Kug73} starting from the Hamiltonian in the atomic limit,
$H_{at}=H_{int}+H_z$, and treating $H_{kin}$ as a perturbation. However,
in the present study we include the {\em full multiplet structure\/} of the
excited states within the $d^8$ configuration which gives corrections of the
order of $J_H$ compared with the earlier results of Refs.
\onlinecite{Kug73,Kug82}.
Knowing the multiplet structure of the $d^8$ intermediate states, the
derivation of the effective Hamiltonian can be done in various ways. The
most straightforward but lengthy procedure is a generalization of the
canonical transformation method used before for the Hubbard\cite{Cha77} and
the three-band\cite{Zaa88} model. A significantly shorter derivation is
possible, however, using the cubic symmetry and starting with the
interactions along the $c$-axis. Here the derivation simplifies tremendously
as one finds only effective interactions which result from the hopping of
holes between the directional $|z\rangle$ orbitals, as shown in Fig.
\ref{virtual}. Next the interactions in the remaining directions can be
generated by the appropriate rotations to the other cubic axes $a$ and $b$,
and applying the symmetry rules for the hopping elements between the
$e_g$ orbitals.\cite{Sla54} The derivation of the spin-orbital model
is given in more detail in Appendix \ref{sec:derivation}.
Depending on whether the initial state is $|z\rangle_i|x\rangle_j$ or
$|z\rangle_i|z\rangle_j$, the intermediate $d_i^{10}d_j^8$ configuration
resulting from the hole-hop $|z\rangle_i \rightarrow |z\rangle_j$, involves
on the $d^8$ site either the interorbital states, the triplet $^3A_2$ and
the singlet $^1E_{\theta}$, or the two singlets built from the states with
doubly occupied orbitals, $^1E_{\varepsilon}$ and $^1A_1$. Of course, the
spins have to be opposite in the latter case, while in the former case also
parallel spin configurations contribute in the triplet channel. Apart from
a constant term, this atomic problem is equivalent to that of the $d^2$
configuration, and thus one might consider instead the spectrum of $d^2$
ions. The eigenstates within the $e_g$ subspace are:
(i) triplet $|^3A_2\rangle$,
(ii) interorbital singlet $|^1E_{\epsilon}\rangle$, and
(iii) bonding and antibonding singlets, $|^1E_{\theta}\rangle$ and
$|^1A_1\rangle$, with double occupancies of both orbitals, where
bonding/antibonding refers to pair hopping term $\propto J_H$ between
$|x\rangle$ and $|z\rangle$ orbital.
The energies of the states $|^3A_2\rangle$ and $|^1E_{\epsilon}\rangle$ are
straighforwardly obtained using ${\vec S}_{ix}\cdot {\vec S}_{iz}=+1/4$ and
${\vec S}_{ix}\cdot {\vec S}_{iz}=-3/4$,
for $S=1$ and $S=0$ states, respectively. The remaining two singlet
energies are found by diagonalizing a $2\times 2$ problem in the subspace
of doubly occupied states. Hence, the resulting spectrum is,\cite{noted8}
\begin{eqnarray}
\label{specd8}
E( ^3A_2 ) &=& U - J_H, \nonumber \\
E( ^1E_{\epsilon} ) &=& U , \nonumber \\
E( ^1E_{\theta} ) &=& U + \case{1}{2}J_H
- \case{1}{2}J_H\left[ 1-(E_z/J_H)^2 \right]^{1/2}, \nonumber \\
E( ^1A_1 ) &=& U + \case{1}{2}J_H
+ \case{1}{2}J_H\left[ 1-(E_z/J_H)^2 \right]^{1/2},
\end{eqnarray}
where $E_z=\varepsilon_x-\varepsilon_z$. At $E_z=0$ it consists of
equidistant states, with a distance of $J_H$ between the triplet
$|^3A_2\rangle$ and the degenerate singlets $|^1E_{\theta}\rangle$ and
$|^1E_{\epsilon}\rangle$ (which form of course an orbital doublet),
as well as between the above singlets and the top singlet $|^1A_1\rangle$.
We emphasize that the simplified Hubbard-like form of electron-electron
interactions (\ref{hint}) which uses two parameters, $U$ and $J_H$, in this
case is an {\em exact representation\/} of the Coulomb interaction in
the $t_{2g}^6e_g^2$ configuration as obtained in the theory of
multiplet spectra, and one finds a one-to-one correspondence between the
energies calculated above, and those found with the Racah parameters $A$,
$B$, and $C$,\cite{Gri71}
\begin{eqnarray}
\label{racah}
E( ^3A_2 ) &=& A - 8B , \nonumber \\
E( ^1E ) &=& A + 2C, \nonumber \\
E( ^1A_1 ) &=& A + 8B + 4C.
\end{eqnarray}
Thus, the parameters used by us are $U=A+2C$ and $J_H=8B+2C$.\cite{noteuj}
We normalize the energies by the Coulomb interaction $U$, and introduce
\begin{equation}
\label{jh}
\eta\equiv J_H/U
\end{equation}
as an energy unit for the Hund's rule exchange interaction. This gives the
excitation energies which correspond to the {\em local excitations\/}
$d^9_id^9_j\rightarrow d^{10}_id^8_j$ on a given bond $(ij)$,
\begin{eqnarray}
\label{ourd8}
\varepsilon( ^3A_2 ) &=& 1 - \eta, \nonumber \\
\varepsilon( ^1E_{\epsilon} ) &=& 1 , \nonumber \\
\varepsilon( ^1E_{\theta} ) &=& 1 + \case{1}{2}\eta
- \case{1}{2}\eta\left[ 1-(E_z/J_H)^2 \right]^{1/2},
\nonumber \\
\varepsilon( ^1A_1 ) &=& 1 + \case{1}{2}\eta
+ \case{1}{2}\eta\left[ 1-(E_z/J_H)^2 \right]^{1/2},
\end{eqnarray}
shown in Fig. \ref{msd8}. We note that the deviation from the equidistant
spectrum at $E_z=0$ becomes significant only for $|E_z|/J_H>1$. Taking the
realistic parameters of the cuprates,\cite{Gra92} one finds for La$_2$CuO$_4$
with $E_z=0.64$ eV that $E_z/J_H\simeq 0.54$, a value representative for
systems that are already far from orbital degeneracy. Since we are
interested here in what happens close to orbital degeneracy, this
allows us to neglect the $E_z$ dependence of the energies of the excited
$d^8$ states, and use the atomic spectrum (\ref{racah}) in the derivation
presented in Appendix \ref{sec:derivation}.
Following the above procedure, we have derived the effective Hamiltonian
${\cal H}$ in spin-orbital space,
\begin{equation}
\label{somcu}
{\cal H} = {\cal H}_J + {\cal H}_{\tau},
\end{equation}
where the superexchange part ${\cal H}_J$ can be most generally written
as follows (a simplified form was discussed recently in Ref.
\onlinecite{Fei97}),
\begin{eqnarray}
{\cal H}_J&=&\sum_{\langle ij\rangle}\left\{
- \frac{t^2}{\varepsilon(^3A_2)}
\left(\vec{S}_i\cdot\vec{S}_j+\frac{3}{4}\right)
{\cal P}_{\langle ij\rangle}^{\zeta\xi}\right. \nonumber \\
&+&\left. \frac{t^2}{\varepsilon(^1E_\epsilon)}
\left(\vec{S}_i\cdot\vec{S}_j-\frac{1}{4}\right)
{\cal P}_{\langle ij\rangle}^{\zeta\xi}\right. \nonumber \\
&+&\left. \left[\frac{t^2}{\varepsilon(^1E_\theta)}
+\frac{t^2}{\varepsilon(^1A_1)}\right]
\left(\vec{S}_i\cdot\vec{S}_j-\frac{1}{4}\right)
{\cal P}_{\langle ij\rangle}^{\zeta\zeta}\right\}.
\label{somj}
\end{eqnarray}
Here $\vec{S}_i$ refers to a spin $S=1/2$ at site $i$, and
${\cal P}_{\langle ij\rangle}^{\alpha\beta}$ are projection operators
on the orbital states for each bond,
\begin{eqnarray}
\label{porbit}
{\cal P}_{\langle ij\rangle}^{\zeta\xi}&=&
(\case{1}{2}+\tau^c_i)(\case{1}{2}-\tau^c_j)+
(\case{1}{2}-\tau^c_i)(\case{1}{2}+\tau^c_j), \nonumber \\
{\cal P}_{\langle ij\rangle}^{\zeta\zeta}&=&
2(\case{1}{2}-\tau^c_i)(\case{1}{2}-\tau^c_j).
\end{eqnarray}
They are either parallel ($P_{i\zeta}=\frac{1}{2}-\tau^c_i$) to the
direction of the bond $\langle ij\rangle$ on site $i$, and perpendicular
($P_{j\xi}=\frac{1}{2}+\tau^c_j$) on the other site $j$, or parallel on both
sites, respectively, and are constructed with the following
orbital operators associated with the three cubic axes ($a$, $b$, $c$),
\begin{eqnarray}
\label{orbop}
\tau^{a}_i & = & -\case{1}{4}(\sigma^z_i - \sqrt{3}\sigma^x_i ),
\nonumber \\
\tau^{b}_i & = & -\case{1}{4}(\sigma^z_i + \sqrt{3}\sigma^x_i ),
\nonumber \\
\tau^c_i & = & \case{1}{2} \sigma^z_i.
\end{eqnarray}
The $\sigma$'s are Pauli matrices acting on the orbital pseudo-spins
$|x\rangle ={\scriptsize\left( \begin{array}{c} 1\\ 0\end{array}\right)},\;
|z\rangle ={\scriptsize\left( \begin{array}{c} 0\\ 1\end{array}\right)}$.
Hence, we find a Heisenberg Hamiltonian for the spins, coupled into an
orbital problem. While the spin problem is described by the continuous
symmetry group $SU(2)$, the orbital problem is clock-model-like, i.e., there
are three directional orbitals: $3x^2-r^2$, $3y^2-r^2$, and $3z^2-r^2$, but
they are not independent. The orbital basis consists of one directional
orbital and its orthogonal counterpart, and we have chosen here
$|z\rangle\equiv 3z^2-r^2$ and $|x\rangle\equiv x^2-y^2$ orbitals.
In general, the energies of these two orbital states, $|x\rangle$ and
$|z\rangle$, are different, and thus the complete effective Hamiltonian of
the $d^9$ model (\ref{somcu})
includes as well the crystal-field term (\ref{hz}) which we write as
\begin{equation}
\label{somez}
{\cal H}_{\tau} = - E_z \sum_i \tau^c_i.
\end{equation}
Here $E_z$ is a crystal field which acts as a "magnetic field" for the
orbital pseudospins, and is loosely associated with an uniaxial pressure
along the $c$-axis. The $d^9$ spin-orbital model (\ref{somcu}) depends thus
on two parameters: (i) the crystal field splitting $E_z$, and (ii) the
Hund's rule exchange $J_H$.
While the first two terms in (\ref{somj}) cancel for the magnetic
interactions in the limit of $\eta\to 0$, the last term favors AF spin
orientation. Although the form (\ref{somj}) might in principle be used for
further analysis, we prefer to make an expansion of the excitation energies
$\varepsilon_n$ in the denominators of Eq. (\ref{somj}) in terms of $J_H$,
and use $\eta=J_H/U$ (\ref{jh}) as a parameter which quantifies the Hund's
rule exchange. This results in the following form of the effective exchange
Hamiltonian in the $d^9$ model (\ref{somcu}),\cite{Fei97,notebugs}
\begin{eqnarray}
\label{somjexp}
{\cal H}_J\! &\simeq& \! J\sum_{\langle ij\rangle} \left[
2\left( {\vec S}_i\cdot{\vec S}_j -\frac{1}{4} \right)
{\cal P}_{\langle ij\rangle}^{\zeta\zeta}
-{\cal P}_{\langle ij\rangle}^{\zeta\xi }\right] \nonumber \\
&-&\! J\eta\sum_{\langle ij\rangle} \left[ {\vec S}_i\cdot{\vec S}_j
\left(\! {\cal P}_{\langle ij\rangle}^{\zeta\zeta}
\! +\! {\cal P}_{\langle ij\rangle}^{\zeta\xi }\!\right)
+\frac{3}{4}{\cal P}_{\langle ij\rangle}^{\zeta\xi }
-\frac{1}{4}{\cal P}_{\langle ij\rangle}^{\zeta\zeta}\right].\nonumber \\
\end{eqnarray}
The first term in Eq. (\ref{somjexp}) describes the AF superexchange
$\propto J=t^2/U$ (where $t$ is the hopping between $|z\rangle$ orbitals
along the $c$-axis), and is obtained when the splittings between different
excited $d^8$ states $\sim J_H$ (Fig. \ref{msd8}) are neglected. As we show
below, in spite of the AF superexchange $\propto J$, {\it no {\rm LRO} can
stabilize in a system described by the spin-orbital model (\ref{somcu}) in
the limit $\eta\to 0$ at orbital degeneracy $(E_z=0)$} because of
the presence of the frustrating orbital interactions which gives a highly
degenerate classical ground state. We emphasize that even in the limit of
$J_H\to 0$ the present Kugel-Khomskii model {\em does not obey\/}
SU(4) symmetry, essentially because of the directionality of the $e_g$
orbitals. Therefore, such an idealized SU(4)-symmetric model\cite{Fri99}
does not correspond to the realistic situation of degenerate $e_g$ orbitals
and is expected to give different answers concerning the interplay of spin
and orbital ordering in cubic crystals.
Taking into account the multiplet splittings, we obtain [second line of
(\ref{somjexp})] again a Heisenberg-like Hamiltonian for the spins coupled
into an orbital problem, with a reduced interaction $\propto J\eta$. It is
evident that the new terms support FM rather than AF spin interactions for
particular orbital orderings. This net FM superexchange originates from the
virtual transitions which involve the triplet state $|^3A_2\rangle$, which
has the lowest energy and thus gives the strongest effective coupling.
We remark in passing that the FM channel is additionally
enhanced for $d^4$ ions when the virtual excitations to double
occupancies in $e_g$ orbitals happen in the presence of partly filled
$t_{2g}$ orbitals, as realized in the manganites.\cite{Shi97,Fei99}
The important feature of the spin-orbital model (\ref{somcu}) is that the
{\it actual magnetic interactions depend on the orbital pattern\/}. This
follows essentially from the hopping matrix elements in $H_{kin}$
(\ref{hkin}) being different between a pair of $|x\rangle$ orbitals, between
a pair of different orbitals (one $|x\rangle$ and one $|z\rangle$ orbital),
and between a pair of $|z\rangle$ orbitals, respectively, and depending on
the bond direction either in the $(a,b)$ planes, or along the
$c$-axis.\cite{Zaa93} We show in Sec. \ref{sec:mfa} that this leads to a
particular competition between magnetic and orbital interactions, and the
resulting phase diagram contains a rather large number of classical
phases, stabilized for different values of $E_z$ and $J_H$.
\section{ Mean-field phase diagram }
\label{sec:mfa}
\subsection{ Anisotropy of antiferromagnetic interactions }
We start the analysis of the $d^9$ spin-orbital (or Kugel-Khomskii) model
(\ref{somcu})--(\ref{somjexp}) by analyzing the MF solution obtained by
replacing the scalar products $\vec{S}_i\cdot\vec{S}_j$ by the Ising term,
$S^z_iS^z_j$. The MF Hamiltonian may be written for the more general
situation where the interaction has uniaxial
anisotropy along the $c$-direction in the 3D lattice as follows,
\begin{eqnarray}
\label{somcumf}
{\cal H}_{\rm MF}\! &\simeq&\! \sum_{\langle ij\rangle} J_{\alpha}\left[
2\left( S_i^zS_j^z -\case{1}{4} \right)
{\cal P}_{\langle ij\rangle}^{\zeta\zeta}
-{\cal P}_{\langle ij\rangle}^{\zeta\xi }\right] \nonumber \\
&-&\! \eta\sum_{\langle ij\rangle} J_{\alpha}\left[ S_i^zS_j^z
\left(\!{\cal P}_{\langle ij\rangle}^{\zeta\zeta}
+{\cal P}_{\langle ij\rangle}^{\zeta\xi }\!\right)
+\case{3}{4}{\cal P}_{\langle ij\rangle}^{\zeta\xi }
-\case{1}{4}{\cal P}_{\langle ij\rangle}^{\zeta\zeta}\right]
\nonumber \\
&-& E_z \sum_i \tau^c_i,
\end{eqnarray}
where $J_a=J_b=J$, and $J_c=J\beta$. For $\beta>1$ the nearest-neighbor bonds
$\langle ij\rangle$ $\parallel c$ are shorter, while for $\beta<1$ these
bonds are longer than the bonds within the $(a,b)$ planes. In the limit of
$\beta\to 0$ the bonds along the $c$-axis may be neglected and the model
reduces to a 2D model, representative for the magnetic interactions between
Cu ions within the CuO$_2$ planes of the high-temperature superconductors.
The presence of AF spin interactions $\propto J$ suggests magnetic
superstructures with staggered magnetization, and we considered several
possibilities, with two- and four-sublattice 3D structures, giving rise to
G-AF and A-AF phases, AF 1D chains
coupled ferromagnetically, and others. The MF Hamiltonian contains
as well an AF {\em interaction between orbital variables\/},
$\sim J\tau^{\alpha}_i\tau^{\alpha}_j$, which suggests that it might be
energetically more favorable to alternate the orbitals in a certain regime of
parameters, and pay thereby part of the magnetic energy. This illustrates the
essence of the {\em frustration\/} of the magnetic interactions present
in the spin-orbital model (\ref{somcu}), as discussed in Sec.
\ref{sec:orbitals}. Therefore, for any classical state the orbitals occupied
by the holes have to be optimized, and we allowed mixed orbitals (MO),
\begin{equation}
\label{mixing}
|i\mu\sigma\rangle=\cos\theta_i|iz\sigma\rangle+\sin\theta_i|ix\sigma\rangle,
\end{equation}
with the values of the mixing angles $\{\theta_i\}$ being variational
parameters to be found from the minimization of the classical energy.
The superexchange in (\ref{somcumf}) depends strongly on the orbital state.
At large positive $E_z$, where the crystal field strongly favors
$|x\rangle$-occupancy over $|z\rangle$-occupancy, one expects that
$\theta_i=\pi/2$ in Eq. (\ref{mixing}), and the holes occupy $|x\rangle$
orbitals on every site. In this case the spins do not interact in the
$c$-direction (see Fig. \ref{virtual}), and there is also no orbital energy
contribution. Hence, the $(a,b)$ planes will decouple magnetically, while
within each plane the superexchange is AF and equal to $9J/4$ along $a$ and
$b$. These interactions stabilize a 2D antiferromagnet, called further AFxx.
The resulting 2D N\'eel state with decoupled $(a,b)$ planes along the
$c$-direction is the well-known classical ground state of the high-$T_c$
superconductors, La$_2$CuO$_4$ and YBa$_2$Cu$_3$O$_6$.\cite{Kam94}
In contrast, if $E_z<0$ and $|E_z|$ is large, $|E_z|/J\gg 1$, then
$\theta_i=0$ in Eq. (\ref{mixing}), and the holes occupy $|z\rangle$
orbitals. The spin system has then strongly anisotropic AF superexchange,
being $4J$ between two $|z\rangle$ orbitals along the $c$-axis, and $J/4$
between two $|z\rangle$ orbitals in the $(a,b)$ planes, respectively. The
corresponding 3D N\'eel state with holes occupying $|z\rangle$ orbitals is
called AFzz. The spin and orbital order in both AF phases is shown
schematically within the $(a,b)$ planes in Fig. \ref{allmfa}.
\subsection{ Antiferromagnetic states in the 3D model }
Assuming an AF classical order in all three directions, the so-called G-AF
state, it is thus obvious that for large $|E_z|$ one finds either the AFxx
or the AFzz phase, depending on whether $E_z>0$ or $E_z<0$, with the
following energies normalized per one site,
\begin{eqnarray}
\label{mfaf}
E_{\rm AFxx}&=&-3J\left(1-{\eta\over 4}\right)-{1\over 2}E_z, \nonumber \\
E_{\rm AFzz}&=&- J\left(1+{\eta\over 4}\right)
-2J\beta\left(1-{\eta\over 2}\right)+ {1\over 2}E_z.
\end{eqnarray}
The AFxx and AFzz phases are degenerate in a 3D system ($\beta=1$) along
the line $E_z=0$, while decreasing $\beta$ moves the degeneracy to
negative values of $E_z$, namely to $E_z=-2J(1-\beta)(1-{\eta \over 2})$.
However, for intermediate values of $|E_z|$ one should allow for mixed
orbitals. Following the argument above about the AF nature of the orbital
interaction, we assume alternating orbitals at two sublattices, A and B.
The alternation should allow the orbitals to compromise between being
identical (optimizing the magnetic energy) and being orthogonal (optimizing
the orbital energy). This is realized by choosing in Eq. (\ref{mixing}) the
angles alternating between the sublattices: $\theta_i=+\theta$ for $i\in A$,
and $\theta_j=-\theta$ for $j\in B$, respectively,
\begin{eqnarray}
\label{orbmoffa}
|i\mu\sigma\rangle&=&\cos\theta|iz\sigma\rangle+\sin\theta|ix\sigma\rangle,
\nonumber \\
|j\mu\sigma\rangle&=&\cos\theta|jz\sigma\rangle-\sin\theta|jx\sigma\rangle.
\end{eqnarray}
The calculation of the energy can be performed either by evaluating the
average values of the operator variables $\{\tau_i^{\alpha}\}$, or by taking
the average values of the orbital projection operators $\{P_{i\alpha}\}$ as
given in Eq. (\ref{fullij}). Using the two-sublattice orbital ordering
(\ref{orbmoffa}), one finds for the bonds $\langle ij\rangle\parallel (a,b)$,
\begin{eqnarray}
\label{orbavab}
\langle P_{i\xi}P_{j\zeta}+P_{i\zeta}P_{j\xi}\rangle&=&
\case{1}{8}(7-4\cos^2 2\theta),
\nonumber \\
\langle 2P_{i\zeta}P_{j\zeta}\rangle&=&
\case{1}{8}(1-2\cos 2\theta)^2,
\end{eqnarray}
and for the bonds $\langle ij\rangle\parallel c$,
\begin{eqnarray}
\label{orbavc}
\langle P_{ix}P_{jz}+P_{iz}P_{jx}\rangle&=&
\case{1}{2}(1- \cos^2 2\theta),
\nonumber \\
\langle 2P_{iz}P_{jz}\rangle&=&
\case{1}{2}(1+\cos 2\theta)^2.
\end{eqnarray}
The classical energy per site as a function of $\theta$ is then given by
\begin{eqnarray}
\label{enemoaaat}
E(\theta)&=&-\frac{J}{4}(1+\frac{\eta}{2})(7-4\cos^2 2\theta)
\nonumber\\
&-&\frac{J}{4} (1-\frac{\eta}{2})(1-2\cos 2\theta)^2
\nonumber\\
&-&\frac{J}{2}\beta(1+\frac{\eta}{2})(1-\cos^2 2\theta)
\nonumber\\
&-&\frac{J}{2}\beta(1-\frac{\eta}{2})(1+\cos 2\theta)^2
\nonumber\\
&+&\frac{1}{2}E_z\cos 2\theta.
\end{eqnarray}
This has a minimum at
\begin{equation}
\label{orbmoaaa}
\cos 2\theta=-{{(1-\frac{\eta}{2})(1-\beta)+\frac{1}{2}\varepsilon_z\over
(2+\beta)\eta}},
\end{equation}
where $\varepsilon_z=E_z/J$, if $\eta\neq 0$, and provided that
$|\cos 2\theta|\le 1$ (a similar condition applies to all the other states
with MO considered below). So, as long as
$2J(\beta-1)-3J(\beta+1)\eta \le E_z \le 2J(\beta-1)+J(5+\beta)\eta$,
there is genuine MO order, while upon reaching the smaller (larger)
boundary value for $E_z$, the orbitals go over smoothly into
$|z\rangle$ ($|x\rangle$), i.e. one retrieves the AFzz (AFxx) phase.
Taking the magnetic ordering in the three cubic directions [$abc$]
as a label to classify the classical phases with MO (\ref{orbmoffa}),
we call the phase obtained in the regime of genuine MO order
MO{\scriptsize AAA}, with classical energy given by
\begin{eqnarray}
\label{enemoaaa}
E_{\rm MO{\scriptsize AAA}}&=&-\left(2+\beta+\frac{3}{4}\eta\right)J
\nonumber \\
&-&J {\left[ (2-\eta)(1-\beta)+\varepsilon_z \right]^2\over
4(2+\beta)\eta}.
\end{eqnarray}
Upon increasing $J_H$, the FM interactions occur which increase the energy
of the AF phases in three dimensions by the term $\frac{3}{4}\eta$ per site
in Eqs. (\ref{mfaf}) (a similar increase of energy occurs also in the
MO{\scriptsize AAA} phase in the region of its existence). This indicates
frustration of magnetic interactions and opens a potential possibility that
other classical phases with FM order along particular directions might be
more stable. We have found a few classical phases when the spins order
ferromagnetically either in particular planes, or along one spatial
direction, and this magnetic order coexists with MO occupied by holes.
For example, the angles in Eq. (\ref{mixing}) can be chosen in such a way
that at least one of the orbitals on two neighboring sites is perpendicular
to the bond direction, e.g. is like $y^2-z^2$ type for a bond along the
$a$-axis. In such a case, the AF superexchange vanishes, and one finds
instead a weaker FM interaction, in agreement with the Goodenough-Kanamori
rules.\cite{Goo63} By this mechanism Kugel and Khomskii\cite{Kug73} proposed
an alternating orbital order to explain the FM planes observed in KCuF$_3$.
Following this argument, let us assume FM order within $(a,b)$ planes, and
the same form (\ref{orbmoffa}) as above for the alternating orbitals at the
two sublattices, $A$ and $B$.
As alternating orbitals can only be arranged to be perpendicular to the
bonds in at most two spatial directions, such an arrangement for the
$(a,b)$ planes forces the orbitals to have nonzero lobes along $c$.
This results in sizable AF superexchange for the bonds
$\langle ij\rangle$ parallel to $c$, which will order the spins
antiferromagnetically in the $c$ direction.
The orbitals may either repeat or stagger along the $c$-axis, and both
states give the same mean-field energy. Taking the magnetic ordering in the
three cubic directions [$abc$] as a label to classify the classical phases
with MO (\ref{orbmoffa}), we call this ground state the MO{\scriptsize FFA}
phase.
With the help of Eqs. (\ref{orbavab}) and (\ref{orbavc}) one obtains the
following classical energy as a function of $\theta$,
\begin{eqnarray}
\label{enemoffat}
E(\theta)&=&-\frac{J}{4}(1+\eta)(7-4\cos^2 2\theta) \nonumber \\
&-&\frac{J}{2}\beta(1+\frac{\eta}{2})(1-\cos^2 2\theta) \nonumber \\
&-&\frac{J}{2}\beta(1-\frac{\eta}{2})(1+\cos 2\theta)^2 \nonumber \\
&+&\frac{1}{2}E_z\cos 2\theta ,
\end{eqnarray}
with a minimum at
\begin{equation}
\label{thetamoffa}
\cos 2\theta={\beta(1-\frac{\eta}{2})-\frac{1}{2}\varepsilon_z
\over 2+(2+\beta)\eta},
\end{equation}
where again the MO exist as long as $|\cos 2\theta|\le 1$. Using Eqs.
(\ref{enemoffat}) and (\ref{thetamoffa}) one finds that the classical
energy of the MO{\scriptsize FFA} phase is given by
\begin{equation}
\label{enemoffa}
E_{\rm MO{\scriptsize FFA}}=-\frac{J}{4}\left(11-7\eta\right)
-\frac{J}{2}{[\beta(1-\frac{\eta}{2})-\frac{1}{2}\varepsilon_z]^2
\over 2+(2+\beta)\eta}.
\end{equation}
As a special case, let us consider first degenerate orbitals ($E_z=0$) in a
3D system ($\beta=1$). Eq. (\ref{thetamoffa}) simplifies in this case to
$\cos 2\theta= (1-\frac{\eta}{2})/(2+3\eta)$. A particularly simple
result is found at $\eta=0$ where $\cos 2\theta= 1/2$, i.e., $\theta=\pi/6$,
and the orbitals stagger like $x^2-z^2$ and $y^2-z^2$, as shown in Fig.
\ref{allmfa}. This staggering was proposed by Kugel and Khomskii as a ground
state of KCuF$_3$;\cite{Kug82} of course, this state is not realized for the
realistic parameters with $\eta\simeq 0.3$, but the optimized orbitals
with $\theta$ given by
(\ref{thetamoffa}) are not so far from this idealized picture.
The energy of the MO{\scriptsize FFA} phase is degenerate with that of the AF
phases at the classical degeneracy point, $M\equiv (E_z/J,\eta)=(0,0)$, and
this phase becomes more stable at $\eta>0$, and $E_z/J\simeq 0$. The magnetic
energy is gained due to relatively strong AF interactions on the bonds
$\langle ij\rangle\parallel c$, and weak FM interactions in the planes
$(a,b)$, perpendicular to the preferred directionality of the MO
(\ref{mixing}) along the $c$-direction, while the orbital energy is gained
due to orbital alternation within the $(a,b)$ planes. Such orbital ordering
remains stable with decreasing $E_z<0$, while two similar states with the
staggering either within the $(b,c)$ or the $(a,c)$ planes, are more stable
for $E_z>0$. Following our convention, these two degenerate MO states stable
at $E_z>0$ are called MO{\scriptsize AFF} and MO{\scriptsize FAF} (see Fig.
\ref{allmfa}), respectively. However, the MO involve in this case the
directional orbital $|\zeta\rangle$ along the AF bonds (i.e.,
$|\zeta_a\rangle \sim 3x^2-r^2$ for MO{\scriptsize AFF} or
$|\zeta_b\rangle \sim 3y^2-r^2$ for MO{\scriptsize FAF},
respectively), and the corresponding orthogonal orbital, $|\xi\rangle$.
Therefore, since the symmetry-breaking field acts on $|z\rangle$ orbitals,
the angles in the two sublattices cannot be exactly equivalent in this case,
unlike in the MO{\scriptsize FFA} phase, and we adopted an ansatz,
\begin{eqnarray}
\label{orbmoaff}
|i\sigma\rangle&=&\cos\theta_+|i\xi \sigma\rangle
+\sin\theta_+|i\zeta\sigma\rangle, \nonumber \\
|j\sigma\rangle&=&\cos\theta_-|i\xi \sigma\rangle
-\sin\theta_-|i\zeta\sigma\rangle,
\end{eqnarray}
where $i\in A$, $j\in B$, and $\theta_{\pm}>0$ for the two
sublattices. Introducing for convenience the new angles,
$\phi=\frac{1}{2}(\theta_++\theta_-)$, and $\delta=\theta_+-\theta_-$, one
finds the following conditions for the energy minimum of the classical
MO{\scriptsize AFF} phase,
\begin{eqnarray}
\label{thetamoaff1}
\cos 2\phi &=&
-\case{1}{4}\left\{\left[(1+{\beta})(2-\eta)
+\varepsilon_z\right]\cos\delta\right. \nonumber \\
&+&\left.\sqrt{3}\varepsilon_z\sin\delta\right\}
\left[1+\beta+(1+2\beta)\eta\right]^{-1} , \\
\label{thetamoaff2}
\tan 2\delta&=&
+\case{1}{2}\sqrt{3}\left[(1+{\beta})(2-\eta)+\varepsilon_z\right]
\varepsilon_z \nonumber \\
&\times&\left\{4\left[1+\beta+(1+2\beta)\eta\right]
+\left[(1+{\beta})(2-\eta)+\varepsilon_z\right]^2\right. \nonumber \\
&-&\left.\case{3}{4}\varepsilon_z^2\right\}^{-1},
\end{eqnarray}
and the energy is given by
\begin{eqnarray}
\label{enemoaff}
&E&_{\rm MO{\scriptsize AFF}}=-\frac{J}{4}\left[
7(1+\eta)+2\beta(1+\cos\delta)\right] \nonumber \\
&-&\frac{J}{32}{\left[[(1+{\beta})(2-\eta)+\varepsilon_z]\cos\delta
+\sqrt{3}\varepsilon_z\sin\delta\right]^2 \over
1+\beta+(1+2\beta)\eta} . \nonumber \\
\end{eqnarray}
Finally, one may consider states in which magnetic energy is gained in
the $c$-direction due to MO with a small admixture of $|z\rangle$ into
orbitals of predominantly $|x\rangle$-character, i.e.,
$\sin\theta_i=1-\epsilon$ in Eq. (\ref{mixing}).
As such a state is a modification of the AFxx phase, the two sublattices
in the $(a,b)$ planes are again physically equivalent, and it suffices to
introduce a single angle $\theta$ to characterize this state. Apart from
(large) energy contributions due to AF order on the bonds in the $(a,b)$
planes, the expansion of the ground state energy contains also (small)
terms depending on the spin order in the $c$-direction,
$\langle S_i^zS_j^z\rangle_{\parallel c}$,
\begin{equation}
\label{mixc}
E=\left(1+\cos2\theta\right)\left(1+\cos2\theta-\eta\right)
\langle S_i^zS_j^z\rangle_{\parallel c}+const,
\end{equation}
which prefers FM order as long as $(1+\cos2\theta)<\eta$. The reason is
that the AF superexchange is a fourth order effect $\sim\epsilon^4$, while
the FM interactions $\propto\eta$ are second order, $\sim\epsilon^2$, and
give a lower energy $E$ as long as the $|z\rangle$ occupancy is small enough.
Following our convention, we call the resulting state the MO{\scriptsize AAF}
phase, with the mixing angle given by
\begin{equation}
\label{orbmoaaf}
\cos2\theta=-{{1-\frac{\eta}{2}+\frac{1}{2}\varepsilon_z
\over \beta(1+\eta)+2\eta}},
\end{equation}
and the classical energy by
\begin{eqnarray}
\label{enemoaaf}
E_{\rm MO{\scriptsize AAF}}&=&-\left(2+\frac{3}{4}\eta\right)J \nonumber \\
&-&\frac{1}{2}\beta(1+\eta)
-J{\left( 2-\eta+\varepsilon_z\right)^2\over 2[\beta(1+\eta)+2\eta]}.
\end{eqnarray}
Therefore, only when the average population of the $|z\rangle$ orbitals,
$\sim\cos^2\theta$, increases sufficiently, one can find a transition to
the AF phase with mixed orbitals, MO{\scriptsize AAA}, discussed above.
By making several other choices of orbital mixing and classical magnetic
order, we have verified that no other commensurate ordering with up to four
sublattices can be stable in the present situation. Although some other
phases could be found, they were degenerate with the above phases only at
the $M$ point, and otherwise had higher energies. Thus, we obtain the
classical phase diagram of the 3D spin-orbital model (\ref{somcu}) by
comparing the energies of the six above phases for various values of two
parameters, $\{E_z/J,J_H/U\}$: two AF phases with two sublattices and pure
orbital character (AFxx and AFzz), three A-AF phases with four sublattices
(MO{\scriptsize FFA} and two degenerate phases: MO{\scriptsize AFF} and
MO{\scriptsize AFF}), one C-AF phase (MO{\scriptsize AAF}), and one G-AF
phase with MO's (MO{\scriptsize AAA}). While the orbital mixing is unstable
at $\eta=0$, the generic sequence of classical phases at finite $\eta$ and
decreasing $E_z/J$ is: AFxx, MO{\scriptsize AAF}, MO{\scriptsize AAA},
MO{\scriptsize AFF}, MO{\scriptsize FFA}, and AFzz, and the magnetic order
is tuned together with the gradually increasing $|z\rangle$ character of
the occupied orbitals.
The result for cubic symmetry ($\beta=1$) is presented in Fig. \ref{mfa3d},
where one finds all six phases, but the MO{\scriptsize AAA} phase does
stabilize only in a very restricted regime of parameters with $J_H/U<0.1$,
before MO{\scriptsize AFF} takes over. Only the first of the above
transitions is a continuous one, and the $|z\rangle$ amplitude
$\sim\cos^2\theta$ increases smoothly from zero and removes the built-in
degeneracy of the 2D AFxx phase with respect to the magnetic order along
the $c$-direction. All the other transition lines in Fig. \ref{mfa3d} are
associated with jumps in the magnetic and in orbital patterns. We emphasize
that all the considered phases with magnetic LRO are degenerate at the point
$M$, with classical energy of $-3J$. In fact $M$ is an infinite-order
quantum critical point, since not only may the spins be chosen to be FM in
certain planes, whence the orbitals have to be tuned to compensate the loss
of the magnetic energy by the orbital energy contributions, as realized in
all MO phases, but also may the orbitals be rotated freely when the spins
are AF in all three directions.We note, however, that the magnetic terms are
essential, and in a purely disordered spin system, with
$\langle S^z_iS^z_j\rangle=0$, a higher energy of $-21J/8$ is found even
with the optimal choice of orbitals with $\cos 2\theta=0$.
We also investigated the phase diagrams for the case of modified hopping
along the $c$-direction ($\beta \neq 1$). One finds that increased hopping
($\beta=1.414$) in the $c$-direction stabilizes the MO phases, and in
particular the MO{\scriptsize AFF} (MO{\scriptsize FAF}) phase [Fig.
\ref{beta}(a)]. By contrast, the MO phases are stable in a narrower range of
$E_z$ for a fixed value of $J_H/U$, if the hopping along the $c$-direction
is decreased below $\beta=1$ [an example of $\beta=0.707$ is shown in Fig.
\ref{beta}(b)]. The decreased stability of the MO{\scriptsize AFF} phase
promotes in this case the AF order with MO in the MO{\scriptsize AAA} phase.
The latter phase is stable only in a relatively narrow range of $E_z$, and
only for small enough $J_H/U$; an increase of $J_H/U$ favors instead FM
order along the $c$-direction. We also note that the orbital mixing sets in
for the MO{\scriptsize AAA} phase (\ref{orbmoaaa}) only at a smaller value
of $E_z$ than in the MO{\scriptsize AAF} phase (\ref{orbmoaaf}).
Interestingly, the point of high degeneracy of the classical states exists
{\em independently of the value of\/} $\beta$, and moves for $\beta\neq 1$
to $E_z=-2J(1-\beta)$. This demonstrates the generic nature of the internal
frustration of spin and orbital interactions in the model, and the crystal
field term just plays here a compensating role for the missing (or enhanced)
magnetic interactions within the $(a,b)$ planes.
Independently of the value of $\beta$, the spin-orbital model (\ref{somcu})
has a universal feature: different classical spin structures become
degenerate at the critical lines in Figs. \ref{mfa3d}-\ref{mfa2d}. This is
also encountered in frustrated 2D magnetic lattices described by simple
Heisenberg Hamiltonians,\cite{Chu91} and may thus be regarded as a
signature of frustration. However, unlike in the purely spin models,
in the present case (\ref{somcu}), the {\em sign\/} of the interactions
changes because of the coupling to the orbital sector, and this
{\em reduces the effective dimensionality\/} for the AF interactions
$\sim J$, with the 3D system behaving like a quasi-1D antiferromagnet.
\subsection{ Phase diagram of a 2D model }
As a special case, we considered the limit of $\beta\to 0$ which gives a 2D
spin-orbital model. The two AF phases with either $|x\rangle$ or $|z\rangle$
orbitals occupied, AFxx and AFzz, are degenerate at $E_z=-2J$. This asymmetry
reflects the large difference between the superexchange interactions for
$|x\rangle$ and $|z\rangle$ orbitals within the $(a,b)$ planes of a 2D system
which has to be compensated by the orbital energy (\ref{somez}).
As the presence of FM planes $\parallel c$-axis is crucial for the ordering
in the MO{\scriptsize AFF} phase (see Fig. \ref{allmfa}), this phase
disappears, while the remaining two phases with AF order within $(a,b)$
planes, MO{\scriptsize AAA} and MO{\scriptsize AAF}, collapse into a single
MO{\scriptsize AA} phase. Hence, one finds in two dimensions a classical
phase diagram with only four phases, which are stable with decreasing $E_z$
and at finite $\eta$ in the following order: AFxx, MO{\scriptsize AA},
MO{\scriptsize FF}, and AFzz (Fig. \ref{mfa2d}). The 2D phase diagram shows
in particular that strong AF superexchange in the $c$-direction is not the
stabilizing factor of the MO{\scriptsize FFA} phase in the 3D model, but
instead these phases are stable due to the orbital interactions which
enforce the orbital alternation shown in Fig. \ref{allmfa}.
For the realistic parameters of La$_2$CuO$_4$ the Cu $d_{x^2-y^2}$ and
$d_{3z^2-r^2}$ orbitals are split, and $E_z\simeq 0.64$ eV.\cite{Gra92}
This material belongs together with Nd$_2$CuO$_4$ to the class of cuprates
with weakly coupled CuO$_2$ planes, and one finds in the present treatment
a 2D AFxx state, as observed in neutron experiments.\cite{Kas98} If however
the orbital splitting is small in a 2D situation, the orbital ordering
couples strongly to the lattice, as the hybrids with alternating phasing on
two sublattices are formed according to Eqs. (\ref{orbmoaff}) The net result
is a quadrupolar distortion as indicated in Fig. \ref{disto}. In fact, using
these arguments Kugel and Khomskii predicted\cite{Kug73} the existence of
such a structural distortion in the MO{\scriptsize FF} phase of a quasi-2D
compound K$2$CuF$_4$. This prediction was confirmed experimentally a few
years later.\cite{Ito76}
The MO{\scriptsize FF} phase of K$_2$CuF$_4$ is magnetically polarized, has
no quantum fluctuations, and is thus well described in a classical theory.
In the next sections we concentrate ourselves on the 3D case, where the
quantum fluctuations are strong and destabilize the classical magnetic
ordering in a particular regime of parameters.
\section{ Elementary excitations }
\label{sec:magnons}
\subsection{General formalism}
\label{sec:genrpa}
The presence of the orbital degrees of freedom in the Hamiltonian
(\ref{somcu}) results in excitation spectra that are qualitatively different
from those of the HAF with a single spin-wave mode. As we have discussed in
the limit of $J_H=0$, the transverse excitations are twofold:
{\em spin-waves\/} and {\em spin-and-orbital waves\/}.\cite{Fei98}
In addition to these two modes there are also {\em longitudinal\/} (purely
orbital) excitations, and thus one finds three elementary excitations for
the present spin-orbital model (\ref{somcu}).\cite{Fei97,Fei98,Bri98} This
gives therefore the same number of modes as found in a 1D SU(4) symmetric
spin-orbital model in the Bethe ansatz method.\cite{Sut75,Fri99}
We emphasize that this feature is a consequence of the dimension (equal
to 15) of the $so(4)$ Lie algebra of the local operators, as explained
below, and is not related to the global symmetry of the Hamiltonian.
Here we present the analysis of the realistic $d^9$ spin-orbital model
for the 3D simple cubic (i.e., perovskite-like) lattice, using linear
spin-wave theory,\cite{Tak89,Aue94} generalized such as to make it
applicable to the present situation.
Before we introduce the excitation operators, it is convenient to rewrite
the spin-orbital model (\ref{somcu}) in a different representation which
uses a four-dimensional space,
$\{|x\! \uparrow\rangle$, $|x\! \downarrow\rangle$,
$|z\! \uparrow\rangle$, $|z\! \downarrow\rangle\}$, instead of a direct
product of the spin and orbital spaces. Hence, we introduce operators which
define purely spin excitations in individual orbitals,
\begin{equation}
S^{+}_{ixx}=d^{\dagger}_{ix\uparrow}d^{}_{ix\downarrow}, \hskip 1.0cm
S^{+}_{izz}=d^{\dagger}_{iz\uparrow}d^{}_{iz\downarrow},
\label{splus}
\end{equation}
and operators for simultaneous spin-and-orbital excitations,
\begin{equation}
K^{+}_{ixz}=d^{\dagger}_{ix\uparrow}d^{}_{iz\downarrow}, \hskip 1.0cm
K^{+}_{izx}=d^{\dagger}_{iz\uparrow}d^{}_{ix\downarrow}.
\label{oplus}
\end{equation}
The corresponding $S^z_{i\alpha\alpha}$ and $K^z_{i\alpha\beta}$ operators
are defined as follows,
\begin{eqnarray}
S^{z}_{ixx}&=&\case{1}{2}(n_{ix\uparrow}-n_{ix\downarrow}), \nonumber \\
S^{z}_{izz}&=&\case{1}{2}(n_{iz\uparrow}-n_{iz\downarrow}), \\
K^{z}_{ixz}&=&\case{1}{2}(d^{\dagger}_{ix \uparrow}d_{iz \uparrow}
-d^{\dagger}_{ix\downarrow}d_{iz\downarrow}),
\nonumber \\
K^{z}_{izx}&=&\case{1}{2}(d^{\dagger}_{iz \uparrow}d_{ix \uparrow}
-d^{\dagger}_{iz\downarrow}d_{ix\downarrow}).
\label{szet}
\end{eqnarray}
The Hamiltonian (\ref{somcu}) contains also purely orbital interactions which
can be expressed using the following orbital-flip ($T_{i\alpha\beta}$)
and orbital-polarization ($n_{i-}$) operators,
\begin{eqnarray}
T_{ixz}&=&\case{1}{2}(d^{\dagger}_{ix \uparrow}d_{iz \uparrow}
+d^{\dagger}_{ix\downarrow}d_{iz\downarrow}),
\nonumber \\
T_{izx}&=&\case{1}{2}(d^{\dagger}_{iz \uparrow}d_{ix \uparrow}
+d^{\dagger}_{iz\downarrow}d_{ix\downarrow}),
\nonumber \\
n_{i-}&=&\case{1}{2}(d^{\dagger}_{ix \uparrow}d_{ix \uparrow}\! +\!
d^{\dagger}_{ix\downarrow}d_{ix\downarrow}\! -\!
d^{\dagger}_{iz \uparrow}d_{iz \uparrow}\! -\!
d^{\dagger}_{iz\downarrow}d_{iz\downarrow} ).
\label{ozet}
\end{eqnarray}
In order to simplify the notation, we also introduce sum operators for
the spin-and-orbital and purely orbital operators,
\begin{eqnarray}
\label{top}
K^{+}_{i}&=&K^{+}_{ixz}+K^{+}_{izx}, \nonumber \\
K^{z}_{i}&=&K^{z}_{ixz}+K^{z}_{izx}, \nonumber \\
T _{i}&=&T _{ixz}+T _{izx} .
\end{eqnarray}
The full set of local operators at a site $i$ constitute an $so(4)$ Lie
algebra. While the spin operators (\ref{splus}) fulfill of course for $x$
and $z$ separately the usual $su(2)$ commutation relations, they also form
collectively a subalgebra of $so(4)$, and the same holds for the
spin-and-orbital operators (\ref{oplus}). However, as we will see below,
for the calculation of the excitations one also needs commutators between
spin and spin-and-orbital operators, so that one cannot avoid considering
the full Lie-algebra structure of $so(4)$, discussed in Appendix
\ref{sec:commute}.
The number of collective modes in a particular phase may be determined as
follows. The $so(4)$ Lie algebra consists of three Cartan operators, i.e.,
operators diagonal on the local eigenstates of the symmetry-broken phase
under consideration (e.g. $S^z_{ixx}$, $S^z_{izz}$, and $n_{i-}$ in the AFxx
phase), plus 12 non-diagonal operators turning the eigenstates into one
another (like $S^{+}_{ixx}$ and $S^{+}_{izz}$ in AFxx). Out of those twelve
operators, six connect two excited states (like $S^{+}_{izz}$ in AFxx),
and are physically irrelevant (at the RPA level), because they give only
rise to 'ghost' modes, modes for which the spectral function vanishes
identically. The remaining six operators connect the local ground state with
an excited state, three of them describing an excitation and three a
deexcitation, and only these six operators are physically relevant. Out of
the three excitations (deexcitations), two are transverse, i.e., change the
spin, and one is longitudinal, i.e., does not affect the spin. For a
classical phase with $L$ sublattices one therefore has $4L$ transverse and
$2L$ longitudinal operators per unit cell. Since the spin-orbital Hamiltonian
(\ref{somcu}) does not couple transverse and longitudinal operators, this
yields also $4L$ transverse and $2L$ longitudinal modes. Because of
time-reversal invariance they all occur in pairs with opposite frequencies,
$\pm \omega^{(n)}_{\vec k}$.
Finally, the $SU(2)$ spin invariance of the Hamiltonian guarantees that the
transverse operators raising the spin are decoupled from those lowering the
spin, and that they are described by the same set of equations of motion, so
that the transverse modes are pairwise degenerate. Such a simplification does
not occur in the longitudinal sector. So, in conclusion, in an $L$-sublattice
phase there are $L$ doubly-degenerate positive-frequency transverse modes and
$L$ nondegenerate positive-frequency longitudinal modes, accompanied by the
same number of negative-frequency modes. This may be compared with the
well-known situation in the HAF, where there is, with only spin operators
involved, only one (not two) doubly-degenerate positive-frequency (transverse)
mode in the two-sublattice N\'{e}el state.
For the actual evaluation it is convenient to decompose the superexchange
terms in the spin-orbital Hamiltonian (\ref{somcu}),
\begin{equation}
{\cal H}_J={\cal H}_{\parallel}+{\cal H}_{\perp},
\label{somlong}
\end{equation}
into two parts which depend on the bond direction:
(i) for the bonds $\langle ij\rangle\parallel (a,b)$,
\begin{eqnarray}
\label{hpara}
{\cal H}_{\parallel}&=&\case{1}{4}J\sum_{\langle ij\rangle{\parallel}}
\left[(1-\case{1}{2}\eta)
(3{\vec S}_{ixx}+{\vec S}_{izz}+\lambda_{ij}\sqrt{3}{\vec K}_{i})
\right. \nonumber \\
& &\left. \hskip 1.2cm
\cdot (3{\vec S}_{jxx}+{\vec S}_{jzz}+\lambda_{ij}\sqrt{3}{\vec K}_{j})
-2\eta {\vec S}_{i}\cdot {\vec S}_{j}
\right. \nonumber \\
& &\left. \hskip .5cm +(1+2\eta)(n_{i-}+\lambda_{ij}\sqrt{3}T_{i})
(n_{j-}+\lambda_{ij}\sqrt{3}T_{j})
\right. \nonumber \\
& &\left. \hskip .5cm -(3+\eta)\right],
\end{eqnarray}
where $\lambda_{ij}=(-1)^{{\vec\delta}{\vec y}}$ with ${\vec y}$ being a unit
vector in the $b$-direction, and
(ii) for the bonds $\langle ij\rangle\perp (a,b)$, i.e., along the
$c$-axis,
\begin{eqnarray}
\label{hperp}
{\cal H}_{\perp}&=&J\sum_{\langle ij\rangle{\perp}}
\left[(4-2\eta){\vec S}_{izz}\cdot{\vec S}_{jzz}
-\eta ({\vec S}_{ixx}\cdot{\vec S}_{jzz}
\right. \nonumber \\
&+&\left. {\vec S}_{izz}\cdot{\vec S}_{jxx})
+(1+2\eta)n_{i-}n_{j-}-\case{1}{4}(3+\eta)\right].
\end{eqnarray}
Here and in the following Sections we consider a 3D model with $\beta=1$.
We note that the orbital interactions (\ref{orbop}) are quite different in
$H_{\parallel}$ and $H_{\perp}$; propagating spin-and-orbital
excitations are possible only within the $(a,b)$ planes, where they are
coupled to the spin excitations, while in the $c$-direction only pure
spin excitations and pure orbital excitations occur, which are decoupled
from one another. This breaking of symmetry between $H_{\parallel}$ and
$H_{\perp}$ is a consequence of the choice of basis as $|x\rangle$ and
$|z\rangle$ orbitals.
In the following Sections we consider transverse and longitudinal
excitations in the various symmetry-broken states. The transverse
excitations, i.e., spin-waves and spin-and-orbital-waves, are calculated
using the spin-changing operators which make a transition to a state
realized in a classical phase at a given site $i$; for example for the
AFxx phase these operators are for $i$ in the A (spin-up) sublattice,
\begin{equation}
S^+_{ixx}=d^{\dagger}_{ix\uparrow}d^{}_{ix\downarrow}, \hskip 1cm
K^+_{ixz}=d^{\dagger}_{ix\uparrow}d^{}_{iz\downarrow}.
\label{excopt}
\end{equation}
The longitudinal excitations without spin-flip are most conveniently
obtained starting from spin-dependent orbital excitation operators,
\begin{equation}
T_{ixz\sigma}=d^{\dagger}_{ix\sigma}d^{}_{iz\sigma}, \hskip 1cm
T_{izx\sigma}=d^{\dagger}_{iz\sigma}d^{}_{ix\sigma}.
\label{excop}
\end{equation}
The commutation relations for these operators are presented in Appendix
\ref{sec:commute}.
\subsection{Antiferromagnetic AFxx phase}
\label{sec:afxxrpa}
The nature and dispersion of elementary excitations in the spin-orbital model
(\ref{somcu}) can be conveniently studied in the leading order of the $1/S$
expansion using the Green function formalism. We note, however, that
equivalent results for the AFxx and AFzz phases can be obtained using
instead an expansion around a classical saddle point with Schwinger
bosons.\cite{Aue94}
We start from the equations of motion for the Green functions generated by
the excitation operators (\ref{excopt}) written in the energy
representation,\cite{Zub60,Hal72}
\begin{eqnarray}
\label{gfafxx1}
E\langle\langle S_{ixx}^+|...\rangle\rangle &=&
{1\over 2\pi}\langle [S_{ixx}^+,...]\rangle +
\langle\langle [S_{ixx}^+,H]|...\rangle\rangle, \\
\label{gfafxx2}
E\langle\langle K_{ixz}^+|...\rangle\rangle &=&
{1\over 2\pi}\langle [K_{ixz}^+,...]\rangle +
\langle\langle [K_{ixz}^+,H]|...\rangle\rangle,
\end{eqnarray}
where the average of the commutator on the right hand side, e.g.
$\langle [S_{ixx}^+,S_{jxx}^-]\rangle$, is evaluated in the classical
ground state. The excitation operators were chosen as leading to the local
states $|ix\!\uparrow\rangle$ realized at one of the sublattices in the
ground state of the AFxx phase. As usually, the commutators in Eqs.
(\ref{gfafxx1}) and (\ref{gfafxx2}) generate higher-order Green functions.
In contrast to the HAF, it does not suffice to consider the spin-flip
Green function $\langle\langle S_{ixx}^+|...\rangle\rangle$, as the
spin-flips may also occur together with an accompanying orbital-flip,
as described by $\langle\langle K_{ixz}^+|...\rangle\rangle$.
We derived the equations of motion for the Green functions generated by
the set of operators $\{S_{ixx}^+,K_{ixz}^+,S_{jxx}^+,K_{jxz}^+\}$, where
$i\in A$ and $j\in B$, and used the random-phase approximation (RPA) for
spinlike operators which linearizes the equations of motion by a decoupling
procedure.\cite{Zub60,Hal72} Thereby, the operators which have nonzero
expectation values in the considered classical state give finite
contributions, e.g. for the first spin-flip Green function one uses
\begin{equation}
\langle\langle S_{ixx}^+S_{mxx}^z|...\rangle\rangle\simeq
\langle S_{mxx}^z\rangle \langle\langle S_{ixx}^+|...\rangle\rangle,
\label{rpas}
\end{equation}
and a similar formula for the mixed spin-and-orbital excitation described by
$\langle\langle K_{ixz}^+|...\rangle\rangle$,
\begin{equation}
\langle\langle K_{ixz}^+S_{mxx}^z|...\rangle\rangle\simeq
\langle S_{mxx}^z\rangle \langle\langle K_{ixz}^+|...\rangle\rangle .
\label{rpak}
\end{equation}
It is crucial that the decoupled operators have different site indices, and
thus the decoupling procedure preserves the local commutation rules given
in Appendix \ref{sec:commute}. Instead, if one uses products of spin and
orbital operators, e.g., $K_{ixz}^+=S_{ixx}^+\sigma_i^+$, one is tempted
to decouple these operators locally\cite{Cas78,Kha97} which would violate
the algebraic structure of the $so(4)$ Lie algebra.
In the present case of the AFxx phase one uses the respective N\'eel state
average values,
\begin{eqnarray}
\langle S_{ixx}^z\rangle&=&-\langle S_{jxx}^z\rangle=\case{1}{2}, \\
\langle n_{i-}\rangle&=& \langle n_{j-}\rangle=\case{1}{2},
\label{avx}
\end{eqnarray}
where $i\in A$ and $j\in B$, and $A$ and $B$ are the two sublattices in a 2D
lattice for the AFxx phase. All the remaining averages vanish, as this phase
has a pure $|x\rangle$-orbital character at every site, which simplifies
significantly the equations of motion which result from the RPA procedure.
The translational invariance of the N\'eel state implies that the
transformed Green functions are diagonal in the reduced Brillouin zone (BZ).
As in the HAF, the Fourier transformed functions are defined for the Green
functions which describe the spin dynamics on a given sublattice, either $A$
or $B$. For instance, the pure spin-flip Green functions are transformed as
follows,
\begin{eqnarray}
\langle\langle S_{{\vec k}xx}^+|...\rangle\rangle_A&=&
\frac{1}{\sqrt{N}}\sum_{i\in A}e^{i{\vec k}{\vec R}_i}
\langle\langle S_{ixx}^+|...\rangle\rangle_A, \nonumber \\
\langle\langle S_{{\vec k}xx}^+|...\rangle\rangle_B&=&
\frac{1}{\sqrt{N}}\sum_{j\in B}e^{i{\vec k}{\vec R}_j}
\langle\langle S_{jxx}^+|...\rangle\rangle_B,
\label{fourier}
\end{eqnarray}
where $N$ is the number of sites in one sublattice. Hence, the problem of
finding the elementary excitations of the considered spin-orbital model
(\ref{somcu}) reduces to the diagonalization of a $4\times 4$ dynamical
matrix at each ${\vec k}$-point, as given in Appendix \ref{sec:dynama}.
The symmetric positive and negative eigenvalues
$\pm\omega_{\vec k}^{(n)}$, with $n=1,2$, solved from the matrix in Eq.
(\ref{gfeq}) may be written in the following form for the AFxx phase,
\begin{eqnarray}
\label{afsw}
[\omega_{\vec k}^{(n)}]^2&=&J^2\left(\lambda_{x}^2+\tau_{x}^2
-Q_{x\vec k}^2-R_{\vec k}^2-2P_{x\vec k}^2\right)
\nonumber \\
&\pm &J^2\left[ (\lambda_{x}^2-\tau_{x}^2)^2
-2(\lambda_{x}^2-\tau_{x}^2)(Q_{x\vec k}^2-R_{\vec k}^2) \right.
\nonumber \\
&-&\left. 4(\lambda_{x}-\tau_{x})^2P_{x\vec k}^2
+(Q_{x\vec k}^2+R_{\vec k}^2+2P_{x\vec k}^2)^2 \right.
\nonumber \\
&-&\left. 4(Q_{x\vec k}R_{\vec k}-P_{x\vec k}^2)^2\right]^{1/2}.
\end{eqnarray}
Here the quantities $\lambda_{\alpha}$ and $\tau_{\alpha}$ play the role
of local potentials and follow from the model parameters, $E_z$ and $J_H$,
\begin{eqnarray}
\label{afxxlambda}
\lambda_x&=&\case{9}{2}-3\eta, \\
\label{afxxtau}
\tau_x&=&\case{7}{2}-4\eta-2-\eta+\varepsilon_z.
\end{eqnarray}
The remaining terms are ${\vec k}$-dependent, and depend on
\begin{eqnarray}
\label{gammap}
\gamma_{+}(\vec k)&=&\case{1}{2}(\cos k_x+\cos k_y), \\
\label{gammam}
\gamma_{-}(\vec k)&=&\case{1}{2}(\cos k_x-\cos k_y), \\
\label{gammaz}
\gamma_{z}(\vec k)&=&\cos k_z.
\end{eqnarray}
The quantities $Q_{x\vec k}$ and $P_{x\vec k}$ for the AFxx phase take the
form,
\begin{eqnarray}
\label{afxxq}
Q_{x\vec k}&=&(\case{9}{2}-3\eta)\gamma_{+}(\vec k), \\
\label{afxxp}
P_{x\vec k}&=&\case{1}{2}\sqrt{3}(3-\eta)\gamma_{-}(\vec k),
\end{eqnarray}
while the last dispersive term,
\begin{equation}
R_{\vec k}=\case{3}{2}\gamma_{+}(\vec k) ,
\label{rdef2}
\end{equation}
carries no index and remains identical for both AF phases (AFxx and AFzz). We
emphasize that the coupling between the spin-wave and spin-and-orbital-wave
excitations occurs due to the terms $\propto P_{x\vec k}$, as seen from Eq.
(\ref{gfeq}). It vanishes in the planes of $k_x=\pm k_y$, but otherwise plays
an important role, as discussed in Sec. \ref{sec:rpa}. In the limit of large
$E_z\to\infty$, Eq. (\ref{afsw}) reproduces the spin-wave excitations for a
2D antiferromagnet with an AF superexchange interaction of
$J(\frac{9}{4}-\frac{3}{2}\eta)$,
\begin{equation}
\label{limitafxx}
\omega_{\vec k}^{(1)}=J\left(\case{9}{2}-3\eta\right)
[1-\gamma_{+}^2(\vec k)]^{1/2},
\end{equation}
while the dispersion of the high-energy spin-and-orbital excitation,
$\omega_{\vec k}^{(2)}\simeq E_z$, becomes negligible.
As explained above, both modes are doubly degenerate.
Consider now the orbital (excitonic) excitations generated by the orbital-flip
operators (\ref{excop}). They are found by considering the equations of motion,
\begin{eqnarray}
\label{gfafxxl1}
E\langle\langle T_{i\alpha\beta \uparrow }|...\rangle\rangle &=&
{1\over 2\pi}\langle [T_{i\alpha\beta \uparrow },...]\rangle
+ \langle\langle [T_{i\alpha\beta \uparrow },H]|...\rangle\rangle,
\nonumber \\ \\
\label{gfafxxl2}
E\langle\langle T_{i\alpha\beta \downarrow}|...\rangle\rangle &=&
{1\over 2\pi}\langle [T_{i\alpha\beta \downarrow},...]\rangle
+ \langle\langle [T_{i\alpha\beta \downarrow},H]|...\rangle\rangle,
\nonumber \\
\end{eqnarray}
and the commutators are calculated using the rules (\ref{excom}).
In general, one finds four different excitation operators at each site.
However, making a Fourier transformations as for the transverse operators
(\ref{fourier}), one may show that only two operators per sublattice suffice
to describe the modes in an antiferromagnet. The structure of the respective
RPA dynamical matrix is given in Appendix \ref{sec:dynama}. The orbital
excitations which follow from Eq. (\ref{gfeqor}) are in general given by
\begin{equation}
\label{genorb}
\zeta_{\vec k}= J \;
\left[u_{\alpha}(u_{\alpha}\pm 2 \rho_{\alpha\vec k})\right]^{1/2},
\end{equation}
yielding two, in general nondegenerate, positive-frequency modes.
In the AFxx phase one finds,
\begin{eqnarray}
\label{afxxu}
u_x&=&\varepsilon_z-3\eta, \\
\label{afxxrho}
\rho_{x\vec k}&=&\case{3}{2}\eta\gamma_{+}(\vec k).
\end{eqnarray}
It is important to realize that the propagation of longitudinal excitations,
being equivalent to a finite dispersion of longitudinal modes, becomes
possible only at $\eta>0$. This follows from the multiplet structure of the
excited $d^8$ states, which allows a spin-flip between the orbitals in the
$|^1E_{\theta}\rangle$ and in the $S^z=0$ component of the
$|^3A_2\rangle$-state
only if $J_H\neq 0$, as illustrated in Fig. \ref{orbex}. The processes
$\sim t_{xz}$ are not shown, as they would lead to a final state shown in
Fig. \ref{orbex}(b), i.e., to a propagation of a spin-and-orbital excitation
which was already considered above. In contrast, the relevant longitudinal
orbital excitation in the symmetry-broken state implies that the exciton
has the same spin as imposed by the N\'eel state of the background; this
state is shown in Fig. \ref{orbex}(c). Therefore, in a perfect N\'eel state
without FM interactions due to $\eta\neq 0$, only local orbital
excitations are possible. These local excitations cost no energy in the limit
of $\varepsilon_z\to 0$ which demonstrates again the frustration of magnetic
interactions at the classical degeneracy point, $\varepsilon_z=\eta=0$.
An example of the excitation spectra is shown in Fig. \ref{modesxx} for the
main directions in the 2D BZ, with $X=(\pi,0)$ and $S=(\pi/2,\pi/2)$.
Near the $\Gamma$ point one finds a (doubly-degenerate) Goldstone mode
$\omega_{\vec k}^{(1)}$ with dispersion $\sim k$ at ${\vec k}\to 0$, as in
the HAF, and a second (doubly-degenerate) transverse mode at higher energy,
$\omega_{\vec k}^{(2)}\simeq \omega_0+ak^2$.
Near $\Gamma$ the Goldstone mode is essentially purely spin-wave, the
second mode purely spin-and-orbital wave. With increasing ${\vec k}$ these
modes start to mix due to the $P_{x{\vec k}}$ term along the $\Gamma-X$
direction. This is best illustrated by the intensity measured in the
neutron scattering experiments, which see only the spin-wave component
in each transverse mode, as explained in more detail in Appendix
\ref{sec:neutrons}. The intensity $\chi({\vec q})$ moves from one mode
to the other along the $\Gamma-X$ direction in the 2D BZ
(Fig. \ref{modesxx}), demonstrating that indeed the lowest (highest)
mode is predominantly spin-wave-like (spin-and-orbital-wave-like) before
the anticrossing point, while this is reversed after the anticrossing of
the two modes. Thus, we make here a specific prediction that {\it two
spin-wave-like modes could be measurable in certain parts of the 2D
BZ}, in particular in the vicinity of an anticrossing,
if only an AFxx phase was realized for parameters not too distant
from the classical degeneracy point. Unfortunately, for the realistic
parameters for the cuprates,\cite{Gra92} one finds $E_z/J\simeq 10$ which
makes the spin-and-orbital excitation and the changes of the spin-wave
dispersion hardly visible in neutron spectroscopy.
The orbital (longitudinal) excitations are found for the parameters of Fig.
\ref{modesxx} at a finite energy, being of the same order of magnitude as
the energy of the spin-and-orbital excitation, $\omega^{(2)}_{\vec k}$. The
weak dispersion of these modes follows from the spin-flip processes in the
{\em excited\/} states, as explained in Fig. \ref{orbex} and discussed above.
We emphasize that the orbital mode has a gap and {\it does not couple\/} to
any spin excitation. At the classical degeneracy point $M$ the orbital mode
falls to zero energy and is dispersionless, expressing that the orbital
can be changed locally without any cost in energy.
\subsection{Antiferromagnetic AFzz phase}
\label{sec:afzzrpa}
The transverse excitations in the AFzz phase are determined by considering the
complementary set of Green functions to that given in Eqs. (\ref{gfafxx1})
and (\ref{gfafxx2}),
\begin{eqnarray}
\label{gfafzz1}
E\langle\langle S_{izz}^+|...\rangle\rangle &=&
{1\over 2\pi}\langle [S_{izz}^+,...]\rangle +
\langle\langle [S_{izz}^+,H]|...\rangle\rangle, \\
\label{gfafzz2}
E\langle\langle K_{izx}^+|...\rangle\rangle &=&
{1\over 2\pi}\langle [K_{izx}^+,...]\rangle +
\langle\langle [K_{izx}^+,H]|...\rangle\rangle,
\end{eqnarray}
with the excitations to the local $|iz\uparrow\rangle$ states.
As usually, the average of the commutator on the right hand side is next
evaluated in the classical ground state. After obtaining the RPA equations,
we thus use the following nonvanishing averages,
\begin{eqnarray}
\langle S_{izz}^z\rangle&=&-\langle S_{jzz}^z\rangle=\case{1}{2}, \\
\langle n_{i-}\rangle&=& \langle n_{j-}\rangle=-\case{1}{2},
\label{avz}
\end{eqnarray}
in the AFzz phase. This leads again to the general form (\ref{gfeq}), with
all the elements except for $R_{\vec k}$ replaced by,
\begin{eqnarray}
\label{afzzlambda}
\lambda_z&=&\case{1}{2}-\eta+2(2-\eta), \\
\label{afzztau}
\tau_z&=&-\case{1}{2}-\eta+2(1-2\eta)-\varepsilon_z, \\
\label{afzzq}
Q_{z\vec k}&=&(\case{1}{2}-\eta)\gamma_{+}(\vec k)
+2(2-\eta)\gamma_{z}(\vec k), \\
\label{afzzp}
P_{z\vec k}&=&\case{1}{2}\sqrt{3}(1-\eta)\gamma_{-}(\vec k) .
\end{eqnarray}
Thus, the transverse excitations have the same form (\ref{afsw}) as in the
AFxx phase, but the above quantities (\ref{afzzlambda})--(\ref{afzzp}) have
to be used.
In the limit of large $E_z\to-\infty$ one finds the spin-wave
for a 3D anisotropic antiferromagnet with strong superexchange equal to
$2J(2-\eta)$ along the $c$-axis, and weak superexchange
$\frac{1}{4}J(1-2\eta)$ within the $(a,b)$-planes,
\begin{eqnarray}
\label{limitafzz}
\omega_{\vec k}^{(1)}&=&J\left\{
\left[(\case{1}{2}-\eta)+2(2-\eta)\right]^2\right. \nonumber \\
& &-\left.\left[(\case{1}{2}-\eta)\gamma_{+}(\vec k)
+2(2-\eta)\gamma_{z}\right]^2\right\}^{1/2},
\end{eqnarray}
while the spin-and-orbital excitation, $\omega_{\vec k}^{(2)}\simeq -E_z$,
is dispersionless. Again, both these transverse modes are doubly degenerate.
The orbital excitations in the AFzz phase are found using
the equations of motion of the form (\ref{gfafxxl1}) and (\ref{gfafxxl2})
which lead to Eq. (\ref{genorb}) with,
\begin{eqnarray}
\label{afzzu}
u_z&=&-\varepsilon_z-3\eta, \\
\label{afzzrho}
\rho_{z,\vec k}&=&-\case{3}{2}\eta\gamma_{+}(\vec k),
\end{eqnarray}
and we find again zero-energy nondispersive modes at $\varepsilon_z=\eta=0$.
The representative excitation spectrum for the AFzz phase is shown in Fig.
\ref{modeszz}. We use the 3D BZ for a $bcc$ lattice with the
standard notation: $W=(\pi,\pi/2,0)$, $L=(\pi/2,\pi/2,\pi/2)$ and
$K=(3\pi/4,3\pi/4,0)$. The transverse modes have qualitatively the same
behavior as in the 2D AFxx phase, and one finds a Goldstone mode
$\omega_{\vec k}^{(1)}$ at the $\Gamma$ point which is spin-wave-like,
accompanied by a finite energy spin-and-orbital mode $\omega_{\vec k}^{(2)}$.
The first one is linear, while the second changes quadratically with
increasing ${\vec k}$. The dispersion in the $\Gamma-X$ direction is, however,
only $\sim 0.7J$, while in the AFxx phase a large dispersion of $\sim 2.5J$
was found (Fig. \ref{modesxx}). This demonstrates the very large difference
between the superexchange in the $(a,b)$-planes in the two AF phases.
Here one should bear in mind, that in a strongly anisotropic antiferromagnet,
such as the AFzz phase, the dispersion of the spin-wave mode in the
$(k_x,k_y)$ plane is roughly $(2 J_{ab} J_c)^{1/2}$, so actually enhanced by
$(J_c/2J_{ab})^{1/2}$ compared with the planar exchange constant. In fact,
there is also strong mixing between spin wave and spin-and-orbital wave along
$\Gamma-X$, depressing at the $X$-point $\omega^{(1)}_X$ by no less than
$0.5 J$ from its pure spin-wave value. The mixing effect is also visible in
the relatively large neutron intensity of the second mode. By contrast,
the transverse excitations are rather pure all along the $W-L$ direction
[where the neutron intensity $\chi({\vec q})$ is larger], except in the
regime where $\omega^{(1)}_{\vec k}\simeq\omega^{(2)}_{\vec k}$ and the
neutron intensity is distributed between the modes. However, owing to the
abruptness of the anticrossing, the range where the modes have
simultaneously appreciable intensity is very narrow, and their energetic
proximity then makes it likely that they would be measured as a single
broad maximum.
The (longitudinal) orbital excitation is found at the $X$ and $L$ points
at the same energy as that of a {\it local\/} excitation from $|z\rangle$
to $|x\rangle$ orbital (see Fig. \ref{modeszz}). It depends only on the
energy difference between the orbitals, and has a weak dispersion by the
same mechanism as described above for the AFxx phase (Fig. \ref{orbex}).
\end{multicols}
\widetext
\subsection{Mixed-orbital FFA phase}
\label{sec:moffarpa}
The excitation operators which couple to the local states in a symmetry-broken
phase with mixed orbitals are linear combinations of the operators considered
in Secs. \ref{sec:afxxrpa} and \ref{sec:afzzrpa}. It is therefore convenient
to make a unitary transformation of the Hamiltonian (\ref{somcu}) to new
orbitals defined as follows for $i\in A$ or $i\in D$ sublattice,
\begin{equation}
\label{newstatei}
\left( \begin{array}{c}
|i\mu\rangle \\
|i\nu\rangle
\end{array} \right) =
\left(\begin{array}{cc}
\cos\theta & \sin\theta \\
-\sin\theta & \cos\theta
\end{array} \right)
\left( \begin{array}{c}
|iz\rangle \\
|ix\rangle
\end{array} \right) ,
\end{equation}
and for $j\in B$ or $j\in C$ sublattice,
\begin{equation}
\label{newstatej}
\left( \begin{array}{c}
|j\mu\rangle \\
|j\nu\rangle
\end{array} \right) =
\left(\begin{array}{cc}
\cos\theta & -\sin\theta \\
\sin\theta & \cos\theta
\end{array} \right)
\left( \begin{array}{c}
|jz\rangle \\
|jx\rangle
\end{array} \right) .
\end{equation}
With these definitions and by choosing the angle $\theta$ at the value which
minimizes the classical energy (\ref{thetamoffa}), we guarantee that
$|i\mu\rangle$ and $|j\mu\rangle$, respectively, are at each site the orbital
state realized in the classical MO{\scriptsize FFA} phase, which is G-type
with respect to the orbital ordering, while $|i\nu\rangle$ and $|j\nu\rangle$
are the excited state, so that one can readily define the excitation operators
pertinent to the symmetry-broken ground state of this phase. Thus the spin,
spin-and-orbital, and orbital operators in terms of the new orbital states
$\{|\mu\rangle,|\nu\rangle\}$ defined by Eqs. (\ref{newstatei}) and
(\ref{newstatej}) are
\begin{eqnarray}
\label{newspin1}
{\cal K}_{i\alpha\beta }^+&=&
|i\alpha \uparrow\rangle\langle i\beta \downarrow| , \\
\label{newspin2}
{\cal K}_{i\alpha\beta }^z&=&\case{1}{2}
(|i\alpha \uparrow\rangle\langle i\beta \uparrow|
-|i\alpha\downarrow\rangle\langle i\beta \downarrow|) , \\
\label{newspin3}
{\cal T}_{i- } &=&\case{1}{2}\sum_{\sigma}
(|i\mu \sigma\rangle\langle i\nu \sigma|
+|i\nu \sigma\rangle\langle i\mu \sigma|) , \\
\label{newspin4}
{\cal N}_{i- } &=&\case{1}{2}\sum_{\sigma}
(|i\mu \sigma\rangle\langle i\mu \sigma|
-|i\nu \sigma\rangle\langle i\nu \sigma|) .
\end{eqnarray}
The new operators: ${\vec {\cal K}}_{i\alpha\beta}$, ${\cal T}_i$ and
${\cal N}_{i-}$ fulfill the same commutation rules as the nontransformed
operators: ${\vec K}_{i\alpha\beta }$, $T_i$, and $n_{i-}$, respectively;
they are given in Appendix \ref{sec:commute}. To simplify the notation we
also introduce total spin and spin-and-orbital operators,
\begin{eqnarray}
\label{newsping1}
{\vec {\cal S}}_{i}&=&{\vec {\cal S}}_{i\mu\mu}+{\vec {\cal S}}_{i\nu\nu}, \\
\label{newsping2}
{\vec {\cal K}}_{i}&=&{\vec {\cal K}}_{i\mu\nu}+{\vec {\cal K}}_{i\nu\mu}.
\end{eqnarray}
The Hamiltonian (\ref{somcu}) has to be transformed by the inverse
transformations to those given by Eqs. (\ref{newstatei}) and
(\ref{newstatej}). For the bonds $\langle ij\rangle\parallel (a,b)$ with
$i\in A$ and $j\in B$ one finds,
\begin{eqnarray}
\label{hparalong}
{\cal H}_{\parallel}&=&\case{1}{4}J\sum_{\langle ij\rangle{\parallel}}
\left\{(1-\case{1}{2}\eta)\left[
((2-\cos 2\theta){\vec {\cal S}}_{i\mu\mu}
+(2+\cos 2\theta){\vec {\cal S}}_{i\nu\nu}
+\sin 2\theta{\vec {\cal K}}_i)
\right.\right. \nonumber \\
& &\left.\left.\hskip 2.5cm
\times((2-\cos 2\theta){\vec {\cal S}}_{j\mu\mu}
+(2+\cos 2\theta){\vec {\cal S}}_{j\nu\nu}
-\sin 2\theta{\vec {\cal K}}_i)
\right.\right. \nonumber \\
& &\left.\left.\hskip 1.2cm
+3(\sin 2\theta({\vec {\cal S}}_{i\mu\mu}-{\vec {\cal S}}_{i\nu\nu})
+\cos 2\theta {\vec {\cal K}}_{i})
(\sin 2\theta({\vec {\cal S}}_{j\mu\mu}-{\vec {\cal S}}_{j\nu\nu})
+\cos 2\theta {\vec {\cal K}}_{j})
\right.\right. \nonumber \\
&+&\left.\left. \!\! \lambda_{ij}\sqrt{3}\left(
((2-\cos 2\theta){\vec {\cal S}}_{i\mu\mu}
+(2+\cos 2\theta){\vec {\cal S}}_{i\nu\nu}
+\sin 2\theta{\vec {\cal K}}_i)
(\sin 2\theta({\vec {\cal S}}_{j\mu\mu}\! -\! {\vec {\cal S}}_{j\nu\nu})
-\cos 2\theta{\vec {\cal K}}_j)
\right.\right.\right. \nonumber \\
& &\left.\left.\left.\hskip 0.4cm
-(\sin 2\theta({\vec {\cal S}}_{i\mu\mu}-{\vec {\cal S}}_{i\nu\nu})
+\cos 2\theta{\vec {\cal K}}_i)
((2-\cos 2\theta){\vec {\cal S}}_{j\mu\mu}
+(2+\cos 2\theta){\vec {\cal S}}_{j\nu\nu}
-\sin 2\theta{\vec {\cal K}}_j)\right)\right]
\right. \nonumber \\
&+&\left. \case{1}{2}\eta\left[
(\cos 2\theta({\vec {\cal S}}_{i\mu\mu}-{\vec {\cal S}}_{i\nu\nu})
-\sin 2\theta{\vec {\cal K}}_i)
(\cos 2\theta({\vec {\cal S}}_{j\mu\mu}-{\vec {\cal S}}_{j\nu\nu})
+\sin 2\theta{\vec {\cal K}}_j)
\right.\right. \nonumber \\
& &\left.\left.\hskip .3cm
-3(\sin 2\theta({\vec {\cal S}}_{i\mu\mu}-{\vec {\cal S}}_{i\nu\nu})
+\cos 2\theta{\vec {\cal K}}_i)
(\sin 2\theta({\vec {\cal S}}_{j\mu\mu}-{\vec {\cal S}}_{j\nu\nu})
-\cos 2\theta{\vec {\cal K}}_j)
\right.\right. \nonumber \\
& &\left.\left. - \lambda_{ij}\sqrt{3}\left(
(\cos 2\theta({\vec {\cal S}}_{i\mu\mu}-{\vec {\cal S}}_{i\nu\nu})
-\sin 2\theta{\vec {\cal K}}_i)
(\sin 2\theta({\vec {\cal S}}_{j\mu\mu}-{\vec {\cal S}}_{j\nu\nu})
-\cos 2\theta{\vec {\cal K}}_j)
\right.\right.\right. \nonumber \\
& &\left.\left.\left.\hskip 1.2cm
+(\sin 2\theta({\vec {\cal S}}_{i\mu\mu}-{\vec {\cal S}}_{i\nu\nu})
+\cos 2\theta{\vec {\cal K}}_i)
(\cos 2\theta({\vec {\cal S}}_{j\mu\mu}-{\vec {\cal S}}_{j\nu\nu})
-\sin 2\theta{\vec {\cal K}}_j)\right)\right]
- 2\eta {\vec {\cal S}}_{i}{\vec {\cal S}}_{j}
\right. \nonumber \\
&+&\left. (1+2\eta)\left[
(\cos 2\theta{\vec {\cal N}}_i-\sin 2\theta{\vec {\cal T}}_i)
(\cos 2\theta{\vec {\cal N}}_j+\sin 2\theta{\vec {\cal T}}_j)
\right.\right. \nonumber \\
& &\left.\left. \hskip 1.0cm
-3(\sin 2\theta{\vec {\cal N}}_i+\cos 2\theta{\vec {\cal T}}_i)
(\sin 2\theta{\vec {\cal N}}_j-\cos 2\theta{\vec {\cal T}}_j)
\right.\right. \nonumber \\
& &\left.\left. -\lambda_{ij}\sqrt{3}\left(
(\cos 2\theta{\vec {\cal N}}_i-\sin 2\theta{\vec {\cal T}}_i)
(\sin 2\theta{\vec {\cal N}}_j-\cos 2\theta{\vec {\cal T}}_j)
\right.\right.\right. \nonumber \\
& &\left.\left.\left. \hskip 1.0cm
+(\sin 2\theta{\vec {\cal N}}_i+\cos 2\theta{\vec {\cal T}}_i)
(\cos 2\theta{\vec {\cal N}}_j+\sin 2\theta{\vec {\cal T}}_j)
\right)\right]-\left(3+\eta\right)\right\},
\end{eqnarray}
while for the bonds $\langle ij\rangle\perp (a,b)$ it takes the form
\begin{eqnarray}
\label{hperplong}
{\cal H}_{\perp}\!&=\!&J\!\sum_{\langle ij\rangle{\perp}}
\! \left\{(1-\case{1}{2}\eta)
((1+\cos 2\theta){\vec {\cal S}}_{i\mu\mu}
+(1-\cos 2\theta){\vec {\cal S}}_{i\nu\nu}
-\sin 2\theta{\vec {\cal K}}_i)
\right. \nonumber \\
& &\left. \hskip 2.1cm \times ((1+\cos 2\theta){\vec {\cal S}}_{j\mu\mu}
+(1-\cos 2\theta){\vec {\cal S}}_{j\nu\nu}
-\sin 2\theta{\vec {\cal K}}_j)
\right. \nonumber \\
&-&\left. \case{1}{4}\eta \left[
((1-\cos 2\theta){\vec {\cal S}}_{i\mu\mu}
+(1+\cos 2\theta){\vec {\cal S}}_{i\nu\nu}
+\sin 2\theta{\vec {\cal K}}_i)
\right.\right. \nonumber \\
& &\left.\left. \hskip .5cm \times ((1+\cos 2\theta){\vec {\cal S}}_{j\mu\mu}
+(1-\cos 2\theta){\vec {\cal S}}_{j\nu\nu}
-\sin 2\theta{\vec {\cal K}}_j)
\right.\right. \nonumber \\
& &\left.\left. \hskip .5cm
+((1+\cos 2\theta){\vec {\cal S}}_{i\mu\mu}
+(1-\cos 2\theta){\vec {\cal S}}_{i\nu\nu}
-\sin 2\theta{\vec {\cal K}}_i)
\right.\right. \nonumber \\
& &\left.\left. \hskip .5cm \times ((1-\cos 2\theta){\vec {\cal S}}_{j\mu\mu}
+(1+\cos 2\theta){\vec {\cal S}}_{j\nu\nu}
+\sin 2\theta{\vec {\cal K}}_j)\right]
\right. \nonumber \\
&+&\left. (1+2\eta)
(\cos 2\theta{\vec {\cal N}}_i-\sin 2\theta{\vec {\cal T}}_i)
(\cos 2\theta{\vec {\cal N}}_j-\sin 2\theta{\vec {\cal T}}_j)
-\case{1}{4}(3+\eta)\right\}.
\end{eqnarray}
Finally, the transformed orbital-anisotropy term reads
\begin{equation}
\label{kk3long}
{\cal H}_{\tau} = E_z \sum_i
(\cos 2\theta{\vec {\cal N}}_i-\sin 2\theta{\vec {\cal T}}_i).
\end{equation}
\begin{multicols}{2
The transverse excitations may be found starting from the relevant raising
operators that lead to the local state $|i\mu\uparrow\rangle$ realized in
one of the sublattices, analogous to those introduced for the AFxx phase
(\ref{excopt}), i.e., the set
$\{{\cal S}_{i\mu\mu}^+,{\cal K}_{i\mu\nu}^+,
{\cal S}_{j\mu\mu}^+,{\cal K}_{j\mu\nu}^+,
{\cal S}_{k\mu\mu}^+,{\cal K}_{k\mu\nu}^+,
{\cal S}_{l\mu\mu}^+,{\cal K}_{l\mu\nu}^+\}$, where
$i\in A$, $j\in B$, $k\in C$, and $l\in D$;
they lead as usual to the orbitals $\{|i\mu\rangle, |j\mu\rangle\}$
(\ref{orbmoffa}) realized in the MO{\scriptsize FFA} phase,
\begin{eqnarray}
\label{gfmoffa1}
E\langle\langle {\cal S}_{i\mu\mu}^+|...\rangle\rangle &=&
{1\over 2\pi}\langle [{\cal S}_{i\mu\mu}^+,...]\rangle +
\langle\langle [{\cal S}_{i\mu\mu}^+,H]|...\rangle\rangle, \\
\label{gfmoffa2}
E\langle\langle {\cal K}_{i\mu\nu}^+|...\rangle\rangle &=&
{1\over 2\pi}\langle [{\cal K}_{i\mu\nu}^+,...]\rangle +
\langle\langle [{\cal K}_{i\mu\nu}^+,H]|...\rangle\rangle.
\end{eqnarray}
We applied the same RPA procedure as in Secs. \ref{sec:afxxrpa} and
\ref{sec:afzzrpa} in order to determine the Green function equations in the
${\vec k}$-space. The longitudinal excitations can be obtained from
operators similar to those used in the AFxx and AFzz phases (\ref{excop}),
\begin{equation}
\label{excop1}
{\cal T}_{i\mu\nu\uparrow}=
d^{\dagger}_{i\mu\uparrow}d^{}_{i\nu\uparrow}, \hskip 1.0cm
{\cal T}_{i\nu\mu\uparrow}=
d^{\dagger}_{i\nu\uparrow}d^{}_{i\mu\uparrow},
\end{equation}
for the $(a,b)$ planes with the $\uparrow$-spins, and the corresponding
${\cal T}_{i\mu\nu\downarrow}$ and ${\cal T}_{i\nu\mu\downarrow}$ for the
$(a,b)$ planes with the $\downarrow$-spins. The commutation operators for
these operators are analogous to those presented in Appendix
\ref{sec:commute} and may be easily obtained. The resulting dynamical
matrices for both transverse and longitudinal excitations are given in
Appendix \ref{sec:dynama}; their numerical diagonalization gave the modes
presented below. There are four doubly-degenerate positive-frequency
transverse modes, and four non-degenerate positive-frequency longitudinal
modes, consistent with the MO{\scriptsize FFA} phase having four sublattices.
An example of the transverse and longitudinal modes in the
MO{\scriptsize FFA} phase is presented in Fig. \ref{modesffa}. The modes are
shown in the respective BZ which corresponds to the magnetic unit cell of
the MO{\scriptsize FFA} phase: The 2D part along $\Gamma-X-S-\Gamma$ is
identical with the AFxx phase (compare Fig. \ref{modesxx}),
reflecting the orbital alternation, while the AF coupling along the $c$-axis
results in the folding of the zone along the $\Gamma-Z$ direction, with
$Z'=(0,0,\pi/2)$ and $S'=(\pi/2,\pi/2,\pi/2)$. One finds one Goldstone mode,
and three other finite-energy modes at the $\Gamma$ point. If no AF coupling
along the $c$-axis is present, similar positive-energy modes describe the
excitation spectrum in the MO{\scriptsize FF} phase in the 2D part of the
BZ (in the region of stability shown in Fig. \ref{mfa2d}), and the
symmetric negative-frequency modes carry then no weight. In contrast, due to
the strong AF interactions in the MO{\scriptsize FFA} phase, the negative
modes give a large energy renormalization due to quantum fluctuations, as
discussed in more detail in Sec. \ref{sec:rpa}.
The spin-wave and spin-and-orbital-wave excitations are well separated along
the $\Gamma-X-S-\Gamma$ path, with a gap of $\sim 0.5J$, as the FM
interactions $\propto J\eta$ are considerably weaker than the orbital
interactions which are $\propto J$. Therefore, the neutron intensity
$\chi({\vec q})$ is found mainly as originating from the lowest energy mode,
$\omega_{\vec k}^{(1)}$, with a small admixture of the higher-energy
spin-and-orbital excitation, $\omega_{\vec k}^{(3)}$. The magnetic
interactions are considerably stronger along the $c$-axis; the modes mix and
the higher-energy excitations, $\omega_{\vec k}^{(n)}$ with $n=3,4$, have
a larger dispersion in the remaining directions with $k_z\neq 0$. Strong
mixing of the modes in this part of the BZ is also visible in the intensity
distribution, with the modes $n=1$ and $n=3$ contributing with comparable
intensities (Fig. \ref{modesffa}). The fact that modes labelled as 2 and 4
have zero intensity is due to the path $\Gamma-Z'-S'-\Gamma$ being in the
high-symmetry BZ plane where $k_x=k_y$ so that $\gamma_{-}(\vec{k})=0$.
Then modes 2 and 4 have equal amplitude but are exactly out-of-phase between
$A$ and $B$ sites as well as between $C$ and $D$ sites, and so their neutron
intensities vanish, and only the companion in-phase modes 1 and 3 are
observable by neutrons. Unfortunately, no experimental verification
of these spectra is possible at present, as the spin excitations
measured in neutron scattering for KCuF$_3$ are consistent with the Bethe
ansatz and thus suggest a spin-liquid ground state with strong 1D AF
correlations instead of the A-AF phase with magnetic LRO.\cite{Ten93}
Interestingly, although the order in the $(a,b)$ planes is FM, the energy
of the Goldstone mode increases {\it linearly in all three directions\/} with
increasing ${\vec k}$, and the slopes are proportional to the respective
exchange interactions. This behavior is a manifestation of the A-AF spin
order; a qualitatively similar spectrum is found experimentally in
LaMnO$_3$,\cite{Hir96} where, however, the excitation spectra describe large
spins $S=2$ of Mn$^{3+}$ ions. The rather small dispersion of the spin-wave
part at low energies is due to small values of the exchange constants for
the actual optimal orientation of orbitals found at $J_H/U=0.3$. We note,
however, that the AF interactions along the $c$-axis are much stronger at
$J_H\to 0$ than in the present case. The AF structure along the $c$-axis may
be easily recognized from the symmetric spin-wave mode in the $\Gamma-Z$
direction with respect to $Z'=(0,0,\pi/2)$, while this mode increases all the
way from the $\Gamma$ to the $X$ point. The fact that only two modes have
nonzero neutron scattering intensity along $\Gamma-Z'-S'-\Gamma$ is due to
this BZ path being in the high-symmetry BZ plane, where $k_x=k_y$ and
$\gamma_{-}(\vec{k})=0$. Then two modes have equal amplitude but are exactly
out-of-phase between $A$ and $B$ sites as well as between $C$ and $D$ sites,
and so their neutron intensities vanish, while only the companion in-phase
modes are visible to neutrons. Unlike in the AF phases, the purely orbital
excitation is here energetically separated from the spin-wave and
spin-and-orbital-wave modes. The dispersion is quite small and decreases
with $\eta$.
\subsection{Mixed-orbital AFF phase}
\label{sec:moaffrpa}
The elementary excitations in the MO{\scriptsize AFF} phase may be obtained
using a similar scheme to that used in Sec. \ref{sec:moffarpa} for the
MO{\scriptsize FFA} phase. First of all, one defines new quantum states
which correspond to the minimum of the classical problem. This is realized
by a unitary transformation of the Hamiltonian to the new orbitals
defined for $i\in A$ sublattice as,
\begin{equation}
\label{affstatei}
\left( \begin{array}{c}
|i\mu_+\rangle \\
|i\nu_+\rangle
\end{array} \right) =
\left(\begin{array}{cc}
\cos\theta_+ & \sin\theta_+ \\
-\sin\theta_+ & \cos\theta_+
\end{array} \right)
\left( \begin{array}{c}
|iz\rangle \\
|ix\rangle
\end{array} \right) ,
\end{equation}
and for $j\in B$ sublattice as,
\begin{equation}
\label{affstatej}
\left( \begin{array}{c}
|j\mu_-\rangle \\
|j\nu_-\rangle
\end{array} \right) =
\left(\begin{array}{cc}
\cos\theta_- & -\sin\theta_- \\
\sin\theta_- & \cos\theta_-
\end{array} \right)
\left( \begin{array}{c}
|jz\rangle \\
|jx\rangle
\end{array} \right) .
\end{equation}
By choosing the angles $\theta_+$ and $\theta_-$ at the values which
minimize the classical energy, given by Eqs. (\ref{thetamoaff1}) and
(\ref{thetamoaff1}), we guarantee that $|i\mu_+\rangle$ and $|j\mu_-\rangle$,
respectively, are at each site the orbital state realized in the classical
MO{\scriptsize FFA} phase, and one may easily define the new excitation
operators with respect to the symmetry-breaking which occurs in this phase;
they are analogous to those given in Eqs. (\ref{newspin1})--(\ref{newsping2}).
Next, the Hamiltonian is rotated to the new representation as described in
Sec. \ref{sec:moffarpa}. We do not present an explicit form of the
spin-orbital Hamiltonian (\ref{somcu}) in this case, as it may be obtained
from Eqs. (\ref{hparalong})--(\ref{kk3long}) by replacing the angle $\theta$
by $\theta_+$ and $\theta_-$ for the sublattice $A$ and $B$, respectively.
Furthermore, due to the degeneracy between the MO{\scriptsize AFF} and
MO{\scriptsize FAF} phases, we had to average the crystal-field between the
two sublattices in the actual calculation.
We have verified that the transverse excitations have a similar dependence
on the ${\vec k}$-vector to those found in the MO{\scriptsize FFA} phase,
and we show the representative data in Fig. \ref{modesaff}. For convenience,
we have rotated the BZ and use just the same notation as in Fig.
\ref{modesffa}. The value of the crystal-field $E_z$ is in the present case
effectively smaller by a factor of two in comparison with the
MO{\scriptsize FFA} phase. This asymmetry is a consequence of the choice of
$|x\rangle$ and $|z\rangle$ states as the orbital basis.
One finds again that the spin-wave and spin-and-orbital-wave excitations are
well separated along the $\Gamma-X-S-\Gamma$ path, and the gap between them
has increased to $\sim 1.2J$. We note a stronger renormalization of the
low-energy modes which follows from weakened FM interactions between the
alternating orbitals in the $(b,c)$-planes in the present case as compared
with those within the $(a,b)$-planes in the MO{\scriptsize FFA} phase.
Although the orbital excitations are still well separated from the remaining
transverse modes, their dispersion is larger than that in Fig. \ref{modesffa}.
\section{ Quantum fluctuations }
\label{sec:rpa}
The size of quantum fluctuation corrections to the classical order
parameters determines the stability of the classical phases. As mentioned
in Sec. \ref{sec:orbitals}, frustration of magnetic interactions leads in
spin models to divergent quantum corrections within the LSW theory. Before
calculating these corrections in the present situation, a generalization
of the usual RPA procedure to a system with several excitations is necessary.
Here we present only the relations needed to calculate the quantum
corrections to the LRO parameter and ground state energy, while more details
will be reported separately.\cite{rvb99}
For that purpose, let us denote here the local operators constituting
the $so(4)$ Lie algebra at site $i$ as Hubbard operators,
$X_i^{\alpha\beta}=|i\alpha\rangle \langle i\beta|$. Using the unity
operator, $\sum_{\beta}X_i^{\beta\beta}=\openone$,
the diagonal operator that refers to the state $|i\alpha\rangle$
{\em realized at site $i$ in the classical ground state\/} of the phase
under consideration may be expanded in terms of the excitation operators,
\begin{equation}
\label{expex}
X_i^{\alpha\alpha}=\openone
-\sum_{\beta\neq\alpha}X_i^{\beta\alpha}X_i^{\alpha\beta},
\end{equation}
while the diagonal operators referring to an {\em excited\/} state
$|i \beta\rangle$ are expressed as
\begin{equation}
\label{expexb}
X_i^{\beta\beta}=X_i^{\beta\alpha}X_i^{\alpha\beta}.
\end{equation}
Applying these equations to the $z$-th spin component
$S^z_i=S^z_{ixx}+S^z_{izz}$ of the total spin at site $i$ in one of the AF
phases with pure orbital character (say AFxx for definiteness), one finds,
for $i$ in the spin-up sublattice,\cite{noteexp}
\begin{eqnarray}
\label{exps}
S_i^z &=& \case{1}{2} (X_i^{x\uparrow,x\uparrow}
- X_i^{x\downarrow,x\downarrow}
+ X_i^{z\uparrow,z\uparrow}
- X_i^{z\downarrow,z\downarrow} ) \nonumber \\
&=& \case{1}{2} \openone
- X_i^{x\downarrow,x\uparrow} X_i^{x\uparrow,x\downarrow}
- X_i^{z\downarrow,x\uparrow} X_i^{x\uparrow,z\downarrow}
\nonumber \\
&=& \case{1}{2} \openone
- S_{ixx}^- S_{ixx}^+ - K_{izx}^- K_{ixz}^+ .
\end{eqnarray}
Taking the average one obtains, with the MF value
$\langle S_i^z\rangle=\frac{1}{2}$,
\begin{eqnarray}
\label{expes}
\langle S_i^z\rangle_{\rm RPA}&=&
\case{1}{2}- \langle S_{ixx}^- S_{ixx}^+\rangle
- \langle K_{izx}^-K_{ixz}^+\rangle \nonumber\\
&=&\case{1}{2}- \langle S_{i}^- S_{i}^+\rangle
- \langle K_{i}^-K_{i}^+\rangle \nonumber\\
&=&\langle S_i^z\rangle-\delta\langle S_i^z\rangle,
\end{eqnarray}
where the second equality is valid because averages like
$\langle S_{ixx}^-S_{izz}^+\rangle$ are zero since they involve `ghost'
modes, so that one may formally replace $S_{ixx}^+$ by $S_{ixx}^+
+S_{izz}^+=S_i^+$, {\it etcetera\/}. The first contribution
$\propto\langle S_i^-S_i^+\rangle$ is the usual renormalization due to spin
waves, while the second term $\propto\langle K_i^-K_i^+\rangle$ stands for
the reduction of $\langle S_i^z\rangle_{\rm RPA}$ due to
spin-and-orbital-wave excitations. Both terms involve a local excitation
preceded by a deexcitation which reproduces the initial local state.
As expected only the transverse excitations contribute to the spin
renormalization. Note that, since Eq. (\ref{exps}) is an {\em exact operator
relation\/}, the present procedure gurantees that Eq. (\ref{expes}) is a
{\em conserving approximation\/} which respects the sum rule for the
occupancies of all states, $\sum_{\beta}\langle X_i^{\beta\beta}\rangle=1$.
The generalization of Eq. (\ref{expes}) to the MO phases using the
operators (\ref{newspin1}), (\ref{newspin2}), or to other order
parameters, like the orbital polarization, is obvious.
The local correlation functions which renormalize the order parameter in
Eq. (\ref{expex}) are determined in the standard way,\cite{Hal72}
\begin{equation}
\label{theorem}
\langle B_i^{\dagger}A_i\rangle=
\frac{1}{N}\sum_{\vec k}\int_{-\infty}^{+\infty}d\omega
{\cal A}_{AB^{\dagger}}({\vec k},\omega)\frac{1}{\exp(\beta\omega)-1},
\end{equation}
where $\beta=1/k_BT$, and
\begin{eqnarray}
\label{weight}
{\cal A}_{AB^{\dagger}}({\vec k},\omega<0)&=&
2{\rm Im}\langle\langle A_{\vec k}|
B^{\dagger}_{\vec k}\rangle\rangle_{\omega-i\epsilon} \nonumber \\
&=&\sum_{\nu<0}{\cal A}_{AB^{\dagger}}^{(\nu)}({\vec k})
\delta(\omega-\omega_{\vec k}^{(\nu)})
\end{eqnarray}
is the respective spectral density for the negative frequencies ($\nu<0$),
and ${\cal A}_{AB^{\dagger}}^{(\nu)}({\vec k})$ are the respective spectral
weights. Therefore, the correlation functions at $T=0$ are found by summing
up the total spectral weight at the negative frequencies,
\begin{equation}
\label{average}
\langle B_i^{\dagger}A_i\rangle=\frac{1}{N}\sum_{\vec k}\sum_{\nu<0}
{\cal A}_{AB^{\dagger}}^{(\nu)}({\vec k}).
\end{equation}
As we show elsewhere,\cite{rvb99} the Hamiltonian of the spin-orbital model
(\ref{somcu}) may be expanded in RPA in terms of the excitation and
deexcitation operators,
\begin{equation}
\label{exph}
{\cal H}\simeq {\cal H}_{\rm MF}+{\cal H}_{\rm RPA},
\end{equation}
where ${\cal H}_{\rm MF}$ is given by Eq. (\ref{somcumf}), and
\begin{eqnarray}
\label{exprpa}
{\cal H}_{\rm RPA}&=&
\sum_{i\in A}\sum_{\mu\mu'}X_i^{\mu\alpha}a^{\mu\mu'}_AX_i^{\alpha\mu'}
+\sum_{j\in B}\sum_{\nu\nu'}X_i^{\nu \beta}a^{\nu\nu'}_BX_i^{ \beta\nu'}
\nonumber \\
&+&\sum_{\langle ij\rangle}\sum_{\mu\nu}
\left( X_i^{\mu\alpha}b_{ij}^{\mu\nu}X_j^{ \beta\nu}
+X_i^{\alpha\mu}b_{ij}^{\mu\nu}X_j^{\nu \beta}\right)
\nonumber \\
&+&\sum_{\langle ij\rangle}\sum_{\mu\nu}
\left( X_i^{\alpha\mu}c_{ij}^{\mu\nu}X_j^{ \beta\nu}
+X_i^{\mu\alpha}c_{ij}^{\mu\nu}X_j^{\nu \beta}\right)
\end{eqnarray}
for a two-sublattice phase (the generalization to the four-sublattice MO
phases is straightforward). The MF part describes the classical problem which
was discussed in Sec. \ref{sec:mfa}. The RPA part (\ref{exprpa}) describes
the many-body problem in a linear approximation, with the fixed indices
$\alpha$ and $\beta$ referring to the symmetry-broken state at site $i$ and
$j$, respectively. This expansion leads, after changing the order of
excitation operators $X_i^{\alpha\beta}$ to normal order, and after making
straightforward transformations, to a compact expression for the average
energy contribution per site,
\begin{eqnarray}
\label{erpa}
E_{\rm RPA}&=&\frac{1}{N}\langle {\cal H}_{\rm RPA}\rangle \nonumber \\
&=&\frac{1}{4}\left[ - {\rm Tr}\{A\}
+\sum_{\nu>0}\frac{2}{N}\sum_{\vec k}\omega_{\vec k}^{(\nu)}\right],
\end{eqnarray}
where $A$ is the matrix of positive on-site coefficients $a_A^{\mu\mu'}$,
$a_B^{\nu\nu'}$, appearing in the first line of Eq. (\ref{exprpa}), and
with the sum running over all modes with positive frequencies (counting
doubly-degenerate modes twice) in the reduced BZ. This expression is seen
to be a direct generalization of the familiar result for the HAF, the
distinction being that more modes contribute here, and so Eq. (\ref{erpa})
represents the energy gain ($E_{\rm RPA}<0$) due to the reduction in
zero-point energy of the propagating modes in comparison with that of the
local excitations. We use Eq. (\ref{erpa}) to calculate the total energy in
RPA,
\begin{equation}
\label{etotal}
E=E_{\rm MF}+E_{\rm RPA}.
\end{equation}
Before discussing the renormalization of the order parameter and the
corresponding energies in RPA, we concentrate ourselves on the behavior of
the transverse excitations when the crossover lines between the classical
phases are approached. As already emphasized in Sec. \ref{sec:magnons}, the
spin-wave and spin-and-orbital-wave excitations couple. As a consequence,
the modes in all considered phases {\em soften\/} when the transition lines
between different classical phases, or classical degeneracy point are
approached. This softening is shown for a representative value of $J_H/U=0.3$
in Fig. \ref{swafreal} for the two AF phases. In the AFxx phase the energy
scales of both excitations are separated for $E_z>4J$, while the
spin-and-orbital mode moves towards zero energy with decreasing $E_z$, and
finally becomes soft at the $X$ point, along $\vec{k}=(\pi,0,k_z)$ and along
equivalent lines in the BZ for $E_z\simeq 1.54J$. A similar mode
softening is found for the AFzz phase at $E_z<0$, with the soft mode along
$\Gamma-X$ and equivalent directions in the BZ at $E_z\simeq -1.84J$. This
pecular softening along lines and not at points in the BZ shows that the
modes behave 2D-like instead of 3D-like: constant-frequency surfaces are
cilinders contracting towards lines, not spheres contracting towards a point.
By making an expansion of Eq. (\ref{afsw}) around the soft-mode lines, one
finds that the (positive) excitation energies are characterized by
{\em finite} masses in the perpendicular directions:
\begin{equation}
\label{massx}
\omega_{\rm AFxx}(\vec{k}) \rightarrow \Delta_x
+ B_x \left( {\bar k}_x^4+14{\bar k}_x^2k_y^2+k_y^4 \right)^{1/2},
\end{equation}
independently of $k_z$ (here ${\bar k}_x=k_x-\pi$), and
\begin{equation}
\label{massz}
\omega_{\rm AFzz}(\vec{k}) \rightarrow \Delta_z
+ B_z \left( k_y^2 + 4k_z^2 \right),
\end{equation}
independently of $k_x$, and similarly along the $\Gamma-Y$ direction with
$k_y$ replaced by $k_x$. As an example we give explicit expressions for the
AFxx phase at $\eta=0$,
\begin{equation}
\label{massxx}
\Delta_x= \frac{9}{2} \frac{\varepsilon_z}{\varepsilon_z+3},
\hskip 1.2cm
B_x=\frac{27}{16}\frac{ 1 }{\varepsilon_z+3},
\end{equation}
where one finds that the gap $\Delta_x\to 0$ when $\varepsilon_z \to 0$,
i.e. upon approaching the $M=(E_z,H_H)=(0,0)$ point at which
the AF order is changed to the AFzz phase. This illustrates a general
principle: $\Delta_i\to 0$ when the crossover line to another phase is
approached, and $B_i\neq 0$ when the modes (\ref{massx}) and (\ref{massz})
soften, making quantum fluctuation corrections to the order parameter
to diverge logarithmically, $\langle\delta S\rangle\sim
\int d^3k/\omega(\vec{k})\sim\int d^2k/(\Delta_i+B_ik^2)\sim\ln\Delta_i$.
We emphasize that for the occurrence of this divergence not only the
finiteness of the mass but also the 2D-like nature of the dispersion is
essential. It enables a 3D system to destabilize LRO by what are essentially
2D fluctuations. So the divergence of the order parameter near the cross-over
lines in the phase diagram and the associated instability of the classical
phases, may be regarded as another manifestation of the effective reduction
of the dimensionality occurring in the spin-orbital model.
We do not present explicitly the softening of the longitudinal modes
which also happens at the transition lines.
A seemingly attractive way to simplify the calculation of the transverse
excitations would be to make a decoupling of the spin-waves and
spin-and-orbital-waves. However, this is equivalent to violating the
commutation rules between the spin and spin-and-orbital operators in
Appendix \ref{sec:commute},\cite{Fei98} and this changes the physics. It
gives the same excitation energies as Eq. (\ref{afsw}), but with
$P_{\alpha\vec k}=0$; the numerical result is given in Fig. \ref{swafpoor}.
Of course, the spin-wave excitation does not depend then on the orbital
splitting $E_z$, and the spin-and-orbital-wave excitation gradually
approaches the line $\omega_{\vec k}=0$ with decreasing $|E_z|$. It has a
weak dispersion which depends on $J_H$ and on the value of $|E_z|$, and
gives an instability at the $\Gamma$ point only, not at lines in the BZ, and
in the phase diagram well beyond the transition lines of Fig. \ref{mfa3d},
i.e., within the MO{\scriptsize FFA} and MO{\scriptsize AFF} phase for
$E_z<0$ and $E_z>0$, respectively. Such spin-wave and spin-and-orbital-wave
modes give, of course, much smaller quantum corrections of the order parameter
and energy than the correct RPA spectra of Fig. \ref{swafreal}.\cite{Fei98}
The spin-waves in the MO{\scriptsize FFA} phase, stable at $E_z<0$, soften
with decreasing $\eta$ (\ref{jh}), as shown in Fig. \ref{swmoffa}. At large
$\eta$ the spin-and-orbital-waves at high energies are well separated from
the spin-wave modes. The latter have a rather small dispersion at $J_H/U=0.3$
which follows from relatively weak FM interactions in the $(a,b)$ planes, and
AF interactions along the $c$-axis. The modes start to mix stronger with
decreasing $\eta$, and finally the gap in the spectrum closes below
$\eta=0.1$. The mode softening occurs again along lines in the BZ, namely
along the $\Gamma-X$ direction. Unfortunately, we could not perform an
analogous analytic expansion of the energies near the softening point to that
in the AFxx and AFzz phases, but the numerical results reported here suggest
a qualitatively similar behavior to these two phases. The MO{\scriptsize AFF}
phase gives an analogous instability at $E_z>0$.
The soft modes in the excitation spectra give a very strong renormalization
of the order parameter $\langle S^z\rangle_{\rm RPA}$ in RPA (\ref{expes})
near the mode softening, as shown in Fig. \ref{szreal}. The quantum
corrections {\em exceed\/} the MF values of the order parameter in the AFxx
and AFzz phases in a region which separates these two types of LRO. Although
one might expect that another classical phase with mixed orbitals and FM
planes sets in instead, and the actual instabilities where
$\delta\langle S_z\rangle\to\infty$ are found indeed beyond the transition
lines to another phase, the lines where
$\delta\langle S^z\rangle=\langle S^z\rangle$ occur still {\em before\/} the
phase boundaries in the phase diagram of Fig. \ref{mfa3d} (see Fig. 1 of Ref.
\onlinecite{Fei97}). This leaves a window where {\em no classical order is
stable\/} in between the G-AF and A-AF spin structures.
The origin of such a strong renormalization of $\langle S^z\rangle$ may be
better understood by decomposing the quantum corrections into individual
contributions as given in Eq. (\ref{expes}) (see Table I). The leading
correction comes from the local spin fluctuation expressed by
$\langle S^-_iS^+_i\rangle$ and enhanced with respect to the the pure spin
model (HAF), while the spin-and-orbital fluctuation,
$\langle K^-_iK^+_i\rangle$, increases rapidly when the instability lines
$\langle S^z\rangle_{\rm RPA}=0$ are approached. Interestingly, the latter
fluctuation is stronger in the AFxx than in the AFzz phase for the same
values of $J_H$ and $|E_z|$ which demonstrates that the AFzz phase is more
robust due to the directionality of the $|z\rangle$ orbitals and the strong
AF bonds along the $c$-axis. This asymmetry is also visible in Fig.
\ref{szreal}, where $\langle S^z\rangle_{\rm RPA}$ decreases somewhat faster
towards zero for $E_z>0$.
In both G-AF phases (AFxx and AFzz) the leading contribution to the
renormalization of $\langle S^z\rangle_{\rm RPA}$ comes from the lower-energy
mode, especially at larger values of $J_H$. In the case of $J_H=0$ one finds,
however, that the contribution from the lower mode either stays approximately
constant (in the AFxx phase), or even decreases (in the AFzz phase) when the
line of the collapsing LRO is approached at $|E_z|\to 0$ (Table I). This
latter behavior shows again that the coupling between the spin-wave and
spin-and-orbital-wave excitations is of crucial importance.\cite{Fei98}
This is further illustrated by Fig. \ref{szpoor}, which shows the
renormalization of $\langle S_z \rangle$ as obtained when spin waves and
spin-and-orbital waves are decoupled in the manner discussed above.
One observes that significant reduction of $\langle S_z \rangle$ then sets
in only very close to the actual divergence.
Also the orbital polarization is renormalized by the quantum fluctuations,
but this is a rather mild effect not showing any instability, since this
renormalization involves only the spin-and-orbital and the orbital
excitation but not the spin excitation, which is the one participating most
strongly in the lowest transverse mode that goes soft. This is seen in Fig.
\ref{nxholes}, where we show $\langle n_x \rangle$, the occupation of the
$|x\rangle$ orbital, again for $J_H/U=0.3$, both at the MF level as well as
including the RPA quantum fluctuations, calculated from an expression
similar to Eq. (\ref{expes}), e.g. in the AFxx phase from
\begin{equation}
\label{expnx}
\langle n_{ix}\rangle = 1-4 \, \langle T_{izx}T_{ixz}\rangle
- \langle K^-_i K^+_i\rangle.
\end{equation}
Especially in the MO{\scriptsize FFA} and MO{\scriptsize AFF} phases the
deviation from the classical value of $\theta$ as given by Eq.
(\ref{thetamoffa}) and by Eqs. (\ref{thetamoaff1}) and (\ref{thetamoaff2}),
respectively, is small. Only in the AFxx phase a significant admixture of
$|z\rangle$ occupancy could occur close to the regime where this phase
becomes unstable due to the divergence of $\langle S^z\rangle_{\rm RPA}$.
The reduction of $\langle S^z\rangle_{\rm RPA}$ in the
MO{\scriptsize FFA}/MO{\scriptsize AFF} phases (Table II), described by a
relation similar to Eq. (\ref{expes}), is in general weaker than that in the
G-AF phases. This is understandable, as the quantum fluctuations contribute
here only from a single AF direction, while the FM order in the planes does
not allow for excitations which involve spin flips and stabilizes the LRO
of A-AF type. For fixed $J_H$ one finds increasing quantum corrections
$\delta\langle S^z\rangle$ when the lines of phase transitions towards the
AF phases are approached. These corrections increase faster with increasing
$|E_z|$ in the MO{\scriptsize FFA} phase, as the increasing occupancy of the
$|z\rangle$-orbital makes the AF interaction stronger there than in the
MO{\scriptsize AFF} phase, where the occupancy of the $|x\rangle$ orbital
increases slower roughly by a factor of two. This qualitative difference
between these two A-AF phases may be seen in Fig. \ref{nxholes}. As in the
G-AF phases, we find that the two lower-energy modes give the larger
contribution to the renormalization of the order parameter. The
spin-and-orbital fluctuation $\langle {\cal K}^-_i{\cal K}^+_i\rangle$
remains almost independent of $E_z$, but increases with decreasing values of
$J_H$. Thus we conclude that the collapse of the LRO in the A-AF (MO) phases
is primarily due to increasing spin fluctuations,
$\langle {\cal S}^-_i{\cal S}^+_i\rangle$, while the spin-and-orbital
fluctuations become of equal importance only when the multicritical point of
the Kugel-Khomskii model $M=(E_z,J_H)=(0,0)$ is approached.
The representative quantum corrections to the ground state energy are given
in Table III. First of all, these corrections are larger by roughly a factor
of two in the G-AF phases (AFxx and AFzz) than in the A-AF phases
(MO{\scriptsize FFA} and MO{\scriptsize AFF}/MO{\scriptsize FAF}). We believe
that this is a generic difference between the quantum corrections in the
A-type and G-type AF phases, with the latter stabilized more due to the spin
fluctuations contributing at all the bonds. Therefore, the G-AF phases win
over the A-AF ones near the transition lines, as for example found at
$J_H/U=2.0$ and $E_z/J=0.2$. However, one should keep in mind that the energy
alone does not suffice for the stability of a particular phase in RPA, since
the MF value of the order parameter, $\langle S^z\rangle$, has to remain
larger than the respective quantum correction, $\delta\langle S^z\rangle$.
Secondly, the 2D AFxx phase is characterized by larger quantum corrections
than the strongly anisotropic AFzz phase at the same values of $J_H/U$ and
$|E_z|/J$. The same observation was made before at the multicritical point
$M=(E_z,J_H)=(0,0)$.\cite{Fei98} This is not surprising since the 2D HAF is
already quite close to the disordered spin state. We note that the energy
gain due to quantum fluctuations of $0.423J$ (obtained for the actual
interactions of $\frac{9}{4}J$ in a 2D HAF) is there considerably smaller
than the values of $\delta E$ of the order of $0.65J$ reported in Table III.
Finally, we note that the dominating contribution to the quantum corrections
to the energy comes from the transverse excitations. The longitudinal
excitations do not contribute at all at $J_H/U=0$, where these modes are
dispersionless. Otherwise, the orbital excitations have always a
significantly smaller dispersion than the value of the orbital gap in the
spectrum, and the resulting quantum corrections are therefore almost
negligible.
\section{ Summary and conclusions }
\label{sec:spinliquid}
Summarizing, we have presented here the case that a generic (Kugel-Khomskii)
model for the dynamics of an orbitally degenerate MHI is characterized by a
number of peculiar features. In this paper we have followed a semi-classical
strategy. Assuming that the ground state exhibits some particular classical
spin- and orbital order, the stability of this order can be
investigated by considering the Gaussian fluctuations around this state.
In this way we find that in various regimes of the zero-temperature
phase-diagram, conventional order is defeated by the quantum fluctuations,
and we expect a qualitative phase diagram as shown in Fig. \ref{artistic}.
In the first place, near the transition lines between the different phases
modes soften, and these soft modes cause the zero-point fluctuations to
diverge. This is not dissimilar from the general theme associated with
the geometrically frustrated quantum spin-models, like the $J_1-J_2-J_3$
model.\cite{Cha88} A significant difference is that in the present case the
source of the problems is distinct: it is associated with the difficulty to
simultaneously satisfy the requirements for a stable spin- and orbital order.
The cause of the frustration is dynamical instead of geometrical.
The most interesting feature is the point at the origin of the phase diagram.
On the classical level it is a point in the zero-temperature phase diagram
where a quasi-1D antiferromagnet (MO{\scriptsize FFA} phase), a 2D
antiferromagnet (AFxx phase), and a mildly anisotropic 3D antiferromagnet
(AFzz phase) become degenerate (Fig. \ref{mfa3d}). In fact, these
possibilities make up only an infinitisimal fraction of the total degeneracy
characterizing this special point. In addition, the orbitals can be freely
rotated on every site, if the spins form a 3D antiferromagnet. Likewise, the
phase diagram of Fig. \ref{artistic} is highly incomplete. Next to $E_z$,
there exist an infinity of other axes emerging from this special point, all
corresponding with distinct ways of explicit {\em local} symmetry breaking
in the orbital sector. One can either call this point an infinite-critical
point, or a point of perfect dynamical frustration, or a point where local
symmetry is dynamically generated.
The obvious problem is that the above wisdom applies only when
quantum-mechanics does not play a role. Physical reality is different, and
since the classical limit is pathological, quantum-mechanics is bound to
take over. Although we have not found a way to make the case precise, it
appears to us that the local symmetry referred to in the previous paragraph
exists only in the classical limit. For this to be active on the quantum
level, it should be that the true ground state is also highly degenerate.
Although we did not prove the uniqueness of the quantum ground state, so
much is clear that the classical local symmetry gets lifted at the moment
that quantum fluctuations become significant: the cancellations occur only
if the spins are fully classical. Regardless the nature of the true ground
state, it is generated by a quantum order-out-of-disorder
mechanism.\cite{Chu91}
The first possibility is a straightforward order-out-of-disorder physics:
the quantum fluctuations affect the energies of the various classical states
in different ways, thereby breaking the classical degeneracy. One of the
saddle points might get uniquely favored and this is what is suggested in
Ref. \onlinecite{Kha97}, where it was argued that the AFzz phase becomes the
ground state at the origin of the phase diagram. Although this is a credible
possibility, one would have to demonstrate that the other possibilities are
less favoured, and moreover, we have showed elsewhere\cite{Fei98} that the
actual calculation by Khaliullin and Oudovenko\cite{Kha97} is flawed.
The case is still open.
Yet another possibility is unconventional spin- and orbital order which is
in a sense dual to the orbital- and spin (anti)ferromagnetism characterizing
the `classical' order: spin-orbital (resonating) valence bond (R)VB states.
We demonstrated before\cite{Fei97} that these straightforward
generalizations of the spin RVB states, well known from the study of quantum
spin-problems, appear as exceptionally stable. In a next publication we will
further elaborate on these matters.\cite{rvb99}
The status of both proposals is rather unsure: they rely at best on the
variational principle and the true vacuum can still be completely different.
In this regard, some recent experiments on the system LiNiO$_2$ are quite
interesting.\cite{Kit98} In this material a Mott-insulator seems to be
realized, characterized by a low spin ($S=1/2$) $e_g$ degenerate Ni(III)
state. One would naively expect this system to be unstable towards a
collective Jahn-Teller distortion, accompanied by spin ordering. This indeed
happens in the closely related system NaNiO$_2$, but in LiNiO$_2$ ordering
phenomena are completely absent,\cite{Hir85} a peculiarity pointed out long
ago.\cite{Bon57} Instead, some quantum-critical
state appears to be present, characterized by power-law behavior of physical
quantities, carrying unusual exponents. Pending the magnitude of the
Li-mediated kinetic exchange ($J_{\rm Li}$), one can view this system as
either disconnected triangular layers of Ni(III) ions (vanishing
$J_{\rm Li}$), or as interpenetrating cubic lattices of these ions which are
described by the Kugel-Khomskii Hamiltonian (large $J_{\rm Li}$).\cite{Fei97}
Hence, the peculiar state seen in the experiments can either originate in
some phenomenon associated with the triangular layers,\cite{whafnium} but
it could also be related to the matters discussed in this paper.
It is easy to settle this issue experimentally. Compare NaNiO$_2$ and
LiNiO$_2$; if the physics of the quantum disorder in the latter has to do
with the (111) layers, one would expect on general grounds that in order to
stabilize an ordered state, the effective dimensionality has to be increased,
of course assuming that the basics of the electronic structure (such like
covalency) do not change appreciably. Hence, in this layer scenario one
would expect stronger layer-layer interactions in NaNiO$_2$ as compared to
LiNiO$_2$, following the standard result of quantum field theory that
fluctuations increase upon lowering dimensionality. This standard wisdom
does not apply to the Kugel-Khomskii model, however. The fluctuations find
their origin in a dynamical frustration, and this frustration is only
present in three space dimensions. Hence, if the disorder in LiNiO$_2$ is
caused by the physics discussed in this paper, its quantum magnetism should
be rather isotropic in 3D space, while NaNiO$_2$ should be more 2D. It is
noticed that according to elementary quantum chemistry Li ions should be
more effective in mediating kinetic exchange than Na ions.
\acknowledgements
We thank P. Horsch, D. I. Khomskii, J. Richter and M. Takano for valuable
discussions. A.M.O. acknowledges the support by the Committee of Scientific
Research (KBN) of Poland, Project No. 2~P03B~175~14.
\end{multicols}
|
\section{Introduction}
There is an increasing evidence for the fact that $M$-theory on anti de
Sitter (AdS) backgrounds can be described by a conformal field theory
at the boundary of AdS, at least in a suitable limit. The low energy
limits involved in the bulk of AdS are the gauged supergravity theories
in various dimensions and the boundary field theories are certain
globally supersymmetric field theories appropriate to the branes
involved.\\
In verifying the AdS/CFT correspondence, the boundary values of the
bulk fields naturally arise. In accordance with the fact that AdS
supersymmetry in a given dimension acts as the conformal
supersymmetry at the boundary of AdS, the boundary values of the bulk
fields are in one-to-one correspondence with the fields of conformal
supergravity defined at the boundary. Thus, it is natural to formulate
the boundary field theory in a conformal supergravity background.
Integration over the boundary (matter) fields should then yield an
effective action involving the conformal supergravity fields, which is
to be compared with the bulk supergravity effective action.\\
This approach was followed in \cite{liu1}, where the coupling of
$N=4,D=4$ super Yang-Mills to $N=4,D=4$ conformal supergravity
\cite{roo1} was studied, as the boundary field theory associated with
gauged supergravity in $AdS_5$. In this spirit, we wish to construct
the coupling of the $(2,0)$ tensor multiplet to $(2,0)$ conformal
supergravity at the six dimensional boundary of $AdS_7$. This result is
expected to provide a convenient framework in studying the $AdS_7/CFT_6$
correspondence. Since the $(2,0)$ tensor multiplet contains a chiral
$2$-form, it is natural to study its field equations rather than an
action from which they may be derivable but which may require the
introduction of additional fields. Thus, we shall primarily study the
covariant field equations, although we shall briefly discuss an action
from which all but the self-duality condition follows, provided that the
self-duality equation is imposed after the variation.\\
The conformal supergravity fields form an off-shell multiplet. We can
treat them as background fields, in which case we need not impose their
equations of motion. However, coupling of conformal supergravity to
$N+5$ copies of the $(2,0)$ tensor multiplet, we can constrain the
scalars fields coming from these tensor multiplets so that they become
a representative of the coset ${SO(N,5)\over SO(N)\times SO(5)}$. As we
will show, this leads to a conformal interpretation of the
$(2,0)$ Poincar\'e supergravity coupled to $N$ tensor multiplets
constructed previously in \cite{romans,ric}.\\
The organization of this paper is as follows. In Sec. 2, we derive the
multiplet of supercurrents and the linearized Weyl multiplet. In Sec.
3, we construct the full $(2,0)$ conformal supergravity theory. As we
did before for the $(1,0)$ case \cite{ber1}, we follow the methods
developed first for $N=1$ in 4 dimensions \cite{superconformN1}. They
are based on gauging the conformal superalgebra \cite{SU221}, which, in
our case, is $OSp(8^*|4)$. As is typical in this method, one then has
to impose constraints on some of the curvatures. In Sec. 4, we find the
complete equations of motions for a single tensor multiplet in a
conformal supergravity background. In Sec. 5, we consider $N+5$ copies
of the tensor multiplet in conformal supergravity background and show
that a geometrical constraint on the scalars leads to the equations of
motion of Poincar\'e supergravity coupled to $N$ tensor multiplets. Further
comments on our results and open problems are collected in the
Conclusions. Our notations and conventions are presented in Appendix A
and the truncation of our results to the $(1,0)$ case \cite{ber1} are
described in Appendix B, as we used that correspondence to obtain our
present results. A superspace description of the tensor, current and
Weyl multiplets is given in Appendix C.
\section{ The $(2,0)$ Supercurrent Multiplet}\la{tm}
In this section we will construct the $(2,0)$ supercurrent multiplet.
Using the invariance of the bilinear couplings between the
currents and the corresponding fields we will derive the {\it linearized}
transformation rules of the (2,0) conformal supergravity multiplet.
In the next section we will extend this to the nonlinear case.\\
Our starting point is the $(2,0)$ tensor multiplet in $D=6$ Minkowski
spacetime describing $8+8$ degrees of freedom\footnote{
In this section the tensor multiplet will only play an auxiliary role
as a means to construct the supercurrent multiplet. Later, in
section~\ref{tm2}, (2,0) tensor multiplet will be introduced as matter multiplets
to be coupled to the conformal supergravity theory constructed in section~\ref{ss:Weylm}.}.
This is the only
on-shell $(2,0)$ matter multiplet in $D=6$. Its field components are
given in Table~\ref{tbl:tensorMultiplet}. It contains a $2$-form
potential $B_{\mu\nu}$ whose self-dual field strength is defined by
\begin{equation}
{H}_{\mu\nu\rho} = 3\partial_{[\mu}B_{\nu\rho]}\ ,
\quad\quad
H_{\mu\nu\rho} = \ft1{3!} \epsilon_{\mu\nu\rho\sigma\lambda\tau}\, H^{\sigma\lambda\tau}\equiv
\tilde H_{\mu \nu \rho }\ .
\label{defHdual}
\end{equation}\\[-.75cm]
It also contains $5$ real scalars $\phi^{ij}\, (i=1,...,4)$ which
transforms as a $5$-plet of $USp(4)$ and a symplectic Majorana-Weyl
spinor $\psi^i$. The basic properties of these fields are tabulated in
Table 1.\\
\fa
The complete rigid superconformal transformations have been given in
\cite{cla1} where the tensor multiplet was studied as the $M5$-brane
worldvolume supermultiplet. For our present purposes we only need the
linearized rigid $Q$-supersymmetry transformation rules
\begin{eqnarray}
\delta B_{\mu\nu} &=& - {\bar\epsilon}\gamma_{\mu\nu}\psi \ ,
\nonumber\w2
\delta\psi^i &=& {\textstyle{1\over 48}} H^{+}_{\mu\nu\rho} \gamma^{\mu\nu\rho} \epsilon^i
+{\textstyle{1\over 4}}\not\! \partial \phi^{ij} \epsilon_j\ ,
\la{tr}
\w2
\delta \phi^{ij} &=& -4 {\bar\epsilon}^{[i}\psi^{j]} -\Omega^{ij}\bar \epsilon \psi \ .
\nonumber
\end{eqnarray}\\[-.75cm]
The self-dual part of the curvature $H^+_{abc}$ transforms as
\begin{equation}
\delta H_{abc}^{+} = -\textstyle{1\over 2}{\bar \epsilon}\not \!\del \gamma_{abc}\psi\ .
\end{equation}\\[-.75cm]
The supersymmetry transformations close provided that the following
linearized field equations are satisfied
\begin{equation}
H^- = \not \!\del \psi^i = \partial^\mu\partial_\mu \phi^{ij} = 0\ .
\la{fe}
\end{equation}\\[-.75cm]
To construct the current multiplet, we start from the Noether currents
of the $(2,0)$ tensor multiplet. These Noether currents are: the
energy-momentum tensor $\theta_{\mu\nu}$, the supersymmetry currents
$J_{\mu i}$ and the ${\rm USp(4)}$ currents $v_\mu^{ij}$. We will use the
improved currents that satisfy the following equations:
\begin{eqnarray}
&& \partial^\mu\theta_{\mu\nu} = 0\ ,\hskip 1.5truecm \theta_{\mu\nu}=\theta_{\nu\mu}\ ,
\hskip 1.5truecm \theta^\mu{}_\mu =0\ ,
\nonumber
\w2
&& \partial^\mu J_{\mu i}= 0\ , \hskip 1.2truecm \gamma^\mu J_{\mu i} = 0\ ,
\nonumber
\w2
&& \partial^\mu v_{\mu }^{ij} = 0\, .
\la{conditions}
\end{eqnarray}\\[-.75cm]
These symmetry properties determine the currents up to constants, which
we have determined by requiring the closure of the rigid linearized
supersymmetry transformations. We thus find the currents
\begin{eqnarray}
\la{currents}
\theta_{\mu\nu} &=& H^+_{\mu ab}H^+_{\nu}{}^{ab}
+ 8 {\bar \psi}\gamma_{(\mu}\partial_{\nu)}\psi + \partial_\mu
\phi^{ij}\partial_\nu\phi_{ij}-\ft1{10} \eta_{\mu\nu} \left
(\partial\phi\right )^2 -{\textstyle{1\over 5}}
\partial_\mu\partial_\nu \phi^2\ ,
\nonumber\w2
J_{\mu i} &=& -\ft23 H^+_{\rho\sigma\tau} \gamma^{\rho\sigma\tau}\gamma_\mu\psi_i
+8\phi_{ij}\stackrel{\leftrightarrow}{\partial_\mu}\psi^j
+\ft85\gamma_{\mu\lambda}\partial^\lambda \left (\phi_{ij}\psi^j\right )\, ,
\w2
v_\mu^{ij} &=& -2\phi_k{}^{(i}{\partial}_\mu\phi^{j)k} +
8{\bar\psi}^i\gamma_\mu\psi^j\ ,
\nonumber
\end{eqnarray}\\[-.75cm]
where $\phi^2 \equiv \phi^{ij}\phi_{ij}$. The currents are conserved
provided that the fields satisfy the free field equations \eq{fe}. When
we apply supersymmetry transformations \eq{tr} on the currents
\eq{currents}, always using the field equations \eq{fe}, we find a full
supermultiplet of operators bilinear in the fields:\footnote{The same
operators have been given in \cite{how1} using the superfield approach.}
\begin{eqnarray}
\delta \theta_{\mu\nu} &=&\ft14 {\bar\epsilon}\gamma_{\rho(\mu}\partial^\rho J_{\nu)}\ ,
\nonumber
\w2
\delta J_\mu^i &=& \gamma^\nu\theta_{\mu\nu}\epsilon^i
-\ft18\left(\gamma^\rho\gamma_{\mu\nu} - {\textstyle{3\over 5}}
\gamma_{\mu\nu}\gamma^\rho\right)\partial^\nu v_\rho^{ij} \epsilon_j
\nonumber
\w2
&& -\ft1{4}\left(\gamma ^{abc}\gamma _{\mu\nu }
+\ft15 \gamma _{\mu\nu}\gamma ^{abc}\right) \partial^\nu t_{abc}^{ij}\epsilon _j\ ,
\nonumber\w2
\delta v_\mu^{ij} &=& -\ft12{\bar\epsilon}^{(i}J_\mu^{j)}
+\ft{15}8{\bar \epsilon}_k \gamma_{\mu\nu}\partial^\nu\lambda^{k(i,j)}\ ,
\la{deltatheta}
\w2
\delta t_{abc}^{ij} &=& -\ft1{24} {\bar\epsilon}^{[i}\gamma^\rho \gamma_{abc} J_\rho^{j]}
+\ft5{32} {\bar\epsilon}^k \not\!\partial \gamma_{abc}\lambda_k^{ij} - ({\rm trace})\, ,
\nonumber
\w2
\delta\lambda^{ij}_k &=&- \ft1{15} \gamma^{abc}t^{ij}_{abc}\epsilon_k - \ft4{15}\gamma^\mu v_{\mu k}{}^{[i}
\epsilon^{j]} -4\not\!\partial d^{ij}_{kl}\epsilon^l - ({\rm trace})\ ,
\nonumber\w2
\delta d^{ij,kl} &=&-\ft{1}8 {\bar\epsilon}^{[i}\lambda^{j],kl}
+ (ij \leftrightarrow kl) -({\rm trace})\ ,
\nonumber
\end{eqnarray}\\[-.75cm]
where we have introduced the operators $\lambda ^{ij}_k$, $t_{abc}^{ij}$and
$d^{ij,kl}$ defined by
\begin{eqnarray}
\la{operators}
t_{abc}^{ij} &=& \ft23 H^+_{abc}\phi^{ij}-\ft43{\bar\psi}^i\gamma_{abc}\psi^j
-\ft13\Omega^{ij} {\bar\psi}\gamma_{abc}\psi\ ,
\nonumber
\w2
\lambda^{ij}_k &=& -\ft{32}{15}\phi^{ij}\psi_k
-\ft{128}{75}\delta^{[i}_k\phi ^{j]\ell }\psi _\ell
+\ft{32}{75} \Omega^{ij}\phi_{k\ell }\psi ^\ell \ ,
\w2
d^{ij}_{k\ell } &=& -\ft1{15}\phi^{ij}\phi_{k\ell } +\ft1{75} \delta^{[i}_{[k}
\delta^{j]}_{\ell ]}\phi ^2 -\ft1{300}\Omega^{ij}\Omega_{k\ell}\phi ^2\ .
\nonumber
\end{eqnarray}\\[-.75cm]
Note that the operator $t_{abc}^{ij}$ is {\it self-dual}. To prove the
transformation rules we have used several $USp(4)$ Schouten identities
such as
\begin{eqnarray}
&& 2\bar \epsilon ^{[i}\phi ^{j]k}\psi _k+2\bar \epsilon ^k\phi _k{}^{[i}\psi ^{j]}
+\phi ^{ij}\bar \epsilon \psi - ({\rm trace})=0\ ,
\nonumber
\w2
&&\gamma^{abc}\epsilon_k\,{\bar\psi}^i \gamma_{abc}\psi^j
-2\gamma^{abc}\epsilon^{[j}\,{\bar\psi}^{i]}\gamma_{abc}\psi_k -({\rm traces}) =0\ .
\la{Schouten}
\end{eqnarray}\\[-.75cm]
The currents \eq{currents} and the operators \eq{operators} constitute
the multiplet of currents. Taking into account the conditions
\eq{conditions} this supermultiplet contains $128 + 128$ degrees of
freedom. Using this current multiplet, the linearized Weyl multiplet is
derived by introducing the Noether coupling
\begin{equation}
\int d^6x \left[ h_{\mu \nu }\theta ^{\mu \nu }+ \bar \psi ^\mu J_\mu
+ V_\mu ^{ij}v^\mu _{ij} + T_{abc}^{ij}t^{abc}_{ij}
+\bar \chi ^{ij}_k\lambda ^k_{ij} +D^{ij,kl} d_{ij,kl}\right]\ .
\la{invariant}
\end{equation}\\[-.75cm]
The independent conformal supergravity fields which have been introduced
above are
\begin{equation}
h_{\mu\nu}\ ,\ \psi_\mu^i\ ,\ V_\mu^{ij}\ ,\ T_{abc}^{ij}\ ,\
\chi^i_{jk}\ ,\ D^{ij, kl}\ .
\la{fc}
\end{equation}\\[-.75cm]
Several properties of these gauge and matter fields are summarized in
Table~\ref{tbl:fieldsWeyl}. \\
\fb
Demanding the invariance of the action \eq{invariant} under rigid
supersymmetry, we find the following linearized Weyl multiplet
transformation rules:
\begin{eqnarray}
\delta h_{\mu\nu} &=& {\bar\epsilon}\gamma_{(\mu}\psi_{\nu)}\ ,
\nonumber\w2
\delta \psi_\mu^i &=& -\ft14 (\partial_\rho h_{\mu\sigma})\gamma^{\rho\sigma}\epsilon^i
+\ft12 V_\mu^{ij}\epsilon_j +\ft1{24}T^{ij}_{abc} \gamma^{abc}\gamma_\mu \epsilon_j\ ,
\nonumber
\w2
\delta V_\mu^{ij} &=&
\ft14{\bar\epsilon}^{(i}\left(\gamma^{\rho\sigma}\gamma_\mu-\ft35\gamma_\mu\gamma^{\rho\sigma}\right)\psi_{\rho\sigma}^{j)}
-\ft4{15}{\bar\epsilon}_k\gamma_\mu\chi^{(i,j)k}\ ,
\nonumber
\w2
\delta T_{abc}^{ij} &=&
\ft18{\bar\epsilon}^{[i}\left(\gamma^{\mu\nu}\gamma_{abc}+\ft15\gamma_{abc}\gamma^{\mu\nu}\right)
\psi_{\mu\nu}^{j]} -\ft1{15}{\bar\epsilon}^k\gamma_{abc}\chi_k^{ij} -({\rm trace})\ ,
\nonumber
\w2
\delta\chi^{ij}_k &=& \ft5{32}
\left(\partial_\mu T^{ij}_{abc}\right)\gamma^{abc}\gamma^\mu \epsilon_k
-\ft{15}8\gamma^{\mu\nu} V_{\mu\nu k}{}^{[i}\epsilon^{j]}
-\ft14 D^{ij}_{k\ell}\epsilon^\ell -({\rm traces})\ ,
\nonumber
\w2
\delta D^{ij,k\ell} &=& -2 {\bar\epsilon}^{[i} \not \!\del \chi^{j],k\ell}
-2 {\bar\epsilon}^{[k} \not \!\del \chi^{\ell],ij} -({\rm trace})\ .
\end{eqnarray}\\[-.75cm]
where the supersymmetry parameter is constant, and we have defined
\begin{eqnarray}
\psi_{\mu\nu} &=& \partial_\mu\psi_\nu -\partial_\nu \psi_\mu\ ,
\nonumber\w2
V_{\mu\nu}^{ij} &=& \partial_\mu V_\nu^{ij} -\partial_\nu V_\mu^{ij}\ .
\end{eqnarray}\\[-.75cm]
In the next section, we generalize the above transformation rules to
obtain the full local superconformal transformation rules.
\section{The (2,0) Conformal Supergravity Theory }\la{ss:Weylm}
In this section, we will construct the nonlinear (2,0) conformal
supergravity theory. Our starting point is the linearized conformal
multiplet constructed in the previous section. This multiplet contains
both gauge fields and matter fields (not to be confused with
the tensor multiplet matter fields that will be coupled in section~\ref{tm2}).
Due to the presence of the matter fields, the nonlinearization cannot be
understood as a straightforward gauging of an underlying superconformal
algebra. To include the matter fields, one must follow a
6-step procedure that has been explained in detail in \cite{ber1}. In the same
reference, this 6-step procedure has been applied to construct the
(1,0) conformal supergravity theory. Here we apply the same procedure
to construct the (2,0) theory.\\
The (2,0) conformal supergravity is based on the superconformal algebra $OSp(8^*|4)$ whose
generators are labeled
\begin{equation}
T_A = P_a\ , Q_{\alpha i}\ , U_{ij}\ , M_{ab}\ , K_a\ , S_{\alpha i}\ , D\ ,
\end{equation}\\[-.75cm]
where $a,b, \cdots$ are Lorentz indices, $\alpha$ is a chiral spinor index
and $i,j = 1, \cdots 4$ are $USp(4)$ indices. $M_{ab}$ and $P_a$ are
the Poincar\'e generators, $K_a$ is the special conformal
transformation, $D$ the dilatation, $Q_{\alpha i}$ and $S_{\alpha i}$ are the
supersymmetry and special supersymmetry generators, respectively, which
are symplectic Majorana-Weyl spinors, 16 real components in total.
Finally, $U^{ij} = U^{ji}$ are the ${\rm USp(4)}$ generators. For more
details on the $OSp(8^*|4)$ algebra and the rigid superconformal
transformations, see \cite{cla1}. \\
The gauge fields corresponding to the above generators are
\begin{equation}
e_\mu{}^a\ ,\psi_\mu^i\ ,\ V_\mu^{ij}\ , \omega_\mu{}^{ab}\ ,
\ f_\mu{}^a\ ,\phi_\mu^i\ ,\ b_\mu\ .
\end{equation}\\[-.75cm]
However, in the realization (Weyl multiplet) which will gauge the
algebra, these fields are not all independent. The independent fields
are given in \eq{fc}, where the first three are the gauge fields
corresponding to the generators $P_a\ , Q_{\alpha i}$ and $U_{ij}$. It is
understood that the linearized gravitational field $h_{\mu\nu}$ has been
replaced by the sechsbein $e_\mu{}^a$. The last three are matter fields
needed for the realization of the superconformal algebra.
The remaining gauge fields are either dependent $(\omega_\mu{}^{ab}\ ,
f_\mu{}^a\ ,\phi_\mu^i)$ (see below), or can be shifted away
($b_\mu$) using $K_a$ invariance.\\
The $(2,0)$ Weyl multiplet describes $128 + 128$ off-shell degrees of
freedom. We first present the result and next explain our notation and
give our definitions. The bosonic transformations of the independent
gauge fields are given by general coordinate transformations and
\begin{eqnarray}
\delta e_\mu{}^a &=& -\Lambda_D e_\mu{}^a - \Lambda^{ab}e_{\mu b}\ ,
\nonumber
\w2
\delta \psi_\mu^i &=& -{\textstyle{1\over 2}}\Lambda_D \psi_\mu^i
+{\textstyle{1\over 2}}\Lambda^i{}_j\psi_\mu^j -{\textstyle{1\over 4}}\Lambda^{ab}\gamma_{ab}\psi_\mu^i\ ,
\nonumber
\w2
\delta V_\mu^{ij} &=& \partial_\mu\Lambda^{ij} +\Lambda^{(i}{}_kV_\mu^{j)k}\, ,\nonumber
\w2
\delta b_\mu &=& \partial _\mu \Lambda_D-2e_\mu ^a\Lambda_{K\,a }\ .
\la{BosTransfWeyl}
\end{eqnarray}\\[-.75cm]
where $\Lambda_D, \,\Lambda_{K\,a },\,\Lambda^{ab}$ and $\Lambda^{ij}$ are the parameters
of dilatation, special conformal, Lorentz and USp(4) transformations,
respectively. The transformation properties of the matter fields,
$T,\chi $ and $D$ under dilatations, Lorentz and USp(4) transformations
follow from the rules (\ref{BosTransfWeyl}) and from
Table~\ref{tbl:fieldsWeyl}. All matter fields are inert under the
special conformal transformations $K$. \\
Following \cite{superconformN1,ber1} we impose the following curvature
constraints
\footnote{Note that in contradistinction to the (1,0) case,
discussed in \cite{ber1}, the matter fields $\chi^i_{jk}$ and $D$ are
absent in the constraints. See also Appendix~\ref{app:truncation}.}
\begin{eqnarray}
R_{\mu\nu}{}^a (P) &=& 0\, ,
\nonumber
\w2
\la{constraints}
R_{\mu\nu}{}^{ab}(M) e^\nu{}_b + {\textstyle{1\over 4}} T_{\mu bc}^{ij}
T^{abc}_{ij} &=& 0\, ,
\w2
\gamma^\mu R_{\mu\nu}^i(Q) &=& 0\, .\nonumber
\end{eqnarray}\\[-.75cm]
The above matter-modified curvatures are defined by
\begin{eqnarray}
R_{\mu\nu}{}^a(P) &=& 2\partial_{[\mu}e_{\nu]}^a + 2b_{[\mu}e_{\nu]}^a
+{\underline { 2\omega_{[\mu}{}^{ab}e_{\nu]b}}} -{\textstyle{1\over 2}}{\bar
\psi}_\mu\gamma^a\psi_\nu\ ,
\nonumber\w2
R_{\mu\nu}{}^{ab}(M) &=&2\partial_{[\mu}\omega_{\nu]}{}^{ab} + 2
\omega_{[\mu}{}^{ac}\omega_{\nu]c}{}^b
-{\underline {8f_{[\mu}{}^{[a}e_{\nu]}{}^{b]}}}
+{\bar\psi}_{[\mu}\gamma^{ab}\phi_{\nu]}
\nonumber
\w2
&&+{\bar\psi}_{[\mu}\gamma^{[a}R_{\nu]}{}^{b]}(Q) +{\textstyle{1\over 2}} {\bar
\psi}_{[\mu}\gamma_{\nu]}R^{ab}(Q) +{\textstyle{1\over
2}}{\bar\psi}_{\mu,i}\gamma_c\psi_{\nu,j} T^{abc,ij}\, ,
\la{curvatures}
\w2
R_{\mu\nu}^i(Q) &=& \left ( 2 \partial_{[\mu}\psi_{\nu]}^i +b_{[\mu}\psi_{\nu]}^i
+ {\textstyle{1\over 2}} \omega_{[\mu}{}^{ab}
\gamma_{ab}\psi_{\nu]}^i - V_{[\mu}^i{}_j \psi_{\nu]}^j\right )\nonumber
\w2
&&+ {\underline { 2\gamma_{[\mu}\phi_{\nu]}^i}} +{\textstyle{1\over 12}}
T^{ij}_{abc}\gamma^{abc}\gamma_{[\mu}\psi_{\nu]j}\ .
\nonumber
\end{eqnarray}\\[-.75cm]
The underlined terms indicate all terms of the form gauge field
$\times$ sechsbein. Since the sechsbein is invertible the corresponding
gauge fields can be solved for from the constraints \eq{constraints}.
Explicitly, the constraints \eq{constraints} enable us to solve for the
gauge fields ($\omega_\mu{}^{ab}\ ,f_\mu{}^a\ ,\phi_\mu^i)$ as follows:
\begin{eqnarray}
\omega_\mu{}^{ab} &=& 2e^{\nu[a}\partial_{[\mu}e_{\nu]}{}^{b]}
- e^{\rho [a}e^{b]\sigma}e_\mu{}^c\partial_\rho e_{\sigma c}
\nonumber
\w2
&& + 2 e_\mu{}^{[a} b^{b]} +{\textstyle{1\over 2}}{\bar\psi}_\mu\gamma^{[a}\psi^{b]}
+ {\textstyle{1\over 4}}{\bar\psi}^a\gamma_\mu\psi^b\ ,
\nonumber
\w2
f_\mu{}^a &=& -{\textstyle{1\over 8}}R^{\prime }_\mu{}^a(M)
+{\textstyle{1\over 80}}e_\mu{}^a R^\prime (M)
+{\textstyle{1\over 32}}T^{ij}_{\mu cd}T_{ij}^{acd}\ ,
\w2
\phi_\mu^i &=&-{\textstyle{1\over 16}}\left (\gamma^{ab}\gamma_\mu
-{\textstyle{3\over 5}}\gamma_\mu\gamma^{ab}\right )R^{\prime }_{ab}{}^i(Q)\ .
\nonumber
\end{eqnarray}\\[-.75cm]
The notation $R^\prime$ indicates that in the corresponding curvature
the underlined term in \eq{curvatures} has been omitted, and
$R'_\mu {}^a=e_b^\nu R'_{\mu \nu }{}^{ba}$.\\
We next give the full non-linear $Q$ and $S$-transformations of the
$(2,0)$ Weyl multiplet:
\begin{eqnarray}
\delta e_\mu{}^a &=&{\textstyle{1\over 2}}\bar\epsilon\gamma^a\psi_\mu\ ,
\nonumber
\w2
\delta b_\mu &=& -\ft12\bar\epsilon\phi _\mu +\ft12\bar \eta \psi_\mu
\nonumber\w2
\delta \psi_\mu^i &=& {\cal D}_\mu\epsilon^i + {\textstyle{1\over 24}}
T^{ij}_{abc}\gamma^{abc}\gamma_\mu\epsilon_j + \gamma_\mu\eta^i\ ,
\nonumber
\w2
\delta V_\mu^{ij} &=& -4{\bar\epsilon}^{(i} \phi_\mu^{j)} -
{\textstyle{4\over 15}}{\bar\epsilon}_k\gamma_\mu\chi^{(i,j)k} -4
{\bar\eta}^{(i}\psi_\mu^{j)}\ ,
\nonumber
\w2
\la{N=4}
\delta T_{abc}^{ij}&=& {\textstyle{1\over 8}}{\bar\epsilon}^{[i}
\gamma^{de}\gamma_{abc}{ R}^{j]}_{de}(Q)
- {\textstyle{1\over 15}} {\bar\epsilon}^k\gamma_{abc}\chi^{ij}_k - ({\rm trace})\ ,
\w2
\delta \chi_k^{ij} &=&{\textstyle{5\over 32}} \left (
{\cal D}_\mu T^{ij}_{abc}\right )\gamma^{abc}\gamma^\mu\epsilon_k -{\textstyle{15\over 16}}
\gamma^{\mu\nu} R_{\mu\nu k}{}^{[i}(V) \epsilon^{j]} -{\textstyle{1\over 4}}D^{ij}_{kl}\epsilon^l
\nonumber
\w2
&&+{\textstyle{5\over 8}} T^{ij}_{abc}\gamma^{abc}\eta_k -({\rm traces})\ ,
\nonumber
\w2
\delta D^{ij,kl} &=& - 2
{\bar\epsilon}^{[i}\not\!\!{\cal D} \chi^{j],kl} + 4 {\bar\eta}^{[i}\chi^{j],kl}
+ (ij\leftrightarrow kl ) -({\rm trace})\ .
\nonumber
\end{eqnarray}\\[-.75cm]
We have used here the following definitions. The covariant derivatives
and $USp(4)$ curvature in (\ref{N=4}) are:
\begin{eqnarray}
{\cal D}_\mu \epsilon^i &=& \partial_\mu\epsilon^i +{\textstyle{1\over 2}}b_\mu\epsilon^i
+{\textstyle{1\over 4}}\omega_\mu{}^{ab}\gamma_{ab}\epsilon^i -{\textstyle{1\over 2}}V_{\mu}{}^i{}_j\epsilon^j\ ,
\nonumber
\w2
R_{\mu\nu}{}^{ij}(V) &=& 2\partial_{[\mu}V_{\nu]}{}^{ij} +
V_{[\mu}{}^{k(i}V_{\nu]}{}^{j)}{}_k +8{\bar\psi}_{[\mu}{}^{(i}
\phi_{\nu]}{}^{j)}
+{\textstyle{8\over 15}}{\bar\psi}_{[\mu,k}\gamma_{\nu]}\chi^{(i,j)k}\ .
\la{de}
\end{eqnarray}\\[-.75cm]
The supercovariant derivatives ${\cal D}_\mu$ of matter fields are defined as
the ordinary derivative $\partial_\mu$ plus a covariantization term which is
always given by minus all the transformation rules of the matter field
with the parameter replaced by the corresponding gauge field. For
example the supercovariant derivative of $T$ is given by
\begin{eqnarray}
{\cal D}_\mu T_{abc}^{ij} &=& \partial_\mu T_{abc}^{ij} +3\omega_{\mu [a}{}^d
T_{bc]d}^{ij} - b_\mu T_{abc}^{ij} +V_\mu^{[i}{}_k T_{abc}^{j]k}
\w2
&& -{\textstyle{1\over 8}}{\bar\psi}_\mu^{[i}
\gamma^{de}\gamma_{abc}{ R}^{j]}_{de}(Q)
+ {\textstyle{1\over 15}} {\bar\psi}_\mu^k\gamma_{abc}\chi^{ij}_k - ({\rm trace})\,
.
\nonumber
\end{eqnarray}\\[-.75cm]
\section{The $(2,0)$ Tensor Multiplet in the Conformal
Supergravity Background}
\la{tm2}
In this section we couple a (2,0) tensor multiplet to conformal
supergravity. Our starting point will be the linearized transformation
rules of the tensor multiplet. The nonlinear rules can then be obtained
by imposing the superconformal algebra via an iterative Noether procedure.
This procedure has been described in detail for the $(1,0)$
case in \cite{ber1}. The same procedure can be applied here.
As an alternative, we will derive the same result by the
requirement that the (2,0) nonlinear tensor multiplet should reproduce,
upon truncation the (1,0) nonlinear tensor multiplet of \cite{ber1}.
The details of this truncation are explained in Appendix~\ref{app:truncation}.\\
Our starting point is the
the linearized equations of motion and the supersymmetry transformations
of the $(2,0)$ tensor multiplet given in section~\ref{tm}.
Applying the truncation procedure described in Appendix~\ref{app:truncation}, we find that
the full nonlinear $Q$ and $S$-transformations of the $(2,0)$ tensor
multiplet are given by:
\begin{eqnarray}
\delta B_{\mu\nu} &=& - {\bar\epsilon}\gamma_{\mu\nu}\psi +{\bar\epsilon}^i
\gamma_{[\mu}\psi_{\nu]}^j\phi_{ij}\ ,
\nonumber
\w2
\delta \psi^i &=& {\textstyle{1\over 48}} { H}^{+}_{\mu\nu\rho}\gamma^{\mu\nu\rho} \epsilon^i
+{\textstyle{1\over 4}}\not\!\! {\cal D} \phi^{ij}
\epsilon_j - \phi^{ij}\eta_j\ ,
\nonumber
\w2
\delta \phi^{ij} &=& -4
{\bar\epsilon}^{[i}\psi^{j]} -({\rm trace})\ .
\la{N=4T}
\end{eqnarray}\\[-.75cm]
The curvature $H_{\mu\nu\rho}$ is defined by
\begin{equation}
H_{\mu\nu\rho} = 3\partial_{[\mu}B_{\nu\rho]}
+3{\bar\psi}_{[\mu}\gamma_{\nu\rho]}\psi
-{\textstyle{3\over2}}{\bar\psi}^i_{[\mu}\gamma_\nu\psi^j_{\rho]} \phi_{ij}\ .
\end{equation}\\[-.75cm]
It satisfies the Bianchi identity
\begin{equation}
{\cal D}_{[a}H_{bcd]} - {\textstyle{3\over 2}} {\bar\psi}\gamma_{[ab} {R}_{cd]}(Q)=0\ .
\end{equation}\\[-.75cm]
The self-dual part of the curvature ${ H}^+_{abc}$ transforms as
\begin{equation}
\delta { H}_{abc}^{+} = -{\textstyle{1\over 2}}{\bar \epsilon}
\not\!\! {\cal D}\gamma_{abc}\psi -3 {\bar\eta}\gamma_{abc}\psi\ .
\end{equation}\\[-.75cm]
Furthermore, we find that the field equations of the $(2,0)$ tensor
multiplet are given by\footnote{
Note that, in contradistinction to the (1,0) case the first field equation
can {\it not} be used to solve for the matter field $T$ in terms of $H^-$
(the scalar $\phi$ is not a singlet under USp(4)). Therefore, the (2,0)
Weyl multiplet has {\it no} alternative formulation containing an antisymmetric
tensor gauge field like the (1,0) Weyl multiplet (see \cite{ber1}).}
\begin{eqnarray}
{\cal F}^-_{abc} &:=& { H}^-_{abc} - {\textstyle{1\over 2}}
\phi_{ij}T_{abc}^{ij} = 0\ ,
\nonumber
\w2
\Gamma^i &:=& \not\!\! {\cal D}\psi^i
-{\textstyle{1\over 15}}\phi^{kl}}\chi^i_{kl} -{\textstyle{1\over 12}
T^{ij}_{abc}\gamma^{abc}\psi_j = 0\ ,
\la{fe2}
\w2
{\cal C}_{ij} &:=& {\cal D}^a {\cal D}_a \phi_{ij} -{\textstyle{1\over
15}}D_{ij}^{kl}\phi_{kl} +{\textstyle{1\over 3}}{ H}^+_{abc} T_{ij}^{abc} +
{\textstyle{16\over 15 }}{\bar \chi}_{ij}^k\psi_k = 0\ ,
\nonumber
\end{eqnarray}\\[-.75cm]
where
\begin{eqnarray}
{\cal D}_\mu \psi^i &=& \left (\partial_\mu - {\textstyle{5\over 2}} b_\mu +
{\textstyle{1\over 4}}\omega_\mu{}^{ab}\gamma_{ab}\right )\psi^i -{\textstyle{1\over
2}}V_\mu^i{}_j\psi^j\nonumber
\w2
&&-{\textstyle{1\over 48}}H^+_{abc}\gamma^{abc}\psi_\mu^i - {\textstyle{1\over 4}}\left (
\not\!\! {\cal D}\phi^{ij}\right )
\psi_{\mu j} + \phi^{ij}\phi_{\mu j}\ ,
\w2
{\cal D}_\mu \phi^{ij} &=& \left (\partial_\mu -2b_\mu\right )
\phi^{ij} + V_\mu{}^{[i}{}_k \phi^{j]k} + 4 \left (
{\bar\psi}_\mu^{[i} \psi^{j]} - {\rm trace}\right )\ .
\nonumber
\end{eqnarray}\\[-.75cm]
To determine the supercovariant d'Alembertian of the scalars we first
calculate the transformation properties of ${\cal D}_a\phi^{ij}$:
\begin{eqnarray}
\delta {\cal D}_a\phi^{ij} &=& 3\Lambda_D {\cal D}_a\phi^{ij}
+ \Lambda^{[i}{}_k {\cal D}_a\phi^{j]k} -\Lambda_a{}^b {\cal D}_b\phi^{ij}
+4\Lambda_{Ka}\phi^{ij}
\nonumber
\w2
&&-4 {\bar\epsilon}^{[i}{\cal D}_a \psi^{j]} +{\textstyle {2\over 15}}\left(
{\bar\epsilon}_l\gamma_a\chi^{(i,k)l}\phi^j {}_k - i\leftrightarrow j\right)
\w2
&&+{\textstyle{1\over 6}}{\bar\epsilon}_k\gamma_a \gamma^{bcd}T_{bcd}^{k[i}
\psi^{j]}
-4 {\bar\eta}^{[i}\gamma_a \psi^{j]} - ({\rm trace})\ .
\nonumber
\end{eqnarray}\\[-.75cm]
{}From this we derive that
\begin{eqnarray}
\la{box}
{\cal D}^a{\cal D}_a \phi^{ij} &=& \partial^a{\cal D}_a \phi^{ij}
-3b^a {\cal D}_a\phi^{ij} + V_a^{[i}{}_k {\cal D}^a\phi^{j]k} +\omega_a{}^{ab}
{\cal D}_b\phi^{ij} -4 f_a{}^a\phi^{ij}\nonumber
\w2
&&+4 {\bar \psi}_a^{[i}{\cal D}^a \psi^{j]} -{\textstyle {2\over 15}}\left (
{\bar\psi}^a_l\gamma_a\chi^{(i,k)l}\phi^j {}_k - i\leftrightarrow j\right )
\w2
&&-{\textstyle{1\over 6}}{\bar\psi}^a_k\gamma_a \gamma^{bcd}T^{k[i}_{bcd}
\psi^{j]}
+4 {\bar\phi}_a^{[i}\gamma^a \psi^{j]} - ({\rm trace})\ .\nonumber
\end{eqnarray}\\[-.75cm]
Note the occurrence of the Riemann curvature scalar in the equation
of motion for the scalar fields through the term $f_a{}^a\phi_{ij}$.
This contains as well graviton as gravitino terms of the supergravity
action, the latter through $f_a^ a=-\ft1{20}R'(M)+\ldots
=\ft1{160}\bar \psi _\mu \gamma ^{\mu \nu \rho }R'_{\nu \rho
}(Q)+\ldots $. Gravitino kinetic terms appear also in the equation of
motion for tensor multiplet fermions through the term
$\not\!\!{\cal D}\psi ^i=\phi^{ij}\gamma ^\mu \phi_{\mu j}+\ldots =
\ft1{10}\phi^{ij}\gamma ^{\mu \nu }R'_{\mu \nu j}(Q)+\ldots$.\\
While it is possible to compute the Green's functions for the tensor
multiplet fields in presence of the conformal supergravity background
by starting from the equations of motion, it
would be convenient to perform such calculations by starting from an
action. To construct a manifestly Lorentz invariant action
requires the introduction of an auxiliary scalar field \cite{pst}. It
has been shown in \cite{cla1} that this can straightforwardly be
implemented in a rigid conformal theory.
We expect that this can be extended for the local superconformal
case.\\
An
alternative approach is to relax the chirality condition on the
$2$-form potential and to write an action which is not invariant, but
whose variation is proportional to $ {\cal F}^-_{abc}$. It
gives the correct equations of motion provided that
the self-duality condition $ {\cal F}^-_{abc}=0$ is imposed after the action is varied
\cite{e,c,ric,p}. Such an action takes the form
\begin{equation}
S=\int d^6 x \left(- \ft16 H^{abc} {\cal F}^-_{abc} -4{\bar\psi}^i\Gamma_i
- {\bar\psi}_\mu^i \gamma^\mu \Gamma^j \phi_{ij}
+\ft14 \phi^{ij} {\cal C}_{ij}\right) \ .
\end{equation}\\[-.75cm]
\section{Poincar\'e Supergravity Coupled to $N$ (2,0) Tensor
Multiplets from the Superconformal Theory}\la{ps}
In this section we construct the matter couplings to $(2,0)$
supergravity by using the superconformal tensor calculus and by
imposing the $SO(N,5)$ symmetry. We hereby closely follow the
procedure of coupling $N=4$, $d=4$ vector multiplets as in
\cite{deRooWag}. This procedure was first introduced in \cite{conformforPoin}
and has been applied to obtain matter couplings in 4 dimensions for $N=1$
\cite{N1YMmsg}, $N=2$ \cite{dWLVP} and $N=4$ \cite{ber2}.
The basic idea is that there is a close relation between
matter-coupled Poincar\'{e} and conformal supergravity theories.
Starting for (a slight generalization of)
the matter-coupled conformal supergravity theory
constructed in the previous section, we simply gauge fix the conformal
scale and S-supersymmetry transformations to reproduce
existing results on D=6 matter-coupled Poincar\'{e} supergravity
\cite{romans,ric}. For a review of this technique, see for example
\cite{review}.\\
We begin by introducing $(N+5)$ copies of the $(2,0)$ tensor multiplets
with fields $B_{\mu\nu}^I\ , \psi^{iI}\ , L_I^{ij}$ where $I=1,...,N+5$
labels the vector representation of $SO(N,5)$. We have denoted the
scalars by $L_I^{ij}$ because they will shortly be constrained. The
constraint will be solved in terms of independent scalar fields which
will again be denoted by $\phi$.\\
The superconformal transformation rules now read
\begin{eqnarray}
\delta B^I_{\mu\nu} &=& - {\bar\epsilon}\gamma_{\mu\nu}\psi^I +{\bar\epsilon}^i
\gamma_{[\mu}\psi_{\nu]}^j L^I_{ij}\ ,
\nonumber
\w2
\delta \psi^{iI} &=& {\textstyle{1\over 48}} { H}^{I+}_{abc}\gamma^{abc} \epsilon^i
+{\textstyle{1\over 4}}\not\!\! {\cal D} L^{Iij}\epsilon_j - L^{Iij}\eta_j\ ,
\nonumber
\w2
\delta L^{Iij} &=& -4 {\bar\epsilon}^{[i}\psi^{j]I} - \Omega^{ij}{\bar\epsilon}\psi^I \ .
\la{5N}
\end{eqnarray}\\[-.75cm]
Since the tensor multiplet fields occur linearly in the full field
equations \eq{fe2}, the latter generalize to the case of $N+5$ tensor
multiplets as
\begin{eqnarray}
&& { H}^{I-}_{abc} - {\textstyle{1\over 2}} L^I_{ij}T_{abc}^{ij} = 0\ ,
\nonumber
\w2
&& \not\!\! {\cal D}\psi_I^i -{\textstyle{1\over 15}}L_I^{kl}}\chi^i_{kl}
-{\textstyle{1\over 12} T^{ij}_{abc}\gamma^{abc}\psi_{Ij} = 0\ ,
\la{fe3}
\w2
&& {\cal D}^a {\cal D}_a L^I_{ij} -{\textstyle{1\over 15}}D_{ij}^{kl}L^I_{kl}
+{\textstyle{1\over 3}}{ H}^{I+}_{abc} T_{ij}^{abc} +{\textstyle{16\over 15 }}{\bar
\chi}_{ij}^k\psi^I_k = 0\ .
\nonumber
\end{eqnarray}\\[-.75cm]
Note that the index $I$ is a global $SO(N+5)$ index and consequently
the derivatives of $L^I_{ij}$ and $\psi^I_i$ occurring in \eq{5N} and
\eq{fe3} are as defined earlier for $\phi_{ij}$ and $\psi$ without any
new connection terms to rotate the index $I$.
To obtain the Poincar\'e supergravity coupled to $N$ copies of the
$(2,0)$ tensor multiplet, we impose the geometrical constraint
\begin{eqnarray}
\eta^{IJ} L_I^{ij} L_{Jk\ell} &=&
-\delta^{[i}_{[k}\delta^{j]}_{\ell ]}+\ft14\Omega^{ij}\Omega_{k\ell}\
\nonumber\w2
&\equiv& \eta^{ij}{}_{k\ell}
\la{c}
\end{eqnarray}\\[-.75cm]
where $\eta_{IJ}$ is a symmetric invariant tensor of $SO(N,5)$ with
signature $(-----++\cdots +)$. The raising and lowering of the
$SO(N,5)$ indices will always be done with the metric $\eta_{IJ}$. The
condition \eq{c}, together with the fact that $L_I^{ij}$ are defined up
to local $USp(4)$ transformations, reduces the number of independent
scalars to $(N+5)\times 5 - 15-10 =5N$, which is the dimension of the
coset ${SO(N,5)\over SO(N)\times SO(5)}$. It is convenient to introduce
an $(N+5)\times N$ matrix $L_I^r\ (r=1,...,N)$ which together with
$L_I^{ij}$ form an $(N+5)\times (N+5)$ matrix $L_I^A$ satisfying the
condition
\begin{equation}
\eta^{IJ} L_I{}^A L_{JB} = \eta^A{}_B\ ,
\end{equation}\\[-.75cm]
where $A=(ij,r)$ and the $\eta^A{}_B$ is the constant metric with
components: $\eta^{ij}{}_{k\ell}\ , \eta^r{}_s=\delta^r{}_s$ and
$\eta^{ij}{}_r=\eta^r{}_{ij}=0$. \\
The constraint \eq{c} is invariant under $S$--supersymmetry. However,
varying it under $Q$--supersymmetry gives the constraint
\begin{equation}
L_I^{ij} \psi^I_k = 0\ .
\la{c2}
\end{equation}\\[-.75cm]
This constraint is easily solved as
\begin{equation}
\psi^i_I = L_I^r \psi^{ri}\ ,
\la{psi}
\end{equation}\\[-.75cm]
where $\psi^{ri}\,(r=1,...,N)$ are the independent fermionic fields.\\
Next, we vary the traceless part of \eq{c2} to obtain the constraint
\begin{equation}
L_I^{ij}{\cal D}_\mu L^I_{k\ell} =
-8{\bar\psi}^{I[i}\gamma_\mu\psi_{I[k}\delta_{\ell]}^{j]} \ ,
\la{v}
\end{equation}\\[-.75cm]
and, making use of \eq{Schouten}, the equation of motion for the
$2$-form potential
\begin{equation}
H_{abc}^{I+} L_I^{ij} =-2{\bar\psi}^i_I \gamma_{abc}\psi^{jI}\ .
\la{he}
\end{equation}\\[-.75cm]
Requiring that the trace part of the constraint \eq{c2} is invariant
under the combined $Q$ and $S$--transformations, and using
\eqs{psi}{he} and performing Fierz re-arrangement, we determine the
$S$--supersymmetry parameter:
\begin{equation}
\eta_i=
-\ft12({\bar\psi}_i^r\gamma^a\psi^k_r)\,\gamma_a\epsilon_k
-\ft1{72} ({\bar\psi}_i^r\gamma^{abc}\psi^k_r)\,\gamma_{abc}\epsilon_k
-\ft1{36}({\bar\psi}^k_r \gamma^{abc}\psi_k^r)\,\gamma_{abc}\epsilon_i \ .
\la{eta}
\end{equation}\\[-.75cm]
Next, we observe that $V_\mu^{ij}$ can be solved from \eq{v} as
\begin{equation}
V_{\mu i}{}^j = 2 L^I_{ik} {\cal D'}_\mu L_I^{jk}
-8{\bar\psi}_i^I\gamma_\mu\psi_I^j\ ,
\end{equation}\\[-.75cm]
where ${\cal D'}_\mu$ is the supercovariant derivative without the
$V_\mu^{ij}$ term. The Weyl multiplet fields $T^{ij}_{abc},\chi^i_{jk}$
and $D^{ij}_{k\ell}$ are also readily solved from
\eq{fe3}. For example,
\begin{equation}
T^{ij}_{abc}= -2 H^{I-}_{abc} L_I^{ij}\ .
\la{t}
\end{equation}\\[-.75cm]
Using $K$--symmetry, we can also set
\begin{equation}
b_\mu=0\ .
\end{equation}\\[-.75cm]
The independent fields we are left with are those of the combined
$(2,0)$ Poincar\'e supergravity plus $N$ tensor multiplet system,
namely:
\begin{equation}
e_\mu{}^a\ , \psi_\mu^i\ , B_{\mu\nu}^I\ , \psi_i^r\ , L_I^A\ .
\end{equation}\\[-.75cm]
The Poincar\'e supersymmetry transformations of these fields can be
found from \eq{N=4} and \eq{N=4T} by using the solutions for the Weyl
multiplet fields and the compensating $S$--supersymmetry transformation
\eq{eta}. We thus find
\begin{eqnarray}
\delta e_\mu{}^a &=&{\textstyle{1\over 2}}\bar\epsilon\gamma^a\psi_\mu\ ,
\nonumber
\w2
\delta \psi_\mu^i &=& {\cal D}_\mu\epsilon^i - {\textstyle{1\over 12}}
L_I^{ij}H^{I-}_{\rho\sigma\tau}\gamma^{\rho\sigma\tau}\gamma_\mu\epsilon_j + \gamma_\mu\eta^i\ ,
\nonumber
\w2
\delta B^I_{\mu\nu} &=& - L^I_r {\bar\epsilon}^i\gamma_{\mu\nu}\psi_i^r
+L^I_{ij}{\bar\epsilon}^i \gamma_{[\mu}\psi_{\nu]}^j \ ,
\la{final}
\w2
\delta \psi^{ir} &=& {\textstyle{1\over 48}} L_I^r { H}^{I+}_{\mu\nu\rho}\gamma^{\mu\nu\rho} \epsilon^i
+{\textstyle{1\over 4}} V_\alpha^{r,ij} \not\!\!{\cal D} \phi^\alpha \epsilon_j
-\delta\phi^\alpha A_\alpha^{rs} \psi_s^i\ ,
\nonumber
\w2
\delta\phi^\alpha &=& 4 V^\alpha_{r,ij}\,{\bar\epsilon}^i\psi^{jr}\ .
\nonumber
\end{eqnarray}\\[-.75cm]
where $\phi^\alpha\ (\alpha=1,...,5N)$ are the scalar fields parametrizing the
coset ${SO(N,5)\over SO(N)\times SO(5)}$ and $\eta^i$ is given in
\eq{eta}. The vielbein $V_\alpha^{r,ij}$, the $SO(N)$ connection
$A_\alpha^{rs}$ and the $USp(4)$ connection $A_\alpha^{ij}$ on this coset are
defined as
\begin{eqnarray}
V_\alpha^{r,ij} &=& L^{Ir}\partial_\alpha L_I^{ij}\ ,
\nonumber\w2
A_\alpha^{rs} &=& L^{Ir}\partial_\alpha L_I^s\ ,
\nonumber\w2
A_{\alpha i}{}^j &=& 2 L^I_{ik} \partial_\alpha L_I^{jk}\ .
\end{eqnarray}\\[-.75cm]
Further definitions are as follows:
\begin{eqnarray}
H^I_{\mu\nu\rho} &=& 3\partial_{[\mu}B^I_{\nu\rho]}
+3{\bar\psi}_{[\mu}\gamma_{\nu\rho]}\psi^r L^I_r
-{\textstyle{3\over2}}{\bar\psi}^i_{[\mu}\gamma_\nu\psi^j_{\rho]} L^I_{ij}\ ,
\nonumber
\w2
{\cal D}_\mu \phi^\alpha &=& \partial_\mu \phi^\alpha -4V^\alpha_{r,ij}\,{\bar\psi}_\mu^i\psi^{jr}\ ,
\la{defs2}
\w2
{\cal D}_\mu \epsilon^i &=& \partial_\mu\epsilon^i +{\textstyle{1\over 4}}\omega_\mu{}^{ab}\gamma_{ab}\epsilon^i
-{\textstyle{1\over 2}}V_{\mu}{}^i{}_j\epsilon^j\ ,
\nonumber
\end{eqnarray}\\[-.75cm]
where the composite connection $V_\mu^{ij}$ is given by
\begin{equation}
V_\mu^{ij} = {\cal D}_\mu \phi^\alpha A_\alpha^{ij} -8{\bar\psi}^i_r\gamma_\mu\psi^{jr}\ .
\end{equation}\\[-.75cm]
Comparing the result \eq{final} with that of \cite{ric}, we find that
all the structures are in agreement except the last term in $\delta
\psi^{ir}$, which is missing in \cite{ric}.\\
The self-duality condition \eq{he} serves as the full field equation
for the $2$-form potential $B_{\mu\nu}^I$. The remaining field equations
follow from the closure of the algebra \eq{final}. The resulting field
equations can be found in \cite{romans,ric}.
Summarizing, in this section we have shown that the (2,0) matter-coupled
Poincar\'{e} theory of \cite{romans, ric} can be reproduced by fixing the
conformal gauges in the (2,0) matter-coupled conformal supergravity
constructed in this paper.
\section{Conclusions}
In this paper we have constructed the local conformal supersymmetry
rules for $(2,0)$ supergravity in 6 dimensions. That includes the
transformation rules for the Weyl multiplet \eq{N=4}, which is the
gauge multiplet of the $OSp(8^*|4)$ superconformal algebra and the
transformation laws \eq{N=4T} of the tensor multiplet. The latter has
field equations given by \eq{fe2}.
These results can be viewed as the quadratic approximation to
the coupling of the full $M5$ brane theory to conformal supergravity in
a physical gauge. It would be interesting to obtain the full coupling
of the $M5$-brane to the $(2,0)$ conformal supergravity.\\
Taking $N+5$ copies of the tensor multiplets and imposing the
constraints described in section~\ref{ps}, reproduces
earlier results on Poincar\'e supergravity theory
coupled to $N$ tensor multiplets \cite{romans,ric}. The generalization
of these results to the case of $N$ coincident $M5$ branes is, of
course, a nontrivial problem. \\
We expect that the results obtained in this paper will have applications
to the study of the $AdS_7/CFT_6$ correspondence. So far, very few results
exist that deal with the calculation of the correlation functions
on the boundary of $AdS_7$ \cite{C1,C2}. Clearly, much remains to be done
to develop a better understanding of this correspondence and
the (2,0) conformal supergravity ought to play a role in this process.\\
Another open problem of interest is the construction of the higher spin
operators of the $(2,0)$ tensor multiplets \cite{rozali} and their
coupling to appropriate higher spin conformal supergravity fields. Of
special interest are the operators which correspond to massless higher
spin fields in the bulk of $AdS_7$. These arise from the product of two
doubleton representations of $OSp(8^*|4)$ \cite{g}. It is natural that
these operators couple to massless higher spin representation of this
group. A field theoretic realization of a higher spin $AdS_7$
supergravity is an interesting and challenging problem at present.
\bigskip\bigskip
\section*{Acknowledgments}
\noindent
We thank Per Sundell for useful discussions, Kostas Skenderis for a
discussion during the July 1998 Amsterdam {\it Workshop on String
Theory and Black Holes}, which triggered this investigation, and
Kor Van Hoof for indicating
corrections to a first version of this work. E.B.
thanks the institutes in Leuven and Texas A\&M, and E.S. and A.V.P.
thank the Institute for Theoretical Physics at Groningen for
hospitality. E.B. and A.V.P. thank the University of Utrecht for
hospitality.
This work was supported by the European Commission TMR
programme ERBFMRX-CT96-0045, in which E.B. is associated to Utrecht.
\bigskip
\begin{appendix} \section{Notations and Conventions}\la{app:notations}
We use the same notations as in \cite{ber1}, apart from the fact that
we now use indices from 0 to 5 with signature $(-+\cdots +)$
rather than the Pauli convention with indices from 1 to 6 with
signature $(+\cdots +)$. Therefore the
Levi--Civita tensor is adapted. We
replace in \cite{ber1}
\begin{equation}
i\epsilon_{abcdef} \rightarrow \epsilon_{abcdef}\ ,
\end{equation}
such that we now have
\begin{equation}
\epsilon_{012345}=1=-\epsilon^{012345}\,,\qquad
\gamma _7=\gamma^0\cdots \gamma ^5=-\gamma _0\cdots \gamma _5 \,.
\end{equation}
The essential formula is as in \cite{ber1}
\begin{equation}
\gamma _{abc}\gamma _7=-\tilde \gamma _{abc}\,,
\label{gamma7dual}
\end{equation}
where the dual is now defined in (\ref{defHdual}).
We raise and lower $USp(4)$ indices
with $\Omega ^{ij}$ as:
\begin{equation}
\lambda^i=\Omega^{ij}\lambda_j\ ,\quad\quad \lambda_i=\lambda^j \Omega_{ji}\ .
\end{equation}\\[-.75cm]
When $USp(4)$ indices are omitted, northwest-southeast contraction is
understood, e.g.
\begin{equation}
{\bar\lambda} \gamma^{(n)} \psi = {\bar\lambda}^i \gamma^{(n)} \psi_i\ ,
\end{equation}\\[-.75cm]
where we have used the following notation
\begin{equation}
\gamma^{(n)} = \gamma^{a_1\cdots a_n} =\gamma^{[a_1}\gamma^{a_2}\cdots \gamma^{a_n]}\ .
\end{equation}\\[-.75cm]
The anti-symmetrizations are always with unit strength. Changing the
order of spinors in a bilinear leads to the following signs
\begin{equation}
{\bar \psi}^{(1)} \gamma^{(n)} \chi ^{(2)} =
t_n
\ {\bar
\chi }^{(2)} \gamma^{(n)} \psi^{(1)}\ ,\qquad
\left\{ \begin{array}{c}
t_n=-1\mbox{ for }n=0,3,4 \\
t_n=1\mbox{ for }n=1,2,5,6
\end{array}\right.
\end{equation}\\[-.75cm]
where the labels $(1)$ and $(2)$ denote any $USp(4)$ representation,
e.g. $(1)=i$ and $(2)=[jk]$.
We frequently use the following Fierz rearrangement formula:
\begin{equation}
\psi_j {\bar\psi}^i= -\ft14({\bar\psi}^i\gamma_a\psi_j) \gamma^a
+\ft1{48}({\bar\psi}^i\gamma_{abc}\psi_j) \gamma^{abc}\ .
\end{equation}\\[-.75cm]
The notation ``- (trace)'' denotes terms that are proportional to
either $\Omega^{ij}$ or $\delta^i_j$ (with ``free'' indices). We use the
notation ``-({\rm traces})'' if both invariant tensors occur. For the
convenience of the reader we give below the explicit expressions of
some trace terms:
\begin{eqnarray}
X^{ij}-({\rm trace}) &=& X^{ij}+\ft14 \Omega^{ij} X^k{}_k\ ,
\nonumber\w2
A^{ij} X_k -({\rm traces}) &=& A^{ij}X_k +\ft45 A^{\ell[i} X_\ell
\delta^{j]}_k -\ft15 \Omega^{ij} A_{k\ell} X^\ell\ ,
\nonumber\w2
S_k{}^{[i} X^{j]} -({\rm traces }) &=& S_k{}^{[i} X^{j]} -\ft15
\delta^{[i}_k S^{j]\ell} X_\ell\ +\ft15 \Omega^{ij} S_k{}^\ell X_\ell\ .
\end{eqnarray}\\[-.75cm]
where $X^i$ and $X^{ij}$ are arbitrary $USp(4)$ tensors, while $A^{ij}$
is an antisymmetric traceless and $S^{ij}$ a symmetric tensor.
\section{The (2,0) $\rightarrow$ (1,0) Truncation}\la{app:truncation}
Many of the formulae for the $(2,0)$ Weyl and tensor multiplet can be
obtained by considering their truncations to the $(1,0)$ case and
comparing with the results of \cite{ber1}. Following \cite{conf98} the
$(2,0)$ Weyl multiplet may also be compared with the N=4, d=4 Weyl
multiplet of \cite{ber2}. \\
We first consider the $(2,0)$ Weyl multiplet. The $(2,0)$ Weyl multiplet
(\ref{N=4}) leads to the $N=2$ Weyl multiplet of \cite{ber1} (see
eq.~(2.26)) upon making the following truncations. We write $i=1,\cdots
,4 = (i =1,2, i^\prime =1,2)$, and we put
\begin{equation}
\Omega ^{ij}=\pmatrix{\epsilon ^{ij}&0\cr 0&\epsilon ^{i'j'}}\ .
\label{OmegaEpsilon}
\end{equation}
The non--vanishing bosonic component
fields are given by
\begin{eqnarray}
V_\mu^{ij} &=& V_\mu^{ij}\ ,
\nonumber
\w2
T_{abc}^{ij} &=& \epsilon^{ij}T_{abc}\, ,\hskip 1truecm T_{abc}^{i^\prime
j^\prime} = -\epsilon^{i^\prime j^\prime}T_{abc}\ ,
\w2
D^{ij}_{kl} &=& \epsilon^{ij}\epsilon_{kl}D\, ,\hskip .9truecm D^{ij}_{k^\prime
l^\prime} = -\epsilon^{ij}\epsilon_{k^\prime l^\prime}D\ ,
\hskip .5truecm
D^{i^\prime j^\prime }_{kl} = -\epsilon^{i^\prime j^\prime }\epsilon_{kl}D\ ,
\nonumber
\w2
D^{i^\prime j^\prime }_{k^\prime l^\prime } &=& \epsilon^{i^\prime j^\prime }
\epsilon_{k^\prime l^\prime} D\, ,\hskip .5truecm
D^{ij^\prime}_{k l^\prime} = - {\textstyle{1\over 2}}
\delta^i_k \delta^{j^\prime}_{l^\prime}D\ .
\nonumber
\end{eqnarray}\\[-.75cm]
For example, the first equation above means that
$V_\mu^{ij'}=0=V_\mu^{i'j'}$. The non-vanishing fermionic component
fields are given by
\begin{eqnarray}
\psi_\mu^i &=& \psi_\mu^i\ ,
\nonumber
\w2
\chi_{ij}^k &=& \epsilon_{ij}\chi^k\ ,\hskip .5truecm
\chi^k_{i^\prime j^\prime} = -\epsilon_{i^\prime j^\prime}\chi^k\ ,
\hskip .5truecm
\chi^{k^\prime}_{i^\prime j} = -{\textstyle{1\over 2}}
\delta^{k^\prime}_{i^\prime}\chi_j\ .
\end{eqnarray}\\[-.75cm]
Thus, for example, $\psi^{i'}=0$. Finally, the non--vanishing
supersymmetry parameters are given by
\begin{equation}
\epsilon^i = \epsilon^i\ ,\hskip 1.5truecm
\eta^i = \eta^i\ ,
\end{equation}\\[-.75cm]
which means that $\epsilon^{i'}=0=\eta^{i'}$. In comparing the truncated
result with the $(1,0)$ Weyl multiplet of \cite{ber1} two remarks are
in order. First of all the $(1,0)$ conventional constraints of
\cite{ber1} contain extra $\chi$-- and $D$-dependent terms which do not
generalize to the $(2,0)$ case. As a consequence the dependent K and S
gauge fields, obtained after truncation, differ from those of
\cite{ber1}. In order to obtain the truncated result one should replace
the K and S gauge fields of \cite{ber1} by the following expressions
\begin{eqnarray}
\la{redefinition}
f_\mu{}^a &\rightarrow& f_\mu{}^a +{\textstyle{1\over 240}}e_\mu{}^aD\ ,
\nonumber
\w2
\phi_\mu^i &\rightarrow& \phi_\mu^i -{\textstyle{1\over 60}}\gamma_\mu\chi^i \ .
\end{eqnarray}\\[-.75cm]
Secondly, in order to remove the $\chi$--dependent term from the
supersymmetry variation of the dilatation gauge field $b_\mu$ (again
this term cannot be generalized to the $(2,0)$ case) one must perform a
field-dependent $K$-transformation on the results of \cite{ber1} with
the following parameter
\begin{equation}
\la{comp}
\lambda_{K\mu} = -{\textstyle{1\over 60}}{\bar\epsilon}\gamma_\mu\chi\ .
\end{equation}\\[-.75cm]
The net effect of these manipulations is that all $\chi$-dependent
terms in the $(1,0)$ theory that cannot be extended to the $(2,0)$ case are
being removed.\\
Next we consider the (2,0) tensor multiplet.
The truncation of this multiplet to the (1,0) case treated in \cite{ber1}
is given by:
\begin{eqnarray}
\phi^{ij} &=& \epsilon^{ij}\sigma\, ,\hskip .5truecm
\phi^{i^\prime j^\prime} = -\epsilon^{i^\prime j^\prime}\sigma\, ,
\hskip .5truecm \phi^{ij^\prime} = 0\, ,\nonumber
\w2
\psi^i &=& \psi^i\, ,\hskip 1truecm \psi^{i^\prime} = 0\, .
\end{eqnarray}\\[-.75cm]
In order to show that the $\phi^{ij}$ field equation (see third
equation of \eq{fe2}) truncates correctly to the $(1,0)$ equation (see
eq.~(3.27) of \cite{ber1}) one has to take special care of the $\phi
D,\, {\bar\psi}_\mu\chi\phi$ and ${\bar\chi}\psi$ terms. Concerning the
$D\phi$ term, starting from the $(1,0)$ case, the redefinition
\eq{redefinition} of $f_a{}^a$ leads to an extra $D\sigma$ term
which is added to the explicit $D\sigma$ term in the equation of motion
(3.27) of \cite{ber1}. As for the ${\bar\psi}_\mu\chi\phi$ terms, the
redefinition of $\phi_\mu$ (see \eq{redefinition}) in the
${\bar\psi}_\mu{{\cal D}}\psi$ term plus the compensating K
transformation given in \eq{comp} lead to two extra
${\bar\psi}_\mu\chi\phi$ terms such that the total contribution cancels.
This is consistent with the fact that the truncation of the
${\bar\psi}_\mu\chi\phi$ term in \eq{box} vanishes identically. Finally,
the redefinition of $\phi^\mu$ in the ${\bar\psi}\phi_\mu$ term in $
{\cal D}^a {\cal D}_a$ (see eq.~(3.30) of \cite{ber1}) leads to
an extra ${\bar\chi}\psi$ term which should be added to the explicit
such contribution in the $\sigma$ field equation. The total then agrees
with the $(1,0)$ truncation of our result \eq{fe2}.
\section{Tensor, Current and Weyl Multiplets in Superspace}
The $(2,0)$ tensor multiplet in flat superspace can be described by a
superfield $\phi^{ij}$ satisfying the constraint \cite{how1}
\begin{equation}
D_\alpha^i \phi^{jk} = \Omega^{i[j}\lambda_\alpha^{k]} - ({\rm trace})\, .
\end{equation}\\[-.75cm]
In flat superspace the current multiplet \eq{deltatheta} is described
by the supercurrent \cite{how1}
\begin{equation}
J^{ij,kl} = \phi^{ij}\phi^{kl} - ({\rm trace})\ ,
\end{equation}\\[-.75cm]
where the superfield $\phi^{ij}$ describes the $(2,0)$ tensor
multiplet.\\
In superspace the Weyl multiplet \eq{N=4} is described by an
anti-selfdual superfield $W_{abc}^{ij}$ in the $\bf 5$ of USp(4), whose
first component is the bosonic field $T_{abc}^{ij}$ and which satisfies
the constraint
\begin{equation}
D_{\alpha i} W_{abc}^{jk} = \delta _i^{[j} \left (\gamma^{de}\gamma_{abc}\chi
^{k]}_{de}\right )_\alpha + \left (\gamma_{abc}\lambda^{jk}_i\right )_\alpha
- ({\rm trace})\ ,
\end{equation}\\[-.75cm]
where $\lambda_i^{jk}$ is in the $\bf 16$ of $USp(4)$.
\end{appendix}
\newpage
|
\section*{}
Many relaxational
processes in macroscopic systems are characterized by a
relaxation function $Q(t)$ that exhibits a stretched exponential
behavior,
\begin{equation}
Q(t)\sim Q(0)\exp[-(t/\tau)^{\beta}],\label{eq1}
\end{equation}
where $0<\beta<1$. Examples include viscoelastic relaxation
\cite{bib01}, dielectric relaxation \cite{bib02}, glassy relaxations
\cite{bib03,bib04,bib05},
relaxation in polymers \cite{bib06,bib07} and long-time decay in trapping
processes \cite{bib08}. Many more examples
\cite{bib09,bib10,bib11,bib12,bib13,bib14,bib15} suggest that (\ref{eq1})
is common to a very wide range of phenomena and macroscopic
materials.
The origin of the stretched exponential is not always clear.
In many cases it is assumed to be the result of a competition
between two exponential processes. In some cases, e.g., trapping
processes at long times, this assumption is well established,
while in others, such as relaxation in glassy materials, this
assumption has been controversially discussed \cite{bib16,bib17} and
alternative models have been also suggested
\cite{bib10,bib18,bib19,bib20}.
We have recently investigated the occurence of stretched exponential
behavior in finite systems, in cases where the relaxation arises due to two
competing exponential processes
\cite{bib21}
We have found that:
(a) the size of the system, although macroscopic, plays a dominant role in
the relaxation time
pattern, leading to an exponential decay sufficiently at long times;
(b) the crossover time, $t_\times$, to the exponential depends
logarithmically on the system size;
(c) the rate of the exponential decay also depends logarithmically on the
system size, and (d) in the special examples of the trapping and the
hierarchically constrained dynamics models the exponential relaxation may
enter before the stretched exponential is reached.
These results are of
relevance to experiments in confined systems, mesoscopic systems and to
Monte-Carlo simulations.
Our theoretical predictions on the finite size effects can serve as an
experimental test for identifying the origin of the mechanism
leading to stretched exponential decay.
We assume that the relaxation function of the whole
system can be represented by an integration over all possible
states $n$, namely,
\begin{equation}
Q(t)=\int_0^{\infty}\Phi(n) Q(n,t)dn.\label{eq2}
\end{equation}
Here, $\Phi(n)$ is the probability that state $n$ is occupied
and $Q(n,t)$ is the dynamic relaxation of the
$n$-th state.
Usually, in the case of
a stretched exponential behavior,
$\Phi(n)$ is assumed to behave as
$\Phi(n)\sim\exp(-an^{\alpha})$, while $Q(n,t)$ decays
exponentially with time as $Q(n,t)\sim\exp(-bt/n^{\gamma})$.
A number of dynamical models that yield a stretched
exponential decay can be formulated in terms of Eq.(\ref{eq2}).
These include the long-time behavior in the trapping
problem \cite{bib08}, the target problem \cite{bib20}, direct energy
transfer \cite{bib20}, trapping of nonidentical interacting
particles \cite{bib23},
hierarchically constrained dynamics \cite{bib16},
models for relaxation in microenulsions and molecular glasses \cite{bib24}
and others.
We now concentrate on three examples:
(i) A particle diffusing in a $d$-dimensional system with
randomly distributed static traps, where we are interested in the
survival probability $Q(t)$ of a particle.
Here the state $n$ represents a particle in a trap-free
region of linear size $n$; $\Phi(n)$ is the probability for
the occurance of a size $n$ trap-free region, and $Q(n,t)$
is the survival probability of the particle in this region \cite{bib08}.
The exponent $\alpha$ is the dimension $d$ of the system,
and $\gamma=2$ due to the diffusional motion.
(ii) A linear system (chain) along which two types of particles ($A$ and $B$)
are diffusing and interacting
via hard core interaction. However, only type $A$ can be
trapped by static traps which are randomly distributed along the
chain.
Here, $Q(t)$ is the survival probability of particles of type $A$,
$\Phi(n)$ is the probability that a free trap region of size $n$ occurs,
and $Q(n,t)$ is the survival probability of a type $A$ particle to survive
in this region. The exponent $\alpha$ is the dimension of the system $\alpha=1$ and
$\gamma=4$ is due to diffusion in the presence of hardcore iterations
\cite{bib25}.
(iii)
Hierarchically constrained dynamics, a model that has been
proposed to account for glassy relaxation \cite{bib16}.
This
model assumes that the relaxation of level $n$ populated
by spins, occurs in stages, and
the constraint imposed by a faster degree of freedom must
relax before a slower degree of freedom can relax. This
implies that the time scale of relaxation in one level is
subordinated to the relaxation below. A possible realization
considered in \cite{bib16} and here is a system with a discrete
series of levels where the relaxation time of level $n$
is
$\tau_n\sim n^\gamma$ (corresponding to the exponential
form of $Q(n,t)$ in (\ref{eq2})), and the weight factor of level $n$,
is $\Phi(n)\sim e^{-an}$ \cite{bib12}, corresponding to $\alpha=1$.
The first
exponential in (\ref{eq2}) is accordingly the probability to occupy level
$n$ and the second exponential represents the decay of that level.
The evaluation of the long time behavior of the integral in
(\ref{eq2}) is performed using the method of steepest descent. The main
contribution to the integral arises from the maximum of the
integrand in (\ref{eq2}), which is obtained from the minimum of the
function, $-an^{\alpha}-bt/n^{\gamma}$, appearing in the
exponent. This yields that the main contribution to (2)
comes from
\begin{equation}
n^{*}\cong (\gamma b t/\alpha a)^{1/(\alpha+\gamma)},\label{eq3}
\end{equation}
leading to \ref{eq1} with $\beta=\alpha/({\alpha+\gamma})<1$,
and $\tau=(\alpha/b\gamma) a^{-\gamma/\alpha}(\gamma/(\gamma+\alpha))^
{1+\gamma/\alpha}$.
However, as shown below, these arguments are valid only in the
thermodynamic limit where the system size is infinite. For a
finite number $N$ of traps (in the trapping system) or a
{\it finite\/} system with a finite number $N$ of spins (in the
hierarchical constraint system) the relaxation function depends explicitly
on $N$. Since our discussion is quite general for systems described
by (\ref{eq2}), in what follows we refer below to traps and spins in the
above examples as elements.
For a single finite
system consisting of $N$ elements, the relaxation function $Q(t)$
represents an
average quantity over the $N$ elements,
\begin{equation}
Q(t)={1\over N}
\sum_{\{n\}}m(n)Q(n,t),\label{eq4}
\end{equation}
where the sum is over all possible states $n$ and $m(n)$
is the number of elements at state $n$, with $\sum_{\{n\}}m(n)=N$.
Since the sum in (\ref{eq4}) is over exponential functions, the value
of $Q(t)$ will fluctuate for different sets of $N$.
There will be a distribution of $Q(t)$, and we are
interested in the typical $Q(t)$, which is around the peak of this
distribution.
In the thermodynamic limit
$N\to\infty$, all states $n$ are occupied,
$m(n)/N$ can be identified with $\Phi(n)$ and (\ref{eq2}) follows.
For $N$ finite, in contrast, there exists a
characteristic "maximum" state
$n=n_{\rm max}(N)$, and this $n_{\max}$
should replace the upper limit ($\infty$) in (\ref{eq2}),
\begin{equation}
Q(t)=\int_0^{n_{\max}} \Phi (n)Q(n,t)\,dn. \label{eq5}
\end{equation}
To estimate how $n_{\max}$ depends on $N$,
we note that the typical number of states $n$
in a sample of $N$
elements is $Z(n)\cong N \Phi(n)\cong N
\exp(-an^{\alpha})$. States with $Z(n)\ll 1$
will not occur in a typical system of $N$ elements, and this yields
\begin{equation}
n_{\max}\cong \left({\ln N\over a}\right)^{1/\alpha}.\label{eq6}
\end{equation}
If $n^*\ll n_{\max}$, the upper limit in (\ref{eq2}) can be approximated
by infinity and thus leads to (\ref{eq1}). However,
if $n^*\gg n_{\max}$ the main contribution to (\ref{eq5}) will
not be from the maximum of the integrand, which is
outside the range of integration, but from $n_{\max}$.
Thus, for $n^*\gg n_{\max}$ we expect
\begin{equation}
Q(t)\cong Q(0)e^{-bt/n_{\max}^{\gamma}}\label{eq7}
\end{equation}
where the time constant of the relaxation, $n_{\rm max}^\gamma$,
scales as $(\ln N)^{\gamma/\alpha}$.
The crossover time from a stretched exponential (\ref{eq1})
to an exponential (\ref{eq7}) can be estimated from the condition
$n^*=n_{\max}$, from which follows
\begin{equation}
t_{\times}\cong{\alpha a\over \gamma b}
\left({\ln N\over a}\right)^{1+\gamma/\alpha}.\label{eq8}
\end{equation}
The striking point in (\ref{eq8}) is the logarithmic dependence
on $N$, which puts $t_{\times}$ in the range of observable
time scales measurable in mesoscopic and even macroscopic systems.
Indeed, the corresponding relaxation value $Q(t_\times)$ scales as
\begin{equation}
Q(t_\times)\sim N^{-\alpha/\gamma},\label{eq9}
\end{equation}
independent of the microscopic parameters $a$ and $b$. For the above
three cases we find:
(i) In the case of the trapping relaxation mechanism where $\alpha=d$
and $\gamma=2$ we obtain,
\begin{equation}
Q(t_\times)/Q(0)\sim N^{-d/2}.\label{eq10}
\end{equation}
(ii) In the non identical case $\alpha=1$ and $\gamma=1$ and thus
\begin{equation}
Q(t_x)/Q(0)\sim N^{-1/4}.\label{eq11}
\end{equation}
(iii) In the hierarchical constraint dynamics
\begin{equation}
Q(t_\times)/Q(0)\sim N^{-1/\gamma}.\label{eq12}
\end{equation}
It is known [8e,23]
that in both examples, for an infinite system,
the stretched exponential behavior of (\ref{eq1}) sets in only at very long
times.. Thus we expect that in the finite system, the crossover will mask
the stretched-exponential pattern.
To test our analytical approach, we performed new Monte Carlo
simulations on two cases (i) and (iii), the trapping model (case (i)) and
the hierarchical constraint model (case (iii)).
In the trapping model, we consider one and two dimensional systems with
a fixed concentration $c=0.5$ of randomly distributed traps,
and vary the size $N/c$ of the system. We calculated numerically the
survival probability $Q(t)$ of a particle as a function of $t$ and
$N$.
In the hierarchical model we have chosen $\tau_n\sim n$ i.e., $\gamma=1$.
We calculated the relaxation function
for system sizes varying from $N=10^2$ to $N=10^5$.
As mentioned earlier, the relaxation function fluctuates for
different sets of $N$. For obtaining the typical behavior
of $Q(t)$, we have considered therefore the "typical"
average $Q(t)_{\rm typ}\equiv\exp(\langle\ln Q(t)\rangle$,
where the brackets denote an average over many sets of $N$ elements.
Note that an arithmetic average over $M$ sets of $N$ elements can not be employed
here, since it leads to a result identical for a larger system with
$M\times N$ elements (see \cite{bib04}). For a discussion of typical
averages see \cite{bib26}.
For simplicity, we shall drop the index "typ" in the following.
\begin{figure}
\epsfig{file=hbk345-fig1.ps,width=6cm,clip=,bbllx=120,bblly=140,bburx=450,bbury=740,angle=-90}
\caption{Plot of $-{\ln[Q(t)/Q(0)]}$ as a function of $t$ in
a double logarithmic presentation for ({\bf a}) the trapping model
in $d=1$ and $d=2$, and ({\bf b}) the hierarchical constraint
model, for several system sizes.
For the trapping model, the system sizes are $N=2\cdot10^3$ (open square),
$2\cdot10^5$ (open circle), $2\cdot10^7$ (open up triangle), $2\cdot10^9$ (open
down triangle) in $d=1$, and $N=9\cdot10^2$ (full square), $9\cdot10^4$ (full
circle), $9\cdot10^6$ (full up triangle) in $d=2$.
For the hierarchical model, the system sizes are $N=10^2$ (full square), $10^3$
(full circle), $10^4$ (full up triangle), $10^5$ (full down triangle).}
\label{fig01}
\label{fig01a}
\label{fig01b}
\end{figure}
Figure 1 shows $-{\ln[Q(t)/Q(0)]}$ as a function of $t$ in
a double logarithmic plot for (i) the trapping model
in $d=1$ and $d=2$, and (iii) the hierarchical constraint
model, both for several system sizes. In all cases, a crossover
from an exponent $\beta < 1$ (at small $t$) towards $\beta=1$ (at large $t$)
can be easily recognized.
The crossover time $t_\times$ shifts towards larger values
when $N$ increases.
\begin{figure}
\epsfig{file=hbk345-fig2.ps,width=6cm,clip=,bbllx=140,bblly=210,bburx=450,bbury=760,angle=-90}
\caption{Plot of the local exponents $\beta$ calculated from the successive
slopes of the corresponding curves in (\ref{fig01}), ({\bf a})
for the trapping model and ({\bf b}) for the
hierarchical model. The horizontal dashed lines represent the
corresponding asymtotic
($N\to \infty, \quad t\to \infty$) values of $\beta$.}
\label{fig02}
\label{fig02a}
\label{fig02b}
\end{figure}
To study the crossover behavior in a more quantitative manner,
we have plotted in Fig.~\ref{fig02} the local exponents $\beta$ obtained from
the local slopes of Fig.~\ref{fig01}, as a function of $t$. In both systems,
for a fixed system size $N$, $\beta$ first decreases with $t$, reaches
a minimum value at a certain time that can be identified with $t_\times$,
and then increases monotonically with time towards $\beta=1$. The figure shows
that the minimum value of $\beta$ has not yet reached its asymptotic
value predicted for infinite systems, i.e., $\beta=1/3$ ($d=1$)
and $\beta=1/2$ ($d=2$) for the trapping system and $\beta=1/2$ for the
hierarchical system.
To show the dependence of the crossover time $t_\times$ on the system size $N$
we have plotted, in Fig.~\ref{fig03}, the values of $t_\times^{\alpha/(\alpha+\gamma)}$
as a function of
$\ln N$. The crossover time was obtained numerically from the
position of the minima of the
curves in Fig.~\ref{fig02}. The resulting straight lines are in
full agreement with the prediction of (\ref{eq8}), supporting our
analytical approach.
\begin{figure}
\epsfig{file=hbk345-fig3.ps,width=6cm,clip=,bbllx=140,bblly=200,bburx=440,bbury=750,angle=-90}
\caption{Plot of $t_\times^{\alpha/(\alpha+\gamma)}$
as a function of $\ln N$,
for ({\bf a}) the trapping model and ({\bf b}) the hierarchical model.
The straight
line supports (\ref{eq8}).
The crossover times $t_\times$ were obtained from
the positions of the minima of Fig.~\ref{fig02}}
\label{fig03}
\label{fig03a}
\label{fig03b}
\end{figure}
In the following we discuss the relevance of our results to
Monte-Carlo simulations and experiments. There exists a
long standing puzzle
in Monte-Carlo simulations of the trapping problem in $d=2$ and 3, that
the predicted stretched exponential could not be observed \cite{bib08}, even
for survival probabilities $Q(t)/Q(0)$ down to
$10^{-21}$ in $d=2$ [8b] and $10^{-67}$ in $d=3$ [8g].
The finding of the logarithmic dependence of $Q(t)$ on the system
size $N$ explains this puzzle.
The Monte-Carlo simulations in $d=2$ and 3 were typically performed
on $10^3$ configurations with about $10^4$ traps, which is
equivalent to having a single system with $N\sim 10^7$ traps.
Using
(\ref{eq10}), we expect for $N=10^7$ traps $Q(t_\times)/Q(0)\cong 10^{-7}$ in
$d=2$. Indeed, for times above $t_\times$ the exponent $\beta$ aproaches unity
as predicted by our theory and seen clearly in Fig.~\ref{fig02a}a. Moreover, for this
system size $\beta$ never reaches the predicted thermodynamic value $\beta =
0.5$, the minimum value of $\beta$ is about $0.65$.
For $d=3$, $Q(t_\times)/Q(0)\cong 10^{-11}$ thus for smaller survival values
($t>t_\times$) one again expects increasing values of $\beta$ approaching
unity.
This explains the exponential decay found
in the early Monte-Carlo simulations. Our results show that this is not
an artefact but due to the finite size of the system.
Moreover, they clearly
indicate that the thermodynamic limit can not even be reached in
one-dimensional macroscopic systems.
It would be of interest to test the above
prediction experimentally by preparing experimental
realizations where size effects can be controlled.
Equations (8) and (10) suggest that the behavior
around the crossover can be measured experimentally.. For
the trapping problem in linear systems, which has been studied experimentally
\cite{bib27,bib28}, we expect for $10^8$ sites and concentrations of traps $c$
between
$10^{-4}$ and $10^{-2}$, that $Q(t_\times)/Q(0)\sim10^{-2}\div10^{-3}$,
which is a survival range that can be detected experimentally.
For the non identical particles (case (iii)), we expect for $10^8$ sites
and concentration of traps $c$ between $10^{-4}$ and $1$ that
$Q(t_x)/Q(0)\sim10^{-1}\div10^{-2}$ which is a survival range
that can be well detected experimentally.
The same arguments are valid for the target problem and therefore a similar
crossover from stretched exponential to exponential decay is expected in
relaxation experiments in low dimensional geometries \cite{bib29}.
Mesoscopic systems such as quantum dots, are also promising
candidates for experiments where the crossover can be relevant.
Identifying
the logarithmic size dependence in experiments
may provide support to
the theories claiming that the observed stretched
exponential is due to competing exponential processes, represented
by (\ref{eq2}).
This work was supported by the German Israeli Foundation (GIF).
|
\section{Introduction}
Almost dating back to the development of QCD itself, supersymmetric
versions of QCD have been closely studied, as tractable laboratories
for extracting exact analytic information about both perturbative and
non-perturbative phenomena in nonabelian gauge theories.
One outstanding puzzle, unresolved since the mid-1980's, concerns the
calculation of the gluino condensate $\Vev{\mathop{\rm tr}\nolimits \lambda^2\over16\pi^2}$
in these models. This is an interesting quantity, as it is a measure
of chiral symmetry breakdown. In pure supersymmetric Yang-Mills (SYM)
theory, by dimensional analysis, one expects
\begin{equation}
\VEV{\mathop{\rm tr}\nolimits \lambda^2\over16\pi^2}\ =\ c\Lambda^3\ ,
\elabel{expect}\end{equation}
where $\Lambda$ is the dynamical scale in the theory (developed by
dimensional transmutation as in QCD), while $c$ is a numerical
constant. Remarkably, there are two approaches in the
literature for calculating
$\Vev{\mathop{\rm tr}\nolimits \lambda^2\over16\pi^2}$, each purporting to be exact (i.e.,
nonrenormalized), but which differ in their predictions of the
constant $c$. This disagreement is especially vexing in light of the
fact that both involve the use of supersymmetric instantons. The first
approach, generally known as ``strong-coupling instanton'' (SCI)
calculations, was developed in
Refs.~\cite{Novikov:1983ee,Rossi:1984bu,Amati:1985uz,Amati:1988ft,Fuchs:1986ft},
while the second approach,
generally known as ``weak-coupling instanton'' (WCI) calculations, was
developed in Refs.~\cite{Fuchs:1986ft,Affleck:1983rr,Novikov:1985ic,Shifman:1988ia};
for self-containedness, both will be reviewed below.
In this paper, we re-examine this old controversy, using our recently
developed methods for studying supersymmetric \it multi\rm-instantons
\cite{MO-I,MO-II,KMS,MO-III}. In particular, by looking at $n$-point
correlators
$\Vev{{\mathop{\rm tr}\nolimits\lambda^2(x_1)\over16\pi^2}\cdots{\mathop{\rm tr}\nolimits\lambda^2(x_n)\over16\pi^2}}$
of the gluino condensate, we will be able to probe arbitrary
topological numbers $k$. In a nutshell, our results cast serious doubt
on the validity of the SCI calculations of the
condensate. Specifically, we will demonstrate that an essential
technical step in the SCI approach, namely the use of cluster
decomposition, is invalid. The important implications of this
observation are as follows. Since cluster decomposition is an
essential requirement of quantum field theories (with very mild
assumptions that are certainly met by SYM theory), the exact quantum
correlators must have this property. That cluster is violated by the
instanton-saturated SCI correlators then means that (contrary to
claims in the literature) the SCI approximation is only giving \it
part \rm of the full answer. Since the SCI correlators obey
supersymmetric perturbative nonrenormalization theorems
\cite{Novikov:1985ic}, it necessarily follows that additional
$non$-perturbative objects must be contributing to the correlators. A
fuller discussion of this point is given in Sec.~VII below; however,
categorizing the nature of these non-perturbative configurations is
beyond the scope of the present paper.\footnote{See however
Ref.~\cite{DHKM} where, in a compactified version of the present
theory, the important role played by monopoles is emphasized.} We
should add that we believe that, in contrast, the WCI correlators are
consistent with cluster decomposition.
In addition, we will address an ingenious, if
controversial,\footnote{See Ref.~\cite{Csaki:1997aw} and the rebuttal
Ref.~\cite{Kogan:1998dt}.} hypothesis of Shifman's, in which the
numerical disagreement between the SCI and WCI results is taken as
circumstantial evidence for the existence of an extra disconnected
vacuum in SYM theory in which chiral symmetry is unbroken
\cite{Kovner:1997im,Shifman:1999mv}. While this so-called
``Kovner-Shifman (KS) vacuum'' can indeed potentially resolve the
disagreement at the 1-instanton sector ($k=1$), we will show that it
fails to do so for the topological sectors with $k>1.$ In other words,
positing a KS vacuum cannot by itself restore the cluster property to
the SCI correlators. This discouraging finding might be viewed as
removing some of the impetus for positing such a vacuum in the first
place.
Finally we will present a novel calculation of
$\Vev{\mathop{\rm tr}\nolimits\lambda^2(x)\over16\pi^2}$ which relates the ${\cal N}=1$
supersymmetric models discussed herein to the exactly soluble
Seiberg-Witten models with ${\cal N}=2$ supersymmetry. This calculation is
of potential pedagogical interest because it bypasses the explicit use
of instantons, and instead relies on functional methods. Not
surprisingly, it recaptures the WCI answer.
Let us sketch in broad strokes the main differences between the SCI
and WCI calculations (a more detailed review will follow). For ${\cal N}=1$
supersymmetric $SU(N)$ gauge theory with no matter, the leading
coefficient of the $\beta$-function is $b_0=3N$, so that
$\Lambda^3$ goes like an ``$N^{\rm th}$ root'' of an instanton:
$\Lambda^3\propto\exp(-8\pi^2/g^2N).$ This means that a na\"\i ve
1-instanton calculation of $\Vev{\mathop{\rm tr}\nolimits\lambda^2\over16\pi^2}\,$---in which
$\lambda$ is simply replaced by its ``classical value'' as an adjoint
fermion zero mode in the instanton background, and all the instanton
collective coordinates, both bosonic and fermionic, are integrated
over---fails; specifically it gives a zero answer, due to unsaturated
Grassmann integrations. In order to perform a sensible 1-instanton
calculation of $\Vev{\mathop{\rm tr}\nolimits\lambda^2\over16\pi^2}$, two alternative, and
necessarily more elaborate, approaches suggest themselves. In the SCI
approach, one calculates the $N$-point correlator of this
condensate, which scales like $\exp(-8\pi^2/g^2),$ and is indeed
nonzero at the 1-instanton level. Furthermore, by a Ward identity
reviewed in the Appendix, it
is independent of the $N$ space-time insertion points $x_i.$ After
performing the requisite collective coordinate integration, one finds:
\begin{equation}
\VEV{{\mathop{\rm tr}\nolimits\lambda^2(x_1)\over16\pi^2}\cdots{\mathop{\rm tr}\nolimits\lambda^2(x_N)\over16\pi^2}}\
=\ {2^N\over(N-1)!\,(3N-1)}\,\Lambda^{3N}\ .
\elabel{SCIans}\end{equation}
In order to extract $\Vev{\mathop{\rm tr}\nolimits\lambda^2\over16\pi^2}$ from the correlator
\eqref{SCIans}, one then invokes cluster decomposition: taking
$|x_i-x_j|\gg\mu^{-1}$ where $\mu$ is the mass gap in this theory, and
remembering the constancy of the correlator, one replaces the
left-hand side of Eq.~\eqref{SCIans} simply by
$\Vev{\mathop{\rm tr}\nolimits\lambda^2\over16\pi^2}^N.$ The net result thus reads:
\begin{equation}
\VEV{\mathop{\rm tr}\nolimits\lambda^2\over16\pi^2}\ =\
{2\over\big((N-1)!\,(3N-1)\big)^{1/N}}\ \Lambda^{3}\,e^{2\pi
iu/N}\qquad\hbox{(SCI result)}\ ,
\elabel{SCIansb}\end{equation}
where $u=0,\ldots,N-1$ indexes the $N$ vacua $\ket{u}$
of the $SU(N)$ theory, and reflects the ambiguity in taking the
$N^{\rm th}$ root of unity. In retrospect (as argued in
Refs.~\cite{Amati:1985uz,Amati:1988ft}),
the reason why the na\"\i ve calculation of
$\Vev{\mathop{\rm tr}\nolimits\lambda^2\over16\pi^2}$ gives zero is that these $N$ vacua
are being averaged over and the phases cancel.
In contrast, in the WCI approach, one modifies the pure gauge theory by adding
matter superfields in such a way that $\Vev{\mathop{\rm tr}\nolimits\lambda^2\over16\pi^2}$
itself (rather than a higher-point function thereof) receives a
nonzero contribution at the 1-instanton level. Next, one decouples
these extraneous matter fields by giving them a mass $M$, and taking the joint
limit $M\rightarrow\infty$ and $\Lambda\rightarrow0$ in the
manner dictated by renormalization group (RG)
decoupling. Matching onto the effective low-energy theory without
matter gives:
\begin{equation}
\VEV{\mathop{\rm tr}\nolimits\lambda^2\over16\pi^2}\ =\
\Lambda^{3}\qquad\hbox{(WCI result)}\ .
\elabel{WCIans}\end{equation}
Note that the RG decoupling procedure forces the low-energy theory
into one of the $N$ degenerate vacua $\ket{u}$, which by
convention we take to be the one with real phase. The nomenclature
``strong coupling'' versus ``weak coupling'' used to designate these
differing approaches refers to the fact that, in the former, as in
QCD, the only scale in the problem is the dynamical scale $\Lambda,$
whereas in the latter, the existence of VEVs ${\rm v}_i$ of the matter
superfields permit a standard semiclassical expansion when the
dimensionless ratios $\Lambda/{\rm v}_i$ are all small. (The holomorphic
properties of SYM theory then permit the analytic continuation of the
answer beyond this regime.)
As mentioned above, it is possible to reconcile the two calculations
\eqref{SCIansb} and \eqref{WCIans} by positing the existence of an extra vacuum
$\ket{S}$ in which the condensate vanishes \cite{Kovner:1997im}.
Specifically, if $p$ and $1-p$ represent
the probability weights in the vacuum sector of the theory for the
standard vacua $\{\ket{u}\}$, and for $\ket{S}$, respectively, and if one
takes
\begin{equation}p\ =\ {2^N\over(N-1)!\,(3N-1)}\ ,
\elabel{Shifmana}\end{equation}
then both the 1-instanton results can be understood. Unfortunately,
the multi-instanton calculations presented below show that the
mismatch between the SCI and the WCI calculations becomes more severe
for higher topological number $k$, and apparently cannot be reconciled
in this way for $k>1.$
This paper is organized as follows. In Secs.~II and III, respectively, we review the SCI
and WCI calculations of
$\Vev{\mathop{\rm tr}\nolimits\lambda^2\over16\pi^2}$ for general gauge
group $SU(N)$. Also in Sec.~III we
present an alternate, non-instanton-based derivation of this
condensate, specific to the gauge group $SU(2)$, which starts from the
Seiberg-Witten solution of the ${\cal N}=2$ model \cite{Seiberg:1994rs}
and flows to the ${\cal N}=1$
model, recapturing the WCI result. In Sec.~IV we discuss cluster
decomposition in more depth, and motivate Shifman's proposal for
reconciling the SCI and WCI calculations by postulating an extra
vacuum state. Our principal results are described in Secs.~V and VI, in which
(extending the SCI approach) we calculate higher-point functions of
the condensate, in the topological sectors $k>1.$ In
Sec.~V we calculate, analytically, the $(kN)$-point functions
$\Vev{{\mathop{\rm tr}\nolimits\lambda^2(x_1)\over16\pi^2}\cdots{\mathop{\rm tr}\nolimits\lambda^2(x_{kN})\over16\pi^2}}$
in $SU(N)$ gauge theory for arbitrary instanton number $k$, but to
leading order in $1/N,$ while in Sec.~VI we calculate, numerically, the
4-point function
$\Vev{{\mathop{\rm tr}\nolimits\lambda^2(x_1)\over16\pi^2}{\mathop{\rm tr}\nolimits\lambda^2(x_2)\over16\pi^2}
{\mathop{\rm tr}\nolimits\lambda^2(x_3)\over16\pi^2}
{\mathop{\rm tr}\nolimits\lambda^2(x_{4})\over16\pi^2}}$ at the 2-instanton level for
gauge group $SU(2)$. In either case our SCI calculations explicitly contradict
the hypothesis of cluster decomposition---both with and without an
extra KS vacuum.\footnote{The numerical calculation is based on a
Monte Carlo integration, which (with our present statistics) is
incompatible with the clustering result at the $5{\textstyle{1\over2}}$ sigma level,
and incompatible with the modified clustering result due to the
incorporation of a KS vacuum (tuned to reconcile the WCI and SCI
1-instanton results), at the $11{\textstyle{1\over2}}$ sigma level.}
Concluding comments are made in Sec.~VII.
\setcounter{equation}{0}
\section{Review of the Strong-Coupling Instanton Calculation}
Let us review the SCI result for
$\Vev{\mathop{\rm tr}\nolimits\lambda^2(x_1)\cdots\mathop{\rm tr}\nolimits\lambda^2(x_N)}$, for pure ${\cal N}=1$
$SU(N)$ gauge theory. The calculation for done originally for the
$SU(2)$ theory in \cite{Novikov:1983ee} and then extended to the
$SU(N)$ theories in \cite{Amati:1985uz} (see also the very comprehensive
review articles \cite{Amati:1988ft,Shifman:1999mv}).
The correlator in question is saturated at the 1-instanton level.
The gauge-invariant collective coordinate integration measure is a
suitable generalization of the Bernard measure \cite{Bernard} to an
${\cal N}=1$ theory,
and reads:\footnote{Our choice of notation is dictated by the
$k$-instanton generalization of this measure, Eq.~\eqref{GImeas}
below. Following Ref.~\cite{Finnell:1995dr}, we correct a factor of two
mistake in the normalization of adjoint fermion zero modes that
pervades much of the literature (e.g.,
Refs.~\cite{Amati:1988ft,Fuchs:1986ft}). Hence our final result for
the $N$-point function, Eq.~\eqref{SCIans}, differs by $2^N$ from
these references.}
\begin{equation}
-{2^{3N+2}\,\pi^{2N-2}\Lambda^{3N}\over(N-1)!\,(N-2)!}\int
d^4a'\,d\rho^2\,(\rho^2)^{2N-4}\,d^2{\cal M}'\,
d^2\zeta\,d^{N-2}\nu\,d^{N-2}\bar \nu\ .
\elabel{measone}\end{equation}
Here $a'_n$ is (minus) the
4-position of the instanton and $\rho$ is its scale size, the
Grassmann spinors ${\cal M}'_\alpha$ and $\zeta_{\dot\alpha}$ parametrize the
supersymmetric and superconformal modes, respectively, of the gluino,
and the Grassmann parameters $\nu_{u'}$ and $\bar \nu_{u'}$,
$u'=1,\ldots,N-2,$ are the superpartners to the iso-orientation modes
which sweep the instanton through $SU(2)$ subgroups of the $SU(N)$
gauge group (note that each $\nu_{u'}$ and $\bar \nu_{u'}$ is a
Grassmann number rather than a Grassmann spinor). The measure includes
the Lambda parameter of the Pauli-Villars (PV) scheme which at the two-loop
level is \cite{Amati:1988ft}
\begin{equation}
\Lambda=g(\mu)^{-2/3}e^{-8\pi^2/(3Ng(\mu)^2)}\mu\ .
\end{equation}
Into this measure
one inserts $\prod_{i=1}^N\mathop{\rm tr}\nolimits\lambda^2(x_i)$ where $\lambda^\alpha(x)$
is the most general classical adjoint fermion zero mode in the
1-instanton background. In terms of these bosonic and fermionic
collective coordinates, one derives (see Eq.~\eqref{corriganid} below):
\begin{equation}\begin{split}\mathop{\rm tr}\nolimits\lambda^2(x)\ =\ -
{\textstyle{1\over4}}\square\Big(&{2\over\rho^2+y^2}\sum_{u'=1}^{N-2}
\bar \nu_{u'}\nu_{u'}\ +\ \zeta_{\dot\alpha}\zeta^{\dot\alpha}\,{y^4\over(\rho^2+y^2)^2}
\\&+\ {\cal M}^{\prime\alpha}{\cal M}'_\alpha\,{2\rho^2+y^2\over(\rho^2+y^2)^2}
\ -\ {\cal M}^{\prime\alpha}y_{\alpha{\dot\alpha}}\zeta^{\dot\alpha}\,{2\rho^2\over(\rho^2+y^2)^2}\Big)
\elabel{insertdef}\end{split}\end{equation}
where
\begin{equation}
y_{\alpha{\dot\alpha}}
\ =\ x_{\alpha{\dot\alpha}}+a'_{\alpha{\dot\alpha}}\ =\
(x_n+a'_n)\sigma^n_{\alpha{\dot\alpha}}\ .
\elabel{ydef}\end{equation}
Now let us carry out the Grassmann integrations in Eq.~\eqref{measone}.
Obviously the $\zeta$ and ${\cal M}'$ Grassmann integrations will be
saturated from the condensates inserted at two points
$\{x_i,x_j\}$ chosen from among the $N$ insertions
$x_1,\ldots,x_N$. For each such pair there are three contributions to
these integrals:
\def{\textstyle{1\over16}}{{\textstyle{1\over16}}}
\begin{subequations}
\begin{align}
&{\textstyle{1\over16}}\square_i\square_j\,{y_i^4(2\rho^2+y_j^2)\over
(\rho^2+y_i^2)^2(\rho^2+y_j^2)^2}\qquad(\zeta^2\hbox{ at } x_i,\
{\cal M}^{\prime2}\hbox{ at }x_j)\ ,\elabel{threecona}\\
&{\textstyle{1\over16}}\square_i\square_j\,{(2\rho^2+y_i^2)y_j^4\over
(\rho^2+y_i^2)^2(\rho^2+y_j^2)^2}\qquad({\cal M}^{\prime2}\hbox{ at } x_i,\
\zeta^{2}\hbox{ at }x_j)\ ,\elabel{threeconb}\\
&{\textstyle{1\over16}}\square_i\square_j\,{2\rho^4\,y_i\cdot y_j\over
(\rho^2+y_i^2)^2(\rho^2+y_j^2)^2}\qquad(\zeta\times{\cal M}'\hbox{ at } x_i,\
\zeta\times{\cal M}^{\prime}\hbox{ at }x_j)\ .\elabel{threeconc}
\end{align}
\end{subequations}
Adding these three contributions gives the simpler expression
\begin{equation}
{-36\rho^8\,(x_i-x_j)^2\over (\rho^2+y_i^2)^4(\rho^2+y_j^2)^4}\ .
\elabel{adding}\end{equation} Now we take advantage of the fact that
this $N$-point function is independent of the $x_i$ (see the
Appendix), to choose these insertion points for maximum simplicity of
the algebra. The simplest conceivable such choice, $x_i=0$ for all
$i,$ turns out to give an ill-defined answer of the form
``$0\times\infty$'' (the zero coming from the Grassmann integrations
as follows from Eq.~\eqref{adding}, and the infinity from divergences
in the $\rho^2$ integration due to coincident poles). In order to
sidestep this ambiguity, one chooses instead:
\begin{equation}
x_1=\cdots=x_{N-1}=0\ ,\qquad x_N=x\ .
\elabel{insertionone}\end{equation} This choice is the simplest one
which gives a well-defined answer with no ``$0\times\infty$''
ambiguity. More ambitiously, one can still perform the calculation
even if all the insertion points are taken to be arbitrary
\cite{Amati:1985uz,Amati:1988ft}; however, we find it convenient for
later to take the minimal resolution provided by
\eqref{insertionone}. From the $(x_i-x_j)^2$ dependence in
Eq.~\eqref{adding}, it follows that the pair of insertions
$\{x_i,x_j\}$ responsible for the $\{\zeta,{\cal M}'\}$ integrations must
include the point $x_N=x$; there are $N-1$ possible such pairs, giving
\begin{equation}
{-36(N-1)\rho^8\,x^2\over
\big(\rho^2+(x+a')^2\big)^4\,(\rho^2+a^{\prime2})_{}^4}
\elabel{addingb}\end{equation}
for these contributions. The remaining Grassmann integrations over
$\{\nu,\bar \nu\}$ are saturated at $x_i=0,$ and give
\begin{equation}
(N-2)!\,\Big({4\rho^2\over
(\rho^2+a^{\prime2})^3}\Big)^{N-2}\ .
\elabel{nuresult}\end{equation}
Combining the denominators in Eqs.~\eqref{addingb}-\eqref{nuresult} with a Feynman
parameter $\alpha$,
\begin{equation}
{1\over(\rho^2+a^{\prime2})^{3N-2}}\,
{1\over(\rho^2+(x+a^{\prime})^2)^4}\ =\
{(3N+1)!\over3!\,(3N-3)!}\int_0^1d\alpha\,{\alpha^3(1-\alpha)^{3N-3}\over\big(
\rho^2+(a'+\alpha x)^2+\alpha(1-\alpha)x^2\big)^{3N+2}}
\elabel{alphadef}\end{equation}
and performing the $d^4a'$ integration then yields:
\begin{equation}\begin{split}
\VEV{\mathop{\rm tr}\nolimits\lambda^2(x_1)\cdots\mathop{\rm tr}\nolimits\lambda^2(x_N)}\ &=\
{2^{3N+2}\,\pi^{2N-2}\Lambda^{3N}\over(N-1)!\,(N-2)!}\int_0^1d\alpha
\int_0^\infty d\rho^2\,(\rho^2)^{2N-4}
\\&\quad\quad\times\
\big(36(N-1)\rho^8x^2\big)(N-2)!\,(4\rho^2)^{N-2}
\\&\quad\quad\times\
{(3N+1)!\over3!\,(3N-3)!}\,{\alpha^3(1-\alpha)^{3N-3}\,\pi^2\over3N(3N+1)\big(
\rho^2+\alpha(1-\alpha)x^2\big)^{3N}}\\&=\
{3(3N-2)\,2^{5N-1}\,\pi^{2N}\Lambda^{3N}\over(N-2)!}\int_0^1d\alpha\,\alpha^2(1-\alpha)^{3N-4}
\\&=\ {2^{5N}\,\pi^{2N}\Lambda^{3N}\over(N-1)!\,(3N-1)}\ ,
\elabel{finalone}\end{split}\end{equation}
in agreement with Eqs.~\eqref{SCIans}-\eqref{SCIansb}.
\setcounter{equation}{0}
\section{Review of the Weak-Coupling Instanton Calculation}
Next, let us review the WCI calculation of the gluino condensate. As
mentioned above, the general WCI strategy is to extend the pure gauge
theory to include matter content, in such a way that
$\Vev{\mathop{\rm tr}\nolimits\lambda^2\over16\pi^2}$ receives a nonzero contribution at
the 1-instanton level. Decoupling the extraneous matter and matching
to the low-energy pure gauge theory is then accomplished using
standard RG prescriptions. Since the precise nature of this extraneous
matter is rather arbitrary, the WCI calculation really stands for a
family of related calculations sharing this basic approach, all of
which give the same result \eqref{WCIans}. Calculations of this type
were done in
\cite{Affleck:1983rr,Novikov:1985ic,Fuchs:1986ft,Shifman:1988ia}
and reviewed in \cite{Shifman:1999mv}.
We will find it efficient to exploit the functional identity (see for
example \cite{Peskin:1997qi}):
\def{\cal W}_{\rm eff}{{\cal W}_{\rm eff}}
\def{\partial\over\partial\tau}{{\partial\over\partial\tau}}
\def{\partial\over\partial\Lambda}{{\partial\over\partial\Lambda}}
\begin{equation}
\VEV{\mathop{\rm tr}\nolimits\lambda^2}\ =\ -8\pi i\VEV{{\partial\over\partial\tau}{\cal W}_{\rm eff}}\ =\
{16\pi^2\over b_0}\VEV{\Lambda{\partial\over\partial\Lambda}{\cal W}_{\rm eff}}\ .
\elabel{fcnlid}\end{equation}
Here ${\cal W}_{\rm eff}$ is the effective superpotential,
\begin{equation}
\tau\ =\ {4\pi i\over g^2}+{\theta\over2\pi}
\elabel{taudef}\end{equation}
is the usual complexified coupling, and
\begin{equation}
\Lambda\ =\ \mu\,e^{2\pi i\tau(\mu)/b_0}
\elabel{Lambdadef}\end{equation}
is the RG-invariant 1-loop dynamical scale of the theory. This result comes
from writing the microscopic gauge theory as
\def{\cal L}{{\cal L}}
\def{\rm Im}{{\rm Im}}
\begin{equation}
{\cal L}\ =\ {1\over4\pi}\,{\rm Im}\big(\tau\int
d^2\theta\,\mathop{\rm tr}\nolimits\,W^{\alpha}W_\alpha\big)\ ,
\elabel{microgt}\end{equation}
where $W^\alpha$ is the gauge field-strength chiral superfield,
and promoting $\tau$ to a ``spurion superfield'',
\defW^a\ =\ \lambda^a+\cdots{W^a\ =\ \lambda^a+\cdots}
\begin{equation}
\tau\ \rightarrow\ T(y,\theta)\ =\
\tau(y)+\sqrt{2}\,\theta^\alpha\chi^\tau_\alpha(y) +\theta^2F^\tau(y)\ .
\elabel{taupromote}\end{equation}
{}From Eqs.~\eqref{microgt}-\eqref{taupromote} it trivially follows that
\def{\cal Z}{{\cal Z}}
\def\,{\Big|}_{T(y,\theta)=\tau}{\,{\Big|}_{T(y,\theta)=\tau}}
\begin{equation}
\VEV{\mathop{\rm tr}\nolimits\lambda^2}\ =\ {8\pi\over{\cal Z}}{\delta\over\delta
F^\tau(x)}\,{\cal Z}\,{\Big|}_{T(y,\theta)=\tau}\ ,
\elabel{Zone}\end{equation}
where
\def{\cal D}{{\cal D}}
\begin{equation}{\cal Z}\ =\ \int{\cal D} W\,e^{i\int d^4x\,{\cal L}}
\elabel{Ztwo}\end{equation}
is the partition function of the
microscopic theory, in the generalized background field
\eqref{taupromote}. In order to derive Eq.~\eqref{fcnlid} from
Eq.~\eqref{Zone}, one assumes that the functional differentiation
indicated in Eq.~\eqref{Zone} formally commutes with the
integrating-out of the microscopic degrees of freedom. In other words,
${\cal Z}$ can be re-expressed in terms of the relevant effective chiral
superfields $\Phi_i$ (whatever these may be\footnote{In the example
culminating in Eq.~\eqref{adsdefc} below, we will find that there are
in fact no residual chiral superfields `$\Phi_i$', so that simply
${\cal W}_{\rm eff}={\cal W}_{\rm eff}(T),$ whereas in the Seiberg-Witten example \eqref{WSWdef}
below, the $\Phi_i$ are the monopole superfields $M$, $\tilde M$ as
well as the dual Higgs $A_D$.}):
\def\Z_{\rm eff}{{\cal Z}_{\rm eff}}
\begin{equation}\VEV{\mathop{\rm tr}\nolimits\lambda^2}\ =\ {8\pi
\over\Z_{\rm eff}}{\delta\over\delta F^\tau(x)}\,\Z_{\rm eff}\,{\Big|}_{T(y,\theta)=\tau}\ ,
\elabel{Zthree}\end{equation}
where
\begin{equation}
\Z_{\rm eff}\ =\ \int{\cal D}\Phi_i\,e^{-i\int d^4x\int
d^2\theta\,{\cal W}_{\rm eff}(\Phi_i,T)}\ .
\elabel{Zfour}\end{equation}
Equation \eqref{fcnlid} then follows from the observation that
$\partial{\cal W}_{\rm eff}/\partial F^\tau=\theta^2\,\partial{\cal W}_{\rm eff}/\partial\tau$.
We now need an explicit expression for the effective superpotential.
Following Affleck, Dine and Seiberg (ADS) \cite{Affleck:1983rr},
it is convenient to start
from $SU(N)$ gauge theory where the number of flavors $N_F$ is fixed
to $N_F=N-1.$ A 1-instanton calculation of the superpotential then
gives:
\defC_{\rm \scriptscriptstyle ADS}{C_{\rm \scriptscriptstyle ADS}}
\def{\tilde Q}{{\tilde Q}}
\begin{equation}{\cal W}_{\rm eff}^{N_F,N}\ \equiv\ {\cal W}_{\rm eff}^{N-1,N}\ =\
C_{\rm \scriptscriptstyle ADS}\,{\Lambda^{b_0}_{N-1,N}\over {\rm det}_{N_F}\big(Q_f{\tilde Q}_{f'}\big)}\ ,
\elabel{adsdef}\end{equation}
where the flavor indices $f,f'=1,\ldots,N_F$ run over the quark
superfields. The coefficient of the $\beta$-function is, for general
$N$ and $N_F,$
\begin{equation}b_0\ =\ 3N-N_F\ .
\elabel{bzerodef}\end{equation}
The normalization constant for the specific case $N_F=N-1$ was fixed
by an explicit 1-instanton calculation, and is simply
\cite{Cordes,Finnell:1995dr} $C_{\rm \scriptscriptstyle ADS}=1$. By
decoupling the quark flavors one at a time, this 1-instanton
expression flows into models with $N_F<N-1$ for which the
superpotential is no longer a 1-instanton phenomenon. In this way one
generalizes Eq.~\eqref{adsdef} to (see e.g.,
Refs.~\cite{Peskin:1997qi,Intriligator:1996au}):
\begin{equation}
{\cal W}_{\rm eff}^{N_F,N}\ =\
C_{\rm \scriptscriptstyle ADS}^{N_F,N}\,\left({\Lambda^{b_0}_{N_F,N}\over
{\rm det}_{N_F}\big(Q_f{\tilde Q}_{f'}\big)}\right)^{1\over N-N_F}\quad(N_F\le
N-1)\ ,
\elabel{adsdefb}\end{equation}
where
\cite{Finnell:1995dr}
\begin{equation}C_{\rm \scriptscriptstyle ADS}^{N_F,N}\ =\ N-N_F\ .
\elabel{cadsdefa}\end{equation}
Starting from this more general superpotential, let us decouple the
remaining quarks, by giving them a common VEV v. Viewing $Q$ as an
$N_F\times N$ matrix, one assumes:
\begin{equation}
\VEV{Q}\ =\
\begin{pmatrix}{\rm v}&{}&0&0&\cdots&0\\{}&\ddots&{}&\vdots&{}&\vdots\\
0&{}&{\rm v}&0&\cdots&0\end{pmatrix}\ ,\quad
\langle{\tilde Q}\rangle\ =\
\begin{pmatrix}\tilde{\rm v}&{}&0\\{}&\ddots&{}\\0&{}&\tilde{\rm v}\\0&\cdots&0\\\vdots&{}&\vdots\\
0&\cdots&0\end{pmatrix}\ .
\elabel{Qvev}\end{equation}
The $D$-flatness condition together with a global gauge rotation gives $\tilde{\rm v}=\bar{\rm v}.$
Taking $|{\rm v}|\rightarrow\infty$ then decouples the quarks as well as a
subset of the gauge fields, leaving a pure $SU(N')$ gauge theory with
$N'=N-N_F$ and $b_0=3N'$. The 1-loop RG matching prescription reads
\cite{Finnell:1995dr}:
\begin{equation}
\left({\Lambda_{N_F,N}\over|{\rm v}|}\right)^{3N-N_F}\ =\
\left({\Lambda_{0,N'}\over|{\rm v}|}\right)^{3N'}\ ,\quad N'=N-N_F\ .
\elabel{matching}\end{equation}
Inputting Eqs.~\eqref{cadsdefa}-\eqref{matching} into Eq.~\eqref{adsdefb} gives:
\begin{equation}{\cal W}_{\rm eff}\ =\
(N-N_F)\,\left({\Lambda^{3N-N_F}_{N_F,N}\over|{\rm v}|^{2N_F}}
\right)^{1\over N-N_F}\ =\ N'\big(\Lambda_{0,N'}\big)^3\ .
\elabel{adsdefc}\end{equation}
The desired result \eqref{WCIans} then follows from Eq.~\eqref{fcnlid}.
Note that the starting-point for this WCI calculation, Eq.~\eqref{adsdef}, is
a \it bona fide \rm \hbox{1-instanton} calculation. The remaining steps
towards the answer involve well-studied path-integral and renormalization group
manipulations (principally Eq.~\eqref{fcnlid}, and Eqs.~\eqref{adsdefb}-\eqref{matching},
respectively). Alternatively, starting again from the functional
identity \eqref{fcnlid}, we can rederive the WCI result \eqref{WCIans} without
any reference to an instanton calculation. Instead, one starts from the
Seiberg-Witten solution of the ${\cal N}=2$ model,\footnote{For the remainder of
the section, we focus on $SU(2)$ gauge theory, and quote well-known
formulae from Seiberg and Witten \cite{Seiberg:1994rs}.}
in the presence of a mass
deformation which breaks the supersymmetry down to ${\cal N}=1.$ In the
strong-coupling domain, in the vicinity of the monopole singularity,
the superpotential looks like:
\def{\cal W}_{\scriptscriptstyle\rm SW}{{\cal W}_{\scriptscriptstyle\rm SW}}
\begin{equation}
{\cal W}_{\scriptscriptstyle\rm SW}\ =\ \sqrt{2}\,\tilde MA_DM\,+\,m\,U(A_D)\ .
\elabel{WSWdef}\end{equation}
Here the chiral superfields $\{M,\tilde M\}$ describe the monopole
multiplet, $A_D$ is the dual Higgs, $U$ is the quantum modulus of the
theory (here, in strong coupling, expressed in terms of $A_D$ rather
than $A$), and $m$ is the mass parameter. The $F$-flatness condition
for the vacuum reads
\begin{equation}
0\ =\ {\partial{\cal W}_{\scriptscriptstyle\rm SW}\over\partial M}\ =\
{\partial{\cal W}_{\scriptscriptstyle\rm SW}\over\partial\tilde M}\ =\ {\partial{\cal W}_{\scriptscriptstyle\rm SW}\over\partial
A_D}\ ,
\elabel{Fflat}\end{equation}
which is solved by
\begin{equation}
a_D\equiv \VEV{A_D}=0\ ,\qquad \VEV{M}=\langle\tilde M\rangle=
\Big(-{m\over\sqrt{2}\,}U'(0)\Big)^{1/2}\ .
\elabel{Fflatsolve}\end{equation}
In the vicinity of this solution, the relationship between $a_D$ and
$u=\Vev{U}$ is given by
\def\Lambda_{\scriptscriptstyle\rm SW}{\Lambda_{\scriptscriptstyle\rm SW}}
\def\Lambda_{\scriptscriptstyle\N=2}{\Lambda_{\scriptscriptstyle{\cal N}=2}}
\begin{equation}
a_D\ =\ {\sqrt{2}\,\over\pi}\int_1^udx\,{\sqrt{x-u}\over\sqrt{x^2-\Lambda_{\scriptscriptstyle\rm SW}^4}}
\elabel{aDdef}\end{equation}
from which it follows that
\begin{equation}
u\ =\ \Lambda_{\scriptscriptstyle\rm SW}^2-2i\Lambda_{\scriptscriptstyle\rm SW} a_D+{\cal O}(a_D^2)\ .
\elabel{useries}\end{equation}
Here the Seiberg-Witten dynamical scale $\Lambda_{\scriptscriptstyle\rm SW}$ is related to the
conventional PV/$\overline{\rm DR}$ scale $\Lambda_{\scriptscriptstyle\N=2}$ via \cite{Finnell:1995dr}
\begin{equation}
\Lambda_{\scriptscriptstyle\rm SW}\ = \sqrt{2}\,\Lambda_{\scriptscriptstyle\N=2}\ .
\elabel{LamSWdef}\end{equation}
Note that the series \eqref{useries} is not an instanton expansion (i.e., an
expansion in $\Lambda_{\scriptscriptstyle\rm SW}^4$); instantons emerge only in
the weak-coupling regime, where $u$ is expanded in terms of $a=\Vev{A}$ rather
than $a_D$.
Applying the identity \eqref{fcnlid} to ${\cal W}_{\scriptscriptstyle\rm SW}$ using
Eqs.~\eqref{useries}-\eqref{LamSWdef} gives the gluino condensate in the vacuum
\eqref{Fflatsolve}:
\begin{equation}
\VEV{\mathop{\rm tr}\nolimits \lambda^2}\ =\ 16\pi^2m\Lambda_{\scriptscriptstyle\N=2}^2\ .
\elabel{SWans}\end{equation}
Next we decouple the adjoint Higgs superfield, by taking
$m\rightarrow\infty.$ In this way we flow to the pure ${\cal N}=1$
supersymmetric $SU(2)$ gauge theory.
The RG matching condition between the scale $\Lambda$ of the ${\cal N}=1$ theory and the
scale $\Lambda_{\scriptscriptstyle\N=2}$ of the mass-deformed ${\cal N}=2$ theory reads \cite{Finnell:1995dr}:
\begin{equation}
m^2\Lambda_{\scriptscriptstyle\N=2}^4\ =\ \Lambda^6\ .
\elabel{SWmatch}\end{equation}
Substituting Eq.~\eqref{SWmatch} into Eq.~\eqref{SWans} once again gives the WCI
answer \eqref{WCIans}.
\setcounter{equation}{0}
\section{Comments on Cluster Decomposition}
In this section, we examine the issue of cluster decomposition in the
context of the gluino condensate. This issue of cluster decomposition
is fundamental to a quantum field theory. The clustering property
requires that for sufficiently large separations $|x_i-x_j|$, compared
with the inverse mass gap,\footnote{For a discussion of clustering and
other references, see Bogolubov {\it et al.\/} \cite{BOG}.}
\begin{equation}
\VEV{\varphi_1(x_1)\cdots\varphi_n(x_n)}\ \rightarrow\
\VEV{\varphi_1}\times\cdots\times\VEV{\varphi_n}\ .
\end{equation}
Generally, this property breaks down when, in a statistical mechanical
sense, the theory is in a mixed phase. In field theory language, this
means there is more than one possible vacuum state. The clustering property
is then restored by restricting the theory to the Hilbert space built
on one of the vacua. In this sense, clustering is violated in a mild
way, and to distinguish this from some other, potentially more serious,
violations uncovered below, we will
say that the theory satisfies a ``generalized notion of clustering''.
Let us consider the calculation of the ${\cal G}_n$, the $n$-point function
of the composite operator $\tr_{N\,}^{}\lambda^2\,$:
\begin{equation}{\cal G}_n(x_1,\ldots,x_n)\ =\ \langle\tr_{N\,}^{}\lambda^2(x_1)\cdots
\tr_{N\,}^{}\lambda^2(x_n)\rangle\ .
\elabel{Gndef}\end{equation}
For present purposes we restrict our attention to pure ${\cal N}=1$ $SU(N)$
gauge theory. Since $\tr_{N\,}^{}\lambda^2$ is the lowest component of a
gauge-invariant chiral superfield (namely $\tr_{N\,}^{} W^2$ where $W^\alpha$
is the field-strength superfield), a well-known identity---reviewed in
the Appendix---says that
\begin{equation}
{\cal G}_n(x_1,\ldots,x_n)\ = \hbox{const.\ ,}
\elabel{Gnconst}\end{equation}
independent of the $x_i$. Next let us consider this constant
correlator in the instanton
approximation. This means
that, at topological level $k$, $\lambda(x)$ is simply to be replaced
by a general superposition of adjoint fermion zero modes in the
general ADHM $k$-instanton background, weighted by Grassmann-valued
parameters (i.e., fermionic collective coordinates). All bosonic and
fermionic collective coordinates are then integrated over, in the
appropriate supersymmetric way reviewed below. It can also be shown
that ${\cal G}_n$ should still be a constant. (The field theory proof of the
constancy of the correlation functions and its extension to the
instanton approximation is discussed in the Appendix.) Now, in $SU(N)$
gauge theory, at the topological level $k$, a multi-instanton has
precisely $2kN$ adjoint fermion zero modes which need to be integrated
over. Let us summarize the rules
for Grassmann integration: if $\xi$ is a Grassmann parameter, then
\begin{equation}
\int d\xi\,\xi=1\ ,\quad\int d\xi\,1=0\ .
\elabel{xiint}\end{equation}
Since $\tr_{N\,}^{}\lambda^2$ is a Grassmann bilinear, it follows that ${\cal G}_n$
is only non-vanishing for $n=kN$. In particular, the one-point function
${\cal G}_1$ always vanishes. In summary, in the instanton
approximation, at topological level $k$, we have the following
selection rule:
\begin{subequations}
\begin{align}
&\langle\tr_{N\,}^{}\lambda^2(x)\rangle\Big|_{k\hbox{-}\rm inst}\
\equiv \ {\cal G}_1\,\Big|_{k\hbox{-}\rm inst}\
=\ 0\quad\hbox{for all}\ k\ ;\elabel{selrulesa}\\
&\langle\tr_{N\,}^{}\lambda^2(x_1)\cdots
\tr_{N\,}^{}\lambda^2(x_n)\rangle\Big|_{k\hbox{-}\rm inst}\ \equiv\ {\cal G}_n\Big|_{k\hbox{-}\rm inst}\ \neq\
0 \quad\hbox{if and only if}\ n=kN\ .\elabel{selrulesb}
\end{align}
\end{subequations}
Notice that these results already indicate a breakdown of clustering
for the correlation functions \eqref{Gndef}, although, as we shall
explain below the breakdown is of the
`mild' variety and can be traced to the fact that in
instanton approximation the theory is in a mixed phase, i.e.~the
instanton approximation samples the theory in a number of distinct
vacua as opposed to a single vacuum.
A general field-theoretic understanding of the selection rule
\eqref{selrulesa}-\eqref{selrulesb} was
suggested in Refs.~\cite{Amati:1985uz,Amati:1988ft}. The suggestion
relies on the fact
that, in ${\cal N}=1$ $SU(N)$ gauge theory, the vacua of the theory come in
an $N$-tuplet \cite{Witten:1982df}. The vacua spontaneously break the
discrete ${\Bbb Z}_{2N}$ anomaly-free remnant of the classical
$U(1)_R$ symmetry to the
${\Bbb Z}_2$ subgroup: $\lambda_\alpha\rightarrow-\lambda_\alpha$.
The vacuum sector therefore consists of:
\begin{equation}\big\{\ \ket{u}\ :\quad0\le u\le N-1\ \big\}\ .
\elabel{ketjdef}\end{equation}
If we define the condensate ${\cal J}$ via
\begin{equation}
{\cal J}\ =\ \bra{u=0}\tr_{N\,}^{}\lambda^2\ket{u=0}\ ,
\elabel{Jdef}\end{equation}
then the $N$-tuple of vacua are related by phase factors, namely the
$N^{\rm th}$ roots of unity:
\begin{equation}
\bra{u}\tr_{N\,}^{}\lambda^2\ket{u}\ =\ {\cal J}\,e^{2\pi iu/N}\ .
\elabel{jvevs}\end{equation}
Now let us see how the selection rule \eqref{selrulesa} comes about. We
define the density matrix
\begin{equation}
\varrho\ =\ {1\over N}\sum_{u=0}^{N-1}\ket{u}\bra{u}\ .
\elabel{densmat}\end{equation}
Since the instanton calculation is ${\Bbb Z}_{2N}$ symmetric, it must
{\it average\/} over all the vacua. This means
\begin{equation}\vev{\tr_{N\,}^{}\lambda^2}=\mathop{\rm Tr}\nolimits\big(\varrho\,\tr_{N\,}^{}\lambda^2\big)
={1\over N}\sum_{u=0}^{N-1}\bra{u}\tr_{N\,}^{}\lambda^2\ket{u}={{\cal J}\over N}
\sum_{u=0}^{N-1} e^{2\pi iu/N}=0\ .
\elabel{firstsel}\end{equation}
Here the capitalized `$\mathop{\rm Tr}\nolimits$' means a trace over the Hilbert space. In
order to check the selection rule \eqref{selrulesb}, we need the additional
assumption of a well-defined clustering limit.
We have:
\begin{equation}\begin{split}
&\langle\tr_{N\,}^{}\lambda^2(x_1)\cdots
\tr_{N\,}^{}\lambda^2(x_n)
\rangle\ =\ \mathop{\rm Tr}\nolimits\big(\varrho\,
\tr_{N\,}^{}\lambda^2(x_1)\cdots
\tr_{N\,}^{}\lambda^2(x_n)\big)\\ &\qquad\qquad\qquad=\
{1\over N}\sum_{u=0}^{N-1}\bra{u}\tr_{N\,}^{}\lambda^2(x_1)\,{\Bbb P}\,\tr_{N\,}^{}\lambda^2(x_2)\,{\Bbb P}\cdots{\Bbb P}\,
\tr_{N\,}^{}\lambda^2(x_n)\ket{u}\ ,
\elabel{npoint}\end{split}\end{equation}
where ${\Bbb P}$ denotes the sum over a complete set of states. At this
point the generalized notion of the clustering assumption enters.
We assume there exists a mass
gap $\mu$ that is dynamically generated in the theory, and we consider
the $n$ insertion points are sufficiently far separated in Euclidean
space compared to this scale: $|x_i-x_j|\gg\mu^{-1}.$ (Since ${\cal G}_n$ is
a constant even in leading semiclassical order, moving to this regime
does not entail any additional approximations.)
In this regime, the generalized cluster decomposition
(in our present usage) is equivalent
to the statement that ${\Bbb P}$ collapses to ${\Bbb P}_0$ where ${\Bbb P}_0$ is the
projection operator onto vacuum states only:
\begin{equation}
{\Bbb P}\rightarrow{\Bbb P}_0\ ,\qquad
{\Bbb P}_0\ =\ \sum_{u=0}^{N-1}\ket{u}\bra{u}\ .
\elabel{Pzerodef}\end{equation}
Using the fact that the operator $\tr_{N\,}^{}\lambda^2$ is diagonal in the
$u$ index,\footnote{This follows from the fact that there should be no mixing
between the sectors of Hilbert space built on each vacuum: they are
{\it super-selection sectors\/}.} it follows that with the replacement \eqref{Pzerodef},
the correlator \eqref{npoint} collapses to
\begin{equation}
{1\over N}\sum_{u=0}^{N-1}\big({\cal J}\,e^{2\pi i u/N}\big)^n\ =\
\begin{cases}
{\cal J}^{kN}\ & n=kN\\ 0\ & \hbox{otherwise}\ .\end{cases}
\elabel{collapse}\end{equation}
Next we consider how this elementary analysis is modified if the
$N$-tuplet of vacua $\{\ket{u}\}$ is supplemented by an extra vacuum
state, the so-called Kovner-Shifman vacuum \cite{Kovner:1997im}, which we denote
$\ket{S}$. A single such vacuum is permissible under the discrete symmetry only if
\begin{equation}
\bra{S}\tr_{N\,}^{}\lambda^2\ket{S}\ =\ 0\ .
\elabel{KScond}\end{equation}
The analysis proceeds just as before, with the obvious modification
that the density matrix $\varrho$ should be replaced by $\varrho',$
defined by
\begin{equation}
\varrho'\ =\ (1-p)\ket{S}\bra{S}\ +\ {p\over N}\sum_{u=0}^{N-1}\ket{u}\bra{u}\ ,
\elabel{varrhopdef}\end{equation}
where the probability $p$ is a real number between 0 and 1.
Proceeding as before, we find that with the generalized clustering
assumption
\begin{equation}
{\Bbb P}\rightarrow{\Bbb P}'_0\ ,\qquad
{\Bbb P}'_0\ =\ \ket{S}\bra{S}\ +\ \sum_{u=0}^{N-1}\ket{u}\bra{u}\ ,
\elabel{Pzerodefa}\end{equation}
one derives
\begin{equation}
\langle\tr_{N\,}^{}\lambda^2(x_1)\cdots\tr_{N\,}^{}\lambda^2(x_n)\rangle\ =\
\begin{cases}p{\cal J}^{kN}\ & n=kN\\ 0\ & \hbox{otherwise}\ .\end{cases}
\elabel{collapseb}\end{equation}
Obviously this modified expression also applies if there are several
distinct KS vacua $\ket{S_i}$ in which $\mathop{\rm tr}\nolimits\lambda^2=0,$ that is
\begin{equation}
\varrho'\ =\ \sum_{i=1}^l q_i\ket{S_i}\bra{S_i}\ +\ {p\over N}\sum_{u=0}^{N-1}\ket{u}\bra{u}\ ,
\elabel{varrhopdefz}\end{equation}
where $p=1-\sum q_i$.
In the following we will calculate these $(kN)$-point correlators,
first analytically for large $N$, then numerically for $N=2$ and $k=2$, and will
find a behavior quite different from either \eqref{collapse} or \eqref{collapseb}.
\setcounter{equation}{0}
\section{Large-$N$ Calculation of Gluino Condensate Correlation Functions}
We now present an explicit evaluation of ${\cal G}_n,$ $n=kN,$ in the limit
$N\rightarrow\infty$ with $k$ held fixed. Our answer turns out to be
incompatible with both Eqs.~\eqref{collapse} and \eqref{collapseb}. The cleanest
way to quantify this disagreement is to consider the $(kN)^{\rm th}$ root,
$({\cal G}_{kN})^{1/kN}.$ In the large-$N$ limit, from
Eq.~\eqref{collapse}, i.e.~clustering without the KS vacuum,
one obtains:
\begin{equation}
\lim_{N\rightarrow\infty}({\cal G}_{kN})^{1/kN}\ =\ {\cal J}(N)\ .
\elabel{rootdef}\end{equation}
We have written ${\cal J}$ as ${\cal J}(N)$ to allow for an unknown $N$
dependence. Using the one instanton expression \eqref{SCIansb} in the
large-$N$ limit one expects
\begin{equation}
{1\over16\pi^2}{\cal J}(N)={2e\over N}\,\Lambda^3\ ,
\elabel{jnclust}\end{equation}
where $e=2.718\cdots.$
The key point is that the right-hand side of \eqref{rootdef} is
independent of the topological number $k$ (as well as of the
space-time insertion points $x_i$). Note that Eq.~\eqref{rootdef} follows,
not only from Eq.~\eqref{collapse}, but also
from Eq.~\eqref{collapseb}, so long as the constant $p$ either has a nonzero
large-$N$ limit, or else vanishes at large $N$
more slowly than exponentially. Alternatively, with the ``Shifman
assumption'' \eqref{Shifmana} for $p$, which vanishes faster than
exponentially at large $N$,
one obtains instead from Eq.~\eqref{collapseb}:
\begin{equation}\lim_{N\rightarrow\infty}({\cal G}_{kN})^{1/kN}\ =\
\Big({2e\over N}\Big)^{1/k}\,{\cal J}(N)\ .
\elabel{rootdefz}\end{equation}
Combining this with the large-$N$ limit of
the 1-instanton expression \eqref{SCIansb},
one extracts instead the expression
\begin{equation}
{1\over16\pi^2}{\cal J}(N)=\ \Lambda^3\
\elabel{jnks}\end{equation}
which now agrees (by construction) with the 1-instanton WCI calculation.
Below we will calculate ${\cal G}_{kN},$ to leading order in $1/N$, but for
all instanton number $k$, and will obtain a markedly different
behavior. Explicitly we will find:
\begin{equation}
{1\over16\pi^2}\lim_{N\rightarrow\infty}({\cal G}_{kN})^{1/kN}\ =\ {2e\over
N}k\Lambda^3+{\cal O}(N^{-2})\ . \elabel{rootdefb}\end{equation} Notice
that for $k=1$ we obviously recover the results
\eqref{rootdef}-\eqref{jnclust} (or \eqref{rootdefz}-\eqref{jnks});
however the linear $k$ dependence is in sharp disagreement with the $k$
dependence of either Eq.~\eqref{rootdef} or Eq.~\eqref{rootdefz}.
This disagreement means that the generalized clustering assumption
\eqref{Pzerodef} is invalid when combined with the instanton
approximation. It also means that that the extension \eqref{Pzerodefa}
of this clustering assumption, in the presence of an extra KS vacuum
state, is likewise invalid.
The large-$N$ calculation proceeds as follows.\footnote{Our
conventions are taken from \cite{MO-III,KMS} which also provide
self-contained reviews of the ADHM formalism for the $SU(N)$ gauge
group.} In supersymmetric
theories, at topological level $k$, the bosonic and fermionic
collective coordinates live, respectively, in complex-valued matrices
$a$ and ${\cal M},$ with elements:
\begin{equation}
a=\begin{pmatrix}w_{uj{\dot\alpha}}\\(a'_{\beta{\dot\alpha}})_{ij}^{}\end{pmatrix}\
,\qquad
{\cal M}=\begin{pmatrix}\mu_{uj}\\({\cal M}'_\beta)_{ij}^{}\end{pmatrix}\ .
\elabel{adef}\end{equation}
The indices run over
\begin{equation}
u=1,\ldots,N\ ,\quad i,j=1,\ldots,k\ ,\quad
{\dot\alpha},\beta=1,2\ ;
\elabel{indexrun}\end{equation}
traces over these indices are denoted `$\mathop{\rm tr}\nolimits_N$', `$\mathop{\rm tr}\nolimits_k$', and
`$\mathop{\rm tr}\nolimits_2$', respectively.
The elements of ${\cal M}$ are Grassmann (i.e., anticommuting)
quantities. The $k\times k$ submatrices $a'_{\beta{\dot\alpha}}\equiv
a'_n\sigma^n_{\beta{\dot\alpha}}$ and ${\cal M}'_\beta$ are subject to the
Hermiticity conditions
\begin{equation}\bar a_n'=a_n'\ ,\quad\bar\M'_\alpha={\cal M}'_\alpha\ .
\elabel{hermcond}\end{equation}
In the instanton approximation, the Feynman path integral is replaced
by a finite-dimensional integration over the degrees of freedom in $a$
and ${\cal M}.$ These $k$-instanton collective coordinates are weighted
according to the integration measure
\cite{meas1,meas2,MO-III,KMS}\footnote{The
reason we have
$2^{k^2/2}$ rather than $2^{-k^2/2}$, as in \cite{MO-III}, is that we
restore Wess and Bagger integration conventions for the ${\cal M}'$
integration: $\int d^2\xi\,\xi^2=1$ rather than 2 where
$\xi^2=\xi^\alpha\xi_\alpha$ is the square of a Grassmann Weyl spinor.}
\defd\mu^{k}_{\rm phys}{d\mu^{k}_{\rm phys}}
\begin{equation}\begin{split}
\intd\mu^{k}_{\rm phys}\ &=\ {2^{k^2/2}(C_1)^k\over{\rm
Vol\,}U(k)}\int d^{4k^2}a'\,d^{2kN}\bar w\,d^{2kN}w\,d^{2k^2}{\cal M}'\,d^{kN}\bar\mu\,d^{kN}\mu
\\&\times\
\prod_{r=1,\ldots,k^2}\Big[\prod_{c=1,2,3}\delta\big({\textstyle{1\over2}}{\rm tr}_k\,T^r(\tr^{}_2\,\, \tau^c \bar a a)\big)
\prod_{{\dot\alpha}=1,2}\delta\left({\rm tr}_k\,T^r(\bar\M a_{\dot\alpha} + \bar a_{\dot\alpha}
{\cal M})\right)\Big]\ ,
\elabel{dmuphysdef}\end{split}\end{equation}
where the two $\delta$-functions enforce the bosonic and fermionic
ADHM constraint conditions, respectively. The integrals over the
$k\times k$ matrices $a'_n$ and ${\cal M}^{\prime}$ are defined as the
integral over the components with respect to a Hermitian basis of
$k\times k$ matrices $T^r$ normalized so that ${\rm
tr}_k\,T^rT^s=\delta^{rs}$. These matrices also provide explicit
definitions of the $\delta$-function factors in the way indicated.
The form of the measure given in Eq.~\eqref{dmuphysdef} is known as
the ``flat measure'', since the bosonic and fermionic ADHM collective
coordinates are integrated over as Cartesian variables, subject to the
nonlinear $\delta$-function constraints. It was uniquely constructed
in Ref.~\cite{meas1} to obey several important consistency
requirements---including cluster decomposition---so that the failure
of cluster uncovered below cannot be attributed to the collective
coordinate measure. In practical applications, however, the flat
measure is not the most useful form available. When
\begin{equation}
N\ \ge\ 2k\ , \elabel{Nkinequality}\end{equation} it is convenient to
switch to the so-called ``gauge-invariant measure,'' involving a new
set of variables in terms of which the arguments of the
$\delta$-functions are linear (and hence trivially implemented)
\cite{MO-III}. This is the form of the measure which we will utilize
in the present section. The restriction \eqref{Nkinequality} is
obviously well suited to the large-$N$ limit. As the name implies, the
gauge-invariant measure can only be used to integrate gauge-invariant
quantities, such as our present focus on correlators formed from
$\tr_{N\,}^{}\lambda^2.$ Alternatively, for the special cases $k\le2,$ it is
easy to solve the nonlinear constraints explicitly without such a
change of variables \cite{Osborn:1981yf,MO-I}.
In order to switch from the flat measure to the gauge-invariant
measure, one trades the collective coordinates $w$ and $\bar w$ (which
transform in the $N$ of $SU(N)$) for the gauge-invariant
bosonic bilinear quantity $W$, defined by \cite{MO-III}
\begin{equation}
\big(W_{\ {\dot\beta}}^{\dot\alpha}\big)_{ij}=\bar w_{iu}^{\dot\alpha}
\,w_{uj{\dot\beta}}\ ,\quad
W^0={\rm tr}_2\,W,\quad W^c={\rm
tr}_2\,\tau^cW, \ \ c=1,2,3\ .
\elabel{Wdef}\end{equation}
The appropriate Jacobian for this change of variables reads:
\begin{equation}
d^{2kN}\bar w\,d^{2kN}w\ \longrightarrow\
c_{k,N}\,\big({\rm det}_{2k}W\big)^{N-2k}d^{k^2}
W^0\prod_{c=1,2,3}d^{k^2}W^c\ ,
\elabel{Jacis}\end{equation}
where
\begin{equation}
c_{k,N}\ =\
{2^{2kN-4k^2+k}\,\pi^{2kN-2k^2+k}\over\prod_{i=1}^{2k}(N-i)!}\ .
\elabel{ckNdef}\end{equation}
Note that the bosonic $\delta$-function in \eqref{dmuphysdef} can be
rewritten in a gauge-invariant way as the condition
\begin{equation}
0=W^c+[\,a'_n\,,\,a'_m\,]\,\tr^{}_2\,\,\tau^c\bar\sigma^{nm}=
W^c - i [\,a'_n\,,\,a'_m\,]\,\bar\eta^c_{nm}
\elabel{adhmredux}\end{equation}
in terms of the gauge-invariant coordinates (here $\bar\eta^c_{nm}$ is
an `t Hooft tensor \cite{tHooft}). As advertised, these constraints are {\it
linear\/} in the new variables $W^c$; consequently the $W^c$ integrals simply remove
the bosonic ADHM $\delta$-functions in Eq.~\eqref{dmuphysdef} (giving rise to the
numerical factor of $2^{3k^2}$ from the ${\textstyle{1\over2}}$'s in the arguments of
the $\delta$-functions).
Next we perform a similar change of variables for the fermions,
letting \cite{MO-III}
\begin{equation}
\mu_{ui}=w_{uj{\dot\alpha}}(\zeta^{{\dot\alpha}
})_{ji}+\nu_{ui},\qquad
\bar\mu_{iu}=(\bar\zeta^{
}_{\dot\alpha})_{ij}\bar w_{ju}^{\dot\alpha}+\bar\nu_{iu}\ ,
\elabel{zetadef}\end{equation}
where $\nu$ lies in the orthogonal subspace to $w$:
\begin{equation}
\bar w_{iu}^{\dot\alpha}\nu_{uj}=0\ ,\qquad \bar\nu_{iu} w^{}_{uj{\dot\alpha}}=0\ .
\elabel{nudef}\end{equation}
One finds:
\begin{equation}
\int d^{kN}\mu\,d^{kN}\bar\mu
\prod_{r=1,\ldots,k^2}\prod_{{\dot\alpha}=1,2}
\delta\left({\rm tr}_k\,T^r(\bar\M a_{\dot\alpha} + \bar a_{\dot\alpha} {\cal M})\right)
\ \longrightarrow
\ 2^{k^2}\int d^{2k^2}\zeta\,d^{kN-2k^2}\nu\,d^{kN-2k^2}\bar \nu\ ,
\elabel{fermiXn}\end{equation}
where the $\delta$-functions have been used to eliminate the
$\bar\zeta$ variables from the problem.
In summary, the gauge-invariant measure is:
\begin{equation}{2^{9k^2/2}\,(C_1)^k\,c_{k,N}\over{\rm Vol\,}U(k)}\int
d^{4k^2}a'\,d^{k^2}W^0\,d^{2k^2}{\cal M}'\,d^{2k^2}\zeta\,
d^{kN-2k^2}\nu\,d^{kN-2k^2}\bar \nu\,({\rm det}_{2k}W)^{N-2k}
\elabel{GImeas}\end{equation}
and the constraint $\delta$-functions have been eliminated for all $k$
satisfying Eq.~\eqref{Nkinequality}.
For $k=1,$ one recovers the expression
\eqref{measone}, a comparison which fixes the
normalization constant $C_1$:
\begin{equation}
C_1\ =\ -2^{N+1/2}\Lambda^{3N}\ .
\elabel{Conedef}\end{equation}
For $k=2$, we recapture the Osborn measure
discussed in Refs.~\cite{Osborn:1981yf,MO-I,meas1}, which we utilize in
Sec.~VI below.
Into this measure we now insert
\begin{equation}
\tr_{N\,}^{}\lambda^2(x_1)\times\cdots\times\tr_{N\,}^{}\lambda^2(x_{kN})\ ,
\elabel{insertinto}\end{equation}
where the gluino $\lambda^\alpha(x)$ is replaced in the instanton
approximation by a general superposition of adjoint fermion zero
modes. In terms of the previously introduced collective coordinates
$a$ and ${\cal M},$ a useful identity states \cite{MO-III}:
\begin{equation}
\tr_{N\,}^{}\lambda^2(x)\ =\ -{\textstyle{1\over4}}
\square\,\tr_{k\,}^{}\bar\M({\cal P}+1){\cal M} f\ ,
\elabel{corriganid}\end{equation}
where the ADHM quantities ${\cal P}$ and $f$ are defined as:
\begin{equation}
{\cal P}=1-\Delta f\bar\Delta\ ,\quad f=(\bar\Delta\Delta)^{-1}\
,\quad\Delta=a+bx\ ,
\elabel{adhmdefs}\end{equation}
and $b$ is the $(N+2k)\times(2k)$ matrix whose lower $2k\times 2k$
part is the identity $\delta_\beta^{\ \alpha}\delta_{il}$ and whose
upper $N\times 2k$ part is zero (quaternionic multiplication is
implied in the product $bx$). As discussed earlier,
${\cal G}_{kN}(x_1,\ldots,x_{kN})$ is actually a constant, independent of
the $x_i$. The $x_i$ can therefore be chosen for maximum simplicity of
the algebra. However, the simplest conceivable choice, $x_i=0$ for all $i$,
results in an ill-defined answer of the form ``$0\times\infty$'' (the
zero coming from unsaturated Grassmann integrations, and the infinity
from divergences in the bosonic integrations due to coincident poles);
we have already noted this fact in the 1-instanton sector in Sec.~II above.
The simplest choice of
the $x_i$ that avoids this problem turns out to be:
\begin{equation}\begin{split}
&x_1=\cdots=x_{kN-k^2}=0\
,\\&x_{kN-k^2+1}=\cdots=x_{kN}=x
\elabel{simple}\end{split}\end{equation}
which we adopt for the remainder of this section.\footnote{As a
nontrivial check on our algebra, we have also numerically integrated
the large-$N$ correlator for insertions other than Eq.~\eqref{simple}, and
verified the constancy of the answer presented below.}
In the large-$N$ limit the large preponderance of the insertions
\eqref{simple} are at $x_i=0,$ and the resulting factor of
$(\tr_{N\,}^{}\lambda^2(0))^{kN-k^2},$ taken together with the Jacobian factor
$({\rm det}_{2k}W)^{N-2k}$ from the measure \eqref{GImeas}, dominate the integral
and can be treated in
saddle-point approximation. Below we will carry out this saddle-point
evaluation in full detail, but we can already quite easily understand
the source of the linear dependence on $k$ in the final result
\eqref{rootdefb}. The chain of argument goes as follows:
\bf1\rm. Let us imagine carrying
out all the Grassmann integrations in the problem. The remaining
large-$N$ integrand will then have the form $\exp\big(-N\Gamma+{\cal O}(\log
N)\big)$ where $\Gamma$ might be termed the ``effective large-$N$ bosonic
instanton action.'' The large-$N$ saddle-points are then the
stationary points of $\Gamma$ with respect to the bosonic collective
coordinates. By Lorentz symmetry, $\Gamma$ can only depend on
the four $k\times k$ matrices $a'_n$ through
even powers of $a'_n$. (This is because the bulk of the insertions
have been chosen to be
at $x_i=0$; otherwise one could form the Lorentz scalar $x^{}_na'_n$ and so
have odd powers of $a'_n$.) It follows that the ansatz
\begin{subequations}
\begin{align}
&a'_n=0\ ,\quad n=1,2,3,4\ ,\elabel{spansatza}\\&W^c=0\ ,\quad
c=1,2,3 \elabel{spansatzb}
\end{align}\end{subequations}
is automatically a stationary point of $\Gamma$ with respect to these
collective coordinates. (Note that \eqref{spansatzb} follows automatically
from \eqref{spansatza} by virtue of the ADHM constraints
\eqref{adhmredux}.) It will actually turn out that, once one assumes these saddle-point
values, $\Gamma$ is independent of the remaining collective
coordinate matrix $W^0$; furthermore we will verify that this
saddle-point is actually a minimum of the Euclidean action.
\bf2\rm. Having anticipated the saddle-point
\eqref{spansatza}-\eqref{spansatzb} using these elementary
symmetry considerations, let us back up to a stage in the analysis
prior to the Grassmann integration, and proceed a little more carefully.
Evaluating the insertions $\tr_{N\,}^{}\lambda^2(x_i)$ on this
saddle-point, one easily verifies that the $\zeta$ modes vanish when $x_i=0$;
consequently the $\zeta$ integrations must be saturated entirely from
the $k^2$ insertions at $x_i=x.$ This leaves the ${\cal M}',$ $\nu$ and
$\bar \nu$ integrations to be saturated purely from the insertions at
$x_i=0$. Moreover, because ${\cal M}'$ carries a Weyl spinor index $\alpha$ whereas
$\nu$ and $\bar \nu$ do not, the $\tr_{N\,}^{}\lambda^2(0)$ insertions
depend on these Grassmann coordinates only through bilinears of the form
$\bar \nu\times \nu$ or ${\cal M}'\times{\cal M}'$; there are no cross terms.
\bf3\rm. Performing all the Grassmann integrations then automatically generates
a combinatoric factor
\begin{equation}
(k^2)!\,(k^2)!\,(kN-2k^2)!\,\begin{pmatrix}kN-k^2\\ k^2\end{pmatrix}\ .
\elabel{combfact}\end{equation}
Here the first three factors account for the indistinguishable
bilinear insertions of the $\zeta,$ ${\cal M}',$ and $\{\nu,\bar \nu\}$ modes,
respectively, while the final factor counts the ways of
selecting the
$k^2$ bilinears in
${\cal M}'$ from the $kN-k^2$ insertions at $x_i=0.$
Multiplying these combinatoric factors together, as well as the
normalization constants $c_{k,N}(C_1)^k$ from Eq.~\eqref{GImeas}, and taking the
$(kN)^{\rm th}$ root yields, in the large-$N$ limit:
\begin{equation}
\lim_{N\rightarrow\infty}\big[
c_{k,N}(C_1)^k\,(k^2)!\,(kN-k^2)!\big]^{1/kN}\ =\
2^{3}\pi^{2}eN^{-1}k\Lambda^3+{\cal O}(N^{-2})\ .
\elabel{kNroot}\end{equation}
Remarkably, apart from a
factor of four, this back-of-the-envelope analysis precisely accounts
for the previously announced final answer, Eq.~\eqref{rootdefb}. Note that
most of the remaining contributions to the saddle-point analysis,
which involve a specific convergent bosonic integral derived below, as
well as the factor $2^{9k^2/2}/{\rm Vol\,}U(k)$ from Eq.~\eqref{GImeas}, reduce
to unity when the $(kN)^{\rm th}$ root is taken in the large-$N$
limit; the missing factor of four will simply come from the leading
saddle-point evaluation of the bosonic integrand.
Here are the details of the large-$N$ calculation of ${\cal G}_{kN}.$
Since the problem has an obvious $U(k)$ symmetry \cite{MO-III}, we will find it
convenient to work in a basis where $W^0$ (which transforms in the
adjoint of the $U(k)$) is diagonal:
\begin{equation}W^0\ =\
\begin{pmatrix}2\rho_1^2&{}&{0}\\{}&\ddots&{}\\{0}&{}&2\rho_k^2\end{pmatrix}\ .
\elabel{Wdiag}\end{equation}
As the notation implies, in the dilute instanton gas limit $\rho_i$
can be identified with the scale size of the $i^{\rm th}$ instanton in
the $k$-instanton sector (see Sec.~II.4 of \cite{MO-III}). The appropriate
change of variables reads:
\begin{equation}
{1\over{\rm Vol}\,U(k)}\int d^{k^2}W^0\ \longrightarrow\ {2^{3k(k-1)/2}\pi^{-k}\over k!}
\int_0^\infty
d\rho_1^2\cdots d\rho_k^2\,\prod_{1\le i<j\le k}(\rho_i^2-\rho_j^2)^2\ .
\elabel{diagcov}\end{equation}
For $k=1$ one has, of course, $\int
dW^0\rightarrow 2\int_0^\infty d\rho^2.$
Now let us consider the Grassmann integrations, beginning with the
$\zeta$ modes. We assume the saddle-point conditions \eqref{spansatza}-\eqref{spansatzb}, in
which case
\begin{equation}
\Delta=\begin{pmatrix}w\\ x\cdot1_{\scriptscriptstyle [k]\times[k]}^{}\end{pmatrix}\ ,\qquad
f=\begin{pmatrix}{1\over\rho_1^2+x^2}&{}&{0}\\ {}&\ddots&{}\\{0}&{}&
{1\over\rho_k^2+x^2}\end{pmatrix}\ ,
\elabel{spmore}\end{equation}
and from Eq.~\eqref{corriganid},
\begin{equation}
\tr_{N\,}^{}\lambda^2(x)\ =\ -\sum_{i,j=1}^k(\zeta_{\dot\alpha})_{ij}
(\zeta^{\dot\alpha})_{ji}\,F_{ij}(x)\ +\ \cdots\ ,
\elabel{zetastuff}\end{equation}
where
\begin{equation}
F_{ij}(x)\ =\
{\textstyle{1\over4}}\square\,{x^4\over(x^2+\rho_i^2)(x^2+\rho_j^2)}
\elabel{Fijdef}\end{equation}
and the omitted terms in Eq.~\eqref{zetastuff} represent dependence on the
other Grassmann modes $\{{\cal M}',\nu,\bar \nu\}$. It is obvious from
Eq.~\eqref{Fijdef} that $F_{ij}(0)=0$, so that the $\zeta$ modes must be
entirely saturated from the $k^2$ insertions at $x_i=x$ as claimed
above. Performing the $\zeta$ integrations then yields
\begin{equation}(-1)^{k^2}(k^2)!\,\prod_{i,j=1}^kF_{ij}(x)\ .
\elabel{zetaans}\end{equation}
Next we consider the insertions at $x_i=0$. Focusing on the ${\cal M}'$ modes
first, one finds from Eq.~\eqref{corriganid}:
\begin{equation}
\tr_{N\,}^{}\lambda^2(0)\ =\
2\sum_{i,j=1}^k({\cal M}^{\prime\alpha})_{ij}({\cal M}'_{\alpha})_{ji}\,
(\rho_i^{-4}+\rho_j^{-4}+\rho_i^{-2}\rho_j^{-2})\
+\ \cdots\ ,
\elabel{Mpstuff}\end{equation}
omitting the $\nu\times\bar \nu$ terms. Hence the ${\cal M}'$ integrations yield
\begin{equation}
\begin{pmatrix}kN-k^2\\ k^2\end{pmatrix}\,(k^2)!\,2^{k^2}\,\prod_{i,j=1}^k
(\rho_i^{-4}+\rho_j^{-4}+\rho_i^{-2}\rho_j^{-2})\ ,
\elabel{Mpans}\end{equation}
where the combinatoric factors in \eqref{Mpans} (as well as in \eqref{zetaans})
have been explained previously.\footnote{One can easily check that
these large-$N$ formulae are consistent with the explicit 1-instanton
calculation presented in Sec.~II which is exact in $N$. In particular,
if one takes $x_i=x$ while $x_j=0,$ then
Eqs.~\eqref{threeconb}-\eqref{threeconc} are suppressed
vis-$\grave{\rm a}$-vis Eq.~\eqref{threecona} by factors of $a'_n$, and
in turn, $a'_n\sim N^{-1/2}$ as follows from Eq.~\eqref{nuans} below.}
Finally we turn to the $\{\nu,\bar \nu\}$ integrations. Since (unlike
the $\zeta$ and ${\cal M}'$ modes) the number of $\nu$ and $\bar \nu$ modes
grows with $N$ as $kN-2k^2,$ it does not suffice merely to plug in the
saddle-point values \eqref{spansatza}-\eqref{spansatzb} and \eqref{spmore}. One must also calculate
the Gaussian determinant about the saddle-point, which provides an
${\cal O}(N^0)$ multiplicative contribution to the answer. Accordingly we
expand about \eqref{spansatza}-\eqref{spansatzb}
to quadratic order in the $a'_n$. The $\nu\times\bar \nu$
contribution to $\tr_{N\,}^{}\lambda^2(0)$ has the form
\begin{equation}
-{\textstyle{1\over2}}\bar \nu_{ju}\nu_{ui}\square f_{ij}{\Big|}_{x=0}\ =\
2\bar \nu_{ju}\nu_{ui}\big(f\cdot\tr^{}_2\, \bar b{\cal P} b\cdot f\big)_{ij}
{\Big|}_{x=0}
\elabel{nunubar}\end{equation}
as follows from Eqs.~\eqref{corriganid}-\eqref{adhmdefs}, and Eq.~(2.63)
of \cite{MO-III}.
Performing
the $\{\nu,\bar \nu\}$ integrations therefore gives
\begin{equation}\begin{split}
&(kN-2k^2)!\,\exp\Big((N-2k)\tr_{k\,}^{}\log\big(2f\cdot\tr^{}_2\, \bar b
{\cal P} b\cdot
f\big){\Big|}_{x=0}\,\Big)\ =
\\ &
(kN-2k^2)!\,\exp\Big((N-2k)\big(\log{\rm det}_k16(W^0)^{-2}\, -\,
\tfrac32\sum_{i,j=1}^k \sum_{n=1}^4a'_{nij}a'_{nji}\,(\rho_i^{-2}+\rho_j^{-2})\,
+\, {\cal O}(a'_n)^4\big)\,\Big)\ .
\elabel{nuans}\end{split}\end{equation}
The negative sign in front of the quadratic term in $a'_n$ confirms
that our saddle-point \eqref{spansatza}-\eqref{spansatzb} is in fact a minimum of the action.
Combining this expression with the measure factor in Eq.~\eqref{GImeas}, namely
\begin{equation}\begin{split}
({\rm det}_{2k}W)^{N-2k}\ &=\ \exp\big((N-2k)\log{\rm det}_{2k}W\big)
\\&=\
\exp\Big((N-2k)\big(\log{\rm det}_k(\tfrac12W^0)^{2}\,+\,{\cal O}(a'_n)^4\,
\big)\Big)\ ,
\elabel{measfact}\end{split}\end{equation}
and performing the Gaussian integrations over $a'_n$, yields:
\begin{equation}
2^{2k(N-2k)}\,(kN-2k^2)!\,\prod_{i,j=1}^k
\Big({2\pi\over3N(\rho_i^{-2}+\rho_j^{-2})}\Big)^2
\ +\ \cdots\ ,
\elabel{nuansb}\end{equation}
where the omitted terms are suppressed by powers of $N$.
Finally one combines Eqs.~\eqref{GImeas}, \eqref{diagcov}, \eqref{zetaans}, \eqref{Mpans} and
\eqref{nuansb} to obtain the leading-order result for the
correlator:
\begin{equation}
\lim_{N\rightarrow\infty}{\cal G}_{kN}\ =\
{2^{5kN+k^2-k+1/2}\,\pi^{2kN-k+1/2}\,e^{kN}\,(k^2)!\,k^{kN-k^2+1/2}\,{\cal
I}_k\,\Lambda^{3kN}\over3^{2k^2}\,N^{kN+k^2-1/2}\,k!}\ ,
\elabel{Gnfinal}\end{equation}
where ${\cal I}_k$ is the convergent integral
\begin{equation}
{\cal I}_k\ =\
\int_0^\infty
d\rho_1^2\cdots d\rho_k^2\,\prod_{1\le i<j\le
k}(\rho_i^2-\rho_j^2)^2\cdot\prod_{i,j=1}^k
F_{ij}(x)\,\big(1-(\rho_j/\rho_i+
\rho_i/\rho_j)^{-2}\big)\ .
\elabel{Ikdef}\end{equation}
Note that ${\cal I}_k$ is independent of $x$ as a simple rescaling argument confirms.
For the simple case $k=1$, the $(\rho_i^2-\rho^2_j)^2$ terms in this
integral are absent; one finds
${\cal I}_1=\tfrac32$ and the expression \eqref{Gnfinal}
agrees---as it must---with the large-$N$ limit of the 1-instanton SCI
result \eqref{finalone}.
\setcounter{equation}{0}
\section{The 4-Point Function of the Gluino Condensate in $SU(2)$ Gauge Theory}
We have seen that cluster decomposition fails (both with and without a
KS vacuum) in the SCI calculation of the gluino condensate, for gauge
group $SU(N)$ in the large-$N$ limit. In this section we focus instead
on the gauge group $SU(2)$. In this case, at the 1-instanton level,
the 2-point function \eqref{SCIans} works out to:
\begin{equation}
\VEV{{\mathop{\rm tr}\nolimits\lambda^2(x_1)\over16\pi^2}\,{\mathop{\rm tr}\nolimits\lambda^2(x_2)\over16\pi^2}}\
=\ \textstyle{4\over5}\,\Lambda^{6}\ .
\elabel{SCItwo}\end{equation}
Here we will calculate the 4-point function, which receives a nonzero
contribution at the 2-instanton level:
\begin{equation}
\VEV{{\mathop{\rm tr}\nolimits\lambda^2(x_1)\over16\pi^2}\ {\mathop{\rm tr}\nolimits\lambda^2(x_2)\over16\pi^2}\
{\mathop{\rm tr}\nolimits\lambda^2(x_3)\over16\pi^2} \
{\mathop{\rm tr}\nolimits\lambda^2(x_4)\over16\pi^2}}\
=\ c\Lambda^{12}\ .
\elabel{SCIfour}\end{equation}
In the absence of a KS vacuum, generalized cluster decomposition
together with Eq.~\eqref{SCItwo}
predicts $c=(4/5)^2=.64$. Alternatively, in the presence of a KS vacuum,
weighted according to Eq.~\eqref{Shifmana} in order to reconcile the SCI and
WCI 1-instanton calculations, one expects $c=4/5=.8$. Instead, we have
calculated $c$ numerically, and find:
\begin{equation}
c\ \simeq\ .500\pm.026\ .
\elabel{cdef}\end{equation}
Here are the details of the calculation.
As mentioned above, for $k=2,$ one can eliminate the $\delta$-function
constraints in Eq.~\eqref{dmuphysdef} without changing variables. Another
simplification for the particular gauge group $SU(2)$ is that one can
adopt a concise quaternionic representation for the ADHM bosonic
collective coordinates, taking advantage of the fact that $SU(2)\cong
Sp(1).$ Specifically,
the $16$ gauge and $8$ gaugino
collective coordinates live, respectively, in the following
matrices:\footnote{See Ref.~\cite{MO-I} for details of notation and
conventions pertinent to Sec.~VI.}
\begin{equation}
a\ =\ \begin{pmatrix}w_1&w_2\\ a'_{11}&a'_{12}\\ a'_{12}&a'_{22}\end{pmatrix}\ ,\qquad
{\cal M}_\gamma\ =\ \begin{pmatrix}\mu_{1\gamma}&\mu_{2\gamma}\\
{\cal M}'_{11\gamma}&{\cal M}'_{12\gamma}\\
{\cal M}'_{12\gamma}&{\cal M}'_{22\gamma}\end{pmatrix}\ ,
\elabel{twoinstmats}\end{equation}
where $a=a_{\alpha{\dot\alpha}}=a_n\sigma^n_{\alpha{\dot\alpha}}$ and the
matrices $a_n$ as well as ${\cal M}_\gamma$ are real-valued (unlike the
complex-valued collective coordinates of the same name introduced in
Eq.~\eqref{adef} which are needed for general $SU(N)$). The resulting
2-instanton ``Osborn measure'' on these collective coordinates is
detailed in Refs.~\cite{Osborn:1981yf,MO-I,meas1}, and reads:
\defd\mu^{2}_{\rm phys}{d\mu^{2}_{\rm phys}}
\begin{equation}
\intd\mu^{2}_{\rm phys}\ =\ 2^{14}\Lambda^6\int
d^4w_1\,d^4w_2\,d^4a'_{11}\,d^4a'_{22}\, d^2\mu_1\,d^2\mu_2\,d^2{\cal M}'_{11}\,
d^2{\cal M}'_{22}\ {\big|\,|a'_3|^2-|a'_{12}|^2\,\big|\over|a'_3|^2}\ .
\elabel{osbdef}\end{equation}
Here the $\delta$-function constraints from the flat measure have been
used to eliminate $a'_{12}$ and ${\cal M}'_{12}$ in terms of the other
collective coordinates, via:
\begin{equation}
a'_{12}\ =\
{1\over4|a'_3|^2}\,a'_3(\bar w_2w_1-\bar w_1w_2)\ ,
\elabel{aonedef}\end{equation}
and
\begin{equation}
{\cal M}'_{12}\ =\ {1\over2|a'_3|^2}\,a'_3\,\big(\,2\bar a'_{12}{\cal M}'_3+\bar w_2\mu_1
-\bar w_1\mu_2\,\big)\ ,
\elabel{Monedef}\end{equation}
where we have defined
\begin{equation}
a'_3\,=\,{\textstyle{1\over2}}(a'_{11}-a'_{22})\
,\qquad{\cal M}'_3\,=\,{\textstyle{1\over2}}({\cal M}'_{11}-{\cal M}'_{22})\ .
\elabel{athreedef}\end{equation}
Into this measure one inserts the 4-point function of the classical condensate,
expressed as a function of the 2-instanton collective coordinates \eqref{twoinstmats}.
The 8-dimensional Grassmann integrations over $\{ \mu_1, \mu_2,
{\cal M}'_{11}, {\cal M}'_{22} \} $ are then accomplished in two steps.
The first step is to expand the integrand in terms of Grassmann variables
using a modified version of the program ``Dill'',
written for {\scshape Mathematica}.\footnote{``Dill'' is
a {\scshape Mathematica} package originally written by Vladan Lucic
\cite{lucic} in 1994
in order to simplify SUSY algebraic expressions.
This program can be modified so that it can
handle the large number of Grassmann variables that we need.}
The second step involves the explicit Grassmann integration, accomplished
using an ``awk-script'' implemented on a UNIX system and
made to perform the symbolic algebra of
Grassmann integration.
The resulting 16-dimensional bosonic integration over $ \{ w_1, w_2,
a'_{11}, a'_{22} \} $, the remaining quaternionic
variables, is carried out
using a standard Monte Carlo integration procedure.
The integrable singularities are handled using the standard procedure:
firstly, dropping a tiny region around the integrable singularities and
then making sure that the contribution from this dropped region is
negligibly smaller than the precision required.
After 450 million points have been sampled, we have obtained the numerical value
\eqref{cdef} given above. As a check on our numerics, we have also verified
the constancy of the answer by comparing different choices for the
four space-time insertion points.
\setcounter{equation}{0}
\section{Discussion}
The mismatch between the strong coupling and weak coupling
calculations is a fascinating puzzle. Previously, only the mismatch at
the one instanton level was known; now we see a mismatch established
at large $N$ for all instanton numbers, and for $N=2$ at the
2-instanton level. Certainly we do not mean to imply that, because of
this mismatch, SCI calculations are all necessarily suspect; indeed an
${\cal N}=4$ supersymmetric version of an SCI calculation performed by some
of us \cite{MO-III} has recently provided a dramatic quantitative and
qualitative verification of Maldacena's conjecture. However in this
case the coupling does not run and the calculation can be performed at
weak coupling, actually small $g^2N$,
where the instanton approximation is fully justified. The
continuation to strong coupling, large $g^2N$, is then accomplished by means of a
non-renormalization theorem. Rather, our
objections to the SCI computation
are more narrow and technical in scope: specifically, our
calculations imply a fundamental breakdown of clustering in the
instanton approximation to the gluino condensate at strong coupling.
One may wonder what the origin for this breakdown is? The usual
justification for the strong coupling calculation is that one can take
$|x_i-x_j|$ much smaller than the scale of strong coupling effects
$\Lambda^{-1}$ and so the theory would be weakly coupled, due to
asymptotic freedom, and the instanton calculation would be
justified. Then, since the correlation functions \eqref{Gndef} are
independent of the positions, the result would be valid at all
distances. This point-of-view has simultaneously been used and
criticized by various authors \cite{Rossi:1984bu,
Novikov:1985ic,Shifman:1988ia}. The asymptotic freedom argument means
that the first-order, second-order, etc., perturbative corrections to
the SCI calculations are small---indeed, for the gluino condensate
correlators discussed herein, these perturbative corrections are
entirely absent due to a nonrenormalization theorem
\cite{Novikov:1985ic}. However,
the asymptotic freedom argument does not guarantee that the zeroth
order instanton calculation is itself complete; there may be
other non-perturbative configurations contributing to the
correlators. Indeed, the breakdown of cluster suggests that such
additional nonperturbative configurations (with
size of order $\Lambda^{-1}$) must be present, and that they must
account for the mismatch between the SCI and WCI calculations.
In contrast, it seems that the WCI
calculation uses a method that has amassed a considerable
pedigree. These kinds of calculations appear to be consistent in all
applications and agree with other non-instanton methods
\cite{Finnell:1995dr}; for example, the two-instanton check of the
Seiberg-Witten approach to ${\cal N}=2$ theories \cite{MO-I,MO-II} and
the latter calculation in Sec.~III. Moreover, in the WCI set-up,
large-scale nonperturbative configurations as just discussed, such as
instanton-antiinstanton pairs of size $\Lambda^{-1}$, would be
exponentially suppressed in the path integral so long as
$\Lambda\ll{\rm v}$. We should further note that, as the separation
between insertions tends to zero, the WCI calculation does \it not \rm
smoothly go over to the SCI calculation as one would naively expect; there are additional
important contributions which will be discussed in a separate publication (work
in progress).
It is unfortunate that, based on our results,
the highly original and intriguing (both theoretically and phenomenologically)
proposal of Shifman, namely the existence of a chirally
symmetric vacuum state, loses much of its {\it raison d'\^etre\/}. Then
again, we have not actually ruled out the existence of such a state.
After the completion of this work, it has been suggested that the
mixing parameter $p$ of the KS vacuum, defined in
Eq.~\eqref{varrhopdef} above, may actually be instanton number
dependent \cite{shifpc}. {\it Prima facie\/}, this appears to be
incompatible with invariance under large gauge transformations
($|k\rangle\rightarrow|k+1\rangle$);
however, if such a counter-intuitive flexibility is
permissible in the definition of the instanton vacuum,
clustering in the presence of the KS vacuum can be saved.
\def\bar{\rm I}{\bar{\rm I}}
Conceptual difficulties with the instanton approximation and cluster
decomposition were pointed out in the context of
pure (non-supersymmetric) QCD some time ago. Since this may have some
bearing on the present discussion, we review some comments of L\"uscher
regarding this issue \cite{Luscher:1979yd}. The
pure instanton (i.e.~no anti-instantons) approximation to QCD
obviously violates parity since
\begin{equation}
\VEV{\mathop{\rm tr}\nolimits F_{nm}{}\,^*F_{nm}}_{\rm inst.}=\rho\neq0\ ,
\end{equation}
where $\rho$ here is the instanton density.
Parity is then recovered by summing over instantons (I) {\it and\/}
anti-instantons ($\bar{\rm I}$); however, in this approximation the cluster property
would not hold. To see this note that
\begin{equation}
\VEV{\mathop{\rm tr}\nolimits F_{nm}{}\,^*F_{nm}(x)\ \mathop{\rm tr}\nolimits F_{pq}{}\,^*F_{pq}(0)}\
\underset{|x|\rightarrow\infty}\rightarrow \ \rho^2\neq0\ ,
\label{sdfds}
\end{equation}
whereas
\begin{equation}
\VEV{\mathop{\rm tr}\nolimits F_{nm}{}\,^*F_{nm}}=\tfrac12\big(\VEV{\mathop{\rm tr}\nolimits F_{nm}{}\,^*F_{nm}}_{\rm
inst.}+\VEV{\mathop{\rm tr}\nolimits F_{nm}{}\,^*F_{nm}}_{\rm anti\hbox{\rm-}inst.}\big)=0\ .
\end{equation}
In order to resolve this clustering conundrum, it is apparent that
additional configurations, which may, in the dilute gas limit, be
thought of as mixtures of
instantons and anti-instantons, would need to be incorporated in the
approximation. In this case the two-point function \eqref{sdfds} would
indeed be zero, the result of summing the I$\,$I, I$\,\bar{\rm I},$ $\bar{\rm I}\,$I and
$\bar{\rm I}\,\bar{\rm I}$ contributions which on average all contribute equally. Away from the dilute instanton gas limit, the identification and physical interpretation of these additional cluster-restoring nonperturbative configurations is necessarily more subtle.
In summary, the results of this paper imply something analogous
in the ${\cal N}=1$ theory: additional
configurations must contribute to the correlators
at strong coupling and resolve the breakdown of clustering (as well as
repairing the mismatch between the SCI and WCI calculations).
In fact, it was suspected some time ago
(see Ref.~\cite{Osborn:1981yf,Belavin:1979fb} and references therein)
that in strongly coupled theories, it may be more appropriate to think
of instantons as composite configurations of some more basic objects:
so-called ``instanton partons''.
The dominant contributions to the path integral at
strong coupling would then arise from the partons themselves.
In Ref.~\cite{DHKM}, we make this piece of folklore more precise by
identifying instanton partons with the monopole configurations of the
supersymmetric Yang-Mills compactified on the cylinder
${\Bbb R}^3 \times S^1$, with the circle having circumference $\beta$.
Each monopole has precisely {\it two\/} gluino zero modes, rather than
four for the instanton. The instanton itself is then identified with a
specific two-monopole
configuration. We calculate the monopole contribution to the gluino
condensate and then, at the end of the day,
take the decompactification limit $\beta \to \infty$.
The value of the gluino condensate obtained in this way,
is precisely the WCI result \eqref{WCIans}.
\section*{Acknowledgments}
We thank Philippe Pouliot, Larry Yaffe, Alex Kovner, Misha Shifman and
Nick Dorey for valuable discussions.
VVK and MPM acknowledge a NATO Collaborative Research Grant.
TJH and VVK acknowledge the TMR network grant FMRX-CT96-0012.
\startappendix
\setcounter{equation}{0}
\Appendix{Constancy of the Correlators}
Since the constancy of the correlation function \eqref{Gndef} plays such an
important role in our analysis, in this appendix, we review the
arguments leading to this result. More importantly, we explain how the
field theory proof remains valid in the instanton approximation.
First the field theory proof \cite{Novikov:1983ee,Rossi:1984bu}.
The argument is completely general and
applies to the correlation functions of any {\it lowest component\/} $A$ of a
gauge invariant chiral superfield $\Phi$:
\begin{equation}
\Phi=A(x)+\sqrt2\theta\psi(x)+\cdots\ .
\end{equation}
In the present discussion, the operator
$\mathop{\rm tr}\nolimits\,\lambda^2$ is the lowest component of the
chiral superfield $\mathop{\rm tr}\nolimits\, W^\alpha W_\alpha$, where $W_\alpha$ is
supersymmetric field strength.
Consider the correlation function
\begin{equation}
\VEV{A_1(x_1)\cdots A_p(x_p)}\ .
\end{equation}
We will show that this is independent of the $x_i$'s. To this end, one
has
\begin{equation}
{\partial\over\partial x^n}A_i(x)={i\over4}\bar\sigma_n^{{\dot\alpha}\alpha}
\{\bar Q_{\dot\alpha},\psi_{i\alpha}(x)\}\ .
\elabel{derivact}\end{equation}
Hence
\begin{equation}\begin{split}
{\partial\over\partial x^n_i}&\VEV{A_1(x_1)\cdots A_p(x_p)}\\
=&\tfrac i4\bar\sigma_n^{{\dot\alpha}\alpha}\langle0\vert A_1(x_1)\cdots A_{i-1}(x_{i-1})
\{\bar Q_{\dot\alpha},\psi_{i\alpha}(x_i)\}A_{i+1}(x_{i+1})\cdots A_p(x_p)\vert0\rangle\\
=&-\tfrac i4\bar\sigma_n^{{\dot\alpha}\alpha}\sum_{j=1}^{i-1}\langle0\vert
A_1(x_1)\cdots[\bar Q_{\dot\alpha},A_j(x_j)]\cdots
A_{i-1}(x_{i-1})\psi_{i\alpha}(x_i)A_{i+1}(x_{i+1})\cdots A_{p}(x_p)\vert0\rangle\\
+&\tfrac
i4\bar\sigma_n^{{\dot\alpha}\alpha}\sum_{j=i+1}^{p}\langle0\vert
A_{1}(x_1)\cdots A_{i-1}(x_{i-1})
\psi_{i\alpha}(x_i)A_{i+1}(x_{i+1})\cdots [\bar
Q_{\dot\alpha},A_{j}(x_j)]
\cdots A_{p}(x_p)\vert0\rangle\ ,
\elabel{commut}\end{split}\end{equation}
where the last line follows by commuting the $\bar Q_{\dot\alpha}$ through the
other insertions, to the left and right, respectively, until it hits
the vacuum which it annihilates. But $[\bar Q_{\dot\alpha},A_{j}(x)]=0$ and
therefore the right-hand side of \eqref{commut} vanishes and
consequently the correlation function is, indeed, independent of the
insertion points. {\it QED\/}.
If the multiple correlator of $\mathop{\rm tr}\nolimits\,\lambda^2$ is constant in the full
field theory, it then becomes an issue as to whether this constancy is
retained in the instanton approximation. That it is, rests upon two
facts. Firstly, the supersymmetry transformations of the fields can be
traded for supersymmetry transformations of the collective coordinates
\cite{Novikov:1983ee,meas1,meas2,KMS,MO-II}. In other words, the
supersymmetry algebra is represented on the collective
coordinates. Specifically, under an infinitesimal supersymmetry
transformation $\xi Q+\bar\xi\bar Q$:
\begin{subequations}
\begin{align}
&\delta a_{\dot\alpha}=\bar\xi_{\dot\alpha}{\cal M},\qquad \delta\bar
a^{\dot\alpha}=-\bar{\cal M}\bar\xi^{\dot\alpha}\elabel{susyta}\\
&\delta{\cal M}=-4ib^\alpha\xi_\alpha,\qquad \delta\bar{\cal M}=4i\xi^\alpha\bar
b_\alpha\ .\elabel{susytb}
\end{align}
\end{subequations}
In particular, \eqref{derivact} will hold, with the fields replaced
by their expression in the instanton background and with the right-hand side
involving the appropriate transformation of the collective
coordinates. {\it Ipso facto\/}, the argument leading to \eqref{commut} will hold
with the transformations acting on the collective coordinates;
moreover $[\bar Q_{\dot\alpha},A]=0$, understood as a transformation of the
collective coordinates. The remaining piece of the proof is the
analogue of the fact that $\bar Q_{\dot\alpha}$ annihilates the vacuum state. In
the instanton approximation, where the functional integral is
approximated by the integral over the collective coordinates, the
analogue of the
statement that the vacuum is a supersymmetry invariant, is the
statement that the measure on the space of collective coordinates is
invariant under the supersymmetry transformations \eqref{susyta}-\eqref{susytb}.
This invariance was proved in \cite{meas1,meas2}.
|
\section{Introduction}
The path integral representation of a quantum system is
very helpful to visualize the quantum dynamics in terms of classical
concepts. In particular, spin coherent states allow for a
representation of a spin as a point on a unit sphere depicting the
dynamics in terms of pseudo-classical spin rotations. Unfortunately, in
the standard spin coherent state path integral, the action contains no
terms quadratic in the velocities, and the typical paths are therefore
not continuous, as it is the case with the familiar Feynman
configuration space path integral. This problem has attracted
considerable attention of mathematical physicists
\cite{klauder}-\cite{bodmann}. In most of these studies a spin in a
constant magnetic field is considered which allows for an explicit
solution. Other work \cite{ellinas}-\cite{cabra} allowing for
time-dependent fields examines the discrete time-lattice version of
the path integral, and it is usually concluded \cite{ercolessi,shibata} that
only formal calculations are possible once the continuous path
integral is employed.
Here we re-examine the semi\-classical propagator of the continuous
time path integral starting from \mbox{Klauder's}
obser\-vation \cite{klauder} that a Wiener regularized coherent state path
integral for the free spin, that is a spin in the absence of a magnetic
field, allows for a well-defined stationary phase
approximation that turns out to be exact. We will show that
essentially the same type of semiclassical approximation leads to the
exact propagator also in the presence of a magnetic field of arbitrary
time dependence.
The paper is organized as follows. In section~2 we introduce the basic
notation and the spin coherent state path integral for a
spin-$\frac{1}{2}$-system. We then present, in section~3,
a spherical Wiener measure regularizing the path integral
and discuss the semiclassical approximation which is shown to
become exact. In section~4 we transform the angle variables to
variables that allow for a more effective calculation of semiclassical
propagators and calculate the spin coherent propagators for two
models. Finally, in section~5, we present our conclusions.
\section{Spin coherent state path integral}
We consider a spin-$\frac{1}{2}$ described by the spin operators
$S_{i},\, (i=x,y,z)$ with the two-dimensional Hilbert space spanned,
e.g., by the eigenvectors $\bigl|\uparrow\bigr>$ and
$\bigl|\downarrow\bigr>$ of $S_z$. For each orientation in real space
characterized by a polar angle $\vartheta$ and an azimuthal angle
$\varphi$ we may introduce a spin coherent state
\cite{perelomov} (we put $\hbar=1$)
\begin{equation}
\bigl|\Omega\bigr> \equiv
\bigl|\vartheta\varphi\bigr> = {\rm e}^{-{\rm i}\varphi S_{z}}
{\rm e}^{-{\rm i} \vartheta S_{y}} \bigl|\uparrow\bigr>.
\label{eq20}
\end{equation}
These states are not orthogonal but form an overcomplete basis in the
Hilbert space. The overlap of two coherent states reads
\begin{equation}
\left<\Omega''|\Omega'\right>=
\cos\left(\frac{\vartheta''}{2}\right)\cos\left(\frac{\vartheta'}{2}\right)
{\rm e}^{\frac{{\rm i}}{2}(\varphi''-\varphi')} +
\sin\left(\frac{\vartheta''}{2}\right)
\sin\left(\frac{\vartheta'}{2}\right)
{\rm e}^{-\frac{{\rm i}}{2}(\varphi''-\varphi')},
\label{eq21}
\end{equation}
and the identity may be represented as
\begin{equation}
I =\frac{1}{2\pi}\int {\rm d}\cos(\vartheta) {\rm d}\varphi
\left|\Omega\right>\left<\Omega\right|.
\label{eq22b}
\end{equation}
Furthermore, the matrix elements of the spin operators take the form
\begin{eqnarray}
\left<\Omega''|S_x|\Omega'\right> &= \frac{1}{2}\left[
\cos\left(\frac{\vartheta''}{2}\right)\sin\left(\frac{\vartheta'}{2}\right)
{\rm e}^{\frac{{\rm i}}{2}(\varphi''+\varphi')} +
\sin\left(\frac{\vartheta''}{2}\right)\cos\left(\frac{\vartheta'}{2}\right)
{\rm e}^{-\frac{{\rm i}}{2}(\varphi''+\varphi')} \right]
\nonumber\\
\left<\Omega''|S_y|\Omega'\right> &= \frac{1}{2{\rm i}}\left[
\cos\left(\frac{\vartheta''}{2}\right)\sin\left(\frac{\vartheta'}{2}\right)
{\rm e}^{\frac{{\rm i}}{2}(\varphi''+\varphi')} -
\sin\left(\frac{\vartheta''}{2}\right)\cos\left(\frac{\vartheta'}{2}\right)
{\rm e}^{-\frac{{\rm i}}{2}(\varphi''+\varphi')} \right]
\nonumber\\
\left<\Omega''|S_z|\Omega'\right> &= \frac{1}{2}\left[
\cos\left(\frac{\vartheta''}{2}\right)\cos\left(\frac{\vartheta'}{2}\right)
{\rm e}^{\frac{{\rm i}}{2}(\varphi''-\varphi')} -
\sin\left(\frac{\vartheta''}{2}\right)\sin\left(\frac{\vartheta'}{2}\right)
{\rm e}^{-\frac{{\rm i}}{2}(\varphi''-\varphi')} \right].
\label{eq22}
\end{eqnarray}
Let us consider a spin in a magnetic field of arbitrary time
dependence described by the Hamiltonian
\begin{equation}
H(t)=B_x(t)S_x+B_y(t)S_y+B_z(t)S_z,
\label{eq22c}
\end{equation}
which gives rise to the unitary time evolution operator
\begin{equation}
U(t)={\cal T}_s\exp\left\{-{\rm i}\int_0^{t}{\rm d} s\,H(s)\right\},
\label{eq22d}
\end{equation}
where ${\cal T}_s$ is the time ordering operator. $U(t)$ can be shown
to be of the form \cite{gilmore}
\begin{equation}
U(t)=\left(\begin{array}{cc}
a(t) & b(t) \\ -b^{*}(t) & a^{*}(t) \\
\end{array}\right),\qquad|a(t)|^2+|b(t)|^2=1,
\label{eq21b}
\end{equation}
where the coefficients obey the linear differential equations
\begin{eqnarray}
\dot a(t)&= -\frac{{\rm i}}{2}B_z(t)a(t)
+\frac{1}{2}\left[{\rm i} B_x(t)+B_y(t)\right]b^{*}(t)\nonumber\\
\dot b(t)&= -\frac{{\rm i}}{2}B_z(s)b(t)
-\frac{1}{2}\left[{\rm i} B_x(t)+B_y(t)\right]a^{*}(t).
\label{eq21c}
\end{eqnarray}
Employing a Trotter decomposition, the propagator may be written as
\begin{equation}
\bigl<\Omega''\bigr|U(t)\bigl|\Omega'\bigr>=
\lim_{\epsilon\to 0}
\int\prod_{k=1}^{n}\frac{{\rm d}\cos(\vartheta_k)
{\rm d}\varphi_k}{2\pi} \prod_{k=0}^{n}
\bigl<\Omega_{k+1}\bigr|{\cal
T}_s\exp\{-{\rm i}\int_{k\epsilon}^{(k+1)\epsilon}
{\rm d} s\, H(s)\}\bigl|\Omega_{k}\bigr>,
\label{eq23}
\end{equation}
where $\epsilon=t/n$, $\Omega_{0}=\Omega'$,
$\Omega_{n+1}=\Omega''$. Now, for $\epsilon\rightarrow 0$ we have
\begin{eqnarray}
\bigl<\Omega_{k+1}\bigr|{\cal
T}_s\exp\{-{\rm i}\int_{k\epsilon}^{(k+1)\epsilon}{\rm d} s\, H(s)\}
\bigl|\Omega_{k}\bigr>
=&\bigl<\Omega_{k+1}\bigr.\bigl|\Omega_{k}\bigr>
\left(1-{\rm i}\epsilon\frac{\bigl<\Omega_{k+1}\bigr|H(k\epsilon)
\bigl|\Omega_{k}\bigr>}
{\bigl<\Omega_{k+1}\bigr.\bigl|\Omega_{k}\bigr>} \right) \nonumber\\
&+ {\cal O}(\epsilon^2), \label{eq24}
\end{eqnarray}
and the right hand side of equation (\ref{eq23}) can be expressed as
\begin{displaymath}
\lim_{\epsilon\to 0} \int\prod_{k=1}^{n}\frac{{\rm d}\cos(\vartheta_k)
{\rm d}\varphi_k}{2\pi} \exp\left\{\sum_{k=0}^{n}
\left[\log{\bigl<\Omega_{k+1}\bigr.\bigl|\Omega_{k}\bigr>}-
{\rm i}\epsilon\frac{\bigl<\Omega_{k+1}\bigr|H(k\epsilon)
\bigl|\Omega_{k}\bigr>}
{\bigl<\Omega_{k+1}\bigr.\bigl|\Omega_{k}\bigr>} \right]\right\}.
\end{displaymath}
With the assumption that for $\epsilon\rightarrow 0$ the paths
$\Omega(s)$ remain continuous, we may expand the terms in the exponent
as
\begin{eqnarray}
\log{\bigl<\Omega_{k+1}\bigr.\bigl|\Omega_{k}\bigr>}&=&
\frac{{\rm i}}{2}\frac{\cos(\vartheta_{k+1})+\cos(\vartheta_k)}{2}
(\varphi_{k+1}-\varphi_{k})\label{eq26}\\
& &-\frac{1}{8}\left[(\vartheta_{k+1}-\vartheta_k)^2
+\sin^2(\vartheta_k)(\varphi_{k+1}-\varphi_k)^2 \right]
+{\cal O}(\delta \Omega^3), \nonumber
\end{eqnarray}
and
\begin{equation}
\epsilon\frac{\bigl<\Omega_{k+1}\bigr|H(k\epsilon)\bigl|\Omega_{k}\bigr>}
{\bigl<\Omega_{k+1}\bigr.\bigl|\Omega_{k}\bigr>} =
\epsilon\bigl<\Omega_{k}\bigr|H(k\epsilon)\bigl|\Omega_{k}\bigr>
+{\cal O}(\epsilon\delta \Omega) \label{eq28}.
\end{equation}
While the term of order $\delta \Omega^2$ in equation (\ref{eq26}) has the
form of the line element on the sphere, this term does not lead
to a Wiener measure in the path integral, since there are no factors
of $\epsilon$ in the denominator. Hence, the assumption of continuous
paths is obsolete, and the resulting continuous path integral
\begin{equation}
\bigl<\Omega''\bigr|U(t)\bigl|\Omega'\bigr>=
\int_{(\vartheta',\varphi')}^{(\vartheta'',\varphi'')}
{\cal D}\cos(\vartheta)\,{\cal D}\varphi
\exp\left\{ {\rm i}\int_{0}^{t}{\rm d} s\left[\frac{1}{2}
\cos(\vartheta)\dot\varphi-H(\vartheta,\varphi,s)\right] \right\},
\label{eq28b}
\end{equation}
where
\begin{eqnarray}
H(\vartheta,\varphi,t)&=&\bigl<\Omega\bigr|H(t)\bigl|\Omega\bigr>
\label{eq28c}\\
&=&\frac{1}{2}\left[B_x(t)\sin(\vartheta)\cos(\varphi)+
B_y(t)\sin(\vartheta)\sin(\varphi)+B_z(t)\cos(\vartheta)\right],
\nonumber
\end{eqnarray}
has only formal meaning.
\section{Wiener regularization and semiclassical approximation}
Following Klauder \cite{klauder} the ill-defined path integral
(\ref{eq28b}) can be turned into a meaningful expression if the
propagator is written as
\begin{equation}
\bigl<\Omega''\bigr|U(t)\bigl|\Omega'\bigr>=\lim_{\nu\to\infty}
\int {\rm d} \mu_{\rm{W}}\,\exp{ \biggl\{ {\rm i}\int_{0}^{t} {\rm d} s
\Bigl[ \frac{1}{2}\cos(\vartheta)\dot\varphi -
H(\vartheta,\varphi,s) \Bigr]\biggr\} } ,
\label{eq29}
\end{equation}
where
\begin{equation}
{\rm d}\mu_{\rm{W}}=N\prod_{s=0}^{t} {\rm d}\cos(\vartheta(s))
{\rm d}\varphi(s) \,\exp{ \biggl\{
- \frac{1}{4\nu}\int_{0}^{t} ds \Bigl[\dot\vartheta^2
+\sin^2(\vartheta)\dot\varphi^2 \Bigr]\biggr\} }
\label{eq30}
\end{equation}
is a Wiener measure on the unit sphere which enforces that only
continuous Brownian motion paths contribute to the path integral. This
amounts to replacing the action of the spin by
\begin{equation}
S_{\nu}[\Omega(s)]= \int_{0}^{t} {\rm d} s
\left\{ \frac{{\rm i}}{4\nu}
\left[ \dot\vartheta^2+\sin^2(\vartheta)\dot\varphi^2\right]
+ \frac{1}{2}\cos(\vartheta)\dot\varphi - H(\vartheta,\varphi,s) \right\}.
\label{eq29b}
\end{equation}
In the limit $\nu\rightarrow\infty$, the $\nu$-dependent terms in the
action vanish, and formally the previous expression (\ref{eq28b}) is
recovered.
Let us now investigate the semiclassical approximation of the path
integral. For finite $\nu$ the Euler-Lagrange equations following from
the action (\ref{eq29b}) read
\begin{eqnarray}
\frac{1}{2}\sin(\vartheta)\dot{\varphi}+\frac{\partial
H}{\partial\vartheta} &=&
-\frac{{\rm i}}{2\nu} \left[\ddot\vartheta -
\sin(\vartheta)\cos(\vartheta)\dot\varphi^2\right]
\nonumber\\
\frac{1}{2}\sin(\vartheta)\dot\vartheta
-\frac{\partial H}{\partial\varphi}&=&
\frac{{\rm i}}{2\nu}\left[ \sin^2(\vartheta)\ddot\varphi
+2\sin(\vartheta)\cos(\vartheta)\dot\vartheta\dot\varphi
\right].\label{eq32}
\end{eqnarray}
For given boundary conditions
$\Omega(0)=\Omega'\equiv (\vartheta',\varphi')$,
$\Omega(t)=\Omega''\equiv (\vartheta'',\varphi'')$ and $t\gg1/ \nu$, these
equations have for small and intermediate times $s \,(s\ll t-1/ \nu)$ a
solution of the form
\begin{equation}
\cos(\vartheta(s))=\cos(\bar\vartheta(s))
+\left[\cos(\vartheta')-\cos(\bar \vartheta')\right]{\rm e}^{-\nu s},
\label{eq33a}
\end{equation}
and
\begin{eqnarray}
\varphi(s)&=& \bar \varphi(s)+ \varphi' - \bar \varphi'
+\frac{{\rm i}}{2}\log\left[\frac{1+\cos(\vartheta')}{1-\cos(\vartheta')}\right]
\nonumber\\
&&-\frac{{\rm i}}{2}\log\left[\frac
{1+\cos(\bar \vartheta')+(\cos(\vartheta')-
\cos(\bar \vartheta')){\rm e}^{-\nu s}}
{1-\cos(\bar \vartheta')-(\cos(\vartheta')-
\cos(\bar\vartheta')){\rm e}^{-\nu s}}\right],
\label{eq34}
\end{eqnarray}
while for intermediate and large times $s \,(s\gg1/ \nu)$ the solution becomes
\begin{equation}
\cos(\vartheta(s))=\cos(\bar\vartheta(s))
+\left[\cos(\vartheta'')-\cos(\bar \vartheta'')\right]{\rm e}^{-\nu(t-s)},
\label{eq33b}
\end{equation}
and
\begin{eqnarray}
\varphi(s)&=&\bar \varphi(s)+ \varphi'' - \bar \varphi''
-\frac{{\rm i}}{2}\log\left[\frac{1+\cos(\vartheta'')}{1-\cos(\vartheta'')}\right]
\nonumber\\
&& +\frac{{\rm i}}{2}\log\left[\frac
{1+\cos(\bar \vartheta'')+(\cos(\vartheta'')
-\cos(\bar \vartheta'')){\rm e}^{-\nu(t-s)}}
{1-\cos(\bar \vartheta'')-(\cos(\vartheta'')
-\cos(\bar \vartheta'')){\rm e}^{-\nu(t-s)}}\right].
\label{eq35}
\end{eqnarray}
Here, $\bar\Omega(s)\equiv(\bar\vartheta(s),\bar\varphi(s))$ is a solution
of the classical equations of motion
\begin{eqnarray}
\frac{1}{2}\sin(\bar\vartheta)\dot{\bar\varphi}&=&
-\frac{\partial H}{\partial\bar\vartheta}
\nonumber \\
\frac{1}{2}\sin(\bar\vartheta)\dot{\bar\vartheta}&=&
\frac{\partial H}{\partial\bar\varphi},
\label{eq45}
\end{eqnarray}
with the boundary conditions $\bar\Omega(0)
=\bar\Omega'\equiv(\bar\vartheta',\bar\varphi')$ and $\bar
\Omega(t)=\bar\Omega''\equiv(\bar\vartheta'',\bar\varphi'')$. Note that the
solution (\ref{eq33a}) and (\ref{eq34}) describes a jump within the time
interval $1/ \nu$ from the initial state $\Omega'$ to the starting
point $\bar\Omega'$ of the classical trajectory
(\ref{eq45}). Likewise, for $s$ near $t$,
the solution (\ref{eq33b}) and (\ref{eq35}) describes a jump from the
endpoint $\bar\Omega''$ of the classical trajectory to the final state
$\Omega''$. Now, in order that the short time and
the long time solutions coincide for intermediate times $1/
\nu\ll s \ll t-1/ \nu$, the boundary conditions of the classical path must
obey the relations
\begin{eqnarray}
\tan\left(\frac{\bar\vartheta'}{2}\right){\rm e}^{{\rm i}\bar\varphi'}
&=& \tan\left(\frac{\vartheta'}{2}\right){\rm e}^{{\rm i}\varphi'}
\label{eq50}\\
\tan\left(\frac{\bar\vartheta''}{2}\right){\rm e}^{-{\rm i}\bar\varphi''}
&=&\tan\left(\frac{\vartheta''}{2}\right){\rm e}^{-{\rm i}\varphi''}.
\label{eq55}
\end{eqnarray}
This determines the size of the jumps of the semiclassical trajectory
near the endpoints.
Inserting the semiclassical path (\ref{eq33a})-(\ref{eq35}) into the
action $S_{\nu}$ and taking the limit $\nu\rightarrow\infty$ one finds
\begin{equation}
\exp\left\{{\rm i} S_{\rm{cl}}\left[\Omega(s)\right]\right\}=
\sqrt{\frac{\sin(\vartheta')\sin(\vartheta'')}
{\sin(\bar\vartheta')\sin(\bar\vartheta'')} }
\exp\left\{{\rm i}\int_{0}^{t}{\rm d} s
\left[\frac{1}{2}\cos(\bar\vartheta)\dot{\bar{\varphi}}
-H(\bar \vartheta,\bar\varphi,s)\right]\right\}.
\label{eq65}
\end{equation}
Klauder has shown that for $H(\Omega)=0$ the expression (\ref{eq65})
coincides with the overlap
$\bigl<\Omega''\bigr.\bigl|\Omega'\bigr>$, so that this ``dominant
stationary phase approximation'' \cite{klauder} without fluctuations
becomes exact for a free spin. In general, for a non-vanishing
Hamiltonian, Klauder has concluded that (\ref{eq65}) ``cannot be
expected to provide the correct result by itself''. In fact, in later
work \cite{daubechies} he has suggested a different definition of the
spin coherent path integral. However, we will prove now that for
any Hamiltonian $H(t)$ the exact propagator is given by
\begin{equation}
\bigl<\Omega''\bigr|U(t)\bigl|\Omega'\bigr>=
\exp\left\{{\rm i} S_{\rm{cl}}\left[\Omega(s)\right]\right\}.
\label{eq66x}
\end{equation}
First we rewrite the overlap between the initial state and the
starting point of the classical trajectory as
\begin{eqnarray}
\bigl<\bar\Omega'\bigr.\bigl|\Omega'\bigr>&=&
\cos\left(\frac{\bar\vartheta'}{2}\right)\cos\left(\frac{\vartheta'}{2}\right)
{\rm e}^{\frac{{\rm i}}{2}(\bar\varphi'-\varphi')} +
\sin\left(\frac{\bar\vartheta'}{2}\right)\sin\left(\frac{\vartheta'}{2}\right)
{\rm e}^{-\frac{{\rm i}}{2}(\bar\varphi'-\varphi')}
\nonumber\\
&=&\frac{
\sqrt{\sin(\vartheta')\sin(\bar\vartheta')}
\left[1+ \tan\left(\frac{\bar\vartheta'}{2}\right)
\tan\left(\frac{\vartheta'}{2}\right){\rm e}^{-{\rm i}(\bar\varphi'-\varphi')} \right]}
{\sqrt{4\tan\left(\frac{\bar\vartheta'}{2}\right)
\tan\left(\frac{\vartheta'}{2}\right){\rm e}^{-{\rm i}(\bar\varphi'-\varphi')}}}.
\label{eq66a}
\end{eqnarray}
Making use of the jump condition (\ref{eq50}), this overlap can be
expressed as
\begin{equation}
\bigl<\bar\Omega'\bigr.\bigl|\Omega'\bigr>=
\sqrt{\frac{\sin(\vartheta')}{\sin(\bar\vartheta')}}.
\label{eq66b}
\end{equation}
Likewise, from (\ref{eq55}) we find for the jump at the endpoint
\begin{equation}
\bigl<\Omega''\bigr.\bigl|\bar\Omega''\bigr>
=\sqrt{\frac{\sin(\vartheta'')}{\sin(\bar\vartheta'')}}.
\label{eq66c}
\end{equation}
Therefore, we have from equation (\ref{eq65})
\begin{equation}
\exp\left\{{\rm i} S_{\rm{cl}}\left[\Omega(s)\right]\right\}=
\bigl<\Omega''\bigr.\bigl|\bar\Omega''\bigr>
\exp\left\{{\rm i}\int_{0}^{t}{\rm d} s
\left[\frac{1}{2}\cos(\bar\vartheta)\dot{\bar{\varphi}}
-H(\bar \vartheta,\bar\varphi,s)\right]\right\}
\bigl<\bar\Omega'\bigr.\bigl|\Omega'\bigr>.
\label{eq66}
\end{equation}
Now, the time evolution operator (\ref{eq21b}) acts on
a coherent state (\ref{eq20}) as
\begin{eqnarray}
U(t)\bigl|\Omega\bigr>&=&\left[ \cos\left(\frac{\vartheta}{2}\right)
{\rm e}^{-\frac{{\rm i}}{2}\varphi}a(t)+\sin\left(\frac{\vartheta}{2}\right)
{\rm e}^{\frac{{\rm i}}{2}\varphi}b(t)\right]\Bigl|\uparrow\Bigr>\nonumber\\
& & +
\left[-\cos\left(\frac{\vartheta}{2}\right)
{\rm e}^{-\frac{{\rm i}}{2}\varphi}b^{*}(t)+\sin\left(\frac{\vartheta}{2}\right)
{\rm e}^{\frac{{\rm i}}{2}\varphi}a^{*}(t)\right]\Bigl|\downarrow\Bigr>.
\label{su0n}
\end{eqnarray}
Apart from a phase factor, the right hand side is again a spin coherent state of
the form (\ref{eq20}). Hence,
\begin{equation}
U(t)\bigl|\Omega\bigr>=\exp\left\{{\rm i}\Phi(t)\right\}\Bigl|\Omega(t)\Bigr>,
\label{su1}
\end{equation}
where $\Omega(t)$ follows from equation (\ref{su0n}) as
\begin{eqnarray}
\vartheta(t)&=&
\arccos\biggl\{\Bigl[|a(t)|^2-|b(t)|^2\Bigr]\cos(\vartheta)\biggr.\nonumber\\
& &+ \biggl.
\Bigl[a^{*}(t)b(t){\rm e}^{{\rm i}\varphi}+
b^{*}(t)a(t){\rm e}^{-{\rm i}\varphi}\Bigr]\sin(\vartheta)\biggr\},
\label{su6}
\end{eqnarray}
and
\begin{eqnarray}
\varphi(t)
&=&-\frac{{\rm i}}{2}\log\left\{
\frac{a^{*}(t)b^{*}(t) - 2\left[a^{*}(t)^2 {\rm e}^{{\rm i}\varphi} -
b^{*}(t)^2 {\rm e}^{-{\rm i}\varphi}\right]\tan(\vartheta)}
{a(t)b(t) - 2\left[a(t)^2 {\rm e}^{-{\rm i}\varphi} -
b(t)^2{\rm e}^{{\rm i}\varphi}\right]\tan(\vartheta) } \right\},
\label{su7}
\end{eqnarray}
and where the phase takes the form
\begin{equation}
\Phi(t)=\frac{1}{2}\varphi(t)-\frac{{\rm i}}{2}
\log\left[\frac{a(t)\cos\left(\frac{\vartheta}{2}\right)
{\rm e}^{-{\rm i}\varphi} + b(t)\sin\left(\frac{\vartheta}{2}\right) }
{a^{*}(t)\cos\left(\frac{\vartheta}{2}\right)+
b^{*}(t)\sin\left(\frac{\vartheta}{2}\right){\rm e}^{-{\rm i}\varphi}
}\right] .
\label{su5}
\end{equation}
In this way we obtain in the spin coherent representation
\begin{eqnarray}
\bigl<\Omega''\bigr|U(t)\bigl|\Omega'\bigr>&=&
a(t)\cos\left(\frac{\vartheta''}{2}\right)\cos\left(\frac{\vartheta'}{2}\right)
{\rm e}^{\frac{{\rm i}}{2}(\varphi''-\varphi')}+
a^{*}(t)\sin\left(\frac{\vartheta''}{2}\right)\sin\left(\frac{\vartheta'}{2}\right)
{\rm e}^{-\frac{{\rm i}}{2}(\varphi''-\varphi')}
\nonumber \\
&& + b(t)\cos\left(\frac{\vartheta''}{2}\right)\sin\left(\frac{\vartheta'}{2}\right)
{\rm e}^{\frac{{\rm i}}{2}(\varphi''+\varphi')}
-b^{*}(t)\sin\left(\frac{\vartheta''}{2}\right)\cos\left(\frac{\vartheta'}{2}\right)
{\rm e}^{-\frac{{\rm i}}{2}(\varphi''+\varphi')}.
\nonumber\\\label{su8}
\end{eqnarray}
Next, let us show that $\Omega(t)$ is a solution of the classical
equations of motion (\ref{eq45}) with initial condition $\Omega(0)=\Omega$.
Inserting equation (\ref{eq21c}) for the time derivatives of the
coefficients $a(t)$ and $b(t)$ we find
\begin{eqnarray}
\frac{\partial\cos(\vartheta(t))}{\partial t}&=&
\frac{{\rm i}}{2}B_x(t)\biggl\{2\Bigl[a^*(t)b^*(t)-a(t)b(t)\Bigr]\cos(\vartheta)
\biggr.\nonumber\\
&& \biggl.-\left[a^*(t)^2 +b(t)^2\right]\sin(\vartheta){\rm e}^{{\rm i}\varphi}
+\left[a(t)^2 +b^*(t)^2\right]\sin(\vartheta){\rm e}^{-{\rm i}\varphi}\biggr\}
\nonumber\\
&&+\frac{1}{2}B_y(t)\biggl\{2\Bigl[a^*(t)b^*(t)+a(t)b(t)\Bigr]\cos(\vartheta)
\biggr.\label{su5b}\\
&&\biggl.-\left[a^*(t)^2 -b(t)^2\right]\sin(\vartheta){\rm e}^{{\rm i}\varphi}
-\left[a(t)^2 -b^*(t)^2\right]\sin(\vartheta){\rm e}^{-{\rm i}\varphi}
\biggr\}
.\nonumber
\end{eqnarray}
Now, using equations (\ref{su6}) and (\ref{su7}), the right hand side
simplifies to give
\begin{equation}
\frac{\partial\cos(\vartheta(t))}{\partial t}=
B_x(t)\sin(\vartheta(t))\sin(\varphi(t))
-B_y(t)\sin(\vartheta(t))\cos(\varphi(t)).
\label{su5c}
\end{equation}
In the same way one derives
\begin{equation}
\frac{\partial\varphi(t)}{\partial t}=
-B_x(t)\frac{\cos(\varphi(t))}{\tan(\vartheta(t))}
-B_y(t)\frac{\sin(\varphi(t))}{\tan(\vartheta(t))}+B_z(t).
\label{su5d}
\end{equation}
The equations (\ref{su5c}) and (\ref{su5d}) are readily shown to coincide
with the equations of motion (\ref{eq45}).
Hence, the time evolution of the labels
$\Omega(t)$ is purely classical.
The phase $\Phi(t)$ in equation (\ref{su1}) may be expressed in classical
terms as well. In order to do so, let us make use of the Schr\"odinger equation for the
operator $U(t)$. Since $\frac{\partial}{\partial t}U(t)=-{\rm i}
H(t)U(t)$, we find from equation (\ref{su1})
\begin{equation}
\Bigl<\Omega(t)\Bigr|\frac{\partial}{\partial t}\Bigl|\Omega(t)\Bigr>
=-{\rm i}\frac{\partial\Phi(t)}{\partial t} -{\rm i}\bigl<\Omega\bigr|H(t)\bigl|\Omega\bigr>.
\label{su2}
\end{equation}
This gives
\begin{eqnarray}
\Phi(t) &=&\int_{0}^{t} {\rm d} s
\Bigl<\Omega(s)\Bigr| {\rm i}\frac{\partial}{\partial s}-H(s)
\Big|\Omega(s)\Bigr> \nonumber\\
&=&\int_{0}^{t}{\rm d} s\left[
\frac{1}{2}\cos(\vartheta)\dot{\varphi} -H(\vartheta,\varphi,s)\right],
\label{su3}
\end{eqnarray}
where the right hand side is just the classical action. Since $\bar
\Omega''=\bar\Omega'(t)$, we have from equations (\ref{su1}) and (\ref{su3})
\begin{equation}
U(t)\bigl|\bar\Omega'\bigr>=\exp\left\{{\rm i}\int_{0}^{t}{\rm d} s
\left[\frac{1}{2}\cos(\bar\vartheta)\dot{\bar{\varphi}}
-H(\bar \vartheta,\bar\varphi,s)\right]\right\}\bigl|\bar\Omega''\bigr>.
\label{su3b}
\end{equation}
Now, we are in the position to rewrite the semiclassical propagator
(\ref{eq65}). Combining equations (\ref{eq66b}), (\ref{eq66c}) and (\ref{su3b}) we find
\begin{equation}
\exp\left\{{\rm i} S_{\rm{cl}}\left[\Omega(s)\right]\right\}=
\bigl<\Omega''\bigr.\bigl|\bar\Omega''\bigr>
\bigl<\bar\Omega''\bigr|U(t)\bigl|\bar\Omega'\bigr>
\bigl<\bar\Omega'\bigr.\bigl|\Omega'\bigr>.
\label{su4}
\end{equation}
On the other hand, using equation (\ref{su3b}) one obtains
\begin{equation}
\bigl<\Omega\bigr|U(t)\bigl|\bar\Omega'\bigr>=
\bigl<\Omega\bigr.\bigl|\bar\Omega''\bigr>
\bigl<\bar\Omega''\bigr|U(t)\bigl|\bar\Omega'\bigr>.
\label{su4b}
\end{equation}
Likewise, with $U(t)=U(-t)^{\dagger}$ one finds
\begin{equation}
\bigl<\bar\Omega''\bigr|U(t)\bigl|\Omega\bigr>=
\bigl<\bar\Omega''\bigr|U(t)\bigl|\bar\Omega'\bigr>
\bigl<\bar\Omega'\bigr.\bigl|\Omega\bigr>,
\label{su4c}
\end{equation}
and equation (\ref{su4}) finally becomes
\begin{equation}
\exp\left\{{\rm i} S_{\rm{cl}}\left[\Omega(s)\right]\right\}
=\bigl<\Omega''\bigr|U(t)\bigl|\Omega'\bigr>.
\label{su9}
\end{equation}
This shows that the dominant stationary phase approximation gives the
exact spin propagator.
To elucidate this point further, we demonstrate
that the semiclassical propagator obeys the Schr\"odinger
equation. From equation (\ref{eq65}) we find for the time rate of change
\begin{eqnarray}
\frac{\partial}{\partial t}
\exp\left\{{\rm i} S_{\rm{cl}}\left[\Omega(s)\right]\right\}&=&
\frac{1}{2}\Biggl\{
\frac{\cos(\bar\vartheta')}{\sin^2(\bar\vartheta')}
\frac{\partial\cos(\bar\vartheta')}{\partial t} +
\frac{\cos(\bar\vartheta'')}{\sin^2(\bar\vartheta'')}
\frac{\partial\cos(\bar\vartheta'')}{\partial t} \Biggr.
\nonumber\\
&&+{\rm i}\cos(\bar\vartheta'')
\left.\frac{\partial \bar\varphi(s,t)}{\partial s}\right|_{s=t}
-2{\rm i} H(\bar\vartheta'',\bar\varphi'',t)\nonumber\\
&&
+{\rm i}\int_{0}^{t}{\rm d} s \left[
-\sin(\bar\vartheta(s,t))\frac{\partial\bar\vartheta(s,t)}{\partial t}
\frac{\partial\bar\varphi(s,t)}{\partial s}
+\cos(\bar\vartheta(s,t))\frac{\partial^2\bar\varphi(s,t)}{\partial
t\partial s}\right.\nonumber\\
&&\Biggl.\left.
-2\frac{\partial H}{\partial\bar\vartheta}
\frac{\partial\bar\vartheta(s,t)}{\partial t}
-2\frac{\partial H}{\partial\bar\varphi}
\frac{\partial\bar\varphi(s,t)}{\partial t}
\right] \Biggr\}
\exp\left\{{\rm i} S_{\rm{cl}}\left[\Omega(s)\right]\right\}.
\label{su10}
\end{eqnarray}
Now, the jump conditions (\ref{eq50}) and (\ref{eq55}) give
\begin{equation}
\cos(\bar\vartheta')=\frac
{\left[1+\cos(\vartheta')\right]{\rm e}^{2{\rm i}\bar\varphi'}
-\left[1-\cos(\vartheta')\right]{\rm e}^{2{\rm i}\varphi'}}
{\left[1+\cos(\vartheta')\right]{\rm e}^{2{\rm i}\bar\varphi'}
+\left[1-\cos(\vartheta')\right]{\rm e}^{2{\rm i}\varphi'}},
\label{su10b}
\end{equation}
and a similar relation for $\cos(\bar\vartheta'')$.
These relations can be used to re-write the first two terms
on the right hand side of equation (\ref{su10}) as
\begin{equation}
\frac{\cos(\bar\vartheta')}{\sin^2(\bar\vartheta')}
\frac{\partial\cos(\bar\vartheta')}{\partial t}
={\rm i}\cos(\bar\vartheta')\frac{\partial\bar\varphi'}{\partial t}
\label{su11a}
\end{equation}
and
\begin{equation}
\frac{\cos(\bar\vartheta'')}{\sin^2(\bar\vartheta'')}
\frac{\partial\cos(\bar\vartheta'')}{\partial t}
=-{\rm i}\cos(\bar\vartheta'')\frac{\partial\bar\varphi''}{\partial t}
\label{su11b}.
\end{equation}
Then, after an integration by parts, equation (\ref{su10}) becomes
\begin{eqnarray}
\frac{\partial}{\partial t}
\exp\left\{{\rm i} S_{\rm{cl}}\left[\Omega(s)\right]\right\}&=&
\Biggl\{
{\rm i}\int_{0}^{t}ds \left[
\frac{1}{2}\sin[\bar\vartheta(s,t)]
\frac{\partial\bar\vartheta(s,t)}{\partial s}
-\frac{\partial H}{\partial\bar\varphi}
\right]\frac{\partial\bar\varphi(s,t)}{\partial t}\Biggr.
\nonumber\\
& &
-{\rm i}\int_{0}^{t}ds \left[
\frac{1}{2}\sin[\bar\vartheta(s,t)]\frac{\partial\bar\varphi(s,t)}{\partial s}
+\frac{\partial H}{\partial\bar\vartheta}
\right]\frac{\partial\bar\vartheta(s,t)}{\partial t}
\nonumber\\
& &
\Biggl. -{\rm i} H(\bar\vartheta'',\bar\varphi'',t) \Biggr\}
\exp\left\{{\rm i} S_{\rm{cl}}\left[\Omega(s)\right]\right\}.
\label{su12}
\end{eqnarray}
Therefore, with the equations of motions (\ref{eq45}), we obtain the
Schr\"odinger equation
\begin{equation}
\frac{\partial}{\partial t}
\exp\left\{{\rm i} S_{\rm{cl}}\left[\Omega(s)\right]\right\}=
-{\rm i} H(\bar\vartheta'',\bar\varphi'',t)\exp\left\{{\rm i} S_{\rm{cl}}\left[\Omega(s)\right]\right\}.
\label{su13}
\end{equation}
Note that the matrix element of the Hamiltonian at the endpoint
$\bar\Omega''$ of the classical trajectory generates the time rate of
change of the semiclassical propagator and not the matrix element at
the final state $\Omega''$. The semiclassical propagator may thus be
written as
\begin{equation}
\exp\left\{{\rm i} S_{\rm{cl}}\left[\Omega(s)\right]\right\}=
\exp\left\{-{\rm i}\int_0^{t}{\rm d} s\,
H(\bar\vartheta''(s),\bar\varphi''(s),s)\right\}
\bigl<\Omega''\bigr|\bigl.\Omega'\bigr>.
\label{su13b}
\end{equation}
To demonstrate that the Schr\"odinger equation (\ref{su13}) generates
the exact quantum dynamics, we start from the equation of motion of $U(t)$
\begin{equation}
\frac{\partial}{\partial t}\bigl<\Omega''\bigr|U(t)\bigl|\Omega'\bigr>
=-{\rm i}\bigl<\Omega''\bigr|H(t)U(t)\bigl|\Omega'\bigr>.
\label{su13c}
\end{equation}
In view of (\ref{su4b}) we have
\begin{eqnarray}
\bigl<\Omega''\bigr|H(t)U(t)\bigl|\Omega'\bigr>&=&
\bigl<\Omega''\bigr|H(t)\bigl|\bar\Omega''\bigr>
\bigl<\bar\Omega''\bigr| U(t)\bigl|\Omega'\bigr>\nonumber\\
&=&\frac{\bigl<\Omega''\bigr|H(t)\bigl|\bar\Omega''\bigr>}
{\bigl<\Omega''\bigr|\bigl.\bar\Omega''\bigr>}
\bigl<\Omega''\bigr|U(t)\bigl|\Omega'\bigr>,
\label{su13d}
\end{eqnarray}
where the first factor in the second line can also be written as
\begin{equation}
\frac{\bigl<\Omega''\bigr|H(t)\bigl|\bar\Omega''\bigr>}
{\bigl<\Omega''\bigr.\bigl|\bar\Omega''\bigr>}
= \bigl<\bar\Omega''\bigr|H(t)\bigl|\bar\Omega''\bigr>
=H(\bar\vartheta'',\bar\varphi'',t).
\label{su14}
\end{equation}
To show this, we represent the matrix elements (\ref{eq22}) of the
spin operators in the form
\begin{eqnarray}
\bigl<\Omega''\bigr|S_x\bigl|\bar\Omega''\bigr>&=&\frac{1}{2}
\frac{
\tan{\frac{\bar\vartheta''}{2}}{\rm e}^{{\rm i}\bar\varphi''} +
\tan{\frac{\vartheta''}{2}}{\rm e}^{-{\rm i}\varphi''}}
{1 + \tan{\frac{\vartheta''}{2}}
\tan{\frac{\bar\vartheta''}{2}}{\rm e}^{{\rm i}(\bar\varphi''-\varphi'')} }
\bigl<\Omega''\bigr.\bigl|\bar\Omega''\bigr>
\nonumber\\
\bigl<\Omega''\bigl|S_y\bigr|\bar\Omega''\bigr>&=&-\frac{{\rm i}}{2}\frac{
\tan{\frac{\bar\vartheta''}{2}}{\rm e}^{{\rm i}\bar\varphi''} -
\tan{\frac{\vartheta''}{2}}{\rm e}^{-{\rm i}\varphi''} }
{1 + \tan{\frac{\vartheta''}{2}}
\tan{\frac{\bar\vartheta''}{2}}{\rm e}^{{\rm i}(\bar\varphi''-\varphi'')} }
\bigl<\Omega''\bigr.\bigl|\bar\Omega''\bigr>
\nonumber\\
\bigl<\Omega''\bigr|S_z\bigl|\bar\Omega''\bigr>&=&\frac{1}{2}\frac{
1 - \tan{\frac{\vartheta''}{2}}
\tan{\frac{\vartheta'}{2}}{\rm e}^{{\rm i}(\bar\varphi''-\varphi'')} }
{1 + \tan{\frac{\vartheta''}{2}}
\tan{\frac{\bar\vartheta''}{2}}{\rm e}^{{\rm i}(\bar\varphi''-\varphi'')} }
\bigl<\Omega''\bigr.\bigl|\bar\Omega''\bigr>,
\label{su15}
\end{eqnarray}
and insert the jump condition (\ref{eq55}) to yield
\begin{eqnarray}
\bigl<\Omega''\bigr|S_x\bigl|\bar\Omega''\bigr>&=&
\frac{1}{2}\sin{\bar\vartheta''}\cos{\bar\varphi''}
\bigl<\Omega''\bigr.\bigl|\bar\Omega''\bigr>
\nonumber\\
\bigl<\Omega''\bigr|S_y\bigl|\bar\Omega''\bigr>&=&
\frac{1}{2}\sin{\bar\vartheta''}\sin{\bar\varphi''}
\bigl<\Omega''\bigr.\bigl|\bar\Omega''\bigr>
\nonumber\\
\bigl<\Omega''\bigr|S_z\bigl|\bar\Omega''\bigr>&=&
\frac{1}{2}\cos{\bar\vartheta''}
\bigl<\Omega''\bigr.\bigl|\bar\Omega''\bigr>.
\label{su16}
\end{eqnarray}
Then, from equation (\ref{eq22c}), the relation (\ref{su14}) is readily
shown, and equations (\ref{su13c}) and (\ref{su13d}) combine again to the
Schr\"odinger equation (\ref{su13}).
\section{Calculation of semiclassical propagators}
In the semiclassical theory described in the previous section
the starting and end points $\bar\Omega'$ and $\bar\Omega''$ of the classical
trajectory $\Omega(s)$ need to be determined by solving a boundary value
problem. This requires usually some effort. The same problem arises
for the coherent state propagator of a simple harmonic oscillator
\cite{klauder} and there it is useful to rewrite the propagator in
terms of the complex Glauber variables \cite{weissman}.
Here, we present a nonlinear transformation of the angle variables of
the semiclassical spin which allows for an
explicit calculation of the spin coherent state propagator
by solving only an initial value type problem. Let us
introduce the variables \cite{perelomov}
\begin{eqnarray}
\zeta&=&\tan\left(\frac{\vartheta}{2}\right){\rm e}^{{\rm i}\varphi}
\nonumber\\
\eta&=&\tan\left(\frac{\vartheta}{2}\right){\rm e}^{-{\rm i}\varphi}.
\label{sp5}
\end{eqnarray}
For real angles $\vartheta$ and $\varphi$ one has
$\eta=\zeta^{*}$, and the transformation corresponds to a
stereographic projection from the south pole of the
unit sphere onto the equatorial plane.
However, usually the semiclassical trajectories
(\ref{eq33a})-(\ref{eq35}) become complex and $\zeta$ and $\eta$ are
independent variables. Using the inverse transformation
\begin{eqnarray}
\vartheta&=&\arccos\left[\frac{1-\zeta\eta}{1+\zeta\eta}
\right] \nonumber\\
\varphi&=&\arctan\left[\frac{\zeta-\eta}{{\rm i}(\zeta+\eta)}
\right],
\label{sp1}
\end{eqnarray}
the Hamiltonian (\ref{eq28c}) takes the form
\begin{equation}
H(\zeta,\eta,t)
=\frac{1}{2}\left[ B_x(t)\frac{\zeta+\eta}{1+\zeta\eta}
-{\rm i} B_y(t)\frac{\zeta-\eta}{1+\zeta\eta}
+B_z(t)\frac{1-\zeta\eta}{1+\zeta\eta}\right],
\label{sp2}
\end{equation}
and the classical action becomes
\begin{eqnarray}
\exp\left\{{\rm i} S_{\rm{cl}}\left[\Omega(s)\right]\right\}&=&
\sqrt{\frac{(1+\zeta(0)\eta(0))(1+\zeta(t)\eta(t))}
{(1+\zeta'\eta')(1+\zeta''\eta'')} }
\left(\frac{\zeta'\eta'\zeta''\eta''}
{\zeta(0)\eta(0)\zeta(t)\eta(t)} \right)^{\frac{1}{4}}
\nonumber\\
& &\times
\exp\left\{\int_{0}^{t}{\rm d} s\left[
\frac{(1-\zeta\eta)(\dot \zeta\eta-\zeta\dot\eta )}{4\zeta\eta(1+\zeta\eta)}
-{\rm i} H(\zeta,\eta,s)\right]\right\}.
\label{sp3}
\end{eqnarray}
The time integral in the exponent may be rewritten as
\begin{eqnarray}
&&\int_{0}^{t}{\rm d} s\left[
\frac{(1-\zeta\eta)(\dot \zeta\eta-\zeta\dot\eta )}{4\zeta\eta(1+\zeta\eta)}
- {\rm i} H(\zeta,\eta,s)\right]\nonumber\\
&&= {\rm i}\int_{0}^{t}{\rm d} s\left[
\frac{{\rm i}}{2}\left(\frac{\dot \zeta\eta-\zeta\dot \eta }{1+\zeta\eta}
-\frac{\partial}{\partial s}\log\left[\frac{\zeta}{\eta}
\right] \right)
-H(\zeta,\eta,s)\right].
\nonumber
\end{eqnarray}
and the jump conditions (\ref{eq50}) and (\ref{eq55}) transform into
the simple boundary conditions
\begin{eqnarray}
\zeta(0)&=&\zeta'\nonumber\\
\eta(t)&=&\eta''.
\label{sp6}
\end{eqnarray}
Thus, we obtain from equation (\ref{sp3})
\begin{eqnarray}
\exp\left\{{\rm i} S_{\rm{cl}}[\Omega(s)]\right\}&=&
\sqrt{\frac{(1+\zeta'\eta(0))(1+\zeta(t)\eta'')}
{(1+\zeta'\eta')(1+\zeta''\eta'')} }
\left(\frac{\zeta'' \eta'}{\zeta' \eta''}\right)^\frac{1}{4}
\nonumber\\
&&\times
\exp\left\{ {\rm i}\int_{0}^{t}{\rm d} s
\left[\frac{{\rm i}}{2}\frac{\dot \zeta\eta-\zeta\dot \eta}{1+\zeta\eta}
-H(\zeta,\eta,s)\right]\right\}.
\label{sp7}
\end{eqnarray}
The classical equations of motion (\ref{eq45}) read in terms of the
new variables
\begin{eqnarray}
\dot \zeta &=&- {\rm i}(1+\zeta\eta)^2\frac{\partial H}{\partial \eta}
\nonumber\\
\dot \eta &=&{\rm i}(1+\zeta\eta)^2\frac{\partial H}{\partial \zeta}.
\label{sp8}
\end{eqnarray}
These equations coincide with the Euler-Lagrange equations of the action
\begin{equation}
S'[\zeta(s),\eta(s)]= \int_{0}^{t}{\rm d} s\left[
\frac{{\rm i}}{2} \frac{\dot \zeta\eta-\zeta\dot \eta }{1+\zeta\eta}
-H(\zeta,\eta,s)\right].
\label{sp8b}
\end{equation}
This action and the associated classical equations of motion have been
studied by several authors \cite{kuratsuji,keski,fukui,funahashi1,ercolessi}.
It is important to note that the
action (\ref{sp8b}) evaluated along the trajectories solving
equations (\ref{sp8}) with boundary conditions (\ref{sp6}) does not
yield the exact quantum mechanical propagator through a relation of
the form (\ref{eq66x}).
Using the explicit form (\ref{sp2}) of the
Hamlitonian the equations (\ref{sp8}) decouple and read explicitly
\begin{eqnarray}
\dot \zeta
&=&-\frac{{\rm i}}{2}B_x(1-\zeta^2)+\frac{1}{2}B_y(1+\zeta^2)+{\rm i} B_z\zeta
\nonumber\\
\dot \eta
&=&\frac{{\rm i}}{2}B_x(1-\eta^2)+\frac{1}{2}B_y(1+\eta^2)-{\rm i} B_z\eta.
\label{sp9}
\end{eqnarray}
Since the solution has to satisfy the conditions (\ref{sp6}), we see
that the boundary-value problem is now reduced to an initial- or final-value
problem. Moreover, the two equations of motion are complex conjugate.
Inserting the equations of motion (\ref{sp9}) into the exponent of
equation (\ref{sp7}), the time integral simplifies and we finally obtain
\begin{eqnarray}
&&\exp\left\{{\rm i} S_{\rm{cl}}[\Omega(s)]\right\}=
\sqrt{\frac{(1+\zeta'\eta(0))(1+\zeta(t)\eta'')}
{(1+\zeta'\eta')(1+\zeta''\eta'')} }
\left(\frac{\zeta'' \eta'}{\zeta'\eta''}\right)^\frac{1}{4}
\nonumber\\
&&\times
\exp\left\{-\frac{{\rm i}}{4}\int_{0}^{t}{\rm d} s
\biggl[B_x\Bigl(\zeta+\eta\Bigr)-
{\rm i} B_y\Bigl(\zeta-\eta\Bigr)+2B_z\biggr]\right\}.
\label{sp10}
\end{eqnarray}
To illustrate the theory we apply it to two specific models.
As a first example we treat the propagator of a two-state system with
Hamiltonian
\begin{equation}
H=\Delta S_x + \epsilon S_z,
\label{ex0}
\end{equation}
which describes a variety of systems. The Hamiltonian
(\ref{ex0}) corresponds to a spin-$\frac{1}{2}$ in a
time-independent magnetic field $\vec B=(\Delta,0,\epsilon)$.
With the method presented above, we express this Hamiltonian as
\begin{equation}
H(\zeta,\eta)=\frac{\Delta}{2}\frac{\zeta+\eta}{1+\zeta\eta}
+\frac{\epsilon}{2}\frac{1-\zeta\eta}{1+\zeta\eta}.
\label{ex1}
\end{equation}
In accordance with the boundary conditions (\ref{sp6}), the equations of
motion (\ref{sp9}) are solved by
\begin{eqnarray}
\zeta(s)&=&-\frac{\epsilon}{\Delta}
-{\rm i}\frac{\omega}{\Delta}\tan\left\{\omega \frac{s}{2}
+\arctan\left[\frac{{\rm i}(\epsilon+\Delta\zeta')}{\omega}\right]
\right\}
\nonumber\\
\eta(s)&=&-\frac{\epsilon}{\Delta}
-{\rm i}\frac{\omega}{\Delta}\tan\left\{\omega\frac{t-s}{2}
+\arctan\left[\frac{{\rm i}(\epsilon+\Delta\eta'')}{\omega}\right]
\right\},
\label{ex2}
\end{eqnarray}
where $\omega=\sqrt{\Delta^2+\epsilon^2}$. The unspecified boundary
values may be written as
\begin{equation}
\zeta(t)=\frac{\omega \zeta'\cos\left(\frac{\omega t}{2}\right)
+{\rm i}(\epsilon\zeta'-\Delta)\sin\left(\frac{\omega t}{2}\right)}
{\omega\cos\left(\frac{\omega t}{2}\right)
-{\rm i}(\Delta\zeta'+\epsilon)\sin\left(\frac{\omega t}{2}\right)},
\label{ex3}
\end{equation}
and\begin{equation}
\eta(0)=\frac{\omega\eta''\cos\left(\frac{\omega t}{2}\right)
+{\rm i}(\epsilon\eta''-\Delta)\sin\left(\frac{\omega t}{2}\right)}
{\omega\cos\left(\frac{\omega t}{2}\right)
-{\rm i}(\Delta\eta''+\epsilon)\sin\left(\frac{\omega t}{2}\right)}.
\label{ex4}
\end{equation}
Now, the time integral in equation (\ref{sp10}) can be readily solved
\begin{eqnarray}
&&\exp\left\{ -\frac{{\rm i}}{2}\int_{0}^{t}{\rm d} s
\left[\frac{1}{2}\Delta\left(\zeta(s)+\eta(s)\right)+\epsilon\right]\right\}=
\nonumber\\
&& \frac{1}{\omega}\left[\omega
\cos\left(\frac{\omega t}{2}\right)
-{\rm i}(\Delta\zeta'+\epsilon)
\sin\left(\frac{\omega t}{2}\right)\right]^{\frac{1}{2}}
\left[\omega
\cos\left(\frac{\omega t}{2}\right)
-{\rm i}(\Delta\eta''+\epsilon)\sin\left(\frac{\omega t}{2}\right)\right]^{\frac{1}{2}}.
\label{ex5}
\end{eqnarray}
Combining these relations we obtain from equation (\ref{sp10})
\begin{eqnarray}
\exp\left\{{\rm i} S_{\rm{cl}}[\Omega(s)]\right\}
&=&
\frac{\left(\frac{\zeta'' \eta'}{\zeta' \eta''}\right)^\frac{1}{4} }
{\sqrt{(1+\zeta'\eta')(1+\zeta''\eta'')}}\biggl[
\left(1+\zeta'\eta''\right)\cos\left(\frac{\omega t}{2}\right) \biggr.
\nonumber\\
&&-\biggl. \frac{{\rm i}\left(\epsilon
-\epsilon\zeta'\eta''+\Delta\zeta'+\Delta\eta''\right)}
{\omega}\sin\left(\frac{\omega t}{2}\right) \biggr].
\label{ex6}
\end{eqnarray}
With the inverse transformation the semiclassical propagator
takes the form of the right hand side of equation (\ref{su8}) with
\begin{equation}
a(t)=\cos\left(\frac{\omega t}{2}\right)-
\frac{{\rm i}\epsilon}{\omega}\sin\left(\frac{\omega t}{2}\right),
\label{ex8}
\end{equation}
and
\begin{equation}
b(t)=-\frac{{\rm i}\Delta}{\omega}
\sin\left(\frac{\omega t}{2}\right),
\label{ex9}
\end{equation}
which coincides with the exact quantum mechanical result.
As a second example, we consider the Landau-Zener problem \cite{shore}
\begin{equation}
H=\omega S_x -\gamma^2t S_z,
\label{ex10a}
\end{equation}
which corresponds to a spin-$\frac{1}{2}$ in the
time-dependent magnetic field $\vec B=(\omega,0,-\gamma^2 t)$.
The Hamiltonian now reads
\begin{equation}
H(\zeta,\eta,t)=\frac{\omega}{2}\frac{\zeta+\eta}{1+\zeta\eta}
-\frac{\gamma^2 t}{2}\frac{1-\zeta\eta}{1+\zeta\eta}.
\label{ex10}
\end{equation}
The equations of motion (\ref{sp9}) are of Riccati form, and for the
present model the transformation
\begin{equation}
\zeta(s)=\frac{a^{*}(s)\zeta -b^*(s)}{b(s)\zeta+a(s)}
\label{ex10b}
\end{equation}
leads to Weber equations for $a(s)$ and $b(s)$ \cite{whittaker} which
are solved in terms of confluent hypergeometric functions
$\Phi(\alpha,\beta,z)$ \cite{gradshteyn}. Accordingly, the solutions
read
\begin{eqnarray}
\zeta(s)&=\frac{D(s)\zeta' + C(s)}{B(s)\zeta'+A(s)}\nonumber\\
\eta(s)&=\frac{
\left[A(t)A(s)-C(t)B(s)\right]\eta'' -\left[A(t)B(s)-B(t)A(s)\right] }
{\left[C(t)D(s)-D(t)C(s)\right]\eta''+\left[A(t)D(s)-B(t)C(s)\right]},
\label{ex11}
\end{eqnarray}
where
\begin{eqnarray}
A(s)&=&\Phi\left(-\frac{{\rm i}}{8}\frac{\omega^2}{\gamma^2},
\frac{1}{2},-\frac{{\rm i}}{2}\gamma^2 s^2\right)
\nonumber\\
B(s)&=&-\frac{{\rm i}}{2}\omega s \,
\Phi\left(-\frac{{\rm i}}{8}\frac{\omega^2}{\gamma^2}+
\frac{1}{2},\frac{3}{2},-\frac{{\rm i}}{2}\gamma^2 s^2\right)
\nonumber\\
C(s)&=&-\frac{{\rm i}}{2}\omega s\,
\Phi\left(-\frac{{\rm i}}{8}\frac{\omega^2}{\gamma^2}+1,
\frac{3}{2},-\frac{{\rm i}}{2}\gamma^2 s^2\right)
\nonumber\\
D(s)&=&\Phi\left(-\frac{{\rm i}}{8}\frac{\omega^2}{\gamma^2}+
\frac{1}{2},\frac{1}{2},-\frac{{\rm i}}{2}\gamma^2 s^2\right).
\label{ex12}
\end{eqnarray}
Now, the unspecified boundary values become
\begin{eqnarray}
\zeta(t)&=\frac{D(t)\zeta' +C(t)}{B(t)\zeta'+A(t)}
\nonumber\\
\eta(0)&=\frac{D(t)\eta''+B(t) }{C(t)\eta''+A(t)}.
\label{ex13}
\end{eqnarray}
To solve the time integral in equation (\ref{sp10}) we make
use of analytic properties of $\Phi(\alpha,\beta,z)$
\cite{gradshteyn} yielding
\begin{eqnarray}
\frac{{\rm d}}{{\rm d} s}A(s)&=&-\frac{{\rm i}}{2}\omega C(s)
\nonumber\\
\frac{{\rm d}}{{\rm d} s}B(s)&=&-\frac{{\rm i}}{2}\omega D(s)
\nonumber\\
\frac{{\rm d}}{{\rm d} s}C(s)&=&-\frac{{\rm i}}{2}\omega A(s) -{\rm i} \gamma^2 s C(s)
\nonumber\\
\frac{{\rm d}}{{\rm d} s}D(s)&=&-\frac{{\rm i}}{2}\omega B(s)-{\rm i} \gamma^2 s D(s).
\label{ex14}
\end{eqnarray}
We then obtain
\begin{equation}
\exp\left\{ -\frac{{\rm i}}{4}\omega\int_{0}^{t}{\rm d} s
\left[\zeta(s)+\eta(s)\right]\right\}=
\sqrt{B(t)\zeta'+A(s)}\sqrt{C(t)\eta''+D(t)}.
\label{ex15}
\end{equation}
Inserting the results into (\ref{sp10}) and expressing the
final and initial states again in terms of angles, the semiclassical
propagator takes the form (\ref{su8}) where
\begin{equation}
a(t)=\exp\left\{\frac{{\rm i}}{4} \gamma^2 t^2\right\}A(t)
\label{ex17}
\end{equation}
and
\begin{equation}
b(t)=\exp\left\{\frac{{\rm i}}{4} \gamma^2 t^2\right\}B(t).
\label{ex18}
\end{equation}
This is again the exact quantum mechanical result.
\section{Conclusions}
We have analyzed the spin coherent state path integral for a
spin-$\frac{1}{2}$ in a magnetic field of arbitrary time
dependence. To obtain a path integral that may be evaluated with
conventional methods, we have introduced a Wiener
regularization. Then, the semiclassical approximation was shown to be
well defined leading to a classical trajectory with jumps at the
endpoints. The action of this trajectory determines the exact quantum
mechanical propagator. Hence, the dominant stationary phase
approximation without fluctuations was shown to become exact for a
spin-$\frac{1}{2}$ system in an arbitrary time-dependent magnetic
field. A non-linear transformation related to the
stereographic projection from the south pole onto the equatorial
plane was found to simplify the explicit determination of the minimal
action trajectory. The method was illustrated by applying it to two
specific models.
The theory presented has a straightforward extension to spin systems
with quantum numbers $s>\frac{1}{2}$ provided the Hamiltonian remains
of the form (\ref{eq22c}) which is, however, no longer the most
general spin Hamiltonian in this case. For other Hamiltonians, for
$s>\frac{1}{2}$, the dominant stationary phase approximation cannot be
expected to remain exact. An interesting extension of the present work
would be the investigation of the semiclassical dynamics of a spin
coupled to other degrees of freedom, e.g., boson
modes. With a proper c-number representation of these modes, the
problem can be described as a spin-$\frac{1}{2}$ in a fluctuating
field, and the method presented here can be applied. This will be
studied in future work.
\acknowledgments
The authors would like to thank Joachim Ankerhold, Andreas Lucke, Phil
Pechukas, Gerhard Stock and Simone Warzel for valuable
discussions. One of us (A~A) is grateful to the Department of
Chemistry of Columbia University, New York, for hospitality during an
extended stay. This work was supported by the Deutsche
For\-schungs\-ge\-mein\-schaft (Bonn) through the Schwer\-punkt\-pro\-gramm
``Zeit\-ab\-h\"angi\-ge Ph\"anomene und Methoden in Quan\-ten\-sys\-te\-men
der Phy\-sik und Che\-mie''. Additional support was
provided by the Deutscher Aka\-de\-mi\-scher Aus\-tausch\-dienst (DAAD).
|
\section{Introduction and main results}
We consider, for arbitrary but fixed $n\in\mathbf{N,}$ $n\geq1,$ the
reaction--diffusion problem
\begin{align}
a_{t} & =a_{xx}-\frac{1}{2}(4ab)^{n}~, \label{system1} \\
b_{t} & =b_{xx}-\frac{1}{2}(4ab)^{n}~, \label{system2}
\end{align}
for $x\in\mathbf{R},$ $t\geq\tau\geq0,$ with initial conditions $a(x,\tau
)=a_{0}(x),$ $b(x,\tau)=b_{0}(x),$ satisfying
\begin{align}
\lim_{x\rightarrow-\infty}a_{0}(x) & =1~, \notag \\
\lim_{x\rightarrow+\infty}b_{0}(x) & =1~, \label{lim002} \\
\lim_{x\rightarrow+\infty}a_{0}(x) & =\lim_{x\rightarrow-\infty}b_{0}(x)=0~.
\notag
\end{align}
The choice of the initial time $t=\tau,$ and a class of initial conditions $%
a_{0},$ $b_{0}$ will be described later on, but for the purpose of this
introduction it is useful to have in mind the ``natural'' case: $\tau=0,$ $%
a_{0}(x)=1$ for $x<0,$ $a_{0}(x)=0$ for $x>0,$ and $b_{0}(x)=1$ for $x>0,$ $%
b_{0}(x)=0$ for $x<0.$
This initial value problem models the time evolution of a chemical system of
two (initially separated) substances $A$ and $B,$ that diffuse in some
substratum and react according to $nA+nB\rightarrow C,$ with a substance $C$
that is supposed not to participate in the reaction anymore. The model is a
mean--field description of such a situation where the functions $a$ and $b$
represent the densities of the substances $A$ and $B.$ For more details see
\cite{Droz}.
Equations (\ref{system1}) and (\ref{system2}) are best studied in terms of
the sum
\begin{equation}
v=a+b~, \label{defv}
\end{equation}
and the difference
\begin{equation}
u=a-b~, \label{defu}
\end{equation}
which satisfy the equations
\begin{align}
u_{t} & =u_{xx}~, \label{equ} \\
v_{t} & =v_{xx}-(v^{2}-u^{2})^{n}~, \label{eqv}
\end{align}
with initial conditions $v_{0}$ and $u_{0}$ (at time $t=\tau)$ that satisfy
\begin{align}
\lim_{x\rightarrow-\infty}u_{0}(x) & =1~, \notag \\
\lim_{x\rightarrow+\infty}u_{0}(x) & =-1~, \label{limitu}
\end{align}
and
\begin{equation}
\lim_{x\rightarrow\pm\infty}v_{0}(x)=1~. \label{limitv}
\end{equation}
For initial conditions $a_{0},$ $b_{0}$ with
\begin{equation}
a_{0}(x)=b_{0}(-x)~, \label{sym001}
\end{equation}
the functions $v_{0}$ and $u_{0}$ are even and odd, respectively, and the
equations (\ref{equ}), (\ref{eqv}) preserve this symmetry. Furthermore, for
the special initial condition
\begin{equation}
u(x,\tau)=-\mu_{1}(x/\sqrt{\tau})~, \label{initsym}
\end{equation}
with $\mu_{1}$ defined by the equation
\begin{equation}
\mu_{1}(y)=\mathrm{erf}(\frac{y}{2})\equiv\frac{2}{\sqrt{\pi}}%
\int_{0}^{y/2}e^{-\sigma^{2}}d\sigma~, \label{defmu}
\end{equation}
equation (\ref{equ}) has the explicit solution
\begin{equation}
u(x,t)=-\mu_{1}(x/\sqrt{t})~. \label{uxt}
\end{equation}
We note that the initial condition (\ref{initsym}) for $u$ (at time $t=\tau$%
) is simply the solution of equation (\ref{equ}) with the ``natural''
initial condition, $u(x,0)=1$ for $x<0,$ $u(x,0)=-1$ for $x>0,$ evaluated at
$t=\tau.$ To keep this paper as simple as possible we now restrict the
discussion to this case, i.e., we consider from now on equation (\ref{eqv})
with initial conditions satisfying (\ref{limitv}), and $u$ given by (\ref
{uxt}). We note, however, that more general (asymmetric) initial conditions
for $u$ could be treated as well. This would lead to corrections to $u$ of
the order $\mathcal{O}(1/t),$ and such corrections do not change in any way
the discussion of the equation for $v$ that follows.
The reaction--diffusion problems considered here develop, in addition to the
built--in diffusive length scale $\mathcal{O}(\sqrt{t}),$ an additional
shorter length scale, on which the reaction takes place. The function $F,$%
\begin{equation}
F=\frac{1}{2}(4ab)^{n}\equiv \frac{1}{2}(v^{2}-u^{2})^{n}~,
\label{reaction001}
\end{equation}
is called the reaction term or reaction front, and we are interested in
describing the asymptotic behavior of the function $F$ for large times. The
knowledge of this behavior is useful, since it appears to be universal, in
the sense that it is largely independent of the choice of the initial
conditions and of the details of the model under consideration. As mentioned
above, if $v_{0}$ is an even function, then $v$ and as a consequence $F$ are
even functions of $x.$ We will see that the critical point of $F$ at $x=0$
is a maximum, and that $F$ decays (rapidly) for large $x.$
Before proceeding any further we note that the factor of $4^{n-1/2}$ in (\ref
{system1}), (\ref{system2}) and (\ref{reaction001}) is just a normalization,
and has been chosen for convenience to make the equation (\ref{eqv}) for $v$
look simple. In fact, any system of the form
\begin{align*}
a_{t} & =D_{a}a_{xx}-k_{a}(ab)^{n}~, \\
b_{t} & =D_{b}b_{xx}-k_{b}(ab)^{n}~,
\end{align*}
with positive $D_{a},$ $D_{b},$ $k_{a},$ and $k_{b},$ and with initial
conditions such that
\begin{align*}
\lim_{x\rightarrow-\infty}a(x,0) & =a_{\infty}>0~, \\
\lim_{x\rightarrow\infty}b(x,0) & =b_{\infty}>0~, \\
\lim_{x\rightarrow+\infty}a(x,0) & =\lim_{x\rightarrow-\infty}b(x,0)=0~,
\end{align*}
can be reduced, by scaling space and time and the amplitudes, to the problem
\begin{align*}
a_{t} & =a_{xx}-\frac{1}{2}(4ab)^{n}~, \\
b_{t} & =Db_{xx}-\frac{1}{2}(4ab)^{n}~,
\end{align*}
with $D>0,$ and with initial conditions such that
\begin{align*}
\lim_{x\rightarrow-\infty}a(x,0) & =1~, \\
\lim_{x\rightarrow\infty}b(x,0) & =\beta>0~, \\
\lim_{x\rightarrow+\infty}a(x,0) & =\lim_{x\rightarrow-\infty}b(x,0)=0~.
\end{align*}
In this paper we have limited the discussion to the case $\beta=1$ and $D=1.
$ The case $\beta\neq1$ leads to a moving reaction front. A change of
coordinates to a moving frame complicates the analysis, but the problem
could still be treated with the methods presented here. Choosing $D=1$ makes
the equations mathematically simpler. As a consequence, as we have seen, the
two equations for $a$ and $b$ can be reduced to just one equation for the
sum $v=a+b,$ since the equation for the difference $u=a-b$ can be solved
explicitly. Even though we do not expect the asymptotic behavior of the
solution to change in any relevant way if $D\neq1$, the strategy of proof
would have to be changed considerably, since the equations can not be
decoupled anymore in that case.
Before we state our results, we briefly discuss the expected dependence of
the results on the parameter $n.$
The case $n=1$ has been studied in detail in \cite{Wittwer}, where it is
proved that in this case the reaction term (\ref{reaction001}) satisfies,
for all $z\in\mathbf{R},$
\begin{equation*}
\lim_{t\rightarrow\infty}t^{2\gamma}F(t^{\alpha}z,t)=\rho(\left| z\right| )~,
\end{equation*}
where $\alpha=1/6,$ and $\gamma=1/3,$ and where $\rho\colon\mathbf{R}_{+}%
\mathbf{\rightarrow R}_{+}$ is a smooth function that decays like $\exp(-%
\mathrm{const.}z^{3/2})$ for large values of $z.$ It follows furthermore
from the results in \cite{Wittwer} that the function $F$ is very small on
the diffusive scale in the sense that for $n=1,$ $y\neq0,$ and all $p\geq0, $%
\begin{equation}
\lim_{t\rightarrow\infty}t^{p}F(\sqrt{t}y,t)=0~. \label{fdiffuse}
\end{equation}
The smallness of $F$ on the diffusive scale is easily understood by
realizing that, for $n=1$ and for positive values of $x$ on the diffusive
scale, i.e., for $x/\sqrt{t}>>1,$ equation (\ref{system1}) essentially
reduces to
\begin{equation}
a_{t}=a_{xx}-\lambda a~, \label{a001}
\end{equation}
with $\lambda>0.$ Therefore, the function $a$ decays exponentially fast to
zero on this scale, and similarly for $b$ for negative values of $x.$
For $n>1,$ however, equation (\ref{system1}) reduces, for $x/\sqrt{t}>>1,$
essentially to
\begin{equation}
a_{t}=a_{xx}-\lambda a^{n}~, \label{a002}
\end{equation}
with $\lambda>0.$ The solution of (\ref{a002}) has an asymptotic behavior
that is radically different from the solution of (\ref{a001}). In
particular, for $n=2$, the solution may even blow up in finite time if $a$
is not a positive function. Note that, for $n$ odd, the nonlinear term in (%
\ref{a002}) is always a ``friction term'', independent of the sign of $a,$
and the case of $n$ odd will therefore turn out to be easier to treat than
the case of $n$ even. It is well known \cite{kupi} that for $n>3$ and small
bounded integrable initial conditions, the nonlinearity in (\ref{a002})
becomes irrelevant for large times in the sense that the solution converges
to a multiple of $\exp (-x^{2}/4t)/\sqrt{t},$ which solves the linear
equation $a_{t}=a_{xx}.$ We would therefore expect that, for $n>3,$ the
function $F$ is of the order $\mathcal{O}(t^{-n/2})$ on the diffusive scale.
This turns out to be wrong. As we will prove below, $F$ is of the order $%
\mathcal{O}(t^{-n/(n-1)})$ for $n>3,$ because $F$ converges on this scale
pointwise to a function that is not integrable at the origin. This
corresponds to a solution of (\ref{a002}) for which the nonlinear term is a
marginal perturbation, i.e., a solution with an amplitude of the order $%
\mathcal{O}(t^{-1/(n-1)})$. We will see that one can take advantage of this
fact, and a diffusive stability bound will be good enough to prove
convergence of $F$ to its limit, but as a consequence, our results will be
limited to the case $n>3.$ The cases $n=2$ and $n=3$ are special and will
not be discussed any further.
The following theorem is our main result.
\begin{theorem}
\label{main}For arbitrary but fixed $n\in\mathbf{N,}$ $n\geq4,$ there exist $%
\tau>0,$ functions $\mu_{1,}$ $\mu_{2},$ $\varphi_{1},$ $\varphi_{2},$ and a
class of initial conditions (specified at $t=\tau),$ such that (\ref{eqv})
has a unique solution $v$ that satisfies for all $t\geq\tau$ the bound
\begin{equation}
\left| v(x,t)-v_{\infty}(x,t)\right| <\frac{\mathrm{const.}}{t^{4\gamma}}~,
\label{bound}
\end{equation}
where
\begin{equation}
v_{\infty}(x,t)=\mu_{1}(\frac{\left| x\right| }{\sqrt{t}})+t^{-\varepsilon
}\mu_{2}(\frac{\left| x\right| }{\sqrt{t}})+t^{-\gamma}\varphi_{1}(\frac{%
\left| x\right| }{t^{\alpha}})+t^{-3\gamma}\varphi_{2}(\frac{\left| x\right|
}{t^{\alpha}})~, \label{vinfty}
\end{equation}
$\gamma=\frac{1}{2n+1},$ $\varepsilon=\frac{1}{n-1}$ and $\alpha=\frac{1}{2}%
-\gamma.$
\end{theorem}
\begin{remark}
This theorem is a local result, in the sense that the class of initial
conditions will be a set of functions in a (small) neighborhood of the
function $v_{\infty,0},$ $v_{\infty,0}(x)=v_{\infty}(x,\tau).$ In
particular, our methods do not allow us to show that the solution with the
``natural'' initial condition $v_{0}\equiv1$ at $t=0$ belongs to this set at
$t=\tau.$ We do expect, however, that this is the case, as has been proved
for $n=1$ in \cite{Wittwer}.
\end{remark}
\begin{remark}
We note that, if an initial condition $v_{0}$ is such that $v_{0}(x)-\left|
u(x/\sqrt{\tau})\right| <0$ for a certain $x,$ then $a_{0}(x)<0$, if $x>0,$
or $b_{0}(x)<0$ if $x<0.$ A priori, we do not need to consider such initial
conditions, since in our model $a$ and $b$ represent particle densities, and
the solutions $a$ and $b$ are positive if the initial conditions $a_{0}$ and
$b_{0}$ are positive. As we will see, for $n\geq4,$ it will not be necessary
to impose that $a_{0}$ and $b_{0}$ be positive everywhere, and it will
neither be necessary to impose that $v_{0}=a_{0}+b_{0}$ be an even function.
\end{remark}
As we will see, the functions $\varphi _{1}$ and $\varphi _{2}$ are small on
the diffusive scale, i.e., for $x\approx \sqrt{t}y,$ $y\neq 0,$ and $t$
large,
\begin{equation}
v_{\infty }(\sqrt{t}y,t)=\mu _{1}(\left| y\right| )+t^{-\varepsilon }\mu
_{2}(\left| y\right| )+\mathcal{O}(t^{-2\varepsilon ^{\prime }})~,
\label{asym001}
\end{equation}
where $\varepsilon ^{\prime }=\varepsilon $ if $n>5,$ and $2\gamma
<\varepsilon ^{\prime }<\varepsilon $ if $n=4,$ $5.$ Using the definition (%
\ref{defv}), (\ref{defu}) for $v$ and $u,$ we therefore find that for $y>0$
and $t$ large,
\begin{equation*}
a(\sqrt{t}y,t)=\frac{1}{2}t^{-\varepsilon }\mu _{2}(y)+\mathcal{O}%
(t^{-2\varepsilon ^{\prime }})~,
\end{equation*}
and similarly for $b,$ for $y<0.$ In contrast to the case $n=1,$ where only
exponentially few particles reach the diffusive scale, the amount of
particles decays only slowly for $n>3.$ Our results imply that, for large
times, the density of the remaining particles is given by the function $\mu
_{2},$ i.e., it is independent of the initial conditions.
As a corollary to Theorem \ref{main} we get a precise description of the
reaction front $F$ on the reactive and the diffusive scale. This description
will be given in Section 4, once we have defined the functions $\mu _{1},$ $%
\mu _{2},$ $\varphi _{1}$ and $\varphi _{2}$ in Section 3. In Section 2 we
explain our strategy for proving Theorem \ref{main}. This strategy is
implemented in Section 5 and Section 6. The Appendix contains the proof of
the existence of the functions $\mu _{2},$ $\varphi _{1}$ and $\varphi _{2}$.
\section{Strategy of the Proof}
Consider functions $v$ of the form
\begin{equation}
v(x,t)=v_{\infty}(x,t)+\psi(x,t)~, \label{ansatz}
\end{equation}
with $v_{\infty}$ as in Theorem \ref{main}, and $\psi(x,\tau)=\psi_{0}(x),$
for some $\tau>>1,$ with $\psi_{0}\in L_{1}\cap L_{\infty}.$ Substituting (%
\ref{ansatz}) into (\ref{eqv}) leads to an equation for the function $\psi$
of the form
\begin{equation}
\dot{\psi}=\psi^{\prime\prime}-V\psi-I-T(\psi)~, \label{eqpsi}
\end{equation}
for certain functions $V$ and $I,$ and for $T$ some nonlinear map. We will
show that if $v_{\infty}$ is defined correctly, $\tau$ large enough and $%
\psi_{0}$ small enough, then $V$ can be chosen positive and $T$ will be
small, so that the solution of equation (\ref{eqpsi}) will be bounded for
large times by the corresponding solution of the inhomogeneous heat equation
$\dot{\psi}=\psi^{\prime\prime}-I.$ We will find that, with the right choice
of $v_{\infty},$%
\begin{equation}
\int dx~\left| I(\sqrt{t}x,t)\right| \leq\mathrm{const.}\text{ }%
t^{-1-4\gamma}~, \label{boundi}
\end{equation}
from which the bound (\ref{bound}) will follow. We note that $4\gamma<\frac
{1}{2}$ for $n\geq4>\frac{7}{2},$ so that contributions of initial
conditions will become irrelevant for large times, i.e., the solution $\psi$
becomes what is called ``slaved to the inhomogeneous term''.
\section{Asymptotic Expansion}
In order to implement the strategy outlined in Section 2, we need a function
$v_{\infty }$ that approximates the solution $v$ for large times
sufficiently well, uniformly in $x.$ Since we would like to control the time
evolution of equation (\ref{eqpsi}) on $L_{1}\cap L_{\infty },$ this
function $v_{\infty }$ needs to satisfy $\lim_{x\rightarrow \pm \infty
}v_{\infty }(x,t)=1$ in order for $v$ to satisfy the boundary conditions (%
\ref{limitv}). Furthermore, the inhomogeneous term $I$ in equation (\ref
{eqpsi}) contains second derivatives of $v_{\infty },$ and the function $I$
can therefore only be in $L_{1}\cap L_{\infty }$ if $v_{\infty }$ is at
least twice differentiable. We now construct a function $v_{\infty }$
satisfying these requirements through a two length--scale asymptotic
expansion.
To simplify the notation later on we use the convention that, unless stated
otherwise,
\begin{equation}
y\equiv\frac{x}{\sqrt{t}}~, \label{defy}
\end{equation}
and
\begin{equation}
z\equiv\frac{x}{t^{\alpha}}~, \label{defz}
\end{equation}
and we will refer to $y$ as the diffusive length scale and to $z$ as the
reactive length scale.
The function $v_{\infty}$ is given by the first and second order terms of a
so called ``matched asymptotic expansion''. The ``matched'' refers to the
fact that such an expansion contains functions that can not be associated
uniquely with one of the length scales and can therefore be used to
``match'' the behavior at large distances of the shorter scale with the
behavior at short distances of the larger scale. Let
\begin{equation}
\delta=\frac{n+2}{n-1}~, \label{defdelta007}
\end{equation}
and let $\gamma,$ $\alpha,$ and $\varepsilon$ be as in Theorem \ref{main}.
Then, the functions $y\equiv t^{-\gamma}z,$ $t^{-\varepsilon}y^{-\delta}%
\equiv t^{-\gamma}z^{-\delta},$ $t^{-\varepsilon}y^{2-\delta}\equiv
t^{-3\gamma }z^{2-\delta}$ and $t^{-2\varepsilon}y^{-2\delta+1}\equiv
t^{-3\gamma }z^{-2\delta+1}$ are of this form and will naturally show up in
the function $v_{\infty}.$ As a consequence, the representation (\ref{vinfty}%
) for $v_{\infty}$ is not unique. If we choose (as we will) to compute the
expansion for $v_{\infty}$ in the order of decreasing amplitudes, i.e., if
we first compute the term of order $\mathcal{O}(t^{-\gamma}),$ then the term
of order $\mathcal{O}(t^{-\varepsilon}),$ and finally the term of order $%
\mathcal{O}(t^{-3\gamma}),$ we get a representation of $v_{\infty}$ of the
form
\begin{equation}
v_{\infty}(x,t)=\mu_{1}(\left| y\right| )+t^{-\gamma}\eta(\left| z\right|
)+t^{-\varepsilon}\left( \mu_{2}(\left| y\right| )-\lambda\left| y\right|
^{-\delta}\right) +t^{-3\gamma}\varphi_{2}(\left| z\right| )~,
\label{repres1}
\end{equation}
where
\begin{equation}
\eta(z)=\varphi_{1}(z)+\lambda z^{-\delta}~, \label{etaphi1}
\end{equation}
with $\lambda$ a certain constant to be determined later.
We note that, by definition, $v_{\infty}$ is a symmetric function, and it is
therefore sufficient to consider positive values of $x$ if we choose
appropriate boundary conditions at $x=0$ to ensure regularity.
Finally, since we will need to describe the asymptotic behavior of various
functions near zero and infinity, we introduce the following notation. Let $%
f $ be a continuous function from $\mathbf{R}_{+}$ to $\mathbf{R,}$ $k$ a
positive integer and $p_{1}<p_{2}<\dots<p_{k}$ real numbers. Then, we say
that
\begin{equation*}
f(x)=\sum_{i=1}^{k}f_{i}~x^{p_{i}}+\dots
\end{equation*}
near $x=0,$ if
\begin{equation*}
\lim_{x\rightarrow0^{+}}\frac{1}{x^{p_{k}}}\left|
f(x)-\sum_{i=1}^{k}f_{i}~x^{p_{i}}\right| =0~,
\end{equation*}
and we say that
\begin{equation*}
f(x)=\sum_{i=1}^{k}f_{i}~x^{-p_{i}}+\dots
\end{equation*}
near $x=\infty,$ if
\begin{equation*}
\lim_{x\rightarrow\infty}x^{p_{k}}\left|
f(x)-\sum_{i=1}^{k}f_{i}~x^{-p_{i}}\right| =0~.
\end{equation*}
\subsection{Equation for $\protect\mu_{1}$}
To lowest order the function $v$ is asymptotic to $\mu_{1}(\left| y\right| )
$, with $\mu_{1}$ as defined in (\ref{defmu}). We note that $\mu_{1}$ has
near $y=0$ the expansion
\begin{equation}
\mu_{1}(y)=\kappa y+\kappa_{3}y^{3}+\dots~, \label{mu1nearzero}
\end{equation}
where $\kappa=\frac{1}{\sqrt{\pi}}$ and $\kappa_{3}=-\frac{1}{12}\kappa$.
Furthermore, $\lim_{y\rightarrow\infty}\mu_{1}(y)=1.$
\subsection{Equation for $\protect\varphi_{1}$}
We make the ansatz $v(x,t)=\mu_{1}(\left| y\right| )+t^{-\gamma}\eta(\left|
z\right| )$ which we substitute into equation (\ref{eqv}). We multiply the
resulting equation with $t^{\gamma+2\alpha},$ and take then the limit $%
t\rightarrow\infty,$ keeping $z$ fixed. This leads to the differential
equation
\begin{equation}
\eta^{\prime\prime}=\left( 2\kappa z\eta+\eta^{2}\right) ^{n}~,
\label{eqeta}
\end{equation}
where $z$ is now considered a variable in $\mathbf{R}_{+}.$ Since $%
\lim_{y\rightarrow\infty}\mu_{1}(y)=1,$ the correct boundary condition for $%
\eta$ at infinity is
\begin{equation}
\lim_{z\rightarrow\infty}\eta(z)=0~, \label{etainfinity}
\end{equation}
and at $z=0$ we impose
\begin{equation}
\eta^{\prime}(0)=-\kappa~, \label{etazero}
\end{equation}
which makes the function $\mu_{1}(\left| y\right| )+t^{-\gamma}\eta(\left|
z\right| )$ twice differentiable at $x=0,$ since $\mu_{1}$ and $\eta$ are
twice differentiable at zero and $\partial_{x}\left( \mu_{1}(\left| y\right|
)+t^{-\gamma}\eta(\left| z\right| )\right) (0)=0.$ A proof of the following
proposition can be found in the appendix.
\begin{proposition}
\label{theoremeta}For $n\in \mathbf{N,}$ $n\geq 2,$ there exists a unique
function $\eta \colon \mathbf{R}_{+}\!\rightarrow \mathbf{R}$ that satisfies
equation (\ref{eqeta}) with the boundary conditions (\ref{etainfinity}) and (%
\ref{etazero}). The function $\eta $ is positive, and has near $z=0$ the
expansion
\begin{equation*}
\eta (z)=\eta _{0}-\kappa z+\eta _{2}z^{2}-\eta _{4}z^{4}+\dots ~,
\end{equation*}
with positive coefficients $\eta _{0},$ $\eta _{2}$ and $\eta _{4}.$ For $z$
large, $\eta $ is of the form
\begin{equation*}
\eta (z)=\frac{\lambda }{z^{\delta }}+\frac{\lambda _{\infty }}{z^{\delta
^{\prime }}}+\dots ~,
\end{equation*}
for a certain constant $\lambda _{\infty },$ with $\delta $ as in (\ref
{defdelta007}),
\begin{equation}
\lambda =\left( \frac{\delta (\delta +1)}{(2\kappa )^{n}}\right) ^{1/(n-1)}~,
\label{deflambda007}
\end{equation}
and
\begin{equation*}
\delta ^{\prime }=\left\{
\begin{array}{ccc}
\left( \sqrt{4n\delta (\delta +1)+1}-1\right) /2 & & 2\leq n\leq 5~, \\
& & \\
2\delta +1 & & n\geq 6~.
\end{array}
\right.
\end{equation*}
\end{proposition}
We note that $3<\delta^{\prime}\leq2\delta+1.$ The constants $\eta_{0},$ $%
\eta_{2},$ $\eta_{4}$ and $\lambda_{\infty}$ are given in the appendix. The
function $\varphi_{1}$ is defined in (\ref{etaphi1}) in terms of $\eta.$
\subsection{Equation for $\protect\mu_{2}\label{smu2}$}
We make the ansatz $v(x,t)=\mu_{1}(\left| y\right| )+t^{-\gamma}\eta(\left|
z\right| )+t^{-\varepsilon}(\mu_{2}(\left| y\right| )-\lambda\left| y\right|
^{-\delta})$ which we substitute into equation (\ref{eqv}). We multiply the
resulting equation with $t^{1+\varepsilon},$ and take then the limit $%
t\rightarrow\infty,$ keeping $y$ fixed. Since $\eta(z)=\eta(t^{\gamma }y)$
and $\lim_{t\rightarrow\infty}t^{\varepsilon-\gamma}\eta(z)-\lambda
y^{-\delta}=0,$ this leads to the differential equation for the function $%
\mu_{2},$
\begin{equation}
\mu_{2}^{\prime\prime}+\frac{1}{2}y\mu_{2}^{\prime}+\varepsilon\mu_{2}=(2%
\mu_{1}\mu_{2})^{n}~, \label{mu2}
\end{equation}
where $y$ is now considered as a variable in $\mathbf{R}_{+}.$ At $y=0$ we
impose the boundary condition
\begin{equation}
\lim_{y\rightarrow0}\mu_{2}(y)y^{\delta}=\lambda~, \label{bmu2-1}
\end{equation}
which removes the leading singularity of the function $\mu_{2}(\left|
y\right| )-\lambda\left| y\right| ^{-\delta}$ at $y=0.$ As we will see, the
sub--leading singularity is proportional to $\left| y\right| ^{2-\delta},$
which is not a twice differentiable function at $y=0$ (except for $n=4$
where $\delta=2).$ This singularity will be cancelled by imposing
appropriate boundary conditions for the function $\varphi_{2}.$
The correct choice of boundary conditions for $\mu _{2}$ at infinity is
somewhat less obvious. In the appendix we show that the condition $%
\lim_{y\rightarrow \infty }\mu _{2}(y)=0$ is not sufficient to single out a
unique function $\mu _{2}.$ If $\mu _{2}$ does converge to zero at infinity,
then it is asymptotic to a solution of the equation
\begin{equation*}
\mu ^{\prime \prime }+\frac{1}{2}y\mu ^{\prime }+\varepsilon \mu =0~.
\end{equation*}
This linear equation is compatible with a (very slow) algebraic decay, $\mu
_{2}(y)\approx y^{-2\varepsilon },$ or with a modified Gaussian decay, $\mu
_{2}(y)\approx \exp (-y^{2}/4)/y^{1-2\varepsilon },$ with the algebraic
decay being the generic case. It will be essential in later sections that $%
\mu _{2}$ decays rapidly at infinity, and we therefore impose the boundary
condition
\begin{equation}
\lim_{y\rightarrow \infty }\mu _{2}(y)y^{2\varepsilon }=0~. \label{bmu2-2}
\end{equation}
A proof of the following proposition can be found in the appendix.
\begin{proposition}
\label{cmu2}For all $n\geq 4,$ there exists a unique positive function $\mu
_{2}\colon \mathbf{R}_{+}\rightarrow \mathbf{R}$ that satisfies equation (%
\ref{mu2}) with the boundary conditions (\ref{bmu2-1}) and (\ref{bmu2-2}).
For $y$ small, the function $\mu _{2}$ is of the form
\begin{equation}
\mu _{2}(y)=\lambda y^{-\delta }+\lambda _{0}y^{2-\delta }+\lambda
_{1}y^{4-\delta }+\dots ~, \label{asymmu1}
\end{equation}
with
\begin{equation}
\lambda _{0}=\frac{1}{2}\frac{\lambda }{\kappa }\frac{-2n\kappa _{3}\delta
(\delta +1)-\kappa (\delta -2\varepsilon )}{\left( n-1\right) \delta (\delta
+1)+2(2\delta -1)}>0~, \label{lambda0}
\end{equation}
with $\lambda _{1}\neq 0$ and with $\lambda $ as in Proposition \ref
{theoremeta}. For $y$ large, the function $\mu _{2}$ decays rapidly in the
sense that
\begin{equation}
\mu _{2}(y)=\exp (-\frac{y^{2}}{4})\left( \frac{C}{y^{1-2\varepsilon }}%
+\dots \right) ~, \label{asymmu2}
\end{equation}
for some constant $C>0.$
\end{proposition}
\subsection{Equation for $\protect\varphi_{2}$}
We make the ansatz $v(x,t)=\mu_{1}(\left| y\right| )+t^{-\gamma}\eta(\left|
z\right| )+t^{-\varepsilon}(\mu_{2}(\left| y\right| )-\lambda\left| y\right|
^{-\delta})+t^{-3\gamma}\varphi_{2}(\left| z\right| )$ which we substitute
into equation (\ref{eqv}). We multiply the resulting equation with $%
t^{3\gamma+2\alpha},$ and take then the limit $t\rightarrow\infty,$ keeping $%
z$ fixed. This leads to the (linear) differential equation for $\varphi
_{2}, $
\begin{equation}
\varphi_{2}^{\prime\prime}+\gamma\eta+\alpha z\eta^{\prime}+(2-\delta
)(1-\delta)\lambda_{0}z^{-\delta}=n\left( 2\kappa z\eta+\eta^{2}\right)
^{n-1}\left[ (2\kappa
z+2\eta)(\varphi_{2}+\lambda_{0}z^{2-\delta})+2\kappa_{3}z^{3}\eta\right] ~.
\label{varphi2}
\end{equation}
In order to compensate the sub--leading singular behavior of $\mu_{2}$ near $%
x=0$ we make the ansatz
\begin{equation}
\varphi_{2}(z)=-\lambda_{0}z^{2-\delta}+h(z)~, \label{newh}
\end{equation}
which we substitute into equation (\ref{varphi2}). For the function $h$ we
get the equation
\begin{equation}
h^{\prime\prime}+\gamma\eta+\alpha z\eta^{\prime}=n\left( 2\kappa z\eta
+\eta^{2}\right) ^{n-1}\left( (2\kappa z+2\eta)h+2\kappa_{3}z^{3}\eta\right)
~. \label{h}
\end{equation}
Since the function $\eta$ is regular near $z=0,$ the solution $h$ turns out
to be regular near $z=0,$ too. Therefore, the function $z\mapsto h(\left|
z\right| )$ is twice differentiable near $x=0$ if we impose at $z=0$ the
boundary condition
\begin{equation}
h^{\prime}(0)=0~. \label{bh-1}
\end{equation}
At infinity we need that $\lim_{z\rightarrow\infty}\varphi_{2}(z)=0.$ We
therefore require that
\begin{equation}
\lim_{z\rightarrow\infty}(h(z)-\lambda_{0}z^{2-\delta})=0~. \label{bh-2}
\end{equation}
A proof of the following proposition can be found in the appendix.
\begin{proposition}
\label{cfi22}For all $n\geq 4,$ there exists a unique function $h\colon
\mathbf{R}_{+}\rightarrow \mathbf{R}$ that satisfies equation (\ref{h}) with
the boundary conditions (\ref{bh-1}) and (\ref{bh-2}). Near $z=0,$ the
function $h$ is of the form
\begin{equation*}
h(z)=h_{0}+h_{2}z^{2}+\dots ~,
\end{equation*}
with certain coefficients $h_{0}$ and $h_{2},$ and for $z$ large $h$ is of
the form
\begin{equation*}
h(z)=\lambda _{0}z^{2-\delta }+\frac{\lambda ^{\prime }}{z^{\delta ^{\prime
}-2}}+\dots ~,
\end{equation*}
with $\lambda _{0}$ as defined in (\ref{lambda0}), for a certain constant $%
\lambda ^{\prime }$, and with $\delta ^{\prime }$ as defined in Proposition
\ref{theoremeta}.
\end{proposition}
\section{The Reaction Front}
Using the properties of the functions $\mu_{1},$ $\mu_{2},$ $\varphi_{1}$
and $\varphi_{2},$ we get from Theorem \ref{main} the following behavior of
the reaction front $F.$
\begin{corollary}
Let $v$ be as in Theorem \ref{main}, and $F$ as defined in (\ref{reaction001}%
). Then, for all $z\in \mathbf{R,}$%
\begin{equation*}
\lim_{t\rightarrow \infty }t^{2n\gamma }F(t^{\alpha }z,t)=\frac{1}{2}%
(2\kappa \left| z\right| \eta (\left| z\right| )+\eta (\left| z\right|
)^{2})^{n}=\frac{1}{2}\eta ^{\prime \prime }(\left| z\right| )=\left\{
\begin{array}{lll}
\eta _{2}-6\eta _{4}\left| z\right| ^{2}+\dots & \text{for} & \noindent
\left| z\right| \approx 0~, \\
& & \\
\frac{1}{2}\left( 2\lambda \kappa \right) ^{n}/\left| z\right| ^{\delta
+2}+\dots & \text{for} & \noindent \left| z\right| >>1~,
\end{array}
\right.
\end{equation*}
and for all $y\neq 0,$%
\begin{equation*}
\lim_{t\rightarrow \infty }t^{n\varepsilon }F(\sqrt{t}y,t)=\frac{1}{2}(2\mu
_{1}~\mu _{2})^{n}(\left| y\right| )=\left\{
\begin{array}{lll}
\frac{1}{2}\left( 2\lambda \kappa \right) ^{n}/\left| y\right| ^{\delta
+2}+\dots & \text{for} & \left| y\right| \approx 0~, \\
& & \\
\exp (-n\left| y\right| ^{2}/4)(2^{n-1}C^{n}/\left| y\right|
^{n(1-2\varepsilon )}+\dots ) & \text{for} & \left| y\right| >>1~.
\end{array}
\right.
\end{equation*}
Here, $\eta _{2},$ $\eta _{4}$ are as defined in Proposition \ref{theoremeta}
and $C$ is as defined in (\ref{asymmu2}).
\end{corollary}
\section{The Equation for $\protect\psi$}
In order to simplify the notation we define the function $\overline{u},$
\begin{equation*}
\overline{u}(x,t)=\mu_{1}(\left| y\right| )~,
\end{equation*}
the function $\mu_{3},$%
\begin{equation*}
\mu_{3}(y)=\mu_{2}(y)-\lambda y^{-\delta}~,
\end{equation*}
the function $\phi,$%
\begin{equation}
\phi(x,t)=t^{-\gamma}\eta(|z|)+t^{-\varepsilon}\mu_{3}(|y|)+t^{-3\gamma
}\varphi_{2}(|z|)~, \label{defphi}
\end{equation}
and the function $\phi_{1},$%
\begin{equation}
\phi_{1}(x,t)=\phi(x,t)-\kappa\frac{\left| x\right| }{\sqrt{t}}~.
\label{defphi1}
\end{equation}
The function $v_{\infty}$ in Theorem \ref{main} and in (\ref{repres1}) can
then be written as $v_{\infty}=\overline{u}+\phi.$
Let now $v=v_{\infty}+\psi.$ Then,
\begin{align*}
\left( v^{2}-u^{2}\right) ^{n} & =\left( \left( \overline{u}+\phi
+\psi\right) ^{2}-u^{2}\right) ^{n}=\left( \left( 2\overline{u}\phi
+\phi^{2}\right) +\left( 2(\overline{u}+\phi)+\psi\right) \psi\right) ^{n} \\
& =\sum_{k=0}^{n}\binom{n}{k}\left( 2\overline{u}\phi+\phi^{2}\right)
^{n-k}\left( 2(\overline{u}+\phi)+\psi\right) ^{k}\psi^{k}~.
\end{align*}
Therefore, substituting the ansatz $v=v_{\infty}+\psi$ into (\ref{eqv})
leads to the following equation for the function $\psi,$%
\begin{equation}
\dot{\psi}=\psi^{\prime\prime}-\widehat{V}\psi-I-\widehat{T}(\psi )~,
\label{equpsi1}
\end{equation}
with the function $\widehat{V},$
\begin{equation}
\widehat{V}=2n~(2\overline{u}\phi+\phi^{2})^{n-1}~(\overline{u}+\phi )~,
\label{vhat}
\end{equation}
the function $I,$%
\begin{equation}
I=-\dot{\phi}+\phi_{1}^{\prime\prime}-\left( 2\overline{u}\phi+\phi
^{2}\right) ^{n}~, \label{equi}
\end{equation}
and the map $\widehat{T},$%
\begin{equation}
\widehat{T}(\psi)=n\left( 2\overline{u}\phi+\phi^{2}\right) ^{n-1}\psi
^{2}+\sum_{k=2}^{n}\binom{n}{k}\left( 2\overline{u}\phi+\phi^{2}\right)
^{n-k}\left( 2(\overline{u}+\phi)+\psi\right) ^{k}\psi^{k}~. \label{that}
\end{equation}
\subsection{The function $V\label{sectionv}$}
The function $\widetilde{\mu},$ $\widetilde{\mu}(y)=\mu_{1}(y)/y$ is
strictly decreasing on $\mathbf{R}_{+},$ and therefore $\mu_{1}(y)/y\geq\mu
_{1}(t^{\gamma}y)/(t^{\gamma}y)$ for $t\geq1.$ Furthermore, the functions $%
\eta$ and $\mu_{1}$ are strictly positive and $\mu_{1}$ is strictly
increasing. These properties imply that, for $t\geq\tau\geq1,$ $\overline
{u}(x,t)+t^{-\gamma}\eta(\left| z\right| )=$ $\mu_{1}(\left| y\right|
)+t^{-\gamma}\eta(\left| z\right| )\geq$ $t^{-\gamma}(\mu_{1}(\left|
z\right| )+\eta(\left| z\right| ))$ $\geq t^{-\gamma}c_{0}>0,$ where $%
c_{0}=\inf_{z>0}(\mu_{1}(z)+\eta(z)).$ Next, since the functions $\mu_{3}$
and $\varphi_{2}$ are bounded and since $3\gamma\geq\varepsilon,$ for $%
n\geq4,$ we have that $\left| t^{-\varepsilon}\mu_{3}(\left| y\right|
)+t^{-3\gamma}\varphi_{2}(\left| z\right| )\right| <\mathrm{const.~}%
t^{-\varepsilon},$ and as a consequence $(\overline{u}+\phi)$ and $(2%
\overline{u}+\phi)$ are positive functions of $x$ for all fixed $t\geq
\tau_{0},$ if $\tau_{0}$ large enough.
\begin{proposition}
\label{propvodd}For $n$ odd, $n\geq 5,$ there exists $\tau _{0}\geq 1,$ such
that for all $t\geq \tau _{0}$ the function $\widehat{V}$ is positive.
\end{proposition}
\begin{proof}
The function $(2\overline{u}\phi +\phi ^{2})^{n-1}$ is positive, for $n$ odd.
\end{proof}
\medskip
As a consequence, for $n$ odd, equation (\ref{equpsi1}) is of the form
indicated in Section 2, with $V=\widehat{V}$ and $T=\widehat{T}.$ The rest
of this section treats the case of $n$ even, which, as indicated in the
introduction, is slightly more delicate. It can be skipped in a first
reading or if the reader is only interested in the case of $n$ odd.
So let $n$ be even. The idea is to split $\widehat{V}$ into its positive
part $V=$ $\widehat{V}_{+}$ and its negative part $V_{1}=\widehat{V}_{-}~$,
and to show that $\widehat{V}_{-}$ is small enough so that it can be treated
together with the nonlinear term. Consider the function $\phi $ defined in (%
\ref{defphi}). The problem is that $\phi $ becomes negative for large values
of $x$, and that therefore $\widehat{V}$ becomes negative for large values
of $x.$ To understand why $\phi $ becomes negative, we note that the leading
order term $t^{-\gamma }\lambda z^{-\delta }$ in the large $z$ asymptotics
of $t^{-\gamma }\eta (z)$ is compensated by the leading order term $%
-t^{-\varepsilon }\lambda y^{-\delta }$ in the large $y$ asymptotics of $%
t^{-\varepsilon }\mu _{3}(y).$ The leading order of $\phi $ at $x$ large is
therefore given by the second order term in the large $z$ asymptotics of $%
\eta $ and the leading term in the large $z$ asymptotics of $\varphi _{2}.$
The first of these terms is proportional to $t^{-\gamma }z^{-\delta ^{\prime
}},$ and the second one is proportional to $t^{-3\gamma }z^{2-\delta
^{\prime }}\equiv t^{-\gamma }y^{2}z^{-\delta ^{\prime }}.$ The
corresponding proportionality constants $\lambda _{\infty }$ and $\lambda
^{\prime }$ can be computed for $n>5$ and turn out to be negative. For $%
n=4,5 $ these constants can not be obtained from asymptotic expansions, but
numerical results show that they are in fact also negative in these cases.
We do not need a proof of this numerical fact, because the following
proposition is also correct for positive $\widehat{V}$.
\begin{proposition}
\label{propV}For $n$ even, $n\geq4,$ there exists $\tau_{1}\geq1$, such that
the function $V_{1}$, satisfies for all $t\geq\tau_{1}$ the bound
\begin{equation}
\sup_{x\in\mathbf{R}}|V_{1}(x,t)|\leq\mathrm{const.}~t^{-\gamma(n-1)(\delta
^{\prime}+1)}~. \label{v1bound}
\end{equation}
\end{proposition}
\begin{proof}
The idea is to write $\phi $ as the sum of a function $\phi _{0}$ that is
positive and a function $\phi _{\infty }$ that absorbs the asymptotic
behavior at infinity. Since $\mu _{3}(y)\approx \lambda _{0}y^{2-\delta }$
for $y$ small, with $\lambda _{0}>0,$ there exists $y_{0}>0$ such that $\mu
_{3}(|y|)\geq 0,$ for all $|y|\leq y_{0}.$ Let $c>0,$ to be chosen below,
and let $\theta $ be the Heaviside step function, i.e., $\theta (x)=1$ for $%
x>0,$ and $\theta (x)=0$ for $x<0.$ Then, we define the function $\phi
_{\infty }$ by the equation
\begin{equation*}
\phi _{\infty }(x,t)=-ct^{-\gamma }\theta (\left| y\right| -y_{0})\left|
y\right| ^{2}\left| z\right| ^{-\delta ^{\prime }}~,
\end{equation*}
and we set $\phi _{0}=\phi -\phi _{\infty }.$ In order to prove that $\phi
_{0}$ is positive, for $c$ large enough, we write $\phi _{0}=\phi
_{0}^{\left( 1\right) }+\phi _{0}^{\left( 2\right) },$ where
\begin{equation*}
\phi _{0}^{(1)}(x,t)=t^{-\gamma }(\eta (\left| z\right| )-\lambda
|z|^{-\delta }\theta (\left| y\right| -y_{0}))+t^{-3\gamma }\varphi
_{2}(\left| z\right| )+c~\theta (\left| y\right| -y_{0})t^{-3\gamma
}|z|^{2-\delta ^{\prime }}~,
\end{equation*}
and
\begin{equation*}
\phi _{0}^{(2)}(x,t)=t^{-\varepsilon }\left( \mu _{3}(\left| y\right|
)+\lambda |y|^{-\delta }\theta (\left| y\right| -y_{0})\right) ~.
\end{equation*}
$\phi _{0}^{(2)}$ is positive for $\left| y\right| >y_{0},$ since in this
case $\phi _{0}^{(2)}(x,t)=\mu _{2}(\left| y\right| )>0,$ and $\phi
_{0}^{(2)}$ is positive for $\left| y\right| <y_{0}$ by definition of $%
y_{0}. $ Next we consider $\phi _{0}^{(1)}.$ For $\left| z\right| <t^{\gamma
}y_{0}$ we have that $\phi _{0}^{(1)}(x,t)=t^{-\gamma }\eta (\left| z\right|
)+t^{-3\gamma }\varphi _{2}(\left| z\right| )$. But $t^{-\gamma }\eta
(z)+t^{-3\gamma }\varphi _{2}(z)>0$ for all $z\geq 0,$ and all $t\geq \tau ,$
if $\tau $ is sufficiently large, since $\eta >0,$ since $\varphi _{2}$ is
bounded, and since $\left| \varphi _{2}(z)\right| <\eta (z)$ for $z$ large
enough. Finally, using the asymptotic properties of $\eta $ and $\varphi
_{2} $ we see that $\phi _{0}^{(1)}>0$ for $\left| z\right| >t^{\gamma
}y_{0} $ if $c$ is chosen large enough.
We now estimate the function $V_{1}.$ From the definition of $\phi _{\infty
} $ we get that
\begin{equation*}
|\phi _{\infty }(x,t)|\leq \mathrm{const.}\text{ }t^{-\gamma (\delta
^{\prime }+1)}~,
\end{equation*}
and therefore, since $\phi _{0}$ is positive, we have the lower bound
\begin{equation*}
\phi (x,t)^{n-1}\geq \overline{c}~t^{-\gamma (n-1)(\delta ^{\prime }+1)}~,
\end{equation*}
for some constant $\overline{c}<0$, from which (\ref{v1bound}) follows.
\end{proof}
\subsection{The function $I$}
\begin{theorem}
\label{thm:Inhomo} Let $I$ be as defined in (\ref{equi}), and let $n\geq4.$
Then, there exists a constant $c_{I}>0,$ such that for all $t\geq1,$%
\begin{equation}
\int_{-\infty}^{\infty}dx~|I(\sqrt{t}x,t)|\leq c_{I}~t^{-1-4\gamma }~.
\label{eqn:thminhomo}
\end{equation}
\end{theorem}
The function $I$ is even, and it is therefore sufficient to bound it for $%
x\geq0.$ The strategy of the proof will be to rewrite the function $I$ as a
sum of functions of the form $t^{-\sigma}G(y)F(z),$ with $\sigma>0,$ and
with $G$ and $F$ functions with appropriate asymptotic behavior at zero and
infinity. Each of the terms in the sum can then be estimated with the help
of Lemma \ref{lem:estimations} below. In order to keep the notation as
simple as possible, we suppress in what follows the arguments of functions
whenever there is no risk of confusion.
\begin{proposition}
For $x\geq0,$ the function $-I$ is of the form
\begin{equation}
-I=\sum_{p=2}^{n}\sum_{q=0}^{p}A_{p,q}+\sum_{i=2}^{8}A_{i}~, \label{repi}
\end{equation}
where
\begin{align*}
A_{2} & =\sum_{p=1}^{n-1}\binom{n-1}{p}t^{-2\gamma(n-1-p)-p%
\varepsilon}T_{1}^{n-1-p}T_{2}^{p}\left( t^{-4\gamma}T_{3}\right) ~, \\
A_{3} & =n(t^{-2\gamma}T_{1}+t^{-\varepsilon}T_{2})^{n-1}\left(
t^{-\gamma(1+\delta^{\prime})}T_{4}+t^{-2\varepsilon}T_{5}+t^{-6\gamma}T_{6}%
\right) ~, \\
A_{4} & =\sum_{p=2}^{n}\binom{n}{p}(t^{-2\gamma}T_{1}+t^{-%
\varepsilon}T_{2})^{n-p}(t^{-4\gamma}T_{3}+t^{-\gamma(1+\delta^{%
\prime})}T_{4}+t^{-2\varepsilon}T_{5}+t^{-6\gamma}T_{6})^{p}~, \\
A_{5} & =-t^{-1-3\gamma}\left( 3\gamma\varphi_{2}+\alpha z\varphi
_{2}^{\prime}\right) ~, \\
A_{6} & =-t^{-2n\gamma-2\gamma}nT_{1}^{n-1}2z^{2-\delta}\eta\left(
\lambda_{0}-\mu_{3}y^{\delta-2}\right) ~, \\
A_{7} & =t^{-2n\gamma+2\gamma-\varepsilon}nT_{1}^{n-1}(2\mu_{1}\mu
_{3}-2\kappa\lambda_{0}y^{3-\delta})-t^{-2n\gamma+2\gamma-\varepsilon
}n(2\kappa\lambda z^{1-\delta})^{n-1}(2\mu_{1}\mu_{3}-2\kappa\lambda
_{0}y^{3-\delta})~, \\
A_{8} & =t^{-2n\gamma+\gamma}nT_{1}^{n-1}2\eta((\mu_{1}-\kappa y)-\kappa
_{3}y^{3})-t^{-2n\gamma+\gamma}n(2\kappa\lambda z^{1-\delta})^{n-1}2\lambda
z^{-\delta}(\mu_{1}-\kappa y-\kappa_{3}y^{3})~, \\
A_{p,q} & =\binom{n}{p}\binom{p}{q}\left( R_{2}^{p,q}-R_{1}^{p,q}\right) ~,
\end{align*}
where
\begin{align*}
R_{1}^{p,q} & =t^{-2n\gamma+2p\gamma-p\varepsilon}(2\kappa\lambda
z^{1-\delta})^{n-p}(2(\mu_{1}-\kappa y)\lambda
y^{-\delta})^{p-q}(2\mu_{1}\mu_{3})^{q}~, \\
R_{2}^{p,q} & =t^{-2n\gamma+2p\gamma-p\varepsilon}T_{1}^{n-p}(2(\mu
_{1}-\kappa y)y^{-\delta}z^{\delta}\eta)^{p-q}(2\mu_{1}\mu_{3})^{q}~,
\end{align*}
and where
\begin{align}
T_{1}(z) & =2\kappa z\eta(z)+\eta(z)^{2}~, \label{t1} \\
T_{2}(y,z) & =2(\mu_{1}(y)-\kappa y)y^{-\delta}z^{\delta}\eta(z)+2\mu
_{1}(y)\mu_{3}(y)~, \label{t2} \\
T_{3}(y,z) & =\left( 2\kappa z+2\eta(z)\right) \varphi_{2}(z)+2\mu
_{3}(y)y^{\delta-2}z^{2-\delta}\eta(z)~, \label{t3} \\
T_{4}(y,z) & =2(\mu_{1}(y)-\kappa y)y^{2-\delta^{\prime}}z^{\delta^{\prime
}-2}\varphi_{2}(z)~, \label{t4} \\
T_{5}(y) & =\mu_{3}(y)^{2}~, \label{t5} \\
T_{6}(y,z) & =2\mu_{3}(y)y^{\delta-2}z^{2-\delta}\varphi_{2}(z)+\varphi
_{2}(z)^{2}~. \label{t6}
\end{align}
\end{proposition}
\begin{proof}
In terms of the functions (\ref{t1})--(\ref{t6}) we get that, for $x>0,$
\begin{equation*}
2\overline{u}\phi +\phi ^{2}=t^{-2\gamma }T_{1}+t^{-\varepsilon
}T_{2}+t^{-4\gamma }T_{3}+t^{-\gamma (1+\delta ^{\prime
})}T_{4}+t^{-2\varepsilon }T_{5}+t^{-6\gamma }T_{6}~,
\end{equation*}
and therefore $(2\overline{u}\phi +\phi ^{2})^{n}=B_{1}+\sum_{i=2}^{4}A_{i}$%
, where
\begin{equation*}
B_{1}=(t^{-2\gamma }T_{1}+t^{-\varepsilon }T_{2})^{n}+n(t^{-2\gamma
}T_{1})^{n-1}(t^{-4\gamma }T_{3})~,
\end{equation*}
and where $A_{2},$ $A_{3}$ and $A_{4}$ are as defined above. Since $-I=\dot{%
\phi}-\phi _{1}^{\prime \prime }+\left( 2\overline{u}\phi +\phi ^{2}\right)
^{n},$ it remains to be shown that
\begin{equation*}
B_{1}-\phi _{1}^{\prime \prime }+\dot{\phi}=\sum_{p=2}^{n}%
\sum_{q=0}^{p}A_{p,q}+\sum_{i=5}^{8}A_{i}~.
\end{equation*}
Using the differential equations for $\mu _{2},$ $\eta $ and $\varphi _{2},$
we find that
\begin{equation*}
B_{1}-\phi _{1}^{\prime \prime }+\dot{\phi}=R_{1}+R_{2}+S_{3}+A_{5}~,
\end{equation*}
where $A_{5}$ as defined above, where
\begin{align*}
R_{1}& =-t^{-n\varepsilon }\left( -(2\kappa \lambda y^{1-\delta })^{n}+(2\mu
_{1}(\lambda y^{-\delta }+\mu _{3}))^{n}\right) ~, \\
R_{2}& =(t^{-2\gamma }T_{1}+t^{-\varepsilon }T_{2})^{n}-(t^{-2\gamma
}T_{1})^{n}~,
\end{align*}
and where
\begin{align*}
S_{3}& =-t^{-2n\gamma -2\gamma }\left[ -n(2\kappa \lambda )^{n}(\frac{\kappa
_{3}}{\kappa }+\frac{\lambda _{0}}{\lambda })z^{-\delta
}+n(T_{1})^{n-1}\left( (\varphi _{2}+\lambda _{0}z^{2-\delta })(2\kappa
z+2\eta )+2\kappa _{3}z^{3}\eta \right) \right] \\
& \hspace{3cm}+n(t^{-2\gamma }T_{1})^{n-1}(t^{-4\gamma }T_{3})~.
\end{align*}
The functions $R_{1}$ and $R_{2}$ can be further decomposed as follows
\begin{align*}
R_{1}& =S_{1}-\sum_{p=2}^{n}\sum_{q=0}^{p}\binom{n}{p}\binom{p}{q}%
R_{1}^{p,q}~, \\
R_{2}& =S_{2}+\sum_{p=2}^{n}\sum_{q=0}^{p}\binom{n}{p}\binom{p}{q}%
R_{2}^{p,q}~,
\end{align*}
where $R_{1}^{p,q}$ and $R_{2}^{p,q}$ are as defined above, and where
\begin{align*}
S_{1}& =-nt^{-2\gamma (n-1)-\varepsilon }(2\kappa \lambda z^{1-\delta
})^{n-1}(2(\mu _{1}-\kappa y)\lambda y^{-\delta }+2\mu _{1}\mu _{3})~, \\
S_{2}& =nt^{-2\gamma (n-1)-\varepsilon }T_{1}^{n-1}T_{2}~.
\end{align*}
It remains to be shown that
\begin{equation*}
\sum_{i=1}^{3}S_{i}=\sum_{i=6}^{8}A_{i}~,
\end{equation*}
but this follows using the definitions.
\end{proof}
\subsubsection{Proof of Theorem \ref{thm:Inhomo}}
In order to characterize the behavior of a function near zero and infinity
we introduce the following family of vector spaces.
\begin{definition}
Let $p$ and $q$ be two real numbers with $p+q\geq0.$ Then, we define $%
\mathcal{V}(p,q)$ to be the vector space of continuous functions $F$ from $%
\mathbf{R}_{+}$ to $\mathbf{R,}$ for which the norm
\begin{equation*}
\left\| F\right\| _{p,q}=\sup_{x\geq0}\left| F(x)\right| \left(
x^{-p}+x^{q}\right)
\end{equation*}
is finite.
\end{definition}
Note that, if a function is in $\mathcal{V}(p,q)$, then it is also in $%
\mathcal{V}(p^{\prime},q^{\prime})$ for any pair of numbers $p^{\prime}\leq
p,q^{\prime}\leq q$ for which$\ p^{\prime}+q^{\prime}\geq0.$ Furthermore, if
$F_{1}$ is in $\mathcal{V}(p_{1},q_{1})$, and $F_{2}$ is in $\mathcal{V}%
(p_{1},q_{2})$, then the product $F_{1}F_{2}$ is in $\mathcal{V}%
(p_{1}+p_{2},q_{1}+q_{2})$.
The following Lemma provides the tool that we use to estimate the terms on
the right hand side of (\ref{repi}).
\begin{lemma}
\label{lem:estimations} \label{lem:bornes} Let $F\in\mathcal{V}(F_{0},F_{1})
$ and $G\in\mathcal{V}(G_{0},G_{1})$, and assume that
\begin{align}
1-G_{1} & <F_{1}~, \label{cb1} \\
1+G_{0} & >-F_{0}~, \label{cb2}
\end{align}
and that
\begin{equation}
F_{1}\neq1+G_{0}~. \label{cb3}
\end{equation}
Then, there is a constant $C>0,$ such that for all $t\geq1,$
\begin{equation}
\int_{0}^{\infty}\left| G(x)F(t^{\gamma}x)\right| ~dx\leq Ct^{-\xi }~,
\label{decay}
\end{equation}
where
\begin{equation}
\xi=\gamma\cdot\min\{F_{1},1+G_{0}\}~. \label{cb4}
\end{equation}
\end{lemma}
\begin{proof}
From (\ref{cb3}) it follows that either $F_{1}<1+G_{0}$ or $F_{1}>1+G_{0}.$
In the first case we get using (\ref{cb1}) that $1-G_{1}<$ $F_{1}<$ $%
1+G_{0}, $ and therefore
\begin{align*}
\int_{0}^{\infty}\left| G(x)F(t^{\gamma}x)\right| ~dx & \leq\left(
\sup_{x\geq0}x^{F_{1}}\left| F(t^{\gamma}x)\right| \right) \int_{0}^{\infty}%
\frac{1}{x^{F_{1}}}\left| G(x)\right| ~dx \\
& \leq\mathrm{const.}~t^{-\gamma F_{1}}~,
\end{align*}
and in the second case we get using (\ref{cb2}) that $-F_{0}<1+G_{0}<F_{1},$
and therefore
\begin{align*}
\int_{0}^{\infty}\left| G(x)F(t^{\gamma}x)\right| ~dx & \leq\left(
\sup_{x\geq0}\frac{1}{x^{G_{0}}}\left| G(x)\right| \right)
\int_{0}^{\infty}x^{G_{0}}\left| F(t^{\gamma}x)\right| ~dx \\
& \leq\mathrm{const.}~t^{-\gamma(1+G_{0})}~.
\end{align*}
\end{proof}
We now show that the right hand side in (\ref{repi}) can be bounded by a sum
of terms of the form $t^{-\sigma}G(y)F(z).$ For each of these terms we then
show that the corresponding functions $G$ and $F$ satisfy the hypothesis of
Lemma \ref{lem:estimations}. This then implies that $\int_{0}^{\infty
}t^{-\sigma}G(x)F(t^{\gamma}x)~dx\leq\mathrm{const.}$ $t^{-(\sigma+\xi)},$
for a certain $\xi$ depending on $F$ and $G.$ It is therefore sufficient to
prove that $\sigma+\xi\geq1+4\gamma$ for all these terms in order to prove
the inequality (\ref{eqn:thminhomo}).
\begin{proposition}
For $y,$ $z>0$ we have the bounds
\begin{align}
T_{2}(y,z) & \leq\widehat{T_{2}}(y)\equiv2\left| \mu_{1}-\kappa y\right|
y^{-\delta}\left( \sup_{z>0}\left| z^{\delta}\eta(z)\right| \right) +2\left|
\mu_{1}\mu_{3}\right| ~, \notag \\
T_{3}(y,z) & \leq\widehat{T_{3}}(z)\equiv\left| \left( 2\kappa
z+2\eta\right) \varphi_{2}\right| +2\left( \sup_{y>0}\left| \mu
_{3}(y)y^{\delta-2}\right| \right) z^{2-\delta}\eta~, \notag \\
T_{4}(y,z) & \leq T_{4,1}(y)\cdot T_{4,2}(z)\equiv\left( 2\left| \mu
_{1}(y)-\kappa y\right| y^{2-\delta^{\prime}}\right) \cdot\left(
z^{\delta^{\prime}-2}\left| \varphi_{2}(z)\right| \right) ~,
\label{inequalities} \\
T_{6}(y,z) & \leq\widehat{T_{6}}(z)\equiv2\left( \sup_{y>0}\left| \mu
_{3}(y)y^{\delta-2}\right| \right) z^{2-\delta}\left| \varphi_{2}\right|
+\left| \varphi_{2}\right| ^{2}~, \notag
\end{align}
and $T_{1}\in\mathcal{V}(0,\delta-1)$, $\widehat{T_{2}}\in\mathcal{V}%
(3-\delta,\delta-1)$, $\widehat{T_{3}}\in\mathcal{V}(2-\delta,\delta^{\prime
}-3)$, $T_{4,1}\in\mathcal{V}(5-\delta^{\prime},\delta^{\prime}-3)$, $%
T_{4,2}\in\mathcal{V}(\delta^{\prime}-\delta,0)$, $T_{5}\in\mathcal{V}%
(4-2\delta,2\delta)$, $\widehat{T_{6}}\in\mathcal{V}(4-2\delta,\delta^{%
\prime }+\delta-4)$.
\end{proposition}
\begin{proof}
The inequalities (\ref{inequalities}) follow by using the triangle
inequality and the asymptotic properties of the functions $\mu_{1}$, $%
\mu_{2} $, $\eta$ and $\varphi_{1}$.
\end{proof}
\paragraph{Bound on the function $A_{2}$.}
We have the bound
\begin{equation*}
|A_{2}|\leq\mathrm{const.}\sum_{p=1}^{n-1}t^{-\sigma}\widehat{T_{2}}%
^{p}~\left( T_{1}^{n-1-p}\widehat{T_{3}}\right) ~,
\end{equation*}
where $\sigma=1+\gamma+3p\varepsilon\gamma.$ The function $G=\widehat{T_{2}}%
^{p}$ is in $\mathcal{V}((3-\delta)p,3\varepsilon p)$, and the function $%
F=T_{1}^{n-1-p}\widehat{T_{3}}$ is in $\mathcal{V}(2-\delta,$ $3\varepsilon
(n-1-p)-3+\delta^{\prime})$. Since $\delta^{\prime}>1,$ the inequalities (%
\ref{cb1}) and (\ref{cb2}) are satisfied and, since $\delta^{\prime}<5$ for $%
n\geq3,$%
\begin{equation*}
\xi/\gamma=\left\{
\begin{array}{ll}
\delta^{\prime}-3\varepsilon p & \text{if}\;p\geq2 \\
3-3\varepsilon p & \text{if}\;p=1
\end{array}
\right. \geq3-3\varepsilon p~.
\end{equation*}
Therefore, $\sigma+\xi\geq$ $1+\gamma+3p\varepsilon\gamma+(3-3p\varepsilon
)\gamma=$ $1+4\gamma\ $as required.
\paragraph{Bound on the function $A_{3}$.}
We have that $A_{3}=t^{-\gamma(1+\delta^{\prime})}B_{3,4}+t^{-2\varepsilon
}B_{3,5}+t^{-6\gamma}B_{3,6},$ where $B_{3,i}=n(t^{-2\gamma}T_{1}+t^{-%
\varepsilon}T_{2})^{n-1}T_{i},$ $i=4,\dots,6$. Since $|T_{2}/T_{1}|\leq%
\mathrm{const.}~t^{3\gamma\varepsilon}$, and $\varepsilon-3\gamma
\varepsilon=2\gamma,$ we have the bound
\begin{equation*}
t^{-\gamma(1+\delta^{\prime})}|B_{3,4}|\leq\mathrm{const.}\text{ }t^{-\sigma
}\left( T_{1}^{n-1}T_{4,2}\right) ~T_{4,1}~,
\end{equation*}
with $\sigma=2\gamma(n-1)+\gamma(1+\delta^{\prime}).$ The function $%
G=T_{4,1} $ is in $\mathcal{V}(5-\delta^{\prime},\delta^{\prime}-3)$ and the
function $F=T_{1}^{n-1}T_{4,2}$ is in $\mathcal{V}(\delta^{\prime}-\delta,3)$%
. Since $\delta^{\prime}>3$ the inequalities (\ref{cb1}) and (\ref{cb2}) are
satisfied and $\xi/\gamma=6-\delta^{\prime}$. Therefore $\sigma+\xi=$ $%
1-3\gamma +\gamma(1+\delta^{\prime})+(6-\delta^{\prime})\gamma=$ $1+4\gamma$
as required.
Similarly, we have that
\begin{equation*}
t^{-2\varepsilon}|B_{3,5}|\leq\mathrm{const.}\text{ }t^{-%
\sigma}T_{1}^{n-1}~T_{5}~,
\end{equation*}
with $\sigma=2\gamma(n-1)+2\varepsilon.$ The function $G=T_{5}$ is in $%
\mathcal{V}(4-2\delta,2\delta)$ and the function $F=T_{1}^{n-1}$ is in $%
\mathcal{V}(0,3)$. The inequalities (\ref{cb1}) and (\ref{cb2}) are
satisfied and $\xi/\gamma=5-2\delta$ $=3-6\varepsilon.$ Therefore, $%
\sigma+\xi=$ $1-3\gamma+2\varepsilon+(3-6\varepsilon)\gamma=$ $1+4\gamma$ as
required.
Finally,
\begin{equation*}
t^{-6\gamma}|B_{3,6}|\leq\mathrm{const.}\text{ }t^{-\sigma}T_{1}^{n-1}%
\widehat{T_{6}}~,
\end{equation*}
where $\sigma=2(n-1)\gamma+6\gamma.$ The function $G\equiv1$ is in $\mathcal{%
V}(0,0)$, and the function $F=T_{1}^{n-1}\widehat{T_{6}}$ is in $\mathcal{V}%
(4-2\delta,3(n-1)\varepsilon+\delta^{\prime}+\delta-4)$. The inequalities (%
\ref{cb1}) and (\ref{cb2}) are satisfied and $\xi/\gamma=1$. Therefore, $%
\sigma+\xi=$ $1-3\gamma+6\gamma+\gamma=$ $1+4\gamma$ as required.
\paragraph{Bound on the function $A_{4}$.}
Since the functions $T_{3}/T_{1}$ and $T_{6}/T_{1}$ are bounded, $%
T_{4}/T_{1}\leq\mathrm{const.}~t^{3\varepsilon\gamma}$ and $T_{5}/T_{1}\leq%
\mathrm{const.}~t^{3\varepsilon\gamma}$ we have that
\begin{equation*}
|A_{4}|\leq\mathrm{const.}\sum_{p=2}^{n}t^{-2\gamma(n+p)}T_{1}^{n}\leq%
\mathrm{const.}~t^{-\sigma}T_{1}^{n}~,
\end{equation*}
where $\sigma=2n\gamma+4\gamma.$ The function $G\equiv1$ is in $\mathcal{V}%
(0,0)$, and the function $F=T_{1}^{n}$ is in $\mathcal{V}(0,3n\varepsilon) $%
. The inequalities (\ref{cb1}) and (\ref{cb2}) are satisfied and $%
\xi/\gamma=1$. Therefore, $\sigma+\xi=$ $\left( 1-\gamma\right)
+4\gamma+\gamma=$ $1+4\gamma$ as required.
\paragraph{Bound on the function $A_{5}$.}
We have the bound
\begin{equation}
\left| A_{5}\right| \leq t^{-\sigma}\left| 3\gamma\varphi_{2}+\alpha
z\varphi_{2}^{\prime}\right| ~,
\end{equation}
where $\sigma=2n\gamma+4\gamma.$ The function $G\equiv1$ is in $\mathcal{V}%
(0,0)$, and the function $F=\left| 3\gamma\varphi_{2}+\alpha z\varphi
_{2}^{\prime}\right| $ is in $\mathcal{V}(2-\delta,\delta^{\prime}-2)$. The
inequalities (\ref{cb1}) and (\ref{cb2}) are satisfied and $\xi/\gamma=1$.
Therefore, $\sigma+\xi=$ $\left( 1-\gamma\right) +4\gamma+\gamma=$ $%
1+4\gamma $ as required.
\paragraph{Bound on the function $A_{6}$.}
We have the bound
\begin{equation*}
|A_{6}|\leq\mathrm{const.}\text{ }t^{-\sigma}\left( T_{1}^{n-1}z^{2-\delta
}\eta\right) ~|\lambda_{0}-\mu_{3}y^{\delta-2}|~,
\end{equation*}
where $\sigma=1+\gamma.$ The function $G=|\lambda_{0}-\mu_{3}y^{\delta-2}|$
is in $\mathcal{V}(2,2)$ and the function $F=T_{1}^{n-1}z^{2-\delta}\eta$ is
in $\mathcal{V}(2-\delta,1+2\delta).$ The inequalities (\ref{cb1}) and (\ref
{cb2}) are satisfied and $\xi/\gamma=3$. Therefore, $\sigma+\xi=$ $%
1+\gamma+3\gamma=$ $1+4\gamma$ as required.
\paragraph{Bound on the function $A_{7}$.}
We have the bound
\begin{equation*}
|A_{7}|\leq\mathrm{const.}\text{ }t^{-\sigma}|T_{1}^{n-1}-(2\kappa\lambda
z^{1-\delta})^{n-1}|~|2\mu_{1}\mu_{3}-2\kappa\lambda_{0}y^{3-\delta}|~,
\end{equation*}
where $\sigma=2n\gamma-2\gamma+\varepsilon$. The function $G=|2\mu_{1}\mu
_{3}-2\kappa\lambda_{0}y^{3-\delta}|$ is in $\mathcal{V}(5-\delta,\delta-3)$
and the function $F=|T_{1}^{n-1}-(2\kappa\lambda z^{1-\delta})^{n-1}|$ is in
$\mathcal{V}(-3,3+\delta^{\prime}-\delta).$ The inequalities (\ref{cb1}) and
(\ref{cb2}) are satisfied and $\xi/\gamma=6-\delta$. Therefore, $\sigma+\xi=$
$\left( 1-\gamma\right) -2\gamma+\varepsilon+(6-\delta)\gamma=$ $1+4\gamma$
as required.
\paragraph{Bound on the function $A_{8}$.}
We have the bound
\begin{equation*}
|A_{8}|\leq \mathrm{const.}\text{ }t^{-\sigma }|T_{1}^{n-1}\eta -(2\kappa
\lambda z^{1-\delta })^{n-1}\lambda z^{-\delta }|~|\mu _{1}-\kappa y-\kappa
_{3}y^{3}|~,
\end{equation*}
where $\sigma =2n\gamma -\gamma .$ The function $G=|\mu _{1}-\kappa y-\kappa
_{3}y^{3}|$ is in $\mathcal{V}(5,-3)$ and the function $F=|T_{1}^{n-1}\eta
-(2\kappa \lambda z^{1-\delta })^{n-1}\lambda z^{-\delta }|\ $is in $%
\mathcal{V}(-3-\delta ,3+\delta ^{\prime }).$ The inequalities (\ref{cb1})
and (\ref{cb2}) are satisfied and $\xi /\gamma =6$. Therefore, $\sigma +\xi
= $ $\left( 1-\gamma \right) -\gamma +6\gamma =$ $1+4\gamma $ as required.
\paragraph{Bound on the functions $A_{p,q}$.}
We have the bound
\begin{equation*}
|A_{p,q}|\leq\mathrm{const.}~t^{-\sigma}~|(z^{\delta}%
\eta)^{p-q}~T_{1}^{n-p}-\lambda^{p-q}(2\kappa\lambda
z^{1-\delta})^{n-p}|~\left( \left| 2(\mu_{1}-\kappa y)y^{-\delta}\right|
^{p-q}\left| 2\mu_{1}\mu_{3}\right| ^{q}\right) ~,
\end{equation*}
where $\sigma=2n\gamma-2p\gamma+p\varepsilon.$ The function $G=\left|
2(\mu_{1}-\kappa y)y^{-\delta}\right| ^{p-q}\left| 2\mu_{1}\mu_{3}\right|
^{q}$ is in $\mathcal{V}(p(3-\delta),3\varepsilon p+q)$ and the function $%
F=|(z^{\delta}\eta)^{p-q}~T_{1}^{n-p}-\lambda^{p-q}(2\kappa z^{1-\delta
})^{n-p}|$ is in $\mathcal{V}(-3\varepsilon(n-p),2+\delta^{\prime
}-3p\varepsilon)$. The inequalities (\ref{cb1}) and (\ref{cb2}) are
satisfied, and
\begin{equation*}
\xi/\gamma=\left\{
\begin{array}{ll}
5-3p\varepsilon & \text{if}\;p=2~, \\
2+\delta^{\prime}-3p\varepsilon & \text{if}\;p\geq3~.
\end{array}
\right.
\end{equation*}
Therefore, $\sigma+\xi=1+4\gamma$, for $p=2$ and $\sigma+\xi=1+\gamma
(1+\delta^{\prime})>1+4\gamma$, for $p\geq3,$ as required.
\bigskip
\noindent This completes the proof of Theorem \ref{thm:Inhomo}. $%
\blacksquare $
\subsection{The Map $T$}
Equation (\ref{equpsi1}) is of the form (\ref{eqpsi}) if we define the map $%
T $ by the equation
\begin{equation}
T(\psi )=\left\{
\begin{array}{ll}
\widehat{T}(\psi ) & \text{for }n\text{ odd}~, \\
\widehat{T}(\psi )+V_{1}\psi ~ & \text{for }n\text{ even}~,
\end{array}
\right. \label{mapt}
\end{equation}
with $\widehat{T}$ as defined in (\ref{that}) and $V_{1}$ as defined in
Section \ref{sectionv}. Using the definitions, we see that $T$ can be
written as,
\begin{equation}
T(\psi )=\sum_{p=1}^{n}\sum_{q=0}^{p}V_{p,q}~\psi ^{p+q}~,
\label{eqn:tpsidef}
\end{equation}
with
\begin{equation}
V_{p,q}=\left\{
\begin{array}{ll}
0 & \text{for }(p,q)=(1,0)\text{ and }n\text{ odd,} \\
V_{1} & \text{for }(p,q)=(1,0)\text{ and }n\text{ even,} \\
\binom{n}{p}\binom{p}{q}(2\overline{u}\phi +\phi ^{2})^{n-p}(2\overline{u}%
+2\phi )^{p-q} & \text{for }p+q\geq 2~.
\end{array}
\right. \label{vpq}
\end{equation}
\begin{proposition}
Let $V_{p,q}$ as in (\ref{vpq}). Then, for all $t\geq1,$
\begin{equation}
\sup_{x\in\mathbf{R}}|V_{p,q}(x,t)|\leq\mathrm{const.}\text{ }t^{-e(p,q)}~,
\label{eqn:Lemmavmp}
\end{equation}
where
\begin{equation}
e(p,q)=\left\{
\begin{array}{ll}
\gamma(n-1)(\delta^{\prime}+1) & \text{for }(p,q)=\left( 1,0\right) ~, \\
2\gamma(n-2)+2\gamma & \text{for}\;(p,q)=(2,0)~, \\
2\gamma(n-p) & \text{for}\;(p,q)\neq(2,0)\;\text{and}\;p+q\geq2~.
\end{array}
\right. \label{mupq}
\end{equation}
\end{proposition}
\begin{proof}
The case $(p,q)=(1,0)$ follows from (\ref{v1bound}). Let now $(p,q)\neq
(1,0).$ Since $\varepsilon -\gamma -\nu \gamma \geq 0,$ for all $\nu ,$ $%
0\leq \nu \leq \delta ,$ we find that
\begin{align*}
\sup_{x\in \mathbf{R}}\left| z\right| ^{\nu }~\left| \phi (x,t)\right| &
\leq t^{-\gamma }(\sup_{z\in \mathbf{R}_{+}}\left| z^{\nu }~\eta (z)\right|
+t^{-(\varepsilon -\gamma -\nu \gamma )}\sup_{y\in \mathbf{R}_{+}}\left|
y^{\nu }~\mu _{3}(y)\right| +t^{-2\gamma }\sup_{z\in \mathbf{R}_{+}}\left|
z^{\nu }~\varphi _{2}(z)\right| ) \\
& \leq \mathrm{const.}\text{ }t^{-\gamma }~.
\end{align*}
Furthermore, since $\mu _{1}(y)=\mathcal{O}(y)$ near $y=0,$%
\begin{equation*}
\left| 2\overline{u}\phi +\phi ^{2}\right| \leq t^{-\gamma }~2\frac{\mu
_{1}(\left| y\right| )}{\left| y\right| }~\left| z\phi \right| +\left| \phi
\right| ^{2}\leq \mathrm{const.}\text{ }t^{-2\gamma }~.
\end{equation*}
Since the function $\left| \overline{u}+\phi \right| $ is bounded, it
follows that $|V_{p,q}(x,t)|\leq \mathrm{const.}$ $t^{-e(p,q)},$ with $%
e(p,q)=2\gamma (n-p).$ For $(p,q)=(2,0)$ we improve this bound using
additional properties of the function $\overline{u}+\phi .$ Namely, since $%
2/(n-2)\leq \delta -1,$ we have that
\begin{align*}
|V_{2,0}(x,t)|& \leq \mathrm{const.}\sup_{x\in \mathbf{R}}\left| (2\overline{%
u}\phi +\phi ^{2})^{n-2}(\overline{u}+\phi )^{2}\right| \\
& \leq \mathrm{const.}\sup_{x\in \mathbf{R}}\left| (2t^{-\gamma }\frac{%
\overline{u}}{y}z\phi +\phi ^{2})^{n-2}(t^{-\gamma }z\frac{\overline{u}}{y}%
+\phi )^{2}\right| \\
& \leq \mathrm{const.}~\sup_{x\in \mathbf{R}}\left| 2t^{-\gamma }\frac{%
\overline{u}}{y}z^{1+2/(n-2)}\phi +\left( z^{2/(n-2)}\phi \right) ~\phi
\right| ^{n-2}~\left| t^{-\gamma }\frac{\overline{u}}{y}\right| ^{2} \\
& \hspace{3cm}+\mathrm{const.}~\sup_{x\in \mathbf{R}}\left| 2t^{-\gamma }%
\frac{\overline{u}}{y}z\phi +\phi ^{2}\right| ^{n-2}~\left( 2t^{-\gamma
}\left| \frac{\overline{u}}{y}\right| \left| z\phi \right| +\left| \phi
\right| ^{2}\right) \\
& \leq \mathrm{const.}~t^{-2(n-2)\gamma -2\gamma }~.
\end{align*}
\end{proof}
\section{Proof of the main result}
For functions $f$ in $\mathcal{J}=L_{1}(\mathbf{R)}\cap L_{\infty }(\mathbf{%
R)}$ we use the norms $\left\| f\right\| _{1}=\int |f(x)|~dx$, $\left\|
f\right\| _{\infty }=\sup_{x\in \mathbf{R}}|f(x)|$ and $\left\| f\right\|
=\left\| f\right\| _{1}+\left\| f\right\| _{\infty }$, and we denote by $%
\mathcal{B}$ the Banach space of functions $\varphi $ in $L_{\infty
}([1,\infty ))\times \mathcal{J}$ for which the norm $\left\| ~~\right\| _{%
\mathcal{B}}$,
\begin{equation*}
\left\| \varphi \right\| _{\mathcal{B}}=\sup_{t\geq 1}t^{4\gamma }\left\|
\varphi (\sqrt{t}~.~,t)\right\| ~,
\end{equation*}
is finite. Let $\tau _{0}$ as in Proposition \ref{propvodd} and $\tau _{1}$
as in Proposition \ref{propV}, and consider, for fixed $\tau >\max \{\tau
_{0},\tau _{1}\},$ functions $\psi $ of the form
\begin{equation*}
\psi (x,t)=\tau ^{-4\gamma }\varphi (x/\sqrt{\tau },t/\tau )~,
\end{equation*}
with $\varphi \in \mathcal{B}.$ Let $K$ be the fundamental solution of the
differential operator $\partial _{t}-\partial _{x}^{2}-\tau V(\sqrt{\tau }%
x,\tau t),$ and let, for given $\nu \in \mathcal{J}$, the map $\mathcal{R}$
be defined by the equation
\begin{equation*}
\mathcal{R}(\varphi )(x,t)=\varphi _{0,1}(x,t)+\varphi _{0,2}(x,t)+\mathcal{N%
}(\varphi )(x,t)~,
\end{equation*}
where
\begin{align*}
\varphi _{0,1}(x,t)& =\int_{\mathbf{R}}K(x,t;y,1)~\nu (y)~dy~, \\
\varphi _{0,2}(x,t)& =\tau ^{4\gamma }\tau \int_{1}^{t}ds\int_{\mathbf{R}%
}dy~K(x,t;y,s)~I(\sqrt{\tau }y,\tau s)~,
\end{align*}
and where
\begin{equation*}
\mathcal{N}(\varphi )(x,t)=\sum_{p=1}^{n}\sum_{q=0}^{p}\mathcal{N}%
_{p,q}(\varphi )(x,t)~,
\end{equation*}
with
\begin{equation*}
\mathcal{N}_{p,q}(\varphi )(x,t)=\tau ^{4\gamma }\tau \int_{1}^{t}ds\int_{%
\mathbf{R}}dy~K(x,t;y,s)~V_{p,q}(\sqrt{\tau }y,\tau s)~\tau ^{-4\gamma
(p+q)}\varphi (y,s)^{p+q}~.
\end{equation*}
The integral equation $\varphi =\mathcal{R}(\varphi )$ is equivalent to the
differential equation (\ref{eqpsi}) with initial condition $\psi
_{0}(x)=\psi (x,\tau )=\tau ^{-4\gamma }\nu (x/\sqrt{\tau }).$ We note that,
since the function $V$ is positive, the kernel $K$ is bounded pointwise by
the fundamental solution $K_{0}$ of the heat equation,
\begin{equation}
K_{0}(x,t;y,s)=\frac{1}{\sqrt{4\pi }}\frac{1}{\sqrt{t-s}}\exp \left( -\frac{1%
}{4}\frac{(x-y)^{2}}{(t-s)}\right) ~.
\end{equation}
The following Proposition makes Theorem \ref{main} precise.
\begin{proposition}
\label{maint}Let $\beta \geq \max \left\{ 1,~3~c_{I}\int_{0}^{1}\left( 1+%
\frac{1}{\sqrt{1-s}}\right) \frac{ds}{s^{1/2+4\gamma }}\right\} ,$ with $%
c_{I}$ as defined in (\ref{eqn:thminhomo}), and let $\tau $ be sufficiently
large. Then, for all $\nu \in \mathcal{J}$ with $\left\| \nu \right\| <\beta
/6$, the equation $\varphi =\mathcal{R}(\varphi )$ has a unique solution $%
\varphi ^{\ast }$ in the ball $\mathcal{U}(\beta )=\{\varphi \in \mathcal{B}%
| $ $\left\| \varphi \right\| _{\mathcal{B}}<\beta \}.$
\end{proposition}
\begin{proof}
Since $4\gamma <1/2,$ the solution of the integral equation will be
dominated by $\varphi _{0,2},$ and, as we will see, $\beta $ has been chosen
such that $\left\| \varphi _{0,2}\right\| _{\mathcal{B}}\leq \beta /3.$ The
idea is therefore to show that, if $\tau $ is large enough to make the
nonlinear part of the map $\mathcal{R}$ small, and if $\left\| \nu \right\|
<\beta /6$, then the map $\mathcal{R}$ contracts the ball $\mathcal{U}(\beta
)$ into itself, which by the contraction mapping principle implies the
theorem. We first show that $\mathcal{R}$ maps the ball $\mathcal{U}(\beta )$
into itself. For the contribution coming from the initial condition we have
\begin{equation*}
\left\| \varphi _{0,1}(\sqrt{t}~.~,t)\right\| \leq \frac{2}{\sqrt{t}}\left\|
\nu \right\| ~,
\end{equation*}
and therefore
\begin{equation*}
\left\| \varphi _{0,1}\right\| _{\mathcal{B}}\leq 2\left\| \nu \right\|
<\beta /3~.
\end{equation*}
We next estimate the norm $\left\| \varphi _{0,2}\right\| _{\mathcal{B}}.$
Let $c(t,s)=\frac{1}{\sqrt{t}}+\frac{1}{\sqrt{t-s}}.$ Then,
\begin{align*}
\left\| \varphi _{0,2}(\sqrt{t}~.~,t)\right\| & \leq \tau ^{4\gamma }\tau
\int_{1}^{t}ds~c(t,s)\int_{\mathbf{R}}dy~\left| I(\sqrt{\tau }y,\tau
s)\right| \\
& =\tau ^{4\gamma }\tau \int_{1}^{t}\sqrt{s}~c(t,s)~ds~\int_{\mathbf{R}%
}dx~\left| I(\sqrt{\tau s}x,\tau s)\right| \\
& \leq c_{I}~\tau ^{4\gamma }\tau \int_{1}^{t}\sqrt{s}~c(t,s)~ds~\left( \tau
s\right) ^{-(1+4\gamma )} \\
& \leq c_{I}~t^{-4\gamma }\int_{0}^{1}c(1,s)~\frac{ds}{s^{1/2+4\gamma }} \\
& \leq \frac{\beta }{3}t^{-4\gamma }~,
\end{align*}
and therefore
\begin{equation*}
\left\| \varphi _{0,2}\right\| _{\mathcal{B}}<\beta /3~.
\end{equation*}
It remains to be shown that the nonlinearity is also bounded by $\beta /3,$
for $\tau $ large enough. For $\varphi \in \mathcal{U}(\beta )$ we have,
\begin{equation}
\left\| \mathcal{N}(\varphi )(\sqrt{t}~.~,t)\right\| \leq \mathrm{const.}%
~\tau ^{4\gamma }\tau \int_{1}^{t}c(t,s)~\sqrt{s}~ds\sum_{p=1}^{n}%
\sum_{q=0}^{p}(\tau s)^{-e(p,q)}~s^{-4\gamma (p+q)}\tau ^{-4\gamma
(p+q)}\left\| \varphi \right\| _{\mathcal{B}}^{p+q}~. \label{normcase}
\end{equation}
For $(p,q)=(1,0)$ we get, since $\delta _{1}\equiv \gamma (n-1)(\delta
^{\prime }+1)-1>0,$
\begin{align*}
\left\| \mathcal{N}_{1,0}(\varphi )(\sqrt{t}~.~,t)\right\| & \leq \mathrm{%
const.}~\tau ^{-\delta _{1}}\beta \int_{1}^{t}c(t,s)~\sqrt{s}%
~ds~s^{-1-\delta _{1}-4\gamma } \\
& \leq \mathrm{const.}~\tau ^{-\delta _{1}}\beta ~t^{-4\gamma }\int_{0}^{1}%
\frac{c(1,s)}{s^{1/2+4\gamma }}~ds~,
\end{align*}
and for $(p,q)=(2,0)$ we get
\begin{align*}
\left\| \mathcal{N}_{2,0}(\varphi )(\sqrt{t}~.~,t)\right\| & \leq \mathrm{%
const.}~\tau ^{4\gamma +1-8\gamma -(2\gamma (n-2)+2\gamma )}\beta
^{2}\int_{1}^{t}c(t,s)~\sqrt{s}~ds~s^{-8\gamma -2\gamma -2\gamma (n-2)} \\
& \leq \mathrm{const.}~\tau ^{-\gamma }\beta ^{2}~t^{-4\gamma }\int_{0}^{1}%
\frac{c(1,s)}{s^{1/2+4\gamma }}~ds~,
\end{align*}
and for the other cases we have
\begin{align*}
\left\| \mathcal{N}_{p,q}(\varphi )(\sqrt{t}~.~,t)\right\| & \leq \mathrm{%
const.}~\tau ^{4\gamma +1-2\gamma (n-p)-4\gamma p-4\gamma q}\beta
^{2n}\int_{1}^{t}c(t,s)~\sqrt{s}~ds~s^{-2\gamma (n-p)-4\gamma p-4\gamma q} \\
& \leq \mathrm{const.}~\tau ^{-\gamma }\beta ^{2n}~t^{-4\gamma }\int_{0}^{1}%
\frac{c(1,s)}{s^{1/2+4\gamma }}~ds~,
\end{align*}
and therefore $\left\| \mathcal{N}(\varphi )\right\| _{\mathcal{B}}\leq
\beta /3$ if $\tau $ is large enough. Using the triangle inequality we get
that $\left\| \mathcal{R}(\varphi )\right\| _{\mathcal{B}}\leq \beta ,$
which proves that $\mathcal{R}\left( \mathcal{U}(\beta )\right) \subset
\mathcal{U}(\beta )$ as claimed. We now show that $\mathcal{R}$ is
Lipschitz. Let $\varphi _{1}$ and $\varphi _{2}$ be in $\mathcal{U}(\beta ).$
We have
\begin{align*}
\left\| \mathcal{N}(\varphi _{1})(\sqrt{t}~.~,t)-\mathcal{N}(\varphi _{2})(%
\sqrt{t}~.~,t)\right\| & \leq \mathrm{const.}~\tau ^{4\gamma }\tau
\int_{1}^{t}ds~c(t,s)~\sqrt{s}\cdot \\
& \sum_{p=1}^{n}\sum_{q=0}^{p}(\tau s)^{-e(p,q)}~s^{-4\gamma (p+q)}\tau
^{-4\gamma (p+q)}~\beta ^{p+q-1}~\left\| \varphi _{1}-\varphi _{2}\right\| _{%
\mathcal{B}}~,
\end{align*}
and therefore we get, using the same estimates as for (\ref{normcase}), that
\begin{equation*}
\left\| \mathcal{R}(\varphi _{1})-\mathcal{R}(\varphi _{2})\right\| _{%
\mathcal{B}}=\left\| \mathcal{N}(\varphi _{1})-\mathcal{N}(\varphi
_{2})\right\| _{\mathcal{B}}\leq \frac{1}{2}\left\| \varphi _{1}-\varphi
_{2}\right\| _{\mathcal{B}}~,
\end{equation*}
provided $\tau $ is large enough. This completes the proof of Theorem \ref
{maint}.
\end{proof}
\section{Appendix}
\subsection{Proof of Proposition \ref{theoremeta}}
We first prove the existence of a unique positive solution of equation (\ref
{eqeta}) satisfying the boundary conditions (\ref{etainfinity}) and (\ref
{etazero}). Then, we derive the results on the asymptotic behavior near zero
and infinity.
\subsubsection{Existence of the function $\protect\eta\label{seta777}$}
\begin{proposition}
Let, for $\rho >0$, $\eta _{\rho }$ be the solution of the initial value
problem on $\mathbf{R}_{+}$,
\begin{align}
\eta ^{\prime \prime }& =(2\kappa z\eta +\eta ^{2})^{n}~, \label{1} \\
\eta ^{\prime }(0)& =-\kappa ~, \notag \\
\eta (0)& =\rho >0~. \notag
\end{align}
Then, there exists a unique $\bar{\rho}$ such that the function $\eta _{%
\overline{\rho }}$ is positive and satisfies $\lim_{x\rightarrow \infty
}\eta _{\bar{\rho}}(x)=0$.
\end{proposition}
\begin{proof}
We first prove that $\overline{\rho}$ is unique. Given a function $\eta$
from $\mathbf{R}_{+}$ to $\mathbf{R}$ we define the function $\mathcal{F}%
(\eta),$ $\mathcal{F}\left( \eta\right) (z)=(\kappa z\eta+\eta^{2})^{n}$.
Assume that there are two values $\rho_{1}>\rho_{2}>0,$ such that the
functions $\eta _{1}\equiv\eta_{\rho_{1}}$ and $\eta_{2}\equiv\eta_{%
\rho_{2}} $ are positive and satisfy $\lim_{x\rightarrow\infty}\eta_{1}(x)=$
$\lim_{x\rightarrow\infty }\eta_{2}(x)=0.$ We first show that the function $%
\eta_{12}=\eta_{1}-\eta_{2}$ is positive for all $x\geq0.$ Namely, if we
assume the contrary, then because $\eta_{12}(0)>0$, there must be a first $%
x_{0}>0$ such that $\eta_{12}(x_{0})=0.$ Furthermore, if $\eta_{12}(x)>0$
then $\eta_{12}^{\prime\prime }(x)=\mathcal{F}\left( \eta_{1}\right) (x)-%
\mathcal{F}\left( \eta _{2}\right) (x)>0,$ and therefore $%
\eta_{12}(x_{0})=\rho_{1}-\rho_{2}+\int_{0}^{x_{0}}dx\int_{0}^{x}dy~%
\eta_{12}^{\prime\prime}(y)>0,$ a contradiction. Therefore $\eta_{12},$ and
as a consequence $\eta_{12}^{\prime\prime},$ are positive for all $x,$ from
which it follows that $\lim_{x\rightarrow\infty}\eta_{12}(x)>0,$ in
contradiction with $\lim_{x\rightarrow\infty}\eta_{12}(x)=\lim_{x\rightarrow%
\infty}\eta _{1}(x)-\lim_{x\rightarrow\infty}\eta_{2}(x)=0.$
To prove the existence of a $\bar{\rho}$ for which $\eta _{\overline{\rho }}$
is positive and for which $\lim_{x\rightarrow \infty }\eta _{\bar{\rho}%
}(x)=0 $, we use the so called shooting method. Note that, for any $\rho >0$%
, the initial value problem (\ref{1}) has a unique solution $\eta _{\rho },$
and since $\eta _{\rho }^{\prime }(0)=-\kappa ,$ the function $\eta _{\rho }$
is strictly decreasing on $[0,x_{\rho })$ for $x_{\rho }$ small enough. We
will show that for small enough $\rho >0$, the graph of $\eta _{\rho }$
intersects the real axis and $\eta _{\rho }$ becomes negative, whereas for $%
\rho $ large enough, $\eta _{\rho }$ has a minimum and then diverges to plus
infinity. The (unique) point between those two sets is $\bar{\rho}$. Define
the two subsets $I_{1}$ and $I_{2}$ of $\mathbf{R}_{+},$
\begin{align*}
I_{1}& =\{\rho \in \mathbf{R}_{+}|~\exists ~x_{1},\eta _{\rho }(x_{1})=0%
\text{{\ \textrm{and\ }}}\eta _{\rho }(x)>0{\ \mathrm{for\ }}x\in \lbrack
0,x_{1})\}~, \\
I_{2}& =\{\rho \in \mathbf{R}_{+}|~\exists ~x_{2},\eta _{\rho }^{\prime
}(x_{2})=0{\ \mathrm{and\ }}\eta _{\rho }^{\prime }(x)<0,\eta _{\rho }(x)>0{%
\ \mathrm{for\ }}x\in \lbrack 0,x_{2})\}~.
\end{align*}
We note that if $\eta _{\rho }^{\prime }(x_{0})=0$ and $\eta _{\rho
}(x_{0})>0,$ for some $x_{0}$, then $\eta _{\rho }^{\prime }>0$ on any
interval $(x_{0},x)$ on which $\eta _{\rho }$ is defined, and a function $%
\eta _{\rho }$ with $\rho \in I_{2}$ can therefore not converge to zero at
infinity. Furthermore, since the function $\eta \equiv 0$ is a solution of
the differential equation (\ref{1}), it follows, since solutions are unique,
that $\eta _{\rho }(x_{0})>0$ if $\eta _{\rho }^{\prime }(x_{0})=0,$ and
therefore the intersection of $I_{1}$ with $I_{2}$ is empty. The sets $I_{1}$
and $I_{2}$ are open, by continuity of the solution $\eta _{\rho }$ as a
function of the initial data $\rho $. We now show that $I_{1}$ is non empty
and bounded, which shows that $\bar{\rho}\equiv \sup I_{1}<\infty .$ This $%
\bar{\rho}$ is neither in $I_{1}$ nor in $I_{2}$, and therefore the function
$\eta _{\bar{\rho}}$ is at the same time strictly positive and strictly
decreasing, and therefore $\lim_{x\rightarrow \infty }\eta _{\bar{\rho}%
}(x)=0 $. To prove that $I_{1}$ is non empty, we fix any $\rho _{1}$
positive and choose $x_{0}>0$ small enough such that on $[0,x_{0}]$ the
solution $\eta _{1}\equiv \eta _{\rho _{1}}$ exists and is strictly
decreasing. Then, $\rho _{1}-\eta _{1}(x_{0})>0.$ Choose now $0<\rho
_{2}<\rho _{1}-\eta _{1}(x_{0})$ and let $\eta _{2}\equiv \eta _{\rho _{2}}$
be the corresponding solution. As before, we have that the function $\eta
_{12}=\eta _{1}-\eta _{2},$ and its second derivative $\eta _{12}^{\prime
\prime },$ are positive on the interval $[0,x_{0}),$ and therefore, since $%
\eta _{2}(x_{0})=$ $\rho _{2}+\int_{0}^{x_{0}}dx\int_{0}^{x}dy~\eta
_{2}^{\prime \prime }(y)$ $=\rho _{2}+\eta _{1}(x_{0})-\rho
_{1}-\int_{0}^{x_{0}}dx\int_{0}^{x}dy~\eta _{12}^{\prime \prime }(y),$ we
find that $\eta _{2}(x_{0})<\rho _{2}-\rho _{1}+\eta _{1}(x_{0}).$ Using the
definition of $\rho _{2}$ we therefore find that $\eta _{2}(x_{0})<0.$
Therefore $\rho _{2}\in I_{1}$. We now prove that $I_{1}$ is bounded. For $%
\rho >0$, let $x_{\rho }$ be the largest value (possibly infinite) such that
on $[0,x_{\rho })$ the solution $\eta _{\rho }$ exists and is strictly
positive. Then, $\eta _{\rho }^{\prime \prime }=\mathcal{F}\left( \eta
_{\rho }\right) $ is positive on $(0,x_{\rho })$ and, therefore $\eta _{\rho
}(x)>\rho -\kappa x$ for $x\in (0,x_{\rho })$. As a consequence, if the
function $\eta _{\rho }$ exists on $[0,\rho /\kappa ],$ then $x_{\rho }\geq
\rho /\kappa .$ Using again that $\eta _{\rho }(x)>\rho -\kappa x$ we then
find that $\eta _{\rho }(x)>\rho /2$ for $x\in \lbrack 0,\rho /2\kappa ]$,
and therefore $\mathcal{F}\left( \eta _{\rho }\right) >(\rho /2)^{2n}$ on $%
[0,\rho /2\kappa ]$, which implies that $\eta _{\rho }^{\prime }(\rho
/2)>-\kappa +(\rho /2)^{2n+1},$ which is positive if $\rho >2\kappa
^{1/2n+1}.$ Therefore $\eta ^{\prime }(x)$ must be equal to zero for some $%
x<\rho /\kappa .$ Any such $\rho $ therefore belongs to $I_{2}.$ If the
function $\eta _{\rho }$ ceases to exist before $x=\rho /\kappa $ it must
have been diverging to plus infinity for some $x<\rho /\kappa $ which again
implies that $\eta _{\rho }^{\prime }(x)$ must have been equal to zero for
some $x<\rho /\kappa ,$ and the corresponding $\rho $ is in $I_{2}.$
\end{proof}
\subsubsection{Asymptotic behavior of the function $\protect\eta$}
The function $\eta $ is regular at zero, and the coefficients of its Taylor
series at zero can be computed recursively. We have $\eta (0)=\eta _{0}>0$
and $\eta ^{\prime }(0)=-\kappa ,$ and therefore we get using the
differential equation that $\eta _{2}=\eta ^{\prime \prime }(0)/2=\eta
_{0}^{2n}/2,$ $\eta ^{\prime \prime \prime }(0)=0$ and $\eta _{4}=-\eta
^{iv}(0)/4!=\frac{n}{12}\eta _{0}^{2n-2}(\kappa ^{2}-\eta _{0}^{2n+1}).$ The
asymptotic behavior of $\eta $ at infinity is obtained as follows. Assuming
that $\eta $ behaves like $\lambda /z^{\delta }$ at infinity we get from the
differential equation that $\delta $ and $\lambda $ are as defined in (\ref
{defdelta007}) and (\ref{deflambda007}), respectively. That this is indeed
the correct leading behavior of $\eta $ at infinity can now be proved by
using standard techniques based on repeated applications of l'H\^{o}pital's
rule. See for example \cite{Brezis}. Since the proof is simple, but lengthy
and quite uninteresting, we do not give the details here.
Once the leading behavior of $\eta $ at infinity has been established we
make the ansatz $\eta (z)=\lambda z^{-\delta }+s(z)$. To leading order we
get for the function $s$ the linear equation
\begin{equation}
s^{\prime \prime }-\frac{n}{\lambda }(2\kappa \lambda )^{n}z^{-2}s=n(2\kappa
\lambda )^{n-1}\lambda ^{2}z^{-3-2\delta }~. \label{eqs}
\end{equation}
There is a certain constant $\lambda _{p},$ such that the function $s_{p},$ $%
s_{p}(z)=\lambda _{p}z^{-1-2\delta }$ is a particular solution of equation (%
\ref{eqs}). The solutions of the homogeneous equation associated with (\ref
{eqs}) are of the form $s_{h}^{\pm }(z)=z^{p_{\pm }}$, and using the
definition (\ref{deflambda007}) for $\lambda $ we find that
\begin{equation}
p_{\pm }=\frac{1}{2}\left( 1\pm \sqrt{1+4n\delta (\delta +1)}\right) ~.
\label{eqn:critexp}
\end{equation}
For $n\geq 6$ we have that $\left| p_{-}\right| >2\delta +1,$ and the
asymptotic behavior of $s$ is therefore for $n\geq 6$ of the form $\lambda
_{\infty }/z^{2\delta +1},$ with $\lambda _{\infty }=\lambda _{p}$, and of
the form $\lambda _{\infty }/z^{\left| p_{-}\right| }$ with some unknown
coefficient $\lambda _{\infty }$ for $n\leq 5$. It is tedious, but not
difficult, to prove that this is indeed the correct second order behavior of
$\eta $ at infinity. We omit the details.
\subsection{Proof of Proposition \ref{cmu2}}
In order to study equation (\ref{mu2}) with boundary conditions (\ref{bmu2-1}%
) and (\ref{bmu2-2}), we make the ansatz $\mu_{2}(x)=m(x)/x^{\delta}.$ For
the function $m$ we get the differential equation
\begin{equation}
m^{\prime\prime}+(\frac{x}{2}-\frac{2\delta}{x})m^{\prime}+\left(
\delta(\delta+1)\frac{1}{x^{2}}-\left( \frac{\delta}{2}-\varepsilon\right)
\right) m=\frac{1}{x^{2}}(2\frac{\mu_{1}(x)}{x}m)^{n}~, \label{eqh}
\end{equation}
and the boundary conditions for $m$ are
\begin{align}
\lim_{x\rightarrow0}m(x) & =\lambda~, \label{bcm0} \\
\lim_{x\rightarrow\infty}m(x)x^{2\varepsilon-\delta} & =0~. \label{bcm1}
\end{align}
\subsubsection{Asymptotic behavior of the function $\protect\mu_{2}$}
As indicated in Section \ref{smu2}, a solution of equation (\ref{mu2}) that
is defined on $\mathbf{R}_{+}$ behaves at infinity either like $%
x^{-2\varepsilon }$ or like $\exp (-x^{2}/4)/x^{1-2\varepsilon }.$ The proof
is similar to the one in \cite{Brezis}. We omit the details. Given the
asymptotic behavior of $\mu _{2}$ at infinity, we find for the function $m$
at infinity either a behavior proportional to $x^{\delta -2\varepsilon }$,
or a behavior proportional to $x^{5\varepsilon }\exp (-x^{2}/4).$ Since $%
\delta -2\varepsilon >0,$ we find that
\begin{equation}
\lim_{x\rightarrow \infty }m(x)=0~, \label{bcm3}
\end{equation}
if and only if the boundary condition (\ref{bcm1}) is satisfied, and we will
impose (\ref{bcm3}) from now on. We now discuss the asymptotic behavior of
the function $m$ near zero. From equation (\ref{eqh}) we see that $m^{\prime
\prime }(0)$ exists if and only if $\delta (\delta +1)m(0)=\left( \kappa
m(0)\right) ^{n}$, i.e., if $m(0)=\lambda ,$ and if $m^{\prime }(0)=0.$ We
then find, that $m^{\prime \prime }(0)/2=\lambda _{0}$, with $\lambda _{0}$
as defined in (\ref{lambda0}). By taking derivatives of equation (\ref{eqh})
we find that $m^{\prime \prime \prime }(0)=0$, and that $m^{iv}(0)/4!=%
\lambda _{1},$ for some constant $\lambda _{1}\neq 0.$ By taking further
derivatives, one can recursively compute the Taylor coefficients of a
solution $m_{0}$ of equation (\ref{eqh}) that is regular (in fact, analytic)
in a neighborhood of zero. The solution $m_{0}$ does however not satisfy the
boundary condition (\ref{bcm3}). The solution of (\ref{eqh}) that does
satisfy (\ref{bcm3}) is of the form
\begin{equation}
m(x)=m_{0}(x)+x^{p}m_{1}(x)~, \label{msing}
\end{equation}
where $p=p_{+}+\delta ,$ with $p_{+}$ as defined in (\ref{eqn:critexp}).
Here, $m_{1}(x)=m_{1}(0)+\dots $, with $m_{1}^{\prime }(0)=0,$ and with $%
m_{1}(0)$ to be determined. The asymptotic form (\ref{msing}) can be
obtained by substituting the ansatz (\ref{msing}) for $m$ into equation (\ref
{eqh}). Since $p>7$ we find from (\ref{msing}) that near zero $%
m_{0}(x)+x^{p}m_{1}(x)=\lambda +\lambda _{0}x^{2}+\lambda _{1}x^{4}+\dots $
. We omit the details of the proof that the asymptotic behavior is as
indicated.
\subsubsection{Existence of the function $\protect\mu_{2}$}
We now prove the existence of a function $m$ that satisfies equation (\ref
{eqh}) with the boundary conditions (\ref{bcm0}) and (\ref{bcm3}). Since the
second derivative of the solution $m$ at zero is positive, and since $m$
converges to zero at infinity, there must be a first $\xi\in\mathbf{R}_{+}, $
such that $m^{\prime}(\xi)=0.$ The basic idea is now to use this position $%
\xi,$ and the value $\rho$ of $m$ at $\xi$, as parameters in shooting
arguments towards zero and infinity. The first shooting argument will allow
us to define a curve $c_{0}$ of initial conditions $(\xi,\rho)$ in $\mathbf{R%
}_{+}^{2},$ for which the boundary condition at zero is satisfied, and the
second shooting argument will allow us to find on this curve an initial
condition for which the boundary condition at infinity is satisfied as well.
So, let $(\xi ,\rho )$ be an initial condition. Locally, i.e., near $\xi ,$
there exists a solution $m_{\xi ,\rho }$ of equation (\ref{eqh}). By
definition, $m_{\xi ,\rho }(\xi )=\rho ,$ $m_{\xi ,\rho }^{\prime }(\xi )=0,$
and therefore we get for the second derivative of $m_{\xi ,\rho }$ at $\xi ,$
\begin{equation*}
m_{\xi ,\rho }^{\prime \prime }(\xi )=\omega _{1}(\xi )\rho ^{n}+\omega
_{2}(\xi )\rho ~,
\end{equation*}
where
\begin{equation}
\omega _{1}(\xi )=\frac{\left( 2\frac{\mu _{1}(\xi )}{\xi }\right) ^{n}}{\xi
^{2}}~, \label{omega1}
\end{equation}
and
\begin{equation}
\omega _{2}(\xi )=\frac{n}{2}\varepsilon -\frac{\delta (\delta +1)}{\xi ^{2}}%
~. \label{omega2}
\end{equation}
For initial conditions such that $\rho =c_{2}(\xi ),$ where
\begin{equation}
c_{2}(\xi )=\left( \frac{\frac{n}{2}\varepsilon }{\left( 2\frac{\mu _{1}(\xi
)}{\xi }\right) ^{n}}\left( \xi _{0}^{2}-\xi ^{2}\right) \right)
^{\varepsilon } \label{c2}
\end{equation}
and $\xi _{0}=\sqrt{\delta (\delta +1)/\left( \frac{n\varepsilon }{2}\right)
},$ we therefore have that $m_{\xi ,\rho }^{\prime \prime }(\xi )=0.$ See
Fig. 1 for the graph of the function $c_{2}.$ The function $c_{2}$ has a
maximum at the point $\xi _{m}$ that satisfies the equation
\begin{equation}
\omega _{1}^{\prime }(\xi _{m})c_{2}(\xi _{m})^{n-1}+\omega _{2}^{\prime
}(\xi _{m})=0~, \label{defxim}
\end{equation}
and the line $c_{2}$ divides the set of initial conditions into two subsets,
a subset $A$ where $m_{\xi ,\rho }^{\prime \prime }(\xi )<0,$ and a subset $%
B $ where $m_{\xi ,\rho }^{\prime \prime }(\xi )>0.$ For initial conditions
on $c_{2}$ we can compute $m_{\xi ,c_{2}(\xi )}^{\prime \prime \prime }(\xi
),$%
\begin{equation*}
m_{\xi ,c_{2}(\xi )}^{\prime \prime \prime }(\xi )=\omega _{1}^{\prime }(\xi
)c_{2}(\xi )^{n}+\omega _{2}^{\prime }(\xi )c_{2}(\xi )~.
\end{equation*}
Comparing with (\ref{defxim}) we find that $m_{\xi _{m},c_{2}(\xi
_{m})}^{\prime \prime \prime }(\xi _{m})=0,$ and we have that $m_{\xi
,c_{2}(\xi )}^{\prime \prime \prime }(\xi )<0$ for $0<\xi <\xi _{m}.$ We now
construct the line $c_{0}$ for $0<\xi <\xi _{m}.$
\begin{proposition}
Fix $\xi,$ $0<\xi<\xi_{m}.$ Then, there exists a unique number $c_{0}(\xi),$
$c_{2}(\xi)>$ $c_{0}(\xi)>\lambda,$ such that $m_{\xi,c_{0}(\xi)}$ is
positive and satisfies $\lim_{x\rightarrow0}m_{\xi,c_{0}(\xi)}(x)=\lambda.$
Furthermore, the function $c_{0}$ is continuous.
\end{proposition}
\begin{proof}
The proof is similar to the one in Section \ref{seta777}. Define the two
subsets $I_{1}$ and $I_{2}$ of the interval $I=(\lambda ,c_{2}(\xi ))$,
\begin{align*}
I_{1}& =\{\rho \in I|~\exists ~0<\xi _{1}<\xi ,~m_{\xi ,\rho }(\xi
_{1})=\lambda \text{ and }\lambda <m_{\xi ,\rho }(x)<c_{2}(x)\text{ for }%
x\in (\xi _{1},\xi )\}~, \\
I_{2}& =\{\rho \in I|~\exists ~0<\xi _{2}<\xi ,~m_{\xi ,\rho }(\xi
_{2})=c_{2}(\xi _{2})\text{ and }\lambda <m_{\xi ,\rho }(x)<c_{2}(x)\text{
for }x\in (\xi _{2},\xi )\}~.
\end{align*}
The intersection of $I_{1}$ with $I_{2}$ is by definition empty, and the
sets $I_{1}$ and $I_{2}$ are open, by continuity of the solution $m_{\xi
,\rho }$ as a function of the initial data $\rho $. We now show that all $%
\rho $ sufficiently close to $\lambda $ are in $I_{1},$ and that all $\rho $
sufficiently close to $c_{2}(\xi )$ are in $I_{2.}$ This implies that $%
c_{0}(\xi )=\sup I_{1}<c_{2}(\xi ),$ and $c_{0}(\xi )$ is neither in $I_{1}$
nor in $I_{2}$, and therefore the function $m_{\xi ,c_{0}(\xi )}$ satisfies $%
\lambda <m_{0}(x)<m_{2}(x)$ for all $0<x<\xi ,$ and therefore $%
\lim_{x\rightarrow 0}m_{\xi ,c_{0}(\xi )}(x)=\lambda ,$ since $%
\lim_{x\rightarrow 0}c_{2}(x)=\lambda .$ So let $(\xi ,\rho )$ be an initial
condition. Then, $m_{\xi ,\rho }$ satisfies the integral equation
\begin{equation}
m_{\xi ,\rho }(x)=\rho +\int_{\xi }^{x}\frac{dy}{p(y)}\int_{\xi
}^{y}p(z)~\Omega (m_{\xi ,\rho }(z),z)~dz~, \label{inteq}
\end{equation}
where
\begin{equation*}
p(z)=\frac{\exp (z^{2}/4)}{z^{2\delta }}~,
\end{equation*}
and where
\begin{equation*}
\Omega (s,z)=\omega _{1}(z)s^{n}+\omega _{2}(z)s~.
\end{equation*}
$\Omega (s,z)$ is strictly negative for $0<z<\xi _{m}$ and $s\approx \lambda
,$ and therefore we find, like in the proof in Section \ref{theoremeta} that
any solution with an initial condition $\rho $ sufficiently close to $%
\lambda $ will cross the line $m\equiv \lambda .$ Similarly, for an initial
condition $(\xi ,\rho )$ close to $(\xi ,c_{2}(\xi ))$ we can use that $%
\Omega (s,z)\approx $ $0,$ and that $\partial _{z}\Omega (c_{2}(z),z)$ is
strictly negative to show that the corresponding solution will cross the
line $c_{2}.$ This completes the proof of the existence of $c_{0}(\xi ).$ To
prove uniqueness it is sufficient to use that $\partial _{s}\Omega (s,z)>0$
for $(s,z)$ in the set $C$ (see Fig. 1), and to integrate the difference of
two solutions from their respective initial condition to zero, which leads
to a contradiction, since both solutions have to be equal to $\lambda $ at
zero. Finally, that $c_{0}$ is a continuous function follows from the
continuity of $m_{\xi ,\rho }$ as a function of $\rho $ and $\xi $ using the
uniqueness of $c_{0}(\xi ).$
\end{proof}
We now prove with a second shooting argument that solutions with initial
conditions $(\xi,c_{0}(\xi)),$ with $\xi\approx0,$ become negative somewhere
in the interval $(\xi,2),$ and that solutions with initial conditions $%
(\xi,c_{0}(\xi))$, with $\xi\approx\xi_{m},$ stay positive and diverge to
plus infinity.
\begin{proposition}
There exists a unique initial condition $(\xi^{\ast},c_{0}(\xi^{\ast}))$
such that the corresponding solution $m_{\xi^{\ast},c_{0}(\xi^{\ast})}$ is
positive and satisfies $\lim_{x\rightarrow\infty}m_{\xi^{\ast},c_{0}(\xi^{%
\ast})}(x)=0.$
\end{proposition}
\begin{proof}
Define the two subsets $I_{1}$ and $I_{2}$ of the interval $I=(0,\xi _{m})$,
\begin{align*}
I_{1}& =\{\xi \in I|~\exists ~\xi _{1}>\xi ,~m_{\xi ,c_{0}(\xi )}(\xi _{1})=0%
\text{ and }m_{\xi ,c_{0}(\xi )}(x)>0,\text{ }m_{\xi ,c_{0}(\xi )}^{\prime
}(x)<0\text{ for }x\in (\xi ,\xi _{1})\}~, \\
I_{2}& =\{\xi \in I|~\exists ~\xi _{2}>\xi ,~m_{\xi ,c_{0}(\xi )}^{\prime
}(\xi _{2})=0\text{ and }m_{\xi ,c_{0}(\xi )}(x)>0,\text{ }m_{\xi ,c_{0}(\xi
)}^{\prime }(x)<0\text{ for }x\in (\xi ,\xi _{2})\}~.
\end{align*}
By definition, the intersection of $I_{1}$ with $I_{2}$ is empty, and the
sets $I_{1}$ and $I_{2}$ are open, by continuity of the solution $m_{\xi
,c_{0}(\xi )}$ as a function of the initial data $\xi $. We now show that
all $\xi $ sufficiently close to $0$ are in $I_{1},$ and that all $\xi $
sufficiently close to $\xi _{m}$ are in $I_{2.}$ This implies that $\xi
^{\ast }=\sup I_{1}<\xi _{m},$ is neither in $I_{1}$ nor in $I_{2}$, and
therefore the function $m_{\xi ^{\ast },c_{0}(\xi \ast )}$ is positive and
decreasing for $x>\xi ^{\ast }$ which implies that $\lim_{x\rightarrow
\infty }m_{\xi ^{\ast },c_{0}(\xi ^{\ast })}(x)=0.$ So let $(\xi ,c_{0}(\xi
))$ be an initial condition with $0<\xi <x_{0},$ with $x_{0}\ll 1.$ The
proof that such an initial condition is in $I_{1}$ is rather lengthy and we
therefore do not give the details here, but on a heuristic level it is easy
to understand why such a solution is in $I_{1}$. Namely, near zero the
asymptotics of the solution $m_{\xi ,c_{0}(\xi )}$ is $m_{\xi ,c_{0}(\xi
)}(x)=\lambda +\lambda _{0}x^{2}+\dots +\left( m_{1}\right) _{\xi ,c_{0}(\xi
)}(0)x^{p}+\dots ,$ where $p=p_{+}+\delta ,$ and where $\left( m_{1}\right)
_{\xi ,c_{0}(\xi )}(0)$ is such that $m_{\xi ,c_{0}(\xi )}^{\prime }(\xi )=0$%
. Neglecting higher order terms we find that $m_{\xi ,c_{0}(\xi )}^{\prime
}(\xi )\approx 2\lambda _{0}\xi +\left( m_{1}\right) _{\xi ,c_{0}(\xi
)}(0)p\xi ^{p-1}$, and we conclude that $\left( m_{1}\right) _{\xi
,c_{0}(\xi )}(0)\approx -\left( 2\lambda _{0}/p\right) /\xi ^{p-2}.$
Therefore, $m_{\xi ,c_{0}(\xi )}^{\prime }(\xi ^{(p-5/2)/(p-1)})\approx
-2\lambda _{0}/\xi ^{1/2}\ll 0$ and $m_{\xi ,c_{0}(\xi )}(\xi
^{(p-5/2)/(p-1)})\approx \lambda ,$ if $\xi $ is small enough. Therefore,
since $m_{\xi ,c_{0}(\xi )}^{\prime \prime }(x)<0$ for all $0<x\ll 1,$ we
find that $m_{\xi ,c_{0}(\xi )}(x)<c_{0}(\xi )-2\lambda _{0}(x-\xi )/\xi
^{1/2},$ and therefore $m_{\xi ,c_{0}(\xi )}(x)=0$ for some $x\leq c_{0}(\xi
)\xi ^{1/2}/\left( 2\lambda _{0}\right) +\xi ,$ as claimed. Next, let $(\xi
,c_{0}(\xi ))$ be an initial condition with $\xi _{m}-x_{1}<\xi <\xi _{m},$
with $0<x_{1}<1$ to be chosen below. We now show that such an initial
condition is in $I_{2}.$ For all $\xi _{0}\geq \xi ^{\prime }\geq \xi $ we
have the lower bounds $m_{\xi ,c_{0}(\xi )}(\xi ^{\prime })\geq m_{\xi
,c_{0}(\xi )}(\xi )k_{1}$ and $p(\xi ^{\prime })m_{\xi ,c_{0}(\xi )}^{\prime
}(\xi ^{\prime })\geq $ $m_{\xi ,c_{0}(\xi )}(\xi )k_{2},$ where
\begin{equation*}
k_{1}=1+\delta (\delta +1)\frac{\exp (\xi _{0}^{2}/4)}{\exp (\xi ^{2}/4)}%
\int_{\xi _{m}-x_{1}}^{\xi _{0}}y^{2\delta }~dy\int_{\xi
_{m}-x_{1}}^{y}z^{-2\delta }~(\frac{1}{\xi _{0}^{2}}-\frac{1}{z^{2}})~dz~,
\end{equation*}
$k_{1}>0,$ and
\begin{equation*}
k_{2}=\delta (\delta +1)\exp (-\xi _{0}^{2}/4)\int_{\xi _{m}-x_{1}}^{\xi
_{0}}z^{-2\delta }~(\frac{1}{\xi _{0}^{2}}-\frac{1}{z^{2}})~dz~,
\end{equation*}
and for $x\geq \xi _{0}$ we therefore have the lower bound
\begin{equation*}
m_{\xi ,c_{0}(\xi )}(x)\geq m_{\xi ,c_{0}(\xi )}(\xi )\left(
k_{1}+k_{2}\int_{\xi _{0}}^{\infty }\frac{dy}{p(y)}\right) ~,
\end{equation*}
and it follows, using again the integral equation (\ref{inteq}), that $%
m_{\xi ,c_{0}(\xi )}$ diverges at (or before) infinity, that therefore $%
m_{\xi ,c_{0}(\xi )}^{\prime }(x)=0$ for some $x>\xi ,$ which implies that $%
\xi \in I_{2},$ provided
\begin{equation}
k_{1}+k_{2}\int_{\xi _{0}}^{\infty }\frac{dy}{p(y)}>0~. \label{cond}
\end{equation}
For $x_{1}$ small enough and for $n$ large enough (\ref{cond}) can be
verified without too much difficulty. With the help of a computer one can
show that (\ref{cond}) is satisfied for the remaining $n\geq 5.$ For $n=4$ (%
\ref{cond}) is not satisfied, since the above bounds on $m_{\xi ,c_{0}(\xi
)}(\xi _{0})$ and $m_{\xi ,c_{0}(\xi )}^{\prime }(\xi _{0})$ are too weak.
Sufficiently good bounds can be obtained by dividing the interval $(\xi ,\xi
_{0})$ in two pieces and by integrating lower bounds on each of the
subintervals. We omit the details. Finally, uniqueness of $\xi ^{\ast }$ can
be proved by integrating the difference of two solutions from $\xi _{0}$ to
infinity, which, using the positivity of $\partial _{s}\Omega (s,z),$ leads
to a contradiction with the fact that both of the solutions converge to zero
at infinity.
\end{proof}
\subsection{Proof of Proposition \ref{cfi22}}
We first proof the existence of a unique solution of equation (\ref{h}) with
the boundary conditions (\ref{bh-1}) and (\ref{bh-2}). Then, we derive the
results on the asymptotic behavior near zero and infinity.
\subsubsection{Existence of the function $\protect\varphi_{2}$}
The equation (\ref{h}) for $h$ is linear. We therefore first construct two
linearly independent solutions $h_{1}$ and $h_{2}$ for the corresponding
homogeneous equation, which we then use to construct, using standard
methods, a solution of (\ref{h}) that satisfies the boundary conditions (\ref
{bh-1}) and (\ref{bh-2}). The homogeneous equation is
\begin{equation}
h^{\prime \prime }-q~h=0~, \label{homh}
\end{equation}
where
\begin{equation}
q(z)=n\left( 2\kappa z\eta (z)+\eta (z)^{2}\right) ^{n-1}(2\kappa z+2\eta
(z))~. \label{defq}
\end{equation}
Since the equation (\ref{homh}) is linear, the integral equation for $h_{1}$%
,
\begin{equation}
h_{1}(x)=1+\int_{0}^{x}~dy\int_{0}^{y}q(z)~h_{1}(z)~dz~, \label{eqh1}
\end{equation}
has a positive solution that exists for all $x$ in $\mathbf{R}_{+}.$ By
definition, we have near $x=0$ the behavior $h_{1}(x)=1+\mathcal{O}(x^{2}).$
At infinity, the solution $h_{1}$ is asymptotic to a solution of the
equation
\begin{equation*}
h^{\prime \prime }(x)-\frac{n}{\lambda }(2\kappa \lambda )^{n}\frac{1}{x^{2}}%
h(x)=0~.
\end{equation*}
This equation is the same as the homogeneous part of equation (\ref{eqs}),
and the leading order behavior of $h_{1}$ at infinity is therefore either
proportional to $x^{p_{+}}$ or to $x^{p_{-}},$ with $p_{\pm }$ as defined in
(\ref{eqn:critexp}). Since $h_{1}$ is positive, we find using (\ref{eqh1}),
that $h_{1}(x)>1$ for all $x$ in $\mathbf{R}_{+},$ and therefore $h_{1}$ is
near infinity of the form $h_{1}(x)=d_{1}x^{p_{+}}+\dots ,$ for some
constant $d_{1}>0$. A second solution of the homogeneous equation (\ref{homh}%
) is
\begin{equation*}
h_{2}(x)=h_{1}(x)~\int_{0}^{x}\frac{1}{h_{1}(y)^{2}}~dy~.
\end{equation*}
Near $x=0$ we have that $h_{2}(x)=x+\dots ,$ and near infinity we find that
\begin{equation}
h_{2}(x)=h_{1}(x)\left( d-d_{2}x^{1-2p_{+}}+\dots \right) ~, \label{h2infty}
\end{equation}
where $d=\int_{0}^{\infty }1/h_{1}(y)^{2}~dy$, and $d_{2}=$ $\left(
1/d_{1}\right) ^{2}/\left( 2p_{+}-1\right) .$ We note that $%
h_{1}h_{2}^{\prime }-h_{1}^{\prime }h_{2}\equiv 1.$ Therefore, the function $%
h_{p}$,
\begin{equation*}
h_{p}(x)=c_{1}(x)~h_{1}(x)+c_{2}(x)~h_{2}(x)~,
\end{equation*}
where
\begin{align*}
c_{1}(x)& =-\int_{0}^{x}h_{2}(y)~f(y)dy~, \\
c_{2}(x)& =\int_{0}^{x}h_{1}(y)~f(y)dy~,
\end{align*}
and where
\begin{equation*}
f(x)=-\gamma \eta (x)-\alpha x\eta ^{\prime }(x)+n\left( 2\kappa x\eta
(x)+\eta (x)^{2}\right) ^{n-1}2\kappa _{3}x^{3}\eta (x)~,
\end{equation*}
satisfies equation (\ref{h}). Near zero, the function $f$ is of the form $%
f(x)=-\gamma \eta _{0}+\dots $, and therefore, using the behavior of $h_{1}$
and $h_{2}$ near zero, we find that $c_{1}$ is near zero of order $\mathcal{O%
}(x^{2})$, and $c_{2}$ is near zero of order $\mathcal{O}(x).$ The function $%
h_{p}$ is therefore of order $\mathcal{O}(x^{2})$ near zero. At infinity,
the function $f$ is of the form $f(x)=f_{\infty }x^{-\delta }+\dots ,$ where
$f_{\infty }=-\gamma \lambda +\alpha \lambda \delta +n(2\kappa \lambda
)^{n-1}2\kappa _{3}\lambda ,$ and therefore the function $c_{2}$ is near
infinity of the form $c_{2}(x)=d_{1}f_{\infty }~x^{p_{+}+1-\delta
}/(p_{+}+1-\delta )+\dots $, and $c_{1}$ is near infinity of the form $%
c_{1}(x)=-d~c_{2}(x)+h_{\infty }+d_{1}d_{2}f_{\infty }~x^{2-p_{+}-\delta
}/(2-p_{+}-\delta )+\dots $, for some constant $h_{\infty }.$ Using these
asymptotic behavior for $c_{1}$, $h_{1}$, $c_{2}$, and $h_{2},$ we find for
the function $h_{p}$ near infinity the behavior,
\begin{align}
h_{p}(x)& =(-d~c_{2}(x)+h_{\infty }+\frac{d_{1}d_{2}f_{\infty }}{%
(2-p_{+}-\delta )}x^{2-p_{+}-\delta }+\dots )h_{1}(x)+c_{2}(x)h_{1}(x)\left(
d-d_{2}x^{1-2p_{+}}+\dots \right) \notag \\
& =h_{\infty }h_{1}(x)-\frac{d_{1}f_{\infty }}{p_{+}+1-\delta }%
d_{1}d_{2}x^{p_{+}+1-\delta }x^{p_{+}}x^{1-2p_{+}}+\frac{d_{1}d_{2}f_{\infty
}}{2-p_{+}-\delta }d_{1}x^{2-p_{+}-\delta }x^{p_{+}}+\dots \notag \\
& =h_{\infty }h_{1}(x)+\frac{f_{\infty }}{2p_{+}-1}\left( \frac{-1}{%
p_{+}+1-\delta }+\frac{1}{2-p_{+}-\delta }\right) x^{2-\delta }+\dots \notag
\\
& =h_{\infty }h_{1}(x)+\lambda _{0}x^{2-\delta }+\dots ~. \label{hpinfty}
\end{align}
In the last equality we have used the definition (\ref{lambda0}) for $%
\lambda _{0}.$ The function $h,$%
\begin{equation*}
h(z)=h_{p}(z)-h_{\infty }~h_{1}(z)~,
\end{equation*}
solves the equation (\ref{h}), satisfies the boundary condition (\ref{bh-1}%
), and since, as we show in the next section, the higher order terms in (\ref
{hpinfty}) converge to zero at infinity, it also satisfies the boundary
condition (\ref{bh-2}).
\subsubsection{Asymptotic behavior of the function $\protect\varphi_{2}$}
By construction, the leading behavior of $h$ at infinity is $h(z)=\lambda
_{0}z^{2-\delta }+\dots $. We therefore make the ansatz $h(z)=\lambda
_{0}z^{2-\delta }+k(z)$, and to leading order, we get for the function $k$
the linear equation
\begin{equation}
k^{\prime \prime }-n\delta (\delta +1)z^{-2}k=c_{k}z^{-\delta ^{\prime }}~,
\label{bvarphi2-2}
\end{equation}
for a certain constant $c_{k}$. The general solution of equation (\ref
{bvarphi2-2}) is
\begin{equation*}
k(z)=\frac{\lambda ^{\prime }}{z^{\delta ^{\prime }-2}}+\mathrm{const.}%
~z^{p_{-}}+\mathrm{const.}~z^{p_{+}}~,
\end{equation*}
with a certain constant $\lambda ^{\prime }$ and with $p_{+}$, $p_{-}$ as
defined in (\ref{eqn:critexp}). Since $\lim_{z\rightarrow \infty
}k(z)/z^{2-\delta }=0$ but $p_{+}>2-\delta ,$ the coefficient of the term
proportional to $z^{p_{+}}$ must be zero. Therefore, since$\left|
p_{-}\right| >\delta ^{\prime }-2$ for all $n,$ the asymptotic behavior of $%
k $ is always given by $\lambda ^{\prime }/z^{\delta ^{\prime }-2}.$ We omit
the details of the proof that this is indeed the correct second order
behavior of $k$ at infinity.\bigskip \bigskip \bigskip
\noindent \textbf{Acknowledgment}
During this work, A.S. was hosted by the University of Chicago and wishes to
thank the Department of Mathematics for its hospitality.
\bigskip
|
\section{Introduction}
It has been suggested recently \cite{add,aadd, add1} that the fundamental
scale of quantum gravity, $M_f$, can be as low as few TeV. The observed
weakness
of gravity is the result of $N$ ( $\geq 2$) new space dimensions in
which gravity can propagate
\footnote{In a different
context an attempt of lowering the string scale to TeV, without lowering
the fundamental Planck scale was considered in \cite{lyk}, based on an
earlier observation in \cite{witten}, see also
\cite{dienes} for lowering the GUT scale. Dynamical localization of
the fields
on a (solitonic) brane embedded in a higher dimensional space-time has
been suggested earlier in the field theoretic context
\cite{localization1/2}, \cite{localization1},\cite{ds}. For
some realizations of this scenario in the $D$-brane context see,
\cite{aadd}, \cite{bw}.}.
The observed (reduced) Planck scale, $M_{P} = (4\pi G_N)^{-1/2}
= 3.4 \cdot 10^{18}$ GeV, where $G_N$ is the Newton constant, is
then related to the reduced Planck scale in $4 + N$ dimensions,
$M_f$ (fundamental scale), by
\begin{equation}
M_{P} = M_f \sqrt{ M_f^N V_N}~,
\label{relation}
\end{equation}
where $V_N \equiv L_1 L_2.... L_N$ is the volume of the extra space,
and $L_i$ is the size of the $i$ - compact dimension.
For definiteness we will assume that the volume has a configuration of
torus in which case $L_i = 2 \pi R_i$, where $R_i$ $(i = 1, 2, ... N)$
are the radii of extra dimensions, so that
\begin{equation}
V_N = (2\pi)^N R_1 R_2 ... R_N~.
\label{volume}
\end{equation}
Using (\ref{relation}) and (\ref{volume}) we get
the constraint on the extra dimension radii:
\begin{equation}
(2\pi)^N R_1 R_2 ... R_N = \frac{M_P^2}{M_f^{N +2}}~.
\label{radii}
\end{equation}
(Notice that in some publications the factor $(2\pi)^N$ is removed
from this relation by redefinition of
the fundamental scale: $M_* = (2\pi)^{N/(N + 2)} M_f$.)
The phenomenological acceptance requires that $N \geq 2$,
since for $N = 1$ the radius would be of the solar system size.
According to present measurement the distance above which the Newtonian
law should not be changed is about 1 mm, and therefore
\begin{equation}
L_i = 2\pi R_i \leq 1~ {\rm mm}~.
\label{mmconstraint}
\end{equation}
For $N = 2$ and $M_f \sim$ TeV one gets from (\ref{relation},\ref{volume})
$R_1 \sim R_2 \sim 0.1$ mm which satisfies (\ref{mmconstraint})
Thus, in theories under consideration it is expected that the Newtonian
$1/r$ law breaks down at the scales smaller than the largest
extra dimension: $L_{max}$. The experimentally most exiting possibility
would be if $L_{max} \sim 1 - 10^{-2}$ mm, that is, in the range of
sensitivity of proposed experiments \cite{measurements}.
As we will argue in this paper the same range
is suggested by neutrino physics, namely, by a solution of the
solar neutrino problem based on existence of
new dimensions.
Usually it is assumed that
all large radii $R_i$ are of the same order of
magnitude. In such a case $ N > 2$ would be well out of sensitivity of any
planned sub-millimeter gravitational measurements.
On the other hand, for $N = 2$
the supernova analysis pushes the lower
bound on $M_f$ to $30$ TeV
\cite{add1} or even to
$50$ TeV \cite{supernova22} implying that $R < 0.01$ mm, which is
again beyond the planned experimental sensitivity.
However, in the absence of any commonly accepted mechanism for
stabilization of large radii\footnote{For some ideas in this direction
see \cite{nimaetal}.},
the requirement of their equality is
unjustified. In the present paper we will
assume that radii may take arbitrary values which satisfy
the fixed over-all volume (\ref{radii}) and phenomenological
(\ref{mmconstraint}) constraints. In such a case the
theory with several extra dimensions
still can be subject of sub-millimeter test, while avoiding
astrophysical and other laboratory bounds.
As we will see, these bounds are sensitive to the shape of extra dimensions.
In this paper we will discuss possible consequences
of the high-dimensional theories for neutrino physics.
In particular, we will suggest new
high-dimensional solution of the solar neutrino puzzle.
This solution
implies that the radius of at least one extra dimension must be
within $0.06 - 0.1 $ mm range.
This observation relies on new high dimension mechanism of neutrino mass
generation suggested in \cite{addm} which we will briefly recall.
According to the framework elaborated in \cite{add,aadd, add1},
all the standard model particles must be localized on a
3-dimensional hyper-surface
('brane') \cite{localization1/2,localization1,ds}
embedded in the bulk of $N$ large extra dimensions.
The same is true for {\it any other} state charged
under the standard model group. The argument is due to the
conservation of the gauge flux, which indicates that no state carrying
a charge under a gauge field localized on the brane, can exist
away from it\cite{ds}\cite{add}.
Thus, all the particles split in two categories:
those that live on the brane, and those
which exist everywhere, 'bulk modes'. Graviton belongs to the
second category. Besides the graviton there can be additional
neutral states propagating in the bulk. In general, the couplings between
the brane, $\psi_{brane}$,
and the bulk $\psi_{bulk}$
modes are suppressed by a volume factor:
\begin{equation}
\frac{1 }{ \sqrt{M_f^N V_N}} \psi_{brane} \psi_{brane}
\psi_{bulk}~.
\label{coupl}
\end{equation}
According to (\ref{relation}) the coupling constant in (\ref{coupl})
equals
\begin{equation}
\frac{M_f}{M_P} =
3 \cdot 10^{-16}\frac{M_f}{ 1 {\rm TeV}}
\end{equation}
and it does not depend on number of extra dimensions.
It was suggested \cite{addm} to use this small
model-independent coupling to explain the smallness of the neutrino mass.
The left handed neutrino, $\nu_L$, having weak isospin and
hypercharge must reside on the brane. Thus, it can get a
naturally small Dirac mass through the mixing with some bulk fermion
and the latter can be interpreted as the right-handed neutrino
$\nu_R$:
\begin{equation}
\frac{h M_f}{M_P} H \bar{\nu}_L \nu_R~.
\label{inter}
\end{equation}
Here $H$ is the Higgs doublet and $h$ is the model-dependent Yukawa
coupling.
After electro weak symmetry breaking the interaction (\ref{inter}) will
generate
the mixing mass
\begin{equation}
m_D = \frac{h v M_f}{M_P}~,
\label{Dmass}
\end{equation}
where $v$ is the VEV of $H$. For $M_f \sim 1$ TeV
and $h = 1$ this mass is about $5.6 \cdot 10^{-5}$ eV.
Being the bulk state, $\nu_R$ has a whole tower of
the Kaluza-Klein (KK) relatives.
For $N$ extra dimensions they can be labeled by a set of
$N$ integers ${n_1, n_2 ... n_N}$ (which determine momenta in extra
spaces):
$\nu_{n_1, n_2 ... n_N ~R}$.
Masses of these states are given by:
\begin{equation}
m_{n_1, n_2 ... n_N} = \sqrt{\sum_i \frac{n_i^2}{R_{i}^2}}~.
\end{equation}
Notice that the masses of the KK states are determined by the {\it radii}
whereas the scale of the Newton law modification
is given by the {\it size} of compact dimensions.
The left handed neutrino couples with all
$\nu_{n R}$ with the same mixing mass (\ref{Dmass}).
This mixing is possible due to the spontaneous breaking of the
translational invariance in the bulk by the brane.
In ref. \cite{ddg} a general case has been
considered with possible universal Majorana mass
terms for the bulk fermions.
Neutrino masses, mixings and vacuum
oscillations have been studied for various
sizes of mass parameters.
In this paper we continue to study the consequences of mixing of the
usual neutrino with bulk fermions in the context
elaborated in Ref. \cite{addm}.
We consider both the neutrino oscillations (in vacuum and medium)
and the resonance conversion.
We show that the resonance conversion
of the electron neutrinos to the bulk states can solve
the solar neutrino problem. This solution
implies $R \sim 0.1$ mm, thus giving connection between neutrino
physics and sub-millimeter gravity measurements.
We also discuss production of the KK- neutrino states in the Early
Universe and in supernovae.\\
\section{Neutrino mixing with the bulk modes. Universality}
Let us first assume that extra dimensions have the hierarchy of radii,
so that only one extra dimension has radius $R$ in sub-millimeter
range and therefore only one tower of corresponding Kaluza-Klein modes is
relevant for the low energy neutrino physics.
The number and the size of other dimensions will be chosen to
satisfy the constraint (\ref{radii}). (We will comment on effects
of two sub-millimeter dimensions in sect. 4 and 5.)
The right handed bulk states, $\nu_{i R}$, form with
the left handed bulk components,
$\nu_{i L}$, the Dirac mass terms
which originate from the quantized internal momenta in extra dimension:
\begin{equation}
\sum_{n = -\infty}^{+\infty} m_n~\bar{\nu}_{n R} \nu_{n L} + h.c.,
~~~~~~~~~
m_n \equiv \frac{n}{R}~.
\label{mass}
\end{equation}
The mass-split is determined by $1/R$.
According to (\ref{inter}, \ref{Dmass})
the bulk states mix with usual left handed neutrino
(for definiteness we will consider the electron neutrino
$\nu_{e L}$) by the Dirac type mass terms
with universal mass parameter:
\begin{equation}
m_D \sum_{n = -\infty}^{+\infty} \bar{\nu}_{n R} \nu_{e L}, ~~~~~~~~~
m_D \equiv \frac{hv M_f}{M_P} \approx
6 \cdot 10^{-5} {\rm eV} h \frac{M_f}{1 {\rm TeV}}~,
\label{mixing}
\end{equation}
where $h$ is the renormalized Yukawa coupling.
The mass terms (\ref{mass},\ref{mixing}) can be rewritten as
\begin{equation}
m_D \bar{\nu}_{0 R}\nu_{e L} +
m_D \sum_{n = 1}^{\infty} (\bar{\nu}_{n R} + \bar{\nu}_{- n R})\nu_{e L}
+ \sum_{n = 1}^{\infty} \frac{n}{R}~
(\bar{\nu}_{n R} \nu_{n L} - \bar{\nu}_{-n R} \nu_{-n L}) + h.c..
\label{mass1}
\end{equation}
Notice that $\nu_{0L}$ decouples from the system. Introducing new states:
\begin{equation}
\tilde \nu_{n L} = \frac{1}{\sqrt{2}}
(\nu_{n L} - \nu_{-n L})~, ~~~~~~~
\tilde \nu_{n R} = \frac{1}{\sqrt{2}}
({\nu}_{n R} + {\nu}_{-n R})
\label{newstates}
\end{equation}
and denoting by $\nu'_{n L}$, $\nu'_{n R}$ the orthogonal
combinations we can write the mass terms in (\ref{mass}) as
\begin{equation}
m_D \bar{\nu}_{0 R}\nu_{e L} +
\sqrt{2} m_D \sum_{n = 1}^{\infty}
\bar{\tilde \nu}_{n R} \nu_{e L}
+ \sum_{n = 1}^{\infty} \frac{n}{R}~
\left(\bar{\tilde \nu}_{n R} \tilde \nu_{n L} +
\bar{\nu}'_{n R} \nu'_{n L}\right)
+ h.c..
\label{mass2}
\end{equation}
Notice that the zero mode has
$\sqrt{2}$ smaller mixing mass with
$\nu_e$ than non-zero modes; the states
$\nu'_{n L}$, $\nu'_{n R}$ decouple from the rest of
the system.
Diagonalization of the mass matrix formed by the mass terms
(\ref{mass2}) in the limit of $ m_D R \ll 1$
gives (see Appendix) the mixing of
neutrino with
$n^{th}$ - bulk mode, $\tilde \nu_{n L}$:
\begin{equation}
\tan \theta_n \approx \frac{\sqrt{2} m_D}{m_n} = \frac{\xi}{n}~,
\label{tan}
\end{equation}
where
\begin{equation}
\xi \equiv \frac{\sqrt{2} h v M_f R}{ M_P}
\end{equation}
determines mixing with the first bulk mode.
The lightest state, $\nu_0$, has the mass $m_0 \approx m_D $ and others,
$\tilde{\nu}_n$,:
\begin{equation}
m_n \approx \frac{n}{R}~.
\end{equation}
According to (\ref{tan}) the electron
neutrino state can be written in terms of the mass eigenstates
as
\begin{equation}
\nu_e \approx \frac{1}{N} \left(\nu_0 + \xi \sum_{n = 1} \frac{1}{n}
\tilde{\nu}_{n}
\right)~,
\label{nue}
\end{equation}
where the normalization factor $N$ equals
\begin{equation}
N^2 = 1 + \xi^2 \sum_{n = 1} \frac{1}{n^2}
= 1 + \frac{\pi^2}{6} \xi^2~.
\label{norm}
\end{equation}
~From the phenomenological point of view the
bulk modes (being the singlets of the SM symmetry group)
can be considered as sterile neutrinos.
Thus, we deal with the coupled system of the electron neutrino and
infinite number of sterile neutrinos mixed. According to (\ref{nue}),
the electron neutrino
turns out to be the coherent mixture of states with increasing mass and
decreasing mixing.\\
The following comment concerning normalization is in order.
The contribution
to the normalization $N^2$ (\ref{norm}) from mass states with
$n = k, k+1, ....$ (starting from number $k$) can be estimated
substituting the sum by the integral:
\begin{equation}
\Delta_k \sim \xi^2 \int_k \frac{dn}{n^2} = \frac{\xi^2}{k}~
\label{norm}
\end{equation}
and for $k \rightarrow \infty$ we get $\Delta_k \rightarrow 0$.
Thus, due to decrease of mixing the effect of heavy states is suppressed.
In real physical processes with energy release $Q$ only
low mass part of the state (\ref{nue}) can be produced.
The states with
$
n > Q R
$
do not appear. This leads to breaking
of universality, that is, to difference of normalization of the
neutrino states produced in processes with different $Q$.
As follows from (\ref{norm}) the contribution of states with $n > Q R$ to
the normalization equals
\begin{equation}
\Delta (Q) = \frac{\xi^2}{Q R} \approx 10^{-10}
\frac{h^2}{Q/1 {\rm MeV}}~.
\label{Qnorm}
\end{equation}
(For $M_f = 10$ TeV and $R^{-1} = 3 \cdot 10^{-3}$ eV.)
Since $Q \gg 1/R$ for sub-millimeter scale the
deviation from
universality is negligible. This is not true for two
extra dimensions of common size \cite{fp} (see later).
For the same reason a change of kinematics of
processes due to emission of
states with $m_n \sim Q$ is unobservable.
The probability of emission of the heavy states is
negligible. \\
\section{Oscillations and Resonance Conversion}
Let us consider the vacuum oscillations of neutrino state
produced as $\nu_e$ (\ref{nue}) to the bulk modes.
The state will evolve with time $t$ as
\begin{equation}
\nu_e(t) = \frac{1}{N} \left(\nu_0 + \xi \sum_{n = 1}
\frac{1}{n} e^{i\phi_n} \tilde{\nu}_{n}
\right)~,
\label{nuet}
\end{equation}
where the phases $\phi_n$ equal
\begin{equation}
\phi_n \approx \frac{(m_n^2 - m_D^2) t}{2 E}~,
\end{equation}
and $E$ is the energy of neutrinos. The survival probability
of the $\nu_e \leftrightarrow \nu_e$ oscillations then equals:
\begin{equation}
P \equiv |\langle \nu_e |\nu_e(t) \rangle|^2
= \frac{1}{N^4}
\left| 1 + \xi^2 \sum_{n = 1} \frac{e^{i\phi_n}}{n^2} \right|^2.
\label{probab}
\end{equation}
Since $\phi_n \propto n^2$, the oscillation picture consists of
interference of the infinite number of modes with increasing
frequencies $\propto n^2$ and decreasing amplitudes $\propto 1/n^2$.
For practical purpose all high frequency modes can be averaged,
so that only a few low frequency oscillations can be observed
depending on the energy resolution of detector. Using
(\ref{probab}) we find the probability averaged over all the modes:
\begin{equation}
\bar{P} = \frac{1}{[1 + (\pi^2/6) \xi^2]^2}
\left[ 1 + \frac{\pi^4}{90}\xi^4 \right].
\label{avera}
\end{equation}
It is smaller than the two neutrino probability
with the same mixing parameter
$\xi$ due to presence of infinite number of mixed
states.
In particular, for $\xi \ll 1$ we get $\bar{P} \approx 1 - (\pi^2/3)
\xi^2$,
whereas in the $2\nu$ case: $\bar{P} \approx 1 - 2\xi^2$.
In the case when only lowest frequency mode (associated to
$\nu_1$) is non-averaged, we get from (\ref{probab})
the survival probability
\begin{equation}
P = \frac{1}{(1 + (\pi^2/6) \xi^2)^2}
\left[ (1 + \xi^2)^2 + \left(\frac{\pi^4}{90} - 1 \right)\xi^4
- 4 \xi^2 \sin^2 \frac{\phi_1}{2}
\right]~.
\label{firstmode}
\end{equation}
According to this equation the depth of oscillations
equals
\begin{equation}
A_P = \frac{4 \xi^2}{[1 + (\pi^2/6) \xi^2]^2}~.
\label{depth}
\end{equation}
Notice that due to presence of (practically) infinite number
of the bulk modes which give just averaged oscillation result,
the depth of oscillations of the lowest mode can not be maximal.
Moreover, the relation between the depth and the average probability
differs from the standard 2$\nu$ - oscillation case (see also \cite{ddg}).
Similarly one can find the probability with two non-averaged modes,
etc..\\
In medium with constant density the mixing with
different bulk states is modified depending on
$m_k^2$ and the potential, $V$, which describes the interaction with
medium:
\begin{equation}
V = G_F \frac{\rho}{m_N} \left( Y_e - \frac{1}{2}Y_n \right)~.
\label{potential}
\end{equation}
Here $G_F$ is the Fermi coupling constant, $m_N$ is the nucleon mass,
$Y_e$ and $Y_n$ are the numbers of electrons and neutrons per
nucleon correspondingly.
The mode for which
the resonance condition \cite{MSW},
\begin{equation}
\frac{m_k^2}{2E} \approx V
\end{equation}
is fulfilled (resonance modes)
will be enhanced: the effective mixing will be enhanced and
the oscillations will proceed with large depth.
The modes with lower frequencies (masses) will be suppressed;
the high frequency modes, $m_k^2/2 E \gg V$,
will not be modified.\\
Let us consider propagation of neutrinos in medium with varying density
$\rho(r)$
keeping in mind applications to solar and supernova neutrinos.
The energies of bulk states do not depend on density:
\begin{equation}
H_i = \frac{m_i^2}{2E}~,
\label{ilevel}
\end{equation}
whereas for the electron neutrino we have
\begin{equation}
H_e \approx V(\rho) ~.
\label{elevel}
\end{equation}
Therefore the level crossing scheme (the ($H - \rho$) - plot which
shows the dependence of the
energies of levels $H_e, H_i$ on density) consists of
infinite number of horizontal parallel lines (\ref{ilevel})
crossed by the electron neutrino line (\ref{elevel}).
In what follows we will concentrate on the case
$\xi \ll 1$, so that $m_D \ll m_n$ for all $n$,
and the crossings (resonances) occur
in the neutrino channels.
The resonance density, $\rho_n$, of $H_e$ crossing with
energy of $n^{th}$ bulk state: $H_n = H_e(\rho_n)$
equals according to (\ref{ilevel}, \ref{elevel})
\begin{equation}
\rho_n = \frac{m_n^2 m_N}{2 E G_F ( Y_e - \frac{1}{2}Y_n)} \propto n^2 ~.
\label{crossing}
\end{equation}
For small mixing ($\xi \ll 1$ ) the resonance layers
for different bulk states
(where the transitions, mainly, take place)
are well separated:
\begin{equation}
\rho_{n +1} - \rho_{n} \gg \Delta \rho_{n R} = \rho_n \frac{2
\xi}{n}~,
\end{equation}
here $\Delta \rho_{n R}$ is the width of the $n^{th}$- resonance layer.
Therefore transformation in each resonance occurs independently
and the interference of effects from different resonances can be
neglected.
In this case the survival probability $\nu_e \rightarrow \nu_e$
after crossing of $k$ resonances is just the product of the
survival probabilities in each resonance:
\begin{equation}
P = P_1 \times P_2 \times ~ ....~\times P_k~.
\end{equation}
Moreover, for $P_i$ we can take the asymptotic formula
which describes transition with
initial density being much larger than the resonance density
and the final density -- much smaller than the resonance density.
As the first approximation we can use the Landau-Zenner formula
\cite{Parke}:
\begin{equation}
P_n \approx \left\{
\begin{array}{ll}
1 & E < E_{n R} \\
{\displaystyle e^{- \frac{\pi}{2} \kappa_n}} & E > E_{n R}
\end{array}
\right. ~,
\label{LZ}
\end{equation}
where $E_{n R}$ is the resonance energy which corresponds
to maximal (initial) density $\rho_{max}$ in the region where neutrinos
are produced:
\begin{equation}
E_{n R} \approx \frac{m_n^2 m_N}{2 E G_F \rho_{max} ( Y_e - \frac{1}{2}
Y_n)}~;
\label{renergy}
\end{equation}
\begin{equation}
\kappa_n = \frac{m_n^2}{2E} \frac{\sin^2 2\theta_n}{\cos 2\theta_n}
\frac{\rho}{d\rho /dr}
\end{equation}
is the adiabaticity parameter \cite{MSW} and
$\sin^2 2\theta_n \approx 4\xi^2/n^2$.
Since $m_n^2 \propto n^2$ whereas $\sin^2 2\theta_n \propto 1/n^2$,
the adiabaticity parameter,
$\kappa_n = \kappa_1 \propto m_n^2 \sin^2 2\theta_n$,
does not depend on $n$ for a given energy.
Using this property we can write
final expression for the survival probability as
\begin{equation}
P \approx
{\displaystyle e^{- \frac{\pi}{2} \kappa_1 f(E)}}~,
\label{surv}
\end{equation}
where
\begin{equation}
\kappa_1 \approx \frac{2 \xi^2}{E R^2} \frac{\rho}{d\rho /dr}~,
\end{equation}
and $f(E)$ is the step-like function
\begin{equation}
f(E) = \left\{
\begin{array}{ll}
0 , & E < E_{1 R} \\
n , & E_{n R} < E < E_{n +1~R}
\end{array}
\right.~.
\label{fff}
\end{equation}
Since high level resonances turn on at higher energies and
$\kappa \propto 1/E$, the effect of conversion decreases with
increase of the order of the resonance, $n$. Moreover,
since in real situation the density is restricted from above
and the energies of neutrinos are in certain range,
only finite number of levels is relevant
and the largest effect is due to the lowest mass resonance.\\
\section{Solution of the Solar Neutrino Problem}
Let us apply the results of previous section to
solution of the solar neutrino problem.
We choose the lowest (non-zero) bulk mass, $m_1 = 1/R$, in such a way
that $\Delta m^2 = 1/R^2$ is in the range of small mixing
MSW solution due to conversion to sterile neutrino
$\nu_e \rightarrow \nu_s$ \cite{BKS}:
\begin{equation}
\frac{1}{R^2} = (4 - 10) \cdot 10^{-6}~ {\rm eV}^2~.
\label{masssq}
\end{equation}
This corresponds to $1/R = (2 - 3)~10^{-3}$ eV or
\begin{equation}
R = 0.06 - 0.10~ {\rm mm}~.
\end{equation}
Using mass squared difference (\ref{masssq}) as well as
maximal density and chemical
composition of the sun we
find from (\ref{renergy})
\begin{equation}
E_{1R} \approx 0.4 \div 0.8~ {\rm MeV}~.
\end{equation}
Thus the $pp$-neutrinos ($E_{pp} < 0.42$ MeV)
do not undergo resonance conversion: $E_{pp} < E_{1 R}$
(see (\ref{surv}, \ref{fff})),
whereas the beryllium neutrinos ($E_{Be} = 0.86$ MeV)
cross the first resonance.
(For smaller $1/R^2$ the
$pp$-neutrinos from the high energy part of their spectrum
can cross the resonance and the flux can be partly
suppressed.)
The energies of the next resonances equal:
$
E_{n R} = n^2 E_{1 R},
$
or explicitly:
$E_{2 R} = 1.6 - 3.2$ MeV,
$E_{3 R} = 3.6 - 7.2$ MeV,
$E_{4 R} = 6.4 - 12.8$ MeV,
$E_{5 R} = 10 - 20 $ MeV,
$E_{6 R} = 14.4 - 28.8 $ MeV. Higher resonances ($n > 6$) turn on at
energies higher than maximal energy of the solar neutrino spectrum and
therefore are irrelevant. The dependence of the
survival probability on energy is shown in the fig. 1.
The dips of the survival probability
at the energies $\sim E_{i R}$
are due to turning the corresponding resonances.
The effects of higher resonances lead to additional
suppression of the survival probability in comparison with
the two neutrino case. Therefore the
parameter $4\xi^2$ which is equivalent to $\sin^2 2\theta$ should be
smaller. We find that
\begin{equation}
4\xi^2 = (0.7 - 1.5)\cdot 10^{-3}
\label{xival}
\end{equation}
gives average suppression of the boron neutrino flux required by the
SuperKamiokande results.
According to (\ref{Dmass}), the value of $\xi$ (\ref{xival})
determines the fundamental scale:
\begin{equation}
M_f = \frac{\xi M_P}{\sqrt{2} hv R} = \frac{1}{h} (0.35 - 0.7)~ {\rm
TeV}~.
\label{scale}
\end{equation}
For small $h$ the scale $M_f$ can be large enough to satisfy
various phenomenological bounds.
Let us consider features of the suggested solution of the solar neutrino
problem.
The solution gives the fit of total rates in all experiments as good as
usual 2$\nu$ flavor conversion does: the $pp$-neutrino flux is
unchanged
or weakly suppressed, the beryllium neutrino flux can be strongly
suppressed,
whereas the boron neutrino flux is moderately suppressed and
this latter suppression can be tuned by small variations
of $\xi$.
\begin{figure}[htb]
\hbox to \hsize{\hfil\epsfxsize=11cm\epsfbox{unnamed.eps}\hfil}
\caption{~~The survival probability as the function of neutrino
energy for the electron neutrino conversion to the bulk
states in the Sun (solid line), $4\xi^2 = 10^{-3}$.
Dot-dashed line shows the survival probability
of the two neutrino conversion for the equivalent mixing
$\sin^2 2\theta = 10^{-3}$. Dashed line
corresponds to the survival probability of the
two neutrino conversion for $\sin^2 2\theta = 4 \cdot 10^{-3}$
which gives good fit of the total rates in all experiments.
}
\label{fbimaxinv}
\end{figure}
Novel feature appears in distortion of the boron neutrino spectrum.
As follows from fig. 1 three resonances turn on in the
energy interval accessible by SuperKamiokande ($E > 5$ MeV).
The resonances lead to the wave-like modulation of the
neutrino spectrum. (Sharp form (\ref{surv}, \ref{fff})
is smeared due to integration over
the production region.) The observation of such a regular wave structure
with $E \propto n^2$ would be an evidence of the
extra dimensions. However, in practice this will be very difficult to
realize.
The SuperKamiokande experiment measures the energy spectrum
of the recoil electrons from the reaction $\nu~ e - \nu~ e$ \cite{SK}.
Integrations over the neutrino energy as well as the
electron energy of the survival probability folded with the
neutrino cross-section and the
electron energy resolution function lead to strong
smearing of the distortion
in the recoil spectrum. As the result, the electron energy
spectrum will have just small positive slope (the larger the energy
the weaker suppression)
with very weak ( $ < 2 - 3$ \%) modulations. It is impossible to
observe such
a modulations with present statistics.
Notice that relative modulations become stronger if mixing,
$\xi$, is larger than $10^{-3}$ and
therefore suppression is stronger. This, however, requires larger
original boron neutrino flux.
The SNO experiment \cite{SNO} will have better sensitivity to
distortion of the spectrum.
The slope of distortion of the neutrino spectrum is substantially smaller
than in the
case of conversion to one sterile neutrino (see fig. 1).
In view of smearing effects due to integration over neutrino and electron
energies (due to finite energy resolution)
we can approximate the step-like function
$f(E)$ in (\ref{fff}) by smooth function $f_{app}(E) \approx
\sqrt{E/E_{1R}}$.
Then the smeared survival probability in the high energy range
can be written as
\begin{equation}
P \approx {\displaystyle e^{-\sqrt{\frac{E_0}{E}}}}~,
\label{pro}
\end{equation}
where $\sqrt{E_0} = \pi \xi^2 \rho /(R^2 d\rho/dr \sqrt{E_{1R}})$.
In the case of two large dimensions with
$R_1 \sim R_2 \sim 0.02 - 0.03~ $ mm (see sect. 5) the number of
bulk states, and therefore the number of relevant resonances
increases quadratically: $n^2$. (Here $n \sim 5 - 6$
is the number of resonances in the energy range of solar neutrinos
in the one dimension case.) Correspondingly, the approximating function
$f_{app}(E)$ will be proportional to $E$.
As the result, $\kappa_1 \cdot f(E) = const$
and the smeared survival probability
will not depend on energy. In this case $P(E) \approx const$ for
$E > E_{1R}$ and there is no distortion of the
recoil electron spectrum. For two different radii: $R_2 < R_1$
one can get any intermediate behaviour of the probability between
that in (\ref{pro}) and $P = const$.
Common signature of both standard $\nu_e - \nu_s$ conversion and
conversion to the bulk modes is the suppression of the neutral current
(NC) interactions. The two can be, however, distinguished using the
following fact.
In the case of the $\nu_e - \nu_s$ conversion there is certain correlation
between suppression of the NC interactions and distortion of the spectrum.
The weaker distortion the weaker suppression of the
NC interactions and vice versa. In the case of $\nu_e - \nu_{bulk}$
conversion a weak distortion can be accompanied by significant
suppression of the NC events.
This can be tested in the SNO experiment.
No significant Day-Night asymmetry is expected due to
smallness of mixing angle.
Thus, the smeared distortion of the energy spectrum
(for $E > 5$ MeV ) is weak or absent in the case of
$\nu_e - \nu_{bulk}$ conversion. However, in contrast to other energy
independent solutions here $pp-$neutrino flux may not be
suppressed, or the energy spectrum of $pp$- neutrinos can be
significantly distorted.\\
Notice that it is impossible to reproduce the large mixing
angle MSW solution of the solar neutrino problem \cite{BKS} in
this context. Indeed, for $\xi \sim O(1)$ and
$1/R^2$ as in (\ref{masssq}) transitions in all low mass resonances
are adiabatic, and therefore the survival probability
has the form:
\begin{equation}
P(E) \sim P_n(E) = \sin^2 \theta_n \approx \frac{\xi^2}{n^2}, ~~~~
E_{n R} < E < E_{n +1~R}.
\label{lma}
\end{equation}
(There are smooth transitions in the regions $E \sim E_{n R}$.)
For a given energy $E$ the probability is determined by
conversion in the nearest $n^{th}$ resonance with
$E_{n R} < E$. According to (\ref{lma})
the probability decreases
monotonously with energy, in
contrast with observations. For instance, if $P \sim 0.5$
in the interval
$E = 0.5 \div 2$ MeV, then it will be
$P \sim 0.2$ for $E = 2 \div 4.5 $ MeV,
$P \sim 0.06$ for $E = 4.5 \div 8$ MeV etc..
If $\xi > 1$, for all the modes $k$ with $k^2/(k +1) < \xi^2$
the resonances will be in the antineutrino channels and
for $k^2/(k +1) > \xi^2$ in the neutrino channels \cite{DS}. \\
Note that solution of the solar neutrino problem
due to the long length vacuum oscillations \cite{VO} into
bulk modes implies
$\Delta m^2 \approx (0.5 - 5) \cdot 10^{-10}$ eV$^2$
and large (maximal) mixing. This leads to the following estimations
\begin{equation}
\frac{1}{R} \sim \sqrt{\Delta m^2} \sim (0.7 - 2)~ 10^{-5} {\rm eV},~~
m_D = \frac{\xi}{R} = (0.5 - 1.5) \cdot 10^{-5} {\rm eV}
\end{equation}
for values $\xi = 0.5 - 0.7 $.
Now the size of the extra dimension equals $L = 6 - 20$ cm
which is excluded by existing tests of the Newton law.\\
The approach opens however another possibility. Suppose that
the radius $R$ is small enough so that KK-excitations
have negligible mixing with usual neutrinos. In this case
the effect of extra dimensions is reduced to
interaction with zero mode,
$\nu_{0R}$, only. Suppose that the
same bulk field
couples with two usual neutrinos:
$\nu_e$ and $\nu_{\mu}$ (or $\nu_{\tau}$) generating the
Dirac mass terms
\begin{equation}
m_{eD} \bar{\nu}_{0R} \nu_e + m_{\mu D} \bar{\nu}_{0R} \nu_{\mu}.
\end{equation}
These terms lead to Dirac neutrino with mass
$m_D = \sqrt{m_e^2 + m_{\mu}^2}$ formed by ${\nu}_{0R}$
and the combination
$(m_{eD} \nu_e + m_{\mu D} \nu_{\mu})/m_D$. The orthogonal
component is massless. In this way the $\nu_e$ and $\nu_{\mu}$
turn out to be mixed with the angle determined
by $\tan \theta = m_{eD}/m_{\mu D}$.
(Similar mechanism of mixing has been considered in Ref. \cite{ddg}.)
For $M_f \sim 1$ TeV and
the original Yukawa couplings with bulk field
$h_e \sim h_{\mu} \sim1 $ we get $m_{eD} \sim m_{\mu D} \sim
10^{-5}$ eV which leads to
$\Delta m^2 \sim 10^{-10}$ eV$^2$ and maximal (large)
$\nu_e - \nu_{\mu}$ mixing. This reproduces values of parameters
required for the
$\nu_e \leftrightarrow \nu_{\mu}$ Just-so oscillation solution of the
solar neutrino problem. The solution has however no
generic signatures of the high dimensional theory
and employs the latter as the source of neutrino mass only.
\section{Solar Neutrinos and Parameters of Extra Dimensions}
As follows from previous section, a solution of the solar neutrino
problem via resonance conversion to the bulk modes implies
the radius of extra dimension $R = 0.06 - 0.1$ mm and
the fundamental scale $M_f > 0.5 - 1$ TeV. To satisfy the relation
(\ref{radii}) we need to introduce more extra dimensions.
Let us assume that second large
dimension exists with radius $R'$. From (\ref{radii}) we get
\begin{equation}
\frac{1}{R'} = \frac{(2\pi)^2 M_f^4 R }{M_P^2} \approx
1.3 \cdot 10^{-3} \left(\frac{M_f}{1 {\rm TeV}}\right)~{\rm eV}.
\end{equation}
For $M_f = 1 $ TeV both extra dimensions will have
radii of the same size. In this case more bulk states are involved in
conversion of solar neutrinos which will lead to absence of the
distortion of the spectrum, as we have discussed in sect. 4.
For $M_f \geq 10 $ TeV the bulk states associated to $1/R'$
dimension do not participate in solar neutrino conversion,
however they are relevant for other
neutrino processes. Let us consider this in more details.
Now the electron neutrino state can be written as
\begin{equation}
\nu_e = \frac{1}{N} \left(\nu_0 +
\xi \sum_{n} \frac{1}{n} \nu_{n,0} +
\xi \sum_{n, k \geq 1}
\frac{1}{\sqrt{n^2 + (R/R')^2 k^2}}
\nu_{n, k}
\right)~,
\label{nue21}
\end{equation}
where index $n$ refers to dimension of larger radius $R$,
while $k$ enumerate the bulk states from
dimension $R'$. The sum over the states
is divided into two parts: the first sum contains the states
with $k = 0$, that is, with small mass split only. This part corresponds
to the one dimensional case discussed in sect. 2 - 4 and
as we have shown in sect. 3,
it does not lead to observable violation of universality.
The second sum contains the towers of states with
both large and small mass splits.
Its contribution to the normalization of the state equals
\begin{equation}
\Delta \equiv
\xi^2 \sum_{n, k} \frac{1}{n^2 + (\frac{R}{R'})^2 k^2}~.
\label{univ}
\end{equation}
We can estimate the sum substituting it by the integral
over $n$ and $k$. Performing first integration over
$n$ from 0 to $\infty$ and then over $k$ from 1 to $(Q R')$ --
the number of states which can be produced in the process with energy
release $Q$, we find:
\begin{equation}
\Delta (Q) =
\frac{\pi}{2} \frac{R'}{R} \xi^2 \ln (Q R') =
\pi
\left(\frac{h v}{M_P}\right)^2 \frac{V_2 M_f^2}{(2\pi)^2} \ln (Q R')~.
\label{univ1}
\end{equation}
The difference of normalizations of the two states produced in
processes with energy releases
$Q_1$ and $Q_2$ gives measure of the universality breaking:
\begin{equation} \Delta_{21} \equiv \Delta (Q_2) -
\Delta (Q_1) =
\frac{\pi}{2} \frac{R'}{R} \xi^2 \ln (Q_2/Q_1).
\end{equation}
Taking $\xi^2 = (2 - 4) \cdot 10^{-4}$ as is implied by the solar neutrino
data, $Q_1\sim 1$ MeV and $Q_2 \sim 100$ GeV as well as $R'/R < 0.1$
we find $\Delta_{12} \sim (2 - 4) \cdot 10^{-4}$ which is below
present sensitivity. Notice, however, that for $R' \sim R$
the violation can be at the level of existing bounds \cite{fp}. \\
Let us consider the case of three extra dimensions with
radii $R \gg R_2, R_3$.
Assuming for simplicity the equality $R_2 = R_3 \equiv R'$,
we find from (\ref{radii}):
\begin{equation}
\frac{1}{R'} = \frac{(2\pi)^{3/2}M_f^2 \sqrt{M_f R }}{M_p}~.
\end{equation}
For $1/R \sim 2.5 \cdot 10^{-3}$ eV and $M_f = 10~$ TeV this equation
gives
$1/R' \sim 3 \cdot 10^{-2}$ GeV.
Performing calculations similar to those for one additional dimension we
find parameter of universality violation
\begin{equation}
\Delta_{21} \approx 2 \left(\frac{h v}{M_P}\right)^2
\frac{V_3 M_f^3}{(2\pi)^3} \frac{(Q_2 - Q_1)}{M_f} =
\frac{(h v)^2 (Q_2 - Q_1)}{M_f^3}
\end{equation}
($V_3 = (2\pi)^3R R'^2$)
which can be as low as $10^{-6}$
even for $Q_2 - Q_1 \sim 100$ GeV
(in this estimation we used $M_f = 10$ TeV and $h = 1$).\\
\section{Astrophysical Bounds. Atmospheric neutrinos}
The important bound on mixing of usual neutrinos with
sterile neutrinos as well
as with bulk states follows from primordial nucleosynthesis:
production of new relativistic degrees of freedom
leads to faster expansion of the Universe and to larger
abundance of $^4 He$. There are two ways of production of the bulk states:
(i) via oscillations and (ii) incoherently via chirality flip.
In the first case the electron neutrino produced as the coherent
combination of mass eigenstates oscillates in medium
to bulk states. Inelastic collision splits the state into
active ($\nu_e$) and to sterile ($\nu_{bulk}$) parts
and after each collision
two parts will oscillate independently.
Oscillations between two collisions average. Production
rate is then
the sum of the averaged oscillation effects over collisions.
The condition that sterile states do not reach equilibrium
puts stringent bound on oscillation parameters \cite{NS}.
The masses squared and mixing angles
of bulk states (\ref{masssq}, \ref{xival})
implied by solution of the solar neutrino problem
satisfy the following relation:
\begin{equation}
\Delta m^2_n \cdot \sin^2 2\theta_n \approx \frac{4\xi^4}{R^2}~=
(4 - 8)\cdot 10^{-9} {\rm eV}^2,
\end{equation}
where $\sin^2 2\theta_n \equiv 4\xi^4/n^2$.
This ``trajectory" in the
$\Delta m^2 - \sin^2 2\theta$ - plot
lies far outside the region excluded by primordial
nucleosynthesis \cite{NS}.
That is, production of the bulk neutrinos
via oscillations is strongly suppressed.
Let us consider incoherent production of the bulk states due to
chirality flip. The production rate of
an individual bulk neutrino ($\Gamma_{1}$) is suppressed
relatively to the production rate
of the left-handed neutrino ($\Gamma_{\nu_e}$) by the chirality - flip
factor
\begin{equation}
{\Gamma_{1}\over \Gamma_{\nu_e}} \sim \left ({m_D \over T}\right)^2~,
\end{equation}
where the temperature $T$ in the denominator comes from the
propagator of a primarily-produced left handed neutrino. The
multiplicity of the final bulk states is
$T R$, so that the total bulk neutrino emission rate is
suppressed as
\begin{equation}
{\Gamma_{bulk}\over \Gamma_{\nu_e}} \sim
\left ({m_D \over T}\right)^2~ T R = \xi \left ({m_D \over T}\right)~.
\label{ratio}
\end{equation}
For the parameters implied by the solar neutrinos and $T \sim 1$ MeV
this ratio is about $3 \cdot 10^{-11}$.
Using (\ref{ratio}) we find the temperature, $T_{bulk}$,
at which
production rate of the bulk states is comparable with expansion rate
of the Universe: $\Gamma_{bulk} = \Gamma_{exp} \sim T^2/M_P$:
\begin{equation}
T_{bulk} = T_{\nu} \sqrt{\frac{T_{\nu}}{\xi m_D}}~,
\label{tempbulk}
\end{equation}
where $T_{\nu} \sim 1$ MeV is the temperature
of the neutrino decoupling. From (\ref{tempbulk})
we find $T_{bulk} \sim 200$
GeV which is much above the ``normalcy" temperature \cite{add1}.
The modes from additional dimensions having smaller radii
give even smaller contribution. \\
The KK-neutrinos as well as the KK-gravitons
produced in stars, in particular, in supernovae, increase the rate of
star cooling \cite{add1}.
No additional sources of cooling have been found from observations of
the SN1987A which put stringent bound on production of the bulk states.
The rate of the incoherent production of the bulk neutrinos in supernovae
is suppressed by the same factor (\ref{ratio}).
For temperature of the core of supernova $T \sim 30 $ MeV
the eq. (\ref{ratio}) gives
$\Gamma_{bulk}/ \Gamma_{\nu_e} \sim 3 \cdot 10^{-12}$.
Then taking into account that bulk neutrinos are emitted from the whole
volume of the core, whereas usual neutrinos are emitted from the
surface (neutrinosphere) we find
that the luminosity in the bulk states is 5 - 6 orders of magnitude smaller
than luminosity in neutrinos.
Production of the bulk states via oscillations is strongly suppressed
by matter effect. Matter suppression is weak or absent for
bulk states with high mass: $m \sim 10^3$ eV. However, their production
is suppressed by very small vacuum mixing.
So, we conclude that parameters required by
solution of the solar neutrino problem satisfy
astrophysical constraints.\\
Let us comment in this connection on solution of the atmospheric neutrino
problem via oscillations of muon neutrinos to the bulk states
$\nu_{\mu} \leftrightarrow \nu_{bulk}$. This solution
requires smaller radius of the extra dimension:
$1/R \approx \sqrt{\Delta m^2_{atm}} \sim (5 - 9)\cdot 10^{-2}$ eV
or $R \sim (2 - 4)\cdot 10^{-3}$ mm and near to maximal mixing:
$\xi \sim 1$. The latter means that the fundamental scale should be
about $M_f \approx (10^3~ {\rm TeV})/ h$.
that is, about 3 orders of magnitude larger than that
for solar neutrinos.
(Notice that approximation $\xi \ll 1$ can not be used
now to diagonalize the mass matrix and results
of sect. 3 and Appendix should be corrected \cite{DS}.
Still mixing of large mass bulk states is suppressed by factor
$1/n$ and these states lead to finite averaged oscillation result.
Only a few low mass states are relevant for non-averaged oscillation
picture.)
The oscillation parameters required by the solution of the
atmospheric neutrino problem violate nucleosynthesis bound.
Indeed, masses and mixing angles of the bulk modes satisfy now the
relation:
\begin{equation}
m_n^2 \cdot \sin^2 2\theta_n \sim \frac{2}{R^2} \sim 6 \cdot 10^{-3}
{\rm eV}^2.
\end{equation}
Nucleosynthesis bound reads \cite{NS}
\begin{equation}
\Delta m^2 < \frac{3 \cdot 10^{-5} {\rm eV}^2 } {\sin^4 2\theta}~.
\end{equation}
{}From these two equation we find that about 20 - 25 lightest
bulk states are in the forbidden region. They turn out to
be in equilibrium, whereas 1 or at most 2 are allowed.
Production of the bulk states via oscillations can be suppressed if there
is substantial ($\sim 10^{-5}$) leptonic asymmetry in the Universe
\cite{FV}.\\
Let us finally consider production of gravitons in the stars.
The generic reason that
saves the bulk gravitons from being ruled out by star cooling
is the infrared-softness of the high dimensional gravity.
On the language of four-dimensional KK modes this can be
visualized as
follows. The rate of each individual KK graviton emission in the star
is suppressed by the universal volume factor
\begin{equation}
\left ({T \over M_P} \right)^2 \sim \left ({T \over M_{f}}
\right)^2
{1 \over M_f^N V_N}~.
\end{equation}
The number of available final states is $ \sim T^NV_N$,
so that the over-all rate is suppressed as
\begin{equation}
\Gamma_{grav} \sim \left({T \over M_P} \right)^2 T^N V_N
\sim \left ({T \over M_f}
\right)^{2 + N}~.
\label{supern}
\end{equation}
According to this expression, the analysis of supernova core
cooling
gives for $N = 2$ the lower bound on the fundamental gravitational
scale $M_f$
about $30$ TeV \cite{add1} - $50$ TeV \cite{supernova22} .
As it is clear from (\ref{supern}), the rate is determined
by the value of $M_{f}$ and it is insensitive to
sizes of individual radii $R_i$ as far as all $1/R_i < T$.
On the other hand, if some radii, $R_k$, do not satisfy this bound, then
the
corresponding KK modes can not be produced in the star and
the corresponding factor $T R_k$ in $T^NV_N$ of Eq. (\ref{supern}) has
to be replaced by 1.
The bottom-line of this discussion is that standard constraint can be
avoided if the extra dimensions have different radii.
For instance, with one
radius $R \sim 0.03$ mm and two others $R_k > T$
(which can be realized even for
$M_f$ as low as several TeV's) the rate becomes
\begin{equation}
\Gamma_{grav} \sim \left ({T \over M_P} \right)^2 (T R)
\approx 10^{-31}~.
\end{equation}
Therefore, some of the radii can be of sub-millimeter size and,
thus, can be a subject of direct experimental search in
future gravitational measurements \cite{measurements}.\\
\section{Conclusions}
We have studied consequences of the neutrino mixing with fermions
propagating in the bulk in the context of theories with
large extra dimensions. The bulk fermions could be
components of bulk gravitino or other singlets of the SM gauge
group.
Phenomenology of this mixing is determined by the following features:
(i) The bulk fermions
can be considered as sterile neutrinos.
(ii) Large number of these sterile neutrinos is
involved in physical processes.
(iii) For $m_D < 1/R$ usual neutrinos are combinations of
mass eigenstates with increasing masses and decreasing admixtures.
The effect of bulk states with large masses is reduced to
averaged oscillations. Low modes can lead to
non-averaged oscillations in vacuum (uniform medium)
or to multi-resonance conversion
in medium with varying density.
The resonance conversion of the electron neutrino
to the bulk states can solve
the solar neutrino problem. Properties of this solution
are similar to those due to conversion to sterile neutrino.
The important difference is that significant suppression of the
boron neutrino flux can be accompanied by weak distortion of the
energy spectrum. Moreover, weak modulation of the boron neutrino
spectrum is
expected due to conversion to several KK-states.
Simultaneous explanation of the atmospheric, solar
and LSND results in terms of neutrino oscillation/conversion
implies existence of sterile neutrino. Moreover, the data
favour $\nu_{\mu} \leftrightarrow \nu_{\tau}$
oscillations as the solution of the atmospheric neutrino problem,
so that the solar neutrinos should be converted
to sterile states. In this connection
one can consider the following possibility.
There is some usual (4 dimensional) mechanism of generation
of the active neutrino masses.
This mechanism produces neutrino
mass pattern with heavy and strongly mixed $\nu_{\mu}$ and
$\nu_{\tau}$ and very light $\nu_e$. Such a pattern
can explain the atmospheric neutrino and LSND results.
Neutrinos
(in general of all flavors) couple also with the bulk fermion.
These couplings (being of the same order for all neutrino species)
generate negligible mixing of the KK-states with heavy
$\nu_{\mu}$ and $\nu_{\tau}$ and large enough mixing
with light $\nu_e$, so that the solar neutrino problem can be solved as
is described in sect. 4.\\
The suggested solution of the solar neutrino problem implies that
the radius of at least one extra dimension is in the range
0.06 - 0.10 mm, that is, in the range of sensitivity of
proposed gravitational measurement.
The fundamental scale should be about
$h M_f \sim 1$ TeV. This mass satisfies the fixed overall
volume condition provided additional large extra dimensions
exist. In the case of one additional extra dimension its radius should be
$1/R' = 1.3 \cdot 10^{-3} (M_f / 1 {\rm TeV})^4$ eV.
For $h \sim 1$ one has $ M_f \sim 1$ TeV,
and $1/R' \sim 2 \cdot 10^{-3}$ eV, so that the
second extra dimension will influence the solar neutrino
data too. If $h \ll 1$, the fundamental scale can be much larger
than 1 TeV, and
$R'$ can be much smaller than 1 mm. For $h = 0.1$ and $M_f = 10$ TeV we
get $1/R' \sim 10$ eV. For two additional dimensions
the common radius equals $1/R' \sim 3 \cdot 10^{-2}$ GeV,
if $h = 1$ and $M_f = 10$ TeV.
For large fundamental scales ($M_{f} > 10 - 20$ TeV),
direct laboratory searches at high energies will be
practically impossible and
neutrinos can give unique (complementary to gravitational measurements)
opportunity to probe the effects of large extra dimensions.\\
\noindent
{\Large \bf Appendix}\\
The mass terms (\ref{mass2}) can be written as
$$
\bar{\nu}_L M \nu_R~,
$$
where ${\nu}_L^T \equiv (\nu_L, \tilde{\nu}_{1 L},
\tilde{\nu}_{2 L} ... )$ and
$\nu_R^T = (\nu_{0 R}, \tilde{\nu}_{1 R}, \tilde{\nu}_{2 R} ... )$;
the modes
$\nu_{0 L}$, $\nu'_{n L}$ $\nu'_{n R}$
decouple from the system, and the mass matrix
$M$ for $k+1$ states equals
$$
M = \left(
\begin{array}{ccccc}
m_D & \sqrt{2}m_D & \sqrt{2}m_D & ... &\sqrt{2}m_D \\
0 & \frac{1}{R} & 0 & ... & 0 \\
0 & 0 & \frac{2}{R} & ... & 0 \\
... & ... & ... & ... & ... \\
0 & 0 & 0 & ... & \frac{k}{R}
\end{array}
\right)~.
$$
Let us consider the matrix $M M^{\dagger}$ which determines
mixing of the left handed neutrinos:
$$
~~~~~~~~~~~~~~~~~~~M M^{\dagger} = \frac{1}{R^2} \left(
\begin{array}{ccccc}
(k + 1/2) \xi^2 & \xi & 2 \xi & ... & k \xi \\
\xi & {1} & 0 & ... & 0 \\
2 \xi & 0 & 4 & ... & 0 \\
... & ... & ... & ... & ... \\
k \xi & 0 & 0 & ... & k^2
\end{array}
\right)~, ~~~~~~~~~~~~~~~~~~~~~~~~~~ (A1)
$$
where $\xi \equiv \sqrt{2} m_D R$.
Notice that the mass matrix (A1)
corresponds to compactification on the circle;
it can be shown that the same matrix follows from
the $Z_2$ - orbifold compactification \cite{ddg}.
The diagonalization of matrix can be performed starting
from the heaviest state $k \gg 1$. Then one can check that
the limit $k \rightarrow \infty$ does not change results.
The rotation by the angle $\theta_k$ in the plane $\nu_{0 L} -
\tilde{\nu}_{k L}$
$$
~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~\tan 2 \theta_k = 2 \frac{\xi}{k}\cdot
\frac{1}{1 - \frac{\xi^2}{k}}
~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~(A2)
$$
diagonalizes corresponding submatrix.
It is easy to perform the diagonalization using $\xi$ as an expansion
parameter: $\xi \ll 1$ as is implied by the
solution of the solar neutrino problem.
The rotation (A1) leads to modification of the
first diagonal element
$
(k + 1/2) \xi^2 \rightarrow (k -1/2) \xi^2 + O (\xi^4)
$
and modification of the mixing terms as
$n \xi \rightarrow \cos \theta_k n \xi$.
For small $\xi$: $\tan \theta_k \approx \xi/k$ and
the eigenvalues equal $ m_k^2 \approx k^2/R^2$.
After $k-1$ subsequent rotations we get for the first
diagonal element: $3/2 \xi^2 + O (\xi^4)$
and for off-diagonal term:
$$
\xi \cos \theta_2\cos \theta_3 ... \cos \theta_k \approx
\xi \left(1 - \frac{1}{2}\xi^2 \sum_{n = 2 ... k} 1/n^2 +...
\right) \approx \xi.
$$
\\
These results can be obtained from the exact characteristic equation
for the mass eigenstates: ${\rm Det}[ M M^{\dagger} - m^2]$ which
can be written explicitly as
$$
\pi \xi^2 \cot(\pi m R) = 2 mR.
$$
which coincides with the characteristic equation in \cite{ddg} for the
same neutrino system.\\
\noindent
{\Large \bf Acknowledgments}\\
One of us (A.S.) is grateful to E. Dudas for useful discussions.
|
\section{Introduction}
The temporal behavior of quantum mechanical systems, being governed
by unitary operators \cite{Dirac}, displays some subtle features at
short \cite{shortt} and long times \cite{longtt}. In order to
discuss the evolution of genuine unstable systems one usually makes
use of the Weisskopf-Wigner approximation \cite{seminal}, which
ascribes the main properties of the decay law to a pole located
near the real axis of the complex energy plane. This yields the
Fermi ``Golden Rule" \cite{Fermigold}. These features of the
quantum evolution are well known and discussed in textbooks of
quantum mechanics \cite{Sakurai} and quantum field theory
\cite{Brown}. For a recent review, see \cite{temprevi}. In this
paper we shall investigate the possibility that the lifetime of an
unstable quantum system can be modified by the presence of a very
intense electromagnetic field. We shall look at the temporal
behavior of a three-level system (such as an atom or a molecule),
where level \#1 is the ground state and levels \#2, \#3 are two
excited states. The system is initially prepared in level \#2 and
if it follows its natural evolution, it will decay to level \#1.
The decay will be (approximately) exponential and characterized by
a certain lifetime, that can be calculated from the Fermi Golden
Rule. But what happens if one shines on the system an intense laser
field, tuned at the transition frequency 3-1? This problem was
investigated in Ref.\ \cite{MPS}, in the context of the so-called
quantum Zeno effect \cite{QZE}. It was found that the lifetime of
the initial state depends on the intensity of the laser field. In
the limit of extremely intense field, the decay should be
considerably slowed down. The aim of this paper is to study this
effect in more detail and discuss a new phenomenon: we shall see
that for physically sensible values of the laser field, the decay
can be {\em enhanced}, rather than hindered. This can be viewed as
an ``inverse" quantum Zeno effect. The whole problem is related to
electromagnetic induced transparency \cite{induced}.
\section{Preliminaries and definitions }
\label{sec-preldef}
\andy{preldef}
We start from the Hamiltonian \cite{MPS}:
\andy{ondarothamdip6}
\begin{eqnarray}}\newcommand{\earr}{\end{eqnarray}
H &=& H_{0}+H_{\rm int}\nonumber\\
&=& \omega_0|2\rangle\langle 2|+\Omega_0|3\rangle\langle 3|
+\sum_{\bmsub k,\lambda}\omega_k a^\dagger_{\bmsub{k}\lambda}
a_{\bmsub{k}\lambda}
+\sum_{\bmsub k,\lambda}\left(\phi_{\bmsub k\lambda}
a_{\bmsub k\lambda}^\dagger|1\rangle\langle2|
+\phi_{\bmsub k\lambda}^* a_{\bmsub
k\lambda}|2\rangle\langle1|\right)\nonumber\\ & &+\sum_{\bmsub
k,\lambda}\left(\Phi_{\bmsub k\lambda}
a_{\bmsub k\lambda}^\dagger|1\rangle\langle3|
+\Phi_{\bmsub k\lambda}^* a_{\bmsub
k\lambda}|3\rangle\langle1|\right),
\label{eq:ondarothamdip6}
\earr
where the first two terms are the free Hamiltonian of the 3-level
system (whose states $|i\rangle$ $(i=1,2,3)$ have energies $E_1=0$,
$\omega_0=E_2-E_1>0$, $\Omega_0=E_3-E_1>0$), the third term is the
free Hamiltonian of the EM field and the last two terms describe
the $1\leftrightarrow2$ and $1\leftrightarrow3$ transitions in the
rotating wave approximation, respectively. The states $|2\rangle$
and $|3\rangle$ are chosen so that no transition between them is
possible (e.g., because of selection rules). The matrix elements of
the interaction Hamiltonian read
\andy{intel}
\begin{eqnarray}}\newcommand{\earr}{\end{eqnarray}
\phi_{\bmsub k\lambda}&=&\frac{e}{\sqrt{2\epsilon_0 V\omega}}
\int d^3x\;e^{-i\bmsub k\cdot\bmsub x}\bm\epsilon_{\bmsub k\lambda}^*\cdot \bm
j_{12}(\bm x) ,
\nonumber \\
\Phi_{\bmsub k\lambda}&=&\frac{e}{\sqrt{2\epsilon_0 V\omega}}
\int d^3x\;e^{-i\bmsub k\cdot\bmsub x}\bm\epsilon_{\bmsub k\lambda}^*\cdot \bm
j_{13}(\bm x),
\label{eq:intel}
\earr
where $-e$ is the electron charge, $\epsilon_0$ the vacuum
permittivity, $V$ the volume of the box, $\omega=|\bm k|$,
$\bm\epsilon_{\bmsub k\lambda}$ the photon polarization and $\bm
j_{\em fi}$ the transition current of the radiant system. For
example, in the case of an electron moving in an external field, we
have $\bm j_{\em fi}=\psi_{\em f}^\dagger \bm\alpha\psi_{\em i}$
where $\psi_{\em i}$ and $\psi_{\em f}$ are the electron
wavefunctions of the initial and final states, respectively, and
the components of $\bm\alpha$ are the usual Dirac matrices. For the
sake of generality we are using relativistic matrix elements
(although our analysis can be performed with nonrelativistic ones
$\bm j_{\em fi}=\psi_{\em f}^*\bm p\psi_{\em i}/m_e$, where $\bm
p/m_e$ is the electron velocity).
We shall concentrate our attention on a 3-level system bathed in a continuous
laser beam, whose photons have momentum
$\bm k_0$ ($|\bm k_0|=\Omega_0$) and polarization $\lambda_0$.
We shall also assume, throughout this paper, that
\andy{noint}
\begin{equation}}\newcommand{\eeq}{\end{equation}
\phi_{\bmsub k_0\lambda_0}=0,
\label{eq:noint}
\eeq
i.e., the laser does not interact with state $|2\rangle$. Also,
since the average number $N_0$ of $\bm k_0$-photons in the total
volume $V$ can be considered very large, we shall perform our
analysis in terms of number (rather than coherent) states of the EM
field \cite{Knight}. In this approximation,
\andy{appros}
\begin{eqnarray}}\newcommand{\earr}{\end{eqnarray}
& & \langle1; 0_{\bmsub k\lambda}, N_0|H_{\rm int} |3; 0_{\bmsub
k\lambda},N_0-1\rangle=\sqrt{N_0}
\Phi_{\bmsub k_0\lambda_0}
\nonumber \\
& & \qquad \qquad \gg
\langle1;1_{\bmsub k\lambda}, N_0-1|H_{\rm int}
|3;0_{\bmsub k\lambda},N_0-1\rangle=
\Phi_{\bmsub k\lambda},
\label{eq:appros}
\earr
$\forall(\bm k,\lambda)\neq(\bm k_0,\lambda_0)$. In the above
equation and henceforth, the vector $|i;n_{\bmsub k \lambda},
M_0\rangle$ represents an atom or a molecule in state $|i\rangle$,
with $n_{\bmsub k \lambda}$ $(\bm k,\lambda)$-photons and $M_0$
laser photons.
In the above approximation, the Hamiltonian (\ref{eq:ondarothamdip6})
can be replaced by
\andy{ondarothamdip7}
\begin{eqnarray}}\newcommand{\earr}{\end{eqnarray}
H \! &\simeq& \! \omega_0|2\rangle\langle
2|+\Omega_0|3\rangle\langle 3| +\sum_{\bmsub k,\lambda}\omega_k
{a}^\dagger_ {\bmsub{k}\lambda} {a}_ {\bmsub{k}\lambda}
+{\sum_{\bmsub k,\lambda}}'\left(\phi_{\bmsub k\lambda} a_{\bmsub
k\lambda}^\dagger|1\rangle\langle2| +\phi_{\bmsub k\lambda}^*
a_{\bmsub k\lambda}|2\rangle\langle1|\right)\nonumber\\ &
&+\left(\Phi_{\bmsub k_0\lambda_0} a_{\bmsub
k_0\lambda_0}^\dagger|1\rangle\langle3| +\Phi_{\bmsub
k_0\lambda_0}^* a_{\bmsub k_0\lambda_0}|3\rangle\langle1|\right),
\label{eq:ondarothamdip7}
\earr
where the prime means that the summation does not include $(\bm
k_0,\lambda_0)$ [due to our hypothesis (\ref{eq:noint})].
The operators
\andy{operatoreN}
\begin{equation}}\newcommand{\eeq}{\end{equation}
{\cal N}=|2\rangle\langle 2|+{\sum_{\bmsub k,\lambda}}'
{a}^\dagger_{\bmsub{k}\lambda}{a}_{\bmsub{k}\lambda},
\qquad
{\cal N}_0=|3\rangle\langle 3|+{a}^\dagger_{\bmsub{k}_0\lambda_0}
{a}_{\bmsub{k}_0\lambda_0}
\label{eq:operatoreN}
\eeq
satisfy
\begin{equation}}\newcommand{\eeq}{\end{equation}
[H,{\cal N}]=[H,{\cal N}_0]=[{\cal
N}_0,{\cal N}]=0,
\eeq
which imply the conservation of the total number of photons plus
the atomic excitation (Tamm-Dancoff approximation
\cite{TammDancoff}). The Hilbert space splits therefore into
sectors that are invariant under the action of the Hamiltonian: in
our case, the system evolves in the subspace labelled by the
eigenvalues ${\cal N}=1$, ${\cal N}_0=N_0$ and the analysis can be
restricted to this sector \cite{Knight}.
\section{Temporal evolution}
\label{tempevol}
\andy{tempevol}
The states of the total system in the sector $({\cal N}, {\cal
N}_0)=(1,N_0)$ read
\label{statesdefin}
\begin{equation}}\newcommand{\eeq}{\end{equation}\label{eq:statesdefin}
|\psi(t)\rangle=x(t)|2;0,N_0\rangle+{\sum_{\bmsub
k,\lambda}}'y_{\bmsub k\lambda}(t)|1;1_{\bmsub
k\lambda},N_0\rangle+{\sum_{\bmsub k,\lambda}}'z_{\bmsub
k\lambda}(t)|3;1_{\bmsub k\lambda},N_0-1\rangle,
\eeq
with the normalization
\andy{normpsi6}
\begin{equation}}\newcommand{\eeq}{\end{equation}\label{eq:normpsi6}
\langle\psi(t)|\psi(t)\rangle=|x(t)|^2+{\sum_{\bmsub k,\lambda}}'|y_{\bmsub k,
\lambda}(t)|^2+{\sum_{\bmsub k,\lambda}}'|z_{\bmsub k,
\lambda}(t)|^2=1,\qquad \forall t.
\eeq
At time $t=0$ we prepare our system in the state
\andy{condinxy6}
\begin{equation}}\newcommand{\eeq}{\end{equation}\label{eq:condinxy6}
|\psi(0)\rangle=|2;0,N_0\rangle\quad\Leftrightarrow\quad
x(0)=1,\;y_{\bmsub k\lambda}(0)=0,\;z_{\bmsub k\lambda}(0)=0,
\eeq
which is an eigenstate of the free Hamiltonian
\begin{equation}}\newcommand{\eeq}{\end{equation}
H_0|\psi(0)\rangle=H_0|2;0,N_0\rangle=\omega_0|2;0,N_0\rangle.
\eeq
We set, without any loss of generality, $E_1+N_0\Omega_0=0$. By
inserting (\ref{eq:statesdefin}) in the time-dependent
Schr\"odinger equation
\begin{equation}}\newcommand{\eeq}{\end{equation}
i\frac{d}{dt}|\psi(t)\rangle= H|\psi(t)\rangle
\eeq
and Laplace transforming with the initial conditions
(\ref{eq:condinxy6}), one readily obtains
\andy{xs,ys,zs}
\begin{eqnarray}}\newcommand{\earr}{\end{eqnarray}
\widetilde x(s)&=&\frac{1}{s+i\omega_0+Q(B,s)},\label{eq:xs}\\
\widetilde y_{\bmsub k\lambda}(s)&=&\frac{-i\phi_{\bmsub k\lambda}(s+i\omega_k)}
{(s+i\omega_k)^2+B^2}\;\widetilde x(s),\label{eq:ys}\\
\widetilde z_{\bmsub k\lambda}(s)&=&-\frac{\sqrt{N_0}\Phi^*_{\bmsub k_0\lambda_0}
\phi_{\bmsub k\lambda}}
{(s+i\omega_k)^2+B^2}\;\widetilde x(s),
\label{eq:zs}
\earr
where the tilde denotes Laplace transform and
\andy{Q(B,s)dipdiscr}
\begin{equation}}\newcommand{\eeq}{\end{equation}
\label{eq:Q(B,s)dipdiscr}
Q(B,s)=\sum_{\bmsub k,\lambda}|\phi_{\bmsub
k\lambda}|^2\frac{s+i\omega_k}{(s+i\omega_k)^2+B^2},
\qquad
B^2=N_0\,|\Phi_{\bmsub k_0\lambda_0}|^2.
\eeq
$B$ is the intensity of the laser field and can be viewed as the
``strength" of the observation performed by the laser beam on level
\#2 \cite{MPS}.
In the continuum limit ($V\rightarrow\infty$), the matrix elements scale
as follows
\andy{chi}
\begin{equation}}\newcommand{\eeq}{\end{equation}\label{eq:chi}
\lim_{V\rightarrow\infty}\frac{V\omega^2}{(2\pi)^3}\sum_\lambda\int d \Omega
|\phi_{\bmsub k\lambda}|^2 \equiv g^2\omega_0\chi^2(\omega) ,
\eeq
where $\Omega$ is the solid angle. The (dimensionless) function
$\chi(\omega)$ and the coupling constant $g$ have the following
general properties:
\andy{chiprop,g2}
\begin{eqnarray}}\newcommand{\earr}{\end{eqnarray}
\chi^2(\omega) & \propto &
\left\{ \begin{array} {ll}
\omega^{2j\mp 1} & \quad \mbox{if $\omega \ll \Lambda$}\\
\omega^{-\beta} & \quad \mbox{if $\omega \gg \Lambda$}
\end{array}
\right. ,
\label{eq:chiprop}\\
g^2 &=& \alpha (\omega_0 /\Lambda )^{2j+1\mp 1} , \label{eq:g2}
\earr
where $j$ is the total angular momentum of the photon emitted in
the $2\rightarrow 1$ transition, $\mp$ represent electric and
magnetic transitions, respectively, $\beta (> 1)$ is a constant,
$\alpha$ the fine structure constant and $\Lambda$ a natural cutoff
(e.g., of the order of the inverse Bohr radius), which determines
the range of the atomic or molecular form factor. The above
equations are due to general properties of the matrix elements
\cite{BLP,FPinduced}.
In order to scale the quantity $B$, we take the limit of very large cavity,
by keeping the density of $\Omega_0$-photons in the cavity constant:
\andy{limterm}
\begin{equation}}\newcommand{\eeq}{\end{equation}\label{eq:limterm}
V\rightarrow\infty,\qquad
N_0\rightarrow\infty,\quad\mbox{with}\quad
\frac{N_0}{V}=n_0=\mbox{const} .
\eeq
We obtain from (\ref{eq:Q(B,s)dipdiscr})
\begin{equation}}\newcommand{\eeq}{\end{equation}
B^2 = n_0 V |\Phi_{\bmsub k_0\lambda_0}|^2 =
(2\pi)^3 n_0 |\varphi_{\lambda_0} (\bm k_0)|^2
\eeq
where $\varphi$ is the scaled matrix element of the 1-3 transition.
As we shall see, in order to affect significantly the lifetime of
level \#2, we shall need a high value of $B$. It is therefore of
interest to consider a 1-3 transition of the dipole type, in which
case the above formula reduces to
\begin{equation}}\newcommand{\eeq}{\end{equation}
B^2 = 2\pi\alpha\Omega_0|\bm{\epsilon}_{\bmsub{k_0}\lambda_0}^*
\cdot\bm x_{13}|^2 n_0,
\eeq
where $\bm x_{13}$ is the dipole matrix element.
\subsection{Laser off}
\label{laseroff}
\andy{laseroff}
Let us first look at the case $B=0$. The laser, tuned at the 1-3
frequency $\Omega_0$, is off and we expect to recover the
well-known physics of a two-level system prepared in an excited
state and coupled to the radiation field \cite{FP1}. In this case,
$Q(0,s)$ is nothing but the self-energy function
\andy{sef}
\begin{equation}}\newcommand{\eeq}{\end{equation}
Q(s)\equiv Q(0,s)=\sum_{\bmsub k,\lambda}|\phi_{\bmsub
k\lambda}|^2\frac{1}{s+i\omega_k}
\label{eq:sef}
\eeq
and becomes, in the continuum limit,
\andy{Q(s)cont}
\begin{equation}}\newcommand{\eeq}{\end{equation}\label{eq:Q(s)cont}
Q(s) \equiv g^2 \omega_0 q(s)\equiv
-i g^2 \omega_0 \int_0^\infty d\omega
\frac{\chi^2(\omega)}{\omega-is},
\eeq
where $\chi$ is defined in (\ref{eq:chi}). The function $\widetilde
x(s)$ in Eq.\ (\ref{eq:xs}) (with $B=0$) has a logarithmic branch
cut extending from 0 to $-i\infty$, no singularities on the first
Riemann sheet (physical sheet) and a simple pole on the second
Riemann sheet. The pole equation is
\begin{equation}}\newcommand{\eeq}{\end{equation}
s+i\omega_0+g^2\omega_0 q_{\rm II}(s)=0,
\eeq
where
\begin{equation}}\newcommand{\eeq}{\end{equation}
q_{\rm II}(s)=q(s e^{-2\pi i})=q(s)+2 \pi\chi^2 (is)
\eeq
is the determination of $q(s)$ on the second Riemann sheet. We note
that $g^2 q(s)$ is $O(g^2)$, so that the pole can be found
perturbatively: by expanding $q_{\rm II}(s)$ around $-i\omega_0$ we
get a power series, whose radius of convergence is $R_c=\omega_0$
because of the branch point at the origin. The circle of
convergence lies half on the first Riemann sheet and half on the
second sheet (Figure \ref{fig:fig1}).
\begin{figure}
\centerline{\epsfig{file=fig1.eps,height=5.5cm}}
\caption{%
Fig.\ 3.1. Cut and pole in the $s$-plane ($B=0$). I and II are the
first and second Riemann sheets, respectively. }
\label{fig:fig1}
\end{figure}
The pole is well inside the convergence circle,
because $|s_{\rm pole}+i\omega_0|\sim g^2\omega_0\ll R_c$, and
we can write
\begin{equation}}\newcommand{\eeq}{\end{equation}
s_{\rm pole}= -i\omega_0-g^2\omega_0 q_{\rm II}(-i\omega_0-0^+)+O(g^4)
=-i\omega_0-g^2 \omega_0 q(-i\omega_0+0^+)+O(g^4),
\eeq
because $q_{\rm II}(s)$ is the analytical continuation of $q(s)$
below the branch cut. By using the formula
\begin{equation}}\newcommand{\eeq}{\end{equation}
\lim_{\varepsilon\rightarrow 0^+}\frac{1}{x\pm i\varepsilon}=
P\frac{1}{x}\mp i\pi\delta(x),
\eeq
one gets from (\ref{eq:Q(s)cont})
\andy{Q(-ieta)}
\begin{eqnarray}}\newcommand{\earr}{\end{eqnarray}
q(-i\eta+0^+)&=&-i\int_0^\infty d\omega\,\chi^2(\omega)\frac{1}
{\omega-\eta-i0^+}\nonumber\\ &=&\pi
\chi^2(\eta)\theta(\eta)-iP\int_0^\infty d\omega\,
\chi^2(\omega)\frac{1} {\omega-\eta}
\label{eq:Q(-ieta)}
\earr
and by setting
\andy{spole}
\begin{equation}}\newcommand{\eeq}{\end{equation}\label{eq:spole}
s_{\rm pole}=-i\omega_0+i\Delta E-\frac{\gamma}{2},
\eeq
one gets
\andy{Fgr}
\begin{equation}}\newcommand{\eeq}{\end{equation}
\gamma=2\pi g^2\omega_0\chi^2(\omega_0)+O(g^4),
\qquad
\Delta E=g^2\omega_0 P \int_0^\infty\frac{\chi^2(\omega)}{\omega-\omega_0}
+O(g^4),
\label{eq:Fgr}
\eeq
which are the Fermi Golden Rule and the second order correction to
the energy $\omega_0$ of level \#2.
\subsection{Laser on}
\andy{laseron}
We now turn our attention to the situation with the laser switched
on, $B\neq0$. The self energy function $Q(B,s)$ in
(\ref{eq:Q(B,s)dipdiscr}) depends on the value of $B$ and can be
written in terms of the self energy function $Q(s)$ in absence of
laser field [Eq.\ (\ref{eq:sef})], by making use of the following
remarkable property:
\andy{propnotev}
\begin{eqnarray}}\newcommand{\earr}{\end{eqnarray}
Q(B,s)&=&\frac{1}{2}\sum_{\bmsub k,\lambda}|\phi_{\bmsub
k\lambda}|^2\left(\frac{1}{s+i\omega_k+iB}+\frac{1}{s+i\omega_k-iB}\right)
\nonumber \\
&=& \frac{1}{2}\left[Q(s+iB)+Q(s-iB)\right].
\label{eq:propnotev}
\earr
Notice, incidentally, that in the continuum limit ($V\to\infty$), due to the above formula, $Q(B,s)$ scales
like $Q(s)$.
The position of the pole $s_{\rm pole}$
(and as a consequence the lifetime $\tau_{\rm
E}\equiv\gamma^{-1}=-1/2\Re e s_{\rm pole}$) depends on the value of $B$.
There are now two branch cuts in the complex $s$
plane, due to the two terms in (\ref{eq:propnotev}). They lie over
the imaginary axis, along $(-i\infty,-iB]$ and
$(-i\infty,+iB]$.
The pole satisfies the equation
\begin{equation}}\newcommand{\eeq}{\end{equation}
s + i\omega_0 + Q(B, s)=0,
\eeq
where $Q(B,s)$ is of order $g^2$, as before, and can again be
expanded in power series around $s=-i\omega_0$, in order to find
the pole perturbatively. However, this time one has to choose the
right determination for the function $Q(B,s)$. There are two cases:
a) The branch point $-iB$ is situated above $-i\omega_0$, so that
$-i\omega_0$ lies on both cuts. See Figure \ref{fig:tagli}(a); b)
The branch point $-iB$ is situated below $-i\omega_0$, so that
$-i\omega_0$ lies only on the upper branch cut. See Figure
\ref{fig:tagli}(b).
\begin{figure}
\centerline{\epsfig{file=fig2.eps,height=5.5cm}}
\caption{%
Fig.\ 3.2. Cuts and pole in the $s$-plane ($B \neq 0$) and
convergence circle for the expansion of $Q(B,s)$ around
$s=-i\omega_0$. I , II and III are the first, second and third
Riemann sheets, respectively. (a) $B<\omega_0$. (b) $B>\omega_0$.}
\label{fig:tagli}
\end{figure}
In case a), i.e.\ for $B<\omega_0$, the pole is on the third
Riemann sheet (under both cuts) and the power series converges in a
circle lying half on the first and half on the third Riemann sheet,
within a convergence radius $R_c = \omega_0-B$, which decreases as
$B$ increases [Figure~\ref{fig:tagli}(a)]. In case b), i.e.\ for
$B>\omega_0$, the pole is on the second Riemann sheet (under the
upper cut only) and the power series converges in a circle lying
half on the first and half on the second Riemann sheet, within a
convergence radius $R_c = B-\omega_0$, which increases with $B$
[Figure~\ref{fig:tagli}(b)]. In both cases we can write, for
$|s_{\rm pole}+i\omega_0|<R_c = |B-\omega_0|$,
\andy{poloB}
\begin{eqnarray}}\newcommand{\earr}{\end{eqnarray}
s_{\rm pole} &=&-i\omega_0-\frac{1}{2}\left\{Q[-i(\omega_0+B)+0^+]
+Q[-i(\omega_0-B)+0^+]\right\}+O(g^4) \nonumber \\
&=&-i \omega_0-\frac{1}{2}g^2 \omega_0\left\{q[-i(\omega_0+B)+0^+]
+q[-i(\omega_0-B)+0^+]\right\}+O(g^4) . \nonumber \\
\label{eq:poloB}
\earr
Our analysis is therefore valid when $|B-\omega_0|$ is larger than
$g^2 \omega_0$, (we remind the reader that we are interested in
large values of $B$). Equation (\ref{eq:poloB}) enables us to
analyze the behavior of the lifetime of level \#2.
\section{Decay rate vs $B$}
We write, as in (\ref{eq:spole}),
$s_{\rm pole}=-i\omega_0+i\Delta E(B)-\gamma(B)/2$. By
substituting (\ref{eq:Q(-ieta)}) into (\ref{eq:poloB}) and taking
the real part, one obtains the following expression for the decay
rate
\andy{gamma(B)}
\begin{equation}}\newcommand{\eeq}{\end{equation}\label{eq:gamma(B)}
\gamma(B)=\pi g^2 \omega_0 \left[\chi^2(\omega_0+B)+
\chi^2(\omega_0-B)\theta(\omega_0-B)\right] +O(g^4),
\eeq
which becomes, taking into account (\ref{eq:Fgr}),
\andy{MPSs}
\begin{equation}}\newcommand{\eeq}{\end{equation}\label{eq:MPSs}
\gamma(B)= \frac{\gamma}{2} \; \frac{\chi^2(\omega_0+B)+
\chi^2(\omega_0-B)\theta(\omega_0-B)}{\chi^2(\omega_0)} +O(g^4).
\eeq
This is our main result and involves no approximations. It
expresses the lifetime $\gamma(B)^{-1}$, when the system is bathed
in an intense laser field $B$, in terms of the ``ordinary" lifetime
$\gamma^{-1}$, when there is no laser field. By taking into account
the general behavior (\ref{eq:chiprop}) of the matrix elements
$\chi^2(\omega)$ and substituting into (\ref{eq:MPSs}), one gets to
$O(g^4$)
\andy{gamma(B)dip}
\begin{equation}}\newcommand{\eeq}{\end{equation}\label{eq:gamma(B)dip}
\gamma(B) \simeq
\frac{\gamma}{2}\left[\left(1+\frac{B}{\omega_0}\right)^{2j\mp 1}
+\left(1-\frac{B}{\omega_0}\right)^{2j\mp 1}
\theta(\omega_0-B)\right], \qquad (B \ll \Lambda)
\eeq
where $\mp$ refers to $1-2$ transitions of electric and magnetic
type, respectively. Observe that, since $\Lambda \sim$ inverse Bohr
radius, the case $B \ll \Lambda$ is the physically relevant one
\cite{FPinduced}. The decay rate is profoundly modified by the
presence of the ``$B$"-field. Its behavior is shown in
Figure~\ref{fig:gamma(B)dip} for a few values of $j$. The case
$j=1$ (transition of electric dipole type) yields a constant value
up to $B=\omega_0$; above this threshold, $\gamma$ increases
linearly with $B$. For $j > 1$ the derivative of $\gamma (B)$ is
continuous. In general, the decay rate $\gamma(B)$ increases with
$B$, so that the lifetime $\gamma(B)^{-1}$ decreases as $B$ is
increased. If one looks at $B$ as the strength of the
``observation" performed by the laser beam on level \#2 \cite{MPS},
one can view this phenomenon as an ``inverse" quantum Zeno effect.
\begin{figure}
\centerline{\epsfig{file=fig3.eps,width=\textwidth}}
\caption{%
Fig.\ 4.1. The decay rate $\gamma(B)$ vs $B$, for electric
transitions with $j=1,2,3$; $\gamma(B)$ is in units $\gamma$ and
$B$ in units $\omega_0$. Notice the different scales on the
vertical axis.}
\label{fig:gamma(B)dip}
\end{figure}
Equation (\ref{eq:gamma(B)dip}) is valid for $B\ll\Lambda$. In the
opposite case, $B\gg\Lambda$, by (\ref{eq:chiprop}) and
(\ref{eq:MPSs}), one gets to $O(g^4$)
\andy{gammaom}
\begin{equation}}\newcommand{\eeq}{\end{equation}\label{eq:gammaom}
\gamma(B) \sim \frac{\gamma}{2} \; \frac{\chi^2(B)}{\chi^2(\omega_0)}
\propto B^{-\beta}.
\qquad (B \gg \Lambda)
\eeq
This is essentially the result obtained in Ref.\ \cite{MPS}. If
such high values of $B$ were experimentally obtainable, the decay
would be considerably hindered (quantum Zeno effect \cite{MPS}).
A thorough investigation of the phenomenon we have just proposed is
in preparation \cite{FPinduced}. It involves a complete discussion
in terms of Fano states and electromagnetic induced transparency
\cite{induced} and an analysis of the decay rates from level \#2 to
the dressed atomic or molecular states.
\medskip
\noindent {\bf Acknowledgments}
We thank E. Mihokova and L.S.\ Schulman for many comments and
useful discussions.
|
\section{Introduction}
Some years ago 't Hooft introduced the concept of Abelian projection
\cite{H1} into non-Abelian gauge theories, in order to explain the confinement
of quarks in four-dimensional $QCD$ as a dual Meissner effect in a dual
superconductor \cite{H02, M}. \\
The Abelian projection allows us, by a careful choice of the gauge, to
describe the physical variables of a non-Abelian $SU(N)$ gauge theory, without
scalar matter fields, as a set of electric charges and magnetic monopoles
interacting via a residual $U(1)^{N-1}$ Abelian gauge coupling. \\
The occurrence of magnetic monopoles into a non-Abelian gauge theory
without matter fields is perhaps the most crucial feature of the Abelian
projection, that furnishes a precise understanding of the structure of the
phases of non-Abelian gauge theories, according to the following
alternatives \cite{H3}. \\
If there is a mass gap, either the electric charge condenses in the vacuum
(Higgs phase) or the magnetic charge does (confinement phase).
If there is no mass gap, the electric and magnetic fluxes coexist
(Coulomb phase). \\
Recently, in an apparently unrelated development \cite{MB}, some mathematical
control was gained over the large-$N$ limit of four-dimensional $QCD$, mapping,
by means of a chain of changes of variables, the function space of the $QCD$
functional integral into an elliptic fibration of Hitchin bundles. \\
Hitchin bundles \cite{Hi} are themselves a fibration of $U(1)$
bundles over spectral branched covers of a Riemann surface, that, in the case
of \cite{MB}, is a torus. \\
In this paper, we point out that the map in \cite{MB} is a version,
in a perhaps global algebraic-geometric setting, of the concept of Abelian
projection \cite{H1}. \\
In fact, the branching points of the spectral cover are identified with the
magnetic monopoles of the Abelian projection, the parabolic points of the cover
with (topological) electric charges and the $U(1)$ gauge group on the cover
with a global version (on the cover) of the $U(1)^{N-1}$ gauge group of the
Abelian projection. \\
The identifications that we have just outlined provide a physical
interpretation of the mathematical construction in \cite{MB}.
Indeed it is precisely this physical interpretation
that explains naturally why the functional integral, once it is expressed as a
functional measure supported over the collective field of the Hitchin fibration,
is dominated by a saddle-point condition in the large-$N$ limit. \\
On the other side, we may think that the mathematical proof, that the
variables of the Abelian projection really capture the
physics of four-dimensional $QCD$ in the large-$N$ limit, relies on the fact
that those variables may be employed to dominate the functional integral in
the large-$N$ limit. \\
The only qualitative feature, in the treatment in \cite{MB}, that was not
already present in the concept of the Abelian projection, is the occurrence
of Riemann surfaces and it is due to the global algebraic-geometric nature
of the methods in \cite{MB}. This, however, makes contact, at least
qualitatively, with another long-standing conjecture about the $QCD$
confinement, the occurrence of string world sheets \cite{GT} and the string
program \cite{Po}. \\
Our last concluding remark is that the electric/magnetic alternative
\cite{H3} and the physical interpretation based on the Abelian projection,
applied in the mathematical framework of \cite{MB}, give us a simple
qualitative criterium to characterize the confinement phase of $QCD$ in the
large-$N$ limit: confinement is equivalent to magnetic condensation, in absence
of electric (parabolic) singularities of the spectral covers. \\
An alternative, compatible interpretation, based on the idea that $QCD$
is equivalent, in the large-$N$ limit, to a theory of strings \cite{GT,Po}
is outlined in the following section.
The rest of the paper is devoted to a technical explanation of the
correspondence between the Abelian projection and the Hitchin fibration
in four-dimensional $QCD$.
\section{The Hitchin fibration as the Abelian projection in the gauge in which
the Higgs current is a triangular matrix}
The Abelian projection, according to \cite{H1}, is really the choice of a
gauge-fixing in such a way that, after the gauge-fixing, the theory is no
longer locally invariant under $SU(N)$ but only under its Cartan subgroup
$U(1)^{N-1}$. The important point about this projection is that it is defined
strictly locally, that is, the gauge rotation $\Omega$ performed at each point
in space-time to implement the gauge-fixing condition, does not depend on the
values of the physical fields in other points of space-time.
This then guarantees that all observables in the new
gauge frame are still locally observable. There are no propagating ghosts.
But $\Omega$ is not completely defined. There is a subgroup,
$U(1)^{N-1}$, of gauge rotations that may still be performed. And
this is why the theory, after the Abelian projection, looks like a local
$U(1)^{N-1}$ gauge theory. \\
If one now tries to gauge-fix this remaining gauge freedom, one
discovers that it cannot be done locally, without encountering
apparent difficulties. But local gauge-fixing is not needed, since
the residual gauge symmetry is the one of a familiar Abelian theory. \\
There may be, however, isolated points, where the local gauge-fixing condition
has coinciding eigenvalues, where the gauge symmetry is not $U(1)^{N-1}$ but a
larger group. Here singularities appear, the magnetic monopoles.
So we see that, topologically, the full theory can only be
topologically equivalent to the $U(1)^{N-1}$ gauge theory if the
latter is augmented with monopole singularities where the $U(1)$
conservation laws for the vortices are broken down into the (less restrictive)
conservation laws of the $SU(N)$ vortices. \\
When we try to gauge-fix completely, we hit upon the Dirac strings, whose
end points are the magnetic monopoles. \\
In addition to the magnetic monopoles, in the $QCD$ case, the gauge-fixed
theory contains also gluon and quark fields, that are charged with respect to
the residual $U(1)^{N-1}$. \\
Therefore we have a set of electric charges and magnetic monopoles
interacting via a residual $U(1)^{N-1}$ Abelian gauge coupling. \\
We now compare this description with the one that arises in \cite{MB}, for
the pure gauge theory without quark matter fields. \\
The functional integral for $QCD$ in \cite{MB} is defined in terms of the
variables $(A_z, A_{\bar z}, \Psi_z, \Psi_{\bar z})$, obtained by means of a
partial duality transformation from $(A_z, A_{\bar z}, A_u, A_{\bar u})$,
where $(z, \bar z, u, \bar u)$ are the complex coordinates on the product of
two two-dimensional tori, over which the theory is defined. \\
$(A_z, A_{\bar z}, \Psi_z, \Psi_{\bar z})$ define the coordinates of an
elliptic fibration of $T^* {\cal A}$, the cotangent bundle of unitary
connections on the $(z, \bar z)$ torus, whose base is the $(u, \bar u)$ torus.
\\
$\Psi_z$ transforms as a field strength by gauge transformations
and it is a non-hermitian matrix. \\
Following Hitchin \cite{Hi}, the gauge is chosen in which $\Psi_z$ is a
triangular matrix, for example lower triangular, that leaves a $U(1)^{N-1}$
residual gauge freedom as in the Abelian projection. \\
The points in space-time where $\Psi_z$ has a pair of coinciding eigenvalues,
correspond to monopoles.
In addition there are the charged components of
$(A_z, A_{\bar z}, \Psi_z, \Psi_{\bar z})$.
We have thus a set of charges and monopoles with a residual $U(1)^{N-1}$,
according to the Abelian projection. \\
In \cite{MB}, however, it is found a dense set in the functional
integral over (the elliptic fibration of) $T^* {\cal A}$, with the property that
the quotient by the action of the gauge group exists as a Hausdorff (separable)
manifold. \\
This dense set is defined in \cite{MB} as the set of
pairs $(A, \Psi)$ that are solutions of the following differential equations
(elliptically fibered over the $(u, \bar u)$ torus):
\begin{eqnarray}
F_A-i \Psi \wedge \Psi &=& \frac{1}{|D|}\sum_p \mu^{0}_{p} \delta_p
i dz \wedge d\bar{z} \nonumber \\
\bar{\partial}_A \psi &=& \frac{1}{|D|}\sum_p \mu_{p} \delta_p
dz \wedge d\bar{z} \nonumber \\
\partial_A \bar{\psi} &=& \frac{1}{|D|}\sum_p \bar{\mu}_{p} \delta_p
d\bar{z} \wedge dz \;
\end{eqnarray}
where $\delta_p$ is the two-dimensional delta-function localized at $z_p$
and $(\mu^{0}_{p},\mu_{p},\bar{\mu}_{p})$ are the set of levels for the
moment maps. The moment maps are the three Hamiltonian densities generating
gauge transformations on $T^* {\cal A}$ that appear in the left hand sides of
Eq.(1) \cite{Hi1}. \\
$\mu^{0}_{p}$ are hermitian traceless matrices, and $\mu_{p}$ are matrices in
the complexification of the Lie algebra of $SU(N)$, that
determine the residues of the poles the Higgs current $\Psi$.
$\psi$ and $\bar{\psi}$ are the $z$ and $\bar z$ components of the one-form
$\Psi$. \\
Eq.(1) defines a dense stratification of the functional integral over
$T^* {\cal A}$ because the set of levels is dense everywhere in function
space, in the sense of the distributions, as the divisor $D$ gets larger and
larger. \\
Eq.(1) defines the data of parabolic $K(D)$ pairs \cite{K} on a torus valued
in the Lie algebra of the complexification of $SU(N)$: a holomorphic
connection $\bar{\partial}_A$ of
a holomorphic bundle, $E$, with a parabolic structure and a parabolic morphism
$\psi$ of the parabolic bundle. The parabolic structure at a point $p$
\cite{MS,K}
consists in the choice of a set of ordered weights, that are positive real
numbers modulo 1, and a flag structure, that
is a collection of nested subspaces $ \cal{F}_{1}
\subset \cal{F}_{2} \subset...\cal{F}_{i}$ labelled by the weights $\alpha_1 \geq \alpha_2
\geq ...\alpha_k$, with the
associated multiplicities defined as: $m_{i+1}=dim \cal{F}_{i+1}-dim
\cal{F}_{i}$.
A parabolic morphism, $\phi$, is a holomorphic map between parabolic
bundles,
$E^1,E^2$, that preserves the parabolic flag structure at each parabolic
point $p$ in the sense that $\alpha^{1}_{i} > \alpha^{2}_{j}$ implies
$\phi( \cal{F}^{1}_{i}) \subset \phi(\cal{F}^{2}_{j+1})$.
We should now explain how a parabolic structure arises from Eq.(1) and how
it follows that $\psi$ is a parabolic morphism with respect to the given
parabolic structure.
Though we are going to choose the gauge in which $\psi$ is a lower triangular
matrix in
most of this paper, we start at an intermediate stage with a
gauge in which $\mu^{0}_{p}$ is diagonal. The eigenvalues of $\mu^{0}_{p}$
modulo $2 \pi$ and divided by $2 \pi$ define the parabolic weights.
Their multiplicities will turn out to be the multiplicities of the yet to
be defined flag structure.\\
Fixed $\mu^{0}_{p}$ and $\mu_p$ in Eq.(1), let $(e_k)$ be an orthonormal
basis of the eigenvectors of $\mu^{0}_{p}$ in decreasing order. This basis
is not necessarily unique if the eigenvalues have non-trivial
multiplicities. However the corresponding flag structure will not be affected
by this lack of uniqueness.
Let $g$ be
the gauge transformation that puts $\mu$ and $\psi$ into lower
triangular form. Let $(g e_k)$ be
the transformed basis and let $\cal F$ be the flag obtained by taking the unions
of subspaces generated by the vectors in the transformed basis
that are the images of eigenvectors of the ordered
eigenvalues with the given multiplicity in such a way that the multiplicities
of the resulting flag are the same as the multiplicities of
the eigenvalues. In addition, by construction, $\psi$ is a parabolic morphism
with respect to the flag since it is holomorphic and lower triangular in the
basis $(ge_k)$. \\
We have thus the data of a parabolic $K(D)$ pair from Eq.(1). \\
There is also a representation theoretic interpretation of Eq.(1). \\
The three equations for the moment maps are equivalent to a vanishing
curvature condition for the non-hermitian connection one-form $B=A+i \Psi$ plus
a harmonic condition for $\psi$ away from the parabolic
divisor \cite{S}. \\
Therefore the set of solution of Eq.(1) can be figured out essentially as a
collection of monodromies around the points of the divisor with values in the
complexified gauge group, that form a representation of the fundamental group
of the torus with the points of the parabolic divisor deleted. \\
't Hooft description of the Abelian projection previously outlined,
applies to $T^* {\cal A}$ and to its dense subset defined by Eq.(1)
a fortiori. In addition, we have just shown that there is an embedding of the
solutions of Eq.(1) into the parabolic $K(D)$
pairs. \\
However, on the parabolic $K(D)$ pairs, 't Hooft concept of Abelian projection
can be carried to its extreme consequences. \\
Indeed, in the global algebraic-geometric framework of the Hitchin fibration
\cite{Hi,K} of parabolic $K(D)$ pairs, it is preferable to concentrate
ourselves on the first eigenvalue and the first eigenstate of the lower
triangular matrix $\Psi_z$, since all the information of the original
parabolic bundle, up to gauge equivalence, can be reconstructed from these
only data \cite{Hi}. \\
The first eigenvalue defines a spectral covering, that is a branched cover
of the two-torus. The eigenspace defines a section of a line bundle, that
determines a $U(1)$ connection on the cover of the torus, instead of the
$U(1)^{N-1}$ bundle on the torus of the Abelian projection. \\
The $U(1)$ connection on the cover, $a$, and the eigenvalue, $ \lambda$, of the
Higgs current can be considered as coordinates of the cotangent bundle of
unitary $U(1)$ connections on the cover, or as parabolic $K(D)$
pairs $(a, \lambda)$ on the cover, valued in the complexification of the Lie
algebra of $U(1)$. \\
The system is now completely abelianized. Correspondingly, not only
the magnetic charges, but also the electric ones can occur only as gauge
invariant topological configurations. \\
The points in space-time where $\Psi_z$ has a pair of coinciding eigenvalues,
that in the Abelian projection correspond to monopoles, are here,
according to Hitchin, simple branching points of the spectral covers, defined
by means of the characteristic equation:
\begin{eqnarray}
Det(\lambda 1-\Psi_{z})=0 ,
\end{eqnarray}
in which the coordinates $(u,\bar{u})$ are kept fixed. \\
All the other branching points can be obtained by collision of these simple
branching points, in the same way monopoles can in the Abelian projection.
The branching points are the end points of string cuts on the Riemann
surfaces, the Dirac strings of the Abelian projection. \\
These Riemann surfaces, the only additional global ingredient with respect
to the Abelian projection, are interpreted as the world sheets of
strings made by electric flux lines. \\
A closed string of electric flux is represented by a Wilson loop of the
$U(1)$ connection $a$ on the cover, along a non-trivial generator of the
fundamental group of the surface. \\
In addition, the Riemann surfaces, defined by the spectral equation,
may posses parabolic points, associated to poles of the eigenvalues
of the Higgs current $\Psi_z$, whose origin is in the parabolic
singularities of the original $su_{c}(N)$-valued $K(D)$ pair, which may be
reflected into a parabolic structure for the $u_{c}(1)$-valued $K(D)$ pair on the
cover. \\
These poles, together with the ones of the $U(1)$ connection, are interpreted
as electric charges. Indeed it is not difficult to see that they are electric
sources, that appear where a boundary-electric loop shrinks to a point. \\
Therefore, the electric charges occur here as topological objects associated
to the parabolic degree \cite{MS} of the $u_{c}(1)$-valued $K(D)$ pair.
On the other side, magnetic topological quantum numbers are associated,
as usual, to the ordinary degree of the $U(1)$ bundle. \\
We should mention however that a subtlety arises in our interpretation
of the Hitchin fibration in terms of the Abelian projection.
As we mentioned in the first part of this section, in the Abelian projection
the gauge-fixing condition leaves a residual non-Abelian gauge symmetry
where a magnetic monopole occurs. This is essentially due to the fact that 't
Hooft chooses to diagonalize a hermitian functional of the fields. On the
contrary, in the case of the dense set
defined by Eq.(1), since $\psi$ is a non-hermitian matrix, it can only be put
in triangular form. This gauge-fixing does not leave in general
a residual compact non-Abelian gauge symmetry even when the eigenvalues
coincide.
However this difficulty can be resolved in the following way, anticipating
somehow some of the conclusions of this paper and the result of \cite{MB2}.
Let us require for the moment that the levels of the non-hermitian moment maps
be nilpotent. Since these are only $N$ conditions at each parabolic point they
do not modify essentially the entropy of the functional integration in the
large-$N$ limit. The true physical meaning of this choice has to do with
confinement and it is explained in \cite{MB2}.
If the residues of the Higgs field are nilpotent, Eq.(1) can be interpreted
as the vanishing condition for the moment maps of the action of the compact
$SU(N)$ gauge group on the pair $(A, \Psi)$ and on the cotangent space of
coadjoint orbits \cite{Ale}:
\begin{eqnarray}
&&F_A-i \Psi \wedge \Psi - \frac{1}{|D|}\sum_p \mu^{0}_{p} \delta_p
i dz \wedge d\bar{z}=0 \nonumber \\
&&\bar{\partial}_A \psi- \frac{1}{|D|}\sum_p n_{p} \delta_p
dz \wedge d\bar{z}=0 \nonumber \\
&&\partial_A \bar{\psi}- \frac{1}{|D|}\sum_p \bar{n}_{p} \delta_p
d\bar{z} \wedge dz=0 \;
\end{eqnarray}
In addition the quotient under the action of the compact gauge group is
hyper-Kahler \cite{K}. By a general result of Hitchin, Karlhede, Lindstr\"{o}m and
Roc\v{e}k \cite{H2}, the hyper-Kahler quotient under the
action of the compact gauge group in Eq.(3) is the same as the quotient
defined by the non-hermitian moment maps:
\begin{eqnarray}
&&\bar{\partial}_A \psi- \frac{1}{|D|}\sum_p n_{p} \delta_p
dz \wedge d\bar{z}=0 \nonumber \\
&&\partial_A \bar{\psi}- \frac{1}{|D|}\sum_p \bar{n}_{p} \delta_p
d\bar{z} \wedge dz=0 \;
\end{eqnarray}
under the action of the complexification of the gauge group.
We can therefore impose a gauge condition
compatible with the compact action in Eq.(3) or a gauge condition compatible
with the action of the complexified group in Eq.(4) getting the same moduli
space.
In the second case we choose the gauge in which
$\psi$ is diagonal. This condition becomes singular
where two or more eigenvalues coincide. In fact it cannot be extended
continuously to the points where the eigenvalues coincide. There it can only
be required that $\Psi_z$ be a triangular matrix. However this condition
leaves now a residual non-Abelian gauge symmetry in the complexification of the gauge
group: the freedom of making triangular gauge transformations, thus
confirming our analogy with 't Hooft definition of magnetic monopoles. \\
To summarize, the ingredients of the Hitchin fibration of the $su_{c}(N)$-
valued $K(D)$ pairs, are the branching points, that are interpreted as
magnetic monopoles, and the $U(1)$ monodromies around closed loops, that are interpreted
as electric lines. In addition, the ordinary degree of the $U(1)$
bundle is interpreted as a (topological) magnetic charge, while the parabolic
degree \cite{MS} of the $U(1)$ bundle is interpreted as a (topological) electric charge.
\\
The difference here, with the letter but not with the spirit of the Abelian
projection, is that the system has been completely abelianized, so
that both the magnetic and the electric charges are topological.
We are thus given a set of charges and monopoles with a $U(1)$ gauge group on
the covering, in analogy with the Abelian projection. \\
We call this description a complete Abelian projection. \\
The string interpretation is as follows.
The spectral covers are the world sheets of strings, made by the electric flux
lines. The confinement condition is equivalent to requiring
that only closed string world sheet occur, since confinement requires
that the flux lines can never break in absence of quarks. \\
If the spectral covers posses parabolic points, the same as
electric charges in the complete Abelian projection, they are,
topologically, Riemann surfaces with boundaries at infinity. \\
For example a sphere with two parabolic points is a topological cylinder. \\
But a cylinder can occur in vacuum string world sheets (we are describing
the contributions to the partition function, the vacuum to vacuum amplitude
indeed) only if open strings propagate. \\
In fact, a closed string that propagates through the torus
breaks into an open one at the parabolic points, since the parabolic points do
not belong to the world sheet. \\
On the contrary, when a closed string meets a
branching point, for example in a once-branched double cover of a torus, the
closed string is pinched into another closed string with the form of a double
loop intersecting at the (simple) branching point. \\
Notice also that the branching points do belong to the world sheet. \\
Thus, the string picture is consistent with the interpretation
of branching points as magnetic charges, where the string electric line can
self-intersect but not break, and of parabolic points as electric
charges, where closed string break into open strings with the parabolic points
as boundaries.
\section{Conclusions}
Our conclusion is that the concept of Abelian projection in \cite{H1} furnishes
a physical interpretation of the structures that appear in the
Hitchin fibration of $K(D)$ pairs, as it is embedded in the $QCD$
functional integral in \cite{MB}. \\
In addition, there is a complementary consistent string interpretation. \\
The most relevant consequence of these interpretations is a criterium
for electric confinement in the framework of \cite{MB}, that is the usual
criterium of magnetic condensation of \cite{H1}.\\
Therefore, if $QCD$ confines the electric charge, the functional measure must
be localized, in the large-$N$ limit, on those parabolic $K(D)$ pairs, whose
image through the Hitchin map, contains monopoles but no charges, that is, in
geometric language, on those spectral covers that are arbitrarily branched,
but that do not posses a parabolic divisor. \\
In turn, this is equivalent to the condition that only spectral covers
spanned by closed strings occur as configurations in the vacuum to vacuum
amplitude. \\
It is amusing to notice that this condition is satisfied by the string
of two-dimensional $QCD$ in the large-$N$ limit \cite{GT}.
\section{Acknowledgements}
We would like to thank Gerard 't Hooft for several clarifying remarks
on the Abelian projection.
|
\section{Introduction}
\label{sec:introduction}
It is generally believed that there must be a regime
in which the gravitational field can be treated as a
classical or ``quasiclassical'' field, but its interaction with
quantum matter fields cannot be neglected.
The standard approach to describe such a regime is the
semiclassical theory of gravity based on the semiclassical
Einstein equation. This is a generalization of the Einstein
equation for a classical metric when the expectation value of the
stress-energy tensor of quantum matter fields is the source of
curvature. The semiclassical theory of gravity is mathematically
consistent and fairly well understood, at least for linear
matter fields \cite{wald94,wald77,fulling,birrell,mostepanenko}.
One expects that semiclassical gravity could be derived
as an approximation of
a fundamental quantum theory of gravity.
However, in the absence of such a fundamental theory, the scope and
limits of the semiclassical theory are less well understood
\cite{wald94,flanagan}.
It has been pointed out, nevertheless, that this semiclassical theory
may not be valid when the matter fields have important quantum
stress-energy
fluctuations \cite{wald94,wald77,birrell,ford82}.
When this is the case, the stress-energy
fluctuations may have relevant back-reaction effects on the spacetime
geometry in the form of induced gravitational
fluctuations \cite{ford82}.
A number of examples have been studied, both in cosmological
and in flat spacetimes, where, for some states
of the matter fields, the stress-energy tensor have significant
fluctuations \cite{kuo93}.
It is thus necessary to extend the semiclassical theory of
gravity to determine the effect of such fluctuations.
To address this problem, different approaches have been
adopted. The aim of the first part of the present paper is to
unify two of these approaches.
One of these approaches relies on the idea, first proposed
by Hu \cite{hu89} in the context of semiclassical cosmology,
of viewing the metric field as the ``system'' of interest and the
matter fields as being part of its ``environment.''
This approach leads naturally to
the influence functional formalism of Feynman and
Vernon \cite{feynman-vernon}.
In this formalism, the integration of the environment
variables in a path integral yields the influence functional, from
which one can define an effective action for the dynamics of the
system \cite{feynman-hibbs,calzettahu,humatacz,husinha,%
caldeira,hu-paz-zhang,hu-matacz94,greiner}.
This approach has been extensively used in the
literature, not only in the framework of semiclassical cosmology
\cite{calzettahu,humatacz,husinha,cv96,lomb-mazz,cv97,ccv97,%
campos-hu,campos-hu2,calver98}, but
also in the context of analogous semiclassical regimes in
quantum mechanics \cite{caldeira,hu-matacz94,hu-paz-zhang2}
and in quantum field theory
\cite{greiner,matacz,morikawa,shaisultanov,gleiser}.
It is based on the observation that the semiclassical equation can be
directly
derived from the effective action of Feynman and Vernon
\cite{calzettahu,greiner,cv96,ccv97,campos-hu,shaisultanov}.
When computing this effective action perturbatively up to quadratic
order in its variables, one usually finds some imaginary terms which
do not contribute to the semiclassical equation.
The key point is then to formally identify the contribution of such
terms in the influence functional with the characteristic functional of
a Gaussian stochastic source. Assuming that, in the semiclassical
regime, this stochastic source interacts with the system variables,
equations of the
Langevin type can be derived for these variables.
However, since this approach relies on a purely formal identification,
doubts may be raised on
the physical significance of the derived equations.
An alternative approach has been introduced in a recent
paper \cite{mv98}. In that work, we proposed a stochastic
semiclassical theory of gravity as a perturbative generalization of
semiclassical gravity to describe the back reaction of the lowest
order stress-energy fluctuations. The idea is in fact quite simple.
One starts realizing that, for
a given solution of semiclassical gravity, the lowest order matter
stress-energy fluctuations can be associated to a
classical stochastic tensor field.
Then, we seek an equation which incorporates in a consistent way
this stochastic tensor as the source of linear perturbations to the
semiclassical metric.
The resulting equation is the semiclassical Einstein-Langevin
equation.
We should emphasize that, even if the metric fluctuations in this
theory are classical (stochastic fluctuations), their
origin is presumably quantum. This is so not only because these metric
fluctuations are induced by the fluctuations of a quantum
operator, but also because they are supposed to describe some remnants
of the quantum gravity fluctuations after some mechanism for
decoherence and classicalization of the metric field
\cite{gell-mann-hartle,hartle,dowker,halliwell,whelan}.
From the formal assumption that such a mechanism is the Gell-Mann
and Hartle mechanism of environment-induced decoherence
of suitably coarse-grained system variables
\cite{gell-mann-hartle,hartle},
one may, in fact, derive the stochastic semiclassical theory
\cite{mv98_2}.
Nevertheless, that derivation is of course formal,
given that, due to the lack of the full quantum theory of gravity, the
classicalization mechanism for the gravitational field
is not understood.
One expects that the stochastic semiclassical theory is valid
when the characteristic time and space scales of variation
of the metric field are well above
its characteristic decoherence scales.
In this regime, the theory can be applied
to compute correlation functions of gravitational perturbations for
points separated by scales larger than these
decoherence scales.
Hence, this theory
may have a number of interesting applications in black hole physics
and in cosmology,
particularly in view of the problem of structure
formation. Some examples of simple applications have already been
given in Refs.~\cite{ccv97,calver98,mv98}.
The purpose of the second part of the paper is to derive
some general results concerning stochastic semiclassical
gravity for stationary and conformally stationary background solutions
of semiclassical gravity (for conformal matter fields in the latter
case). We analyze two issues: the existence of a
fluctuation-dissipation relation and the creation of particles by
stochastic metric perturbations.
Under very general conditions,
a fluctuation-dissipation relation is known to exist in
models of quantum mechanics, and also in some models of quantum
many-body systems or quantum fields
in the presence of classical fields
\cite{landau,kubo,grabert,schwinger61,weber,kubo85,martin,weber,%
jackiw}.
This is a relation between quantum fluctuations of a system
in a state of thermal equilibrium and the dissipative properties of
this system caused by classical linear perturbations on it.
The idea of a fluctuation-dissipation relation in the theory of
quantum fields in curved spacetimes and in the semiclassical
back-reaction problem was already present in some
early papers \cite{sciama,mottola,hu89}.
A fluctuation-dissipation relation has been found
in some of the previous derivations of semiclassical
Langevin-type equations
\cite{husinha,cv96,campos-hu,campos-hu2}.
Some authors believe that such a relation should always be
present and embody the physics of the back reaction of matter fields
on the gravitational field
\cite{husinha,campos-hu,campos-hu2,hu99,hu-raval-sinha}.
It is also believed that noise and dissipation must be
related to the creation of particles by stochastic metric
perturbations
\cite{hu89,calzettahu,humatacz,husinha,cv97,ccv97,hu99,hu-raval-sinha}.
In stationary and conformally stationary spacetimes
(for conformal fields in the latter case), one can define a state of
thermal equilibrium for the matter fields.
When the background solution of semiclassical
gravity is of one of these types, we can identify a dissipation
kernel in the corresponding semiclassical Einstein-Langevin equation
which is related to the fluctuations of the stochastic source by a
fluctuation-dissipation relation.
We also study the production of
particles by stochastic metric perturbations to such backgrounds:
we relate particle
creation to the vacuum stress-energy fluctuations and
we show that the
mean value of created particles is enhanced by the
presence of metric fluctuations.
The plan of the paper is the following.
In Sec.~\ref{sec:E-L}, we construct the stochastic semiclassical
theory of gravity to describe the back reaction of the stress-energy
fluctuations on the spacetime.
In Sec.~\ref{sec:influence action}, we show that the
semiclassical Einstein-Langevin equation obtained in
Sec.~\ref{sec:E-L} can actually be formally derived with the
functional approach. This connection
clarifies the physical meaning of the
Langevin-type equations previously derived by
functional methods
\cite{calzettahu,humatacz,husinha,cv96,lomb-mazz,cv97,ccv97,%
campos-hu,campos-hu2,calver98},
since it shows that the formally introduced stochastic source
is directly related to the matter stress-energy fluctuations.
We then use the functional approach to write the Einstein-Langevin
equation in an explicit form, which is more suitable
for specific calculations.
In Sec.~\ref{sec:stationary}, we derive the
fluctuation-dissipation relation for stationary and conformally
stationary backgrounds and the results for particle creation by
stochastic metric perturbations.
Finally, in Sec.~\ref{sec:conclu}, we summarize our main conclusions.
Throughout this paper we use the $(+++)$ sign conventions and the
abstract index notation of Ref.~\cite{wald84},
and we work with units in which $c=\hbar =1$.
\section{Stochastic semiclassical
gravity}
\label{sec:E-L}
In this section, we construct the stochastic semiclassical
theory of gravity as a perturbative extension of semiclassical
gravity to describe the back reaction of quantum stress-energy
fluctuations on the gravitational field.
Let us begin with a brief overview of the semiclassical
theory of gravity interacting with linear matter fields.
Let $({\cal M},g_{ab})$ be a globally hyperbolic four-dimensional
spacetime and consider a linear
quantum field $\Phi$ on it. For the sake of definiteness, we will
take $\Phi$ as a real scalar field, but all the analysis of this
section is valid for any kind of linear quantum field or
for a set of linear independent quantum fields. Throughout this
section we shall work in the Heisenberg picture.
The field operator in this picture, $\hat{\Phi}$, is an
operator-valued distribution solution of
the Klein-Gordon equation,
\begin{equation}
\left( \Box -m^2- \xi R \right) \hat{\Phi}=0,
\label{Klein-Gordon}
\end{equation}
where $m$ is the mass,
$\Box \!\equiv\! \bigtriangledown_{\!a}\bigtriangledown^{a}$,
with $\bigtriangledown_{\!a}$ being the covariant derivative
associated to
the metric $g_{ab}$, and $\xi$ is a dimensionless parameter coupling
the field to the scalar curvature $R$. To indicate that the field
operator is a functional of the metric $g_{ab}$, we will
write $\hat{\Phi}[g](x)$.
The classical stress-energy tensor is
obtained by functional derivation of
the classical action for the field in a background
spacetime $({\cal M},g_{ab})$
with respect to the metric. This tensor is a functional
$T_{ab}[g,\Phi]$ of the metric $g_{ab}$ and of the classical field
$\Phi$. For a real scalar field, it is
\begin{equation}
T_{ab}[g,\Phi]=\bigtriangledown_{\!a}\Phi
\bigtriangledown_{\!b}\!\Phi- {1\over 2}\, g_{ab}
\bigtriangledown^{c}\!\Phi \bigtriangledown_{\!c}\!\Phi
-{1\over 2}\, g_{ab}\, m^2 \Phi^2
+\xi \left( g_{ab} \Box
-\bigtriangledown_{\!a}\!\! \bigtriangledown_{\!b}
+\, G_{ab} \right) \Phi^2,
\label{class s-t}
\end{equation}
where $G_{ab}$ is the Einstein tensor.
The next step is to define a stress-energy tensor operator
$\hat{T}_{ab}[g](x)$. In a naive way, one would replace
the classical field $\Phi$ in the functional
$T_{ab}[g,\Phi]$ by its corresponding quantum operator
$\hat{\Phi}[g]$. However, since the field
operator is well-defined only as a distribution on spacetime and this
procedure involves taking the product of two distributions at the same
spacetime point, the formal expression for
$\hat{T}_{ab}[g]$ is ill-defined and we need a regularization
procedure.
We may formally think of a regularized ``operator''
$\hat{T}_{ab}[g](x;\Omega)$,
depending on some regulator $\Omega$, defined by giving
a precise prescription for computing its
matrix elements for physically acceptable states of the field.
These states are assumed to be Hadamard states on the
Fock space of a Hadamard vacuum state \cite{wald94}.
The states may have to be regularized also in some way and the
procedure may involve some analytic continuation in the values of the
regulator.
Of course, if we remove the regularization in the results for these
matrix elements, we would obtain infinite quantities.
Once the regularization prescription has been introduced,
a renormalized and regularized stress-energy
``operator'' $\hat{T}^{R}_{ab}[g](x;\Omega)$ may be defined as
\begin{equation}
\hat{T}^{R}_{ab}[g](x;\Omega)=\hat{T}_{ab}[g](x;\Omega)+
F^{C}_{ab}[g](x;\Omega)\, \hat{I},
\label{renorm s-t}
\end{equation}
where $\hat{I}$ is the identity operator and $F^{C}_{ab}[g]$
are some symmetric tensor counterterms,
which can be written in terms of the regulator $\Omega$ and local
functionals of the metric $g_{cd}(x)$.\footnote{In the
point-splitting regularization method, for instance,
one introduces a point $y$ in a normal neighborhood of the point $x$,
so some non-local dependence on the metric is explicitly introduced in
the regularized stress-energy operator and then also in the
counterterms. Using this regularization technique, the
regulator can be taken as the vector $\sigma^a(x,y)$, which is the
tangent vector at the point $x$ to the geodesic joining $x$ and $y$
with length equal to the arc length along this geodesic. In this case,
the counterterms can be written in terms of the vector $\sigma^a(x,y)$
and tensors which are local
functionals of the metric $g_{ab}(x)$ \cite{fulling,christensen}.}
These counterterms can and must be chosen in such a way
that, for any pair of physically acceptable states
$|\psi\rangle$ and $|\varphi\rangle$, the matrix element
of the renormalized operator $\hat{T}^{R}_{ab}[g]$, defined by
\begin{equation}
\langle\psi|\hat{T}^{R}_{ab} |\varphi\rangle \equiv
\lim_{\Omega \rightarrow \Omega_p}
\langle\psi|\hat{T}^{R}_{ab} |\varphi\rangle (\Omega),
\label{renorm s-t 2}
\end{equation}
where $\Omega_p$ means the ``physical value'' of the regulator,
is finite (well defined as a distribution) and satisfies Wald's
axioms \cite{fulling,wald77}.
Using the point-splitting
or the dimensional regularization methods, these
counterterms can be extracted from the singular part of a
Schwinger-DeWitt series \cite{fulling,christensen,bunch}.
The choice
of these counterterms is not unique, each different choice is
called a ``renormalization scheme,'' and this leads to some ambiguity
in the definition of the renormalized stress-energy tensor operator.
But this ambiguity can be
absorbed into the renormalized coupling constants
appearing in the equations of motion for the gravitational field.
Thus, the ambiguity is only a mathematical artifact of the separation
of the action into a gravitational part and a matter part, but the
physically relevant equations are in fact unique
\cite{fulling,fulling74}.
The semiclassical Einstein equation for the metric $g_{ab}$ can then
be written as
\begin{equation}
{1\over 8 \pi G} \left( G_{ab}[g]+ \Lambda g_{ab} \right)-
2 \left( \alpha A_{ab}+\beta B_{ab} \right)\hspace{-0.3ex}[g]=
\langle \hat{T}^{R}_{ab} \rangle [g],
\label{semiclassical Einstein eq}
\end{equation}
where $\langle\hat{T}^{R}_{ab} \rangle [g]$ is the
expectation value of $\hat{T}^{R}_{ab}[g]$
in some physically acceptable state of the quantum field on the
spacetime $({\cal M},g_{ab})$. The notation
$\langle\hat{T}^{R}_{ab}\rangle [g]$ is used
to indicate that this expectation value is a functional of the metric
$g_{cd}$, not only because the stress-energy tensor operator depends
on the metric, but also because the state of the matter field
depends on the spacetime
(in general, such
state depends on the global structure of the spacetime manifold).
In Eq.~(\ref{semiclassical Einstein eq}),
$G$, $\Lambda$, $\alpha$ and
$\beta$ are renormalized coupling constants, respectively, the
Newtonian gravitational constant, the cosmological constant and two
dimensionless coupling constants. These constants may be seen
as the result of ``dressing'' the bare coupling constants in a
suitably regularized version of the gravitational part of the action,
\begin{equation}
S_{g}[g] \equiv
\int \! d^4 x \, \sqrt{- g} \left[{1\over 16 \pi G_{B}}
\left(R-2\Lambda_{B}\right)
+ \alpha_{B} C_{abcd}C^{abcd}+\beta_{B} R^2 \right],
\label{grav action}
\end{equation}
where $C_{abcd}$ is the Weyl tensor and the subindex
${\scriptstyle B}$ in the
coupling constants means ``bare.''
These renormalized coupling constants are supposed to be determined
experimentally (for the specific renormalization scheme that one has
chosen and for the characteristic scales of the physics under
consideration).
The tensors $A_{ab}$ and $B_{ab}$ in
Eq.~(\ref{semiclassical Einstein eq}) come from the functional
derivatives with
respect to the metric of the terms
quadratic in the curvature in $S_{g}[g]$, which are needed to ensure
the renormalizability of the theory. These tensors are explicitly
given by
\begin{eqnarray}
A^{ab} \equiv {1\over\sqrt{- g}} \frac{\delta}{\delta g_{ab}}
\int\! d^4 x \, \sqrt{- g}\, C_{cdef}C^{cdef}
&=& {1\over2}\,g^{ab}C_{cdef} C^{cdef}
-2R^{acde}{R^b}_{cde}
+4R^{ac}{R_c}^b \nonumber \\
&& -\,{2\over3}\,RR^{ab}
-2\hspace{0.2ex} \Box \hspace{-0.2ex} R^{ab}
+{2\over3} \bigtriangledown^{a}\!\bigtriangledown^{b} R
+{1\over3}\,g^{ab} \hspace{0.2ex} \Box \hspace{-0.2ex} R,
\label{A}
\end{eqnarray}
and
\begin{equation}
\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!
B^{ab} \equiv {1\over\sqrt{- g}} \frac{\delta}{\delta g_{ab}}
\int\! d^4 x\,\sqrt{- g}\, R^2
= {1\over2}\,g^{ab} R^2-2 R R^{ab}
+2 \bigtriangledown^{a}\!\bigtriangledown^{b} R
-2 g^{ab} \hspace{0.2ex} \Box \hspace{-0.2ex} R,
\label{B}
\end{equation}
where $R_{abcd}$ is the Riemann tensor and $R_{ab}$ is the Ricci tensor.
Note that each of the terms in Eq.~(\ref{semiclassical Einstein eq})
has vanishing divergence.
Notice also that we could add a classical stress-energy
tensor to the right hand side of
Eq.~(\ref{semiclassical Einstein eq}), if we had a classical matter
source, but, for simplicity, we shall ignore such a term.
As long as the gravitational field is assumed to be described by a
classical Lorentzian metric $g_{ab}$, the semiclassical
Einstein equation seems to be the only physically
plausible dynamical equation for this metric.
The reason is that, in classical general relativity, the metric
$g_{ab}$ couples to matter through the stress-energy
tensor. For a field quantized on the spacetime $({\cal M},g_{ab})$
and for a given state of this field, the expectation value of the
renormalized stress-energy tensor operator is the only physically
observable (up to the ambiguity mentioned above) c-number
stress-energy tensor that we can construct.
A solution of semiclassical gravity consists of a
spacetime $({\cal M},g_{ab})$, a quantum field operator
$\hat{\Phi}[g]$ satisfying Eq.~(\ref{Klein-Gordon}),
and a physically acceptable state $|\psi\rangle[g]$
for this field (which can also be a mixed state characterized by a
density operator),
such that Eq.~(\ref{semiclassical Einstein eq}) is
satisfied when the expectation value in the state $|\psi\rangle[g]$
of the renormalized operator $\hat{T}^{R}_{ab}[g]$
is put on the right hand side.
Let us now introduce stress-energy fluctuations.
Given a solution of semiclassical gravity, the stress-energy
tensor will in general have quantum fluctuations. To lowest order,
such fluctuations are described by the bi-tensor, which shall be
called noise kernel, defined by
\begin{equation}
8 N_{abcd}(x,y) \equiv \lim_{\Omega \rightarrow \Omega_p}
\bigl\langle \bigl\{
\hat{t}_{ab}(x) , \,
\hat{t}_{cd}(y)
\bigr\} \bigr\rangle [g](\Omega),
\label{noise}
\end{equation}
where $\{ \; , \: \}$ means the anticommutator and
$\hat{t}_{ab}(x;\Omega) \equiv \hat{T}_{ab}(x;\Omega)-
\langle \hat{T}_{ab}(x) \rangle (\Omega)$.
Note that we have
defined this noise kernel in terms of the unrenormalized
``operator'' $\hat{T}_{ab}[g](x;\Omega)$.
For a linear quantum field, this can be done
because the ultraviolet singular
behavior of $\langle\hat{T}_{ab}(x)
\hat{T}_{cd}(y)\rangle (\Omega)$ is the same as that of
$\langle\hat{T}_{ab}(x)\rangle (\Omega)
\langle\hat{T}_{cd}(y) \rangle (\Omega)$, so
$N_{abcd}(x,y)$ is free of ultraviolet divergencies.
One can trivially see from the substitution of (\ref{renorm s-t})
into (\ref{noise}) that
we can replace $\hat{T}_{ab}[g](x;\Omega)$ by
the renormalized operator $\hat{T}^{R}_{ab}[g](x)$,
and omit the limit
$\Omega \!\rightarrow \! \Omega_p$, in the last expression.
The result is obviously
independent of the renormalization scheme that one chooses
to define $\hat{T}^{R}_{ab}$.
As a perturbative correction to semiclassical gravity,
we want now to introduce an equation in which the
stress-energy fluctuations described by (\ref{noise})
are the source of classical gravitational fluctuations.
Thus, we assume that the gravitational field is
described by $g_{ab}+h_{ab}$,
where $h_{ab}$ is a linear perturbation
to the background metric $g_{ab}$, solution of
Eq.~(\ref{semiclassical Einstein eq}).
The renormalized stress-energy
operator and the state of the quantum field
may be denoted by $\hat{T}^{R}_{ab}[g+h]$ and
$|\psi\rangle[g+h]$, respectively, and
$\langle\hat{T}^{R}_{ab} \rangle [g+h]$
is the corresponding expectation value.
Let us introduce a Gaussian stochastic tensor field $\xi_{ab}$
defined by the following correlators:
\begin{equation}
\langle\xi_{ab}(x) \rangle_c = 0,
\hspace{6ex}
\langle\xi_{ab}(x)\xi_{cd}(y) \rangle_c = N_{abcd}(x,y),
\label{correlators}
\end{equation}
where
$\langle \hspace{1.5ex} \rangle_c$ means statistical
average.
In general, the two-point correlation function of a stochastic tensor
field $\xi_{ab}$ must be a symmetric,
in the sense that
$\langle\xi_{ab}(x)\xi_{cd}(y) \rangle_c =
\langle\xi_{cd}(y)\xi_{ab}(x) \rangle_c$, and
positive semi-definite real bi-tensor field.
Since the renormalized operator $\hat{T}^{R}_{ab}$ is
self-adjoint,
it is easy to see from the definition
(\ref{noise}) that $N_{abcd}(x,y)$ satisfies all these conditions.
Therefore, the relations (\ref{correlators}), with the cumulants of
higher order taken to be zero, do truly characterize a stochastic
tensor field $\xi_{ab}$.
The simplest equation which can incorporate in a consistent way the
stress-energy fluctuations described by $N_{abcd}(x,y)$
as the source of classical metric fluctuations is
\begin{equation}
{1\over 8 \pi G} \Bigl( G_{ab}[g+h]+
\Lambda\left(g_{ab}+h_{ab}\right) \Bigr)-
2 \left( \alpha A_{ab}+\beta B_{ab} \right)\hspace{-0.3ex}
[g+h]=\langle
\hat{T}^{R}_{ab}\rangle [g+h] +2 \xi_{ab} ,
\label{Einstein-Langevin eq}
\end{equation}
which must be understood as a dynamical equation for $h_{ab}$
to linear order.
Eq.~(\ref{Einstein-Langevin eq})
is the semiclassical Einstein-Langevin equation, which
gives a first order correction to semiclassical gravity.
One could also seek equations describing higher order corrections,
which would involve higher order stress-energy fluctuations, but,
for simplicity, we shall stick to the lowest order.
In order to check the consistency of
Eq.~(\ref{Einstein-Langevin eq}), note that
the term $\xi_{ab}$ does not depend on $h_{cd}$,
since it is completely determined from the solution of
semiclassical gravity by the correlators
(\ref{correlators}). Even so, this term must be considered as
of first order in perturbation theory around
semiclassical gravity. As shown in Ref.~\cite{mv98},
$\xi_{ab}$
is covariantly conserved up to first order in this perturbation
theory, in the sense that $\bigtriangledown^{a} \xi_{ab}$
behaves deterministically as the zero vector field
on ${\cal M}$
($\bigtriangledown^{a}$ is the covariant derivative
associated to the background metric $g_{ab}$).
It is thus consistent to include the
term $\xi_{ab}$ in the right hand side of
Eq.~(\ref{Einstein-Langevin eq}).
It was also shown in Ref.~\cite{mv98} that
for a conformal field, {\it i.e.}, a field whose classical action is
conformally invariant ({\it e.g.}, a massless conformally coupled
scalar field), $\xi_{ab}$
is ``traceless'' up to first order in perturbation theory,
since $g^{ab}\xi_{ab}$ behaves deterministically as a vanishing
scalar.
Hence, in the case of a conformal matter field,
the trace of
the right hand side of Eq.~(\ref{Einstein-Langevin eq})
comes only from the trace anomaly.
Since Eq.~(\ref{Einstein-Langevin eq}) is a linear stochastic
equation for $h_{ab}$ with an inhomogeneous term $\xi_{ab}$,
a solution can be formally written as a functional
$h_{ab}[\xi]$.
Such a solution can be characterized by
the whole family of its correlation functions.
From the average of
Eq.~(\ref{Einstein-Langevin eq}), the average of the
metric, $g_{ab}+\langle h_{ab} \rangle_c$,
must be a solution of the semiclassical Einstein equation linearized
around $g_{ab}$. The fluctuations of the metric around this average
can be described by the moments of order higher than one of the
stochastic field
$h_{ab}^{\rm f}[\xi] \equiv h_{ab}[\xi]
-\langle h_{ab} \rangle_c$.
Finally, for the solutions of Eq.~(\ref{Einstein-Langevin eq})
we have the gauge freedom
$h_{ab} \rightarrow
h'_{ab}\equiv h_{ab}+ \bigtriangledown_{\!a}
\zeta_{b}+\bigtriangledown_{\!b} \zeta_{a}$,
where $\zeta^{a}$ is any stochastic vector field on ${\cal M}$ which
is a functional of $\xi_{cd}$,
and $\zeta_{a} \equiv g_{ab}\zeta^{b}$.
Note that the tensors which appear in
Eq.~(\ref{Einstein-Langevin eq}) transform as
$R_{ab}[g\!+\!h']\!=\! R_{ab}[g\!+\!h]\!+\!
{\cal \pounds}_{\mbox{}_{\! \zeta}} R_{ab}[g]$
(to linear order in the perturbations),
where ${\cal \pounds}_{\mbox{}_{\! \zeta}}$ is the Lie
derivative with respect to $\zeta^a$.
If we substitute $h_{ab}$ by $h'_{ab}$ in
Eq.~(\ref{Einstein-Langevin eq}), we get
Eq.~(\ref{Einstein-Langevin eq}) plus the Lie derivative
of a combination of the tensors which appear in
Eq.~(\ref{semiclassical Einstein eq}). This last tensorial
combination vanishes when Eq.~(\ref{semiclassical Einstein eq}) is
satisfied. Thus, it is necessary that the
set $({\cal M},g_{ab},\hat{\Phi}[g],|\psi\rangle[g])$ be a solution of
semiclassical gravity to ensure that the Einstein-Langevin equation
(\ref{Einstein-Langevin eq}) is gauge invariant.
\section{Derivation from an influence action}
\label{sec:influence action}
The purpose of this section is to derive the semiclassical
Einstein-Langevin equation (\ref{Einstein-Langevin eq}) by
a method based on functional techniques.
The same method has been in fact used
in the literature to derive
Langevin-type equations in the context of semiclassical cosmology
\cite{calzettahu,humatacz,husinha,cv96,lomb-mazz,cv97,ccv97,campos-hu,%
campos-hu2,calver98}
and of analogous semiclassical regimes for systems of
quantum mechanics \cite{caldeira,hu-matacz94,hu-paz-zhang2}
and of quantum field theory
\cite{greiner,matacz,morikawa,shaisultanov,gleiser}.
Using these functional techniques, we also work out the
Einstein-Langevin equation
more explicitly, in a form more suitable for specific
calculations. Here, we consider again
the simplest case of a linear real scalar field $\Phi$.
These functional techniques are based on
the closed time path (CTP) functional formalism, due
to Schwinger and Keldysh \cite{schwinger61,schwinger62}.
This formalism is designed to obtain expectation
values of field operators in a direct way
and it is suited to derive dynamical equations for expectation values;
see Refs.~\cite{ctp,cv94,campos-hu} for detailed reviews.
In our case, this formalism will be useful to obtain an
expression for the expectation value
$\langle \hat{T}^{ab}\rangle [g\!+\!h]$ as an expansion in
the metric perturbation.
When the full quantum system consists of a distinguished subsystem
(the ``system'' of interest)
interacting with an environment (the remaining degrees of freedom),
the CTP functional formalism
turns out to be related
\cite{calzettahu,greiner,cv96,campos-hu,morikawa,shaisultanov,mv98_2}
to the
influence functional formalism of Feynman and Vernon
\cite{feynman-vernon}.
In this latter formalism, the integration of the environment
variables in a CTP path integral yields the influence functional, from
which one can define an effective action for the dynamics of the
system \cite{feynman-hibbs,calzettahu,humatacz,husinha,%
caldeira,hu-paz-zhang,hu-matacz94,greiner}.
Applying this influence functional formalism to our problem, the
semiclassical Einstein-Langevin equation will be formally derived in
subsection \ref{subsec:formal derivation}.
In our case, we consider
the metric field $g_{ab}(x)$ as the ``system'' degrees of freedom,
and the scalar field $\Phi(x)$
and also some ``high-momentum'' gravitational modes
\cite{whelan} as the ``environment'' variables.
Unfortunately, since the form of a complete
quantum theory of gravity interacting with matter is unknown,
we do not know what these ``high-momentum'' gravitational modes are.
Such a fundamental quantum theory might not even be a field theory,
in which case the metric and scalar fields would not be
fundamental objects \cite{hu99}.
Thus, in this case, we cannot attempt to evaluate
the influence action of Feynman and Vernon
starting from the fundamental quantum theory and
performing the path integrations in the environment variables.
Instead, we introduce the influence action for
an effective quantum field theory of gravity and matter
\cite{donoghue,humatacz},
in which such ``high-momentum'' gravitational
modes are assumed to have been already ``integrated out.''
Adopting the usual procedure of
effective field theories \cite{weinberg,donoghue},
one has to take the effective action for the metric and the scalar
field of
the most general local form compatible with general covariance:
$S[g,\Phi] \!\equiv \! S_g[g]+S_m[g,\Phi]+ \cdots$, where $S_g[g]$
is given by (\ref{grav action}),
\begin{equation}
S_m[g,\Phi] \equiv -{1\over2} \int\! d^4x \, \sqrt{- g}
\left[g^{ab}\partial_a \Phi \hspace{0.2ex} \partial_b \Phi
+\left(m^2+ \xi R \right)\Phi^2 \right],
\label{scalar field action}
\end{equation}
and the dots
stand for terms of order higher than two
in the curvature and in the number of
derivatives of the scalar field [because of the
Gauss-Bonnet theorem in four spacetime dimensions,
no further terms of second order in the curvature are needed in the
gravitational action (\ref{grav action})].
In this paper, we shall neglect the higher order terms as well as
self-interaction terms for the scalar field.
The second order terms are necessary to renormalize
one-loop ultraviolet divergencies of the scalar field
stress tensor.
Since ${\cal M}$ is a globally hyperbolic manifold,
we can foliate it by a family of $t\!=\! {\rm constant}$ Cauchy
hypersurfaces $\Sigma_{t}$. We denote by
${\bf x}$ the coordinates on each of these hypersurfaces, and by
$t_{i}$ and $t_{f}$ some initial and final times, respectively.
The integration domain for the action terms must be understood
as a compact region ${\cal U}$ of the manifold ${\cal M}$, bounded by
the hypersurfaces $\Sigma_{t_i}$ and $\Sigma_{t_f}$.
Assuming the form (\ref{scalar field action}) for
the effective action which couples the scalar and the metric fields,
we can now
introduce the corresponding influence functional.
This is a functional of two copies of the metric field that we denote
by $g_{ab}^+$ and $g_{ab}^-$.
Let us assume that, in the quantum effective theory,
the state of the full system
(the scalar and the metric fields) in the Schr\"{o}dinger picture
at the initial time $t\! =\! t_{i}$ can be described by
a factorizable
density operator, {\it i.e.}, a density operator which can be written
as the tensor product of two operators on the Hilbert spaces
of the metric and of the scalar field.
Let $\hat{\rho}^{\rm \scriptscriptstyle S}(t_{i})$ be the
density operator describing the initial state of the
scalar field.
If we consider the theory of a scalar field quantized in the
classical background spacetime $({\cal M},g_{ab})$ through the action
(\ref{scalar field action}), a
state in the Heisenberg picture described by a density operator
$\hat{\rho}[g]$ corresponds to this state.
Let $\left\{ |\varphi(\mbox{\bf x})\rangle^{\rm \scriptscriptstyle S}
\right\}$
be the basis of eigenstates of the scalar
field operator $\hat{\Phi}^{\rm \scriptscriptstyle S}({\bf x})$
in the Schr\"{o}dinger picture:
$\hat{\Phi}^{\rm \scriptscriptstyle S}({\bf x})
\, |\varphi\rangle ^{\rm \scriptscriptstyle S}=
\varphi(\mbox{\bf x})
\, |\varphi\rangle^{\rm \scriptscriptstyle S}$.
The matrix elements of
$\hat{\rho}^{\rm \scriptscriptstyle S}(t_{i})$ in this basis will be
written as
$\rho_{i} \!\left[\varphi,\tilde{\varphi}\right] \equiv
\mbox{}^{\rm \scriptscriptstyle S}
\langle \varphi|\,\hat{\rho}^{\rm \scriptscriptstyle S}(t_{i})
\, |\tilde{\varphi}\rangle^{\rm \scriptscriptstyle S}$.
We can now introduce the influence functional as the following
path integral over two copies of the scalar field:
\begin{equation}
{\cal F}_{\rm IF}[g^+,g^-] \equiv
\int\! {\cal D}[\Phi_+]\;
{\cal D}[\Phi_-] \;
\rho_i \!\left[\Phi_+(t_i),\Phi_-(t_i) \right] \,
\delta\!\left[\Phi_+(t_f)\!-\!\Phi_-(t_f) \right]\:
e^{i\left(S_m[g^+,\Phi_+]-S_m[g^-,\Phi_-]\right) }.
\label{path integral}
\end{equation}
The above double path integral can be rewritten as a closed time path
(CTP) integral, namely, as an integral over a single copy of field
paths with two different time branches, one going forward in time from
$t_i$ to $t_f$, and the other going backward in time from $t_f$
to $t_i$.
From this influence functional,
the influence action, $S_{\rm IF}[g^+,g^-]$, and the effective action of
Feynman and Vernon, $S_{\rm eff}[g^+,g^-]$, are defined by
${\cal F}_{\rm IF}[g^+,g^-] \equiv
e^{i S_{\rm IF}[g^+,g^-]}$ and
$S_{\rm eff}[g^+,g^-]\equiv S_{g}[g^+]-S_{g}[g^-]
+S_{\rm IF}[g^+,g^-]$.
Expression (\ref{path integral}) is ill-defined,
it must be regularized to get a meaningful
influence functional. We shall assume that we can use
dimensional regularization, that is, that we can give sense
to Eq.~(\ref{path integral}) by
dimensional continuation of all the quantities that appear in this
expression.
We should point out, nevertheless, that
for this regularization method to work one must be able to perform an
analytic continuation to Riemmanian signature
\cite{wald79}.
Thus, we
substitute the action $S_m$ in (\ref{path integral})
by some generalization to $n$ spacetime
dimensions, which may be chosen as
\begin{equation}
S_m[g,\Phi_{n}] = -{1\over2} \int\! d^n x \, \sqrt{- g}
\left[g^{ab} \partial_a \Phi_{n} \partial_b \Phi_{n}
+\left(m^2+ \xi R \right)\Phi_{n}^2 \right],
\label{scalar action}
\end{equation}
where we use a notation in which a subindex $n$ is attached to
these quantities that have different physical dimensions than the
corresponding physical quantities in
four dimensions. A quantity with the subindex $n$ can always
be associated to another without this subindex by means of a
mass scale $\mu$; thus, for the scalar field
$\Phi_{n}\!=\! \mu^{(n-4)/2} \,\Phi$.
We also need to substitute the action (\ref{grav action}) by
some suitable generalization to $n$ spacetime dimensions. We
take
\begin{equation}
S_g[g]=\mu^{n-4} \!\int \! d^n x \,\sqrt{- g}
\left[{1\over 16 \pi G_{B}}
\left(R-2\Lambda_{B}\right)+ {2\over 3}\,\alpha_{B}
\left(R_{abcd}R^{abcd}-
R_{ab}R^{ab} \right)+\beta_{B} R^2 \right].
\label{grav action in n}
\end{equation}
By the Gauss-Bonnet theorem, this action gives for $n \!=\! 4$
the same equations of motion as the action (\ref{grav action}).
The form of (\ref{grav action in n}) is suggested by the
Schwinger-DeWitt analysis of the ultraviolet divergencies in the
matter stress-energy tensor using dimensional regularization
\cite{bunch}.
Using (\ref{scalar action}) and (\ref{grav action in n}), one can
write the effective action of Feynman and Vernon,
$S_{\rm eff}[g^+,g^-]$, in dimensional regularization.
Since the action terms (\ref{scalar action}) and
(\ref{grav action in n})
contain second order derivatives of the metric, one should also add
some boundary terms \cite{wald84,humatacz}.
The effect of these
terms is to cancel out the boundary terms which appear
when taking variations of $S_{\rm eff}[g^+,g^-]$ keeping the value
of $g^+_{ab}$ and $g^-_{ab}$ fixed on the boundary of ${\cal U}$.
Alternatively, in order to obtain the equations of motion for
the metric in the semiclassical regime, we can work with the action terms
(\ref{scalar action}) and (\ref{grav action in n}) (without boundary
terms) and neglect all boundary terms when taking variations with
respect to $g^{\pm}_{ab}$. From now on, all the functional derivatives
with respect to the metric will be understood in this sense.
\subsection{The semiclassical Einstein
equation in dimensional regularization}
\label{subsec:semiclassical in n}
From the action
(\ref{scalar action}), we can define the stress-energy tensor
functional in the usual way
\begin{equation}
T^{ab}[g,\Phi_{n}](x) \equiv {2\over\sqrt{- g(x)}} \,
\frac{\delta S_m[g,\Phi_{n}]}{\delta g_{ab}(x)},
\label{s-t functional}
\end{equation}
which yields (\ref{class s-t}).
Working in the Heisenberg picture, we can now formally introduce the
regularized stress-energy tensor operator as
\begin{equation}
\hat{T}_{n}^{ab}[g] \equiv T^{ab}[ g,\hat{\Phi}_{n}[g]],
\hspace{5 ex}
\hat{T}^{ab}[g] \equiv \mu^{-(n-4)}\, \hat{T}_{n}^{ab}[g],
\label{regul s-t}
\end{equation}
where $\hat{\Phi}_{n}[g](x)$ is the field operator,
which satisfies the Klein-Gordon equation
(\ref{Klein-Gordon}) in $n$ spacetime dimensions,
and where we use a symmetrical ordering (Weyl ordering) prescription
for the operators. Using the Klein-Gordon equation,
the stress-energy operator can be written as
\begin{equation}
\hat{T}_{n}^{ab}[g] = {1\over 2} \left\{
\bigtriangledown^{a}\hat{\Phi}_{n}[g]\, , \,
\bigtriangledown^{b}\hat{\Phi}_{n}[g] \right\}
+ {\cal D}^{ab}[g]\, \hat{\Phi}_{n}^2[g],
\label{regul s-t 2}
\end{equation}
where ${\cal D}^{ab}[g]$ is the differential operator
\begin{equation}
{\cal D}^{ab}_{x} \equiv \left(\xi-{1\over 4}\right) g^{ab}(x)
\Box_{x}+ \xi
\left( R^{ab}(x)- \bigtriangledown^{a}_{x}
\bigtriangledown^{b}_{x} \right).
\label{diff operator}
\end{equation}
From the definitions (\ref{path integral}),
(\ref{s-t functional}) and (\ref{regul s-t}), one
can see that
\begin{equation}
\langle \hat{T}_n^{ab}(x) \rangle [g] =
\left. {2\over\sqrt{- g(x)}} \,
\frac{\delta S_{\rm IF}[g^+,g^-]}
{\delta g^+_{ab}(x)} \right|_{g^+=g^-=g},
\label{s-t expect value}
\end{equation}
where the expectation value is taken in the $n$-dimensional spacetime
generalization of the state described by
$\hat{\rho}[g]$.
Therefore, differentiating
$S_{\rm eff}[g^+,g^-]= S_{g}[g^+]-S_{g}[g^-]
+S_{\rm IF}[g^+,g^-]$
with respect to $g^+_{ab}$, and then setting
$g^+_{ab}=g^-_{ab}=g_{ab}$, we get
the semiclassical Einstein
equation in dimensional regularization:
\begin{equation}
{1\over 8 \pi G_{B}} \left( G^{ab}[g]+ \Lambda_{B} g^{ab} \right)-
\left({4\over 3}\, \alpha_{B} D^{ab}
+2 \beta_{B} B^{ab}\right)\! [g]
= \mu^{-(n-4)}
\langle \hat{T}_{n}^{ab}\rangle [g],
\label{semiclassical eq in n}
\end{equation}
where
\begin{eqnarray}
&&D^{ab} \equiv {1\over\sqrt{- g}} \frac{\delta}{\delta g_{ab}}
\int \! d^n x \,\sqrt{- g} \left(R_{cdef}R^{cdef}-
R_{cd}R^{cd} \right)
= {1\over2}\, g^{ab} \! \left( R_{cdef} R^{cdef}-
R_{cd}R^{cd}+\Box \hspace{-0.2ex} R \right)
-2R^{acde}{R^b}_{cde}
\nonumber \\
&& \hspace{56.5ex}
-\, 2 R^{acbd}R_{cd}
+4R^{ac}{R_c}^b
-3 \hspace{0.2ex}\Box \hspace{-0.2ex} R^{ab}
+\bigtriangledown^{a}\!\bigtriangledown^{b}\! \hspace{-0.2ex} R,
\label{D}
\end{eqnarray}
and $B^{ab}$ is defined as in (\ref{B}) but
for $n$ spacetime dimensions, although its explicit expression
in terms of the metric and curvature tensors is the
same.
When $n\!=\!4$, one has that $D^{ab}\!=\!(3/2) A^{ab}$,
where $A^{ab}$ is the tensor defined in (\ref{A}).
From equation
(\ref{semiclassical eq in n}), renormalizing the coupling
constants to eliminate the ``divergencies'' in
$\mu^{-(n-4)} \langle \hat{T}_{n}^{ab}\rangle [g]$,
and taking the limit $n\!\rightarrow \! 4$, we
get the physical semiclassical Einstein equation
(\ref{semiclassical Einstein eq}).
\subsection{A formal derivation of
the semiclassical Einstein-Langevin
equation}
\label{subsec:formal derivation}
In the spirit of the previous section, we now seek a dynamical
equation for a linear perturbation $h_{ab}$ to a semiclassical
metric $g_{ab}$, solution of
Eq.~(\ref{semiclassical eq in n}) in $n$ spacetime dimensions.
From the result of the previous subsection, if such equation were
simply a linearized semiclassical Einstein equation, it could be
obtained from an expansion of the effective action
$S_{\rm eff}[g+h^+,g+h^-]$. In particular, since, from
Eq.~(\ref{s-t expect value}), we have that
\begin{equation}
\langle \hat{T}_n^{ab}(x) \rangle [g+h]=
\left. {2\over\sqrt{-\det (g\!+\!h)(x)}} \,
\frac{\delta S_{\rm IF}
[g\!+\!h^+,g\!+\!h^-]}{\delta h^+_{ab}(x)}
\right|_{h^+=h^-=h},
\label{perturb s-t expect value}
\end{equation}
the expansion of $\langle \hat{T}_{n}^{ab}\rangle [g\!+\!h]$
to linear order in $h_{ab}$ can be obtained from an expansion of the
influence action $S_{\rm IF}[g+h^+,g+h^-]$ up to second order
in $h^{\pm}_{ab}$.
To perform the expansion of the influence action,
we have to compute the first and second order
functional derivatives of $S_{\rm IF}[g^+,g^-]$
and then set $g^+_{ab}\!=\!g^-_{ab}\!=\!g_{ab}$.
If we do so using the path integral representation
(\ref{path integral}), we can interpret these derivatives as
expectation values of operators.
The relevant second order derivatives are
\begin{eqnarray}
\left. {1\over\sqrt{- g(x)}\sqrt{- g(y)} } \,
\frac{\delta^2 S_{\rm IF}[g^+,g^-]}
{\delta g^+_{ab}(x)\delta g^+_{cd}(y)}
\right|_{g^+=g^-=g} \!\!
&=& -H_{\scriptscriptstyle \!{\rm S}_{\scriptstyle n}}^{abcd}[g](x,y)
-K_n^{abcd}[g](x,y)+
i N_n^{abcd}[g](x,y), \nonumber \\
\left. {1\over\sqrt{- g(x)}\sqrt{- g(y)} } \,
\frac{\delta^2 S_{\rm IF}[g^+,g^-]}
{\delta g^+_{ab}(x)\delta g^-_{cd}(y)}
\right|_{g^+=g^-=g} \!\!
&=& -H_{\scriptscriptstyle \!{\rm A}_{\scriptstyle n}}^{abcd}
[g](x,y)
-i N_n^{abcd}[g](x,y),
\label{derivatives}
\end{eqnarray}
where
\begin{eqnarray}
&&N_n^{abcd}[g](x,y) \equiv
{1\over 8} \left\langle \bigl\{
\hat{t}_n^{ab}(x) , \,
\hat{t}_n^{cd}(y)
\bigr\} \right\rangle [g],
\hspace{8ex}
H_{\scriptscriptstyle \!{\rm S}_{\scriptstyle n}}^{abcd}
[g](x,y) \equiv
{1\over 4}\:{\rm Im} \left\langle {\rm T}^{\displaystyle \ast}\!\!
\left( \hat{T}_n^{ab}(x) \hat{T}_n^{cd}(y)
\right) \right\rangle \![g],
\nonumber \\
&&H_{\scriptscriptstyle \!{\rm A}_{\scriptstyle n}}^{abcd}
[g](x,y) \equiv
-{i\over 8} \left\langle
\bigl[ \hat{T}_n^{ab}(x), \, \hat{T}_n^{cd}(y)
\bigr] \right\rangle \![g],
\hspace{5ex}
K_n^{abcd}[g](x,y) \equiv
\left. {-1\over\sqrt{- g(x)}\sqrt{- g(y)} } \, \left\langle
\frac{\delta^2 S_m[g,\Phi_{n}]}{\delta g_{ab}(x)\delta g_{cd}(y)}
\right|_{\Phi_{n}=\hat{\Phi}_{n}}\right\rangle \![g],
\nonumber \\
\mbox{}
\label{kernels}
\end{eqnarray}
with $\hat{t}_n^{ab} \equiv \hat{T}_{n}^{ab}-
\langle \hat{T}_{n}^{ab} \rangle$, and
using again a Weyl ordering prescription
for the operators in the last of these expressions.
Here, $[ \; , \: ]$ means the commutator, and we
use the symbol ${\rm T}^{\displaystyle \ast}$
to denote that, first,
we have to time order the field operators $\hat{\Phi}_{n}$ and then
apply the derivative operators which appear in each term
of the product $T^{ab}(x) T^{cd}(y)$, where $T^{ab}$ is
the functional (\ref{class s-t}). For instance,
\begin{equation}
{\rm T}^{\displaystyle \ast}\!\! \left(\hspace{-0.07ex}
\bigtriangledown^{a}_{\!\!\! \mbox{}_{x}}
\hspace{-0.1ex}\hat{\Phi}_{n}(x)\!
\bigtriangledown^{b}_{\!\!\! \mbox{}_{x}}\!\hat{\Phi}_{n}(x)\!
\bigtriangledown^{c}_{\!\!\! \mbox{}_{y}}\!\hat{\Phi}_{n}(y)\!
\bigtriangledown^{d}_{\!\!\! \mbox{}_{y}}\!\hat{\Phi}_{n}(y)\!
\right)\! =\!\!\!\!\lim_{
x_1,x_2 \rightarrow x_{\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!
\!\!\!\!\!
\mbox{}_{\mbox{}_{\mbox{}_{\mbox{}_
{\mbox{}_{\scriptstyle x_3,x_4 \rightarrow y}}}}}} }\!\!\!
\bigtriangledown^{a}_{\!\!\! \mbox{}_{x_1}}\!\!\hspace{0.02ex}
\bigtriangledown^{b}_{\!\!\! \mbox{}_{x_2}}\!\!
\bigtriangledown^{c}_{\!\!\! \mbox{}_{x_3}}\!\!
\bigtriangledown^{d}_{\!\!\! \mbox{}_{x_4}}\!
{\rm T}\! \left(\hat{\Phi}_{n}(x_1)\hat{\Phi}_{n}(x_2)
\hat{\Phi}_{n}(x_3)\hat{\Phi}_{n}(x_4) \right)\!,
\label{T star}
\end{equation}
where ${\rm T}$ is the usual time ordering.
This ${\rm T}^{\displaystyle \ast}$ ``time ordering'' arises because
we have path integrals containing products of derivatives of the
field, which can be expressed as derivatives of the path
integrals which do not contain such derivatives.
Notice, from the definitions
(\ref{kernels}), that all the kernels which appear
in expressions (\ref{derivatives}) are real and that
$H_{\scriptscriptstyle \!{\rm A}_{\scriptstyle n}}^{abcd}$ is also
free of ultraviolet divergencies in the limit
$n \!\rightarrow \!4$.
From (\ref{derivatives}) and (\ref{kernels}),
it is clear that the imaginary part of the
influence action, which does not contribute to the semiclassical
Einstein equation (\ref{semiclassical eq in n})
because the expectation value of
$\hat{T}_{n}^{ab}[g]$ is real, contains information on the
fluctuations of this operator. From (\ref{s-t expect value}) and
(\ref{derivatives}), taking into account that
$S_{\rm IF}[g,g]=0$ and that
$S_{\rm IF}[g^-,g^+]=
-S^{ {\displaystyle \ast}}_{\rm IF}[g^+,g^-]$, we can write the
expansion for the influence action
$S_{\rm IF}[g\!+\!h^+,g\!+\!h^-]$ around a background
metric $g_{ab}$ in terms of the kernels (\ref{kernels}).
Taking into account that
these kernels satisfy the symmetry relations
\begin{equation}
H_{\scriptscriptstyle \!{\rm S}_{\scriptstyle n}}^{abcd}(x,y)=
H_{\scriptscriptstyle \!{\rm S}_{\scriptstyle n}}^{cdab}(y,x),
\hspace{3 ex}
H_{\scriptscriptstyle \!{\rm A}_{\scriptstyle n}}^{abcd}(x,y)=
-H_{\scriptscriptstyle \!{\rm A}_{\scriptstyle n}}^{cdab}(y,x),
\hspace{3 ex}
K_n^{abcd}(x,y) = K_n^{cdab}(y,x),
\label{symmetries}
\end{equation}
and introducing a new kernel
\begin{equation}
H_n^{abcd}(x,y)\equiv
H_{\scriptscriptstyle \!{\rm S}_{\scriptstyle n}}^{abcd}(x,y)
+H_{\scriptscriptstyle \!{\rm A}_{\scriptstyle n}}^{abcd}(x,y),
\label{H}
\end{equation}
this expansion can be finally written as
\begin{eqnarray}
S_{\rm IF}[g\!+\!h^+,g+h^-]
&=& {1\over 2} \int\! d^nx\, \sqrt{- g(x)}\:
\langle \hat{T}_{n}^{ab}(x) \rangle [g] \,
\left[h_{ab}(x) \right] \nonumber \\
&&
-\,{1\over 2} \int\! d^nx\, d^ny\, \sqrt{- g(x)}\sqrt{- g(y)}\,
\left[h_{ab}(x)\right]
\left(H_n^{abcd}[g](x,y)\!
+\!K_n^{abcd}[g](x,y) \right)
\left\{ h_{cd}(y) \right\} \nonumber \\
&&
+\,{i\over 2} \int\! d^nx\, d^ny\, \sqrt{- g(x)}\sqrt{- g(y)}\,
\left[h_{ab}(x) \right]
N_n^{abcd}[g](x,y)
\left[h_{cd}(y) \right]+0(h^3),
\label{expansion 2}
\end{eqnarray}
where we have used the notation
\begin{equation}
\left[h_{ab}\right] \equiv h^+_{ab}\!-\!h^-_{ab},
\hspace{5 ex}
\left\{ h_{ab}\right\} \equiv h^+_{ab}\!+\!h^-_{ab}.
\label{notation}
\end{equation}
We are now in the position to carry out the formal derivation of the
semiclassical Einstein-Langevin equation.
The procedure is well known
\cite{calzettahu,humatacz,husinha,cv96,lomb-mazz,cv97,ccv97,%
campos-hu,campos-hu2,calver98,caldeira,hu-matacz94,hu-paz-zhang2,%
greiner,matacz,morikawa,shaisultanov,gleiser},
it consists of deriving a new
``improved'' effective action using the
the following identity:
\begin{equation}
e^{-{1\over 2} \!\int\! d^nx\, d^ny \, \sqrt{- g(x)}\sqrt{- g(y)}\,
\left[h_{ab}(x) \right]\, N_n^{abcd}(x,y)\, \left[h_{cd}(y)\right] }=
\int\! {\cal D}[\xi_n]\: {\cal P}[\xi_n]\, e^{i \!\int\! d^nx \,
\sqrt{- g(x)}\,\xi_n^{ab}(x)\,\left[h_{ab}(x) \right] },
\label{Gaussian path integral}
\end{equation}
where ${\cal P}[\xi_n]$ is the probability distribution
functional of a Gaussian stochastic tensor $\xi_n^{ab}$
characterized by the correlators
\begin{equation}
\langle\xi_n^{ab}(x) \rangle_{c}\!= 0,
\hspace{10ex}
\langle\xi_n^{ab}(x)\xi_n^{cd}(y) \rangle_{c}\!=
N_n^{abcd}[g](x,y),
\label{correlators in n}
\end{equation}
with $N_n^{abcd}$ given in (\ref{kernels}),
and where
the path integration measure is assumed to be a scalar under
diffeomorphisms of $({\cal M},g_{ab})$. The above identity follows
from the identification of the right hand side of
(\ref{Gaussian path integral}) with the characteristic functional for
the stochastic field $\xi_n^{ab}$.
In fact, by differentiation of this expression
with respect to $\left[h_{ab}\right]$, it can be checked that this is
the characteristic functional of a stochastic field characterized by the
correlators (\ref{correlators in n}).
When $N_n^{abcd}(x,y)$ is strictly positive definite, the
probability distribution functional for $\xi_n^{ab}$ is explicitly
given by
\begin{equation}
{\cal P}[\xi_n]=
\frac{e^{-{1\over2}\!\int\! d^nx\, d^ny \, \sqrt{-g(x)}\sqrt{-g(y)}\,
\xi_n^{ab}(x) \, N^{-1}_{n\,abcd}(x,y)\, \xi_n^{cd}(y)}}
{\int\! {\cal D}\bigl[\bar{\xi}_n\bigr]\:
e^{-{1\over2}\!\int\! d^nz\, d^nw \, \sqrt{-g(z)}\sqrt{-g(w)}\,
\bar{\xi}_n^{ef}(z) \, N^{-1}_{n\,efgh}(z,w)\, \bar{\xi}_n^{gh}(w)}},
\label{gaussian probability}
\end{equation}
where $N^{-1}_{n\,abcd}[g](x,y)$ is the inverse of
$N_n^{abcd}[g](x,y)$ defined by
\begin{equation}
\int\! d^nz \, \sqrt{- g(z)}\, N_n^{abef}(x,z) N^{-1}_{n\,efcd}(z,y)=
{1\over2} \left(\delta^a_c \delta^b_d+\delta^a_d \delta^b_c \right)
{\delta^n(x\!-\!y) \over \sqrt{- g(x)}}.
\label{inverse}
\end{equation}
Using the identity
(\ref{Gaussian path integral}), we can write the modulus of the
influence functional in the approximation (\ref{expansion 2}) as
\begin{equation}
\bigl|\hspace{0.2ex}{\cal F}_{\rm IF}[g+h^+,g+h^-]
\hspace{0.2ex}\bigr|=
e^{-{\rm Im}\, S_{\rm IF}[g+h^+,g+h^-]}=
\left\langle e^{i \int\! d^nx \,
\sqrt{- g(x)}\,\xi_n^{ab}(x)\,\left[h_{ab}(x) \right] }
\right\rangle_{\! c}
\label{infl funct modulus}
\end{equation}
where $\langle \hspace{1.5ex} \rangle_c$ means statistical average
over the stochastic tensor $\xi_n^{ab}$.
Thus, the effect of the imaginary part of the influence action
(\ref{expansion 2}) on the corresponding influence
functional is equivalent to the averaged effect of the stochastic
source $\xi_n^{ab}$ coupled linearly
to the perturbations $h_{ab}^{\pm}$.
The influence functional, in the approximation
(\ref{expansion 2}), can be written as a statistical average over
$\xi_n^{ab}$:
\begin{equation}
{\cal F}_{\rm IF}[g+h^+,g+h^-]= \left\langle
e^{i {\cal A}^{\rm eff}_{\rm IF}[h^+,h^-;g;\xi_n]}
\right\rangle_{\! c},
\label{infl funt as average}
\end{equation}
with
\begin{equation}
{\cal A}^{\rm eff}_{\rm IF}[h^+,h^-;g;\xi_n] \equiv
{\rm Re}\, S_{\rm IF}[g\!+\!h^+,g\!+\!h^-]+\!
\int\! d^nx \,
\sqrt{- g(x)}\,\xi_n^{ab}(x)\left[h_{ab}(x) \right]+0(h^3),
\label{eff influence action}
\end{equation}
where ${\rm Re}\, S_{\rm IF}$ can be read from the expansion
(\ref{expansion 2}).
Note that the stochastic term in this action contains the information
of the imaginary part of $S_{\rm IF}$.
Introducing a new ``improved'' effective action
\begin{equation}
{\cal A}_{\rm eff}[h^+,h^-;g;\xi_n] \equiv S_{g}[g+h^+]-S_{g}[g+h^-]+
{\cal A}^{\rm eff}_{\rm IF}[h^+,h^-;g;\xi_n],
\label{stochastic eff action}
\end{equation}
where $S_{g}[g+h^{\pm}]$ has to be expanded up to second order
in the perturbations $h_{ab}^{\pm}$,
the equation of motion for
$h_{ab}$ can be derived as
\begin{equation}
\left. {1\over\sqrt{-\det (g\!+\!h)(x)}} \,
\frac{\delta {\cal A}_{\rm eff}[h^+,h^-;g;\xi_n]}{\delta h^+_{ab}(x)}
\right|_{h^+=h^-=h}=0.
\label{eq of motion}
\end{equation}
From (\ref{perturb s-t expect value}), taking into account that
only the real part of the influence action contributes to the
expectation value of the stress-energy tensor, we get,
to linear order in $h_{ab}$,
\begin{equation}
{1\over 8 \pi G_{B}}\biggl( G^{ab}[g\!+\!h]+
\Lambda_{B} \left(g^{ab}\!-\!h^{ab}\right) \biggr)-
\left({4\over 3}\, \alpha_{B} D^{ab}
+ 2 \beta_{B} B^{ab} \right)\![g \!+\! h] = \mu^{-(n-4)}
\langle \hat{T}_{n}^{ab} \rangle [g\!+\!h]
+2 \mu^{-(n-4)} \xi_n^{ab},
\label{Einstein-Langevin eq in n}
\end{equation}
where
$h^{ab}\!\equiv\! g^{ac}g^{bd}h_{cd}$, that is,
$g^{ab}\!- h^{ab}\!+ 0(h^2)$ is
the inverse of the metric $g_{ab}\!+\!h_{ab}$.
This last equation is the semiclassical Einstein-Langevin
equation in dimensional regularization.
As we have pointed out in section \ref{sec:E-L},
the two-point correlation function of the stochastic source
in this equation [see Eq.~(\ref{correlators in n})],
given by the
noise kernel defined in (\ref{kernels}), is
free of ultraviolet divergencies in the limit $n \!\rightarrow \!4$.
Therefore, in the Einstein-Langevin equation
(\ref{Einstein-Langevin eq in n}),
one can perform exactly the same renormalization procedure
as for the semiclassical Einstein equation
(\ref{semiclassical eq in n}).
After this,
Eq.~(\ref{Einstein-Langevin eq in n}) will yield the physical
semiclassical Einstein-Langevin equation
(\ref{Einstein-Langevin eq}).
The derivation presented in this paper clarifies the
physical meaning of the stochastic source formally introduced in the
effective action (\ref{eff influence action}) by the identification
(\ref{infl funct modulus}), since it links its two-point correlation
function to the stress-energy fluctuations
by Eqs.~(\ref{correlators in n}) and (\ref{kernels}).
There is also a connection
between the equations obtained by this formal functional
method and the equations derived from the (in general, also formal)
assumption that
decoherence and classicalization of suitably coarse-grained
system variables is achieved
through the mechanism proposed by Gell-Mann
and Hartle \cite{gell-mann-hartle} in the consistent histories
formulation of a quantum theory. This last approach allows to evaluate
the probability distribution associated to such decoherent variables,
given by the diagonal elements of a decoherence functional, and, under
some approximations, to derive effective quasiclassical equations of
motion for them. These effective equations of motion
can be shown to
coincide \cite{mv98_2} with the semiclassical equations for the
background and the
Langevin-type equations for perturbations obtained from the above
functional method.
Taking this connection into account,
we can also conclude that, if one formally assumes that the
Gell-Mann and Hartle mechanism works for the metric field, one is
lead to the semiclassical Einstein equation and the semiclassical
Einstein-Langevin equation for the background metric and for the
metric perturbations, respectively \cite{mv98_2}.
We end this subsection with some comments on the relation between the
semiclassical Einstein-Langevin equation (\ref{Einstein-Langevin eq})
and the Langevin-type equations for stochastic metric perturbations
recently derived in the literature
\cite{calzettahu,humatacz,husinha,cv96,lomb-mazz,cv97,ccv97,%
campos-hu,campos-hu2,calver98}.
In these previous derivations, one starts with the influence functional
(\ref{path integral}), with the state of the scalar field assumed to
be an ``in'' vacuum or
an ``in'' thermal state, and computes explicitly the expansion for the
corresponding influence action around a specific metric background.
One then applies the above formal method to
derive a Langevin equation for the perturbations to this background.
However, most of these derivations start with a
``mini-superspace'' model and, thus, the metric perturbations are
assumed from the beginning to have a restrictive form. In those cases,
the derived Langevin equations do not correspond exactly to
our equation,
Eq.~(\ref{Einstein-Langevin eq}), but to a ``reduced''
version of this equation, in which only some components of the noise
kernel in Eq.~(\ref{correlators}) (or some particular combinations of
them) influence the dynamics of the metric perturbations.
Only those
equations which have been derived starting from a completely general
form for the metric perturbations
\cite{cv96,lomb-mazz,campos-hu,campos-hu2}
are actually particular cases of the
semiclassical Einstein-Langevin equation
(\ref{Einstein-Langevin eq}).
Note, however, that the stochastic equation derived in
Refs.~\cite{campos-hu,campos-hu2} do not correspond exactly to
Eq.~(\ref{Einstein-Langevin eq}), since the background
(Minkowski spacetime and a scalar field in a thermal state)
is not a solution of semiclassical gravity.
In this case, for the reasons explained in Sec.~\ref{sec:E-L}, the
equation for the metric perturbations is not gauge invariant.
\subsection{Explicit linear form
of the Einstein-Langevin equation}
\label{subsec:explicit}
We can write Eq.~(\ref{Einstein-Langevin eq in n}) in a more explicit
form by working out the expansion of
$\langle \hat{T}_{n}^{ab}\rangle [g\!+\!h]$
up to linear order in the
perturbation $h_{ab}$.
From Eq.~(\ref{perturb s-t expect value}),
we see that this expansion can
be easily obtained from (\ref{expansion 2}).
Noting, from (\ref{kernels}), that
\begin{equation}
K_n^{abcd}[g](x,y)= -{1\over 4} \,
\langle \hat{T}_{n}^{ab}(x) \rangle [g] \,
{g^{cd}(x)\over\sqrt{- g(y)}}\,
\delta^n(x\!-\!y)-{1\over 2}\,{1\over\sqrt{- g(y)}}
\left\langle \left.
\frac{\delta T^{ab}[g,\Phi_{n}](x)}{\delta g_{cd}(y)}
\right|_{\Phi_{n}=\hat{\Phi}_{n}}\right\rangle \![g],
\label{K}
\end{equation}
we get
\begin{equation}
\langle \hat{T}_n^{ab}(x) \hspace{-0.1ex}\rangle
[g\!+\!h] =
\langle \hat{T}_n^{ab}(x) \hspace{-0.1ex}\rangle
[g] +
\langle
\hat{T}_n^{{\scriptscriptstyle (1)}\hspace{0.1ex} ab}
[g;h](x) \hspace{-0.1ex} \rangle [g] -
2 \!\int\! \hspace{-0.2ex} d^ny \,
\sqrt{- g(y)} \hspace{0.2ex} H_n^{abcd}[g](x,y) \hspace{0.2ex}
h_{cd}(y) + 0(h^2),
\label{s-t expect value expansion}
\end{equation}
where the operator
$\hat{T}_n^{{\scriptscriptstyle (1)}\hspace{0.1ex} ab}$ is defined
from the term of first order in the expansion of
$T^{ab}[g+h,\Phi_{n}]$ as
\begin{equation}
T^{ab}[g\!+\!h,\Phi_{n}]=T^{ab}[g,\Phi_{n}]+
T^{{\scriptscriptstyle (1)}\hspace{0.1ex} ab}[g,\Phi_{n};h]
+0(h^2), \hspace{3.5 ex}
\hat{T}_n^{{\scriptscriptstyle (1)}\hspace{0.1ex} ab}
[g;h]\equiv
T^{{\scriptscriptstyle (1)}\hspace{0.1ex} ab}[g,\hat{\Phi}_{n}[g];h],
\label{T(1)}
\end{equation}
using, as always, a Weyl ordering prescription for the operators in the
last definition. Note that the third term on the right hand side of
Eq.~(\ref{s-t expect value expansion}) is a consequence of
the dependence
on $h_{cd}$ of the field operator $\hat{\Phi}_{n}[g+h]$ and of the
density operator $\hat{\rho}[g+h]$.
Substituting (\ref{s-t expect value expansion}) into
(\ref{Einstein-Langevin eq in n}), and taking into account that
$g_{ab}$ satisfies the semiclassical Einstein equation
(\ref{semiclassical eq in n}), we can write the Einstein-Langevin
equation (\ref{Einstein-Langevin eq in n}) as
\begin{eqnarray}
&&{1\over 8 \pi G_{B}}\left(
G^{{\scriptscriptstyle (1)}\hspace{0.1ex} ab}
[g;h](x)-
\Lambda_{B}\, h^{ab}(x) \right) -
{4\over 3}\, \alpha_{B} D^{{\scriptscriptstyle (1)}\hspace{0.1ex} ab}
[g;h](x)
-2\beta_{B} B^{{\scriptscriptstyle (1)}\hspace{0.1ex} ab}
[g;h](x) \nonumber \\
&&- \,\mu^{-(n-4)} \langle
\hat{T}_n^{{\scriptscriptstyle (1)}\hspace{0.1ex} ab}
[g;h](x) \rangle [g]+
2 \!\int\! d^ny\, \sqrt{- g(y)}\,\mu^{-(n-4)}
H_n^{abcd}[g](x,y)\, h_{cd}(y)
\hspace{-0.2ex}=
2 \mu^{-(n-4)} \xi_n^{ab}(x).
\label{Einstein-Langevin eq 2}
\end{eqnarray}
In the last
equation we have used the superindex ${\scriptstyle (1)}$ to denote
the terms of first order in the
expansion in $h_{ab}$ of the tensors $G^{ab}[g+h]$,
$D^{ab}[g+h]$ and $B^{ab}[g+h]$. Thus,
for instance, $G^{ab}[g+h]\!=\!G^{ab}[g]+
G^{{\scriptscriptstyle (1)}\hspace{0.1ex} ab}[g;h]+ 0(h^2)$.
The explicit expressions for the tensors
$G^{{\scriptscriptstyle (1)}\hspace{0.1ex} ab}[g;h]$,
$D^{{\scriptscriptstyle (1)}\hspace{0.1ex} ab}[g;h]$ and
$B^{{\scriptscriptstyle (1)}\hspace{0.1ex} ab}[g;h]$ can be
found in the Appendix of Ref.~\cite{mv98_2},
and
$T^{{\scriptscriptstyle (1)}\hspace{0.1ex} ab}[g,\Phi_{n};h]$
is given in
Appendix \ref{sec:expansions of tensors}.
From
$T^{{\scriptscriptstyle (1)}\hspace{0.1ex} ab}[g,\Phi_{n};h]$, we can
write an explicit expression for the operator
$\hat{T}_n^{{\scriptscriptstyle (1)}\hspace{0.1ex} ab}$.
In fact,
using the Klein-Gordon equation, and
expressions (\ref{regul s-t 2}) and (\ref{diff operator}) for the
stress-energy operator, we have
\begin{equation}
\hat{T}_n^{{\scriptscriptstyle (1)}\hspace{0.1ex} ab}
[g;h]=\left({1\over 2}\, g^{ab}h_{cd}-\delta^a_c h^b_d-
\delta^b_c h^a_d \right) \hat{T}_{n}^{cd}[g]
+{\cal F}^{ab}[g;h]\, \hat{\Phi}_{n}^2[g],
\label{T(1) operator}
\end{equation}
where ${\cal F}^{ab}[g;h]$ is the differential operator
\begin{eqnarray}
{\cal F}^{ab} &\equiv& \left(\xi\!-\!{1\over 4}\right)\!\!
\left(h^{ab}\!-\!{1\over 2}\, g^{ab} h^c_c \right)\! \Box+
{\xi \over 2} \left[
\bigtriangledown^{c}\! \bigtriangledown^{a}\! h^b_c+
\bigtriangledown^{c}\! \bigtriangledown^{b}\! h^a_c-
\Box h^{ab}-
\bigtriangledown^{a}\! \bigtriangledown^{b}\! h^c_c-
g^{ab}\! \bigtriangledown^{c}\! \bigtriangledown^{d} h_{cd}
\right. \nonumber \\
&&+\left. g^{ab} \Box h^c_c
+\left( \bigtriangledown^{a} h^b_c+
\bigtriangledown^{b} h^a_c-\bigtriangledown_{\! c}
\hspace{0.2ex} h^{ab}-
2 g^{ab}\! \bigtriangledown^{d}\! h_{cd} +
g^{ab}\! \bigtriangledown_{\! c} \! h^d_d
\right)\! \bigtriangledown^{c}
-g^{ab} h_{cd} \bigtriangledown^{c}\! \bigtriangledown^{d}
\right].
\label{diff operator F}
\end{eqnarray}
It is understood that indices are raised
with the background inverse metric $g^{ab}$ and that all the
covariant derivatives are associated to the metric $g_{ab}$.
Substituting expression (\ref{T(1) operator}) into
Eq.~(\ref{Einstein-Langevin eq 2}),
and using the semiclassical equation
(\ref{semiclassical eq in n}) to get an expression for
$\mu^{-(n-4)}
\langle \hat{T}_{n}^{ab}\rangle [g]$, we can
finally write the semiclassical Einstein-Langevin equation in
dimensional regularization as
\begin{eqnarray}
&&{1\over 8 \pi G_{B}}\Biggl[
G^{{\scriptscriptstyle (1)}\hspace{0.1ex} ab}\!-\!
{1\over 2}\, g^{ab} G^{cd} h_{cd}+ G^{ac} h^b_c+G^{bc} h^a_c+
\Lambda_{B} \left( h^{ab}\!-\!{1\over 2}\, g^{ab} h^c_c \right)
\Biggr](x)
\nonumber \\
&&
- \,
{4\over 3}\, \alpha_{B} \left( D^{{\scriptscriptstyle
(1)}\hspace{0.1ex} ab}
-{1\over 2}\, g^{ab} D^{cd} h_{cd}+ D^{ac} h^b_c+D^{bc} h^a_c
\right)\! (x)
-2\beta_{B}\left( B^{{\scriptscriptstyle (1)}\hspace{0.1ex} ab}\!-\!
{1\over 2}\, g^{ab} B^{cd} h_{cd}+ B^{ac} h^b_c+B^{bc} h^a_c
\right)\! (x) \nonumber \\
&&- \, \mu^{-(n-4)}\, {\cal F}^{ab}_x
\langle \hat{\Phi}_{n}^2(x) \rangle [g]
+2 \!\int\! d^ny \, \sqrt{- g(y)}\, \mu^{-(n-4)}
H_n^{abcd}[g](x,y)\, h_{cd}(y)
=2 \mu^{-(n-4)} \xi^{ab}_n(x),
\label{Einstein-Langevin eq 3}
\end{eqnarray}
where the tensors $G^{ab}$, $D^{ab}$ and
$B^{ab}$ are computed from the semiclassical metric $g_{ab}$,
and where we have omitted the functional dependence on $g_{ab}$ and
$h_{ab}$ in $G^{{\scriptscriptstyle (1)}\hspace{0.1ex} ab}$,
$D^{{\scriptscriptstyle (1)}\hspace{0.1ex} ab}$,
$B^{{\scriptscriptstyle (1)}\hspace{0.1ex} ab}$ and
${\cal F}^{ab}$ to simplify the notation.
Notice that, in Eq.~(\ref{Einstein-Langevin eq 3}),
all the ultraviolet divergencies in
the limit $n \!\rightarrow \!4$, which must be removed by
renormalization of the coupling constants, are in
$\langle \hat{\Phi}_{n}^2(x) \rangle$ and the
symmetric part
$H_{\scriptscriptstyle \!{\rm S}_{\scriptstyle n}}^{abcd}(x,y)$ of the
kernel $H_n^{abcd}(x,y)$, whereas the
kernels $N_n^{abcd}(x,y)$ and
$H_{\scriptscriptstyle \!{\rm A}_{\scriptstyle n}}^{abcd}(x,y)$ are
free of ultraviolet divergencies.
These two last kernels can be
written in terms of
$F_{n}^{abcd}[g](x,y) \equiv
\left\langle \hat{t}_n^{ab}(x)\,\hat{t}_n^{cd}(y)
\right\rangle \![g]$
as
\begin{equation}
N_n^{abcd}[g](x,y))=
{1\over 4}\,{\rm Re} \, F_{n}^{abcd}[g](x,y),
\hspace{7ex}
H_{\scriptscriptstyle \!{\rm A}_{\scriptstyle n}}^{abcd}[g](x,y)=
{1\over 4}\,{\rm Im} \, F_{n}^{abcd}[g](x,y),
\label{finite kernels}
\end{equation}
where we have used that
$2 \left\langle \hat{t}_n^{ab}(x)\, \hat{t}_n^{cd}(y) \right\rangle=
\left\langle \left\{ \hat{t}_n^{ab}(x), \, \hat{t}_n^{cd}(y)
\right\}\right\rangle +
\left\langle \left[ \hat{t}_n^{ab}(x),
\, \hat{t}_n^{cd}(y)\right]\right\rangle$,
and the fact that the first term on the right hand side of this
identity is real, whereas the second one is pure imaginary.
Once we perform the renormalization procedure
in Eq.~(\ref{Einstein-Langevin eq 3}), setting
$n \!= \!4$ will yield the physical semiclassical
Einstein-Langevin equation.
Note that, due to the presence of the kernel $H_n^{abcd}(x,y)$,
this equation will be usually non-local in the metric perturbation.
\subsection{The kernels for a vacuum state}
\label{subsec:vacuum}
We conclude this section by considering the case in which the
expectation values that appear in the Einstein-Langevin equation
(\ref{Einstein-Langevin eq 3}) [see Eqs.~(\ref{kernels})]
are taken in a vacuum state $|0 \rangle$
(for a field quantized on $({\cal M},g_{ab})$ in the Heisenberg
picture), such as, for instance, an ``in'' vacuum.
In this case we can go further and write these
expectation values in terms of the Wightman and Feynman
functions, defined as
\begin{equation}
G_n^+(x,y) \equiv \langle 0| \,
\hat{\Phi}_{n}(x) \hat{\Phi}_{n}(y) \,
|0 \rangle [g],
\hspace{5 ex}
i G\!_{\scriptscriptstyle F_{\scriptstyle \hspace{0.1ex} n}}
\hspace{-0.2ex}(x,y)
\equiv \langle 0| \,
{\rm T}\! \left( \hat{\Phi}_{n}(x) \hat{\Phi}_{n}(y) \right)
\hspace{-0.2ex}
|0 \rangle [g].
\label{Wightman and Feynman functions}
\end{equation}
These expressions for the kernels in the Einstein-Langevin
equation will be very useful for explicit
calculations.
To simplify the notation, we omit the functional
dependence on the semiclassical metric $g_{ab}$, which will be
understood in all the expressions below.
From (\ref{finite kernels}), we see that the kernels
$N_n^{abcd}(x,y)$ and
$H_{\scriptscriptstyle \!{\rm A}_{\scriptstyle n}}^{abcd}(x,y)$
are the real and imaginary parts, respectively, of
$F_{n}^{abcd}(x,y) \!=\!
\langle 0| \, \hat{T}_n^{ab}(x)\, \hat{T}_n^{cd}(y)\, |0 \rangle
\!-\!
\langle 0| \, \hat{T}_n^{ab}(x)\,|0 \rangle
\langle 0| \,\hat{T}_n^{cd}(y)\, |0 \rangle$. Since, from
(\ref{regul s-t 2}), we can write the operator $\hat{T}_n^{ab}$ as a
sum of terms of the form $\left\{ {\cal A}_x \hat{\Phi}_{n}(x),
\,{\cal B}_x \hat{\Phi}_{n}(x)\right\}$, where ${\cal A}_x$ and
${\cal B}_x$ are some differential operators, we can express
$F_{n}^{abcd}(x,y)$ in
terms of the Wightman function using
\begin{eqnarray}
&&\left\langle
\left\{ {\cal A}_x \hat{\Phi}_{n}(x),
{\cal B}_x \hat{\Phi}_{n}(x)\right\} \hspace{-0.4ex}
\left\{ {\cal C}_y \hat{\Phi}_{n}(y),
{\cal D}_y \hat{\Phi}_{n}(y)\right\} \hspace{-0.1ex}
\right\rangle
\!-\! \left\langle
\left\{ {\cal A}_x \hat{\Phi}_{n}(x),
{\cal B}_x \hat{\Phi}_{n}(x)\right\} \hspace{-0.1ex} \right\rangle
\left\langle
\left\{ \hspace{-0.1ex} {\cal C}_y \hat{\Phi}_{n}(y),
{\cal D}_y \hat{\Phi}_{n}(y)\right\} \hspace{-0.1ex}
\right\rangle \hspace{20ex}
\nonumber \\
&& \hspace{41ex}
= 4 \,{\cal A}_x {\cal C}_y G_n^+(x,y)\,
{\cal B}_x{\cal D}_y G_n^+(x,y)+
4 \,{\cal A}_x {\cal D}_y G_n^+(x,y)\,
{\cal B}_x{\cal C}_y G_n^+(x,y),
\label{Wightman expression}
\end{eqnarray}
where ${\cal C}_x$ and ${\cal D}_x$ are also
some differential operators and where the expectation values are
taken in the vacuum $|0 \rangle$.
This identity
can be easily proved using Wick's theorem or by
writing the operator $\hat{\Phi}_{n}(x)$ in terms of the creation
and annihilation operators of the Fock representation corresponding to
the vacuum $|0 \rangle$. Using a Schwinger-DeWitt expansion for the
Wightman function $G_n^+(x,y)$, one can actually see that the two
terms on the right hand side of the last expression are free of
ultraviolet divergencies in the limit $n\!\rightarrow \! 4$.
Finally, we find
\begin{eqnarray}
F_{n}^{abcd}(x,y)
&=&\bigtriangledown^{a}_{\!\!\! \mbox{}_{x}}\!
\bigtriangledown^{c}_{\!\!\! \mbox{}_{y}}\! G_n^+(x,y)
\bigtriangledown^{b}_{\!\!\! \mbox{}_{x}}\!
\bigtriangledown^{d}_{\!\!\! \mbox{}_{y}} G_n^+(x,y)
+\bigtriangledown^{a}_{\!\!\! \mbox{}_{x}}\!
\bigtriangledown^{d}_{\!\!\! \mbox{}_{y}}\! G_n^+(x,y)
\bigtriangledown^{b}_{\!\!\! \mbox{}_{x}}\!
\bigtriangledown^{c}_{\!\!\! \mbox{}_{y}} G_n^+(x,y)
\hspace{32.5ex}
\nonumber \\
&&
+\, 2\, {\cal D}^{ab}_{\!\! \scriptscriptstyle x} \bigl(
\bigtriangledown^{c}_{\!\!\! \mbox{}_{y}} G_n^+(x,y)
\bigtriangledown^{d}_{\!\!\! \mbox{}_{y}}\! G_n^+(x,y) \bigr)
+2\, {\cal D}^{cd}_{\!\! \scriptscriptstyle y} \bigl(
\bigtriangledown^{a}_{\!\!\! \mbox{}_{x}} G_n^+(x,y)
\bigtriangledown^{b}_{\!\!\! \mbox{}_{x}}\! G_n^+(x,y) \bigr)
+2\, {\cal D}^{ab}_{\!\! \scriptscriptstyle x}
{\cal D}^{cd}_{\!\! \scriptscriptstyle y} \bigl(
G_n^{+ 2}(x,y) \bigr),
\label{Wightman expression 2}
\end{eqnarray}
where ${\cal D}^{ab}_{\!\! \scriptscriptstyle x}$ is the differential
operator (\ref{diff operator}). From this expression and the relations
(\ref{finite kernels}), we get expressions for the kernels
$N_n^{abcd}(x,y)$ and
$H_{\scriptscriptstyle \!{\rm A}_{\scriptstyle n}}^{abcd}(x,y)$ in
terms of the Wightman function $G_n^+(x,y)$.
The kernel
$H_{\scriptscriptstyle \!{\rm S}_{\scriptstyle n}}^{abcd}(x,y)$,
defined in (\ref{kernels}), can be written in terms of the Feynman
function noting that, from Wick's theorem,
\begin{eqnarray}
&&{\rm Im} \, \Bigl\langle {\rm T}^{\displaystyle \ast}\! \Bigl(
\left\{ {\cal A}_x \hat{\Phi}_{n}(x),
{\cal B}_x \hat{\Phi}_{n}(x)\right\} \hspace{-0.3ex}
\left\{ {\cal C}_y \hat{\Phi}_{n}(y),
{\cal D}_y \hat{\Phi}_{n}(y)\right\} \Bigr)
\Bigr\rangle
\nonumber \\
&&\hspace{20ex} =- 4 \, {\rm Im} \Bigl[ {\cal A}_x {\cal C}_y
G\!_{\scriptscriptstyle F_{\scriptstyle \hspace{0.1ex} n}}
\hspace{-0.2ex}(x,y)\,
{\cal B}_x{\cal D}_y
G\!_{\scriptscriptstyle F_{\scriptstyle \hspace{0.1ex} n}}
\hspace{-0.2ex}(x,y)+
{\cal A}_x {\cal D}_y
G\!_{\scriptscriptstyle F_{\scriptstyle \hspace{0.1ex} n}}
\hspace{-0.2ex}(x,y) \,
{\cal B}_x{\cal C}_y
G\!_{\scriptscriptstyle F_{\scriptstyle \hspace{0.1ex} n}}
\hspace{-0.2ex}(x,y) \Bigr],
\label{Feynman expression}
\end{eqnarray}
where, again,
${\cal A}_x$, ${\cal B}_x$, ${\cal C}_x$ and ${\cal D}_x$ are
real differential operators and the expectation value is
in the vacuum $|0 \rangle$. The kernel
$H_{\scriptscriptstyle \!{\rm S}_{\scriptstyle n}}^{abcd}(x,y)$ is then
obtained by adding up the contribution of all the differential
operators which appear in the product $T^{ab}(x) T^{cd}(y)$, where
$T^{ab}$ is the functional (\ref{class s-t}).
After a long calculation, we get
\begin{eqnarray}
H_{\scriptscriptstyle \!{\rm S}_{\scriptstyle n}}^{abcd}(x,y)=
- {1 \over 4} \, {\rm Im} \Bigl[ &&
\bigtriangledown^{a}_{\!\!\! \mbox{}_{x}}\!
\bigtriangledown^{c}_{\!\!\! \mbox{}_{y}}\!
G\!_{\scriptscriptstyle F_{\scriptstyle \hspace{0.1ex} n}}
\hspace{-0.2ex}(x,y)
\bigtriangledown^{b}_{\!\!\! \mbox{}_{x}}\!
\bigtriangledown^{d}_{\!\!\! \mbox{}_{y}}
G\!_{\scriptscriptstyle F_{\scriptstyle \hspace{0.1ex} n}}
\hspace{-0.2ex}(x,y)
+\bigtriangledown^{a}_{\!\!\! \mbox{}_{x}}\!
\bigtriangledown^{d}_{\!\!\! \mbox{}_{y}}\!
G\!_{\scriptscriptstyle F_{\scriptstyle \hspace{0.1ex} n}}
\hspace{-0.2ex}(x,y)
\bigtriangledown^{b}_{\!\!\! \mbox{}_{x}}\!
\bigtriangledown^{c}_{\!\!\! \mbox{}_{y}}
G\!_{\scriptscriptstyle F_{\scriptstyle \hspace{0.1ex} n}}
\hspace{-0.2ex}(x,y) \nonumber \\
&&
-\,g^{ab}(x) \bigtriangledown^{e}_{\!\!\! \mbox{}_{x}}\!
\bigtriangledown^{c}_{\!\!\! \mbox{}_{y}}
G\!_{\scriptscriptstyle F_{\scriptstyle \hspace{0.1ex} n}}
\hspace{-0.2ex}(x,y)
\bigtriangledown_{\!\!e}^{ \mbox{}_{x} }\!
\bigtriangledown^{d}_{\!\!\! \mbox{}_{y}}
G\!_{\scriptscriptstyle F_{\scriptstyle \hspace{0.1ex} n}}
\hspace{-0.2ex}(x,y)
-g^{cd}(y) \bigtriangledown^{a}_{\!\!\! \mbox{}_{x}}\!
\bigtriangledown^{e}_{\!\!\! \mbox{}_{y}}
G\!_{\scriptscriptstyle F_{\scriptstyle \hspace{0.1ex} n}}
\hspace{-0.2ex}(x,y)
\bigtriangledown^{b}_{\!\!\! \mbox{}_{x}}
\bigtriangledown_{\!\!e}^{ \mbox{}_{y} }
G\!_{\scriptscriptstyle F_{\scriptstyle \hspace{0.1ex} n}}
\hspace{-0.2ex}(x,y) \nonumber \\
&&
+\,{1 \over 2}\, g^{ab}(x) g^{cd}(y)
\bigtriangledown^{e}_{\!\!\! \mbox{}_{x}}\!
\bigtriangledown^{f}_{\!\!\! \mbox{}_{y}}
G\!_{\scriptscriptstyle F_{\scriptstyle \hspace{0.1ex} n}}
\hspace{-0.2ex}(x,y)
\bigtriangledown_{\!\!e}^{ \mbox{}_{x} }\!
\bigtriangledown_{\!\!f}^{ \mbox{}_{y} }
G\!_{\scriptscriptstyle F_{\scriptstyle \hspace{0.1ex} n}}
\hspace{-0.2ex}(x,y)
+{\cal K}^{ab}_{\! \scriptscriptstyle x} \bigl(
2 \hspace{-0.2ex} \bigtriangledown^{c}_{\!\!\! \mbox{}_{y}}\!
G\!_{\scriptscriptstyle F_{\scriptstyle \hspace{0.1ex} n}}
\hspace{-0.2ex}(x,y)
\bigtriangledown^{d}_{\!\!\! \mbox{}_{y}}\!
G\!_{\scriptscriptstyle F_{\scriptstyle \hspace{0.1ex} n}}
\hspace{-0.2ex}(x,y)
\nonumber \\
&&
-\, g^{cd}(y) \bigtriangledown^{e}_{\!\!\! \mbox{}_{y}}\!
G\!_{\scriptscriptstyle F_{\scriptstyle \hspace{0.1ex} n}}
\hspace{-0.2ex}(x,y)
\bigtriangledown_{\!\!e}^{ \mbox{}_{y} }\!
G\!_{\scriptscriptstyle F_{\scriptstyle \hspace{0.1ex} n}}
\hspace{-0.2ex}(x,y) \bigr)
+{\cal K}^{cd}_{\! \scriptscriptstyle y} \bigl(
2 \hspace{-0.2ex} \bigtriangledown^{a}_{\!\!\! \mbox{}_{x}}\!
G\!_{\scriptscriptstyle F_{\scriptstyle \hspace{0.1ex} n}}
\hspace{-0.2ex}(x,y)
\bigtriangledown^{b}_{\!\!\! \mbox{}_{x}}\!
G\!_{\scriptscriptstyle F_{\scriptstyle \hspace{0.1ex} n}}
\hspace{-0.2ex}(x,y)
\nonumber \\
&&
-\, g^{ab}(x) \bigtriangledown^{e}_{\!\!\! \mbox{}_{x}}\!
G\!_{\scriptscriptstyle F_{\scriptstyle \hspace{0.1ex} n}}
\hspace{-0.2ex}(x,y)
\bigtriangledown_{\!\!e}^{ \mbox{}_{x} }\!
G\!_{\scriptscriptstyle F_{\scriptstyle \hspace{0.1ex} n}}
\hspace{-0.2ex}(x,y) \bigr)
+2\, {\cal K}^{ab}_{\! \scriptscriptstyle x}
{\cal K}^{cd}_{\! \scriptscriptstyle y} \bigl(
G\!_{\scriptscriptstyle F_{\scriptstyle \hspace{0.1ex} n}}^{\;\: 2}
\hspace{-0.2ex}(x,y) \bigr) \Bigr],
\label{Feynman expression 2}
\end{eqnarray}
where ${\cal K}^{ab}_{\! \scriptscriptstyle x}$ is the differential
operator
\begin{equation}
{\cal K}^{ab}_{\! \scriptscriptstyle x} \equiv
\xi \left( g^{ab}(x) \Box_{\! \scriptscriptstyle x}
-\bigtriangledown^{a}_{\!\!\! \mbox{}_{x}}\!
\bigtriangledown^{b}_{\!\!\! \mbox{}_{x}}+\, G^{ab}(x) \right)
-{1 \over 2}\, m^2 g^{ab}(x).
\label{diff operator K}
\end{equation}
An alternative expression for
$H_{\scriptscriptstyle \!{\rm S}_{\scriptstyle n}}^{abcd}(x,y)$,
which is more similar to expression (\ref{Wightman expression 2}),
can be obtained taking into account that
$G\!_{\scriptscriptstyle F_{\scriptstyle \hspace{0.1ex} n}}
\hspace{-0.2ex}(x,y)$ is a Green function of the Klein-Gordon
equation in $n$ spacetime dimensions, which satisfies
\begin{equation}
\left( \Box_{\! \scriptscriptstyle x} -m^2- \xi R(x) \right)
G\!_{\scriptscriptstyle F_{\scriptstyle \hspace{0.1ex} n}}
\hspace{-0.2ex}(x,y)={\delta^n(x\!-\!y) \over \sqrt{- g(x)}},
\label{Green function eq}
\end{equation}
and using that in dimensional regularization
$\left[\delta^n(x\!-\!y) \right]^2=0$.
Finally, note that, in the vacuum
$|0 \rangle$, the term
$\langle \hat{\Phi}_{n}^2 (x) \rangle$ in
equation (\ref{Einstein-Langevin eq 3}) can also be written as
$\langle \hat{\Phi}_{n}^2(x) \rangle=
i G\!_{\scriptscriptstyle F_{\scriptstyle \hspace{0.1ex} n}}
\hspace{-0.2ex}(x,x)=G_n^+(x,x)$.
It is worth noting that, when the points $x$ and $y$ are spacelike
separated, $\hat{\Phi}_{n}(x)$ and $\hat{\Phi}_{n}(y)$ commute and,
thus, $G_n^+(x,y) \!=\!
i G\!_{\scriptscriptstyle F_{\scriptstyle \hspace{0.1ex} n}}
\hspace{-0.2ex}(x,y) \!=\!
(1/2) \langle 0| \,
\{ \hat{\Phi}_{n}(x) , \hat{\Phi}_{n}(y) \} \, |0 \rangle$, which is
real. Hence, from the above expressions, we have that
$H_{\scriptscriptstyle \!{\rm A}_{\scriptstyle n}}^{abcd}(x,y) \!=\!
H_{\scriptscriptstyle \!{\rm S}_{\scriptstyle n}}^{abcd}(x,y) \!=\!
0$.
This fact is not surprising since, from the causality
of the expectation value of the stress-energy operator, we know that
the non-local dependence on the metric perturbation in the
Einstein-Langevin equation must be causal.
\section{Fluctuations
in stationary and conformally stationary backgrounds}
\label{sec:stationary}
In this section, we derive a number of results concerning the
stochastic semiclassical theory of gravity for two classes of
background solutions of semiclassical gravity. The first class
consists of a stationary spacetime and a scalar field in thermal
equilibrium or in its vacuum state.
In the second class, the spacetime is
conformally stationary, the scalar field is massless and conformally
coupled, and its state is the conformal vacuum or a
thermal state built on the conformal vacuum.
In subsections \ref{subsec:f-d in stationary} and
\ref{subsec:f-d in conformally stationary}, we identify a kernel in
the corresponding Einstein-Langevin equations which is related to the
noise kernel by a fluctuation-dissipation relation.
In subsection \ref{subsec:particle creation}, we study the creation of
particles by stochastic metric perturbations and see that
this phenomenon can be related to the vacuum noise kernel.
We show that the mean value
of created particles is enhanced by the presence of metric
fluctuations with respect to the same quantity
in the ``perturbed'' semiclassical spacetime
$({\cal M},g_{ab} \!+\! \langle h_{ab} \rangle_c)$.
Let us assume that the semiclassical
spacetime $({\cal M},g_{ab})$ is stationary, {\it i.e.}, that
it possesses a global timelike Killing vector field
$\zeta^a$, ${\cal \pounds}_{\mbox{}_{\! \zeta}}g_{ab}=0$, where
${\cal \pounds}_{\mbox{}_{\! \zeta}}$ is the Lie
derivative with respect to $\zeta^a$.
Writing the Killing vector as $\zeta^a=(\partial / \partial t)^a$,
this spacetime can be
foliated by a family of Cauchy hypersurfaces $\Sigma_{t}$, labeled by
the Killing time $t$,
so we can give coordinates $(t,{\bf x})$ to each
spacetime point, where ${\bf x}$ are the space
coordinates on each of these hypersurfaces.
Using this foliation, we can construct a
Hamiltonian operator
$\hat{H}[g]$ in the way described in Appendix \ref{sec:Hamiltonian}.
This is a time independent, {\it i.e.},
independent of the Cauchy hypersurface $\Sigma_{t}$, Hamiltonian
operator, so it represents the Hamiltonian operator
in both the Heisenberg and the Schr\"{o}dinger pictures.
In this case, there is a natural Fock representation based
on a decomposition of the field operator $\hat{\Phi}_{n}[g]$ in
a complete set of modes of positive frequencies $\omega_{k}$
with respect to $\zeta^a$, and their complex
conjugates.\footnote{In some cases,
additional restrictions may be necessary to
avoid infrared divergencies, such as that the scalar field is massive,
$m^2 \!\neq\! 0$, or that the norm of the Killing vector is not
arbitrarily small \cite{wald94,kay}.}
This defines a natural Fock space, the many-particle states of which
are eigenstates of the Hamiltonian $\hat{H}[g]$. Thus,
the notion of particles is
physically well defined in this spacetime
\cite{wald94,kay,mostepanenko}.
The Hamiltonian operator in this Fock representation, renormalized by
normal ordering, is given by
$\hat{H}[g]\!=\!\sum_{k} \omega_{k} \,\hat{a}_k^{\dag} \hat{a}_k$,
where $\hat{a}_k^{\dag}$ and $\hat{a}_k$
are the creation and annihilation
operators on the Fock space. Here,
the summation must be understood as
representing either a sum over a set of discrete indices or an
integral with some suitable measure (or a combination of these two
possibilities).
The time-evolution operator corresponding to this Hamiltonian operator
is then given by
$\hat{\cal U}[g](t,t^{\prime})
\! \equiv \exp \hspace{0.2ex}\bigl(-i \hat{H}[g] \,
(t\!-\!t^{\prime})\bigr)$.
In this section, even if we sometimes write $t_i$ or $t_f$, we shall
always consider these initial and final times in the limit
$t_i \!\rightarrow \! -\infty$ and $t_f \!\rightarrow \! +\infty$
(we assume that such limits can be taken).
\subsection{The fluctuation-dissipation
relation in a stationary background}
\label{subsec:f-d in stationary}
For a real scalar field quantized on the stationary spacetime
$({\cal M},g_{ab})$, we can define a state of thermal equilibrium
at temperature $T$.
This state is described in the Heisenberg picture
by the density operator of the grand canonical ensemble:
\begin{equation}
\hat{\rho}[g]={e^{-\beta \hat{H}[g]} \over
{\rm Tr} \! \left(e^{-\beta \hat{H}[g]} \right) },
\label{thermal state}
\end{equation}
where $\beta \equiv 1/k_B T$ and $k_B$ is Boltzmann's constant
(there are no chemical potential terms
because we deal with a real scalar field).
This kind of thermal states for fields in stationary curved
backgrounds was first considered in
Refs.~\cite{dowker77,gibbons78}.
Since the density operator (\ref{thermal state})
commutes with the time-evolution operator
$\hat{\cal U}[g](t,t^{\prime})$, the corresponding
initial density operator in the Schr\"{o}dinger picture is simply
$\hat{\rho}^{\rm \scriptscriptstyle S}(t_i)\!=\!\hat{\rho}[g]$.
Given any pair of operators in the Heisenberg picture,
$\hat{P}[g](x)$ and $\hat{Q}[g](x)$,
the expectation value
$\langle \hat{P}(x) \hat{Q}(x^{\prime })
\rangle_{\mbox{}_{\mbox{}_{\! \scriptscriptstyle T}}}[g]$
depends on $t$ and $t^{\prime}$ only through the difference
$t-t^{\prime}$, since
\begin{equation}
\langle \hat{P}(x) \hat{Q}(x^{\prime })
\rangle_{\mbox{}_{\mbox{}_{\! \scriptscriptstyle T}}}=
{\rm Tr} \! \left[ \hat{\rho}\, \hat{P}(x)
\hat{Q}(x^{\prime }) \right]=
{\rm Tr} \! \left[ \hat{\rho}\,
\hat{P}^{\rm \scriptscriptstyle S}({\bf x})
e^{-i \hat{H}(t-t^{\prime})}
\hat{Q}^{\rm \scriptscriptstyle S}({\bf x^{\prime}})
e^{i \hat{H}(t-t^{\prime})} \right],
\end{equation}
where $\hat{P}^{\rm \scriptscriptstyle S}({\bf x})$ and
$\hat{Q}^{\rm \scriptscriptstyle S}({\bf x})$ are
the operators in the Schr\"{o}dinger picture corresponding to
$\hat{P}(x)$ and $\hat{Q}(x)$, respectively, and we use
$\langle \hspace{1.5ex} \rangle
_{\mbox{}_{\mbox{}_{\! \scriptscriptstyle T}}}$ to denote an
expectation value in the state described by
(\ref{thermal state}).
In particular, with the choice
$\hat{\rho}^{\rm \scriptscriptstyle S}(t_i)\!=\!\hat{\rho}[g]$,
the kernels
$N_{n }^{abcd}[g](x,x^{\prime })$,
$H_{\scriptscriptstyle \!{\rm S}_{\scriptstyle n
}}^{abcd}[g](x,x^{\prime })$ and
$H_{\scriptscriptstyle \!{\rm A}_{\scriptstyle n
}}^{abcd}[g](x,x^{\prime })$
depend on the time coordinates
as a function of $t-t^{\prime}$.
Therefore, we
can introduce Fourier transforms in the time coordinate as
\begin{equation}
K(x,x^{\prime})= \int^{\infty}_{-\infty} {d\omega \over 2 \pi}\,
e^{-i \omega (t-t^{\prime})}\,
\tilde{K}(\omega;{\bf x},{\bf x^{\prime}}),
\label{Fourier transform}
\end{equation}
where $K(x,x^{\prime})$ is any function which depends on time only
through $t\!-\!t^{\prime}$.
As it is shown in Appendix \ref{sec:Hamiltonian}, Wick's theorem can
be generalized for thermal $N$-point functions, defined as expectation
values
of products of the field operator in the state described by
(\ref{thermal state}). It is then easy to see that the expressions
found in subsection \ref{subsec:vacuum} also hold for the kernels
$N_{n }^{abcd}[g](x,x^{\prime })$,
$H_{\scriptscriptstyle \!{\rm S}_{\scriptstyle n
}}^{abcd}[g](x,x^{\prime })$ and
$H_{\scriptscriptstyle \!{\rm A}_{\scriptstyle n
}}^{abcd}[g](x,x^{\prime })$ at finite $T$ if
we replace the Wightman and Feynman functions
(\ref{Wightman and Feynman functions}) by the analogous thermal
expectation values.
In this case, a simple relationship (in the form of a
fluctuation-dissipation relation) exists
between the kernels
$N_{n }^{abcd}[g](x,x^{\prime })$ and
$H_{\scriptscriptstyle \!{\rm A}_{\scriptstyle n
}}^{abcd}[g](x,x^{\prime })$.
In fact, from (\ref{finite kernels}), we can write these kernels
as
\begin{equation}
8\, N_{n }^{abcd}(x,x^{\prime })=
F_{n}^{abcd}(x,x^{\prime })+
F_{n}^{cdab}(x^{\prime },x),
\hspace{5ex}
8 i\, H_{\scriptscriptstyle \!{\rm A}_{\scriptstyle n
}}^{abcd}(x,x^{\prime })=
F_{n}^{abcd}(x,x^{\prime })-
F_{n}^{cdab}(x^{\prime },x),
\label{rel}
\end{equation}
where we omit the functional dependence on $g_{ab}$.
In terms of the Fourier transforms
(\ref{Fourier transform}), these relations are
\begin{eqnarray}
8\, \tilde{N}_{n }^{abcd}
(\omega;{\bf x},{\bf x^{\prime}})&=&
\tilde{F}_{n}^{abcd}
(\omega;{\bf x},{\bf x^{\prime}})+
\tilde{F}_{n}^{cdab}
(-\omega;{\bf x^{\prime}},{\bf x}), \nonumber \\
8 i\, \tilde{H}_{\scriptscriptstyle \!{\rm A}_{\scriptstyle n
}}^{abcd}
(\omega;{\bf x},{\bf x^{\prime}})&=&
\tilde{F}_{n}^{abcd}
(\omega;{\bf x},{\bf x^{\prime}})+
\tilde{F}_{n}^{cdab}
(-\omega;{\bf x^{\prime}},{\bf x}).
\label{relations}
\end{eqnarray}
By analytically continuing $t$ to complex
values in $F_{n}^{abcd}(x,x^{\prime })$,
one can derive a symmetry relation for this bi-tensor which involves
different values of this complex time. Taking into account that the
time evolution of the operator $\hat{t}_n^{ab}$ is given in this
stationary case by
$\hat{t}_n^{ab}(t+\Delta t,{\bf x})=
e^{i \hat{H} \Delta t}\, \hat{t}_n^{ab}(t,{\bf x}) \,
e^{-i \hat{H} \Delta t}$, and using the
cyclic property of the trace, we get
$F_{n}^{abcd}(t,{\bf x};
t^{\prime },{\bf x}^{\prime })= F_{n}^{cdab}
(t^{\prime },{\bf x}^{\prime };t+i \beta,{\bf x})$,
or, equivalently, in terms of its Fourier transform,
\begin{equation}
\tilde{F}_{n}^{abcd}
(\omega;{\bf x},{\bf x^{\prime}})= e^{\beta \omega}
\tilde{F}_{n}^{cdab}
(-\omega;{\bf x^{\prime}},{\bf x}).
\end{equation}
This relation is known as the Kubo-Martin-Schwinger relation
\cite{kubo,martin}.
From this last expression and (\ref{relations}), we obtain the
following simple relation between
$\tilde{N}_{n }^{abcd}$ and
$\tilde{H}_{\scriptscriptstyle \!{\rm A}_{\scriptstyle n
}}^{abcd}$:
\begin{equation}
\tilde{H}_{\scriptscriptstyle \!{\rm A}_{\scriptstyle n
}}^{abcd}
(\omega;{\bf x},{\bf x^{\prime}})=
-i \, \tanh\! \left( {\beta \omega \over 2} \right)
\tilde{N}_{n }^{abcd}
(\omega;{\bf x},{\bf x^{\prime}}),
\label{f-d relation}
\end{equation}
which can also be written as
\begin{equation}
H_{\scriptscriptstyle \!{\rm A}_{\scriptstyle n
}}^{abcd}
(t,{\bf x};t^{\prime },{\bf x}^{\prime })=
\int^{\infty}_{-\infty} dt^{\prime \prime}\,
K_{\scriptscriptstyle {\rm FD} }
(t-t^{\prime \prime})\,
N_{n }^{abcd}
(t^{\prime \prime},{\bf x};t^{\prime },{\bf x}^{\prime }),
\label{f-d relation 2}
\end{equation}
with
\begin{equation}
K_{\scriptscriptstyle {\rm FD} }(t)
\equiv -\int^{\infty}_{0} {d\omega \over \pi}\,
\sin (\omega t)\hspace{0.2ex}
\tanh\! \left( {\beta \omega \over 2} \right)=
-k_B T \: {\rm P} \hspace{-0.3ex} \left[\hspace{0.2ex}
{\rm cosech}\, (\pi k_B T \hspace{0.2ex} t) \right],
\end{equation}
where ${\rm P}$ denotes a Cauchy principal value distribution.
Since, as we have pointed out above, the kernels
$H_{\scriptscriptstyle \!{\rm A}_{\scriptstyle n
}}^{abcd}$ and
$N_{n }^{abcd}$ are free of ultraviolet
divergencies in the limit
$n\!\rightarrow \! 4$, we can define
\begin{equation}
H_{\scriptscriptstyle \!{\rm A} }^{abcd}
(x,x^{\prime })
\equiv \lim_{n \rightarrow 4}\, \mu^{-2 (n-4)}
H_{\scriptscriptstyle \!{\rm A}_{\scriptstyle n
}}^{abcd}(x,x^{\prime }),
\hspace{5ex}
N ^{abcd}(x,x^{\prime }) \equiv
\lim_{n \rightarrow 4}\, \mu^{-2 (n-4)}
N_{n }^{abcd}(x,x^{\prime }),
\label{physical kernels}
\end{equation}
which are the kernels that appear in the physical semiclassical
Einstein-Langevin equation, Eq.~(\ref{Einstein-Langevin eq}), after
performing the renormalization procedure in
Eq.~(\ref{Einstein-Langevin eq 3}).
These physical kernels will also satisfy the relation
(\ref{f-d relation 2}) or, equivalently, their Fourier transforms will
satisfy (\ref{f-d relation}).
These results are independent of the
regularization method used.
The relation (\ref{f-d relation}) can be written in
an alternative way.
Introducing a new kernel (this is actually a family of
kernels) defined by
$\tilde{H}_{\scriptscriptstyle \!{\rm A}_{\scriptstyle n
}}^{abcd}
(\omega;{\bf x},{\bf x^{\prime}}) \equiv -i \omega \,
\tilde{\gamma}_{n }^{abcd}
(\omega;{\bf x},{\bf x^{\prime}})$, that is,
$H_{\scriptscriptstyle \!{\rm A}_{\scriptstyle n
}}^{abcd}(x,x^{\prime})
= \partial
\gamma_{n }^{abcd}(x,x^{\prime})
/ \partial t$, Eq.~(\ref{f-d relation}) yields
\begin{equation}
\tilde{N}_{n }^{abcd}
(\omega;{\bf x},{\bf x^{\prime}}) =
\omega \:
{\rm cotanh }\! \left( {\beta \omega \over 2} \right)
\tilde{\gamma}_{n }^{abcd}
(\omega;{\bf x},{\bf x^{\prime}}),
\label{f-d relation 3}
\end{equation}
or, equivalently,
\begin{equation}
N_{n }^{abcd}
(t,{\bf x};t^{\prime },{\bf x}^{\prime })=
\int^{\infty}_{-\infty} dt^{\prime \prime}\,
J_{\scriptscriptstyle {\rm FD} }
(t-t^{\prime \prime})\,
\gamma_{n }^{abcd}
(t^{\prime \prime},{\bf x};t^{\prime },{\bf x}^{\prime }),
\label{f-d relation 4}
\end{equation}
where
\begin{equation}
J_{\scriptscriptstyle {\rm FD} }(t)
\equiv \int^{\infty}_{0} {d\omega \over \pi}\,
\cos (\omega t)\: \omega \:
{\rm cotanh }\! \left( {\beta \omega \over 2} \right).
\end{equation}
This integral gives a distribution which is singular at
$t \!=\! 0$ and for $t \! \neq \! 0$ reduces to
$J_{\scriptscriptstyle {\rm FD} }(t)
\! = \! - \pi \, [k_B T \: {\rm cosech}( \pi k_B T \, t)]^2$.
The relations (\ref{f-d relation}) or
(\ref{f-d relation 2}) [or the equivalent forms (\ref{f-d relation 3})
or (\ref{f-d relation 4})] have the same form as the
fluctuation-dissipation relations which appear in quite general
models of quantum mechanics
\cite{landau,kubo,grabert,schwinger61,weber,kubo85}.
The derivation of these relations is usually done in the framework of
linear response theory, in which one considers the response of a
quantum system, which is
initially at thermal equilibrium, when an external classical
time-dependent linear perturbation is ``switched on.''
When evaluating the change in the
expectation value of the relevant operator (the operator which
couples to the perturbation) induced by the presence
of the perturbation,
a dissipative term can be identified as the term
which changes the sign under a time reversal
transformation in the perturbation.
This term is characterized
by a kernel called the dissipation kernel. It can be shown that the
dissipation kernel is related to the fluctuations in equilibrium (in the
absence of the perturbation) of the relevant operator
by a relation which is exactly the same as (\ref{f-d relation 2}) or
(\ref{f-d relation}). This is the fluctuation-dissipation relation.
Using this linear response theory approach, the same
fluctuation-dissipation relation has also been derived for
some models of quantum many-body systems \cite{martin,kubo85} or quantum
fields \cite{weber,jackiw} coupled to external classical fields.
This fluctuation-dissipation relation appears also in the context of
quantum Brownian motion (or ``semiclassical'' Brownian motion),
in which one is interested in
the dynamics of a macroscopic particle
in interaction with a heat bath environment, usually modelized by an
infinite set of quantum harmonic oscillators.
In these models, when the variable representing the center of mass
position of the macroscopic particle decoheres,
it can be effectively described as a
classical stochastic variable.
The equation of motion for this stochastic variable is
a linear Langevin equation with a Gaussian
stochastic source. The classical variable introduced in linear
response theory can be envisaged as the position of the Brownian
particle, but now this variable becomes a dynamical stochastic
variable.
The dissipative term in this Langevin equation is the responsible for
the irreversible dynamics of the Brownian particle.
This term contains a
dissipation kernel which is related to the correlator
of the stochastic source by the relations (\ref{f-d relation 2}) or
(\ref{f-d relation}) \cite{caldeira,hu-paz-zhang}.
This is again the fluctuation-dissipation relation.
There are also some models
in which a purely quantum description of the Brownian particle is
considered \cite{lindenberg,kac}.
The dynamics of this particle is then described by a quantum
operator in the Heisenberg picture. By elimination of all the
environment degrees of freedom in the equation of motion for this
operator, one finds a quantum Langevin equation with quantum
fluctuating and dissipative terms. These terms are again related by a
fluctuation-dissipation relation of the form (\ref{f-d relation 2}) or
(\ref{f-d relation}).
These analogies allow us to identify the equivalent relations
(\ref{f-d relation 2}) and (\ref{f-d relation}),
and the analogous relations
for the physical kernels (\ref{physical kernels}), as the
fluctuation-dissipation relation in our context. Because of this
relation, the kernel
$H_{\scriptscriptstyle \!{\rm A} }^{abcd}
(x,x^{\prime })$
shall be called the dissipation kernel.
The same fluctuation-dissipation relation was derived by Mottola
\cite{mottola} in the context of quantum field theory in curved
spacetime using the linear response theory approach. This author
considered the case in which the background spacetime is static, but
his result is easily generalized to a stationary background.
In this paper, we have derived the
same relation in the context of a Langevin equation for stochastic
metric perturbations, which would presumably describe the
effective dynamics of gravitational fluctuations after a process of
decoherence.
For the particular case of a massless scalar field in a Minkowski
background, this fluctuation-dissipation relation was
derived in Refs.~\cite{campos-hu,campos-hu2}
from an explicit evaluation of the kernels.
It is clear that the kernel
$N ^{abcd}(x,x^{\prime })$ describes
fluctuations in exactly the same sense as the quantum-mechanical
models described above. In fact, as it was pointed out by Mottola
\cite{mottola} from the point of view of linear
response theory, it gives the fluctuations in equilibrium of the
stress-energy operator. Alternatively, as we have shown in the previous
sections, it gives the two-point correlation
function of the Gaussian stochastic source in the semiclassical
Einstein-Langevin equation.
However, the term containing the ``dissipation'' kernel
$H_{\scriptscriptstyle \!{\rm A} }^{abcd}
(x,x^{\prime })$
in the Einstein-Langevin equation does not generally
change sign under a time-reversal transformation in the metric
perturbations.
\subsubsection{Zero temperature limit}
\label{subsub:zero T}
A state of the scalar field which is of special interest is that
described by $\hat{\rho}^{\rm \scriptscriptstyle S}(t_i)\!=\!
\hat{\rho}[g]\! =\! |0 \rangle \langle 0|$, where $|0 \rangle$ is the
vacuum state. This vacuum
state can be obtained as the zero temperature limit,
$T\!\rightarrow \! 0$, of the previous thermal state.
The fluctuation-dissipation relation for this state is
easily obtained by setting $T=0$ in expression (\ref{f-d relation})
or (\ref{f-d relation 2}). We find
$\tilde{H}_{\scriptscriptstyle \!{\rm A}_{\scriptstyle n
}}^{abcd}
(\omega;{\bf x},{\bf x^{\prime}})=
-i \, {\rm sign}\, \omega \,
\tilde{N}_{n }^{abcd}
(\omega;{\bf x},{\bf x^{\prime}})$,
or, equivalently, it has the form (\ref{f-d relation 2}),
with
\begin{equation}
K_{\scriptscriptstyle {\rm FD} }(t)
= -i \int^{\infty}_{-\infty} {d\omega \over 2 \pi}\,
e^{-i \omega t} \,{\rm sign}\, \omega =
-{1 \over \pi}\,
{\rm P}\!\hspace{-0.1ex} \left( {1 \over t} \right).
\end{equation}
This fluctuation-dissipation relation in the alternative form
(\ref{f-d relation 3}) reads
$\tilde{N}_{n }^{abcd}
(\omega;{\bf x},{\bf x^{\prime}}) =
\omega \:
{\rm sign}\, \omega \:
\tilde{\gamma}_{n }^{abcd}
(\omega;{\bf x},{\bf x^{\prime}})$,
or, it has the form (\ref{f-d relation 4}), with
\begin{equation}
J_{\scriptscriptstyle {\rm FD} }(t)
= \int^{\infty}_{0} {d\omega \over \pi}\,
\cos (\omega t)\: \omega =
- {1 \over \pi}\,
{\cal P}\!f \!\hspace{-0.1ex} \left( {1 \over t^2} \right),
\end{equation}
where ${\cal P}\!f ( 1/ t^2 )$
denotes a Hadamard finite part distribution, which is related to
${\rm P}( 1/t )$ by
${\cal P}\!f ( 1/ t^2 ) = - (d/dt) \hspace{0.2ex} {\rm P}( 1/t )$
(the definitions of these distributions can be found in
Refs.~\cite{schwartz}).
\subsubsection{High temperature limit}
Let us now consider the high temperature limit. This limit can only be
performed when there exists a cutoff frequency $\Omega$, such that
$\tilde{N}_{n }^{abcd}
(\omega;{\bf x},{\bf x^{\prime}})$ vanishes for
$\omega\! >\!\Omega$ (by (\ref{f-d relation}),
$\tilde{H}_{\scriptscriptstyle \!{\rm A}_{\scriptstyle n
}}^{abcd}(\omega;{\bf x},{\bf x^{\prime}})$
will also vanish for these values of $\omega$). Such a cutoff
frequency is usually related to a characteristic cutoff frequency of
the environment degrees of freedom. The high temperature limit
corresponds to the limit in which $k_B T \!\gg \! \hbar \Omega$.
In this limit (keeping only the leading order contributions), we
expect that thermal fluctuations dominate over quantum fluctuations.
To study this limit, it is
convenient to restore the dependence in $\hbar$ in the previous
results. For this, one has to multiply the constants
$\alpha_{B}$ and $\beta_{B}$ by
$\hbar$ and the kernel $H_n^{abcd}$ by $1/ \hbar$ in the
Einstein-Langevin equation (\ref{Einstein-Langevin eq 3}),
and change
the combination $\beta \omega$ by $\hbar\beta \omega$ in the previous
expressions. In this limit, we can approximate
$\tanh ( \hbar\beta \omega / 2 ) \simeq \hbar\beta \omega / 2$,
and the fluctuation-dissipation relation
reduces to
\begin{equation}
{1 \over\hbar} \, \tilde{H}_{\scriptscriptstyle \!{\rm A}
_{\scriptstyle n }}^{abcd}
(\omega;{\bf x},{\bf x^{\prime}})=
-i \: {\omega \over 2 k_B T} \,
\tilde{N}_{n }^{abcd}
(\omega;{\bf x},{\bf x^{\prime}}),
\label{classical f-d relation}
\end{equation}
or, equivalently, since in this case
$K_{\scriptscriptstyle {\rm FD} }(t)
\simeq (\hbar / 2 k_B T) \, (d / dt)\, \delta (t)$,
\begin{equation}
{1 \over\hbar} \, H_{\scriptscriptstyle \!{\rm A}_{\scriptstyle n
}}^{abcd}
(t,{\bf x};t^{\prime },{\bf x}^{\prime })=
{1 \over 2 k_B T}\, {\partial \over \partial t}\,
N_{n }^{abcd}
(t,{\bf x};t^{\prime },{\bf x}^{\prime }).
\label{classical f-d relation 2}
\end{equation}
Note that
$(1 /\hbar) \, H_{\scriptscriptstyle \!{\rm A}_{\scriptstyle n
}}^{abcd}$ is the kernel that appears in the
Einstein-Langevin equation (\ref{Einstein-Langevin eq 3}) when one
writes the dependence in $\hbar$ explicitly. This relation has the
same form as the classical Green-Kubo fluctuation-dissipation
relation which appears either in a classical theory of linear response
\cite{green,kubo}
or in a classical theory of Brownian motion \cite{kac,zwanzig}.
Notice, from (\ref{classical f-d relation 2}), that in this high
temperature limit we can simply take
$\gamma_{n }^{abcd}(x,x^{\prime})
= (\hbar / 2 k_B T) \,
N_{n }^{abcd}(x,x^{\prime})$.
\subsection{The fluctuation-dissipation relation
for conformal fields in a conformally stationary background}
\label{subsec:f-d in conformally stationary}
In the case of a massless conformally coupled scalar field ($m\!=\!0$
and $\xi\!=\!1/6$) and a conformally stationary solution of
semiclassical gravity [for instance, a Robertson-Walker (RW)
spacetime], the fluctuation-dissipation relation derived in the previous
subsection can be generalized
when the state of the field in the background solution
is the conformal vacuum or a thermal state built on the conformal
vacuum.
In this case, the action (\ref{scalar field action}) for
the scalar field is conformally invariant. It is convenient to
preserve this conformal invariance when working in dimensional
regularization. This can be done by changing in all the previous
expressions which involve dimensional regularization the parameter
$\xi$ by the function $\xi(n)\!\equiv \!(n-2)/[4(n-1)]$ and, of
course, taking $m\!=\!0$. In this way, the dimensional regularized
stress-tensor operator (\ref{regul s-t}) is traceless.
Let $({\cal M},\overline{g}_{ab})$ be a $n$ dimensional
conformally stationary spacetime, that is, a spacetime with a global
timelike conformal Killing vector field
$\zeta^a$:
${\cal \pounds}_{\mbox{}_{\! \zeta}}\overline{g}_{ab}
=(2/n) \, \hspace{-1.5ex}\mbox{ }^{\mbox{ ^{c}\hspace{-0.5ex}\zeta_{c}
\:\overline{g}_{ab}$, where $\hspace{-1.5ex}\mbox{ }^{\mbox{ _{\!a}$
is the covariant derivative associated to $\overline{g}_{ab}$.
This means that the metric $\overline{g}_{ab}$ is conformally related
to a stationary metric $g_{ab}$:
$\overline{g}_{ab}(x)=e^{2 \varpi(x)} g_{ab}(x)$, where
$\varpi(x)$ is a scalar function.
As previously, writing $\zeta^a\!=\!(\partial / \partial t)^a$,
the semiclassical spacetime can be foliated by Cauchy hypersurfaces
$\Sigma_{t}$ and coordinates $(t,{\bf x})$ can be assigned to the
spacetime points.
There is a ``natural'' Fock representation based
on a decomposition of the field operator
$\hat{\Phi}_{n}[\bar{g}]$ in terms
of a complete set of modes
$\{ \bar{u}_{k_{\mbox{}_{\scriptstyle n}}}\hspace{-0.4ex}(x) \}$,
solution of the Klein-Gordon equation
with metric $\overline{g}_{ab}$,
of the form
$\bar{u}_{k_{\mbox{}_{\scriptstyle n}}}\hspace{-0.4ex}(x)=
e^{-(n-2)\varpi(x)/2} \hspace{0.2ex}
u_{k_{\mbox{}_{\scriptstyle n}}}\hspace{-0.4ex}(x)$,
where
$\{ u_{k_{\mbox{}_{\scriptstyle n}}}\hspace{-0.4ex}(x) \}$
is a complete set of mode solutions
of the Klein-Gordon equation in $( {\cal M}, g_{ab})$
which are of positive frequencies $\omega_{k}$
with respect to $\zeta^a$.
Hence, in this sense, we can write the field operator as
$\hat{\Phi}_{n}[\bar{g}]=
e^{-(n-2)\varpi/2} \hspace{0.2ex} \hat{\Phi}_{n}[g]$,
where $\hat{\Phi}_{n}[g]$ is the
field operator in the stationary spacetime
$( {\cal M}, g_{ab})$.
Assuming that no infrared divergencies are present, so that
this quantum field theory construction is well defined,
the conformal vacuum $|0 \rangle$ is defined as the vacuum
state of the Fock space corresponding to this representation.
If $\hat{a}_k^{\dag}$ and $\hat{a}_k$ are the creation and annihilation
operators on this Fock space, this state satisfies
$\hat{a}_k |0 \rangle\!=\!0$.
As shown in Appendix \ref{sec:Hamiltonian}, in this case we can
construct a conserved energy operator which can be identified with the
Hamiltonian of a field quantized on $( {\cal M}, g_{ab})$:
$\hat{E}[\bar{g}] \!=\!
\hat{H}[g]\!=\!
\sum_{k} \omega_{k} \,\hat{a}_k^{\dag} \hat{a}_k$.
This energy operator, however, is not a
time-evolution generator for the field operator
$\hat{\Phi}_{n}[\bar{g}]$, it
generates the time-evolution of the conformally related operator
$\hat{\Phi}_{n}[g]$.
The many-particle states of the Fock space built on the conformal
vacuum are eigenstates of this energy operator.
From this energy operator, a state of thermal equilibrium
for a conformal scalar field quantized on
$({\cal M},\overline{g}_{ab})$ can be defined using
the density operator (\ref{thermal state}).
Thermal equilibrium states defined in this way
were first proposed by Gibbons and Perry
\cite{gibbons78}. These authors were inspired in a result by
Israel \cite{israel72} in the framework of
relativistic kinetic theory, who found that thermal equilibrium
distribution functions can be defined for massless particles
in conformally stationary spacetimes.
A number of applications have been developed in the literature to
study finite-temperature effects of quantum conformal fields
in RW universes \cite{cooke77}
and in two-dimensional spacetimes \cite{cramer98}.
Let us begin with a solution of the semiclassical
Einstein equation (\ref{semiclassical eq in n})
consisting of a quantum conformal scalar field in
a conformally stationary spacetime
$({\cal M},\overline{g}_{ab})$, in the thermal state
(\ref{thermal state}).
Taking into account that the
action (\ref{scalar action}) with $m\!=\!0$ and
$\xi\!=\!\xi(n)$ satisfies
$S_m[\bar{g},\bar{\Phi}_{n}]\!=\!S_m[g,\Phi_{n}]$, where
$\bar{\Phi}_{n} \!\equiv\! e^{-(n-2)\varpi/2} \Phi_{n}$,
it is easy to see that
$\hat{T}_{n}^{ab}[\bar{g}] =e^{-(n+2)\varpi} \hat{T}_{n}^{ab}[g]$.
Therefore, the kernels evaluated in the thermal state at a
temperature $T$ can be related to the corresponding kernels
for the stationary background
$({\cal M},g_{ab})$. For the noise kernel, we have
\begin{equation}
N_{n }^{abcd}[\bar{g}](x,x^{\prime })=
e^{-(n+2)\varpi(x)}\, e^{-(n+2)\varpi(x^{\prime })}\,
N_{n }^{abcd}[g](x,x^{\prime }),
\label{conformal relations for kernels}
\end{equation}
and the same relation holds for the kernels
$H_{\scriptscriptstyle \!{\rm S}_{\scriptstyle n
}}^{abcd}$ and
$H_{\scriptscriptstyle \!{\rm A}_{\scriptstyle n
}}^{abcd}$.
Since the kernels $N_{n }^{abcd}[g]$ and
$H_{\scriptscriptstyle \!{\rm A}_{\scriptstyle n
}}^{abcd}[g]$ satisfy the relation
(\ref{f-d relation 2}) [or, equivalently, (\ref{f-d relation})],
this leads to a fluctuation-dissipation relation
between the kernels
$N_{n }^{abcd}[\bar{g}]$ and
$H_{\scriptscriptstyle \!{\rm A}_{\scriptstyle n
}}^{abcd}[\bar{g}]$. The same relation holds
for the physical kernels obtained by taking the limit
$n\!\rightarrow \! 4$ as in Eq.~(\ref{physical kernels}).
For the conformal vacuum state, which corresponds to
$T\!=\!0$, the fluctuation-dissipation relation follows directly
from the result of subsection \ref{subsub:zero T}.
In the particular case of a spatially flat
RW solution of semiclassical gravity,
this conformal vacuum fluctuation-dissipation relation
was obtained before in Ref.~\cite{cv96}
after an explicit calculation of the corresponding
kernels.
The same relation was derived in Ref.~\cite{husinha}
in the framework of a ``reduced'' version of the Einstein-Langevin
equation inspired in a Bianchi-I type ``mini-superspace'' model.
\subsection{Particle creation}
\label{subsec:particle creation}
Let us now return to the case in which $({\cal M},g_{ab})$ is
stationary, the scalar field has arbitrary mass $m$ and arbitrary
coupling parameter $\xi$, and consider the stochastic perturbation
$h_{ab}$.
Note that
$({\cal M},g_{ab}\!+\!h_{ab})$ can be viewed as representing an
ensemble of spacetimes distributed according to some
probability distribution functional. We are in fact considering a
scalar field quantized on each of these spacetimes, described by the
operator $\hat{\Phi}[g\!+\!h]$, and the family of states
of the field, described by $\hat{\rho}[g\!+\!h]$.
Let $h_{ab}[\xi]$ be a solution to the
semiclassical Einstein-Langevin equation,
Eq.~(\ref{Einstein-Langevin eq}), whose moments vanish for times
$t<t_{\rm \scriptscriptstyle I}$ or, at least, they vanish
``asymptotically'' in the remote past
($t \! \rightarrow \! - \infty$). This means that there is a
``remote past epoch'' ($t<t_{\rm \scriptscriptstyle I}$ or
$t \! \rightarrow \! - \infty$) in which $h_{ab}$ behaves
deterministically as a zero tensor.
In that case, if we take
$\hat{\rho}^{\rm \scriptscriptstyle S}(t_i)\!=\!
|0 \rangle \langle 0|$, where $|0 \rangle$ is the vacuum of the
natural Fock space for the field quantized on $({\cal M},g_{ab})$,
and we consider the limit $t_i \! \rightarrow \! - \infty$,
we have $\hat{\rho}[g+h]\!=\! |0,{\rm in} \rangle
\langle 0,{\rm in}|$, where
$|0,{\rm in} \rangle$
represents the family of
``in'' vacua for the field quantized on
$({\cal M},g_{ab}\!+\!h_{ab})$.
Treating $h_{ab}$ as a classical ``external'' perturbation, one could
construct a Hamiltonian operator $\hat{H}[g+h](t)$
in the Heisenberg picture for which
$|0,{\rm in} \rangle$ would be the ground state
in the ``remote past epoch.''
However, at later times, due to the presence of the
perturbation $h_{ab}$, this
``in'' vacuum state will generally
not be the ground state of the Hamiltonian.
One then says that ``particles'' are created in the ``in''
vacuum.
Physically meaningful many particle ``out'' states,
in particular, an ``out'' vacuum
$|0,{\rm out} \rangle$
for the scalar field in each of the spacetimes
$({\cal M},g_{ab}\!+\!h_{ab})$ can be defined if there is also a
``far future epoch'' for which $h_{ab}$ vanishes (in the same
statistical sense as above), either in an exact way for
$t > t_{\rm \scriptscriptstyle F}$ or ``asymptotically'' for
$t \! \rightarrow \! + \infty$. When this is the case,
the vacuum persistence amplitude
$\langle 0,{\rm out}|0,{\rm in}\rangle [g\!+\!h]
\equiv e^{i W[g+h]}$ is given by the
following path integral:
\begin{equation}
e^{i W[g+h]}=
\int \! {\cal D}[\Phi_n]\,
\langle 0,{\rm out}|\Phi_n(t_2),t_2 \rangle [g\!+\!h]\;
\langle \Phi_n(t_1),t_1 |0,{\rm in}\rangle [g\!+\!h] \;
e^{i S_{m}[g+h,\Phi_n] },
\label{vacuum persistence amplitude}
\end{equation}
where
$|\varphi_n,t_{1}\rangle$ and $|\varphi_n,t_{2}\rangle$
denote, respectively, eigenstates of the field
operator $\hat{\Phi}_n[g\!+\!h](t,{\bf x})$ at some arbitrary times
$t\!=\!t_1$ and $t\!=\!t_2$, where $t_2>t_1$, with
eigenvalue $\varphi_n({\bf x})$, and where the integration domain for
the action is between $t_1$ and $t_2$. The wave functionals
$\langle \varphi_n,t_{1}|0,{\rm in}\rangle$ and
$\langle \varphi_n,t_{2}|0,{\rm out}\rangle$ have in general a
dependence on the metric, which we have indicated
in (\ref{vacuum persistence amplitude}).
In the limit
$t_1 \!\rightarrow \! -\infty$ and
$t_2 \!\rightarrow \! +\infty$, these wave functionals
do not depend on the perturbation $h_{ab}$.
The total
probability of particle creation is given by \cite{cespedes}
\begin{equation}
P[h;g] = 2 \lim_{n \rightarrow 4}\, {\rm Im}\, W[g\!+\!h].
\label{probability of particle creation}
\end{equation}
One can show that ${\rm Im}\, W$ is free of ultraviolet
divergencies in the limit $n \!\rightarrow\! 4$, and that it is always
positive or zero, so that the probability $P$ is well defined by this
expression.
As we have done in the previous section for the influence action, we
can now expand the action $W[g+h]$ in the perturbation $h_{ab}$.
In order to do so, one has to evaluate the functional derivatives
of $W[g+h]$ in the background metric $g_{ab}$.
Using (\ref{vacuum persistence amplitude}), these derivatives
can be related to ``in-out'' matrix elements of operators
in the background. Since $g_{ab}$ is stationary,
the ``in'' and ``out'' vacua in the background
must be identified with the natural vacuum $|0 \rangle$. Therefore,
these background ``in-out'' matrix elements become expectation values
in the state $|0 \rangle$. It is then easy to see that
the expansion of $W[g+h]$ in the metric perturbation $h_{ab}$ is equal
to that of $S_{\rm IF}[g+h^+,g+h^-]$ with
$h^+_{ab}=h_{ab}$ and
$h^-_{ab}=0$, and taking the expectation values in
$|0 \rangle$. In particular, from
the imaginary part of this expansion
[see Eq.~(\ref{expansion 2})], we get
\begin{equation}
P[h;g]=\int\! d^4x\, d^4y \, \sqrt{- g(x)}\sqrt{- g(y)}\: h_{ab}(x)\,
N^{abcd}[g](x,y)\,
h_{cd}(y)+0(h^3),
\label{stochastic probability}
\end{equation}
where $N^{abcd}$ is the zero temperature
physical noise kernel defined in (\ref{physical kernels}).
This physical noise kernel is
related to the lowest order
quantum stress-energy fluctuations in vacuum
by (\ref{noise}).
Note that the higher order corrections
in (\ref{stochastic probability}) would contain higher order
vacuum stress-energy fluctuations.
Eq.~(\ref{stochastic probability}) is a generalization of an
expression derived by Sexl and Urbantke \cite{sexl} for the total
probability of particle creation by
metric perturbations around Minkowski spacetime.
Eq.~(\ref{stochastic probability}) gives also the expectation value of
the number operator for ``out'' particles in the ``in'' vacuum,
computed to lowest order in the metric perturbation.
In order to show this, let us expand the scalar field
action as the action in the stationary
background plus interaction terms (the terms containing the metric
perturbation).
The interaction term to lowest order in $h_{ab}$ is
$S^{(1)}\!=\! \int \! d^nx \, {\cal L}_n^{(1)}[\Phi_{n},h;g]$, with
${\cal L}_n^{(1)} \!=\!
(1/2) \, \sqrt{- g} \: T^{ab}[g,\Phi_{n}] \hspace{0.2ex} h_{ab}$.
In order to construct the $S$-matrix operator, we need the interaction
Hamiltonian density operator in the interaction picture.
Note that the field and canonical momentum operators in the
interaction picture can be identified with the operators
$\hat{\Phi}_{n}[g]$ and $\hat{\Pi}_{n}[g]$, respectively.
Following Appendix \ref{sec:Hamiltonian}, we can obtain
the canonical Hamiltonian density for the metric
$g_{ab}\!+\!h_{ab}$ and
work out the interaction term to first order
in the metric perturbation.
Although in this case the interaction Lagrangian density depends on
the derivatives of the field, we find that, to first order in
$h_{ab}$, the interaction Hamiltonian density operator in the
interaction picture is given by
$- {\cal L}_n^{(1)}[\hat{\Phi}_{n}[g],h;g]$.
Hence, to first order in the metric perturbation, the
$S$-matrix operator is given by
$\hat{S} = 1 \!+\! \hat{S}^{(1)} \!+\! O(h^2)$, where
\begin{equation}
\hat{S}^{(1)} = {i \over 2} \int \! d^nx \,
\sqrt{- g} \: \hat{T}_n^{ab}[g] \, h_{ab}.
\label{S matrix}
\end{equation}
The expectation value of the ``out'' particle number operator,
$\hat{N}^{\rm out}$,
in the ``in'' vacuum (in the Heisenberg picture) is given by
$N[h;g] \equiv \langle 0,{\rm in}| \hspace{0.2ex}
\hat{N}^{\rm out} \hspace{0.2ex} |0,{\rm in} \rangle
= \langle 0 | \hat{S}^{\dag} \hat{N} \hat{S} |0 \rangle$,
where $\hat{N}$ is the particle number operator in the background
$\hat{N} \!\equiv \! \sum_{k} \hat{a}_k^{\dag} \hat{a}_k$.
To lowest order, we have
\begin{equation}
N[h;g] = \sum_{k,p}
\left| \langle 1_k, 1_p | \hat{S}^{(1)} |0 \rangle \right|^2
+ O(h^3),
\label{number of particles}
\end{equation}
where $| 1_k, 1_p \rangle$ is the two-particle state
$| 1_k, 1_p \rangle \equiv
\hat{a}_k^{\dag} \hat{a}_p^{\dag} \, |0 \rangle$.
Clearly, since $\hat{S}^{(1)}$ is quadratic in the field operator,
at this order $N$ can also be written as
$N/2 =
\sum_n \langle 0| \hat{S}^{(1)\, \dag} |n \rangle
\langle n| \hat{S}^{(1)} |0 \rangle
-\langle 0| \hat{S}^{(1)\, \dag} |0 \rangle
\langle 0| \hat{S}^{(1)} |0 \rangle+ O(h^3)$, where $\{ |n \rangle \}$
represents the complete orthonormal basis of the Fock space.
Using (\ref{S matrix}), this last expression can be written in terms
of the vacuum noise kernel
$N_{n }^{abcd}[g](x,y)$ [see (\ref{kernels})].
Taking the limit $n \!\rightarrow \! 4$, we see that
the expression for one half of
the number of created particles $N[h;g]/2$ to lowest order in the metric
perturbation coincides with that for $P[h;g]$ in
Eq.~(\ref{stochastic probability}).
The energy of the created particles, defined as
$E[h;g] \equiv \langle 0,{\rm in}| \hspace{0.2ex}
\sum_k \omega_k \hspace{0.2ex}
\hat{N}_k^{\rm out} \hspace{0.2ex} |0,{\rm in} \rangle
= \langle 0 | \hat{S}^{\dag} \hat{H} \hat{S} |0 \rangle$,
where $\hat{N}_k^{\rm out}$ is the ``out'' number operator in the
$k$ mode and $\hat{H}\!=\!
\sum_{k} \omega_{k} \,\hat{a}_k^{\dag} \hat{a}_k$
is the Hamiltonian operator
in the background, is similarly given by
\begin{equation}
E[h;g] = {1 \over 2} \sum_{k,p} (\omega_k + \omega_p)
\left| \langle 1_k, 1_p | \hat{S}^{(1)} |0 \rangle \right|^2
+ O(h^3).
\label{energy of particles}
\end{equation}
Comparison of (\ref{energy of particles})
with (\ref{number of particles}) and
(\ref{stochastic probability}), suggests that it may be possible
in some cases to write
this last expression (in the limit
$n \!\rightarrow \! 4$) in terms of the
Fourier transform of the vacuum noise kernel.
As an example, let us consider
the case when $({\cal M},g_{ab})$ is
$({\rm I\hspace{-0.4 ex}R}^{4},\eta_{ab})$
\cite{mv98,paperII}, which is the trivial solution of semiclassical
gravity. Working in a global
inertial coordinate system $\{x^\mu \}$, in this case
the kernels depend only on the difference $(x\!-\!y)^\mu$ and, thus,
we can define their Fourier transforms as
$K(x \!-\! y) \equiv (2\pi)^{-4} \!
\int \! d^4 p \,
e^{i p \cdot (x-y)}\, \tilde{K}(p)$, where
$p \hspace{-0.2ex} \cdot\! x \equiv \eta_{\mu \nu} p^\mu x^\nu$.
Introducing the Fourier
transform of $h_{ab}(x)$ in a similar way
[note that $\tilde{h}_{ab}(-p) \!=\!
\tilde{h}_{ab}^{\displaystyle \ast}(p)$],
(\ref{stochastic probability}) can be written as
\begin{equation}
P[h;\eta] = \int \! {d^4 p \over (2\pi)^4 } \,
\tilde{N}^{abcd}(p) \,
\tilde{h}_{ab}^{\displaystyle \ast}(p) \,
\tilde{h}_{cd}(p) + O(h^3).
\label{prob in Minkowski}
\end{equation}
On the other hand, the energy of the created particles is given by
\cite{frieman}
\begin{equation}
E[h;\eta] = 2 \int \! {d^4 p \over (2\pi)^4 } \:
p^0 \, \theta(p^0) \,
\tilde{N}^{abcd}(p) \,
\tilde{h}_{ab}^{\displaystyle \ast}(p) \,
\tilde{h}_{cd}(p) + O(h^3).
\end{equation}
The vacuum noise and dissipation kernels for a Minkowski background
can be written in terms of two pairs of scalar kernels,
$N_r(x \!-\! y)$ and $D_r(x \!-\! y)$, respectively,
with $r \!=\! 1,2$ \cite{paperII} (see also Ref.~\cite{mv98} for a
particular case in which $N_2\!=\!D_2\!=\!0$).
Each pair of kernels $(N_r, D_r)$ satisfies the
fluctuation-dissipation relation found in
subsection \ref{subsub:zero T}.
One finds \cite{cespedes,frieman} that
\begin{equation}
\tilde{N}^{abcd}(p) \,
\tilde{h}_{ab}^{\displaystyle \ast}(p) \,
\tilde{h}_{cd}(p) =
\tilde{C}^{{\scriptscriptstyle (1)}}_{abcd}(p)\,
\tilde{C}^{{\scriptscriptstyle (1)}
{\displaystyle \hspace{0.1ex}\ast \hspace{0.1ex}}
abcd}(p) \, \tilde{N}_1(p)+
\left| \hspace{0.1ex}
\tilde{R}^{{\scriptscriptstyle (1)}}(p)
\hspace{0.1ex}\right|^2 \tilde{N}_2(p),
\label{relation in Minkowski}
\end{equation}
where $\tilde{C}^{{\scriptscriptstyle (1)}}_{abcd}(p)$,
$\tilde{R}^{{\scriptscriptstyle (1)}}(p)$ and $\tilde{N}_r(p)$
are, respectively, the
Fourier transforms of the linearized Weyl tensor, the scalar
curvature and the
kernels $N_r(x \!-\! y)$, $r \!=\! 1,2$;
$\tilde{N}_r(p)$ depend only on
$p^2 \equiv \eta_{\mu \nu} p^\mu p^\nu$. It is then easy to see, using
the fluctuation-dissipation relation, that
\begin{equation}
E[h;\eta] = i \int \! {d^4 p \over (2\pi)^4 } \:
p^0 \left[ \tilde{C}^{{\scriptscriptstyle (1)}}_{abcd}(p)\,
\tilde{C}^{{\scriptscriptstyle (1)}
{\displaystyle \hspace{0.1ex}\ast \hspace{0.1ex}}
abcd}(p) \, \tilde{D}_1(p)+
\left| \hspace{0.1ex}
\tilde{R}^{{\scriptscriptstyle (1)}}(p)
\hspace{0.1ex}\right|^2 \tilde{D}_2(p) \right]
+O(h^3).
\label{energy of particles in Minkowski}
\end{equation}
Hence, in the case of a Minkowski background, the energy
of the created particles can be expressed in terms of the dissipation
kernels $D_1$ and $D_2$ for the Minkowskian vacuum.
It is not clear, however, that, for other
stationary backgrounds, the energy
of the created particles can be related to dissipation in vacuum in a
similar way.
The probability of particle creation
(\ref{stochastic probability}) is
a fluctuating quantity, due to the functional dependence
on the stochastic perturbation $h_{ab}$.
We may compute its averaged
value $\langle P[h;g] \rangle_c$, which [neglecting the higher order
corrections in Eq.~(\ref{stochastic probability})] is given by
\begin{equation}
\langle P[h;g] \rangle_c = P[\langle h \rangle_c;g]+
\int\! d^4x\, d^4y \, \sqrt{- g(x)}\sqrt{- g(y)}\:
N^{abcd}[g](x,y)\,
\langle h^{\rm f}_{ab}(x) h^{\rm f}_{cd}(y) \rangle_c,
\label{mean P}
\end{equation}
where $h_{ab}^{\rm f} \equiv h_{ab} -\langle h_{ab} \rangle_c$.
The first term in the right hand side of Eq.~(\ref{mean P})
is the probability of particle creation
(or one half of the number of created particles) that one would
obtain in the spacetime
$({\cal M},g_{ab} \!+\! \langle h_{ab} \rangle_c)$.
The second term will be greater than zero when stress-energy
fluctuations are present\footnote{Except in some rare cases,
for which
$N^{abcd}(x,y)$ is not strictly positive
definite and $\langle h^{\rm f}_{ab}(x) h^{\rm f}_{cd}(y) \rangle_c$
is such that it ``hits'' the zero eigenvalue.}
since, from the Einstein-Langevin equation, this implies
$\langle h^{\rm f}_{ab}(x) h^{\rm f}_{cd}(y) \rangle_c \!\neq \! 0$.
Note that, when this is the case, from the fluctuation-dissipation
relation of subsection \ref{subsub:zero T}, the vacuum
dissipation kernel will be also non-vanishing.
Hence, metric fluctuations induced by matter stress-energy
fluctuations generally increase the mean value of the number of
created particles with respect to the same quantity
in the ``perturbed'' semiclassical spacetime
$({\cal M},g_{ab} \!+\! \langle h_{ab} \rangle_c)$.
The above result for the total probability of particle creation
and number of created particles can
be easily generalized to the case of a massless conformally coupled
scalar field and
a conformally stationary semiclassical background.
When this background is a spatially
flat RW universe \cite{cv96},
performing conformal transformations in the metric
perturbations and in the kernels as in
Eq.~(\ref{conformal relations for kernels}),
one gets expressions analogous to (\ref{prob in Minkowski}),
(\ref{relation in Minkowski}) and
(\ref{energy of particles in Minkowski}) with
$N_2 \!=\! D_2 \!=\! 0$ (see Refs.~\cite{cv94,cv96,mv98}
for more details).
\section{Conclusions}
\label{sec:conclu}
In the first part of this paper,
we have shown how a consistent stochastic semiclassical
theory of gravity can be formulated. This theory is a perturbative
generalization of semiclassical gravity which describes the back
reaction of the lowest order stress-energy fluctuations of quantum
matter fields on the gravitational field through the semiclassical
Einstein-Langevin equation. We have shown that this equation
can be formally derived with a method
based on the influence functional of Feynman and Vernon, where one
considers the metric field as the ``system'' of interest and the
matter fields as part of its ``environment'' \cite{hu89}.
Our approach clarifies the physical meaning of the semiclassical
Langevin-type equations previously derived with the same functional
method
\cite{calzettahu,humatacz,husinha,cv96,lomb-mazz,cv97,ccv97,%
campos-hu,campos-hu2,calver98}, since it links the source of stochastic
fluctuations to quantum matter stress-energy fluctuations, and allows
to formulate the theory in a general way.
At the same time, we have also developed a method to compute the
semiclassical Einstein-Langevin equation using dimensional
regularization. This provides an alternative and more direct way of
computing the equation with respect to the previous calculations,
based on a specific evaluation of the effective action of Feynman and
Vernon \cite{calzettahu,humatacz,husinha,cv96,lomb-mazz,cv97,ccv97,%
campos-hu,campos-hu2,calver98}.
In a subsequent paper \cite{paperII}, we shall apply this method to
solve the Einstein-Langevin equation around some simple solutions of
semiclassical gravity.
The second part of the paper was devoted to the existence
of fluctuation-dissipation relations and to particle creation in
the context of stochastic semiclassical gravity.
When the background solution of semiclassical gravity consists of a
stationary spacetime and a scalar field in a thermal equilibrium state,
we have identified a dissipation kernel in the Einstein-Langevin
equation which is related to the noise kernel by a
fluctuation-dissipation relation. The same relation was
previously derived by Mottola \cite{mottola} using a linear response
theory approach. We have also generalized this result to the case of a
conformal scalar field in a conformally stationary background solution
of semiclassical gravity.
Our analysis seems to indicate that for a fluctuation-dissipation
relation to be present in stochastic semiclassical gravity, the
semiclassical background solution must satisfy certain conditions.
In this paper we have just analyzed the simplest cases for which such
a relation exists. Further work must be done to investigate whether a
similar relation is present in other situations of physical
interest,
such as black hole backgrounds \cite{hu99,hu-raval-sinha,campos-hu2},
or non-conformal fields in RW backgrounds in the instantaneous
vacua or the thermal states defined in Ref.~\cite{weiss86}.
We have also studied particle creation by stochastic metric
perturbations in stationary and conformally stationary (for conformal
matter fields in this latter case) background solutions of
semiclassical gravity. We have expressed the
total probability of particle creation and the number of created
particles (the expectation value of the number operator for
``out'' particles in the ``in'' vacuum) in terms of the vacuum noise
kernel. We have shown that the averaged value of those quantities
is enhanced by the presence of stochastic metric fluctuations.
In the particular cases of a Minkowski
background and a conformal field in a spatially flat RW background,
the energy of the created particles can
be expressed in terms of the vacuum dissipation kernels.
It should be stressed that the concept of particle creation is only
well defined when the solutions of the Einstein-Langevin equation
vanish in the ``remote past''
and in the ``far future'' (at least, ``asymptotically'').
However,
there can be physically meaningful solutions of the Einstein-Langevin
equation that do not satisfy these rather strong conditions.
In this case,
vacuum noise and dissipation in stochastic semiclassical gravity can
include effects that are not associated to particle creation.
\acknowledgments
We are grateful to Esteban Calzetta, Jaume Garriga,
Bei-Lok Hu, Ted Jacobson and Albert Roura
for very helpful suggestions and discussions.
This work has been partially supported by the
CICYT Research Project number
\mbox{AEN95-0590}, and the European Project number
\mbox{CI1-CT94-0004}.
|
\section{Introduction}
The Rossi X-Ray Timing Explorer (RXTE) has uncovered strong
oscillations of X-ray flux during X-ray bursts in several low mass
X-ray binaries (Strohmayer et al. 1996). Current interpretation
favors a rotation mechanism for the burst oscillations: asymmetric
nuclear burning leaves a `hot spot' which rotates with the neutron
star and produces a strong modulation (Strohmayer et al. 1998). The
frequency of the burst oscillation is then the spin frequency of the
neutron star, or twice the spin frequency for two spots (Miller 1999).
Oscillations have been discovered in X-ray bursts from the following
systems: 4U 1728-34 (363 Hz, Strohmayer et al. 1996; Strohmayer, Zhang
\& Swank 1997), KS~1731-260 (524 Hz, Smith, Morgan \& Bradt 1996), a
source near the galactic center (589 Hz, Strohmayer, Jahoda \& Giles
1997), Aql X-1 (549 Hz, Zhang et al. 1998), 4U 1636-536 (581/290 Hz,
Strohmayer et al. 1998, Miller 1999), and 4U 1702-429 (330 Hz,
Markwardt, Strohmayer \& Swank 1999). The observed frequencies are
close to the 401 Hz spin frequency of the accreting millisecond pulsar
SAX J1808.4-3658 (Wijnands \& van der Klis 1998), further
strengthening the identification of these frequencies with the neutron
star spin.
The detailed energy dependence of these burst oscillations is one
avenue that remains to be explored. Here I show that the low energy
photons in a burst oscillation from Aql X-1 lag the high energy
photons by roughly 15\% of the oscillation period. Lags of the same
sign and similar magnitudes have also been detected in other fast
signals from low mass X-ray binaries: the kilohertz quasi-periodic
oscillations, QPOs (Vaughan et al. 1997, 1998, Kaaret et al. 1999) and
the SAX J1808.4-3658 pulsed emission (Cui, Morgan \& Titarchuk 1998).
A simple mechanism of Doppler shifted emission may explain these lags.
Strong Doppler effects are expected to be important since the fast
spin rates imply high speeds ($\beta=v/c \sim0.1$). As a hot spot on
the spinning neutron star approaches the observer (at early phases)
the emission is Doppler boosted and blue shifted, as it recedes (at
later phases) the emission is deboosted and red shifted. At early
phases the spectra are also attenuated due to the smaller projected
area. The result is that low energy photons are preferentially
emitted after the high energy photons. A quantitative test of this
Doppler delay scenario matches the observed low energy lags in Aql~X-1
well. The possibility of Doppler effects and the fact that they may
manifest in pulse phase spectroscopy has been noted before by
Strohmayer et al. (1998).
In the next section I present the measurement of the lag in the X-ray
burst from Aql X-1. In Section~3 I describe a simple model for the
relativistic effects and compare the predicted delays to those
observed. Section~4 discusses these results in a broader context.
\section{Measurements}
For this analysis I consider the X-ray burst from Aql~X-1 starting 1
1997 March 1 23:27:40 UTC (see Zhang et al. 1998 for a report of this
burst). I use data from the RXTE Proportional Counter Array (PCA) in
an `event' mode with high time resolution (122$\mu$sec) and high
energy resolution (64 channels). A section of the lightcurve is shown
in Figure~1 (top). There are gaps in the event mode data since the
required telemetry rate is high. Within the 4 second time window
shown in Figure~1 (top), the power density spectrum for all the
channels shows a strong oscillation at 549.7 Hz (Figure~1, bottom). In
the following I calculate Fourier transforms within this time window.
Phase delays in a signal between two energy bands are quantified by
means of cross spectral analysis (van der Klis et al. 1987; for more
information see Vaughan et al. 1994; Nowak et al. 1999). The cross
spectrum is defined as $C(j) = X_1^{*}(j) X_2(j)$, where $X$ are the
measured complex Fourier coefficients for the two energy bands at a
frequency $\nu_j$. The phase lag between the signals in the two bands
is given by the argument of $C$ (its position angle in the complex
plane). The error in the phase lag is calculated here from the
coherence function uncorrected for counting statistics (Nowak et
al. 1999). The cross correlation code used here has been employed to
calculate phase lags in black hole candidates (Ford et al. 1999), SAX
J1808.4-3658 and kilohertz QPOs and matches the results reported in
the literature.
Figure~2 shows the resulting phase lags from the cross spectra of the
4 seconds of data described above. Negative numbers indicate that the
oscillations in the low energy band (3.5--5.7 keV) lag those in the
higher energy bands. The lags are calculated by averaging the signal
in the range 549.6--550.1 Hz. The delays in each band up to 30 keV
(where background dominates) are 3$\sigma$ significant. The delay
between 3.5--5.7 keV and the entire 5.7--43.6 keV band is
$0.93\pm0.18$ rad, 5$\sigma$ significant.
Deadtime effects can in principle affect the measured phase lag. The
data considered here are in the tail of the burst (rate of 9280 c/s,
full energy band) where deadtime is less important. One method of
correcting for deadtime is to subtract a cross vector averaged over
high frequencies where no correlation is expected (van der Klis et
al. 1987). Employing this correction does not change the values
measured here.
Due to the data gaps it is not possible to perform cross correlations
on long stretches of data earlier in the burst. Cross correlations on
0.5 sec intervals of data earlier in the burst return large errors on
phase delays with inconclusive results.
\section{Model}
As a simple model for the lags I consider discrete hot spots on the
surface of the rotating neutron star. The rest frame emission of the
clump is a blackbody. The observed spectrum at frequency $\nu$ at
spin phase $\theta$ is:
$$ F_\nu(\nu) = A_0 cos\delta [\gamma^{-1} (1-\beta\mu
cos\theta)^{-1}]^3 $$
$$~~~~~~~~~ \times ~~ \nu^3 [exp(\nu/kT)-1]^{-1} $$ where $\beta=v/c$,
$\gamma$ is the Lorentz factor, $kT=kT_0 \gamma^{-1} (1-\beta\mu
cos\theta)^{-1}$ (with $kT_0$ the rest frame temperature), $\mu$ is
the sine of the angle between the spin axis and the line of sight, and
$A_0$ is a normalization. The above formula is a relativistic
transformation of the blackbody which shifts $kT$ and modifies the
normalization such that $F/ \nu^3$ is conserved (see Rybicki \&
Lightman 1979). The $cos\delta$ term is an area projection factor,
with $\delta$ the angle between the normal and line of sight in the
rest frame ($\delta \sim \pi-\theta$). The phase angle, $\theta$, is
defined such that phase zero is with the spot approaching the observer
directly. The spots are considered small and isotropically emitting in
the rest frame.
I take $kT_0$=1 keV, $\beta=0.1$. These are values appropriate for the
neutron star; a more exact value of $kT_0$ is in principle possible
from the spectral fits but this depends on the fraction of the surface
contributing to the modulated hot spot emission. A more exact value of
$\beta$ depends on the neutron star radius. I also take $\mu=1$,
i.e. a line of sight through the equator. The spin frequency is 275
Hz and two antipodal hot spots produce an oscillation at 550 Hz. Such
a geometry, where the $\sim$550 Hz signal is a harmonic of the spin,
is suggested by recent results on other burst oscillations (Miller
1999). The resulting spectra are blackbodies whose temperature shifts
by 10\% over the period. Averaged over phase, the spectrum is
approximately blackbody in shape with $kT$ within 1\% of the input
$kT_0$.
The spectra as a function of $\theta$, folded through the RXTE
response matrix, yield lightcurves of count rates in various energy
bands. From these lightcurves I calculate the phase lag in the 550 Hz
signal with the FFT and cross-correlation programs used in the
measurements above. The results of this calculation are shown with the
data in Figure~2. There are no free parameters, only the assumptions
taken above. The calculated lags will decrease if $kT_0$ is increased
or $\beta$ is decreased. The smaller delay for higher $kT_0$ happens
since the peak of the lightcurves comes later in phase for higher
energy photons, corresponding to a smaller delay between high and low
energy photons. Observing at higher inclinations (decreased $\mu$)
will also decrease the lag. The lightcurves that yield these predicted
lags generally have maxima at earlier phases for higher energies and
are more sharply peaked in shape at higher energies.
This simple model neglects general relativistic effects
(e.g. Strohmayer 1992, Miller \& Lamb 1998). Two main factors from GR
will effect the observed light curves. Gravitational bending makes the
spots observable at $\theta<0$ or $\theta>\pi$, stretching the
pulse. Light travel time delays, longer for more extreme bending, will
also shift the pulse. These effects depend on the compactness of the
star. Given the quality of the present data, a more detailed treatment
including these effects is not justified. An overall gravitational
redshift also means that $kT$ in the local frame is higher, as in
X-ray burst spectral models.
\section{Discussion}
The previous sections show that low energy photons lag high energy
photons in the oscillation signal of an X-ray burst from Aql X-1. The
sign and magnitude of the lags are in agreement with the simple model
considered in Section~3 of two hot spots on the neutron star producing
a Doppler boosted and shifted spectra as the star rotates.
This Doppler delay mechanism for producing low energy lags may
describe not only the lags in the X-ray burst oscillations but also
the lags in the accreting millisecond pulsar SAX J1808.4-3658 (Cui,
Morgan \& Titarchuk 1998) and the (lower frequency) kilohertz QPOs
(Vaughan et al. 1997, 1998; Kaaret et al. 1999). Both show a lag of
low energy photons relative to high energy photons with magnitudes of
roughly $\sim$100 $\mu$sec ($\sim0.3$ rad) for SAX J1808.4-3658 and
$\sim$30 $\mu$sec ($\sim0.2$ rad) for the kilohertz QPOs in similar
energy bands to those considered here. Some models link the frequency
of the kilohertz QPOs to a Keplerian motion in the disk (Miller, Lamb
\& Psaltis 1998, Stella \& Vietri 1999; but see Titarchuk, Lapidus \&
Muslimov 1998). If any of the kilohertz QPOs is a result of Keplerian
motion, one might expect a soft lag due to Doppler delays. Such lags
have been observed in what is likely the lower frequency of the two
QPOs.
Doppler delays are an alternative to previous mechanisms invoked to
produce lags. Comptonization has been one process used to explain low
energy lags in SAX J1808.4-3658 (Cui, Morgan \& Titarchuk 1998). Low
energy lags are produced if high energy photons are injected into a
relatively cool Comptonizing cloud. This is the opposite of the
situation normally considered: Comptonization by a hot cloud in the
same region. A hot cloud produces a lag of high energy photons, as
shown quantitatively for fast signals by Lee \& Miller (1998).
Another mechanism suggested for low energy delays is an extended,
cooling hot spot with lower energy photons from the outer regions
(Cui, Morgan \& Titarchuk 1998).
More measurements of phase lags in X-ray burst oscillations are
clearly needed, in particular in the $\sim$350 Hz oscillations which
are likely from single spots. Improved statistics will also yield a
better test of the predicted energy dependence of the lags.
I thank Michiel van der Klis, Jan van Paradijs, Mariano M\'endez and Walter
Lewin for helpful comments. I thank Katja Pottschmidt and coworkers
at the University of Tuebingen for comparisons of our cross
correlation codes. I acknowledge support by the Netherlands
Foundation for Research in Astronomy with financial aid from the
Netherlands Organization for Scientific Research (NWO) under contract
numbers 782-376-011 and 781-76-017 and by the Netherlands
Researchschool for Astronomy (NOVA).
|
\section{Introduction}
An optical vortex soliton appears as a stationary self-trapped beam in a
self-defocusing optical medium that carries a phase singularity on an
electromagnetic field, so that the beam intensity vanishes at a certain
point, and the field phase changes by $2\pi m$ ($m$ being integer) along
any closed loop around the zero-intensity
point. If such an object is created in a linear bulk medium
\cite{nye,nye_book}, it
preserves the singularity but expands due to diffraction.
However,
in a nonlinear medium, the diffraction-induced expansion of the vortex
core can be compensated for by a nonlinearity-induced change in the
refractive index of a nonlinear medium, thereby creating a stationary
self-trapped structure, {\em an optical vortex soliton.} Such nonlinear
localized waves carrying a singularity
were first introduced as stationary solutions of the nonlinear
Schr{\"o}dinger (NLS) equation in the pioneering paper by Ginzburg and
Pitaevsky \cite{gin_pit} to describe topological excitations in
superfluids, but the same objects appear in many other fields \cite{pismen}
including nonlinear optics \cite{dark_review}.
The parametric interactions may provide an efficient way of vortex
transformation. In particular, by mixing
waves of different frequencies, one can change the vortex topological
charge $m$ and even the vortex polarization. Recently, the first experimental
results on the vortex generation in the presence of two-wave parametric
mixing have been reported in nonlinear optics, including the second-harmonic
generation (SHG) \cite{soskin,dho} and more general types of frequency
conversion \cite{berz} and sum-frequency mixing \cite{lithuania} where the
generation
of higher-order ($|m| >1)$ linear vortices in the case of negligible
spatial walk-off between harmonics was demonstrated.
To the best of our knowledge, no theory of parametric optical vortices
in
the presence of both diffraction and nonlinearity has been developed so
far. In a nonlinear regime, an interplay between diffraction and
parametric coupling of the harmonic fields is expected to lead to the
formation of stationary structures -
{\em parametric vortex solitons} - supported by three- or four-wave
mixing between
the phase-matched waves of different frequencies. Stability of such multi-
frequency
vortex
solitons is a key issue. For example, in the problem of
SHG in a diffractive bulk medium, vortex
solitons
are expected to be unstable due to parametric modulational instability
of
the two-wave background field \cite{MI_bk}. Recently, it has been
suggested
\cite{tristram} that taking into account {\em a weak defocusing cubic
nonlinearity} one can eliminate the
development of parametric modulational instability allowing stable dark
solitons
to exist. Some examples of stable two-wave parametric dark
solitons have been presented in Ref. \cite{tristram}, and it has
been pointed out that, in the problem of SHG, a stable vortex soliton of
the lowest possible charge
($|m| = 1$) can exist describing a $2\pi$-phase twist of
the fundamental wave and $4\pi$-phase twist in the second-harmonic field.
In the present paper we suggest a general approach to the analysis of
{\em multi-component vortex solitons} resulting from parametric wave mixing.
The general theory is then developed in detail in the no-walkoff case for
{\em two examples}:
(i) parametric interaction of the first and second
harmonics in a medium with competing quadratic and cubic nonlinearity,
and
(ii) parametric interaction between the first and third
harmonics in a medium with a cubic nonlinear response.
In both the cases we find different classes of vortex solitons as (2+1)-
dimensional dark solitons
of circular symmetry carrying a phase singularity,
and investigate their stability to propagation and modulational stability of
the supporting two-component background waves.
The paper is organized as follows. In Sec. II we briefly present two models of
parametric wave interaction that describe a phase-matched coupling
between
the fundamental frequency mode and its harmonic field, in the case of
phase-matched wave mixing and no walk-off. The further analysis of the
asymptotic structure of stationary localized solutions for parametric
vortex solitons is rather general,
and it is presented in Sec. III for both the models. Section IV is
devoted
to the analysis of vortex solitons in the model of competing
nonlinearities.
We find numerically the profiles of two-component vortex solitons and
investigate their stability to propagation. In particular,
we
reveal the existence of novel classes of dark-soliton solutions of
radial
symmetry, including {\em a ring-vortex soliton}, that consists of a
vortex
core in the harmonic field surrounded by a bright ring of its
fundamental
frequency, and
{\em a halo-vortex}, a two-wave vortex soliton with nonmonotonic tails.
The corresponding results are also obtained for the problem of
the third-harmonic generation in Sec. V. Finally, Sec. VI gives the
summary of our results and briefly discusses some related issues including
the comments on experimental verifications and a link with other problems.
\section{Models of two-wave parametric interaction}
\subsection{Competing Nonlinearities}
First, we consider the model of competing quadratic and cubic
nonlinearities introduced earlier for the (1+1)-dimensional case
in Ref. \cite{compet} and recently generalized to the case of
(2+1)-dimensional bright
solitons of radial symmetry in a bulk medium \cite{compet_3D}.
We assume that a beam of a
fundamental harmonic (FH) with the frequency $\omega$ is launched into a
medium possessing combined
quadratic [or $\chi^{(2)}$] and cubic [or $\chi^{(3)}$] nonlinear
response under the condition of phase-matched SHG.
The FH beam generates a second harmonic (SH) wave, and such a
two-wave mixing process in a bulk medium is described
by a system of two coupled nonlinear equations,
\begin{equation}
\label{compphys}
\begin{array}{l}
{\displaystyle 2ik_{1}\frac{\partial E_{1}}{\partial z} +
\nabla^{2}_{\perp} E_{1} + \frac{8\pi\omega^{2}}{c^{2}}\chi^{(2)}
E_{2}E_{1}^{*}e^{-i\Delta kz}
+ } \\*[9pt]
{\displaystyle
\frac{12\pi\omega^{2}}{c^{2}}\chi^{(3)}\left(\left| E_{1}\right|^{2}+
\rho\left|E_{2}
\right|^{2}\right) E_{1} = 0,} \\*[9pt]
{\displaystyle 4ik_{1}\frac{\partial E_{2}}{\partial z} +
\nabla^{2}_{\perp}
E_{2}
+ \frac{16\pi\omega^{2}}{c^{2}}\chi^{(2)}E_{1}^{2}e^{i\Delta kz}
+} \\*[9pt]
{\displaystyle
\frac{48\pi\omega^{2}}{c^{2}}\chi^{(3)}\left(\left| E_{2}\right|^{2}+
\rho\left|E_{1}
\right|^{2}\right) E_{2} = 0,} \\*[9pt]
\end{array}
\end{equation}
where $E_{1}$ and $E_{2}$ are the complex amplitude envelopes of FH
($\omega_{1} = \omega$) and SH ($\omega_{2} = 2\omega$) waves,
respectively; $k_{1} = k(\omega)$ and $k_{2} = k(2 \omega)$ are the
corresponding wave numbers; $\Delta k \equiv (2k_{1} - k_{2})$ is the
wave-vector
mismatch between the harmonics, $\rho$ (which we take $\rho = 2$) is
the cross-phase-modulation coefficient, and the coefficients $\chi^{(2)}$ and
$\chi^{(3)}$ are proportional to the second- and third-order susceptibility
tensor elements and they
characterize the combined nonlinear response of an optical medium.
Adopting a similar set of scaling transformations as in Ref.
\cite{compet_3D}, we measure the transverse coordinates in the units of
the
beam radius $R_{0}$, and the propagation coordinate, in the units of the
beam diffraction length $R_{d} = 2k_{1}R_{0}^{2}$. Then, applying the
transformations
\begin{equation}
\label{scaling}
\nonumber
\begin{array}{l}
{\displaystyle E_{1} = \beta
c^{2}(16\pi\omega^{2}\chi^{(2)}R^{2}_{0})^{-1} e^{i\beta z} u(x,y,z),}
\\*[9pt]
{\displaystyle E_{2} = \beta
c^{2}(8\pi\omega^{2}\chi^{(2)}R_{0}^{2})^{-1} e^{i[(2\beta +
\Delta)z]} w(x,y,z),}
\end{array}
\end{equation}
where the parameter
$\beta$ stands for the nonlinearity-induced change of the beam
propagation
constant and $\Delta = 2 k_{1} R_{d}^{2} \Delta k$, we obtain a system
of
normalized equations for $u$ and $w$,
\begin{equation}
\label{normalcomp}
\begin{array}{l}
{\displaystyle i\frac{\partial u}{\partial z} + s \nabla^{2}_{\perp} u-
u +
wu^{\ast} + \chi \Big(\frac{|u|^{2}}{2\sigma} + \rho |w|^{2}\Big) u =
0,} \\*[9pt]
{\displaystyle i\sigma\frac{\partial w}{\partial z} + s
\nabla^{2}_{\perp} w- \alpha w + \frac{u^{2}}{2} + \chi (2\sigma |w|^{2}
+ \rho |u|^2) w = 0,}
\end{array}
\end{equation}
where $\alpha = (2\beta +\Delta)\sigma/\beta$, $s \equiv {\rm
sign} \beta$, and the coordinates are rescaled as follows $z \rightarrow
z/\beta$ and $(x,y) \rightarrow (x,y)/\sqrt{|\beta|}$. For the
spatial beam propagation we take $\sigma = 2$. Parameter
$\chi$ describes a competition between quadratic and cubic
nonlinearities, and it is defined as
\begin{equation}
\label{chi}
\chi = \beta \frac{3
c^{2}}{16\pi\omega_{1}^{2}R_{0}^{2}}\frac{\chi^{(3)}}
{[\chi^{(2)}]^{2}}.
\end{equation}
Stationary solutions are then described by Eqs. (\ref{normalcomp})
with the $z$-derivatives omitted.
To look for radially symmetric solutions carrying
a phase singularity, we use
the polar coordinates $r = \sqrt{x^{2}+y^{2}}$, $\phi =
\tan^{-1}(x/y)$, and make the following substitutions,
\begin{equation}
\label{ansatz}
u(r,\phi) = U(r) e^{im \phi}, \;\;\; w(r,\phi) = W(r) e^{i N m
\phi},
\end{equation}
where $U(r)$ and $W(r)$ are real functions and, for parametric interaction
between the fundamental
and second harmonics, $N = 2$ whereas $m$ is an integer number that
characterises the vortex charge.
Substituting Eqs. (\ref{ansatz}) into Eqs. (\ref{normalcomp}), we
obtain
\begin{equation}
\label{station}
\begin{array}{l}
{\displaystyle \frac{d^{2} U}{d r^{2}} + \frac{1}{r}\frac{d U}{d r} -
\frac{m^{2} U^{2}}{r^{2}} + s \frac{\partial F}{\partial U} =0,}
\\*[9pt]
{\displaystyle \frac{d^{2} W}{d r^{2}} + \frac{1}{r}\frac{d W}{d r} -
\frac{m^{2} N^{2} W^{2}}{r^{2}} + s \frac{\partial F}{\partial W} =0,}
\end{array}
\end{equation}
where the function $F$ has the meaning of an effective potential, and
it is defined as
\begin{equation}
\label{F1}
\begin{array}{l}
{\displaystyle F = F_{1}(U,W) = - \frac{1}{2} U^{2} +
\frac{1}{2} U^{2} W - \frac{\alpha}{2}W^{2} } \\*[9pt]
{\displaystyle + \chi \Big(\frac{1}{16} U^4 + W^4 +
\frac{1}{2} \rho W^2 U^2 \Big).}
\end{array}
\end{equation}
\subsection{Third-Harmonic Generation}
A similar type of two-wave parametric interaction occurs
under the condition of the third-harmonic generation (THG).
Bright and dark solitary waves in a waveguide geometry
(i.e. with one transverse dimension) have been analyzed
in Ref. \cite{thg}. In this case, the parametric interaction occurs
between the fundamental beam ($\omega_{1} = \omega$)
and its third harmonic ($\omega_{3} = 3 \omega$), and the corresponding
physical model of the parametric wave mixing in a bulk can be described by
a system of two
coupled equations,
\begin{equation}
\label{thgphys}
\begin{array}{l}
{\displaystyle 2 ik_{1}\frac{\partial E_{1}}{\partial z} + s \nabla^{2} E_{1} -
} \\*[9pt]
{\displaystyle \chi \left[\left(|E_{1}|^{2} + 2 |E_{3}|^{2}\right) E_{1} +
E_{1}^{\ast 2}E_{3} e^{-i\Delta kz}\right] = 0,} \\*[9pt]
{\displaystyle 2ik_{3}\frac{\partial E_{3}}{\partial z} + s \nabla^{2} E_{3} -
} \\*[9pt]
{\displaystyle 9\chi \left[\left(|E_{3}|^{2} + 2 |E_{1}|^{2}\right) E_{3} +
\frac{1}{3}E_{1}^{3}E_{3} e^{i\Delta kz}\right] = 0,} \\*[9pt]
\end{array}
\end{equation}
where $E_{1}$ and $E_{3}$ are the slowly varying envelopes of the
first and third harmonic fields, respectively, with corresponding
wave numbers $k_{1} = k(\omega)$ and $k_{3} = k(3\omega)$;
$\Delta k = 3k_{1}-k_{3}$ is the wave-vector mismatch between the harmonics
and $\chi = (3\pi \omega^{2}/c^{2})|\chi^{(3)}|$ is the nonlinearity
parameter, which is assumed here to be always positive, whereas
$\chi^{(3)} < 0$.
We follow a normalisation procedure similar to that used above for the
competing nonlinearity model. Again, the transverse coordinate is measured
in units of the beam width $R_{0}$ and the propagation coordinate, in units of
the diffraction length $R_{d} = 2k_{1}R_{0}^{2}$. Using the transformations
of
Ref. \cite{thg}
\begin{equation}
\label{thgtrans}
\begin{array}{l}
{\displaystyle E_{1} = \left(\sqrt{\beta}/3\sqrt{k_{1}R_{0}^{2}\chi}\right)
e^{i\beta z}
u(x,y,z),} \\*[9pt]
{\displaystyle E_{2} = \left(\sqrt{\beta}/\sqrt{k_{1}R_{0}^{2}\chi}\right)
e^{i(3\beta +\Delta)z}w(x,y,z),} \\*[9pt]
\end{array}
\end{equation}
the physical equations (\ref{thgphys}) can be written in the following
normalised form [cf. Eqs. (\ref{normalcomp})],
\begin{equation}
\label{thgnormal}
\begin{array}{l}
{\displaystyle i\frac{\partial u}{\partial z} + s \nabla^{2} u - u -
\frac{s}{3} u^{\ast 2}w - s \left(\frac{|u|^{2}}{9} + 2 |w|^{2}\right) u
= 0,}
\\*[9pt]
{\displaystyle i\sigma\frac{\partial w}{\partial z} + s \nabla^{2} w
- \alpha w - \frac{s}{9}u^{3} - s (9 |w|^{2} + 2 |u|^2) w = 0,}
\end{array}
\end{equation}
where $u$ and $w$ are the normalised amplitudes of the fundamental harmonic
field and its third harmonic,
$\alpha = \sigma (3 \beta + \Delta)/\beta$, $\Delta =
2k_{1}R_{0}^{2}\Delta k$,
$s \equiv {\rm sign} \beta$, the transverse and propagation coordinates have
been rescaled
in terms of the nonlinearity-induced change of the propagation constant $\beta$,
$z \rightarrow z/\beta$ and $(x,y) \rightarrow (x,y)/\sqrt{|\beta|}$, and, for
spatial solitons, we take $\sigma = 3$.
Importantly, everywhere below we consider only
{\em defocusing cubic nonlinearity} searching for vortex-type
solitary waves on a modulationally stable nonvanishing background.
Stationary radially symmetric localized solutions of
Eqs. (\ref{thgnormal}) have the form (\ref{ansatz}) with $N = 3$, and
they
satisfy Eqs. (\ref{station}) with the potential $F$, this time
defined as
\begin{equation}
\label{F2}
\begin{array}{l}
{\displaystyle F = F_{2}(U,W) = - \frac{1}{2} U^{2} -
\frac{\alpha}{2} W^{2} } \\*[9pt]
{\displaystyle - s \Big(\frac{1}{9} U^{3} W +
\frac{1}{36} U^4 + \frac{9}{4} W^4 + W^2 U^2 \Big).}
\end{array}
\end{equation}
Thus, in both the cases, stationary vortex-like structures
are described by the same system of equations (\ref{station})
with different types of the potential $F$.
This observation allows us to perform further analytical
calculations in a rather general form, and, therefore, most of them
are universal and can be applied to other models.
\section{General theory of parametric vortex solitons}
\subsection{Stationary Solutions}
Stationary radially symmetric solutions of Eqs. (\ref{normalcomp}) [Eqs.
(\ref{thgnormal})] are given by Eqs. (\ref{station}) with the potential
function $F$ defined in Eq. (\ref{F1}) [Eq. (\ref{F2})] and $N = 2$
[$N = 3$]. It is important to note that the parametric coupling
between the modes brings {\em several new features} in the vortex
structure and properties. Indeed, as follows from Eqs. (\ref{ansatz})
and (\ref{station}), a vortex with the charge $m$ in the fundamental
mode is always coupled to a vortex of the charge $N m$ ($N = 2,3$) in
the harmonic component. This makes parametric vortices very different
from
all types of vortex solitons analyzed earlier in the systems of two
incoherently coupled NLS equations (see, e.g., Ref. \cite{sheppard1} and
references therein).
\subsection{Analysis of Vortex Asymptotics}
We are interested in the localized solutions supported by a two-component
finite-amplitude background wave. For $r \rightarrow \infty$, the background
amplitudes ($U_0,W_0$) satisfy the coupled algebraic equations:
\begin{equation}
\label{back}
\frac{\partial F}{\partial U} = 0, \hspace{1cm} \frac{\partial
F}{\partial W} = 0,
\end{equation}
which may have one or more nontrivial solutions.
Importantly, due to the self-action effect we always have a special
solution of the form
$(0,W_0)$, that corresponds to an excited harmonic field only.
A vortex soliton is a localized nonlinear
mode that asymptotically approaches the background
$(U_0,W_0)$ for $r \rightarrow \infty$, but its intensity
vanishes for $r \rightarrow 0$ to keep the terms
$\sim (m^{2}/r^{2}) U$ and $\sim (m^{2} N^{2} /r^{2}) W$
in Eqs. (\ref{station}) finite.
This implies that we can find the vortex asymptotics in a rather
general form. For $r \rightarrow 0$, we look for solutions
of Eqs. (\ref{station}) in the form:
\begin{equation}
\label{AABB}
\begin{array}{l}
{\displaystyle U = U_{0} - \frac{A}{r^{2}} - \frac{A_{2}}{r^{4}} +
\ldots,} \\*[9pt]
{\displaystyle W = W_{0} - \frac{B}{r^{2}} - \frac{B_{2}}{r^{4}} +
\ldots,} \\*[9pt]
\end{array}
\end{equation}
where $(U_0,W_0)$ is a solution of Eqs. (\ref{back}) for the background
amplitudes.
Keeping in Eqs. (\ref{station}) only the asymptotic terms up to the order
of $\sim 1/r^{2}$,
we obtain,
\begin{equation}
\label{a_b}
\begin{array}{l}
{\displaystyle s m^{2} U_{0}
+ \Big(\frac{\partial^{2} F}{\partial U^{2}}\Big)_{0} A
+ \Big(\frac{\partial^{2} F}{\partial U \partial W}\Big)_{0} B =0,}
\\*[9pt]
{\displaystyle s N^{2} m^{2} W_{0}
+ \Big(\frac{\partial^{2} F}{\partial W^{2}}\Big)_{0} B
+ \Big(\frac{\partial^{2} F}{\partial W \partial U}\Big)_{0} A =0,}
\end{array}
\end{equation}
where the index $'0'$ stands for the values calculated at $U = U_0$
and $W = W_0$. Solutions of the linear equations (\ref{a_b}) for $A$
and $B$ can be easily found analytically; they define the
asymptotics of the vortex solitons for different values of the vortex
charge $m$ in terms of the background amplitudes $U_{0}$ and $W_{0}$
defined by Eqs. (\ref{back}).
The analysis of the asymptotics gives us important information
about the vortex structure. If both the products $A U_0$ and $B W_0$ are
positive [see Eqs. (\ref{AABB})], the vortex has a standard profile with
the intensity in the core growing monotonically and always lower
than the background intensity. However, if one of these products
is negative, somewhere across the vortex the intensity becomes higher
than the asymptotic background intensity. That implies that the vortex
core is surrounded by a bright ring of higher intensity. We call
such structures {\em `halo-vortices'}. In both the cases mentioned
above, such vortex solitons may exist on a modulationally stable
background, and some examples are given below in Sections \ref{comp_non}
and \ref{third_non}.
\subsection{Vortex Soliton as a Waveguide}
The concept of light guiding light (see e.g., Ref. \cite{phys_today} and
references therein) is based on a simple observation that a spatial optical
soliton
(e.g., vortex)
creates an effective optical waveguide in a nonlinear medium that can
guide a wave of different frequency or polarization. It is
clear that a vortex soliton creates a waveguide of radial symmetry
which can guide a fundamental mode (no nodes) of the other wave. For
the case of two incoherently coupled NLS equations describing two
orthogonal polarizations, the guiding properties of vortex solitons
have been analyzed by Haelterman and Sheppard
\cite{sheppard1}. The first demonstration of an optically written waveguide
based on an optical vortex has been recently reported by Truscott {\em et al}
\cite{experiment}. However, the theory developed in Ref.
\cite{sheppard1}
is not valid for the case of the resonant interactions and parametrically
coupled waves.
Indeed, the parametric interaction forces
the harmonic field to vanish for $r \rightarrow 0$, {\em trapping a singularity}
of the order of $N m$. Therefore, a parametric vortex
{\em cannot guide a fundamental mode}. To analyse the guiding properties
of
parametric vortex solitons, we note that Eqs. (\ref{back})
with the potential $F$ defined by Eqs. (\ref{F1}) and (\ref{F2})
have the solution ($U_{0} = 0$, $W_{0} \neq 0$).
Therefore, we consider a vortex soliton created by
a harmonic field $W$, with a stationary profile described
by the nonlinear equation,
\[
\frac{d^2 W}{d r^{2}} + \frac{1}{r} \frac{d W}{d r} -
\left(\frac{N^2 m^2}{r^2} + s \alpha\right) W - \gamma W^3
= 0,
\]
where $\gamma = - 4 s \chi$, for the model (\ref{normalcomp}),
and $\gamma = 9$, for the model (\ref{thgnormal}).
This equation always has a solution in the form of a vortex
soliton with the charge $N m$ provided $\gamma > 0$ and $s \alpha < 0$.
Now, an eigenvalue equation
for a linear mode guided by the vortex $W(r)$ follows from
the first equation of the system (\ref{station}). Assuming
$U \ll {\rm max}(W)$, we obtain,
\begin{equation}
\label{guide}
\frac{d^2 U}{d r^{2}} + \frac{1}{r} \frac{d U}{d r} -
\left[\frac{m^2}{r^2} + s - s G(r)\right] U = 0,
\end{equation}
where
\begin{equation}
\nonumber
G(r) = \Big(\frac{d^2 F}{d U^{2}}\Big)
\bigg|_{{U=0}, \;}.
\end{equation}
Equation (\ref{guide}) is a standard eigenvalue problem of
the linear waveguide theory, and it can be studied analytically, e.g.
by means of variational methods (see, e.g.,
Ref. \cite{snyder_love} and references therein). To make some analytical
estimates, we present $G(r)$ in an approximate form and obtain
\[
\frac{d^2 U}{d r^{2}} + \frac{1}{r} \frac{d U}{d r} -
\frac{m^2}{r^2} U - E U + \frac{C^2}{(D^2 + r^{2})} U = 0,
\]
where $E$, $C$, and $D$ are, in general, functions of $\alpha$ and
$\gamma$. The parameters are chosen to provide the best
approximation
of the effective potential $G(r)$. Using the standard variational method (or
Ritz optimisation approach) and looking for a bifurcation of a linear mode
taken in a trial form, $f(r) = r \exp{(-\kappa r)}$, we obtain an implicit
expression to determine the mode cutoff
$\alpha$,
\begin{equation}
\nonumber
E = (2 C^2 - 1 - 2 m^2)^3/(36 C^4 D^2),
\end{equation}
which we analyse below for some particular cases.
\subsection{Modulational Instability}
Stability of the stationary vortex solitons described by the system
(\ref{station}) is an important issue. In general, the stability analysis of
vortices in nonlinear models is a complicated and, generally speaking,
unsolved problem. Instability can develop due to the presence of unstable
eigenmodes
localized near the vortex core
and, in the one-dimensional case, this type of instability of
{\em dark solitons} leads to the soliton motion, i.e. it is {\em a drift
instability} (see, e.g., Ref. \cite{dark_review} and references
therein). Since moving vortices with nonzero minimum intensity (similar
to grey solitons) do not exist, similar drift instability is not
observed for vortices. The main instability which is usually associated
with a vortex
soliton originates from the instability of the nonlocalized background
field.
The analysis of modulational instability of the background field can be
carried out in a
general form. First, we write Eqs. (\ref{normalcomp}) and
Eqs. (\ref{thgnormal}) in the form
\begin{equation}
\label{nonstat}
\begin{array}{l}
{\displaystyle i \frac{d u}{d z} + s \nabla^{2} u +
\frac{\partial {\cal F}}{\partial u^{*}} =0,} \\*[9pt]
{\displaystyle i \sigma \frac{d w}{d z} + s \nabla^{2} w +
\frac{\partial {\cal F}}{\partial w^{*}} =0,}
\end{array}
\end{equation}
with ${\cal F}$ defined as
\begin{equation}
\label{complexF}
\begin{array}{l}
{\displaystyle {\cal F} \rightarrow {\cal F}_{1} = - |u|^{2} +
\frac{1}{2} (u^{2} w^{*} + u^{*2} w) - \alpha |w|^{2} + } \\*[9pt]
{\displaystyle \chi \Big(\frac{1}{8} |u|^4 + 2 |w|^4 +
\rho |w|^2 |u|^2 \Big),} \\*[9pt]
\end{array}
\end{equation}
for the model of competing nonlinearities, or
\begin{equation}
\begin{array}{l}
{\displaystyle {\cal F} \rightarrow {\cal F}_{2} = - |u|^{2} +
\frac{s}{9} (u^{3} w^{*} + u^{*3} w) - \alpha |w|^{2} + } \\*[9pt]
{\displaystyle + s \left(\frac{1}{18} |u|^4 + \frac{9}{2} |w|^4
+ 2 |w|^2 |u|^2\right),}
\end{array}
\end{equation}
for the model of the third-harmonic generation.
We look for stability
of the background wave solution $(U_{0},W_{0})$ defined by
Eqs. (\ref{back}), and linearize Eqs. (\ref{nonstat}) around
this stationary solution substituting:
\begin{equation}
\label{modulation}
\begin{array}{l}
{\displaystyle u = U_{0} + a e^{i {\bf \vec{k} \cdot \vec{r}} + i \omega z}
+
b e^{-i {\bf \vec{k} \cdot \vec{r}} - i \omega z},} \\*[9pt]
{\displaystyle w = W_{0} + c e^{i {\bf \vec{k} \cdot \vec{r}} + i \omega z}
+
d e^{-i {\bf \vec{k} \cdot \vec{r}} - i \omega z}.}
\end{array}
\end{equation}
As a result, we obtain a system of linear equations for $a$, $b^{*}$,
$c$, and $d^{*}$ leading to the characteristic equation:
\begin{equation}
\label{modulat_equat}
\nonumber
\begin{vmatrix}
A_{u^{*},u^{*}}-\Omega & A_{u^{*},u} & A_{u^{*},w^{*}} & A_{u^{*},w} \\
A_{u,u^{*}} & A_{u,u} + \Omega & A_{u,w^{*}} & A_{u,w} \\
A_{w^{*},u^{*}} & A_{w^{*},u} & A_{w^{*},w^{*}}-\Omega & A_{w^{*},w} \\
A_{w,u^{*}} & A_{w,u} & A_{w,w^{*}} & A_{w,w} + \Omega \end{vmatrix}
= 0.
\end{equation}
Here $A_{n,m} \equiv (\partial^{2}{\cal F}/ \partial n \; \partial
m)|_{(u = U_{0}, w = W_{0})}$, where $m,n = (u, u^{*}, w,
w^{*})$ ($n \neq m$), and
for the $\alpha = \beta$ case
\[
A_{n,n} \equiv \left(\frac{\partial^2{\cal F}}{\partial n
^2}\right)\big|_{(u = U_{0}, w = W_{0})} - |\vec{\bf k}|^{2}.
\]
Solving the characteristic equation with respect to $\Omega$, we
conclude the modulational instability analysis: purely real $\Omega$
solutions for all positive $|\vec{\bf k}|^{2}$ (with other parameters fixed)
indicate a modulationally stable background for this fixed set of
the parameters. Below we present the results of the modulational
instability analysis for two cases of the parametric two-wave
interaction.
\section{Competing nonlinearities}
\label{comp_non}
Analysis of modulational instability for the system (\ref{normalcomp})
has been briefly presented in Ref. \cite{tristram}. Below we repeat the
main steps of that analysis for the completeness of this paper.
Solutions of Eqs. (\ref{normalcomp}) for background waves can be found
by solving the coupled algebraic equations:
\[
\begin{array}{l}
{\displaystyle 12 \chi W_{0}^{3}+ 12 W_{0}^{2}+W_{0}(\alpha-8+ 2\chi^{-1}) =
2\chi^{-1},} \\*[9pt]
{\displaystyle U_{0}^{2} = 4(1-W_{0})\chi^{-1} - 8 W_{0}^{2},}
\end{array}
\]
for real $U_{0}$ and $W_{0}$. There exist up to {\em three such solutions}
with
both amplitudes $U_{0}$ and $W_{0}$ being nonzero. Performing the
analysis
of modulational instability for each of the three solutions at $s=\pm
1$, we find that there exists only {\em one modulationally stable}
mode. The~parameter domains where such a solution exists are
presented in Figs.~\ref{MI_comp}(a,b). Note, that $\rm sign (s \chi) =
\rm sign \chi^{(3)}$ and thus modulationally stable solutions exist only
for $\chi^{(3)} < 0$.
Importantly, the amplitude of the modulationally stable background
diverge
in the limit $\chi \rightarrow 0$ so that the stable background solution
exists exclusively due to {\em mutual action of quadratic and cubic
nonlinearities}. Other nonlinear modes are modulationally unstable in the
whole domain of their existence and they are not presented in
Figs.~\ref{MI_comp}(a,b). Modulational stability in the limit of large
negative $\chi^{(3)}$ [e.g., for $s= -1$ and $\chi >0$, see Fig.
\ref{MI_comp}(a)], is not surprising because stable dark solitons are
known to exist in a defocusing Kerr medium without quadratic
nonlinearity. Here, we are interested in the case when the effective
nonlinearity is predominantly quadratic, i.e. $|\chi U_{0}| \sim |\chi
W_{0}|~\ll~1$. We found that this condition can only be satisfied for
$s = +1$ where modulationally stable background waves of moderate
amplitudes exist for relatively small values of negative $\chi$ [see Fig.
1(b)].
Using the numerical relaxation technique, we have found that a
continuous family of two-component vortex solitons exists in the whole region
of the existence of modulationally stable background waves shown in
Figs. \ref{MI_comp}(a, b). Figures \ref{example}(a) and \ref{example}(b)
present an example of such a vortex soliton.
Analysis of the vortex asymptotics demonstrates that halo-vortices can
exist in Eqs. (\ref{normalcomp}) only if $s = -1$ and $\chi > 0$, in
the narrow domain
shown in Fig. \ref{halocomp}. Both terms with $B$ and
$B_{2}$ factors in the asymptotic expansion (\ref{AABB}) contribute to
the formation of the halo (i.e. both $B W_0$ and $B_{2} W_0$ products
are negative).
{\em Ring-vortex solitons} can also exist in the model
(\ref{normalcomp}), see Figs. 4(a,b) and Figs. 5(a,b). Variational
analysis allows us to find
an approximate expression for the bifurcation curve where such solutions
appear,
$(2-\alpha) =\sqrt{\alpha/\chi}$. As $\alpha$ increases,
the maximum of the bright-ring amplitude approaches the value $U_0$ of
two-wave modulationally stable parametric plane waves.
At the values of $r$ where $U$ approaches $U_0$, the second
component,
$W$, also approaches the corresponding plane wave amplitude $W_0$ [see
Fig. \ref{vb_examp_comp}(b)].
Such ring-vortex solitons can be unstable due to
modulational instability of the background wave $U_0 = 0$,
$W_0^2 = \alpha/4 \chi$. For example, for $s = -1$, $\alpha >
0$, modulational stability is defined by the condition
$(\alpha - 2) > \sqrt{\alpha/\chi}$. The regions of the existence
of modulationally stable one-component plane waves and ring-vortex
solitons of Eqs. (\ref{normalcomp}) are presented in Fig. \ref{major}.
Existence of two-component stable vortex solitons composed of
parametrically coupled fields suggests that such vortices can be
excited
in the process of the harmonic generation. In Fig. 7 we present the
numerical simulation results supporting this idea. We launch a mode of
the fundamental frequency without a seeded second harmonic assuming the
condition of phase-matching. The vortex soliton dynamics is simulated
using a split-step beam propagation method (BPM). To solve the problem
of the
vortex phase geometry, we simulate a system of four vortices, arranged
such that horizontally and vertically adjacent vortices are opposite in
charge. Lines of equal phase are chosen to correspond to the lines of
the force of an equivalent system of electrostatic point charges. With
the periodic
boundary conditions imposed by BPM, this configuration means that, in fact,
an infinite array of vortices is simulated. Figure 7 shows the
vortex generation by an input fundamental mode with single-charged
vortices. Due to phase matching with the second harmonic, we observe a
generation of double-charge vortices in the harmonic field (see the
plots at $z=1$) and then periodic oscillations of the two-component
background and the vortex profiles near a stationary state corresponding
to a lattice of two-component vortex solitons (see the plots at $z=10$
as an example of
such dynamics).
\section{Third-harmonic generation}
\label{third_non}
Vortex solitons of Eqs. (\ref{thgnormal}) have fewer parameters in
comparison with the parametric vortices described by Eqs. (\ref{normalcomp}),
and thus they can be analysed much more easily numerically.
These vortex solitons are found in the whole region
of the existence of modulationally stable plane waves,
i.e. for $\alpha < \alpha_{\rm th} \approx 14.509$.
Examples of such vortex solitons
are shown in Figs. \ref{example_thg}(a, b).
A third-harmonic component of the vortex solitons
has a nonmonotonic tail for
$10.85 < \alpha < 14.509$, however it can only be called
a halo vortex for the interval $11.26 < \alpha < 14.509$, where
the absolute value of a local extremum in the structure of the
vortex tail is greater than the corresponding plane wave
background value $W_0$. The halo is becoming more
pronounced as $\alpha \rightarrow 14.509$.
An example of a halo-vortex soliton of Eqs. (\ref{thgnormal})
is shown in Fig. \ref{halo_thg}.
Ring-vortex solitons have also been found
for the model (\ref{thgnormal}), see Fig. 10.
In this case, a variational analysis allows us to find
an approximate analytical result for the bifurcation point ($s = -1$)
where a ring-like mode is guided by the vortex:
\begin{equation}
\label{explicit}
\nonumber
\alpha_{\rm bif} = \frac{(\gamma/\delta)}{\{1-\frac{2}{9} \left[1-
\frac{\gamma}{2 \delta N^2 m^2} (1 + 2 m^2) \right]^3\}}.
\end{equation}
For $\gamma = 9$, $\delta = 2$, $N = 3$, and $m = 1$ this
gives $\alpha_{\rm bif} \approx 4.52$, which agrees well with
numerical data. We also find that, in general, coupled ring-vortex
solitons exist for $\alpha > 4.5$, and for each such value of $\alpha$
(except $\alpha = \alpha_{\rm bif}$) there exist {\em two different types
of
ring-vortex solitons} (see Fig. \ref{maximum}). As the parameter $\alpha$
decreases,
the maximum of the bright ring in the fundamental mode approaches
the value of $U_0$ of the two-wave modulationally stable background field.
Again, as it has been observed for the model of competing
nonlinearities, at values of $r$ where $U$ approaches $U_0$, the vortex
component $W$ deforms significantly approaching the corresponding plane-wave
amplitude $W_0$. We note that {\em all these ring-vortex solitons are
modulationally stable}, because, in the framework of Eqs. (\ref{thgnormal}),
modulational instability does not occur for one-component plane wave solutions.
\section{Concluding Remarks}
We have analyzed two-component vortex solitons supported by parametric wave
mixing in a nonlinear optical medium. We have considered two classes of
such vortex solitons. In the first case, we have studied the existence,
structure, and stability of vortex solitons supported by
phase-matched interaction between the
fundamental and second-harmonic waves in a quadratic medium, and the effect
of the next-order cubic nonlinearity has been taken into account for
suppressing modulational instability of the supporting plane-wave
background. In the second case, we have considered how the vortex parameters,
structure, and stability are modified due to the process of
third-harmonic generation when the phase-matched wave interaction
generates a corresponding multi-charge vortex component in a harmonic
field. In particular, we have predicted the so-called
`halo-vortex' consisting of a two-wave vortex core surrounded by a
bright
ring on a non-vanishing background. Additionally, we have analyzed the
waveguiding properties of a vortex soliton in the case when it guides a
harmonic field due to a phase-matched parametric interaction.
A rigorous analysis of the stability of these parametric vortex solitons
is
still an open problem, as well as the effect of walk-off on the vortex
existence and stability.
As for experimental verifications of the vortex solitons described above,
we would like to mention that, at least in the low-intensity regime,
parametric vortices have already been observed in nonlinear optics. A
possibility of SHG by a beam with a vortex was first mentioned and
experimentally verified in Ref. \cite{soskin}, where a vortex of the
topological charge $m=2$ was found in the second-harmonic wave when the
fundamental wave contained a vortex of the topological charge $m=1$. The
similar results on SHG have been presented by Dholakia {\em et al} in
Ref. \cite{dho}, whereas more complicated processes of sum-frequency
mixing with beams carrying phase singularities were reported by
Ber\v{z}anskis {\em et al} \cite{berz,lithuania}. It is worth noticing that
in all of
those observations the different harmonics experienced noticeable walk-off
that makes the stationary structures difficult to observe, also introducing
novel features in the vortex dynamics. In particular, for a collinear
type I phase-matched SHG with an input beam carrying a single-charge vortex,
Matijo\v{s}ius {\em et al} \cite{matij} observed two intensity zeroes in
the second-harmonic field with the separation of two SHG vortices due to
walk-off. Thus, we can expect that stationary two-component vortex solitons
discussed above can be observed in typical upconversion experiments when a
high-intensity beam undergoes frequency doubling simultaneously with the
creation of a phase singularity produced by a phase mask at the input,
similar to the experiments mentioned above which were performed at
moderate powers. {\em Stability} of those vortex solitons requires small
(or zero) walk-off and a small defocusing Kerr nonlinearity of an optical
material at both (or at least fundamental wave) frequencies.
Additionally, we would like to mention that the parametrically coupled
equations
of competing nonlinearities, similar to Eqs. (\ref{normalcomp}) analysed
above, have been recently introduced by Heinzen {\em et al} \cite{heinzen}
to describe the dynamics of {\em coupled atomic and molecular Bose-Einstein
condensates}, leading to a kind of ``super-chemistry'' in which the
formation of molecules is a controlled parametric quantum process. In spite
of the fact that both atomic and molecular condensates should be considered
in a trapping external potential \cite{BEC}, many of the features of the
coupled stationary states, including all the types of the vortex states
introduced above, are expected to exist in the model of atom-molecular
condensates as well, providing a much broader view of the phenomenology of
parametric vortex solitons.
At last, we expect that the concept of the two-component parametric vortices,
generated and supported by the third-harmonic generation process, can be
important in the so-called {\em third-harmonic microscopy} (see, e.g.,
\cite{THGspectra}) where an image is rendered using a series of
cross-sectional images produced by third-harmonic generation within the
specimen. Vortices can then be formed due to the development of caustics
\cite{nye_book} in the reflected harmonic field, indicating the regions of
highly concentrated inhomogeneities. This technique is based on the fact that
the nonlinear susceptibility of solid media vary over many orders of
magnitude, compared with linear refractive index changes that vary in a
relatively small range.
\section*{Acknowledgments}
Yuri Kivshar thanks O. Bang, P. DiTrapani, B. Ivanov, L. Pismen, A.
Piskarskas, Y. Silberberg and V. Smilgevi\v{c}ius
for useful discussions of the properties of multi-component and parametric
vortices and harmonic generation.
The work has been supported by the Australian Photonics
Cooperative Research Centre and the Australian Research Council. A
brief
summary of the
results has been presented at the OSA Annual Meeting in
Baltimore, USA
(October 4-9, 1998).
|
\section{}
With the development of the accelerator and radioactive beam technique,
a lot of new experiments using the radioactive beams with large
neutron or proton excess become possible. The degree of freedom
of isospin of nuclear matter is becoming important for research.
It offers the possibility to study the properties of nuclear matter
in the range from symmetrical nuclear matter to pure neutron matter.
Some theoretical investigations to the equation of state, chemical
and mechanical instabilites as well as liquid-gas phase transition for
isospin asymmetrical nuclear matter were performed already.
In addition, the isospin dependent nucleon - nucleon cross section is
also an important subject due to its significant effects on the
dynamical process of heavy ion reactions induced by radioactive beams.
Some new phenomena stemmed from the isospin have been revealed. For
examples, the isospin dependences of the preequilibrium nucleon
emission, the nuclear stopping, the nuclear collective flow,
total reaction cross section, radii of neutron-rich nuclei
and subthreshold pion production have been studied by
several groups \cite{Bali98,Macpl,Xucpl}.
However, more experimental and theoretical
studies are still needed for understanding the isospin physics. As a
trial, the isospin effects were investigated with the lattice gas
model in this letter.
The lattice gas model of Lee and Yang \cite{Yang52}, in which the
grancanonical partition function of a gas with one type of
atoms is mapped into the canonical ensemble of an Ising model for
spin 1/2 particles, has uccessfully described the liquid-gas
phase transition for atomic system. The same model has already
been applied to nuclear physics for isospin symmetrical systems
in the grancanonical ensemble \cite{Biro86} with an approximate
sampling \cite{Mull97} of the canonical ensemble
\cite{Jpan95,Jpan96,Camp97,Gulm98},
and also for isospin asymmetrical nuclear matter in the mean field
approximation \cite{Sray97}. In this model, $A$ nucleons with an
occupation number $s$ which is defined as $s$ = 1 (-1) for a proton
(neutron) or $s$ = 0 for a vacancy, are placed in the $L$ sites of
lattice. Nucleons in the nearest neighbouring sites have
interaction with an energy $\epsilon_{s_i s_j}$. The hamiltonian
is written by
\begin{equation}
E = \sum_{i=1}^{A} \frac{P_i^2}{2m} - \sum_{i < j} \epsilon_{s_i s_j}s_i s_j
\end{equation}
The interaction constant $\epsilon_{s_i s_j}$ is fixed to reproduce
the binding energy of the nuclei, $\epsilon_{nn,pp}$ =
$\epsilon_{-1-1,11}$ = 0. MeV, $\epsilon_{pn,np}$ =
$\epsilon_{1-1,-11}$ = -5.33 MeV. We use a three-dimension cubic
lattice L with a size l, a number of nucleons $A = N + Z$ and a
temperature T. The freeze-out density of disassembling system is
$\rho_f$ = $\frac{A}{L} \rho_0$ where $\rho_0$ is the normal nucleon
density. The disassembly of the system is to be calulated
at $\rho_f$, beyond which nucleons are too far apart to interact.
$A$ nucleons are put in $L$ cubes by Monte Carlo sampling using the
Metropolis algorithm \cite{Metr53}. Once the nucleons have been placed, their
momentum is generated by a Monte Carlo sampling of Maxwell Boltzmann
distribution. Various observables can be calculated in a straightforward
fashion.
One of the basic measurable quantities is the distribution of
fragment mass. In this lattice gas model, two neighboring nucleons
are viewed to be in the same fragment if their relative kinetic energy
is insufficient to overcome the attractive bond:
$P_r^2/2\mu + \epsilon_{s_i s_j} < 0 $. This method is similiar to the so-called
Coniglio-Klein's prescription \cite{Coni80}. In this letter, we use
the above condition to construct the fragments and their distributions.
We chose several isotopes of Xe as examples of the study of isospin
effects in the lattice gas model. Their isospin parameter
($\frac{N-Z}{A}$) is 0.11, 0.16, 0.21 and 0.26 for $^{122}$Xe,
$^{129}$Xe, $^{137}$Xe
and $^{146}$Xe, respectively. The freeze-out density $\rho_f$
has been chosen to be close to 0.39 $\rho_0$, extracted from the analysis of
Ar + Sc \cite{Jpan95} and
$^{35}$Cl + Au and $^{70}$Ge + Ti \cite{Beau96} with the
same model. There is also good support from experiment that the value
of $\rho_f$ is significantly below 0.5$\rho_0$ \cite{Agos96}.
We use the 343 cubic lattice with size of 7 which results that
the freeze-out density $\rho_f$/$\rho_0$ of $^{122,129,137,146}$Xe is
0.36, 0.38, 0.40, and 0.43, respectively. The other input
parameter is the temperature, we perform the calculation from 4 to 7
MeV. For each isotope 1000 events are accumulated at each temperature.
Fig.1 shows the mass distribution of fragments at T = 4, 5, 6 and 7 MeV
for $^{129}$Xe.
Clearly the disassembling mechanism evolves with the nuclear
temperature. A few light particles and fragments are emitted and the
big residue reserves at T = 4 MeV which indicates typical
evaporation mechanism.
With the increasing temperature, the shoulder of mass distribution
occurs due to the competition between the fragmentation and the evaporation.
This shoulder disappears and the mass distribution becomes power law shape
at T = 6 MeV,
corresponding to the multifragmentation region. When the temperature becomes
much higher, the mass distribution becomes
steeper indicating that the disassembling process becomes more violent.
The power law fit, Y(A) $\propto$ $A^{-\tau}$, for these mass distribution
can be introduced here. It
has already been observed that a minimum of power law parameter
$\tau_{min}$ exists for most systems if the critical behavior
takes place. The lines in Fig.1 represent the power law fit.
Fig.2 displays the several physical quantities
as a function of temperature for Xe nuclei with the different isospin.
The minimums of $\tau$ parameters in Fig.2a locate
closely at 5.5 MeV for all the systems, which illustrates
its minor dependence on the isospin. In other words, there is a universal
mass distribution regardless of the size of disassembling source when
the critical phenomenon takes place. However, the $\tau$ parameters show
different values outside the
critical region for nuclei with different isospin, eg., $\tau$
decreases with isospin when T $>$ 5.5 MeV (multifragmentation
region). Similiarly, the mean multiplicity of
intermediate mass fragment $N_{IMF}$, defined as the number of
fragments with 3$\leq$ Z $\leq$16 here, has analogous characters in
Fig.2b \cite{Maprc,Zhengcpl}. There are the maximums for Xe systems near to 5.5 MeV. When the
temperature becomes higher, the larger the source, the higher the $N_{IMF}$.
Fig.2c plots the information entropy H as a function of
temperature for Xe isotopes. The information entropy was
introduced by Shannon in information theory first\cite{Denb85}.
It is defined as
\begin{equation}
H = -\sum_{i} {p_i ln(p_i)} ,
\end{equation}
where $p_i$ is the probability having "i" produced particles in each event,
the sum is taken over all multiplicities of products from the
disassembling system. H reflects the capacity of the information or
the extent of disorder. We introduce this entropy into the nuclear
disassembly here. As expected, the entropy H reveals the peak close
to 5.5 MeV for all isotopes. These peaks indicate that the opening
of the phase space and the number of the states at the critical
point is the largest. In the other words, the
systems at the critical point have the largest
fluctuation which leads to the largest disorder.
After the critical point, the entropy H increases with the
isospin and/or the source size.
In Fig.2d we give the temperature dependences of Campi's second moment
of the mass distribution \cite{Camp88}, which is defined as
\begin{equation}
S_2 = \frac{ \sum_{i \neq Amax} {A_i^2 \times n_i(A_i)}}{A} ,
\end{equation}
where
$n_i(A_i)$ is the number of clusters with $A_i$ nucleons and the
sum excludes the largest cluster $A_{max}$, $A$ is the mass of the system.
At the percolation
point $S_2$ diverges in an infinite system and is at maximum in
a finite system. Fig.2d gives the maximums of $S_2$ around
5.5 MeV for different isotopes, respectively. Again, the critical behavior
occurs in the same temperature as other observables.
In conclusion, the critical behaviors are explored for Xe isotopes
in the lattice gas model, namely the minimum of power-law parameter $\tau$
of mass distribution, the rise and fall of mean multiplicity of IMF,
information entropy and Campi's second moment.
In a narrow region of
critical point, the features of the above quantities show no
dependence on the isospin of the disassembling system. It reflects
that a universal law exists for the same element in the critical
point. On the contrary, these quantities have isospin dependence at
the same temperature outside the critical
region. Noting that the information entropy is introduced into such an
analysis for the first time, and it seems to be useful for the
searching of critical phenomena in nuclear physics.
It will be interesting and meaningful to have some experiments
to compare our conclusion in the near future.
We would like to thank Dr. Pan Jicai and Subal Das Gupta for helps and fruitful
discussions. Ma Yu-gang would like
to thank NSFC for the receipt of National Distinguished Young
Investigator Fund.
|
\section{Introduction}
The presence of
abundance gradients in spiral galaxies is well established from
observations of their H\,II regions (e.g. Pagel \& Edmunds \cite{pe81}).
From emission line strengths, abundances
of He, N, O, Ne \& S can be measured, and typically O/H
decreases with galactocentric distance (e.g. Shaver et al. \cite{shav};
Walsh \& Roy \cite{wr89}). The
situation with regard to early--type galaxies is however more
complex. There is little interstellar medium
and hence no well--dispersed H\,II regions; resort must be made
to the line of sight integrated properties of the starlight.
Both sets of abundance indicators present their own
advantages and disadvantages: emission line regions can be
subject to local enrichments making them atypical of the
general interstellar medium; stellar indicators, whilst
providing line of sight abundances, are subject to
contributions from stars all over the HR diagram, having a
range in metallicity and age. In addition even a
contemporaneous stellar population can appear to
possess an abundance spread of light elements such as C, N
and O as found for globular cluster giants (Kraft \cite{kraf}).
Stellar abundances are measured from the strength of absorption
lines, such as Mg~I and molecular bands
such as CN and TiO (Gorgas et al. \cite{gorg}, Davies et al.
\cite{dav} or colours (Visvanathan \& Sandage \cite{visa}
and Peletier et al. \cite{pel}).
Synthesis techniques (Tinsley \cite{tins}, Faber \cite{fab} and
O'Connell \cite{oco}) are required to provide
abundance determinations. Abundances of various elements, such as
Fe, Mg, Ca, Na etc can be obtained (see e.g. Vazdekis et al.
\cite{vaz}). Their accuracy is
limited by the fact that only strong lines can be used, because of
the considerable velocity broadening in the galaxies.
Stellar colour and absorption line variations across
the faces of early--type galaxies are generally observed and
interpreted by an outwardly--decreasing metallicity (eg.
Davies et al. \cite{dav}, Bica \cite{bic}) although
significant gradients in age are also sometimes claimed (Worthey
et al. \cite{wor}, Trager et al. \cite{trag}).
Globular clusters cannot be used to determine the stellar population
distribution of ellipticals, since their colours and line strengths
are not representative of the main stellar component. For
example Bridges et al. (\cite{brig}) found that the average metallicity
for the globular clusters in M~104 is [Fe/H]=$-$0.7, about a
factor 6 - 10 lower than for the stars in the central region
(Vazdekis et al. \cite{vaz}).
Studies pioneeered by Jacoby, Ciardullo, Ford and
co-workers have shown that in early-type systems, planetary
nebulae (PN) can be useful both as distance indicators
(e.g. Jacoby \& Ciardullo \cite{jac92}) and as
probes of the galaxy kinematics (e.g. Hui et al. \cite{huic}).
PN, easily detected from their
strong emission line spectra, can also be used as abundance
tracers in a way comparable to H\,II regions, since some
elements, O in particular, but also Ne and S, are not generally
affected by the nucleosynthetic processing in most PN
progenitor central stars. Measuring abundances of PN
provides a unique way to determine the abundance spread of the
old stellar population in distant ellipticals, as predicted from star
formation theories (Arimoto \& Yoshii \cite{ari})
and population synthesis (Bica \cite{bic}). Direct measurement of
the stellar abundance spread from the ground is limited to local group
galaxies such as M31 and M32 where single stars can feasibly be
resolved.
NGC~5128 (Centaurus-A) is the closest giant elliptical
($\sim$3.5 Mpc, Hui et al. \cite{huia}; morphological type S0,pec
Sandage \& Tammann \cite{sata}), and has a
projected size on the sky of over 1$^{\circ}$, making it ideal
for spatial studies.
HII regions, star formation and interstellar matter are found in the
inner regions, possibly arising from a merger with a more metal poor
galaxy (see for example Quillen et al. \cite{quil}), so cannot
provide reliable abundance
diagnostics for the old stellar system. 784 PN have been
detected in an area 20$\times$10kpc (Hui et al. \cite{huib})
and the luminosity function within the brightest 1.5
magnitudes was used to determine the distance. The radial velocities
of 433 of these PN have been measured (from the brightest line,
that of [O~III]5007\AA) in order to study the dynamics of the halo
(Hui et al. \cite{huic}). Using the PN as test particles, the
gravitational potential of the galaxy can be studied: Cen-A
was found to have a tri-axial potential with the galaxy minor axis offset
from the rotation axis by 40$^\circ$. Measuring the rotation
and velocity dispersion of the PN system, it was shown that M/L increases
with radius suggesting that dark matter was present in the galaxy
halo (Hui et al. \cite{huic}). This is also in agreement with globular cluster
velocity measurements (Hui et al. \cite{huic} for NGC~5128 and
Bridges et al. \cite{brig} for M~104)
So far no spectroscopy of these PN has been obtained and their
relation to the PN population in the Milky Way, which may have
a different metallicity and certainly a different star formation
history to Cen-A,
is not known. In particular it is not known if the brightest PN observed
are exceptional (perhaps of Type I) and what is the effect of line of sight
extinction on the PN luminosity function. In Cen-A, it is known that there
is a jet and extended emission line regions along the jet (e.g.
Morganti et al. \cite{morg}) as well as H\,II regions in the vicinity of the
dust lane, all of
which could be included in the PN census at some level. Spectroscopy can
therefore be seen as important both in terms of the PN population and its
use as a distance indicator and in terms of probing the chemical history of
the host galaxy. The long-term goal is to study abundance gradients and the
spread in abundances at a given radius, from
large numbers of PN observed with multi-object techniques.
In this paper the results of deep spectroscopic integrations with a 3.6m
telescope of a few selected
PN in NGC~5128 are presented. Section 2 summarises the observations and
Sect. 3 details the data reduction and presents the results. In Sect. 4
we discuss the data and the relevance of these observations for abundance
determination, as well as the prospects for future such observations with
8-10m class telescopes.
\section{Observations}
Long-slit spectra centred on three of the brightest PN in the catalogue
of Hui et al. (\cite{huib}) were observed with EFOSC1 (Buzzoni et
al. \cite{buzz})
on the ESO 3.6m telescope. Table 1 lists the three PN with the ID numbers,
J2000 coordinates and 5007\AA\ magnitudes (from Hui et al. \cite{huib})
and Fig. 1 shows the positions of the
PN\#5601 is the brightest PN observed in NGC~5128 (by 0.1mag.).
The targets were chosen to cover a range in galactocentric radius and
where possible the slit orientation was chosen so that at least one
other PN from the
Hui et al. (\cite{huib}) catalogue would lie on the slit. For the
slit length centred on PN\#5601 two other PN were included in the slit
(listed in Table 1);
for PN\#4001 one of the desired PN (4013) was however missed
by the slit. No target very close to the high surface brightness central
region of the galaxy was chosen in order to minimize the contribution of
galaxy continuum to the PN spectra. In total, spectra of five
PN in Cen-A were detected and
are all listed in Table 1 together with their radial distance from
the galaxy nucleus.
Figure 1 shows the positions of these five PN against
the POSS image, with the 1.425 radio contours overlayed
in black (from Condon et al. \cite{cond}).
\begin{figure*}
\resizebox{\hsize}{!}{\includegraphics{H1150f1.ps}}
\caption{The POSS image of NGC~5128 is shown with the positions of the
five observed planetary nebulae indicated by white
crosses; the target designations are from Hui et al. (\cite{huib}).
The contours are from the 1.425 GHz radio continuum map of Condon et al.
(\cite{cond}). J2000 coordinates are shown.}
\end{figure*}
For each target an [O~III] narrow
band filter image (ESO\#686, $\lambda_{CEN}$ 5013\AA, $\Delta \lambda$
56\AA) and an off-emission filter (ESO\#714, $\lambda_{CEN}$ 5483\AA,
$\Delta \lambda$ 182\AA, Str\"{o}mgren $y$) image were obtained with
EFOSC1 in order to acquire the PN on the slit. Exposure times were
usually 5min for the [O~III] image and 2min for the continuum band.
Visual `blinking' of the two images confirmed the PN and the object was
then centred in the slit. At least one more [O~III] image was obtained
during each sequence of exposures on the same target to ensure that the
source was well-centred in the slit.
\begin{table*}
\caption[]{Planetary Nebulae Observed in NGC~5128}
\begin{flushleft}
\begin{tabular}{llcrl}
Central & Target & RA ~(2000)~ Dec & Radial$^{\dag}$ & m$_{5007}$ \\
PN ID & Name$^{\ast}$ & $h$ ~~$m$ ~~$s$ ~~~~~$^\circ$ ~~$'$ ~~$''$ &
dist ($''$) & (mag.)$^{\ddag}$ \\
\hline
5601 & 5601 & 13 25 53.52 $-$43 08 54.7 & 547~~ & 23.51 \\
& 5621 & 13 25 53.81 $-$43 08 37.5 & 535~~ & 25.64 \\
& 5619 & 13 25 55.31 $-$43 07 12.4 & 474~~ & 25.70 \\
& & & \\
1902 & 1902 & 13 26 40.78 $-$42 49 33.6 & 1061~~ & 24.01 \\
4001 & 4001 & 13 25 41.12 $-$42 54 35.0 & 418~~ & 23.89 \\
\end{tabular}
\end{flushleft}
$^{\ast}$ PN designation from Hui et al. (\cite{huib})\\
$^{\dag}$ Measured from the J2000 position of the nuclear radio source
(13$^{h}$ 25$^{m}$ 27.7$^{s}$ $-$43$^\circ$ 01$'$ 06$''$,
Wade et al. \cite{wad}). For reference 1$''$=17pc \\
$^{\ddag}$ $ m_{5007} = -2.5log F_{5007} - 13.74 $,
where $F_{5007}$ is the [O~III]5007\AA\ line flux in ergs cm$^{-2}$ s$^{-1}$
(Ciardullo et al. \cite{cia89}).
\end{table*}
The slit position centred on PN\#1902 also included an emission
line filament (approximate J2000 position 13$^{h}$ 26$^{m}$ 49$^{s}$
$-$42$^\circ$ 49$'$ 25$''$). This is part of the system of filaments
associated with the jet in NGC~5128 some of which have been
spectroscopically studied by Morganti et al. (\cite{morg}).
The detector of EFOSC1 was a Tek 512$\times$512 thinned CCD (ESO\#26,
TK512CB) with 27$\mu$m pixels, which project to 0.61$''$. Given the
seeing encountered of 1-1.5$''$, slit widths of 1.5$''$ were employed
to ensure a good balance between receiving the majority of the flux
from the point-like PN without a severe penalty of sky and galaxy
background light, and of ensuring
optimal sampling at the detector. The actual slit width employed
is listed in Table 2. All the spectra were obtained with the B300 grism,
which covers the wavelength range 3640 to 6860\AA\ at a dispersion of
6.3A/pixel; the resulting resolution of the spectra was about 14\AA.
Bias frames, dome flat fields and spectroscopic sky flats were obtained
to correct the CCD pixel response; neon and argon lamps spectra for
correction of distortions and wavelength calibration; and broad slit
(5$''$) spectra of spectrophotometric standard stars EG~54 (Oke \cite{oke}),
EG~274 (Hamuy et al. \cite{ham}) and LTT~3864 (Hamuy et al. \cite{ham})
for flux calibration.
\begin{table}
\caption[]{Log of EFOSC1 observations}
\begin{flushleft}
\begin{tabular}{cccrc}
Target & Slit width & Date & Exp. & ZD's \\
& ($''$) & & (s)~ & ($^\circ$) \\
\hline
5601 & 1.5 & 1995 Apr 03 & 2400 & 18.3, 14.8, \\
+ 5619 & & & & 15.3, 19.7, \\
+ 5621 & & & & 26.2, 36.0, \\
& & & & 43.4 \\
& 5.0 & & 600 & 30.9 \\
& & & & \\
1902 & 1.5 & 1995 Apr 04 & 2400 & 44.3, 36.8, \\
& & & & 29.5 \\
& & & & \\
5601 & 1.5 & 1995 Apr 04 & 2400 & 19.7, 15.3, \\
+ 5619 & & & & 14.8, 18.5, \\
+ 5621 & & & & 25.9, 34.1, \\
& & & & 41.4 \\
& & & & \\
1902 & 1.5 & 1995 Apr 05 & 2400 & 43.0, 35.1, \\
& & & & 27.8 \\
& & & & \\
4001 & 1.5 & 1995 Apr 05 & 2400 & 21.4, 17.7, \\
& & & & 14.4, 16.0, \\
& & & & 21.3, 28.1, \\
& & & & 35.9, 42.1 \\
\end{tabular}
\end{flushleft}
Resulting total exposure times per target field are: \\
PN\#5601+5619+5621 - 9.50 hrs \\
PN\#1902 - 4.00 hrs \\
PN\#4001 - 5.33 hrs
\end{table}
The slit orientation was kept fixed for each set of observations of a given
target. This was necessary to avoid the delays resulting from many
replacements of the slit on the target which would have resulted from
the conventional tracking of the parallactic angle by slit rotations.
In addition, keeping a fixed slit orientation facilitated the subtraction
of the underlying (galaxy) continuum and the detection of several PN along the
slit lengths. However imposing a fixed orientation leads to differential
loss of light with wavelength, largest at the higher zenith distances, as
the parallalactic angle differs from the slit position angle. The requirement
of long integration times and only three allocated nights forced us to observe
the targets for some (small) fraction of the time at zenith distances exceeding
40$^\circ$, where the differential
atmospheric refraction between 3700 and 6700\AA\ exceeds the slit width
(e.g. Fillipenko \cite{fill}).
In order to attempt to control the amount of wavelength-dependent slit loss through
differential refraction, we adopted an observing strategy of always
including a bright stellar source on the slit. The spectrum of this source
could then be used to monitor, and correct, the differential
refraction losses.
Comparison of the extracted spectra of the star, corrected for atmospheric
extinction, at high and low zenith distance should allow this correction
to be applied
to the spectra of the PN. One exposure of PN\#5601 with the star on the slit
was also made with a broad slit (5$''$) in order to test the validity of this
technique with a spectrum essentially free from any differential slit losses.
\section{Reductions and Results}
\subsection{Reductions}
All the spectra were reduced in the usual way, using the
spectroscopic packages in IRAF. A super-bias frame was formed by averaging
many individual bias frames and this
bias image was subtracted from all frames. A mean flat field image was formed
from many exposures to a tungsten
lamp shining off a reflector in the dome and was used to rectify
the pixel-to-pixel variations. The sky flat was employed to map the
response of the system in the cross dispersion direction, thus correcting
any vignetting in the optical system or variation in slit transmission.
The exposures of the Neon and
Argon lamps were employed to fit the known wavelengths of the
comparison lines and the spectra were rebinned into
channels of constant wavelength width by fitting third
order polynomials. Figure 2 shows the long slit spectrum centred on PN\#5601
formed by averaging (with cosmic ray rejection) the first four exposures
on April 03 (Table 2). The other two PN (5619 and 5621) are clearly
visible from their [O~III] line emission. In addition there are a number
of continuum sources detected, some of which are probably stellar
clusters in NGC~5128. There is a faint red star displaced less than
one seeing disc from the PN\#5601 and visible in Fig. 2. The bright
continuum source closest to PN\#5619 was the one used in attempting to
correct for differential refraction slit losses.
\begin{figure*}
\resizebox{\hsize}{!}{\includegraphics{H1150f2.ps}}
\caption{The 2-D longslit spectrum centred on PN\#5601 (see
Table 1) formed by averaging the first four exposures on April 03
(see Table 2) is shown. Wavelength increases from bottom to top and
the left edge corresponds to south. The position of the
[O~III]5007\AA\ line is indicated as are the positions of the three
detected PN (see Table 1).}
\end{figure*}
Spectra of the PN, and the reference star, were extracted from each
image using
optimal weighting after subtracting the mean sky from the vicinity of the
objects. The atmospheric
extinction was corrected for each exposure and absolute flux calibration
applied from the observation of the spectrophotometric standard.
The background contained a substantial contribution of
galaxy light, especially for PN\#4001 so it was necessary to
restrict the background region to be close to the PN (with
typically at least three times as many pixels in the sky as in the
extracted PN). After extraction and removal of cosmic rays,
the individual spectra were flux calibrated.
The spectra of the reference continuum source from each exposure
were compared. The spectra showed a very large difference.
Spectra taken before transit of the star indicated an upward
correction to the blue fluxes and a downward correction to red fluxes
relative to the spectrum taken at the lowest airmass.
For the spectra taken after transit of the star a downward
correction to the blue fluxes and an upward correction to red fluxes,
relative to the spectrum taken at the lowest airmass, was found.
However applying such corrections to the extracted PN spectra gave
inconsistent line fluxes in the sense that the correction
factors were too steep with wavelength and resulted in discordant
spectra from before and after meridian passage. The explanation for the
derivation of such unrealistically large corrections is unclear.
The continuum source could be extended (e.g. be a cluster in NGC~5128
itself) and have different colours in different regions; however this
seems unlikely since exactly the same behaviour was exhibited by
the reference spectra for the other targets. The most probable
explanation is that the slit rotates slightly during the course of the
observations so that the flux received in the slit tracks across the
image in opposite directions on either side of the meridian,
exaggerating the effects of differential atmospheric extinction.
It was found that the line ratios of the extracted (and extinction
corrected) PN spectra did not vary systematically with airmass
beyond the errors of measurement.
In addition the line ratios in the extracted spectra did not differ
from those of the broad slit exposure of PN\#5601 by more than the
errors, although flux determination of the H$\alpha$ line was
hampered by the broadened sky lines. Since emission lines were only
detected over the wavelength range 4500 to 6700\AA, and the
differential atmospheric refraction is 0.70$''$ over this range
for an airmass of 1.4, then in 1.5$''$ seeing with a 1.5$''$ slit
the differential flux loss was at maximum 30\% (see Fig. 1
of Jacoby \& Kaler \cite{jaka}). Thus overall only small losses
in spectrophotometric
integrity of the combined spectra should result. The fluxed spectra
for each PN were averaged (using weights based on exposure time)
on a case by case basis excluding the
last exposure at highest airmass to form the final PN spectra.
The extracted spectrum of the jet filament was treated similarly.
\subsection{Results}
Figure 3 shows the mean spectra of the five PN observed in Cen-A.
The red star continuum under PN\#5601 could not be effectively subtracted.
The emission lines were interactively fitted by Gaussians and the
flux in the lines are listed in Tables 3 and 4 for each PN and the
filament. In Table 3 the line flux data for the three brighter PN
are collected. The errors in Table 3 take into account the continuum
under the line and the photon noise in the sky-subtracted spectra;
the errors on the H$\beta$ flux have been propagated to the
other line flux errors. The measured signal-to-noise on the
[O~III]5007\AA\ line flux for the brightest PN (5601) is 55.
The reddening correction was calculated by comparing the observed
H$\alpha$/H$\beta$ ratio to the Case B value using the Seaton
(\cite{seat}) Galactic reddening law and is listed in Table 3.
The dereddened line fluxes (the error on the extinction
was not propagated to the dereddened line errors) together
with the observed H$\beta$ flux are listed in Table 3. In Table 4
the fluxes are presented for the two fainter PN and the filament
near PN\#1902. The errors are substantially larger than for the
data presented in Table 3, since the PN are fainter; for
example for PN\#5619, the signal-to-noise on the measurement of the
[O~III]5007\AA\ line is 9.
\begin{figure*}
\resizebox{\hsize}{!}{\includegraphics{H1150f3.ps}}
\caption{The spectrum of the five PN detected in
NGC~5128 is shown. The designation of each object is indicated
(from Hui et al. \cite{huib}).}
\end{figure*}
\begin{table*}
\caption[]{Emission line fluxes of three brightest PN observed in NGC~5128}
\begin{flushleft}
\begin{tabular}{cl|rrrr|rrrr|rrrr}
PN\# & & \multicolumn{4}{c}{1902} & \multicolumn{4}{c}{4001} & \multicolumn{4}{c}{5601} \\
Ident. & $\lambda$ & I$_O^\ast$ & $\pm$ & I$_D^\dag$ & $\pm$ & I$_O$ & $\pm$ & I$_D$ & $\pm$ & I$_O$ & $\pm$ & I$_D$ & $\pm$ \\
& (\AA) & & & & & & & & & & & & \\
\hline
He~II & 4686 & 64 & 23 & 63 & 23 & 12 & 3 & 12 & 3 & 20 & 16 & 20 & 17 \\
H$\beta$ & 4861 & 100 & 0 & 100 & 0 & 100 & 0 & 100 & 0 & 100 & 0 & 100 & 0 \\
{[O~III]} & 4959 & 498 & 93 & 502 & 93 & 330 & 76 & 331 & 76 & 414 & 51 & 416 & 51 \\
{[O~III]} & 5007 & 1185 & 216 & 1197 & 219 & 960 & 218 & 965 & 219 & 1268 & 153 & 1268 & 153 \\
He~I & 5876 & 31 & 7 & 33 & 7 & & & & & 30 & 16 & 31 & 17 \\
H$\alpha$ & 6563 & 261 & 49 & 286 & 54 & 274 & 57 & 286 & 62 & 271 & 39 & 286 & 41 \\
{[N~II]} & 6583 & 135 & 28 & 147 & 30 & 18 & 5 & 17 & 5 & 56 & 11 & 59 & 12 \\
{[S~II]} & 6716+31 & 58 & 27 & 63 & 30 & & & & & 24 & 9 & 25 & 10 \\
\hline
c & & \multicolumn{4}{c}{-0.12$\pm$0.23} & \multicolumn{4}{c}{-0.06$\pm$0.28} &
\multicolumn{4}{c}{-0.07$\pm$0.19} \\
Log$_{10}$F(H$\beta$)$^\ddag$ & & \multicolumn{4}{c}{-16.3} & \multicolumn{4}{c}{-16.4} & \multicolumn{4}{c}{-16.2} \\
\end{tabular}
\end{flushleft}
\noindent
$^\ast$ Observed flux normalised to I(H$\beta$)=100 \\
$^\dagger$ Dereddened flux normalised to I(H$\beta$)=100 \\
$^\ddag$ Absolute observed H$\beta$ flux
\end{table*}
\begin{table*}
\caption[]{Emission line fluxes of fainter PN and filament observed in NGC~5128}
\begin{flushleft}
\begin{tabular}{cl|rrrr|rrrr|rrrr}
PN\# & & \multicolumn{4}{c}{5619} & \multicolumn{4}{c}{5621} & \multicolumn{4}{c}{Filament} \\
Ident. & $\lambda$ & I$_O^\ast$ & $\pm$ & I$_D^\dag$ & $\pm$ & I$_O$ & $\pm$ & I$_D$ & $\pm$ & I$_O$ & $\pm$ & I$_D$ & $\pm$ \\
& (\AA) & & & & & & & & & & & & \\
\hline
He~II & 4686 & & & & & 46 & 38 & 48 & 38 & & & & \\
H$\beta$ & 4861 & 100 & 0 & 100 & 0 & 100 & 0 & 100 & 0 & 100 & 0 & 100 & 0 \\
{[O~III]} & 4959 & 482 & 190 & 484 & 190 & 365 & 153 & 358 & 150 & 452 & 136 & 441 & 133 \\
{[O~III]} & 5007 & 1484 & 480 & 1489 & 482 & 904 & 360 & 879 & 355 & 1074 & 313 & 1036 & 302 \\
H$\alpha$ & 6563 & 278 & 96 & 286 & 99 & 365 & 127 & 286 & 100 & 382 & 113 & 286 & 85 \\
{[N~II]} & 6583 & & & & & 252 & 100 & 197 & 80 & 251 & 73 & 186 & 54 \\
{[S~II]} & 6716+31 & & & & & 67 & 30 & 51 & 23 & & & & \\
\hline
c & & \multicolumn{4}{c}{-0.04} & \multicolumn{4}{c}{0.33} &
\multicolumn{4}{c}{0.40} \\
Log$_{10}$F(H$\beta$)$^\ddag$ & & \multicolumn{4}{c}{-17.1} &
\multicolumn{4}{c}{-16.7} & \multicolumn{4}{c}{-16.6} \\
\end{tabular}
\end{flushleft}
\noindent
$^\ast$ Observed flux normalised to I(H$\beta$)=100 \\
$^\dagger$ Dereddened flux normalised to I(H$\beta$)=100 \\
$^\ddag$ Absolute observed H$\beta$ flux
\end{table*}
\section{Discussion}
\subsection{Planetary Nebula spectra}
The five PN observed in NGC~5128 show spectra entirely typical of PN;
the spectra are not obviously distinguishable from those of Galactic PN.
Although the signal-to-noise is not high, the range of line fluxes,
absolute H$\beta$ fluxes and [O~III]5007\AA/H$\beta$ ratios is similar
to that for high excitation Galactic PN. There were no low excitation
PN spectra among the five, but this is probably not surprising given that
the source detection was performed in [O~III] and the emphasis here was
on the brightest objects. The range in [O~III]
brightness covered is 7.5 (from the photometry of Hui et al. \cite{huib})
and 8 in H$\beta$ flux from the long slit observations (Tables 3 and 4).
From the standpoint of the evolution of low mass stars it is not
surprising that the spectra are typical, but on the other hand
these are among the brightest PN in a whole galaxy. Doubt had been expressed
that the PN at the peak of the luminosity function may not have been
typical of the general PN population and that distance estimates which
relied on the peak of the luminosity function could suffer from
systematic bias (Bottinelli et al. \cite{bott}; Tammann, \cite{tam}).
These effects have been carefully refuted (Feldmeier et al. \cite{feld},
McMillan et al. \cite{mcmi} and Jacoby \cite{jac96}) and together
with the spectra shown in Fig. 3 and tabulated in Tables 3 and
4 amply demonstrate that the brightest PN in a galaxy {\bf are} typical.
That they are the brightest is simply due to the fact that they are
observed whilst at their peak luminosity. The 5007\AA\ luminosity is
generally higher for higher mass progenitor stars and also peaks in the
later stages of evolution of lower mass stars (Sch\"{o}nberner \& Tylenda
\cite{scho}). However for PN with high core masses, above $\sim$0.65M$_\odot$,
the high nitrogen abundance can decrease the efficacy of cooling by oxygen
emission, reducing the 5007\AA\ flux by $\sim$ 0.5mag (Kaler \&
Jacoby \cite{kaja}).
The extinction correction for the three brightest objects indicates
a slightly negative value. The Galactic component of reddening to
NGC~5128 given by Burstein \& Heiles \cite{buhe} is E$_{B-V}$ = 0.123
(c=0.18). This value is similar to that (E$_{B-V}$ = 0.10) adopted by
van den Bergh (\cite{berg}) and that derived most recently from
DIRBE dust maps
by Schlegel et al. \cite{schl} (E$_{B-V}$ = 0.115); however
Jablonka et al. (\cite{jabl})
measured values of E$_{B-V}$ as low as 0.03 from spectrophotometry
of globular clusters in NGC~5128. The extinction values for
the PN are consistent with these values within the errors, except
perhaps for PN\#5601. However it is puzzling
that the values are systematically low; any local extinction in
NGC~5128, or dust within the nebulae themselves, would increase the
value above the baseline for the Galactic line of sight extinction.
Slit lossses through atmospheric refraction should not alone
account for bias. Errors may have arisen in subtraction of the
underlying stellar continuum whereby emission line flux is lost
to stellar absorption lines, although the effect would be to produce
an increased extinction on account of the generally higher H$\beta$
absorption equivalent width compared with H$\alpha$. However one of the
PN (5621) does show an extinction above the Galactic value, although
with a substantial error, as does the filament, which is however an
extended object. The extinction to
the filament is in the range of values for the extinction determined
by Morganti et al. (\cite{morg}); the filament is in the same
vicinity as their Field 2 (see their Fig. 2) although about ten
times lower
surface brightness (compare their Table 4 for spectra). The
[O~III]5007\AA/H$\beta$ ratio measured here is also similar to
the values measured by Morganti et al. (\cite{morg}).
The most probable explanation of the depressed reddening values
is that the central wavelength of the guiding camera
lies at one end of the range H$\beta$ to H$\alpha$; by tracking on
the image in the vicinity of the wavelength around H$\beta$, flux is
systematically lost from the slit at H$\alpha$ for airmasses much
greater than 1.0. Subsequent to this conclusion we were informed
that the guiding camera of the ESO 3.6m at the time of the
observations was sensitive over the wavelength range 3700 to 5000\AA.
The dereddened fluxes were formed employing the observed reddening,
even if negative; this serves to compensate for the losses of the
red part of the spectra. No specific correction for foreground
(Galactic) reddening was employed.
It is apparent that the three brightest PN show no evidence for
intrinsic reddening (within the substantial measurement errors), which
is probably not surprising given that they are the the brightest PN
observable in the galaxy. Any extinction would move them to
lower observed fluxes; PN\#5621 is for example as intrinsically
bright as PN\#4001. The effect of local galactic extinction and
dust intrinsic to the PN must play a role in shaping the PN
luminosity function (Jacoby \cite{jac89}).
With spectroscopy of the brightest PN, the effect of dust on the
luminosity function, and hence on the distance estimate through
fitting of this function (Ciardullo et al. \cite{cia89}; see also
Mendez et al. \cite{men}), can be directly quantified.
However if dust is associated with PN dependent on their
luminosity it would be expected to have a strong effect on the PN
luminosity function. Jacoby \& Ciardullo (\cite{jaccia}) in their
study of PN in M~31 have found a weak correlation between extinction
and 5007\AA\ luminosity, which also exists for PN in the LMC. The
surprising net effect on the PN luminosity function is that the apparent
peak brightness is nearly independent of absolute peak brightness.
On the basis of the [O~III]5007\AA/H$\beta$ and He~II~4686\AA/H$\beta$
ratios the excitation class can be defined (Dopita \& Meatheringham
\cite{dom90}). A least squares fit of the effective temperature
from photoionization models (Dopita \& Meatheringham \cite{dom91b})
against excitation class for a uniform set of observations and
models of Magellanic Cloud PN allows estimation of effective temperature.
Use of this data set could be criticized since the LMC and SMC have
low metallicities compared to Galactic or more metal rich
galaxies, but the modelling of
AGB evolution shows no strong dependence of stellar temperature on
metallicity (Dopita et al. \cite{djv}).
Table 5 lists the excitation class and indicative temperature of
the PN in Cen-A. PN\#5619 could not have a reliably assigned
stellar temperature since its excitation class is high (based on
its high [O~III]/H$\beta$ ratio) yet it was too faint to detect
He~II4686\AA\ (excitation class above 5.0 requires the He~II/H$\beta$
ratio). These temperatures should be seen as upper limits, since
if the nebulae are optically thin the high ionization emission
is enhanced; given the large line ratio errors the likely errors
are at least $\pm$10000K.
\begin{table*}
\caption[]{Parameters of the NGC~5128 Planetary Nebulae}
\begin{flushleft}
\begin{tabular}{lrrrrr}
Parameter & PN\#1902 & PN\#4001 & PN\#5601 & PN\# 5619 & PN\# 5621 \\
\hline
Excit. Class & 7.8 & 5.0 & 5.4 & 6.7 & 7.0 \\
T$_{\ast}$(K) & 180000 & 100000 & 110000 & $\sim$140000 & 155000 \\
Log L$_{\ast}$ & 4.0 & 3.9 & 4.2 & - & 3.7 \\
M$_{\ast}$ (M$_\odot$) & 0.68 & 0.64 & 0.83 & - & 0.62 \\
T$_e$(K) & 13000 & 14000 & 13000 & 14000 & 12500 \\
Z(Z$_\odot$) & -0.3 & -0.6 & -0.4 & -0.6 & -0.4 \\
O$^{++}$/H $\times$10$^{5}$ & 26 & 13 & 21 & 14 & 25 \\
12$+$Log$_{10}$(O/H) & $<$8.5 & 8.2 & 8.5 & $>$8.4 & $<$8.3 \\
$[O/H]$ & $>$-0.4 & -0.7 & -0.4 & $<$-0.5 & $>$-0.6 \\
N/O & 0.4 & 0.3 & 0.4 & - & 0.5 \\
\hline
\end{tabular}
\end{flushleft}
\end{table*}
\subsection{Abundances of the Planetary Nebulae}
Even for the three brightest PN observed in NGC~5128, the weak
diagnostic forbidden lines were not detected; thus it is not
possible to measure accurate electron temperatures (from the
[O~III]5007/4363\AA\ ratio) or densities (for example from the
[S~II]6716/6731\AA\ ratio). Nevertheless an attempt was made to
estimate the oxygen abundance in order to compare it with other
abundance diagnostics (e.g. from stellar absorption lines).
Dopita et al. (\cite{djv}) presented a diagnostic diagram of
PN metallicity {\em v.} effective temperature, in which the
electron temperture can be determined from the [O~III]/H$\beta$
ratio. This plot was derived from photoionization models
of a grid of optically thick PN; the effective temperature
being determined from the fit to the Magellanic Cloud data
(Dopita \& Meatheringham \cite{dom91b}).
In Table 4 the estimated values of the electron temperature
are listed; for PN\#1902 and 5621 the range of effective
temperatures are out of range of the diagnostic plot, but
were extrapolated. Row 6 of Table 4 lists the
derived metallicities of the PN (all element abundances
scaled except Helium), based on the Dopita et al. (\cite{djv})
calibration. From the electron temperature estimates (Table 4
row 5), the empirical O$^{++}$ abundances were determined
from the [O~III]/H$\beta$ ratios and are listed in row 7.
A correction for the presence of O$^{3+}$ was made using the
ionization correction factor derived from the He/He$^{++}$
ratio (Kingsburgh \& Barlow \cite{kiba}); the He/H ratio
was assumed fixed at 0.15. The He~I 5876\AA\
line indicates He$^{+}$/H$^{+}$ ratios as high as 0.2 but
this line suffers from proximity to the strong Na~I telluric
lines, and the difficulty in good sky subtraction leads to
a large error on the line measurement. Assuming that the N/O ratio
for the PN is the same as the mean value for the Galactic PN
(0.28; Kingsburgh \& Barlow \cite{kiba}), the fraction of
O$^{+}$/H$^{+}$ was estimated and thus the total oxygen
abundance. The uncertainties introduced in
correcting for the presence of O$^{3+}$ and O$^{+}$ are
between 20 and 40\%. Where the [N~II]6583\AA\ line was
strong it was assumed that the N/O ratio was higher than the
Galactic mean value and the O$^{+}$
contribution was included as an upper limit. Row 8 lists the
derived logarithmic O/H abundances and row 9 the oxygen
abundances compared to solar.
In addition to these empirical estimations of the nebular parameters,
the photoionization modelling package CLOUDY (Ferland \cite{ferl})
was used to model the spectra matching the [O~III] luminosity and
relative line strengths; the carbon abundance was assumed as
12$+$Log$_{10}$(C/H)=8.7. The derived parameters were generally in
good agreement with those empirically derived; the stellar temperatures
were about 10000K lower however.
The fair agreement between the
empirical O/H abundance estimates and those from photoionization
models in Table 5 provides assurance that the use of empirical relations
calibrated from lower metallicity Magallanic Cloud planetary nebulae
does not seriously affect (within $\sim$0.15dex) the resulting
abundance estimates.
In rows 3 and 4 of Table 3, the derived
stellar luminosity and core mass (from Sch\"{o}nberner \cite{sch81},
\cite{sch83} tracks) are also listed. From the models, the diagnostically
useful N/O ratio was calculated and is listed in row 10. Due to the limited
spectroscopic constraints on the models, the nitrogen abundances are
uncertain; the N/O ratios could be a factor of 2 -- 3 higher, but are not
likely to be much lower if the PN are optically thick (which appears
probable both from the photoionization models and their high luminosity).
Whilst the values
of N/O are moderately high (the mean value for non-Type I Galactic PN is
0.28 - Kingsburgh \& Barlow \cite{kiba}), only PN\#5621 satisfies the
criterion of N/O $>$ 0.5 for classification as a Type I PN,
considered to arise from higher mass progenitor stars (Peimbert \&
Torres Peimbert \cite{pepe}). This object in addition displays
He~II emission strong relative to H$\beta$,
so may be a bona fide
Type I nebula (He/H$\geq$0.125). A direct measurement of the
N/O ratio (such as from [N~II]6583/[O~II]3727\AA\ line ratio) would be
required to confirm this classification. However the two brightest
PN (5601 is the brightest PN in the galaxy detected by Hui et al.
\cite{huia}) are not obviously Type I PN; this is consistent with
observations of PN in the Magellanic Clouds (Dopita \& Meatheringham,
\cite{dom91b}) that Type I PN are not the most luminous in a
population on account of their fast evolution to high effective
temperatures and hence lower luminosities. Type I PN could be more
optically thin than their lower mass counterparts as suggested by
Mendez et al. (\cite{men}), thereby leading to lower observed [O~III]
luminosities than expected from the luminosities of the central stars.
Type I PN may also be more copious producers of dust, hence further
lowering their observed luminosities.
The quality of the oxygen abundance determinations is not high
enough to investigate any evidence of a metallicity gradient;
the O/H abundance {\em v.} projected radius shows no trend, perhaps
even a suggestion of increasing with increased galactocentric radius.
The sample is too small and the quality of the O/H determinations
too low to draw any conclusions. The mean [O/H] abundance appears
to be -0.5 for the PN in Cen-A; this can be compared with the
mean value for 42 non-Type I Galactic PN of -0.24 (Kingsburgh \& Barlow
\cite{kiba}). It is surprising that the oxygen abundance
is lower than characteristic for Galactic PN, given that the
metallicity is expected to be higher in this high luminosity elliptical
galaxy. The range of galactocentric radii
probed by the five PN is however 7.1 to 18.0 Kpc (Table 1), which,
by analogy with the Milky Way would show a lower metallicity
than the core (by [O/H] $\sim$ -0.3 at 7Kpc e.g. Shaver et al.
\cite{shav}).
There does not appear to be a direct metallicity determination for the
large scale stellar content of this galaxy.
However the metallicity of NGC~5128 can be estimated using the
tight relations between velocity dispersion, luminosity, and
Mg$_2$ index for giant ellipticals (Faber \& Jackson \cite{faja};
Terlevich et al. \cite{terl}). The Mg$_2$ index can then be
converted to metallicity using stellar population models.
NGC~5128 has a central velocity dispersion between 150 and 200 km s$^{-1}$
(Wilkinson et al. \cite{wilk}). This might however be a lower limit,
since the inner regions are obscured by the large central dust lane.
Wilkinson et al. (\cite{wilk}) conclude that M$_B \approx -$20.5 mag.;
this then corresponds to an Mg$_2$ index of 0.31. To convert this
value into a metallicity, a value for the age of NGC~5128 must be
assumed. For an age of 17 Gyr Vazdekis et al. (\cite{vaz}) give
a metallicity slightly higher than solar, or [Fe/H]=0.4 for an age of
6 Gyr. These numbers again depend slightly on the IMF chosen, but it
is fairly certain that the stellar indicators show an abundance
which is about solar or higher. In addition the globular clusters
have a higher mean metallicity than for the Milky Way. In NGC~5128
Harris et al. (\cite{har92}) determined a mean metallicity
$<[$Fe/H$]>$ of -0.8 for 62 globulars from Washington photometry
whilst the mean for all globulars in the Milky Way is -1.35. Clearly
a determination of the stellar metallicity variation with radius is
required for NGC~5128 to contrast with the measurements from the
planetary nebulae. It should also be a priority
for future observations to study PN near the galaxy core (but
avoiding the dust lane) in order to search for high abundance PN.
A detailed investigation of the discrepancy between the stellar
and PN abundances is beyond the scope of this paper involving as
it does stellar evolution, chemical enrichment processes, mergers
and elliptical galaxy formation. However a similar discrepancy
between the stellar and PN metallicity
is seen with abundance data from PN in the bulge of M~31, where
a mean [O/H]$\sim$-0.5 is found (Jacoby \& Ciardullo \cite{jaccia}) in strong
contrast to the apparently super-solar stellar abundances.
One obvious reason for such discrepancies could be that the PN and
stellar abundances do not refer to the same stars; for example the
absorption line spectra would be weighted by the most luminous
stars. In the optical the more metal rich stars are generally fainter
than lower metallicity ones so that the stellar indicators
(e.g. Mg$_{2}$ index) are weighted to older stars, except
in young populations (less than a few Gyr). However
if the chemical enrichment
proceeds monotonically with time, then the younger stars are
more metal rich; although enrichments of 0.5 dex in a few Gyr might
require epochs of star formation rather than steady evolution.
If mergers with lower luminosity (and metallicity) galaxies
contributed substantially to the stellar population this would
decrease or even reverse the trend of increasing metallicity
with time. NGC~5128 appears to have suffered a recent merger.
Unlike most other ellipticals it has a prominent twisted disk of
gas containing numerous HII regions, and lying approximately along
the galaxy minor axis. The velocities of the PN (Hui et al.
\cite{huic}) indicate that the inner disk, containing the HII
regions, is rapidly rotating or is the remnant of a precessing,
nearly polar, gas disk in an axisymmetric potential (Sparke
\cite{spark}). The five PN observed could have originated from
the smaller, lower metallicity, infalling galaxy; however only
a fraction of the total number of PN in NGC~5128 could have been
so produced since the luminosity specific PN density for NGC~5128
is similar to that for other early-type galaxies (Hui et al.
\cite{huia}).
A second reason for the low O abundance from the PN in comparison
with the expected high stellar (Fe peak) abundance is suggested
by the known anti-correlation between (Galactic) stellar Fe/H and O/Fe
(King \cite{king}). If the stellar population is super metal
rich when considered from the viewpoint of Fe abundances it
is not from the perspective of O abundances, which the PN
reflect, independent of any stellar O/Fe calibration.
However such a relation may be rather specific to
the chemical evolution history of the Galaxy which must be
very different from that of NGC~5128. But ellipticals
tend to show Mg/Fe larger than solar (e.g. Worthey et al. \cite{wor})
and the stellar Mg abundance should follow O (Faber et al.
\cite{fab92}) which brings back the discrepancy between the PN and
the stellar abundances. The extensive data
on PN in different galaxies collected by Stasinska et al. (\cite{stas})
show that O/H deduced from the PN is dependent on their luminosity
(viz. stellar core mass), being larger for high luminosity PN.
In addition the more luminous PN may be younger and should then
probe the
interstellar medium at later epochs. Although their sample does
not encompass massive early-type galaxies, it implies that the
low mean O/H of the five PN in NGC~5128 is difficult to understand.
The luminosity specific PN number density is also high for
NGC~5128, consistent with the bluer colour of NGC~5128
(Peimbert \cite{peim}; Hui et al. \cite{huia}) suggesting
a lower metallicity than the giant ellipticals in Virgo,
such as NGC~4472. Population age may also be a consideration for
PN number density, since it appears from
the relatively few number of PN in Galactic globular clusters
that old stars produce relatively few PN (Jacoby et al. \cite{jac97}).
Contamination of the PN sample by halo objects, which, by
comparison with the Galactic PN sample (e.g. Howard et al.
\cite{how}) have notably low metallicity, could be reduced by
considering the kinematics. From the NGC~5128 PN
kinematic survey (Hui et al. \cite{huic}), PN\#5601 and
1902 could be halo objects but their O abundances are on the
high side of the mean (Table 5). The effect of high metallicity
on PN formation is also not known - for example an
enhanced mass loss rate on the AGB could result in dispersal of the
envelope before the central star has had time to heat up enough
to ionize the nebula (the AGB manqu\'{e} channel Greggio \&
Renzini \cite{gren}; see also Ferguson \& Davidson \cite{feda})
This phenomenon would give rise to fewer PN at higher metallicity
explaining the lower average metallicity of the PN. It is expected
that the brightest PN come from a slightly metal poor
population (Ciardullo \& Jacoby \cite{cija}) as predicted
by the models of Dopita et al (\cite{djv}). In this
context it would be useful to search for Galactic bulge PN
with super-solar metallicity in order to reach a clearer
understanding of the role of metallicty on PN evolution.
In addition more and better data on the
PN and the stellar populations in metal rich systems, not
confined to the bulges of spirals, are required for an
understanding of what controls PN evolution in such environments
before detailed evolutionary scenerios in particular galaxies
can be developed.
\subsection{Exploring Abundance Gradients with PN}
In order for the PN to be reliable tracers, their abundances
must reflect those of the gas from which the stars were formed
and not solely be a consequence of nuclear reprocessing. From studies
of Galactic PN, the O, Ne, S and Ar gradient (Maciel \&
K\"{o}ppen \cite{mac}) matches that of the H\,II
regions (Shaver et al. \cite{shav}) as does the He gradient (Peimbert
\& Serrano \cite{pei}). This applies to the
(common) Type~II nebulae, not to the minority Type~I PN,
originating from higher mass progenitors and having enhanced
He, N and Ne. The Type II PN and H\,II regions in the lower
metallicity environment of the Magellanic Clouds also indicate
similar abundances (see Clegg \cite{cleg}). Richer
(\cite{rich}) has arrived at the important conclusion
that the brightest PN in ellipticals have the same status as
abundance indicators as H\,II regions in spirals. Spectra
comparable to or better than the ones presented here for
PN\#1902, 4001 and 5601 are required to distinguish the Type II
from Type I nebulae and to determine improved oxygen abundances.
For the brighest PN it will be feasible to detect the [O~III]4363\AA\
line and thus determine O abundances to $\pm$0.2dex or better;
for lower luminosity PN, or galaxies more distant than NGC~5128,
empirical abundance determination and photoionization modelling,
as performed here, is required
and the derived abundances are of lower individual weight. However
average oxygen abundances at least comparable in accuracy to
those for the integrated stellar population as a function of
radius, and for individual globular clusters, are achievable.
A rather large sample of PN is required to distinguish a
trend in the abundance, since a PN at a given effective radius
may reside at a large distance from the galaxy, due to
projection. The radial velocity data could be used to give
a partial answer to distinguish halo PN from body PN. Multi-object
spectroscopy techniques are required to obtain spectra
of the requisite numbers of PN to distinguish a trend and to
sample the line of sight abundance spread, excluding halo
objects. However given that most of the PN are projected against
a strong stellar continuum, then multi-slit rather than multi-fibre
instruments are required to provide accurate background subtraction
of the spectra. Coherent fibre bundles, one for each PN,
could alternatively be employed to allow effective 2-D background
subtraction. However single fibres are adequate for
radial velocity work based on the brightest line of 5007\AA.
Since the orientation of a slit must be kept
fixed to provide a good background subtraction and to allow several PN
to be observed per slit, then the use of an Atmospheric Dispersion
Corrector is highly advantageous to ensure spectrophotometry over the
requisite large wavelength range ($\sim$3700 - 6800\AA\ for He, N, O,
Ne and S abundance determinations). When considering observation
of PN in galaxies more distant
than NGC~5128, the issue of background subtraction will become more
crucial as larger variations in galaxy continuum will be included
in the slit or aperture. Spectrophotometry of extra-galactic PN
is a field where multi-object techniques on 8-10m telescopes will
bring a rich harvest of data to bear on the history of chemical
enrichment in galaxies of all types.
\section{Conclusions}
The first spectra of planetary nebulae in the nearby early-type galaxy
NGC~5128 have been presented. The spectra of five PN from the
catalogue of Hui et al. (\cite{huib}) have been
observed over an observed emission line brightness range of a
factor 8 and galactocentric radius range from 7 to 18 kpc. The spectra show
characteristic high ionization emission lines
similar to Galactic PN and confirming that the brightest PN in a galaxy
are entirely typical. The mean [O/H] of the five PN, determined
by empirical methods and modelling, is $-$0.5 with a spread of
0.3dex. This low metallicity contrasts with that of the assumed metal
rich stellar population of NGC~5128.
\begin{acknowledgements}
We would like to thank M. Richer for stimulating comments on
the subject of probing galactic abundances from the PN
population.
\end{acknowledgements}
\noindent
{\bf Note added in proof} \\
Harris et al. (AJ 116, 2866, 1998 and AJ 117, 855, 1999)
have obtained HST photometry of a globular cluster and the
field halo stars in NGC~5128, situated at a distance of $\sim$21 kpc
from the galaxy centre. For the halo stars they estimate a mean
metallicity of $<[Fe/H]>$ = $-$0.4, but with a broad range.
Although the PN observed in this paper were at smaller
galactocentric distances, there is interesting agreement
between the metallicity of the PN from the [O/H] determinations
presented here and those for
the Red Giant Branch stars observed by Harris et al.
|
\section{Introduction}
Recently we presented a model to fit the high energy cosmic ray spectrum
using the hypothesis that the electron neutrino is a
tachyon.\cite{Ehrlich} A good fit to the spectrum was obtained using
$|m_\nu| \equiv \sqrt{-m^2} = 0.5\pm 0.25$ eV/c$^2.$ The signature
prediction of the model is the existence of a neutron flux `spike' in
the cosmic rays centered on $E = 4.5 \pm 2.2$ PeV, and having a width
$\Delta\log E = 0.1$ (FWHM). Although the existence of neutral cosmic
rays from point sources remains a highly controversial subject, we
report here that an examination of the published literature on cosmic
rays from Cygnus X-3 reveals just such a hitherto unreported neutral
particle spike centered on E = 4.5 PeV with a level of statistical
significance of $6\sigma.$ An additional prediction of the model that
the integrated flux of neutrons above 0.5 EeV should be 0.048 percent
that above 2 PeV is also consistent with results from two out of three
experiments.
Although few physicists have taken tachyons seriously since they were
first proposed in 1962\cite{Bilaniuk}, their existence is clearly an
experimental question. In 1985 Chodos, Hauser and
Kosteleck\'{y}\cite{Chodos85}, suggested that neutrinos were tachyons --
an idea that is consistent with experiments used to determine the
neutrino mass. Chodos et al.\cite{Chodos92,Chodos94} also suggested a
remarkable empirical test of the tachyonic neutrino hypothesis, namely
that stable particles should decay when they travel
with sufficiently high energies. Consider, for example, the
energetically forbidden decay $p\rightarrow n + e^++ \nu_e.$ In order to
conserve energy in the CM frame the neutrino would need to have
$E < 0$. But tachyons with $m^2 <0$ have $E<p,$ and therefore the
sign of their energy in the lab frame $E_{lab} = \gamma(E + \beta p
cos\theta)$ will be positive for a proton velocity $\beta > \beta_{th}
\equiv -E/p cos\theta.$ With the aid of a little kinematics it can
easily be shown that the threshold energy for proton decay is
$E_{th}\approx 1.7|m_{\nu}|^{-1}$ PeV, with $|m_{\nu}|$ in eV.
Thus, if neutrinos are tachyons, energetically forbidden decays become
allowed when the parent particle has sufficient energy -- in seeming
contradiction with the principle of relativity that whether or not a
process occurs should not depend on the observer's reference frame.
That contradiction is only an apparent one, however, because what
appears to the lab observer as a proton decay emitting a neutrino
appears to the CM observer as a proton absorbing an antineutrino from a
background sea.
\section{Cosmic Rays}
Since cosmic rays bombard the Earth with energies far in excess of what
can be achieved in present day accelerators, it is natural to ask
whether any evidence for a process such as proton decay exists there at
very high energies. One striking feature of the cosmic ray spectrum is
the ``knee" or change in power law that occurs at $E \approx 4$ PeV.
Various two-source mechanisms have been suggested to account for this spectral
feature, but some researchers have identified it as arising from a single
type of source.\cite{Erlykin} In 1992 Kosteleck\'{y}\cite{Kostelecky}
suggested that for a tachyonic neutrino mass $|m_\nu| \approx 0.3 eV,$
the proton decay threshold energy occurs at the knee of the cosmic ray
spectrum, and could explain its existence. The idea is that cosmic ray
nucleons on their way to Earth would lose energy through a chain of
decays $p\rightarrow n\rightarrow p\rightarrow n\rightarrow \cdots,$
which would deplete the spectrum at energies above $E_{th}.$ However,
Kosteleck\'{y} regarded the existence of the knee by itself as
insufficient evidence for the tachyonic neutrino hypothesis in view of
other more conventional explanations of the knee of the cosmic ray
spectrum. He also did not attempt to model the spectrum, nor mention
the signature neutron spike.
Recently this author has developed a tachyonic neutrino model that fits
a number of features of the cosmic ray spectrum in addition to the
knee.\cite{Ehrlich} These include the existence and position of the
``ankle" (another change in power law at $E \approx 6$ EeV), the
specific changes in power law at the knee and ankle, the changes in
composition of cosmic rays with energy, and the ability of cosmic rays
to reach us above the conjectured GZK ``cutoff."\cite{Greisen,Takeda}
Although the fit to the cosmic ray spectrum was a good one, the model is
highly speculative, because it is at variance with
conventional wisdom about cosmic rays and it arbitrarily assumed
that the decay rate for protons (for $E>E_{th}$) was far greater than
that for neutrons.
Nevertheless, the model did make the striking prediction of a cosmic ray
neutron flux in a narrow range of energies just above $E_{th}$ -- a
neutron ``spike." The pile up of neutrons in a narrow interval just
above $E_{th}$ is a consequence of the fractional energy loss of the
nucleon in proton decay becoming progressively smaller, the closer the
proton energy gets to $E_{th}$. The position of the predicted cosmic ray
neutron spike depends on the value assumed for $|m_\nu|$. From the fit
to the cosmic ray spectrum we found $|m_\nu| = 0.5 \pm 0.25$ eV/c$^2,$ and
hence we predicted a neutron spike at $E = 4.5 \pm 2.2$ PeV. In fact
the model predicted that most nucleons should be neutrons for
$E > E_{th},$ because it was assumed that as nucleons lose energy in the
$p\rightarrow n\rightarrow p\rightarrow \cdots$ decay chain, the
lifetime and hence the decay mean free path for neutrons is far
greater than for protons, and so nucleons above $E_{th}$ would spend
nearly all of their time en route as neutrons.\cite{assumption} But, the model also
predicts that for energies above the spike the neutron component does
not become an appreciable fraction of the total cosmic ray flux until
around 1 EeV. While neutrons might reach Earth at EeV energies in
conventional cosmic ray models, it would be difficult to understand any
sizable neutron component at energies as low as E=4.5 PeV, where the
neutron mean free path before decay would be only about 100 ly. In the
present model, however, A = 1 cosmic rays can travel very many neutron
decay lengths and still arrive as neutrons because many steps of the
$p\rightarrow n\rightarrow p\rightarrow \cdots$ decay chain occur for
nucleons having energies above $E_{th}.$
\section{Cygnus X-3 Data}
One way to look for a neutron flux would be to find a cosmic ray signal
that points back to a specific source, since neutrons are unaffected by
galactic magnetic fields. Starting in 1983 a number of cosmic ray
groups did, in fact, report seeing signals in the PeV range from
Hercules X-1 and Cygnus X-3. At the time these signals were
believed to be either gamma rays or some hitherto unknown long-lived
neutral particle, since neutrons, as already noted, should not live long
enough to reach Earth (except in the present model). Some
of the experiments coupled detection of extensive air showers with
detection of underground muons.\cite{Samorski,Marshak} The observed
high muon intensity was found to be consistent with hadrons but not with
showers induced by gamma rays.\cite{Marshak,Stanev} It was widely
believed that the mass of the neutral particle was $m\approx 1
$ GeV/c$^2.$\cite{Cudell} Thus, all the observed or conjectured
properties of these particles were consistent with neutrons: neutral
strongly interacting particles with $m\approx 1$ GeV/c$^2.$
Following a period of excitement in the 1980's, many researchers began
to look critically at some of the observations of ultra-high energy
cosmic rays from point sources. This skepticism was based in part on
the inconsistencies between results reported in different experiments.
As Chardin and Gerbier have noted\cite{Chardin}, a number of papers used
data selection procedures that made direct comparisons difficult, e.g.,
using different phase intervals to make cuts, variously reporting the
total flux or only the flux in a particular phase bin, and reporting
only ``muon-poor" events. Also, some papers appeared to
inflate the statistical significance of their results.
But, the most serious challenge to the idea of neutral particles in the
PeV range from Cygnus X-3 and other point sources came from a trio of
high sensitivity experiments\cite{Alexandreas,Aglietta,Cronin} that
reported seeing no signals from point sources claimed earlier. In the
most sensitive experiment of the three, the upper limit on the flux of
neutral particles from Cygnus X-3 above 1.175 PeV was far below the
fluxes reported by those experiments claiming signals
earlier.\cite{Borione} There seems to be only two possibilities: either
{\it all} the earlier experiments claiming signals were in error, or
Cygnus X-3 and other reported sources all had turned off about the time
improved instrumentation became available. Table I offers some support
for the latter possibility, because (a) the phases of the signals are in
rough agreement in three experiments, and (b) the integrated flux above
a PeV does appear to systematically decrease over time taking all
experiments together. (Among those claiming signals only those claiming
more than $4\sigma$ have been listed, and among those citing upper
limits only those giving upper limits on the flux above a PeV have been
listed.) The suggestion that signals from Cygnus X-3 have fallen with
time was first raised by N. C. Rana et al. based on X-ray and gamma ray
data in four different wavelength regions.\cite{Rana} In what follows,
we make the ``optimistic" assumption that earlier experiments were
seeing real signals, and we consider to what extent those reports of
signals from Cygnus X-3 support the prediction of a 4.5 PeV neutron
spike.
In the 1980's there were eight cosmic ray groups that cited fluxes in
the PeV range of signals pointing back to Cygnus X-3, (some which were
inconsistent as mentioned earlier.) In nearly all cases limited
statistics required reporting the flux integrated over energy in only
one or at most two energy intervals.
\begin{table}[hbt]
\begin{center}
\begin{tabular}{ r r r r r r}
Ref & Years & E in PeV & Flux & Stat. sig. & Phase\\
\hline
\cite{Samorski} & 76-79 & $>2$ & $7.4\pm 3.2$ & $4.4\sigma$ & 0.1-0.3 \\
\cite{Hayashida}& 78-81 & $>1$ & $<3$ & & \\
\cite{Lloyd} & 79-82 & $>3$ & $1.5\pm0.3$ & $5\sigma$ & 0.225-0.25\\
\cite{Muraki} & 86-88 & $>1$ & $2.7\pm0.5$ & $4.7\sigma$ & 0.25-0.30\\
\cite{Cronin} & 89 & $>1$ & $<23$ & & \\
\cite{Borione} & 90-95 & $>1.175$ & $<0.1$ & & \\
\end{tabular}
\end{center}
\caption{Experiments reporting integrated fluxes (or upper
limits) in units of $\times 10^{-14}$ particles cm$^{-2}$ sec$^{-1}$ for
Cygnus X-3 for PeV energies. Only experiments reporting nonsporadic
signals claimed to be at a level of more than $4\sigma$ have been
listed. The van der Klis and Bonnet-Bidaud ephemeris has been used for
finding the phase interval in each case.} \end{table}
One group (Lloyd Evans et al.\cite{Lloyd}), however, had good enough
statistics to report fluxes in eight energy bins spanning the location
of the predicted 4.5 PeV neutron spike, and it had an energy acceptance
threshold near $E_0$ = 1 PeV, which could give one energy bin before the
spike itself. The signal seen by Lloyd Evans et al. from Cygnus X-3 did
not appear until the data is selected on the basis of orbital phase
determined from the X-ray binary's 4.79 h orbital period, and the time
of signal arrival. Lloyd-Evans et al. found that if they looked at the
number of counts in 40 phase bins, one of these bins showed a sizable
excess (73 counts when the average was 39). The information in Table II
is taken from Lloyd-Evans et al.\cite{Lloyd}, with the last column added by
this author. Fig. 1 displays the data in that last column. We would
expect a flat distribution on the basis of chance, assuming that the
signal were just a statistical fluctuation. In fact, averaged over all
phases, the distribution must be flat and zero height, regardless of
whether the signal is real or not. Note, that a spike appears centered
on the value predicted by the tachyonic neutrino model, and that all the
remaining bins have a flux consistent with zero. The gaussian curve
drawn with arbitrary height in the figure shows what would be predicted
by the model given a neutron spike of width $\Delta\log E = 0.1 (FWHM)$
and a 50 percent energy resolution ($\Delta\log E = \pm 0.176$).
According to Lloyd-Evans, the actual resolution
was probably around 50 percent, and very likely less than 100
percent\cite{informal} We estimate the statistical
significance of this spike occurring by chance by dividing the excess
number of events in the two bins straddling 5 PeV by the square root of
the expected number of events in those two bins:
$28.4/\sqrt{22.6}=6.0\sigma.$ It is interesting that in their article,
Lloyd-Evans et al. displayed only the integrated flux $I(>E)$ versus
energy, and hence failed to mention the spike. Instead, they simply
noted that the integrated spectrum appeared to steepen right after 10
PeV.
How can we be sure that the spike seen in Lloyd-Evans et al. data is not an
artifact of the data analysis or a statistical fluctuation? Six standard
deviations
may seem interesting, but the original peak in their phase plot was far
less impressive, particularly allowing for a ``trials factor" of 40,
since such a peak might have been seen in any one of the 40 phase bins.
Suppose that in fact the original peak in the phase plot were a
statistical fluctuation, how could one then get a $6\sigma$ peak in the
flux versus energy distribution for events in a specific phase bin?
Clearly, such a peak would require some correlation between energy and
phase. This could in principle occur, because observed cosmic ray
energy is correlated with declination angle, and hence with time of day.
However, all cosmic rays in a given phase bin arrive at one of five,
i.e., 24/4.79, times throughout the day, and those arrival times slowly
advance from day to day, since the Cygnus X-3 period is not exactly
divisible into 24 hours. Thus, over the years of data-taking each phase
bin would sample times of the day with an almost uniform distribution,
making it difficult to see how a phase-energy correlation could occur.
\begin{table}[hbt]
\begin{center}
\begin{tabular}{ r r r r}
E (in PeV) & Observed & Expected & Excess $\pm\sqrt{Expected}$\\
\hline
1-3 & 16 & 13.9 & 2.1$\pm$ 3.7 \\
3-5 & 34 & 16.4 & 17.6$\pm$ 4.0 \\
5-11 & 17 & 6.2 & 10.8$\pm$ 2.5 \\
11-18 & 4 & 2.4 & 1.6$\pm$ 1.6 \\
18-36 & 3 & 4.3 & $-$1.3$\pm$ 2.1 \\
36-72 & 6 & 3.4 & 2.6$\pm$ 1.8 \\
72-140 & 2 & 0.8 & 1.2$\pm$ 0.9 \\
$>$140 & 0 & 0.6 & $-$0.6$\pm$ 0.8 \\
\end{tabular}
\end{center}
\caption{Observed and expected event counts reported by Lloyd-Evans et
al. in differential energy bins for the phase interval 0.225-0.250. The
last column has been added by the author. The ``Expected" counts for
each energy interval are based
on the average over all phases.}
\end{table}
\begin{figure}[hbt]
\begin{center}
\leavevmode
\epsfxsize=3.25in
\epsffile{output7.ps}
\caption{Data points from the last column of Table II plotted at the
middle of each interval in log E in PeV. The gaussian curve
centered on 4.5 PeV is what one would expect to find in Lloyd-Evans
data, given a neutron spike of width $\Delta\log E = 0.1$(FWHM), and a
50 percent energy resolution ($\Delta\log E = \pm 0.176$)}
\end{center}
\end{figure}
(It could be that at their source
the phase and energy of cosmic rays are correlated, but in that case we would
be dealing with a real source, not a statistical fluctuation, as
hypothesized above.)
Ideally, one would want to combine the Lloyd-Evans et al. data with that
of other experiments in the PeV region to see if
the spike either is destroyed or enhanced. Several problems arise with
the other existing data, in which a signal is claimed from Cygnus X-3:
one experiment used only ``muon-poor" events\cite{Kifune}, two
experiments reported only the integral flux above some energy (no energy
bin defined)\cite{Kirov,Baltrusaitis}, two reported the flux in an
energy bin three times the width used by
Lloyd-Evans\cite{Samorski,Tonwar}, and none was contemporaneous with
Lloyd-Evans, thereby severely diminishing their utility.
Aside from the spike, one other prediction of the tachyonic neutrino
model is that neutrons should also be seen as a significant and rising
fraction of the cosmic ray flux above around 1.0 EeV. In fact, two
cosmic ray groups have reported seeing neutral particles from Cygnus X-3
having energies above 0.5 EeV with fluxes of $1.8\pm 0.7$\cite{Teshima},
and $2.0\pm 0.6$\cite{Cassiday}, while a third group reporting merely an
upper limit to the flux $<0.4$\cite{Lawrence} -- all in units of
$10^{-17}$ particles cm$^{-2}$ s$^{-1}$.
These measured fluxes above 0.5 EeV can be compared directly with the
neutron flux predictions from the tachyonic neutrino
model.\cite{Ehrlich} As noted previously, the ratio of the integral flux
of neutrons above 0.5 EeV to that above 2 PeV is predicted to be $R =
4.8\times 10^{-4}.$ The predicted neutron flux for $E > 0.5$ EeV is
then $R$ times the measured flux reported by Lloyd-Evans et al. for $E >
2$ PeV, or: $R\times 7.4\pm 3.2 \times 10^{-14} = 3.5 \pm
1.5 \times 10^{-17}$ particles cm$^{-2}$ s$^{-1}$, which is in quite
good agreement with the two groups that measured a flux, rather than an
upper limit. Although subsequent data accummulation by these two groups
failed to show a signal from Cygnus X-3\cite{informal1}, that only adds
additional support to the hypothesis that the source faded over time.
If it is true that Cygnus X-3 and other point sources were active in the
early 1980's and subsequently have turned off, is there any way to check
whether there really is a 4.5 PeV neutron spike without waiting for
specific sources to come back on? Without knowing where the sources
are, the model can make no prediction of the
anisotropy or the the angular
distribution of sources of high energy cosmic rays. However, recall that the model
predicts that {\it all} the cosmic rays include a 4.5 PeV neutron spike,
not just those pointing back to the handful of possible sources looked
at so far. Thus, if one selects events in a narrow energy band centered
on 4.5 PeV, one could look at their arrival directions on the two
dimensional map of the sky, and see if there is a noticeable clustering
of points, which would indicate neutral particles coming from specific
sources. Moreover, if those sources were episodic, one should observe a
nonuniform distribution in arrival times for events for a given source.
Consider a specific example. The integrated flux in the 4.5 PeV spike
is 0.1 neutrons per m$^2$-sr-s, which would give around 3 million counts
over 5 years for an array of area 250,000 m$^2$. If the array had an
energy resolution of 100 percent, it would also record a background
count rate roughly four times as great in the energy bin centered on 4.5
PeV. Suppose the angular resolution were $\Delta\theta = 0.01$ rad, which
would allow up to $4/{\Delta\theta}^2 = 4\times 10^4$ solid angle bins
to be defined. Each bin would then have on the average 400 background
counts. Further suppose that the cosmic rays reaching Earth came from N
point sources, then those solid angle bins pointing back to sources
would have an average signal to background ratio: $10^4/N.$
Identification of sources should then be possible, unless N were larger
than the number of solid angle bins, and no subset of sources were
appreciably brighter than others.
\section{Summary}
In summary, a highly speculative tachyonic neutrino model\cite{Ehrlich},
which fits the cosmic ray spectrum well, predicts a spike of neutrons at
an energy where, given the neutron lifetime and distance to likely
sources, very few should appear. A search through the literature for
sources of neutral cosmic rays has identified a particular experiment
with a favorable energy acceptance threshold, good enough statistics,
and enough energy bins spanning the region of the neutron spike to test
the prediction. The data do show a $6\sigma$ spike located right at the
predicted energy, which was not identified in the original work. The
failure of other subsequent more sensitive experiments to see a signal
from Cygnus X-3 would seem to require that this source has since turned
off -- a possibility given some support by both time trends of data from
different experiments, and data within the same experiments. The
characteristics of the neutral particles from Cygnus X-3 seem to be
consistent with neutrons rather than gamma rays, based on muon data from
various experiments. For the EeV region, where the model also predicts
neutrons (though not a spike), two out of three experiments show a
positive signal from Cygnus X-3, and they report a flux whose magnitude
(relative to the flux in the spike) is well-predicted by the model. The
hypothesis that the electron neutrino is a tachyon would seem to be
supported, and it can be further tested without waiting for specific
point sources to come back on.
|
\section{Introduction}
The recent production and observation of antihydrogen ($\overline{\rm{H}}$)
\cite{oelert,mandel}
opens new possibilities for
precision tests of CPT symmetry.
The two-photon 1S-2S transition frequency
has been measured
to $3.4$ parts in $10^{14}$
in an atomic beam of hydrogen (H)
\cite{hansch}
and to about one part in $10^{12}$
in trapped H.\cite{cesar}
It is hoped that
an eventual measurement of the line center
to about $1$ mHz,
corresponding to a resolution of one part in $10^{18}$,
would be possible.\cite{hanschICAP}
If such precisions could also be achieved
in the spectroscopy of $\overline{\rm{H}}$,
comparisons of corresponding frequencies in
H\ and $\overline{\rm{H}}$\
could yield stringent tests of CPT symmetry.
Current proposals for
$\overline{\rm{H}}$\ spectroscopy
involve both beam and trapped-atom
techniques,\cite{mandel2,gab2}
and are faced with a number of
outstanding challenges
including
the issue of achieving these precisions
in {\em trapped} H\ and
$\overline{\rm{H}}$.\cite{ce}
We consider the
theoretical prospects for placing appropriate bounds
on CPT and Lorentz violation
in experiments involving the
spectroscopy of free
or magnetically trapped H\ and $\overline{\rm{H}}$.
All local Lorentz-invariant
quantum field theories of point
particles,
including the standard model
and quantum electrodynamics (QED),
are invariant under the discrete symmetry
CPT.\cite{cpt}
Attempts to produce a fundamental theory
involving gravity often involve
string theory
and the spontaneous breaking of these symmetries\cite{kps}
and,
in these investigations,
the status of CPT symmetry is far less clear.
Observable effects of CPT breaking
are already known to be small,
and so it is reasonable to assume they
would be suppressed by at least one power
of the low-energy scale to Planck scale ratio.
Thus,
their detection could occur
only in extremely sensitive experiments.
In this proceedings,
we show that effects of this type
can appear in H\ and $\overline{\rm{H}}$\ spectra
at zeroth order in the fine-structure constant.
In addition,
these effects are theoretically detectable not only
in 1S-2S lines but also in hyperfine transitions.
The framework of our analysis is
an extension of the standard model and QED\cite{ck}
that includes spontaneous CPT and Lorentz
breaking at a more fundamental level.
Desirable features of this microscopic theory
appear to include energy-momentum conservation,
gauge invariance,
renormalizability,
and microcausality.\cite{ck}
Analyses in the context of this theoretical framework
have been done for
photon properties,\cite{ck}
neutral-meson experiments,\cite{kps,ckpv,expt,ak}
Penning-trap tests,\cite{bkr}
and baryogenesis.\cite{bckp}
\section{Free H\ and $\overline{\rm{H}}$}
We first consider
the spectra of \it free \rm
H\ and $\overline{\rm{H}}$.
For H,
the electron of mass $m_e$
and charge $q = -|e|$
in the proton Coulomb potential
$A^\mu = (|e|/4 \pi r, 0)$
is described by a modified Dirac equation
arising from
the standard-model extension.
Taking $i D_\mu \equiv i \partial_\mu - q A_\mu$,
the four-component electron field $\psi$
satisfies
\begin{equation}
\left( i \gamma^\mu D_\mu - m_e - a_\mu^e \gamma^\mu
- b_\mu^e \gamma_5 \gamma^\mu
- {\textstyle{1\over 2}} H_{\mu \nu}^e \sigma^{\mu \nu}
+ i c_{\mu \nu}^e \gamma^\mu D^\nu
+ i d_{\mu \nu}^e \gamma_5 \gamma^\mu D^\nu \right) \psi = 0
\label{dirac}
\end{equation}
in units with $\hbar = c = 1$.
CPT is violated by
the two terms involving the couplings
$a_\mu^e$ and $b_\mu^e$,
while CPT is preserved by
the three terms involving
$H_{\mu \nu}^e$, $c_{\mu \nu}^e$, and $d_{\mu \nu}^e$.
Lorentz invariance is broken by
all five couplings,
which are assumed to be small.\cite{ck}
Free protons are also described by
a modified Dirac equation\cite{bkr}
with corresponding couplings
$a_\mu^p$, $b_\mu^p$, $H_{\mu \nu}^p$, $c_{\mu \nu}^p$,
and $d_{\mu \nu}^p$.
It is possible to eliminate
various combinations of these quantities
through suitable field redefinitions.
In the following,
we keep all couplings,
thus showing explicitly that these expressions
are unobservable.\cite{ck}
Observable effects in the spectra of
free H\ and $\overline{\rm{H}}$\
can be studied using
perturbative calculations
in the context of relativistic quantum mechanics.
In this calculation,
the unperturbed hamiltonians
and their eigenfunctions are
identical for H\ and $\overline{\rm{H}}$.
In addition,
all perturbative effects
from conventional quantum electrodynamics
are also the same in both systems.
However,
the perturbations arising from
the CPT- and Lorentz-breaking couplings
for the electron in H\
can differ from those
for the positron in $\overline{\rm{H}}$.
These perturbations are obtained
from Eq.\ \rf{dirac}
by a standard method involving charge conjugation
(for $\overline{\rm{H}}$)
and field redefinitions.\cite{bkr}
Similarly,
additional energy perturbations
are generated by
the CPT- and Lorentz-breaking couplings
for the proton and antiproton,
and can be obtained to leading order via
relativistic two-fermion techniques.\cite{D2}
Let the (uncoupled) electronic and nuclear angular momenta
be denoted by
$J=1/2$ and $I=1/2$
respectively,
with third components $m_J$, $m_I$.
Using a perturbative calculation,
the energy corrections for the basis states $\ket{m_J,m_I}$
can be found.
For protons or antiprotons,
we find that the leading-order energy corrections
for spin eigenstates
have the same mathematical form as those
for electrons or positrons,
except for the replacement of superscripts $e$ with $p$
on the CPT- and Lorentz-violating couplings.
In H,
we find that the
leading-order energy shifts in
the 1S level are identical to those
in the 2S level.
Taking $m_p$ for the proton mass,
the shifts are
\begin{eqnarray}
\Delta E^{H} (m_J, m_I)
& \approx &
(a_0^e + a_0^p - c_{00}^e m_e - c_{00}^p m_p)
\cr
&&
+ (-b_3^e + d_{30}^e m_e + H_{12}^e) {m_J}/{|m_J|}
\cr
&&
+ (-b_3^p + d_{30}^p m_p + H_{12}^p) {m_I}/{|m_I|} ~ .
\label{EHJI}
\end{eqnarray}
Similarly,
for $\overline{\rm{H}}$,
the leading-order energy shifts
$\Delta E^{ \overline{H}}$
in the 1S levels
are identical to those in
the 2S levels,
and are
given by the expression \rf{EHJI}
with the substitutions
$a_\mu^e \rightarrow - a_\mu^e$,
$d_{\mu \nu}^e \rightarrow - d_{\mu \nu}^e$,
$H_{\mu \nu}^e \rightarrow - H_{\mu \nu}^e$;
$a_\mu^p \rightarrow - a_\mu^p$,
$d_{\mu \nu}^p \rightarrow - d_{\mu \nu}^p$,
$H_{\mu \nu}^p \rightarrow - H_{\mu \nu}^p$.
We note that because Eq.~(\ref{EHJI}) contains
spatial components of the couplings,
it would be necessary to take into account the geometry
when comparing results from different
experiments.
For example,
measurements taken at different times of the day
would be sensitive to different projections
of the couplings due to the rotation of the Earth.
The electron and proton spins in H\
are coupled by the hyperfine interaction,
and this is also the case for the
positron and antiproton spins
in $\overline{\rm{H}}$.
The total angular momentum $F$
must be considered,
and the appropriate basis states become
linear combinations $\ket{F,m_F}$
of the $\ket{m_J,m_I}$ states.
The allowed two-photon 1S-2S transitions
satisfy the selection rules
$\Delta F = 0$ and $\Delta m_F = 0$.\cite{cagnac}
There are thus four allowed transitions
for both H\ and $\overline{\rm{H}}$,
those for which the spins remain unchanged.
However,
no leading-order effects appear in the frequencies
of any of these transitions,
because according to Eq.\ \rf{EHJI}
the 1S and 2S states with identical spin configurations
have identical leading-order energy shifts.
Thus in the present theoretical context,
there are no signals of
Lorentz or CPT violation in
free H\ or in free
$\overline{\rm{H}}$\ at leading-order in 1S-2S
spectroscopy.
This agrees with results found previously\cite{bkr}
for the Penning trap,
showing that observable CPT-violating effects must also involve
CT violation and a spin-flip.
To overcome this limitation,
one could consider
the dominant subleading energy-level shifts
involving the CPT- and Lorentz-breaking couplings
in free H\ and $\overline{\rm{H}}$.
These would be hard to detect because they
arise as relativistic corrections of order $\alpha^2$.
They do, however,
differ for some of the 1S and 2S levels
and therefore observable effects
could in principle occur.
An example is the term proportional to
$b_3^e$ in Eq.\ \rf{dirac},
which produces a frequency shift in the
$m_F = 1 \rightarrow m_{F^\prime} = 1$ line
relative to the $m_F = 0 \rightarrow m_{F^\prime} = 0$ line
(which remains unshifted),
given by
\begin{equation}
\delta \nu^H_{1S-2S} \approx - \alpha^2 b_3^e / 8 \pi
\quad .
\end{equation}
A similar suppression by a factor
at least of order
$\alpha^2 \simeq 5\times 10^{-5}$
would occur in
the proton-antiproton corrections.
As a result of these suppressions,
Penning-trap $g-2$ experiments
are likely to be more sensitive to
some of the CPT- and Lorentz-violating quantities
than experiments involving 1S-2S spectroscopy
in free H\ and $\overline{\rm{H}}$.
In fact,
the estimated attainable bound\cite{bkr}
on $b_3^e$
obtained
with existing technology
in anomaly-frequency comparisons
with electron-positron Penning-trap experiments
would suffice to place a bound
of $\delta \nu^H_{1S-2S} \mathrel{\rlap{\lower4pt\hbox{\hskip1pt$\sim$} 5$ $\mu$Hz
on observable shifts of the 1S-2S frequency
in free H\ from the electron-positron sector.
This is beyond the resolution of 1S-2S spectroscopy.
For the proton-antiproton quantities
in the standard-model extension,
experiments have not yet been performed,
but bounds attainable would
also yield tighter constraints on these
parameters than would be possible in 1S-2S spectroscopy.
It is relevant to ask why
$g-2$ experiments
are potentially more sensitive to
observable effects than
comparisons of 1S-2S
transitions in free H\ and $\overline{\rm{H}}$.
This is surprising because
the conventional figure of merit for CPT breaking
in electron-positron $g-2$ experiments,\cite{pdg}
\begin{equation}
r_g = |g_{e^-} - g_{e^+}|/g_{\rm av} \mathrel{\rlap{\lower4pt\hbox{\hskip1pt$\sim$} 2
\times 10^{-12}
\quad ,
\end{equation}
is six orders of magnitude weaker
than the idealized resolution
of the 1S-2S line,
$\Delta \nu_{1S-2S}/\nu_{1S-2S} \simeq 10^{-18}$.
However,
the figure of merit $r_g$ in Penning-trap
$g-2$ experiments is inappropriate
in the present theoretical context.\cite{bkr}
The point is that the
experimental sensitivity
to CPT- and Lorentz-violating effects
is determined by
the absolute frequency resolution
for unsuppressed transitions.
The idealized 1S-2S line-center resolution
is about 1 mHz,
which would appear to be better than
the 1 Hz absolute frequency resolution
in $g-2$ measurements.
However,
$g-2$ experiments are
directly sensitive to $b_3^e$
because they involve spin-flip transitions,
whereas the 1S-2S transitions
in free H\ or $\overline{\rm{H}}$\
are sensitive only to the suppressed combination
$\alpha^2 b_3^e/8\pi$.
As a result,
the bound on $b_3^e$
from electron-positron $g-2$ experiments
is thus about two orders of magnitude sharper
than that from 1S-2S comparisons.
In addition to the 1S-2S transition,
there are certainly others
available in H\ and $\overline{\rm{H}}$.
The above discussion suggests
that transitions
between states with different spin configurations
might yield tighter bounds.
Such experiments would require external fields
to select particular spin states.
\section{Trapped H\ and $\overline{\rm{H}}$}
We next consider spectroscopy of
H\ or $\overline{\rm{H}}$\
in the presence of a uniform magnetic field.
A way to do this is by confining the particles
in a magnetic trap
such as an Ioffe-Pritchard trap,\cite{ip}
and imposing an axial bias magnetic field.
The situation is directly relevant to
proposed experiments.\cite{gab2}
In the following,
we denote each of the 1S and 2S hyperfine Zeeman levels
in order of increasing energy
in a magnetic field $B$ by
$\ket{a}_n$, $\ket{b}_n$, $\ket{c}_n$, $\ket{d}_n$,
with $n=1$ or $2$,
for both H\ and $\overline{\rm{H}}$.
In the case of H,
the four states expressed in terms of the
basis states $\ket{m_J,m_I}$ are
\begin{eqnarray}
\ket{d}_n &=& \ket{{\textstyle{1\over 2}}, {\textstyle{1\over 2}}} \quad,
\nonumber \\
\ket{c}_n &=& \sin \theta_n \ket{-{\textstyle{1\over 2}},{\textstyle{1\over 2}}} +
\cos \theta_n \ket{{\textstyle{1\over 2}},-{\textstyle{1\over 2}}}
\quad ,
\nonumber\\
\ket{b}_n &=& \ket{-{\textstyle{1\over 2}}, -{\textstyle{1\over 2}}} \quad,
\nonumber \\
\ket{a}_n &=& \cos \theta_n \ket{-{\textstyle{1\over 2}},{\textstyle{1\over 2}}} -
\sin \theta_n \ket{{\textstyle{1\over 2}},-{\textstyle{1\over 2}}}
\quad .
\label{a}
\end{eqnarray}
The mixing angles $\theta_n$
are functions of the magnetic field,
and are different for the 1S and 2S states:
\begin{equation}
\tan 2 \theta_n \approx \fr{(51 {\rm ~mT})}{n^3B}
\quad .
\end{equation}
The states
$\ket{c}_1$ and $\ket{d}_1$
are low-field seekers,
and in principle remain confined
near the magnetic-field minimum of the trap.
However,
a population loss occurs
due to spin-exchange collisions
$\ket{c}_1 + \ket{c}_1 \rightarrow \ket{b}_1 + \ket{d}_1$
of the $\ket{c}_1$ states over time,
so that primarily $\ket{d}_1$ states
are confined.
A transition that would seem natural to consider
is that between the unmixed-spin states
$\ket{d}_1$ and $\ket{d}_2$
because it is field independent
for practical values of the magnetic field.
The idea would be to compare
the frequency $\nu^H_d$
for the 1S-2S transition $\ket{d}_1 \rightarrow \ket{d}_2$
in H\
with the frequency $\nu^{\overline{H}}_d$
for the corresponding spectroscopic line in $\overline{\rm{H}}$.
But,
in H\ the spin configurations of the
$\ket{d}_1$ and $\ket{d}_2$ states are the same,
so any shifts occurring are again suppressed.
The same is true for $\overline{\rm{H}}$,
and so we find
\begin{equation}
\delta \nu^H_d = \delta \nu^{\overline{H}}_d \simeq 0
\end{equation}
at leading order.
Another transition of theoretical interest
would be the 1S-2S transition
$\ket{c}_1 \rightarrow \ket{c}_2$ in H\
and the analogous $\overline{\rm{H}}$\ transition.
The point would be to exploit
the spin mixing of these states in a nonzero magnetic field.
An unsuppressed frequency shift
would arise because the hyperfine splitting
depends on $n$,
thus producing a spin difference
between the 1S and 2S levels
in this 1S-2S transition
between $\ket{c}_1$ and $\ket{c}_2$:
\begin{equation}
\delta \nu_c^H \approx
-\kappa (b_3^e - b_3^p - d_{30}^e m_e
+ d_{30}^p m_p - H_{12}^e + H_{12}^p)/2\pi ~.
\label{nucH}
\end{equation}
In this expression,
$\kappa$ is a spin-mixing function
given by
\begin{equation}
\kappa\equiv \cos 2\theta_2 - \cos 2\theta_1
\quad .
\end{equation}
This function is always less than one,
so to avoid losing sensitivity
the optimal situation would involve the largest
possible value.
This maximum is $\kappa \simeq 0.67$
and occurs at $B \simeq 0.011$~T,
as illustrated in Figure~\ref{kapfunc}.
\begin{figure}[t]
\hspace*{1.4cm}
\centerline{
\psfig{figure=kappas.ps,height=3in,width=3in}}
\caption{The dimensionless functions $\kappa$ and $\hat{\kappa}$.
For $\kappa$, the maximum value of approximately 0.67 occurs
at about 0.011~T.
The function $\hat{\kappa}$ increases to within
about two percent of its asymptotic value (one)
as the magnetic field is increased from zero
to $0.25$ Tesla.
\label{kapfunc}}
\end{figure}
The corresponding 1S-2S shift
$\delta \nu_c^{\overline{H}}$ for $\overline{\rm{H}}$\
in the same magnetic field can also be found.
Relative to a fixed magnetic field,
the hyperfine states in $\overline{\rm{H}}$\
have opposite positron and antiproton spins
compared to the electron and proton spins in H.
As a result,
the expression for
$\delta \nu_c^{\overline{H}}$ is
identical to that for $\delta \nu_c^H$ in Eq.\ \rf{nucH}
except that the signs of $b_3^e$ and $b_3^p$ are changed.
The frequencies $\nu_c^H$ and $\nu_c^{\overline H}$
depend on spatial components of Lorentz-violating couplings
and would therefore vary diurnally
in the comoving Earth frame.
Another effect would be an istantaneous difference
\begin{equation}
\Delta \nu_{1S-2S,c} \equiv \nu_c^H
- \nu_c^{\overline{H}} \approx - \kappa (b_3^e - b_3^p)/\pi
\label{delcc}
\end{equation}
for measurements made in the same magnetic trapping fields.
The transition $\ket{c}_1 \rightarrow \ket{c}_2$
when compared with
the transition $\ket{d}_1 \rightarrow \ket{d}_2$
is theoretically more sensitive to
CPT and Lorentz violation
by a factor of order
$4/\alpha^2 \simeq 10^5$.
However,
the 1S-2S transition
$\ket{c}_1 \rightarrow \ket{c}_2$
in H\ and $\overline{\rm{H}}$\
depends on the magnetic field,
and the resultant Zeeman broadening
due to the inhomogeneous trapping fields
would have to be overcome.
Even at a temperature of $100 \mu$K,
the transition
in both H\ and $\overline{\rm{H}}$\
would be broadened to over 1 MHz
for $B\simeq 10$ mT.
This would severely hinder
the experimental attainment of
resolutions on the order of the natural line width.
Figure \ref{1s2sfig} illustrates one case for the conventional and
perturbed frequencies in the four 1S-2S transitions.
In this figure, $b_3^p>0$ and all the other couplings are zero.
\begin{figure}[t]
\hspace*{3cm}
\centerline{
\psfig{figure=brnu1S2S.eps,height=3.4in}}
\caption{Conventional and perturbed frequencies
for the 1S-2S transition as a function of magnetic field.
The vertical scale is the shift in the
usual Bohr-model 1S-2S frequency of
about $2.5\times 10^{15}$~Hz.
The bold lines are for the conventional frequencies,
the fainter solid line is for the perturbed hydrogen transition
frequencies, and the dashed line is for
the perturbed antihydrogen frequencies.
We have taken $b_3^p>0$, with all other couplings zero.
The upper set of three lines represents the
$\ket{a}_1 \rightarrow \ket{a}_2$
transition,
and the lower set the
$\ket{c}_1 \rightarrow \ket{c}_2$
case.
The single straight line is for the
$\ket{b}_1 \rightarrow \ket{b}_2$
and
$\ket{d}_1 \rightarrow \ket{d}_2$ cases,
showing how these transitions are
field independent and
unperturbed by the $b_3^p$ coupling.
\label{1s2sfig}}
\end{figure}
\section{Hyperfine Transitions}
We now consider the possibilities for
spectroscopy of the hyperfine 1S levels.
Motivated by the fact that
transitions between the $F = 0$ and $F^\prime = 1$
hyperfine states can be measured with accuracies
better than $1$ mHz in a hydrogen maser,\cite{ramsey}
hyperfine transitions in masers and
in trapped H\ and $\overline{\rm{H}}$\
are worth considering
for tests of CPT and Lorentz symmetry.
The energy levels of all four hyperfine states
in the ground state of hydrogen
are shifted due to
CPT- and Lorentz-violating effects.
All the shifts contain an identical contribution
$a_0^e + a_0^p -c_{00}^e m_e -c_{00}^p m_p$
that leaves energy differences unaffected.
The remaining spin-dependent terms
are
\begin{eqnarray}
\Delta E_a^H &\simeq&
\hat\kappa (b_3^e - b_3^p - d_{30}^e m_e
+ d_{30}^p m_p - H_{12}^e + H_{12}^p)
\quad ,
\nonumber\\
\Delta E_b^H &\simeq&
b_3^e + b_3^p - d_{30}^e m_e
- d_{30}^p m_p - H_{12}^e - H_{12}^p
\quad ,
\nonumber\\
\Delta E_c^H &\simeq& -\Delta E_a^H
\quad , \qquad
\Delta E_d^H \simeq - \Delta E_b^H
\quad ,
\label{abcd}
\end{eqnarray}
where $\hat\kappa \equiv \cos2 \theta_1$.
If there is no magnetic field, then $\hat\kappa =0$
and the energies of $\ket{a}_1$ and $\ket{c}_1$ are unshifted.
However,
equal and opposite energy shifts occur for
$\ket{b}_1$ and $\ket{d}_1$.
The degeneracy of the three $F=1$ ground-state
hyperfine levels is therefore removed even for
$B=0$.\footnote{No conflict with Kramer's theorem occurs
in the breaking of the
$\ket{b}$-$\ket{d}$ degeneracy at zero field,
because the Lorentz-violating coefficients
in Eq.~\rf{abcd} break time-reversal symmetry.
A possible method of detecting the splitting might
involve looking directly for a difference frequency.}
For instance,
the
$\ket{d}_1 \rightarrow \ket{a}_1$
and $\ket{b}_1 \rightarrow \ket{a}_1$
transitions differ in their frequencies
by the unsuppressed and diurnally varying
quantity
\begin{equation}
|\Delta \nu_{d-b}^H| \approx
|b_3^e + b_3^p - d_{30}^e m_e - d_{30}^p m_p
- H_{12}^e - H_{12}^p|/\pi
\quad .
\end{equation}
In the presence of a magnetic field,
all four hyperfine Zeeman energy levels are shifted.
For the $\ket{a}_1$ and $\ket{c}_1$ states,
the spin-mixing function $\hat\kappa$
controls the shifts.
As $B$ increases from zero,
$\hat\kappa$ increases,
attaining $\hat\kappa \simeq 1$ when $B \simeq 0.3$ T.
The function $\hat\kappa$ is illustrated in
Fig.~\ref{kapfunc}.
The shifts in the energy levels as given in
Eq.~\rf{abcd} are partially illustrated in
Figure~\ref{fig1}.
\begin{figure}[t]
\vskip 2cm
\hspace*{2cm}
\centerline{
\psfig{figure=hyperfine.eps,height=3.0in}}
\caption{Hyperfine levels for 1S states
versus magnetic field.
The vertical axis represents the
shift in energy (in frequency units) relative to the
usual Bohr-model $n=1$ energy of $-13.6$~eV.
The bold solid line is for the unperturbed case,
the finer solid line and the dashed lines are
for hydrogen and antihydrogen respectively.
We have taken $b_3^p>0$
and all other couplings zero.
\label{fig1}}
\end{figure}
The usual H\ maser employs
a small ($B \mathrel{\rlap{\lower4pt\hbox{\hskip1pt$\sim$} 10^{-6}$ T) magnetic field
and works with the
field-independent $\sigma$ transition
$\ket{c}_1 \rightarrow \ket{a}_1$.
The leading-order effects from
CPT and Lorentz violation in high-precision
measurements of this line
$\ket{c}_1 \rightarrow \ket{a}_1$ are suppressed,
because for this situation $\hat\kappa \mathrel{\rlap{\lower4pt\hbox{\hskip1pt$\sim$} 10^{-4}$.
However,
a shift $\Delta \nu_{d-b}^H$ does occur in
the frequency difference between the
field-dependent transitions $\ket{d}_1 \rightarrow \ket{a}_1$
and $\ket{b}_1 \rightarrow \ket{a}_1$
relative to the conventional value,
and the associated diurnal variations
would provide an unsuppressed signal
of CPT and Lorentz violation.
The resolution of this difference
would be reduced by
broadening due to field inhomogeneities.
In addition, it would be necessary to
distinguish it from possible backgrounds
due to residual Zeeman splittings.
The direct comparison of transitions
between hyperfine Zeeman levels in H\ and $\overline{\rm{H}}$\
could address the issue of background splittings.
Moreover,
the magnetic-field dependence of the frequency
could be eliminated to first order
by working at an appropriate value of the field.
One option might be to consider
high-resolution spectroscopy
at the field-independent transition point $B \simeq 0.65$~T
on the $\ket{d}_1 \rightarrow \ket{c}_1$ transition
in trapped H\ and $\overline{\rm{H}}$.
Experimental hurdles would include
Doppler broadening and
potentially larger field inhomogeneities
due to the relatively high bias field.
Obtaining frequency resolutions of order 1 mHz
would be a challenge,
requiring cooling to temperatures of order 100 $\mu$K
with a good signal-to-noise ratio
and a stiff box shape for the trapping potential.
At this bias-field strength,
the electron and proton spins
in state $\ket{c}_1$
interact more strongly with the field
than with each other
and are highly polarized with
$m_J = 1/2$ and $m_I = - 1/2$.
Thus, the transition $\ket{d}_1 \rightarrow \ket{c}_1$
is in essence a proton spin-flip.
For this transition,
we obtain frequency shifts
\begin{eqnarray}
\delta \nu_{c \rightarrow d}^H &\approx&
(-b_3^p + d_{30}^p m_p + H_{12}^p)/\pi
\quad ,
\nonumber\\
\delta \nu_{c \rightarrow d}^{\overline{H}} &\approx&
(b_3^p + d_{30}^p m_p + H_{12}^p)/\pi
\end{eqnarray}
for H\ and
$\overline{\rm{H}}$\ respectively.
One way to detect such terms
would be to search for diurnal variations
in the frequencies
$\nu_{c \rightarrow d}^H$
and
$\nu_{c \rightarrow d}^{\overline{H}}$.
Another possibility would be to consider
their instantaneous difference,
\begin{equation}
\Delta \nu_{c \rightarrow d} \equiv
\nu_{c \rightarrow d}^H - \nu_{c \rightarrow d}^{\overline{H}}
\approx - 2 b_3^p / \pi
\quad .
\label{nudiff}
\end{equation}
This difference
could provide a direct, clean, and sharp test
of the CPT-violating coupling $b_3^p$ for the
proton.
We can introduce
dimensionless figures of merit
appropriate for experiments investigating
various direct and diurnal-variation signals.
This is done in analogy with definitions
made for similar tests in Penning traps.\cite{bkr}
As an example,
a figure of merit for the signal in Eq.\ \rf{nudiff}
could be chosen as
\begin{eqnarray}
r^H_{rf,c \rightarrow d} & \equiv &
{|({\cal E}_{1,d}^H - {\cal E}_{1,c}^H)
- ({\cal E}_{1,d}^{\overline{H}} - {\cal E}_{1,c}^{\overline{H}})|}/
{{\cal E}_{1,{\rm av}}^H}
\nonumber \\
&\approx &
2\pi |\Delta \nu_{c \rightarrow d}| /m_H
\quad .
\label{rrf}
\end{eqnarray}
Here, ${\cal E}_{1,d}^H$, ${\cal E}_{1,c}^H$ and the
corresponding quantities for $\overline{\rm{H}}$\
are relativistic energies in ground-state hyperfine levels,
and $m_H$ is the atomic mass of H.
If, for example,
a frequency resolution of 1 mHz were attained,
this would correspond to an upper bound
of about
$r^H_{rf,c \rightarrow d} \mathrel{\rlap{\lower4pt\hbox{\hskip1pt$\sim$} 5 \times 10^{-27}$.
The CPT- and Lorentz-violating coupling $b_3^p$
would then be limited to $|b_3^p| \mathrel{\rlap{\lower4pt\hbox{\hskip1pt$\sim$} 10^{-18}$~eV.
This is about three orders of magnitude better
than estimated attainable bounds\cite{bkr}
from $g-2$ experiments in Penning traps
and more than four orders of magnitude better
than the limit attainable from 1S-2S
transitions.
We also note that the frequency resolution
of high-precision clock-comparison experiments,
which can also bound Lorentz violation,
lies below 1 $\mu$Hz.\cite{hd}
In these experiments,
leading-order bounds are obtained on $b_3^p$
in combination with other couplings.\cite{kla}
Since the nuclei involved are relatively complex,
the theoretical analysis prevents
$b_3^p$ from being isolated.
The experiments discussed here are sensitive
only to spatial components of CPT-violating couplings.
A boost would be needed to be sensitive to
timelike components such as
$b_0^e$,
and would also
enhance CPT- and Lorentz-violating
effects.\cite{ak}
This would be an advantage of the proposed experiments
\cite{mandel2}
measuring the fine structure and Lamb shift
with a relativistic beam of $\overline{\rm{H}}$.
Although they would probably have poorer resolutions
than the others discussed here,
constraints on $b_0^e$ and $b_0^p$ may be possible.
In conclusion,
we have shown that 1S-2S transitions
involving the mixed-spin $\ket{c}$ states
as well as the
spin-flip
$\ket{d}_1 \rightarrow \ket{c}_1$
hyperfine transition
could give rise to signals of
Lorentz and CPT violation
in magnetically confined
H\ or $\overline{\rm{H}}$\ atoms.
These signals would not be suppressed by powers
of the fine-structure constant.
They would indicate observable and
qualitatively new physics originating at the Planck scale.
\section*{Acknowledgments}
I thank Robert Bluhm and Alan Kosteleck\'y,
who collaborated on this work.
Partial support was provided by the U.S.\ D.O.E.\
under grant number DE-FG02-91ER40661 and by the N.S.F.\
under grant number PHY-9503756.
\section*{References}
|
\section{Introduction}
In a previous communication\cite{r1} we had described a cosmological scheme
which is consistent with observations and yet does not invoke dark matter. It
is ofcourse well known that the puzzle of galactic rotational velocities can
be explained by the dark matter hypothesis\cite{r2,r3}. Briefly put, using the
well-known relation for rotation under gravitation,
\begin{equation}
\frac{mV^2}{r} = \frac{GMm}{r^2}\label{e1}
\end{equation}
we would expect that from (\ref{e1}) the rotational velocities $V$ at the edges
of galaxies would obey the relation
\begin{equation}
V^2 = \frac{GM}{r}\label{e2}
\end{equation}
where $M$ is the galactic mass, $r$ the radius of the galaxy and $G$, the
gravitational constant. That is the velocities would fall off with increasing
distance from the centre of the galaxy. However, it is observed that these
velocities tend to a constant\cite{r3},
\begin{equation}
V \sim 300 Km/sec\label{e3}
\end{equation}
Alternatively, it is observed that the mass of the universe obeys the law\cite{r4},
\begin{equation}
M \propto R^n, n \approx 1\label{e4}
\end{equation}
These discrepancies can be explained if there is unobserved or unaccounted, that
is missing or dark matter, whose gravitational influence is nevertheless present.
This would also close the universe, that is the expansion would halt and a
collapse would ensue. While no dark matter has yet been discovered, it must
be mentioned that one candidate is a massive neutrino\cite{r5}. Recently, the
Superkamiokande experiments have yielded the first evidence for this\cite{r6},
but it is recognized that this mass, roughly of the order of a billionth that
of the electron is far too small to be the missing or dark matter.\\
On the other hand, latest observations of distant supernovae by different teams
of observers show that the universe would continue to expand for ever\cite{r7,r8}.\\
The cosmological scheme considered in reference \cite{r1} (cf. also ref.\cite{r9})
predicts precisely such a behaviour and moreover, explains (\ref{e4}) without
invoking dark matter. In this scheme, particles, typically pions are
fluctuationally created out of a background ZPF. (Other mysterious, hitherto
empirical relations, like Dirac's large number equations or the inexplicable
Weinberg pion-Hubble constant relation are deduced in this theory).\\
We will now show that in this cosmological scheme, not only the puzzling
galactic rotation relation (\ref{e3})is explained, but also the fact that
structures like galaxies and superclusters would naturally arise.
\section{Galactic Rotation}
We first observe that for a typical galaxy, the mass $'M'$ which is about
$10^{11}$ solar masses, is given by
\begin{equation}
M = Nm = 10^{70}.m\label{e5}
\end{equation}
where $'m'$ is the mass of a typical elementary particle, which in the
literature has been taken to be a pion\cite{r10} and $'N'$ their number.\\
We next observe that the size $'r'$ of a typical galaxy (about a $100000$ light
years) is given by,
\begin{equation}
r = \sqrt{N}l\label{e6}
\end{equation}
where $'l'$ is the pion Compton wavelength and $'N'$ is given in (\ref{e5}).\\
Finally in the cosmological scheme referred to (cf.ref.\cite{r1,r9}), we have,
\begin{equation}
G = a/\sqrt{\bar N}\label{e7}
\end{equation}
where $\bar N \sim 10^{80}$ is the number of pions in the universe and a
$\sim 10^{32}$.\\
Introducing (\ref{e5}), (\ref{e6}) and (\ref{e7}) into (\ref{e2}), we get for
the rotational velocity, because as we go to the edge, the number of particles
$\to N$, the relation (\ref{e3}), consistent with observations.
\section{Large Scale Structures}
It is quite remarkable that equation (\ref{e6}) is true for the universe itself,
as originally pointed out by Eddington, and also for superclusters, as can
be verified. Moreover (\ref{e6}) is a very general relation in the theory of
Brownian motion - it shows that the system under consideration could be
thought of as a collection of these elementary particles in random motion\cite{r11,r12}.
Further, (\ref{e6}) also shows that these structures have an overall flat or
two-dimensional character, which is indeed true\cite{r13}. In particular,
galaxies have vast flat discs and superclusters have a cellular character\cite{r2}.
It must also be pointed out that recent observations do indeed suggest such an
anisotropy\cite{r14}. Finally, the recently discovered neutrino masses are
small, in which case the particles have relativistic speeds, and this is known
to imply the above type of flat structures\cite{r3}.\\
On the other hand, (\ref{e6}) is not valid for stars. At these distance scales,
gravitation is strong enough so that the Brownian approximation (\ref{e6}) is
no longer valid.
\section{Conclusion}
We have thus explained without invoking dark matter, the galactic rotational
relation (\ref{e3}) and also have obtained a rationale for the existence of
structures like galaxies and superclusters.
|
\section{Introduction}
The experimental program of the LEP accelerator at CERN was foreseen to
proceed in two steps: a first period of running around the energy of the Z
boson, and a
second period at higher energy, having as main goals the production of W
boson pairs and the search for new particles.\par
W bosons can be studied at LEP in a unique environment. Fundamental ingredients
of the Standard Model\cite{sem} as carriers of the charged electroweak
interaction, these particles were discovered in 1983 in $p\bar{p}$ collisions
by the UA1 and UA2 collaborations at CERN\cite{ua1},\cite{ua2},\cite{wzdisc}.
Further,
more precise measurements were performed by the CDF and D0 experiments running
at the Tevatron collider at Fermilab \cite{cdfd0}.\par
At LEP it is therefore the
first time that W bosons are produced in the clean environment of leptonic
interactions. In the hadronic case the most copious source
of Ws is the Drell-Yan mechanism, with production and subsequent decay of
single Ws. Due to the large QCD background to the hadronic decay channel,
most of the measurements performed are relative to the cleaner decay channels
$W\to e\bar{\nu_e}$ and $W\to \mu\bar{\nu_\mu}$. In $e^+e^-$ interactions,
W bosons are mainly produced in pairs, and according to their decay WW
events can be classified as fully leptonic, semileptonic and fully hadronic.
Above the WW production threshold, all decay channels can be studied with
small background contamination, giving a broader picture of the physics of
these particles.\par
Measurements of WW and single W production cross sections can be performed,
as well as W decay branching ratios, providing a test of lepton universality
for charged current interactions.\par
As it will be discussed in more detail in the next sections, the W mass is
one of the fundamental parameters of the Standard Model. Its actual value
depends via radiative corrections from unknown parameters like the mass of the
Higgs boson, or on the presence of physics beyond the Standard Model.
The error on this quantity from the measurements performed at LEP2 is
presently the same as that coming from hadronic interactions, and
being still dominated by statistics, it will improve in the next years of
data taking. Preliminary studies \cite{yrep} have shown that with the target
luminosity of 500 $pb^{-1}$ LEP2 can reach a precision on this quantity
$\Delta M_W=50$ MeV, with a factor 2 improvement with respect to the present
measurements.\par
The main limitations to the accuracy achievable on the mass are coming from
the LEP energy measurement and to final state interactions leading to a
distortion of the reconstructed W mass. Since these effect can produce
sensible mass shifts, as well as modification in other observables, several
models have been proposed and tested with the available data.\\
The full reconstruction of final states is extremely important for the study of
Trilinear Gauge Couplings. S-channel production of W bosons occurs via diagrams
involving $\gamma WW$ and $ZWW$ vertices. A deviation from the standard model
for these vertices can modify WW production cross section, as well as
distributions of W production and decay angles. Constraints on anomalous
couplings are coming from the combined studies of WW events, as well as
single-W and single-$\gamma$.\par
If standard model couplings are assumed, W hadronic branching fractions are
proportional to squares of CKM matrix elements $|V_{ab}|^2$. The matrix element
$|V_{cs}|$ is in particular presently known with worse precision with respect
to the others; assuming the unitarity of the matrix and the knowledge of the
other matrix elements, a more precise determination of this quantity can be
derived from the WW hadronic branching fraction. A less precise but more
direct determination can be obtained from a charm tagging of the jets from
W decays, exploiting heavy-quark characteristics of charm in an environment
with small contamination from b quarks.
\section{Tree-level relations for gauge bosons}
The unified theory of weak and electromagnetic interaction is based
on the invariance of the Lagrangian under transformations of the
$SU(2)_L \times U(1)$ symmetry group. The three fields $(W^1_\mu,W^2_\mu,W^3_\mu)$
are connected to the weak isospin T, and the field $B_\mu$ to
the weak hypercharge Y. These quantum numbers are related to the electric
charge Q by $Q=T_3+Y/2$, where $T_3$ is the third component of the
weak isospin. The physical fields, the carriers of the charged (W$^\pm$) and
neutral (Z) weak currents, and of the electromagnetic current (A) are
linear combinations of the above
\[W^\pm_\mu=\frac{1}{\sqrt{2}}(W^1_\mu\mp iW^2_\mu)\]
\[\binom{A_\mu}{Z_\mu}=\begin{pmatrix}\cos\theta_W&\sin\theta_W\\
-\sin\theta_W&\cos\theta_W\end{pmatrix}\binom{B_\mu}{W^0_\mu}\]
The weak angle $\sin\theta_W$ introduced relates the coupling constants
of the $SU(2)_L$ and $U(1)$ interactions to the electric charge
\[g=\frac{e}{\sin\theta_W}\]
\[g'=\frac{e}{\cos\theta_W}\]
At low energies, the electroweak theory is equivalent to the Fermi
theory of the weak interactions. The Fermi constant can be expressed as
\[ G_F=\frac{g^2}{4\sqrt{2}M_W^2}=\frac{\pi\alpha_{QED}}{\sqrt{2}M_W^2
\sin^2\theta_W}\]
Since vector bosons behave as real particles, a gauge-invariant kinetic term
must be added to the Lagrangian describing electroweak interactions.
Due to the non-Abelian structure of the gauge field, the commutators of the
covariant derivatives involved do not vanish, but produce terms leading to
the self-interaction of the gauge bosons:
\[{\cal L}_W^{kin}=-\frac{1}{4}W^j_{\mu\nu}W_j^{\mu\nu}\]
\[W^j_{\mu\nu}=\partial_\mu W^j_\nu-\partial_\nu W^j_\mu+
g\epsilon^j_{km}W^k_\mu W^m_\nu\]
The last term involves the product of two W fields, so in the Lagrangian
terms
for trilinear and quadrilinear gauge boson interactions are present.\par
In the standard model the vector bosons acquire a mass by the spontaneous
breaking of the symmetry group, introducing a Higgs \cite{higgs} doublet
\[\binom{\Phi^+}{\Phi^0}=\binom{\Phi_1+i\Phi_2}{\Phi_3+i\Phi_4} \]
and choosing a vacuum expectation value $|\Phi^0|^2=v^2/2>0$.
The masses of the vector
bosons are then determined by this vacuum expectation value and the coupling
constants:
\[m_W=\frac{gv}{2},\hspace{2cm}m_Z=\frac{\sqrt{g^2+{g'}^2}}{2}v,\hspace{2cm}
m_\gamma=0\]
which leads to a relation between the masses of the vector bosons and the
weak angle
\[\cos^2\theta_W=\frac{m^2_W}{\rho m^2_Z}\]
The $\rho$ parameter is equal to unity at tree level. Deviations from
this value can arise from radiative corrections.\par
\section{Indirect determination of the W mass}
To obtain accurate predictions, the tree-level relations presented above
are no longer sufficient, but the evaluation of higher orders is needed,
especially given the accuracy of the available data.
An example is the possibility of extracting the W
mass from the standard model relations without directly measuring this
quantity.
From the equations shown in the previous section, it is possible to derive
\[\sqrt{2}G_FM^2_W(1-\frac{M^2_W}{\rho M^2_Z})=\pi\alpha\]
To include higher-order effects, different schemes can be used. The most
common approach \cite{ybooklep1} is to keep $\rho$ to its tree level
value 1 and include
all corrections in a quantity $\Delta r$, that accounts for both weak
and electromagnetic effects:
\[\sqrt{2}G_FM^2_W(1-\frac{M^2_W}{M^2_Z})=\frac{\pi\alpha}{1-\Delta r}\]
Since $\alpha$, $G_F$ and $M_Z$ are experimentally well known
quantities, it is possible to derive the W mass from
\[M_W^2=\frac{1}{2}M_Z^2(1+\sqrt{1-\frac{4A_0^2}{M_Z^2}\frac{1}{1-\Delta r}})\]
with
\[A_0^2=\frac{\pi\alpha}{\sqrt{2}G_F}\]
In the standard model, vertex and propagator corrections can be
decoupled into an
electromagnetic part, due to the running of the coupling constant
$\alpha_{QED}$, and a weak part, that contains terms showing a quadratic
dependence on the top mass and a logarithmic dependence on the Higgs mass:
\[\Delta r=\Delta\alpha+\frac{\cos^2\theta_W}{\sin^2\theta_W}\frac{3 G_\mu
m_t^2}{8\pi^2\sqrt{2}}+\frac{\sqrt{2}G_\mu M_W^2}{16\pi^2}[\frac{11}{3}
(\log\frac{m_H^2}{M_W^2}-\frac{5}{6})]+... \]
The mass of those particles enters therefore as a parameter to the indirect
determination of the W mass, as can be seen in figure \ref{fig:mwmtew99}
(full curve), where a clear correlation emerges between the indirect
determinations of the masses of the W and the top quark, as extracted from
a fit to precision electroweak data (mainly coming from LEP1 measurements)
available in winter 1999 \cite{ewmeas99}.\par
\begin{figure}[tb]
\begin{center}
\includegraphics[width=8cm]{m99_mt_mw_contours.eps}
\end{center}
\caption{Comparison between the direct measurements of the masses of W
and top mass and the predictions from the electroweak fit, for various
values of the Higgs mass} \label{fig:mwmtew99}
\end{figure}
The plot also shows as a dotted ellipse the direct measurement of both
masses, in good agreement with the indirect predictions. It is possible then
to also include the measured value of $m_t$ and $m_W$ in the fit, and get
some indication on the mass of the Higgs boson.\par
It is therefore very important to improve the precision on the direct
determination of the W mass, to further constraint the standard model, and
get stronger bounds on the allowed range for the Higgs mass.\par
\section{Models for Trilinear Gauge Coupling}
We have already seen that in the standard model vertices involving gauge
bosons only derive from the request of gauge invariance of the kinetic term.
Since trilinear couplings are extensively studied at LEP2, possible
deviations from the standard model value will be discussed.\par
The most general Lorenz-invariant effective Lagrangian, expressing the
coupling of two oppositely charged
and one neutral vector bosons is the following\cite{ybooktgc}:
\[ {\cal L}_{WWV}^{eff}/g_{WWV}=ig_1^V(W_{\mu\nu}^\dagger W^\mu V^\nu-
W_\mu^\dagger V_\nu W^{\mu\nu})+i k_V W_\mu^\dagger W_\nu V^{\mu\nu}\]
\[+\frac{i\lambda_V}{m_W^2}W_{\mu\nu}^\dagger W^\mu_\rho V^{\nu\rho}-
g_4^V W_\mu^\dagger W_\nu(\partial^\mu V^\nu+\partial^\nu V^\mu)\]
\[+g_5^V\epsilon^{\mu\nu\rho\sigma}(W^\dagger_\mu \overrightarrow{\partial_\rho}
W_\nu)V_\sigma+i \tilde{k}_V W_\mu^\dagger W_\nu\tilde{V}^{\mu\nu}\]
\[+\frac{i\tilde{\lambda}_V}{m_W^2}W_{\mu\nu}^\dagger W_\rho^\mu \tilde{V}^{\nu\rho}\]
Here V stands for either a photon or a Z boson ($V=\gamma,Z$), and W for the
W field. $g_{WWV}$ are fixed to
\[ g_{WW\gamma}=-e \hspace{3cm} g_{WWZ}=-e \cot \theta_W.\]
At tree level, the SM predicts $g_1^Z=g_1^\gamma=k_Z=k_\gamma=1$, with all
other couplings vanishing. The terms $g_1^V, k_V$ and
$\lambda_V$ conserve C and P separately, while $g_5^V$ violates both C and P
conserving CP. The coupling between W and photons can be related to intuitive
physical quantities; in particular the terms conserving C and P correspond to
the lowest-order terms in a multipole expansion of the interactions between Ws
and photons; thus they can be related to the magnetic moment $\mu_W$ and the
electric quadrupole moment $Q_W$:
\[\mu_W=\frac{e}{2 m_W}(1+k_\gamma+\lambda_\gamma)\]
\[Q_W=-\frac{e}{m_W^2}(k_\gamma-\lambda_\gamma)\]
The two parity-violating couplings $\tilde{k}_\gamma$ and $\tilde{\lambda}
_\gamma$ respect charge-conjugation invariance, and are related to the electric
dipole moment $d_W$ and to the magnetic quadrupole moment $\tilde{Q}_W$:
\[d_W=\frac{e}{2m_W}(\tilde{k}_\gamma+\tilde{\lambda}_\gamma)\]
\[\tilde{Q}_W=-\frac{e}{m_W^2}(\tilde{k}_\gamma-\tilde{\lambda}_\gamma).\]
Due to the relatively limited statistics available in the present experimental
facilities, the set of free parameters present in the general Lagrangian
quoted above is too large for practical uses. This set of parameters can be
reduced under a certain number of assumptions, depending on the way the
effective Lagrangian quoted above is made gauge-invariant, i.e. what kind of
new physics is expected to generate the couplings. If a light Higgs boson is
present, and considering only the C- and P- conserving operators, the effective
Lagrangian can take the gauge-invariant form
\[{\cal L}^{TGC}=ig'\frac{\alpha_{B\Phi}}{m_W^2}(D_\mu \Phi)^\dagger B^{\mu\nu}
(D_\nu \Phi)+ig\frac{\alpha_{W\Phi}}{m_W^2}(D_\mu \Phi)^\dagger \tau W^{\mu\nu}
(D-\nu\Phi)+g\frac{\alpha_W}{6m_W^2}W^\mu_\nu(W^\nu_\rho\times W^\rho_\mu)\]
with g and g' the SM couplings of the $SU(2)_L$ and $U(1)_Y$ symmetries.
When the Higgs field is replaced by its vacuum expectation value
$(0,v/\sqrt{2})$, the following relations can be written:
\[\Delta g^Z_1=g^Z-1=\frac{\alpha_{W\Phi}}{\cos^2\theta_W}\]
\[\Delta k_\gamma=(k_\gamma-1)=-\frac{\cos^2\theta_W}{\sin^2\theta_W}
(\Delta k_Z-\Delta g_1^Z)=\alpha_{W\Phi}+\alpha_{B\Phi}\]
\[\lambda_\gamma=\lambda_Z=\alpha_W\]
and is then natural to use the above relations and express all measured
quantities as a function of $(\Delta g_1^Z,\Delta k_\gamma,\lambda_\gamma)$ or the $\alpha$ parameters.
\par
In the absence of a light Higgs, a non linear approach can be followed to
make the effective Lagrangian gauge-invariant. In this scheme, it is convenient
to express the couplings as a function of the lowest-dimension operators
$(\Delta g_1^Z,\Delta k_\gamma,\Delta k_Z)$,
while the parameters $\lambda_\gamma$ and $\lambda_Z$ are usually set to zero.
\par
\section{Four-fermion production in $e^+e^-$ collisions}
Processes involving W pair production in e$^+$e$^-$ collisions are a subset
of a larger set of diagrams contributing to four-fermion final state and
interfering with each other, so all of them have to be considered when dealing
with W events. Processes contributing to 4-fermion final states can be
divided into charged current (CC) and neutral current (NC). The first class
comprises production of (up, antidown) and (down-antiup)-type fermion pairs,
where each pair has the same generation index. Final states produced via
W production belong to this class. Neutral current events are those where
two fermion-antifermion pairs are produced, and they
are mediated by the neutral gauge bosons. Obviously these two classes
overlap for certain final states. The number of Feynman diagrams
contributing to the charged current class is shown in table \ref{tab:cc}
for the possible combinations of final states. Three different cases occur
(shown in the table by different character types):
\begin{itemize}
\item {\bf CC11 family (boldface)}: for two different fermion pairs, none
of which is an electron, or electron neutrino, no identical particles are
involved, and there is at maximum 11 diagrams.
\item CC20 family (normal): one e$^\pm$ and one $\nu_e$ are in the final
state, so additional diagrams with t-channel exchange of the gauge boson are
present
\item {\it CC43/mix43 CC56/mix56 (italics)}: two mutually charge
conjugate pairs are produced, so these diagrams can proceed via both
charged and neutral gauge boson exchange
\end{itemize}
\begin{table}\begin{center}
\begin{tabular}{|c|c|c|c|c|c|}\hline
&$\bar{d}u$&$\bar{s}c$&$\bar{e}\nu_e$&$\bar{\mu}\nu_\mu$&$\bar{\tau}\nu_\tau$\\ \hline
$d\bar{u}$&{\it 43}&{\bf 11}&20&{\bf 10}&{\bf 10}\\
$e\bar{\nu_e}$&20&20&{\it 56}&18&18\\
$\mu\bar{\nu_\mu}$&{\bf 10}&{\bf 10}&18&{\it 19}&{\bf 9}\\ \hline
\end{tabular}
\caption{Number of diagrams for Charged Current final states}
\label{tab:cc}
\end{center}\end{table}
\par
\begin{figure}[p]
\begin{center}
\includegraphics[width=14cm,bbllx=40,bblly=100,bburx=550,bbury=685]{feyn.ps}
\end{center}
\caption{Feynman diagrams involved in $e^+e^-\to u\bar{d}
\mu\bar{\nu_\mu}$ final
state. Graphs in light grey (1, 6 and 7) correspond to W pair production
(CC03)} \label{fig:feyncc11}
\end{figure}
\par
In figure \ref{fig:feyncc11} the 10 diagrams contributing at tree level to the
e$^+$e$^-\to u\bar{d} \mu\bar{\nu_\mu}$ events are shown.
All semileptonic decays (except those where an
electron is present in the final state) are produced through the same set
of diagrams, with the proper redefinition of the final state particles.
The graphs 1, 6 and 7 are the only ones where a W pair is produced;
they are often referred to as CC03 processes.
Contributions from single- and non-resonant processes are particularly
large for $l\nu l\nu$ and $qqe\nu$ final states, leading to an ambiguity in
the definition of the signal. The approach followed by the LEP collaborations
is slightly different:\begin{itemize}
\item DELPHI, L3: consider efficiencies on signal inside generator level
cuts, and apply multiplicative factors for translating the cross section
measured for the full set of diagrams into a cross section relative to
W-pair production only
\item ALEPH, OPAL: consider the additional diagrams as a background,
neglecting the interference between these processes and the W pair diagrams
\end{itemize}
It was shown that these procedures give the same results within a 1\% accuracy.
\section{Machine parameters, schedule and calibration}
Due to the strong increase in synchrotron radiation, the only possibility to
operate the LEP machine well above the Z resonance is to considerably increase
the LEP1 accelerating power. Since the machine layout cannot be changed,
this can only be achieved raising the accelerating gradient in the straight
sections. The 128 five-cell copper cavities constituting the accelerating system
of LEP in the first phase were able to deliver a peak RF power corresponding
to a voltage of 400 MV
per revolution, clearly inadequate for LEP2 needs (over 3000 MV). The big
increase in performances was only possible due to the operation of
superconducting cavities, that because of their very high quality factor could
provide as much as 6 MV/m of accelerating gradient.\par
The installation of those cavities proceeded in several steps, and so did
the energy of operation of the machine. The machine schedule in the period
1996-1997 is shown in table \ref{tab:sched}, where only the data-taking periods
with total energy above the Z peak have been considered. Apart from the runs above
the WW threshold, relevant to this report, a short run at 130-136 GeV has been
taken, to clarify possible anomalies in the 4-jet production at that energy.
The results of that run are summarized in \cite{4jet}.
\begin{table}
\begin{center}
\begin{tabular}{|l|l|l|l|l|}\hline
Period&N. SC cavities&N. Cu cavities&$\int{\cal L}$ (pb$^{-1}$)&
Energy (GeV)\\ \hline
1996 a&144&182&12.1&161\\
1996 b&176&150&11.3&172\\
1997&240&86&63.8&183\\
1997 b&240&86&7.2&130-136\\
1998 & 272&48&196.4&189\\ \hline
\end{tabular}
\end{center}
\label{tab:sched}
\caption{Characteristics and performances of the LEP machine in the years 1996-1998. Only runs above the Z peak have been listed.}
\end{table}
\par
Since all fitting methods use the centre-of-mass energy as a kinematic
constraint, the uncertainty on the knowledge of the LEP beam energy directly
reflects into a systematic error on the value of the W mass:
\[\Delta M_W=\frac{\Delta E_b}{E_b} M_W\]
To fulfill the
requested precision on the mass, the LEP energy has to be known with an
accuracy better than 20 MeV. At LEP1, energy calibration was performed via
resonant depolarization (RDP)\cite{respol}. This method can not be used at
LEP2, since there is no possibility to have polarization at physics energies.
Several RDP measurements have been however performed at lower energies,
and extrapolated. The main error on this method comes from the extrapolation
itself, leading to a total error of about 20 MeV at 189 GeV.\par
As an independent approach, a spectrometer is planned for LEP, to be
operational in 1999 and beyond. This will consist in a fully equipped dipole
magnet, giving a precision of about $1 \mu m$ on the beam position, extracting
the particle momentum out of their curvature in the dipole magnetic field.
\par
\section{WW cross section and branching ratios}
WW events can have very different topologies, depending on the different
decays of the two W bosons. As two extreme cases, the decay of two W bosons
could produce a high-multiplicity four-jet event, as well as a low-energy
imbalanced event with only two charged leptons seen in the detector.\par
Accordingly, the selection criteria, the backgrounds and the possible
systematic uncertainties in the selections can be very different. All LEP
collaborations have different analysis for the possible WW decay channels.
They are here grouped into three main categories: fully leptonic, semileptonic
and fully hadronic decays.
\subsection{Fully leptonic events}
Fully leptonic events $WW\to l\nu l\nu$ are usually characterized by:
\begin{itemize}
\item two high-energy acoplanar leptons
\item missing momentum not pointing to the beam pipe, due to the undetected
neutrinos
\end{itemize}
An example of a $l\nu l\nu$ event detected in the DELPHI detector is shown
in figure \ref{fig:delpevlnln}. The energy distribution of the most energetic
lepton in L3 for $l\nu l\nu$ candidates is in figure \ref{fig:l3elept}.\par
Events are usually classified into lepton-lepton, lepton-jet and jet-jet
categories, where here lepton stands for either electron or muon, and narrow
jets are considered, to account for hadronic $\tau$ decays.\par
\begin{figure}[tb]
\begin{minipage}{.45\linewidth}
\begin{center}
\includegraphics[width=6cm,bb=30 220 500 640,clip]{delphevwwll.ps}
\caption{A $WW\to \tau\nu \mu\nu$ event in Delphi. A narrow jet originated
from $\tau$ decay (on the left), and a muon (traversing the whole detector)
are widely acoplanar.} \label{fig:delpevlnln}
\end{center}
\end{minipage}
\begin{minipage}{.45\linewidth}
\begin{center}
\includegraphics[width=7cm]{l3lnln.eps}
\caption{Lepton energy distribution in L3 for fully leptonic events} \label{fig:l3elept}
\end{center}
\end{minipage}
\end{figure}
\par
Due their topology, the main backgrounds to these processes are:
\begin{itemize}
\item Two-fermion events from $Z/\gamma$ decays (especially $\tau$ pair
production), Bhabha scattering events
\item high-energy $\gamma-\gamma$ interactions
\item $ZZ\to ll\nu\nu$ events (mainly for $\sqrt{s} > 184$ GeV)
\end{itemize}
\par
Since the leptons are required to be acoplanar, the most dangerous background
from two fermion events is represented by radiative $Z/\gamma$ decays.
Events with a high-energy isolated photon are rejected, since this is a
clear indication of radiative $Z/\gamma$ decays. Also events with missing
momentum pointing at very low angle are rejected, since in this case the
radiated photon could have been lost in the beam pipe or in a badly-instrumented
sector of the detector.
The typical selection efficiencies for this channel are higher for the case
in which two stable leptons are produced, and lower for the jet-jet case, due
to the stronger cuts needed to suppress the larger background.
Overall efficiencies are around 70\%.
\par
\subsection{Semileptonic events}
The semileptonic channels $WW\to q\bar{q}l\bar{\nu}_l$, in particular those
with an electron or a muon in the final
state, have quite similar topology. These events are characterized by two
hadronic jets, a high-energy lepton and large (and similar) missing energy
and momentum due to the neutrino. $qq\tau\nu$ events are usually more balanced
due to the additional neutrinos produced in $\tau$ decays, and the missing
energy is larger. The lepton from $\tau\to e$ and $\tau\to\mu$ decays is
softer, than that produced directly from the W, while hadronic $\tau$ decays
produce a narrow jet.\par
The background is mainly coming from hadronic Z radiative decays, but is
different for the three channels. For the $qqe\nu$ case, the main baground
comes from Z radiative events with the photon in the detector, associated to a
nearby track, or converted into a $e^+ e^-$ pair. This is particularly
true in the forward region of the detector, where most of the radiative photons
are emitted, and where usually the tracking capabilities of the detector are
not optimal. Background to the $qq\mu\nu$ channel is mainly coming from
semileptonic decays of b quarks in $Z\to qq (\gamma)$ events, or from $ZZ\to
qq\mu\mu$ processes, where one of the muons is not identified, and mimics
missing momentum. Having less strong signature, the $qq\tau\nu$ channel has
usually more complicated selections, and its main background arises from
radiative Z hadronic decays where the photon gets undetected and a third
jet fakes that coming from $\tau$ decays.\par
For the final selection, DELPHI and L3 use a cut-based approach, requiring
good isolation for the lepton and high invariant mass for the decay products
of the two Ws. ALEPH and OPAL combine the informations coming from similar
variables using an event probability function. Typical efficiencies are
of the order of 80\% for the electron and muon channels, and 50\% for the
$\tau$ channel.\par
\begin{figure}[tb]
\begin{center}
\mbox{\includegraphics[width=6cm,bb=100 228 534 652,clip]{evqqen.eps}}
\end{center}
\caption{A $qqe\nu$ event in L3. The electron is visible as a large ``tower''
in the electromagnetic calorimeter, in the bottom part of the event. The
two hadronic jets are opposite to each other.}
\label{fig:qqlnfig}
\end{figure}
\begin{figure}
\begin{center}
\includegraphics[width=14cm]{opalqqln.eps}
\end{center}
\caption{Distributions of measured lepton energies for events selected
in OPAL in the three semileptonic channels} \label{fig:qqlndist}
\end{figure}
\par
\subsection{Hadronic W decays}
The channel $WW\to q\bar{q} q\bar{q}$ has about the same branching ratio as
the sum of three semileptonic ones, i.e. about half of the total number of
WW decays. It is characterized by four high-energy well separated hadronic
jets, coming from hadronization of the quarks from the W decays.\par
There are two main sources of background:\begin{itemize}
\item $Z\to q\bar{q}$ events with hard gluon radiation
\item $ZZ\to q\bar{q} q\bar{q}$ events.
\end{itemize}
The latter is almost irreducible, since well-isolated high energy jets
are produced, and the only difference with respect to the signal is the
slightly higher jet-jet invariant mass. Z decays have a larger cross section,
but since the two additional jets are coming from gluon radiation, they are
usually less energetic and closer to the emitting quark.\par
All experiment try to combine all available informations in an optimized way.
This is done combining several variables (event shape, invariant
masses angles between jets etc.), using a likelihood discriminator (OPAL) or
a neural network (DELPHI,ALEPH, L3).
The variables used by the OPAL collaboration are shown in figure
\ref{fig:opalqqqq}, together with the resulting likelihood and the value of the
cut. In order to have an additional gain in statistical power and have a
further cross-check on the background, ALEPH and L3 fit the neural network
output distribution by a linear combination of distributions for signal and
background.
\begin{figure}[p]
\begin{center}
\includegraphics[width=14cm]{opalqqqq.eps}
\end{center}
\caption{Distributions of the variables used by OPAL in the $WW\to q\bar{q}
q\bar{q}$ analysis. The points indicate the data, the open histogram
represents the MonteCarlo expectations for the signal, and the hatched
histogram shows the background estimate.} \label{fig:opalqqqq}
\end{figure}
\par
\begin{figure}[tb]
\begin{center}
\includegraphics[width=8cm,angle=270,bb=40 30 570 800,clip]{al4jcom.ps}
\end{center}
\caption{A $WW\to q\bar{q} q\bar{q}$ event in ALEPH. The four jets coming
from W decays are clearly separated.}
\label{fig:al4j}
\end{figure}
The uncertainties related to the QCD modeling of the fragmentation process,
in particular of the final state interactions, are the main sources of
systematic errors for this channel. These uncertainties will be discussed in
more detail in the section about W mass measurements.
\par
\section{Determination of WW cross section and branching fractions}
The cross sections for the individual WW decay channels measured as above
are combined to extract a value of the total WW production cross section
and branching ratio. To make use of physical assumption like i.e. lepton
universality, likelihood-based fit are used by the different collaborations.
For a given channel $i$, the expected number of events, $\mu_i$, is computed
accounting for the background and the cross-efficiencies among that channel
and all the others ($\epsilon_{ij}$):
\[\mu_i=L\times(\Sigma_j \epsilon_{ij}\sigma_j+\sigma_i^{bg})\]
where L is the collected luminosity and the sum runs on all channels.
The cross sections are extracted maximizing the likelihood
\[{\cal L}=\Pi_i P(N_i,\mu_i)\]
where P is the Poisson probability of observing $N_i$ events in a given
channel i, with $\mu_i$ expected.\par
This likelihood can be maximized leaving different free parameters, according
to the physics assumptions. If for instance no assumptions are made, the
cross sections $\sigma_j$ for all channels are left free. On the other hand,
it is possible to extract the total cross section, imposing the knowledge of
the W standard model branching fractions. In this case, the above cross
sections are expressed as
the product of the WW cross section $\sigma$ (which is now the only free
parameter in the fit) times the branching ratio for the
corresponding channel.\par
The total WW production cross section for the four experiments at $\sqrt{s}$
of 183 and 189 GeV are listed in table \ref{tab:totxsec}. All results from the
run at 189 GeV are preliminary, and taken from the contributions of the
various collaborations to the winter conferences\cite{xsecmor99}.\par
\vspace{1cm}
\begin{table}
\begin{center}
\begin{tabular}{|l|l|l|}
\hline
\multicolumn{3}{|c|}{Cross-section $\sigma_{\tt CC03}$ (pb)}\\
\hline
Experiment & $\sqrt{s}=$183~GeV & $\sqrt{s}=$189~GeV\\
\hline
ALEPH & 15.57 $\pm$ 0.68$^n$ & 15.64 $\pm$ 0.43 \\
DELPHI & 15.86 $\pm$ 0.74$^n$ & 15.79 $\pm$ 0.49 \\
L3 & 16.53 $\pm$ 0.72$^p$ & 16.20 $\pm$ 0.46 \\
OPAL & 15.43 $\pm$ 0.66$^p$ & 16.55 $\pm$ 0.40 \\
\hline
LEP & 15.83 $\pm$ 0.36 & 16.07 $ \pm$ 0.23 \\
SM & 15.70 $\pm$ 0.31 & 16.65 $ \pm$ 0.33 \\
\hline
\multicolumn{3}{|c|}{$^p$ Published $^n$ New Preliminary}\\
\hline
\end{tabular}
\end{center}
\label{tab:totxsec}
\end{table}
\par
\vspace{1cm}
\par
The combined LEP cross section only considers statistical errors. The combined
value for all LEP2 energies is shown in figure \ref{fig:xsec}. In addition to
the curve predicted from the standard model, this figure shows the WW cross
section in the two cases where the ZWW vertex has zero coupling, and where
WW production occurs only via the neutrino exchange diagram. In both cases
the WW production cross section diverges for large values of $\sqrt{s}$,
and is also incompatible with the values measured at LEP. Therefore, the
simple cross section measurement represents a confirmation of the non-Abelian
structure of the electroweak interactions.
\begin{figure}[p]
\begin{center}
\includegraphics[width=6cm]{wwxs.eps}
\end{center}
\caption{Total WW production cross section} \label{fig:xsec}
\end{figure}
\begin{figure}[p]
\begin{minipage}{.45\linewidth}
\begin{center}
\includegraphics[width=6cm]{brall.eps}
\caption{Leptonic branching ratios} \label{fig:lbr}
\end{center}
\end{minipage}
\begin{minipage}{.45\linewidth}
\begin{center}
\includegraphics[width=6cm]{brhad.eps}
\caption{Hadronic branching ratios} \label{fig:hbr}
\end{center}
\end{minipage}
\end{figure}
If the cross section is not fixed, it is possible to
determine the W decay branching ratios, leaving them as free parameters for
the fit. The results for the different leptonic ratios is shown in figure
\ref{fig:lbr}, showing a direct verification of th lepton universality
in charged current weak interactions. The hadronic branching ratio is shown
in figure \ref{fig:hbr}. This value can be expressed in terms of the CKM
matrix elements using the following expression:
\[\frac{Br(W\to q\bar{q})}{1-Br(W\to q\bar{q})}=(1+\frac{\alpha}{\pi})\Sigma
|V_{ij}|^2\]
that yields, using the combined LEP value:
\[\Sigma |V_{ij}|^2=2.10\pm 0.08\]
The experimental knowledge of all elements of the CKM matrix is quite good,
apart from $V_{cs}$, suffering from large uncertainties (about 20\%) of both
experimental and theoretical nature\cite{pdg}.
Imposing unitarity of the CKM matrix
and considering the measurements of the other matrix elements, the previous
result can be reinterpreted as a determination of $|V_{cs}|$:
\[|V_{cs}|=1.002\pm 0.0016 (stat) \pm 0.002 (syst)\]
A completely independent technique to determine this quantity will be presented
in section 12.\par
\par
\section{W mass}
As mentioned in the introduction, the W mass is one of the most important
measurements of the LEP2 program. At energies close to the WW production
threshold, the highest sensitivity is reached deriving the mass from the
cross section measurement, an approach conceptually similar to that used
at LEP1 to measure the mass of the Z boson. At higher energies, the
W mass is derived directly from the measured invariant mass of the W decay
products.\par
\subsection{W mass from threshold cross section}\par
\begin{figure}[tbh]
\begin{center}
\includegraphics[width=8cm,clip]{w_mass_161_final.eps}
\end{center}
\label{fig:mass161}
\caption{W mass from the combined LEP cross section at $\sqrt{s}=161$ GeV}
\end{figure}\par
Assuming validity of the SM, the W mass can be extracted from the cross
section, for a fixed centre-of-mass energy. The sensitivity of this approach
is maximal close to threshold, due to the steep rising of the cross section
in that region; for this reason this method was used to determine the mass
from the cross section of the first run at $\sqrt{s}=161$ GeV. In figure
\ref{fig:mass161} the dependence of the cross section on the W mass is
shown, together with the combined measurement of the mass from the LEP
experiments.\par
\subsection{W mass from direct reconstruction}
At higher energies the dependence of the cross section on the W mass is
negligible, so the derivation of the mass from the cross section can no
longer be used. On the other hand, the
WW cross section is much larger than at threshold, and it is possible to use
the direct reconstruction method, i.e. the W mass is extracted with a fit to
the invariant mass distribution of the W decay products. The calculation of
these masses is only trivial in the semileptonic case, where two jets are
coming from a W and the system of lepton and neutrino from the other.
In the fully leptonic case, the system is underconstrained, due to the
presence of at least two undetected neutrinos. The ALEPH collaboration has
been the only one so far to use this channel for mass fits, using the energy
of the two leptons, with small statistical power. In the fully hadronic
case, since at least four jets are present in the detector, several mass pairs
could be formed. Criteria based on reconstructed masses, angles etc.
are used to get the best pairing, with efficiencies of the order of 80\%;
however including the other pairings with smaller weight can help increasing
the mass sensitivity.\par
The main problem to measure the W mass from reconstructed distribution is
to account for all distortions coming from detector effects and selection
biases. Two main approaches are used:\begin{itemize}
\item convolution
\item MC reweighting
\end{itemize}
All LEP collaborations use both methods, quoting one for the final results
and the other as a cross-check.\par
In both cases the final error on the W mass will be determined by the total
number of candidates as well as the detector resolution in measuring invariant
masses. In order to improve the detector performances, it is possible to
impose some physical constraint to any single event. Since energy and momentum
are conserved in the collision, it is possible to perform a fit to energies
and angles of the final states particles, such that the fit results satisfy
the kinematical constraint and are as close as possible to
the measured ones. This procedure largely increases the sensitivity to the W
mass, but requires a precise knowledge of the LEP energy since this value is
directly used in imposing the energy conservation.\par
The L3 and OPAL collaboration also exploit the additional kinematical
constraint that the masses of the two produced W bosons must be
equal within the W width.\par
\subsection{Convolution}
\begin{figure}[tbh]
\begin{center}
\includegraphics[width=9cm]{ideo4jcom.ps}
\end{center}
\caption{Likelihood contours in the $M^1_W-M^2_W$ plane for a four-jet
event in DELPHI. The three maxima correspond to the three possible
jet pairings.}
\label{fig:delpev}
\end{figure}
In the convolution method (DELPHI), the theoretical W line-shape curve
(depending on the W mass) is convoluted with an analytical function
describing detector effects. The experimental line-shape is compared to the
convoluted curve, determining a likelihood function, having as a free
parameter the W mass. Maximizing the likelihood it is possible to extract
the value of the W mass for which the curve obtained smearing the theoretical
distribution mostly resembles the experimental curve.\par
The main difficulty of this method is in the modelization of the the detector
response, that must be included an analytical form. Morover, the reconstructed
mass has a bias that must be corrected comparing with the MC. On the other
hand, the method allows the use of different event weights depending on the
detector resolution for each data event, thus improving the accuracy of the
measurement.\par
In particular, DELPHI fits the masses in the $M_W^1$, $M_W^2$ plane, using
all three combinations for the qqqq channel (see figure \ref{fig:delpev}).
\par
\subsection{MC Reweighting}
\begin{figure}[p]
\begin{minipage}{.45\linewidth}
\begin{center}
\includegraphics[width=6cm]{l3mass.eps}
\caption{W mass distribution in L3} \label{fig:l3wmass}
\end{center}
\end{minipage}
\begin{minipage}{.45\linewidth}
\begin{center}
\includegraphics[width=6cm]{l3zmass.eps}
\caption{Z mass distribution in L3} \label{fig:l3zmass}
\end{center}
\end{minipage}
\end{figure}
\begin{figure}[p]
\begin{center}
\includegraphics[width=6cm]{mw_ffff_172-189.eps}
\caption{Results for the W mass from direct reconstruction in the four
LEP experiments. ALEPH-L3: preliminary results including the run at 189 GeV;
OPAL-DELPHI: final results from the runs at 172 and 183 GeV of centre of mass
energy}
\label{fig:wmassres}
\end{center}
\end{figure}
\par
This method, used by the other collaborations, uses a large number of Monte
Carlo events to establish the correspondence between generated and
reconstructed masses.
These events are generated with a given value of the W mass;
an analytical code is used to reweight them according to the mass that best
fits the data, using an iterative procedure based on a likelihood similar to
that used for the convolution method.\par
As a cross-check for the validity of the method, the L3 collaboration has
presented a fit to the Z mass, performed on radiative $Z\to q\bar{q}\gamma$
events, using the same method as the one used to fit the mass of the W.
Since the value of the Z mass is known with very high precision from LEP1
measurements, the good agreement of the fitted value with the expectation
is a good test of the complete mass analysis method. In figures
\ref{fig:l3wmass} and \ref{fig:l3zmass}, the reconstructed invariant mass
distributions for WW (all channels) and radiative Z events are shown. It is
interesting to notice that for hadronic WW events the two best jet pairing
are included, so a large but flat background from incorrect jet pairing is
present.\par
Both method can be used to perform a two-parameter fit, where both $M_W$ and
$\Gamma_W$ are left free. The correlation of the two measurements is quite
small, and a statistical error of about 200 MeV per experiment on the width
measurement can be obtained.\par
The W mass results from direct reconstruction are shown in figure
\ref{fig:wmassres}. In this plot, all results refere to data taken at center
of mass energies between 172 and 189 GeV\cite{massval}.\par
The present combined value from LEP is $M_W=80.368\pm0.065$ GeV (statistical
error only), as precise as the combined value obtained from hadron machines.
\par
\subsection{Systematic uncertainties on the mass measurement}
Given the importance of the measurement of the W mass, as discussed in
the introduction, and the good statistical accuracy reached by the measurement,
the understanding of the systematic
uncertainties associated to this measurement are crucial to fully exploit the
potentiality of LEP. Systematic uncertainties can come
from several sources:\begin{itemize}
\item beam energy\par
this value is used as a global normalization factor in the kinematic fit,
so its uncertainty directly reflects into an uncertainty on the mass
\item ISR-FSR\par
the incomplete simulations of initial and final state radiation can be
estimated comparing mass results obtained using different MonteCarlo
implementations of these effects
\item detector effects\par
errors due to a non-perfect simulation of the detector response can be
estimated varying resolution and energy scale in reasonable ranges
\item technical effects\par
the finite precision at which the accuracy of the mass fitting method is
tested, as well as the limited MonteCarlo statistics;
\item background\par
the cross section and energy shape of the background is varied, leading to
some small modifications of the measured values of the W mass
\item QCD final state interactions\par
a significant bias to the W mass measured at LEP in the 4-jet channel could
come from QCD interactions in the final state such as color reconnection or
Bose-Einstein effects. Theoretical models for both effects give quite different
results, so presently the experiments assign large systematic errors, comparing
the mass results obtained with the different methods. A more detailed
description of these effects and of some experimental ways to discriminate
among the various models will be discussed in section 11.
\end{itemize}
In table \ref{tab:systmass}, typical values of uncertainties for the sources
of systematic errors on the W mass listed above are quoted. These numbers
have to be considered as indicative, since they can vary even substantially
from an experiment to another.\par
\begin{table}[h]
\begin{center}
\begin{tabular}{|c|c|}\hline
Effect&Systematic error (MeV)\\ \hline
Beam Energy&20\\
ISR-FSR&10\\
Detector&10\\
Technical&20\\
Background&10\\
Fragmentation&30\\
Final State&30\\ \hline
Total&55\\ \hline
\end{tabular}
\vspace{1cm}
\caption{\label{tab:systmass}Typical valus of uncertainties for the
various systematic sources.}
\end{center}
\end{table}
Presently, the LEP collaborations quote systematic uncertainties larger than
50 MeV (even higher in the 4-jet channel), similar to the present combined
statistical accuracy. Much work is in progress to lower the systematic
unctertainties, in order to fuly profit from the increase in statistics
expected in the next years.\par
\section{Trilinear Gauge Couplings}
\begin{figure}[h]
\begin{center}
\includegraphics[width=7cm,bbllx=10,bblly=410,bburx=550,bbury=640,clip]{ancoup.ps}
\end{center}
\label{fig:tgc}
\caption{Three gauge boson vertex}
\end{figure}
The effect of anomalous trilinear gauge couplings is a modification of the
WW production cross section and a distortion of the event kinematics.\par
In particular, the study of these couplings is performed with a combined fit
to the total
WW production cross section as well as the distribution of the production and
decay angle of each W boson (see figure \ref{fig:tgcang}).
\begin{figure}[h]
\begin{center}
\includegraphics[width=7cm,bbllx=30,bblly=190,bburx=570,bbury=600,clip]{tgcangles.eps}
\end{center}
\label{fig:tgcang}
\caption{Production and decay angles in a WW event}
\end{figure}
\par
If no jet charge algorithm is used, the W production angle can only be
unambiguously determined in semileptonic events, where the charge of the
lepton reflects the charge of the parent W. In the fully leptonic case,
all production and decay angles can be determined with a two-fold ambiguity
using kinematic criteria if initial state radiation and W width are neglected
\cite{lnlnreco}. In the four-jet channel the ambiguity on the angles can
only be solved using the jet charge. Algorithms based on the Feynman-Fields
approach \cite{feynfields} have correct charge identification probability
around 70\%.\par
Detector effects in angle resolution and charge confusion are accounted for
using reweighting algorithms similar to those used for the measurement of
the W mass. An alternative approach is that based on Optimal Observables
\cite{optobs}. Instead of fitting the kinematic distributions, the
differential cross section parametrized as a quartic function of the anomalous
couplings:
\[\frac{d\sigma}{d\Omega}=c_0(\Omega)+c_1(\Omega)\Psi+c_2(\Omega)\Psi^2\]
where $\Omega$ are the phase space variables, and $\Psi$ is one of the
couplings allowed to be different from the standard model, while the others
are kept to zero. Assuming that all couplings are small, it is possible to
neglect the term $c_2$, and apart from this approximation the observable
\[O_1=\frac{c_1(\Omega)}{c_0(\Omega)}\]
contains all the information carried by the distributions $d\sigma/d\Omega$,
allowing the extraction of the couplings from a 1-dimensional fit.\par
\begin{figure}[p]
\begin{center}
\includegraphics[width=14cm,bb=45 215 525 735,clip]{l3tgcpic.ps}
\end{center}
\label{fig:tgcvar}
\caption{Distributions of production and decay angles in L3 WW events at 189 GeV}
\end{figure}
\begin{figure}[p]
\begin{center}
\includegraphics[width=14cm,bb=17 153 560 715]{leptgc1d.ps}
\end{center}
\label{fig:leptgc1}
\caption{Values of single coupling variables obtained combining the four LEP
experiments, assuming the other variables set to the standard model value. The
four curves in each plot correspond to the results of the four LEP
experiments.}
\end{figure}
\begin{figure}[p]
\begin{center}
\includegraphics[width=14cm,bb=0 120 550 700]{leptgc2d.ps}
\end{center}
\label{fig:leptgc2}
\caption{Combined allowed contours for TGC variables when one is set to the
standard model value and the other are left as free parameters of the fit.}
\end{figure}
\par
\begin{figure}[tbh]
\begin{center}
\includegraphics[width=7cm]{l3singph.eps}
\end{center}
\label{fig:l3phot}
\caption{Energy spectrum of single photon
events in L3 for 189 GeV data}
\end{figure}
\par
Diagrams involving the interaction of three vector bosons are not only present
in WW final state events, but also in other kind of processes, e.g. production
of single W and single photons (figure \ref{fig:l3phot}. With respect to WW
production, these processes
are more sensitive to $\gamma WW$ couplings, $\Delta \kappa_\gamma$ and
$\Delta \lambda_\gamma$.\par
As discussed above, all analyses performed to study trilinear gauge couplings
are based on reconstructed distributions of W decay products (or photon).
The final sensitivity of the result strongly depends on the reconstruction
performances for these quantities, therefore also in this case kinematic fits
are widely used. The main sources of systematics are coming from the accuracy
of the MonteCarlo modeling of detector effects, as well as the effects in
fragmentation as well as final-state hadronic interactions, as already
discussed for the case of the W mass measurement.
\par
Final results for the couplings can be expressed in terms of one variable
constraining the others to the standard model values, or fitting two variables
at the same time.
The LEP combined results following these two approaches are
shown in figures \ref{fig:leptgc1} and \ref{fig:leptgc2}. Results from the
single experiments can be found in the references \cite{tgcww}.
\par
\section{QCD effects}
Hadronic interactions occur between W decay products, and can lead to
modification of the observed final states. In particular the understanding
of these effects is very important for the W mass measurement, since they can
lead to biases in the 4-jet channel far larger than the target accuracy for
this measurement. In particular, Bose-Einstein and color reconnection effects
will be discussed.\par
\subsection{Bose-Einstein effects}
Bose-Einstein correlations enhance the production of identical bosons close
in direction and momentum. These effects have already been observed in
nucleus-nucleus and hadron-hadron interactions\cite{beh}, as well as in
hadronic Z decays at LEP1 \cite{belep1}.\par
In WW events, Bose-Einstein correlations can occur:\begin{itemize}
\item between bosons within same jet
\item between bosons from same W
\item between bosons from different Ws
\end{itemize}
Only the last case is important for W mass studies, since it generates
distortions in the reconstructed invariant mass spectrum.\par
\begin{figure}[t]
\begin{minipage}{.45\linewidth}
\begin{center}
\includegraphics[width=5cm,bb=63 206 510 725,clip]{delbepic.ps}
\caption{R(Q) distribution from DELPHI data at 183 and 189 GeV. (a): in
semileptonic events. (b) in fully hadronic events, compared to a model with
full BEC. (c) same data as in (b), compared to a model with BEC only inside
the same jet. Data seem favoring models with full BEC.} \label{fig:delbe}
\end{center}
\end{minipage}
\begin{minipage}{.45\linewidth}
\begin{center}
\includegraphics[width=6cm]{albefig.eps}
\caption{Data/MC ratio of R(Q) distributions in ALEPH. Data (full dots) are
in better agreement with MonteCarlo where BEC occur only inside the same W
(stars) than in models where they occur between different Ws (open dots)}
\label{fig:albe}
\end{center}
\end{minipage}
\end{figure}
\par
To model this effect, the correlation function between two equal bosons is
assumed to be Gaussian:
\[R(Q)=(1+\lambda e^{-Q^2R^2})\]
where $\lambda$ and R are the amplitude and the radius
of the effect, and Q is the four-momentum difference between the two identical
bosons $Q^2=(p_1-p_2)^2$.\par
\par
From the experimental point of view, most of the effect shows up between
pions of the same charge inside hadronic jets, while pions of different charge
are not affected. Bose-Einstein correlations produces an enhancement in the
ratio of the Q distribution between same-sign ($\rho^{\pm\pm}$) and
opposite-sign ($\rho{\pm\mp}$) pions:
\[R(Q)=\frac{\rho^{\pm\pm}(Q)}{\rho^{\pm\mp}(Q)}\]
The effect can either be seen from the R(Q) distribution itself (OPAL),
in the double ratio $R(Q)_{DATA}/R(Q)_{MC}$ (ALEPH, L3), or defining
\[R(Q)=\frac{\rho^{\pm\pm}(Q)}{\rho^{\pm\pm}_{MC}(Q)}\]
where the Monte Carlo sample has no Bose-Einstein correlations (DELPHI).\par
Delphi results are slightly in favor of the presence of Bose-Einstein
Correlations between particles from different Ws (figure \ref{fig:delbe}),
while the results from ALEPH seem disfavoring (by 2.7 $\sigma$) such
correlations (figure \ref{fig:albe}).\par
The preliminary values for the amplitude of the effect between pions coming
from different W $\lambda^{diff W}$ are listed in table \ref{tab:bec}
\cite{beexp}. To be
noticed that ALEPH and L3 results are preliminary, and the data sample used
are widely different. With the present available information the presence of
Bose-Einstein correlations between different Ws is still unclear.\par
\begin{table}[h]
\begin{center}
\begin{tabular}{|c|c|}\hline
Experiment&$\lambda^{diff W}$\\ \hline
ALEPH&$0.15\pm0.18$\\
DELPHI&$-0.20\pm 0.22$\\
L3&$0.75\pm 1.80$\\
OPAL&$0.22\pm 0.53$\\ \hline
\end{tabular}
\vspace{1cm}
\caption{\label{tab:bec}Fitted values for $\lambda$}
\end{center}
\end{table}
\par
\subsection{Colour reconnection}
String effects between jets coming from different Ws (but from partons with
opposite colour) are another source of distortion of the mass distribution.
They can occur since the distance in space between the two W decay vertices
is of the order of 0.1 fm, while the hadronic scale is of the order of 1 fm.
\par
\begin{figure}[H]
\begin{center}
\includegraphics[width=6cm]{fcolrea.eps}
\end{center}
\end{figure}
\par
Perturbative contributions are expected to be small,\cite{cr2} but the non perturbative
part can lead to mass shifts of the order of some hundreds MeV, depending on
the model.
The experimental study of these effects exploits the fact that in addition
to W mass shifts, colour reconnections produce modifications in the topology
of the events.
The most important observable used to discriminate among the different models
is the charged multiplicity in four-jet events (often expressed as difference
or ratio between charged particle production in $qqqq$ and $2\times qql\nu$
events, which are not affected by CR)\cite{cr1}. Typically,
models predicting shifts in the W mass of the order of several hundred MeV
also predict a difference in
the number of charged particles between semileptonic and hadronic events of
about 10\%, while for models leading to smaller mass shifts this
difference is few per cent. As shown if figure \ref{fig:opalcr}, it is very
difficult with present data to discriminate among the models implemented
in the event generators \cite{crall}, since
the predicted shift is similar and quite small.\par
\begin{figure}[tb]
\begin{center}
\includegraphics[width=8cm]{opalcrfig2.eps}
\end{center}
\caption{Number of charged particles in 4-jet hadronic events compared to
different color reconnection models. Apart from the VNI model, that fails to
reproduce the thrust distribution (plot c), the other models predict very
small shifts in the charged multiplicity, and the present data are not able to
discriminate among them.}\label{fig:opalcr}
\end{figure}
\par
Table \ref{tab:cr} shows the difference (the ratio for DELPHI) between then
charged multiplicity of 4-jet events and twice the one of semileptonic events
\cite{crexp}.
Present errors on this quantity are still too large to be compared to the
existing models.\par
\begin{table}[h]
\begin{center}
\begin{tabular}{|c|c|}\hline
Experiment&$\Delta<n_{ch}>$\\ \hline
ALEPH&$0.47\pm0.44\pm0.26$\\
L3&$-.0\pm0.8\pm0.5$\\
OPAL&$0.7\pm0.8\pm0.6$\\ \hline
&$<N_{ch}^{qqqq}>/2<N_{ch}^{qq}>$\\ \hline
DELPHI&$0.977\pm0.017\pm0.027$\\ \hline
\end{tabular}
\vspace{1cm}
\caption{\label{tab:cr}Difference (ratio for DELPHI) in charged multiplicity
between four-jet and two-jet events. Results from L3 and OPAL do not include
189 GeV data. Only OPAL results are final.}
\end{center}
\end{table}
\par
Other observables for studying color reconnection studies are:\begin{itemize}
\item charged multiplicity in the low momentum region
\item track characteristic distributions (momentum, rapidity, $p_t$,...)
\item event shape (thrust,...)
\item heavy hadron multiplicity (K,p with $0.2 GeV< p < 1.4$ GeV)
\end{itemize}
Also for these observables no clear indications for color reconnection can be
derived.
\par
\section{Charm production in W decays}
For real W bosons the production of b quarks is either forbidden by energy
conservation (as in the case $W\to tb$) or strongly Cabibbo-suppressed
(as in the case of the decay $W\to cb$). For this reason, charm is the
heaviest quark largely produced in W decays.
Using its heavy-quark characteristics in an almost b-free environment, it is
possible to measure the charm production branching ratio in W decays.
In the standard model this value is precisely determined by the unitarity
of the CKM matrix:
\[\frac{|V_{cd}|^2+|V_{cs}|^2+|V_{cb}|^2}{|V_{cd}|^2+|V_{cs}|^2+|V_{cb}|^2+|V_{ud}|^2+|V_{us}|^2+|V_{ub}|^2}\
\par
\begin{figure}[tb]
\begin{center}
\includegraphics[width=8cm]{alvcsfig.eps}
\end{center}
\caption{Distribution of the Fisher Discriminant for (a) semileptonic
events (b) fully-hadronic events (c) the sum of the two classes}
\label{fig:alephfd}
\end{figure}
\par
For this reason, a measurement of the charm fraction in W decays is a direct
test of the unitarity of the CKM matrix.
Furthermore, using the precise determinations of the other elements, it is
possible to convert this measurement into a determination of $|V_{cs}|$.
\par
This approach to the determination of this matrix element is less precise
than the derivation from the hadronic branching ratio, but more direct.
For charm tagging, the three experiments performing this measurement use
different experimental techniques:
\begin{itemize}
\item ALEPH (172+183 GeV data) use a neural network with 12 input
variables or a Fisher discriminator (see figure \ref{fig:alephfd}).
Variables used come from b-tagging, event
shapes, exclusive decays etc. A combination of the two methods is used for the
final result.
\item L3 (183 GeV data) splits jets into four categories, depending whether
an inclusive lepton ( $e,\mu$ ), a $D^*\to D^0\pi^\pm$
or none of the above is found. Each category is then analyzed by a separate
neural network.
\item DELPHI (172 GeV data) exploits its RICH detector for kaon
identification, thus directly measuring $V_{cs}$ through an s-tag in addition
to a charm tag performed in a similar way as the other experiments.
\end{itemize}
The results obtained are summarized in table \ref{tab:vcs}\cite{vcsexp}.
\begin{table}[H]
\begin{center}
\begin{tabular}{|c|c|}\hline
&$|V_{cs}|$\\ \hline
$D\to Kl\nu$(PDG)&$1.04\pm 0.16$\\ \hline
ALEPH&$1.00\pm0.10\pm0.06$\\
L3&$0.98\pm0.22\pm0.08$\\
DELPHI&$0.91\pm0.14\pm0.05$\\ \hline
LEP direct&$0.96\pm0.09$\\ \hline
LEP BR(W$\to$qq)&$1.03\pm0.04$\\ \hline
CKM unitarity&$0.9745\pm0.0005$\\ \hline
\end{tabular}
\vspace{1cm}
\caption{\label{tab:vcs}Determinations of $|V_{cs}|$}
\end{center}
\end{table}
\par
\section{Conclusions}
After three years of data taking above the WW threshold, W physics at LEP has
reached the realm of precision measurements. Data have been collected at
161, 172, 183 and 189 GeV of center of mass, with increasing integrated
luminosity. The selection of WW events provides measurements of the WW
total cross section and of the branching ratios of all decay channels.
All these measurement are in agreement with the predictions from the standard
electroweak theory. A sector where deviation from the standard theory could be
expected is the study of the trilinear gauge coupling. Fits to the cross
section and to the kinematics of the events so far are in agreement with the
expectations, so limits on the presence of anomalous couplings are established.
Charm production in W decays provides a direct determination of $|V_{cs}|$,
while an indirect measurement can be derived from the hadronic branching
fraction. The most important measurement in W physics at LEP is the W mass.
After a derivation from the threshold cross section in the first run, the mass
is now measured using the direct reconstruction of the invariant mass of the
W decay products. The present combined result $M_W=80.368\pm0.065$ is as
precise as the one obtained from hadronic machines, and its accuracy will
further improve in the next years. The question is still open on whether the
accuracy on the W mass will be finally dominated by systematic uncertainties.
Some systematic errors are likely to considerably improve with more study, but
the large uncertainties associated with QCD interactions in the four jet
channel may be not so easy to reduce, and limit the final accuracy of the
result.\par
LEP2 will run for two more years, aiming to reach the design integrated
luminosity of 500 $pb^{-1}$ and trying to reach the centre-of-mass energy of
200 GeV. W physics will fully profit from
more data and from more energy points, and the results from the four LEP
experiments will increase their sensitivity both in the measurement of
the parameters of the standard theory and in the search for new phenomena.
\par
\newpage
\section{Acknowledgements}
I would like to thank M. Pohl for having introduced me to the field of W
physics.
Many thanks also to M.Grunewald, L.Malgeri and A.Tonazzo and for
careful reading of the manuscript and A.Rubbia for the encouragement.
\nonumsection{References}
\noindent
|
\section{Introduction}
The Standard Model (SM) fermions forming complete multiplets of a
single gauge group, and the unification of the SU(3)$_C$, SU(2)$_W$,
and U(1)$_Y$ gauge couplings of the Minimal Supersymmetric Standard
Model (MSSM) at $\sim 10^{16}$ GeV are strong suggestions that there
is a grand unified gauge group (SU(5) or bigger) at a very high energy
scale. However, the successful prediction of the gauge coupling
unification and the proton decay constraint require the triplet
partners of the two light Higgs doublets to have masses of the order
of the grand unification scale. The ``doublet-triplet splitting''
problem is the most problematic aspect of a grand unified theory
(GUT). There exist several solutions to this problem. However, to get
a complete grand unified model which incorporates these solutions in
a simple and appealing way is not easy.
One of the most appealing solutions to the ``doublet-triplet
splitting'' problem is the pseudo-Goldstone bosons (PGB)
mechanism~\cite{PGB,BDM,BCR}, where the Higgs doublets remain light
because they belong to the pseudo-Goldstone multiplets coming from
breaking of the enlarged global symmetry of the Higgs superpotential.
Here we briefly review it. The model is based on the gauge group
SU(6). The Higgs sector consists of an adjoint ({\bf 35}), $\Sigma$,
and a pair of fundamental ({\bf 6}) and anti-fundamental ($\bar{\bf
6}$), $H$ and $\bar{H}$. Provided no cross coupling exists between
$\Sigma$ and $H$, $\bar{H}$, there is an effective SU(6)$_\Sigma
\times$SU(6)$_H$ symmetry of the Higgs sector. The $\Sigma$ and $H$,
$\bar{H}$ Higgses develop the following vacuum expectation values (vevs),
\beq
\label{vevs}
\vev{\Sigma}&=& {1\over \sqrt{12}}{\rm diag}
(1, 1, 1, 1, -2, -2) v_\Sigma ,\\
\label{vevh}
\vev{H}&=&\langle\bar{H}\rangle =(1, 0, 0, 0, 0, 0) v_H.
\eeq
These two SU(6)'s are then broken down to
SU(4)$\times$SU(2)$\times$U(1) and SU(5) respectively, while the SU(6)
gauge symmetry is broken down to the SM gauge group
SU(3)$_C\times$SU(2)$_\times$U(1)$_Y$. The successful prediction of
the $\sin^2 \theta_W$ is preserved if $v_H > v_\Sigma$. After
counting the number of the Goldstone modes and the broken gauge
generators, one finds that there are two electroweak doublets not
eaten by the gauge bosons and hence left massless. They are linear
combinations of the electroweak doublets coming from $\Sigma$ and $H,
\bar{H}$ fields,
\be
h_u={v_H H_{\Sigma} - \sqrt{3 \over 4}v_\Sigma H_H \over \sqrt{v_H^2
+{3\over 4}v_{\Sigma}^2}}, \,\,\,
h_d={v_H \bar{H}_{\Sigma} - \sqrt{3\over 4} v_\Sigma \bar{H}_{\bar{H}} \over
\sqrt{v_H^2 +{3\over 4}v_{\Sigma}^2}},
\ee
which parametrize the flat direction of the relative orientations of
the $\Sigma$ and $H$, $\bar{H}$ vevs. After including the soft
supersymmetry (SUSY) breaking terms, the vevs will shift by an amount
of the order of soft SUSY breaking parameters. This shift naturally
generates a $\mu$ term, $\mu h_u h_d$, of the same order as the soft
SUSY breaking terms from the superpotential. The Higgs potential
\be
V(h_u, h_d)= m_1^2 h_d^\dagger h_d + m_2^2 h_u^\dagger h_u
+ m_3^2 (h_u h_d + \mbox{h.c.}) + D\mbox{-terms},
\ee
(where $m_1^2=m_d^2+\mu^2$, $m_2^2=m_u^2+\mu^2$, $m_3^2=B\mu$,
$m_u^2$, $m_d^2$ are soft SUSY breaking mass terms for the
up- and down-type Higgses and $B$ is the soft SUSY breaking
parameter associated with the $\mu$-term,) satisfies the
boundary condition
\be
m_1^2(M_G)=m_2^2(M_G)=-m_3^2(M_G)
\ee
at the GUT scale $M_G$, so that there are two massless doublet bosons
corresponding to the Goldstone modes. This provide a constraint on the
phenomenology of this model~\cite{BDM,CR}. The effective
SU(6)$\times$SU(6) symmetry is explicitly broken by the couplings of
matter fields to both $\Sigma$ and $H\bar{H}$. The radiative corrections
from these couplings lift the flat direction and it is possible to
obtain the desired electroweak symmetry breaking Higgs potential after
running down to the low scale~\cite{BDM,CR}.
The problem of this model is that the cross coupling $H\Sigma \bar{H}$ is
allowed by the gauge symmetry. If it exists, it destroys the
SU(6)$_\Sigma\times$SU(6)$_H$ global symmetry of the Higgs sector and
therefore the PGB mechanism for the light Higgs doublets. Some extra
discrete symmetries or larger gauge symmetry are needed to forbid this
coupling~\cite{BDM,BCR,Be}. Besides, as one expects from the quantum
gravity effects, all higher dimensional operators suppressed by the
Planck scale ($M_{\rm Pl}$), which are allowed by symmetries, might be
present and have ${\cal O}(1)$ coefficients. If that is true, then
because $M_{\rm GUT}/M_{\rm Pl}$ is not a big suppression factor, the extra
symmetries have to forbid the cross couplings between $\Sigma$ and
$H$, $\bar{H}$ to very high orders. This may require some unappealing
symmetries or charge assignments. It would be desirable to have some
better ways to suppress these unwanted couplings. As we will see in
the next section, it can be naturally achieved if there are extra
dimensions in which the gauge and some matter fields can propagate
while $\Sigma$ and $H$, $\bar{H}$ are localized on two separate
branes. This provides a different mechanism to forbid the unwanted
interactions from symmetry reasons.
Another problem of this model is how to obtain the SM fermion masses.
In SU(6) models, a family of light matter fields (quarks and leptons)
can be contained in $\bf 15\, +\, \ov{6}\, +\, \ov{6}$, which is the
smallest anomaly-free combination of chiral representations. However,
there is no renormalizable Yukawa coupling between the light fermions
belonging to $\bf 15\, +\, \ov{6}\, +\, \ov{6}$ and the light
Higgses. In order to get the large top Yukawa coupling, one can
introduce a $\bf 20$, a pseudo-real representation, which contains the
top quark (${\bf 10}_3$ of SU(5)), then the top quark is naturally the
only one which can receive an ${\cal O}(1)$ coupling from the
interaction ${\bf 20}\,\Sigma \,{\bf 20}$. Other fermions can get
masses from the nonrenormalizable operators and therefore are
naturally suppressed. However, if all nonrenormalizable operators
consistent with the gauge symmetry exist, a realistic fermion mass
pattern is not obtained. Therefore, one also has to introduce extra
discrete symmetries and assume that the higher dimensional operators
are generated by integrating out some heavy vector-like
fields~\cite{BDSBH,Be} in order to obtain a realistic pattern of
fermion masses and mixings. In section~\ref{fermion} we will find
that the geometry of extra dimensions and branes for the PGB mechanism
can also help to explain the fermion mass hierarchies without
appealing to flavor symmetries.
\section{Doublet-triplet splitting in extra dimensions}
Now let us discuss how the doublet-triplet splitting and fermion mass
hierarchies can naturally arise in the scenario with extra dimensions
and branes. We assume that the SU(6) gauge field propagates in the
bulk of a $4+n$ dimensional space-time with $n$ dimensions of space
compactified with a radius $R$. The two kinds of Higgses $\Sigma$ and
$H$, $\bar{H}$ on the other hand are localized on two parallel 3-branes
separated by a distance $r$ in the $4+n$ dimensions, so there is no
direct interaction between them. Extra dimensions with
compactification radius larger than the Planck (or string) length have
been considered in string theory~\cite{An,Ly,HW,Co}; they have been
used to lower the unification scale~\cite{DDG}; a very large
compactification radius can even push the fundamental Planck scale,
$M_*$, down to ${\cal O}$(TeV), providing an alternative solution to
the hierarchy problem~\cite{ADD,ADD2,ADM}. In this paper we consider the
compactification of extra dimensions occurs at high energies, around
the GUT scale, so that the successful gauge coupling unification still
works in the traditional way.\footnote{In fact, the simple SUSY GUT
prediction of the strong coupling constant is a little higher than
the experimental value. If $1/R$ is smaller than $M_{\rm GUT}$, the
contribution from the Kaluza-Klein states of the gauge fields will
lower the prediction of the strong coupling constant, so it may be
favorable to have $1/R$ a little bit lower than $M_{\rm GUT}$. We will not
discuss this in details. See the Refs.~\cite{DDG,Carone} for the
discussions of gauge coupling unification.} We assume that the
distance between these two 3-branes, $r$, is much smaller than the
compactification radius $R$, but larger than the fundamental Planck
distance $1/M_*$, so that we can still use the field theory
description without dealing with the full quantum gravitational
theory.\footnote{The validity of the field theory description may be
questioned at the scale very close to $M_*$. However, without knowing
how to describe the full quantum gravitational theory, we assume that
just beneath $M_*$, physics can be described by the usual field theory
with the gravity effects included in the higher dimensional
interactions suppressed by $M_*$.}
Therefore we have
\be
M_{\rm GUT} \, \mathop{}_{\textstyle \sim}^{\textstyle <} \,\frac{1}{R}\, < \,\frac{1}{r}\, < \,M_* \,< \,M_{\rm Pl},
\ee
where $M_{\rm Pl} \simeq 2.8\times 10^{18}$GeV is the effective
4-dimensional Planck scale. If there are no additional large extra
dimensions in which gravitons propagate,
$M_{\rm Pl}$ is related to $M_*$ and $R$ by~\cite{ADD,ADD2}
\be
\frac{M_{\rm Pl}^2}{M_*^2} = M_*^n R^n .
\ee
This relation can be modified if there are additional large
dimensions in which gravitons propagate.
On these two branes, we assume for simplicity that the Higgs
superpotential takes the simple form,
\beq
W_1&=&m_\Sigma \Sigma^2 +\Sigma^3, \\
W_2&=& S(H\bar{H} -v_H^2),
\eeq
where $S$ is a singlet field, so that $\Sigma$ and $H$, $\bar{H}$
acquire vevs of the form given by Eqs.(\ref{vevs}),(\ref{vevh}).
To preserve the gauge coupling unification we need
\be
M_{\rm GUT}\, = \, v_\Sigma \, < \, v_H \, (<\, M_*),
\ee
so that the light Higgses are predominantly contained in $\Sigma$.
How exactly they acquire such vevs is not important, and they may
be generated dynamically~\cite{GUTscale}.
Most of the SM matter fields as well as some additional heavy
vector-like fields live in the bulk. We assume that the extra
dimensions are compactified on an orbifold so that the unwanted zero
modes are projected out and we can get chiral multiplets in four
dimensions.
At low energies, the four dimensional effective Lagrangian obtained
from integrating out the extra dimensions (or equivalently from a
pure four dimensional point of view, integrating out the heavy
Kaluza-Klein towers of the bulk fields) will contain light fields
coming from both branes and the bulk.
The SU(6)$_\Sigma\times$SU(6)$_H$ global symmetry on these
two branes is broken by the couplings of the matter fields living in
the bulk to the Higgses on both branes. Nevertheless, if we assign a
matter parity (which is equivalent to the $R$-parity of the MSSM)
$-1$ to all fields living in the bulk, and $+1$ to the Higgses
$\Sigma$ and $H$, $\bar{H}$, then any interaction between the $\Sigma$
or $H$, $\bar{H}$ and the bulk fields must contain at least two fields
with parity $-1$.
One can easily see that any diagram having both $\Sigma$ and
$H$, $\bar{H}$ as only external lines contains a loop at least.
By the non-renormalization theorem, no direct superpotential couplings
between $\Sigma$ and $H$, $\bar{H}$ (and containing no matter fields) at
any order can be generated after integrating out the extra
dimensions.
Thus, the PGB mechanism for the doublet-triplet splitting
can work naturally in this scenario.
\section{Fermion masses}
\label{fermion}
Before getting into the details of the fermion masses and mixings in
the Standard Model, we first discuss in general the possible
suppressions of couplings we may get in such a scenario. In addition
to the usual Planck mass suppression for the higher dimensional
operators, the suppressions may also come from the large volume factor
of the extra dimensions and from integrating out the vector-like bulk
fields and their Kaluza-Klein excitations.
The couplings of an (external) bulk field (after integrating
out the extra dimensions and heavy fields) to the brane fields
are suppressed by the volume factor of the extra dimensions.
The zero modes contain an $R^{-n/2}$ factor after the Fourier
decomposition to match the mass dimensions of fields in different
space dimensions, so the dimensionless coefficients of the couplings
are naturally suppressed by~\cite{ADD2,AD}
\be
\epsilon \equiv (M_* R)^{-\frac{n}{2}} \left( = \frac{M_*}{M_{\rm Pl}},
\; \mbox{if no additinal large dimensions for gravitons}\right).
\ee
This may explain the weakness of the unified gauge coupling at the GUT
scale. To get ${\cal O}(1)$ Yukawa coupling for the top quark, we
therefore assume that the $\bf 20$ (denoted by $\eta$, with matter
parity $-1$) containing the top quark lives on the same brane in which
$\Sigma$ resides, then the $\eta\, \Sigma\, \eta$ interaction which
contains the top Yukawa coupling is naturally ${\cal O}(1)$. All
other matter and vector-like fields are assumed to live in the bulk.
Light fermion masses come from higher dimensional operators. Higher
dimensional operators can already be present in the fundamental
Lagrangian (suppressed by powers of $M_*$) if they involve fields in
the bulk and on one brane only. They can also be generated by
integrating out the heavy vector-like fields in the bulk and extra
dimensions if they contain fields on both branes. This is somewhat
similar to the Froggatt-Nielsen mechanism~\cite{FN}. However, the
suppression of these higher dimensional operators is different. It
also depends on the transverse distance $r$ between the two branes and
the number of extra dimensions, as there is a tower of the
Kaluza-Klein states of the vector-like fields. The case when there are
vector-like scalars connecting two branes is discussed in
Ref.~\cite{AD}. It is simply the Yukawa potential (or the propagator
of the vector-like field) in the $n$ transverse direction. The
generalization to the supersymmetric case is straightforward. The
propagator in the transverse direction is
\be
\Delta_V (r)= \int d^n \kappa \; e^{i\kappa r} {-i\kappa+m_V
\over \kappa^2+m_V^2}.
\ee
One gets an exponential suppression ($e^{-m_V r}$) if the mass of the
vector-like fields $m_V$ is larger than $1/r$, and a power suppression
($r^{1-n}$) if $m_V$ is smaller than $1/r$. For one extra dimension
and $m_V < 1/r$, there can even be no suppression. We will
parametrize the dimensionless suppression factor (in the unit of
$M_*$) by $\delta_V$ ({\it e.g.}, $(M_* r)^{-a}$ in the case of power
suppression). We will find that to obtain successful fermion masses
some of the suppression factor from integrating out the vector-like
fields should be ${\cal O}(1)$, so it is favorable to have just one
extra dimension.
In the bulk, there are three sets of
${\bf 15\, +\, \ov{6}\, +\, \ov{6}}$ chiral matter
multiplets, denoted by $\psi_i({\bf 15}),\; \bar{\phi}_i({\bar{\bf 6}}),\;
\bar{\phi}'_i({\bar{\bf 6}}),\; i=1, 2, 3$. In addition, we assume
that there are 3 pairs of vector-like fields of the SU(6)
representations $({\bf 20}_1,\, {\bf 20}_2)$, ${\bf (6,\, \ov{6})}$,
and ${\bf (70,\, \ov{70})}$
with masses (of the zero modes) $m_{20}$, $m_6$, and $m_{70}$
respectively. They all have matter parity $-1$. The field content
is summarized in Table~\ref{content}.
\begin{table}[htb]
\begin{center}
\begin{tabular}{|c|c|c|} \hline
Brane 1 & Bulk & Brane 2 \\ \hline
$\Sigma$ & SU(6) gauge field,
$\psi_i,\, \bar{\phi}_i, \, \bar{\phi}'_i, \, i=1, 2, 3$ &
$H, \, \bar{H}$ \\
$\eta$ & ${\bf 20}_1, {\bf 20}_2, \, {\bf 6, \ov{6}, \, 70,
\ov{70}}$ & $S, \; (N)$ \\
\hline
\end{tabular}
\end{center}
\caption{Field content in the bulk and on the two branes.
As it will be discussed later, the singlet field $N$ is included
if we need to generate the neutrino mass to account for the
atmospheric neutrino oscillation.
\label{content}}
\end{table}
In terms of the usual SU(5)$_{\rm GUT}$ subgroup, these
representations decompose into:
\beq
{\bf 6} &=& {\bf 1+5, \quad \ov{6}=1+\ov{5},} \nonumber \\
{\bf 15} &=& {\bf 5+10,} \nonumber \\
{\bf 20} &=& {\bf 10 + \ov{10},} \nonumber \\
{\bf 35} &=& {\bf 1+5+\ov{5}+24,} \nonumber \\
{\bf 70} &=& {\bf 5+10+15+\ov{40}, \quad \ov{70}=\ov{5}+\ov{10}+\ov{15}+40,}
\eeq
Integrating out the vector-like fields and the extra dimensions, we
obtain the operators appearing in the low energy effective four dimensional
theory. The dimensionless coefficient (after factorizing out powers of
$M_*$ of the dimensionful coupling) of an operator is suppressed by a
power of $\epsilon$ for each external bulk field, and by $\delta_V$
if it is generated by integrating out the vector-like fields $V,
\ov{V}$. For Yukawa couplings coming from nonrenormalizable
operators, they will also be suppressed by $v_H/M_*$ or
$v_\Sigma/M_*$. If the light Higgs doublets come from $H$, $\bar{H}$,
there is a further suppression of the mixing angle $\sim v_\Sigma
/v_H$. In the following we discuss these effective operators and the
SM fermion masses and mixings arising from them. For simplicity and
the organization purpose, the operators are written in terms of
the SU(6) language. What we really mean are the operators involving
the light fields contained in those operators, since the heavy fields
($\sim M_{\rm GUT}$) in these operators should
have been integrated out too.
For example,
$\eta \Sigma \psi_2 H$ means ${\bf 10}_{\eta} {h_u}_{\Sigma}
{\bf 10}_{\psi_2} \langle H \rangle$ (and the higher order
term ${\bf 10}_{\eta}
\langle \Sigma \rangle {\bf 10}_{\psi_2} {h_u}_ H$).
\begin{description}
\item [Operators which decouple the extra states] (Fig.~\ref{heavy}):
\begin{figure}[htbp]
\centerline{\psfig{file=figdiag1.ps,width=0.9\textwidth}}
\caption{Diagrams which decouple the extra states. Fields on the
left of the left interaction point are on brane 1. Fields on the
right of the right interaction point are on brane 2. Fields between
the interaction points are in the bulk. }
\label{heavy}
\end{figure}
\begin{itemize}
\item
$\eta H \psi_3$ (diagram (a)): We can define $\psi_3$
with this operator by using the rotation freedom among $\psi_i$'s,
and ${\bf 20}_1$ to be the one which couples to $\eta$. The ${\bf 10}$
(of SU(5)$_{\rm GUT}$) in $\psi_3$ and $\ov{\bf 10}$ in $\eta$
become heavy due to $\langle H \rangle = v_H$, leaving only
three {\bf 10}'s (in $\eta, \, \psi_2 , \, \psi_1$) in the low energies.
\item
$\psi_i \bar{H} \bar{\phi}'_j$ (diagram (b)): The {\bf 5}'s in
$\psi_i$ and $\ov{\bf 5}$'s in $\bar{\phi}'_j$ are married by
$\langle \bar{H} \rangle$, leaving only three $\ov{\bf 5}$'s (in
$\bar{\phi}_i$) in the low energies.
\end{itemize}
Because of the suppresion factors involved, some decoupled states
will have masses a little bit lower than $M_{\rm GUT}$. However, they are
complete SU(5) multiplets, so they do not affect the coupling
unification. We can see that $\psi_3$ is completely decoupled, so we
will drop it in the following discussion. In SU(5) notation, the three
light generations are contained in the ${\bf 10}$'s of $\eta, \,
\psi_2, \, \psi_1$, and $\ov{\bf 5}$'s of $\bar{\phi}_3, \,
\bar{\phi}_2, \, \bar{\phi}_1$. They are the only SM non-singlets
matter fields left massless at this stage. (The SU(5) singlets can
also be decoupled. It will be seen when we discuss neutrino masses.)
\item [Up-type quark masses] (Fig.~\ref{up}):
\begin{figure}[htbp]
\centerline{\psfig{file=figdiag2.ps,width=0.9\textwidth}}
\caption{Diagrams which generate the up-type quark masses.}
\label{up}
\end{figure}
\begin{itemize}
\item
$\eta \Sigma \eta$ (diagram (c)): It contains ${\bf 10}_3
{\bf 5}_{\Sigma} {\bf 10}_3$ in SU(5) notation.
There is no suppression and therefore
it gives an ${\cal O}(1)$ Yukawa coupling to the top quark.
\item
$\eta \Sigma \psi_2 H$ (diagram (d)): One can rotate $\psi_i,\;
i=1,2$, so that only $\psi_2$ couples to $\ov{\bf 70}$ and $H$.
It generates the 23 and 32 elements of the up Yukawa matrix
of ${\cal O}(\epsilon \delta_{70} ({v_H \over M_*}))$.
\item
$\eta \psi_i H (H \bar{H})$ (diagram (e)): It does not contain $\Sigma$,
so the light Higgs has to come from $H$, which causes a
$({v_{\Sigma}\over v_H})$ mixing suppression. It generates 13, 31, 23, 32
elements of ${\cal O}(\epsilon \delta_{20} ({v_H\over M_*})^2
({v_{\Sigma} \over v_H}))$.
\item
$\Sigma \psi_2 \psi_i H H$ (diagram (f)): It generates 22, 12, 21
elements of ${\cal O}(\epsilon^2 \delta_{20} \delta_{70}
({v_H \over M_*})^2)$.
\item
$\Sigma \psi_i \psi_j H H (H \bar{H})$ (diagram (g)): One can always
attach a pair of $(H \bar{H})$ to the interaction on the brane 2,
which will be suppressed by an extra $(v_H/M_*)^2$. This gives
the leading contribution to the 11 element.
\end{itemize}
In the leading order the up-type Yukawa matrix looks like
\be
\lambda_U \sim \left( \begin{array}{ccc}
u_4 & u'_3 & u_2 \\
u'_3 & u_3 & u_1 \\
u_2 & u_1 & 1
\end{array} \right),
\ee
where
\beq
u_1 &\sim & \epsilon \delta_{70} \left({v_H \over M_*}\right),\\
u_2 &\sim & \epsilon \delta_{20} \left({v_H \over M_*}\right)^2
\left({v_{\Sigma} \over v_H}\right),\\
u_3,\, u'_3 &\sim & \epsilon^2 \delta_{20} \delta_{70}
\left({v_H \over M_*}\right)^2,\\
u_4 &\sim & \epsilon^2 \delta_{20} \delta_{70}
\left({v_H \over M_*}\right)^4.
\eeq
If we take $\epsilon \sim {1\over 3}$, ${v_H \over M_*} \sim
{1\over 5}$, ${v_{\Sigma} \over v_H} \sim {1\over 3}$,
$\delta_{70} \sim {1\over 2}$, and $\delta_{20} \sim {1\over 40}$,
then we have at the GUT scale,
\beq
\lambda_t &\sim & 1 ,\\
\lambda_c &\sim & u_1^2 \sim \epsilon^2 \delta_{70}^2
\left({v_H \over M_*}\right)^2 \sim 10^{-3} ,\\
\lambda_u &\sim & u_4, \, {{u'}_3^2\over u_1^2} \,
\mbox{(two comparable contributions)} \sim (2-3)\times 10^{-6} .
\eeq
Remember that the light fermion Yukawa couplings will increase
in renormalization group (RG) running to low energies while the
top Yukawa coupling will roughly approach some fixed point.
The mass ratios of light quarks to the top quark will be enhanced
by a factor of 5--10 relative to those at the GUT scale.
After taking into account the RG effect, the above numbers give
a good approximation to the up-type quark masses.
In diagonalizing the mass matrix, the 23 rotation angle $U_{23}
\sim u_1 \sim 3\times 10^{-2}$ is about the same order as
$V_{cb}$. Other rotation angles are much smaller than the
corresponding Cabibbio-Kobayashi-Maskawa (CKM) matrix elements,
so they have to be generated from the down sector.
\item [Down-type quark and charged lepton masses] (Fig.~\ref{down}):
\begin{figure}[htbp]
\centerline{\psfig{file=figdiag3.ps,width=0.9\textwidth}}
\caption{Diagrams which generate the down-type quark and
charged lepton masses.}
\label{down}
\end{figure}
\begin{itemize}
\item
$\eta \Sigma \bar{\phi}_3 \bar{H} \bar{H}$ (diagram (h)): We can use
the rotation freedom among $\bar{\phi}_i$'s to define $\bar{\phi}_3$
with this operator. This gives the 33 elements of the down and charged
lepton Yukawa matrix. In leading order, the Higgs doublet to which the
fermions couple comes from $\Sigma$. They are ${\cal O}(\epsilon
\delta_{70} ({v_H \over M_*})^2)$ and the same for the down-type
quark and the charged lepton, so we have appproximate $b-\tau$
unification.
\item
$\Sigma \psi_2 H \bar{\phi}_3 \bar{H} \bar{H}$ (diagram (i)): $\psi_2$
and $\bar{\phi}_3$ have been defined before. This operator
contributes to the 23 elements of the down and lepton Yukawa
matrices and is ${\cal O}(\epsilon^2 \delta_{70}^2
({v_H\over M_*})^3)$.
\item
$\langle\Sigma\rangle \bar{\phi}_i \psi_j \bar{H}$ (diagram (j)):
This operator only redefines $\bar{\phi}'_i$ and is irrelevant
for fermion masses~\cite{BDSBH}.
\item
$\langle\Sigma\rangle \bar{\phi}_i \psi_j \bar{H} (H \bar{H})$ (diagram (k)):
Attaching $(H \bar{H})$ to the previous diagram, we can get a diagram
contributing to the fermion masses. We can rotate $\bar{\phi}_i$
to have only $\bar{\phi}_2,\, \bar{\phi}_3$ in this operator.
If we had not defined $\psi_i$'s in the up sector, we could also
have defined $\psi_2$ by this operator, then it would have contributed
only to the 22 and 23 elements of the mass matrices. The rotation
angle between the two bases will in general be ${\cal O}(1)$,
which accounts for why the Cabibbo angle is large. In the basis
used for the up sector, it contributes to the 12, 13, 22, 23
elements of the down and charged lepton mass matrices (with
12, 13 elements smaller than 22, 23 elements by $\sim \sin\theta_C
\sim 0.2$). An important fact of this operator is that the
intermediate states $({\bf 6},\, \ov{\bf 6})$ do not contain
${\bf 10}, \, \ov{\bf 10}$ of SU(5)$_{\rm GUT}$, so the light Higgs
doublet has to come from $\bar{H}$. The contribution of this
operator to the Yukawa couplings is therefore ${\cal O}
(\epsilon^2 \delta_6 ({v_{\Sigma}\over M_*})({v_H \over M_*})^2
({v_{\Sigma}\over v_H}))$. The vev of $\Sigma$ gives a ratio of
$1 : -2$ to the down-type quark and the charged lepton Yukawa
matrix elements. Since it is the dominant term
to the 22 elements and hence the leading contribution to the
second generation masses, this offers an explanation of the
discrepancy between $m_s$ and $m_{\mu}$ from the simple unification
relation.
\end{itemize}
Other matrix elements and non-leading contributions can be
obtained by attaching more $(H \bar{H})$ or $\Sigma$ to previous
diagrams. In the following we only discuss the leading
contributions.
\begin{itemize}
\item
$\eta \Sigma \bar{\phi}_i \bar{H} \bar{H} (H \bar{H})$ (diagram (l)):
This gives the leading contribution to the 31, 32 elements of
the down and lepton mass matrices of ${\cal O}(\epsilon
\delta_{70} ({v_H \over M_*})^4)$.
\item
$\Sigma \psi_2 H \bar{\phi}_i \bar{H} \bar{H} (H \bar{H})$ (diagram (m)):
This gives the leading contribution to the
21 elements of the down and lepton mass matrices of
${\cal O}(\epsilon^2 \delta_{70}^2 ({v_H \over M_*})^5)$.
\item
$\langle\Sigma\rangle^2 \bar{\phi}_i \psi_j \bar{H} (H \bar{H})$ (diagram (n)):
This gives the leading contribution to the 11 elements of
${\cal O}(\epsilon^2 \delta_6 ({v_{\Sigma}\over M_*})^2
({v_H \over M_*})^2 ({v_{\Sigma}\over v_H}))$.
\end{itemize}
In the leading order the down-type Yukawa matrix looks like
\be
\lambda_D \sim \left( \begin{array}{ccl}
d_6 & sd_3 & sd'_3 \\
d_5 & d_3 & d_2(+d'_3) \\
d_4 & d_4 & d_1
\end{array} \right),
\ee
where
\beq
d_1 &\sim & \epsilon \delta_{70} \left({v_H \over M_*}\right)^2, \\
d_2 &\sim & \epsilon^2 \delta_{70}^2 \left({v_H \over M_*}\right)^3, \\
d_3,\, d'_3 &\sim & \epsilon^2 \delta_{6}
\left({v_{\Sigma} \over v_H}\right)^2 \left({v_H \over M_*}\right)^3, \\
d_4 &\sim & \epsilon \delta_{70} \left({v_H \over M_*}\right)^4, \\
d_5 &\sim & \epsilon^2 \delta_{70}^2 \left({v_H \over M_*}\right)^5, \\
d_6 &\sim & \epsilon^2 \delta_{6} \left({v_{\Sigma} \over v_H}\right)^3
\left({v_H \over M_*}\right)^4,
\eeq
and we have explicitly put in the Cabibbo angle $s\sim 0.2$.
Again, taking the previous assumed suppression factors,
$\epsilon \sim {1\over 3}$, ${v_H \over M_*} \sim
{1\over 5}$, ${v_{\Sigma} \over v_H} \sim {1\over 3}$,
$\delta_{70} \sim {1\over 2}$, $\delta_{20} \sim {1\over 40}$,
with $\delta_6 \sim 1$, we have the following relations at
the GUT scale,
\beq
{\lambda_b \over \lambda_t} &\sim & \epsilon \delta_{70}
\left({v_H \over M_*}\right)^2 \sim 10^{-2}, \\
{\lambda_s \over \lambda_b} &\sim & {d_3\over d_1} \sim
\epsilon {\delta_6 \over \delta_{70}} \left({v_{\Sigma}
\over v_H}\right)^2 \left({v_H\over M_*}\right) \sim
1.5\times 10^{-2}.
\eeq
In running down to low energies, we get enhancements of
$\sim 3-5$ for ${\lambda_b(m_b)\over \lambda_t(m_t)}$
and $\sim 2$ for ${\lambda_s(1{\rm GeV})\over \lambda_b(m_b)}$.
There are several comparable leading contributions to $\lambda_d$
after diagonalization, {\it e.g.,} $d_6,\, {(sd_3 d_2 d_4)/ (d_1 d_3)},
\, {(sd_3 d_1 d_5)/ (d_1 d_3)}$. In terms of ${\lambda_d \over
\lambda_s}$, they are
\beq
{d_6 \over d_3} &\sim & \left({v_{\Sigma}\over v_H}\right)
\left({v_H\over M_*}\right) \sim {1\over 15}, \\
{sd_3 d_2 d_4\over d_1 d_3^2} &\sim & {s \delta_{70}^2
\left({v_H\over M_*}\right)^2 \over \delta_6 \left(
{v_{\Sigma} \over v_H}\right)^2} \sim 2\times 10^{-2}, \\
{sd_3 d_1 d_5\over d_1 d_3^2} &\sim & {s \delta_{70}^2
\left({v_H\over M_*}\right)^2 \over \delta_6 \left(
{v_{\Sigma} \over v_H}\right)^2} \sim 2\times 10^{-2}.
\eeq
The smallness of $\lambda_e$ may due to a mild cancellation
among these contributions. We can see that we also get a very
good approximation of the down quark and charged lepton masses
(for small $\tan \beta$).
The rotation angles for diagonalizing the down quark mass
matrix are
\beq
D_{23} &\sim & {d_2\over d_1} \sim \epsilon \delta_{70}
\left({v_H\over M_*}\right) \sim 3\times 10^{-2}, \\
D_{13} &\sim & {sd'_3 \over d_1} \sim s \epsilon
{\delta_6 \over \delta_{70}} \left({v_{\Sigma} \over v_H}\right)^2
\left({v_H\over M_*}\right) \sim 3\times 10^{-3}, \\
D_{12} &\sim & s \sim 0.2 \;({\cal O}(1)).
\eeq
$U_{23}$ and $D_{23}$ are comparable and their combination gives
$V_{cb}$. Other `CKM matrix elements are dominated by the rotation
of the down sector and they are all generated at the right
magnitudes. It is quite remarkable that without any extra flavor
symmetry and allowing most general operators, the SM fermion
masses and mixings pattern is naturally obtained provided the various
mass scales are such that they produce the appropriate suppression
factors.
\item [Neutrino masses] (Fig.~\ref{neutrino}):
\begin{figure}[htbp]
\centerline{\psfig{file=figdiag4.ps,width=0.9\textwidth}}
\caption{Diagrams which generate the neutrino masses.}
\label{neutrino}
\end{figure}
Majorana masses of the left handed neutrinos can be generated by
diagram (o). This diagram also decouples the SU(5) singlets in
$\bar{\phi}_i$ (and also in $\bar{\phi}'_i$ by replacing
$\bar{\phi}_i$ by $\bar{\phi}'_i$). Because there is no distinction
among the three generations of $\bar{\phi}_i$, in general we have
large mixings among the neutrinos. The neutrino masses generated by
this diagram are of the order $\sim \epsilon^2 ({v_{\Sigma} \over
v_H})^2 {v_{\rm EW}^2\over M_*} \sim 10^{-5}-10^{-6}$~eV. This is in
the right range of explaining the solar neutrino problem through the
``just-so'' vacuum oscillation solution~\cite{justso}, but too small
to account for the atmospheric neutrino problem, which requires
$\delta m_{\rm atm} \sim 3\times 10^{-2} -
10^{-1}$~\cite{atm,kamio}. To accommodate a larger neutrino mass, one
can introduce a singlet field $N$ (with matter parity $-1$) on brane
2 (which contains $H\bar{H}$). Then, from diagram (p), one neutrino mass
of the order $\sim \epsilon^2 ({v_{\Sigma} \over v_H})^2 {v_{\rm
EW}^2\over M_N}$ can be generated. It will be in the right range for
the atmospheric neutrino problem if the mass of the singlet mass $M_N$
is $\sim 10^{13}-10^{14}$~GeV. The next neutrino mass obtained from
attaching $(H\bar{H})$'s to diagram (p) will be suppressed by $({v_H\over
M_*})^4 \sim 10^{-3}$, close to that required for the vacuum
oscillation solution of the solar neutrino problem.
\end{description}
\section{Conclusions}
In conclusion, extra dimensions and fields localized on branes provide
a new way to understand the absence or smallness of some couplings
without symmetry arguments~\cite{AD,ADDM,RS,AS}. This kind of idea
has been used to obtain small fermion masses in the Standard Model and
to suppress proton decay~\cite{AS}. Here we find that by localizing
two kinds of Higgses on two separate branes, the most difficult
``doublet-triplet splitting'' problem of the grand unified theory is
naturally solved by the pseudo-Goldstone boson mechanism. In addition,
after including several vector-like fields in the bulk, and allowing
the most general interactions consistent with the background geometry
and with their natural strength, all Standard Model fermion masses and
mixings can be correctly produced without any flavor symmetry. The
neutrino masses and mixings required for the solar and atmospheric
neutrino problems can also be easily accommodated. It is very
interesting that the complicated picture of the Standard Model can be
realized by such a simple model. Extra dimensions at such high
energies will not give us the exciting new collider signatures such as
production of the graviton Kaluza-Klein states. Nevertheless, it
gives a simple realization of the grand unified theory and the fermion
masses with the pseudo-Goldstone boson solution to the
``doublet-triplet splitting'' problem. If it is true, the boundary
condition of the Higgs parameters should be verified in the future
experiments.
{\bf Acknowledgements}
The author would like to thank N. Arkani-Hamed and B.A. Dobrescu
for discussion.
Fermilab is operated by Universities Research Association, Inc.,
under contract DE-AC02-76CH03000 with U.S. Department of Energy.
\newcommand{\Journal}[4]{{#1} {\bf #2} {(#3)} {#4}}
\newcommand{Ap. J.}{Ap. J.}
\newcommand{Can. J. Phys.}{Can. J. Phys.}
\newcommand{Nuovo Cimento}{Nuovo Cimento}
\newcommand{Nucl. Phys.}{Nucl. Phys.}
\newcommand{Mod. Phys. Lett.}{Mod. Phys. Lett.}
\newcommand{Phys. Lett.}{Phys. Lett.}
\newcommand{Phys. Rev.}{Phys. Rev.}
\newcommand{Phys. Rep.}{Phys. Rep.}
\newcommand{Phys. Rev. Lett.}{Phys. Rev. Lett.}
\newcommand{Prog. Theor. Phys.}{Prog. Theor. Phys.}
\newcommand{Sov. J. Nucl. Phys.}{Sov. J. Nucl. Phys.}
\newcommand{Z. Phys.}{Z. Phys.}
|
\section{Searches using the liquid scintillator
subdetector}
De R\'{u}jula and Glashow have calculated the light yield of nuclearites
traversing transparent materials on the basis of the black-body radiation
emitted along the heated track \cite{ruhula}. The light yield per unit
track length is (in natural units, $\hbar = c = 1$)
\begin{equation}
\frac{dL}{dX}=\frac{\sigma}{6 \pi^{2}\sqrt{2}}\omega^{5/2}_{max}(m/n)^{3/2}
v^{2},
\end{equation}
where $m$ is the mass of a molecule of the traversed material,
$n$ is the number of
submolecular
species in a molecule, $\omega _{max}$ is the maximum frequency for which
the material is transparent,
$\sigma$ is the nuclearite cross section and $v$ its velocity.
In Ref.~\cite{prl92} this formula was applied
to our liquid scintillator. It was shown that the scintillator
subdetector is sensitive even to very small nuclearite masses and to low
velocities ($\beta \simeq 5
\times 10^{-5}$). The
light yield is
above the 90\% trigger efficiency threshold of the MACRO scintillator
slow-particle trigger system for most nuclearite masses
($dL/dX > 10^{-2}$ \units{MeV \ cm^{-1}}).
The scintillators are therefore sensitive not
only to galactic
($\beta \sim 10^{-3}$) or extragalactic nuclearites
(higher velocities),
but also to those
possibly trapped in our solar system ($\beta \sim 10^{-4}$).
The nuclearite detection efficiency in the scintillator subdetector is
assumed to be similar (or larger) to that for magnetic monopoles;
the selection criteria used to search for monopole events are
also applicable for nuclearites.
No saturation effects of the detectors, electronics, or reconstruction
procedure are expected to reduce the detection efficiency of the
liquid scintillator subdetector.
Different monopole triggers and analysis procedures have been used
in the search for cosmic ray strangelets in different velocity domains following
the evolution of the detector.
The relevant parameters of each
actual search for nuclearites using the scintillator subdetector are
presented in Table~1. As no candidate satisfied all the requirements,
the resulting flux upper limits at 90\% confidence level (C.L.) are listed
in Table~1 and presented in Fig.~2.
Some details on the present searches are given below.
The slow monopole trigger which includes the analog Time Over Half Maximum
(TOHM) electronics and the digital Leaky Integrator (LI) electronics
\cite{nim93,pub96-2}
recognizes wide pulses or long trains of single photoelectrons generated
by slow particles, rejecting large and short pulses produced by muons
or radioactive decay products.
When a trigger occurs, the wave forms of
both the
anode and the dynode (for the 1989-91 run period) for
each photomultiplier tube are separately
recorded by two Wave Form Digitizers (WFD). A visual scan
is then performed on the selected events.
This procedure was applied to the searches for
nuclearites with $ 10^{-5} < \beta < 3.5 \times 10^{-3}~\cite{prl92,icrc97}$.\par
The Fast Monopole Trigger (FMT)
is based on the time of flight between
two layers of
scintillators. A slow coincidence between two layers is
vetoed by a fast coincidence
between them. Additional wave form analysis is performed on the selected events.
This procedure was applied to the search for intermediate
velocity nuclearites
($2.5 \times 10^{-3} < \beta < 1.5 \times 10^{-2}$) \cite{prl92}.
The scintillator muon trigger. A fast nuclearite should produce a light
yield ($dL/dx$) at least three orders of magnitude larger than that from a
typical muon. It was
checked that no negative effects arise
on the detecting system from the larger pulse heights.
No event was found having a $dL/dx$ in both walls greater than 10 times that of
a muon.
This technique
was applied in the early analyses for high velocity nuclearites
($ 1.5 \times 10^{-2} < \beta < 1$) \cite{prl92}.
The Energy Reconstruction Processor (ERP) is a
single-counter energy threshold trigger \cite{phr1,phr2}.
The ERP analysis requires triggers in two different scintillator
planes, separated
in the vertical direction by at least 2 \units{m}, insuring a time of flight long
enough for accurate velocity measurements. The energy deposition must be
at least 600 \units{MeV} in each counter.
The ERP analysis was used to search for fast nuclearites ($\beta > 0.1$)
\cite{icrc97}.
The raw triggering efficiency for nuclearites with $\beta > 0.1$
is essentially 100\%.
Pulse Height Recorder and Synchronous
Encoder (PHRASE)
is a system designed primarily for the detection of supernova
neutrinos \cite{phr1,phr2}. The event selection requires a coincidence between
two scintillator planes, with no more than 2 contiguous hits in each plane,
with an energy release of at least 10 \units{MeV} in each layer. It was checked that
no negative effects arise from larger pulse heights. A minimum separation of
2 \units{m} is required for hits in the two counters,
while a software cut ($\beta \leq
0.1$) is imposed in order to reject the tail of the cosmic ray muon
distribution. The particle velocity is reconstructed using the scintillator
time information. The PHRASE search for nuclearites
covers a large velocity range: $1.2 \times 10^{-3} < \beta < 10^{-1}$
\cite{icrc97}. The lower
limit corresponds to the threshold for the detection of bare
monopoles with unit Dirac magnetic charge ($g = g_D$);
as the light yield produced by nuclearites is larger than that
of monopoles, the nuclearite search might be extended to lower velocities.
For candidates with $\beta \leq 5 \times 10^{-3}$ we compare the duration
of the scintillation light pulse (measured by the PHRASE WFD) with the one
computed using the particle velocity; candidates with
$5 \times 10^{-3} \leq \beta < 0.1$
are cross-checked on the basis of the measured energy loss. All the candidates
with $\beta \simeq 5 \times 10^{-3}$ are examined using both techniques, in
order to ensure the continuity of the analysis.
\section{Searches using the nuclear track subdetector}
The nuclear track detector is located horizontally in the
middle of the lower MACRO structure, on the vertical east
wall and on the lower part of the vertical north wall. It is organised in
modules (``wagons'') of $\sim 25 \times 25 \units{cm^2}$; a ``wagon'' contains
three layers of CR39, three layers of lexan and 1 \units{mm} thick aluminium
absorber. Details of the
track-etch subdetector are given in Ref.~\cite{pub94-2}; the total area
is 1263 \units{m^2}.
At the point that this analysis ended we had etched
227 \units{m^2} of CR39 with an average exposure
time of 7.6 years.
In Ref. \cite {cr39} it was shown that the formation of an etchable track
in CR39 is related to the Restricted Energy Loss (REL) which is the fraction
of the total energy loss which remains localized
in a cylindrical region with about 10 \units{nm} diameter
around the particle trajectory \cite{benton}.
There are two contributions to REL:
the electronic energy loss ($S_e$), which represents the energy transferred
to the electrons, and the nuclear energy loss ($S_n$),
which represents the energy
transferred to the nuclei in the material. In Ref. \cite {cr39}
it was
shown that $S_n$ is as effective as $S_e$ in producing etchable tracks in our
CR39. This result was confirmed in Ref. \cite{yudong} for different
types of CR39.
In the case of nuclearites the REL is practically equal to $S_n$; thus Eq.~1
may be used for calculating REL.
In Fig.~3 we present the energy loss of nuclearites in CR39;
the calculation assumes that the
energy is transferred to the traversed material by displacing the matter
in the nuclearite path by elastic or quasi-elastic collisions. Such processes
would produce the
breaking of the CR39 polymeric bonds, leaving etchable latent
tracks, if the energy loss is above the detector threshold.
For the MACRO CR39 the ``intrinsic" threshold is
about 20 \units{MeV\ g^{-1} cm^2}
in the condition of a chemical etching in 8N NaOH water solution at
80\hbox{${}^\circ$}\units{C}; this is
shown in Fig.~3 as the lower horizontal line. The dotted line in Fig.~3
represents the REL for $g =g_D$ bare magnetic monopoles in CR39
\cite{derkaoui98}.
Several ``tracks'' were observed, mainly due to recoil protons
from neutron
interactions or due
to polymerization inhomogeneities.
In the conditions of the average exposure time in MACRO the number of
background tracks is about 0.5 \units{{/}m^2} of CR39.
When we required that the observed etch cones were
present on at least four CR39 surfaces and were consistent with
being from the same particle track,
all of the candidates were ruled out.
From Fig.~3 it is apparent that our CR39 is sensitive to nuclearites
of
any mass and with
$\beta > 1.5 \times 10^{-5}$.
Nuclearites with mass larger than $\sim 10^{15}$ \units{GeV/c^2}
can be detected even for velocities
as small as $\beta = 10^{-5}$.
As a consequence, the 90\% C.L. limit for $\beta \sim 1$
monopoles established
by the nuclear track subdetector ($6.8 \times 10^{-16}$ \units{
cm^{-2}s^{-1}sr^{-1}}) applies also to an isotropic flux of $ M >
5.6 \times 10^{22} \units{GeV/c^2}$ nuclearites.
This limit is presented in Fig.~2 as curve ``F"
and is included in Table~1.
For lower mass nuclearites the 90\% C.L. flux limit is twice this
value ($1.4 \times 10^{-15}$ \units{cm^{-2}s^{-1}sr^{-1}}) because
of the solid angle effect shown in Fig.~1.
\section{Discussions and conclusions}
No nuclearite candidate was found in any of the reported searches.
The 90\% C.L. flux upper limits for an isotropic
flux of nuclearites are presented in Fig.~2.
Because either the scintillator or the CR39 can give us a credible
nuclearite detection, we sum the independent parts of the individual
exposures
to obtain the global limit denoted as ``MACRO'' in
Fig.~2. This procedure ensures the 90\%\ C.L. significance of the
global limit.
All limits presented in Fig.~2 refer to the flux of nuclearites
at the level of the MACRO detector, i.e., below an average rock thickness
of 3700 \units{hg/cm^2}. To compare our limit with the limits
published by different experiments and with the limit calculated from DM
density in our galaxy,
we integrated the energy loss equation for a path
corresponding to the averaged rock thickness and for different velocities
at the detection level. Thus we obtained a relation between the nuclearite
velocities at the level of the detector and at the ground level.
Similar calculations were made for other underground experiments
\cite{orito,price}.
Fig.~4 shows the 90\% C.L. MACRO upper limit for a flux of downgoing
nuclearites
compared with the limits reported in Refs.
\cite{nakamura} (``Nakamura"), \cite{orito} (``Orito"),
the indirect mica limits \cite{price,ghosh} and with the DM limit,
assuming a velocity
at ground level of $\beta = 2 \times 10^{-3}$. At $\beta = 2 \times 10^{-3}$
the 90\% C.L. MACRO limit for an isotropic flux of nuclearites is
$2.7 \times 10^{-16}$ \units{cm^{-2}s^{-1}sr^{-1}}.
In Fig.~4 we extended the MACRO limit above the
DM bound, in order to show the transition to an isotropic
flux for nuclearite masses larger than $\simeq 6 \times 10^{22}$
\units{GeV/c^2}.
\par
\bigskip\par
\vskip 10 pt
\noindent{\bf Acknowledgements.}
\vskip 0.5 truecm
We gratefully acknowledge the support of the director and of the staff of the
Laboratori Nazionali del Gran Sasso and the invaluable assistance of the
technical staff of the Institutions participating in the experiment. We thank
the Istituto Nazionale di Fisica Nucleare (INFN), the U.S. Department of
Energy and the U.S. National Science Foundation for their generous support of
the MACRO experiment. We thank INFN, ICTP (Trieste) and World Laboratory
for providing fellowships and grants for non Italian citizens.\par
|
\section{Introduction}
\label{sec:intro}
There is a long history of studies of the magnetic response of an electron gas,
confined to a finite boundary. Starting with Bohr and van Leeuwen
\cite{vanvleck}, to Landau's finite diamagnetism in the quantum regime
\cite{landau,dingle}. The problem is still of current
theoretical an experimental interest \cite{rev}, in particular due to
the realization that for most geometries of the confining boundary one
can find classically chaotic behavior \cite{studies1}-\cite{levy}.
Most previous studies of this problem have assumed that the external
magnetic field is static and they have concentrated in calculating
the static magnetic susceptibility, except for the dynamic magnetic
field experimental work of Reulet et al. \cite{reulet}.
In an earlier paper \cite{paper1} (referred to as I hereafter),
we investigated the classical dynamics and the quantum
signatures of classical chaos, for {\it one electron} confined to a
circular quantum dot structure. The dot was subjected to uniform $d.c.$
($B_{dc}$) plus $a.c.$ ($B_{ac}f(t)$), with periodic
$f(t)=f(t+2\pi/\omega_0)$) perpendicular magnetic fields. There, we
established an approximate phase boundary in the parameter space spanned by
$(\epsilon= B_{ac}/B_{dc},\tilde{\omega_c}=\omega_c/\omega_0)$ that separates
the classically regular from the chaotic regimes, where
${\omega_c}$ is the Larmor frequency of the $d.c.$ field.
The phase diagram shown in Fig. \ref{fig1},
which we shall often use in our analysis here, separates the
quasi-integrable from chaotic regimes. In I we established clear
correspondences between the transitions in the classical behavior
and their corresponding quantum signatures. From the statistical
properties of the quasienergy spectrum of the one-period evolution operator,
going from Poisson to Gaussian orthogonal ensemble, to the semiclassical
phase space correspondences via the Husimi quasienergy eigenfunction
distribution functions.
In this paper we present a detailed quantum mechanical study of the zero
temperature magnetic response of a noninteracting electron
gas confined to a circular boundary and subject to the same combination
of a $d.c.$ plus $a.c.$ magnetic fields. Here we are interested
in considering the magnetic response of this model for an $N$ electron system
that satisfies the Pauli exclusion principle.
Another basic question, first addressed in this paper, is how does
the transition from regular to chaotic behavior in the classical case,
where the particles are indistinguishable, affect
their fermionic quantum nature. Most previous studies of the quantum
manifestations of classical chaos have centered on one-particle
problems. Here we only address the important particle-statistics
many-particle problem, and leave for a future study the relevant
effects of electronic interactions. As we show below there
are indeed clear manifestations of the particle statistics, which
are different if we are in the classically integrable regime from
those where the system is chaotic.
The organization of the rest of the paper is as follows: In section
II we briefly recapitulate the main elements of the single particle model
studied in I, together with expressions for the matrix elements of the
operators needed in our analysis. Next we outline our method to calculate
the matrix elements of multi-electron operators in a basis of properly
(anti)symmetrized eigenfunctions. In section III we present our main
results for the magnetization, orbital currents and energy. We calculated
both the time evolution of such operators and their time-averages as a
function $N$ and the parameters $(\epsilon,\tilde{\omega_c})$. We also
include a perturbative calculation, fully described in the Appendix,
that quantitatively explains our numerical results for the magnetization
in the quasi-integrable weak-field regime. Finally, in section IV
we present a summary of our conclusions, with an estimate of a few
experimental parameters that may give an idea of the regimes in frequency
and fields where the transition between integrable and chaotic regimes
discussed in this paper could be tested.
\section{The Model}
\label{sec:model}
\subsection{One-Electron Wavefunction}
We start by recalling the main features of the single particle formalism,
as explained in paper I, and next its extension to the N non-interacting
electron problem. The model we consider here is that of electrons
confined to a disk, and subject to a steady ($B_{dc}$)
and a time-periodic ($B_{ac}$) magnetic field.
After scaling to appropriate dimensionless units, the model
Hamiltonian considered here is,
\begin{equation}
\label{eq:5-1a}
\tilde H = \tilde H_{dc} + \tilde H_1(\tau),
\end{equation}
which in polar coordinates reads,
\begin{equation}
\label{eq:5-1b}
\tilde H_{dc} = -\frac{{\tilde\hbar}^2}{2}
\left( \frac{d^2}{dr^2} + \frac{1}{r}\frac{d}{dr}
\right) + \frac{\ell^2 {\tilde\hbar}^2}{2 r^2} + \frac{1}{2}
\left(\frac{{\tilde\omega_c}}{2}\right)^2 r^2 +
\ell\, {\tilde\hbar} \frac{{\tilde\omega_c}}{2} ,
\end{equation}
and with the time-dependent kick component
\begin{equation}
\label{eq:5-ic}
\tilde H_1(\tau) = \frac{1}{2}\ \eta\ r^2
\sum_{n=-\infty}^{\infty} \delta (\tau-n).
\end{equation}
The dimensionless units are defined as,
\begin{eqnarray}
\label{5-all2}
\label{eq:5-2a}
r = \frac{\rho}{R_0}, \quad 0\le &r& \le 1;\quad
\tau = \frac{t}{T_0} \equiv \frac{\omega_0}{2\pi}\,t, \\
\label{eq:5-2b}
{\tilde\omega_c} = \frac{\omega_c}{\omega_0}, \quad
{\tilde\hbar} = \frac{\hbar}{m^*\omega_0 R_0^2}, \quad \epsilon &=&
\frac{B_{ac}}{B_{dc}} = \frac{\omega_{ac}}{\omega_c},
\quad{\rm and}\quad\eta = \left(\frac{\epsilon\,
{\tilde\omega_c}}{2}\right)^2.
\end{eqnarray}
Here $R_0$ is the radius of the disk quantum dot assumed to have
rigid walls. $T_0$ is the drive period of the $a.c.$ field,
$\omega_c=e^*B_{dc}/(m^*c)$ is the static Larmor frequency, in
terms of the effective electron mass $m^*$ $ (\sim 0.067m_e)$,
the screened electronic charge $e^* $ $(\sim 0.3e)$ \cite{benaaker}, and the
dynamic frequency $\omega_{ac}=e^*B_{ac}/(m^*c)$.
The exact eigenfunctions of the static Hamiltonian
$\tilde H_{dc}$ are given in terms of the Whittaker $M$
functions \cite{dingle},
\begin{equation}
\label{eq:5-a}
\tilde\psi_{n\ell}(r) = \sqrt{\frac{2}{N_{n\ell}}}\;
\frac{1}{r}\;M_{\chi_{n\ell},{\mid \ell\mid}/2}(\frac{f}{2} r^2),
\end{equation}
with $n$ the principal quantum number, $\ell $ the angular
momentum eigenvalue, and $N_{n\ell}$ a normalization constant.
The frustration parameter, $f$, that measures the number of
flux quanta in the disk, is defined by,
\begin{equation}
\label{eq:5-b}
f = \frac{\Phi}{\Phi_0} \equiv
\frac{B_{dc}\pi R_0^2}{\left(hc/2e^*\right)} \equiv
\frac{\tilde{\omega_c}}{\tilde\hbar}=(\frac{R_0}{\ell _B})^2,
\end{equation}
with $\ell_B=(\frac{\hbar c}{eB_{dc}})^{1/2}$ the magnetic
length and $\Phi_0 = \frac{hc}{2e^*}$ the quantum of flux.
The eigenenergies
\begin{equation}
\label{eq:5-c}
\tilde E_{n\ell} = 2 ( \chi_{n\ell} + \ell ),
\end{equation}
are determined by the requirement that the wavefunction
vanishes at the boundary i.e., by the zeros of the Whittaker function,
$M_{\chi_{n\ell},{\mid \ell\mid}/2}(\frac{f}{2}) = 0$.
We calculated the energy eigenvalues $\tilde E_{n\ell}$ for the static
problem in a basis of Whittaker functions as a function of $B_{dc}$,
and checked our numbers by fully reproducing the results of
Ref.\cite{studies1}. The Whittaker functions have the advantage
of being valid over the entire
range of parameters, however, they are numerically difficult to evaluate
for the full time-dependent problem. For convenience
when calculating the time-dependent problem,
we decided also to expand the total (single particle) wavefunction
in a Fourier Sine-basis. In this case
\begin{eqnarray}
\label{5-all3}
\label{eq:5-3a}
\langle r |\tilde H_{dc}\, |\tilde\psi_{n\ell}(\phi)\rangle &=&
\tilde E_{n\ell}\,
\tilde\psi_{n\ell}(r) \frac{e^{i \ell \phi}}{\sqrt{2 \pi}},\\
\label{eq:5-3b}
\langle r|\tilde\Psi(\phi)\rangle &=& \sum_{n=1}^{\infty}
\sum_{\ell=-\infty}^{\infty}
\tilde\psi_{n\ell}(r) \frac{e^{i \ell \phi}}
{\sqrt{2 \pi}},\\
\label{eq:5-3c}
\tilde\psi_{n\ell}(r = 1) &=& 0, \qquad \qquad
\int_{0}^{1} \tilde\psi_{n\ell}^{2}(r) \, r \,dr = 1,\\
\label{eq:5-3d}
{\rm and} \,\,\,
\langle r|\tilde \psi_{n\ell}\rangle &=& \sqrt{\frac{2}{r}}
\sin(n\pi r).
\end{eqnarray}
This basis set is properly orthonormalized, and automatically
satisfies the boundary conditions. To calculate the spectrum
of the static problem, we used, nonetheless, the exact eigenvalues
of $\tilde H_{dc}$, given by the zeros of the Whittaker functions.
Doing this allowed us also to check the reliability of our Sine-basis
numerical method.
We then computed the required matrix elements of the operators
we are interested in, within the Sine-basis method.
For example, for the magnetization operator
\begin{equation}
\label{eq:5-4}
\mbox{\boldmath{$\mu$}} = \frac{e^*}{2m^*c}
\left( \mbox{\boldmath{$\cal L$}}
- \frac{e^*}{c}~\mbox{\boldmath{r}}\times
{\rm{\bf A}}(\mbox{\boldmath{r}})\right),
\end{equation}
where \mbox{\boldmath{$\cal L$}} is the angular momentum operator and
${\rm{\bf A}}(\mbox{\boldmath{r}})$ is the electromagnetic
vector potential in normalized coordinates. In the present case, we take the
magnetic field perpendicular to the plane, then the z-component
of the magnetization operator is
\begin{equation}
\label{eq:5-5}
\tilde M_z(r) = \frac{\hat M_z}{\mu_B} = -\frac{\hat L_z}{\tilde\hbar} -
\frac{f}{2} \hat r^2,
\end{equation}
with $L_z$ the z-component of the angular momentum,
$\mu_B = \frac{|e^*|\hbar}{2m^*c}$ the Bohr magneton, and $\hat M_z$
the magnetization operator along the z-axis. The matrix elements of
$\tilde M_z$ in the Fourier Sine-basis are given by,
\begin{eqnarray}
\label{alleq5-6}
\langle m|\tilde M_z|n\rangle = &-&\Big\{\ell
+ \frac{f}{2}\left(\frac{1}{3}-\frac{1}
{2n^2\pi^2}\right) \Big\}~\delta_{mn} \nonumber \\
&-&\Big\{ \frac{f}{2} ~\frac{(-)^{m+n}}{\pi^2}\frac{8mn}
{(m^2-n^2)^2} \Big\}~(1-\delta_{mn}).
\end{eqnarray}
Similarly, starting from the definition of the current density operator
\begin{equation}
\label{eq:5-7}
{\rm {\bf J}} = \frac{1}{2m^*}\left(-i\hbar
\mbox{\boldmath{$\nabla$}} -
\frac{e^*}{c}{\rm{\bf A}}\right) + c.c.,
\end{equation}
(where $c.c.$ stands for complex conjugate),
we have the following expression for the azimuthal current
densities: $J_{\phi} = J_{\phi}^{(para)} + J_{\phi}^{(dia)}$, where
\begin{equation}
\label{eq:5-8}
J_{\phi}^{(para)} = -\frac{i\tilde \hbar}{2}
\frac{1}{r}\frac{\partial}{\partial\phi}
+ c.c. \quad {\rm and}, \quad
J_{\phi}^{(dia)} = \frac{\tilde \omega_c}{2} r,
\end{equation}
are the paramagnetic and diamagnetic current densities, respectively.
In the Fourier Sine-basis, the matrix elements of the
current densities
(in units of $\tilde \hbar$) are given by
\begin{eqnarray}
\label{alleq5-9a}
\label{eq:5-9}
\langle m|J_{\phi}^{(para)}|n\rangle &=&
\ell{\tilde \hbar}\cdot
\left\{\begin{array}{ll}
- {\rm Ci}[2n\pi] + \gamma_E + {\rm ln}(2n\pi) & \mbox{$(m=n)$} \\
- {\rm Ci}\left[(m+n)\pi\right] + {\rm Ci}\left[(m-n)\pi\right] &
\mbox{$(m\ne n),$}\end{array}
\right. \\
\label{eq:5-9b}
\langle m|J_{\phi}^{(dia)}|n\rangle &=&
\frac{f}{2}{\tilde \hbar}\cdot
\left\{\begin{array}{ll}
\frac{1}{2} & \mbox{$(m=n)$} \\
\frac{(-)^{m+n}-1}{\pi^2} \frac{4mn}{(m^2-n^2)^2} &
\mbox{$(m\ne n)$}\end{array}
\right.
\end{eqnarray}
where $\gamma_E = 0.57721~566649\dots$ is Euler's gamma number,
and $\rm{Ci}(x)$ is the Cosine integral.
Finally, the expression for the one-period time-evolution
operator $U_{\ell}(\tau,\tau_0)$,
for the single particle Hamiltonian, that satisfies the
dynamical equation
\begin{equation}
\label{eq:5-10}
i {\tilde\hbar}\,\frac{\partial}{\partial\tau}\,U_{\ell}
(\tau,\tau_0) = ( {\tilde H_{dc}} + {\tilde H_1(\tau)} )\,
U_{\ell}(\tau,\tau_0),
\end{equation}
is
\begin{equation}
\label{eq:5-11}
U_\ell(1,0) = \exp\left(-\frac{i}{{\tilde\hbar}}\,\frac{1}{2}
\eta\,r^2\right)\,\exp\left(-\frac{i}{{\tilde\hbar}}\,
{\tilde H_{dc}}\right) .
\end{equation}
The total (single particle) wavefunction at any integer multiple
$N_T$ of the period (hereafter taken to be 1), is given by
repeated applications of $U_{\ell}$ to the initial wavefunction:
\begin{equation}
\label{eq:5-12}
|\Psi_{\ell}(r,\phi,N_T)\rangle = U_{\ell}^{N_T}\,
|\Psi_{\ell}(r,\phi,0)\rangle .
\end{equation}
\subsection{Many-Electron Wavefunctions}
One can directly generalize the above single-electron
formalism to the many-electron case. Take the initial $N$-electron
wavefunction to be
\begin{equation}
\label{eq:5-13}
|\Phi(\mbox{\boldmath{$r$}}_1,
\mbox{\boldmath{$r$}}_2\dots ,
\mbox{\boldmath{$r$}}_N)\rangle \equiv
|\Phi(1,2\dots ,N)\rangle,
\end{equation}
which is antisymmetric under exchange of an odd number of
particles (the Pauli exclusion principle) :
\begin{equation}
\label{eq:5-14}
|\Phi(1,\dots i,\dots j,\dots ,N)\rangle =
-|\Phi(1,\dots j,\dots i,\dots ,N)\rangle.
\end{equation}
Let the $i$-th single-particle eigenstate satisfy
the equation
\begin{equation}
\label{eq:5-15}
\tilde H^{(i)}|\Psi_{n_{i}\ell_{i}}(i)\rangle =
E_{n_{i}\ell_{i}}^{(i)}|\Psi_{n_{i}\ell_{i}}(i)\rangle.
\end{equation}
We know that the Slater antisymmetrization
procedure for
the non-interacting $N$-electron state can be written as
the following tensor product \cite{huang}:
\begin{eqnarray}
\label{alleq5-16}
{\rm{\bf\hat A}}&\cdot&|\Phi(1,2\dots ,N)\rangle
= \frac{1}{\sqrt{N!}}\left|\begin{array}{llll}
|\Psi_{n_1\ell_1}(1)\rangle &
\otimes|\Psi_{n_2\ell_2}(1)\rangle &
\cdots &
\otimes|\Psi_{n_N\ell_N}(1)\rangle \\
|\Psi_{n_1\ell_1}(2)\rangle &
\otimes|\Psi_{n_2\ell_2}(2)\rangle &
\cdots &
\otimes|\Psi_{n_N\ell_N}(2)\rangle \\
\vdots &
\vdots &
\ddots &
\vdots \\
|\Psi_{n_1\ell_1}(N)\rangle &
\otimes|\Psi_{n_2\ell_2}(N)\rangle &
\cdots &
\otimes|\Psi_{n_N\ell_N}(N)\rangle
\end{array}\right| \nonumber \\
&=& \frac{1}{\sqrt{N!}}\sum_{P} \delta_{P}~\Bigg[
|\Psi_{P\{n_1\ell_1\}}(1)\rangle\otimes
|\Psi_{P\{n_2\ell_2\}}(2)\rangle\cdots
\otimes|\Psi_{P\{n_N\ell_N\}}(N)\rangle\Bigg]
\end{eqnarray}
where ${\rm {\bf \hat A}}$ is the antisymmetrization operator, and
$P$ is the permutation operator
\begin{eqnarray}
\label{5-all17}
\label{eq:5-17a}
P\{1,2,\dots ,N\} &=& \{P1,P2,\dots ,PN\},\\
\label{eq:5-17b}
\delta_{P} &=& \left\{\begin{array}{ll}
+1 & \mbox{(even $P$)} \\
-1 & \mbox{(odd $P$)}
\end{array} \right. .
\end{eqnarray}
The summation runs over all possible permutations. Furthermore,
the trace of any sum of $N$-body operators
${\hat O} = \sum_{i=1}^N {\hat O_i}$
can be written as
\begin{eqnarray}
\label{alleq5-18}
{\rm Tr}\{{\hat O}\} &=& \langle\Phi(1,\dots ,N)|
{\hat O}|\Phi(1,\dots ,N)\rangle \nonumber \\
&=& \frac{1}{N}\sum_{i,j=1}^{N}
\langle\Psi_{n_{i}\ell_{i}}(j)|{\hat O_{j}}
|\Psi_{n_{i}\ell_{i}}(j)\rangle.
\end{eqnarray}
Now consider the time evolution of such a system.
For simplicity, we take as the initial state the lowest energy
(ground) state of the {\it unperturbed} system
({\sl i.e.,} without the $a.c.$ field) allowed
by the Pauli Principle :
\begin{equation}
\label{eq:5-19}
|\Phi_{1,\dots, N}(t=0)\rangle = {\rm{\bf\hat{A}}}\cdot\left\{
|\Psi_{n_{1}\ell_{1}}(1)\rangle\otimes
|\Psi_{n_{2}\ell_{2}}(2)\rangle
\cdots\otimes |\Psi_{n_{N}\ell_{N}}(N)\rangle\right\}.
\end{equation}
The one-period time-evolution operator for the N-electron system
is given by the tensor product,
\begin{equation}
\label{eq:5-20}
\mbox{\boldmath{$U$}} = U_1(1,0)\otimes U_2(1,0)\cdots
\otimes U_N(1,0).
\end{equation}
Since the antisymmetrization and time-evolution operators commute,
the state after $N_T$ periods is simply given by
\begin{eqnarray}
\label{eq:5-21}
|\Phi_{1,\dots, N}(N_T)\rangle &=&
\mbox{\boldmath{$U$}}^{N_T}
\cdot|\Phi_{1,\dots, N}(t=0)\rangle \nonumber \\
&=& {\rm{\bf\hat{A}}}\cdot\left\{U_1^{N_T}
|\Psi_{n_1\ell_1}(1)\rangle \cdots\otimes
U_N^{N_T}|\Psi_{n_N\ell_N}(N)\rangle\right\},
\end{eqnarray}
where we've used the notation $U_i(1,0)\equiv U_i,$ for $i=1,\ldots,N$.
We can now generalize Eq. (\ref{alleq5-18}) for the trace
of an operator at any integer multiple $N_T$ of the period as,
\begin{eqnarray}
\label{eq:5-22}
{\rm Tr}\{{\hat O}\}(t=N_T)
&=& \langle\Phi_{1,\dots, N}(t=N_T)|{\hat O}
|\Phi_{1,\dots, N}(t=N_T)\rangle \nonumber \\
&=& \frac{1}{N}\sum_{i,j=1}^{N}\langle\Psi_{n_{i}\ell_{i}}(j,t=0)|
(U_j^{\dagger})^{N_T}
{\hat O_{j}}U_j^{N_T}|\Psi_{n_{i}\ell_{i}}(j,t=0)\rangle.
\end{eqnarray}
In particular, for the average quantum time-dependent magnetization
{\it per electron} we have
\begin{equation}
\label{eq:5-23}
\frac{\langle{\tilde M_z}\rangle(N_T)}{N} =
-\frac{L}{N} + \frac{1}{N^2}
\sum_{i,j=1}^N \langle\Psi_{n_{i}\ell_{i}}(j)|
(U_j^{\dagger})^{N_T}
(-\frac{f}{2} {r}_j^2)U_j^{N_T}|\Psi_{n_{i}\ell_{i}}(j)\rangle,
\end{equation}
where $L=\sum_{i=1}^{N} \ell_i$. Similarly, we can write the corresponding
expressions for the time-dependent orbital currents, and the
total time-dependent Hamiltonian average, which we term the averaged energy.
For example, the {\it time-averaged magnetization},
$\langle\langle M_z \rangle\rangle$, is defined by
\begin{equation}
\label{eq:5-23b}
\langle\langle M_z \rangle\rangle = \lim_{N_T\rightarrow\infty}
\frac{1}{N_T} \sum_{n=1}^{N_T} \langle M_z \rangle(n).
\end{equation}
\section{Results}
\label{sec:res}
We now come to the discussion of the main results of this paper,
that are concerned with the dynamic and time--averaged properties of
different relevant operators. The most striking
features are observed in the magnetization of the system, which we
shall discuss first as a function of the number $N$ of electrons
and ($\epsilon,\tilde\omega_c$).
We also give results for the orbital current as well as
interesting results for the time-averaged energy as a function
of $N$ in different ($\epsilon,\tilde\omega_c$) parameter regimes.
\subsection{Magnetization, orbital currents and energy}
We start with the time--dependent dynamics of the magnetization
and its corresponding power spectra for a single electron.
The power spectrum $S(\nu)$ is the
square of the Fourier transform of the expectation value of the
magnetization operator $\langle\tilde M_z\rangle(t)$.
(For notational simplicity, we write $\langle\tilde M_z\rangle(t)$ by
$\langle M\rangle(t)$.) We keep $\tilde\omega_c$
fixed, and sweep through values of $\epsilon$, from small to large.
As can be seen from the phase diagram in Fig. \ref{fig1},
for fixed $\tilde \omega_c$, as $\epsilon$ increases the
underlying classical dynamics changes from quasi-integrable
to chaotic.
In Fig. \ref{fig2}(a) we see that $\langle M \rangle$ is diamagnetic
in the regular regime and oscillates periodically with time, reflected
in the very strong peak in its power spectrum shown in Fig. \ref{fig2}(b).
As we increase the values of $\epsilon$ ({\sl i.e.}, as we approach
the chaos border in the phase diagram), the intermediate dynamics gets
more complex, as shown in Figs. \ref{fig3}(a) and \ref{fig3}(b).
We see in Fig. \ref{fig3}(b), that there are two peaks in $S(\nu)$,
which are due to the quasi-beats seen in Fig. \ref{fig3}(a). In the chaotic
region, $<M>(t)$ shows essentially irregular behavior, Fig. \ref{fig4}(a),
while $S(\nu)$ has a broad background, shown in Fig. \ref{fig4}(b).
Note that the average value of $<M>(t)$ increases in magnitude,
becoming steadily more diamagnetic in the chaotic regime.
The one-electron behavior changes significantly with the addition of
more electrons. The pattern of change from regular to chaotic
is similar as in the one-electron case as we sweep through
the same $\epsilon$-values as above. The spectral function
develops more resonances in the regular region,
whereas in the chaotic regime it has a broad band spectrum. As we
continue to increase the number of electrons there are more ``beats"
in the time--dependence of the magnetization and more peaks in the
spectral function.
It is then more convenient to consider the time-averaged properties of
the magnetization, the orbital current, or the total energy, as a function
of the number of electrons. It is in this type of function that we can
see important qualitative differences that represent the changes from
the classical regular to chaotic behavior in the quantum dynamics.
In Fig. \ref{fig5}(a) we see that for two-electrons the time-averaged
magnetization in the regular regime is {\it paramagnetic}, whereas for three
electrons becomes diamagnetic again. We note that the dia- to para-magnetic
changes are non-monotonic as a function of $N$. For example,
for $N=4$ it switches back to paramagnetic, but remains diamagnetic for
both $N=5$ and $N=6$. A similar situation occurs with the orbital
current as shown in Fig. \ref{fig5}(b).
We mention that the specific value of the frustration parameter ($f =
\frac{{\tilde\omega_c}}{\tilde\hbar}$) determines
if the magnetization flips from dia- to para-magnetic as we
keep adding electrons. Basically, this phenomenon occurs when
$f\sim O(1)$, i.e. when we add one flux quantum to the dot.
In all other cases, the magnetization remains diamagnetic
and monotonically increasing in magnitude as the number
of electrons increases. We provide a theoretical perturbation
theory explanation of these dia- to para-magnetic transitions result
in the next subsection.
As we increase the value of $\epsilon$, we enter the chaotic regime.
There we find that for {\it all} electron numbers
the magnetization is {\it always diamagnetic}, at least up to the
maximum number of electrons we considered ($\sim 25)$.
In Fig. \ref{fig5}(c) we show the time-averaged magnetization in
the chaotic regime, which also oscillates as a function of $N$, but it
is always negative and of larger magnitude than in
the quasi-regular regime. A similar situation
occurs for the orbital current (as shown in Fig. \ref{fig5}(d),
although it has less sharp changes as a function of $N$
than does $<<M>>$.)
In Fig. \ref{fig6}(a) we consider the time-averaged magnetization for a
fixed value of $N=1$, $\epsilon =0.1$ and $\tilde \hbar =0.1$
as a function of $\omega_c$. In this quasi-integrable regime we see
that $<<M>>$ is diamagnetic and decays quadratically as a function of
$\omega_c$. The situation changes in the chaotic regime,
shown in Fig. \ref{fig6}(c), where there is also decay with $\omega_c$
but now the behavior is not as smooth as in the quasi-integrable regime.
We show in Fig. \ref{fig6}(b) the behavior of the time averaged
energy as a function of the number $N$ of electrons. Here we see a clear
quadratic growth as a function of $N$.
The situation is remarkably different when the single-particle
classical dynamics is chaotic. In this case, shown in Fig. \ref{fig6}(d),
the time-averaged energy grows clearly {\it linearly} with $N$.
This implies that the classically chaotic solutions do have a significant
quantum signature in the averaged energy, that changes the quadratic
quasi-integrable regime behavior to a linear $N$ dependence in the
chaotic regime. We now present a simple heuristic argument
as to why the change over between quadratic and linear
$N$ behavior is actually directly related to the Pauli exclusion principle.
We note that in the zero magnetic field case, each of
the $N$ electrons in the circular dot of radius $R_0$
occupies an exclusion principle space of order $R_0/N^{1/2}$,
while the static free particle kinetic energy changes like $N/R_0^2$.
We expect that this situation does not change much when we are
in the quasi-integrable regime, for finite fields and low frequencies.
In the classically chaotic regime, in the presence of stronger magnetic
fields or higher frequencies, the magnetic field will tend, on the average,
to localize more the electrons to Larmor orbits inside the dot and in the
boundaries. When the field is larger, so that the Larmor radius and $R_0$
are comparable, the Landau levels have to be taken into account. In this
case the electrons will not necessarily feel the presence of the boundary
and they will remain localized in their ``chaotic'' Landau orbits due to the
time-dependent kicks. In this limit the contribution from
the kinetic energy is much less relevant, and the Larmor orbit radius
will be less dependent of $N$ and $R_0$.
\subsection{Perturbative evaluation of the magnetization in the
quasiregular regime}
In this subsection we present a perturbative analysis that provides
an explanation for the magnetization oscillations as a function
of $N$ in the quasi-integrable regime. Let us first consider the
time-independent part of the one-electron Hamiltonian ${\tilde H}$ ,
and write it as,
\begin{equation}
\label{eq:5-24}
\tilde H_{dc} = \tilde H_{0} + \tilde V,
\end{equation}
where,
\begin{equation}
\label{eq:5-25}
\tilde H_{0} = -\frac{{\tilde\hbar}^2}{2}
\left( \frac{d^2}{dr^2} + \frac{1}{r}\frac{d}{dr}
\right) + \frac{\ell^2 {\tilde\hbar}^2}{2 r^2} +
\ell\, {\tilde\hbar} \frac{{\tilde\omega_c}}{2} ,
\end{equation}
and,
\begin{equation}
\label{eq:5-26}
{\tilde V} = \frac{1}{2}\left(\frac{{\tilde\omega_c}}{2}\right)^2 r^2.
\end{equation}
We will consider the limiting case of very small $B_{dc}$ field, {\sl i.e.},
${\tilde\omega_c}\ll 1$. As was first shown by Dingle,
one can write the eigenvalues to first order in ${\tilde\omega_c}$, by
considering the zero field basis functions of the disk Bessel eigenfunctions:
\begin{equation}
\label{eq:5-27}
{\tilde H_0} |\psi_{n\ell}^{(1)}\rangle \simeq E_{n\ell}^{(1)}
|\psi_{n\ell}^{(1)}\rangle ,
\end{equation}
where the normalized eigenvalues are \cite{dingle},
\begin{equation}
\label{eq:5-28}
E_{n\ell}^{(1)} = \frac{\alpha_{n\ell}^2}{2f} + \ell + \frac{f}{12}
\left\{1 + \frac{2(\ell^2-1)}{\alpha_{n\ell}^2} \right\} +
O({\tilde\omega_c}^2).
\end{equation}
Here $\alpha_{n\ell}$ is the $n$th zero of the Bessel function
$J_{\ell}(x)$,
and the {\it unperturbed} basis functions are given by
(we consider only the radial part, the angular part is clear)
\begin{equation}
\label{eq:5-29}
\langle r|\psi_{n\ell}^{(0)}\rangle =
\frac{\sqrt{2}}{J_{\ell+1}(\alpha_{n\ell})}
\,J_{\ell}\left(\alpha_{n\ell}r\right) .
\end{equation}
The matrix elements of the perturbation are then given by,
\begin{equation}
\label{eq:5-30}
\langle\psi_{m\ell}^{(0)}|{\tilde V}|\psi_{n\ell}^{(0)}\rangle \equiv
{\tilde V_{mn}} = \frac{1}{2}\left(\frac{{\tilde\omega_c}}{2}\right)^2
\left\{\begin{array}{cc} \frac{1}{3}\left[1+\frac{2(\ell^2-1)}
{\alpha_{n\ell}^2}\right] & m=n \\
\strut{
\frac{8\alpha_{m\ell}\alpha_{n\ell}}{(\alpha_{m\ell}^2
-\alpha_{n\ell}^2)^2} } & m\ne n \end{array} \right.
\end{equation}
The perturbed non degenerate eigen-functions are obtained from
standard perturbation analysis,
\begin{eqnarray}
\label{all5-31}
|\psi_{n\ell}^{(1)}\rangle &=& |\psi_{n\ell}^{(0)}\rangle + \sum_{m\ne n}
\frac{{\tilde V_{mn}}}{E_{m\ell}^{(0)}-E_{n\ell}^{(0)}}
\,|\psi_{m\ell}^{(0)}\rangle
+ O\left({\tilde V_{mn}}^2\right) \nonumber \\
&=& |\psi_{n\ell}^{(0)}\rangle + 2f{\tilde\omega_c}^2
\sum_{m\ne n} \frac{\alpha_{m\ell}\alpha_{n\ell}}
{(\alpha_{m\ell}^2-\alpha_{n\ell}^2)^3}|\psi_{m\ell}^{(0)}\rangle
+ O\left(\alpha^{-8}\right),
\end{eqnarray}
where we've used the unperturbed energy levels
$E_{m\ell}^{(0)}-E_{n\ell}^{(0)} =
\frac{1}{4}(\alpha_{m\ell}^2-\alpha_{n\ell}^2)$.
Using Eq.(\ref{all5-31}), and the definition of the
magnetization operator, Eq.(\ref{eq:5-5}), the
leading first order matrix element contribution to ${\tilde M_z}$ is given by
\begin{eqnarray}
\label{all5-32}
\langle\psi_{n\ell}^{(1)}|{\tilde M_z}|\psi_{n\ell}^{(1)}\rangle \equiv
{\langle{\tilde M_z}\rangle}_{mn} &=& -\ell - 2f
\langle\psi_{n\ell}^{(0)}|r^2|\psi_{n\ell}^{(0)}\rangle +
O({\tilde\omega_c}^2) \\
&=& -\ell - \frac{2f}{3}\left[1+\frac{2(\ell^2-1)}
{\alpha_{n\ell}^2}\right] + O({\tilde\omega_c}^2) .
\end{eqnarray}
Once again, if we take as the initial state the lowest
energy state allowed by the Pauli Principle,
\begin{equation}
\label{eq:5-33}
|\Psi_{1,\dots ,N}(0)\rangle = {\rm{\bf\hat{A}}}\cdot\left\{
|\psi_{n_{1}\ell_{1}}(1)\rangle\otimes
|\psi_{n_{2}\ell_{2}}(2)\rangle
\cdots\otimes |\psi_{n_{N}\ell_{N}}(N)\rangle\right\},
\end{equation}
we can generalize Eq.(\ref{all5-32}) to the $N$-electron case. We
find that the averaged magnetization {\it per electron}, to
first order approximation, is
\begin{equation}
\label{eq:5-34}
\frac{\langle{\tilde M_z}\rangle}{N} = -\frac{2f}{3} - \frac{1}{N}
\sum_{i=1}^N \left\{\ell_i + \frac{4f}{3}\,\frac{(\ell_i^2-1)}
{\alpha_{n_i\ell_i}^2} \right\} + O({\tilde\omega_c}^2) .
\end{equation}
Next, we perform a linear-response theory analysis of the {\it full}
time-dependent problem, assuming that $\epsilon\ll 1$.
We show in the Appendix that within the perturbative approximation
for a single electron, the average magnetization at time
$N_T$ is given by,
\begin{eqnarray}
\label{eq:eq5-35}
\langle{\tilde M_z}\rangle(N_T) &\simeq& -\left\{\ell_i +
\frac{f}{2}\langle r^2\rangle_{n_i,n_i}\right\} \nonumber \\
&+& \frac{f}{{2\tilde\hbar}}\left(
\frac{\epsilon{\tilde\omega_c}}{2}\right)^2\sum_{p=1}^{N_T}\sum_{n\ne n_i}
\langle r^2\rangle_{n,n_i}^2 \sin\left\{\omega_{n_i,n}(N_T - p)\right\} ,
\end{eqnarray}
where
\begin{equation}
\label{eq:5-36}
\langle r^2\rangle_{n,n_i} = \frac{1}{3}\left[1+\frac{2(\ell_i^2-1)}
{\alpha_{n_i\ell_i}^2}\right] + O({\tilde\omega_c}^2) .
\end{equation}
Clearly, the last term in Eq.(\ref{eq:eq5-35}) is of
$O({\tilde\omega_c}^3)$, so it can be ignored within the current
approximation, which means that to lowest order, the time-dependence
plays no significant role in determining the average magnetization.
To test the approximation for $\langle{\tilde M_z}\rangle/N$,
Eq.(\ref{eq:5-34}), we show in Fig. \ref{fig7}, a comparison of
results from the perturbative and numerically exact calculations.
The perturbative results agree remarkably well with the
numerical calculations. We can now understand why the
averaged magnetization oscillates in sign in the regular regime, and it is
because $\langle{\tilde M_z}\rangle/N$ depends most strongly
on $\sum_{i=1}^N \ell_i$. Whenever this sum of angular momentum
quantum numbers flips sign, so does the magnetization. For example, for
the parameters shown in Fig. \ref{fig7}, the values of $F(N) = \sum_{i=1}^N \ell_i$ for
successive values of $N$ are
\begin{equation}
\label{eq:5-37}
F(N=1,2,\ldots,12) = 0,-1,0,-2,0,0,-3,0,-1,0,-4,0,
\end{equation}
corresponding to
\begin{equation}
\ell_i = (0,-1,1,-2,2,0,-3,3,-1,1,-4,4),
\end{equation}
i.e., the $\ell$-values of the twelve-electron (unperturbed) ground state.
This is a selection rule associated with the symmetries present in
the system.
What is interesting is that in the chaotic region, there is no
such flipping of the sign. This difference may constitute an
experimentally accessible signature of chaos in the
quantum system. Clearly, such behavior is
exclusively a consequence of the Pauli principle, for we would
not observe any change in the response per electron without it,
since ignoring it in the multi-electron case
would lead to a trivial rescaling of the single electron results.
\section{Conclusions}
\label{sec:conc}
To summarize, we have studied a model of a non-interacting $N$-electron
system, confined to a circular structure with rigid boundaries,
and subjected to perpendicular constant and time-periodic magnetic fields.
We studied the magnetization and orbital currents as a function of
time, as well as the time-averaged magnetization
$\langle\langle {\tilde M_z}\rangle\rangle$, and energy as a function
of electron number $N$. We can make a strong connection between the
dynamic response of $\langle{\tilde M_z}\rangle(t)$ (or it's power
spectrum), and the underlying classical dynamics -- as the classical
system makes a transition to chaos as we vary the applied magnetic fields,
the dynamics changes from being harmonic to essentially noisy. There are
three central significant conclusions: first, the Pauli Principle affects
the behavior of this non-interacting system significantly, e.g. in
terms of oscillations of $\langle\langle {\tilde M_z}\rangle\rangle$ as
a function of $N$. This behavior is directly related to the Pauli
principle that allows the electrons to optimally reduce their averaged
$\langle\langle M\rangle\rangle$ at specific values of the total angular
momentum. Second, while these oscillations in the quasi-integrable
regime cause the system to flip back and forth between dia- and
para-magnetic behavior, the system remains diamagnetic at all times
in the chaotic regime. We also found a very interesting change in the
time-averaged energy as a function of $N$, going from quadratic in
the quasi-integrable regime to linear in the chaotic one. We provided a
simple heuristic explanation of this behavior related to Pauli's
exclusion principle.
In this paper we have not considered the effects of Coulomb interactions that
can significantly complicate the analyses. There are static studies that
have considered the changes in the classical dynamics due to
interactions. What has been found in some examples is that
if the system of non-interacting particles was non-chaotic, as the
interaction parameter increases the dynamics can become chaotic
\cite{inter1}. In the quantum
regime the Random Matrix Theory
statistics can exhibit a transition from
Poisson to orthogonal ensemble
as the interaction strength increases \cite{inter2}. What happens in the
time-dependent case considered in this paper that deals with quasi-energy
statistics is not known at present.
Here we have considered a circular disk in the presence of a
time dependent magnetic field. It is only when we have the ac component
of the field added to the dc one
that chaos appears. In contrast, if the field is static
but the geometry is changed one can have chaotic classical solutions.
The relevance of the Pauli principle as seen in the zero temperature
magnetization has been studied, for example, by \cite{studies1}. At
present we do
not know what happens when
the geometry is not circular and we have a gas of Pauli electrons in the
presence of a dc+ac magnetic field. We expect to consider the two problems
mentioned above in the future.
To conclude, we briefly give some estimates in terms of physical
units of the field strengths and frequencies required to observe
the effects predicted by our model calculations. In a GaAs-AlGaAs
semiconductor the radius $R_0$ of a quantum dot device
\cite{marcus1},\cite{levy} can be between 0.1 and 10$\mu$m, a sheet
electron density $n\sim 10^{11}$ cm$^{-2}$, a mobility
$\mu\sim 265\,000$ cm$^2$/V$\cdot$s, and a characteristic
level spacing $\Delta\epsilon \sim 0.05$ meV or $\sim 0.5$ K.
In the ballistic electronic motion regime the elastic mean free path
$\l_\phi\sim 10\mu$m, with phase coherence length varying
between 15 and 50 $\mu$m. Typically the power injected is
smaller than $1$ nW, which is necessary to avoid electron heating.
For a dot radius of $R_0\sim 1\mu$m, the kick frequency $\omega_0$
can be obtained from Eqs.(\ref{eq:5-2b}) as
$
\omega_0 = \frac{\hbar}{m^*R_0^2}\,\frac{1}{{\tilde\hbar}}
\simeq \frac{2}{{\tilde\hbar}} \,{\rm GHz}.$
Then the required $B_{dc}$ and $B_{ac}$ magnetic fields have the values:
$B_{dc} = \frac{\omega_0 m^* c}{e^*}\,{{\tilde\omega_c}}
\simeq 20 \frac{{\tilde\omega_c}}{{\tilde\hbar}}$ {\rm gauss}, and
$B_{ac} = \epsilon B_{dc} \simeq
20 \frac{\epsilon{\tilde\omega_c}}{{\tilde\hbar}}
\,{\rm gauss}.$ The $a.c.$ Larmor frequency is
$\omega_{ac} = \epsilon {\tilde\omega_c} \simeq
20 \,\frac{\epsilon{\tilde\omega_c}}{{\tilde\hbar}}\,
{\rm MHz}.$
With these values, in the quasi-integrable regime,
with parameters $(\epsilon,{\tilde\omega_c})^{(reg)} = (0.1,0.1)$,
we get
$
\omega_0^{(reg)} \simeq 20 \,{\rm GHz}$ and
$B_{ac}^{(reg)} \simeq 20 \,{\rm gauss}.
$
In the chaotic regime we take the parameters
$(\epsilon,{\tilde\omega_c})^{(chaos)} = (2.0,2.0)$.
which leads to
$\omega_0^{(chaos)} \simeq 20 \,{\rm GHz}$
and $B_{ac}^{(chaos)} \simeq 800 \,{\rm gauss}$.
These results for the regular and chaotic regimes are within
experimental reach.
\section*{Acknowledgments}
This work has been supported in part by CONACYT 3047P and by NSF grant
DMR-9521845.
\newpage
|
\section{Single Atom in a Cavity}
\label{sec1}
In the experiments of \cite{mabuchi1996a,hood1998a,mabuchi1998b} very
cold Cesium atoms are dropped into tiny single-mode Fabry-Perot
cavities. The master equation for the system is well known. The
Hamiltonian for a two level atom interacting with a single driven mode
of the electromagnetic field in an optical cavity using the electric
dipole and rotating wave approximations (in the interaction picture
with respect to the driving laser frequency) is
\begin{eqnarray}
H &=&\frac{{\bf p}^{2}}{2m}+V({\bf r})+\hbar (\omega _{0}-\omega _{L})\sigma
_{+}\sigma _{-}+\hbar (\omega _{c}-\omega _{L})a^{\dagger }a+ \nonumber \\
&&\hbar g_{0}\psi ({\bf r})(a^{\dagger }\sigma _{-}+\sigma _{+}a)+\hbar
E(a^{\dagger }+a). \label{hamiltonian}
\end{eqnarray}
The first term is the kinetic energy of the atom, the second describes
an external potential which in subsequent sections we will allow to
confine the atom in the absence of a cavity field. The next two terms
are the energy in the internal state of the atom and the cavity
excitation. The fifth term describes the position dependent
interaction of the cavity mode and the atomic dipole. The strength of
this interaction is determined by the single photon Rabi frequency
$g_{0}$ which depends on the cavity mode volume and the dipole matrix
elements for the relevant atomic transition. The final term describes
the driving of the cavity by a coherent (laser) driving field of
amplitude $E$, chosen here to be real. The atomic transition
frequency is $\omega _{0}$, the cavity has a resonance at the
frequency $\omega _{c}$ and the driving frequency is $\omega _{L}$.
The cavity mode function is $\psi ({\bf r})=\cos (k_{L}x)\exp (-(y^{
2}+z^{2})/w_{0}^{2})$, describing the Gaussian standing wave
structure of the field in the Fabry-Perot cavity, the optical
wavelength $\lambda _{L}=852.359$~nm for the Cesium transition
employed ($k_{L}=2\pi /\lambda _{L}$).
Dissipation in the system is due to cavity losses and spontaneous
emission. By treating the modes external to the cavity as heat
reservoirs at zero temperature it is possible to derive the standard
master equation for the density operator of the system
\cite{carmichael1993a}, $\rho $,
\begin{equation}
\dot{\rho}=\frac{1}{i\hbar }[H,\rho ]+2\kappa
{\cal D}[a]\rho +2\gamma {\cal D}[\tilde{\sigma}_{-}]\rho .
\label{mastereqn}
\end{equation}
The superoperator ${\cal D}[c]$ acting on a density matrix $\rho $ is ${\cal %
D}[c]\rho =c\rho c^{\dagger }-\frac{1}{2}c^{\dagger }c\rho
-\frac{1}{2}\rho c^{\dagger }c.$ The dipole decay rate is $\gamma $,
while the cavity field decay constant is $\kappa $. The third term
describes the effect of spontaneous emission and $\tilde{\sigma}_{-}$
is an operator describing both the change of internal state and the
momentum kick on the atom due to a single spontaneous emission.
These experiments represent an important improvement on previous work
in that each atom remains in the cavity for many Rabi cycles. The
cooling of the atoms prior to entering the cavity effects a separation
of timescales of the dynamics of the external degrees of freedom of
the atom and the other degrees of freedom in the problem. The
variation of the coupling due to the atomic motion, frequency with
which the atom passes through wavelengths of the standing wave in the
cavity, is, at least initially, much less than the other frequencies
involved,
\begin{equation}
\left|{\bf p} \cdot \nabla \psi({\bf r})\right| \ll g_{0},\kappa ,\gamma .
\end{equation}
However the driving of the cavity field and the interaction of this
field with the atom leads to disturbance of the motion which can be
understood in terms of a semi-classical theory of the mechanical
effects of light in the cavity. In particular the effect of the dipole
force in trapping atoms near the antinodes of the cavity field was
observed in \cite{hood1998a}. However the rapid exchange of excitation
between the atom and the cavity mode leads to an increased momentum
diffusion or heating in this system over a free-space standing-wave.
The effect of heating was probably very significant in
\cite{mabuchi1996a,mabuchi1998b}. These semi-classical parameters can
be calculated numerically, as a function of atomic velocity, through a
matrix continued fraction calculation just as in the free-space
theory. Simulations of the classical trajectories of the atoms inside
the cavity can be performed in three dimensions using a Langevin
equation approach with the semiclassical force, friction and momentum
diffusion acting on a classical point particle
\cite{doherty1997a,mabuchi1998b}. It was found that the atom is in the
cavity long enough to be significantly heated and that only a few
atoms will be sufficiently slow that their motion along the standing
wave can be tracked. The heating of the motion means that the atoms
will eventually boil out of even the very deep potential wells that
can be set up by the dipole force due to the very large field
gradients inside the cavity.
One way to reduce the noise on the atom is to move into a highly
detuned regime, the dispersive limit of CQED, in which the atom
induces a phase shift on the field and the the dipole force provides a
nearly conservative potential for the atom. This corresponds to the
far off resonance trapping of atoms in optical lattices. In this limit
both the cavity field and the atomic internal state can be
adiabatically eliminated and a master equation written for the quantum
mechanical motional state alone \cite{doherty1998a}
\begin{mathletters}
\begin{eqnarray}
\dot{\rho} &=&\frac{1}{i\hbar }[H^{\prime },\rho ]+\frac{2g_{0}^{4}E^{2}}{%
\kappa ^{3}\Delta ^{2}}{\cal D}[\cos ^{2}(k_{L}x)]\rho \\
H^{\prime } &=&\frac{p_{x}^{2}}{2m}-\hbar \frac{g_{0}^{2}E^{2}}{\kappa
^{2}\Delta }\cos ^{2}(k_{L}x).
\end{eqnarray}
The Lindblad term describes heating due to light scattering caused by
cavity assisted spontaneous emission. Essentially this is an extra
contribution to the light scattering heating present in far off
resonant optical lattices and takes place even though the atom is in
principle never excited. Such a regime does provide the hope of long
trapping times but there are technical difficulties associated with
attaining sufficiently high detunings to fully realize the model,
although effects such as light scattering due to free space
spontaneous emission could easily be included in this treatment.
There are several reasons to have some other means of trapping the
atom in the cavity. Firstly even in the far detuned regime driving the
cavity field does not give a particularly slow heating environment for
the atom due to the increased light scattering out through the cavity
mirrors. Secondly, quantum computing and other interesting schemes for
this atom cavity system tend to require that the cavity field is
initially in the vacuum state and is not driven and that the motion of
the atom is undisturbed during the action of the gate. So it is
enticing to consider loading the cavity and trapping the atom with an
optical lattice, perhaps another --- lower finesse --- mode of the
Fabry-Perot, or with an ion trap. We anticipate that in any practical
realization of these models there will be a means of confining the
atom other than the cavity field alone.
It is important that the heating rate $\gamma _{\text{heat}}$ of the
atom in this potential be slow compared to the other dynamics of the
system and we anticipate that as in the dispersive regime considered
above the system will preserve the situation present initially in
current experiments, that the frequency $\omega $ associated with the
atomic motion is small
compared to
\kappa ,\gamma $ which are in turn small compared to the coherent
coupling
g.$ \ Conditions which will ultimately realize unitary evolution of
the atomic internal states and the cavity field will therefore be,
\end{mathletters}
\begin{eqnarray*}
g &\gg &\kappa ,\gamma , \\
g &\gg &\omega \gg \gamma _{\text{heat}}.
\end{eqnarray*}
As we will see below there is also a regime in which the mechanical
motion of the atom is much faster than the internal state evolution
and therefore effectively decouples from it. However the heating of
the atom would still have to be negligible for at least a few Rabi
cycles so that a gate could be performed before the motional state had
to be reset. This requires
\[
\omega \gg g\gg \gamma _{\text{heat}}
\]
which implies very much larger quality factors than current or near
future optical lattice or ion trap technology could provide. Finally
the requirement that the action of the gate not affect the motion of
the atom too greatly will mean that transitions between harmonic
oscillator states require more energy than is provided by the atom
emitting or absorbing a photon. Thus we also require that the recoil
frequency $\omega _{r}=\hbar k_{L}^{2}/2m$ for the atom and transition
under consideration is small compared to the motional frequency
\[
\omega \gg \omega _{r}.
\]
\section{Quantum Computing in CQED}
\label{sec2}
There are several proposals for realizing quantum gates in CQED with point
dipoles. We wish to consider here the unavoidable effects of the motion and
thus understand the level of control of the motion which will be
necessary
to realize these schemes with a high fidelity. In particular we wish to
compare a system based on controlling the times for which a given
interaction is turned on and off with one which relies on an adiabatic
passage through eigenstates of a time-dependent Hamiltonian and for which
the interaction time is not critical.
\subsection{Raman Scheme}
A model of the first type is given by van Enk {\em et al}
\cite{vanenk1997a} which employs a Raman transition in a cavity to
effect a two bit quantum gate. The procedure obtains conditional
dynamics for the two atomic internal states through the exchange of a
cavity photon between the two atoms, which are imagined to be confined
to the antinode of the cavity field. It is also assumed the atoms can
each be driven through the side of the cavity by a separate laser.
Each atom has two states $|0\rangle _{i},|1\rangle _{i}$which form the
qubit, an auxillary level $|r\rangle _{i}$ which is coupled to the
cavity and an excited
state
|e\rangle _{i}$ from which the laser driving is detuned. A Raman
transition is employed since this reduces the effect of spontaneous
emission. The interaction Hamiltonian for the atom cavity interaction
with the excited state adiabatically eliminated is
\begin{equation}
H=\sum_{j=1,2}\frac{gf_{j}(t)}{2}|1\rangle _{jj}\langle r|a+\text{H.c.}
\end{equation}
where the atoms have the level structure shown in figure
(\ref{level}). The function $\ f_{i}(t)<1$ describes some laser
driving pulse shape and the constant $g$ describes the effective Raman
coupling. Since both atoms interact with the cavity mode this
Hamiltonian can be used to build a quantum gate. This gate can be
designed so that population in $|0\rangle $ rarely has to interact
with the cavity thereby reducing the errors. In \cite {vanenk1997a} a
sequence of pulses is described which realizes the universal two-bit
gate
\begin{eqnarray*}
|0\rangle _{1}|0\rangle _{2} &\rightarrow &|0\rangle _{1}|0\rangle
_{2};\;|1\rangle _{1}|0\rangle _{2}\rightarrow -|1\rangle
_{1}|0\rangle _{2}
\\
|0\rangle _{1}|1\rangle _{2} &\rightarrow &|0\rangle _{1}|1\rangle
_{2};\;|1\rangle _{1}|1\rangle _{2}\rightarrow |1\rangle _{1}|1\rangle
_{2}.
\end{eqnarray*}
\subsection{Adiabatic Passage via Dark State }
A model of the second type is given by Pellizzari {\em et al} \cite
{pellizzari1995a} who show how to perform a controlled-NOT\ and
various other quantum gates by encoding two qubits onto four levels of
a single atom and employing laser driving to achieve the conditional
dynamics. Information about one of the qubits is transferred back and
forth between the two atoms in the gate by transferring coherences
between the ground states of one atom to the other through an
adiabatic passage involving excitation of the cavity field. This
adiabatic passage is through a dark state based on the single atom
dark states discussed in \cite{parkins1993a} and thus suppresses
spontaneous emission without employing a Raman transition since the
excited states of the atoms are in principle never occupied. The
interaction Hamiltonian is very similar to the one for the previous
system
\begin{equation}
H=\sum_{j=1,2}\frac{g}{2}|e\rangle _{jj}\langle r|A+\frac{\Omega
_{j}(t)}{2} |e\rangle _{jj}\langle 1|A+\text{H.c.}
\end{equation}
and this Hamiltonian has the dark states
\begin{eqnarray*}
|D_{0}\rangle &=&|r,r,0\rangle \equiv |r\rangle _{1}|r\rangle
_{2}|0\rangle
_{c} \\
|D_{1}\rangle &\propto &\Omega _{1}g|r,1,0\rangle +\Omega
_{2}g|1,r,0\rangle -\Omega _{1}\Omega _{2}|r,r,1\rangle .
\end{eqnarray*}
Here we have labeled the states in the same way as in the previous
model although $|r\rangle $ could in fact be used as the logical zero
of the qubit. With the second atom initially prepared in $|r\rangle$,
switching on the laser driving the second atom, so that $\Omega_{2}$
is initially large, and then slowly increasing the driving on the
first atom while decreasing the driving on the second, transfers the
state of the
first atom on
\{|r\rangle ,|1\rangle \}$ to the second atom. With appropriate
driving on the transition $|r\rangle \leftrightarrow |0\rangle$ the
logical state of the first atom is transfered to the logical state of
the second atom. In order to perform a gate it is necessary to
consider more than these three levels on atoms and the two qubits can
be transferred to coherences between four ground states of the second
atom on which the gate can be performed through Raman transitions
between the ground states. Since this laser driving can be achieved in
such a way that all of the field gradients in the vicinity of the atom
will be small we can disregard motional effects for this part of the
evolution. What is of interest is the motional state dependence of the
actual adiabatic transfer of coherence through the cavity field
described here so we will restrict ourselves to the simpler three
level system and investigate the behavior of the dark state when
motional states and the position dependence of the coupling to the
cavity are included.
\subsection{Errors in Quantum Computers}
Several strategies for overcoming the effects of unwanted couplings to
the environment --- including the motion of the atom --- are possible.
One approach, that of quantum error correcting codes, first described
by Shor and Steane \cite{shor1995a,steane1996a}, and encoded gate
operations, follows from the realization that errors during the
storage of a quantum state or its manipulation can be corrected by
coding the qubits in larger Hilbert spaces made up of several
identical quantum systems. This would correspond in our case to
several atoms in a single cavity or perhaps several cavities each with
their own atoms in order to achieve the redundancy necessary to
overcome the effects of environmental noise. In this case it can be
shown that given sufficient resources any quantum computation can be
performed as long as the fundamental error rate for each of the
systems is below some threshold value \cite{knill1998a}. If this
approach is taken the critical thing to know for a particular
candidate system is the fundamental error rate due to a given coupling
to the environment and what can be done to modify this rate. Other
strategies are system specific and revolve around characterizing the
errors that occur in a given physical realization of a quantum
computer and correcting for those errors alone, perhaps to all orders.
Work along these lines is very far advanced in the case of CQED
quantum computing, \cite{vanenk1997a} and references therein.
In the work of van Enk {\em et al} \cite{vanenk1997a} it i s necessary
that a stationary property holds for the coupling to the environment.
This essentially requires that the evolution operators which entangle
the basis vectors of the computational subspace with the environment
depend only on the length of time it takes to operate the gate and
that they commute with each other. This allows schemes which
symmetrize the effect of the noise on all of the basis states of the
qubits by allowing them to interact with the environment at different
times. Studying the effects of motion on the CQED quantum computer is
important because, unlike photon absorption, spontaneous emission and
systematic errors in the driving times, errors induced by the motion
do not obey this stationary property. This is because the free
evolution term and the position dependent coupling term do not
commute. In any situation where the motion of the atom is comparable
to the coupling $g,$ so that the evolution operators associated with
these terms in the Hamiltonian will be significantly non-commuting,
there will be a departure from the conditions required for the
correction schemes of \cite{vanenk1997a} to work ideally. Therefore
the rate of errors due to the motion will be one contribution to the
noise processes which limit the performance of the CQED quantum
computer with currently available simplified approaches to error
correction.
\subsection{Model for Motion of Atoms}
We need a model to describe the fundamental effects of motion in CQED
models such as those discussed above. The model we will use is
motivated by the differences of timescales discussed in the previous
section and the knowledge of correction schemes for photon losses and
spontaneous emission to all orders. The important rates are therefore
the coherent coupling rate achieved for the gate, the frequency
associated with the motion of atom and the recoil frequency which
describes the effect of the emission or absorption of a single photon
on the motion of the atom. Position dependence of the coupling
combined with the initial spread of the motional state and motion of
the atom in the trapping potential will be sources of noise in the
computation. We will assume that the particle is trapped in a harmonic
potential with a low heating rate. If the atom is cold and confined in
a far off resonance dipole trap then this will be the most significant
contribution to the effects of the motion although the oscillation
frequency may depend on the internal state of the atom. If the
oscillation rate in the potential is small compared to the effective
coupling frequency $g$ then in effect we are just adjusting the length
scale of the initial state of motion and the precise shape of the
potential and any dependence of the potential on the internal state
will not significantly change the results. We will leave the
consideration of heating of the atom during the gate action for future
work although in the limit in which the motion is rather slower than
the coupling and the heating rate is slower again than this then the
only significant effects of heating will be to lead to initially
thermal states of the motion which we will consider in the following.
Assuming then that each atom is trapped in a standing wave cavity in a
harmonic potential that is the same regardless of the internal state
of the atom we have the Hamiltonian
\begin{equation}
H_{i}=\frac{g_{i}}{2}\cos \left( \eta \left( A_{i}+A_{i}^{\dagger }\right)
\right) \left( \tilde{\sigma}_{i}+\tilde{\sigma}_{i}^{\dagger }\right)
+\omega A_{i}^{\dagger }A_{i}, \label{motqg}
\end{equation}
where we have defined the operator $\tilde{\sigma}_{i}=|1\rangle
_{ii}\langle r|a$ and $A_{i}$ is the lowering operator for the
motional state of the atom. The recoil frequency $\omega _{r}=\hbar
k_{L}^{2}/2m$ and
the oscillation frequency $\omega $ define a Lamb-Dicke parameter $\eta =$ $%
\sqrt{\omega _{r}/\omega }.$ In the ideal situation the atom is
tightly confined compared to the wavelength of the light and this
Lamb-Dicke parameter is very small. We assume that the couplings to
the lasers driving the atoms are position independent, this makes
sense because these lasers reach the atom through the sides of the
cavity and the beam width will in general be much larger than a
wavelength. This then is the model depicted schematically in figure
(\ref{schem}). There will in fact be some interaction between errors
caused by the motion and by cavity decay, which is the most important
feature left out of our model, however the purpose of this work is to
identify the ultimate sources of error due to the motion in the
situation in which the effects of cavity decay can in principle be
reversed to all orders as in \cite{vanenk1997a}.
\subsection{Characterising Imperfect Gates}
We also need a means of characterizing the success and failure of a
gate. In essence we need a measure of the distance between the actual
evolution and the ideal one. Several such measures of the distance
between superoperators have been used or proposed in related work
\cite {poyatos1997a,giedke1998a,aharanov1998a}. In this case we are
just interested in a simple measure which is physically motivated for
quantum gates. We will employ a simple modification the entanglement
fidelity introduced in \cite{schumacher1996a}. This is related to the
overlap or
fidelity of a state $\rho $ to some desired pure state $|\psi \rangle$, $%
F=\langle \psi |\rho |\psi \rangle$. $F$ is one if and only if $\rho
=|\psi
\rangle \langle \psi |$. The entanglement fidelity for a noisy evolution $%
{\cal E}$ on some state $\rho$ of a system $Q$ is
\begin{equation}
F_{e}(\rho ,{\cal E})=\langle \psi ^{RQ}|\left( {\cal I}^{R}\otimes {\cal E}%
\right) \left( |\psi ^{RQ}\rangle \langle \psi ^{RQ}|\right) |\psi
^{RQ}\rangle
\end{equation}
where $|\psi ^{RQ}\rangle $ is a pure state of $Q$ and a fictional
auxiliary system $R$ such that Tr$_{R}(|\psi ^{RQ}\rangle \langle \psi
^{RQ}|)=\rho $ and ${\cal I}^{R}$ is the identity superoperator on
$R.$ It is shown in \cite{schumacher1996a} that $F_{e}$ is independent
of the particular purification $|\psi ^{RQ}\rangle $ chosen. The
entanglement fidelity can be thought of as
characterizing how well the state and its entanglement are preserved by $%
{\cal E}$. It is shown in \cite{nielsen1996a} that
\[
F_{e}\left( \rho ,{\cal E}\right) =\min_{\rho ^{RQ},{\cal E}^{\prime
}}F\left( \left( {\cal E}^{\prime }\otimes {\cal I}^{Q}\right)
(\rho ^{RQ}),\left( {\cal E}^{\prime }\otimes {\cal E}\right) \left(
\rho ^{RQ}\right) \right)
\]
where $F$ is the fidelity of mixed states defined in \cite{jozsa1994a}
and describes how close two density matrices are to each other. $\rho
^{RQ}$ is
an extension of the state $\rho $ to the combined system such that Tr$%
_{R}(\rho ^{RQ})=\rho $ and ${\cal E}^{\prime }$ is an arbitrary
evolution on the auxiliary space $R$. Thus the entanglement fidelity
corresponds to the worst possible fidelity of the system state after
the evolution ${\cal E} $ to its initial state regardless of how the
system is entangled with the environment and of what dynamics ${\cal
E}^{\prime }$ the environment is undergoing. The entanglement
fidelity provides a good measure of the preservation of a state in the
memory of a quantum computer which could be entangled with many other
qubits in the computer and where these qubits could be undergoing
arbitrary evolutions as part of the computation. Moreover if $\rho
=\sum p_{i}|\psi _{i}\rangle \langle \psi _{i}|$ then the
entanglement fidelity is less than or equal to the average fidelity under $%
{\cal E}$ of the ensemble making up $\rho ,$ $F_{e}\leq \sum
p_{i}\langle \psi _{i}|{\cal E}\left( |\psi _{i}\rangle \langle \psi
_{i}|\right) |\psi _{i}\rangle .$ But on the other hand if the
fidelity of all of the pure states $|\psi \rangle $ with support on
$\rho $ is close to one then the entanglement fidelity is close to one
also \cite{knill1997a}.
Motivated by these considerations we will use a gate entanglement
fidelity which measures how close ${\cal E}$ is to the ideal unitary
evolution $U$ over the whole computational subspace of $\{|0\rangle
_{1}|1\rangle _{1}|0\rangle _{2}|1\rangle _{2}\}$ by
\begin{equation}
F_{eg}\left( {\cal E},U\right) =\langle \psi ^{RQ}|U^{\dagger }\left( {\cal I%
}^{R}\otimes {\cal E}\right) \left( |\psi ^{RQ}\rangle \langle \psi
^{RQ}|\right) U|\psi ^{RQ}\rangle
\end{equation}
where the $\rho $ is the completely mixed state on the computational
subspace Tr$_{R}(|\psi ^{RQ}\rangle \langle \psi ^{RQ}|)=\rho ={\cal I}%
^{C}/4.$ Thus if $F_{eg}$ is close to one then the gate is close to
ideal for all initial states of the two qubits regardless of how they
are entangled with the other qubits in computer and of how these other
qubits are being manipulated during the gate operation. This measure
has the property of measuring not just how close the evolution is to
the ideal evolution for any pure state on the computational subspace
but also how well the evolution preserves entanglement between the
state of the system and the state of \ other systems which may be part
of the quantum computer.
\section{Gate Fidelity for Raman Scheme}
\label{sec3}
Position dependence of the coupling and motion of the atom in the
trapping potential will be a source of noise. In order that the atom
in fact be localized near the antinode of the cavity it should occupy
a motional state of low excitation in the potential and have a recoil
frequency $\omega _{r}=\hbar k_{L}^{2}/2m$ much smaller than the
oscillation frequency of the atom. Assuming that this is the case we
can perform time-dependent perturbation theory in order to find the
fidelity of the fundamental entangling evolution. If we do this to
$O(\omega _{r}^{2}/\omega ^{2})\,$we can consider the simplified
Hamiltonian
\begin{mathletters}
\begin{eqnarray}
H &=&H_{0}+V, \\
H_{0} &=&\frac{g}{2}\left( \tilde{\sigma}+\tilde{\sigma}^{\dagger }\right)
+\omega A^{\dagger }A, \\
V &=&-\frac{\eta ^{2}g}{4}\left( A+A^{\dagger }\right) ^{2}\left( \tilde{%
\sigma}+\tilde{\sigma}^{\dagger }\right)
\end{eqnarray}
\end{mathletters}
By moving into the interaction picture defined by $H_{0}$ it is
possible to do time-dependent perturbation theory to calculate
approximate states for the system and therefore the entanglement
fidelities and motional excitation. If we restrict our interest just
to one atom in the cavity for the moment then we can calculate the
Schr\"{o}dinger picture ket where the internal and cavity states are
initially $|1\rangle _{1}|0\rangle _{\text{c}}$ and then leave the
laser on such that in the idealized (point-dipole) case we end in the
state $|r\rangle _{1}|1\rangle _{\text{c}}$. We wish to leave open the
possibility of tailoring the length of the laser pulse such that the
fidelity of the final state is optimized by using a pi-pulse
appropriate to the mean-squared position of the atom. Thus we choose
the interaction
time
t=\pi (1+\delta )/g$ where $\delta$ will be of order $\omega
_{r}/\omega $ and will be chosen to maximize the fidelity of the final
state. We will consider initially just a number state of the atom and
perform the thermal average at the end of the calculation.
The overall Schr\"{o}dinger picture state after the evolution is
\begin{eqnarray*}
&&\left(
\begin{array}{c}
1-\frac{1}{2}\left( \frac{\pi \eta ^{2}}{4}\left( 2n+1\right) -\frac{\pi
\delta }{2}\right) ^{2} \\
-\left( \frac{\eta ^{2}g}{4\omega }\sin (\pi \omega /g)\right) ^{2}\left(
n^{2}+n+1\right)
\end{array}
\right) |r\rangle _{1}|1\rangle _{\text{c}}|n\rangle _{\text{m}1} \\
&&+i\left( \frac{\pi \eta ^{2}}{4}\left( 2n+1\right) -\frac{\pi \delta }{2}%
\right) |1\rangle _{1}|0\rangle _{\text{c}}|n\rangle _{\text{m}1} \\
&&+\left( \frac{\eta ^{2}g}{8\omega }\sqrt{(n+1)(n+2)}\left( 1-e^{-2i\pi
\omega /g}\right) \right) |1\rangle _{1}|0\rangle
_{\text{c}}|n+2\rangle _{\text{m}1} \\
&&-\left( \frac{\eta ^{2}g}{8\omega }\sqrt{n(n-1)}\left( 1-e^{2i\pi \omega
/g}\right) \right) |1\rangle _{1}|0\rangle _{\text{c}}|n-2\rangle
_{\text{m}1}.
\end{eqnarray*}
where we have only retained those terms which turn out to affect the
fidelity and entropy up to fourth order in the Lamb-Dicke parameter
and have disregarded an overall phase. Clearly the majority of the
population is in the desired final state, there is also population
left in the original internal state and a superposition of motional
states.
We assume that the initial motional state is in fact a thermal state
of average excitation $\bar{n}.$ The thermal averaging can be
performed by summing the series for the terms in the reduced density
matrix of the internal and cavity states resulting from each of the
individual initial number states since these are just geometric series
or their derivatives. The fidelity for this interaction with the
cavity is
\begin{eqnarray}
F &=&1-\frac{\pi ^{2}\delta ^{2}}{4}+\frac{\pi ^{2}\eta \delta }{4}%
\left( 2\bar{n}+1\right) -\left( \frac{\pi \eta ^{2}}{4}\right)
^{2}\left( 8\bar{n}^{2}+8\bar{n}+1\right) \nonumber \\
&&-\left( \frac{\eta ^{2}g}{2\omega }\right) ^{2}\sin ^{2}(\pi \omega
/g)\left( \bar{n}^{2}+\bar{n}+\frac{1}{2}\right) .
\end{eqnarray}
We may have sufficient control over the length of the laser pulse that
we can choose $\delta $ so as to maximize this quantity, thus giving
us the best possible fidelity of the evolution. Setting $\delta =\eta
^{2}\left( 2\bar{n}+1\right) /2$ gives us
\begin{eqnarray}
F_{\text{opt}} &=&1-\left( \frac{\pi \eta ^{2}}{2}\right) ^{2}\left
( \bar{n}^{2}+\bar{n}\right) \nonumber \\
&&-\left( \frac{\eta ^{2}g}{2\omega }\right) ^{2}\sin ^{2}(\pi \omega
/g)\left( \bar{n}^{2}+\bar{n}+\frac{1}{2}\right) .
\end{eqnarray}
These expressions show the basic behavior of the system in a number of
regimes. In the most relevant limit that the harmonic oscillation is
much
slower than the internal dynamics $\omega \ll g$ we get $F_{\text{opt}%
}\simeq 1-2\left( \pi \eta ^{2}/2\right) ^{2}\left( \bar{n}^{2}+\bar{n}+%
\frac{1}{4}\right)$. In this case the motion of the atom is irrelevant
during the time for a pi-pulse and so in this case the parameters
$\eta $ and $\bar{n}$ essentially just define the initial position
spread and coherence length of the atomic motional state. This is the
fidelity that would be achieved for any such initial state regardless
of the details of the atomic motion on longer timescales. The opposite
limit of very fast motion $g\ll \omega \,\ $amounts to a rotating wave
approximation for the mechanical motion in which the atom oscillates
in the potential many times during a single operation,
$F_{\text{opt}}\simeq 1-\left( \pi \eta ^{2}/2\right) ^{2}\left(
\bar{n}^{2}+\bar{n}\right) .$ This limit is attractive since it
suggests that if the oscillator is sufficiently cold then the effects
of motion could be overcome simply by modifying the naive length for a
pi-pulse of the system. However this still requires the assumption
that the heating rate of the motion $\omega _{\text{heat}}\ll g$,
which implies an enormous quality factor for the mechanical motion.
The laser power required to achieve $g\ll \omega$ in a far off
resonant optical trap would probably be prohibitive in any case. Noise
in current ion trap experiments would result in heating rates that
were at least comparable with the couplings $g$ so such a regime would
appear to be unfeasible with near future technology.
Neither will the computing operations leave the motional state
unmodified. The state will be heated until eventually it will become
necessary to cool the motion of the atom. An indication of this can be
found by calculating the excitation of the motional state after one
exchange on excitation between the atom and the cavity. This is
entirely due to contributions resulting from transitions between
motional states at some stage during the evolution and as such depends
on trigonometric functions of the ratio between coupling and
mechanical oscillation frequencies and is independent of small changes
in the length of the laser driving,
\begin{equation}
\langle A^{\dagger }A\rangle -\bar{n}=\left( \frac{\eta ^{2}g}{2\omega }%
\right) ^{2}\sin ^{2}(\pi \omega /g)\left( 2\bar{n}+1\right) .
\end{equation}
So that in the rotating wave regime $g\ll \omega $ the effective
decoupling of the internal state and cavity dynamics from the motion
means that the motional state is unaffected by the action of the gate.
On the other hand in the more realistic situation $\omega \ll g,$
$\langle A^{\dagger} A\rangle -\bar{n}\simeq \left( \pi \eta
^{2}/2\right) ^{2}\left ( 2\bar{n}+1\right)$.
It is straightforward to extend this calculation to the full evolution
of the quantum gate with two atoms in the cavity described above and
to evaluate the entanglement fidelity for the gate operation
\begin{eqnarray}
\label{entfid}
F_{eg} &=&1-\pi ^{2}\eta ^{4}\left( \bar{n}^{2}+\bar{n}+\frac{1}{8}\right)
\\
&&-\frac{\eta ^{2}g^{2}}{4\omega }\sin ^{2}(\pi \omega /g)\left( 1+\cos
^{2}(\pi \omega /g)\right) \left( \bar{n}^{2}+\bar{n}+\frac{1}{2}\right)
\nonumber
\end{eqnarray}
which we give here for the situation in which the driving is not
optimized. In this limit of fidelity close to one the entanglement
fidelity essentially reduces the to average of the fidelities of the
gate operation on each of the four basis states of the computational
subspace, although this is not true in general.
The motion of the atom will be excited by the action of the gate
depending on the actual initial state of the gate. However in the
operation of the gate this initial state could be any superposition of
these and could be entangled with the state of other qubits in the
computer. As a
measure of the overall heating of the motion we will calculate $n_{i}=$Tr$%
\left( A_{i}^{\dagger }A_{i}{\cal E}\left( |\psi ^{RQ}\rangle \langle
\psi ^{RQ}|\right) \right) $ for a purification on the
computational
subspace of
\rho ^{C}={\cal I}^{C}/4$ as discussed above. This basically assumes
no knowledge of the internal state and therefore averages the effect
of the motion for each of the four basis states of the computational
subspace. These excitation parameters can be calculated
\begin{mathletters}
\begin{eqnarray}
n_{1}-\bar{n} &=&\left( \frac{\eta ^{2}g}{2\omega }\right) ^{2}\sin
^{2}(\pi
\omega /g)\left( \allowbreak 2\bar{n}+1\right), \label{anheat} \\
n_{2}-\bar{n} &=&\left( \frac{\eta ^{2}g}{4\omega }\right) ^{2}\sin
^{2}(2\pi \omega /g)\left( \allowbreak 2\bar{n}+1\right).
\end{eqnarray}
The different dependence on $\omega /g$ is due to the motional states
of the atoms being excited at different times during the action of the
gate.
We have also performed numerical simulations of the Hamiltonian
(\ref{motqg} ) with laser pulses as described above to all orders in
the Lamb-Dicke parameter, employing a number state expansion of the
motional operators. Here we will plot results for initial states
where both the cavity and the atomic motion are initially in the
ground states $|\psi ^{RQ}\rangle =|\psi
^{RC}\rangle |0\rangle _{c}|0\rangle _{1m}|0\rangle _{2m}$ where Tr$%
_{C}\left( |\psi ^{RC}\rangle \langle \psi ^{RC}|\right) ={\cal
I}^{C}/4.$ In figure (\ref{entfid1}) the gate entanglement fidelity
for this procedure is plotted as a function of the Lamb-Dicke
parameter along with the analytic approximation resulting from
equation (\ref{entfid}). This approximation is seen to hold up to
reasonably large values of $\eta .$ We considered recoil frequencies
sufficiently small that the entanglement fidelity and motional
excitation were effectively independent of $\omega $ except through
$\eta $, the actual values used in these simulations were $g=1,\omega
_{r}=0.0005$. Once $\eta \simeq 0.4,$ errors are in the region of
$10\%$. The motional excitation of the first atom is plotted in figure
(\ref{heat1}) along with the approximation of equation (\ref{anheat}).
\section{Adiabatic Passage and Motion} \label{sec4}
A comparison of the previous scheme with one involving adiabatic
passage is motivated by the fact that adiabatic passage schemes do not
depend on the pulse area of the laser pulses. The position variation
of the coupling means that different parts of the wave function see
different pulse areas and so may be under or over-rotated by the
driving laser. All that is required for the adiabatic theorem to hold
is that the Hamiltonian is varied sufficiently slowly that
non-adiabatic transitions, the rate of which depend on the energy
separation between the eigenstates, do not occur, see for example
\cite{messiah1962a}. The energy spacing of the eigenstates from
neighboring eigenstates is determined by the size of the coupling $g$
so it might be hoped that it would be practical to perform the
transfer sufficiently slowly that all the population in a wide area
around the antinode of the field was transferred through the dark
state with high fidelity.
Another way of seeing this is to consider the eigenstates of the
Hamiltonian with motion included. Dark states exist which can effect
the transfer as long as $\omega \ll g$ --- a kind of Raman-Nath regime
for the gate. The values $i,j,k,l,m$ in the ket $|i,j,k,l,m\rangle
\equiv |i\rangle_{1} |j\rangle_{2} |k\rangle_{\text{c}}
|l\rangle_{\text{m}1} |m\rangle_{\text{m}2}$ refer to the internal
state of the first atom, of the second atom, the cavity state, the
motional state of the first atom and of the second atom, respectively.
The states
\end{mathletters}
\begin{eqnarray*}
|D_{1}\rangle &\propto &\Omega _{1}g|r,1,0,n_{1},n_{2}\rangle +\Omega
_{2}g|1,r,0,n_{1},n_{1}\rangle \\
&&-\Omega _{1}\Omega _{2}|r,r,1,n_{1},n_{2}\rangle \\
&&-\eta ^{2}\Omega _{1}\Omega _{2}\left( A_{1}^{\dagger }+A_{1}\right)
^{2}|r,r,1,n_{1},n_{2}\rangle \\
&&-\eta ^{2}\Omega _{1}\Omega _{2}\left( A_{2}^{\dagger }+A_{2}\right)
^{2}|r,r,1,n_{1},n_{2}\rangle
\end{eqnarray*}
are eigenstates of the Hamiltonian including atomic motion up to
$O(\eta ^{4})$. Thus terms in the Hamiltonian of $O(\eta ^{2})$ do not
cause errors in the adiabatic passage as they do in the Raman scheme.
As a result we could expect fidelities for the process differing from
one by numbers of $O(\eta ^{8}).$
We simulated the adiabatic transfer of coherence discussed above for $%
g=1,\omega _{r}=0.0001$. We used Gaussian pulse profiles
$f_{i}(t)=\exp \left( -\left( g\left( t-t_{i}\right) /40\right)
^{2}/2\right) $ where the two pulses were separated by a time
$\Delta t=t_{2}-t_{1}=80/g.$ This choice resulted in transfer with
high fidelity and little population of the atomic excited states for
the point dipole atom. Figure (\ref{fid2}) plots the fidelity of the
transfer for different values of $\eta .$ Any gate built on this
principle will be limited by the fidelity of this procedure. Laser
pulses into the side of the cavity will have very much less affect on
the motion and will be achievable with high fidelity presuming the
atoms can be addressed separately by the lasers. Thus we do not show a
full gate entanglement fidelity for the gate described in
\cite{pellizzari1995a} here but it will be of the same order as the
fidelity for the transfer if laser operations are performed
accurately. The striking feature of figure (\ref {fid2}) is that for a
much larger range of $\eta $ the fidelity is essentially undisturbed
by the motion. As $\eta $ is increased the transfer takes place with
increased cavity excitation and motional excitation, but without any
increase in the excited state population, as suggested by the
approximate dark state above. Thus for $\eta \simeq 0.4$ errors are
still only 0.2\% although the fidelity has begun a sharp decline.
Simulation of the internal and external degrees of freedom for two
atoms as well as the cavity mode requires a very large Hilbert space
which is computationally intensive and so we have not explored values
of $\eta $ beyond those plotted here. So that the point at which the
fidelity becomes unusefully small is yet to be established.
\section{Conclusions} \label{sec5}
We have investigated the effect of motion on experiments in cavity
QED. In particular we discussed the necessity of good control over the
motional state in order to realize the dynamics expected for models
involving point dipole atoms. The ultimate limitations that the
motional state places on quantum computing in CQED systems was
discussed for both a Raman scheme and one involving adiabatic passage
via a dark state to transfer information between the atoms. The scheme
involving adiabatic passage was found to be extremely robust to the
precise nature of the atomic motion which may be an important
consideration in future experimental implementations of similar
schemes.
\section*{Acknowledgments}
This work is supported by the Marsden Fund of the Royal Society of New
Zealand.
\begin{figure}[tbp]
\caption{Level structure of the atoms in cavity QED quantum computing
models. Information is typically encoded on the states $\{|0\rangle
,|1\rangle \}$.}
\label{level}
\end{figure}
\begin{figure}[tbp]
\caption{Schematic of imagined CQED quantum gate. Two laser beams drive
atoms which are harmonically trapped at antinodes of a high finesse
microcavity}
\label{schem}
\end{figure}
\begin{figure}[tbp]
\caption{Entanglement fidelity for quantum gate with motion as a
function of the Lamb-Dicke parameter $\protect\eta $. Both atoms are
initially in the ground states of their motion. The solid line is from
numerical
calculations to all orders of $\protect\eta $ while the dotted line
represents an analytical approximation up to $O(\protect\eta ^{4}).$}
\label{entfid1}
\end{figure}
\begin{figure}[tbp]
\caption{Graph of average excitation $\langle a^{\dagger }a\rangle $ of one
atom after the action of the quantum gate as a function of the
Lamb-Dicke parameter $\protect\eta .$ The atoms are initially in the
ground state of their motion. The solid line represents the results
of numerical computations to all orders of $\protect\eta $ while the
dotted line represents an analytical approximation to
$O(\protect\eta ^{4}).$}
\label{heat1}
\end{figure}
\begin{figure}[tbp]
\caption{Fidelity of the adiabatic transfer of from one atom to another as
a function of the Lamb-Dicke parameter with atoms initially in the
ground state of their motion. Note the improved performance
compared to figure (\ref{entfid1}).}
\label{fid2}
\end{figure}
\vspace{-0.3cm}
|
\section{Introduction}
The motivation for this work has its origin in recent experimental
data by DiMauro et al~\cite{sheehy:1998:ema}
who studied high order harmonic generation
(HOHG) and above threshold ionization (ATI) in potassium driven
by strong radiation in the wavelength range 3200--3900 nm.
Although HOHG and ATI have been and continue to be studied
extensively, the bulk of the data and theory have concentrated on
the noble gases. Experimental convenience has been one of the
reasons for this preference, but, at least as far as HOHG is
concerned, their relatively high ionization potentials and
resistance to ionization have also tilted attention in their
direction. The alkali atoms belong to an entirely different
class, when it comes to their behavior under strong field
excitation. For atomic numbers comparable to the respective
noble gas (potassium versus argon in our context), their
ionization potential is lower by more than a factor of three.
Their excited states and distribution of the oscillator strength
for transitions from the ground state are also considerably
different. The energy of the first excited state in potassium
is much closer to the ground state than it is in argon.
As a consequence, if we consider, for example, 12-photon
ionization in potassium, five of the photons reach above the first
excited state, and the rest seven must be absorbed within the manifold
of its excited states. In contrast, for 12-photon ionization
of argon, it takes nine photons to reach the energy range of
the first excited state and only the remaining three will involve excitation
within the manifold of excited states. In addition, the
wavelength needed for potassium (about 3000 nm)
is longer by a factor of three than that needed
for argon
(about 1100 nm).
One might thus expect that, in the process of ionization, an
extensive manifold of excited and Rydberg states will be
strongly driven and perhaps populated. This should lead to
a structure in the ATI energy spectrum different in appearance
from what we are accustomed to. One might also anticipate
that the behavior should have similarities with that observed
in Rydberg states driven by microwave fields. Resolution
requirements in photoelectron energy analysis do not allow the
observation of individual ATI peaks in that case, although it
is quite feasible to resolve such peaks in potassium driven
from its ground state by radiation at mid-infrared wavelengths;
as Sheehy et al.~\cite{sheehy:1998:ema}
have shown. It is the exploration of
possible links and similarities between these two situations that
induced us to undertake this work. Clearly, the requirements
on intensity for saturation are expected to be lower in
potassium than in argon. Moreover, given the expected
participation of manifolds of excited states in potassium, the Keldysh
parameter as a criterion for the departure from multiphoton
ionization may not be as valid as it should be in argon where
most of the energy interval from the ground state to the
continuum is empty of excited states; imitating thus better the
Keldysh model which is essentially based on a ground state
and an ionization threshold.
The outline of this paper is as follows: The theory, namely the model
used to describe the atom, and the two propagation methods used to
solve the time-dependent Schr\"odinger equation are briefly presented
in section~(\ref{sec:Method}). Section~(\ref{sec:Results}), starts
with a presentation of the parameter range of the simulations that
follow and a demonstration of convergence by comparing two different
methods. It then moves on to present results in the 12-, 13-, and
14-photon ionization range, and discuss a low-energy plateau in the
ATI spectrum. 12-photon ionization of Hydrogen starting from the 2s
is compared to the results from Potassium. We conclude in
section~(\ref{sec:Conclusions}), by summarizing the main findings of
our results. In the appendix we present some details of the atomic
structure model, and compare it with an alternative approach.
\section{Method}
\label{sec:Method}
Potassium, being an alkali atom, can be considered a single
electron system for most of the phenomena in which double excitation
does not play an essential role. The first ionization threshold is
about 4.34 eV above the ground state; 18.8 eV are needed to
reach the lowest doubly excited state, which leaves us with enough
room to study single electron dynamics. The simplest way to do this
is by using a model potential that incorporates the effect of the
core electrons, and thus reduces the dynamics to a single electron
scheme. Different implementations of this basic idea have been used
so far, with some success, in the Single Active Electron approach,
pioneered by Kulander~\cite{Kulander:1988:tdt}, the frozen core
calculations~\cite{tang:1991:ntd}, and other model potential
calculations~\cite{lambrecht:1998:pca}. For the purpose of studying
the dynamics in the mid-infrared, it is sufficient to use a
simple form of the potential, proposed by
Hellmann~\cite{szasz:1985:pta}, which in atomic units is given as:
\begin{equation}
V_m = - \frac{1}{r} + A \frac{e^{-K r}}{r},
\label{eq:hellmann:definition}
\end{equation}
where $A$ is 1.989 and $K$ is 0.898. All formulas that follow are given
in atomic units. The Hellmann potential with the above mentioned
parameters, results in a ground state energy of 38950 wavenumbers from
the first ionization threshold, which is lower than the energy of
the actual
ground state (35010 wavenumbers) by 11\%. Owing to the difference between the
ground state energies of the model potential and of the real atom, we
also use a scaled wavelength, namely we rescale the energy of the
photon needed in an experiment that studies the same
process, by the ratio of the model to the theoretical ground state
energy. An alternative approach is to correct the energy of the
ground state, and probably some matrix elements to satisfy the
oscillator strength sum rules, but this leads to nonlocal
modifications in the Hamiltonian and is difficult to implement
effectively when the propagation is not done in an eigenbasis expansion.
The dynamical part of the problem is treated by solving the resulting
time-dependent Schr\"odinger equation in the dipole approximation:
\begin{equation}
\label{schroedinger_equation}
{\mathrm{i}} \partial_t \Psi(\vect{r}, t)
=
\left[
H_a + D(t)
\right]
\Psi(\vect{r}, t),
\end{equation}
where $\Psi$ is the wavefunction describing the outer electron, and
depends on the spatial electronic coordinates and on time, $H_a$ is
the time-independent field-free atomic Hamiltonian, and $D(t)$ is the
dipole interaction of the atom with the field. We only use the
velocity form of the interaction operator, following detailed studies
on the convergence properties of the solution~\cite{cormier:1996:ogg},
which have shown that the expansion of the wavefunction in terms of
spherical harmonics can be shortened dramatically if the velocity
gauge is used instead of the length gauge. In the velocity gauge, the
dipole operator can be cast in the following form:
\begin{equation}
\label{DT}
D(t) = - \alpha {\vect{A}(t)} \cdot \vect{p},
\end{equation}
where $\alpha$ is the fine structure constant, $\vect{p}$ is the
momentum operator, and ${\vect{A}(t)}$ is the vector potential which, within the
dipole approximation, has no spatial dependence. We choose a
convenient form for the pulse envelope, namely a $\cos^2$, avoiding
the long tails of a Gaussian that make the numerics more difficult,
without significantly affecting the results. The explicit form is:
\begin{equation}
A(t)=
\frac{{\cal E}_{0}}{\omega}
\cos(\frac{\pi t}{\tau})^{2}
\cos(\omega t),
\quad
\text{with }
t \in [\frac{-\tau}{2},\frac{\tau}{2}],
\end{equation}
where $A(t)$ is the amplitude of the vector potential, $\tau/2$ is the
Full Width at Half Maximum (FWHM), ${\cal E}_{0}$ and $\omega$ the
maximum field strength and fundamental frequency respectively.
The solution of the resulting time-dependent equation is written in a
system of
spherical coordinates and expanded in terms of radial functions
and angular spherical harmonics. This choice is dictated by
the central symmetry of the atomic system and has the advantage of
requiring the discretization of only one coordinate. Note that this
choice leads to efficient algorithms only if the linearly polarized
field is not too strong (compared to the coulomb field) in which case
the global symmetry of the entire system would rather be
cylindrical. Writing the solution in the cylindrical system would
require discretization of 2 coordinates~\cite{maragakis:1997:tdh} greatly increasing the
numerical cost of the algorithm. The problem is therefore treated
within "a box" (a sphere in the present case) whose radius is chosen
sufficiently large to contain the expanded atom during the
interaction. Part of the
procedure for testing convergence consists in ascertaining
that the ATI spectrum is
insensitive to the radius of the box.
We have used two methods of propagating the TDSE, namely a propagation
onto eigenstates in a
box~\cite{lambropoulos:1998:tea,lambropoulos:1998:asi}, and a
propagation on a $B$-Splines
basis~\cite{cormier:1996:ogg,cormier:1997:ati}.
The expansion of the time-dependent wavefunction on an eigenbasis set
reads:
\begin{equation}
\label{eq:psi_basis}
\Psi (t) =\sum_{l,n} b_{l,n} (t) \Phi^{l}_{n(E)},
\end{equation}
where $\Phi^{l}_{n(E)}$ are the field-free box eigenstates of the atom of
angular momentum $l$. Since we use linearly polarized light, we only need
the $m=0$ magnetic quantum number, as the initial state has
$m=0$, and dipole transition selection rules forbid mixing of other
magnetic sublevels. The time-dependent Schr\"{o}dinger equation is
transformed into:
\begin{equation}
\mbox{i} \frac{d}{dt}b_{l,n}(t)=\sum_{n^{\prime}l^{\prime}}(E_{nl}
\delta_{nn^{\prime}}\delta_{ll^{\prime}}-D_{nl,n^{\prime}l^{\prime}}(t))
b_{l^{\prime},n^{\prime}}(t),
\end{equation}
with the initial condition $|b_{l=0,n=1}(0)|^{2}=1$. $E_{nl}$ are the
eigenvalues in the box; $D_{nl,n^{\prime}l^{\prime}}$ are the dipole
matrix elements. Thus the problem has been transformed to a set of
ordinary differential equations for the coefficients $b_{l,n}(t)$ of
the wavefunction, which are solved using a high order, explicit
propagation technique, namely a fifth-order and sixth-order
Runge-Kutta-Verner method.
The ionization yield is calculated by adding up
the occupation probabilities of all discretized continuum states at the end
of the pulse; the above threshold ionization (ATI) spectrum is
obtained by the window operator projection
technique~\cite{kulander:1992:tds,gavrila:1992:ail}. Bound state
populations are given directly by the square of the norm of the
coefficients of equation~(\ref{eq:psi_basis}). For the construction
of the box-eigenstates that are used as our basis, we use an
expansion onto $B$-Splines, a method that is gaining momentum in many
parts of atomic physics as was pointed out
in~\cite{sapirstein:1996:ubs}. The codes we use are based on ideas
published in~\cite{tang:1991:ntd,lambropoulos:1998:tea},
which have been expanded to accommodate the need for large boxes.
It should be stressed that after the basis has been
constructed (i.e.\ energies and matrix elements have been
calculated), the rest of the procedure is neutral to the
technique for the construction of the basis.
The second approach rests on expanding the radial part of the
time-dependent wavefunction directly onto
$B$-splines~\cite{cormier:1996:ogg,cormier:1997:ati}:
\begin{equation}
\label{eq:Psi_BSplines}
{\Psi(\R, t)}
=
\llimSum{i l} b_i^{l}(t) \frac{B_i^{(k)}(r)}{r} \YLOTF
\end{equation}
where in addition to the spherical harmonics, $B_i^{(k)}(r)$ is the
$i$-th $B$-spline of order $k$ depending only on the radial coordinate
and $b_i^{l}(t)$ are time-dependent coefficients to be determined by
the solution of the TDSE. Again, only $m=0$ magnetic quantum numbers
are relevant. The major difference in this approach is that
we need not prediagonalize anything other than the initial state.
Substitution of equation~(\ref{eq:Psi_BSplines}) into the
Schr\"odinger equation~(\ref{schroedinger_equation}) leads to a banded
system of differential equations, which is solved by implicit
propagation techniques, currently involving a Bi-conjugate gradient
method with preconditioner. This second method of propagation scales
only linearly with increasing box size: it is thus more efficient
when we need a large box. Although our original exploratory
calculations have been made with the eigenbasis expansion method, which
works quite well for small boxes, most of the results presented in
what follows have been obtained through the direct expansion of the
time-dependent
Hamiltonian onto $B$-Splines.
\section{Results \& Discussion}
\label{sec:Results}
\subsection{General Considerations}
\label{sec:sub:general_considerations}
The only guideline as to what to expect in our
study are the experimental data by Sheehy et
al.~\cite{sheehy:1998:ema}, in which 12-, 13-, and 14-photon
ionization of Potassium has been studied, with 3.2 $\mu$m, 3.6 $\mu$m,
and 3.9 $\mu$m, 1.5 ps pulses respectively, and intensities close to
saturation. The 1.5 ps pulse is computationally impractical
in the TDSE
framework, thus we have chosen to place our study in the short pulse
regime, with the total pulse duration $\tau$ of the $\cos^2$ pulse
being 20 optical cycles, which (depending on the wavelength)
corresponds to a width of 96 fs to 136 fs for the results that we
present. The wavelengths we use are scaled, to compensate for
the inaccuracy of the ground state energy, and are such that 12-, 13-
and 14-photon ionization takes place. An intensity range estimate is
obtained by calculating the generalized cross-section through a scaling
technique~\cite{lambropoulos:1985:mmi,lambropoulos:1987:u}, using the
energy (0.295 Hartree), and radius (5.24 a.u.) obtained with the
general Hartree-Fock code published by
Froese-Fischer~\cite{fischer:1987:ghf}. From the cross-section, we
obtain a
saturation intensity estimate by solving $\Gamma
t_{\text{eff}} = 1$, where $\Gamma$ is the ionization width, and
$t_{\text{eff}}$ an effective pulse duration, of the order of its
FWHM~\cite{charalambidis:1997:mis}. After the first time-dependent
calculations, it was established that scaling was underestimating
saturation intensity
by more than an order of magnitude. We also
calculate the Ammosov, Delone, Krainov (ADK) rate of tunneling
ionization~\cite{ammosov:1986:tic}, which does not depend on the
wavelength. Solving $\Gamma t_{\text{eff}} = 1$, we
obtain a saturation intensity estimate of about $4 \times 10^{12}$
W/cm$^2$, in agreement with the results of the simulations.
For some characteristic wavelengths, corresponding to 12-photon
ionization scaled from the experimental wavelength, and the limits of
the 13-photon ionization range, we show, in
table~(\ref{tab:parameter:all}), the FWHM duration $\tau/2$, the
scaled-theory saturation intensity I$_{\mathrm{s}}$, the ADK theory
saturation intensity estimate I$_{\mathrm{ADK}}$, and the upper
limit E$_{\mathrm{c}}$ of the converged ATI spectrum
for a 3000 atomic units box.
E$_{\mathrm{c}}$ is calculated by estimating the energy needed for a
free electron originally placed at the nucleus to reach the boundary
of the box during the pulse; it is a useful simple estimate of the box
size needed to study ATI spectra, as has been shown in section (5)
of~\cite{cormier:1997:ati}. Next, and for two intensities, we show
the ponderomotive energy $U_p$, which is the major component of the
shift of the Rydberg states and the
continuum~\cite{avan:1976:ehf,mittleman:1984:kmi,freeman:1986:pea}.
It is given by:
\begin{equation}
U_p = \frac{I}{4 \omega_0^2} \sim I \lambda^2
\label{eq:ponderomotive}
\end{equation}
where $I$ is the laser intensity, $\omega_0$ the photon energy and
$\lambda$ the corresponding wavelength. For the highest intensities
that we use,
the ponderomotive energy is a multiple of the photon energy, which is
around a third of an eV for the wavelengths in the table. We also
present the Keldysh tunneling parameter
$\gamma$~\cite{keldysh:1965:u}, defined by:
\begin{equation}
\label{eq:gammasqr}
\gamma^2 = \frac{I_p}{2 U_p}
\end{equation}
where $I_p$ is the ionization potential, and $U_p$ the ponderomotive
potential. Note that for all wavelengths, and for intensities up to
$2 \times 10^{12}$ W/cm$^2$, $\gamma$ is larger than one. Thus,
according to the Keldysh theory of tunneling ionization, the process
lies in the multiphoton regime, and it is meaningful to refer to
the order of the transitions involved.
\end{multicols}
\begin{table}
\caption{Parameters for some of the calculations in Potassium.}
\label{tab:parameter:all}
\begin{tabular}{ddccdcdd}
$\lambda$(nm) & $\tau/2$(fs)& $I_{\mathrm{s}}$(W/cm$^2$) &
$I_{\mathrm{ADK}}$(W/cm$^2$) &
$E_{\mathrm{c}}$(eV) & $I$(W/cm$^2$) & $U_p$(eV) & $\gamma$ \\
\hline
2880 & 96 & $1.5 \times 10^{11}$ & $4.1 \times 10^{12}$ &7.76 & $10^{11}$ & 0.073 & 5.23 \\
& & & & & $10^{12}$ & 0.733 & 1.65 \\
3125 & 104 & $1.3 \times 10^{11}$ & $4.0 \times 10^{12}$ & 6.59 & $10^{11}$ & 0.086 & 4.82 \\
& & & & & $10^{12}$ & 0.863 & 1.52 \\
3300 & 110 & $1.2 \times 10^{11}$ & $4.0 \times 10^{12}$ &5.91 & $10^{11}$ & 0.096 & 4.56 \\
& & & & & $10^{12}$ & 0.961 & 1.44 \\
\end{tabular}
\end{table}
\begin{multicols}{2}
\begin{figure}
{\epsfxsize=12cm \epsfbox{levelPhoton.3300.ps}}
\caption{Truncated atomic structure of Potassium (calculated by the
Hellmann potential), and quantum paths leading to 13-photon ionization}
\label{fig:level-photon:3300nm}
\end{figure}
In figure~(\ref{fig:level-photon:3300nm}) we show a visual
representation of Potassium in a 3300 nm field, corresponding to the
lowest energy photons in the 13-photon ionization range. A truncated
part of the bound atomic levels, of angular momentum up to 4, is
shown, together with the quantum paths leading to 13-photon
ionization. Graphs of this type prove to be useful tools in the
qualitative analysis of the processes involved in multiphoton
phenomena.
In our study, we analyze the wavefunction at the end of the pulse, and
thus obtain information on the electron spectrum and ionization.
Depending on the physical quantity we are looking for, the parameters
needed to achieve convergence vary, making it easier to obtain the
value of angle and energy integrated ionization, than the ATI
spectrum. We have ensured the convergence of our results, by varying
the box size and the grid sampling density, in a way similar to what
has been presented in~\cite{cormier:1997:ati}, and by relying on
empirical findings such as the definition of E$_{\text{c}}$, or the
needed density of discretized continuum states per photon energy, to
guide our parameter choice. We have in addition conducted a
further independent
test of the numerics involved, by comparing the
two different methods, i.e.\ the expansion in terms of
box-eigenstates, or the
direct expansion of the radial part in a
$B$-Splines basis. A sample result, for a demanding quantity such as
the ATI spectrum in the 13-photon range, is shown in
figure~(\ref{fig:ATI:basis-bsp:compare:7e11}), where the results of
the two simulations at 3300 nm, $7 \times 10^{11}$ W/cm$^2$ and 110 fs
are displayed on top of each other. The direct $B$-Splines
method involves a box of 3000 a.u., with 3000 linearly sampled $B$-splines
of order 7, for each angular momentum up to $l=20$, parameters that
have proven more than sufficient for the method to converge in the
range shown. After the propagation is over, a projection to
scattering states is used to obtain the ATI spectrum. The eigenstates
expansion involves a box of 2500 a.u., with 2500 linearly sampled
$B$-Splines of order 9 for each angular momentum up to $l=15$ for
constructing the eigenbasis. This basis was then truncated to the lowest
1000 basis elements per angular momentum; using only the 1000 lowest
states proved to be sufficient for the energy range presented here,
since the higher energy discretized continuum states play a role
in the high energy, low-yield part of the spectrum. The
window-operator technique~\cite{kulander:1992:tds} was used to obtain
the photoelectron spectrum after the simulation, and the spectrum was
renormalized for the comparison with the $B$-Splines spectrum.
\begin{figure}
{\epsfxsize=12cm \epsfbox{ATI.basis-bsp.compare.7e11.eps}}
\caption{Two different, time-dependent methods of calculation of the
dynamics of the Potassium model potential, for a 3300 nm, $7
\times 10^{11}$ W/cm$^2$, 110 fs, $\cos^2$ pulse. The full line
is the ATI spectrum obtained through the $B$-splines expansion
method, and the dashed line is the ATI spectrum obtained through
the eigenbasis calculation.}
\label{fig:ATI:basis-bsp:compare:7e11}
\end{figure}
Despite the differences in the methods used, we observe an excellent
agreement of the results in
figure~(\ref{fig:ATI:basis-bsp:compare:7e11}), even on a logarithmic
scale. The $B$-Splines method shows richer structure at the minima
between the peaks, demonstrating its superiority at the finest parts
of the results. In detailed analysis of the ATI results (too long to
be
presented here), we have studied the behavior of the side peaks with
intensity and we have noted that different peaks show different shifts with
intensity, which helps us in classifying them as either Freeman
resonances~\cite{freeman:1987:ati} or Bardsley
fringes~\cite{bardsley:1988:u,cormier:1996:ogg}.
Note further the clean formation of a plateau in the
ATI-peak heights, showing up in both methods, and extending over the
first 5 to 6 peaks; we study this plateau later in the paper.
Since, in theory, the two methods are related by a unitary
transformation of the basis from a spatially localized ($B$-Splines)
to a field-free diagonal (eigenbasis) representation, the agreement of
the results is expected; however, practice has shown that convergence
is strongly
affected by the underlying numerics, especially when it comes to ATI
spectra that stress the subtle parts of our codes.
It is the first time we have used such an elaborate procedure to ascertain
the accuracy of our results.
All results presented from now on are from
calculations with the direct expansion onto $B$-splines, which is more
efficient both in computer space and time, when the scale of the
simulation increases. We have compared the results of both methods
for all intensities at 3300 nm (13-photon range) and some intensities
in the 12-photon range (2880 nm). The convergence of all other
calculations presented was ensured within the $B$-Splines propagation
method only, using well documented techniques~\cite{cormier:1997:ati}:
variation of box-size, $B$-Splines density and order, and variation of
the number of angular momenta.
\subsection{Behavior of ion yields as a function of intensity}
\begin{figure}
{\epsfxsize=12cm \epsfbox{yield.2880nm.fit.eps}}
\caption{The ion yield versus intensity, and a fit of the
low-intensity data to a power law. A 2880 nm, 96 fs,
cos $^2$ pulse is used. The dashed line shows the ADK-theory
prediction.}
\label{fig:yield:2880nm:fit}
\end{figure}
We begin the presentation of our results with a study of ion
yields, beginning with
the 12-photon
process at a 2880 nm pulse, of 96 fs temporal width, and
intensities ranging from $10^{11}$ W/cm$^2$ to $10^{13}$ W/cm$^2$. A
2000 a.u.\ box with 2000 linearly sampled $B$-Splines of order 7, for
each angular momentum up to $l=20$, proves sufficient for obtaining
the ion yield. In figure~(\ref{fig:yield:2880nm:fit}) we present the
results, together with a power-law fit to the low-intensity part of
the spectrum. In the same figure we also show the ionization yield
estimate obtained by the ADK-theory: in this and the following
figures where the ADK predictions are displayed, we have integrated
the ADK-rate over a square pulse of maximum intensity equal to the
FWHM of the pulse used.
Saturation sets in at about $3\times 10^{12}$ W/cm$^2$,
substantially higher than the scaled estimate of $10^{11}$ W/cm$^2$,
and in very good agreement with the ADK prediction.
The low-intensity spectrum seems
to follow a multiphoton perturbative behavior, in agreement with the
value of the tunneling parameter $\gamma$ in
table~(\ref{tab:parameter:all}), and thus the power-law; but the least
squares fit yields a slope of 7.5, substantially smaller than the
lowest-order perturbation theory expectation of 12. The same behavior
appears in the experimental results~\cite{sheehy:1998:ema}, where in
the 12-photon ionization curve, a slope of 7 has been measured. Note
that the ADK-theory, which is often used in comparisons to
experiments due to its simplicity, markedly fails to predict the
low-intensity yield.
\end{multicols}
\begin{figure}
{\epsfxsize=16cm \epsfbox{combine-K-H.eps}}
\caption{Potassium under 2880 nm radiation (left), and Hydrogen starting
from the 2s under 4090 nm radiation (right). A truncated part of
the atomic structure is shown, together with the quantum paths
leading to ionization.}
\label{fig:compare:K:H2s}
\end{figure}
\begin{multicols}{2}
A scaled system, namely the 12-photon ionization process in Hydrogen
starting from the metastable 2s state and interacting with 4090 nm
light, is used as a test for these results. These two different systems
are compared in figure~(\ref{fig:compare:K:H2s}), where we show bound
states of Potassium and Hydrogen for the 4 lowest angular momenta,
together with the 12-photon quantum paths leading to ionization, in
energy units scaled to the photon energy. The 2s state is chosen
instead of the ground state of Hydrogen as the initial state, so that
the ionization threshold energy and the distribution of the bound
states, other than the degeneracies, resemble those of Potassium.
\begin{figure}
{\epsfxsize=12cm \epsfbox{H2s-yield.4090nm.eps}}
\caption{Ion yield for Hydrogen starting from the 2s in the
12-photon regime, and a fit of
the low-intensity part to a power law. A 4090 nm, 136 fs, cos$^2$
pulse is used. The dashed line shows the ADK-theory prediction.}
\label{fig:H2s:yield:2880-scal}
\end{figure}
\vspace{1cm}
The resulting ion yield for a 136 fs pulse is shown in
figure~(\ref{fig:H2s:yield:2880-scal}). A 3500 a.u.\ box with 3500
linearly sampled $B$-Splines of order 7 for each of the 21 angular
momenta is used; this is certainly an overkill just for obtaining the
ionization spectrum, but we also analyzed the ATI spectra which need
such large boxes due to the long propagation times involved. We see
the same kind of structureless saturated spectrum; albeit this time
the yield is higher for the same intensities (notice the scales in the
figures), and the saturation intensity is smaller by a factor of 2, in
accordance with the scaling relations that predict a higher
cross-section for Hydrogen starting from the 2s.\modify{You can
mention Lars here}
The ADK theory (shown as the dashed line in the figure) departs at the
lower part of the spectrum, and predicts a smaller saturation intensity.
The low-intensity part again shows a linear
dependence in the log-log plot, and this time the slope is 5.7, even
less than what it is in Potassium. The slopes in both Potassium and
Hydrogen 2s, roughly equal the order needed to ionize from the first
exited state, the 4p and 3s or 3d respectively, yet, no unambiguous
model conforming to all of our data could be constructed. Working in
a related context, Pont et al.~\cite{pont:1990:lft} have
constructed a theory describing multiphoton ionization in a strong
field of low frequency $\omega$, obtaining an asymptotic expansion of
the ac quasienergy in powers of $\omega^2$. Their paper contains
results for the rate of ionization from the 1s state of hydrogen in a
circularly polarized 1064 nm field. When translated into a log-log
plot of the rate vs intensity, the rate seems to increase roughly like
the eighth power of the intensity, instead of the twelfth power as
should be expected if the rate was perturbative.
\begin{figure}
{\epsfxsize=12cm \epsfbox{yield.3300nm.eps}}
\caption{Ion yield versus intensity. A 3300 nm, 110 fs,
cos$^2$ pulse is used. The dashed line shows the ADK-theory
prediction.}
\label{fig:yield:3300nm}
\end{figure}
We move on to 13-photon ionization, where the spectrum shows an
unexpected feature. In figure~(\ref{fig:yield:3300nm}) we show the
ion yield versus intensity, using 3300 nm, 110 fs pulses,
calculated within a box of 3000
a.u.\ with 3000 $B$-Splines per angular momentum up to $l=20$; for
comparison we also display the predictions of the ADK theory. The
saturation intensity is similar to the one in the 12-photon case, and
again a power-law behavior of the signal with intensity holds for the
lowest intensities. We observe, however, a ``knee'' in
the ion-yield curve, for intensities around $8\times10^{11}$ W/cm$^2$.
This feature resembles a dynamic resonance, rather
unexpected given the very high order of the processes involved.
\end{multicols}
\begin{figure}
\begin{center}
{\epsfxsize=14cm \epsfbox{bounds.3300nm.7e11.s-f.eps}}
\end{center}
\caption{Bound state distribution at the end of a 3300 nm, 110 fs, cos$^2$
pulse. We present the population probability of bound states with
respect to their main quantum number, for the four lowest angular
momenta $l$. The lines have no physical meaning, and are only meant
as a guide to the eye.}
\label{fig:bound:dist}
\end{figure}
\begin{multicols}{2}
In figure~(\ref{fig:bound:dist}) we plot the distribution of the final
bound states for an intensity of $7\times10^{11}$ W/cm$^2$. We plot
the population probability, versus the principal quantum number for a
few of the lowest angular momenta. Most of the population is still in
the ground state at this intensity; most of the excited
population is concentrated in the $4p$ state, and this is true for all
lower intensities. Structure appears in the low-angular
momentum, low-excited states, then a smooth, flat part, and a
bump at the highest excitations. This bump turns out to be
artificial, created by the pseudostates lying between the true bound
states and the discretized continuum states. This was confirmed by
observing that by making the box smaller, and thus changing the number
of supported bound states, the same effect always appeared in the
region just before the continuum. The smooth region just before the
bump is expected, given the small energy differences of these states,
compared with the shifts experienced during the pulse.
\begin{figure}
{\epsfxsize=12cm \epsfbox{yield.scan.13.wl.eps}}
\caption{The ion yield versus wavelength is shown. A 20 cycles
cos$^2$ pulse is used. We use boxes of 2000 a.u., with
2000 linearly sampled $B$-Splines of order 7, for each of the 21
lowest angular momenta.}
\label{fig:yield:scan:13:wl}
\end{figure}
In order to shed more light on the behavior of the ``knee'', we
performed a series of calculations, for different wavelengths,
scanning the entire 13-photon range, from 3125 nm, to 3300 nm, with a
step of 25 nm, and with pulses of 20 optical cycles, whose widths
correspond to 104 fs for the 3125 nm wavelength, and to 110 fs for 3300 nm.
In figure~(\ref{fig:yield:scan:13:wl}), we plot the ion yield as a
function of the wavelength, for a series of intensities ranging from
$10^{11}$ W/cm$^2$, to $3\times10^{12}$ W/cm$^2$. The curves separate
in a natural way, since for a fixed wavelength a higher intensity
yields more ionization, and we note that for the highest intensities,
above about $10^{12}$ W/cm$^2$, all wavelengths result in practically
the same yield. In the graph we label only the intensities of
interest
that are in the range of 2--8$\times 10^{11}$ W/cm$^2$, where
we see a broad resonance shifting to higher wavelengths with
increasing intensity; this suggests a resonance shifting downwards by
increasing the intensity.
In figure~(\ref{fig:ion-exc:3150-3300}) we plot (for a few, selected
wavelengths) the ionization curves, together with the total
excitation, which is defined as the population in all bound states
other than the ground state at the end of the pulse. The excitation
is essentially dominated by the population of the 4p state, for most
of the low- to mid-intensity range. Note the smooth variation of the
curves with the wavelength; the results for the other wavelengths of
figure~(\ref{fig:yield:scan:13:wl}), essentially interpolate the ones
shown here. All curves exhibit a ``knee'' which is more pronounced
in the excitation spectrum; after that, ionization mimics excitation
in its behavior, whereas for the lower intensities --- see especially
3150nm and 3200 nm --- their behavior may differ substantially. One
can argue that the low intensity part is typical in the multiphoton
picture, where the difference in the orders of the processes involved
is expected to show up as a difference in excitation compared to
ionization. In a Floquet picture, few selected avoided crossings
would describe the bulk of the dynamics. As the intensity increases,
and higher excitations play a more important role, a regime is
reached, where ionization and excitation are linked, similar to the
transition to the classical chaos regime, which has been discussed in
the microwave context, for example in~\cite{blumel:1987:mih}.
\end{multicols}
\begin{figure}
{\epsfxsize=16cm \epsfbox{ion-exc.3150nm-3300nm.eps}}
\caption{The ion yield and the total excitation for (left to right,
and top to bottom) 3150, 3200, 3250, and 3300 nm radiation and a
20 cycles pulse. A box of 2000 a.u. is used.}
\label{fig:ion-exc:3150-3300}
\end{figure}
\begin{multicols}{2}
For all wavelengths, and all intensities below the ``knee'', the bound
state distribution is qualitatively similar to the one in
figure~(\ref{fig:bound:dist}), with the 4p being by far the most
populated excited state. On the basis of
figure~(\ref{fig:level-photon:3300nm}), this could happen assuming the
shifts of the 4s and 4p to be such that the two states are brought
closer together when the field is on. Indeed, by calculating the
lowest order shift in the presence of the field and at the wavelengths
of interest, it turns out that the 4s state shift is negative, and
relatively small, whereas the 4p state, repelled by 5s, shifts down by
more than three times as much. Thus, for the larger energy
photons in the 13-photon range, a resonant excitation of the
4p state during the pulse occurs; for the lower energy photons of the
same range, the shift pushes the states towards each other easing a
near-resonant transfer. The phenomenon of a low excitation playing an
essential role in a 13-photon ionization process, is easier to imagine
in an atom like Potassium, than in Hydrogen starting from the ground
state, since the excitation lies at less than half the energy that
is needed to reach ionization. It should be noted here that
13-photon ionization simulations in Hydrogen, starting from the 2s at
a scaled 13-photon wavelength of 4474 nm, display no corresponding
characteristic, which can be attributed to the difference in the lowest
excitation, as figure~(\ref{fig:compare:K:H2s}) shows.
\begin{figure}
{\epsfxsize=12cm \epsfbox{yield.3510nm.eps}}
\caption{The ion yield for a 14-photon ionization process. A 3510
nm, 117 fs, cos$^2$ pulse is used. The dashed line shows the ADK
theory prediction.}
\label{fig:yield:3510nm}
\end{figure}
We close our discussion of ionization curves, by presenting in
figure~(\ref{fig:yield:3510nm}) the results of 14-photon ionization
calculations, at a 3510 nm wavelength, with a 117 fs pulse. A 3000
a.u.\ box was used, with 3000 $B$-Splines of order 7 per angular
momentum, up to angular momentum $l=20$. We also present the ADK
theory predictions: the departure at the lower intensities is not so
dramatic as it was at the shorter wavelengths.
\subsection{Photoelectron energy spectra and ATI}
\label{sec:sub:photoelectron_ATI}
We move on to the presentation of the ATI spectra, which, as seen in
figure~(\ref{fig:ATI:basis-bsp:compare:7e11}), exhibit a clean plateau
in the low energy range. This plateau shows up at all wavelengths
we have checked, in Potassium as well as in the Hydrogen simulations
starting from the 2s,
and it may thus be considered a global feature for mid-infrared
wavelengths. In figure~(\ref{fig:ATI:plateau:3300nm:all}) a series of
ATI spectra at 3300 nm, 110 fs $\cos^2$ pulses, and for selected
intensities between $10^{11}$ W/cm$^2$ and $10^{12}$ W/cm$^2$ is shown; the
higher the intensity, the higher the signal shown in the figure. A
box of 3000 a.u., with 3000 $B$-splines of order 7 for each of the
angular momenta up to $l=20$ is used. The vertical axis in the figure
corresponds to the ionization probability density in units of 1/eV.
The shift of the
ATI peaks to lower energies with increasing intensity is linear with
intensity, and, as expected, is well described by the ponderomotive
shift of the ionization threshold. We note that the extent of the
plateau grows, almost in proportion to the intensity.
\begin{figure}
{\epsfxsize=12cm \epsfbox{ATI-many.3300nm.eps}}
\caption{ATI spectra for (bottom to top) 1, 2, 3, 4, 6, $9 \times
10^{11}$ W/cm$^2$.
A 110 fs, 3300 nm, cos$^2$ pulse is used.}
\label{fig:ATI:plateau:3300nm:all}
\end{figure}
\begin{figure}
{ \epsfxsize=12cm \epsfbox{ATI.plateau.peaks.3150nm.eps}}
\caption{The height of the first few ATI peaks as a function of
intensity. A 3150 nm, 100 fs, cos$^2$ pulse is used. The lines
have no physical meaning and are only used as a guide to the eye.
We notice the convergence of the peaks which indicates the plateau
formation.}
\label{fig:ATI:plateau:peaks}
\end{figure}
A clearer way of viewing this phenomenon is presented in
figure~(\ref{fig:ATI:plateau:peaks}). We plot the height of the first
few ATI peaks versus the intensity of the pulse, this time for a 3150
nm wavelength and a width of 105 fs, corresponding again to a 20-cycle
pulse. Note the power law (linear dependence on a log-log
plot) of the first few peak-heights with the intensity. This
indicates that a perturbative process is taking place. As the
intensity increases, more peaks enter the plateau range and the
perturbative picture ceases to hold. This is demonstrated in
figure~(\ref{fig:ATI:plateau:peaks}) by the
merging of the curves, which implies that for a range of peaks their
heights are basically the same. As the intensity increases, the
plateau expands and more curves merge.
\vspace{1cm}
\begin{figure}
{\epsfxsize=12cm \epsfbox{ATI.plateau-range.fit.3300nm.eps}}
\caption{ATI spectra plateau range as a function of intensity. A
3300 nm, 110 fs, cos$^2$ pulse is used. Hand-read definition of
plateau range.}
\label{fig:ATI:plateau:3300nm:fit}
\end{figure}
In figure~(\ref{fig:ATI:plateau:3300nm:fit}), we show the plateau
range plotted versus the intensity for the calculation shown in
figure~(\ref{fig:ATI:plateau:3300nm:all}). In this figure, the
plateau range is defined as the abscissa of the interception of a
horizontal line joining the first few peaks, and a line over the
decreasing peaks of the spectrum. The fit was made by hand, to
manually remove the effect of accidental resonances in the first peaks
of the spectrum. We notice the sharp linear dependence of the plateau
range on the intensity. On the same figure, we show a fit of the
selected points to a linear function of intensity. The least squares
fit gives a plateau range of $(2.3 \times 10^{-12} \pm 4 \times
10^{-14})$eV cm$^2$/W$\times I + (0.28 \pm 0.03)$eV. Measured in
units of
$U_p$ for the 3300 nm wavelength, the same equation reads: $(2.4 \pm
0.04) U_p + (0.28 \pm 0.03)$ eV. This shows a plateau scaling in
proportion to $2.4 U_p$.
\end{multicols}
\vspace{1cm}
\begin{figure}
{\epsfxsize=17cm \epsfbox{ATI.plateau-range.all.eps}}
\caption{ATI spectra plateau range for a wealth of intensities and
wavelengths. The horizontal axis shows the intensity in units of
W/cm$^2$; the vertical axis shows the extent of the plateau in
units of the ponderomotive energy. The large error-bars present
in the lowest intensities originate from the rigorous definition
of the plateau range described in the text.}
\label{fig:ATI:plateau:all}
\end{figure}
\begin{multicols}{2}
To clarify the relation of the plateau to the ponderomotive energy, we
combine all of the data from the simulations in the 12-, and 13-photon
range. We simplify the definition of the extent of the plateau to a
rigorous one, whose extraction from the data is automated, namely, we
use the abscissa of the first peak at which the exponential fall of
the peak heights begins. The results are shown in
figure~(\ref{fig:ATI:plateau:all}), where for each intensity the
plateau range measured in units of $U_p$ is shown for all wavelengths. The
error-bars reflect the imprecision due to cases in which the plateau
ends between two consecutive peaks. From the definition of the
plateau range that we use, it follows that the real plateau has the
same probability to be located anywhere within our error-bars. Notice
the huge uncertainty (larger than $U_p$) at small intensities which
reflects the fact that $U_p$ is smaller than the peak spacing. At
higher intensities, the ponderomotive shift grows until it is
considerably larger than the peak-spacing, thus shrinking the error-bars
to less than $U_p$. The data is fitted to a normal
distribution with an average value of 2.8 $U_p$ and a standard
deviation of 0.5 $U_p$.
This result relates to the classical theory of the
ionization process, as has been first developed in the simpleman's
model~\cite{heuvell:1988:lce,gallagher:1988:ati,corkum:1989:ati}. The
main arguments have recently been clearly restated in the appendix
of~\cite{lagattuta:1998:qmh}, although in the context of a
rescattering picture~\cite{walker:1996:ers}. The idea rests upon
a free-electron maximally gaining 3.17 $U_p$ within a
field, returning to its original position. This happens within the
first cycle, or the first period of the electronic motion, subsequent
cycles lead to 2.4 $U_p$, and then to less until the maxima gradually
converge to 2 $U_p$ over many cycles, and the average kinetic energy
becomes $U_p$. A relevant study has been made more than 10 years ago
by Gallagher in the microwave regime~\cite{gallagher:1988:ati},
where ionization of Na Rydberg
states was studied, and the spectra were explained using the above
mentioned theory. In that paper, a plot of the extend of the spectrum
starting from the 40 s state of Na, shows approximate scaling with
3.45 $U_p$. An energy of $U_p$ should be subtracted when compared to
short-pulse experiments, as the electrons do not keep the
extra ponderomotive energy since they cannot sample the spatial
gradient of the field. After the observation, in the optical regime,
of the long-range plateau extending from 2 to 8
$U_p$~\cite{paulus:1994:pat,hansch:1997:rhe}, the theory has been
extended to include backscattering from the nucleus, having thus
provided
answers to pertinent questions~\cite{walker:1996:ers,hu:1997:pat}. It
should be noted that other than the original Gallagher experiment, all
of these experiments and theories exhibit a steep fall in the region up
to $2U_p$, which then stabilizes or falls with a lower slope in the
region from 2 to 8 $U_p$. This is in contrast to our findings in the
mid-infrared regime that show a smooth, almost flat region in the
low energy spectra, and an increased downward slope afterwards. Due
to the numerical demands on convergence, in the present study we do
not analyze the region up to and beyond 8 $U_p$, and cannot therefore
establish the presence or not of a second plateau.
\begin{figure}
{\epsfxsize=12cm \epsfbox{H2s.ATI.1.2.4.7e11.eps}}
\caption{ATI spectra of Hydrogen starting from the 2s, in a 4090 nm
field. The spectra for 1, 2, 4, and $7\times10^{11}$ W/cm$^2$ are
shown. A box of 3500 a.u.\ with 3500 $B$-Splines per angular
momentum is used.}
\label{fig:H2s:ATI:1-2-4-7e11}
\end{figure}
We have also confirmed that the plateau characteristic has no relation to the
atomic structure involved, by examining the ATI spectrum in the scaled
problem, namely 12-photon ionization of Hydrogen starting from the 2s
state. The data are from the simulations corresponding to the
wavelength displayed in figure~(\ref{fig:compare:K:H2s}), and the
parameters used are the same as those used to plot
figure~(\ref{fig:H2s:yield:2880-scal}). Few selected ATI spectra are
shown in figure~(\ref{fig:H2s:ATI:1-2-4-7e11}), all of which display the
plateau, whose range is estimated as $2.6 \pm 0.3 U_p$, close
to what we obtained from the cumulative data in Potassium.
\section{Conclusions}
\label{sec:Conclusions}
In summary, we have conducted a theoretical study of the dynamics of
Potassium interacting with a high intensity, mid-infrared, short,
laser pulse.
We chose Potassium because it has
recently been studied experimentally by Sheehy et
al.~\cite{sheehy:1998:ema}. We have performed time dependent
calculations in terms of
both a direct expansion of the
Schr\"odinger equation onto $B$-Splines, as well as
an expansion onto
field-free eigenstates within a box, and obtained remarkable agreement
between the two methods.
We have studied a 12-photon ionization process, and have observed that
the low-intensity, power-law
behavior of the ion yield has an exponent much lower than the
perturbative expectation of 12.
We have obtained the same behavior from
the study of a scaled system, namely Hydrogen starting from the 2s in
the 12-photon range.
In the 13-photon ionization of Potassium, we noted a ``knee''
structure in the ion yield, and linked it to a similar, more
pronounced, behavior in the excitation.
We interpret this feature
as a broad dynamical resonance with the lowest excited state.
The ATI
spectra, in all cases that we have studied, display a clean formation
of a plateau in the first few peak-heights. Although the extension of
a plateau in the ATI spectrum at optical wavelengths from 2 to 8 times the
ponderomotive energy has been discussed in the literature, the
existence of a clear low-energy plateau has (to the best
of our knowledge) not been observed in other studies in the optical or
UV regime, but has been measured in the microwave regime and interpreted
by the simpleman's theory of ionization~\cite{gallagher:1988:ati}.
Through the analysis of the cumulative data from our simulations, we
have determined the extent of the plateau to scale with the
ponderomotive energy $U_p$ as $(2.8 \pm 0.5) U_p$, which is compatible
with the predictions of the classical theory.
Given the much longer wavelength and therefore longer optical period,
an initially launched wavepacket will have more time to spread before
it backscatters from the core. This raises the question as to whether
backscattering would play an important role at this wavelength. Recall
that backscattering for shorter wavelengths has been associated with
changes of the slope of the ATI spectrum up to around 10 $U_p$. A
related question of course is whether the initially launched wavepacket
might be narrower as a result of which their might not be substantial
additional spreading by the time it reaches the core. A definitive
evaluation of this aspect would require much more extensive
calculations which may be worth undertaking in the future.
\section{Acknowledgements}
We would like to thank Dr.~L.~F.~DiMauro for making available their
experimental results~\cite{sheehy:1998:ema} prior to publication.
Computer time, space and support from the Rechnenzentrum Garching, and
especially John Cox and Dr.~R.~Volk is gratefully acknowledged. One
of us (P.~M.) would like to thank Dr.~L.~A.~A.~Nikolopoulos for useful
discussions.
\section{Appendix}
In the appendix we discuss the model potential used, by comparing it
to the accepted structure of Potassium and to the alternative frozen
core Hartree-Fock model. The atomic structure of Potassium can be
found in several places. Since we cannot in a straightforward
manner incorporate relativistic effects in our theory, we use the
weighted energy levels, which are calculated as:
\begin{equation}
E_{\text{av}}
=
\frac{1}{\sum\limits_{J_i}(2 J_i + 1)}
\sum\limits_{J_i} (2 J_i + 1) E_{J_i}
\end{equation}
The atomic eigenenergies $E_{J_i}$ are taken
from~\cite{smith:1995:asl}. For easier comparison with the numerical
results, we measure the energies from the ionization threshold which
is at 35009.814 wavenumbers according to Sugar and
Corliss~\cite{sugar:1985:ael}.
The multiplet oscillator strengths are calculated using the
approximate formula~\cite{wiese:1969:atp}:
\begin{equation}
f_{ik}^{\text{multiplet}}
=
\frac{1}{\sum\limits_{J_i}(2 J_i + 1)}
\sum\limits_{J_i J_k}(2 J_i + 1) f(J_i, J_k)
\end{equation}
The data needed in this formula are taken from~\cite{smith:1995:asl}.
Note that in this database the quantity that is given is $\log((2 J_i
+1) f(J_i, J_k))$. The multiplet oscillator strengths are shown in
the second column of table~(\ref{tab:Potassium:fosc}). A few
oscillator strengths are also presented in page 300
of~\cite{sobelman:1992:asr}. They are in agreement with the ones
shown in the table.
\end{multicols}
\begin{table}
\caption{Potassium energy levels measured from first ionization
threshold. E$_{\text{Exp}}$ (cm$^{-1}$) is the weighted
nonrelativistic value, E$_{\text{HF}}$ is the Frozen Core
Hartree-Fock result, and E$_{\text{H}}$ is the Hellman Potential
value.}
\label{tab:Potassium:energies}
\begin{tabular}{dddd}
State & E$_{\text{Exp}}$ (cm$^{-1}$) &
E$_{\text{HF}}$ (cm$^{-1}$) & E$_{\text{H}}$ (cm$^{-1}$) \\ \hline
$4s$ & -35010 & -32253 & -38947 \\
$5s$ & -13983 & -13376 & -15761 \\
$6s$ & -7560 & -7326 & -8331 \\
$4p$ & -21986 & -20972 & -22647 \\
$5p$ & -10296 & -10000 & -10696 \\
$6p$ & -6005 & -5876 & -6209 \\
$3d$ & -13474 & -12755 & -11952 \\
$4d$ & -7612 & -7213 & -6735 \\
$5d$ & -4824 & -4600 & -4322 \\
$4f$ & -6882 & -6860 & -6852 \\
$5f$ & -4403 & -4390 & -4384 \\
$6f$ & -3057 & -3049 & -3045 \\
$5g$ & -4393 & -4390 & -4389 \\
\end{tabular}
\end{table}
\begin{multicols}{2}
An alternative approach to the model potential, that originally seemed
appealing, is the Frozen Core Hartree-Fock method. Its main merit is
its {\em ab initio} nature, and the comparatively better description
of atomic structure. Extensive theoretical discussion exists on the
Frozen Core Hartree-Fock method.
\modify{[NOTE: A short (one-sentence) historical note is appropriate]}
We used an implementation that is documented in~\cite{chang:1993:bsb}.
The work in that paper was concerned with
the application of
the method to configuration interaction on the Frozen Core Hartree
Fock basis, whereas here we are interested only in the single outer electron
case. Atomic structure is described quite well, as we see in the
third column of Table~(\ref{tab:Potassium:energies}) where we show
selected bound state energies, and in columns 3 and 4 of
Table~(\ref{tab:Potassium:fosc}), where we present bound-bound
oscillator strengths in the length and velocity gauge starting from s,
p, and d states respectively. The energies are measured from the
first ionization threshold and are expressed in wavenumbers for direct
comparison with the available data. A 500 a.u.\ box, with
500 $B$-Splines of order 9 was used for the calculations.
We notice that the agreement between oscillator strengths calculated
in the length and velocity gauges is quite good for the cases
displayed in the tables.
The major disadvantage of the method is the inconvenient scaling of the time
needed to calculate the basis, and the actual size of the basis
calculated, which limits us to small cases, up to 1000 a.u.. The
time needed by the frozen core Hartree Fock scales as $N^4$, although
in principle it is limited by a $N^3$ factor. The size scales with $N^2$
per angular momentum, so that for a $N=3000$ basis of 20
angular momenta, we need approximately 2 GB and 35 CPU days in our
workstation cluster to perform the structure calculations.
Thus it is mainly numerical considerations that dictate the use of
the pseudopotential method.
\end{multicols}
\begin{table}
\caption{Some Potassium multiplet oscillator strengths, starting from few
lowest $s$, $p$, $d$ states. f is the exact multiplet oscillator
strengths, f$_{\text{HF}}$ is the
Hartree-Fock result calculated in the length gauge,
and f$_{\text{H}}$ is the Hellmann potential result.}
\label{tab:Potassium:fosc}
\begin{tabular}{cdddd}
State & f & f$_{\text{HF}}$(len) &
f$_{\text{HF}}$(vel) & f$_{\text{H}}$\\ \hline
$4s \rightarrow 4p $ & 1.01 & 1.07 & 1.02 & 0.96 \\
$4s \rightarrow 5p $ & 0.009 & 0.01 & 0.008 & 0.026 \\
$4s \rightarrow 6p $ & 0.0009 & 0.0012 & 0.0008 &0.0055 \\
$5s \rightarrow 5p $ & 1.49 & 1.52 & 1.49 & 1.42 \\
$5s \rightarrow 6p $ & 0.031 & 0.026 & 0.024 & 0.075 \\
$6s \rightarrow 6p $ & 1.92 & 1.96 & 1.94 & 1.82 \\
$4p \rightarrow 5s $ & 0.18 & 0.18 & 0.17 & 0.19 \\
$4p \rightarrow 6s $ & 0.019 & 0.018 & 0.017 & 0.009 \\
$4p \rightarrow 3d $ & 0.89 & 0.93 & 0.97 & 0.90 \\
$4p \rightarrow 4d $ & 0.0003 & 0.012 & 0.015 & 0.092 \\
$4p \rightarrow 5d $ & 0.003 & 0.0002 & 0.0005 & 0.028 \\
$5p \rightarrow 6s $ & 0.31 & 0.32 & 0.31 &0.34 \\
$5p \rightarrow 4d $ & 1.20 & 1.25 & 1.29 & 1.00 \\
$5p \rightarrow 5d $ & 0.0077 & 0.032 & 0.036 & 0.14 \\
$5p \rightarrow 6d $ & 1.48$ \times 10^{-5}$ & 0.0037 & 0.0046 & 0.048 \\
$3d \rightarrow 5p $ & 0.14 & 0.16 & 0.18 & 0.14 \\
$3d \rightarrow 6p $ & 0.0066 & 0.0066 & 0.0078 & 6.75$ \times 10^{-9}$ \\
$3d \rightarrow 4f $ & 0.76 & 0.88 & 0.88 & 1.061 \\
$3d \rightarrow 5f $ & 0.17 & 0.16 & 0.16 & 0.14 \\
$3d \rightarrow 6f $ & 0.067 & 0.062 & 0.062 & 0.049 \\
$4d \rightarrow 6p $ & 0.30 & 0.34 & 0.35 & 0.25 \\
$4d \rightarrow 5f $ & 0.39 & 0.17 & 0.17 & 0.97 \\
$4d \rightarrow 6f $ & 0.14 & 0.62 & 0.62 & 0.18 \\
\end{tabular}
\end{table}
\begin{multicols}{2}
The last columns in tables~(\ref{tab:Potassium:energies}),
and~(\ref{tab:Potassium:fosc}) are the results of the Hellmann
pseudopotential, as presented in
equation~(\ref{eq:hellmann:definition}), calculated with a box of 2500
a.u., with 2500 $B$-Splines of order 9. Since the potential is
$l$-independent, length and velocity gauge oscillator strengths agree,
without the need of introducing a correction to the dipole
operator~\cite{norcross:1973:pc}. We calculated the standard
deviation of the length and velocity absorption oscillator strengths
starting from a specific state. For all bound states and up to the
lower part of the continuum spectrum, which is what interests us, this
was less than $10^{-5}$, which reassures us of the completeness of our
description. The model potential represents the atom less accurately
than the Hartree-Fock does, both energies and oscillator
strengths. Nevertheless, its excellent numerical properties, originating
from its simplicity, make the large calculations feasible.
Scaling techniques help to map the model atom results onto real
experiments, but our main aim is to study general features pertaining
in the mid-infrared range, thus making the actual atom used of
secondary importance.
\end{multicols}
|
\section{Introduction}
The work of recent years has given us an extensive understanding
of the entanglement of pure bipartite quantum states.
While there are
still many open questions about the entanglement of finite collections
of quantum states \cite{JP}, a rather complete understanding of
`asymptotic' entanglement, that of a large number of copies of a
quantum state, has emerged:
a pure state
is unentangled if and only if the state can be written in a product
form $\ket{\Psi}=\ket{\Psi_A}\ket{\Psi_B}$. The single good
quantitative measure of entanglement is
$E=S(\rho_A)=S({\rm Tr}_B\proj{\Psi})$, where $S$ is the von Neumann
entropy. And, a collection of bipartite pure states with total
entanglement $E$ can be reversibly interconverted into any other
collection of pure states with entanglement $E$ by purely local
operations.
However, despite much recent effort, we cannot claim to have such a
complete understanding of quantum entanglement for bipartite mixed
states. Its characterization has a remarkably greater complexity and
richness than the pure-state case: There is {\em not}, except in very
simple cases, an unambiguous way to say if a mixed state is entangled
or not. There is {\em no} single good quantitative measure of
mixed-state entanglement. And, it seems that entanglement is
irreversibly lost when one attempts to convert it from one mixed-state
embodiment to another.
Much of this difficulty can be traced to the basic fact
\cite{Sch,Houst} that there is no single way of viewing a mixed
quantum state as an ensemble of pure states. In fact, we know that
there are infinitely many such representations, and we have previously
noted that in general these pure-state ensembles exhibit entirely
different entanglement properties. For example, the completely mixed
state of two qubits is equally well described as an ensemble of
product basis states (no entanglement) or as an ensemble of the four
Bell states (all maximally entangled).
It seems that no single measure of entanglement for mixed states is
correct, but many different ones are useful depending on the
situation. The ensemble decomposition of a mixed state with the
maximal entanglement is useful in situations where the two parties
holding the state can be given aid by a third party to extract a pure
state with the greatest entanglement; we have termed the average
pure-state entanglement of this ensemble the ``entanglement of
assistance'' \cite{div:fuc:mab:smo:tha:uhl:98}. Another,
operational characterization of entanglement is the ``distillable
entanglement'' $D$ \cite{ben:ber:pop:sch:96,bdsw}, the average number
of maximally entangled singlet pairs that can be extracted from many
copies of the mixed state by local operations and classical
communication. Yet another way of quantifying entanglement related to
$D$ has been proposed \cite{ved:ple:rip:kni:97} in which the minimum
distance from the set of separable mixed states is taken as the
measure of entanglement.
Finally, the entanglement measure on which we focus in this paper is
the ``entanglement of formation'' \cite{ben:ber:pop:sch:96,bdsw}. It
is the average pure-state entanglement of the ensemble which has {\em
minimal} entanglement that describes the mixed state. Thus this is
dual, in an operational sense, to the entanglement of assistance. It
plays several other roles as well: it is converse, in some sense, to
the distillable entanglement, in that it gives the number of EPR
singlet pairs needed to create the mixed state by local operations.
It, like the measure of entanglement in
Ref. \cite{ved:ple:rip:kni:97}, provides bounds on the distillable
entanglement, and thus on other quantities such as the quantum
capacity of noisy channels that are of great current interest in
quantum information theory.
Thus, we believe that a complete characterization of the mathematical
properties of the entanglement of formation should be valuable in the
continued development of quantum information theory. In this paper we
give new results on one particular feature of the entanglement of
formation, the least number of states needed to make up a
minimal-entanglement ensemble of a mixed state. (In \cite{uhl} such
optimal decompositions of mixed quantum states have already been
considered, but with respect to a function related to, but different
from, the entanglement of formation.) Determining such optimal
decompositions gives information about the minimal-complexity
procedures for creating a mixed state from a supply of EPR singlets.
But besides the operational significance of our results, we believe
that the characterizations we provide are of greater significance on
account of the light they shed on the complexity and richness of the
mathematical structure of this important concept in quantum
information theory.
\section{Entanglement Basics}
Let $\rho$ be a density matrix on the bipartite Hilbert space ${\cal H}_A
\otimes {\cal H}_B$ and let ${\cal E}_{\rho}=\{p_i,\ket{\psi_i}\}$
with $p_i>0$ be an ensemble into which $\rho$ can be decomposed:
\begin{equation}
\rho=\sum_i p_i \ket{\psi_i} \bra{\psi_i}.
\end{equation}
The entanglement of formation \cite{bdsw} of $\rho$ is defined as \cite{foot1}
\begin{equation}
E(\rho)=\min_{{\cal E}_{\rho}}\sum_i p_i E(\ket{\psi_i}\bra{\psi_i}),
\label{eform}
\end{equation}
where
\begin{equation}
E(\ket{\psi}\bra{\psi})=S({\rm Tr}_A \ket{\psi}\bra{\psi})=S({\rm Tr}_B \ket{\psi}\bra{\psi}),
\end{equation}
where $S(.)$ is the von Neumann entropy:
\begin{equation}
S(\rho)=-\mbox{Tr}\,\rho \log \rho.
\end{equation}
The minimization in Eq. (\ref{eform}) makes an analytical computation
of the entanglement of formation of mixed states a nontrivial task.
Only in a bipartite Hilbert space $2 \otimes 2$ has
the problem of determining the entanglement of formation of any
density matrix been completely solved, in the work of Wootters
\cite{woot}. Uhlmann \cite{uhl} has shown that every bipartite
density matrix $\rho$ admits an optimal decomposition, that is, the
one that achieves the entanglement of formation $E$ of $\rho$
(Eq. (\ref{eform})), with at least $\rank{\rho}$ and at most
$\rank{\rho}^2$ different pure states, where $\rank{\rho}$ is the rank
of the density matrix. We call the number of different pure states in an
ensemble that forms a decomposition of a density matrix $\rho$ the
{\em cardinality} of the ensemble. We say that the {\em optimal ensemble
cardinality} of a separable state $\rho$, which we denote by
$\enscar{\rho}$, is $k$ if at least $k$ different pure states are
required for a separable decomposition of $\rho$. Since the number of
states must at least be sufficient to span the support of $\rho$,
\begin{equation}
\enscar{\rho} \geq \rank{\rho},
\label{excee}
\end{equation}
directly giving Uhlmann's lower bound. However, it has been a open
question whether there are states for which the optimal decomposition
has more than $\rank{\rho}$ different pure states. Note that the
states in the decomposition of a density matrix $\rho$ are always in
the range of $\rho$. This means that if $\enscar{\rho} > \rank{\rho}$
the states in the optimal decomposition will be linearly dependent.
In this paper we present two sets of examples of separable bipartite
states $\rho$ for which we prove that the cardinality of the optimal
decomposition of $\rho$ exceeds $\rank{\rho}$. These are the first
examples of such states. Both types of examples can be found in
principle in arbitrary high dimensions.
It is useful to classify states according to their behavior under
partial transposition. Let $\PT{B}{\rho}=({\bf 1}_A \otimes T) \rho$ where
$T$ is transposition of a matrix in a chosen basis. $\rho$ is
{\em positive under partial transposition} (PPT) if $\PT{B}{\rho}$ is a
density matrix, i.e., it has no negative eigenvalues. If $\rho$ is {\em negative under partial
transposition} (NPT) then $\PT{B}{\rho}$ has at least one negative
eigenvalue. It is known that for $2 \otimes 2$ and $ 2 \otimes 3$
systems, PPT is a necessary and sufficient condition for separability
\cite{horo1}. For a bipartite state in a Hilbert space of arbitrary
dimension, PPT is a necessary condition for separability
\cite{per:96}.
\section{Separable States At The Boundary}
In this section we show that if a separable state and its partial
transpose have unequal ranks, then one of them must have its optimal
ensemble cardinality greater than its rank. From this result we prove that
partial transposes of full-rank separable states that lie on the
boundary of the set of PPT states have optimal ensemble cardinality greater
than their ranks. Finally we give examples of such states for any $n
\otimes n$ system.
We start with the following straightforward observation:
\begin{lem}
Let $\rho$ be a separable state on ${\cal H}_A \otimes {\cal
H}_B$. Let $\PT{B}{\rho}=({\bf 1}_A \otimes T) \rho$ where $T$ is
transposition in a chosen basis. Then
\begin{equation}
\enscar{\rho}=\enscar{\PT{B}{\rho}}.
\end{equation}
\label{lemequal}
\end{lem}
{\em Proof\ } We prove that $\enscar{\PT{B}{\rho}} \leq \enscar{\rho}$
and $\enscar{\PT{B}{\rho}} \geq \enscar{\rho}$. Since $\rho$ is
separable, its optimal decomposition involves only product states:
\begin{equation}
\rho=\sum_{i=1}^{\enscar{\rho}} p_i \pproj{\psi_i}{\phi_i}.
\label{dec1}
\end{equation}
Then it follows that
\begin{equation}
\PT{B}{\rho}=\sum_{i=1}^{\enscar{\rho}} p_i \pproj{\psi_i}{\phi_i^*},
\label{dec2}
\end{equation}
and thus $\enscar{\PT{B}{\rho}}$ is at most $\enscar{\rho}$. By
performing the partial transpose again on $\PT{B}{\rho}$ we can prove
the inequality in the other direction.
$\Box$
We have seen that the optimal ensemble cardinality is invariant under
partial transposition. The rank of a density matrix $\rho$ is not
necessarily invariant under partial transposition. We can draw the
following conclusion:
\begin{theo}
Let $\rho$ be a separable state on ${\cal H}_A \otimes {\cal H}_B$. If
\begin{equation}
\rank{\rho} \neq \rank{\PT{B}{\rho}} \enspace,
\label{note}
\end{equation}
then either $\rho$ has the property that $\enscar{\rho} > \rank{\rho}$ or
$\PT{B}{\rho}$ has the property that $\enscar{\PT{B}{\rho}} > \rank{\PT{B}{\rho}}$.
\label{theorank}
\end{theo}
{\em Proof\ } This follows directly from Lemma \ref{lemequal} and
Eq. (\ref{excee}).
$\Box$
Where do we find separable states $\rho$ with the property
Eq. (\ref{note})? For this we look at full-rank separable states that
lie on the boundary of the set of PPT states. The following lemma,
illustrated by Fig. \ref{fig1}, looks into this:
\begin{lem}
Let $\rho$ be a separable state on ${\cal H}_A \otimes {\cal H}_B$ with
full rank, $\rank{\rho}=\dim {\cal H}_A \otimes {\cal H}_B$. If
$\rho$ lies on the boundary of the set of PPT states, then
\begin{equation}
\rank{\rho} > \rank{\PT{B}{\rho}}.
\end{equation}
\label{boundfull}
\end{lem}
{\em Proof\ }
The set of PPT states ${\cal S\dnn{PPT}}=\{\rho \mid \PT{B}{\rho} \ge 0 \}$
is a closed convex set. For the separable states $\rho$ on the boundary of
this set, the state $\PT{B}{\rho}$ has at least one eigenvalue which is zero,
as $\rho$ is arbitary close to entangled states $\rho_E$ for which
$\PT{B}{\rho_E}$ has at least one {\em negative} eigenvalue.
Thus $\PT{B}{\rho}$ does not have full rank, and for full-rank states $\rho$ this implies $\rank{\rho} > \rank{\PT{B}{\rho}}$. $\Box$
One can remark the following: In $2 \otimes 2$ and $2 \otimes 3$ all
entangled density matrices have the property that $\PT{B}{\rho}
\not\geq 0$ \cite{horo1}. Therefore any separable density matrix
$\rho$ that is on the boundary of the set of separable states and has
full rank, fulfills the conditions of Lemma \ref{boundfull}. With
Theorem \ref{theorank} it follows that the partial transposes of these
density matrices have the desired property, i.e.,
$\enscar{\PT{B}{\rho}} > \rank{\PT{B}{\rho}}$.
A final comment about the results of this section: Eqs. (\ref{dec1})
and (\ref{dec2}) show that for separable $\rho$, $\rho=\PT{B}{\rho}$
if all the states $\phi_i$ are real in some local basis, so obviously
the ranks of $\rho$ and $\PT{B}{\rho}$ are equal. Thus, any state
$\rho$ satisfying Lemma \ref{boundfull} must have complex state
vectors in any separable decomposition in any local basis. Note that
even real density matrices $\rho$ sometimes have optimal
decompositions which require complex vectors \cite{bdsw1}. Readers
may find it surprising that even for complex vectors, there exist sets
$\ket{\psi_i}\ket{\phi_i}$ for which the set
$\ket{\psi_i}\ket{\phi_i^*}$ spans a space of a different dimension;
but this is exactly the consequence of Lemma \ref{boundfull}.
\subsection{Examples}
\label{ex1}
The generalized Werner state \cite{horodistill} in $n \otimes n$ is defined
as
\begin{equation}
\rho\dnn{W}(f)={f}\proj{\Psi^+}+\frac{1-f}{n^2} {\bf 1}_{n^2} \enspace ,
\label{genwern}
\end{equation}
where $\ket{\Psi^+}=\frac{1}{\sqrt{n}}\sum_{i=0}^{n-1} \ket{ii}$ is
the maximally entangled state in $n \otimes n$ and $0 \leq f \leq 1$.
Let
\begin{equation}
\rho(f)=\PT{B}{\rho\dnn{W}(f)}$, on$\ {\cal H}_n\otimes{\cal H}_n.\ \
\end{equation}
It has
been shown by Horodecki and Horodecki \cite{horodistill} that the state
$\rho\dnn{W}(f)$ is separable for $0\le f\le 1/n$. On the other hand,
for $1/n<f\le 1$, $\rho(f)$ is not positive semidefinite. Therefore the
state $\rho\dnn{W}(1/n)$ lies at the PPT-NPT boundary. Upon inspection of
Eq. (\ref{genwern}) we see that the rank of $\rho\dnn{W}(1/n)$ is full,
$\rank{\rho\dnn{W}(1/n)}=n^2$. Thus we can use Theorem \ref{theorank} and
Lemma \ref{boundfull}.
It is easy to show by direct calculation that $\rho(1/n)$ has exactly
$n(n-1)/2$ zero eigenvalues and hence its rank is $n(n+1)/2$.
Therefore by Theorem \ref{theorank} and Lemma \ref{boundfull}, we find
that $\rho(1/n)$ has an optimal ensemble cardinality of at least $n^2$
whereas the $\rank{\rho(1/n)}=n(n+1)/2$. For this state, the ratio
$\frac{\enscar{\rho(1/n)}} {\rank{\rho(1/n)}}$ can be as large as 2
when $n$ tends to $\infty$. This ratio could be made higher if there
exists a $\rho$ in $n \otimes n$ for which $\PT{B}{\rho}$ can have
more than $n(n-1)/2$ eigenvalues; we have no indication that this is
possible. But for the state $\rho(1/n)^{\otimes k}$, $k$ tensor
copies of the above state, the ratio of ${\cal L}_{{\cal E}}$ to
${\cal R}$ can be made arbitrarily large (going to $2^k$ for
$n\rightarrow\infty$).
\section{Barely Completable Sets of Product States}
In \cite{upb1,upb2} the notions of an unextendible product
basis and an uncompletable product basis were introduced. A product
basis is a set of $k$ separable orthogonal pure states in $n\otimes
m$. Considering the case $k<nm$, the basis is unextendible if there
are no additional pure product states orthogonal to all the members of
the basis; it is uncompletable (in $n\otimes m$) if the number of such
additional states is less than $nm-k$. In \cite{upb1,upb2} it is shown
that the completely mixed state $\rho$ on the Hibert space
complementary to the space spanned by the unextendible product basis
is entangled, but has the property that ${\rho^{\rm T_B}}$ is positive
semidefinite. In the following, we will need the notion of a local
extension of a bipartite Hilbert space ${\cal H}={\cal H}_A \otimes
{\cal H}_B$: a local extension of ${\cal H}$ is a Hilbert space ${\cal
H}'=({\cal H}_A \oplus {\cal H}'_A) \otimes ({\cal H}_B \oplus {\cal
H}'_B)$.
In \cite{upb1} and \cite{upb2} the notions of an unextendible product
basis and an uncompletable product basis were introduced
\cite{expl}. It was shown how to construct, from an unextendible
product basis, a bipartite entangled state $\rho$ for which
$\PT{B}{\rho}$ is positive semidefinite. We will need the notion of a
local extension of a bipartite Hilbert space ${\cal H}={\cal H}_A
\otimes {\cal H}_B$: A local extension of ${\cal H}$ is a Hilbert
space ${\cal H}'=({\cal H}_A \oplus {\cal H}'_A) \otimes ({\cal H}_B
\oplus {\cal H}'_B)$.
\begin{theo}
Let $\{\ket{\alpha_i} \otimes \ket{\beta_i}\}_{i=1}^{|S|}$ be a partial
product basis $S$ in ${\cal H}={\cal H}_A \otimes {\cal H}_B$.
If $S$ is uncompletable in ${\cal H}$, but $S$ is completable in some
local extension ${\cal H}'$ of ${\cal H}$, then $\rho_S$ defined as
\begin{equation}
\rho_S=\frac{1}{\dim {\cal H}-|S|}\left({\bf 1}-\sum_{i=1}^{|S|}
\pproj{\alpha_i}{\beta_i}\right),
\end{equation}
has the property that
\begin{equation}
\enscar{\rho_S} > \rank{\rho_S}.
\end{equation}
\label{uncomprank}
\end{theo}
{\em Proof\ } As the set of states $S$ is completable in a local
extension of ${\cal H}$, the state $\rho_S$ is separable, by Lemma 2
of \cite{upb1}. The idea of this Lemma 2 is that the completion of the
set $\{\ket{\alpha_i} \otimes \ket{\beta_i}\}_{i=1}^{|S|}$ in ${\cal
H}'$ give rise to a separable state
\begin{equation}
\rho_S'=\frac{1}{\dim {\cal H'}-|S|}\left({\bf 1'}-\sum_{i=1}^{|S|}
\pproj{\alpha_i}{\beta_i}\right).
\end{equation}
But $\rho_S$ is obtained from $\rho_S'$ by local projections on to
${\cal H}_A$ and ${\cal H}_B$ and therefore $\rho_S$ is separable
as well. However, since $\rho_S$ is uncompletable in ${\cal
H}$, $\rho_S$ cannot be represented as an ensemble of
orthogonal product states of cardinality $\rank{\rho_S}$. Thus any optimal
decomposition of $\rho_S$ must use non-orthogonal product states. The
von Neumann entropy of $\rho_S$ is equal to $S(\rho_S)=\log
\rank{\rho_S}$ as $\rho_S$ is the identity on a space of dimension
$\rank{\rho_S}$. In order to achieve this entropy, the optimal
decomposition of $\rho_S$ has to use more than $\rank{\rho_S}$ product
states, or
\begin{equation}
\enscar{\rho_S} > \rank{\rho_S},
\end{equation}
because any density matrix $\rho$ which is a mixture of only $n$
non-orthogonal states has entropy strictly less than $\log n$
bits. $\Box$
\subsection{An Example}
In \cite{upb1} an example was given of a set of orthogonal product
states on $3 \otimes 4$ that is not completable in $3 \otimes 4$, but
is completable in $3 \otimes 5$. We reproduce the states here:
Consider the states $\vec{v}_i\otimes \vec{w}_i,\;\; i=0,\ldots,4$
with $\vec{v}_i$ defined as
\begin{equation}
\vec{v}_i=N (\cos{{ 2 \pi i \over 5}},\sin{{2 \pi i}\over 5},h),\;\;
i=0,\ldots,4,
\label{defP}
\end{equation}
with $h={1 \over 2} \sqrt{1+\sqrt{5}}$ and
$N=2/\sqrt{5+\sqrt{5}}$. The states $\vec{w}_j$ are defined as
\begin{eqnarray}
\nonumber\vec{w}_j&=N'(\sqrt{\cos(\pi/5)} \cos(2j\pi/5),\sqrt{\cos(\pi/5)} \sin(2j\pi/5),\\
&\sqrt{\cos(2\pi/5)}\cos(4j\pi/5),\sqrt{\cos(2\pi/5)}\sin(4j\pi/5)),
\end{eqnarray}
with normalization $N'=\sqrt{2/\sqrt{5}}$. Note that $\vec{w}_j^T
\vec{w}_{j+1}=0$ (addition mod $5$). One can show that this set,
albeit extendible on $3 \otimes 4$, is not completable: One can at
most add three vectors like $\vec{v}_0 \otimes
(\vec{w}_0,\vec{w}_1,\vec{w}_4)^{\perp}$, $\vec{v}_3 \otimes
(\vec{w}_2,\vec{w}_3,\vec{w}_4)^{\perp}$ and
$(\vec{v}_0,\vec{v}_3)^{\perp} \otimes
(\vec{w}_1,\vec{w}_2,\vec{w}_4)^{\perp}$. The completion of this set
in $3 \otimes 5$ is particularly simple, being given by the following
ten states:
\begin{equation}
\begin{array}{lr}
(\vec{v}_1,\vec{v}_4)^{\perp} \otimes \vec{x}_0, & \vec{v}_0 \otimes (\vec{w}_0^{\perp} \in \mbox{span}(\vec{x}_4,\vec{x}_1)), \\
(\vec{v}_0,\vec{v}_2)^{\perp} \otimes \vec{x}_1, & \vec{v}_1 \otimes (\vec{w}_1^{\perp} \in \mbox{span}(\vec{x}_0,\vec{x}_2)), \\
(\vec{v}_1,\vec{v}_3)^{\perp} \otimes \vec{x}_2, & \vec{v}_2 \otimes (\vec{w}_2^{\perp} \in \mbox{span}(\vec{x}_1,\vec{x}_3)), \\
(\vec{v}_2,\vec{v}_4)^{\perp} \otimes \vec{x}_3, & \vec{v}_3 \otimes (\vec{w}_3^{\perp} \in \mbox{span}(\vec{x}_2,\vec{x}_4)), \\
(\vec{v}_0,\vec{v}_3)^{\perp} \otimes \vec{x}_4, & \vec{v}_4 \otimes (\vec{w}_4^{\perp} \in \mbox{span}(\vec{x}_3,\vec{x}_0)).
\end{array}
\label{complPO}
\end{equation}
The state $\rho_S$ on $3 \otimes 4$ has rank seven, but the
separable decomposition consists of ten non-orthogonal states obtained
by projecting the orthogonal states of the completion,
Eq. (\ref{complPO}), back into the $3 \otimes 4$ Hilbert space. It
is not known whether there exists a separable decomposition with more
than seven but with fewer than ten states.
\section{Discussion}
The results presented here on the minimum cardinality of optimal ensembles
raises a large number of tantalizing questions; we would like to know
this cardinality for all possible mixed states. So far, our rigorous
results apply only to separable states. There is some empirical
evidence that for entangled mixed states as well (in fact, for states
arising in the theory of unextendible product bases), the optimal
ensembles can have a cardinality greater than the rank \cite{JSpc}. But we
have found no techniques for proving any results for inseparable mixed
states. We would also like to know whether there are cases for which
the Uhlmann upper bound of ${\cal R}(\rho)^2$ is attained. The states
shown above are still far from this; for the states at the end of
Sec. \ref{ex1} with $n=2$ and any $k$, ${\cal L}_{{\cal E}}={\cal
R}^{{\log 4}\over{\log 3}}$. Finally we note that all the rigorous
results we have pertain to cases where ${\cal L}_{{\cal E}}$, while
greater than ${\cal R}$, never exceeds the Hilbert space dimension.
Is there some reason that ${\cal L}_{{\cal E}} $ can never exceed this
dimension?
In conclusion, we have shown two different families of unentangled
mixed states for which the minimal number of pure states in an optimal
minimal-entanglement ensemble is provably greater than the rank of the
mixed state. In both cases the proofs are possible because the mixed
state is marginally separable, in the first case because the partial
transpose of the state has zero eigenvalues, and in the second because
the state is defined as the complement of a barely completable product
basis.
\section*{acknowledgments}
We are grateful to Peter Shor for the construction of the barely
uncompletable product-basis example used here. We thank Charles
Bennett, Richard Jozsa, John Smolin, and Armin Uhlmann
for helpful discussions. AVT
would like to thank David D. Awschalom for his invaluable support
without which it would have been impossible to work in this exciting
field. This work has been supported in part by the Army Research
Office under contract numbers DAAG55-98-C-0041 and
DAAG55-98-1-0366. AVT and BMT would like to thank IBM Research for
logistical support during their visits to the IBM Thomas J. Watson
Research Center.
|
\section{Introduction}
Since the first data taking in 1992, the HERA high energy $e p$ collider
has shown to be a powerful tool for the study of the strong interaction, in
particular to test
the domain of applicability and the relevance of several approximations of
perturbative QCD (pQCD) in the field of diffraction.
\subsection {Total Cross Section and Diffraction}
Two major experimental discoveries were made at HERA for the understanding
of strong interactions and of hadron structure.
First, the observation that, in the deep inelastic scattering (DIS) domain,
the $\gamma^* p$ cross section increases rapidly with energy.
This is attributed to an enhancement of the number of gluons in the proton,
the gluon structure function
$x \cdot G(Q^2,x)$ thus growing fast as $x$ decreases ($x$ is the Bjorken
scaling variable: $x = Q^2 / 2 p \cdot q$, where $p$ and $q$ are, respectively,
the proton and the intermediate photon four-momenta and $Q^2 = - q^2$;
the $\gamma^* p$ centre of mass energy $W$ is given by $W^2 = Q^2 / x - Q^2$).
This ``hard'' behaviour differs from that of the total and the elastic
hadron-hadron cross sections (closely related through the optical theorem),
which are characterised by a ``soft'' energy dependence.
In the framework of Regge theory~\cite{Collins}, elastic scattering
is attributed at high energy to the exchange between the incoming hadrons of
a colourless object, the pomeron ${I\!\!P}$.
The energy dependence of the total cross section is proportional to
$W^{2[1-\alpha_{{I\!\!P} }(t)]}$, where the pomeron trajectory $\alpha_{{I\!\!P}}(t)$
is parameterised as~\cite{DoLa,cudell_fit}
\begin{equation}
\alpha_{{I\!\!P}}(t) = \alpha_{{I\!\!P}}(0) + \alpha^\prime \cdot t
\simeq 1.08 + 0.25 \ t ,
\label{eq:soft_pom}
\end{equation}
$t$ being the square of the four-momentum transfer.
The second major discovery in DIS at HERA is the substantial contribution
($8 - 10 \% $)
of events formed of two hadronic subsystems separated by a large gap in
rapidity, devoid of hadronic activity~\cite{Derrick}.
This process is similar to diffractive scattering in hadron-hadron
interactions, where the incoming hadrons are excited without colour
exchange.
Diffraction thus forms an extension of elastic scattering and is dominated
at high energy by pomeron exchange with the ``soft'' behaviour of
eq.~(\ref{eq:soft_pom}).
The interesting feature at HERA was to observe diffraction as a leading twist
process in DIS.
\subsection {``Soft'' Vector Meson Production}
An important case of diffractive scattering is that of vector meson (VM)
production, in particular when the proton remains intact in the reaction:
$e + p \rightarrow e + p + VM$.
\begin{figure}[htb] \unitlength 1.0 cm
\vspace{-0.6cm}
\begin{center}
\epsfig{file=hz.vm.gp.lf.eps,width=8.0cm,height=8.0cm}
\vspace{-0.5cm}
\caption{Total photoproduction and VM photoproduction cross sections,
for fixed target and HERA experiments.
The solid curves are obtained from the ``soft'' pomeron
parameterisation ~\protect\cite{DoLa}, with a decreasing contribution of
reggeon exchange at low energy; the energy dependence
noted $W^{0.22}$ is obtained from eq.~(\ref{eq:soft_pom}), taking into
account the $t$ distribution of the events.
The ``hard'', $W^{0.8}$, dependence is also shown in the case of \mbox{$J/\psi$}\
photoproduction.}
\label{fig:Wdep}
\vspace{-0.2cm}
\end{center}
\end{figure}
In the vector meson dominance (VDM) approach, the $J^{PC} = 1^{--}$ photon
is modelled as the superposition of the lightest VM's ($\rho$, $\omega$,
$\phi$).
The total $\gamma p$ cross section is thus expected to present the
characteristic ``soft'' behaviour of hadron-hadron interactions.
The production of light VM's, which is directly related to
elastic scattering (with a differential absorption by the target proton of
some of the hadronic components of the photon) is also expected to
present a ``soft'' energy dependence.
The gross features of this interpretation are supported by a huge quantity
of data accumulated by fixed target experiments~\cite{bauer,CHIO,NMC,E665}.
At high energy, the HERA experiments have measured the total
cross section in photoproduction
($Q^2 \approx 0$)~\cite{z_sigmatot,h_sigmatot} and the cross section for
diffractive photoproduction of
$\rho$~\cite{zeus_phot,h_rho_phot,zeus_phott,zeus_phottt},
$\omega$~\cite{z_om_phot}, $\phi$~\cite{z_phi_phot}.
They exhibit the ``soft'' energy dependence described by
parameterisation (\ref{eq:soft_pom}), as shown on Fig.~\ref{fig:Wdep}
(at low energy, a contribution from reggeon exchange, decreasing with $W$, is
present for $\sigma_{tot}$, \mbox{$\rho$}\ and $\omega$).
The W dependence of \mbox{$\rho$}\ photoproduction, studied as a function of $t$,
has also allowed measuring the slope $\alpha^\prime$ of the pomeron
trajectory~\cite{z_rh_alphaprim}.
\begin{center}
\begin{figure}[htb] \unitlength 1.0 cm
\vspace{-0.7cm}
\begin{picture}(12.0,8.0)
\put(-0.1,0.7){\epsfig{file=z_hight_rho_traj_plots.ps,%
bbllx=20pt,bblly=154pt,bburx=523pt,bbury=667pt,%
width=6.0cm,height=6.0cm}}
\put(6.5,1.1){\epsfig{file=z_hight_rho_traj.ps,%
bbllx=20pt,bblly=154pt,bburx=523pt,bbury=667pt,%
width=5.2cm,height=4.8cm}}
\end{picture}
\vspace{-0.8cm}
\caption{a) $W$ dependence of \mbox{$\rho$}\ meson photoproduction, in several bins in
$t$~\protect\cite{z_rh_alphaprim}.
b) Slope of the pomeron trajectory for the reaction
$\gamma p \rightarrow \rho p$ as obtained from a); the
dotted curve represents the parameterisation of~\protect\cite{DoLa}.}
\label{fig:alphaprim}
\vspace{-0.6cm}
\end{figure}
\end{center}
\subsection {``Hard'' Vector Meson Production and QCD}
At HERA, the main interest is for the production of light VM's at high \mbox{$Q^2$}\
or high \mbox{$|t|$}, and for the production of heavy vector mesons.
This is because two far-reaching questions can be raised:
1. Is the ``soft'', hadron-like behaviour observed in light VM
photoproduction also observed in the presence of a ``hard'' scale: high \mbox{$Q^2$},
high \mbox{$|t|$}\ or large quark mass ($c$, $b$) ?
2. In the presence of a ``hard'' scale, what are the relevant assumptions
and approximations in pQCD calculations required to describe diffractive
VM production ? Can this shed light on the partonic nature of the pomeron ?
A large number of experimental studies have thus been performed at HERA to
investigate these questions.
Data have been collected, in the presence of the scales $Q^2$, $m_q$ and $t$,
on the production of
$\rho$, $\omega$, $\phi$, $\rho^\prime$, $J/\psi$, $\psi(2s)$ and $\Upsilon$
mesons, with studies of the differential $Q^2$, $W$ and $t$ distributions, of the
polarisation characteristics, of the cross section ratio between several
VM production and of the mass shape.
Only a small fraction of these results will be presented here.
They are largely based on results presented at the
29th Int. Conf. on HEP held at Vancouver, Canada, in July, 1998.
\begin{center}
\begin{figure}[htb] \unitlength 1.0 cm
\vspace{-2.5cm}
\begin{picture}(10.0,4.5)
\put(2.7,-0.6){\epsfig{file=fey_seq_tem.ps,%
bbllx=74pt,bblly=321pt,bburx=535pt,bbury=503pt,%
height=2.4cm,width=6.7cm}}
\put(6.0,1.6){\oval(0.67,0.65)}
\end{picture}
\vspace{-0.3cm}
\caption{Schematic representation of VM production in pQCD.}
\label{fig:factor}
\vspace{-0.5cm}
\end{figure}
\end{center}
A large number of theoretical papers based on pQCD has also
been published, presenting predictions for VM production under various
assumptions and approximations
(see e.g.~\cite{kopel,ryskin,Brodsky,fks,mrt,ivanov,royen}).
A general feature of these approaches is that, at high energy, the amplitude
is factorised in three contributions, characterised by very different time
scales (see Fig.~\ref{fig:factor}):
\begin{equation}
A = \Psi^*_{\gamma^* \rightarrow q \bar q}
\otimes M_{q \bar q + p \rightarrow q \bar q + p}
\otimes \Psi_{q \bar q \rightarrow V} .
\label{eq:factor}
\end{equation}
The first factor corresponds to the amplitude for a long distance fluctuation
of the photon into a $q \bar q$ pair.
The second factor describes the (short-time) scattering amplitude of this
hadronic state with the proton.
The exchange is generally modelled as a gluon pair (i.e. a colour singlet
system), with $M \propto |x \cdot G(K^2,x)|^2$, the square of the gluon
density in the proton.
The order of magnitude of the scale $K^2$ at which the gluon structure
function is probed is
$K^2 \simeq 1 / 4 \ (Q^2 + m_{V}^2 + \mbox{$|t|$})$,
since these three variables contribute to the ``resolution'' of the
process; the factor $1 / 4$
comes from the sharing of the momenta between the two quarks.
The third factor in eq.~(\ref{eq:factor}) accounts for the recombination of
the scattered hadronic state in the VM wave function.
However, as stressed e.g. in~\cite{fks}, theoretical calculations are
affected by significant uncertainties concerning the choice of the QCD scale,
of the gluon distribution and of the VM wave function, in particular the
effects of Fermi motion of the quarks within the meson.
\section{Differential Distributions}
\subsection{$W$ Dependence}
The most striking manifestation of pQCD features in VM production is
to be expected in the $W$ dependence of the cross section, since the latter
is related to the square of the gluon density in the proton, which increases
rapidly with $W$ in the presence of a hard scale.
\begin{center}
\begin{figure}[htb] \unitlength 1.0 cm
\begin{picture}(12.0,6.0)
\vspace{-0.8cm}
\put(0.0,0.0){\epsfig{file=hz.jpsi.fit.eps,%
width=6.0cm,height=6.0cm}}
\put(6.5,0.0){\epsfig{file=eps_q2_a0.eps,
width=5.7cm,height=5.0cm}}
\end{picture}
\vspace{-0.8cm}
\caption{a) $W$ dependence of \mbox{$J/\psi$}\ meson photo- and electroproduction;
the curves represent predictions of a pQCD model~\protect\cite{fks} for
different gluon distribution functions in the proton and
$m_c = 1.5$ \mbox{\rm GeV}~\protect\cite{h_jpsi_DIS}.
b) \mbox{$Q^2$}\ dependence of the intercept $\alpha_{I\!\!P}(0)$ for \mbox{$\rho$}\ meson
electroproduction~\protect\cite{h_rho_DIS};
the dashed lines represent the range of values for the ``soft''
pomeron intercept, as derived from fits to the total and elastic
hadron-hadron cross section measurements~\protect\cite{DoLa,cudell_fit}.}
\label{fig:Wdephard}
\end{figure}
\vspace{-0.5cm}
\end{center}
A ``hard'' behaviour is observed in photoproduction of \mbox{$J/\psi$}\
mesons~\cite{z_jpsi_phot,h_jpsi_phot,h_jpsi_photoprod}, as shown in
Figs.~\ref{fig:Wdep} and~\ref{fig:Wdephard}a.
When the $W$ dependence of the cross section is parameterised as
$\propto W^{\delta}$ (in a Regge approach,
$\delta = 4 \alpha_{I\!\!P}(\langle t \rangle)$),
one finds $\delta \simeq 0.8$ for \mbox{$J/\psi$}\ photoproduction.
The contrast is thus manifest with the ``soft'' behaviour of light VM
photoproduction, for which $\delta = 0.20 - 0.25$ (this value is in
agreement with the parameterisation of eq.(\ref{eq:soft_pom}), taking into
account the $t$ distribution).
A similar behaviour is observed for \mbox{$J/\psi$}\
electroproduction~\cite{h_jpsi_DIS,z_rho_jpsi_DIS}
(Fig.~\ref{fig:Wdephard}a).
The curves on this figure represent the predictions of a model
based on pQCD calculations~\cite{fks}, for different parameterisations of
the gluon distribution in the proton.
The agreement of these predictions with the data, especially as to the
shape of the distribution (the absolute values are
sensitive to the input charm quark mass), supports the modelisation of the
pomeron as a colour-singlet gluon pair.
For \mbox{$\rho$}\ and \mbox{$\phi$}\ meson electroproduction, the ``hard'' scale is related
to \mbox{$Q^2$}.
Although the precision of the data is still limited, an indication is
present of a steeper $W$ dependence of the $\gamma^* p$ cross section
as $Q^2$ increases for the \mbox{$\rho$}~\cite{z_rho_jpsi_DIS,h_rho_DIS}
and for the \mbox{$\phi$}~\cite{z_phi_DIS}.
This is shown for the \mbox{$\rho$}\ on Fig.~\ref{fig:Wdephard}b, where the pomeron
intercept $\alpha_{{I\!\!P}}(0)$ is plotted.
\subsection{$Q^2$ Dependence}
The cross section for \mbox{$\rho$}\ production in the DIS domain is presented
as a function of \mbox{$Q^2$}\ on Fig.~\ref{fig:Q2_b}a for the
ZEUS~\cite{z_rho_jpsi_DIS} and H1~\cite{h_rho_DIS} experiments, which
are in agreement.
The \mbox{$Q^2$}\ dependence is well parameterised in this domain as
${\rm d} \sigma / {\rm d} Q^2 \propto 1 / {(Q^2+m_V^2)}^n$,
with $n \simeq 2.28 \pm 0.06$ (combined value).
This behaviour is expected from pQCD calculations, which give for the
(dominant - see below) longitudinal cross section~\cite{Brodsky}:
$\sigma_L \propto [\alpha_s(Q^2) \cdot x G(Q^2,x)]^2 / Q^6$, when
taking into account the $Q^2$ dependence in $\alpha_s(Q^2)$ and in $x
G(Q^2,x)$, as well as other uncertainties affecting the
calculations~\cite{fks}.
Over the full measurement range, including photoproduction, the \mbox{$Q^2$}\
dependence of \mbox{$\rho$}\ cross section is best described by the QCD based
model of ref.~\cite{royen}.
For \mbox{$\phi$}\ production~\cite{z_phi_DIS}, a value similar to that for the \mbox{$\rho$}\ is
found.
For \mbox{$J/\psi$}\ production, the values $n = 2.24 \pm 0.19$~\cite{h_jpsi_DIS} and
$n = 1.58 \pm 0.25$~\cite{z_rho_jpsi_DIS} are obtained.
\subsection{$t$ Dependence}
For not too large \mbox{$|t|$}\ values,
the $t$
distribution of VM production can reasonably well be parameterised in the
exponential form ${\rm d} \sigma / {\rm d} t \propto e^{- b |t|}$.
In an optical model approach of diffraction, the slope parameter $b$ is
related to the convolution of the sizes of the interacting objects:
$b \propto R_p^2 + R_{q \bar q}^2$, with the proton radius $R_p$ giving a
contribution of the order of $4 - 5$~\mbox{${\rm GeV}^{-2}$}.
As observed in Fig.~\ref{fig:Q2_b}b, the slope $b$ for \mbox{$\rho$}\
production decreases when \mbox{$Q^2$}\ increases, in agreement with the decrease of
the transverse size of the virtual $q \bar q$ pair expected in pQCD
calculations.
For \mbox{$J/\psi$}\ photo- and electroproduction, a slope of the order of
$b \simeq 4-5$ \mbox{${\rm GeV}^{-2}$}\ is
measured~\cite{h_jpsi_phot,z_jpsi_phot,z_rho_jpsi_DIS,h_jpsi_DIS},
confirming the small size
of the \mbox{$J/\psi$}\ meson.
\begin{figure}[htbp] \unitlength 1.0 cm
\begin{center}
\vspace{-1.0cm}
\begin{picture}(12.0,6.0)
\put(0.0,0.0){\epsfig{file=fin_sect_q2_papier.eps,%
height=5.8cm,width=5.8cm}}
\put(6.0,0.1){\epsfig{file=b_q2_paper.eps,%
height=5.5cm,width=5.8cm}}
\end{picture}
\end{center}
\vspace{-0.5cm}
\caption{a)~\mbox{$Q^2$}\ dependence of \mbox{$\rho$}\ production cross
section~\protect\cite{z_rho_jpsi_DIS,h_jpsi_DIS}; the superimposed
curve is for $n = 2.24$.
b)~\mbox{$Q^2$}\ dependence of the slope parameter $b$ for elastic \mbox{$\rho$}\ production:
fixed target measurements of ref.~\protect\cite{CHIO,NMC,E665} and
HERA measurements of
ref.~\protect\cite{z_rho_jpsi_DIS,H1_rho_94,h_rho_DIS} . }
\label{fig:Q2_b}
\vspace{-0.3cm}
\end{figure}
Fig.~\ref{fig:Q2_b}b also suggests that, at low \mbox{$Q^2$}, the slope parameter
$b$ for \mbox{$\rho$}\ production increases from the fixed target to the HERA energy
range.
This behaviour, known as ``shrinkage'' and expected in Regge theory, is
related to the non-zero slope $\alpha^\prime$ of the ``soft'' pomeron
trajectory.
In contrast, no shrinkage is expected in a pQCD approach for asymptotically
high values of the QCD scale ($\alpha^\prime \approx 0$).
However, no significant measurement has been possible so far using the HERA
experiments only, neither for \mbox{$\rho$}\ production at high \mbox{$Q^2$}\ nor for \mbox{$J/\psi$}\
production, and the conclusions to be drawn from comparisons between
fixed target and HERA data remain
controversial~\cite{z_rh_alphaprim,h_shrinkage_jpsi}.
\section{Polarisation}
The measurement of VM decay angular distributions allows the
determination of the spin density matrix elements, which are related to the
helicity amplitude $T_{\lambda_{V} \lambda_{\gamma}}$, where
$\lambda_{V}$ and $\lambda_{\gamma}$ are the helicities of the VM
and of the photon, respectively~\cite{schilling-wolf}.
In the case of $s$-channel helicity conservation (\mbox{{\rm SCHC}}), the helicity of
the photon is retained by the VM and the matrix elements containing
helicity changing amplitudes ($\lambda_{V} \neq \lambda_{\gamma}$)
are thus zero.
Measurements of the full set of matrix elements have been performed
for \mbox{$\rho$}\ as a function of \mbox{$Q^2$}\ (Fig.~\ref{fig:matqsq}a), $W$ and
$t$~\cite{h_rho_DIS,z_ang}, and for $\phi$ mesons~\cite{z_ang}.
As is visible on Fig.~\ref{fig:matqsq}, the data are compatible with \mbox{{\rm SCHC}},
except for a small but significant deviation from zero of the
matrix element \mbox{$r_{00}^{5}$}.
The helicity flip amplitude $T_{\lambda_{\rho} \lambda_{\gamma}} = T_{01}$
is thus determined to be $8 \pm 3 \%$ of the non-flip amplitudes
$\sqrt {T_{00}^2 + T_{11}^2}$.
This value is of the order of magnitude of that found at lower energy and
lower \mbox{$Q^2$}~\cite{CHIO,joos}.
Neglecting the small violation of \mbox{{\rm SCHC}}\ (which would affect the value of $R$
by $2.5 \pm 1.5 \%$), the matrix element \mbox{$r_{00}^{04}$}\
can be used to extract the ratio $R$ of cross sections for
\mbox{$\rho$}\ production by longitudinal and transverse virtual photons:
$R = \sigma_L / \sigma_T = \mbox{$r_{00}^{04}$} / \varepsilon \cdot (1-\mbox{$r_{00}^{04}$})$,
where $\varepsilon$ is the polarisation parameter
($\langle \varepsilon \rangle = 0.99$ at HERA).
Fig.~\ref{fig:matqsq}b shows that R rises steeply at small \mbox{$Q^2$}, and that
the longitudinal $\gamma^*p$ cross section dominates over the transverse
cross section for \mbox{$Q^2$}\ $\mathrel{\rlap{\lower4pt\hbox{\hskip1pt$\sim$}$ 2~\mbox{${\rm GeV}^2$}.
However, the rise is non-linear, with a weakening dependence at large \mbox{$Q^2$}\
values, and $R$ is $\approx$ 3 for \mbox{$Q^2$}\ $\mathrel{\rlap{\lower4pt\hbox{\hskip1pt$\sim$} 10-20$~\mbox{${\rm GeV}^2$}.
This feature is not reproduced by numerous models based on VDM or QCD, which
predict a linear increase of $R$ with \mbox{$Q^2$}.
However, the model of ref.~\cite{royen}, based on QCD, gives a good
description of $R$ over the full \mbox{$Q^2$}\ range, as does also a model based
on GVDM~\cite{sss}.
Another model based on QCD~\cite{mrt} predicts a moderate increase of $R$
with \mbox{$Q^2$}\ in the DIS domain.
\begin{center}
\vspace*{1.9cm}
\begin{figure}[htbp] \unitlength 1.0cm
\begin{picture}(0.0,0.0)
\put(-0.5,0.0){\epsfig{file=graph_mat_ele_qeda.ps,%
bbllx=55pt,bblly=189pt,bburx=450pt,bbury=693pt,%
height=8.5cm,width=6.5cm}}
\put(0.2,7.85){\large \mbox{$r_{00}^{04}$}}
\put(2.1,7.85){\large {\rm Re} \mbox{$r_{10}^{04}$}}
\put(4.0,7.85){\large \mbox{$r_{1-1}^{04}$}}
\put(5.9,7.85){\large \mbox{$r_{00}^{1}$}}
\put(0.2,5.82){\large \mbox{$r_{11}^{1}$}}
\put(2.1,5.82){\large {\rm Re} \mbox{$r_{10}^{1}$}}
\put(4.0,5.82){\large \mbox{$r_{1-1}^{1}$}}
\put(5.9,5.82){\large {\rm Im} \mbox{$r_{10}^{2}$}}
\put(0.2,3.8){\large {\rm Im} \mbox{$r_{1-1}^{2}$}}
\put(2.4,3.8){\large \mbox{$r_{00}^{5}$}}
\put(4.0,3.8){\large \mbox{$r_{11}^{5}$}}
\put(5.9,3.8){\large {\rm Re} \mbox{$r_{10}^{5}$}}
\put(0.2,1.8){\large \mbox{$r_{1-1}^{5}$}}
\put(2.1,1.8){\large {\rm Im} \mbox{$r_{10}^{6}$}}
\put(4.0,1.8){\large {\rm Im} \mbox{$r_{1-1}^{6}$}}
\put(5.7,1.2){\Large H1}
\put(5.7,0.5){\large \mbox{$Q^2$}~[\mbox{${\rm GeV}^2$}] }
\end{picture}
\put(7.3,1.0){\epsfig{file=r_q2_paper.eps,%
height=5.0cm,width=5.cm}}
\vspace*{-0.4cm}
\caption{a) Spin density matrix elements for elastic electroproduction
of \mbox{$\rho$}\ mesons as a function of \mbox{$Q^2$}~\protect\cite{h_rho_DIS};
the dashed lines indicate the expected null values in the case of SCHC.
b) The ratio $R$ of cross sections for elastic \mbox{$\rho$}\ meson electroproduction
by longitudinal and transverse photons, as a function of \mbox{$Q^2$}.}
\label{fig:matqsq}
\end{figure}
\vspace*{-0.5cm}
\end{center}
It is also found~\cite{h_rho_DIS} that the longitudinal and transverse
amplitudes are nearly in phase ($\cos \delta = 0.93 \pm 0.03$), assuming
\mbox{{\rm SCHC}}\ and natural parity exchange. This is similar to lower energy
measurements~\cite{CHIO,joos,delpapa}
A QCD based calculation~\cite{ivanov} predicts for the amplitudes the
hierarchy
\begin{equation}
|T_{\lambda_{\rho} \lambda_{\gamma}}| =
|T_{00}| > |T_{11}| > |T_{01}| > |T_{10}| > |T_{1-1}|,
\label{eq:hierarchy}
\end{equation}
which is supported by the measurement of the matrix elements, and also
the magnitude of the element \mbox{$r_{00}^{5}$}~\cite{h_rho_DIS,z_ang}.
Values of the matrix elements close to those for the \mbox{$\rho$}\ are obtained
for \mbox{$\phi$}\ mesons~\cite{z_ang}.
For \mbox{$J/\psi$}, the ratio $R$ of cross sections increases from values compatible
with zero in photoproduction~\cite{h_jpsi_phot,z_jpsi_phot}
to $\approx 0.4$ for $\langle Q^2\rangle \simeq 4$
\mbox{${\rm GeV}^2$}~\cite{z_rho_jpsi_DIS,h_jpsi_DIS}; this is smaller than
for \mbox{$\rho$}\ production at the same $\langle Q^2\rangle$, but is of the same
order if compared at the same value of $Q^2 / m_V^2$.
\section{Other Features}
\subsection{VM Production Ratio}
Predictions are obtained in pQCD for the cross section ratio of different VM
production~\cite{kopel,fks}.
As apparent in eq. (\ref{eq:factor}), this ratio is determined by the photon
coupling to the $q \bar q$ pairs, i.e. the charge of the quarks in the VM's,
and the effects of the wave functions.
For $\phi / \rho$~\cite{z_phi_phot,z_phi_DIS,H1_phi,H1_rho_95}, the
ratio increases with \mbox{$Q^2$}\ towards the value $2 / 9$
obtained from quark counting (see Fig.~\ref{fig:VM_ratio}a).
For $\psi / \rho$, the ratio is about a factor $1/200$ in photoproduction
in the HERA energy range, but flavour symmetry is restored within a factor 2
for \mbox{$Q^2$}\ above 10 \mbox{${\rm GeV}^2$}~\cite{z_rho_jpsi_DIS,H1_rho_94}.
\begin{center}
\begin{figure}[htbp] \unitlength 1.0 cm
\vspace{-0.8cm}
\begin{picture}(12.0,6.0)
\put(-0.3,0.0){\epsfig{file=z_phi_rho.eps,%
width=6.5cm,height=6.0cm}}
\put(6.2,-0.2){\epsfig{file=h_psi2s_psi.eps,%
width=6.5cm,height=6.0cm}}
\end{picture}
\vspace{-0.7cm}
\caption{Ratio $R$ of cross sections for a) $\phi$ and \mbox{$\rho$} ;
b) $\psi(2s)$ and \mbox{$J/\psi$}\ meson production, as a function of \mbox{$Q^2$}.}
\label{fig:VM_ratio}
\end{figure}
\vspace{-0.7cm}
\end{center}
The case of the $\psi(2s) / \psi$ ratio illustrates the interesting
phenomenon of the ``scanning'' of the VM wave function as \mbox{$Q^2$}\ varies.
Because of the node in the $\psi(2s)$ wave function, which induces
approximately cancelling contributions in the production amplitude, the
photoproduction of $\psi(2s)$ mesons is small.
As \mbox{$Q^2$}\ increases, the transverse size of the $q \bar q$ pair decreases,
thus avoiding the cancellation effect.
The resulting increase with \mbox{$Q^2$}\ of the cross section ratio is illustrated
in Fig.~\ref{fig:VM_ratio}b~\cite{h_jpsi_DIS}, the asymptotic limit being
computed to be of the order of 0.5~\cite{kopel,fks}.
\subsection{Mass Distribution}
For \mbox{$\rho$}\ photoproduction, the ($\pi,\pi$) mass distribution is
distorted with respect to a (relativistic) Breit-Wigner distribution,
with an excess of events at small masses and a deficit at large masses.
This phenomenon, known as ``skewing'', is attributed to the interference
between resonant \mbox{$\rho$}\ production and non-resonant pion pair production, the
interference changing sign at the resonance pole~\cite{soding}.
The skewing is observed to decrease in photoproduction as \mbox{$|t|$}\
increases~\cite{zeus_phottt} (see Fig.~\ref{fig:skewing}a).
The skewing also decreases with increasing
\mbox{$Q^2$}\ as seen in Fig.~\ref{fig:skewing}b for two different
parameterisations~\cite{soding,ross_sto}.
\begin{center}
\begin{figure}[htbp] \unitlength 1.0 cm
\vspace{-0.8cm}
\begin{picture}(12.0,6.0)
\put(0.0,-0.2){\epsfig{file=z_rho_skew_t_plots.eps,%
width=7.0cm,height=6.0cm}}
\put(6.3,-0.2){\epsfig{file=skew_paper.eps,%
width=5.7cm,height=5.6cm}}
\end{picture}
\vspace{-0.5cm}
\caption{a) Mass distribution $M_{\pi \pi}$ for \mbox{$\rho$}\ photoproduction,
for different values of \mbox{$|t|$}~\protect\cite{zeus_phottt}; the dashed curves
represent the Breit-Wigner distribution, the dotted curves the interference
with non-resonant pion pair, and the full curves the sum.
b) Two parameterisations~\protect\cite{soding,ross_sto} of the skewing of
the $M_{\pi \pi}$ mass distribution for \mbox{$\rho$}\ production, as a function of
\mbox{$Q^2$}~\protect\cite{E665,zeus_phot,zeus_phottt,h_rho_phot,z_rho_jpsi_DIS,h_rho_DIS}.}
\label{fig:skewing}
\vspace{-1.0cm}
\end{figure}
\end{center}
\section{Conclusions}
Abundant data have been collected at HERA on diffractive production of
light and heavy vector mesons, in the presence of the scales $Q^2$, $m_q$
and $t$.
A strong energy dependence of the cross section is observed for \mbox{$J/\psi$}\
production; an indication is found for a similar behaviour for \mbox{$\rho$}\
mesons at high \mbox{$Q^2$}.
In the light of perturbative QCD, with the pomeron modelled as a gluon pair,
these features are interpreted as reflecting the strong increases of the
gluon distribution in the proton at high energy, and quantitative agreement is
reached for \mbox{$J/\psi$}\ production.
The \mbox{$Q^2$}\ dependence of VM production is also qualitatively explained in pQCD
approaches.
The ratio of the longitudinal to transverse photon cross sections for \mbox{$\rho$}\
production increases rapidly with \mbox{$Q^2$}, but this increase is non-linear for
$Q^2 \mathrel{\rlap{\lower4pt\hbox{\hskip1pt$\sim$} 2$ \mbox{${\rm GeV}^2$}.
This behaviour has been reproduced recently by a model based on QCD.
More generally, the full set of \mbox{$\rho$}\ meson spin density matrix elements has
been measured.
The correct hierarchy between scattering amplitudes and the magnitude
of the dominant helicity-flip amplitude are also qualitatively reproduced
in a QCD approach.
In summary, great progress has been made in the understanding of VM
production at high energy when a hard scale is present ($m_c$, \mbox{$Q^2$}).
This contributes significantly to the understanding of
diffraction in a QCD framework.
\section*{Acknowledgements}
It is a pleasure to thank the organisors for a pleasant and fruitful
Symposium, and my colleagues in H1 and ZEUS, in particular B. Clerbaux and
P. Newmann, for numerous interesting discussions on diffraction.
\section*{References}
|
\section{Introduction}
\label{sec:intro}
The ongoing LEP run at energies above the
$W^+W^-$ threshold has made it possible to study directly the
non-Abelian structure of the electroweak Standard Model (SM) in the
clean environment of $e^+e^-$ collisions.
LEP2 has not only confirmed the existence
of triple gauge-boson vertices, inferred either
indirectly~\cite{indirect} or from observations at $p\bar{p}$
colliders~\cite{tev}, but also established constraints on them~\cite{LEP}.
The goal of the present studies at LEP2, and at future
colliders, is to test the structure of the bosonic sector with
a precision comparable to that achieved for the fermion-vector boson
couplings. Such precise measurements of gauge vector-boson
couplings will not only provide stringent tests of the gauge structure
of the SM, but also probe for new physics.
Within the Standard Model, triple and quartic vector-boson
interactions are intimately related to the gauge structure of the
model and therefore are completely determined. Of course, radiative
corrections within the SM modify the tree-level couplings. However,
such corrections are quite small. In particular for the $CP$-violating
couplings they are expected to be exteremely small and unmeasurable in
near future since $e.g.$ the electric dipole moment of the $W$ boson
vanishes to two loops\cite{KP}. On the other hand, any theory
incorporating new physics may conceivably induce much larger (already
at the one-loop) deviations in some of the couplings. Corrections at a
permille level can be expected in multi-Higgs or supersymmetric
extensions\cite{hisusy}. Models with dynamical breaking of electroweak
symmetry by new strong forces, could produce even larger corrections
\cite{dynmodel}. The concomitant $CP$ violation could then manifest
itself in non-zero $CP$-violating gauge boson couplings, observation of
which would be a clear signal of beyond-SM physics.
Owing to the general perception that the $CP$-violating couplings are
severely constrained by the data on the neutron electric dipole moment
(EDM)~\cite{edmn}, in phenomenological analyses of the physics
potential of future colliders, their sensitivity to these couplings
has received less attention than that accorded to the $CP$-conserving
ones. However, with these constraints being the subject to naturalness
assumption, there is no substitute for a direct measurement.
Furthermore, since they depend on different combinations of anomalous
couplings, the direct measurements are complementary to the indirect
analyses. In the present paper we wish to study the sensitivity of
future $e^+e^-$ linear colliders (LC) to $CP$-violating couplings, in
particular we will consider the process $e^+e^-\rightarrow \nu\bar{\nu}
\gamma$ as a means to test $WW\gamma$ interactions. The motivation for
our study, of course, is that the origin of $CP$ violation remains
unexplained and it should be experimentally investigated wherever
possible. We find that at an $e^+e^-$ 500 GeV linear collider with an
integrated luminosity of ${\cal L}=125 \: {\rm fb}^{-1}$, an analysis of the
differential cross-section allows us to derive the following 95\% C.L.
limits $|\tilde{\kappa}_{\gamma}|<0.18$,
$|\tilde{\lambda}_{\gamma}|<0.069$. With a high luminosity option of
500 fb$^{-1}$, more stringent limits $|\tilde{\kappa}_{\gamma}|<0.13$,
$|\tilde{\lambda}_{\gamma}|<0.049$ can be established. For comparison,
we find that the LEP2 experiments at $\sqrt{s}=192$ GeV and ${\cal L}=2
\: {\rm fb}^{-1}$ can reach a sensitivity of the order
$|\tilde{\kappa}_{\gamma}| \sim |\tilde{\lambda}_{\gamma}|\sim 2$
\section{Anomalous Couplings}
\label{sec:anom}
It is convenient to describe the phenomenology at scales well below
the scale of new physics in a model independent way. It is, by now,
standard to introduce an effective low-energy Lagrangian that contains
only SM fields. This assumes that the physics responsible for any
deviations is not directly observable, but can manifest itself
through virtual corrections. This formalism provides a simple
parametrization of the triple gauge-boson couplings (TGC). In purely
phenomenological terms, the effective Lagrangian for the $WW\gamma$
and the $WWZ$ interactions can be expressed in terms of seven
parameters each~\cite{hpzh} {\em viz.}
\begin{equation}
\begin{array}{rl} \displaystyle
{\cal L}_{\it eff}^{WWV} = &
\displaystyle -i g_{V} \Bigg[ \hspace*{0.3em}
( 1 + \Delta g^V_1 )
\left( W^\dagger_{\mu \nu} W^\mu
- W^{\dagger\mu} W_{\mu \nu}
\right) V^\nu +
( 1 + \Delta \kappa_V)
W^\dagger_{\mu} W_\nu V^{\mu\nu}\\[1.5ex]
& \hspace*{2.3em} \displaystyle
+ \frac{\lambda_V}{M_W^2}
W^\dagger_{\mu \nu} {W^\nu}_\sigma
V^{\sigma\mu} -i g^V_5
\epsilon^{\mu\nu\rho\sigma}(W^\dagger_\mu
\partial_{\rho} W_{\nu}- W_{\nu}\partial_{\rho}W^\dagger_{\mu})
V_\sigma
\\[1.5ex]
&\hspace*{2.3em} \displaystyle
+
i g_{4}^{V}W_{\mu}^{\dagger}W_{\nu}(\partial^{\mu}V^{\nu}
+\partial^{\nu}V^{\mu})
+ \tilde{\kappa}_{V}W_{\mu}^{\dagger}W_{\nu}\tilde{V}^{\mu\nu}+
\frac{\tilde{\lambda}_{V}}{m_{W}^2}W_{\lambda\mu}^{\dagger}W_{\ \
\nu}^{\mu}\tilde{V}^{\nu\lambda}\;\Bigg{]}\:,
\end{array}
\label{lagrangian}
\end{equation}
where $V^{\mu}$ is a neutral vector boson field, i.e. either the $\gamma$
or the $Z$ field. The $W^{\mu}$ ($W^{\mu\dagger}$) stands for the
$W^{-}$($W^{+}$) field, respectively, and
$V_{\mu\nu}=\partial_{\mu}V_{\nu}-\partial_{\nu}V_{\mu}$,
$W_{\mu\nu}=\partial_{\mu}W_{\nu}-\partial_{\nu}W_{\mu}$,
$\tilde{V}_{\mu\nu}=\frac{1}{2}\varepsilon_{\mu\nu\rho\sigma}V^{\rho\sigma}$.
The overall normalization is such that the coupling $g_{V}$
is defined as in the SM, $viz.$,
\begin{equation}
g_\gamma = e, \qquad g_Z = e \cot \theta_W \ ,
\end{equation}
with $\theta_W$ the weak mixing angle.
In the SM we have, at the tree level,
\begin{equation}
\Delta g_1^V = \Delta\kappa_V = \lambda_V = \tilde{\kappa}_V
= \tilde{\lambda}_V = g_4^V = g_5^V = 0 \
\end{equation}
Non zero values of the above, usually called anomalous gauge
couplings, would indicate new physics.
The three couplings, $\Delta g_1^V$, $ \Delta\kappa_V $ and $\lambda_V$,
are even under both $C$ and $P$ transformations. Of the remaining four,
two $\tilde{\kappa}_V $ and $ \tilde{\lambda}_V$ violate $P$ but conserve $C$, $g_4^V$ respects $P$
but violates $C$, and $g_5^V$ violates both $P$ and $C$.
Eq.~(\ref{lagrangian}) represents the most general $WWV$ Lagrangian
consistent with Lorentz- and gauge-invariance.
Higher derivative terms can be absorbed
into the above couplings provided these are treated as form factors
and not constants. It is thus important to bear in mind the fact that the
strength of the various terms in the vertex would vary (in general,
independently) with the momentum scales of the process being
considered. The imaginary parts of the form factors are essentially
the absorptive parts of the $WWV$ vertex functions, and, as such, are
small in the SM or MSSM. Although absorptive parts that arise
from the same sector of new physics as the anomalous couplings
themselves (as for example in the
models of Ref.~\cite{dynmodel}) need not be suppressed,
we will assume here that the anomalous couplings are real.
To date, the only direct limits on $CP$-violating $WW\gamma$ couplings
are \\[1mm]
($i$) \mbox{$-0.92<\tilde{\kappa}_{\gamma}<0.92$,}\ \mbox{$-0.31<\tilde{\lambda}_{\gamma}<0.30$}
from an analysis of $p\bar{p}\rightarrow W\gamma + X$ events done by the
D0 Collaboration at Tevatron~\cite{D0:Wgamma}, and
\\[1mm]
($ii$) $\tilde{\kappa}_{\gamma} = 0.11^{+0.71}_{-0.88}\pm 0.09$,
$\tilde{\lambda}_{\gamma}=0.19^{+0.28}_{-0.41}\pm0.11$ from the analysis of
$e^+e^- \rightarrow W^+W^-$ and $ We\nu$ data collected by DELPHI
Collaboration~\cite{Delphi}.
A competitive indirect limit $|\tilde{\kappa}_\gamma|<0.6$ has been derived
recently~\cite{gb} from the $b\rightarrow s \gamma$ CLEO data~\cite{cleo},
whereas indirect constraints based on neutron electric dipole moment
(EDM) put a very strong limit $|\tilde{\kappa}_\gamma|\:\raisebox{-0.5ex}{$\stackrel{\textstyle<}{\sim}$}\: 2 \times
10^{-4}$~\cite{edmn}. By investigating $e^+e^-\rightarrow W^+W^-$ at a future
500 GeV linear collider, it has been shown that the constraint
$|\tilde{\kappa}_{\gamma}|\leq 0.1$ can be established~\cite{likeli}. Similar
conclusion has been reached for the process $p\bar{p}\rightarrow W\gamma$ at
an upgraded Tevatron~\cite{dawson}. Better limits
of the order of $10^{-3}$ can be reached in the $e\gamma$ and
$\gamma\gamma$ modes of the linear colliders with the polarized
Compton back-scattered photon beams \cite{eg-gg}. As for $WWZ$ couplings, the
reaction $e^+e^-\rightarrow\nu\bar{\nu} Z$ has been examined~\cite{rindani}
with the resulting limit of the order $|g^{Z}_4|< 0.1$.
Given the tight (but subject to theoretical assumptions)
indirect bound from EDM, it seems
unlikely that a non-zero $CP$-violating
couplings will be observed directly at future colliders.
Nevertheless, the EDM and the direct observables that we study in this
paper depend on different combinations of the anomalous couplings
and consequently provide complementary information.
We argue therefore that experimental searches,
wherever possible, should be attempted, if only to overdetermine
the system.
\section{Why $e^+e^- \rightarrow \nu\bar{\nu}\gamma$?}
\label{sec:nunug}
The process $e^+e^-\rightarrow W^+W^-$ necessarily involves both
$WW\gamma$ and $WWZ$ vertices and consequently all 14 couplings. On
the other hand, the reaction \cite{vvg}
\begin{equation}
e^- (p_1) + e^+(p_2) \longrightarrow
\nu(p_3) + \bar{\nu}(p_4) + \gamma(p_5),
\label{nunug}
\end{equation}
with the Feynman diagrams shown in Fig.~\ref{fig:feyn}, has the
advantage\footnote{The same is also true for both $p\bar{p}\rightarrow W\gamma$ and
$e^\pm\gamma \rightarrow W^\pm\nu$.}
that only the $WW\gamma$ vertex is present.
Therefore the process with ``a photon + missing
energy'' in $e^+e^-$ annihilation can probe $WW\gamma$ couplings
independently of $WWZ$,
reducing greatly the number of unknown couplings to be determined
experimentally.
\begin{figure}[htb]
\input{fig_nng.axo}
\vspace*{2em}
\caption{\em The Feynman diagrams responsible for
$e^+ e^- \rightarrow \bar \nu \nu \gamma$.}
\label{fig:feyn}
\end{figure}
Moreover, the number of unknown couplings is further reduced to
$ \Delta\kappa_\gamma$, $\lambda_\gamma$, $\tilde{\kappa}_\gamma$ and
$\tilde{\lambda}_\gamma$, since
electromagnetic gauge invariance requires that, for on-shell
photons, $\Delta g^\gamma_1 =g^\gamma_4 =g^\gamma_5=0 $, though these
can assume other values for off-shell photons, a fact often missed in
the literature.
For the process (\ref{nunug}), being formally of
higher order in the
electroweak coupling than $W$ pair production, one may expect a
reduced sensitivity. However, since the total cross section for this
process increases\footnote{Actually the rapid rise of the cross
section allows for generous kinematical cuts to suppress possible
backgrounds, as we will demonstrate in this paper.} with incoming
energy \cite{jk_dc} until fairly large $\sqrt{s}$,
while that for $W^+W^-$ decreases strongly with
$\sqrt{s}$, the reduction in sensitivity is less and less severe at
higher energy. Coupled with $W^+W^-$ and $W\gamma$ measurements,
performed at the different momentum transfers, the process
(\ref{nunug}) would allow the possibility of studying the form factor
nature of anomalous couplings.
The sensitivity of $e^+e^-\rightarrow \nu\bar{\nu}\gamma$ to $C$ and $P$
conserving anomalous couplings ($\Delta\kappa_{\gamma}$ and
$\lambda_{\gamma}$) at future linear
colliders\footnote{First experimental results from this process
at LEP2 have been published recently~\protect\cite{lep:photon}.},
has been recently
studied in detail in Ref.~\cite{jk_dc}. Here we will study the
sensitivity to $CP$-violating couplings, namely to $\tilde{\kappa}_{\gamma}$ and
$\tilde{\lambda}_{\gamma}$. In the static limit they are related to the electric
dipole moment $d_W=e(\tilde{\kappa}_\gamma+\tilde{\lambda}_\gamma)/2m_W$ and magnetic quadrupole
moment $Q_W=-e(\tilde{\kappa}_\gamma-\tilde{\lambda}_\gamma)/m^2_W$ of the $W^+$.
Since, in the process $e^+e^-\rightarrow\nu\bar{\nu}\gamma$, the only
kinematical variables at our disposal are the energy $E_\gamma$ and
the polar angle $\theta_\gamma$ of the produced photon, no truly
$CP$-odd observables can be constructed. In the absence of the phases of
$\tilde{\kappa}_{\gamma}$ and $\tilde{\lambda}_\gamma$ we can only look for
the quadratic effects that the $CP$-violating anomalous couplings
induce in the differential distribution of photons. It is therefore
possible to exploit the same $CP$-conserving observables and to follow
the methods of Ref.~\cite{jk_dc} to derive limits on $CP$-violating
photon couplings at future $e^+e^-$ linear colliders.
\section{The SM expectations for $e^+e^-\rightarrow \nu\bar{\nu}\gamma$}
\label{sec:SM}
The SM cross-section for the process (\ref{nunug})
is best calculated using helicity amplitudes
and the relevant expressions can be found in Ref.~\cite{AKS}.
Experimentally, the signal comprises a single energetic photon
plus missing momentum carried by invisible neutrinos. The energy and
direction of the photon can be measured with high accuracy. Note
however, that only diagram \mybox{5} of Fig.~\ref{fig:feyn}
can contribute to the signal, while all diagrams (including \mybox{5})
contribute to the background.
By applying simple kinematical cuts,
some of these contributions can be suppressed and the
sensitivity to TGC enhanced significantly.
For example, diagrams \mybox{3} and \mybox{4}
are responsible for an enhancement of the
cross-section at both small photon energy and small emission
angles. To eliminate these, we impose
\begin{equation}
25^{\circ}<\theta_{\gamma}<155^{\circ}
\label{ang_cut}
\end{equation}
as well as\footnote{The same energy cut has been used in the
recent analysis by ALEPH \cite{lep:photon}.}
\begin{equation}
E_{\gamma} > 0.1 \sqrt{s} \ .
\label{energy_cut}
\end{equation}
Note that cut (\ref{energy_cut}) is different from that of
Ref.~\cite{jk_dc}, wherein a $\sqrt{s}$--independent cut of
$E_\gamma > 25 \: {\rm GeV} $ was imposed. This modification obviously
implies that the selected events
must have higher transverse momentum thus avoiding
potential background contributions
from processes such as $e^+e^- \rightarrow \gamma\gamma$ where one
photon disappears down the beam pipe.
This is especially true for larger $\sqrt{s}$.
Similarly, to eliminate events where an
on-shell $Z$ boson decays to a $\nu\bar{\nu}$ pair
(diagrams \mybox{1} and \mybox{2} with radiative
$Z$-return), we require here that the photon energy satisfies
\begin{equation}
\left| E_\gamma - \frac{s - m_Z^2}{2 \sqrt{s}}
\right|
> 5\Gamma_{Z} \ ,
\label{minv_cut}
\end{equation}
where $M_{Z}$ and $\Gamma_{Z}$ are the mass and the width of the $Z$
boson. With these cuts, the SM cross section
(summed over neutrino flavours) is
\begin{equation}
\Maroon
\sigma_{\rm SM}(\bar \nu \nu \gamma)
= \left\{
\begin{array}{ll}
0.469 \: {\rm pb} & \qquad \sqrt{s} = 350 \: {\rm GeV} \\
0.437 \: {\rm pb} & \qquad \sqrt{s} = 500 \: {\rm GeV} \\
0.361 \: {\rm pb} & \qquad \sqrt{s} = 800 \: {\rm GeV} \ .
\end{array}
\right.
\label{nngam_cs}
\Black
\end{equation}
For CM energy in the range (200--1000 GeV), the cross-section falls
almost linearly (see Fig.~\ref{fig:rt_s_sm}). This is in marked contrast
to Fig.~(2$a$) of Ref.~\cite{jk_dc} where the cross-section was shown to
{\em increase} with $\sqrt{s}$. The difference obviously lies in the
\begin{figure}[hb]
\vspace*{-0.0cm}
\hspace*{-0.5cm}
\centerline{
\epsfxsize=7.0cm\epsfysize=6.5cm
\epsfbox{fig_rts_sm.ps}
}
\caption{\em The energy dependence of the SM
cross section.
}
\label{fig:rt_s_sm}
\end{figure}
stronger form of the energy cut (\ref{energy_cut}) that we use. It
might seem that the loss of statistics that such a cut entails
might reduce the sensitivity. However, as we shall show in
section~\ref{sec:sensit}, this is not really the case.
\section{The anomalous contribution}
\label{sec:anom_contr}
A non-zero value for any one of the anomalous couplings in
eq.~(\ref{lagrangian}) would imply additional terms in the matrix
element arising from diagram \mybox{5}. It is easy to see
that the contributions due to $\tilde{\kappa}_\gamma$ and
$\tilde{\lambda}_\gamma$ do not interfere
with the SM amplitude. This is but a reflection of the fact
that one cannot construct a $CP$-odd observable for the process
of eq.~(\ref{nunug}). Denoting $d_{ij} \equiv p_i \cdot p_j$, the anomalous
contribution to the spin-summed (averaged) matrix-element-squared
is then
\begin{equation}
\begin{array}{rcl}
\displaystyle \left( \frac{e g^2}{P_{13} P_{24} } \:
\right)^{-2} \;
|{\cal M}|^2_{\rm anom}
& = & \displaystyle
\tilde{\kappa}^2 {\cal C}
+
\frac{4 \tilde{\lambda}^2}{m_W^4} d_{13} d_{24}
\left\{ {\cal C} + (d_{13} - d_{24})
(d_{15} d_{35} - d_{25} d_{45})
\right\}\\[1.5ex]
& + & \displaystyle
\frac{2 \tilde{\kappa} \tilde{\lambda}}{m_W^2}
\Big\{ (d_{13} + d_{24}) {\cal C}
+ (d_{13} - d_{24})
(d_{24} d_{15} d_{35}
- d_{13} d_{25} d_{45})\\[2ex]
& & \displaystyle \hspace*{3em}
- 2 d_{15} d_{25} d_{35} d_{45}
\Big\} \ ,
\end{array}
\label{anom_contr}
\end{equation}
where we have suppressed the subscript $\gamma$ in $\tilde{\kappa}$ and $\tilde{\lambda}$
and
\[
\begin{array}{rcl}
P_{13} & \equiv & 2 d_{13} + m_W^2
\\
P_{24} & \equiv & 2 d_{24} + m_W^2
\\
{\cal C} & \equiv &
d_{14} d_{25} d_{35} + d_{23} d_{15} d_{45}
\ .
\end{array}
\]
The numerical value of the extra contribution can be
obtained by integrating $|{\cal M}|^2_{\rm anom}$
over the appropriate phase space
volume. In Fig.~\ref{fig:rt_s_anom}, we display this quantity
as a function of $\sqrt{s}$ for unit values of $\tilde{\kappa}$ and $\tilde{\lambda}$
(and $n=0$; for $n>1$ see below). The
generalization to arbitrary values is trivial. In contrast to the SM
case (Fig.~\ref{fig:rt_s_sm}), the anomalous contribution {\em grows}
with the CM energy, the effect being more pronounced for dimension
6 operator, $i.e.$ the non-zero $\tilde{\lambda}$ coupling
(note the different scales on vertical axes).
This is but a consequence of the lack of unitarity for such theories.
\begin{figure}[htb]
\vspace*{-0.0cm}
\hspace*{-0.5cm}
\centerline{
\epsfxsize=6.cm\epsfysize=7.0cm
\epsfbox{fig_rts_kapsq.ps}
\vspace*{-0.0cm}
\hspace*{-1.0cm}
\epsfxsize=6.cm\epsfysize=7.0cm
\epsfbox{fig_rts_lamsq.ps}
\vspace*{-0.0cm}
\hspace*{-1.0cm}
\epsfxsize=6.cm\epsfysize=7.0cm
\epsfbox{fig_rts_intf.ps}
\vspace*{-0.3cm}
}
\caption{\em The energy dependence of the {\em extra}
pieces in the cross section for unit values of the
anomalous couplings. The solid, short-dashed, long-dashed and
dot-dashed curves correspond to $n = 0, 1, 2, 4$ in
eqn.(\protect\ref{unitarity}). The scale ($\Lambda$) of new physics
has been assumed to be $1 \: {\rm TeV} $.
}
\label{fig:rt_s_anom}
\end{figure}
Tree-level unitarity may be restored by postulating a
form-factor behaviour for $\tilde{\kappa}$ and $\tilde{\lambda}$. To wit,
\begin{equation}
\tilde{\kappa} = \tilde{\kappa}_0 \left[ \frac{\Lambda^4}
{(\Lambda^2 + 2 d_{13})
(\Lambda^2 + 2 d_{24})}
\right]^n
\label{unitarity}
\end{equation}
and similarly for $\tilde{\lambda}$~\cite{edmn}.
In eqn.(\ref{unitarity}), $n$ is an integer and
$\Lambda$ is the scale where new physics manifests itself.
While $n \geq 1$ ensures that the cross-section falls off
for sufficiently high $\sqrt{s}$, the effect is not so marked
for the regime of interest, even for relatively low values
of $\Lambda$ (see Fig.~\ref{fig:rt_s_anom}).
A related consequence of this lack of partial wave unitarity
is that the high-energy end of the photon spectrum gets
disproportionately populated. In fact, the strongest
dependence on anomalous TGC appears towards the end of the
photon energy spectrum, {\em i.e.}, in the region which is seriously affected
by the cut~(\ref{minv_cut}) designed to eliminate $\gamma Z$ events.
Fortunately though, this overpopulation persits for somewhat
lower values of $E_\gamma$ as well thus allowing us to draw
relatively strong constraints on such couplings.
\section{An estimate of the sensitivity}
\label{sec:sensit}
In assessing the sensitivity of a future LC we will use the double
differential distribution
${\rm d}^2 \sigma/{\rm d} E_{\gamma}\,{\rm d} \cos\theta_\gamma $
with the phase space divided into a number of bins.
Choosing a simple $\chi^{2}$ test to
derive 95\% C.L. boundaries in the two-parameter space of
($\tilde{\kappa}_{\gamma}$, $\tilde{\lambda}_{\gamma}$), we define
\begin{equation}
\chi^{2}=\sum_{i}^{{\rm \# \ of\ bins}} \left |
{\frac{N_{SM}(i)-N_{AN}(i)}{\Delta N_{SM}(i)}} \right |^{2} \:,
\label{chiform}
\end{equation}
with $N_{SM}(i)$ and $N_{AN}(i)$ being the
number
of events in the bin $i$ expected within the SM and a theory with the
anomalous TGC, respectively. The error
$\Delta N_{SM}$ is defined as a combination of statistical and
systematic errors ({\it cf.}~\cite{jk_dc})
\begin{equation}
\Delta N_{SM}=\sqrt{(\sqrt{N})^2+(\delta_{syst}N)^{2}} \: .
\label{delta}
\end{equation}
For our numerical analysis, we use a few
sets of machine parameters ( {\em i.e.}\ luminosities and CM energies)
considered in the current ECFA-DESY workshop on future
LC~\cite{ecfa-desy},
\begin{equation}
{\cal L} =
\left\{
\begin{array}{rrrc rlll}
25, & 75, & 100 & \& & 300 & \: {\rm fb}^{-1} & \qquad \sqrt{s} = 350 \: {\rm GeV} \\
75, & 125, & 300 & \& & 500 & \: {\rm fb}^{-1} & \qquad \sqrt{s} = 500 \: {\rm GeV} \\
125, & 200, & 500 & \& & 800 & \: {\rm fb}^{-1} & \qquad \sqrt{s} = 800 \: {\rm GeV} \
\end{array}
\right. \ .
\label{lumin}
\end{equation}
For comparison, we also consider the LEP2 environment with $\sqrt{s}=192$
GeV and ${\cal L}=0.5$ and 2 fb$^{-1}$.
We divide the entire range of angular acceptance
(see eq.~(\ref{ang_cut})) into 26 equal-sized bins of $5^\circ$ each,
while, for $E_\gamma$, we assume uniform bins of $10 \: {\rm GeV} $ each.
It might seem counterintuitive to use more bins for large $\sqrt{s}$
in view of the smaller SM cross-section. However, this decrease in
cross-section is more than compensated for by the large increase in the
proposed luminosity.
\begin{figure}[htb]
\vspace*{-0.0cm}
\hspace*{-0.5cm}
\centerline{
\epsfxsize=8.3cm\epsfysize=8cm
\epsfbox{fig_cont_192.ps}
\vspace*{-0.0cm}
\hspace*{-0.3cm}
\epsfxsize=8.3cm\epsfysize=8cm
\epsfbox{fig_cont_350.ps}}
\vspace*{-0.5cm}
\centerline{
\vspace*{-1cm}
\hspace*{-0.cm}
\epsfxsize=8.3cm\epsfysize=8cm
\epsfbox{fig_cont_500.ps}
\vspace*{-0.0cm}
\hspace*{-0.3cm}
\epsfxsize=8.3cm\epsfysize=8cm
\epsfbox{fig_cont_800.ps}
\vspace*{-2.3cm}
}
\caption{\em 95\% exclusion contours for different
energies and luminosities.
It is assumed that there is no form-factor suppression for the
couplings ($n = 0$ or $\Lambda \rightarrow \infty$ in
eq.~(\protect\ref{unitarity})).}
\label{fig:contours}
\end{figure}
The systematic error
$\delta_{syst}$ arises mainly from the detector
parameters (e.g. uncertainty in the luminosity measurement and
the detector efficiency).
We take $\delta_{syst}$ to be 2\%, which is commonly considered as a
fairly conservative assumption. As it turns out, the error in
$\Delta N_{SM}$ is dominated by statistical errors and the
results are insensitive to small changes in $\delta_{syst}$.
For a theory with 2 variables, 95\% C.L. corresponds to
$\chi^{2}=6$ and the corresponding contours are shown in
Fig.~\ref{fig:contours}. Clearly, a marked improvement
accompanies an increase in luminosity. This reflects the already
stated fact of $\delta_{syst}$ in eq.~(\ref{delta}) being dominated by
the statistical errors.
Also easy to discern is that a higher CM energy results
in stronger constraints even for the same luminosity. This is
expected since such couplings lead to a rapid growth in the
number of events with $\sqrt{s}$ (see Fig.~\ref{fig:rt_s_anom}).
By the same reasoning, the improvement along the $\tilde{\lambda}$ axis
is more pronounced than that along the $\tilde{\kappa}$ axis.
The contours in Fig.~\ref{fig:contours} are nearly elliptical and
with very little tilt. This of course implies that there is
only a small correlation between the constraints on $\tilde{\kappa}$ and $\tilde{\lambda}$,
a feature that we should have expected from a comparison of
the relative strengths of the $\tilde{\kappa}^2$, the $\tilde{\lambda}^2$ and the $\tilde{\kappa} \tilde{\lambda}$
pieces in the cross-section (see Fig.~\ref{fig:rt_s_anom}). If one of the
coupling is known to be identically zero
(or determined by another experiment),
then the individual bounds on the other can be obtained
by determining $\chi^2 = 1 $ (68.3\% C.L.) or $\chi^2 = 3.8$ (95\% C.L.)
as the case may be. These bounds
are summarised in Table~\ref{95results}.
\input{table.1}
Until now, we have sidestepped a few issues, namely the role of
beam polarization, possible form-factor dependence of the couplings
and the role of the minimum energy cut. The first is easy to estimate.
Since the signal receives contributions only from left-handed electrons,
polarizing the beam so is expected to help. But this would also increase
the background by almost as
much\footnote{Right-handed electrons participate only in diagrams
\mybox{1} and \mybox{2} of Fig.~\protect\ref{fig:feyn}. But these
are almost totally eliminated by the cut of
eq.~(\protect\ref{minv_cut}).}.
Consequently, the improvement is akin to that resulting from a somewhat
higher luminosity.
As for the form-factor dependence, clearly the anomalous contribution
decreases with increasing $n$ (or smaller $\Lambda$). More importantly,
the high-energy part of the photon spectrum gets depleted faster. This
implies weaker constraints as evinced by Fig.~\ref{fig:other_contours}($a$).
Again, the effect is more pronounced along the $\tilde{\lambda}$ axis.
Finally, we come to the role of the cut on photon energy
eq.~(\ref{energy_cut}). As we commented earlier, one role of
this cut is to eliminate contributions from diagrams \mybox{3} and
\mybox{4} of Fig.~\ref{fig:feyn}. Strictly speaking, this is not
necessary as the $\chi^2$ function would simply assign a low weight
to such bins. In fact, in the absence of other backgrounds,
this cut (as also the other two) only succeeds in rejecting
additional small but positive contributions to the $\chi^2$.
Fortunately, this is not a severe loss as Fig.~\ref{fig:other_contours}($b$)
demonstrates clearly. More importantly, eq.~(\ref{energy_cut}) serves
to eliminate other backgrounds such as $e^+ e^- \rightarrow \gamma \gamma$
where one of the photons is missed by the electromagnetic calorimeter.
In a real experimental setup, the precise nature and utility
of such cuts would be dictated by the detector design and hence
it is premature to dwell on it at further detail.
\begin{figure}[htb]
\vspace*{-0.0cm}
\hspace*{-0.5cm}
\centerline{
\epsfxsize=7.5cm\epsfysize=7.0cm
\epsfbox{fig_cont_unitar.ps}
\vspace*{-0.0cm}
\hspace*{-0.15cm}
\epsfxsize=7.5cm\epsfysize=7.0cm
\epsfbox{fig_cont_egam.ps}
\vspace*{-0.3cm}
}
\caption{\em
(a) The effect of form-factor behaviour
(see eq.~(\protect\ref{unitarity})) on the exclusion
contour for a given energy and luminosity.
(b) The effect of the minimum energy requirement
(see eq.~(\protect\ref{energy_cut})) on the exclusion
contour for a given energy and luminosity.
Backgrounds other than those from eq.~(\protect\ref{nunug})
are deemed to be absent.}
\label{fig:other_contours}
\end{figure}
\section{Conclusions}
\label{sec:concl}
We have investigated the process $e^+e^-\rightarrow\nu\bar{\nu}\gamma$
as a means to derive limits on $CP$-violating couplings
$\tilde{\kappa}_\gamma,\:\tilde{\lambda}_\gamma$ at a future linear $e^+e^-$ collider.
Being sensitive only to the $WW\gamma$ couplings, it permits their
independent evaluation. In addition, probing different kinematical
configurations (real photon,
space-like $W$) as compared to that of $e^+e^- \rightarrow W^+W^-$
(real $W$, time-like photon),
it allows us to probe the form factor behaviour of
the couplings over a distinct region of the momentum transfer. We
have shown that using the differential distribution
$d\sigma/dE_{\gamma}\,d\cos\theta_\gamma$ and appropriate kinematical cuts,
constraints comparable with those expected from $W\gamma$ production at
Tevatron or $W^+W^-$ production at a 500 GeV LC can be obtained. As this
study makes use of all attainable physical information, a
potential improvement of the results can be only achieved by applying
other statistical methods of testing the consistency with the SM.
\newpage
\begin{center}
ACKNOWLEDGMENTS
\end{center}
\noindent AK wishes to thank James Stirling for many
illuminating discussions.
The early stage of AK's
work was partially supported by the TEMPUS mobility project MJEP 9006.
JK has been partially supported by the Polish Committee for
Scientific Research Grant No. 2 P03B 030 14.
|
\section{INTRODUCTION}
Broad-band spectra of gamma-ray bursts (GRBs)
pose a difficult challenge to any theoretical model trying to
explain them. Looking only at a limited range of energy, as, for example, each
of the different instruments on board the {\it Gamma Ray Observatory} (GRO)
does individually, results in
a featureless power law perhaps with some curvature. However, a broad-band spectrum, ranging over
many decades in energy, typically contains interesting features like
peaks, curvature and breaks. Such features will be diagnostic of the
physical processes in the burst fireball and the spectra can be used
to directly test models of burst emission. Only a few broad-band spectra
have been produced (Schaefer et al. 1998; Greiner et al. 1995; Hurley
et al. 1994), as only bright bursts detected by multiple instruments
on board the GRO have a wide enough range of available data. The
brightest such burst is GRB 930131 which reached a peak flux of
$105 \mbox{\ ph} \mbox{\ s}^{-1} \mbox{\ cm}^{-2}$ (Meegan et al. 1996). This burst has
BATSE trigger number 2151 and has been called the ``Superbowl Burst'' after
its time of occurence. EGRET and COMPTEL spectra have already appeared
in the literature (Sommer et al. 1994; Ryan
et al. 1994), but no BATSE spectrum has been presented due to
severe deadtime problems.
This paper is organized in the following way. In \S 2, we provide general
methods for combining spectra obtained by the same instrument during
different time intervals (2.1.),
as well as for combining spectra taken by different instruments covering the
same time interval (2.2.).
These methods can
be used in many common GRB applications, provided the necessary requirements
are met. In \S 3, we carry out the construction of the broad-band spectrum
of GRB 930131 from 20 keV to 200 MeV.
First, we describe how the individual (BATSE, COMPTEL, and
EGRET) spectra have been obtained (3.1.).
Then, we argue why these independently reduced spectra
can be combined with the method of \S 2, where we point out
the non-obliging nature of this procedure in the present case. After
presenting the resulting spectrum (3.2.), we compare this to theoretical
models of the GRB emission mechanism (3.3.). Subsequently, we discuss
evidence for spectral evolution (3.4.). Finally, \S 4 summarizes the
spectral properties of this remarkable burst.
\section{COMBINING SPECTRA}
Often the problem occurs to combine individual spectra into either a
time-averaged or an instrument-averaged spectrum. The second case arises
in cross-calibrating spectral information from instruments that are
sensitive in different energy ranges.
In this section, it is
assumed that the observed count spectra have already been reduced into
photon spectra.
In the following, we describe the method of combining spectra and
give the relevant formulae, which are then applied to the case of
GRB~930131 in Section 3.
\subsection{Combining Across Time}
Suppose the time over which one wants to average is divided up into
smaller time intervals $k$ with respective livetimes $\tau_{k}$. For
each time interval $k$ and energy bin $i$ the photon flux
(in units of photons/area/energy/time) is
$\left(\frac{dn}{de}\right)_{ik}$ with standard deviation $\sigma_{ik}$.
Then, constructing the time-averaged spectrum is straightforward. With
the total livetime given by
$\tau_{total}=\sum_{k}\tau_{k}$, the time-averaged photon flux in
energy bin $i$ is
\begin{equation}
\left(\frac{dN}{dE}\right)_{i}=\tau^{-1}_{total}\sum_{k}\left
(\frac{dn}{de}\right)_{ik}
\tau_{k} \mbox{\ \ ,}
\end{equation}
and the resulting standard deviation is
\begin{equation}
\sigma_{i}^{2}=\tau^{-1}_{total}\sqrt{\sum_{k}\left(\sigma_{ik}\tau_{k}
\right)^{2}}
\mbox{\ \ .}
\end{equation}
\subsection{Combining Across Different Instruments}
Spectra from different instruments can be combined just as can spectra
from multiple detectors on the same instrument. We here assume that the
combination process is robust, i.e., that the resulting spectrum is not
greatly obliging (cf., Section 3.2.).
This has to be justified on a case by case basis.
Another requirement
is that either the input spectra are for identical time
intervals, or they cover the entire burst. This
combination can be described as a four-step process:
{\it Step A}:\\
The spectra from different instruments are divided into energy bins in
different ways. Therefore, as a first step, all the bin boundaries
($E^{low}$ and $E^{high}$)
from all the instruments are put into increasing order and then used
to define subbins. Assume that after the ordering, the following
sequence arises:\,
$\mbox{...}<E_{k-1}<E_{k}<E_{k+1}<\mbox{...}$\,
Then define the $k$th subbin to cover an energy interval between $E_{k}$ and
$E_{k+1}$. Figure 1 illustrates this procedure for the case of two
instruments.
{\it Step B}:\\
It is preferable to conduct the combining in $\nu F_{\nu}$-space,
where $\nu F_{\nu}\propto\left(\frac{dN}{dE}\right)E^{2}$. Then the
spectrum is roughly constant over a given energy bin, as opposed
to the usual steep decline in ordinary $\frac{dN}{dE}$-space. Now, for
energy bin $i$ of instrument $m$,
having lower and higher energies $E_{mi}^{low}$ and $E_{mi}^{high}$,\,
respectively, define the energy flux
per logarithmic energy interval
\begin{equation}
\left(\frac{d\varphi}{dE}\right)_{mi}\equiv\left(\frac{dN}{dE}\right)_{mi}\left(E_{mi}^{mid}\right)^{2} \pm \sigma_{mi}\mbox{\ \ ,}
\end{equation}
where $E_{mi}^{mid}=\sqrt{E_{mi}^{low}\cdot E_{mi}^{high}}$ and
$\sigma_{mi}$ is the uncertainty of $\left(\frac{d\varphi}{dE}\right)_{mi}$.
Our procedure presumes that $\left(\frac{d\varphi}{d\varphi}\right)_{mi}$
changes little across each energy bin, as is the case for energy bins
that are small compared to either the detector resolution or the structure
in the spectrum. This covers
virtually all GRB applications, although a simple interpolation scheme
might be appropriate for a particularly steep spectrum observed with
very broad bins.
{\it Step C}:\\
Now, we want to cross-combine the
spectra of different instruments.
In constructing the spectrum for subbin $k$, we first determine whether
a given instrument $m$ has an energy bin $i$ overlapping the subbin.
If this is the case, we set
\begin{equation}
\left(\frac{d\varphi}{dE}\right)_{mk}=\left(\frac{d\varphi}{dE}\right)_{mi}
\mbox{\ \ \ \ \ and\ \ \ \ \ } \sigma_{mk}=\sigma_{mi}\mbox{\ \ .}
\end{equation}
Figure 1 shows the case of two instruments having overlapping energy
bins with subbin $k$. The energy flux
of the cross-combined spectrum is the weighted average of
all contributing spectra:
\begin{equation}
\left(\frac{d\phi}{dE}\right)_{k} = \sigma_{k}^{2} \cdot \sum_{m}
\frac{1}{\sigma_{mk}^{2}}\left(\frac{d\varphi}{dE}\right)_{mk}\mbox{\ \ ,}
\end{equation}
where
\begin{equation}
\sigma_{k}=\left(\sum_{m}\sigma_{mk}^{-2}\right)^{-1/2}\mbox{\ \ .}
\end{equation}
{\it Step D}:\\
As a last step, put together the subbins into larger bins of width
appropriate for the spectral resolution and features.
Rebinning, e.g., two subbins $k$ and $k+1$ into a larger bin $l$
with boundaries $E_{l}^{low}$ and $E_{l}^{high}$ is accomplished by the following:
\begin{equation}
\left(\frac{d\Phi}{dE}\right)_{l}=w_{k} \left(\frac{d\phi}{dE}\right)_{k}
+ w_{k+1} \left(\frac{d\phi}{dE}\right)_{k+1}\mbox{\ \ ,}
\end{equation}
where one has for the respective weights
\begin{equation}
w_{k}=
\frac{E_{k+1}-E^{low}_{l}}{E^{high}_{l}-E^{low}_{l}}\mbox{\ \ ,}
\end{equation}
and
\begin{equation}
w_{k+1}=
\frac{E^{high}_{l}-E_{k+1}}{E^{high}_{l}-E^{low}_{l}}\mbox{\ \ .}
\end{equation}
The resulting standard deviation is
\begin{equation}
\sigma^{2}_{l}=w_{k}^{2}\sigma^{2}_{k}
+ w_{k+1}^{2}\sigma_{k+1}^{2}\mbox{\ \ .}
\end{equation}
If the output bin covers more than two subbins, then equations (7)-(10) can be
easily generalized or used repeatedly.
\placefigure{fig1}
\section{THE SPECTRUM OF GRB 930131}
\subsection{The Individual Spectra}
For all 3 instruments (BATSE, EGRET, COMPTEL), their photon spectra
have been obtained by the traditional forward-folding technique
(Loredo \& Epstein 1989). This technique assumes a variety of spectral
models $M$, and convolves them with the respective detector response
matrix (DRM), symbolically $C_{model}=\mbox{DRM}\ast M$, where $C_{model}$ is
the count spectrum predicted by the model. The parameters of the model are then
adjusted to obtain the best fit to the observed count spectrum,
$C_{obs}=\mbox{DRM}\ast P_{true}$, where $P_{true}$ is the true (photon)
spectrum of the source. Alternatively, a model-independent inverse technique
could have been adopted, where $P_{true}=\mbox{DRM}^{-1}\ast C_{obs}$.
Attempts at doing so have proven unconvincing, and the nearly universal
practice in gamma-ray astronomy is to use forward-folding techniques.
One exception is the direct inversion
method of Pendleton et al. (1996), which has only been applied to low
resolution (4-channel) data and introduces considerable additional
error (10-15\%).
\subsubsection{\it BATSE Spectrum}
The ``Superbowl Burst'' suffers from severe deadtime effects,\,which is the
reason why the original discovery paper (Kouveliotou et al.\,1994) does
not present a spectrum for the BATSE energy range. For this bright burst, most of
the flux arrives in the first 0.06~seconds, a situation which saturates the
BATSE Large Area Detectors (LADs), whereas the smaller but thicker Spectroscopy
Detectors (SDs) can reliably record the intense photon flux. In constructing our
spectrum, we have selected the two burst-facing Spectroscopy Detectors (SD 4 and 5), for
which there are available the well suited
STTE-data (SD Time-Tagged Events), which cover
the first $\sim$~1.5~s of the burst and which have a time resolution
of $128\mbox{\,} \mu$s. Therefore, we can correct for the deadtime effects by
subdividing the total time into 53 individual spectra with a duration
of as short as a few ms around the first, intense peak. For each time interval,
the photon spectrum is obtained by following the procedure described in
Schaefer et al. (1994). In carrying out the forward-folding, we assume a
single power-law spectral model.
Then, by applying the methods of
Section 2.1., we constructed time-averaged spectra for SD 4 and 5
, which were then in turn combined (as described in Section 2.2.) to
give the overall spectrum for the BATSE
energy-range (21 keV to 1.18 MeV, above which the flux-errors
exceed 100\%).
\subsubsection{\it COMPTEL And EGRET Spectra}
The COMPTEL and EGRET spectra have previously been published (Ryan et al. 1994,
and Sommer et al. 1994, respectively) and we refer the reader to these
papers for details. We have chosen to work with the spectrum reported
by the EGRET Total Absorption Shower Counter (TASC), since the EGRET spark
chamber is too severely affected by deadtime effects. The TASC spectrum
covers an energy range from 1 MeV to 180 MeV.
The overlap region between BATSE and
TASC is nicely covered by the COMPTEL instrument, where the COMPTEL Telescope
spectrum covers the range from 0.75 Mev to 30 MeV. Both spectra have
been obtained by the forward-folding technique with a power-law model, and
are corrected for deadtime effects.
\subsection{The Combined Spectrum}
To construct the combined spectrum with the method described
in Section 2.2., we first have to ascertain the robustness of this procedure.
It is well known (Fenimore et al. 1983) that the resulting spectral shape
can possibly depend sensitively on the details
of the fitting technique (i.e., that
the spectra might be ``obliging''). In principle, it could make a big
difference whether the low- and high-energy parts, covered by different
instruments, are first unfolded separately and only then combined together, or
whether the unfolding is done simultaneously to all instruments. The
physical reason for this is that high-energy photons might masquerade
as low-energy ones, and that, consequently, the low-energy part of
the spectrum cannot be accurately
unfolded independently of the high-energy part.
For the present case, however, this problem does not occur. It has
been convincingly shown that the BATSE Spectroscopy Detectors are
non-obliging (Schaefer et al. 1994; cf., their Figures 11 and 52).
This is primarily due to their thickness, which largely minimizes photon
energies being underreported.
The TASC and COMPTEL spectra, on the other
hand, are not affected by the lower energy BATSE range. Finally, treating
the COMPTEL and TASC spectra independently of each other is rendered
possible by the fact that the model fitting leads to almost
identical results ($\frac{dN}{dE}\propto E^{-2}$). We are
therefore justified in combining the independently obtained spectra
from the 3 GRO
instruments (BATSE-EGRET-COMPTEL) into the overall, broad-band spectrum
of GRB 930131.
This combination is carried out with
the method of section 2.2., where we have been careful
to construct our BATSE spectrum such that it exactly matches the
time coverage of the EGRET TASC instrument, and approximately that
of COMPTEL. To evaluate how well the instruments agree in the mutual
overlap region around 1 MeV, we compare the fluxes at 1 MeV for the
3 instruments (in units of $10^{-3}$photons cm$^{-2}$ sec$^{-1}$ keV$^{-1}$):
BATSE 2$\pm$2,
COMPTEL 8$\pm$3, and TASC $2\pm$0.5. The agreement between BATSE and TASC
is good, although the BATSE errors approach
100\% at these high energies. The COMPTEL flux is
somewhat high, but due to its uncertainties it does not
contribute significantly to the weighted average of
the final, combined spectrum.
Table 1 and Figure 2 present
the $\nu F_{\nu}$ ($\propto \left(\frac{dN}{dE}\right)E^{2}$) spectrum
in units of (photons s$^{-1}$ cm$^{-2}$ keV$^{-1}$)$\ast (E^{mid}/100
\mbox{\ keV})^{2}$.
The resulting spectrum is remarkably flat, as compared
to other published broad-band spectra, which have a much more peaked
appearance (cf., Schaefer et al. 1998). In the following section we ask, whether
this rather unusual spectral shape is consistent with the model of
shocked synchrotron emission, which successfully fits the characteristics
of other broad-band GRB spectra.
Subsequently, we investigate
whether the flat spectrum of GRB 930131 can be understood as a result of
spectral evolution.
\placefigure{fig2}
\placetable{spec}
\subsection{Model-Fits}
We fit our combined spectrum to the shocked synchrotron model
of Tavani (1996a, b),\,which gives the following analytical expression
for the energy flux:
\begin{equation}
\psi_{model}\equiv\left(\frac{d\Phi}{dE}\right)=\nu F_{\nu}=C\nu\left[I_{1} + \frac{1}{e}I_{2}\right]
\end{equation}
\begin{equation}
I_{1}=\int_{0}^{1}y^{2}e^{-y}F\left(\frac{\nu}{\nu_{c}^{\ast}y^{2}}\right)\mbox{d}y
\end{equation}
\begin{equation}
I_{2}=\int_{1}^{\infty}y^{-\delta}F\left(\frac{\nu}{\nu_{c}^{\ast}y^{2}}\right)\mbox{d}y\mbox{\ \ ,}
\end{equation}
where $F(x)\equiv x\int_{x}^{\infty}\mbox{K}_{\frac{5}{3}}(w)\mbox{d}w$ is
the usual synchrotron spectral function with $\mbox{K}_{\frac{5}{3}}$ being
the modified Bessel-function of order $\frac{5}{3}$ and $e=2.718...$.\,The
normalization constant $C$ has units of specific flux.\,
Equations
(12) and (13) are summing up the synchrotron emission from a Maxwellian
distribution of electron energies which breaks to a power law at high energies.\,
Here, $\delta$ is the index of the supra-thermal power-law distribution of
particles, resulting from relativistic shock-acceleration. The critical
frequency $\nu_{c}^{*}$ describes where most of the synchrotron power is emitted.
We apply the Levenberg-Marquardt method of non-linear $\chi^{2}$ fitting
(cf., Numerical Recipes, Press et al. 1992) to minimize
\begin{equation}
\chi^{2}=\sum_{i=1}^{N}\left(\frac{\psi_{i}-\psi_{model}(\nu_{i}^{mid};C,\delta,\nu_{c}^{\ast})}
{\sigma_{i}}\right)^{2}\mbox{\ \ .}
\end{equation}
Our observed spectrum with flux $\psi_{i}=\left(\frac{d\Phi}{dE}\right)_{i}$ and uncertainty $\sigma_{i}$ contains $N=37$ data points.
\,Our best-fit parameters are:
\begin{equation}
C=104 \pm 8 \mbox{\ \ erg cm$^{-2}$ sec$^{-1}$ Hz$^{-1}$}
\end{equation}
\begin{equation}
\delta=3.3 \pm 0.1
\end{equation}
\begin{equation}
h\nu_{c}^{\ast}=98 \pm 14 \mbox{\ \ keV}
\end{equation}
The fit has a chi-squared of $\chi^{2}=38$ with 34 degrees of freedom.
Therefore, we can conclude that the spectrum of
GRB 930131 is consistent with the Tavani-model. At low
energies,\,the spectrum is asymptotically approaching
$\nu F_{\nu} \propto \nu^{4/3}$, as is usual for burst spectra (Schaefer
et al. 1998). This behavior
is predicted by optically thin synchrotron theory (Katz 1994).
\subsection{Spectral Evolution}
All of the published broad-band spectra (Schaefer et al. 1998; Greiner
et al. 1995; Hurley et al. 1994) are
strongly peaked and fall off steeply above the peak energy. GRB~930131, on
the other hand, has a spectrum which remains constant (within a factor of 4)
over four orders of magnitude in energy. Can this
behavior be understood as the result of a superposition of many spectra, which
individually show the usual, strongly peaked shape and whose peak energy
evolves with time? For the BATSE energy range, the number of received photons is
sufficiently large to allow the construction of time-resolved spectra, whereas
for COMPTEL and EGRET, the dearth of photons renders this detailed treatment
impossible.
In Figure 3, we present the resulting BATSE spectra for 4 different times. The
lightcurve of GRB 930131, as amply documented in the literature (Kouveliotou et
al. 1994; Ryan et al. 1994; Sommer et al. 1994), shows a sharp, intense first
pulse,\,lasting for $\sim$ 0.06 s after the BATSE trigger, followed by a second, less
intense and less sharp pulse, lasting from $\sim$ 0.75 s to $\sim$ 1.00 s after
the trigger. In between, the ``interpulse'' region of Figure 3, there
is significant yet faint
flux. Finally, there is again relatively little flux subsequently to the second pulse (lasting
for another 50 s). In Figure 3, the first pulse is further subdivided into the
spectrum for the time before the maximum flux is reached (0.00 - 0.03 s) and
that for the time after the maximum (0.03 - 0.06 s).
Since these time-resolved spectra cover only the low-energy range, a meaningful fit
to the Tavani-model (cf., Section 3.3.) cannot be done, since the value of the power-law
extension $\delta$ and the location of the peak energy $h\nu^{\ast}_{c}$ are mostly
constrained by the high-energy regime. Both the
spectra for the first and second pulses
are consistent, though, with the spectral fit (besides the normalization $C$) obtained
for the overall spectrum (cf., Figure 2). Consequently, there is no evidence
that the unusual flat morphology of the ``Superbowl-Burst'' spectrum is caused
by the superposition of individually strongly-peaked, time-variable spectra.
The spectrum between pulses is inconsistent with the average burst spectral
shape. The observed $\nu F_{\nu}$ is close to $\nu^{0}$ from 21~keV to 1~MeV
with no significant curvature or maximum. The extreme brightness of
GRB~930131 allows for this unique measure of the interpulse spectrum.
\placefigure{fig3}
\section{SUMMARY AND CONCLUSIONS}
After having given the relevant formulae for combining individual spectra, we
applied these methods to construct the broad-band spectrum of GRB 930131. With
appropriate deadtime corrections we first obtained the spectrum for the BATSE energy range,
which we then combine with the already published spectra from the COMPTEL and
EGRET TASC instruments. Broad-band spectra are fortunate occurences (multiple
instruments on board the GRO have to see a bright burst), available for only
a handful of bursts.
Within the general framework of an expanding relativistic
fireball, impacting on a surrounding medium (M\'{e}sz\'{a}ros \& Rees 1993), an attractive
model for the production of the $\gamma$-ray photons is synchrotron emission from
a shocked and highly magnetized plasma (Tavani 1996a, b).
This model is successful in fitting the strongly peaked
spectral shapes (in $\nu F_{\nu}$-space)
of the GRBs for which broad-band spectra have been obtained. Since our resulting spectrum
is so unusually flat, it poses an interesting challenge to the Tavani-model. As
described in Section 3.3., the model does fit well, although with a
value for the power-law component, which lies at the extreme end of the typically
encountered range, $3<\delta<6$. In the
BATSE energy-range, we were able to construct time-resolved spectra, which show
no evidence for significant evolution.
\acknowledgments{We thank D. Palmer for his suggestions
concerning the severe deadtime problem
in the BATSE data, as well as M. Kippen
and E. Schneid for their helpful discussions.
}
\vfill\eject
|
\section{Introduction}
Cepheid variables constitute one of the
most important primary distance calibrators.
Indeed, they obey a Period-Luminosity (PL) relation:
\begin{equation}
\langle M_V \rangle = \delta \ \log P + \rho
\end{equation}
from which the absolute magnitude $\langle M_V \rangle$
can be determined just from the measurement of the period,
provided that the slope $\delta$ and the zero-point $\rho$ are known.
The slope of the PL relation seems very well established
from ground-based observations in the Large Magellanic Cloud (LMC) because
the population incompleteness bias pointed out for more distant galaxies
(Lanoix et al. 1999a) seems negligible in the LMC.
The slope of the PL relation is easier to obtain from an external galaxy
because, all Cepheids being at the same distance, the slope can be
determined by using apparent magnitudes instead of absolute magnitudes.
A reasonable value for the photometric V-band is
$\delta = -2.77 \pm 0.08$ (see for instance Gieren et al. 1998, Tanvir 1997, Caldwell \& Laney 1991, Madore
\& Freedman 1991).
In the present study we will adopt this value and will discuss further the effect
of a change of it.
The establishment of the zero-point still remains a
major goal. Today, thanks to the HIPPARCOS satellite
\footnote{HIPPARCOS parallaxes are ten times better
than those obtained from ground-based observations
(i.e., $\sigma_{\pi} \approx 1$ milliarcsec).},
the trigonometric parallaxes of galactic Cepheids are accessible,
allowing a new determination of $\rho$.
After the first release of HIPPARCOS data, a
calibration of the Cepheid PL relation
was published by Feast \& Catchpole (1997, hereafter FC).
This work gave a distance for the LMC galaxy larger than
the one generally assumed.
However, some papers (Madore \& Freedman 1998, Sandage \& Tammann 1998)
argued that this calibration is only brighter than previous ones
at the level of $\le 0.1$ mag.
An independent study of the calibration of the PL relation
based on the same data also led to a long distance scale (Paturel et al. 1996)
and to a large LMC distance modulus of 18.7 (Paturel et al. 1997).
All these studies may be affected by statistical biases due either
to the cut of negative parallaxes or to the method used for bypassing
these cuts. This justifies that we want to analyze deeper these
results.
HIPPARCOS parallaxes $\pi$ may have large standard deviations $\sigma_{\pi}$
leading sometimes to negative parallax so that the distance $d$(pc)=1/$\pi$ cannot
be calculated. Anyway, it is a biased estimate of the true distance
(Brown et al. 1997). Thus, it seems impossible to use it for a direct
calculation of the zero-point.
On the other hand, rejecting negative parallaxes
generates a Lutz-Kelker bias type
(Lutz \& Kelker 1973) while rejecting parallaxes with
large $\sigma_{\pi}/{\pi}$
generates another bias (Brown et al. 1997).
In order to bypass this problem, FC suggest
calculating $\rho$ from the weighted mean of the function:
\begin{equation}
10^{0.2\rho} = 0.01 \pi 10^{0.2(\langle V_0 \rangle - \delta \log P)}
\label{rho}
\end{equation}
This treatment assumed that the exponent of a mean is identical to the
mean of the exponents. FC justify it by saying that ``the scatter about
the PL(V) relation is relatively small''.
They chose a weighting and compute the mean
of $10^{0.2 \rho}$, from which they derive $\rho$. \\
As a matter of fact, they use a Period-Color (PC) relation for dereddening their magnitudes.
Because of the near degeneracy of the reddening slope and the colour term in a
Period-Luminosity-Color relation, this technique will have much the same narrowing effect
on the PL relation as including a color term would. For a Cepheid of known distance
the scatter is reduced from 0.2 down to about 0.1.
However since the HIPPARCOS parallaxes may have large errors, we see from equation
\ref{rho} that the scatter in $10^{0.2\rho}$ could be increased in this manner.
Precisely, we would like to answer the following questions:
\begin{itemize}
\item Can we obtain a good result by rejecting poor parallaxes?
\item Is the dispersion small enough to justify the calculation of $\rho$ using the
mean of $10^{0.2 \rho}$?
\item Is the final result biased or not?
\item Is it possible to adopt another weighting than that of FC?
\end{itemize}
In section 2 we use the HIPPARCOS sample of Cepheids to confirm that
rejecting negative parallaxes or parallaxes with a poor $\sigma_{\pi} / \pi$
gives a biased zero-point and to test the FC method with different
weighting systems.
This suggests making a simulated sample for
which the zero-point is {\it a priori} known and then to apply the same
treatment to it.
In section 3 we explain how the simulated sample is built in order to
reproduce all the properties of the true HIPPARCOS sample.
Then, in section 4 we give the result of the FC method applied to the simulated
sample with different weightings. This shows that the calculated zero-points
and the associated standart deviations depend on
the adopted weigthing.
In section 5, the previous results are discussed
and explained. The consequences are drawn for estimating
the best zero-point from the HIPPARCOS Cepheid sample for both V and I bands.
\section{Use of the HIPPARCOS Cepheid sample}
The complete Cepheid sample is extracted from the catalogue
HIPPARCOS (1997). Among all variable stars, we keep only those labelled DCEP
(classical $\delta-$type Cepheids) and DCEPS (first overtone pulsators),
and then obtained a total of 247 Cepheids. The period of the 31 overtone pulsators
is converted to the fundamental period $P$ according to Alcock et al. (1995):
\begin{equation}
P_{1}/P=0.716 - 0.027 \log P_{1}
\end{equation}
The $B$ and $V$ photometry is available from the David Dunlap Observatory Galactic Cepheid
Database (Fernie et al. 1995), except for nine Cepheids
(CK Cam, BB Gem, KZ Pup, W Car, DP Vel, BB Her, V733 Aql, KL Aql and V411 Lac)
which were excluded from the present study.
Therefore, the final sample (table 4) is made of 238 Cepheids (31 overtones).
The color excess is then calculated using the FC method,
i.e. calculation of the intrinsic color $\langle B \rangle _0- \langle V
\rangle _0$ from a linear relation
color vs. $\log P$, according to Laney \& Stobie (1994):
\begin{equation}
\langle B \rangle _0- \langle V \rangle _0 = 0.416 \log P +0.314.
\label{laney}
\end{equation}
We use the relation from Laney \& Stobie (1993) to compute the V extinction :
\begin{equation}
R_{V} = 3.07 + 0.28 (\langle B \rangle _0- \langle V \rangle _0) + 0.04 E_{(B-V)}
\label{laney2}
\end{equation}
Figure \ref{cut} shows how the quantity $10^{0.2 \rho}$ varies with the apparent
magnitude $V$. This quantity is directly
needed for the calculation of the zero-point $\rho$. Clearly, the dispersion
increases with the magnitude, but the distribution is quite symetrical
around a given value.
If a cut is applied on the sample to reject negative parallaxes
(filled triangles in figure \ref{cut}) the mean of $10^{0.2 \rho}$ is
overestimated. If one uses only measurements with
$0 < \sigma_{\pi} / \pi < 0.5 $
(open circles in figure \ref{cut}), again, $10^{0.2 \rho}$ is overestimated.
Thus, as claimed by Brown et al. (1997), a bias is clearly
confirmed if one cuts the sample. We will no more consider cuts involving parallaxes
as a way of obtaining a valuable result.
\begin{figure}
\epsfxsize=8.5cm
\epsfbox{cut.eps}
\caption{Effects of cuts. The horizontal line corresponds to the
zero-point value $\rho = -1.43$.
If one rejects negative parallaxes (filled triangles) or keeps parallaxes with
$0 \le \sigma_{\pi} / \pi \le 0.5$ (open circles), the mean is overestimated.
}
\label{cut}
\end{figure}
Figure \ref{cut} does not exhibit
a small dispersion. So, we do not know if the FC's
procedure leads to the proper value of $\rho$.
For the calculation of the mean of $10^{0.2 \rho}$ they
use individual weights taken as the reciprocal of the square of the
standard error of the second term of equation \ref{rho}.
For a given Cepheid, the weight is given by:
\begin{equation}
\omega_{i} \approx [10^{-2} \sigma_{\pi_{i}} 10^{0.2(\langle V_{0_{i}} \rangle - \delta \log P_{i})}]^{-2}
\label{wfc}
\end{equation}
because the error on the term $10^{0.2(\langle V_{0_{i}} \rangle
- \delta \log P_{i})}$ is negligible as shown by FC.
This weighting is mathematically the most rigorous. However, some
other empirical weightings may be worthy of interest.
Since the error on $\rho$ is mainly due to the large uncertainty $\sigma_{\pi}$,
we will test a weight in $\sigma_{\pi_{i}}^{-2}$ and in $(\sigma_{\pi_{i}}/\pi_i)^{-2}$.
Further, we will also use an unweighted mean because the dispersion looks quite
symetrical around a mean value and a $V^{-2}$ weighting because the dispersion increases
with V.
We then repeated the FC tests as well as the other weightings and found the results
given in table \ref{weight}.
\begin{table}
\caption{Values of $\rho$ calculated with different weightings and different
cuts in $V$ magnitude. The standard deviation of each value is given in
parenthesis.}
\begin{tabular}{lrrr}
\hline
Weighting & All $V$ & $V \leq 8$ & $V \leq 6$ \\
\hline
F\&C & $ -1.45(0.10)$ & $ -1.47(0.10)$ & $ -1.45(0.08)$ \\
$\sigma_{\pi_{i}}^{-2}$ & $ -1.04(0.37)$ & $ -1.38(0.22)$ & $ -1.45(0.16)$ \\
$(\sigma_{\pi_{i}}/\pi_i)^{-2}$ & $ 1.19(0.57)$ & $ - $ & $ - $ \\
No weight & $ -0.19(0.74)$ & $ -1.38(0.22)$ & $ -1.40(0.17)$ \\
$V^{-2}$ & $ -0.64(0.65)$ & $ -1.41(0.22)$ & $ -1.42(0.13)$ \\
\hline
\label{weight}
\end{tabular}
\end{table}
From this table we see that, when all Cepheids are used, the calculated
zero-point $\rho$ strongly depends on the adopted weighting.
The instability of this result can be explained by the very large dispersion
at large $V$. This large dispersion quite justifies the second question of
section 1.
According to the shape of figure \ref{cut}, we see that the dispersion can be
reduced by cutting the sample at a given apparent magnitude. Table \ref{weight} shows that such
a cut gives a more stable result.
Moreover, the weighting adopted by FC gives the lowest dispersion.
For instance, keeping the brightest 11 Cepheids, we obtain $\rho=-1.45$
with a very small standard deviation of 0.05 ($V \leq 5.5$). We also try to
keep only stars with the highest weights (whatever the weighting system). However, that leads us to
the same results with slightly higher dispersions.
In practice, we have no means of knowing if a bias has been introduced
as long as the observed sample is used because the true zero-point is
not known. Only a simulated sample, with a zero-point {\it a priori} known,
can provide the answer to the third question of section 1.
This justifies the construction of simulated samples.
\section{Construction of simulated samples}
To build a simulated sample only three quantities have to be drawn independently:
\begin{itemize}
\item The parallax $\pi$
\item The logarithm of the period $\log P$
\item The column density of interstellar matter along the line of sight.
\end{itemize}
\subsection{The simulated ``true parameters''}
Assuming a homogeneous 3D distribution of galactic
Cepheids (this is justified owing to
relatively small depth of HIPPARCOS survey regarding the
depth of the galactic disk),
we draw at random the x,y,z coordinates over the range
[-2100, 2100] pc. We keep only
Cepheids within a radius of 2100 pc and then deduce the true parallax:
\begin{equation}
\pi = 1/ \sqrt{x^2+y^2+z^2}
\end{equation}
250 true parallaxes are drawn in such a way. Each point will
be a Cepheid in our simulated sample.\\
Then, for each Cepheid we draw $\log P$ following a distribution which reproduces
the observed distribution of periods (Fig. \ref{h_logp}a and \ref{h_logp}b).
We then calculate the absolute magnitude $\langle M_V \rangle$
from the relation:
\begin{equation}
\langle M_V \rangle = \delta_V \log P + \rho_V + \Delta
\end{equation}
where $\delta_V = -2.77$ is the adopted slope as said in the introduction,
$\rho_V= -1.30 $ is the
arbitrarily fixed zero-point and $\Delta$ is a Gaussian
intrinsic dispersion ($\langle \Delta \rangle =0$ ; $\sigma(\Delta)=0.2$)
which reflects the width of the instability strip.
The absolute magnitude in B-band $\langle M_B \rangle$ is calculated
in the same way using
the same intrinsic dispersion multiplied by 1.4.
We reproduce in this manner the correlation of the residuals
as well as the dispersion of the true CP relation related to
the color variation across the instability strip. We chose
$\delta_B=\delta_V+0.416$ and $\rho_B=\rho_V + 0.314$,
so that it implies the relation between the intrinsic
color $\langle B \rangle _0 - \langle V \rangle _0 $ and
$\log P$ from Laney \& Stobie (1994):
\begin{equation}
\langle B \rangle _0 - \langle V \rangle _0 = (\delta_B - \delta_V) \log P
+ \rho_B - \rho_V
\label{color}
\end{equation}
The true intrinsic color $\langle B \rangle _0 - \langle V \rangle _0$
is calculated from this linear relation. We then reduce the dispersion
of the PL relation down to 0.1 as already explained in the introduction.
The relation of $E_{(B-V)}$ versus the calculated photometric
distances (adopting, for instance, distances from Fernie et al. 1995) shows (Fig. \ref{ebmv_dist})
that the observed Cepheids are located in
a sector. All line of sight directions have extinction (no point below
the dashed line). In slightly obscured directions (dashed line) one can see
stars up to $\approx 5000$ pc, while in very obscured regions (dotted line)
the closest Cepheids are detected not farther than $\approx 1100$ pc.
\begin{figure}
\epsfxsize=8.5cm
\hbox{\epsfbox{ebmv_dist.eps}}
\caption{Color exces versus photometric distances from Fernie (Fernie et al.
1995) for HIPPARCOS Cepheids.
The slope $E_{(B-V)} / distance $ measures the density of the interstellar
medium. In slightly obscured directions (dashed line)
one can see stars up to $\approx 5000$ pc, while in very obscured regions
(dotted line) the closest Cepheids are detected not farther than $\approx 1100$ pc.
}
\label{ebmv_dist}
\end{figure}
The slope $E_{(B-V)} / distance $ is a
measure of the density of the interstellar
medium in a given direction. This density varies over a large range
due to the patchiness of the galactic extinction, but, for a given
line of sight, the extinction, and thus the color excess, is assumed to be
proportional to the distance. This figure is used to obtain the extinction
for each Cepheid. We draw at random the slope $E_{(B-V)} / distance $ over
the range defined by the dashed and dotted lines (Fig. \ref{ebmv_dist}).
Using the true distance $1/\pi$ we then deduce the true color excess
$E_{(B-V)}$, and the true extinctions:
\begin{equation}
A_V= R_V E_{(B-V)}
\end{equation}
\begin{equation}
A_B= R_B E_{(B-V)}
\end{equation}
with $R_V$ = 3.3 and $R_B = 4.3$.
\subsection{The simulated ``observed parameters''}
Now we calculate the parameters which would be observed.
First, the apparent $B$ and $V$ magnitudes
are simply:
\begin{equation}
\langle V \rangle = 5 \log (1/\pi) - 5 + \langle M_V \rangle + A_V + \epsilon_V
\end{equation}
\begin{equation}
\langle B \rangle = 5 \log (1/\pi) - 5 + \langle M_B \rangle + A_B + \epsilon_B
\end{equation}
where $\epsilon_V$ and $\epsilon_B$ are two independent Gaussian variables
which reproduce measurement uncertainties (the intrinsic
scatter of the PL relation
is already counted in $\langle M_V \rangle $ and $\langle M_B \rangle $). We adopted for both:
$\langle \epsilon \rangle = 0.0$ and $\sigma_{\epsilon} = 0.005$.
The parallax which would be observed is calculated from the true one
and an associated $\sigma_{\pi}$ obtained through
the figure \ref{spi_v}a. This figure shows two populations: one below the
dotted line, the other about the dotted line.
First, we draw the membership to one of these families in the right proportion.
Then, from the linear relationships of the corresponding family
and the $V$ magnitude already computed, we calculate
$\log \sigma_{\pi}$ (i.e. $\sigma_{\pi}$). Finally, the
observed $\pi$ is obtained
by drawing one occurence in the Gaussian distribution $(\pi, \sigma_{\pi})$.\\
Concerning the observed color excess, it will simply
be deduced from the relation:
\begin{equation}
E_{(B-V)}= \langle B \rangle - \langle V \rangle -
(\langle B \rangle _0 - \langle V \rangle _0)
\label{exces}
\end{equation}
with $\langle B \rangle _0 - \langle V \rangle _0$ deduced from the PC
relation \ref{color} as we did in section 2.
We also need to determine the observed value of the coefficient $R_V$.
We draw its value according to a Gaussian distribution centered on the chosen
true value (3.3) with a dispersion of 0.05. So, we suppose that the observed
value has no systematic shift with respect to the true value.
Finally, in order to reproduce selection effects like the Malmquist bias (Malmquist 1920)
we reject the Cepheids which could not be observed according to their
apparent magnitudes (i.e. their probability to be detected). We draw a
random parameter $t \in [0, 1]$ and compute the quantity:
\begin{equation}
t_0 = \frac{1}{1+\exp^{\alpha (\langle V \rangle - \langle V_{lim} \rangle)}}
\end{equation}
Whenever $t \le t_0$ the star may be observed by HIPPARCOS and we keep
it in our sample, and in the other case it will be rejected. We assume
$\alpha = 1$ and $\langle V_{lim} \rangle = 12.5$. Moreover, whenever
$ \langle V \rangle \le 1.9$, the Cepheid would be too bright
(unrealistic apparent magnitude) and then rejected. The number of
simulated Cepheids is then almost equal to the true one.
In order to show that the simulated sample is comparable to the
true HIPPARCOS one, we plot for one simulated sample the
same figures (Fig. \ref{h_logp} to \ref{rho_v}) as those produced
with the true HIPPARCOS sample. Note that the figures from the
simulated sample are made from a single drawing which is not necessarily
an optimal representation of the true sample.
\section{RESULTS}
The result may depend on the particular sample we draw. In order to reduce the
uncertainty due to this choice, we made 1000 different random drawings
(each of them with about 240 Cepheids) and adopted the mean result.
We obtain the result shown in the table 2 (let us recall that the input zero-point
is $\rho_{V}=-1.30$).
\begin{table}
\caption{Values of $\rho$ calculated using 1000 simulated samples. We used different
weightings and different cuts in $V$ magnitude as in the study made with the true sample.
The standard deviation of each value is given in parenthesis.}
\begin{tabular}{crrr}
\hline
weighting & All $V$ & $V \leq 8$ & $V \leq 6$ \\
\hline
true zero-point & $ -1.30 $ & $ -1.30 $ & $ -1.30 $ \\
\hline
F\&C & $ -1.31(0.14)$ & $ -1.30(0.15)$ & $ -1.31(0.21)$ \\
$\sigma_{\pi_{i}}^{-2}$ & $ -1.33(0.21)$ & $ -1.31(0.22)$ & $ -1.31(0.26)$ \\
$(\sigma_{\pi_{i}}/\pi_i)^{-2}$ & $ 0.03(0.47)$ & $ - $ & $ - $ \\
No weight & $ -1.36(0.43)$ & $ -1.33(0.39)$ & $ -1.32(0.34)$ \\
$V^{-2}$ & $ -1.33(0.33)$ & $ -1.32(0.32)$ & $ -1.31(0.29)$ \\
\hline
\end{tabular}
\end{table}
The simulation clearly confirms that the weighting in
$(\sigma_{\pi_{i}}/\pi_i)^{-2}$ is meaningless.
Again, it is confirmed that a cut in magnitude gives more stable results
because the method of averaging $10^{0.2 \rho}$ to get $\rho$ is better
justified with small dispersion.
This answers the second question of section 1.
The simulation also confirms that the FC weighting leads to the lowest dispersion
and that the results are too low at only a 0.02 or 0.01 mag. level.
In order to analyze the effect of the measurement errors,
we progressively reduce the observational
errors (but not the intrinsic dispersion) introduced in our simulation. The reduction
is made from their realistic values down to zero. We compute the mean value of the
distribution of $\rho$ as we go along, and plot the results in figure~\ref{rho_err}.
It appears that the zero-point values comes closer to the
real value $\rho = -1.30$. Moreover, the FC weighting gives clearly the
more stable result.
The trends of figure \ref{rho_err} (decreasing of $\rho$ with increasing errors)
can be explained solely by errors on $\pi_{i}$ because they disappear when
$\sigma_{\pi_{i}}$ is forced to zero.
Further, we checked that removing both the measurement errors and the intrinsic dispersion
removes the residual shift for all kinds of weighting and gives back the initial
value $\rho = -1.30$.
This proves that our simulation procedure works well.
\begin{figure}
\epsfxsize=8.5cm
\hbox{\epsfbox{rho_err.eps}}
\caption{Zero-point values for each weighting when the observational error
is progressively reduced from its realist value down to zero.}
\label{rho_err}
\end{figure}
\section{DISCUSSION}
The results of the previous section allow us to answer the
questions of section 1:
a cut in apparent magnitude reduces the dispersion and
gives reliable results because averaging $10^{0.2 \rho}$ works
better with small dispersion.
Whatever the weighting adopted, the zero-point is not biased by more than $0.03$ mag.
The FC weighting gives the
smallest standard deviation, and the systematic shift never exceeds $0.01$ mag.
Let us analyze the main effects which are responsible for a shift. Two effects are
present: effect of averaging in $10^{0.2 \rho}$ and Malmquist effect. We will see
that they work in two opposite directions.
Consider two Cepheids comparable in every aspect,
i.e. located at the same distance in two
directions with the same interstellar absorption, measured with the same
$\sigma_{\pi}$ so that they have the same observed parallaxes,
and both with the same periods, but one located near one edge of the
instability strip whereas the second is located at the opposite edge.
Their absolute magnitudes would then be:
\begin{equation}
\langle M_V^1 \rangle = \delta \log P + \rho + \zeta
\end{equation}
\begin{equation}
\langle M_V^2 \rangle = \delta \log P + \rho - \zeta
\end{equation}
where $\zeta$ is the actual value of the intrinsic dispersion
($\langle \Delta \rangle =0$ ; $\sigma(\Delta) \approx 0.2$) across
the instability strip. \\
When using these two Cepheids to compute the zero-point of the PL relation
directly from the parallax we would obtain:
\begin{equation}
\rho_1 = \langle V_{0} \rangle + \zeta' + 5 \log \pi -10 - \delta \log P
\label{rho1}
\end{equation}
\begin{equation}
\rho_2 = \langle V_{0} \rangle - \zeta' + 5 \log \pi -10 - \delta \log P
\label{rho2}
\end{equation}
with $\zeta' \le \zeta$ because of dereddening method.
The mean of the two values corresponds then to the true value $\rho$ since:
\begin{equation}
\frac{\rho_1 + \rho_2}{2} = \rho_{true}
\end{equation}
However we have shown why such a direct mean cannot be used with HIPPARCOS data.
So we compute the mean (or the weighted mean) of the two quantities
$10^{0.2 \rho_1}$ and $10^{0.2 \rho_2}$.
The mean zero-point $\overline{\rho}$ can be expressed as:
\begin{equation}
\overline{\rho} = \rho_{true} +
5 \ \log [\frac{{\omega_1}/{\omega_2}10^{0.2 \zeta'} + 10^{- 0.2 \zeta'}}
{1 + {\omega_1}/{\omega_2}} ]
\end{equation}
where $\omega_1$ and $\omega_2$ are the weights of the two quantities.
If we adopt $\omega_1/\omega_2=1$ (i.e. no weighting or same weights) and $\zeta'=0.2$
(overvalued in order to highlight the way $\rho$ is biased)
we obtain $\overline{\rho} = \rho_{true} + 0.01$. The observed $\rho$ slighly increases
in this manner.
The Malmquist effect works in the other direction. The biased absolute magnitude $\overline{M'}$
is too bright (Teerikorpi 1984) :
\begin{equation}
\overline{\langle M' \rangle} = \overline{\langle M \rangle} - 1.38 \sigma^{2}
\end{equation}
where $\overline{\langle M \rangle}$ is the unbiased magnitude. This formula gives the global
correction, not the correction for individual Cepheid. Assuming a pessimistic value $ \sigma = 0.2$
(once again, since the use of the PC relation as a narrowing effect, $ \sigma $ is surely lower than
this value) the shift would be at worst $-0.055$. Then the observed $\rho$ diminishes.
Finally, the net shift would be $-0.04$ or less. However, figure \ref{rho_err} which reproduces both
effects with realistic uncertainties gives a shift of $\rho_{observed} = \rho_{true} - 0.01$
when the FC weighting is used. This shift takes in account these two effects. One will then apply
it on the value deduced from HIPPARCOS data.
We investigate now the effect of a change in the adopted slope. We adopted $\delta = -2.77 \pm 0.08$.
What would be the change in the PL relation if the true slope was different from this value?
In table \ref{slope} we give the values of the mean $\langle M_{V} \rangle $ deduced from our simulation, the
input relations being :
\begin{equation}
\langle M_V \rangle + 4.07 = -2.77 (\log P - 1)
\end{equation}
or
\begin{equation}
\langle M_V \rangle + 3.74 = -2.77 (\log P - 0.88)
\end{equation}
We note that the absolute magnitude at $\log P = 1$ (or $\log P = 0.88$) doesn't change very much
(less than $0.03$) as far as the $\log P$ does not change from the mean of calibrating Cepheids.
\begin{table}
\caption{Effect of the chosen slope on the final magnitudes computed at the mean $\log P$ ($0.88$)
and at $\log P = 1$}
\begin{tabular}{lll}
\hline
slope & $\langle M_V \rangle$ at $\log P$ mean & $\langle M_V \rangle$ at 10 d \\
\hline
-2.60 & -3.76 & -4.07 \\
-2.70 & -3.75 & -4.08 \\
-2.77 & -3.75 & -4.08 \\
-2.80 & -3.75 & -4.08 \\
-2.90 & -3.74 & -4.09 \\
-3.00 & -3.74 & -4.10 \\
\hline
Reference values& & \\
\hline
-2.77 & -3.74 & -4.07 \\
\hline
\label{slope}
\end{tabular}
\end{table}
\section{CONCLUSION}
The conclusion is that the intrinsic dispersion (even Gaussian and symetrical) of
the instability strip is responsible for too low values of $\rho$ and may lead
to a slightly biased result as long as the zero-point $\rho$ is deduced
by averaging $10^{0.2 \rho}$. However it is compensated by the Malmquist bias, and,
using a PC relation for dereddening the individual Cepheids,
the final effect is globally very small. Indeed, our simulation
shows that it is almost negligible (Fig. \ref{rho_err}) even when we
account for measurement errors.
With realistic measurement errors the bias is about $-0.01$.
A cut in apparent magnitude reduces the uncertainty on the zero-point.
The best unbiased zero-point is obtained by cutting the sample at $V \leq 5.5$ mag.
The result is (after correction of the residual shift of $-0.01$):
\begin{equation}
\rho = -1.44 \pm 0.05 \ \ \ \ \ \ (n=11)
\label{result}
\end{equation}
for a slope $\delta_V = -2.77 \pm 0.08$ and a weighted
mean $\langle \log P \rangle=0.82 $. The adopted V-band PL relation is then
$ \langle M_V \rangle = -2.77 \pm 0.08 \ \log P - 1.44 \pm 0.05$ or
$ \langle M_V \rangle + 4.21 = -2.77 \ (\log P - 1)$.
\section*{Acknowledgements}
We thank L. Szabados for communicating his list of binary Cepheids and the referee
for his valuable comments.
\section*{APPENDIX A}
At present, the Hubble Space Telescope has observed Cepheids in 19 galaxies
(see Lanoix et al. 1999b for an extensive compilation). These observations
are made in two bandpasses (V and I), so that we need a calibration of the PL relation both
in V and I to apply a dereddening procedure (see Freedman et al. 1994, for instance) and compute the distance
moduli of these galaxies.
With this aim in view for a future paper, we then perform the I calibration based on HIPPARCOS parallaxes
in the light of our V calibration.
The major problem is that there's no homogeneous I photometry available for each Cepheid
of the calibrating sample, and that a selection may induce a biased result.
As a matter of fact we found I (Cousins) photometry for 174 Cepheids of the sample from Caldwell \& Coulson
(1987). We apply to these values
a tiny correction (0.03 mag) in order to convert them into intensity averaged magnitudes.
The I magnitudes of these stars are listed in table 4 when available.
Since the selection doesn't come from a rough cut in the HIPPARCOS sample, it will not necessarily lead us
to a biased result.
We then apply the same selection to the V sample and compute again the visual
zero-point. From these 174 cepheids we obtain:
\begin{equation}
\rho = -1.49 \pm 0.10
\end{equation}
This result is almost identical to the one obtained with the complete sample (Eq. \ref{result}), so that we conclude that this selection
implies a little bias of $0.04$ with respect to the complete sample and only $0.05$ with respect to the
adopted final value. We will take it into account to determine
the associated I zero-point.
The residuals of the I and the V PL relations are correlated so that we will apply the same procedure as we do for the
V band and obtain the same narrowing effect of the instability strip. We then need the slope of the I PL relation
as well
as the I ratio of total to selective absorption. Concerning the slope that is well determined, we choose
$\delta_I = -3.05$ (see Gieren et al. 1998, Madore \& Freedman 1991 for instance). Let's recall that the influence
of a variation of the slope is very weak.
Concerning $R_{I}$, we choose according to Caldwell \& Coulson (1987):
\begin{equation}
R_{I} = 1.82 + 0.20 (\langle B \rangle _0- \langle V \rangle _0) + 0.02 E_{(B-V)}
\end{equation}
The calculus leads to :
\begin{equation}
\rho_{I} = -1.84 \pm 0.09
\end{equation}
\begin{figure}
\epsfxsize=8.5cm
\hbox{\epsfbox{rho_i.eps}}
\caption{Position of the 174 remaining Cepheids in the I diagram zero-point vs. magnitude. The horizontal line
corresponds to $\rho_{I} = -1.84$.}
\label{rho_i}
\end{figure}
Figure \ref{rho_i} shows that these 174 Cepheids still have an almost symetrical
distribution around a mean value, and that only faint stars with low weights
have been rejected from the sample.
That may explain why the result is only slightly biased.
Keeping in mind that the instability strip is narrower in I than in V band, the bias du to the selection of this
sample should be less than $0.04$ mag. Finally we adopt:
\begin{equation}
\rho_{I} = -1.81 \pm 0.09
\end{equation}
for a slope $\delta_I = -3.05$.
\section*{APPENDIX B}
We also investigated the effect of binarity as pointed out by Szabados (1997).
We indeed found that the dispersion of the zero-point is reduced when
only non-binary Cepheids are used. However, we interpreted this effect
by the fact that confirmed non-binary Cepheids are brighter. Actually,
using either non-binary (Evans 92) or binary (Szabados private communication)
Cepheids does not affect significantly the value of the zero-point.
|
\section{Introduction} \label{s1}
In order to facilitate a description of the content
of this paper we briefly introduce the notation
used throughout this manuscript. The open
complex upper half-plane is abbreviated by
${\mathbb{C}}_+=\{z\in{\mathbb{C}}\,|\,\Im(z)>0\};$ the symbols ${\mathcal H},
{\mathcal K}$ represent complex separable Hilbert spaces and
$I_{\mathcal H}$ the corresponding identity operator in ${\mathcal H}$.
Moreover, we denote by
${\mathcal B}({\mathcal H})$,
${\mathcal B}_\infty({\mathcal H})$, the Banach spaces of bounded and
compact operators in ${\mathcal H}$ and by ${\mathcal B}_p({\mathcal H})$,
$p\geq 1$ the standard Schatten-von Neumann trace ideals
(cf., \cite{GK69}, \cite{Si79}). Real and
imaginary parts of an operator $T$ with
$\text{\rm{dom}}(T)=\text{\rm{dom}}(T^*)$ are
defined as usual by $\Re(T)=(T+T^*)/2$ and $\Im(T)=
(T-T^*)/(2i);$ the spectrum, essential and absolutely
continuous spectrum of $T$ is abbreviated by
$\text{\rm{spec}}(T)$, $\text{\rm{ess.spec}}(T)$, and $\text{\rm{ac.spec}}(T)$, respectively.
For a self-adjoint operator $H=H^*$ in
${\mathcal H}$, the associated family of strongly
right-continuous orthogonal spectral
projections of $H$ in ${\mathcal H}$ is denoted by
$\{E_H(\lambda)\}_{\lambda\in{\mathbb{R}}}$.
Before describing the content of each section, we briefly
summarize the principal results of this paper. Let $H_0$ and
$H=H_0+V$ be self-adjoint operators in ${\mathcal H}$ with
$V=V^*\in{\mathcal B}_1({\mathcal H})$ and denote by $\xi(\lambda,H_0,H)$
Krein's spectral shift function associated with the pair
$(H_0,H)$, uniquely defined for a.e.~$\lambda\in{\mathbb{R}}$ by
$\xi(\cdot,H_0,H)\in L^1({\mathbb{R}};d\lambda)$ and
\begin{equation}
\tr((H-z)^{-1}-(H_0-z)^{-1})=-\int_{\mathbb{R}} d\lambda\,
\xi(\lambda,H_0,H)(\lambda -z)^{-2}, \quad
z\in{\mathbb{C}}\backslash{\mathbb{R}}. \label{1.1}
\end{equation}
Denoting by $\text{\rm{index}} (P,Q)$ the index of a Fredholm pair of
orthogonal projections in ${\mathcal H}$, one of our principal
results represents $\xi(\lambda,H_0,H)$ as an averaged
index for the Fredholm pair of projections
$(E_{J+A(\lambda)+tB(\lambda)}((-\infty,0)),
E_J((-\infty,0)))$, $t\in{\mathbb{R}}$, as follows (cf.
Theorem~\ref{main}),
\begin{align}
&\xi(\lambda, H_0,H)=\frac{1}{\pi} \int_{\mathbb{R}} dt
\frac{\text{\rm{index}} \big(E_{J+A(\lambda)+tB(\lambda)}
((-\infty,0)),E_J((-\infty,0))\big )}{1+t^2}
\label{1.2}
\end{align}
for a.e.~$\lambda\in{\mathbb{R}}$. Here $V=H-H_0$ is decomposed as
\begin{equation}
V=KJK^*, \quad K\in{\mathcal B}_2({\mathcal H}), \,\, J=\text{\rm{sgn}}(V) \label{1.3}
\end{equation}
and
\begin{align}
&A(\lambda)=\nlim_{\varepsilon \downarrow 0}
\Re(K^*(H_0-\lambda-i\varepsilon)^{-1}K), \label{1.4} \\
&B(\lambda)=\nlim_{\varepsilon \downarrow 0}
\Im(K^*(H_0-\lambda-i\varepsilon)^{-1}K)
\text{ for a.e. } \lambda\in{\mathbb{R}}. \label{1.5}
\end{align}
In the special sign-definite case where $J=\pm I_{\mathcal H}$, formula
\eqref{1.2} implies a result of Pushnitskii \cite{Pu97}
(cf. Corollary~\ref{c5.6}).
Next, observing that the trace class-valued operator
$K^*(H_0-z)^{-1}K$, $z\in{\mathbb{C}}_+$ has nontangential boundary
values $K^*(H_0-\lambda-i0)^{-1}K$ for a.e.~$\lambda\in{\mathbb{R}}$
in ${\mathcal B}_p({\mathcal H})$-topology for each $p>1$, but in general
not in the trace norm ${\mathcal B}_1({\mathcal H})$-topology, we
introduce the notion of a trindex, $\text{\rm{trindex}} (\cdot,\cdot)$,
for a pair of operators $(A,Q)$ in ${\mathcal H}$, where
$A\in{\mathcal B}({\mathcal H})$ is bounded and $Q=Q^*=Q^2$ is an
orthogonal projection in ${\mathcal H}$, as follows: the pair
$(A,Q)$ is said to have a {\it trindex}, denoted by $\text{\rm{trindex}}
(A,Q)$, if there exists an orthogonal projection $P$ in
${\mathcal H}$ such that $(A-P)\in{\mathcal B}_1({\mathcal H})$ and $(P,Q)$ is a
Fredholm pair of orthogonal projections in ${\mathcal H}$. In this
case one then defines,
\begin{equation}
\text{\rm{trindex}} (A,Q)=\tr(A-P)+\text{\rm{index}} (P,Q). \label{1.6}
\end{equation}
Introducing the spectral shift operator
$\Xi(T)=(1/\pi)\Im(\log(T))$ for a bounded dissipative
operator $T$, with $T^{-1}$ also bounded in ${\mathcal H}$, our second
principal result (cf. Theorem~\ref{t5.8}) identifies
$\xi(\lambda,H_0,H)$ with the trindex of the pair
$(\Xi(J+K^*(H_0-\lambda-i0)^{-1}K),\Xi(J))$, that is,
\begin{equation}
\xi(\lambda,H_0,H)=\text{\rm{trindex}}(\Xi(J+K^*(H_0-
\lambda-i0)^{-1}K), \Xi(J)) \text{ for a.e. }
\lambda \in{\mathbb{R}}. \label{1.7}
\end{equation}
The trindex representation \eqref{1.7} paves the way for
introducing a generalized
spectral shift function to be discussed in
detail in Section~\ref{s5}. We also show that
the averaging formula
\eqref{1.2} should be viewed as a generalized Birman-Schwinger
principle
\begin{equation}
\xi(\lambda, H_0,H)= \frac{1}{\pi}\int_{\mathbb{R}} dt\,
\frac{\widehat\xi(0_-,J+A(\lambda)+tB(\lambda),J)}{1+t^2}
\text{ for a.e. } \lambda \in {\mathbb{R}}, \label{1.5a}
\end{equation}
where $\widehat\xi(\cdot,J+A(\lambda)+tB(\lambda),J)$
denotes the
generalized spectral shift function associated with the pair
$(J+A(\lambda)+tB(\lambda),J)$, $t\in {\mathbb{R}}$.
Next we turn to a description of the content of each
section. In Section~\ref{s2} we briefly review basic
properties of the index of a Fredholm pair of projections in
${\mathcal H}$ (mainly following \cite{ASS94}, see also
\cite{AS94}, \cite{Ka97}, \cite{Ka55}) and then present a
discussion of the notion of trindex and some of its
properties. Section~\ref{s3} is devoted to the concept of a
$\Xi$-operator as recently introduced in \cite{GMN99} and
further discussed in \cite{GM99}. More precisely, if $T$ is
a bounded dissipative operator with $T^{-1}\in{\mathcal B}({\mathcal H})$,
then $\Xi(T)$ is defined by
\begin{equation}\label{1.8}
\Xi(T)=\frac{1}{\pi} \Im (\log (T)).
\end{equation}
Section~\ref{s3} also studies sufficient conditions on
$A=A^*\in{\mathcal B}_\infty({\mathcal H})$ and $0\leq B\in
{\mathcal B}_1({\mathcal H})$ to
guarantee $(\Xi(S+A+iB)-\Xi(S+A))\in{\mathcal B}_1({\mathcal H})$ for
given $S=S^*\in{\mathcal B}({\mathcal H})$. Section~\ref{s4}, the
technical core of this paper, provides a discussion of
averaged Fredholm indices. Introducing the family of normal
trace class operators,
\begin{equation}
{\mathcal A} (z)=B^{1/2}(S+zB)^{-1}B^{1/2}, \quad
z\in{\mathbb{C}}\backslash{\mathbb{R}}, \label{1.9}
\end{equation}
where $S=S^*\in{\mathcal B}({\mathcal H})$, $S^{-1}\in{\mathcal B}({\mathcal H})$, $0\leq
B\in{\mathcal B}_1({\mathcal H})$,
associated with the (dissipative) family of operator-valued
Herglotz functions
\begin{equation}
T(z)=S+zB, \quad z\in{\mathbb{C}}_+ \label{1.10}
\end{equation}
(i.e., $T$ is analytic in ${\mathbb{C}}_+$ and $\Im(T(z))\geq 0$
for $z\in{\mathbb{C}}_+$), we prove
\begin{equation}\label{1.11}
\tr (\log(T(z))-\log(S))=
\sum_{k=1}^\infty m_k \int_0^1 d\tau\, \frac{z
\lambda_k}{1+\tau z \lambda_k},
\end{equation}
(cf.~Theorem~\ref{tt.2}), where $\{\lambda_k \}_{k\in{\mathbb{N}}}
\subset {\mathbb{R}}$ is the set of eigenvalues with associated
multiplicities $\{m_k\}_{k\in{\mathbb{N}}}$ of the self-adjoint trace
class operator $B^{1/2}S^{-1}B^{1/2}$. This then yields our
principal result (Theorem~\ref{ttr.8}) relating the trindex
of the pair $(\Xi(S+A+iB), \Xi(S))$ and the averaged Fredholm
index $n(t)=\text{\rm{index}} (\Xi(S+A+tB),\Xi(S))$, $t\in{\mathbb{R}}$, as
\begin{equation}\label{1.12}
\text{\rm{trindex}}
(\Xi( S+A+iB),
\Xi(S))=
\frac{1}{\pi}\int_{\mathbb{R}} \frac{dt\,n(t)}{1+t^2}.
\end{equation}
Moreover, we provide a version of the Birman-Krein
formula relating the left-hand side of
\eqref{1.12} and the determinant of the abstract scattering matrix
\begin{equation}
\exp (-2\pi i\, \text{\rm{trindex}}
(\Xi( S+A+iB),
\Xi(S)))=\det (I_{\mathcal H}-2iB^{1/2}(S+A+iB)^{-1}B^{1/2}).
\end{equation}
Section~\ref{s5} finally presents the applications of our
formalism to Krein's spectral shift function
$\xi(\lambda,H_0,H)$ as discussed in \eqref{1.2},
\eqref{1.7}, and \eqref{1.5a}.
In conclusion, we note that Krein's spectral shift function
\cite{Kr53}--\cite{KJ81}, a concept originally
introduced by
Lifshits \cite{Li52}, \cite{Li56}, continues to generate
considerable interest. Without repeating the extensive
bibliography recently provided in \cite{GMN99}, we remark
that the spectral shift function plays a fundamental role in
scattering theory, (relative) index theory, spectral
averaging and its application to localization properties of
random Hamiltonians, eigenvalue counting functions and
spectral asymptotics, semi-classical approximations, and
trace formulas for one-dimensional Schr\"odinger and Jacobi
operators. A very selective list of recent pertinent
references includes, for instance, \cite{BP98},
\cite{CHM96}, \cite{EP97}--\cite{GM99},
\cite{GS96}, \cite{KS98}, \cite{Mu98},
\cite{Pu97}--\cite{Pu98a}, \cite{Ro98}, \cite{Sa97},
\cite{Si95},
\cite{Si98}. For many more references the interested
reader can
consult the 1993 reviews by Birman and Yafaev \cite{BY93},
\cite{BY93a}, and
\cite{GM99},
\cite{GMN99}.
\section{Index of a Pair of Projections and Trindex}
\label{s2}
In this section we recall the main properties of the
index of a Fredholm pair of orthogonal projections in
${\mathcal H}$ and discuss a closely related notion of a trindex
of a
pair of operators, one of which is a bounded
operator in
${\mathcal H}$ and the other is an orthogonal projection in
${\mathcal H}$. Finally, we introduce the notion of a generalized
trace for a pair of bounded operators in ${\mathcal H}$.
Let $P$ and $Q$ be orthogonal projections in a
complex separable Hilbert space ${\mathcal H}$. The pair
$(P,Q)$ is
said to be a Fredholm pair if the map
\begin{equation}
QP\big|_{\text{\rm{ran}}(P)}:\text{\rm{ran}}(P) \to \text{\rm{ran}} (Q) \label{2.1}
\end{equation}
is a Fredholm operator from the Hilbert space
$\text{\rm{ran}} (P)$ to the Hilbert space $\text{\rm{ran}} (Q)$. In this case
one defines the index of the pair
$(P,Q)$ as the Fredholm
index of the operator $QP\big|_{\text{\rm{ran}}(P)}$
\begin{equation}\label{tr.1}
\text{\rm{index}} (P,Q)=\text{\rm{index}}(QP\big|_{\text{\rm{ran}}(P)}).
\end{equation}
The following three results, Lemmas~\ref{l2.1} and \ref{ltr.4}
and Theorem~\ref{ttr.3}, recall well-known
results for Fredholm pairs of projections. We refer, for
instance, to \cite{AS94}, \cite{ASS94}, \cite{Ka97},
\cite{Ka55} and the references cited therein.
We start with two important criteria for a pair $(P,Q)$ of
self-adjoint projections to be a Fredholm pair.
\begin{lemma}\label{l2.1}
(i) A necessary and sufficient condition that $(P,Q)$ be
a Fredholm pair is that $P-Q=F+D$,
where $F,D$ are self-adjoint,
$\|D\|<1$, and $F$ is a finite-rank operator.\\
(ii) $(P,Q)$ is a Fredholm pair if and only if $+1$
and $-1$
do not belong to the essential spectrum of $(P-Q)$
\begin{equation}\label{tr.2}
\pm 1\notin\text{\rm{ess.spec}} (P-Q).
\end{equation}
In this case $\ker(P-Q\pm I_{\mathcal H})$ are both finite
dimensional and
\begin{equation}\label{tr.3}
\text{\rm{index}} (P,Q)=\text{\rm{dim}} (\ker (P-Q-I_{\mathcal H}))-
\text{\rm{dim}} (\ker (P-Q+I_{\mathcal H})).
\end{equation}
In particular, if either
\begin{equation}\label{tr.4}
(P-Q)\in {\mathcal B}_\infty({\mathcal H}),
\end{equation}
or
\begin{equation}\label{tr.5}
\|P-Q\|<1,
\end{equation}
then $(P,Q)$ is a Fredholm pair.
\end{lemma}
The following result summarizes some of the most important
properties of the index of a Fredholm pair of projections.
\begin{theorem}\label{ttr.3}
(i) Let $(P,Q)$ be a Fredholm pair of projections in
${\mathcal H}$.
Then so is $(Q,P)$ and
\begin{equation}\label{tr.6}
\text{\rm{index}} (P,Q)=-\text{\rm{index}} (Q,P).
\end{equation}
(ii) Let $(P,Q)$ and $(Q,R)$ be Fredholm pairs in
${\mathcal H}$ and
either $(P-Q)\in {\mathcal B}_\infty({\mathcal H})$ or $(Q-R)\in
{\mathcal B}_\infty({\mathcal H})$. Then $(P,R)$ is a Fredholm pair
and one
has the chain rule
\begin{equation}\label{tr.7}
\text{\rm{index}} (P,R)=\text{\rm{index}} (P,Q)+\text{\rm{index}} (Q,R).
\end{equation}
(iii) If $\|P-Q\|<1$ then
\begin{equation}\label{tr.8}
\text{\rm{index}} (P,Q)=0.
\end{equation}
(iv) If $(P-Q)\in {\mathcal B}_1({\mathcal H})$ then
\begin{equation}\label{tr.9}
\text{\rm{index}} (P,Q)=\tr (P-Q).
\end{equation}
\end{theorem}
As shown in \cite{ASS94}, the compactness assumption in
connection with \eqref{tr.7} cannot be dropped in general.
Theorem~\ref{ttr.3}\,(ii)
combined with Lemma~\ref{l2.1}\,(i) implies a
stability result for the index for a Fredholm pair of
projections under small perturbations.
\begin{lemma}\label{ltr.4}
Let $P, $ $P_1$, and $Q$ be orthogonal projections
in ${\mathcal H}$.
Assume that
\begin{equation}\label{tr.10}
(P-Q)\in {\mathcal B}_\infty({\mathcal H})
\end{equation}
and
\begin{equation}\label{tr.11}
\|P-P_1\|<1.
\end{equation}
Then $(P,Q)$ and $(P_1, Q)$ are Fredholm pairs in
${\mathcal H}$ and
\begin{equation}\label{tr.12}
\text{\rm{index}} (P,Q)=\text{\rm{index}} (P_1,Q).
\end{equation}
\end{lemma}
\begin{proof}
That $(P,Q)$ is a Fredholm pair follows from
\eqref{tr.10} and Lemma~\ref{l2.1}\,(ii). Since
\eqref{tr.10} and \eqref{tr.11} hold, the difference
$P_1-Q$
can be represented as a sum of a contraction (with norm
strictly less than one)
and a finite-rank operator. Thus
$(P_1,Q)$ is a Fredholm pair by Lemma~\ref{l2.1}\,(i).
Since \eqref{tr.10}
holds one
can apply Theorem~\ref{ttr.3}\,(ii) to
conclude
that
\begin{equation}\label{tr.13}
\text{\rm{index}} (P,P_1)=\text{\rm{index}} (P,Q)+\text{\rm{index}} (Q,P_1).
\end{equation}
By Theorem~\ref{ttr.3}\,(iii) and \eqref{tr.11}
one gets
\begin{equation}\label{tr.14}
\text{\rm{index}} (P,P_1)=0.
\end{equation}
Combining \eqref{tr.6}, \eqref{tr.13}, and
\eqref{tr.14} one arrives at \eqref{tr.12}.
\end{proof}
\begin{definition} \label{d2.4} Let $A\in {\mathcal B}({\mathcal H})$
and $Q$ be
an orthogonal projection in ${\mathcal H}$.
We say that the pair $(A,Q)$ has a {\it trindex}, denoted
by $\text{\rm{trindex}}(A,Q)$, if there
exists an orthogonal projection $P$ in ${\mathcal H}$ such that
$(A-P)\in {\mathcal B}_1({\mathcal H})$ and $(P,Q)$ is a Fredholm pair
of orthogonal projections in
${\mathcal H}$. In this case the trindex of the pair $(A,Q)$ is
defined by
\begin{equation}\label{tr.17}
\text{\rm{trindex}} (A,Q)=\tr(A-P)+\text{\rm{index}} (P,Q).
\end{equation}
\end{definition}
The following result shows that the trindex of the pair
$(A,Q)$ is well-defined, that is, it is independent of
the choice of the projection $P$
satisfying the conditions in Definition~\ref{d2.4}.
\begin{lemma}\label{ltr.5}
Let $A\in {\mathcal B}({\mathcal H})$, and $P_1$, $P_2$, and $Q$
be orthogonal projections in ${\mathcal H}$ such that
\begin{equation}\label{tr.18}
(A-P_j) \in {\mathcal B}_1({\mathcal H}), \quad j=1,2,
\end{equation}
and
\begin{equation}\label{tr.19}
(P_j,Q) \text{ is a Fredholm pair, } \quad j=1,2.
\end{equation}
Then
\begin{equation}\label{tr.20}
\tr(A-P_1)+\text{\rm{index}}(P_1,Q)=\tr(A-P_2)+\text{\rm{index}}(P_2,Q).
\end{equation}
\end{lemma}
\begin{proof}
By \eqref{tr.18} one concludes that $(P_1-P_2)\in
{\mathcal B}_1({\mathcal H})$ and hence by Lemma~\ref{l2.1}\,(ii),
the pair $(P_1,P_2)$ is a Fredholm pair. In particular,
Theorem~\ref{ttr.3}\,(iv) implies
\begin{equation}
\tr(A-P_1)=\tr(A-P_2)+\tr(P_2-P_1)
=\tr(A-P_2)+\text{\rm{index}}(P_2,P_1).\label{tr.21}
\end{equation}
Hence, Theorem~\ref{ttr.3}\,(ii) yields
\begin{align}
&\tr(A-P_1)+\text{\rm{index}}(P_1,Q)
=\tr(A-P_2)+\text{\rm{index}}(P_2,P_1)+\text{\rm{index}}(P_1,Q) \nonumber \\
&=\tr(A-P_2)+\text{\rm{index}}(P_2,Q)\label{tr.22}
\end{align}
proving \eqref{tr.20}.
\end{proof}
\begin{remark} \label{r2.5}
Our motivation for introducing the concept of a trindex
for
a pair $(A,Q)$ lies in the following two facts.\\
(i) If $A=P$, with $(P,Q)$ a Fredholm pair of
projections in
${\mathcal H}$, then
\begin{equation}
\text{\rm{trindex}}(P,Q)=\text{\rm{index}}(P,Q). \label{2.19}
\end{equation}
(ii) If $A\in{\mathcal B}({\mathcal H})$ and $Q$ is an
orthogonal projection in ${\mathcal H}$ with
$(A-Q)\in{\mathcal B}_1({\mathcal H})$, then
\begin{equation}
\text{\rm{trindex}}(A,Q)=\tr(A-Q). \label{2.20}
\end{equation}
\end{remark}
The stability result for the trindex of a pair $(A,Q)$
analogous to Lemma~\ref{ltr.4} then reads as follows.
\begin{lemma} \label{l2.6}
Let $A\in {\mathcal B}({\mathcal H})$, and $Q$, $Q_1$ be orthogonal
projections in ${\mathcal H}$ such that $(Q-Q_1)\in
{\mathcal B}_\infty({\mathcal H})$ and
\begin{equation}
\|Q-Q_1\|<1.
\end{equation}
If $(A,Q)$ has a trindex, then $(A,Q_1)$ has a trindex and
\begin{equation}
\text{\rm{trindex}}(A,Q)=\text{\rm{trindex}}(A,Q_1).
\end{equation}
\end{lemma}
\begin{proof}
It suffices to combine Lemma~\ref{ltr.4} and \eqref{tr.17}.
\end{proof}
\begin{definition} \label{sp.4} Let $A, B\in {\mathcal B}({\mathcal H})$
and $(A-B)\in{\mathcal B}_\infty({\mathcal H})$.
We say that the pair $(A,B)$ has a {\it generalized trace,}
denoted by $\text{\rm{gtr}}(A,B)$, if there
exists an orthogonal projection $Q$ in ${\mathcal H}$ such that
both pairs, $(A,Q)$ and $(B,Q)$ have a trindex.
In this case the generalized trace of the pair
$(A,B)$ is defined by
\begin{equation}\label{sp17}
\text{\rm{gtr}}(A,B)=\text{\rm{trindex}}(A,Q)-\text{\rm{trindex}} (B,Q).
\end{equation}
\end{definition}
The following result shows that $\text{\rm{gtr}}(A,B)$ is well-defined,
that is,
it is independent of the choice of the orthogonal
projection $Q$
satisfying the conditions in Definition~\ref{sp.4}.
\begin{lemma} \label{l2.9}
Let $A,B\in{\mathcal B}({\mathcal H})$, $(A-B)\in{\mathcal B}_\infty({\mathcal H})$,
and $Q_1,Q_2$ orthogonal projections in ${\mathcal H}$ such that
$(A,Q_j)$ and $(B,Q_j)$, $j=1,2$ have a trindex. Then
\begin{equation}
\text{\rm{trindex}}(A,Q_1)-\text{\rm{trindex}}(B,Q_1)=\text{\rm{trindex}}(A,Q_2)-\text{\rm{trindex}}(B,Q_2).
\label{2.27}
\end{equation}
\end{lemma}
\begin{proof}
By hypothesis, there exist orthogonal projections $P_{j,A},
P_{j,B}$ in ${\mathcal H}$ such that $(A-P_{j,A}),(B-P_{j,B})\in
{\mathcal B}_1({\mathcal H})$ and the pairs $(P_{j,A},Q_j)$ and
$(P_{j,B},Q_j)$, $j=1,2$ are Fredholm pairs of projections.
Since $(A-B)\in{\mathcal B}_\infty({\mathcal H})$ one infers $(P_{j,A}
-P_{k,B})\in{\mathcal B}_\infty({\mathcal H})$, $j,k\in\{1,2\}$ and hence
$(P_{j,A},P_{k,B})$, $j,k\in\{1,2\}$ are Fredholm pairs.
Moreover, $(P_{2,A}-P_{1,A}),(P_{2,B}-P_{1,B})
\in{\mathcal B}_1({\mathcal H})$. Thus,
\begin{align}
&\text{\rm{trindex}}(A,Q_1)-\text{\rm{trindex}}(B,Q_1)-(\text{\rm{trindex}}(A,Q_2)-\text{\rm{trindex}}(B,Q_2))
\nonumber \\
&=\tr(A-P_{1,A})+\text{\rm{index}}(P_{1,A},Q_1)-\tr(A-P_{2,A})
-\text{\rm{index}}(P_{2,A},Q_2) \nonumber \\
&-\tr(B-P_{1,B})-\text{\rm{index}}(P_{1,B},Q_1)+\tr(B-P_{2,B})
+\text{\rm{index}}(P_{2,B},Q_2) \nonumber \\
&=\tr((A-B-P_{1,A}+P_{1,B})-(A-B-P_{2,A}+P_{2,B}))
+\text{\rm{index}}(P_{1,A},Q_1) \nonumber \\
&-\text{\rm{index}}(P_{1,B},Q_1) +\text{\rm{index}}(P_{2,B},Q_2)
-\text{\rm{index}}(P_{2,A},Q_2) \nonumber \\
&=\tr(P_{2,A}-P_{1,A})-\tr(P_{2,B}-P_{1,B}) +
\text{\rm{index}}(P_{1,A},P_{1,B})+\text{\rm{index}}(P_{2,B},P_{2,A}) \nonumber \\
&=\text{\rm{index}}(P_{2,A},P_{1,A}) +\text{\rm{index}}(P_{1,A},P_{1,B})
-\text{\rm{index}}(P_{2,B},P_{1,B}) +\text{\rm{index}}(P_{2,B},P_{2,A}) \nonumber \\
&=\text{\rm{index}}(P_{2,A},P_{1,B}) +\text{\rm{index}}(P_{1,B},P_{2,A}) =0 \label{2.28}
\end{align}
by repeatedly using \eqref{tr.6} and \eqref{tr.7}.
\end{proof}
\begin{remark} \label{r2.10}
If $A,B\in{\mathcal B}({\mathcal H})$ with $(A-B)\in{\mathcal B}_1({\mathcal H})$,
then
\begin{equation}
\text{\rm{gtr}}(A,B)=\tr(A-B). \label{2.29}
\end{equation}
\end{remark}
We were somewhat hesitant to introduce concepts
such as {\it trindex}, $\text{\rm{trindex}}(\cdot,\cdot)$ and {\it
generalized trace},
$\text{\rm{gtr}}(\cdot,\cdot)$ as additional entities to such familiar
quantities like the Fredholm index, $\text{\rm{index}}(\cdot,\cdot)$ and
the trace, $\tr(\cdot)$. However, it will become clear from
the remainder of this paper, that both objects seem to be
very natural in the context of Krein's spectral shift
function (cf. Corollary~\ref{iff} and Remark~\ref{rindex}).
\section{The $\Xi$ Operator} \label{s3}
Suppose $T$ is a bounded dissipative operator in the
Hilbert
space ${\mathcal H}$ (i.e., $\Im(T)\geq 0)$ and
$L$ is the minimal self-adjoint dilation
of $T$ (cf.~\cite[Ch.~III]{SF70}) in the Hilbert space
${\mathcal K}\supseteq
{\mathcal H}$.
We define the $\Xi$-operator associated with the dissipative
operator $T$ by
\begin{equation} \label{x.1}
\Xi(T)= P_{{\mathcal H}} E_L((
-\infty,0))P_{{\mathcal H}}\vert_{{\mathcal H}},
\end{equation}
where $P_{{\mathcal H}}$ is the orthogonal projection in
${\mathcal K}$ onto ${\mathcal H}$ and
$\{ E_L(\lambda)\}_{\lambda\in {\mathbb{R}}}$ represents the
family of strongly right-continuous
orthogonal spectral projections of $L$ in ${\mathcal K}$.
In particular, if $T=T^*$, the $\Xi$-operator coincides
with the spectral projection of $T$
corresponding to the negative semi-axis $(-\infty, 0)$,
\begin{equation}\label{x.3}
\Xi(T)= E_T(( -\infty,0)),
\end{equation}
since in this case the minimal self-adjoint dilation of
$T$ coincides with $T$.
\begin{remark} \label{r3.1}
(i) By \eqref{x.1}, $\Xi$ is a nonnegative contraction,
\begin{equation}\label{x.4}
0\le \Xi(T)\le I_{{\mathcal H}}.
\end{equation}
(ii) If $T\in {\mathcal B}({\mathcal H})$ is a bounded
dissipative operator and $T^{-1}\in {\mathcal B}({\mathcal H})$
then $\Xi(T)$ can be expressed in terms of the operator
logarithm of $T$ by
\begin{equation}\label{x.2}
\Xi(T)=\pi^{-1} \Im (\log (T)),
\end{equation}
where $\log(T)$ is defined by
\begin{equation}\label{2.17}
\log (T)=-i\int_0^\infty d \lambda \,
((T+i\lambda)^{-1}-(1 +i\lambda)^{-1}I_{\mathcal H})
\end{equation}
in the sense of a ${\mathcal B}({\mathcal H})$-norm convergent Riemann
integral (cf. the extensive treatment in \cite{GMN99}).
Without going into details (these may be found in
\cite{GMN99}), we remark that \eqref{x.2} resembles the
exponential Herglotz representation
for scalar-valued Herglotz functions
studied in detail by
Aronszajn and Donoghue \cite{AD56}.
Indeed, for any Herglotz
function $t(z)$ (i.e., $t:{\mathbb{C}}_+\to{\mathbb{C}}_+$ analytically),
$\log(t(z))$ is also a Herglotz function admitting the
representation,
\begin{equation}
\log(t(z))=c+\int_{{\mathbb{R}}}d\lambda \,
\xi(\lambda)((\lambda-z)^{-1}-\lambda
(1+\lambda^2)^{-1}), \quad z\in{\mathbb{C}}_+, \label{2.18}
\end{equation}
where $c\in {\mathbb{R}}$ and
\begin{equation}\label{2.19a}
0\leq \xi \leq 1 \text{ and } \xi
(\lambda)=\pi^{-1}\lim_{\varepsilon
\downarrow 0}\Im (\log (t(\lambda+i\varepsilon)))
\text{ a.e.}
\end{equation}
\end{remark}
At this point a natural question arises. Suppose $S$ is
a bounded dissipative
operator in ${\mathcal H}$ and $T=S+A+iB$, $A=A^*\in{\mathcal B}({\mathcal H})$,
$0\leq B\in{\mathcal B}({\mathcal H})$ its dissipative bounded
perturbation. Can one expect an interesting relationship
between $\Xi(T)$ and $\Xi(S)$? The following is a first
result in this direction.
\begin{lemma}\label{lx.0}
Let $S,T \in {\mathcal B}({\mathcal H})$ be dissipative operators such
that
$S^{-1}, T^{-1} \in {\mathcal B}({\mathcal H})$ and assume
$(T-S)\in {\mathcal B}_1({\mathcal H})$. Then
\begin{equation}\label{x.4b}
(\Xi(T) -\Xi(S))\in {\mathcal B}_1({\mathcal H}).
\end{equation}
\end{lemma}
\begin{proof}
Since $S$ and $T$ have a bounded inverse, $\log(T)$ and
$\log(S)$ are well-defined and (cf.~\cite{GMN99})
\begin{align}
&\log(T)-\log(S)=
-i\int_0^\infty dt\,
((T+it)^{-1}-(S+it)^{-1}) \nonumber \\
&=i\int_0^\infty dt\, ((T+it)^{-1}(T-S)(S+it)^{-1}).
\label{x.4c}
\end{align}
Using standard estimates for resolvents
$(T+it)^{-1}$ and $(S+it)^{-1}$ $(t\ge 0)$
of the dissipative operators $T$ and $S$ entering
\eqref{x.4c} (cf.~Lemma~2.6 in \cite{GMN99})
one concludes
\begin{equation}\label{x.4d}
(\log(T)-\log(S) )\in {\mathcal B}_1({\mathcal H}),
\end{equation}
if $(T-S)\in {\mathcal B}_1({\mathcal H})$. Thus,
$(\Xi(T)-\Xi(S)) \in {\mathcal B}_1({\mathcal H})$ by \eqref{x.2} and
\eqref{x.4d}.
\end{proof}
We assume the following hypothesis in the sequel.
\begin{hypothesis}\label{hx.1}
Assume $S=S^*\in{\mathcal B}({\mathcal H})$,
$A=A^*\in{\mathcal B}_\infty({\mathcal H})$,
$0\leq B\in {\mathcal B}_1({\mathcal H})$, and
$(S+A+\tau_0 B)^{-1}\in{\mathcal B}({\mathcal H})$ for some
$\tau_0\in {\mathbb{R}}$.
\end{hypothesis}
\begin{remark}\label{rx.1}
Under Hypothesis~\ref{hx.1} one infers
\begin{equation}\label{ess}
0\notin \text{\rm{ess.spec}}(S)
\end{equation}
by the stability of the essential spectrum under compact
perturbations. Moreover, by the analytic Fredholm
theorem (cf.~\cite[Sect.~VI.5]{RS80}),
$(S+A+zB)^{-1}\in{\mathcal B}({\mathcal H})$,
$z\in\overline {{\mathbb{C}}_+}$ except for $z$ in a discrete set
${\mathcal D}\subset{\mathbb{R}}$, with $\pm\infty$ the only
possible accumulation points of ${\mathcal D}$. In particular,
\begin{equation}\label{bound}
(S+A+\varepsilon B)^{-1} \in {\mathcal B}({\mathcal H}) \text{ for }
\varepsilon \text{ sufficiently small, } \varepsilon \ne 0.
\end{equation}
\end{remark}
\begin{lemma}\label{lx.1}
Assume Hypothesis~\ref{hx.1} with $A=0$. Then
\begin{equation}\label{x.5}
(\Xi(S+t B)-\Xi(S))\in {\mathcal B}_1({\mathcal H}) \text{ for all }
t \in
{\mathbb{R}}
\end{equation}
and
\begin{equation}\label{x.6}
\|\Xi(S+t B) -
\Xi(S)\|_{{\mathcal B}_1({\mathcal H})}=O(t) \text{ as } t \downarrow 0.
\end{equation}
Moreover, $S+zB$, $z\in{\mathbb{C}}_+$ has a bounded inverse,
\begin{equation}\label{x.6a}
(S+zB)^{-1} \in {\mathcal B}({\mathcal H}) \text{ for all } z\in{\mathbb{C}}_+
\end{equation}
and
\begin{equation}\label{x.6b}
(\Xi(S+z B)-\Xi(S))\in {\mathcal B}_1({\mathcal H}) \text{ for all } z\in
\overline {{\mathbb{C}}_+}\,.
\end{equation}
\end{lemma}
\begin{proof} Since $(S+\tau_0 B)^{-1}\in{\mathcal B}({\mathcal H})$ for
some $\tau_0\in {\mathbb{R}}$ and $B\in{\mathcal B}_\infty ({\mathcal H})$, one
infers that
$0\notin \text{\rm{ess.spec}} (S)$. Thus,
there exists a closed interval
$\Delta$, $0\in \Delta$, such that
$\text{\rm{dim}}(\text{\rm{ran}} (E_{S+t B}(\Delta)))<\infty$ for all $t\in {\mathbb{R}}$.
Hence,
for given
$t\in{\mathbb{R}}$, one can find a clockwise oriented bounded contour
$\Gamma_t$
encircling $(\text{\rm{spec}}(S+t B)\cup\text{\rm{spec}}(S))\cap
(-\infty,0)$ such that
\begin{equation}\label{x.7}
\Xi(S+t B)=E_{S+t B}((-\infty,0))=\frac{1}{2\pi i}
\oint_{\Gamma_t}
d\zeta\, (S+t B-\zeta)^{-1}, \quad t\in{\mathbb{R}}
\end{equation}
and
\begin{equation}\label{x.8}
\Xi(S)=E_{S}((-\infty,0))=\frac{1}{2\pi i}\oint_{\Gamma_t}
d\zeta\, (S-\zeta)^{-1}.
\end{equation}
The second resolvent identity for $S+t B$ and $S$ then
implies
$((S+tB-\zeta)^{-1}-(S-\zeta)^{-1})\in {\mathcal B}_1({\mathcal H})$,
$\zeta\in \Gamma_t$, proving \eqref{x.5}.
Since $B\geq 0$, there exists a closed interval
$\Delta$, $ \Delta\subset (-\infty,0)$, such that
\begin{equation}\label{x.9}
\bigcup_{t\in [0, \varepsilon]}\text{\rm{spec}}
(S+t B) \cap (-\infty,0) \subset \Delta \text{ for }
\varepsilon >0 \text{ sufficiently small.}
\end{equation}
Thus, choosing the contour
\begin{equation}\label{x.10}
\Gamma=\{\zeta\in {\mathbb{C}} \, | \, \text{dist}
(\zeta, \Delta)=\frac{1}{2} \text{dist}(0,\Delta)\},
\end{equation}
the representation \eqref{x.7} is valid
for all $t\in [0,\varepsilon]$, for sufficiently small
$\varepsilon >0$. Using the second resolvent
identity again and the standard estimate
$\|(A-\zeta)^{-1}\|\le (\text{dist}(\zeta,
\text{\rm{spec}}(A)))^{-1}$
for every self-adjoint operator $A$ in ${\mathcal H}$, one
obtains
\begin{align}
\|\Xi(S+t B)-\Xi(S) \|_{{\mathcal B}_1({\mathcal H})}&=
\|E_{S+t B}((-\infty, 0))
-E_{S}((-\infty, 0)) \|_{{\mathcal B}_1({\mathcal H})}
\nonumber \\
&
\le (2/\pi)t (\text{dist}(0,\Delta))^{-2}
\| B\|_{{\mathcal B}_1({\mathcal H})} |\Gamma|,
\quad t \in [0,\varepsilon], \label{x.11}
\end{align}
where $|\Gamma|$ denotes
the length of the contour $\Gamma$. This proves
\eqref{x.6}.
Next, consider the operator-valued Herglotz function
\begin{equation}\label{x.12}
M(z)=S+zB, \quad z\in{\mathbb{C}}_+
\end{equation}
(i.e., $\Im(M(z))\geq 0$ for all $z\in{\mathbb{C}}_+$). By
hypothesis
there exists a $\tau_0\in{\mathbb{R}}$ such that
$(S+\tau_0 B)^{-1}\in{\mathcal B}({\mathcal H})$. Thus, for sufficiently
small values of $|z-\tau_0|$, $z\in {\mathbb{C}}_+$, the operator
$M(z)$,
being a small perturbation of the invertible operator
$S+\tau_0 B$, has a bounded inverse.
By Lemma 2.3 in \cite{GMN99} and \eqref{x.5}, $M(z)$ is
invertible for
all $z\in{\mathbb{C}}_+$, proving \eqref{x.6a}.
Since $S+zB$, $z\in {\mathbb{C}}_+, $ and $S+\tau_0 B$ are boundedly invertible
(dissipative) operators, Lemma~\ref{lx.0} implies
$(\Xi(S+zB)-\Xi(S+\tau_0 B)) \in {\mathcal B}_1({\mathcal H})$.
By \eqref{x.5}, $(\Xi(S+\tau_0 B)-\Xi(S))\in
{\mathcal B}_1({\mathcal H})$
implying \eqref{x.6b}.
\end{proof}
\begin{remark}\label{rkr.1}
We note that condition \eqref{ess} is crucial in connection
with
\eqref{x.5}. Indeed, there exists
$S=S^*\in {\mathcal B}({\mathcal H})$
with
$
0\in \text{\rm{ess.spec}}(S)
$
and a self-adjoint rank-one operator $B$,
such that
\begin{equation}\label{notin}
\Xi(S+B)-\Xi(S)=E_{S+B}((-\infty,0))-
E_{S}((-\infty,0))\notin {\mathcal B}_1({\mathcal H})
\end{equation}
as implied by a result of Krein \cite{Kr53}.
\end{remark}
\begin{corollary}\label{ct.1}
Assume the hypotheses of Lemma~\ref{lx.1}. Then
\begin{equation} \label{c.1}
\tr (\Xi(S+tB)-\Xi(S))=0 \text{ for } t>0
\text{ sufficiently small}
\end{equation}
and
\begin{equation}\label{c.2}
\tr (\Xi(S-tB)-\Xi(S))=\text{\rm{dim}} (\ker(S)) \text{ for } t>0
\text{ sufficiently small.}
\end{equation}
\end{corollary}
\begin{proof}
Eq. \eqref{c.1} follows from \eqref{x.6}
and Theorem~\ref{ttr.3}\,(iii),(iv). Next one notes that
for $t>0$ sufficiently small, $E_{S-tB}(\{0\})=0$ by
Remark~\ref{rx.1}, and
hence
\begin{align}
&\Xi(S-tB)-\Xi(S)=
E_{S-t B}((-\infty, 0))- E_S((-\infty,0)) \nonumber \\
&=E_S([0, \infty))-E_{S-t B}([0, \infty))=
E_S([0, \infty))-E_{S-tB}((0, \infty)) \nonumber \\
&=E_S(\{0\})+ E_S((0,\infty))-E_{S-tB}((0, \infty)) \nonumber \\
&=E_S(\{0\})+ E_{-S}((-\infty, 0))-E_{-S+t B}((-\infty, 0))
\nonumber \\
&=E_S(\{0\})+\Xi(-S) -\Xi(-S+tB).
\end{align}
Since by \eqref{c.1} (replacing $S$ by $-S$),
\begin{equation}\label{c.3}
\tr (\Xi(-S) -\Xi(-S+tB))=0 \text{ for sufficiently small }
t>0,
\end{equation}
one obtains \eqref{c.2}.
\end{proof}
Lemma~\ref{lx.0} yields $(\Xi(T)-\Xi(S))\in{\mathcal B}_1({\mathcal H})$
for $S=S^*\in{\mathcal B}({\mathcal H})$ and $T$ a dissipative
operator with $(T-S)\in {\mathcal B}_1({\mathcal H})$.
In the case of general compact perturbations with trace
class imaginary parts, Hypothesis~\ref{hx.1} is not
sufficient
for the difference $(\Xi(S+A+iB)-\Xi(S))$ to be of
trace class
(see Remark~\ref{rce} below).
The following result, however, shows that the pair
$(\Xi(S+A+iB),\Xi(S))$ has a trindex.
\begin{lemma}\label{ltr.6}
Assume Hypothesis~\ref{hx.1}.~Then
\begin{equation}\label{x.20}
(\Xi( S+A+iB) -\Xi( S+A))\in {\mathcal B}_1({\mathcal H})
\end{equation}
and the pair $(\Xi( S+A), \Xi( S))$ is a Fredholm
pair of orthogonal projections. Thus,
$(\Xi( S+A+iB), \Xi( S))$ has a trindex and
\begin{align}
&\text{\rm{trindex}} (\Xi( S+A+iB) ,\Xi( S)) \nonumber \\
&=\tr ( \Xi( S+A+iB) -\Xi( S+A)))+
\text{\rm{index}} (\Xi(S+A), \Xi(S)) . \label{tr.23}
\end{align}
\end{lemma}
\begin{proof}
Since by hypothesis $(S+A+\tau_0 B)^{-1}\in{\mathcal B}({\mathcal H})$
for some $\tau_0 \in {\mathbb{R}}$, \eqref{x.6b} implies
\eqref{x.20}.
Due to the fact that $0\notin \text{\rm{ess.spec}} (S)$
(cf.~Remark~\ref{rx.1}) one can represent the spectral
projections $ \Xi(S+A)=E_{S+A}((-\infty,0))$ and
$\Xi(S)=E_{S}((-\infty,0))$ by the Riesz integrals
\eqref{x.7}, \eqref{x.8} and arguing as in the proof of
Lemma~\ref{lx.1} then yields
\begin{equation}\label{x.21}
(\Xi(S+A)-\Xi(S))\in {\mathcal B}_\infty({\mathcal H}).
\end{equation}
Thus, by Lemma~\ref{l2.1}\,(ii), the pair
$(\Xi(S+A),\Xi(S))$ is a
Fredholm pair of orthogonal projections, which together
with \eqref{x.20} proves \eqref{tr.23}.
\end{proof}
\begin{lemma}\label{lcom.2}
Let $S$ be
a signature operator, that is, $S=S^*=S^{-1}$ and
$A=A^*\in {\mathcal B}_2({\mathcal H})$. Then
\begin{equation}\label{ce.2}
(\Xi(S+A)-\Xi(S))\in {\mathcal B}_1({\mathcal H})
\end{equation}
if and only if
\begin{equation}\label{comm}
[A,S]=(AS-SA)\in {\mathcal B}_1({\mathcal H}).
\end{equation}
\end{lemma}
\begin{proof} Define two orthogonal projections
$P=S_+$ and $Q=S_-$ such that $P+Q=I_{\mathcal H}$ and
$
S=P-Q,
$
where $S_{\pm}=(|S|\pm S)/2$ (taking into
account that $\ker(S)=\{0\}$).
Next, we note
\begin{align}
&\frac{1}{2\pi i}\oint_{\Gamma}d\zeta
(S-\zeta)^{-1}A(S-\zeta)^{-1}
\nonumber \\
&
=PAP\frac{1}{2\pi i}\oint_{\Gamma}d\zeta(1-\zeta)^{-2}
+QAQ\frac{1}{2\pi i}\oint_{\Gamma}d\zeta(-1-\zeta)^{-2}\nonumber \\
&+(PAQ+QAP)\frac{1}{2\pi i}\oint_{\Gamma}d\zeta(1-\zeta)^{-1}
(-1-\zeta)^{-1}
=\frac{1}{4}S[S,A], \label{ce.4}
\end{align}
where the clockwise oriented contour $\Gamma$
encircles $\text{spec}(S+A)\cap
(-\infty,0)$.
On the other hand,
\begin{align}
&\Xi(S+A)-\Xi(S)=E_{S+A}((-\infty,0))
-E_{S}((-\infty,0))\nonumber \\
&=-\frac{1}{2\pi i}\oint_{\Gamma}d\zeta (S-\zeta)^{-1}
A(S-\zeta)^{-1}\nonumber \\
&\quad +\frac{1}{2\pi i}\oint_{\Gamma}d\zeta
(S+A-\zeta)^{-1}A(S-\zeta)^{-1}A(S-\zeta)^{-1}\nonumber \\
&=-\frac{1}{4}S[S,A]+\frac{1}{2\pi i}\oint_{\Gamma}d\zeta
(S+A-\zeta)^{-1}A(S-\zeta)^{-1}A(S-\zeta)^{-1}
\label{ce.3}
\end{align}
using \eqref{ce.4}.
Since $A\in {\mathcal B}_2({\mathcal H})$, the last term in \eqref{ce.3}
is a trace class operator and, hence,
$(\Xi(S+A)-\Xi(S))\in {\mathcal B}_1({\mathcal H})$ if and only
if $[S,A]\in
{\mathcal B}_1({\mathcal H})$.
\end{proof}
Combining the results of Lemmas~\ref{ltr.6}
and \ref{lcom.2}
we get the following result.
\begin{corollary}\label{iff}
Assume $A=A^*\in {\mathcal B}_2({\mathcal H})$, $0\le B\in
{\mathcal B}_1({\mathcal H})$, and
$S=S^*=S^{-1}$. Then
\begin{equation}\label{ce.5}
(\Xi(S+A+iB)-\Xi(S))\in {\mathcal B}_1({\mathcal H})
\end{equation}
if and only if
\begin{equation}\label{ce.6}
[A,S]\in {\mathcal B}_1({\mathcal H}).
\end{equation}
\end{corollary}
\begin{remark}\label{rce}
Corollary~\ref{iff} illustrates why Hypothesis~\ref{hx.1}
is insufficient to garantee \eqref{ce.5}. Indeed,
choosing ${\mathcal H}={\mathcal K} \oplus {\mathcal K}$, $S=I_{\mathcal K}
\oplus (-I_{\mathcal K})$, and $A=\left(\begin{smallmatrix}
0& a \\ a & 0
\end{smallmatrix}\right)$, where
$a\in{\mathcal B}_2({\mathcal K})\backslash
{\mathcal B}_1({\mathcal K})$ with $||a||<1/2$ for some (necessarily
infinite
dimensional)
complex separable Hilbert space ${\mathcal K}$, one infers the
tripple $(S,A,B)$
satisfies
Hypothesis~\ref{hx.1} for any $0\le B\in{\mathcal B}_1({\mathcal H})$ and
$[A,S]\notin{\mathcal B}_1({\mathcal H})$.
Consequently, in this case,
\begin{equation}\label{ce.7}
(\Xi(S+A+iB)-\Xi(S))\notin {\mathcal B}_1({\mathcal H})
\end{equation}
by Corollary~\ref{iff}.
\end{remark}
The next result concerns the continuity of
$\text{\rm{trindex}}(\Xi( S+A+iB), \Xi(S))$ under small perturbations
of $A$ and $B$ in the operator and trace norm topology,
respectively.
\begin{theorem}\label{ttr.7}
Assume Hypothesis~\ref{hx.1}. Let
$A_j=A^*_j\in {\mathcal B}_\infty({\mathcal H})$,
$0\leq B_j \in {\mathcal B}_1({\mathcal H})$, $j\in {\mathbb{N}}$, such that
$\lim_{j\to\infty}\|A_j-A\|=0$ and
$\lim_{j\to\infty}\|B_j-B\|_{{\mathcal B}_1({\mathcal H})}=0$. Then
\begin{equation}
\lim_{j\to \infty}\text{\rm{trindex}} (\Xi(S+A_j+iB_j),
\Xi(S)))=\text{\rm{trindex}} (\Xi(S+A+iB),
\Xi(S))).\label{tr.26}
\end{equation}
\end{theorem}
\begin{proof}
By hypothesis, $(S+A+\tau_0 B)^{-1}\in{\mathcal B}({\mathcal H})$
for some $\tau_0\in{\mathbb{R}}$. Since
$\|A_j- A\|\to 0$ and $\|B_j -B\|\to 0$ as $j\to \infty$,
the
operators $S+A_j+\tau_0 B_j$ are also invertible for
$j\ge j_0$, $j_0$ sufficiently large. Since
the operator-valued logarithm is a continuous function of
its (dissipative) argument in the ${\mathcal B}({\mathcal H})$-topology,
\begin{equation}\label{tr.27}
\nlim_{j\to \infty} \log(S+A_j+\tau_0 B_j)=
\log(S+A+\tau_0 B)
\end{equation}
and hence by \eqref{x.2},
\begin{equation}\label{tr.28}
\nlim_{j\to \infty} \Xi(S+A_j+\tau_0 B_j)=
\Xi(S+A+\tau_0 B).
\end{equation}
By Remark~\ref{rx.1}, $0\notin \text{\rm{ess.spec}} (S)$ and
thus
\begin{equation}\label{tr.29}
(\Xi(S+A_j+\tau_0 B_j)-\Xi(S))\in {\mathcal B}_\infty({\mathcal H})
\end{equation}
and
\begin{equation}\label{tr.30}
(\Xi(S+A+\tau_0 B)-\Xi(S))\in {\mathcal B}_\infty({\mathcal H}).
\end{equation}
Thus the difference $(\Xi(S+A_j+\tau_0 B_j)
-\Xi(S+A+\tau_0 B))$ is a compact operator and hence
by \eqref{tr.28} and Lemma~\ref{ltr.4} one gets
\begin{equation}
\lim_{j\to \infty}\text{\rm{index}} (\Xi(S+A_j+\tau_0 B_j)
,\Xi(S))=\text{\rm{index}} (\Xi(S+A+\tau_0 B)
,\Xi(S))). \label{tr.31}
\end{equation}
The estimate ($t\geq 0$)
\begin{align}
& \|(S+A+iB+it)^{-1}-(S+A+\tau_0 B+it)^{-1} \nonumber \\
&-(S+A_j+iB_j+it)^{-1}+(S+A_j+\tau_0 B_j+it)^{-1}
\|_{{\mathcal B}_1({\mathcal H})} \nonumber \\
&=(1+t^2)^{-1}o(1) \text{ as } j\to \infty, \label{tr.33}
\end{align}
with remainder term $o(1)$ uniform with respect to
$t\geq 0$,
then yields
\begin{align}
\lim_{j\to \infty}\| &\log(S+A_j+iB_j)-\log(S+A_j
+\tau_0 B_j)
\nonumber \\
& -\log(S+A+iB)-\log(S+A+\tau_0
B)\|_{{\mathcal B}_1({\mathcal H})}=0. \label{tr.32}
\end{align}
Hence
\begin{align}
\lim_{j\to \infty}
\|&\Xi(S+A_j+iB_j)-\Xi(S+A_j+\tau_0 B_j) \nonumber \\
& -\Xi(S+A+iB)+\Xi(S+A+\tau_0
B)\|_{{\mathcal B}_1({\mathcal H})}=0. \label{x.32}
\end{align}
Combining \eqref{tr.31} and \eqref{x.32} yields
\begin{align}
\lim_{j\to \infty} &(\tr (\Xi(S+A_j+iB_j)-
\Xi(S+A_j+\tau_0 B_j))
+\text{\rm{index}} ( \Xi(S+A_j+\tau_0 B_j), \Xi(S))) \nonumber \\
&= \tr (\Xi(S+A+iB)-\Xi(S+A+\tau_0 B))
+\text{\rm{index}} ( \Xi(S+A+\tau_0 B), \Xi(S)) \nonumber \\
&=\text{\rm{trindex}} (\Xi(S+A+iB),\Xi(S)),
\end{align}
proving \eqref{tr.26}.
\end{proof}
\section{Averaged Fredholm Indices and the Birman--Krein Formula} \label{s4}
Assume Hypothesis~\ref{hx.1} with $A=0$.
By Remark~\ref{rx.1} the operators $S+zB$, $z\in {\mathbb{C}}$
have a bounded inverse except for $z$ in a discrete set
$ {\mathcal D}\subset {\mathbb{R}}$ with $\pm\infty$ the only possible
accumulation points of ${\mathcal D}$. Introducing the family
of normal trace class operators
\begin{equation}\label{t.45}
{\mathcal A}(z)=B^{1/2}(S+zB)^{-1}B^{1/2},
\quad z \in {\mathbb{C}} \backslash {\mathcal D}.
\end{equation}
one verifies the resolvent-type identities
\begin{equation}\label{t.46}
{\mathcal A}(z_1)-{\mathcal A}(z_2)
=(z_2 -z_1) {\mathcal A}(z_1){\mathcal A}(z_2), \quad
\frac{d}{dz}{\mathcal A}(z)=-{\mathcal A}(z)^2.
\end{equation}
Thus, ${\mathcal A}(z_1)$ and
${\mathcal A}(z_2)$ commute for all $z_1,z_2\in{\mathbb{C}}\backslash
{\mathcal D}$
and have a common complete orthogonal system of
eigenvectors,
denoted by $\{\varphi_k \}_{k=1}^\infty$.
Making use of \eqref{t.46} one immediately gets the
following
result.
\begin{lemma}\label{lt.4}
The operator ${\mathcal A}(z)$, $z\in{\mathbb{C}}\backslash {\mathcal D}$ has the
eigenvalue $z^{-1}$ with multiplicity
$\text{\rm{dim}} (\ker(S))$. In particular, if $S^{-1}\in{\mathcal B}({\mathcal H})$,
then
\begin{equation}
z^{-1} \notin \text{\rm{spec}}({\mathcal A}(z)). \label{4.3}
\end{equation}
Let $\varphi$ be an eigenvector of ${\mathcal A}(z_1)$
corresponding to the eigenvalue $\mu(z_1)$.
Then $\varphi$ is an eigenvector of ${\mathcal A}(z_2)$
corresponding to the eigenvalue
\begin{equation}\label{t.47}
\mu(z_2)=\frac{\mu(z_1)}
{1-\mu(z_1) (z_1-z_2)}
\end{equation}
of the same multiplicity and,
\begin{equation}\label{t.48}
1-\mu(z_1) (z_1-z_2)\ne 0.
\end{equation}
\end{lemma}
Under Hypothesis~\ref{hx.1} with $A=0$, the operator
${\mathcal A}(\varepsilon)$
is well-defined for small (real) values of $\varepsilon$,
$\varepsilon \ne 0$,
${\mathcal A}(\varepsilon)\in {\mathcal B}_1({\mathcal H})$
(cf.~Remark~\ref{rx.1}) and hence
$\arctan ({\mathcal A}(\varepsilon))\in {\mathcal B}_1({\mathcal H})$.
\begin{theorem}\label{tt.bb}
Assume Hypothesis~\ref{hx.1} with $A=0$. Then
\begin{equation}\label{lt.6-}
\lim_{\varepsilon \downarrow 0}
\tr( \arctan ({\mathcal A}(\varepsilon) ))
=\int_{\mathbb{R}} \frac{dt\,n^*(t)}{1+t^2},
\end{equation}
where
$n^*(t)$ is the left-continuous function.
\begin{equation}\label{t.66a}
n^*(t)=\begin{cases} \sum_{s\in [0,t)}\,\text{\rm{dim}}(\ker (sB-S)),
&t >0, \\
0, &t=0, \\
-\sum_{s\in [t,0)}\text{\rm{dim}}(\ker (sB-S)), &t < 0. \end{cases}
\end{equation}
\end{theorem}
\begin{proof}
Fix a $\delta_0\in {\mathbb{R}}$, $\delta_0\notin{\mathcal D}$. Denote by
$\{\mu_k(\delta_0)\}_{k\in{\mathbb{N}}}\subset
\text{\rm{spec}} ({\mathcal A}(\delta_0))\backslash\{\delta_0^{-1}\}$ the
eigenvalues of ${\mathcal A}(\delta_0)$ different from
$\delta_0^{-1}$ with corresponding multiplicities
$\{ m_k\}_{k\in{\mathbb{N}}}$.
First we prove the following representation,
\begin{equation}\label{t.49}
\lim_{\varepsilon \downarrow 0}
\tr( \arctan ({\mathcal A}(\varepsilon)))=
\sum_{k=1}^\infty m_k\arctan
(\lambda_k)+
\frac{\pi}{2}\text{\rm{dim}} (\ker (S)),
\end{equation}
with
\begin{equation}\label{t.50a}
\lambda_k=\frac{\mu_k(\delta_0)}{1-\mu_k(\delta_0)\delta_0},
\quad k\in {\mathbb{N}}.
\end{equation}
Since
${\mathcal A}(\delta_0)\in{\mathcal B}_\infty({\mathcal H})$, there exists a
$\gamma >0$ such that
\begin{equation}
\text{\rm{spec}}({\mathcal A}(\delta_0))\cap (\delta_0^{-1} - \gamma,
\delta_0^{-1} + \gamma)
=\begin{cases}
\emptyset, & \text{\rm{dim}}(\ker (S))=0, \\
\{\delta_0^{-1} \}, & \text{\rm{dim}}(\ker (S))>0.
\end{cases} \label{t.50}
\end{equation}
For sufficiently small values of $\varepsilon \in {\mathbb{R}}$,
$\varepsilon \ne 0$,
the self-adjoint operator ${\mathcal A}(\varepsilon)$
is well-defined and the trace of
$\arctan({\mathcal A}(\varepsilon))$ reads
\begin{equation}\label{t.51}
\tr( \arctan ({\mathcal A}(\varepsilon)))=
\sum_{k=1}^\infty m_k \arctan(\mu_k(\varepsilon))+
\text{\rm{dim}}(\ker (S)) \arctan (\varepsilon^{-1}),
\end{equation}
where by Lemma~\ref{lt.4},
\begin{equation}\label{t.52}
\mu_k(\varepsilon)=\frac{\mu_k(\delta_0)}{1-\mu_k(\delta_0)
(\delta_0-\varepsilon)}, \quad k\in{\mathbb{N}}.
\end{equation}
For $\varepsilon<\frac{\gamma \delta_0}{2}
\| {\mathcal A}(\delta_0) \|^{-1}$ one obtains
\begin{align}
&\bigg |\sum_{k=1}^\infty m_k
\arctan (\mu_k(\varepsilon))
- \sum_{k=1}^\infty m_k\arctan (\lambda_k ) \bigg | \nonumber \\
&\leq \sum_{k=1}^\infty m_k |\mu_k(\delta_0) |\,
|(1-\mu_k(\delta_0)\delta_0)^{-1} -
(1-\mu_k(\delta_0)(\delta_0-\varepsilon))^{-1}| \nonumber \\
&\leq 2 \varepsilon (\gamma\delta_0)^{-2}
\sum_{k=1}^\infty m_k
|\mu_k(\delta_0) |^2 \le 2 \varepsilon
(\gamma\delta_0)^{-2}
\|{\mathcal A}(\delta_0)
\|^2_{{\mathcal B}_2({\mathcal H})} . \label{t.54}
\end{align}
By \eqref{t.54}, using the dominated convergence theorem,
one can perform the limit
$\varepsilon \downarrow 0$ in \eqref{t.51} and upon
combining \eqref{t.50a} and \eqref{t.52} one arrives at
\eqref{t.49} using Lemma~\ref{lt.4}.
Since the multiplicities
$m_k=\text{\rm{dim}} (\ker (B^{1/2}(S+\delta_0 B)^{-1}B^{1/2}-
\mu_k(\delta_0)I_{\mathcal H}))$
of the eigenvalues $\mu_k(\delta_0)$ can also be computed as
\begin{equation}\label{t.61}
m_k=\text{\rm{dim}} (\ker (B-\lambda_k S)),
\end{equation}
where $\lambda_k$ are given by \eqref{t.50a}, the absolutely
convergent series $\sum_{k=1}^\infty m_k
\arctan(\lambda_k)$ can be represented as the
Lebesgue integral
\begin{equation}\label{t.62}
\sum_{k=1}^\infty m_k\arctan(\lambda_k)=
\sum_{k=1}^\infty m_k\int_0^{\lambda_k}\frac{dt}{1+t^2}=
\int_{\mathbb{R}} \frac{dt\,m(t)}{1+t^2},
\end{equation}
where $m(t)$ is the following eigenvalue counting function
\begin{equation}
m(t)=\begin{cases}
\sum_{\lambda\in (t,\infty)}\,\text{\rm{dim}}(\ker
(B-\lambda S)), &t> 0, \\
0, & t=0, \\
-\sum_{\lambda\in (-\infty, t)}\text{\rm{dim}}(\ker
(B-\lambda S)),&t < 0. \end{cases}
\label{t.63}
\end{equation}
Making the change of variables $t\to 1/t$ (separately on
$(-\infty,0)$ and $(0,\infty)$)
\begin{equation}\label{t.64}
\int_{\mathbb{R}} \frac{dt\, m(t)}{1+t^2}=\int_{\mathbb{R}}
\frac{dt\, m(1/t)}{1+t^2},
\end{equation}
one infers by \eqref{t.62} and \eqref{t.49} that
\begin{equation}
\lim_{\varepsilon \downarrow 0}\tr (
\arctan ({\mathcal A}(\varepsilon)))
=\int_{\mathbb{R}} \frac{dt\, m(1/t)}{1+t^2}+
\text{\rm{dim}} (\ker(S))\int_0^\infty \frac{dt}{1+t^2}
=\int_{\mathbb{R}} \frac{dt\, n(t)}{1+t^2}, \label{t.65}
\end{equation}
where
\begin{equation}\label{t.66}
n(t)=m(1/t) + \text{\rm{dim}}(\ker(S))\frac{1+\text{\rm{sgn}}(t)}{2},
\quad t\ne 0.
\end{equation}
Combining \eqref{t.63}
and \eqref{t.66} one concludes that $n(t)=n^*(t)$ for a.e.
$t\in{\mathbb{R}}$, where $n^*(t)$ is a left-continuous function
given by \eqref{t.66a}.
\end{proof}
The following result is one of the main technical tools
in our paper.
\begin{theorem}\label{tt.2}
Let $S=S^*\in {\mathcal B}({\mathcal H})$, $S^{-1}\in {\mathcal B}({\mathcal H})$,
$0\leq B\in {\mathcal B}_1({\mathcal H})$ and introduce
\begin{equation}\label{t.6}
T(z)=S+zB, \quad z\in\overline {{\mathbb{C}}_+}.
\end{equation}
Define ${\mathcal Z}={\mathbb{C}}_+\cup \{ x\in {\mathbb{R}} \, |
\, (S+\tau x B)^{-1} \in {\mathcal B}({\mathcal H}) $
for all $\tau \in [0,1] \}$. Then
\begin{equation}\label{t.7}
(\log(T(z))-\log(S)) \in {\mathcal B}_1({\mathcal H}), \quad z\in {\mathcal Z}
\end{equation}
and
\begin{equation}\label{t.8}
\tr (\log(T(z))-\log(S))=
\sum_{k=1}^\infty m_k \int_0^1 d\tau\, \frac{z
\lambda_k}{1+\tau z \lambda_k},\quad z\in {\mathcal Z},
\end{equation}
where $\{\lambda_k \}_{k\in{\mathbb{N}}}\subset {\mathbb{R}}$ is the
set of eigenvalues with associated multiplicities
$\{m_k\}_{k\in{\mathbb{N}}}$ of
the self-adjoint trace class operator
$B^{1/2}S^{-1}B^{1/2}$.
\end{theorem}
\begin{proof}
First one notes that $T(z)$, $z\in {\mathcal Z}$ is invertible.
For $z\in
{\mathbb{R}}$ this holds by hypothesis and
for $\Im(z) >0$ this holds since $S^{-1}\in{\mathcal B}({\mathcal H})$.
Thus, $\log(T(z))$ and $\log(S)$ are well-defined.
In the following let $z\in {\mathcal Z}$. Since
$B\in {\mathcal B}_1({\mathcal H})$, the representation
\begin{equation}\label{t.11}
\log(T(z))-\log(S)=
-i\int_0^\infty dt\,((S+z B+it)^{-1}-(S+it)^{-1}),
\end{equation}
and estimates for the resolvents $(S+iB+it)^{-1}$ and
$(S+it)^{-1}$, $t\ge 0$, analogous to those in the
proof of Lemma 2.6 of \cite{GMN99} yield
\eqref{t.7}. Therefore,
\begin{equation}
\tr(\log(T(z))-\log(S))
=-i\tr \bigg(\int_0^\infty dt\,((S+ z B+it)^{-1}
-(S+it)^{-1})\bigg ).
\label{t.12}
\end{equation}
Based on the estimates in the proof of Lemma~2.6
in \cite{GMN99}
mentioned above and the fact that $T=\int_0^\infty ds \,
T(s)\in{\mathcal B}_1({\mathcal H})$ and
\begin{equation}
\tr (T)=\int_0^\infty ds\,\tr( T(s)), \label{t.12a}
\end{equation}
whenever $T(s)$
is continuous with respect to $s\in [0,\infty)$ in
${\mathcal B}_1({\mathcal H})$-topology and $\|T(s)\|_{{\mathcal B}_1({\mathcal H})}
\leq C(1+s)^{-1-\varepsilon}$ for some $\varepsilon >0$,
one can interchange the integral and the trace in
\eqref{t.12} and obtain
\begin{equation}
\tr(\log(T(z))-\log(S))
=-i\int_0^\infty dt \,\tr((S+z B+it)^{-1}
-(S+it)^{-1})).
\label{t.13}
\end{equation}
Next, using the fact that $((S+\tau zB+it)^{-1}
-(S+it)^{-1})$
is differentiable with respect to $\tau$ in trace norm for
$(\tau,t)\in [0,1]\times [0,\infty)$ and
\begin{align}
&(d/d\tau)((S+\tau zB+it)^{-1}-(S+it)^{-1}) \nonumber \\
&=-(S+\tau zB+it)^{-1}zB(S+\tau zB+it)^{-1} \label{4.29}
\end{align}
in trace norm, one concludes
\begin{align}
&(d/d\tau)\tr((S+\tau zB+it)^{-1}-(S+it)^{-1}) \nonumber \\
&=-\tr((S+\tau zB+it)^{-1}zB(S+\tau zB+it)^{-1}) \nonumber \\
&=-\tr((S+\tau zB+it)^{-2}zB) \label{4.30}
\end{align}
and hence
\begin{equation}
\tr ((S+z B+it)^{-1}-(S+it)^{-1})=
-\int_0^1d\tau \tr(S+\tau
zB+it)^{-2}zB), \quad t\ge 0,
\label{t.14}
\end{equation}
integrating \eqref{4.30} from $0$ to $1$ with respect to
$\tau$. Combining
\eqref{t.13} and
\eqref{t.14} one obtains
\begin{equation}\label{t.19}
\tr (\log(T(z))-\log(S))=i\int_0^\infty dt
\int_0^1 d\tau\,\tr((S+\tau zB+it)^{-2}zB).
\end{equation}
Using the estimate
\begin{equation}
|\tr((S+\tau zB+it)^{-2}B)| \le
\|(S+\tau zB+it)^{-1}\|^2 \|B\|_{{\mathcal B}_1({\mathcal H})}
\le C(1+t^2)^{-1}, \label{t.20}
\end{equation}
which holds uniformly with respect to $\tau\in [0,1]$,
Fubini's theorem implies
\begin{equation}\label{t.21}
\tr (\log(T(z))-\log(S))=i\int_0^1 d\tau\,
\int_0^\infty dt\,\tr((S+\tau zB+it)^{-2}zB).
\end{equation}
Applying \eqref{t.12a} again, \eqref{t.20} implies
\begin{equation}
\int_0^\infty dt
\tr((S+\tau z B+it)^{-2}z B)
=\tr\bigg(\int_0^\infty dt\,
(S+\tau z B+it)^{-2} z B\bigg).\label{t.22}
\end{equation}
Using
\begin{equation}\label{t.23}
\int_0^\infty dt\,
(S+\tau zB+it)^{-2}=-i(S+\tau zB)^{-1},
\end{equation}
and combining \eqref{t.21}--\eqref{t.23} one finally gets
\begin{align}
\tr (\log(T(z))-\log(S))&=\int_0^1 d \tau\,
\tr ((S+\tau zB)^{-1}z B)
\nonumber \\
& =\int_0^1 d\tau\, \tr (z B^{1/2} (S+\tau
z B)^{-1} B^{1/2}). \label{t.24}
\end{align}
The trace of $zB^{1/2}(S+\tau zB)^{-1}B^{1/2}$,
$\tau\in [0,1]$, can easily be computed in terms of the
eigenvalues
$\{ \lambda_k \}_{k\in{\mathbb{N}}}$ of
$B^{1/2}S^{-1}B^{1/2}$. By Lemma~\ref{lt.4},
$zB^{1/2}(S+\tau zB)^{-1}B $ has the eigenvalues
\begin{equation}\label{t.25}
\mu_k(\tau,z)=\frac{z\lambda_k}{1+\tau z\lambda_k},
\quad k\in {\mathbb{N}},
\end{equation}
with associated multiplicities $\{m_k\}_{k\in{\mathbb{N}}}$. By
Lidskii's theorem (cf.~\cite[Ch.~3]{Si79})
\begin{equation}\label{t.26}
\tr(zB^{1/2} (S+\tau zB)^{-1} B^{1/2})
=\sum_{k=1}^\infty m_k\mu_k(\tau,z).
\end{equation}
By \eqref{t.24} and \eqref{t.25}
\begin{equation}\label{t.27}
\tr(\log(T(z))-\log(S))=
\int_0^1d \tau \sum_{k=1}^\infty m_k\mu_k(\tau,z).
\end{equation}
Since $B^{1/2}S^{-1}B^{1/2}\in{\mathcal B}_1({\mathcal H})$ is
self-adjoint, one concludes that $\{\lambda_k\}_{k\in
{\mathbb{N}}}\subset {\mathbb{R}}$ and that the series $\sum_{k=1}^\infty
|\lambda_k|$ converges. Hence, applying the dominated
convergence theorem, one can interchange the sum and the
integral in \eqref{t.27}, arriving at \eqref{t.8}.
\end{proof}
\begin{corollary}\label{ctt.2}
Under the assumptions of Theorem~\ref{tt.2},
\begin{equation}
\tr( \Im (\log(S+iB)-\Im(\log(S)))=
\tr( \arctan (B^{1/2} S^{-1} B^{1/2})).
\end{equation}
\end{corollary}
\begin{proof}
Pick $z=i$ in Theorem~\ref{tt.2}.
Taking the imaginary part of both sides of \eqref{t.8},
an explicit computation of the integrals in \eqref{t.8}
yields
\begin{equation}
\Im( \tr(\log(S+iB))-\tr(\log(S)))=
\sum_{k=1}^\infty m_k \arctan (\lambda_k)=
\tr(\arctan (B^{1/2}S^{-1}B^{1/2})).
\end{equation}
\end{proof}
\begin{remark}
Corollary~\ref{ctt.2} is
an operator analog of the following elementary
fact
\begin{equation}\label{tttt.1}
\Im(\log (a+ib))-\Im (\log(a))=
\arctan(b^{1/2}a^{-1}b^{1/2}), \quad
a\in{\mathbb{R}}\backslash\{0\}, \, b>0,
\end{equation}
where $\log(\cdot)$ and $\arctan(\cdot)$ denote the
corresponding principal branches, that is,
\begin{equation}\label{tttt.2}
\Im(\log(\lambda))=0, \quad
\lambda>0 \text{ and } -\frac{\pi}{2}< \arctan
(\lambda) <\frac{\pi}{2}, \, \lambda \in {\mathbb{R}}.
\end{equation}
\end{remark}
Theorem~\ref{tt.2} for $z=1$,
has important consequences when computing the
trace of $(\Xi(S+B)-\Xi(S))$ (the case of self-adjoint
perturbations).
We start with the simplest case of self-adjoint
perturbations where $(S+tB)^{-1}\in{\mathcal B}({\mathcal H})$ for all
$t\in [0,1]$.
\begin{lemma}\label{lt.2a}
Let $S=S^*\in {\mathcal B}({\mathcal H})$,
$0\leq B\in {\mathcal B}_1({\mathcal H})$. Assume
\begin{equation}\label{t.2a}
(S+tB)^{-1}\in {\mathcal B}({\mathcal H}) \text{ for all }
t\in [0,1].
\end{equation}
Then
\begin{equation} \label{t.4a}
\tr(\Xi(S+B)-\Xi(S))=0.
\end{equation}
\end{lemma}
\begin{proof}
Under hypothesis \eqref{t.2a} one can apply
Theorem~\ref{tt.2}
for $z=1$. Thus, $\tr(\Xi(S+B)-\Xi(S))\in{\mathbb{R}}$
and therefore
\eqref{t.4a} holds due to \eqref{x.2} and the fact
that the left-hand side of \eqref{t.8} is real.
\end{proof}
Next we relax the condition \eqref{t.2a} of
invertibility of $S+tB$ for all
$t\in [0,1]$, still assuming, however, that $S+\tau_0 B$ has
a bounded inverse for some $\tau_0\in {\mathbb{R}}$. The
following result
is concerned with the situation where the map $t\mapsto
(S+tB)^{-1}$, $t\in [-1,1]$ is singular at some points.
\begin{theorem}\label{tt.4c}
Let $S=S^*\in {\mathcal B}({\mathcal H})$, $0\leq B\in {\mathcal B}_1({\mathcal H})$
and assume $(S+\tau_0 B)^{-1}\in {\mathcal B}({\mathcal H})$ for some
$\tau_0\in {\mathbb{R}}$. Then
\begin{align}
\tr(\Xi(S+B)-\Xi(S))&=
-\sum_{s\in (0,1]} \text{\rm{dim}}(\ker (S+sB)), \label{t.2c} \\
\tr(\Xi(S-B)-\Xi(S))&=
\sum_{s\in (-1,0]} \text{\rm{dim}}(\ker (S+sB)). \label{t.2cc}
\end{align}
\end{theorem}
\begin{proof}
Since by hypothesis, $(S+\tau_0 B)^{-1}\in{\mathcal B}({\mathcal H})$,
Remark~\ref{rx.1} implies that
$(S+t B)^{-1}\in {\mathcal B}({\mathcal H})$ for all $t\in [0,1]$
except possibly at a finite number
of points $0=t_0 < t_1< t_2 < ... <t_n<t_{n+1}=1$.
Introducing the notation
\begin{equation}\label{t.3c}
E(t)=\Xi(S+tB), \quad t\in[0,1]
\end{equation}
one obtains for $\delta>0$ sufficiently small,
\begin{align}
&E_{S+B}((-\infty, 0))-E_S((-\infty, 0))=E(1)-E(0) \nonumber \\
&=\sum_{k=1}^{n+1} (E(t_k)-E(t_k-\delta))+
\sum_{k=1}^{n+1}(E(t_k-\delta)-E(t_{k-1}+\delta)) \nonumber \\
&+\sum_{k=1}^{n+1}(E(t_{k-1}+\delta)-E(t_{k-1})).
\label{t.4c}
\end{align}
By Lemma~\ref{lt.2a},
\begin{equation}\label{t.5c}
\tr \bigg(\sum_{k=1}^{n+1}(E(t_k-\delta)-E(t_{k-1}
+\delta))\bigg)=0,
\end{equation}
and by Corollary~\ref{ct.1} (for $\delta >0$ sufficiently
small),
\begin{equation}\label{t.6c}
\tr\bigg(\sum_{k=1}^{n+1}(E(t_{k-1}+\delta)
-E(t_{k-1}))\bigg)=0,
\end{equation}
while
\begin{align}
&\tr \bigg(\sum_{k=1}^{n+1} (E(t_k)-E(t_k-\delta))\bigg)=
-\sum_{k=1}^{n+1}\text{\rm{dim}}(\ker(S+t_kB)) \nonumber \\
&=-\sum_{s\in (0,1]}\text{\rm{dim}}(\ker (S+sB)).\label{t.7c}
\end{align}
Combining \eqref{t.4c}--\eqref{t.7c} proves \eqref{t.2c}.
Setting $W=S-B$ one gets by \eqref{t.2c},
\begin{align}
&\tr\big( \Xi(S)-\Xi(S-B) \big )=
\tr\big(\Xi(W+B)-\Xi(W) ) \nonumber \\
&=-\sum_{s\in (0,1]} \text{\rm{dim}}(\ker (W+sB))=
-\sum_{s\in (0,1]} \text{\rm{dim}}(\ker (S+(s-1)B)) \nonumber \\
&=-\sum_{s\in (-1,0]}\text{\rm{dim}}(\ker (S+sB)),
\label{t.8c}
\end{align}
proving \eqref{t.2cc}.
\end{proof}
As an immediate consequence one has the following result.
\begin{corollary}\label{ctt.1}
Assume the hypotheses of Theorem~\ref{tt.4c}. Then
\begin{align}
&\tr(\Xi(S+tB)-\Xi(S)) \nonumber \\
&=\text{\rm{index}} (\Xi(S+tB),\Xi(S))=
\begin{cases}
-\sum_{s\in (0,t]}\text{\rm{dim}}(\ker (S+sB)), &t> 0, \\
0, &t=0, \\
\sum_{s\in (t,0]}\,\text{\rm{dim}}(\ker (S+sB)) ,&t < 0.
\end{cases}\label{t.9c}
\end{align}
\end{corollary}
\begin{remark}\label{safr}
The trace formula \eqref{t.9c}
can be interpreted as follows. The Fredholm index of the pair
of spectral projections $(\Xi(S),\Xi(S+B))$
coincides with the number of eigenvalues of
$S+sB$ which cross the point
$0$ from the left to the right as the coupling
constant $s$ increases from $0$ to $1$.
\end{remark}
Now, we are prepared to prove our principal result.
\begin{theorem}\label{ttr.8}
Assume Hypothesis~\ref{hx.1}.
Then the pair $( \Xi( S+A+iB),
\Xi(S))$
has a trindex and
\begin{equation}\label{tr.35}
\text{\rm{trindex}}
(\Xi( S+A+iB),
\Xi(S))=
\frac{1}{\pi}\int_{\mathbb{R}} \frac{dt\,n(t)}{1+t^2},
\end{equation}
where
\begin{equation}
n(t)=\text{\rm{index}} (\Xi(S+A+tB),\Xi(S)).
\label{tr.36}
\end{equation}
\end{theorem}
\begin{proof}
By Hypothesis~\ref{hx.1}, $(S+A+t B)^{-1}\in{\mathcal B}({\mathcal H})$
for all $t > 0$ sufficiently small. By
Corollary~\ref{ctt.2},
\begin{align}
&\tr( \Xi(S+A+tB +iB )-\Xi(S+A+t B)) \nonumber \\
&=(1/\pi)\tr (\arctan
( B^{1/2}(S+A+t B)^{-1}B^{1/2})). \label{y.1}
\end{align}
By Lemma~\ref{lx.1} one concludes that
\begin{equation}\label{y.2}
\lim_{t \downarrow 0}\|\Xi(S+A+t B)
-\Xi(S+A) \|_{{\mathcal B}_1({\mathcal H})}=0
\end{equation}
and from standard properties of the operator logarithm
(cf.~\cite{GMN99}) one also infers
\begin{equation}\label{y.3}
\lim_{t \downarrow 0}\|\Xi(S+A+iB+t B)
-\Xi(S+A+iB) \|_{{\mathcal B}_1({\mathcal H})}=0.
\end{equation}
Combining \eqref{y.1}--\eqref{y.3} one obtains
\begin{equation}
\tr(\Xi(S+A+iB)-\Xi(S+A))=(1/\pi)\lim_{t \downarrow 0}
\tr( \arctan (B^{1/2}(S+A+t B)B^{1/2})) \label{y.4}
\end{equation}
and by Theorem~\ref{tt.bb} one infers
\begin{equation}
\tr(\Xi(S+A+iB)-\Xi(S+A))=\frac{1}{\pi}
\int_{\mathbb{R}} \frac{dt\,n^*(t)}{1+t^2}=
\frac{1}{\pi}\int_{\mathbb{R}} \frac{dt\,n^*(-t)}{1+t^2}, \label{y.5}
\end{equation}
where
\begin{equation}\label{y.6}
n^*(-t)=\begin{cases}
\sum_{s\in [0,t)}\,\text{\rm{dim}}(\ker (sB-S-A)), &t> 0, \\
0, &t=0, \\
-\sum_{s\in [t,0)}\text{\rm{dim}}(\ker (sB-S-A)),&t < 0. \end{cases}
\end{equation}
By Corollary~\ref{ctt.1},
\begin{equation}
n^*(t)=\text{\rm{index}}(\Xi(S+A+tB),(\Xi(S+A)) \label{4.67}
\end{equation}
and therefore,
\begin{equation}
\tr(\Xi(S+A+iB)-\Xi(S+A))=
\frac{1}{\pi}\int_{\mathbb{R}} dt\, \frac{\text{\rm{index}} (\Xi(S+A+tB),
\Xi(S+A))}{1+t^2}.
\label{y.7}
\end{equation}
Since by \eqref{x.21}, $(\Xi(S+A)-\Xi(A))\in
{\mathcal B}_\infty({\mathcal H})$, one concludes that
$(\Xi(S+A),\Xi(A))$ is a Fredholm pair of orthogonal
projections. Moreover, $(\Xi(S+A+iB)-\Xi(S))\in
{\mathcal B}_1({\mathcal H})$ implies that the pair
$(\Xi(S+A+iB),\Xi(S))$ has a trindex and hence
\begin{align}
&\text{\rm{trindex}} (\Xi(S+A+iB),\Xi(S)) \nonumber \\
&=\tr (\Xi(S+A+iB)-\Xi(S+A))+
\text{\rm{index}} (\Xi(S+A), \Xi(S)). \label{y.8}
\end{align}
Now \eqref{tr.35} follows from the chain rule
\eqref{tr.7} for the index of a pair of projections,
\begin{align}
&\text{\rm{index}} (\Xi(S+A+tB),\Xi(S+A))+\text{\rm{index}}(\Xi(S+A),
\Xi(S)) \nonumber \\
&=\text{\rm{index}} (\Xi(S+A+tB),\Xi(S)),
\end{align}
and from the fact that the measure $(1/\pi)
(1+t^2)^{-1}dt$ is a probability measure on ${\mathbb{R}}$.
\end{proof}
\begin{remark} \label{r4.10}
(i) If $B=0$, $n(t)$ is independent of $t$
and \eqref{tr.35} together with
$(1/\pi)\smallint_{\mathbb{R}} dt\,(1+t)^{-2}=1$ then imply
\begin{equation}
\text{\rm{trindex}}(\Xi(S+A),\Xi(S))=\text{\rm{index}}(\Xi(S+A),\Xi(S)), \label{4.70}
\end{equation}
consistent with \eqref{2.19}.\\
(ii) The integral \eqref{tr.35} carries
out a ``smoothing'' of the integer-valued function $n(t)$
resulting in the expression of the trindex of a pair of
$\Xi$-operators. \\
(iii) Theorem~\ref{ttr.8} shows, in particular,
that under Hypothesis~\ref{hx.1} the difference
\begin{equation}\label{tr.37}
(\Xi( S+A+iB)-\Xi(S))\in {\mathcal B}_1({\mathcal H})
\end{equation}
if and only if
\begin{equation}\label{tr.38}
(E_{S+A}((-\infty,0)) -
E_{S}((-\infty,0))) \in {\mathcal B}_1({\mathcal H}).
\end{equation}
Under hypothesis \eqref{tr.38} one then obtains
\begin{align}
&\text{\rm{trindex}} ( \Xi( S+A+iB),
\Xi(S))=
\tr ( \Xi( S+A+iB)- \Xi(S)).
\label{tr.39}
\end{align}
\end{remark}
\begin{remark}
Theorem~\ref{ttr.8} is an operator analog of the fact
\begin{align}
&\Im(\log (a+ib))-\Im (\log(a))=
\int_{{\mathbb{R}}}dt(1+t^2)^{-1}
(\chi_{(-\infty,0)}(a+tb)-\chi_{(-\infty,0)}(a)), \nonumber \\
&\hspace*{8.5cm} a\in{\mathbb{R}}\backslash\{0\}, \, b>0,
\end{align}
where
$\chi_\Omega(\cdot)$ denotes the characteristic function
of $\Omega\subset{\mathbb{R}}$.
\end{remark}
There are two important special cases when
\eqref{tr.38} holds. For instance, if $S=I_{\mathcal H}$
or $S=-I_{\mathcal H}$ and $A=A^*\in{\mathcal B}_\infty({\mathcal H})$, the
difference \eqref{tr.38} is even a finite-rank operator,
\begin{align}
\Xi(S+A)-\Xi(S) &=E_{S+A}((-\infty,0))
-E_{S}((-\infty,0)) \nonumber \\
&=\begin{cases}
E_A((-\infty, -1)), & S=I_{\mathcal H}, \\
-E_A([1, \infty)), & S=-I_{\mathcal H}. \end{cases}
\label{tr.40}
\end{align}
\begin{lemma}\label{ltrr.9}
Assume Hypothesis~\ref{hx.1} and $S=I_{\mathcal H}$ or
$S=-I_{\mathcal H}$. Then for $S=I_{\mathcal H}$
\begin{equation}\label{tr.42}
\Xi( I_{\mathcal H}+A+iB)\in {\mathcal B}_1({\mathcal H})
\end{equation}
and
\begin{equation}\label{tr.43}
\tr (\Xi( I_{\mathcal H} +A+iB))=
\frac{1}{\pi}\int_{\mathbb{R}} \frac{dt\,n_-(t)}{1+t^2},
\end{equation}
where
\begin{equation}\label{tr.44}
n_-(t)=\text{\rm{rank}} (E_{A+tB}((-\infty, -1)))
\end{equation}
is a decreasing right-continuous function.
For $S=-I_{\mathcal H}$ one has
\begin{equation}\label{tr.45}
(\Xi( -I_{\mathcal H}+A+iB)-I_{\mathcal H}) \in {\mathcal B}_1({\mathcal H})
\end{equation}
and
\begin{equation}\label{tr.46}
\tr(\Xi( -I_{\mathcal H} +A+iB)-I_{\mathcal H})
=
-\frac{1}{\pi}\int_{\mathbb{R}} \frac{dt\,n_+(t)}{1+t^2},
\end{equation}
where
\begin{equation}\label{tr.47}
n_+(t)=\text{\rm{rank}} (E_{A+tB}([1,\infty)))
\end{equation}
is an increasing right-continuous function.
\end{lemma}
\begin{proof}
Let $S=I_{\mathcal H}$. Then $\Xi(S)=\Xi(I_{\mathcal H})=0$ and
\begin{equation}
\text{\rm{index}} ( \Xi(I_{\mathcal H} +A+tB),\Xi(I_{\mathcal H}) )=
\tr(\Xi(I_{\mathcal H} +A+tB))
=\text{\rm{rank}} (E_{A+tB}((-\infty, -1))), \label{tr.48}
\end{equation}
prove \eqref{tr.43} and \eqref{tr.44}. Next, let
$S=-I_{\mathcal H}$. Then $\Xi(S)=\Xi(-I_{\mathcal H})=I_{\mathcal H}$ and
\begin{align}
&\text{\rm{index}}( \Xi(-I_{\mathcal H} +A+tB), \Xi(-I_{\mathcal H}))=
\tr(( \Xi(-I_{\mathcal H}+A+tB)-I_{\mathcal H}) \nonumber \\
&=-\tr( E_{-I_{\mathcal H} +A+tB}([0, \infty)))=
-\text{\rm{rank}} (E_{A+tB}([1,\infty))) \label{tr.49}
\end{align}
prove \eqref{tr.46} and \eqref{tr.47}.
\end{proof}
Theorem~\ref{ttr.8} admits the following immediate
extension.
\begin{corollary}\label{ctr.8}
Assume that the triples $(S,A_1,B_1)$ and $(S,A_2,B_2)$
satisfy
Hypothesis~\ref{hx.1}. Then the pair $( \Xi( S+A_1+iB_1),
\Xi(S+A_2+iB_2))$ has a generalized trace and
\begin{equation}\label{spr.35}
\text{\rm{gtr}}(\Xi( S+A_1+iB_1),\Xi((S+A_2+iB_2))=
\frac{1}{\pi}\int_{\mathbb{R}} \frac{dt\,n(t)}{1+t^2},
\end{equation}
where
\begin{equation}
n(t)=\text{\rm{index}} (\Xi(S+A_1+tB_1),\Xi((S+A_2+tB_2)).
\label{spr.36}
\end{equation}
\end{corollary}
\begin{proof}
By Theorem~\ref{ttr.8} the pairs $(\Xi( S+A_1+iB_1),
\Xi(S))$
and
$(\Xi( S+A_2+iB_2), \Xi(S))$ have a trindex and hence
the pair
$( \Xi( S+A_1+iB_1),\Xi(S+A_2+iB_2))$ has a generalized
trace and
\begin{align}
&\text{\rm{gtr}}(\Xi( S+A_1+iB_1),\Xi(S+A_2+iB_2)) \nonumber \\
&=\text{\rm{trindex}}( \Xi( S+A_1+B_1),\Xi(S))-\text{\rm{trindex}}(\Xi( S+A_2+iB_2),
\Xi(S)). \label{sprr.34}
\end{align}
Moreover, the following representations hold
\begin{align}
\text{\rm{trindex}} (\Xi( S+A_1+iB_1),\Xi(S))&=
\frac{1}{\pi}\int_{\mathbb{R}} \frac{dt\,\text{\rm{index}} (\Xi(S+A_1+tB_1),
\Xi(S))}{1+t^2}, \label{sprr.35} \\
\text{\rm{trindex}} (\Xi( S+A_2+iB_2),\Xi(S))&=
\frac{1}{\pi}\int_{\mathbb{R}} \frac{dt\,\text{\rm{index}}
(\Xi(S+A_2+tB_2),\Xi(S))}{1+t^2}. \label{sprr.36}
\end{align}
Under Hypothesis~\ref{hx.1},
$(\Xi(S+A_j+tB_j)-\Xi(S))\in{\mathcal B}_\infty({\mathcal H})$, $j=1,2$
and
hence, by Theorem~\ref{ttr.3}\,(i),
\begin{align}
&\text{\rm{index}} (\Xi(S+A_1+tB_1),\Xi(S))-\text{\rm{index}}
(\Xi(S+A_2+tB_2),\Xi(S))
\nonumber \\
&=\text{\rm{index}} (\Xi(S+A_1+tB_1),\Xi(S+A_2+iB_2)). \label{sprr.37}
\end{align}
Combining \eqref{sprr.34}--\eqref{sprr.37} proves
\eqref{spr.35} and \eqref{spr.36}.
\end{proof}
Finally, we turn to a version of the Birman-Krein
formula \cite{BK62}.
\begin{theorem} \label{ttr.10}
Under Hypothesis~\ref{hx.1}, the operator
\begin{equation} \label{tr.53}
{\bf S}=I_{\mathcal H} -2iB^{1/2} (S+A+iB)^{-1}B^{1/2}
\end{equation}
is unitary, $({\bf S}-I_{\mathcal H}) \in {\mathcal B}_1({\mathcal H})$ and its
Fredholm determinant can be represented as follows
\begin{equation} \label{tr.54}
\det ({\bf S})=\exp(-2\pi i\,\text{\rm{trindex}} (\Xi(S+A+iB),\Xi(S))).
\end{equation}
\end{theorem}
\begin{proof}
Introduce the family of trace class operators
\begin{equation}\label{tr.55}
{\mathcal A}(z)=B^{1/2} (S+A+zB)^{-1}B^{1/2}, \quad \Im(z)>0.
\end{equation}
Then,
\begin{equation}\label{tr.56}
{\bf S}=I_{\mathcal H}-2i{\mathcal A}(i), \quad ({\bf S}-I_{\mathcal H})\in
{\mathcal B}_1({\mathcal H})
\end{equation}
and hence the Fredholm determinant of ${\bf S}$ is
well-defined.
By the analytic Fredholm theorem the set of
$z\in {\mathbb{C}}$ such that
$S+A+zB$ does not have a bounded inverse is discrete
and therefore there
exists a $\delta>0$ such that ${\mathcal A}(\delta)$ is
well-defined.
By Lemma~\ref{lt.4},
the operator ${\mathcal A}(\delta)$ has the eigenvalue
$\delta^{-1}$ if and only if
$\text{\rm{dim}} (\ker(S+A))\ne 0$ with associated multiplicity equal
to $\text{\rm{dim}} (\ker(S+A))$.
Let $\{\mu_k(\delta)\}_{k\in{\mathbb{N}}}=\text{\rm{spec}}({\mathcal A}(\delta))
\backslash\{\delta^{-1}\}$ different from the eigenvalue
$\delta^{-1}$ with multiplicities $\{m_k\}_{k\in{\mathbb{N}}}$.
By Lemma~\ref{lt.4}, ${\mathcal A}(i)$ has the eigenvalues
\begin{equation}\label{tr.59}
\mu_k(i)=\frac{\mu_k(\delta)}
{1-\mu_k(\delta) (\delta-i)}, \quad k\in{\mathbb{N}},
\end{equation}
with multiplicities $\{m_k\}_{k\in{\mathbb{N}}}$ and, in addition,
the eigenvalue $-i$ of multiplicity
$\text{\rm{dim}} (\ker (S+A))$ (if $(S+A)$ has a nontrivial kernel).
Moreover, ${\mathcal A}(i)$ has no other eigenvalues different
from zero.
Therefore, by \eqref{tr.56},
\begin{equation}\label{tr.60}
\det ({\bf S})=(-1)^{\text{\rm{dim}} (\ker (S+A))}\prod_{k=1}^\infty
\bigg (\frac{1-\mu_k(\delta) \delta -i \mu_k(\delta)}
{1-\mu_k(\delta) \delta +i \mu_k(\delta)}
\bigg )^{m_k}.
\end{equation}
Moreover,
\begin{align}
&\text{\rm{trindex}} (\Xi( S+A+iB), \Xi(S)) \nonumber \\
&=\tr(\Xi( S+A+iB)- \Xi(S+A) )+\text{\rm{index}}(\Xi( S+A),
\Xi(S)) \nonumber \\
&=(1/\pi) \lim_{\varepsilon \downarrow 0}
\tr(\arctan (B^{1/2} (S+A+\varepsilon B)^{-1} B^{1/2}))+
\text{\rm{index}}(\Xi( S+A), \Xi(S)) \nonumber \\
&= (1/\pi)\sum_{k=1}^\infty m_k\arctan
(\mu_k(\delta)(1-\mu_k(\delta)\delta)^{-1})+
(1/2)\text{\rm{dim}} (\ker (S+A)) +n, \label{tr.61}
\end{align}
for some $n\in{\mathbb{Z}}$. Combining \eqref{tr.60} and
\eqref{tr.61}
results in \eqref{tr.54}.
\end{proof}
To avoid additional technicalities we only treated
the case where $S$ is bounded. It is clear
that our formalism in Section~\ref{s3} extends to unbounded
dissipative operators $T$ in ${\mathcal H}$, but such an
extension will be discussed elsewhere.
\section{Some Applications} \label{s5}
The main purpose of this section is to obtain
new representations for Krein's spectral shift function
associated with a pair of self-adjoint operators
$(H_0,H)$ and to provide
a generalization
of the classical Birman-Schwinger principle, replacing
the traditional eigenvalue counting functions by
appropriate spectral shift functions.
We start with our representation of Krein's spectral shift
function and temporarily assume the
following hypothesis.
\begin{hypothesis}\label{h3.1}
Let $H_0$ be a self-adjoint operator in ${\mathcal H}$
with domain $\text{\rm{dom}} (H_0)$, $J$ a bounded self-adjoint
operator
with $J^2=I_{{\mathcal H}}$, and $K\in {\mathcal B}_2({\mathcal H})$ a
Hilbert-Schmidt operator.
\end{hypothesis}
Introducing
\begin{equation}\label{3.1}
V=KJK^*
\end{equation}
we define the self-adjoint operator $H$ in ${\mathcal H}$ by
\begin{equation}\label{3.2}
H=H_0+V, \quad \text{\rm{dom}}(H)=\text{\rm{dom}}(H_0).
\end{equation}
We could have easily incorporated the case where $K$ maps
between different Hilbert spaces but we omit the
corresponding
details. Moreover, we introduce the following bounded
operator
in ${\mathcal H}$,
\begin{equation}\label{4.28}
\Phi (z)
=J+K^*(H_0-z)^{-1}K:{\mathcal H} \rightarrow {\mathcal H}, \quad
z\in {\mathbb{C}}\backslash {\mathbb{R}}.
\end{equation}
One easily verifies (cf.~\cite[Sect.~3]{GMN99}) that
$\Phi(z)$
is an operator-valued Herglotz function in ${\mathcal H}$ (i.e.,
$\Im(\Phi(z))\geq 0$
for all $z\in{\mathbb{C}}_+$) and that
\begin{equation}\label{3.8}
\Phi (z)^{-1}=J-JK^*(H-z)^{-1}KJ, \quad z\in{\mathbb{C}}
\backslash {\mathbb{R}}.
\end{equation}
In the following it is convenient to choose
\begin{equation}
J=\text{\rm{sgn}} (V),
\end{equation}
where in the present context the sign function is
defined by
$\text{\rm{sgn}} (x)=1$ if $x\geq 0$ and $\text{\rm{sgn}} (x)=-1$ if $x<0$.
Next, let $\{P_n\}_{n\in{\mathbb{N}}}$ be a family of finite-rank
spectral projections of $V$ satisfying,
\begin{equation}\label{4.5}
\slim_{n\to \infty}P_n=I_{{\mathcal H}}.
\end{equation}
Introducing the finite-rank operators
\begin{equation}
K_n=KP_n, \quad
V_n=(K_n)J(K_n)^*, \quad n\in {\mathbb{N}} \label{4.4},
\end{equation}
one infers (cf.~e.g., \cite{Gr73})
\begin{equation}\label{4.5a}
\lim_{n\to \infty} \|V_n-V\|_{{\mathcal B}_1({{\mathcal H}})}.
\end{equation}
Together with the operator-valued Herglotz function
$\Phi(z)$ given by
\eqref{4.28}, we introduce the family of operator-valued
Herglotz functions
\begin{equation}\label{4.37}
\Phi_n (z)
=J+K_n^*(H_0-z)^{-1}K_n:{\mathcal H} \rightarrow {\mathcal H},
\end{equation}
and its finite-rank restriction
\begin{equation}\label{4.38}
\Psi_n (z)=P_n\Phi_n(z)P_n :P_n{\mathcal H} \rightarrow P_n{\mathcal H}.
\end{equation}
One computes as in \eqref{3.8},
\begin{equation} \label{4.39}
\Phi_n^{-1} (z)=J-JK_n^*(H_n-z)^{-1}K_nJ,
\end{equation}
where
\begin{equation} \label{4.40}
H_n=H_0+V_n, \quad \text{\rm{dom}} (H_n)=\text{\rm{dom}} (H_0).
\end{equation}
Consequently,
\begin{equation}\label{4.41}
\Psi_n^{-1} (z)=\Phi_n^{-1} (z)|_{P_n{\mathcal H}} \text{ in }
P_n{\mathcal H}.
\end{equation}
Since $\Psi_n(z)$, $z\in {\mathbb{C}}_+$ is invertible in
$P_n{\mathcal H}$,
one infers from \cite{GT97} (see also, \cite{Ca76}) the
existence of a family of operators $\{\Xi_n (\lambda)\}$
defined for (Lebesgue) a.e.~$\lambda\in {\mathbb{R}}$, satisfying
\begin{equation}\label{4.42}
0\le \Xi_{n}(\lambda)\le I_n
\text{ for a.e. } \lambda\in {\mathbb{R}},
\end{equation}
where $I_n$ denotes the identity operator in $P_n{\mathcal H}$,
and
\begin{align}
\log (\Psi_n(z))&=C_n +\int_{\mathbb{R}} d\lambda \, \Xi_n(\lambda)
((\lambda-z)^{-1}-\lambda (1+\lambda^2)^{-1}), \quad
z\in {\mathbb{C}}\backslash {\mathbb{R}}, \label{4.43} \\
C_n &=C_n^*=\Re(\log \Psi_n(i)), \label{4.44} \\
\Xi_n(\lambda)&=\pi^{-1}\lim_{\varepsilon \downarrow 0}
\Im(\log (\Psi_n(\lambda+i\varepsilon ))
\text{ for a.e. } \lambda \in {\mathbb{R}}. \label{4.45}
\end{align}
Next, we briefly recall the notion of Krein's spectral
shift
function for a pair of self-adjoint operators in ${\mathcal H}$
(cf.~e.g., \cite[Sect.~19.1]{BW83}, \cite{BK62},
\cite{BP98},
\cite{BY93}, \cite{BY93a}, \cite{Ka78}, \cite{KS98},
\cite{Kr62},
\cite{Kr83}, \cite{Kr89}, \cite{KJ81}, \cite{Pu97},
\cite{Pu98a}, \cite{Si75}, \cite[Ch.~8]{Ya92} and the
literature cited therein), a concept originally
introduced by
Lifshitz \cite{Li52}, \cite{Li56}. Assuming $\text{\rm{dom}}(H_0)
=\text{\rm{dom}}(H)$ and $(H-H_0)\in{\mathcal B}_1({\mathcal H})$ (this could be
considerably relaxed but suffices for our present purpose),
Krein's (real-valued) spectral shift function
$\xi(\lambda,H_0,H)$ is uniquely defined for~a.e.
$\lambda\in{\mathbb{R}}$ by
\begin{align}
&\xi(\cdot,H_0,H)\in L^1({\mathbb{R}};d\lambda), \nonumber \\
& \tr((H-z)^{-1}-(H_0-z)^{-1})=-\int_{\mathbb{R}} d\lambda\,
\xi(\lambda,H_0,H)(\lambda -z)^{-2}, \quad
z\in{\mathbb{C}}\backslash{\mathbb{R}}. \label{4.45aa}
\end{align}
\begin{lemma}\label{l4.7}
Denote by $\xi(\lambda,H_0,H_n)$ the spectral shift
function
associated with the pair $(H_0,H_n)$. Then
\begin{equation}
\xi(\lambda,H_0,H_n)=
\tr_{P_n{\mathcal H}} ( \Xi_n(\lambda))-N_n \text{ for a.e. }
\lambda \in {\mathbb{R}}, \label{4.45a}
\end{equation}
where
\begin{equation}
N_n=\# \{ \text{of strictly negative eigenvalues of } V_n,
\text{\,counting multiplicity\,}\}. \label{4.45b}
\end{equation}
\end{lemma}
\begin{proof} Differentiating $\tr_{P_n{\mathcal H}}
(\log (\Psi_n(z)))$
with respect to $z$ (c.f., \cite[Sect.~IV.1]{GK69})
\begin{equation}\label{4.46}
(d/dz)\tr_{P_n{\mathcal H}}(\log(\Psi_n(z))=
\tr_{P_n{\mathcal H}}(\Psi_n^{-1}(z)\Psi^{\prime}_n(z)),
\end{equation}
one obtains by \eqref{4.37} and \eqref{4.41}
\begin{equation}
\tr_{P_n{\mathcal H}}(\Psi_n^{-1}(z)\Psi^{\prime}_n(z))=
\tr_{{\mathcal H}}(P_n\Phi_n^{-1}(z)P_nP_n\Phi^{\prime}_n(z)P_n)=
\tr_{{\mathcal H}}(\Phi_n^{-1}(z)\Phi^{\prime}_n(z)),
\label{4.47}
\end{equation}
since $(d/dz)\Phi_n(z)=(d/dz)P_n\Phi_n(z)P_n$.
However, $\tr_{{\mathcal H}}(\Phi_n^{-1}(z)\Phi^{\prime}_n(z))$
can be computed explicitly,
\begin{align}
& \tr_{{\mathcal H}}((\Phi_n^{-1}(z)\Phi^{\prime}_n(z)) \nonumber \\
&=\tr_{{\mathcal H}}(J-JK_n^*(H_n-z)^{-1}K_nJ)
K_n^*(H_0-z)^{-2}K_n)
\nonumber \\
&=\tr_{{\mathcal H}}(K_n(J-JK_n^*(H_n-z)^{-1}
K_nJ)K_n^*(H_0-z)^{-2})
\nonumber \\
&=\tr_{{\mathcal H}}(H_0-z)^{-1}K_n(J-JK_n^*(H_n-z)^{-1}
K_nJ)K_n^*(H_0-z)^{-1})
\nonumber \\
&=\tr((H_0-z)^{-1}K_nJK_n^*(H_0-z)^{-1}) \nonumber \\
&-\tr_{{\mathcal H}}(
(H_0-z)^{-1}K_nJK_n^*(H_n-z)^{-1}K_nJK_n^*(H_0-z)^{-1})
\nonumber \\
&=\tr_{{\mathcal H}}((H_0-z)^{-1}V_n(H_0-z)^{-1}-
(H_0-z)^{-1}V_n(H_n-z)^{-1}V_n(H_0-z)^{-1}) \nonumber \\
&=-\tr_{{\mathcal H}}( (H_n-z)^{-1}-(H_0-z)^{-1})
\quad z\in {\mathbb{C}}\backslash {\mathbb{R}}, \label{4.49}
\end{align}
iterating the second resolvent identity.
Combining \eqref{4.46}--\eqref{4.49} one infers
\begin{equation}\label{4.50}
(d/dz)\tr_{P_n{\mathcal H}}(\log(\Psi_n(z))=
-\tr_{{\mathcal H}}( (H_n-z)^{-1}-(H_0-z)^{-1}).
\end{equation}
Taking traces in \eqref{4.43} one gets
\begin{equation}\label{4.52}
\tr_{P_n{\mathcal H}}(\log (\Psi_n(z)))=
\tr_{P_n{\mathcal H}}(C_n) +
\int_{\mathbb{R}} d\lambda \,\tr_{P_n{\mathcal H}}( \Xi_n(\lambda))
((\lambda-z)^{-1}-\lambda (1+\lambda^2)^{-1}),
\end{equation}
and thus, differentiating \eqref{4.52} with respect to $z$,
\begin{equation}\label{4.53}
(d/dz)\tr(\log (\Psi_n(z)))=
\int_{\mathbb{R}} d\lambda \,\tr( \Xi_n(\lambda))
(\lambda-z)^{-2}.
\end{equation}
Comparing \eqref{4.53} and \eqref{4.50} one arrives at
the trace formula
\begin{equation}
\int_{\mathbb{R}} d\lambda \,\tr( \Xi_n(\lambda))
(\lambda-z)^{-2}=-\tr( (H_n-z)^{-1}-(H_0-z)^{-1}),
\quad z\in{\mathbb{C}}\backslash {\mathbb{R}}, \label{4.54}
\end{equation}
and hence up to an additive constant, $\tr( \Xi_n(\lambda))$
coincides with the spectral shift function
$\xi(\lambda,H_0,H_n)$ associated with the pair $(H_0,H_n)$
for a.e.~$\lambda \in {\mathbb{R}}$.
Next we determine this constant. Introducing
$J_n=J\big|_{P_n{\mathcal H}}$, $J_n^2=I_n$, one obtains
\begin{align}
\log(J_n)&=i\Im(\log(J_n)) \nonumber \\
&=\pi^{-1}\int_{\mathbb{R}} d\lambda \,\Im(\log(J_n) )
((\lambda-z)^{-1}-\lambda (1+\lambda^2)^{-1}). \label{4.43a}
\end{align}
Moreover,
\begin{equation}\label{4.43b}
\pi^{-1}\tr_{P_n{\mathcal H}}(\Im
(\log ((J_n)))=N_n,
\end{equation}
where $N_n$ denotes the number of strictly negative
eigenvalues
of $J_n$, counting multiplicity. Thus, $N_n$ coincides
with the
number of strictly negative eigenvalues of $V_n$. Combining
\eqref{4.43a}, \eqref{4.43b}, and using \eqref{4.43}
results in
\begin{align}
&\log (\Psi_n(z))-\log(J_n) \nonumber \\
&=C_n +\int_{\mathbb{R}} d\lambda \, (\Xi_n(\lambda)-\pi^{-1}\Im
(\log ((J_n)))((\lambda-z)^{-1}-\lambda (1+\lambda^2)^{-1})
\label{4.44b}
\end{align}
and hence in
\begin{equation}
\tr_{P_n{\mathcal H}}(\Im(\log (\Psi_n(z))-\log(J_n)))
=\int_{\mathbb{R}} d\lambda \,(\tr( \Xi_n(\lambda))-N_n)
\frac{\Im(z)}{(\lambda-\Re(z))^{^2}+\Im(z)^2}. \label{4.56}
\end{equation}
Since $\vert\vert (H_0-iy)^{-1}\vert\vert =O(\vert y
\vert^{-1}) $ as $ y \uparrow +\infty$, one concludes
\begin{equation}\label{4.57}
y\vert\vert
\log (\Psi_n(iy))-\log(J_n) \vert\vert=O
(1) \text{ as } y \uparrow + \infty
\end{equation}
and hence that
$y\Im(\tr_{P_n{\mathcal H}} (\log(\Psi_n(iy))-\log (J_n)))$
is bounded $\text{ as } y \uparrow +\infty$.
In particular, \eqref{4.56} and \eqref{4.57} imply that
\begin{equation}\label{4.58}
\xi_n(\lambda)=\tr_{P_n{\mathcal H}} (\Xi_n(\lambda))-N_n
\end{equation}
is integrable,
\begin{equation}\label{4.60}
\xi_n(\cdot) \in L^1({\mathbb{R}};d\lambda).
\end{equation}
Since
\begin{equation}
\int_{\mathbb{R}} d\lambda(\lambda-z)^{-2}=0 \text{ for all }
z\in{\mathbb{C}}, \, \Im(z)\ne 0,
\end{equation}
\eqref{4.58} and \eqref{4.54} yield the trace formula
\begin{align}
&\int_{\mathbb{R}} d\lambda \,\xi_n(\lambda)(\lambda-z)^{-2}=
\int_{\mathbb{R}} d\lambda \,\tr_{P_n{\mathcal H}}( \Xi_n(\lambda))
(\lambda-z)^{-2} \nonumber \\
&=-\tr_{{\mathcal H}}( (H_n-z)^{-1}-(H_0-z)^{-1}),\quad z\in
{\mathbb{C}}\backslash {\mathbb{R}},
\label{4.62}
\end{align}
which together with \eqref{4.60} proves
\eqref{4.45a}, \eqref{4.45b}.
\end{proof}
\begin{theorem}\label{t5.8}
Assume Hypothesis~\ref{h3.1} and fix a $p>1$. Moreover,
let
$V=KJK^*$, where $J=\text{\rm{sgn}} (V)$. Then the spectral shift
function
$\xi(\lambda,H_0,H)$ associated with the pair $(H_0,H)$,
$H=H_0+V$ admits the representation
\begin{equation}
\xi(\lambda,H_0,H)=\text{\rm{trindex}}(\Xi(J+K^*(H_0-
\lambda-i0)^{-1}K), \Xi(J)) \text{ for a.e. }
\lambda \in{\mathbb{R}}, \label{5.52}
\end{equation}
where $K^*(H_0-\lambda-i0)^{-1}K$ is defined as
\begin{equation}
\lim_{\varepsilon\downarrow 0}\|K^*(H_0-\lambda-i0)^{-1}K-
K^*(H_0-\lambda-i\varepsilon)^{-1}K\|_{{\mathcal B}_p({\mathcal H})}=0
\text{ for a.e. } \lambda \in {\mathbb{R}}. \label{4.1}
\end{equation}
\end{theorem}
\begin{proof}
First of all, one notes that the boundary values
$K^*(H_0-\lambda-i0)^{-1}K$ and $K^*(H-\lambda-i0)^{-1}K$
exist
$\lambda$ a.e. in the topology ${\mathcal B}_p({\mathcal H})$ for
every $p>1$ (but in general not for $p=1$,) \cite{Na89},
\cite{Na90} (see also \cite[Ch.~3]{BW83},
\cite{BE67}, \cite{de62}). By
\eqref{3.8}, the operator
$J+K^*(H_0-\lambda-i0)^{-1}K$ has a bounded inverse
for a.e.
$\lambda\in{\mathbb{R}}$. Moreover (see, e.g.,
\cite[Sect.~I.3.4]{BW83}),
\begin{align}
&\lim_{\varepsilon \downarrow 0} \|\Im(K^*(H_0-
\lambda-i\varepsilon)^{-1}K)-\Im(K^*(H_0-
\lambda-i0)^{-1}K)\|_{{\mathcal B}_1({\mathcal H})}=0 \label{5.1} \\
& \hspace*{8.6cm} \text{ for a.e. } \lambda \in {\mathbb{R}}. \nonumber
\end{align}
Thus, there exists a set $\Lambda\subset{\mathbb{R}}$ with
$|{\mathbb{R}}\backslash\Lambda|=0$ ($|\cdot|$ denoting Lebesgue
measure on ${\mathbb{R}}$) satisfying the following properties.
For any $\lambda\in \Lambda:$
\noindent (i) The boundary values
$\Phi(\lambda+i0)=\lim_{\varepsilon
\downarrow 0}\Phi(\lambda+i\varepsilon)$ exist in
${\mathcal B}_p({\mathcal H})$-topology (cf. \eqref{4.28}).
\noindent (ii) The operator $\Phi(\lambda+i0)$ has a
bounded
inverse.
\noindent (iii) $\Im(\Phi(\lambda +i\varepsilon))$
converges
to $\Im(\Phi(\lambda+i0))$ as $\varepsilon \downarrow 0$
in ${\mathcal B}_1({\mathcal H})$-topology.
For any $n\in {\mathbb{N}}$, $\lambda\in \Lambda$ introduce
the function
\begin{equation}\label{4.6}
\xi_n(\lambda)=\text{\rm{trindex}}( \Xi(P_n\Phi(\lambda+i0)P_n,
\Xi(P_nJP_n)).
\end{equation}
Since $P_n$ commute with $J$ and the subspace $P_n {\mathcal H}$
is invariant for
$J+P_nK^*(H_0-\lambda-i0)^{-1}KP_n$ one
concludes by \eqref{4.6} that (cf. \eqref{2.20})
\begin{align}
\xi_n(\lambda)&=\tr ( \Xi(J+P_n
K^*(H_0-\lambda-i0)^{-1}KP_n)-\Xi(J)) \nonumber \\
&=\text{\rm{trindex}} ( \Xi(J+P_n
K^*(H_0-\lambda-i0)^{-1}KP_n), \Xi(J)), \quad
\lambda \in \Lambda. \label{4.7}
\end{align}
On the other hand, by Lemma~\ref{l4.7}, \eqref{4.37},
and \eqref{4.38} one infers
that the function $\xi_n(\lambda)$ coincides with the
spectral shift function
$\xi(\lambda, H_0, H_n)$ associated with the pair
$(H_0, H_n)$
\begin{equation}\label{4.79}
\xi_n(\lambda)=\xi(\lambda, H_0, H_n),\text{ a.e. }
\lambda \in {\mathbb{R}},
\end{equation}
where $H_n$ is given
by \eqref{4.40}.
By a result of Gr\"umm \cite{Gr73}, properties
(i)--(iii), and \eqref{4.5}, one obtains for
$\lambda\in \Lambda$,
\begin{align}
&\lim_{n\to \infty}
\|(\Re (P_nK^*(H_0-\lambda-i0)^{-1}KP_n)
-\Re ( K^*(H_0-\lambda-i0)^{-1}K)\|_{
{\mathcal B}_p({\mathcal H})}=0, \label{4.8} \\
&\lim_{n\to \infty}
\|(\Im ( P_nK^*(H_0-\lambda-i0)^{-1}KP_n)-\Im (
K^*(H_0-\lambda-i0)^{-1}K)\|_{{\mathcal B}_1({\mathcal H})}=0.
\label{4.9}
\end{align}
Applying the approximation Theorem~\ref{ttr.7} then yields
\begin{equation}\label{4.10}
\lim_{n\to \infty} \xi_n(\lambda)=
\text{\rm{trindex}}( \Xi(\Phi(\lambda+i0), \Xi(J)), \quad
\lambda \in \Lambda \text{ pointwise}.
\end{equation}
Convergence \eqref{4.5a} of $V_n$ to $V$ in trace norm
implies the convergence of the corresponding spectral
shift
functions $\xi_n(\lambda)$
to the spectral shift function $\xi(\lambda,H_0,H)$ in
$L^1({\mathbb{R}})$. This in turn implies the existence of a
subsequence
$\{ \xi_{n_k}(\lambda)\}_{k\in{\mathbb{N}}}$ converging pointwise
a.e. Together with \eqref{4.10} this proves \eqref{5.52}.
\end{proof}
\begin{remark}\label{rindex}
Let $\Lambda=\{\lambda\in{\mathbb{R}}\,|\, \text{s.\,t. } A(\lambda)
\text{ and } B(\lambda)
\text{ exist, } A(\lambda)\in{\mathcal B}_2({\mathcal H}),
B(\lambda)\in{\mathcal B}_1({\mathcal H}), \text{ and }
(J+A(\lambda)+iB(\lambda))^{-1}\in{\mathcal B}({\mathcal H}) \}$. Then,
as shown in Corollary~\ref{iff}, the condition
\begin{equation}
[A(\lambda),J]\notin {\mathcal B}_1({\mathcal H}) \label{5.49}
\end{equation}
is necessary and sufficient for the validity of
\begin{equation}
(\Xi(J+A(\lambda)+iB(\lambda))-\Xi(J))\notin
{\mathcal B}_1({\mathcal H}).
\label{5.53}
\end{equation}
Next, note that the following three conditions,
\begin{align}
(i) \, &\text{\rm{rank}} (E_V((-\infty,0))=\text{\rm{rank}} (E_V((0,
\infty))=\infty, \\
(ii)\, & \lambda\in \text{\rm{ess.spec}} (H_0), \\
(iii)\, & A(\lambda)\notin {\mathcal B}_1({\mathcal H}),
\end{align}
are a consequence of condition \eqref{5.49}.
Thus, if at least one of the conditions (i)--(iii) is
violated,
the $\xi$-function (c.f., \eqref{5.52}) can be
represented in
the simple form
\begin{equation}
\xi(\lambda,H_0,H)=\tr(\Xi(J+K^*(H_0-
\lambda-i0)^{-1}K)- \Xi(J)),
\label{rind.1}
\end{equation}
and the concept of a trindex becomes redundant in this case.
On the other hand, there are of course examples (c.f.,
Remark~\ref{rce}), where
\begin{equation}
|\{\lambda \in {\mathbb{R}} \, |\, (\Xi(J+|V|^{1/2}(H_0-
\lambda-i0)^{-1}|V|^{1/2})- \Xi(J))\notin
{\mathcal B}_1({\mathcal H}) \}|
>0,
\label{5.53a}
\end{equation}
with $|\cdot |$ denoting Lebesgue measure on ${\mathbb{R}}$. A
concrete
example illustrating \eqref{5.53a} can be constructed as
follows.
Consider an infinite dimensional complex separable
Hilbert space
${\mathcal K}$, an operator $0\leq
k=k^*\in{\mathcal B}_2({\mathcal K})\backslash{\mathcal B}_1({\mathcal K})$, with
$\ker(k)=\{0\}$, and a self-adjoint operator $h_0$ in
${\mathcal K}$ such
that
\begin{equation}
a(\lambda)=\nlim_{\varepsilon\to 0}\Re(k(h_0-\lambda
-i\varepsilon)^{-1}k)\notin{\mathcal B}_1({\mathcal K}) \text{ for~a.e. }
\lambda\in{\mathbb{R}}. \label{5.54}
\end{equation}
Existence of such ${\mathcal K}$, $k$, and $h_0$ can be
inferred from
\cite{Na89}. Next, define ${\mathcal H}={\mathcal K}\oplus\,{\mathcal K}$ and
introduce $H_0=\left(\begin{smallmatrix} h_0& 0 \\ 0 & 0
\end{smallmatrix}\right)$, $V=
i\left(\begin{smallmatrix} 0& k^2
\\ -k^2 & 0 \end{smallmatrix}\right)$. Then $J=\text{\rm{sgn}}(V)=
i\left(\begin{smallmatrix} 0& I_{\mathcal K} \\ -I_{\mathcal K} & 0
\end{smallmatrix}\right)$, $|V|^{1/2}=
\left(\begin{smallmatrix}
k & 0 \\ 0 & k \end{smallmatrix}\right)$, and
\begin{align}
A(\lambda)=\nlim_{\varepsilon\to
0}
\Re(|V|^{1/2}(H_0-\lambda-i\varepsilon)^{-1}|V|^{1/2})
&=\begin{pmatrix}a(\lambda) & 0 \\ 0
& -(1/\lambda)k^2 \end{pmatrix} \label{5.55}
\end{align}
for~a.e.~$\lambda\in{\mathbb{R}}\backslash\{0\}$. Moreover, one computes
\begin{equation}
[A(\lambda),J]=i\begin{pmatrix} 0& a(\lambda)
+(1/\lambda)k^2 \\ a(\lambda)+(1/\lambda)k^2 & 0
\end{pmatrix} \notin {\mathcal B}_1({\mathcal H}) \text{ for~a.e. }
\lambda\in{\mathbb{R}}, \label{5.56}
\end{equation}
since $k^2\in{\mathcal B}_1({\mathcal K})$.
\end{remark}
As a consequence of Theorems~\ref{ttr.8} and \ref{t5.8}
one has the following representation for the spectral
shift function via the integral of the index of a
Fredholm pair of spectral projections.
\begin{theorem} \label{main}
Assume Hypothesis~\ref{h3.1} and introduce $V=KJK^*$. In
addition, for a.e.~$\lambda\in{\mathbb{R}}$, let
$A(\lambda)+iB(\lambda)$ be the normal boundary values
of the
operator-valued Herglotz function $K^*(H_0-z)^{-1}K$
on the real axis, that is,
\begin{equation}
A(\lambda)=\nlim_{\varepsilon \downarrow 0}
\Re(K^*(H_0-\lambda-i\varepsilon)^{-1}K)
\text{ for a.e. } \lambda\in{\mathbb{R}}
\end{equation}
and
\begin{equation}
B(\lambda)=\nlim_{\varepsilon \downarrow 0}
\Im(K^*(H_0-\lambda-i\varepsilon)^{-1}K)
\text{ for a.e. } \lambda\in{\mathbb{R}}.
\end{equation}
Then
\begin{align}
&\xi(\lambda, H_0,H)=\frac{1}{\pi} \int_{\mathbb{R}} dt
\frac{\text{\rm{index}} \big(E_{J+A(\lambda)+tB(\lambda)}
((-\infty,0)),E_J((-\infty,0))\big )}{1+t^2}
\label{pus} \\
&\hspace{8.1cm} \text{for a.e. } \lambda\in{\mathbb{R}}. \nonumber
\end{align}
\end{theorem}
In the particular case of
sign-definite perturbations, that is,
$J=I_{\mathcal H}$ or $J=-I_{\mathcal H}$, applying Lemma~\ref{ltrr.9}
yields the following result originally due to
Pushnitski~\cite{Pu97}, representing the spectral shift
function in terms of an integrated eigenvalue
counting function.
\begin{corollary}\label{c5.6} \mbox{\rm
(Pushnitski~\cite{Pu97}.)}
Let $0\le V\in {\mathcal B}_1({\mathcal H})$ and $H_0=H_0^*$. Then the
spectral shift
function $\xi(\lambda, H_0, H_0\pm V)$ associated
with the pair $(H_0, H_0\pm V)$ admits the representation
\begin{equation} \label{puss}
\xi(\lambda, H_0,H_0\pm V)=\pm\frac{1}{\pi} \int_{\mathbb{R}} dt
\frac{\text{\rm{rank}} (E_{\mp (A(\lambda)+
tB(\lambda)}((1, \infty)))}
{1+t^2}.
\end{equation}
\end{corollary}
\begin{remark} \label{r5.5}
(i) Strictly speaking, a direct application of
Lemma~\ref{ltrr.9} in the case
of nonpositive perturbations would give the representation
\begin{equation}
\xi(\lambda, H_0,H_0- V)=-\frac{1}{\pi} \int_{\mathbb{R}} dt
\frac{\text{\rm{rank}} (E_{ A(\lambda)+tB(\lambda)}
([1, \infty)))}
{1+t^2},
\end{equation}
which, however, yields the same result as in
\eqref{puss}, since
\begin{equation}
\int_{\mathbb{R}} dt
\frac{\text{\rm{rank}} (E_{ A(\lambda)+tB(\lambda)}(\{1\}))}
{1+t^2}=0 \text{ for a.e. }\lambda\in {\mathbb{R}}.
\end{equation}
(ii) In the special case where $\lambda \in
{\mathbb{R}}\backslash\{\text{\rm{spec}} (H_0)\cup \text{\rm{spec}} (H)\}$,
\eqref{pus} turns into
\begin{align}
\xi(\lambda, H_0,H)&=\text{\rm{index}}(E_{J+A(\lambda)}((-\infty, 0)),
E_J((-\infty, 0)))
\nonumber \\
&=\tr (E_{J+A(\lambda)}((-\infty, 0))-
E_J((-\infty, 0)))
. \label{5.52a}
\end{align}
In the particular cases of sign-definite perturbations
($J=\pm I_{\mathcal H}$) one obtains
\begin{equation}
\xi(\lambda, H_0,H_0\pm V)=\pm\text{rank}
(E_{\mp A(\lambda)}((1, \infty))), \label{sobo}
\end{equation}
since $\text{rank} (E_{\mp A(\lambda)}(\{1\}))=0$ for
a.e.~$\lambda\in{\mathbb{R}}$.
The result \eqref{sobo} is due to Sobolev \cite{So93}.
\end{remark}
\vspace*{3mm}
The trindex representation \eqref{5.52} for $\xi(\lambda,H_0,H)$
enables us to introduce a new generalized spectral
shift function outside the trace-class perturbation scheme
under rather weak assumptions on $H-H_0$. First we recall the
following exponential representation for operator-valued Herglotz
functions partially proven in \cite{GMN99}.
\begin{theorem}\label{t5.8a}
Suppose $M:{\mathbb{C}}_+ \rightarrow {\mathcal B}({\mathcal H})$ is an
operator-valued Herglotz function
and $M(z_0)^{-1} \in {\mathcal B}({\mathcal H})$ for some {\rm (}and hence
for all{\rm \,)} $z_0\in{\mathbb{C}}_+$. Then
there exists a family of bounded self-adjoint
weakly {\rm (}Lebesgue{\rm \,)} measurable operators
$\{\widehat\Xi(\lambda) \}_{\lambda\in {\mathbb{R}}}\subset {\mathcal B}({\mathcal H})$,
\begin{equation}\label{5.64}
0\le \widehat\Xi(\lambda)\le I_{\mathcal H} \text{ for a.e. }
\lambda\in {\mathbb{R}}
\end{equation}
such that
\begin{equation}\label{5.65}
\log(M(z))=C+
\int_{\mathbb{R}} d \lambda \, \widehat\Xi(\lambda)
((\lambda-z)^{-1}-\lambda(1+\lambda^2)^{-1}),
\quad z\in {\mathbb{C}}_+
\end{equation}
the integral taken in the weak sense, where $C=C^*=\Re(\log(M(i))) \in
{\mathcal B}({\mathcal H})$.
Moreover, suppose there exists a measurable subset $\Lambda\in{\mathbb{R}}$,
$|\Lambda|\neq 0$ such that
\begin{equation}
\nlim_{\varepsilon\to 0}M(\lambda+i\varepsilon)
=M(\lambda+i0)\in{\mathcal B}({\mathcal H}) \text{ for a.e. }
\lambda\in\Lambda \label{5.66}
\end{equation}
such that $\Im(M(\lambda+i0))\geq 0$ and
$M(\lambda+i0)^{-1}\in{\mathcal B}({\mathcal H})$ for a.e.~$\lambda\in\Lambda$. Then
\begin{equation}
\widehat \Xi(\lambda) =\Xi(M(\lambda+i0)) \text{ for a.e. }
\lambda\in\Lambda. \label{5.67}
\end{equation}
\end{theorem}
\begin{proof}
Since \eqref{5.64} and \eqref{5.65} have been proven in \cite{GMN99},
we focus on \eqref{5.67}. Let $\{P_n\}_{n\in{\mathbb{N}}}$ be an increasing
family of orthogonal projections of rank $n$, that is, $\text{\rm{rank}}(P_n)=n$,
$P_n{\mathcal H}\subset P_{n+1}{\mathcal H}$, with
\begin{equation}
\slim_{n\to \infty} P_n=I_{\mathcal H}. \label{5.67a}
\end{equation}
Combining the norm continuity of the logarithm of bounded dissipative
operators as discussed in Section~2 of \cite{GMN99} with the
exponential Herglotz
representation for $P_n\log(M(z))P_n$ (i.e., the finite-dimensional analog
of \eqref{5.65} as in \eqref{4.43}--\eqref{4.45}), one infers for each
$n\in{\mathbb{N}}$ the existence of a subset $\Lambda_n\subseteq\Lambda$,
$|\Lambda\backslash\Lambda_n|=0$ such that
\begin{equation}
P_n\widehat\Xi(\lambda)P_n=P_n\Xi(M(\lambda+i0))P_n \text{ for all }
\lambda\in\Lambda_n. \label{5.67b}
\end{equation}
Thus,
\begin{equation}
P_n\widehat\Xi(\lambda)P_n=P_n\Xi(M(\lambda+i0))P_n \text{ for all }
\lambda\in\bigcap_{m\in{\mathbb{N}}}\Lambda_m, \,\, n\in{\mathbb{N}}. \label{5.67c}
\end{equation}
Since $|\Lambda\backslash\bigcap_{n\in{\mathbb{N}}}\Lambda_n|=0$ one concludes
\eqref{5.67}.
\end{proof}
Next, assuming $H_0$ and $V$ to be self-adjoint operators in ${\mathcal H}$
with corresponding domains $\text{\rm{dom}} (H_0)$ and $\text{\rm{dom}}(V)$, such
that
\begin{equation}
\text{\rm{dom}}(|V|^{1/2})\supseteq \text{\rm{dom}}(|H_0|^{1/2}), \label{5.68}
\end{equation}
and introducing the signature operator
\begin{equation}
J=\text{\rm{sgn}}(V) \text{ with } J|_{\ker(V)}=I_{\mathcal H}|_{\ker(V)},
\label{5.69}
\end{equation}
we may define the bounded operator-valued Herglotz function
$\phi(z)\in{\mathcal B}({\mathcal H})$
\begin{equation}
\phi(z)=J+\overline{|V|^{1/2}(H_0-z)^{-1}|V|^{1/2}}\, , \quad z\in{\mathbb{C}}_+.
\label{5.70}
\end{equation}
(here the bar denotes the operator closure in ${\mathcal H}$.)
In addition, we suppose that
\begin{equation}
\phi(z)^{-1}\in{\mathcal B}({\mathcal H}) \text{ for some (and hence for all) }
z\in{\mathbb{C}}_+. \label{5.70a}
\end{equation}
Applying Theorem~\ref{t5.8a}, one
concludes that $\phi(z)$ admits the representation
\begin{align}
&\log (\phi(z))=C +\int_{\mathbb{R}} d\lambda \, \widehat \Xi(\lambda,H_0,H)
((\lambda-z)^{-1}-\lambda (1+\lambda^2)^{-1}), \quad
z\in {\mathbb{C}}\backslash {\mathbb{R}}, \label{5.71} \\
&C =C^*=\Re(\log \phi(i)), \quad
0\le \widehat\Xi(\lambda,H_0,H)\le I_{\mathcal H}
\text{ for a.e. } \lambda\in {\mathbb{R}}. \label{5.72}
\end{align}
Here our notation $\widehat\Xi(\lambda,H_0,H)$ emphasizes the underlying
pair $(H_0,H)$, where $H$``$=$''$H_0+V$ formally represents the
perturbation of $H_0$ by $V$. We will return to a discussion of this
point in Remark~\ref{r5.9a} below.
\begin{definition}\label{d5.9}
In addition to \eqref{5.68}--\eqref{5.72} assume the existence of
a (Lebesgue) measurable set $\Lambda$, $|\Lambda|\neq 0$, such that the
pair
$(\widehat\Xi(\lambda,H_0,H),\Xi(J))$ has a trindex for
a.e.~$\lambda\in\Lambda$. Then the {\it generalized spectral shift
function} $\widehat\xi(\cdot,H_0,H)$
associated with the pair $(H_0,H)$ is defined by
\begin{equation}
\widehat\xi(\lambda,H_0,
H)=\text{\rm{trindex}}(\widehat\Xi(\lambda,H_0,H), \Xi(J)) \text{ for a.e. }
\lambda\in \Lambda. \label{5.73}
\end{equation}
\end{definition}
\begin{remark} \label{r5.9a}
A close look at $\phi(z)$ and $\widehat\Xi$ in \eqref{5.71} and \eqref{5.72}
reveals that both objects depend on the self-adjoint operators $H_0$ and
$V$. Thus, a logical choice of notation for $\widehat\Xi$ would have
indicated its dependence on the pair $(H_0,V)$. We decided against that
choice since in practical applications, $\widehat\Xi$ in \eqref{5.71} is
associated with a pair of self-adjoint operators $(H_0,H)$, where $H$
results as an additive perturbation of $H_0$ by $V$ and hence resorted
to the more familiar notation $\widehat\Xi(\lambda,H_0,H)$. But this raises
the question of how to define such a self-adjoint operator $H$, given
$H_0$ and $V$. Perhaps the most natural solution of this problem in our
context goes back to Kato \cite{Ka66} (see also \cite{KK66}) and proceeds
as follows. One defines the resolvent of the self-adjoint operator $H$ in
${\mathcal H}$ (and hence $H$ itself) by
\begin{align}
(H-z)^{-1}&=(H_0-z)^{-1} \label{5.73c} \\
& \quad -(|V|^{1/2}(H_0-\overline z)^{-1})^*
\phi(z)^{-1}\overline {|V|^{1/2}(H_0-z)^{-1}}, \quad
z\in{\mathbb{C}}_+. \nonumber
\end{align}
A detailed discussion of this point of view can be found in Yafaev's
monograph \cite[Sects.~1.9, 1.10]{Ya92}.
\end{remark}
\vspace*{3mm}
By Theorem~\ref{t5.8}, the generalized spectral shift function coincides
with Krein's spectral shift function in the case of trace class
perturbations, that is,
\begin{equation}
\widehat\xi(\lambda,H_0,H)=\xi(\lambda,H_0,H) \text{ for a.e. }
\lambda\in{\mathbb{R}}, \label{5.73a}
\end{equation}
using the standard factorization of $(H-H_0)$ into
$(H-H_0)=|V|^{1/2}\text{\rm{sgn}}(V)|V|^{1/2}\in{\mathcal B}_1({\mathcal H})$.
\begin{lemma}\label{l5.10}
Let $S$ be a signature operator, $S=S^*=S^{-1}$,
$A=A^*\in {\mathcal B}_{\infty}({\mathcal H})$, and $\Lambda={\mathbb{R}}\backslash
\{\text{\rm{spec}}(S+A)\cup \{-1,1\}\}$. Then the generalized spectral
shift function $\widehat\xi (\lambda, S+A,S)$ associated with
the pair $(S+A,S)$ is well-defined for a.e.
$\lambda\in\Lambda$. Moreover, $\widehat\xi (\lambda,
S+A,S)$ has a continuous representative on $\Lambda$ \rm{(}still denoted
by
$\widehat\xi (\lambda,S+A,S)$\rm{)} such that
\begin{equation}
\widehat\xi (\lambda, S+A,S)=\text{\rm{index}}(E_{S+A}((-\infty,\lambda )),
E_{S}((-\infty,\lambda)) ),\quad \lambda\in \Lambda.
\label{5.74}
\end{equation}
In particular, taking $\lambda\uparrow 0$,
\begin{equation}
\widehat\xi (0_-, S+A,S)=\text{\rm{index}}(E_{S+A}((-\infty,0 )), E_{S}((-\infty, 0)) ),
\quad \lambda\in \Lambda.
\label{5.75}
\end{equation}
\end{lemma}
\begin{proof}
Since $\text{\rm{spec}}(S)\subseteq \{-1,1\}$ and $A\in{\mathcal B}_\infty({\mathcal H})$, the
spectrum of $S+A$ is a discrete set with only possible accumulation
points at
$\pm 1$. Next, the normal boundary values
$|A|^{1/2}(S+A-\lambda+i0)^{-1}|A|^{1/2}=
|A|^{1/2}(S+A-\lambda)^{-1}|A|^{1/2}$ exist in norm for all
$\lambda\in\Lambda$. Moreover,
\begin{equation}
(\text{\rm{sgn}}(-A)+|A|^{1/2}(S+A-\lambda)^{-1}|A|^{1/2})^{-1}\in{\mathcal B}({\mathcal H}),
\quad \lambda\in\Lambda, \label{5.76}
\end{equation}
which can be seen as follows: suppose that \eqref{5.76} is false,
then by compactness of $A$ there exists an $f\in{\mathcal H}$ such that
\begin{equation}
(\text{\rm{sgn}}(-A)+|A|^{1/2}(S+A-\lambda)^{-1}|A|^{1/2})f=0. \label{5.76a}
\end{equation}
Multiplying \eqref{5.76a} by $\text{\rm{sgn}}(-A)$ one infers that
\begin{equation}
(I_{\mathcal H}+\text{\rm{sgn}}(-A)|A|^{1/2}(S+A-\lambda)^{-1}|A|^{1/2})f=0.
\label{5.76b}
\end{equation}
Since
$A\in{\mathcal B}_\infty({\mathcal H})$ and
$\text{\rm{spec}}(CD)\backslash\{0\}=\text{\rm{spec}}(DC)\backslash\{0\}$ for any
$C,D\in{\mathcal B}({\mathcal H})$, one concludes that there is a $g\in{\mathcal H}$ such that
$(I_{\mathcal H} -(S+A-\lambda)^{-1}A)g=0$. Thus, $(S-\lambda)g=0$ and hence
$\lambda\in\{-1,1\}$, which
contradicts the fact that $\lambda\in\Lambda$. This proves \eqref{5.76}.
By Theorem~\ref{t5.8a},
\begin{equation}
\widehat\Xi(\lambda,S+A,S)=
\Xi(\text{\rm{sgn}}(-A)+|A|^{1/2}(S+A-\lambda)^{-1}|A|^{1/2})
\quad \text{ for a.e. } \lambda\in\Lambda. \label{5.78}
\end{equation}
Moreover, the pair
\begin{equation}
(\text{\rm{sgn}}(-A)+|A|^{1/2}(S+A-\lambda)^{-1}|A|^{1/2}, \text{\rm{sgn}}(-A)) \label{5.77}
\end{equation}
is a Fedholm pair. Hence the generalized spectral shift function is
well-defined and given by
\begin{align}
&\widehat\xi(\lambda,
S+A,S)=\text{\rm{trindex}}(\Xi(\text{\rm{sgn}}(-A)+|A|^{1/2}(S+A-
\lambda)^{-1}|A|^{1/2}),\Xi(-A)) \nonumber \\
&=\text{\rm{index}}(E_{\text{\rm{sgn}}
(-A)+|A|^{1/2}(S+A-\lambda)^{-1}|A|^{1/2}}((-\infty, 0)),
E_{\text{\rm{sgn}}(-A)}((-\infty, 0))) \label{5.79} \\
& \hspace*{8.5cm}\text{ for a.e. } \lambda\in\Lambda. \nonumber
\end{align}
Since the right-hand side of \eqref{5.79} is continuous on $\Lambda$
by Theorem~\ref{ttr.7}, $\widehat\xi (\lambda,S+A,S)$ has a continuous
representative on $\Lambda$.
Next, assume $A\in {\mathcal B}_1({\mathcal H})$. Then Krein's spectral shift function
$\xi(\lambda,S+A,S)$ associated with the pair $(S+A,S)$ coincides with
the right-hand side of \eqref{5.74} for a.e.~$\lambda\in{\mathbb{R}}$ (see, e.g.,
\cite{BP98}), proving \eqref{5.74} for $A\in {\mathcal B}_1({\mathcal H})$ applying
\eqref{5.73a}.
The general case of compact operators $A\in{\mathcal B}_\infty({\mathcal H})$ can be
handled using an appropriate approximation argument. Denoting by
$\{\lambda_n\}_{n\in {\mathbb{Z}}}$ the eigenvalues of $A$ and by
$\{P_n\}_{n\in {\mathbb{Z}}}$
the corresponding spectral projections associated with
$\lambda_n$, and introducing the family of the self-adjoint operators
\begin{equation}
A_\rho=\sum_{n\in {\mathbb{Z}}}\rho^{-|n|}\lambda_nP_n, \quad \rho\in (0,1),
\label{5.80}
\end{equation}
one concludes that $A_\rho\in {\mathcal B}_1({\mathcal H})$, $\rho\in (0,1)$, and
\begin{equation}
\nlim_{\rho\uparrow 1} \|A_\rho-A\|=0. \label{5.81}
\end{equation}
Given $\lambda\in \Lambda$, there exists a $\rho_0\in (0,1)$, such that
for all $\rho\in (\rho_0, 1)$ the point
$\lambda\in\Lambda_\rho$, $\Lambda_\rho={\mathbb{R}}\backslash\{\text{\rm{spec}}
(S+A_\rho)\cup \{-1,1\}\}$, and therefore, by \eqref{5.74} (for
$A\in{\mathcal B}_1({\mathcal H})$), for such $\rho$ we have the representation
\begin{equation}
\xi (\lambda, S+A_\rho,S)
=\text{\rm{index}}(E_{S+A_\rho}((-\infty,\lambda )), E_{S}((-\infty, \lambda)) ),
\quad \rho\in (\rho_0,1). \label{5.82}
\end{equation}
Here $\xi (\lambda, S+A_\rho,S)$ denotes the continuous representative
of (the piecewise constant) Krein's spectral shift function on
$\Lambda_\rho$.
Applying Theorem~\ref{ttr.7} once again,
one can pass to the limit $\rho \uparrow 1$ to obtain
\begin{equation}
\lim_{\rho\uparrow 1}\xi (\lambda, S+A_\rho,S)=
\text{\rm{index}}(E_{S+A}((-\infty,\lambda )), E_{S}((-\infty, \lambda)) ,
\quad \lambda\in \Lambda. \label{5.83}
\end{equation}
By \eqref{5.79} we also have
\begin{align}
&\xi(\lambda, S+A_\rho,S)
\nonumber \\
&=\text{\rm{index}}(E_{\text{\rm{sgn}} (-A_\rho)+|A_\rho|^{1/2}(S+A_\rho-
\lambda)^{-1}|A_\rho|^{1/2}}((-\infty, 0)),
E_{\text{\rm{sgn}}(-A_\rho)}((-\infty, 0)) )\label{5.84}.
\end{align}
Taking into account that
$ \text{\rm{sgn}}(-A_\rho)=\text{\rm{sgn}}(-A)$, $\rho\in (0, 1)$ \eqref{5.84}
implies
\begin{align}
&
\lim_{\rho\uparrow 1} \xi(\lambda, S+A_\rho,S)
\nonumber \\
&=
\text{\rm{index}}(E_{\text{\rm{sgn}} (-A)+|A|^{1/2}(S+A-
\lambda)^{-1}|A|^{1/2}}((-\infty, 0)),
E_{\text{\rm{sgn}}(-A)}((-\infty, 0)) ),\quad \lambda\in \Lambda \label{5.85}
\end{align}
by Theorem~\ref{ttr.7}. The right-hand side of \eqref{5.85}
coincides with the continuous representative of the generalized spectral
shift function $\widehat\xi(\lambda, S+A,S)$, $\lambda\in\Lambda$, which
together with \eqref{5.83} proves \eqref{5.74}. Finally, \eqref{5.75}
is a consequence of \eqref{5.74} and the left continuity of
$\text{\rm{index}}(E_{S+A}((-\infty,\lambda )), E_{S}((-\infty, \lambda)) )$ on
${\mathbb{R}}\setminus \{-1,1\}$.
\end{proof}
Combining Theorem~\ref{main} and Lemma~\ref{l5.10}, one can
finally reformulate Theorem~\ref{main} as follows
using the concept of the generalized spectral shift function.
\begin{theorem}\label{BSCHW}
Under the assumptions of Theorem~\ref{main},
the spectral shift function $\widehat\xi(\lambda, H_0,H)$ associated
with the pair $(H_0,H)$ admits the representation
\begin{equation}\label{bsc}
\widehat\xi(\lambda, H_0,H)= \frac{1}{\pi}\int_{\mathbb{R}} dt\,
\frac{\widehat\xi(0_-,J+A(\lambda)+tB(\lambda),J)}{1+t^2}, \text{ for
a.e. } \lambda \in {\mathbb{R}},
\end{equation}
where $\widehat\xi(\cdot,J+A(\lambda)+tB(\lambda),J)$ is the continuous
representative of the generalized spectral shift function
associated with the pair
$ (J+A(\lambda)+tB(\lambda),J)$ for a.e. $\lambda\in{\mathbb{R}}$, $ t\in {\mathbb{R}}$.
\end{theorem}
\begin{proof}
The assertion is a direct consequence of Theorem~\ref{main} and
Lemma~\ref{l5.10}.
\end{proof}
\begin{remark}\label{principle}
Suppose $H, H_0, V$ are self-adjoint in ${\mathcal H}$, with
$V\in{\mathcal B}_1({\mathcal H})$ and $H=H_0+V$. Suppose $H_0$ has a spectral
gap and $\lambda\in \Lambda$, with $\Lambda$ a joint spectral gap of
$H_0$ and $H$. Then
\eqref{bsc} turns into
\begin{equation}\label{bsn}
\xi(\lambda, H_0,H)=
\xi(0,J+|V|^{1/2}(H_0-\lambda)^{-1}|V|^{1/2},J), \quad
\lambda\in\Lambda,
\end{equation}
where $\xi(\lambda,H_0,H)$
($\xi(0,J+|V|^{1/2}(H_0-\lambda)^{-1}|V|^{1/2},J)$) denotes the
continuous representative of Krein's spectral shift function associated
with the pair $(H_0,H)$ ($(J+|V|^{1/2}(H_0-\lambda)^{-1}|V|^{1/2},J)$)
on $\Lambda$. In particular, if $H_0$ is bounded from below and
$\lambda<\inf(H_0)$, and the perturbation $V$ is non-positive (i.e.,
$V\le0)$, the equality \eqref{bsn} has the following meaning: the number
of eigenvalues of the operator $H=H_0+V$, located to the left of the
point $\lambda $, $\lambda<\inf(H_0)$, coincides with the number
of eigenvalues of $|V|^{1/2}(H_0-\lambda)^{-1}|V|^{1/2}$ which are greater
than $1$. Therefore, in this special case where
$\lambda<\inf\text{\rm{spec}}(H_0)$, \eqref{bsn} represents the classical
{\it Birman--Schwinger principle} (a term coined by Simon, see, e.g.,
\cite{Si79}) as originally introduced by Birman
\cite{Bi66} (see also \cite[Ch.~7]{Sc61}). Thus, \eqref{bsn} should be
interpreted as the {\it Birman-Schwinger principle in a gap} and hence
\eqref{bsc} as the {\it generalized Birman-Schwinger principle.} We
emphasize that Theorem~\ref{BSCHW} introduces a new twist in connection
with the (generalized) Birman-Schwinger principle: the role of
eigenvalue counting
functions in the traditional formulation of the
Birman-Schwinger principle (see \cite{BS91}
for a modern formulation of the principle) is now replaced by the more
general concept of the (generalized) spectral shift function
$\widehat\xi(\lambda, H_0,H)$ and an appropriate
average over $\widehat\xi(0_-,J+A(\lambda)+tB(\lambda),J)$.
\end{remark}
|
\section{Introduction}
The next generation of mobile systems will most likely be B-ISDN
compatible. This paper considers buffer fill distribution in a
mobile ATM environment. The ATM protocol standard specifies fixed
size 53 byte cells consisting of 48 bytes of payload and a 5 byte header.
Because the ATM cell sizes are relatively small, there will
be a moderate amount of cut-through. This means that there is an increase
in performance because ATM cells will
be available for use immediately after being de-encapsulated from the
wireless media carrying the ATM cells. Also, mobile systems will be able
to take advantage
of the standardized QoS parameters as well
as having an end-to-end standards based protocol with fixed networks.
This will lead to standards based integration with fixed networks and
integration of voice, data, and video.
M/D/1 analysis of fixed size ATM cells provides optimistic results
because M/D/1 assumes that sources are Poisson.
A technique which does not make that assumption provides a more
accurate analysis and is extended in this paper to a mobile
environment.
It appears that there has been little work done concerning the effects
of mobility on ATM. This paper will attempt to build a foundation for
analyzing mobile ATM networks by extending previous work for fixed
ATM environments. This analysis
would be useful for determining the base station queue fill distribution
and probability of cell loss in a mobile environment. It would also be
useful for simplifying mobile CBR cell simulations.
\section{Mobile Systems Analysis}
The equilibrium buffer fill distribution can be described
by a set of differential equations assuming sources alternate
asynchronously between exponentially distributed periods in ``on'' and
``off'' states. Figure \ref{cbr} shows such a source. Note that the
``on'' to ``off'' probability is normalized to one and that the
transitions represented are intensities.
\begin{figure}[htbp]
\centerline{\psfig{file=figures/cbr.eps,width=3.25in}}
\caption{Constant Bit Rate ``on''-``off'' ATM Source.}
\label{cbr}
\end{figure}
In addition, the probabilities
that mobile sources have links to a given buffer are included. The sources
represent
mobile user nodes which are transmitting and receiving ATM traffic, and the
buffer represents a switch as shown in
Figure \ref{mobcbr1}. The details of such a mobile ATM system
implementation are described in the Rapidly Deployable Radio Networks
(RDRN) Network Architecture \cite{BushRDRN} and \cite{BushICC96}.
\begin{figure*}[htbp]
\centerline{\psfig{file=figures/mobcbr1.ps,width=5.25in}}
\caption{Mobile ATM Sources before Handoff.}
\label{mobcbr1}
\end{figure*}
Figure \ref{mobcbr2} illustrates the handoff of a remote node from
one switch to another. Note that the ``on''-``off'' CBR Source model is
similar to the ``connected''-``disconnected'' status of the remote
nodes as they handoff from one base station to another. This observation
is used in developing the analysis for mobile ATM systems.
\begin{figure*}[htbp]
\centerline{\psfig{file=figures/mobcbr2.ps,width=5.25in}}
\caption{Mobile ATM Sources after Handoff.}
\label{mobcbr2}
\end{figure*}
In a fixed network, the queue fill distribution is determined for multiple
constant rate on-off sources. However, it assumes that the number of
sources remains constant over a sufficient period of time for the
equilibrium probabilities to be valid. There are at least two ways
of extending the analysis to a mobile cellular environment.
Consider the ATM cell queue fill distribution at the base station.
The simplest, but least accurate extension is to determine the average
number of channels used per cell area as $t \rightarrow \infty$. However,
there is nothing to stop mobile units from concentrating in a small
number of cell areas at some time.
There is a hard limit on the number of sources a base station will
accept because each base station has a limited number of ports.
Once this number is reached, further handoffs into
such a cell will cause their connections to terminate. Thus, determining
how many codes to assign to a base station is a critical design choice,
\cite{Lee}.
Note that code assignment can be dynamic, but this will not be
considered here in order to help simplify the analysis. Also, cell areas
can be designed to overlap \cite{Lee}. Although this can increase the
probability of interference, it has
a beneficial effect on handoff. When a mobile unit determines that a
handoff is likely to occur and the cell it will enter has no channels
available, the mobile unit can continue to use the cell within which
it currently resides, and queue the handoff to the next cell. If a channel
becomes available in the destination cell before the mobile leaves its
current cell, the handoff can take place successfully. It would be
interesting to see the effect of queuing the handoffs. Again, in order
to keep the computation simple for this paper, this will not be
considered.
This paper makes the simplifying assumption that each mobile unit is a
single ATM source multiplexed at the base station. In general, each mobile
unit would be a set of sources; however, this could again be a future
extension of this paper.
\subsection{Mobile Node Analysis}
This paper makes use of the analysis and notation in \cite{Hong}.
There are two probability distributions that need to be developed:
the channel holding time and the equilibrium probability of the
number of channels used per base switch. The channel holding time is
the probability that a particular base station's channel will be in
use at a given time, or equivalently, that a particular source still exists.
The equilibrium probability of channels in use
for a given base switch is useful in this analysis as shown later.
The first of many simplifying assumptions is that there is a known average
number of new calls per second per unit area. Let this be $\lambda_{R}$
where $R$ is the radius of the particular cell area. Handoffs are attempted
at an average rate per cell, $\lambda_{Rh}$. The ratio of handoff attempts
to new call attempts will be $\gamma_{o} \stackrel{\rm \Delta}{=} \frac{\lambda_{Rh}}{\lambda_{R}}$.
\newcommand{T_{h}}{T_{h}}
\newcommand{T_{M}}{T_{M}}
\newcommand{f_{T_M}(t) = \mu_M e^{-\mu_M t}}{f_{T_M}(t) = \mu_M e^{-\mu_M t}}
\newcommand{T_{Hn}}{T_{Hn}}
\newcommand{T_{Hh}}{T_{Hh}}
\newcommand{T_{n}}{T_{n}}
Let $P_{B}$ be the average number of new call attempts which are blocked. Then
new calls are accepted at an average rate of $\lambda_{Rc} = \lambda_{R} (1-P_{B})$.
Similarly, let $P_{fh}$ be the average number of handoff attempts which are
blocked. Then handoff calls are accepted at a rate $\lambda_{Rhc} = \lambda_{Rh} (1-P_{fh})$.
The ratio of the average accepted handoffs to the average number of new
calls accepted is $\gamma_{c} \stackrel{\rm \Delta}{=} \frac{\lambda_{Rhc}}{\lambda_{Rc}}$. The
channel holding time, $T_{h}$ , is a random variable defined as the time
beginning when a channel is accessed, either via a new call or handoff,
until the channel is released, via handoff or call completion. In order
to define this, another random variable, $T_{M} $ is defined. $T_{M}$ is
the time duration of a call, regardless of handoff or blocking.
It is simplified as an exponential with average value, $\frac{1}{\mu_M}$.
Thus the pdf is
\begin{equation}
\label{ftm}
f_{T_M}(t) = \mu_M e^{-\mu_M t}
\end{equation}
The strategy for determining the channel holding
time distribution is to consider the time remaining for a call which
has not been handed off yet, $T_{Hn} $, and the time remaining after
a handoff, $T_{Hh}$. Since the call duration, $T_{M}$ is memoryless,
the time remaining for a call after handoff has the same distribution
as the original call duration. Let $T_{n}$ be the time the mobile
unit remains in the original cell area, and $T_{h}$ be the time the
mobile resides in the cell area after handoff. $T_{Hn}$ is the minimum
of the call duration, $T_{M}$, or the dwell time in the originating
cell area, $T_{n}$. A similar reasoning applies to the cell area into
which a mobile unit has moved after a handoff; $T_{Hh}$ is the minimum
of the call duration, $T_{M}$, or the dwell time in the cell area after
handoff, $T_{h}$.
\newcommand{F_{T_{Hn}}}{F_{T_{Hn}}}
\newcommand{F_{T_{M}}}{F_{T_{M}}}
\newcommand{F_{T_{n}}}{F_{T_{n}}}
\newcommand{F_{T_{h}}}{F_{T_{h}}}
\newcommand{F_{T_{Hh}}}{F_{T_{Hh}}}
Thus,
\begin{eqnarray}
\label{FTH}
F_{T_{Hn}}(t) & = & F_{T_{M}}(t) + F_{T_{n}}(t)(1 - F_{T_{M}}(t)) \nonumber \\
F_{T_{Hh}}(t) & = & F_{T_{M}}(t) + F_{T_{h}}(t)(1 - F_{T_{M}}(t))
\end{eqnarray}
where $(1-F_{T_{M}}(t))$ is the probability that a call does not complete
within the current cell area.
\newcommand{\frac{\lrc}{\lrc+\lrhc}\Fthn (t)+\frac{\lrhc}{\lrc + \lrhc}\Fthh (t)}{\frac{\lambda_{Rc}}{\lambda_{Rc}+\lambda_{Rhc}}F_{T_{Hn}} (t)+\frac{\lambda_{Rhc}}{\lambda_{Rc} + \lambda_{Rhc}}F_{T_{Hh}} (t)}
The distribution of channel holding time in a particular cell area is
a weighted function of the equations shown in \ref{FTH} above,
\begin{equation}
F_{T_H}(t) = \frac{\lrc}{\lrc+\lrhc}\Fthn (t)+\frac{\lrhc}{\lrc + \lrhc}\Fthh (t)
\end{equation}
\newcommand{1-\eum+\frac{\eum}{1+\gc}(\Ftn(t)+\gc\Fth(t))}{1-e^{-\mu_{M}t}+\frac{e^{-\mu_{M}t}}{1+\gamma_{c}}(F_{T_{n}}(t)+\gamma_{c}F_{T_{h}}(t))}
Substituting the values from Equation \ref{ftm},
\begin{equation}
\label{FTHone}
F_{T_H}(t) = 1-\eum+\frac{\eum}{1+\gc}(\Ftn(t)+\gc\Fth(t))
\end{equation}
and differentiating to get the pdf,
\begin{eqnarray}
\label{fTH}
f_{T_H}(t) & = & \mu_{M} e^{-\mu_{M}t} \nonumber \\
& & + \frac{e^{-\mu_{M}t}}{1 + \gamma_{c}} \left[ f_{T_{n}} (t)+ \gamma_{c} f_{T_{h}} (t) \right]
\nonumber \\
& & - \frac{e^{-\mu_{M}t} }{1 + \gamma_{c}} \left[ F_{T_{n}} (t) + \gamma_{c} F_{T_{h}} (t) \right]
\end{eqnarray}
To determine the equilibrium probability of the number of mobile
hosts using a base station, approximate the channel holding time as
simply an exponential distribution. The birth-death
Markov chain can be used to find the equilibrium probability of
the number of sources in each cell area. The {\em up rates} are
$\lambda_{R} + \lambda_{Rh}$ and the {\em down rates} are multiples
of the mean channel holding time.
Putting the Markov chain in closed form,
\begin{equation}
\label{EPj}
P_j = \frac{\left( \lr+\lrh \right)^{j}}{j!\mu_{H}^{j}}P_{0}
\end{equation}
where,
\begin{equation}
P_0 = \frac{1}{\sum_{k=0}^{C} \frac{\left( \lr+\lrh \right)^{k}}{k!\mu_{H}^{k}}}
\end{equation}
Note that $C$ is the total number of channels for a base station and
handoffs will fail with probability $P_C$, i.e. all channels
in that cell area are currently in use.
\subsection{Mobile CBR Source Analysis}
Assume that the number of mobile hosts in a cell area is independent of
whether its CBR source is on or off. We can now modify the probability
that $i$ sources are on and the queue fill is less than $x$ by incorporating
the probability that there are at least $i$ sources,
shown in Equation \ref{mobmod}.
\begin{equation}
\label{mobmod}
P_{i_{mobile}}(t,x) \stackrel{\rm \Delta}{=} P_{j \geq i}\ and\ P_i(t,x)
\end{equation}
$P_i(t,x)$ is the probability that at time $t$, $i$ sources are on, and
the number of items in the buffer does not exceed $x$. $P_{j \geq i}$ is the
probability that there are at least $i$ sources sending data to the buffer.
$P_{j \geq i}$ can be found from Equation \ref{EPj} as follows,
\begin{equation}
\Pngi{i} = \sum_{j=i}^C P_j
\end{equation}
The buffer fill distribution as defined in \cite{Anick1982} is
\begin{eqnarray}
\label{Pbd}
\Pbd
\end{eqnarray}
Now that the channel equilibrium probabilities have been determined,
we can account for the fact that the sources
are mobile. Since the channel equilibrium probabilities have no
dependence on time, the method of solution in \cite{Anick1982} can
be used with minor modifications,
\begin{eqnarray}
\label{Pbdmobile}
\Pbdmobile
\end{eqnarray}
From \cite{Anick1982}, $F_i(x)$ is the equilibrium probability that $i$
sources are on, and the buffer content does not exceed $x$. Thus
$F_i(\infty)$ is the probability that $i$ out of $N$ sources are
simultaneously on. In the mobile environment, this is now,
\begin{equation}
F_i(\infty) = \sum_{j=1}^{C} P_{j} {C \choose j} \left(\frac{\lambda}{1+\lambda}\right)^j\left(\frac{1}{1+\lambda}\right)^{C-j}
\end{equation}
The mobile extension from Equation \ref{mobmod} carries through \cite{Anick1982}
for example, equation (13) in \cite{Anick1982} is now,
\begin{equation}
\phi_{i_{mobile}} \stackrel{\rm \Delta}{=} \phi_{i}P_{j=i}
\end{equation}
and
\begin{equation}
\Phi(1) = \sum_{i=0}^{C}\phi_{i}P_{j=i}
\end{equation}
$\phi_{i}$ is the right eigenvector of
\begin{equation}
zD\phi = M\phi
\end{equation}
where D and M are matrices used to represent the differential equation
in Equation \ref{Pbd}.
$\Phi(x)$ is the generating function of $\phi$. These values are useful
in \cite{Anick1982} for solving the equilibrium buffer fill differential
equation. The remainder of the solution is straight forward from
\cite{Anick1982}. Thus it has been shown how an analysis of constant bit
rate on-off sources which model fixed length ATM packet sources, is
extended to a mobile environment.
Note that the analysis uses a technique which is more accurate than
$M/D/1$ for the fixed size ATM cells, yet uses a memoryless analysis
for the channel holding time distribution. This is a reasonable approach
since the variable length channel hold times can be accurately modeled
by a memoryless analysis, while the $M/D/1$ analysis yields optimistic
results which can be replaced by the more accurate method in \cite{Anick1982}
as this section has described.
\label{extension}
\subsection{Example}
The following is a simple example of the analysis using the same
parameters as the simulation in the next section. The parameters
required are:
\begin{itemize}
\item $\lambda_{R} = 0.06$ calls/sec/square mile
\item $T_{M} = 40$ secs
\item $V_{max} = 0.03$ miles/sec
\item $C = 3$ channels per base station
\end{itemize}
From the equations in the previous section, we can develop an analytical
solution for the remaining parameters. Using basic probability the integral
of Equation \ref{fTH} should be one. Also, $\mu_{H}$ is an exponential
approximation of Equation \ref{FTHone}, which can be found by taking the
integral of the difference of $F_{T_H}(t)$ from Equation \ref{FTHone} and
$\mu_{H}$ and setting the result to zero,
\begin{equation}
\int_0^\infty f_{TH}(t) dt = 1
\end{equation}
\begin{equation}
\int_0^\infty \left[ F_{TH}(t) - e^{\mu_{H} t} \right] dt = 0.
\end{equation}
This provides two equations and two unknowns which provide the solution
for $\lambda_{Rh}=2.16$ and $\mu_{H}=9.48$. These values can be
used to determine the $P_{j}$ which can then be used in our
extension of \cite{Anick1982} as discussed previously. In the graph of $P_{j}$
shown in Figures \ref{Pj6_graph}, it appears that there
will be a high probability of blocking, since $P_{B} = P_{C} = P_{3}$.
This is compared with an arrival rate of 0.01 calls/sec/unit
area in Figure \ref{Pj01_graph}, which has a maximum at one
channel per station, and a lower blocking probability.
\begin{figure}[htbp]
\centerline{\psfig{file=figures/mobexam6.ps,width=3.25in}}
\caption{Channel Usage Prob. Density Function for 0.6 Call/Sec.}
\label{Pj6_graph}
\end{figure}
\begin{figure}[htbp]
\centerline{\psfig{file=figures/mobexam01.ps,width=3.25in}}
\caption{Channel Usage Prob. Density Function for 0.01 Call/Sec.}
\label{Pj01_graph}
\end{figure}
Figures \ref{Pj6_graph} and \ref{Pj01_graph} are in agreement with the
simulation results in Figures \ref{changraph6} and \ref{changraph01}
as additional verification.
\subsection{Simulation and Results}
The mobile communications system model\footnote{A mobile cellular telephone
system library comes with the BONeS software. As much as possible of that
library is used as a basis
for this simulation.} is shown in figure \ref{mobile-cellular-telephone-system}.
It is an open system; mobile hosts are generated at rate with inter-arrival
time specified by {\bf Exp Pulse Mean}, initiate a call for an average
exponential duration specified by {\bf Mean Session Length}, and exit the
system when either the call is complete or the mobile moves out of all cell
areas.
\begin{figure*}[htbp]
\centerline{\psfig{file=figures/mobsys.ps,width=5.25 in}}
\caption{Top Level Mobile System.}
\label{mobile-cellular-telephone-system}
\end{figure*}
The first step is to create the base stations and their cell areas.
The total number of base stations is:
\begin{equation}
{\bf Base\ Station\ Matrix\ Size}^2
\end{equation}
and they are located in a square array. They all have the same number
of channels, {\bf Total Channels for Base Station}. Each cell area can
be approximated as a circle with radius:
\begin{equation}
\frac{{\bf Distance\ Between\ Base\ Stations}}{2}
\end{equation}
Mobile hosts are created in {\bf Create Mobile Users}. All the
mobile parameters are uniformly distributed, except the session duration
which is exponential and agrees with equation \ref{ftm} of our analysis.
New mobiles enter the system with an interarrival time of
{\bf Exp Pulse Mean}. The mobile host will arrive at a location
uniformly distributed anywhere in the area covered by all cells.
Since a mobile makes one call in its lifetime, a mobile host represents
a single connection. Thus,
\begin{equation}
\lambda_{R} = \frac{1}{({\bf Exp\ Pulse\ Mean})(Total\ Cell\ Area)(B)}
\end{equation}
where $B$ is the number of base stations.
The following modules act upon the mobile hosts throughout their lifetime.
The mobile hosts are assigned the nearest base station
{\bf Assign Base Station to Mobiles Users}, and an available channel from that
base station {\bf Assign Channel to Mobile Users}. Then all
mobile hosts dwell in their cell areas for time {\bf Delta Time}
{\bf Delay Mobile Users}. The mobile host then moves to its next location
which may be uniformly chosen from {\bf Direction of Motion Options} and
may lead to a new cell or completely outside the cellular system
{\bf Move Mobile Users}.
After moving, the quality of signal is checked
and if below a given criteria\footnote{In this
case, if the distance between a mobile host and its currently assigned
base station is greater than $\frac{2}{3} \sqrt{D^2-\left({\frac{D}{2}}
\right)^2}$
where $D$ is the distance between base stations, then
the channel quality is considered unacceptable.}
the channel is released ({\bf Release Channel}) and
the mobile host is reassigned to a new base station
({\bf Assign Base Station to Mobile Users}).
Note that the
direction of travel by a mobile user is uniformly chosen from North,
South, East, or West. A more
sophisticated analysis of speed and direction is contained in \cite{Leung}.
The simulation results are shown in Figures \ref{changraph6} and
\ref{changraph01}. These results agree with the results of the analysis
shown in Figures \ref{Pj6_graph} and \ref{Pj01_graph}.
\begin{figure}[htbp]
\centerline{\psfig{file=figures/sgraph01.ps,width=3.5in}}
\caption{Channel Usage Probability for 0.6 Call/Sec.}
\label{changraph6}
\end{figure}
\begin{figure}[htbp]
\centerline{\psfig{file=figures/sgraph6.ps,width=3.5in}}
\caption{Channel Usage Probability for 0.01 Call/Sec.}
\label{changraph01}
\end{figure}
\section{Summary}
This paper attempted to extend \cite{Anick1982} to a mobile
environment. It also presented the results of a simulation
of a mobile environment in preparation for simulating the
extension. The mobile environment adds several new dimensions
to fixed communications analysis, such as cell area, speed,
direction, channel holding time, channels used at a base station,
frequency of handoffs.
This simulation concentrated on finding channel hold time, $T_H$,
and the average number of channels used, $P_j$, which are required
for the extension of \cite{Anick1982}.
This paper also suggested areas for further research such as
extending \cite{Anick1982} to mobiles which are treated as
multiple CBR sources, analyzing queued handoffs, and enhancing
the channel hold time and number of channels used by using a
PDE for the speed and direction analysis as in \cite{Leung}.
|
\section{Introduction}
\par
There is considerable current interest in trying to isolate the lightest
glueball.
Several experiments have been performed using glue-rich
production mechanisms.
One such mechanism is Double Pomeron Exchange (DPE) where the Pomeron
is thought to be a multi-gluonic object.
Consequently it has been
anticipated that production of
glueballs may be especially favoured in this process\cite{closerev}.
\par
The WA102 experiment at the CERN Omega Spectrometer
studies centrally produced exclusive final states
formed in the reaction
\noindent
\begin{equation}
pp \longrightarrow p_{f} X^{0} p_s,
\label{eq:1}
\end{equation}
where the subscripts $f$ and $s$ refer to the fastest and slowest
particles in the laboratory frame respectively and $X^0$ represents
the central system.
\section{A partial wave analysis of the $K \overline K$ system}
\par
The isolation of the reaction
\begin{equation}
pp \rightarrow p_{f} (K^+ K^-) p_{s}
\label{eq:b}
\end{equation}
has been described in detail in a previous publication\cite{re:kkpap}.
A Partial Wave Analysis (PWA) of the centrally produced \mbox{$K^+K^-$ } system has been
performed,
using the reflectivity basis\cite{chung},
in 40~MeV intervals of the \mbox{$K^+K^-$ }
mass spectrum using an event-by-event maximum likelihood
method\cite{re:kkpap}.
The $S_0^-$ and $D_0^-$-Waves
from the physical solution are shown in fig.~\ref{fi:1}.
\begin{figure}[h]
\vspace{7.0cm}
\begin{center}
\special{psfile=kktalk13.eps voffset=-210 hoffset=10
hscale=60 vscale=60 angle=0}
\end{center}
\caption{\it The $S_0^-$ and $D_0^-$-Waves
resulting from a partial wave analysis of
the $K^+K^-$ system.}
\label{fi:1}
\end{figure}
\par
The $S_0^-$-wave shows a threshold enhancement; the peaks at 1.5 GeV and
1.7~GeV are interpreted as being due to the
$f_0(1500)$ and $f_J(1710)$ with J~=~0.
A fit has been performed to the $S_0^-$ wave using
three interfering Breit-Wigners to describe the $f_0(980)$, $f_0(1500)$
and $f_J(1710)$ and a background
of the form
$a(m-m_{th})^{b}exp(-cm-dm^{2})$, where
$m$ is the
\mbox{$K^+K^-$ }
mass,
$m_{th}$ is the
\mbox{$K^+K^-$ }
threshold mass and
a, b, c, d are fit parameters.
The resulting fit is shown in fig.~\ref{fi:1} and gives
for the $f_0(980)$ M~=~985~$\pm$~10~MeV, $\Gamma$~=~65~$\pm$~20~MeV,
for the $f_0(1500)$ M~=~1497~$\pm$~10~MeV, $\Gamma$~=~104~$\pm$~25~MeV and
for the $f_0(1710)$ M~=~1730~$\pm$~15~MeV, $\Gamma$~=~100~$\pm$~25~MeV
parameters which are consistent with the PDG\cite{PDG98} values for these
resonances.
\par
The $D_0^-$-wave shows peaks in the 1.3 and 1.5~GeV regions,
presumably due to the $f_2(1270)/a_2(1320)$ and $f_2^\prime(1525)$ and
a wide structure above 2 GeV. There is no evidence for
any significant structure in the D-wave in the region of the
$f_J(1710)$. In addition, there are no statistically significant
structures in any of the other waves.
A fit has been performed to the $D_0^-$ wave above 1.2~GeV using
three incoherent relativistic spin 2 Breit-Wigners
to describe the $f_2(1270)/a_2(1320)$,
$f_2^\prime(1525)$ and the peak at 2.2 GeV and
a background of the form described above.
The resulting fit is shown in fig.~\ref{fi:1} and gives
for the $f_2(1270)/a_2(1320)$ M~=~1305~$\pm$~20~MeV,
$\Gamma$~=~132~$\pm$~25~MeV,
for the $f_2^\prime(1525)$ M~=~1515~$\pm$~15~MeV, $\Gamma$~=~70~$\pm$~25~MeV
and for the
$f_2(2150)$ M~=~2130~$\pm$~35~MeV, $\Gamma$~=~270~$\pm$~50~MeV.
\par
A study has also been made of the centrally produced \mbox{$K^0_SK^0_S$ }
channel\cite{re:kkpap}.
This channel has lower statistics than the \mbox{$K^+K^-$ } channel but has
the advantage that only even spins can contribute, which also means that
there are only two ambiguous solutions to the PWA.
The physical solution resulting from the PWA is
the same as for the \mbox{$K^+K^-$ } final state; namely
that
the $S_0^-$-wave shows a threshold enhancement and peaks at 1.5 GeV and
1.7~GeV interpreted as being due to the
$f_0(1500)$ and $f_J(1710)$ with J~=~0.
\section{A partial wave analysis of the $\pi \pi$ system}
\begin{figure}[h]
\vspace{9.0cm}
\begin{center}
\special{psfile=pipitalk8.eps voffset=-10 hoffset=20
hscale=50 vscale=40 angle=0}
\end{center}
\caption{\it a), b), c) The $S_0^-$ wave d) the $P_0^-$ wave and
e) the $D_0^-$ wave
resulting from a partial wave analysis of
the $\pi^+\pi^-$ system.}
\label{fi:2}
\end{figure}
\par
The isolation of the reaction
\begin{equation}
pp \rightarrow p_{f} (\pi^+ \pi^-) p_{s}
\label{eq:c}
\end{equation}
has been described in detail in a previous publication\cite{re:pipipap}.
The resulting centrally produced \mbox{$\pi^+\pi^-$ } system
consists of 2.87 million events.
A PWA of the centrally produced \mbox{$\pi^+\pi^-$ } system has been
performed,
using the reflectivity basis\cite{chung},
in 20~MeV intervals of the \mbox{$\pi^+\pi^-$ }
mass spectrum using an event-by-event maximum likelihood
method\cite{re:pipipap}.
The $S_0^-$, $P_0^-$ and $D_0^-$-Waves
from the physical solution are shown in fig.~\ref{fi:2}.
The $S_0^-$-wave spectrum
shows a clear threshold enhancement followed by a sharp
drop at 1~GeV. There is clear evidence for the $\rho(770)$ in the $P_0^-$ wave
and for the $f_2(1270)$ in the
$D_0^-$ wave.
\par
An interesting feature of the $D_0^-$ wave is the presence of a structure
below 1~GeV.
In order to see if this effect is due to acceptance problems or
problems due to non-central events, we have reanalysed the data
using a series of different cuts
but after acceptance
correction no cut has been found that can remove
the low mass structure.
In order to investigate any systematic effects we have also analysed the
central $\pi^0\pi^0$ data and a similar structure is also
found\cite{re:pi0pi0pap}.
This structure does indeed seem to be a real effect
which is present in other centrally produced $\pi \pi$ systems\cite{cenprod}.
\par
In order to obtain a satisfactory fit
to the $S_0^-$ wave from threshold to 2~GeV it has been found to be
necessary to use
three interfering Breit-Wigners to describe the $f_0(980)$, $f_0(1300)$
and $f_0(1500)$ and a background
of the form
$a(m-m_{th})^{b}exp(-cm-dm^{2})$, where
$m$ is the
\mbox{$\pi^+\pi^-$ }
mass,
$m_{th}$ is the
\mbox{$\pi^+\pi^-$ }
threshold mass and
a, b, c, d are fit parameters.
The fit is shown in fig.~\ref{fi:2}a) for the entire mass range
and in fig.~\ref{fi:2}b) for masses above 1 GeV.
The resulting parameters are
for the $f_0(980)$ M~=~982~$\pm$~3~MeV, $\Gamma$~=~80~$\pm$~10~MeV, for the
$f_0(1300)$ M~=~1308~$\pm$~10~MeV, $\Gamma$~=~222~$\pm$~20~MeV and for the
$f_0(1500)$ M~=~1502~$\pm$~10~MeV, $\Gamma$~=~131~$\pm$~15~MeV
which are consistent with the PDG~\cite{PDG98} values for these
resonances.
As can be seen, the fit describes the data well for masses below 1~GeV.
It was not possible to describe the data above 1~GeV without the addition
of both the $f_0(1300)$ and $f_0(1500)$ resonances.
However, even with this fit using
three Breit-Wigners it can be seen that the fit does not
describe well the 1.7 GeV region.
This could be due to a \mbox{$\pi^+\pi^-$ } decay mode of the $f_J(1710)$ with J~=~0.
Including a fourth Breit-Wigner in this mass region decreases the
$\chi^2$ from 256 to 203 and yields
for the $f_J(1710)$ M~=~1750~$\pm$~20~MeV and $\Gamma$~=~160~$\pm$~30~MeV
parameters which are consistent with the PDG~\cite{PDG98} values for the
$f_J(1710)$.
The fit is
shown in fig.~\ref{fi:2}c) for masses above 1 GeV.
\section{A Glueball-$q \overline q$ filter in central production ?}
The WA102 experiment studies mesons produced in double exchange processes.
However, even in the case of pure DPE
the exchanged particles still have to couple to a final state meson.
The coupling of the two exchanged particles can either be by gluon exchange
or quark exchange. Assuming the Pomeron
is a colour singlet gluonic system if
a gluon is exchanged then a gluonic state is produced, whereas if a
quark is exchanged then a $q \overline q $ state is produced\cite{closeak}.
In order to describe the data in terms of a physical model,
Close and Kirk\cite{closeak},
have proposed that the data be analysed
in terms of the difference in transverse momentum ($dP_T$)
between the particles exchanged from the
fast and slow vertices.
The idea being that
for small differences in transverse momentum between the two
exchanged particles
an enhancement in the production of glueballs
relative to $q \overline q$ states may occur.
\begin{figure}
\vspace{7.0cm}
\begin{center}
\special{psfile=dptplot_new.eps voffset=-25 hoffset=60
hscale=50 vscale=40 angle=0}
\end{center}
\caption{\it The ratio of the amount of resonance with
$dP_T$~$\leq$~0.2 to the amount with
$dP_T$~$\geq$~0.5~GeV.
}
\label{fracratio}
\end{figure}
\par
The contribution of each resonance as a function
of $dP_T$ has been calculated.
Figure~\ref{fracratio} shows the ratio of the number of events
for $dP_T$ $<$ 0.2 GeV to
the number of events
for $dP_T$ $>$ 0.5 GeV for each resonance considered.
It can be observed that all the undisputed $q \overline q$ states
which can be produced in DPE, namely those with positive G parity and $I=0$,
have a very small value for this ratio ($\leq 0.1$).
Some of the states with $I=1$ or G parity negative,
which can not be produced by DPE,
have a slightly higher value ($\approx 0.25$).
However, all of these states are suppressed relative to the
the glueball candidates the
$f_0(1500)$, $f_J(1710)$, and $f_2(1930)$,
together with the enigmatic $f_0(980)$,
which have
a large value for this ratio.
\begin{figure}
\vspace{7.0cm}
\begin{center}
\special{psfile=phiang.eps voffset=-40 hoffset=60
hscale=50 vscale=50 angle=0}
\end{center}
\caption{\it The azimuthal angle between the fast and
slow protons ($\phi$) for various final states.
}
\label{fi:phidep}
\end{figure}
\section{The azimuthal angle between the outgoing protons}
\par
The azimuthal angle ($\phi$) is defined as the angle between the $p_T$
vectors of the two protons.
Naively it may be expected that this angle would be flat irrespective
of the resonances produced.
Fig.~\ref{fi:phidep} shows the $\phi$ dependence for two
$J^{PC}$~=~$0^{-+}$ final states (the $\eta$ and $\eta^\prime$),
two $J^{PC}$~=~$1^{++}$ final states (the $f_1(1285)$ and $f_1(1420)$) and
two $J^{PC}$~=~$2^{++}$ final states
(the $\phi \phi$ and $K^*(892) \overline K^*(892)$ systems).
The $\phi$ dependence is clearly not flat and considerable variation
is observed between final states with different $J^{PC}$s.
\section{Summary}
\par
In conclusion, a partial wave analysis of the centrally
produced $K \overline K$ system has been performed.
The striking feature is the
observation of peaks in the $S_0^-$-wave corresponding to
the $f_0(1500)$ and $f_J(1710)$ with J~=~0.
In addition, a partial wave analysis of a
high statistics sample of centrally produced \mbox{$\pi^+\pi^-$ } events
shows that the $S_0^-$-wave is composed of
a broad enhancement at threshold, a sharp drop
at 1 GeV due to the interference between the $f_0(980)$
and the S-wave background, the $f_0(1300)$, the $f_0(1500)$ and
the $f_J(1710)$ with J~=~0.
\par
A study of centrally produced pp interactions
show that there is the possibility of a
glueball-$q \overline q$ filter mechanism ($dP_T$).
All the
undisputed $q \overline q $ states are observed to be suppressed
at small $dP_T$, but the glueball candidates
$f_0(1500)$, $f_J(1710)$, and $f_2(1930)$ ,
together with the enigmatic $f_0(980)$,
survive.
In addition, the production cross section for different
resonances depends strongly on the azimuthal angle between the
two outgoing protons.
|
\section{Introduction}
Reaction-diffusion processes are the subject of much research
\cite{fitzhugh}
\cite{nagumo}
\cite{pearson_9}
\cite{pearson_10}
\cite{gray_8}
\cite{turing_7},
a reaction-diffusion process occurs as reactants in a solution diffuse in the
liquid and react amongst themselves. A common approach to reaction-diffusion
processes is to consider the density fields of the different reactants
participating in the reactions. This approach stands in contrast to the more
naive approach of tracking the locations of the different reactants, or
computing the wave functions of the different reactants.
Whatever approach is taken the interest in a reaction-diffusion system is
usually in its spatio-temporal evolution. The density field approach is especially adept
for this purpose, since the actual location of specific reactants is, usually
of no interest. In the density fields approach the spatio-temporal evolution is modeled
through partial differential equations (PDE's).
Another approach to reaction-diffusion processes that we have suggested is the
microscopic approach. In this approach we consider the number of reactants at
discrete lattice points, where the lattice models space. The main difference
from the density field approach is that rather than using continuous
densities in a continuous space as in the density field approach, we
use discrete densities in discrete space.
The microscopic simulation approach is closer to the real simulated system when there are only trace
densities of the different reactants. This is because it is in this situation
that the discrete nature of the reactants comes into play. Consequently the PDE
approach describes the system with less accuracy than when there are many reactants.
Reaction-Diffusion processes are not restricted to describing chemical
systems. Indeed reaction-diffusion processes have even been used extensively in
population biology \cite{maynard}. We have also used a reaction-diffusion
model in a marketing context. We have seen that discretization was crucial in
the behavior of the modeled market. Thus showing that microscopic
simulation could be of use to researchers who need to model real-life systems.
\section{The Density Field Approach}
\subsection{Analytical approach}
In this section we shall describe the density field approach \cite{maynard}.
As
mentioned in the introduction, The density field approach considers the
evolution of the the density fields of the reactants participating in the
reaction-diffusion system. Say that reactants numbered 1 to n are participating in the reactions
(possibly as reactants or as products). Then we could label the density
fields by \(\rho_{i}(\vec{x},t)\). The density is defined as:
\begin{equation}
\label{eq:density_definition}
\rho_{i}(\vec{x},t)=\frac{1}{n_{0}}\lim_{V(A) \to 0}\frac{N(i,A,\vec{x},t)}{V(A)}
\end{equation}
A is a box, \(N(1,A,\vec{x},t)\) is the number of molecules of type \(i\) that are in
the box A located about \(\vec{x}\). \(V(A)\) is the volume of box \(A\). And
\(n_{0}\) is a constant that serves the same purpose as Avogadro's
constant. It should be noted that, if a smooth density function is wanted,
the limit should not be taken to zero
literally, but rather should be taken down to a scale much larger than one that
shows the discretization of the reactants, and much smaller than the scale of
the macroscopic spatial patterns.
Let us assume \(l\) possible reactions, where reaction i is of the form:
\begin{equation}
\label{eq:typical_reaction}
S_{j_{1,i}}+...+S_{j_{m_{i},i}}\longrightarrow{}S_{k_{1,i}}+...+S_{k_{n_{i},i}}.
\end{equation}
Where \(S_{r}\) denotes the \(r^{th}\) reactant. As an example of such a
reaction let us look at:
\begin{equation}
\label{eq:example_reaction}
S_{1}+S_{1}+S_{2}\longrightarrow{}S_{1}+S_{1}+S_{1}.
\end{equation}
this reaction is of the form (\ref{eq:typical_reaction}). An interpretation of this reaction is that
two reactants of species 1 can cause a reactant of species 2 to turn into a
reactant of species 1.
Let us consider a system which has two chemicals, which we shall denote, as
usual, by \(S_{1}\) and \(S_{2}\). These chemicals can diffuse with diffusion
coefficients of \(D_{1}\) and \(D_{2}\) respectively. These chemicals can also
react according to the reaction scheme (\ref{eq:example_reaction}).
The time evolution of the fields, \(\rho_{1}(\vec{x},t)\)
and \(\rho_{2}(\vec{x},t)\), is given by:
\begin{equation}
\label{eq:time_evol_ex1}
\frac{\partial{}\rho_{1}}{\partial{}t}=D_{1}\nabla^{2}\rho_{1}+k\cdot\rho_{1}^{2}\rho_{2}
\end{equation}
\begin{equation}
\label{eq:time_evol_ex2}
\frac{\partial{}\rho_{2}}{\partial{}t}=D_{2}\nabla^{2}\rho_{2}-k\cdot\rho_{1}^{2}\rho_{2}
\end{equation}
Equation \ref{eq:time_evol_ex1} has two terms on the right hand side (RHS), let us turn our
attention first to the second term. The term contains the expression
\(\rho_{1}^{2}\rho_{2}\). This is proportional to the probability that two
reactants of species 1 and one reactant of species 2 meet in a small region in
space (the volume of that region is a given). The coefficient \(k\) is the
probability that, once the reactants met in the small region in space, they
will react with one another. The Diffusion term is the
familiar term, which originates from the ``random walk'' of the reactants.
In general, when we have \(l\) reactions, each of the form (\ref{eq:typical_reaction}),
\begin{equation}
\label{eq:l_reactions}
S_{j_{1,i}}+...+S_{j_{m_{i},i}}\longrightarrow{}S_{k_{1,i}}+...+S_{k_{n_{i},i}}.
\end{equation}
If we denote by \(\mathbf{N}_{p,r}\) the number of reactants of species \(\mathbf{r}\) created or annihilated by reaction \(\mathbf{p}\) then we have the following rate equations:
\begin{eqnarray}
\label{eq:general_rate_PDE}
\frac{\partial\rho_{r}}{\partial{}t}=
\sum_{p=1}^{l}k_{p}N_{p,r}
\prod_{s=1}^{s=m_{p}}\rho_{j_{s,p}}+
D_{r}\nabla^{2}\rho_{r}
\end{eqnarray}
\subsection{Simulation by finite difference}
The former section introduced the formalism of the density field approach
which is essentially analytic, but there is no data structure on a computer
that can hold an arbitrary continuous field. So space is discretized in the computer
simulation. The other problem which arises is the need to integrate the
differential equations over time. This is done again by discretization but the
solution now is to discretize time. The most naive way to integrate using the
differential equation is by Euler integration. This method's main drawback is
the computation time that it requires to get accurate results. But
fundamentally it is no different than other more sophisticated methods such as
the runga-cutta method. We shall outline this finite difference approach using
Euler integration below.
In the finite difference approach using Euler integration one replaces the
fields \(\rho_{r}(\vec{x},t)\) with
\(\vec{x}\in{}R^{d}\) and \(t\in{}R\) by
\(\rho_{r}^{*}(\vec{x},t)\) where
\(\vec{x}\in{}\omega^{d}\) \(t\in{}\omega\) (\(\omega\)
being the natural numbers). Let us suppose that the following equations hold
for the fields \(\rho_{r}\):
\begin{equation}
\label{eq:rate_equals_f}
\frac{\partial\rho_{r}}{\partial{}t}=f(\rho_{1},...,\rho_{n})+D\nabla^{2}\rho_{r},
\end{equation}
In order to make this transition from \(\rho\) to
\(\rho^{*}\), space is conceptually divided to a discrete d-dimensional lattice of spacing
\(\Delta{}x\) and time is divided to a discrete 1-dimensional lattice (actually
a series) of spacing \(\Delta{}t\). Now the aim is to
make the following equality be a good approximation:
\begin{equation}
\label{eq:finite_differ_approx}
\rho_{r}^{*}\left(\left(n_{1},n_{2},...,n_{d}\right),m\right)
\approx{}
\rho_{r}\left(\Delta{}x\cdot\left(n_{1},n_{2},...,n_{d}\right),m\cdot{}\Delta{}t\right)
\end{equation}
The way to make this approximation good is to let \(\Delta{}x\) and
\(\Delta{}t\) be small and to let \(\rho^{*}\) follow the dynamics:
\begin{eqnarray}
\label{eq:dynamics_finitedif_f}
\rho_{r}^{*}(\vec{x},t+1)-\rho_{r}^{*}(\vec{x},t)=\Delta{}t\cdot{}f(\rho_{1}^{*},...,\rho_{n}^{*})+\frac{D_{r}\Delta{}t}{\Delta{}x^{2}}
\nonumber\\
(\rho_{r}^{*}({\vec{x}}+(1,0,...0),t)+\rho_{r}^{*}({\vec{x}}+(-1,0,...0),t)+\nonumber\\
\rho_{r}^{*}({\vec{x}}+(0,1,...0),t)+\rho_{r}^{*}({\vec{x}}+(0,-1,...0),t)+ ...\nonumber\\
\rho_{r}^{*}({\vec{x}}+(0,0,...1),t)+\rho_{r}^{*}({\vec{x}}+(0,0,...-1),t)
-2d\rho_{r}^{*}({\vec{x}},t))
\end{eqnarray}
The first term on the right hand side of the equation is the first order
approximation of the difference between \(\rho_{r}(\vec{x},t)\)
and \(\rho_{r}(\vec{x},t+\Delta{}t)\), after a time
interval of \(\Delta{}t\) has passed assuming the dynamics
(\ref{eq:rate_equals_f}).
The second term accounts for diffusion
and includes the discretization of the \(\nabla^{2}\) operator. This term includes
positive and negative terms. The positive terms are contributions to the density
at site \(\vec{x}\) from densities at neighboring
sites. Neighboring sites are those sites which have the same coordinates as
\(\vec{x}\) but for a single coordinate which must be only
one lattice point away. This contribution is due the fact that diffusion
causes chemicals to move from one location to another in space. The negative
term accounts for the chemicals leaving site \(\vec{x}\).
Now, if we replace \(f\) with the terms from the density field approach to
reaction-diffusion, we get:
\begin{eqnarray}
\label{eq:full_finite_difference}
\rho_{r}^{*}(\vec{x},t+1)-\rho_{r}^{*}(\vec{x},t)=\nonumber\\
\Delta{}t\sum_{p=1}^{l}k_{p}N_{p,r}
\prod_{s=1}^{s=m_{p}}\rho_{j_{s,p}}^{*}(\vec{x},t)
\nonumber\\
+\frac{D_{r}\Delta{}t}{\Delta{}x^{2}}
(\rho_{r}^{*}({\vec{x}}+(1,0,...0),t)+\rho_{r}^{*}({\vec{x}}+(-1,0,...0),t)+\nonumber\\
\rho_{r}^{*}({\vec{x}}+(0,1,...0),t)+\rho_{r}^{*}({\vec{x}}+(0,-1,...0),t)+ ...\nonumber\\
\rho_{r}^{*}({\vec{x}}+(0,0,...1),t)+\rho_{r}^{*}({\vec{x}}+(0,0,...-1),t)
-2d\rho_{r}^{*}({\vec{x}},t))
\end{eqnarray}
\section{Microscopic Simulation of Reaction-Processes}
\subsection{Fundamentals of the Approach}
In the previous section we have seen the prevailing approach for dealing with
reaction-diffusion processes. Another approach that can be used is the
microscopic simulation approach. This approach takes discretization one step
further in the sense that the fields are discretized themselves, but takes a
welcomed step backwards in the sense that time is not discretized. This
approach is useful because it is closer to reality. Chemicals
are discrete entities (at least in the classical approach which is a good
approximation in solutions).
Again we have a field \(\rho_{r}^{**}(\vec{x},t)\) where
\(\vec{x}\in{}\omega^{d}\) \(t\in{}\omega\),
but this time \(\rho_{r}^{**}(\vec{x},t)\in\omega\). We
have said that time is not considered discrete in the microscopic simulation
model, but nevertheless we have \(t\in{}\omega\). This is not a contradiction
it is simply an expression of the fact that reactions occur at discrete time
points. Let us denote simulation discrete time with \(t^{*}\) and real time
with \(t\). Say the reactions occur, in the real system, at times \(t_{n}\).
Then when the simulation is at time \(t^{*}=n\) it is supposed to approximate
the real system at time \(t_{n}\). Simulating the real system between the times
\(t_{n}\) is of no use since nothing happens, except for diffusion which we
should also treat as a reaction. But the problem is that diffusion, in the real system, is not a
process that occurs at time points, rather it is a continuous process. On the
other hand we have modeled space by discrete sites. Diffusion is modeled by
chemicals hopping from one site to another. This process is discrete since chemicals
hop at discrete time points. So amongst the times \(t_{n}\) there are times
at which the reaction which takes place is diffusion, that is to say hopping of chemicals
to neighboring sites. We should stress that the time interval between
\(t_{n}\) and \(t_{n+1}\) is not a constant. So the real time is not approximated by
\(t^{*}\cdot\Delta{}t\) for some \(\Delta{}t\).
The time interval between \(t_{n}\) and \(t_{n+1}\) is large
when the time interval between successive reactions, in the real system, is
large. Roughly speaking this happens when there are not many reactants, or many
inert reactants. This is also when
the microscopic simulation is at its best (in terms of the simulation's speed),
since the simulation's
single step covers a lot of time. We shall give a quantative estimate for this
time interval later.
Let us now turn to the relation of \(\rho^{**}\) to the real system that it is
supposed to approximate. We assume space of dimensions d. Let us denote by
\(N_{r}(\vec{x},l,t)\) the
number of reactants, in the real system, of species \(r\), located in a box of
length \(l\) around \(\vec{x}\) at time \(t\). The
approximation relation is given below:
\begin{equation}
\label{eq:mic_react_approx}
\rho^{**}_{r}(\vec{x},t^{*})
\approx{}N_{r}(\Delta{}x\cdot\vec{x},\Delta{}x,t_{t^{*}})
\approx{}\rho^{*}_{r}(\vec{x},t_{t^{*}})\cdot\Delta{}x^{d}n_{0}
\end{equation}
Let us imagine a grid of spacing \(\Delta{}x\) dissecting the real system, so
that space is divided into little boxes. This division is not physical, but
mental. Now each such box is simulated as a site on the simulation lattice. The
number of reactants at each lattice point should approximate the number of
reactants in a the little boxes imagined in the real system. This is the nature
of the first approximate equality in equation (\ref{eq:mic_react_approx})
. The second approximate equality in equation
(\ref{eq:mic_react_approx}) is due to
the approximation of the finite difference approach to the real system.
\subsection{The Monte-Carlo Method}
Now we shall introduce the dynamics of
\(\rho^{**}\). Since the microscopic simulation is a Monte-Carlo simulation,
\(\rho^{**}\)'s dynamics follow the following rules:
\begin{enumerate}
\item{} Choose \(\rho^{**}(\vec{x},t^{*}+1)\) from a
probability space, \(\Omega_{0}(t^{*})\).
\item{} Calculate the new probability space \(\Omega_{0}(t^{*}+1)\).
\end{enumerate}
\(\Omega_{0}(t)\) associates a probability for every possible
\(\rho^{**}(\vec{x},t+1)\), but actually we are intent
on performing one reaction at a time. So \(\Omega_{0}(t)\) will give non-zero
probabilities only for those \(\rho^{**}(\vec{x},t+1)\)
that differ from \(\rho^{**}(\vec{x},t)\) by a single
reaction (or diffusive hopping). So we can look at \(\Omega_{0}(t)\) as associating a probability for
every possible reaction. So let us construct a probability space \(\Omega(\Re,t)\)
which associates a probability for each possible reaction.
Let us speak of a reaction, \(\Re\). This reaction can potentially take place
anywhere in real space. In particular the reaction can fall within any one of the
little boxes that we have discussed in the previous sub-section.
We shall denote reaction \(\Re\) taking place at a little box corresponding to
site \(\vec{x}\) by \(\Re_{\vec{x}}\).
Our task
now is to find out, given some initial conditions, what is the probability that
the reaction that will take place next is \(\Re_{\vec{x}}\).
Let us expand a bit on the stochastic process that the real system
undergoes. The stochastic process is comprised of events (reaction and
diffusion) occurring stochastically at
discrete time points. The events we are considering are the real
reactions and the movement of reactants from one little box to an adjacent
one. Let \(dt\) be a small time interval, then for \(\Re_{\vec{x}}\) there is a chance
\(P(\Re_{\vec{x}},t^{*})dt\) that this reaction will occur in the time interval \(dt\). The
probability density, \(P(\Re_{\vec{x}},t^{*})\),
is called the reaction rate for reaction \(\Re_{\vec{x}}\), and it is exactly
what we used in order to formulate the PDE for the real system, as we shall
soon see. First let us notice that this probability density has reciprocal time
as units, which is consistent with the name 'rate'.
The connection of \(P(\Re_{\vec{x}},t^{*})dt\) to the PDE's will be useful in calculating
\(P(\Re_{\vec{x}},t^{*})\).
So let us explore this connection. Reaction \(\Re\) is of the
form (1), that is to say:
\begin{equation}
\label{eq:reaction_re}
S_{j_{1,\Re}}+...+S_{j_{m_{\Re},\Re}}\longrightarrow{}S_{k_{1,\Re}}+...+S_{k_{n_{\Re},\Re}}.
\end{equation}
A crucial assumption for the
validity of the PDE's is that there is some time scale, \(dt\), during which
\(P(\Re_{\vec{x}},t^{*})\) does not change much and still for the same time scale, \(dt\), many
reactions occur. Under these assumptions it can be shown that the number of
reaction and diffusion events that occur at the time interval \(dt\) in the box
around \(\vec{x}\) is, simply, \(P(\Re_{\vec{x}},t^{*})dt\). Let us
look at species \(r\) ( recall the notation used in equation
(\ref{eq:reaction_re}) ). Now let\footnote{\#S denotes the number of elements
in S.}
\begin{equation}
\label{eq:s_imbalance}
N_{\Re,r}=\#\{n|n\leq{}n_{\Re},k_{n,\Re}=r\}-\#\{n|n\leq{}m_{\Re},j_{n,\Re}=r\}.
\end{equation}
This \(N_{\Re,r}\) gives the number of reactants of species \(r\) created in the reaction
\(\Re\) ( a negative number indicates that the species is annihilated in the
reaction). So the number of reactants of
species \(r\) formed, at the box around \(\vec{x}\), in the time interval \(dt\) is:
\begin{equation}
\label{eq:formed_interval}
P(\Re_{\vec{x}},t^{*})\cdot{}dt\cdot{}N_{\Re,r}
\end{equation}
The number of reactants of species \(r\) at the box which
corresponds to site \(\vec{x}\) at time \(t_{n}\) is approximated by
\(\rho^{**}_{r}(\vec{x},n)\). So the rate at
which \(\rho^{**}_{r}(\vec{x},n)\) changes due to reaction \(\Re_{\vec{x}}\)
is given by:
\begin{equation}
\label{eq:rate_of_reaction}
\frac{\partial\rho^{**}_{r}}{\partial{}t}_{\Re_{\vec{x}}}=P(\Re_{\vec{x}},t^{*})\cdot{}N_{\Re,r}.
\end{equation}
Where the index \(\Re_{\vec{x}}\) denotes that the reference is to the rate of change due
to reaction \(\Re_{\vec{x}}\) alone. We have already seen the rate of change in
the finite difference case (equation (\ref{eq:full_finite_difference})). There we expressed
\(\frac{\partial\rho^{*}}{\partial{}t}\) as a sum of terms, each expressing a
rate due to different reactions. Another term was due to diffusion. Realizing
that the terms appearing in the finite difference case express the same thing
as the rate expressed at (\ref{eq:rate_of_reaction}) modulo the approximation
relation (\ref{eq:mic_react_approx}) we can find an expression for \(P(\Re_{\vec{x}},t)\), this
is given by:
\begin{equation}
\label{eq:expression_for_P}
P(\Re_{\vec{x}},t^{*})=\Delta{}x^{d}{}n_{0}\cdot{}k_{\Re}\prod_{s=1}^{s=m_{\Re}}\left(\rho^{**}_{j_{s,\Re}}\frac{1}{\Delta{}x^{d}{}n_{0}}\right)
\end{equation}
This assumes that the reaction in question is a real reaction, that is not
diffusion. For the case of diffusion of species \(l\) we have the following equation:
\begin{equation}
\label{eq:Pexpression_diffusion}
P(\Re_{\vec{x}},t^{*})=2d\cdot{}D_{l}\cdot\frac{\rho^{**}_{l}}{\Delta{}x^{d}}
\end{equation}
This equation is derived considering the last term in equation
(\ref{eq:full_finite_difference}),
\(-\frac{2dD_{r}\Delta{}t}{\Delta{}x^{2}}\rho_{r}^{*}(\vec{x},t)\), which is due to
reactants hopping from site \(\vec{x}\) to neighboring sites. And then
considering that \(P(\Re_{\vec{x}},t^{*})\), in equation
(\ref{eq:Pexpression_diffusion}), is the probability density for
hopping to neighboring sites. After using the approximation relation
(\ref{eq:mic_react_approx}) we get (\ref{eq:Pexpression_diffusion}).
Now that we have an expression for the rates \(P(\Re_{\vec{x}},t^{*})\) of the
different reactions, we still face the task of finding the probability
associated with each possible reaction (including diffusion) by
\(\Omega\). Again let us point out that the probability that \(\Omega\)
associates with reaction \(\Re_{\vec{x}}\) at time \(t^{*}\) is the probability that this
reaction will come next in the sequence of reactions. Now between the times
\(t_{t^{*}}\) and \(t_{t^{*}+1}\) nothing happens in the reaction chamber,
apart from reactants moving inside the little boxes we have imagined. In this
time interval the reactants don't cross the boundaries of the boxes. If the rate of
\(\Re_{\vec{x},1}\) is twice that of
\(\Re_{\vec{x},2}\) then we should
expect that reaction \(\Re_{\vec{x},1}\) has twice the chance to be the next reaction
that occurs than \(\Re_{\vec{x},2}\). Let us expand a bit on the nature of the
assumption we made in the previous statement. We assume that the probability
distribution associated with \(\Omega\) is dependent only on the situation of
the current configuration of the reaction chamber and is independent of the
time that passed since the last reaction-diffusion event. So it is
actually this assumption (that the stochastic process has no memory) that allows us to compute the probability of different
reaction and diffusion events associated by \(\Omega\) as a function of the rates.
The preceding argument leads to the following relation, given two
reaction-diffusion events \(\Re_{\vec{x},1}\) and \(\Re_{\vec{x},2}\):
\begin{equation}
\label{eq:one_P_all_P}
\frac{\Omega(\Re_{\vec{x},1},t^{*})}{\Omega(\Re_{\vec{x},2},t^{*})}=\frac{P(\Re_{\vec{x},1},t^{*})}{P(\Re_{\vec{x},2},t^{*})}
\end{equation}
Taking into account the normalization of probability distribution to one, we now
can compute the probability associated by \(\Omega\) to a reaction-diffusion
event \(\Re_{\vec{x}},t^{*}\).
The solution is:
\begin{equation}
\label{eq:finally_omega}
\Omega(\Re_{\vec{x},j},t^{*})=\frac{P(\Re_{\vec{x},j},t^{*})}{\sum_{\Re_{\vec{x},i}}{P(\Re_{\vec{x},i},t^{*})}}.
\end{equation}
\section{Anderson Localization in a Reaction-Diffusion System}
\subsection{The Reaction-Diffusion System}
Anderson localization is a phenomenon associated with electron-transport
behavior in disordered materials. The disorder in the material induces a phase
transition of the electron eigen-functions from an unlocalized state in which Ohm's
law is valid into a localized state in which the material behaves as an insulator.
The wave functions of an electron under the influence of a
periodic potential is periodic. A question arises, in the context of
disordered materials,
of how this periodic
wave function is influenced by disturbances to the periodicity of the
potential. One might think that the eigenfunction of a slightly perturbed
potential (that is to say perturbed from an originally periodic potential),
would be eigen-fucntions slightly perturbed from a periodic function. This
intuition proves misleading in the metal-insulator transition case. It is seen that some of
the eigen-functions exhibit a marked departure from periodic functions, even for
small disturbance of the potential. These functions are seen to be
localized. Meaning that there are small ``islands'' in which the wave function
has high modulus and there are large spaces between these islands where the wave-function has low
modulus. Indeed the wave functions that depart from periodicity have the
approximate form :
\begin{equation}
\label{eq:localized_state}
e^{-\frac{\left(x-x_{0}\right)}{\xi}}
\end{equation}
\(\xi\) is the localization length of the eigenfunction. This
approximation is good for the tails of the eigenfunction.
This Anderson localization effect is also observed in reaction diffusion
system. Let us take the example given by Shnerb and Nelson. They presented a
reaction-diffusion system described the schematics:
\begin{eqnarray}
\label{eq:anderson_schematics}
A+F\longrightarrow{}A+A+F
\nonumber\\
A+A\longrightarrow{}\emptyset
\end{eqnarray}
\(\mathbf{A}\) is the only reactant to undergo diffusion.\\
Where \(\emptyset\) signifies that the reaction has no products (alternatively
it can be interpreted that the products of the reaction are
inert). These reactions together with diffusion can be seen as the schematics
of the population biology of bacteria. The first reaction accounts for the
reproduction of the bacteria. The rate at which the bacteria reproduce is
controlled also by the concentration of F. F can represent, for example, the intensity of
light that falls on the bacteria \cite{shnerb_1}. But can also represent
any factor that controls the rate of bacteria reproduction, with the provision
that this factor is not
variable in time, and in particular non-exhaustible. The second reaction accounts for the
dying of bacteria due to overcrowding. The reaction dynamics has the species
\(\mathbf{A}\)
facilitating the production of more \(\mathbf{A}\)'s. This property
\(\mathbf{A}\) is called autocatalysis. Autocatalysis is an important concept in pattern
formation \cite{turing_7}, the origins of life
\cite{dyson}
\cite{segre}
\cite{kauffman_model}
\cite{kauffman}
\cite{oparin}
and economy \cite{sorin_5} (where autoctalysis is the underlying concept of
multiplicative dynamics).
To see the similarity of this problem to an eigenfunction problem in quantum
mechanics, such as Anderson localization, let us write down the PDE's which are
associated with the reaction dynamics above:
\begin{equation}
\label{eq:full_shnerb}
\frac{\partial{}\rho_{A}}{\partial{}t}=D_{A}\nabla^{2}\rho_{A}+\rho_{A}\cdot{}\rho_{F}-\rho_{A}^{2}
\end{equation}
which can be reformulated into:
\begin{equation}
\label{eq:full_near_operator}
\frac{\partial{}\rho_{A}}{\partial{}t}=(D_{A}\nabla^{2}+\rho_{F}-\rho_{A})(\rho_{A})
\end{equation}
Now if we drop the last term, we have:
\begin{equation}
\label{eq:linear_near_operator}
\frac{\partial{}\rho_{A}}{\partial{}t}=(D_{A}\nabla^{2}+\rho_{F})(\rho_{A})
\end{equation}
which is quite analogous to the time dependent Schr\"{o}dinger equation:
\begin{equation}
\label{eq:schrodinger}
i\frac{\partial{}\psi}{\partial{}t}=(\frac{1}{2m}\nabla^{2}+U(x,t))(\psi)
\end{equation}
Where we have \(U\) playing the part of \(\rho_{F}\), \(\frac{1}{2m}\)
playing the part of \(D_{A}\) and \(\psi\) playing the part of \(\rho_{A}\). We
have only to remember that we have dropped the quadratic term and that there is
an additional coefficient, -\(i\), in the Schr\"{o}dinger equation \footnote{\(i\) here
is \(\sqrt{-1}\)}. Consequently, the reaction-diffusion PDE, without the
quadratic term, can be seen as a Schr\"{o}dinger equation with imaginary time.
Notice that we did not write an equation for
\(\frac{\partial\rho_{F}}{\partial{t}}\) since \(\rho_{F}\) does not undergo
any dynamics and therefore does not change in time. The phenomena of Anderson localization
determines that if \(\rho_{F}\) is not constant in space, but rather is
stochastic, then we shall have these localized states that we have described
above, both for the quantum mechanics case and for the reaction-diffusion
case. In the quantum mechanics case, a constant or periodic potential,
\(U(x,t)\), entails a periodic field, \(\psi\). In the reaction-diffusion
case, a constant potential, \(\rho_{F}\), entails a constant field,
\(\rho_{A}\).
A constant, stochastic potential naturally arises when speaking of
disordered materials \cite{efros}. But in reaction-diffusion systems, such a potential is
more of a stretch. We have given the example of light intensity as a stochastic
constant potential. But in many more cases the reaction rate associated with
the reproduction of bacteria, or some chemical, is dynamic.
This leads us to deal with a dynamic potential, \(U(x,t)\). Nelson and
Shnerb have already discussed a time-dependent potential of the form
\(U(x-vt)\) (actually they have looked at the case where the media is moving,
which is equivalent to a moving potential in the opposite direction).
\subsection{Results of microscopic simulations for Anderson localization}
We simulated (using microscopic simulation) the reaction dynamics associated with
Anderson localization .
The reaction schematics (together with the reaction rates) is given
below\footnote{The numbers in parentheses at the right of the reactions denote
their corresponding rates}:
\begin{enumerate}
\item \(\mathbf{A} + \mathbf{C} \rightarrow \mathbf{A} + \mathbf{A} + \mathbf{C} (\mathbf{15})\)
\item \(\mathbf{A} + \mathbf{A} \rightarrow \emptyset (\mathbf{3})\)
\item \(\mathbf{A} \rightarrow \emptyset (\mathbf{60})\)
\end{enumerate}
The size of the reaction chamber is 300x300 lattice points. The average number of
molecules of type \(\mathbf{C}\) per site is 5. They are dispersed in the beginning of the simulation
randomly. That is to say that the positions of the 5x300x300 molecules of type
\(\mathbf{C}\)
are chosen at
random. The spatial distribution that resulted can be seen in figure (\ref{fig:Anderson_food}).
The initial distribution of \(\mathbf{A}\) is similarly dispersed but with and
average of 2 molecules of type \(\mathbf{A}\) per site. Only \(\mathbf{A}\) molecules diffuse. The simulation's diffusion rate is 3. After awhile a steady state is
reached. Off-course \(\mathbf{C}\) is not dynamic, and it remains as in figure
(\ref{fig:Anderson_food}). On the other hand \(\mathbf{A}\) is dynamic and the
steady state is depicted in figure (\ref{fig:Anderson_localization})
\begin{figure}
\frame{
\vspace {0.01in}
\epsfxsize=3.1in
\epsfbox{Lehmann1.eps}
}
\caption{
Snapshot of molecules of type \(\mathbf{C}\).
}
\label{fig:Anderson_food}
\end{figure}
\begin{figure}
\frame{
\vspace {0.01in}
\epsfxsize=3.1in
\epsfbox{Lehmann2.eps}
}
\caption{
Snapshot of molecules of type \(\mathbf{A}\).
}
\label{fig:Anderson_localization}
\end{figure}
\subsection{Anderson Localization - in the search of dynamic clustering }
Shnerb and Nelson\cite{shnerb_1} have explored the consequences of changing the potential
\(U(x)\) by a moving potential \(U(x-vt)\), what would happen if instead of
having the potential drift at a constant speed the potential would undergo
diffusion itself? This would correspond to replacing (\ref{eq:full_shnerb}) by
the two coupled equations:
\begin{eqnarray}
\label{eq:potential_diffuse}
\frac{\partial{}\rho_{A}}{\partial{}t}=D_{A}\nabla^{2}\rho_{A}+\rho_{A}\cdot{}\rho_{F}-\rho_{A}^{2}
\nonumber\\
\label{eq:moving_potential}
\frac{\partial{}\rho_{F}}{\partial{}t}=D_{F}\nabla^{2}\rho_{F}
\end{eqnarray}
We have found no treatment in the literature for this kind of moving potential for
a good reason: if \(\rho_{F}\) is treated as non-negative continuous variable
as is usually done in the context of differential equations and since the only
dynamics imposed on \(\rho_{F}\) by (\ref{eq:potential_diffuse}) is diffusion,
\(\rho_{F}\) converges to a steady spatially-uniform state \(\rho_{F}=const\),
giving rise to the trivial solutions of a system with no potential fluctuations
at all. But if the potential is treated as a discrete variable diffusion does
not necessarily induce the uniform distribution of the potential and
localization effects still have a chance to prevail. This situation leads to
clear differences between the different simulation approaches we have
described, in further sections we will show other examples in which
discretization leads to different results in different simulation methods.
Extending the analogy to population biology we could look at \(\mathbf{A}\) as representing
a population of parasites dependent on a host species \(\mathbf{F}\) for a-sexual
reproduction. The population represented by \(\mathbf{F}\) is not affected by
the parasites and performs random diffusion in space. The phenomena
we are most interested in is dynamic clustering or grazing. Dynamic clustering would under
our analogy represent parasite herds moving in space due to changes in the
spatial distribution of the species they need in order to reproduce.
Translating our situation into a reaction-diffusion system will give the same
results as (\ref{eq:anderson_schematics}) but in this case both \(\mathbf{A}\)
and \(\mathbf{F}\) undergo diffusion. We have added to
(\ref{eq:anderson_schematics}) the reaction:
\begin{equation}
\label{eq:A_dying}
A\longrightarrow{}\emptyset
\end{equation}
representing the dying of \(\mathbf{A}\) not due to overcrowding.
Since (\ref{eq:anderson_schematics})
describes no creation or elimination of \(\mathbf{F}\) the total number of
\(\mathbf{F}\)s in all the lattice sites is constant throughout the simulation and
\(<\mathbf{F}>\) - the average concentration of \(\mathbf{F}\) is
constant as well. Given a constant value for \(D_{A}\) the results of
simulations of (\ref{eq:move_F}) are controlled by \(<\mathbf{F}>\) and \(D_{F}\)
We have simulated (using microscopic simulation) the following
reaction-diffusion system with different values of \(<\mathbf{F}>\) and
\({D_{F}}\) and keeping \(D_{A}=6\):
\footnote{Numbers in parentheses at the right of reactions are the
corresponding reaction-rates.}
\begin{eqnarray}
\mathbf{A} + \mathbf{F} \rightarrow \mathbf{A} + \mathbf{A} + \mathbf{F} (\mathbf{13})
\nonumber\\
\mathbf{A} + \mathbf{A} \rightarrow \emptyset (\mathbf{0.01})
\nonumber\\
\mathbf{A} \rightarrow \emptyset (\mathbf{8})
\nonumber\\
\label{eq:move_F}
\end{eqnarray}
The size of the simulation is 128x128. The system was seeded with 3 reactants
of species \(\mathbf{F}\) per cell and 1 reactant of species \(\mathbf{A}\) per
cell (the reactants were placed at random locations).
\begin{figure}
\frame{
\vspace {0.01in}
\epsfxsize=3.1in
\epsfbox{Lehmann3.eps}
}
\caption{
Snapshot of \(\mathbf{A}\)'s concentration, simulation parameters are
\(D_{A}=6\),\(D_{F}=4\),\(<\mathbf{F}>\)=0.15) .
}
\label{fig:mesh159a}
\end{figure}
\begin{figure}
\frame{
\vspace {0.01in}
\epsfxsize=3.1in
\epsfbox{Lehmann4.eps}
}
\caption{
Snapshot of the same simulation as in (\ref{fig:mesh159a}) at a later time,
the clusters have moved. The result of this simulation falls into the \(\beta\)
category in our notation.
}
\label{fig:mesh159b}
\end{figure}
We have divided the results of the simulation into 3 categories:
\begin{itemize}
\item\(\alpha\)- The simulation results in \(\mathbf{A}\) filling the whole
simulation space.
\item\(\beta\)- The simulation results in \(\mathbf{A}\) forming dynamic clusters.
\item\(\gamma \)-The simulation results in the total extinction of
\(\mathbf{A}\) from the simulation space.
\end{itemize}
\begin{figure}
\frame{
\vspace {0.01in}
\epsfxsize=3.1in
\epsfbox{Lehmann5.eps}
}
\caption{
\(\mathbf{A}'s\) fill up the simulation space, this corresponds to a \(\gamma\)
situation in
our notation.
}
\label{fig:fillup}
\end{figure}
\begin{figure}
\frame{
\vspace {0.01in}
\epsfxsize=3.1in
\epsfbox{Lehmann6.eps}
}
\caption{
Results of simulations of the system(\ref{eq:move_F}) for different values of
\(<\mathbf{F}>\) (x-axis) and \(D_{F}\)(y-axis). Notice that dynamic
clustering occurs also in the realistic range of: \(0.5D_{A}<D_{F}<1.5D_{A}\).
Circles denote simulation resulting in \(\alpha\) situations, Asterisks denote
simulations resulting in \(\beta\) situations
and squares denote simulations resulting in \(\gamma\) situations.
}
\label{fig:phase}
\end{figure}
\begin{figure}
\frame{
\vspace {0.01in}
\epsfxsize=3.1in
\epsfbox{Lehmann7.eps}
}
\caption{
Results of simulations of the system(\ref{eq:move_F}) for \(D_{F}=300\)
(y-axis) as a function of \(<F>\)(x-axis),
circles denote simulations ending in an \(\alpha\) situation and squares denote
simulations ending in a \(\gamma\)
situation. The change in behavior is abrupt and occurs in the 0.55-0.61 region.
Simulations in that parameter region resulted in \(\beta\) situations.
}
\label{fig:phase300}
\end{figure}
In fig(\ref{fig:phase300}) we can see an abrupt phase-transition between
\(\alpha\) states to \(\gamma\) states for high \(D_{F}\) values around \(<F>=0.61\).
We shall give a theoretical explanation to this phase change.
For high \(D_{F}\) values \(\mathbf{F}\) reactants take shorter times to pass
between different parts of the simulation space. Therefore we can assume that \(\mathbf{A}\) reactants are
affected by all \(\mathbf{F}\) reactants in the simulation space and the
mean-field approximation:
\begin{equation}
\label{eq:meanapprox}
F=<F>
\end{equation}
is valid.
Assuming (\ref{eq:meanapprox}) we are now looking for solutions to:
\begin{equation}
\label{eq:constantF}
\frac{\partial{}\rho_{A}}{\partial{}t}=D_{A}\nabla^{2}\rho_{A}+k_{1}\rho_{A}\cdot{}<F>-k_{2}\rho_{A}^{2}-k_{3}\rho_{A}
\end{equation}
Keeping in mind that \(\mathbf{A}\) can take only integral values and dropping
the diffusion term from (\ref{eq:constantF}) we find that:
\begin{eqnarray}
\label{eq:solution_node}
<F>=\frac{k_{2}+k_{3}}{k_{1}}
\nonumber\\
\mathbf{A}=1
\end{eqnarray}
is an unstable node fixed point of (\ref{eq:constantF}). This leads to predict
the death of \(\mathbf{A}\)s for \(<F>\) smaller than \(\frac{k_{2}+k_{3}}{k_{1}} \) and that
\(\mathbf{A}\)s fill up the space for \(<F>\) larger than \(\frac{k_{2}+k_{3}}{k_{1}} \).
In the case of the simulated set of reactions (\ref{eq:move_F}) we have:
\begin{equation}
\frac{k_{2}+k_{3}}{k_{1}}\approx0.61
\end{equation}
The preceding argument explains the sharp transition between \(\alpha\) and
\(\gamma\) states for the high \(D_{F}=300\) value as can be seen in fig(\ref{fig:phase300}).
One might be tempted to think that this mean-field argument will be sufficient
in explaining the emergence of grazing at low \(D_{F}\) values: for low
diffusion coefficients persistent spatial discrepancies in \(\mathbf{F}\)'s concentration
prevail throughout the simulation space, some areas would be affected from a
local \(\mathbf{F}\) concentration larger than 0.61 and would sustain a
population of \(\mathbf{A}\)s and some areas with a lower local \(\mathbf{F}\)
concentration would be empty of \(\mathbf{A}\)s. These spatial discrepancies
change with time causing our clusters to move across the simulation space.
If this was the sole mechanism leading to the grazing behavior one would expect
that the \(\beta\) simulation results would appear for low \(D_{F}\) values
equally distributed on both sides of the \(<F>\)=0.61 asymptote. Clearly,
fig(\ref{fig:phase}) shows us that this is not the case, the \(\beta\)
situations are concentrated around lower and lower \(<F>\) values as \(D_{F}\) decreases.
Therefore we should search for another mechanism in order to explain the
behavior seen in fig(\ref{fig:phase}). For lower \(D_{F}\) values, \(D_{A}\) is not
negligible and clusters are not only supported by an influx of \(\mathbf{F}\)
reactants, but are also supported by their own ability to move and ``find''
areas of high \(\mathbf{F}\) concentration in their surroundings. This
mechanism of ``searching'' for \(\mathbf{F}\) reactants should be part of an
explanation for the behavior seen in fig(\ref{fig:phase}) at low \(D_{F}\) values.
\section{Microscopic simulation of the Gray-Scott model}
\subsection{The Gray-Scott model - a pattern formation mechanism}
The Gray-Scott model\cite{gray_8} was first designed as a model of glycolysis
and as in the simplest form of Turing\cite{turing_7} pattern formation it involves two
reactants one of them enhancing the auto-catalysis of the other, but the
geometrical patterns resulting from this model are different from the ones
observed in the Turing pattern case and unlike the Turing pattern case pattern formation occurs when diffusion
coefficients are equal as well. The model is given by:
\begin{eqnarray}
\label{eq:gray-scott}
\frac{\partial\rho_{C}}{\partial{}t}=
\nabla^{2}\rho_{C}-\rho_{C}\rho_{A}^{2}+F(1-\rho_{C})\quad ,
\nonumber\\
\frac{\partial\rho_{A}}{\partial{}t}=
D_{A}\nabla^{2}\rho_{A}+\rho_{C}\rho_{A}^{2}-(F+k)\rho_{A}
\end{eqnarray}
where \(\rho_{A}(\vec{x},t)\) and \(\rho_{C}(\vec{x},t)\) are the concentration
fields for two chemical reactants \(\mathbf{A}\) and \(\mathbf{C}\)
respectively. The terms \(F(1-\rho_{C})\) and \(-(F+k)\rho_{A}\) in (\ref{eq:gray-scott}) describe the
system as being in contact with an external reservoir
kept at \(\rho_{C}=1\) and \(\rho_{A}=0\). Indeed (\ref{eq:gray-scott}) accepts
the solution:
\begin{eqnarray}
\label{eq:solution}
\rho_{C}=1
\nonumber\\
\rho_{A}=0
\end{eqnarray}
Consider the equations that result from (\ref{eq:gray-scott}) by dropping
the diffusion terms. The above mentioned stable fixed point exists throughout
parameter space but for some range of \(\mathbf{F}\) and \(\mathbf{k}\)
Pearson\cite{pearson_10} has found another fixed point. Fixing \(\mathbf{k}\), increase
or decrease of \(\mathbf{F}\) causes the second steady state to be lost. The
process of losing or gaining fixed-points as a function of the system's
parameters is called bifurcation. Looking at the non-diffusive system's
behavior under change in the external parameters can be very useful in order
to understand the general behavior of the diffusive-system under microscopic simulation.
Two types of bifurcations are usually distinguished - saddle-node bifurcation
and Hopf bifurcation.
In a saddle-node bifurcation either a fixed point appears and splits into two
fixed points or two fixed points become one and then disappear. The main feature
of such bifurcations is the nature of the single fixed point at the
bifurcation. The linearized system must have one zero eigenvalue and one
non-zero eigenvalue at that point. In other words if
\begin{equation}
\label{eq:linearized_system}
\frac{\partial\vec{f(t)}}{\partial{}t}=\mathbf{M}\vec{f(t)}
\end{equation}
is the linear expansion of the system around the fixed point, with
\(\mathbf{M}\) being the 2x2
coefficient matrix, then
at the bifurcation point \(det(\mathbf{M})=0\) and \(tr(\mathbf{M})\neq0\).
Hopf bifurcation occurs when an unstable (stable) focus goes through the
creation of a limit cycle and becomes stable (unstable). At the bifurcation
point both eigenvalues are purely imaginary and the Real part of the
eigenvalues is positive (negative) before the bifurcation point and negative
(positive) after the bifurcation point.
In the case of our system, given \(\mathbf{k}\) the second fixed point is lost
through saddle-node bifurcation as \(\mathbf{F}\) is increased and by Hopf bifurcation as
\(\mathbf{F}\) is decreased.
Pearson \cite{pearson_9} considers
\(\rho_{C}\) as the density of a liquid fuel and \(\rho_{A}\) as a temperature field. Fuel is
constantly fed from an external reservoir kept at a constant concentration.
The fixed point in (\ref{eq:solution}) is stable therefore small perturbations in
the temperature field around the zero will tend to die out and return back to the
zero, but what if the perturbation was big enough (a match is thrown into the
fuel) to cause the term \(\rho_{C}\rho_{A}^{2}\) to be significant?
Here the auto-catalytic nature of (\ref{eq:gray-scott}) would cause temperature and
fuel consumption to increase inside the ignited area and the fire starts
spreading across space, leaving regions with low concentrations of fuel. Given
the feed parameters \(\mathbf{F}\) and \(\mathbf{k}\) the behavior is
controlled by the value of \(D_{A}\). Large values of \(D_{A}\) would cause fast moving fire wavefronts and
small values of \(D_{A}\) cause stable standing spots of fire fueled by the
fast moving \(\mathbf{C}\). One of the more interesting patterns arising from
(\ref{eq:gray-scott}) is that of replicating spots predicted in simulations
\cite{pearson_10} and confirmed in experiment \cite{experimental_spots}. Given the right parameter
values an initial large perturbation breaks up into standing spots, the fuel
concentration at the middle of the spot decreases and causes the spot to divide
into two smaller expanding replicas and so on.
Analytic spot-like solutions to (\ref{eq:gray-scott}) were found in
\cite{pearson_9} but only in the one-dimensional case and in the
\(D_{A}<<1\) limit.
The different two-dimensional patterns resulting from the Euler integration of (\ref{eq:gray-scott})
were classified in \cite{pearson_10}.
\subsection{Reproducing the replicating spots phenomenon in microscopic simulations.}
The following equations :
\begin{eqnarray}
\label{eq:PDEspots}
\frac{\partial\rho_{C}}{\partial{}t}=
2\cdot10^{-5}\nabla^{2}\rho_{C}-\rho_{C}\rho_{A}^{2}+0.018(1-\rho_{C})
\nonumber\\
\frac{\partial\rho_{A}}{\partial{}t}=
10^{-5}\nabla^{2}\rho_{A}+\rho_{C}\rho_{A}^{2}-0.074\rho_{A}
\end{eqnarray}
fall in the parameter space domain described by Pearson\cite{pearson_10} to produce the replicating
spots behavior. Pearson has used lattice-spacing of 0.01 and has started his
simulation in the trivial state: \(\rho_{A}=0\), \(\rho_{C}=0\) apart from the
perturbated square put at: \(\rho_{A}=0.25\), \(\rho_{C}=0.5\).
The reaction dynamics corresponding to equation (\ref{eq:PDEspots}) are:
\begin{enumerate}
\item \(\mathbf{A} + \mathbf{A} + \mathbf{C} \rightarrow \mathbf{A} + \mathbf{A} + \mathbf{A} (\mathbf{1})\)
\item \(\mathbf{A} \rightarrow \emptyset (\mathbf{0.74})\)
\item \(\mathbf{C} \rightarrow \emptyset (\mathbf{0.018})\)
\item \( \emptyset \rightarrow \mathbf{C} (\mathbf{0.018})\)
\end{enumerate}
The simulation space (of size 64x64 sites) is seeded with 1000 \(\mathbf{C}\)s and no \(\mathbf{A}\)s, except for the
perturbated square (of size 20x20 sites) that is seeded with 250 \(\mathbf{A}\)s per
cell and 500 \(\mathbf{C}\)s per cell. \(n_{0}\) was chosen to be \(10^7\) and
lattice spacing was chosen to be \(0.01\) (so a concentration of \(1\)
corresponds to \(1000\) reactants per lattice site).
\begin{figure}
\frame{
\vspace {0.01in}
\epsfxsize=3.1in
\epsfbox{Lehmann8.eps}
}
\caption{
Replicating spots of \(\mathbf{A}\) created by the microscopic simulation.
Spots in the process of dividing are shown.
}
\label{fig:spotsA}
\end{figure}
\begin{figure}
\frame{
\vspace {0.01in}
\epsfxsize=3.1in
\epsfbox{Lehmann9.eps}
}
\caption{
The same simulation as in
fig(\ref{fig:spotsA}) at a later time, the original spots have split in two.
}
\label{fig:spotsC}
\end{figure}
\begin{figure}
\frame{
\vspace {0.01in}
\epsfxsize=3.1in
\epsfbox{Lehmann10.eps}
}
\caption{
The same simulation as in
fig(\ref{fig:spotsC}) and (\ref{fig:spotsA}) but at a later time. This figure
shows species \(\mathbf{C}\).
}
\label{fig:spotsD}
\end{figure}
\begin{figure}
\frame{
\vspace {0.01in}
\epsfxsize=3.1in
\epsfbox{Lehmann11.eps}
}
\caption{
This figure shows results of PDE simulation of the replication spots.
Again showing species \(\mathbf{C}\).
}
\label{fig:spotsE}
\end{figure}
Although we have been successful in reproducing the general replicating spots
behavior, a closer look at the results as shown in figures (\ref{fig:spotsC})
and (\ref{fig:spotsA}) shows that the randomness introduced in the microscopic
simulation has the effect of breaking the square symmetry preserved by the PDE
simulations. Pearson originally achieved this symmetry breaking by introducing noise
in the initial conditions, but the same effect is created by the microscopic simulation.
Figure (\ref{fig:spotsD}) shows the results of the simulation at a later
time, replication has formed more spots. Figure (\ref{fig:spotsE}) shows the results of a PDE simulation of the
same system. We can see the similarity in the results of the two simulations,
due to the fact that \(n_{0}\) is large. We shall see, in the next subsection,
that when other parameters are chosen (including \(n_{0}\)), the two systems
give crucially different results.
\subsection{The uniqueness of persistently dynamic reaction-fronts to
microscopic simulation. }
We now present an example in which microscopic simulations of the Gray-Scott model
create different results than the ones obtained by the PDE approach. The
phenomenon we are interested in is the presence persistent reaction fronts
propagating through the simulation space. Much research has been devoted to
reaction fronts
\cite{oscillations}\cite{traveling_RNA}\cite{CIMA_1}\cite{CIMA_2}\cite{CIMA_3}.
Patterns which are
non-stationary in time and inhomogeneous in space at the same time are rare in a homogeneous
medium. We suggest a system where the reaction-fronts are spatially and
temporally non-stationary within a homogeneous medium (all the participating
reactants have non-zero equal diffusion rates, so the system is not equivalent
to a non-homogeneous medium system).
Our persistent reaction-fronts are
created by the following mechanism: A localized small \(\mathbf{A}\) area
consumes \(\mathbf{C}\) reactants in its surroundings and creates reaction-fronts of high
\(\mathbf{A}\) concentration propagating across the simulation space cleaning
areas from the presence of \(\mathbf{C}\)s. The reaction-front then runs out of
high \(\mathbf{C}\) concentration areas and decays, giving rise to the renewal
of \(\mathbf{C}\)s and new wavefronts produced by the surviving \(\mathbf{A}\)s and so on. We have
simulated microscopically the following reaction-diffusion system:
\begin{enumerate}
\item \(\mathbf{A} + \mathbf{A} + \mathbf{C} \rightarrow \mathbf{A} + \mathbf{A} + \mathbf{A} (\mathbf{8})\)
\item \(\mathbf{A} \rightarrow \emptyset (\mathbf{0.3})\)
\item \(\mathbf{C} \rightarrow \emptyset (\mathbf{0.02})\)
\item \( \emptyset \rightarrow \mathbf{C} (\mathbf{0.1})\)
\end{enumerate}
The diffusion coefficients are \(D_{A}=1\) and
\(D_{C}=1\). The simulation size chosen was 80x80. \(n_{0}=1\) and
\(\Delta{}x=1\). The simulation space was seeded with 1 \(\mathbf{C}\) per
site, and no \(\mathbf{A}\)'s were seeded except for a perturbed square of size
64x64 sites where 5 reactants of species \(\mathbf{A}\) per site were seeded.
The simulation shows persistent reaction-fronts of the type seen
in fig(\ref{fig:mic_fronts}).
\begin{figure}
\frame{
\vspace {0.01in}
\epsfxsize=3.1in
\epsfbox{Lehmann12.eps}
}
\caption{
Clear high \(\mathbf{A}\) concentration reaction-fronts can be seen during microscopic simulations. These
reaction-fronts are persistent, being supported by surviving \(\mathbf{A}\) reactants.
}
\label{fig:mic_fronts}
\end{figure}
\begin{figure}
\frame{
\vspace {0.01in}
\epsfxsize=3.1in
\epsfbox{Lehmann13.eps}
}
\caption{
The initial reaction-front seen when integrating the corresponding PDE system
with \(\Delta{}x=1\) and \(n_{0}=1\). The reaction-front dies out and the
system is driven to the constant \(\rho_{A}=0\) steady-state.
}
\label{fig:pde_fronts}
\end{figure}
Using Euler Integration to simulate the system we get different results, the initial square perturbation does propagate in the form of
a reaction-front shown in fig(\ref{fig:pde_fronts}), but after the initial front cleans the space from high
\(\mathbf{C}\) concentration areas the front decays and gives way to the total
elimination of \(\mathbf{A}\)s from the simulation space. Unlike the
microscopic case in which the discretization and randomness leave behind some islands of
\(\mathbf{A}\) from which the next reaction-front can emerge, the Euler
Integration drives the system towards the \(\rho_{A}=0\)
steady-state. Introducing noise in the initial conditions does not change the
situation the system tends to smooth out these noises and is driven to the
constant steady state.
We have tried to reproduce the persistent reaction-fronts pattern using Euler
Integration with no success, although it is plausible that the same behavior
will appear for high values of \(n_{0}\).
\section{Microscopic simulation of Marketing models}
\subsection{Using rate equations to describe the marketing of products}
The application of physical sciences methods to economic and financial
research, nowadays common practice among the scientific community\cite{bouchaud_6}, was initiated by the
work of Bachelier\cite{bachelier_3} and Mandelbrot\cite{mandelbrot_2}. The
basic situation is similar to the other areas of research we have previously discussed, global ``macroscopic'' economic phenomena are generated by the underlying
``microscopic'' process of buy and sell. Under these circumstances it is
natural to try and use microscopic simulation of market models in order to
reproduce as an example we could take the
spreading of steam engines or gunpowder starting from localized innovative centers
and sweeping across wide regions of the globe.
When using microscopic simulation to describe product
marketing one can regard a lattice-space element in two different ways:
\begin{enumerate}
\item As a geographical region - neighborhood, town or country.
\item As a business entity - company or corporate.
\end{enumerate}
In the first case neighboring elements represent geographically close
regions, in the latter case they represent companies that are in business contact.
The reactants could stand for any kind of valuable passing from one place or
business entity to another. One could have reactants representing
money, products, ideas, manpower or technology. Reactions could represent economic
processes in which valuables are transformed, lost or created. Diffusion stands
for the transfer of these valuables from one location or business entity to
another. Although the use of microscopic simulations in the context of marketing seems natural
enough there still are some flaws to point at and points to defend. The
microscopic simulation
is probabilistic in nature while some of the processes we try to
represent are deterministic. As we explained in previous sections rates
represent the occurrence rate of an underlying Poisson process for some of the
economic processes described we have no reason to think that they are
Poissonic in nature. Furthermore the microscopic simulation's lattice-space elements are always
connected with exactly 4 neighboring sites, thus disabling the simulation of
environments in which the degree of ``connectedness'' varies from site to
site. If we consider sites to be business entities as we have previously
suggested, we certainly would like to consider different amounts of
connectivity between business entities - a feature not supported by the
microscopic simulation system.
\subsection {The ``Tamagotchi'' model}
In collaboration with J. Goldenberg and D. Mazursky\footnote{Hebrew
University School of Business Management} we have devised the
following set of reactions:\footnote{Numbers at the right of reactions are the
corresponding reaction-rates.}
\begin{enumerate}
\item \(\mathbf{B} + \mathbf{C} \rightarrow \mathbf{A} + \mathbf{B} + \mathbf{B} (\mathbf{8})\)
\item \(\mathbf{A} \rightarrow \emptyset (\mathbf{0.7})\)
\item \(\mathbf{B} \rightarrow \emptyset (\mathbf{1.5})\)
\item \(\emptyset \rightarrow \mathbf{C} (\mathbf{0.1})\)
\item \(\mathbf{C} \rightarrow \emptyset (\mathbf{0.02})\)
\end{enumerate}
The inspiration to these reaction dynamics comes from the wave of
``Tamagotchi'' games we have been subjected to during a short period of 1996.
These reaction dynamics were simulated in a system of size 128x128 sites seeded
with 1 reactant of species \(\mathbf{C}\) only. A square of size 64x64 sites was
seeded with 1 reactant of species \(\mathbf{C}\) per cell and 5 reactant of
species \(\mathbf{B}\) per cite. Other coefficients are: \(n_{0}=1\), \(\Delta{}x=1\).
\begin{itemize}
\item \(\mathbf{A}\) - represents a product (``Tamagotchi'').
\item \(\mathbf{B}\) - represents the idea or concept of the product.
\item \(\mathbf{C}\) - represents money.
\end{itemize}
In this model neighboring locations should be interpreted in the geographical sense.
The only diffusing reactant is \(\mathbf{B}\) with diffusion coefficient
\(\mathbf{1}\) but since \(\mathbf{B}\)
represents an idea or concept the diffusion it performs is replicative. By
replicative we mean that instead of hopping from one cell to a neighboring
cell a copy of the original reactant is created and planted in the neighboring
cell. This replicative diffusion represents the fact that ideas or concepts
do not physically pass from one location to another, instead a location exposed
to a new idea or concept in a neighboring location creates its own copy of the
original idea. In terms of the microscopic simulation algorithm the only change is that when
executing the diffusion of \(\mathbf{B}\), a \(\mathbf{B}\) unit is added to the
target location but none is subtracted in the original location. The first
reaction represents the following process: A person exposed to the product
concept that is in possession of money spends the money on buying the product
and a new concept or idea of the product is ``born'' in that persons mind. The
second reaction represents the loss of products due to old age. The third
reaction represents the fact that ideas tend to be forgotten. The fourth
reaction represents an influx of money and the fifth reaction represents the
spending of money on products other than \(\mathbf{A}\). The ``replicative''
diffusion that \(\mathbf{B}\) undergoes represents the influence ideas have on
neighboring locations.
\subsection{Wave fronts the ``Tamagotchi'' model}
During the simulations of the ``Tamagotchi'' model we have observed long
lasting wavefronts of high \(\mathbf{A}\) concentrations moving across the
simulation space. The process giving rise to such waves seems to be the
following: Starting with a small amount of \(\mathbf{A}\) in a \(\mathbf{C}\)
rich area the concepts created by this small \(\mathbf{A}\) concentration
propagate to neighboring sites and induce the decrease of \(\mathbf{C}\)
concentration and increase in \(\mathbf{A}\) concentration in these sites. The
areas to which this \(\mathbf{B}\) influence arrives then become \(\mathbf{C}\)
poor areas stopping the increase in \(\mathbf{A}\)'s concentration giving rise
to their decrease until \(\mathbf{C}\)'s concentration is high enough to
sustain the next wavefront.
\begin{figure}
\frame{
\vspace {0.01in}
\epsfxsize=3.1in
\epsfbox{Lehmann14.eps}
}
\caption{
Snap shot of \(\mathbf{A}\)'s concentration clear wavefronts can be seen. Sites
with concentration above 2 are drawn black.
}
\label{fig:mesh131A}
\end{figure}
In order to estimate the propagation speed of the wavefronts we have used the
following correlation integral. Let \(\Omega\) denote our two-dimensional
reaction chamber and let \(\rho_{A}(\vec{x},t)\) denote the
concentration of \(\mathbf{A}\) at location \(\vec{x}\) and time
\(\mathbf{t_{0}}\) for \(\vec{x}\in\Omega\) and \(\mathbf{t_{0}}\geq\mathbf{0}\). For \(\mathbf{r}\geq\mathbf{0}\) let:
\begin{equation}
\label{eq:volume}
\textit{V}(\Omega)=\int_{\Omega}1d\vec{x}
\end{equation}
\begin{eqnarray}
\label{eq:correlations1D}
\mathbf{C}(r,t_{0},t)=\frac{1}{\textit{V}(\Omega)}\int^{2\pi}_0\int_{\Omega}\rho_{A}(\vec{x},t_{0})\rho_{A}(\vec{x}+
\left(
\begin{array}{c}
r\cos\theta\\r\sin\theta
\end{array}
\right)
,t_{0}+t)d\vec{x}d\theta - \nonumber\\
\frac{1}{\textit{V}(\Omega)^{2}}\int_{\Omega}\rho_{A}(\vec{x},t_{0})d\vec{x}\int_{\Omega}\rho_{A}(\vec{y},t_{0}+t)d\vec{y}
\end{eqnarray}
Given \(\mathbf{t}\geq\mathbf{0}\) and \(\mathbf{r}\geq\mathbf{0}\)
(\ref{eq:correlations1D}) is a measure for the
correlation between \(\mathbf{A}\)'s concentration at time \(\mathbf{t_{0}}\) and
\(\mathbf{A}\)'s concentration at time \(\mathbf{t_{0}+t}\) at locations with
distance \(\mathbf{r}\). If \(\rho_{A}\) is static \(C(r,t_{0},t)\)
is maximal for \(\mathbf{r}=\mathbf{0}\) for all \(\mathbf{t}\)s. If on the
other hand \(\rho_{A}\) is moving at speed \(\mathbf{v}\) we expect for a given
\(\mathbf{t}\) that C is maximal for \(\mathbf{r}\)=\(\mathbf{v}\mathbf{t}\).
\begin{figure}
\vspace {0.01in}
\epsfxsize=3.1in
\epsfbox{Lehmann15.eps}
\caption{
\(C(r,t_{0},t)\) plotted as a function of \textit{r} for \textit{t}=3,4,5, the
higher the plot the lower is \textit{t}. A clear 'hump' can be seen to advance in
\textit{r} as \textit{t} increases, corresponding to the evolution of
the wavefront.
}
\label{fig:speeds131A}
\end{figure}
\begin{figure}
\vspace {0.01in}
\epsfxsize=3.1in
\epsfbox{Lehmann16.eps}
\caption{
\(C(r,t_{0}+50,t)\) plotted as a f unction of \textit{r} for \textit{t}=3,4,5, the
higher the plot the lower is \textit{t}. The same 'hump' as in (\ref{fig:speeds131A})
although \(t_{0}\) has increased by 50.
}
\label{fig:speeds131B}
\end{figure}
In both (\ref{fig:speeds131A}) and (\ref{fig:speeds131B}) we can see an
increase of 2 in \(\mathbf{r}\) as \(\mathbf{t}\) increases by 2 giving rise
to an estimate of:
\begin{equation}
v\approx{}1
\end{equation}
When speeds are measured in units of lattice-sites per simulation-time.
The similar approximations due to the different choices of \(t_{0}\) indicate
the constant speed of advance of the wavefront.
\section{Acknowledgment}
We wish to thank Sorin Solomon for instructing us in this work. We thank Jacob
Goldenberg and David Mazursky their help in finding the application
to marketing. We also thank Nadav Shnerb for fruitful discussions.
|
\section{Introduction - Astronomy of the Next Century and
Gravitational Physics}
Two major directions of astronomy in the next century are {\em high
energy ($x$-ray, $\gamma$-ray) astronomy} and {\em gravitational wave
astronomy.} The former is driven by observations by $x$- and
$\gamma$-ray satellites, e.g., CGRO, AXAF, XTE, HETE II,
GLAST~\cite{missions}, current or planned for the next few years.
High energy radiation is often emitted in regions of strong
gravitational fields, near black holes (BHs) or neutron stars (NSs).
One of the biggest mysteries of modern astronomy, $\gamma$-ray bursts,
is likely to be generated by events involving NSs or BHs. For the full
description of strong, dynamic gravitational fields, we need
Einstein's theory of general relativity.
The second major direction, gravitational wave astronomy, involves
directly the dynamical nature of spacetime in the Einstein theory of
gravity. The tremendous recent interest in this frontier is driven by
the gravitational wave observatories presently being built or planned
in US, Europe and outer space, e.g., LIGO, VIRGO, GEO600, LISA, LAGOS
~\cite{nasa}, and the Lunar Outpost Astrophysics Program~\cite{nasa}.
The American LIGO and its European counterparts VIRGO and GEO600 are
scheduled to be on line in a few years~\cite{LIGOweb}, making
gravitational wave astronomy a reality. These observatories provide a
completely new window on the universe: existing observations are
mainly provided by the electromagnetic spectrum, emitted by individual
electrons, atoms or molecules, easily absorbed, scattered and
dispersed. Gravitational waves are produced by coherent bulk motion
of matter and travel nearly unscathed through space, coming to us
carrying the information of the strong field regions where they were
originally generated.~\cite{thorne96} This new window will provide very
different information about our universe that is either difficult or
impossible to obtain by traditional means.
The numerical determination of the gravitational waveform is crucial
for gravitational wave astronomy. Physical information in the data is
to be extracted through template matching techniques~\cite{Cutler93},
which {\em presupposes} that reliable example waveforms are
known.~\cite{template} Gravitational waveforms are important both as
probes of the fundamental nature of gravity, and for the unique
physical and astronomical information they carry. The information
would be difficult to obtain otherwise, ranging from nuclear physics
(e.g., the EOS of NSs~\cite{Cutler93}) to cosmology (e.g., direct
determination of the Hubble constant~\cite{schutz86}). In most situations, the
gravitational waveforms cannot be calculated without full scale
general relativistic numerical simulations.
In short, both of these frontiers of astronomy call for
numerical simulations based on the Einstein theory of gravity.
If astrophysicists are to fully understand the non-linear and
dynamical gravitational fields involved in these observational data,
detailed modeling taking dynamic general relativity into full account
must be carried out.
\section{Challenges of Computational General Relativistic Astrophysics}
The application of the Einstein theory of gravity to
{\it realistic} astrophysical systems needs computational power in the
range of (at least) multi-TeraFlop/TeraByte, and corresponding
capabilities in visualization, networking and storage.
\noindent $\bullet$ Computational challenges due to the complexity of
the physics involved: The Einstein equations are probably the most
complex partial differential equations in all of physics, forming a
system of dozens of coupled, nonlinear equations, with thousands of
terms, of mixed hyperbolic, elliptic, and even undefined types in a
general coordinate system. The evolution has elliptic constraints
that should be satisfied at all times. In simulations without
symmetry, as would be the case for realistic processes, it involves
hundreds of 3D arrays, and ten of thousands of operations per grid
point per update. Moreover, for simulations of
astrophysical processes, we need to integrate numerical relativity
with traditional tools of computational astrophysics, including
hydrodynamics, nuclear astrophysics, radiation transport and
magneto-hydrodynamics, which govern the evolution of the source terms
(i.e., the right hand side) of the Einstein equations. This
complexity demands massively parallel computation.
\noindent $\bullet$ The object under numerical construction being the
spacetime itself presents unique challenges: According to the
singularity theorems of general relativity, region of strong gravity
often generate spacetime singularities. Due to the need to avoid
spacetime singularities~\cite{numrel}, and to obtain long term
stability in the numerical simulations, sophisticated control of the
coordinate system is needed for the construction of a
numerical spacetime. This dynamic interplay between the
spacetime being constructed and the computational coordinate choice
itself (``gauge choice'') is a unique feature of general relativity
that makes the numerical simulations much more demanding. Beside
extra computational power, advanced visualization tools, preferably
real time interactive ``window into the oven'' visualization, are
particularly useful in the numerical construction.
\noindent $\bullet$ The multi-scale problem: Astrophysics of strongly
gravitating systems inherently involves many length and time scales.
The microphysics of the shortest scale (the nuclear force), controls
macroscopic dynamics on the stellar scale, such as the formation and
collapse of neutron stars (NSs). On the other hand, the stellar scale
is at least 10 times {\it less} than the wavelength of the
gravitational waves emitted, and many orders of magnitude less than
the astronomical scales of their accretion disk and jets; these larger
scales provide the directly observed signals. Numerical studies of
these systems, aiming at direct comparison with observations,
fundamentally require the capability of handling a wide range of
dynamical time and length scales. While such multi-scale problems can
be handled with advanced 3D AMR techniques, it leads to further
requirements on computation power and (3D AMR) visualization.
In short, in order to meet the challenges of Computational General
Relativistic Astrophysics we need to push not only the frontier of the
computation power for number crunching. The visualization requires
basically as much computer power as what generates the data. The
highly multi-disciplinary nature of the research demands collaborative
code development. The large amount of data, visualization needs, and
collaborative effort require high performance networking and
meta-computing. In the following section we use a specific sample
problem to illustrate the requirements on Flop rate, memory, disk and
storage sizes, which in turns determine the base line of visualization
and networking requirements. Where we stand at present will also be
discussed briefly.
\section{Neutron Star Coalescence As An Example on Computational Requirements}
We use the problem of coalescing binary neutron stars to show the
computational requirements in general relativistic astrophysics. The
reason that the coalescence of neutron stars is a meaningful example
is many-fold: It is a significant problem in astrophysics and
astronomy; it involves many ingredients in general relativistic
astrophysics; and it is a problem attracting much current research
effort both nationally and internationally.
\medskip
\noindent $\bullet$ Coalescing neutron star binary systems are common in the
Universe, with the well known Hulse-Taylor binary pulsar PSR1913+16
being an example. The coalescence events are expected
to be detectable by LIGO, with an observation rate of 29 yr$^{-1}$ for
$h=0.5$ and 43 yr$^{-1}$ for $h=0.8$. ~\cite{finn95}
\noindent $\bullet$ The physical information
in LIGO data is to be extracted through the standard template matching
technique~\cite{Cutler93}. For this we need to determine the waveforms
of the gravitational radiation generated by the coalescence events,
which can only be obtained through large scale simulations.
\noindent $\bullet$ A very enticing reason for studying the
coalescence event lies in the fact that observations of such events by
gravitational wave observatories may allow us to determine
cosmological parameters like $H_0$ and $q_0$, without going through the
cosmic distance ladder, and is independent of the optical
identification of the source and the evolution of the source rate
density with redshift.\cite{schutz86,finn95}
\noindent $\bullet$ Gravitational wave signals from coalescing
binaries may reveal important information on the equation of state of
dense nuclear matter, including the nuclear compression modulus, the
hadronic effective masses, the relative hyperon-nucleon and
nucleon-nucleon coupling constants, possible kaon condensation and a
quark/hadron phase transition.
\noindent $\bullet$ Study of coalescing neutron star
binaries may also answer other long standing questions in nuclear
astrophysics.
NSNS binary mergers may eject
\cite{ls76} extremely neutron-rich matter which decompresses,
beta-decays and neutron captures, forming the classical r-process
\cite{lmrs77,mbc92}. Detailed numerical simulations of
the shock heating and mass ejection process are needed.
\noindent $\bullet$ Coalescing neutron star binaries are among the
most popular candidates of gamma ray bursts.~\cite{paczyn86,piran95}
In order to evaluate the feasibility of the model, detailed studies
taking the full general relativistic effects are needed to determine the
maximum possible energy released, heating and mass ejection in the
coalescence process.
\subsection{Minimum Configuration}
\noindent $\bullet$ Description of the Physical System:
Two 1.4 solar mass neutron stars in head-on collision falling in from
infinity. General relativistic simulation begins when the two stars are
$4 R$ apart, with $R=$ radius of star. Simulation covers $10ms$ in time for
the dynamics of the merging and ringdown phases, and $20R$ in space for resolution
and boundary considerations.
\noindent $\bullet$ Purposes: Study the general relativistic dynamics of the merging and ringdown phases
of head-on collision.
\noindent $\bullet$ Grid Setup: Resolution=25 gridpoints/$R$, Total Grid Size = $500 ^ 3 = 10 ^ 8 $
\noindent $\bullet$ Memory Requirement: $180GBytes$
\noindent $\bullet$ Floating Point Operations:
Flops/gridpt/time step = $10 ^4 ~~~~$ (With only weak coordinate control)
Total number of time steps = $10 ^4 ~~~~$
Total flops = $10 ^ {16} ~~~~$
Run time = 3 hours ~~~~ (With 1 TeraFlops sustained)
\noindent $\bullet$ Disk:
Run time disk size = 800 GBytes ~~(Output 10 functions with 1/100 sampling)
Storage = 8 TeraBytes ~~~~(with 10 runs for comparison studies)
\noindent $\bullet$ Present Status:
Code for carrying out this simulation is {\it currently} available. A
code constructed for the NASA Neutron Star Grand Challenge Project which is
capable of solving the full Einstein equations coupled to general
relativistic hydrodynamics has recently been released.~\cite{ourwebpage} This
code has been tested on a 1024 node T3E-1200 (provided for
the neutron star project for performance tests, though not available for
production runs), achieving 142GFLops and linear scaling up to 1024 nodes.
A summary of the test results are given below.
(The NSF Black Hole Grand Challenge Project is also constructing
massively parallel code for solving the Einstein equations, see
~\cite{matzner} for present status.)
\noindent {\em \underline{Code tested}}:
NASA Neutron Star Grand Challenge GR3D Einstein Spacetime (ADM)
coupled to MAHC HYPERBOLIC\_HYDRO (code tested with the released
version, without special tuning for this 1024 node machine.)
\noindent {\em \underline{Date tested}}: May 10, 1998
\begin{verbatim}
32 bit 64 bit
--------------------------------------------------------------------
Grid Size per Processor 84x84x84 66x66x66
Processor topology 8 x8 x16 8 x8 x16
Total Grid Size 644 x 644 x 1284 500 x 500 x 996
Single Proc MFlop/sec 144.35 118.33
Aggregate GFlop/sec 142.2 115.8
Scaling efficiency 96.
--------------------------------------------------------------------
\end{verbatim}
\subsection{Medium Configuration}
\noindent $\bullet$ Description of Physical System:
Two 1.4 solar mass neturon stars in inspiral coalescence. Full
general relativistic simulation begins when stars enter the last 8
orbits. Simulation covers $60ms$ in time and $22R$ in space.
\noindent $\bullet$ Purposes:
Study the general relativistic inspiral dynamics
beginning with the 3PN breakpoint. This enables reliable initial
data to be set. Study the effects of the angular momentum and
gravitational radiation backreaction on shock heating in the merger phase.
\noindent $\bullet$ Grid Setup:
Resolution=50 gridpoints/$R$, Total Grid Size = $10 ^ 9 $
\noindent $\bullet$ Memory Requirement: $1.8 TBytes$
\noindent $\bullet$ Floating Point Operations:
Flops/gridpt/time step = $10 ^4 ~~~~$ (With only weak coordinate control)
Total number of time steps = $10 ^5 ~~~~$
Total flops = $10 ^ {18} ~~~~$
Run time = 300 hours ~~~~(With 1 TeraFlops sustained)
\noindent $\bullet$ Disk:
Run time disk size = 20 TBytes ~~~(Output 10 functions with 1/400 sampling)
Storage = 100 TeraBytes (with 5 runs for comparison studies)
\noindent $\bullet$ Visualization: Need parallel visualization engine.
\noindent $\bullet$ Present Status:
Code basically ready for pilot studies. Tests of the effects of the
implementation of weak coordinate control to be performed.
\subsection{Preferred Configuration}
\noindent $\bullet$ Description of Physical System:
Two 1.4 solar mass neturon stars in inspiral coalescence. Full
general relativistic simulation begins when stars enter the last 8
orbits. Simulation covers $60ms$ in time and $40R$ in space (one
wavelength for gravitational wave with period 1ms).
\noindent $\bullet$ Purposes: Study the same system with strong coordinate control and more reliable wavefrom
extraction.
\noindent $\bullet$ Grid Setup:
Resolution=50 gridpoints/$R$, Total Grid Size = $10 ^ {10} $
\noindent $\bullet$ Memory Requirement: $18 TBytes$
\noindent $\bullet$ Floating Point Operations:
Flops/gridpt/time step = $10 ^5 ~~~~$ (With strong coordinate control)
Total number of time steps = $10 ^5 ~~~~$
Total flops = $10 ^ {20} ~~~~$
Run time = 3,000 hours ~~~~(With 10 TeraFlops sustained)
\noindent $\bullet$ Disk:
Run time disk size = 200 TBytes ~~(Output 10 functions with 1/400 sampling)
Storage = 1000 TeraBytes
(with 5 runs for comparison studies, template preparation not included)
\noindent $\bullet$ Need to push the frontiers on computation, storage, visualization, and networking.
\noindent $\bullet$ Present Status:
Code basically ready for pilot studies. Efficient control
the coordinate system to be investigated.
\section{Acknowledgements}
I thank S. Finn, K. Blackburn, M. Miller, L. Smarr, B. Sugar,
M. Tobias, J. Towns, C. Will, and J. York for useful input in
preparing this document.
The gereral relativistic astrophysics code "GR3D" discussed in Sec. 4
is developed by the NCSA-Potsdam-Wash U numerical relativity
collaboration, with support from the NSF Gravitational Physics Program
Grant No. Phy-96-00507, NASA HPCC/ESS Grand Challenge Applications
Grant No. NCCS5-153, NSF NRAC Allocation Grant no. MCA93S025, and
the Albert Einstein Institute.
|
\section{Introduction}
The dynamics of elastic objects
driven through an environment with random pinning
interactions is complex and its present understanding
still incomplete. Processes of this type occur in
microscopic systems such as charge density waves
driven by an electric field \cite{CDW,FLR},
superconductors subject to external magnetic fields
\cite{LRK}, and polymers moving through a
inhomogeneous environment. Related processes on a
macroscopic scale occur in the motion of geological
faults \cite{BK,CL,FDR}, or in sliding friction between
an elastic object and a rigid one \cite{VG}. Theoretical
studies of these phenomena are usually modeled
by an elastic network moving in a rigid environment
or over a rigid substrate with random interactions
between them. In the simplified case where one of
the two objects is homogeneous, one may distinguish
between $\it random\;network$ models, where the
random interactions are assigned to sites on the elastic
network while the medium or substrate are
homogeneous, and the $\it random \;
substrate $, models where the random interactions are
assigned to the sites of the substrate or the embedding
medium while the elastic network is homogeneous.
These two type of models are illustrated in Fig. 1
Random network models were invoked to describe
randomly pinned charge density waves (CDW) in
solids \cite{DF}-\cite{P} and were widely studied.
Most recent work on random substrate models
concentrated on the high velocity limit, where the
distortions of the network induced by the random
interactions are mild \cite{GLD}-\cite{BAL}, while the
dynamical properties near the critical external field for
depinning received less attention.
We focus on the simplest $\it discrete$ realization of
the random substrate model containing no ad-hoc
features such as velocity dependent forces \cite{CL}, and
compare it to the analogous random network model. The
study shows that the dynamics of the random substrate
model above depinning threshold is interesting and
complex, featuring several distinct regimes.
In contrast, in the random network case the elastic
object follows a simple steady state motion.
The pinned state of the random substrate model below
threshold is characterized by a wide distribution of $\it
strain\; avalanches$. Similar avalanches occur in the
random network model, however the critical exponents
describing the distributions of avalanche sizes are
definitely different.
The properties of the discrete random substrate model
depend on three parameters: The external field, the ratio
of pinning strength to the elastic stiffness, and the
characteristic size of the local pinning regions. We find
that whereas in some parts of the relevant phase
diagram the two models are approximately equivalent,
through most of it there are pronounced differences
between them, and the critical behavior belongs to
distinct universality classes.
\section{The Model}
Sliding elastic objects are represented here by a discrete
chain of particles connected by springs. (see Fig. 1)
In the random network limit each particle is attributed a
random pinning force, while the substrate is
homogeneous.
For an external force $E$ switched on at $t=0$ acting
uniformly on all particles, the dynamics of the
displacements of the positions of the particles from their
initial equilibrium positions: $\xi_i = x_i(t)-x_i(0)$ is
described by the following overdamped equations of
motion:
$\it random\;network$
\begin{equation}
{\partial{\xi_i}}/{\partial{t}}=
\max(0,\kappa({\xi}_{i+1}+\xi_{i-1}-2\xi_i)+E-
V(i))
\end{equation}
This realization of the random network model, also
known as the $\it ratchet\;model$
\cite{W87} successfully describes the dynamical phase
transition of charge density waves from a pinned state to
a DC current conducting state at a critical external field
$E_c$ \cite{W87,puri,MF,MS,FLRM}.
In the random substrate model the particles interact with
the substrate via local pinning forces $ V(x_i)$ ,
whereby the i'th particle at position $x_i $ (in the
substrate frame of reference) moves if the total force
acting on it is greater than $V(x_i)$. As before, the
dynamics for the displacements: $\xi_i =
x_i-{x_i}(0) $, is described by the following equations:
$\it random\;substrate$
\begin{equation}
{\partial{\xi_i}}/{\partial{t}}=
\max(0,\kappa({\xi}_{i+1}+\xi_{i-1}-2\xi_i)+E-
V({\xi_i}+i))
\end{equation}
Initially, the particles positions form a 1D lattice with
unit spacing:
${x_i}(0)=i$.
Pinning forces are local, with a characteristic range
$\Delta $. In our study the pinning force is constant
inside lattice unit cells, ($\Delta=1 $), and
varies randomly from one cell to the next.
We chose a binary distribution for $V$ of the type:
$\rho(V) = p\delta(V-V_1)+(1-p)
\delta(V-V_0)$, usually with p=0.5.
We adopt periodic boundary conditions, so that the
argument of $V$ in equation (1) is $
{\xi_i\;mod}(L)+i $.
Both models also depict a discrete one dimensional
elastic interface moving in the $\xi$ direction in a $ \it
two\;dimensional$ $(i,\xi)$ plane \cite{puri}, where
the interface is initially $\it flat$.
In the random substrate model a pinning $\it site$ on
the $i$ axis,
transforms into lines of pinning sites of slope unity in the
$(i,\xi)$ plane, as described in Fig. 2.
In the random network case, pinning sites form
$\it columns$ of fixed pinning strength $V_i$ in the
same plane.
$\it Approximate\;scaling\;properties$
The dependence of the general behavior of the random
substrate model on the elastic
stiffness $\kappa$, and on the size $\Delta$ of regions
where pinning forces do not vary, can be simplified by
the following scaling relation \cite{valid} (provided
$\max(\kappa,\kappa\Delta)<1$):
\begin{equation}
\xi(\kappa,\Delta, t)= {\kappa^{-1}}\xi(1,\kappa \Delta,
t/\kappa )
\end{equation}
Extending this scaling relation to $\Delta >1$ results in
a hybrid random network-random substrate model, so
that the pure random substrate model is not preserved.
However, in the limit $\Delta>>1$ the pure random
network model is approached, so that the random
network model and the random substrate model approach
each other in the very high stiffness limit, as $\kappa
\rightarrow\infty$.
\section{Dynamical properties}
In the random network model above the
threshold $E_c$ all particles eventually move with the
same finite velocity as a $ \it rigid$ distorted object.
Adding the $N$ equations of motion Eq. (1) the elastic
forces sum-up to zero, so that the velocity of
the center of mass is:
\cite{W87}
\begin{equation}
{U_{RN}}=E-E_c
\end{equation}
where for a chain of $N$ particles $E_c={1\over
N}{\sum V_i}$
The dynamical behavior of the {\it random substrate}
model is more complex, and two distinct regimes
appear as the driving force is varied above the depinning
threshold $E_1$:
$\it stick-slip\; regime$ : $E_1< E < E_2$ ;
The center of mass of the system moves with a roughly
constant velocity, modulated by fast fluctuations,
although at any time $ \it a\; finite\; fraction$
of the particles are not in motion.
The center of mass velocity $U$ obtained numerically
for several values of stiffness $\kappa$ is shown in Fig. 3.
Each curve corresponds to a single representative
random configuration for chains N=100-1000.
At the threshold there is a small discontinuity, followed
by a linear $E$ dependence.
The fractions of particles that are in motion
at a given instant corresponding to the velocity curves in
Fig 3. appear in the inset.
Deriving the CM velocity by summing the individual
equations Eq. (2) is more subtle in the random substrate
case, since the equations for particles with zero velocity
are actually $\it inequalities$. Time averages over
sufficiently long periods are equivalent to those obtained
from a linearized version of Eq. (1,2) where the ratchet
condition is omittted, leading to:
\begin{equation}
{U_{RS}}= E -\sum n(V_i,E) V_i
\end{equation}
where $n(V_i,E)$ is the time averaged fraction of
particles occupying the i'th unit cell. Since the residence
time of particles over a strong pinning area is longer
than over a weak pinning one, the mean occupation
fractions $n(V_i,E)$ are monotonously increasing
functions on $ V_i$ , so that for two systems sharing
the same distribution of random pinning strengths, the
mean velocity of the random network is always greater
than that of the random substrate system, asymptotically
approaching it from below in the limit of high velocity.
The threshold $E_1$ is the solution of the equation:
\begin{equation}
E_1 -\sum n(V_i,E_1) V_i
\end{equation}
so that the inequality $E_1>E_c$ is always valid.
The center of mass velocity of the random substrate
model can be expressed as a driving force minus a
velocity dependent drag force $F(U)$:
\begin{equation}
U_{RS}=E-F(U_{RS})
\end{equation}
Analysis of the velocity data shows that $F(U)$ is equal to the
static friction force $E_1$ in the limit of zero velocity,
{\it increases} with increasing velocity above $E_1$ reaching some
plateau, eventually tending to $E_c$ for high velocities.
This weakening of the kinetic friction force with increasing
velocity is generated by a detachment instability where the
fraction of immobile particles abruptly vanishes (Fig. 3 inset).
Qualitatively similar features were observed in experiments
where an elastic membrane is dragged over a
rigid substrate, performing stick-slip motion. \cite{VG}
The center of mass velocity for $E$ just above the
depinning transition scales as $(E-E_1)^\beta $. For
the random network model, trivially, $\beta=1$. Our
results for the random substrate model also follow a
linear $E-E_1$ dependence, i.e. $\beta=1$. This is in
contrast with Ref. \cite{CH}, where $\beta =0.47$ is
reported. The differences in exponents may be due to
differences between our model and the specific
realization of the random substrate model in
Ref.\cite{CH}. This is supported by the fact that
in nonlinear dynamical systems with quenched
randomness, universality is often weaker than in
equilibrium critical phenomena, so that some critical
exponents may depend on details on short length scales.
A striking demonstration of this for a directed
polymer in a random medium, a problem closely related
to the one considered here, was published recently
\cite{LZ}.
$\it free\; motion\; regime$: $E> E_2$ ;
In this regime all particles possess a finite velocity at all
times (See Fig. 3, inset). The instantaneous velocity of
the center of mass is made up of a constant part plus a
random $\it washboard $ like modulation induced by
motion over a fluctuating pinning landscape. In the limit:
$E>>E_2$, the velocity tends to that of the analogous random
network Eq. (3),
and the relative fluctuating component of the velocity
diminishes (Fig. 3). The drag force $F(U)$ approaches
$E_c$.
Our model assumes the presence of local friction
forces randomly distributed along the interface, while
their microscopic origin are outside its scope.
The stick-slip dynamics in the random substrate model
is a consequence of the random arrangement of simple
local interactions together with the cooperative effect of
the elastic forces. While the decrease of macroscopic
kinetic friction with velocity in real materials may have
various causes, comparison with our results suggests
surface inhomogeneity or randomness may play a
significant role.
\section{Scaling Properties of the Pinned Sate}
For both models, interfaces evolving by the dynamics
described in Eq (1) or Eq (2), starting from a flat initial
configuration, ($\xi_i=0$ at t=0) and subject to a
constant drive $ E $ below the pinning threshold,
eventually reach a static state of strongly strained
domains, or $\it
strain\;avalanches$
\cite{PIN}, seperated by $ \it virgin $ particles which
never moved. For $E$ approaching $E_c$ from below,
the random network model undergoes a second order
dynamical phase transition \cite{W87,puri}.
The following numerical results show that the random
substrate model follows a qualitatively similar critical
behavior, but of a different universality class.
Fig. 2 shows a typical pinned
interface for
the random substrate model.
Here interfaces are hindered by
the tilted pinning lines, leading to characteristic
triangular structures with a roughness exponent
$\zeta=1$. In contrast, in the
random network model, distortion of domains is
much more pronounced and the roughness exponent is
$\zeta=3/2$
\cite{W87,puri}.
Similar to the random network model and to many other
non-equilibrium systems (e.g. the scaling of avalanche
sizes in self-organized criticality: \cite{BTW}),
the $\it number $ of domains of size $\ell$ per
unit length in the random substrate model has the
generic scaling form:
\begin{equation}
n(\ell)\sim {\lambda^{\sigma}}{{x}^{-\tau}}
\Phi(x)
\end{equation}
where $x=\ell/{\lambda(\epsilon)}$, and $\Phi(x)$ is a
slowly varying function for $x<1$ with
a steep cutoff at $ x>1$. The typical size of the
largest domains, $\lambda$, diverges as
$\epsilon^{-\nu}$ where $ \epsilon=E-E_1 $. The
number of domains is not fixed. Close to the depinning
threshold the number of virgin particles vanishes, and
the whole system is tiled by strained domains. This
global condition implies that the first moment of $n(l)$
is a slowly varying function of $\epsilon$ that does not
become singular at criticality, leading to $ {\sigma}=2 $
for $\tau>1$, while for $\tau<1$, $\sigma=1-\tau $.
The exponents of the random network model
are different: $ \tau=3/2$, $\sigma=2$, $\nu=2$
\cite{W87}.
We obtained $n(\ell)$ for the random substrate model by
solving Eq (1) numerically for an ensemble of 40
random realizations of chains of $10^4$ particles. Fig.
4 shows the rescaled histograms. The collapse of the
data for different values of $\epsilon$ was achieved by
rescaling using $\nu = 2 $ and $\tau=0.9 \pm 0.1$.
The analysis of higher moments of $ n(\ell) $ does not
show significant deviations from single parameter
scaling. The values of critical exponents are
siginificantly different than those of the random network
model, confirming that the two models belong to
distinct universality classes.
\section{The Phase Diagram}
The properties of the random substrate model depend on
the network's elastic stiffness. Intuition suggests that
when the network becomes very rigid, the differences
between the two systems should vanish.
Fig. 5 represents a ($\kappa\; E$) phase diagram,
obtained from numerical studies of chains of N=100.
As the stiffness increases, the random
substrate threshold $E_1$ approaches $E_c$. The
second threshold $E_2$ is of the order of $V_{max}$, the
upper bound of the pinning strength distribution, and depends only
weakly on $\kappa$.
Within the pinned regimes of both models, in region I of
the phase diagram, configurations of strained domains of
both models are roughly the same. The dark circles
denotes values of $\kappa, E$ where the relative
Hamming distance $D$ \cite{HAM} between locations
of strained domains for pairs of systems from each
model with the same set of pinning strengths $\{V_{i}\}$,
is smaller than 0.06.
For the random network model the largest displacement
${\xi_{max}}(\epsilon)$ is bounded by
$\lambda(\epsilon)^{3/2}\sim \epsilon^{-3}$.
The scaling of the random substrate model with stiffness
given by Eq (3), implies that for large
$\kappa > \xi_{max} $ the two models track.
This is a sufficient condition , so it yields a lower bound
for the boundary of region I where the
two models coincide, given by values of $\kappa,E$:
\begin{equation}
E_c-E \sim {1\over \kappa^{1/3}}
\end{equation}
An upper bound for $E(\kappa)$ is given by the relation:
$E_c-E \sim 1/\kappa$ \cite{valid}.
The corresponding boundary is also shown in Fig. 5.
The fact that for stiff networks the two models are roughly
equivalent, means that aspects of friction between very stiff
solid bodies
which are related to inhomogeneities on the interfaces can be
described by the random network model, or by models
based on it, elaborated to better conform to real solids.
|
\section{Introduction}
Purely collisional systems
were among the first to
be studied by molecular-dynamics simulation\cite{Alder57}.
These systems include hard spheres or
hard ellipsoids, which undergo purely elastic collisions,
and square-well fluids, in
which an attractive impulse force at
a particular interparticle position is present in addition to the
hard-core interactions. The algorithms
for such systems are exact, to within roundoff error, and consist of
free particle motion punctuated by exact resolution of the impulsive
collisions---the resulting phase space trajectory is discontinuous.
On the other hand, the vast majority of current molecular-dynamics
simulations are performed on systems with continuous potentials.
For such systems, the trajectory must be
approximated using a numerical timestepping scheme
such as the popular Verlet algorithm\cite{Verlet67}.
There
exist, however, systems that are neither purely collisional,
nor continuous, but are hybrids of the two.
Important examples of such systems
are the restricted primitive model (RPM) for electrolyte solutions and
dipolar hard spheres\cite{Hansen86}. In addition, the use of hard-core
potentials
with attractive continuous tails is common in perturbative treatments
of liquids\cite{Weeks71}. Since the the algorithms
for simulating impulsive and continuous systems are fundamentally
different from one another, the construction of hybrid methods for
mixed systems is non-trivial and little studied. Consequently, the
vast majority of studies on such systems have utilized
Monte Carlo simulation techniques, eliminating the possibility of
obtaining dynamical information.
In this paper, we present a new method for
mixed hard-core/continuous potentials,
which we call the Impulsive Verlet algorithm.
This algorithm is suitable for any continuous
potential, is less likely than alternatives to miss collisions,
and exhibits good stability and
energy conservation in long time simulation. In the construction
of this new method we have been guided by recent work in the use
of Hamiltonian splitting methods for the development of efficient and
stable molecular-dynamics algorithms\cite{Tuckerman92,Sanz-Serna95}.
A few {\em ad hoc} hybrid methods have been constructed
for mixed (hard/soft) systems\cite{Stratt81,McNeil82,Heyes82,Suh90}.
All of these methods are rather similar, in that the
particles are advanced according to the continuous forces by a time
step using a standard algorithm for continuous potentials, usually some
variant of the Verlet algorithm, and the trajectories are checked for
the existence of particle overlaps at the end or during the step.
If no overlaps occur, the procedure is repeated for the next step. If
overlaps (collisions) do occur, the system is returned to its state
before the step and then is advanced without momentum
modification by the forces to the time of collision, and the
momenta are then modified according to the rules of elastic collision. This
process
is repeated until all collisions have been resolved
and the end of the time step is reached. (One major difference between
the algorithms is whether overlaps are checked only at the end of
each step, or throughout the step. In the former case\cite{McNeil82}, it
is possible that glancing collisions are missed during the dynamics.)
Heyes\cite{Heyes82} and Suh, {\it et al.}\cite{Suh90} apply such algorithms
to the restricted primitive model for electrolytes (hard-sphere with
embedded charges in a dielectric continuum) with some apparent
success. Unfortunately, as with the other papers on algorithms
for mixed systems, no quantitative discussion on the stability or
accuracy of the algorithm is given, making it difficult to evaluate
the quality of the methods.
The Impulsive Verlet method is developed in the next two sections,
followed by a discussion of certain
numerical experiments on two model systems,
comparing our scheme with the algorithm used in
Suh,{\em et al.}\cite{Suh90}.
\section{Splitting Methods for Mixed Dynamics}
Consider a system of $N$ particles with instantaneous positions
$\mbox{\boldmath$q$} = (q_1,q_2,...,q_N)$ in $d$ dimensions
interacting according to
a continuous potential $V_c(\{{\bf q_i}\})$, assumed for simplicity
to be spherically symmetric and pairwise additive, that is,
\begin{equation}
V_{\rm c}({\bf q}) = \sum_{i=1}^{N} \sum_{j > i} \phi_{\rm c}(q_{ij}) \; ,
\end{equation}
where $q_{ij} \equiv \, \mid{\bf q_j} - {\bf q_i}\mid$, and $\phi$ is
any smooth function of one variable.
In addition, suppose the particles to have a hard core of diameter
$\sigma$; that is, when
the distance between two particles is $\sigma$ an elastic collision
occurs that reflects the momentum of
each particle along the collision vector. Such a hard-sphere core
can be represented formally by a discontinuous pair potential of the form
\begin{equation}
\phi_{\rm hs}(q_{ij}) = \left\{\begin{array}{cc} \infty & q_{ij} \le \sigma, \\
0, & q_{ij} > \sigma \;.
\end{array}
\right.
\label{hs_pot}
\end{equation}
We will define the energy function of the mixed system by analogy with
continuous dynamics as the sum of the kinetic and formal potential energies:
\begin{equation} H(\mbox{\boldmath$q$}, \mbox{\boldmath$p$}) = T(\mbox{\boldmath$p$}) +
V_{\rm hs}(\mbox{\boldmath$q$}) + V_{\rm c}(\mbox{\boldmath$q$}),
\label{hs_Hamil}
\end{equation}
where,
\[
V_{\rm hs} = \sum_{i=1}^N\sum_{j>i} \phi_{\rm hs}(q_{ij})
\]
and
\begin{equation}T(\mbox{\boldmath$p$}) = \frac{1}{2}\mbox{\boldmath$p$}^T \mbox{\boldmath$M$}^{-1}
\mbox{\boldmath$p$},\label{kinetic} \end{equation}
is the kinetic energy ($\mbox{\boldmath$M$}$ is the mass matrix)
and $\mbox{\boldmath$p$}=(p_1,p_2,...,p_N)$, where each $p_i$
is a $d$-dimensional vector. Despite appearances,
this energy function is not, properly speaking,
a Hamiltonian. Nonetheless, we can view the dynamics
of the hard-sphere fluid as the limiting dynamics in repulsive inverse-power
potentials
of the form $V_{\rm sw}(r) = 1/r^\beta$, with $\beta$ a large positive integer.
In this sense and for the purpose of constructing numerical methods,
we can interpret the formal energy $H$ as representing
a very hard repulsive inverse-power Hamiltonian. We will often
refer to $H$ as
the {\em pseudo-Hamiltonian}.
We define the flow map as the generator of the phase
space trajectory,
\begin{equation}
\left(\begin{array}{c} {\bf q}(\tau+t) \\ {\bf p}(\tau+t) \end{array} \right) =
\psi_{t,H}
\left(\begin{array}{c} {\bf q}(\tau) \\ {\bf p}(\tau) \end{array} \right) \; .
\end{equation}
The family of flow maps is closed under composition,
\begin{equation}
\psi_{t_1,H} \circ \psi_{t_2,H} = \psi_{t_2,H} \circ \psi_{t_1,H} = \psi_{t_1 + t_2,H},
\label{concatenation}
\end{equation}
for any times $t_1$ and $t_2$.
A continuous Hamiltonian system
can often be split into integrable
subproblems with Hamiltonians $H_1$ and $H_2$\cite{Sanz-Serna95}:
\begin{equation} H(\mbox{\boldmath$q$}, \mbox{\boldmath$p$}) = H_1(\mbox{\boldmath$p$}) + H_2(\mbox{\boldmath$q$})\; .
\label{splitHamilt}
\end{equation}
The flow map of the full Hamiltonian can then be approximated as the
concatenation of flow maps for the subproblems. There are a variety of
ways of doing this, but the most common is based on a Trotter factorization
\begin{equation}
\psi_{h,H} = \psi_{\frac{h}{2},H_2} \circ \psi_{h,H_1} \circ \psi_{\frac{h}{2},H_2}
+ {\cal O}(h^3)\; ,
\label{trotter}
\end {equation}
where $h$ is the time step.
For a separable Hamiltonian such as Eq.~\ref{hs_Hamil}
with $V_{\rm hs} = 0$, this factorization reduces to the usual velocity-Verlet
algorithm\cite{Swope82} when $H_1 = T({\bf p})$ and $H_2 = V_{\rm c}({\bf q})$.
The splitting framework for continuous Hamiltonians
suggests a means of constructing integrators
for mixed impulsive/continuous systems. A natural
splitting for the pseudo-Hamiltonian is
to let $H_1 = T({\bf p}) + V_{\rm hs}({\bf q})$ and $H_2 = V_{\rm c}({\bf q})$. (Note that in
this case $H_1$ is a function of both ${\bf p}$ and ${\bf q}$, but since this
represents a system with free particle motion punctuated by elastic
collisions, it is exactly integrable.)
This gives
\begin{equation}
\left(\begin{array}{c} {\bf q}^{n+1} \\ {\bf p}^{n+1} \end{array} \right) =
\psi_{\frac{h}{2},V_c} \circ \psi_{h,T + V_{hs}} \circ \psi_{\frac{h}{2},V_c}
\left(\begin{array}{c} {\bf q}^n \\ {\bf p}^n \end{array} \right) \; ,
\end{equation}
where ${\bf q}^n$ and ${\bf p}^n$ are the approximations to the phase
space variables after the $n$-th time step.
In other words, the momenta are adjusted at the beginning of each time
step by one-half step according to the continuous forces (``kick'').
The positions
are next advanced for one time step, resolving all elastic collisions, but
without further momentum modification by the continuous forces (``push'').
At the
end of the step, the momenta are advanced again by a half step using
the forces calculated from the new positions (another ``kick'').
This is nearly identical to the
algorithm of Suh, {\em et al.}\cite{Suh90} except that there
momenta are only defined at half steps and a
leap-frog formulation is used:
\begin{equation}
\left(\begin{array}{c} {\bf q}^{n+1} \\ {\bf p}^{n+1/2} \end{array}
\right)_{\mbox{Suh}} =
\psi_{h,T + V_{hs}} \circ \psi_{h,V_c}
\left(\begin{array}{c} {\bf q}^n \\ {\bf p}^{n-1/2} \end{array} \right) \;.
\end{equation}
Viewing the hard-sphere potential as being approximated by a very hard
inverse-power repulsive potential, we see that either of the above two
splitting methods is
symmetric (i.e. time-reversible). From Eq.~\ref{trotter}
we naively expect that such a
method (applied to the inverse-power potential approximation)
is second order accurate, meaning that in one step a {\em local error}
of size $O(h^3)$ is introduced; on a finite fixed time interval, these
errors accumulate, but the total growth
or {\em global error} is at most $O(h^2)$.
However,
the demonstration of third-order local error requires a $C^3$ solution, and
this assumption will break down in the limit of hard-sphere dynamics,
in particular during a collision step. In fact,
the local error introduced during
a collision is really $O(h)$.
We illustrate this point with the simple example of a nonlinear
``impact oscillator'' with one degree-of-freedom pseudo-Hamiltonian
\begin{equation}
H = \frac{p^2}{2} + \phi_{\rm hs}(q) + \phi_{\rm c}(q),
\label{h1deg}
\end{equation}
describing a point mass acted on by some
potential $\phi_c$ in collisional dynamics
with a hard wall at $q=\delta$. The particle moves in the continuous
potential $\phi_{\rm c}$ according to Newton's equations, until an impact,
when $q=\delta$, then the momentum changes sign and the motion continues from
the impact point.
Consider a numerical step
from the point $(q_0,p_0)$ at time $t=0$
for a timestep of size $h$ during which the
particle motion includes a single collision event. (We mostly
use subscripts to index particle number and superscripts for timestep,
but for the discussion that follows we need to indicate powers of the momenta;
so for this one-particle model, we will use subscripts for the timestep index.)
We need to compute the local energy error contribution during a
collisional step for this single degree-of-freedom model problem.
The sequence of computations is
\begin{eqnarray}
\hat{p} & = & p_0 - \frac{h}{2} \phi_{\rm c}'(q_0), \label{jump1}\\
h_{\#} & = & -\frac{q_0-\delta}{\hat{p}}, \label{jump2}\\
h_{\flat} & = & h-h_{\#}, \label{jump3}\\
\tilde{p} & = & - \hat{p}, \label{jump4}\\
q_1 & = & \delta - h_{\flat} \hat{p}, \label{jump5}\\
p_1 & = & \tilde{p} - \frac{h}{2} \phi_{\rm c}'(q_1)\; , \label{jump6}
\end{eqnarray}
where $h_{\#}$, and $h_{\flat}$ are the time to the next collision and
the time from that collision to the end of the time step, respectively.
Substituting the endpoint values into the energy relation, we quickly
find
\begin{eqnarray*}
\Delta H = H(q_1,p_1)-H(q_0,p_0) & = &\frac{1}{2}(\tilde{p} - \frac{h}{2} \phi_{\rm c}'(q_1)
)^2
+ \phi_{\rm c}(q_1) - \frac{1}{2} p_0^2 - \phi_{\rm c}(q_0),\\
& = & \frac{1}{2}(-p_0 + \frac{h}{2} \phi_{\rm c}'(q_0) - \frac{h}{2}
\phi_{\rm c}'(q_1))^2\\
& & \hspace{0.3in}
+ \phi_{\rm c}(q_1) - \frac{1}{2} p_0^2 - \phi_{\rm c}(q_0).
\end{eqnarray*}
Expand $\phi_{\rm c}$ in a Taylor series about $q=\delta$, substitute, and
cancel like terms to obtain
\begin{eqnarray*}
\Delta H & = & - \frac{h}{2} p_0 (\phi_{\rm c}'(q_0) - \phi_{\rm c}'(q_1))
+ \frac{h^2}{8} (\phi_{\rm c}'(q_0) - \phi_{\rm c}'(q_1))^2\\
& & \hspace{0.3in}
+ \phi_{\rm c}'(\delta)(q_1-q_0) + \frac{1}{2} \phi_{\rm c}''(\delta)( (q_1-\delta)^2 - (q_
0-\delta)^2) + E_h.
\end{eqnarray*}
The remainder $E_h$ contains terms of order the third power of $h$ or
higher, i.e. $|E_h/h^3|$ is bounded for all $h<1$ such that the step contains
a collision.
Indeed, $\frac{h^2}{8} (\phi_{\rm c}'(q_0) - \phi_{\rm c}'(q_1))^2$ is also of this
order, since $q_1-q_0$ is proportional to $h$. From the equations
\[
q_1 = \delta - h_{\flat} \hat{p}, \hspace{0.5in}
q_0 = \delta - h_{\#} \hat{p},
\]
and the use of a Taylor series expansion of $\phi_{\rm c}'$, we arrive after discarding
terms of order three or higher at,
\[
\Delta H = \frac{h(h_{\#}-h_{\flat})}{2} \phi_{\rm c}''(\delta) \hat{p}^2
- \phi_{\rm c}'(\delta)(h_{\flat}-h_{\#})\hat{p} + \frac{1}{2} \phi_{\rm c}''(\delta)
( h_{\flat}^2 - h_{\#}^2)\hat{p}^2 + \tilde{E}_h.
\]
with $\tilde{E}_h$ again of third order.
This finally leads to
\begin{eqnarray*}
\Delta H & = & \frac{(h_{\#}-h_{\flat})}{2} (h - (h_{\#}+h_{\flat})) \phi_{\rm c}''(\delta
) \hat{p}^2
- \phi_{\rm c}'(\delta)(h_{\flat}-h_{\#})\hat{p} + \tilde{E}_h\\
& = & - \phi_{\rm c}'(\delta)(h_{\flat}-h_{\#})\hat{p} + \tilde{E}_h.
\end{eqnarray*}
Therefore, the expected energy error introduced in this one collisional
step is
\begin{equation} \label{ejump}
H(q_1,p_1)-H(q_0,p_0) = -\phi_{\rm c}'(\delta)(h_{\flat}-h_{\#})\hat{p} + \tilde{E}_h,
\end{equation}
where the quantity $|\tilde{E}_h/h^3|$ is
bounded independent of $h$. (A similar
result would hold for the solution error.)
Technically speaking, it is incorrect to say that the energy jump in one step
is $O(h)$ since if we decrease the timestep $h$ sufficiently, there will
be no collision event within the particular step, and so the error will revert
to $O(h^3)$.
Nonetheless, in any timestepping simulation in which there are collision
events, these steps will introduce errors proportional to $h$. If we define
the {\em local approximation error} $e_{loc}$
as the maximum of magnitudes of the local
errors introduced, then $e_{loc}$
is of first order in $h$,
not third order as we would expect in the continuous case.
Since
there are, in general, a finite number of such collisions in any
finite interval, the accumulation is bounded and
the global error is also $O(h)$. The apparent contradiction of an odd-order
symmetric method is just one of several anomalies that result from the
complex transition from the smooth problem to the discontinuous limit.
In another terminology, we could say that the splitting
method undergoes an {\em order reduction} for stiff potential wells.
From this discussion and Eq.~\ref{ejump}, we expect the
naive splitting method to give rather poor energy conservation,
except in three special cases:
\begin{description}
\item[Case 1] Collisions do not occur {\em within}
timesteps but precisely {\em at} the timesteps, so third order is
recovered.
\item[Case 2] The collisions occur at precisely the middle of
a timestep, so that the first order term in the error formula vanishes
and third order local energy drift is again recovered.
\item[Case 3] Third order will be recovered if
the {\em derivative of the continuous pair potential vanishes for
two spheres in contact}.
\end{description}
To illustrate this last point, we apply the method to one degree-of-freedom
anharmonic ``impact oscillator" with a continuous potential, $\phi_c(q) =
\frac{1}{2}q^2 + \frac{1}{4}q^4.$ We show in Fig.~\ref{orderdepWall} the
maximum total energy error as a function of the time step when the wall
is placed at $q = 0.00$ and $q = -4.00$.
One can see that the naive splitting is a second order method when the derivative
at the wall vanishes .
Because it is only applicable for a relatively limited class of
potentials the naive splitting method is not a candidate for a viable general
technique, however, it does provide a good starting point for the development
of a general method, which we call the Impulsive Verlet (IV) algorithm.
\section{Impulsive Verlet}
To develop our method, we deliberately exploit two of the special cases in
the naive algorithm
for which third order can be expected, namely Cases 1 and 3 mentioned at the
end of the previous section. (Case 2, the situation that collisions
occur at the midpoint of the time interval, does not appear to be of
practical use.) We
begin by introducing an artificial splitting
of the continuous potential, $\phi_{\rm c}(q_{ij})$, into
into a short-ranged part, $\phi_1(q_{ij})$, and a
long-range part, $\phi_2(q_{ij})$, according to
\begin{equation}
\phi_{\rm c}(q_{ij}) = \phi_1(q_{ij}) + \phi_2(q_{ij})\;.
\label{splitPotential}
\end{equation}
(This decomposition is similar to that invoked in
multiple timestepping\cite{Tuckerman92,Tildesley78,Windemuth91,Skeel94}
molecular-dynamics algorithms.)
For the reasons discussed above, the long-range
(and therefore most expensive to
calculate) part of the potential is defined so that the derivative vanishes
at the hard-core separation. We define $\phi_2(q_{ij})$ as follows:
\begin{equation} \phi_2(q) = \left\{ \begin{array}{ll} P(q_1), & q < q_1,
\\
P(q), & q_1 \le q < q_2, \\
\phi_c(q), & q \ge q_2\; ,
\label{V_2}
\end{array} \right.
\end{equation}
where, $q_1$ and $q_2$ are parameters, and $ P(r) = A_o + A_1 r +A_2r^2 + A_3 r^3$
is a Hermite interpolant
introduced so that the two potentials are smooth to the order $C^1$ for any
continuous potential. From Eqns.~\ref{splitPotential}) and~\ref{V_2}),
$\phi_1(q)$ is given by
\begin{equation}
\phi_1(r_{ij}) = \left\{ \begin{array}{cr} \phi_{c}(q) - P(q_1) & q < q_1 \\
\phi_{c}(q) - P(q) &q_1 \le q < q_2
, \\
0 & q \ge q_2.
\end{array} \right.
\label{V_1}
\end{equation}
The continuity condition, $P(r_2) = \phi_{c}(r_2)$ , and the smoothness
conditions, $P^{\prime}(q_2) = \phi_c^{\prime}(q_2)$, $P^{\prime}(q_1) = 0$,
and $P^{\prime \prime}(q_1) = 0$, allow us to calculate the coefficients
of the Hermite interpolant, giving
\begin{equation} A_3 = \frac{\phi_c^{\prime}(q_2)}{6r_1(q_1-q_2)+3(q_2^2-q_1^2)}
\; \mbox{for $q_1 \neq q_2$,} \label{A3}
\end{equation}
\begin{equation} A_2 = -3 q_1 A_3, \label{A2} \end{equation}
\begin{equation} A_1 = 3 q_1^2 A_3, \label{A1} \end{equation}
\begin{equation} A_0 = -(A_1 q_2 + A_2 q_2^2 + A_3q_2^3) + V_{c}(q_2)\;.
\label{A0}
\end{equation}
(An example of this potential splitting for an inverse-sixth-power attractive
potential, $\phi_c(q)=-\epsilon(\sigma/q)^6$, is shown in
Fig.~\ref{splitPotInv6_1.1_1.2}.)
Next, we define $N$-body potentials $V_1$ and $V_2$ as a sum of pair contributions from $\phi_1$ and $\phi_2$, respectively.
We then split the total Hamiltonian in the following way:
\begin{equation} H_1(\mbox{\boldmath$q$} ,\mbox{\boldmath$p$}) = T(\mbox{\boldmath$p$}) +
V_{\rm hs}(\mbox{\boldmath$q$}) + V_1(\mbox{\boldmath$q$}) \label{H1exp}
\end{equation}
and
\begin{equation} H_2(\mbox{\boldmath$q$}) = V_2(\mbox{\boldmath$q$})\;.
\label{H2exp}
\end{equation}
The Trotter factorization (Eq.~\ref{trotter}) is now applied to this
splitting. The problem now is that $H_2$ is not integrable and its flow
map must be approximated. This is done is the following way:
\begin{equation}
\psi_{H_2,h} \approx \prod_{i=1}^{n_c+1} \psi_{V_{\rm hs}}\circ \psi_{V_1,\tau_i^{(c)}/2}
\circ \psi_{T,\tau_i^{(c)}} \circ \psi_{V_1,\tau_i^{(c)}/2} \;,
\end{equation}
where $n_c$ is the number of hard-sphere collisions between during the time
step $h$, $\tau_i^{(c)}$ is the time between each collision (with $\tau_1^{(c)}$ being
measured from the beginning of the time step until the first collision
and $\tau_{n+1}^{(c)}$ measured from the last collision to the end of the time
step), and $\psi_{V_{\rm hs}}$ is an
operator representing the resolution of each elastic collisions. This is
essentially the
execution of a Verlet step of length $\tau^{(c)}$ between each elastic collision.
The collision times can be calculated since the Verlet step generates a quadratic
trajectory, which together with the collision condition for two particles $i$ and $j$
can be written as
\begin{equation}
\|\mbox{\boldmath$q$}_i(\tau^{(c)}) - \mbox{\boldmath$q$}_j(\tau^{(c)})\|^2 -
\sigma^2 = 0,\label{coll_cond} \;
\end{equation}
generates a quartic equation for $\tau^{(c)}$.
We describe below the algorithm for the Impulsive Verlet molecular-dynamics
simulation in more detail.
\vspace{0.2in}
\begin{center}
\fbox{
\begin{minipage}[t]{4.5in}
\footnotesize
\begin{center}
\underline{Impulsive Verlet Timestepping Algorithm}
\end{center}
\noindent
\begin{itemize}
\item[] ${\bf p}_i^{n+1/2,0} = {\bf p}^{n,0} + \frac{1}{2}
{\bf F}_{2,i}({\bf q}^{n,0}) h$
\item[] do $i_c = 1,n_c$
\begin{itemize}
\item[] ${\bf p}_i^{n+1/2,i-1/2} = {\bf p}^{n+1/2,i-1} + \frac{1}{2}
{\bf F}_{1,i}({\bf q}^{n,i-1}) \tau_c^{i}$
\item[] ${\bf q}^{n,i} = {\bf q}^{n,i-1} + {\bf M}^{-1}
{\bf p}^{n+1/2,i_c} \tau_c^{i}$
\item[]$\tilde{\bf p}_i^{n+1/2,i} = {\bf p}^{n+1/2,i-1/2} +
\frac{1}{2} {\bf F}_{1,i}({\bf q}^{n,i}) \tau_c^{i}$
\item[]${\bf p}^{n+1/2,i} = \psi_{V_hs} \left( \begin{array}{c}
{\bf q}^{n,i_c}\\ \tilde{\bf p}^{n+1/2,i} \end{array} \right )$
\end{itemize}
\item[] end do
\item[]${\bf p}_i^{n+1/2,n_c+1/2} = {\bf p}^{n+1/2,n_c} + \frac{1}{2}
{\bf F}_{1,i}({\bf q}^{n,n_c}) (h - \sum_{i=1}^{n_c} \tau_c^{i})$
\item[] ${\bf q}^{n+1,0} = {\bf q}^{n,n_c} + {\bf M}^{-1}
{\bf p}^{n+1/2,n_c+1/2} (h - \sum_{i=1}^{n_c} \tau_c^{i})$
\item[]${\bf p}_i^{n+1/2,n_c+1} = {\bf p}^{n+1/2,n_c+1/2} + \frac{1}{2}
{\bf F}_{1,i}({\bf q}^{n+1,0}) (h - \sum_{i=1}^{n_c} \tau_c^{i})$
\item[]${\bf p}_i^{n+1,0} = {\bf p}^{n+1/2,n_c} + \frac{1}{2}
{\bf F}_{2,i}({\bf q}^{n+1,0}) h$
\end{itemize}
\end{minipage}
}
\end{center}
\vspace{0.2in}
To make sure that no collisions are missed it is necessary to ensure
that the quartic equation (Eq.~\ref{coll_cond}) is accurately solved to give the
nearest root to zero. This is not a trivial problem as the solution
becomes increasingly unstable as smaller time steps are used (i.e. when the
time to collision is small). To ensure the inaccuracies are not large
enough to affect the overall accuracy and order of the method, we employ
Laguerre's method\cite{NumRec} to find all roots of the quartic and take
the smallest, positive real root, which is then refined using Newton-Raphson.
This proved to be sufficient at all but the very smallest time steps studied.
There is a small probability that the Impulsive Verlet method can miss a
grazing collision, since the trajectories that are followed in
determining collisions are quadratic approximations. However, this probability
is greatly reduced in comparison to the method of Suh, {\it et al.} or any other algorithm
that uses linear motion to determine the collisions.
\section{Numerical Experiments}
We test the Impulsive Verlet algorithm using as our continuous potentials,
$\phi_c(q)$, the Lennard-Jones potential
\begin{equation}
\phi_{c,LJ} = 4\epsilon \left [\left (\frac{\sigma}{q} \right )^{12} -
\left (\frac{\sigma}{q} \right )^6 \right ] \; .
\end{equation}
and an attractive inverse-sixth-power potential
\begin{equation}
\phi_{c,6} = -\epsilon \left (\frac{\sigma}{q} \right )^6 \; .
\end{equation}
In both potentials $\sigma$ is the same as the hard-core diameter.
We truncate both potentials at the distance $q^*_c = q_c/\sigma = 2.5$ and,
to ensure their continuity, they are shifted so that the value of the
potential at the cutoff is zero.
In implementing the Impulsive Verlet algorithm,
we split each potential as prescribed in Eq.~\ref{splitPotential}-~\ref{A0},
with $q_1$ and $q_2$ as input parameters.
For the Lennard-Jones potential there is, of course, a natural splitting,
namely that of Weeks, Chander and Anderson (WCA)\cite{Weeks71}, where the
potential is split at the minimum with $q_1^* = q_2^* = 2^{1/6}$, which gives the
following splitting:
\begin{equation} \phi_{1,LJ}(q;\mbox{WCA}) = \left\{ \begin{array}{cc} 4\epsilon
[ (\frac{\sigma}
{q})^{-12} - (\frac{\sigma}{q})^{-6}]
+ \epsilon & q < 2^{\frac{1}{6}}\sigma, \\
0, &q \ge 2^{\frac{1}{6}}\sigma.
\end{array} \right.
\label{V_1LJ}
\end{equation}
\begin{equation} \phi_{2,\mbox{LJ}}(q,\mbox{WCA}) = \left\{ \begin{array}{cc} -\epsilon, &
q< 2^{\frac{1}{6}}\sigma , \\
4\epsilon [ (\frac{\sigma}{q})^{-12} - (\frac{\sigma}{q})^{-6}] , & q \ge 2^{\frac{1}{6}}\sigma,
\end{array} \right.
\label{V_2LJ}
\end{equation}
The MD simulations were carried out on systems of 108 particles.
The system of reduced units was
chosen so that all quantities are dimensionless. So, as units of distance and
energy we used the potential parameters $\sigma$ and $\epsilon$, respectively,
and the mass of one atom as the unit mass. The unit of time is
$(m \sigma^2/\epsilon)^{1/2}$. An asterisk superscript indicates reduced units.
Except were otherwise indicated all simulations are performed using a reduced density
$\rho^{\ast} = \rho \sigma^3 = 0.9$ and reduced temperature $T^{\ast} = kT/\epsilon = 2.5$.
In addition, a cubic box with periodic boundary conditions is used.
For greater efficiency, the MD program incorporates three neighbor lists \cite{Allen87} for the evaluation
of the short-range force, the long-range force, and the collision times.
The results of the Impulsive Verlet on the instantaneous total energy for the
Lennard-Jones and the attractive inverse sixth continuous potentials are
illustrated in Fig.~\ref{LJ_ImpVlet_Suh_4E-3} and \ref{Inv6_ImpVlet_Suh_4E-3}.
A comparison to the naive splitting algorithm of
Suh, {\it et al.}\cite{Suh90} is also made for both potentials.
The superiority in energy conservation and stability of
the Impulsive Verlet algorithm over the naive splitting method is
striking.
We study in Fig.~\ref{ljOrder} the order of the method
while varying $q_1$ and $q_2$. The order is
obtained by plotting (on a log-log) the maximum energy error for a
fixed-length simulation versus the time step. A comparison with a straight
line of slope two tells us that the method is of second order for
various values of $q_1^*$ and $q_2^*$. (Note the slight deviation of the slope at
very small time steps from the theoretical value of 2.0 is due to the
difficulty in solving the quartic equation for the collision times when
the time to collision is very small. This is not a real problem in practice
since the goal of molecular-dynamics simulation is to use the largest
time steps possible.)
Finally, to demonstrate the ability of the Impulsive-Verlet method
to yield relevant dynamical quantities, we show in Fig.~\ref{vAuto} the result for the normalized velocity
autocorrelation function ,
$ C(t) = \langle {\bf v}(t)\cdot{\bf v}(0)\rangle / \langle {\bf v}(0)\cdot{\bf v}(0)\rangle$,
for the Lennard-Jones system (108 particles) with $\rho^* = 0.9$ and $T^* = 0.9$. In this calculation
we use a splitting with $q_1^* = 1.122$ and $q_2^* = 1.5$.
\section{Conclusion}
We have introduced a molecular-dynamics method for mixed hard-core/continuous
potentials, which we refer to as the Impulsive Verlet algorithm.
This algorithm is produced by extending general potential splitting methods
to the specific case of mixed potentials. In addition to providing a
mechanism for generating the Impulsive Verlet method, the potential
splitting formalism helps to understand the failings of previous methods.
The Impulsive Verlet algorithm uses a
quadratic trajectory between collisions and does not miss any collisions of
the approximate trajectory.
As a result the algorithm is suitable for any type of continuous potential, is
second order, has good energy preservation, and is far more
stable over long time simulation than previously integrators for such systems.
(A detailed theoretical analysis of the algorithm is the subject of current research.)
\section{Acknowledgements}
The authors were supported in this work by NSF Grant DMS-9627330. In addition,
the simulations reported herein were performed on computers provided by
the Kansas Institute for Theory and Computational
Science (KITCS) and the Kansas Center for Advanced Scientific Computing
(KCASC).
The authors thank Steve Bond for helpful discussions.
|
\section{INTRODUCTION}
It is clearer every day the contribution that first-principles calculations
are making to several fields in physics, chemistry, and recently
geology and biology. The steady increase in computer power and the progress
in methodology have allowed the study of increasingly more complex and
larger systems \cite{rmp}. It has been only recently that the scaling
of the computation expense with the system size has become an important issue
in the field. Even efficient methods, like those based on
Density-Functional theory (DFT), scale like $N^{2\mbox{-}3}$, being $N$ the
number of atoms in the simulation cell. \cite{rmp} This problem stimulated the
the first ideas for methods which scale linearly with system size
\cite{Ordejon95}, a field that has been the subject of important efforts
ever since \cite{review}.
The key for achieving linear scaling is the explicit use of locality,
meaning by it the insensitivity of the properties of a region of the
system to perturbations sufficiently far away from it \cite{Kohn96}.
A local language will thus be needed for the two different problems
one has to deal with in a DFT-like method: building the self-consistent
Hamiltonian, and solving it. Most of the initial effort was dedicated
to the latter \cite{Ordejon95,review} using empirical or semi-empirical
Hamiltonians. The {\sc Siesta} project \cite{rc,mrs,ijqc}
started in 1995 to address the former. Atomic-orbital basis sets
were chosen as the local language, allowing for arbitrary basis sizes,
what resulted in a general-purpose, flexible linear-scaling DFT program
\cite{ijqc,prbprep}. A parallel effort has been
the search for orbital bases that would meet the standards of
precision of conventional first-principles calculations, but keeping
as small a range as possible for maximum efficiency.
Several techniques are presented here.
Other approaches pursued by other groups are also
shortly reviewed in section II. All of them are based on local bases
with different flavors, offering a fair variety of choice between
systematicity and efficiency. Our developments of atomic bases for
linear-scaling are presented in section III.
{\sc Siesta} has been applied to quite varied systems during these years,
ranging from metal nanostructures to biomolecules. Some of the
results obtained are briefly reviewed in section IV.
\section{METHOD AND CONTEXT}
{\sc Siesta} is based on
DFT, using local-density \cite{rmp} and generalized-gradients functionals
\cite{pbe}, including spin polarization, collinear and non-collinear
\cite{noncol}.
Core electrons are replaced by norm-conserving pseudopotentials \cite{tm2}
factorized in the Kleinman-Bylander form \cite{kb}, including
scalar-relativistic effects, and non-linear partial-core corrections
\cite{pcec}. The one-particle problem is then solved using linear combination
of atomic orbitals (LCAO). There are no constraints either on the
radial shape of these orbitals (numerically treated), or
on the size of the basis, allowing for the full
quantum-chemistry know-how \cite{huzinaga} (multiple-$\zeta$,
polarization, off-site, contracted, and diffuse orbitals).
Forces on the atoms and the stress tensor are obtained from the
Hellmann-Feynman theorem with Pulay corrections \cite{rc}, and are used for
structure relaxations or molecular dynamics simulations of different
types.
Firstly, given a Hamiltonian, the
one-particle Schr\"odinger equation is solved yielding the energy
and density matrix for the ground state. This task is performed either
by diagonalization (cube-scaling, appropriate for systems under a hundred
atoms or for metals) or with a linear-scaling algorithm.
These have been extensively reviewed elsewhere \cite{review}.
{\sc Siesta} implements two $O(N)$ algorithms
\cite{Ordejon95,kim} based on localized Wannier-like
wavefunctions.
Secondly, given the density matrix, a new Hamiltonian matrix is
obtained. There are different ways proposed in the literature to perform
this calculation in order-$N$ operations.
$(i)$ Quantum chemists have explored algorithms for Gaussian-type orbitals
(GTO) and related technology \cite{huzinaga}. The long-range Hartree potential
posed an important problem that has been overcome with Fast Multipole
Expansion techniques plus near-field corrections \cite{head}.
Within this approach, periodic boundary conditions for extended systems
require additional techniques that are under current development \cite{pbcscu}.
$(ii)$ Among physicists tradition favors
more systematic basis sets, such as plane-waves and variations thereof.
Working directly on a real-space grid was early proposed as
a natural possibility for linear scaling \cite{Hernandez95}.
Multigrid techniques allow efficient
treatment of the Hartree problem, making it very attractive. However,
a large prefactor was found \cite{Hernandez95} for the linear
scaling, making the order-$N$ calculations along this line not so practical
for the moment. The introduction of a basis of localized functions on the
points of the grid (blips) was then proposed as an operative method within
the original spirit \cite{Hernandez97}. It is probably more expensive
than LCAO alternatives, but with the advantage of a systematic basis.
Another approach \cite{haynes} works with spherical Bessel
functions confined to (overlapping) spheres wisely located within
the simulation cell. As for plane-waves, a kinetic energy cutoff
defines the quality of the basis within one sphere. The number,
positioning, and radii of the spheres are new variables to consider,
but the basis is still more systematic than within LCAO.
$(iii)$ There are mixed schemes that use atomic-orbital bases but evaluate
the matrix elements using plane-wave or real-space-grid techniques.
The method of Lippert {\it et al.} \cite{hutter1} uses GTO's and
associated techniques for the computation of the matrix
elements of some terms of the Kohn-Sham Hamiltonian. It uses
plane-wave representations of the density for the calculation of the
remaining terms. This latter method is conceptually very similar
to the one presented earlier by Ordej\'on {\it et al.} \cite{rc},
on which {\sc Siesta} is based.
The matrix elements within {\sc Siesta} are also calculated in two
different ways \cite{ijqc}: some Hamiltonian terms in a real-space grid
and other terms (involving two-center integration) by very efficient,
direct LCAO integration \cite{sankey}. While {\sc Siesta} uses numerical
orbitals, Lippert's method works with GTOs, which allow analytic
integrations, but require more orbitals.
Except for the quantum-chemical approaches, the methods mentioned
require smooth densities, and thus soft pseudopotentials. A recent
augmentation proposal \cite{hutter2} allows a substantial improvement in
grid convergence of the method of Lippert {\it et al.} \cite{hutter1},
possibly allowing for all-electron calculations.
\section{ATOMIC ORBITALS ADAPTED TO LINEAR SCALING}
The main advantage of atomic orbitals is their efficiency (fewer orbitals
needed per electron for similar precision)
and their main disadvantage is the lack of systematics for optimal
convergence, an issue that quantum chemists have been working on for
many years \cite{huzinaga}. They have also clearly shown that there
is no limitation on precision intrinsic to LCAO.
{\it Orbital range.}
The need for locality in linear-scaling algorithms imposes
a finite range for matrix elements, which has a strong influence on the
efficiency of the method. There is a clear challenge ahead for finding
short-range bases that still give a high precision.
The traditional way is to neglect matrix elements
between far-away orbitals with values below a tolerance.
This procedure implies a departure from the original Hilbert space
and it is numerically unstable for short ranges. Instead, the use
of orbitals that would strictly vanish beyond a certain radius was proposed
\cite{sankey}. This gives sparse matrices consistently within the
Hilbert space spanned by the basis, numerically robust even for small
ranges.
In the context of {\sc Siesta}, the use of pseudopotentials imposes
basis orbitals adapted to them. Pseudoatomic orbitals (PAOs) are
used, i.e., the DFT solution of the atom with the pseudopotential.
PAO's confined by a spherical infinite-potential wall \cite{sankey},
has been the starting point for our bases.
Fig.~1 shows $s$ and $p$ confined PAOs for oxygen.
Smoother confining potentials have been proposed as a better converging
alternative \cite{horsfield}.
A single parameter that defines the confinement radii of different
orbitals is the orbital {\it energy shift} \cite{jose},
$\Delta E_{\small \rm PAO}$, i.e., the energy increase
that each orbital experiences when confined to a finite sphere. It
defines all radii in a well balanced way, and allows the systematic
convergence of physical quantities to the required precision.
Fig.~2 shows the convergence of geometry and cohesive
energy with $\Delta E_{\small \rm PAO}$ for various
systems. It varies depending on the system and physical quantity, but
$\Delta E_{\small \rm PAO} \approx 100 - 200$ meV gives
typical precisions within the accuracy of current GGA functionals.
{\it Multiple-$\zeta$.}
To generate confined multiple-$\zeta$ bases,
a first proposal \cite{projection} suggested the use of the
excited PAOs in the confined atom. It works well for short ranges,
but shows a poor convergence with $\Delta E_{\small \rm PAO}$,
since some of these orbitals are unbound in the free atom.
In the split-valence scheme, widely used in quantum chemistry,
GTOs that describe the tail of the atomic orbitals are
left free as separate orbitals for the extended basis.
Adding the quantum-chemistry \cite{huzinaga} GTOs' tails
to the PAO bases gives flexible bases, but the confinement control
with $\Delta E_{\small \rm PAO}$ is lost.
The best scheme used in {\sc Siesta} calculations so far is based
on the idea \cite{joseluis} of adding, instead of a GTO, a numerical
orbital that reproduces the tail of the PAO outside a radius
$R_{\rm DZ}$, and continues smoothly towards the origin as $r^l(a-br^2)$,
with $a$ and $b$ ensuring continuity and differenciability at $R_{\rm DZ}$.
This radius is chosen so that the norm of the tail beyond has a given
value. Variational optimization
of this {\it split norm} performed on different systems
shows a very general and stable performance for values around
15\% (except for the $\sim 50\%$ for hydrogen). Within exactly the same
Hilbert space, the second orbital can be chosen as the difference between
the smooth one and the original PAO, which gives a basis orbital strictly
confined within the matching radius $R_{\rm DZ}$, i.e., smaller than the
original PAO. This is illustrated in Fig.~1. Multiple-$\zeta$ is
obtained by repetition of this procedure.
{\it Polarization orbitals.}
A shell with angular momentum $l+1$ (or more shells with higher $l$)
is usually
added to polarize the most extended atomic valence orbitals ($l$), giving
angular freedom to the valence electrons. The (empty) $l+1$ atomic orbitals
are not necessarily a good choice, since they are typically too extended.
The normal procedure within quantum chemistry \cite{huzinaga} is using
GTOs with maximum overlap with valence orbitals.
Instead, we use for {\sc Siesta} the numerical orbitals resulting from the
actual polarization of the pseudoatom in the presence of a small electric
field \cite{jose}. The pseudoatomic problem is then exactly solved
(within DFT), yielding the $l+1$ orbitals through comparison with
first order perturbation theory. The range of the polarization
orbitals is defined by the range of the orbitals they polarize.
It is illustrated in Fig.~3 for the $d$ orbitals
of silicon.
The performance of the schemes presented here has been tested for various
applications (see below) and a systematic study will be presented elsewhere
\cite{prbprep}. It has been found in general that double-$\zeta$, singly
polarized (DZP) bases give precisions within the accuracy of GGA functionals
for geometries, energetics and elastic/vibrational properties.
{\it Other possibilities.}
Scale factors on orbitals are also used, both for orbital contraction and for
diffuse orbitals. Off-site orbitals can be introduced. They serve for
the evaluation of basis-set superposition errors
\cite{maider}. Spherical Bessel functions are also included,
that can be used for mixed bases between our approach
and the one of Haynes and Payne \cite{haynes}.
\section{BRIEF REVIEW OF APPLICATIONS}
{\it Carbon Nanostructures.}
A preliminary version of {\sc Siesta} was first applied to study the
shape of large hollow carbon fullerenes \cite{rc} up to C$_{540}$,
the results contributing to establish
that they do not tend to a spherical-shape
limit but tend to facet around the twelve corners given by the pentagons.
{\sc Siesta} has been also applied to carbon nanotubes. In a first study,
structural, elastic and vibrational properties were characterized
\cite{tubephonons}. A second work was dedicated to their
deposition on gold surfaces, and the STM images that they originate
\cite{stmprl}, specially addressing experiments on finite-length tubes.
A third study has been dedicated to the opening of single-wall nanotubes
with oxygen, and the stability of the open, oxidized tubes for intercalation
studies \cite{marioprl}.
{\it Gold Nanostructures.}
Gold nanoclusters of small sizes (up to Au$_{75}$) were found \cite{auprl}
to be amorphous, or nearly so, even for sizes for which very favorable
geometric structures had been proposed before. In a further study the
origin of this striking situation is explained in terms of local stresses
\cite{auprep}.
Chains of gold atoms have been studied addressing the experiments
\cite{auexp} which show them displaying remarkably long interatomic spacings
(4 - 5 \AA). A first study \cite{tosatti} arrives at the conclusion that a
linear gold chain would break at interatomic spacings much smaller than the
observed ones. It is illustrated in Fig.~4 \cite{auhilos}. A possible
explanation of the discrepancy is reported elsewhere.\cite{auhilos}
{\it Surfaces and Adsorption.}
A molecular dynamics simulation was performed \cite{gabriel} on the clean
surface of liquid silicon close to the melting temperature, in which
surface layering was found, i.e., density oscillations of roughly atomic
amplitude, like what was recently found to happen in the surface of other
liquid metals \cite{layering}. Unlike them, though, the origin for silicon
was found to be orientational, reminescent of directed octahedral bonding.
Adsorption studies have also been performed on solid silicon surfaces,
Ba on Si(100) \cite{basi} and C$_{60}$ on Si(111) \cite{c60si}.
Both works study adsorption geometries and energetics. For Ba, interactions
among adsorbed atoms and diffusion features are studied. For C$_{60}$,
STM images have been simulated and compared to experiments.
{\it Nucleic Acids.}
Feasibility tests on DNA were performed in the early stages of the
project, by relaxing a dry B-form poly(dC)-poly(dG) structure with a
minimal basis \cite{mrs,ijqc}. In preparation of realistic calculations,
a thorough study \cite{maider} of 30 nucleic acid pairs has been
performed addressing the precision of the approximations
and the DZP bases, and the accuracy of the GGA functional \cite{pbe},
obtaining good results even for the hydrogen bridges.
Based on that, a first study of dry A-DNA
has been performed, with a full relaxation of the structure, and an
analysis of the electronic characteristics \cite{dnaprep}.
\section{CONCLUSIONS}
The status of the {\sc Siesta} project has been briefly reviewed,
putting it in context with other methods of liner-scaling DFT, and
briefly describing results obtained with {\sc Siesta} for a variety of
systems. The efforts dedicated to finding schemes for atomic bases adapted
to linear-scaling have been also described. A promising
field still very open for future research.
\vspace{10pt}
{\it Acknowledgments.}
We are grateful for ideas, discussions, and support of
Jos\'e L. Martins, Richard M. Martin, David A. Drabold, Otto F. Sankey,
Julian D. Gale, and Volker Heine. EA is very grateful to the Ecole
Normale Sup\'erieure de Lyon for its hospitality.
PO is the recipient of a Sponsored Research Project from
Motorola PCRL. EA and PO acknowledge travel support of the $\Psi_k$
network of ESF.
This work has been supported by Spain's DGES grant PB95-0202.
|
\subsection{Notation}
\begin{eqnarray}
\essprod{p}{q}{1}{2} &\equiv& \duess{p}{1}{3} \ddess{q}{2}{3} \\
\essdot{p}{q} &\equiv& \ess{p}{1}{2} \ddess{q}{1}{2} \\
\tessprod{p}{q}{r}{t}{1}{2} &\equiv& \duess{p}{1}{3} \udess{q}{4}{3} \duess{r}{4}{5} \ddess{t}{2}{5} \\
(s_{p}. s_{q} . s_{r} . s_{t}) &\equiv& \ess{p}{2}{3} \udess{q}{4}{3} \duess{r}{4}{5} \ddess{t}{2}{5} \\
\ddess{p}{1}{1} &\equiv& s_{\mu} \sigma^{\mu}_{\grksp{1}\ifdot{\grksp{1}}} \\
s_{p\ifdot{\grksp{1}}\grksp{1}} &\equiv& s_{\mu} \bar{\sigma}^{\mu}_{\ifdot{\grksp{1}}\grksp{1}}
\end{eqnarray}
\begin{eqnarray}
\left( \ugsp{pqr}{} \dgsp{nlm}{} \right) &\equiv& \ugsp{pqr}{1} \dgsp{nlm}{1}
\\
\ugsp[i]{pqr}{1} &\equiv& \uth[i]{pq}{2} \duinvess{pq}{2}{1} - \uth[i]{qr}{2} \duinvess{qr}{2}{1}
\\
\left( \ugsp{p}{} \ess{q}{}{} \ess{r}{}{} \dgsp{t}{} \right) &\equiv& \ugsp{p}{1} \udess{q}{3}{1} \duess{r}{3}{2} \dgsp{t}{2}
\end{eqnarray}
\subsection{Useful Relations}
\begin{eqnarray}
\ugsp{pqr}{1} &=& - \ugsp{rqp}{1} \\
\ugsp{pqr}{1} &=& \ugsp{prq}{2} \udess{pr}{3}{2} \duinvess{pq}{3}{1} \\
\ugsp{pqr}{1} &\propto& \ugsp{\langle pqr \rangle}{1},
\end{eqnarray}
where $\langle pqr \rangle$ is the ordered form of $pqr$ ( e.g. $\langle 312 \rangle = 123$).
\begin{eqnarray}
\essprod{p}{q}{1}{2} &=& \left( \eta^{\mu\nu} \eps{1}{2} + \sigma^{\mu\nu}_{\grksp{1}\grksp{2}} \right) s_{p\mu} s_{q\nu} \\
\ess{p}{1}{1} &=& s_{p}^{\ifdot{\grksp{1}}\grksp{1}}
\end{eqnarray}
\beq \begin{array}{rcl}
\eqal{
\duess{12}{1}{1} \ddess{12}{2}{1} }{=}{ - \frac{1}{2} \sess{12} \scfeps{1}{2}
}
\leqal{
\udess{12}{1}{1} \ddess{12}{1}{2} }{=}{ - \frac{1}{2} \sess{12} \scfepsb{1}{2} \elab{l9spsqu},
}
\end{array} \eeq
\begin{eqnarray}
\duess{12}{1}{2} \ddess{23}{3}{2} \udess{12}{3}{1} &=& \essdot{12}{23} \ddess{12}{1}{1} - \frac{1}{2} \sess{12} \ddess{23}{1}{1}. \elab{l9spswap}
\end{eqnarray}
\begin{eqnarray}
{(s_{2} . s_{1})}_{\alpha_{2}\alpha_{1}} &=& - {(s_{1} . s_{2})}_{\alpha_{1}\alpha_{2}}
\end{eqnarray}
\begin{eqnarray}
{(s_{1} . s_{2})}_{\alpha_{2}\alpha_{1}} &=& {(s_{1} . s_{2})}_{\alpha_{1}\alpha_{2}} + (s_{1}s_{2})\, \epsilon_{\alpha_{1}\alpha_{2}}
\end{eqnarray}
\begin{eqnarray}
2 \left( s_{1}.s_{2}.s_{3}.s_{4} \right)_{\alpha_{1}\alpha_{2}} &=&
- {(s_{3} . s_{4})}_{\alpha_{1}\alpha_{2}}\,(s_{1}s_{2}) +
{(s_{2} . s_{4})}_{\alpha_{1}\alpha_{2}}\,(s_{1}s_{3}) - \nonumber \\ & &
{(s_{2} . s_{3})}_{\alpha_{1}\alpha_{2}}\,(s_{1}s_{4}) - \nonumber \\ & &
\left( {(s_{1} . s_{4})}_{\alpha_{1}\alpha_{2}} +
\epsilon_{\alpha_{1}\alpha_{2}}\,(s_{1}s_{4}) \right) \,(s_{2}s_{3}) + \nonumber \\ & &
\left( {(s_{1} . s_{3})}_{\alpha_{1}\alpha_{2}} +
\epsilon_{\alpha_{1}\alpha_{2}}\,(s_{1}s_{3}) \right) \,(s_{2}s_{4}) - \nonumber \\ & &
\left( {(s_{1} . s_{2})}_{\alpha_{1}\alpha_{2}} +
\epsilon_{\alpha_{1}\alpha_{2}}\,(s_{1}s_{2}) \right) \,(s_{3}s_{4}) - \nonumber \\ & &
\epsilon_{\alpha_{1}\alpha_{2}}\,(s_{1}.s_{2}.s_{3}.s_{4}) \elab{essfourexp}
\end{eqnarray}
\beq \begin{array}{rcl}
\eqal{
\tessprod{p}{q}{r}{t}{1}{2} }{=}{ \frac{1}{2} \essdot{p}{r} \essprod{q}{t}{1}{2} - \frac{1}{2} \essdot{p}{q} \essprod{r}{t}{1}{2}
}
\leqal{}{}{ \mbox{} - \frac{1}{2} \essdot{q}{r} \essprod{p}{t}{1}{2} + i \epsilon^{\mu\nu\rho\kappa} s_{p\mu} s_{q\nu} s_{r\rho} {\sigma_{\kappa\grksp{1}}}^{\ifdot{\grksp{1}}} \ess{t}{2}{1}
}
\end{array} \eeq
\section{#1}
\renewcommand{\thesection}{\Alph{section}}}
\newlength{\appwidth}
\settowidth{\appwidth}{Appendix A }
\makeatletter
\newcommand{\newsec}{
\renewcommand{\l@section}[2]{%
\addpenalty{\@secpenalty}%
\addvspace{1.0em \@plus\partial@}%
\setlength\@tempdima{1.5em}%
\addtolength{\@tempdima}{\appwidth}%
\begingroup
\parindent \z@ \rightskip \@pnumwidth
\parfillskip -\@pnumwidth
\leavevmode \bfseries
\advance\leftskip\@tempdima
\hskip -\leftskip
##1\nobreak\hfil \nobreak\hbox to\@pnumwidth{\hss ##2}\par
\endgroup}
}
\makeatother
\newcommand{\ensuremath{\Lambda}}{\ensuremath{\Lambda}}
\newcommand{\ensuremath{\bar{\Lambda}}}{\ensuremath{\bar{\Lambda}}}
\newcommand{\delmat}[2]{\ensuremath{{\delta_{#1}}^{#2}}}
\newcommand{\deldotmat}[2]{\ensuremath{{\delta_{\dot{#1}}}^{\dot{#2}}}}
\newcommand{\spc}[2]{\ensuremath{{w_{#1}}^{#2}}}
\newcommand{\viel}[2]{\ensuremath{{e_{#1}}^{#2}}}
\newcommand{\inviel}[2]{\ensuremath{{e^{#1}}_{#2}}}
\newcommand{\pee}[1]{\ensuremath{P_{#1}}}
\newcommand{\emm}[2]{\ensuremath{M_{#1#2}}}
\newcommand{\ing}[1]{\ensuremath{T_{#1}}}
\newcommand{\que}[2]{\ensuremath{{Q_{#1}}^{#2}}}
\newcommand{\beps}[1]{\ensuremath{\bar{\epsilon}_{#1}}}
\newcommand{\sigdd}[3]{\ensuremath{ \sigma^{#1}_{#2 \dot{#3}}}}
\newcommand{\sigdu}[3]{\ensuremath{ {\sigma^{#1}_{#2}}^{ \dot{#3}}}}
\newcommand{\sigud}[3]{\ensuremath{ {\sigma^{#1 #2}}_{\dot{#3}}}}
\newcommand{\siguu}[3]{\ensuremath{ \sigma^{#1 #2 \dot{#3}}}}
\newcommand{\dsigdd}[3]{\ensuremath{ \sigma_{#1 #2 \dot{#3}}}}
\newcommand{\dsigdu}[3]{\ensuremath{ {\sigma_{#1}_{#2}}^{ \dot{#3}}}}
\newcommand{\dsigud}[3]{\ensuremath{ {\sigma_{#1 #2}}_{\dot{#3}}}}
\newcommand{\dsiguu}[3]{\ensuremath{ \sigma_{#1 #2 \dot{#3}}}}
\newcommand{\bsigdd}[3]{\ensuremath{ \bar{\sigma}^{#1}_{#2 \dot{#3}}}}
\newcommand{\bsigdu}[3]{\ensuremath{ {\bar{\sigma}^{#1}_{#2}}^{ \dot{#3}}}}
\newcommand{\bsigud}[3]{\ensuremath{ {\bar{\sigma}^{#1 #2}}_{\dot{#3}}}}
\newcommand{\bsiguu}[3]{\ensuremath{ \bar{\sigma}^{#1 #2 \dot{#3}}}}
\newcommand{\twsigdd}[3]{\ensuremath{ \sigma^{#1}_{#2 {#3}}}}
\newcommand{\twsigdu}[3]{\ensuremath{ {\sigma^{#1}_{#2}}^{ {#3}}}}
\newcommand{\twsigud}[3]{\ensuremath{ {\sigma^{#1 #2}}_{{#3}}}}
\newcommand{\twsiguu}[3]{\ensuremath{ \sigma^{#1 #2 {#3}}}}
\newcommand{\dtwsigdd}[3]{\ensuremath{ \sigma_{#1 #2 {#3}}}}
\newcommand{\dtwsigdu}[3]{\ensuremath{ {\sigma_{#1}_{#2}}^{ {#3}}}}
\newcommand{\dtwsigud}[3]{\ensuremath{ {\sigma_{#1 #2}}_{{#3}}}}
\newcommand{\dtwsiguu}[3]{\ensuremath{ \sigma_{#1 #2 {#3}}}}
\newcommand{\btwsigdd}[3]{\ensuremath{ \bar{\sigma}^{#1}_{#2 {#3}}}}
\newcommand{\btwsigdu}[3]{\ensuremath{ {\bar{\sigma}^{#1}_{#2}}^{ {#3}}}}
\newcommand{\btwsigud}[3]{\ensuremath{ {\bar{\sigma}^{#1 #2}}_{{#3}}}}
\newcommand{\btwsiguu}[3]{\ensuremath{ \bar{\sigma}^{#1 #2 {#3}}}}
\newcommand{\bthtu}[1]{\ensuremath{ \bar{\theta}^{\dot{#1}}}}
\newcommand{\bthtd}[1]{\ensuremath{ \bar{\theta}_{\dot{#1}}}}
\newcommand{\bepsu}[1]{\ensuremath{ \bar{\epsilon}^{\dot{#1}}}}
\newcommand{\bepsd}[1]{\ensuremath{ \bar{\epsilon}_{\dot{#1}}}}
\newcommand{\bchiu}[1]{\ensuremath{ \bar{\chi}^{\dot{#1}}}}
\newcommand{\bchid}[1]{\ensuremath{ \bar{\chi}_{\dot{#1}}}}
\newcommand{\blamu}[1]{\ensuremath{ \bar{\lambda}^{\dot{#1}}}}
\newcommand{\blamd}[1]{\ensuremath{ \bar{\lambda}_{\dot{#1}}}}
\newcommand{\bup}[2]{\ensuremath{ \bar{#1}^{\dot{#2}}}}
\newcommand{\bdn}[2]{\ensuremath{ \bar{#1}_{\dot{#2}}}}
\newcommand{\bDu}[1]{\ensuremath{ \bar{D}^{\dot{#1}}}}
\newcommand{\bDd}[1]{\ensuremath{ \bar{D}_{\dot{#1}}}}
\newcommand{\bQu}[1]{\ensuremath{ \bar{Q}^{\dot{#1}}}}
\newcommand{\bQd}[1]{\ensuremath{ \bar{Q}_{\dot{#1}}}}
\newcommand{\bWu}[1]{\ensuremath{ \bar{W}^{\dot{#1}}}}
\newcommand{\bWd}[1]{\ensuremath{ \bar{W}_{\dot{#1}}}}
\newcommand{\bqu}[1]{\ensuremath{ \bar{q}^{\dot{#1}}}}
\newcommand{\bqd}[1]{\ensuremath{ \bar{q}_{\dot{#1}}}}
\newcommand{\bpd}[1]{\ensuremath{ \bar{\partial}_{\dot{#1}}}}
\newcommand{\bpu}[1]{\ensuremath{ \bar{\partial}^{\dot{#1}}}}
\newcommand{\ptd}[1]{\ensuremath{ \partial_{#1}}}
\newcommand{\ptu}[1]{\ensuremath{ \partial^{#1}}}
\newcommand{\ensuremath{ \bar{\phi}}}{\ensuremath{ \bar{\phi}}}
\newcommand{\rest}[2]{\ensuremath{ \left. #1 \right|_{#2}}}
\newcommand{\resttz}[1]{\ensuremath{ \left. #1 \right|_{\theta = 0}}}
\newcommand{\ensuremath{(1 \leftrightarrow 2)}}{\ensuremath{(1 \leftrightarrow 2)}}
\newcounter{sub}
\renewcommand{\thesub}{\alph{sub}}
\newenvironment{subfig}{
\setcounter{sub}{0}
}{}
\newcommand{\subfigitem}{
\refstepcounter{sub}
(\alph{sub})
}
\newcommand{\grkend}[1]{
\ifthenelse{\equal{#1}{1}}{\mu}{}
\ifthenelse{\equal{#1}{2}}{\nu}{}
\ifthenelse{\equal{#1}{3}}{\rho}{}
\ifthenelse{\equal{#1}{4}}{\sigma}{}
\ifthenelse{\equal{#1}{5}}{\epsilon}{}
\ifthenelse{\equal{#1}{6}}{\kappa}{}}
\newcommand{\grksp}[1]{
\ifthenelse{\equal{#1}{1}}{\alpha}{}
\ifthenelse{\equal{#1}{2}}{\beta}{}
\ifthenelse{\equal{#1}{3}}{\gamma}{}
\ifthenelse{\equal{#1}{4}}{\delta}{}
\ifthenelse{\equal{#1}{5}}{\epsilon}{}
\ifthenelse{\equal{#1}{6}}{\kappa}{}}
\newcommand{\scfm}[2]{\ensuremath{M_{\grkend{#1} \grkend{#2}} }}
\newcommand{\scfeps}[2]{\ensuremath{\epsilon_{\grksp{#1} \grksp{#2}} }}
\newcommand{\scfepsb}[2]{\ensuremath{\epsilon_{\dot{\grksp{#1}} \dot{\grksp{#2}} } }}
\newcommand{\scfe}[2]{\ensuremath{\eta_{\grkend{#1} \grkend{#2}} }}
\newcommand{\scfp}[1]{\ensuremath{P_{\grkend{#1}} }}
\newcommand{\ensuremath{D }}{\ensuremath{D }}
\newcommand{\scfk}[1]{\ensuremath{K_{\grkend{#1}} }}
\newcommand{\scfq}[1]{\ensuremath{Q_{\grksp{#1} } }}
\newcommand{\scfqb}[1]{\ensuremath{\bar{Q}_{\dot{\grksp{#1}}} }}
\newcommand{\scfs}[1]{\ensuremath{S_{\grksp{#1}} }}
\newcommand{\scfsb}[1]{\ensuremath{\bar{S}_{\dot{\grksp{#1}}} }}
\newcommand{\ensuremath{R }}{\ensuremath{R }}
\newcommand{\algc}[3]{\ensuremath{ \left[ #1, #2 \right] &=& #3 }}
\newcommand{\alga}[3]{\ensuremath{ \left\{ #1, #2 \right\} &=& #3 }}
\newcommand{\scfsig}[4]{\ensuremath{{\sigma_{\grkend{#1} \grkend{#2} \grksp{#3} }}^{ \grksp{#4} } }}
\newcommand{\scfsigb}[4]{\ensuremath{{\bar{\sigma}_{\grkend{#1} \grkend{#2} \dot{\grksp{#3}} }}^{ \dot{\grksp{#4}} } }}
\newcommand{\dscfsig}[4]{\ensuremath{\sigma^{\grkend{#1} \grkend{#2} }_{ \grksp{#3} \grksp{#4} } }}
\newcommand{\dscfsigb}[4]{\ensuremath{\bar{\sigma}^{\grkend{#1} \grkend{#2} }_{\dot{\grksp{#3} } \dot{\grksp{#4}} } }}
\newcommand{\hspace{-6 ex}}{\hspace{-6 ex}}
\newcommand{\hspace{-4500\unitlength}}{\hspace{-4500\unitlength}}
\newcommand{\beprop \beprop}{\hspace{-4500\unitlength} \hspace{-4500\unitlength}}
\newcommand{\bqprop \bqprop}{\beprop \beprop \beprop \beprop}
\newcommand{\bhprop \bhprop }{\bqprop \bqprop \bqprop \bqprop }
\newenvironment{lbrce}{\left\{}{\right\}}
\newcommand{ \begin{lbrce} }{ \begin{lbrce} }
\newcommand{\left[}{\left[}
\newcommand{\left(}{\left(}
\newcommand{\end{lbrce} }{\end{lbrce} }
\newcommand{\right]}{\right]}
\newcommand{\right)}{\right)}
\newlength{\ifdota}
\newlength{\ifdotb}
\settowidth{\ifdotb}{\mbox{$a()$}}
\newcommand{\ifdot}[1]{
\settowidth{\ifdota}{\mbox{\ensuremath{a(#1)}}}
\ifthenelse{\lengthtest{\ifdota = \ifdotb}}{ }{ \ensuremath{\dot{#1}}}
}
\newcommand{\ifbrack}[1]{
\settowidth{\ifdota}{\mbox{\ensuremath{a(#1)}}}
\ifthenelse{\lengthtest{\ifdota = \ifdotb}}{ }{ \ensuremath{(#1)}}
}
\newcommand{\swp}[2]{\ensuremath{P_{\grksp{#1} \ifdot{\grksp{#2}}} }}
\newcommand{\ensuremath{D }}{\ensuremath{D }}
\newcommand{\swk}[3][{}]{\ensuremath{K^{\ifbrack{#1}\grksp{#2}\ifdot{\grksp{#3}}} }}
\newcommand{\lbswk}[3][{}]{\ensuremath{K_{0}^{#1\grksp{#2}\ifdot{\grksp{#3}}} }}
\newcommand{\mlswk}[3]{\ensuremath{K_{#3}^{\grksp{#1}\ifdot{\grksp{#2}}} }}
\newcommand{\ensuremath{\mathbf{q} }}{\ensuremath{\mathbf{q} }}
\newcommand{\ensuremath{T }}{\ensuremath{T }}
\newcommand{\swq}[2][{}]{\ensuremath{Q_{\grksp{#2} \ifbrack{#1}} }}
\newcommand{\swqb}[2][{}]{\ensuremath{\bar{Q}^{\ifbrack{#1}}_{\ifdot{\grksp{#2}}} }}
\newcommand{\sws}[2][{}]{\ensuremath{S^{\grksp{#2} \ifbrack{#1}} }}
\newcommand{\swsb}[2][{}]{\ensuremath{\bar{S}^{\ifdot{\grksp{#2}}}_{\ifbrack{#1}} }}
\newcommand{\ensuremath{R }}{\ensuremath{R }}
\newcommand{\pdv}[2]{\ensuremath{\frac{\partial #1}{\partial #2}}}
\newcommand{\sdrv}[3]{\ensuremath{\partial_{#1\grksp{#2}\ifdot{\grksp{#3}}}}}
\newcommand{\tdrv}[3][{}]{\ensuremath{\partial^{#1}_{\grksp{#2}}}}
\newcommand{\dq}[1][{}]{\ensuremath{\delta q_{#1}}}
\newcommand{\brkess}[5]{\ensuremath{ {s_{#1}^{ \left(\grksp{#4}\right. }
}^{ \ifdot{\grksp{#3}} } s_{#2}^{ \left. \grksp{#5}\right)
\ifdot{\grksp{#3}} } }}
\newcommand{\brkessbar}[5]{\ensuremath{
s_{#1}^{ \grksp{#3}\left( \ifdot{\grksp{#4}}\right. }
{s_{#2 \grksp{#3}}}^{ \left. \ifdot{\grksp{#5}}\right) } }}
\newcommand{\ess}[3]{\ensuremath{ s_{#1}^{ \grksp{#2} \ifdot{\grksp{#3}} } }}
\newcommand{\kay}[2]{\ensuremath{ k_{ \grksp{#1} \ifdot{\grksp{#2}} } }}
\newcommand{\sess}[1]{\ensuremath{s_{#1}^{ 2} }}
\newcommand{\duess}[3]{\ensuremath{ {s_{#1\grksp{#2}}}^{ \ifdot{\grksp{#3}} } }}
\newcommand{\udess}[3]{\ensuremath{ { {s_{#1}}^{\grksp{#2}} }_{ \ifdot{\grksp{#3}}} }}
\newcommand{\ddess}[3]{\ensuremath{ s_{#1 \grksp{#2} \ifdot{\grksp{#3}}} }}
\newcommand{\invess}[3]{\ensuremath{{s^{-1}_{#1}}^{ \grksp{#2} \ifdot{\grksp{#3}}} }}
\newcommand{\duinvess}[3]{\ensuremath{ {s^{-1}_{#1\grksp{#2}}}^{ \ifdot{\grksp{#3}} } }}
\newcommand{\udinvess}[3]{\ensuremath{ { {s^{-1}_{#1}}^{\grksp{#2}} }_{ \ifdot{\grksp{#3}}} }}
\newcommand{\ddinvess}[3]{\ensuremath{s^{-1}_{#1 \grksp{#2} \ifdot{\grksp{#3}}} }}
\newcommand{\essdot}[2]{\ensuremath{(\ess{#1}{}{} . \ddess{#2}{}{}) }}
\newcommand{\essprd}[4]{\ensuremath{(\ess{#1}{}{} . \ddess{#2}{}{} . \ess{#3}{}{} . \ddess{#4}{}{}) }}
\newcommand{\tessprd}[6]{\ensuremath{(\ess{#1}{}{} . \ddess{#2}{}{} . \ess{#3}{}{} . \ddess{#4}{}{})_{\grksp{#5} \grksp{#6} } }}
\newcommand{\esspair}[4]{\ensuremath{(\ess{#1}{}{} . \ddess{#2}{}{})_{\grksp{#3} \grksp{#4} } }}
\newcommand{\sth}[2][{}]{\ensuremath{\theta_{#2}^{\ifbrack{#1}2} }}
\newcommand{\dth}[3][{}]{\ensuremath{\theta^{\ifbrack{#1}}_{#2 \grksp{#3}} }}
\newcommand{\uth}[3][{}]{\ensuremath{\theta^{\ifbrack{#1}\grksp{#3}}_{#2} }}
\newcommand{\mth}[3][{}]{\ensuremath{\theta^{\ifbrack{#1}#3}_{#2} }}
\newcommand{\Th}[1]{\ensuremath{\Theta^{\grksp{#1}} }}
\newcommand{\Ch}[2]{\ensuremath{\bar{\chi}_{#1\ifdot{\grksp{#2}}} }}
\newcommand{\uCh}[2]{\ensuremath{\bar{\chi}_{#1}^{\ifdot{\grksp{#2}}} }}
\newcommand{\sgsp}[2][{}]{\ensuremath{
\bar{\Lambda}_{#2}^{\ifbrack{#1} 2}
}}
\newcommand{\dgsp}[3][{}]{\ensuremath{{{\bar{\Lambda}_{#2\ifdot{\grksp{#3}}}}}^{\ifbrack{#1}} }}
\newcommand{\ugsp}[3][{}]{\ensuremath{
\bar{\Lambda}_{#2}^{\ifbrack{#1}\ifdot{\grksp{#3}}}
}}
\newcommand{\ptl}[3]{\ensuremath{ \partial_{#1}^{ \grksp{#2} \ifdot{\grksp{#3}} } }}
\newcommand{\sptl}[1]{\ensuremath{\partial_{#1}^{ 2} }}
\newcommand{\duptl}[3]{\ensuremath{ {\partial_{#1\grksp{#2}}}^{ \ifdot{\grksp{#3}} } }}
\newcommand{\udptl}[3]{\ensuremath{ { {\partial_{#1}}^{\grksp{#2}} }_{ \ifdot{\grksp{#3}}} }}
\newcommand{\ddptl}[3]{\ensuremath{ \partial_{#1 \grksp{#2} \ifdot{\grksp{#3}}} }}
\newcommand{\ensuremath{{q_{0}}}}{\ensuremath{{q_{0}}}}
\newcommand{\ensuremath{\mathcal F}}{\ensuremath{\mathcal F}}
\newcommand{\ensuremath{\mathcal H}}{\ensuremath{\mathcal H}}
\newcommand{\ttz}[1]{\ensuremath{\left. #1 \right|_{\theta_{2}=0}}}
\newcommand{n}{n}
\newcommand{\hspace{-18 pt}}{\hspace{-18 pt}}
\newcommand{\xpx}[1]{\ensuremath{ X \frac{\partial }{\partial X} \left( #1 \right)}}
\newcounter{lisno}
\renewcommand{\thelisno}{(\alph{lisno})}
\newcommand{\partial}{\partial}
\newcommand{\begin{array}}{\begin{array}}
\newcommand{\end{array}}{\end{array}}
\newcommand{\es}[1]{\ensuremath{#1}}
\newcommand{\su}[1]{\subsection*{#1}}
\newcommand{\alpha}{\alpha}
\newcommand{\theta}{\theta}
\newcommand{\epsilon}{\epsilon}
\newcommand{\Lambda}{\Lambda}
\newcommand{\z}[1]{\dot{#1}}
\section{Introduction}
\slab{introone}
The symmetries of Maxwell's classical equations have played a defining role in modern physics. Their importance for relating observers which moved at constant velocity with respect to each other was realised by Lorentz, prior to the development of special relativity by Einstein in 1905. Their gauge symmetry was discovered by Weyl in the 1920's and was extended to construct the standard model in
1967, However, it is only more recently that the true importance of
electromagnetic duality and conformal invariance has become apparent.
In fact, the conformal invariance of the classical Maxwell equations was realised as
long ago as 1909 \bibnum{pcwone}. Unfortunately, quantum effects in Maxwell's
theory coupled to electrons and in all other four-dimensional
theories which were subsequently studied for many years, lead to
violations of their conformal symmetries. The corresponding anomaly
is directly related to the appearance of infinities in quantum
field theory. Despite this, in the 1960's and 1970's there was
a revival of interest in four dimensional
conformal symmetry \bibnum{pcwtwo} and it was found that the two and three point
Green's functions could
be determined up to constants by conformal symmetry \bibnum{pcwthr}.
With the discovery of supersymmetry,
examples of conformally invariant four dimensional quantum field
theories were found. The first such example \bibnum{pcwfou} being the
$N=4$ Yang-Mills theory. Subsequently, it was realised that
there were an infinite number of $N=2$ theories \bibnum{pcwfiv} and even some
$N=1$ theories \bibnum{pcwsix}.
More recently other examples of conformally invariant
supersymmetric theories have been found \bibnum{pcwsev}.
Supersymmetric theories are most naturally formulated in terms of
superfields, since only then is their supersymmetry manifest and,
as a result, can their quantum properties be most systematically
studied. However, the superfields which describe physical
quantities are always subject to constraints.
For example, the Wess-Zumino model and the field strengths of $N=1$
and
$N=2$ Yang-Mills theory are described by chiral superfields. In fact,
these constraints,
which imply that these superfields
in effect live on only a subspace of the usual Minkowski superspace,
are directly responsible for the well known non-renormalisation
theorems in supersymmetric theories \bibnum{pcweig}. As a result of the pattern
found when calculating the chiral Green's functions in two dimensional
$N=2$ superconformal minimal models \bibnum{pcwnin} it
was proposed that the constraints on the superfields when combined
with superconformal invariance could also lead to results in four
dimensions which were stronger than those that were generically found
in conformally invariant, but non-supersymmetric theories. The first
such result was the realisation that the relation between the $R$-weight and the dilation weight of any chiral superfield could be used to
determine its dilation weight in a superconformal theory \bibnum{l9scfpap7}. In
reference \bibnum{l9scfpap8} the chiral Ward identities for any $N$ were given and
it was shown that there were no superconformal chiral
invariants. In a subsequent series of papers \bibnum{pcwtwe} it was also
realised that theories that involved harmonic superfields, such as
$N=4$ Yang-Mills theory would have very strong constraints placed on
them as a result of their superconformal invariance.
An early dicussion of three-point functions in $N=1$ superconformal theories appears in reference \bibnum{mpsetal}. In reference \bibnum{l9scfpap7} an expression for the three-point Green's function
for
$N=1$ chiral superfields was given. Unfortunately,
this expression was not correct and
was subsequently corrected by
the authors of the present paper in the thesis of reference \bibnum{apthesis}.
Although the work in this thesis was made available to some
workers, and some of its results were reviewed in reference \bibnum{pcwfoutnb},
it is not available to most workers in the field. A discussion of the superconformal group was given in references \bibnum{pcwfoutna}, \bibnum{conthesis} and \bibnum{pcwrig}. In this paper we
give some of the results of reference \bibnum{apthesis}
and extend them by calculating
the most general expression for the three-point chiral Green's function for
any $N$. The result can be succinctly summarised as
\begin{eqnarray}
G_{3}^{(N)} &=& \left( \frac{\sess{12}\sess{23}}{\sess{13}} \right)^{(N-2)} \left( \frac{\sess{12}}{\sess{13}} \right)^{\frac{(4-N)q_3}{N} }
\left( \frac{\sess{23}}{\sess{13}} \right)^{\frac{(4-N)q_1}{N}} \prod_{i=1}^{N} \sgsp[i]{123}, \elab{introgenres}
\end{eqnarray}
where $\sum_i q_i = N$.
Additionally, we show that \bibnum{apthesis}, contrary to na\"{\i}ve expectations, this does not in general imply that the chiral Green's functions higher than three-points
are determined up to constants. In fact, as a direct consequence of the nilpotent properties of these Green's functions, we find that we can not uniquely determine any solution above three-points when the total $R$-charge of the Green's function, denoted by \ensuremath{{q_{0}}}, is greater than one. However, in the particular case when $\ensuremath{{q_{0}}} = 1$, we find that the $N=1$ four-point solution is uniquely specified up to four constants by the superconformal Ward identities and we show how to construct the appropriate solution once these constants have been specified.
\section{General Properties of Solutions}
\slab{introtwo}
As discussed in \bib{l9scfpap8}, the superconformal Ward identities for translations, dilations and special conformal transformations acting on chiral Green's functions, $G$, are
\begin{eqnarray}
\swp{1}{1} \, G &=& \sum_{p=1}^{n} \begin{lbrce} \partial_{\alpha \dot \alpha} \end{lbrce} G=0, \elab{l9scwp}
\end{eqnarray}
\begin{eqnarray}
\ensuremath{D } \, G &=& \sum_{p=1}^{n} \begin{lbrce} s ^{\alpha \dot \alpha } \partial_{\alpha \dot \alpha} + {1\over 2} \theta ^{\alpha j} \partial_{\alpha j } + q{(4-N)\over N} \end{lbrce} G=0, \elab{l9scwd}
\end{eqnarray}
\begin{eqnarray}
\swk{2}{2} \, G &=& \sum_{p=1}^{n} \begin{lbrce} s ^{\alpha \dot \beta } s ^{\beta \dot \alpha } \partial_{\alpha \dot \alpha} +s ^{\alpha \dot \beta } \theta ^{\beta j} \partial_{\alpha j } + q{(4-N)\over N}s^{\beta \dot \beta} \end{lbrce} G=0, \elab{l9scwk}
\end{eqnarray}
For supersymmetry, they are
\begin{eqnarray}
\swq[i]{1} \, G &=& \sum_{p=1}^{n} \begin{lbrce} {\partial_{\alpha i} }\end{lbrce} G=0, \elab{l9scwq}
\end{eqnarray}
\begin{eqnarray}
\swqb[i]{1} \, G &=& \sum_{p=1}^{n} \begin{lbrce} \theta^{\alpha i} {\partial_{\alpha \dot \alpha} } \end{lbrce} G =0. \elab{l9scwqb}
\end{eqnarray}
For the internal symmetry, with traceless parameter $ E_j^{\ i}$, we
have the corresponding Ward identity
\begin{eqnarray}
\ensuremath{T } \, G &=& \sum_{p=1}^{n} \begin{lbrce} \theta ^{\alpha j} E_j^{\ i} \partial _{\alpha i} \end{lbrce} G=0, \elab{l9scwi}
\end{eqnarray}
and finally those for $S$-supersymmetry are given by
\begin{eqnarray}
\swsb[i]{1} \, G &=& \sum_{p=1}^{n} \begin{lbrce} s ^{\beta \dot \alpha }\partial _{ \beta i} \end{lbrce} G=0, \elab{l9scwsb}
\end{eqnarray}
\begin{eqnarray}
\sws[i]{2} \, G &=& \sum_{p=1}^{n} \begin{lbrce} s ^{\beta \dot\alpha } \theta ^{\alpha i}\partial _{\alpha \dot \alpha} - \theta ^{\beta i} \theta ^{\alpha j}\partial_{\alpha j} + q{(4-N)\over N}\theta ^{\beta i} \end{lbrce} G=0. \elab{l9scws}
\end{eqnarray}
In the above equations the sum is over $p$, however, this index is
suppressed on the coordinates and on $q$. We have used the shorthand notation
$\partial_{\alpha \dot \alpha}={\partial \over \partial s^{\alpha
\dot \alpha}}$ and
$\partial _{ \alpha i}= {\partial \over \partial \theta
^{\alpha i}}$.
For $N\neq 4$ we also have $R$ symmetry, with the corresponding
Ward identity
\begin{eqnarray}
\ensuremath{R } \, G &=& \sum_{p=1}^{n} \begin{lbrce} \theta ^{\alpha j} {\partial_{\alpha j}} -2q \end{lbrce} G=0. \elab{l9scwr}
\end{eqnarray}
The operators \[ \begin{lbrce} \swp{1}{1}, \swk{1}{1}, \ensuremath{D }, \swq{1}, \swqb{1}, \sws{1}, \swsb{1}, \ensuremath{R } \end{lbrce} , \] in the above, obey the superconformal algebra.
The variable \ess{p}{1}{1} is a chiral variable and takes the form
\begin{eqnarray}
\ess{}{1}{1} &=& x^{\alpha \dot\alpha} - \frac{i}{2} {\uth{}{1}}^{j} \bar{\theta}^{\dot\alpha}_{j}.
\end{eqnarray}
To begin with, we consider the case $N=1$, where \eqn{l9scwi} is trivially realised since $E$ is zero.
It is well known that the solution of the constraint in \eqn{l9scwp} implies that the Green's functions are functions of the differences \ess{pq}{1}{2} only, where $q=p+1$. This is easy to see if we consider our independent variables to be the differences \ess{pq}{1}{2}, where $q=p+1$, along with the sum
\begin{eqnarray}
\ess{0}{1}{2} &\equiv& \sum_{p}^{n} \ess{p}{1}{2} .
\end{eqnarray}
Clearly, any function of \ess{p}{1}{1} can be written in terms of these variables instead.
It follows from the chain rule that, for any Green's function $G$ obeying \eqn{l9scwp}, we may write
\begin{eqnarray}
\pdv{G}{\ess{0}{1}{1}}
&=& \sum_{p}^{n} \pdv{G}{\ess{p}{1}{1}}
\, \, \, \, \, = \, \, \, \, \, 0,
\end{eqnarray}
and the result follows.
A similar argument shows that the same is true for the $\uth{p}{1}$ variables. Defining
\begin{eqnarray}
\Th{1} &=& \sum_{p}^{n} \uth{p}{1} ,
\end{eqnarray}
we see that
\begin{eqnarray}
\pdv{G}{\Th{1}}
&=& \sum_{p}^{n} \pdv{G}{\uth{p}{1}}
\, \, \, \, \, = \, \, \, \, \, 0,
\end{eqnarray}
and thus $G$ is independent of $\Th{1}$ from \eqn{l9scwq}.
Given these simplifications, one might wonder whether any of the other operators in the algebra may be expressed as the derivative of a single variable by a suitable choice of independent variables. In particular, we consider \eqn{l9scwsb}, as it has a very simple form. We use the fact that $G$ is a function of \ess{p,p+1}{1}{1} and \uth{p,p+1}{1} only, and as a result, we may use the chain rule to write \swsb{1} as
\begin{eqnarray}
\swsb{1} &=&
\remark{\sum_{p=1}^{n-1} \ess{p}{2}{1} \pdv{\uth{p,p+1}{3}}{\uth{p}{2}} \pdv{G}{\uth{p,p+1}{3}} + \sum_{p=2}^{n } \ess{p}{2}{1} \pdv{\uth{p-1,p}{3}}{\uth{p}{2}} \pdv{G}{\uth{p-1,p}{3}}
\nonumber \\ &=& \sum_{p=1}^{n-1} \ess{p}{2}{1} \pdv{G}{\uth{p,p+1}{2}} - \sum_{p=2}^{n } \ess{p}{2}{1} \pdv{G}{\uth{p-1,p}{2}}
\nonumber \\ &=& }
\sum_{p=q-1}^{n-1} \ess{pq}{2}{1} \pdv{G}{\uth{pq}{2}} . \elab{l9swsb2}
\end{eqnarray}
However, we wish to go further and write it as
\begin{eqnarray}
\swsb{1} \, = \, \pdv{G}{\Ch{0}{1}} &=& \sum_{p=q-1}^{n-1} \pdv{\uth{pq}{2}}{\Ch{pq}{1}} \pdv{G}{\uth{pq}{2}} ,
\end{eqnarray}
for some variable \Ch{0}{1}, which must be a function of \ess{p,p+1}{1}{1} and \uth{p,p+1}{1}, where we also have
\begin{eqnarray}
\Ch{0}{1} &=& \sum_{p=1}^{n-1} \Ch{p, p+1}{1} \elab{l9chidef}
\end{eqnarray}
for some variables \Ch{pq}{1}.
Comparing this with \eqn{l9swsb2} we find
\begin{eqnarray}
\pdv{\uth{pq}{2}}{\Ch{pq}{1}} &=& \ess{pq}{2}{1}, \, \, \,\, \, \,\, \, \, p = q-1,
\end{eqnarray}
which implies
\begin{eqnarray}
\Ch{pq}{1} &=& \uth{pq}{2} \ddinvess{pq}{2}{1}, \elab{l9chipqdef}
\end{eqnarray}
and therefore
\begin{eqnarray}
\Ch{0}{1} &=& \sum_{p=q-1}^{n-1} \uth{pq}{2} \ddinvess{pq}{2}{1}. \elab{l9spindep}
\end{eqnarray}
Clearly, one can write any function of \ess{p,p+1}{1}{1} and \uth{p,p+1}{1} in terms \ess{p,p+1}{1}{1}, \Ch{0}{1} and \dgsp{p, p+1, p+2}{1}, defined by
\begin{eqnarray}
\ugsp{pqr}{1} &=& \uCh{pq}{1} - \uCh{qr}{1}.
\end{eqnarray}
From \eqn{l9spindep}, $G$ is independent of \Ch{0}{1} and must therefore be a function of the remaining independent variables, namely the \dgsp{pqr}{1} and the \ess{p,p+1}{1}{1}.
In summary, any arbitrary function, $G( \ugsp{pqr}{1}, \ess{pq}{1}{1} )$, obeys Equations \eqnnum{l9scwp}, \eqnnum{l9scwq} and \eqnnum{l9scwsb}, which leaves the Ward identities \eqnnum{l9scwd}, \eqnnum{l9scwk}, \eqnnum{l9scwqb}, \eqnnum{l9scws} and \eqnnum{l9scwr} to be solved.
\section{A Particular Three-Point Solution}
\slab{partic}
In the case of the three-point function, we see immediately from the above that there is only one independent spinor, namely \ugsp{123}{1}. We can see that a solution proportional to \ugsp{123}{1} alone is not possible even if we consider non-scalar solutions. The general form of such a solution would have to be
\begin{eqnarray}
G_{3}' &=& \ugsp{123}{1} \, h_{\ifdot{\grksp{1}} \grksp{2}}(\ess{pq}{}{}).
\end{eqnarray}
The action of \swqb{3} on this function yields two linearly independent terms, which can each be set to zero using the Ward identity in \eqnnum{l9scwqb}. These are :
\begin{eqnarray}
\sth{12} \, \, \Rightarrow \,\,\, \sdrv{12}{3}{3} \left( \invess{12}{3}{1} h_{\ifdot{\grksp{1}} \grksp{2}} \right) &=& 0, \\
\sth{23} \, \, \Rightarrow \,\,\, \sdrv{23}{3}{3} \left( \invess{23}{3}{1} h_{\ifdot{\grksp{1}} \grksp{2}} \right) &=& 0,
\end{eqnarray}
from which it is clear that $h_{\ifdot{\grksp{1}} \grksp{2}} = 0$, and thus there is no solution of this form.
For now, we restrict our attention to scalar solutions and deduce that the scalar three-point Green's function must be of the form
\begin{eqnarray}
G_{3} &=& \frac{1}{2} f(\ess{12}{1}{1}, \ess{23}{1}{1}) \, \sgsp{123} \elab{l93ptgf},
\end{eqnarray}
and the dependence on \uth{p}{1} is completely fixed. We must now determine the scalar function $f(\ess{12}{1}{1}, \ess{23}{1}{1})$ such that the remaining Ward identities are obeyed.
First note that the Ward identity of \eqn{l9scwd} involves the operator $\ess{}{1}{1} \partial_{\alpha \dot\alpha}$ which merely counts the overall power of \ess{pq}{1}{1} in $G$. The Ward identity of \eqn{l9scwr} involves the operator $\uth{j}{1} \partial_{\alpha j}$ which does the same for $\uth{j}{1}$ and thus \ensuremath{R }-symmetry fixes the value of
\begin{eqnarray}
\ensuremath{{q_{0}}} &\equiv& \sum_{p=1}^{n} q_{p}. \elab{l9qsum}
\end{eqnarray}
In this case, \eqn{l93ptgf} shows that
$\ensuremath{{q_{0}}} = 1$ is the only possibility.
One can then see by inspection that the dilation operator $\ensuremath{D }$ constrains $f$ to be of degree $-2$ in \ess{pq}{1}{1}.
Consider next the Ward identity of \eqn{l9scwqb}. This includes the
action of the operator \swqb{1} on $G_{3}$, which gives
\begin{eqnarray}
\swqb{1} \left[ \frac{1}{2} f \sgsp{123} \right] &=& - f \dgsp{123}{1} \left[ \frac{\sth{12}}{\sess{12}} - \frac{\sth{23}}{\sess{23}} \right] + \frac{1}{2} \sgsp{123} \, \swqb{1} f \nonumber \\
&=& \frac{2 f}{\sess{12}\sess{23}} \left[ \sth{12} \uth{23}{1} \ddess{23}{1}{1} - \sth{23} \uth{12}{1} \ddess{12}{1}{1} \right] + \frac{1}{2} \sgsp{123} \, \swqb{1} f. \elab{l9qbong3}
\end{eqnarray}
We note that any scalar function of \ess{12}{1}{1} and \ess{23}{1}{1} can be written in terms of the three independent variables
\begin{eqnarray}
a \, = \, \sess{12}, \, \,
& b \, = \, \essdot{12}{23},& \, \,
c \, = \, \sess{23},
\end{eqnarray}
where we have used the shorthand notation
\begin{eqnarray}
\essdot{12}{23} &=& \ess{12}{1}{1} \ddess{23}{1}{1}.
\end{eqnarray}
This follows from the equations \eqnnum{l9spsqu} and \eqnnum{l9spswap} given in \app{notapp}, which can be used to reduce any scalar expression at 3 points to a function of $a$, $b$ and $c$.
Using this we can rewrite \eqn{l9qbong3} as
\begin{eqnarray}
\! \! \! \! \! \! \! \! \! \! \! \! \swqb{1} \left[ \frac{1}{2} f \sgsp{123} \right] &=& \frac{(\sth{12} \uth{23}{1} \ddess{12}{1}{1} + \sth{23} \uth{12}{1} \ddess{23}{1}{1}) }{\sess{12}\sess{23}} \left[ 1 + a \pdv{}{a} + b \pdv{}{b} + c \pdv{}{c} \right] f
\end{eqnarray}
This implies that $f$ is a function of degree $-1$ in $a$, $b$, $c$ (i.e. degree $-2$ in \ess{12}{}{} and \ess{23}{}{}), which we know already from the dilation operator \ensuremath{D }. So, in this case \swqb{1} does not give any new constraints on $f$.
Without loss of generality, the function $f(a,b,c)$ may be considered as an arbitrary function of degree $0$ multiplied by any particular function of degree $-1$ in its arguments, with numerator $1$. Let us choose this function to be $1/\sess{13}$. Any degree zero function is in general a function of two independent ratios of $a$, $b$ and $c$, such as $(a+ 2 b + c)/a$ and $c/a$. Thus, we may write $G_{3}$ as
\begin{eqnarray}
G_{3} &=& \frac{1}{2} f' \left(\sess{13}/\sess{23} , \sess{12}/\sess{23}\right) \, \frac{\sgsp{123}}{{\ess{13}{}{}}^{2}} \elab{l93ptgf2},
\end{eqnarray}
which is the most general solution to Equations \eqnnum{l9scwp}, \eqnnum{l9scwd},
\eqnnum{l9scwq},
\eqnnum{l9scwqb}, \eqnnum{l9scwsb} and
\eqnnum{l9scwr}, for a scalar three point function.
We will find the most general chiral Green's
function in \sect{uniq}, however, as a step in this direction
we will now show that we can choose $f'=1$, and show by explicit calculation that $G_{3}$ is also a solution to \eqn{l9scws}, and so all the superconformal Ward identities. The full calculation is too long to reproduce here, however, we note that whilst $G$ has to be a function of the differences in the coordinates, i.e. \ess{p,p+1}{1}{1} and \uth{p,p+1}{1},
the Ward identity of \eqn{l9scws} involves $\sws{1} \, G$, which is not.
$\sws{1} \, G_{3}$ can therefore be expressed in terms of coefficients of \uth{p}{1} which are either of the form $\sth{p}\uth{q}{1}$ or $\uth{1}{1} \uth{2}{2} \uth{3}{3}$. It is the latter case which gives most difficulty, so we shall only discuss the coefficient of $\sth{1}\uth{3}{1}$ here.
Explicit calculation reveals that
\begin{eqnarray}
\scfs{2} \left[ \frac{\sgsp{123}}{2 \sess{13}} \right] &=& \frac{2\sth{1}\uth{3}{1} \ess{3}{2}{1} \ddess{13}{1}{1}}{\sess{12}{\ess{13}{}{}}^{4}}
- \frac{4\sth{1}\uth{3}{4} \ess{1}{2}{1} \ddess{23}{1}{1}}{\sess{12}\sess{23}\sess{13}} \nonumber \\ & &
+ \frac{4\sth{1}\uth{3}{4} \ess{1}{2}{1} \ddess{12}{1}{1} \ess{12}{1}{3} \ddess{23}{3}{4}}{{\ess{12}{}{}}^{4}\sess{23}\sess{13}}
+ \frac{4\sth{1}\uth{3}{4} \ess{1}{2}{1} \ddess{13}{1}{1} \ess{12}{1}{3} \ddess{13}{3}{4}}{\sess{12}\sess{23}{\ess{13}{}{}}^{4}}
\nonumber \\ & &
+ (1 - 3q_{1})\frac{2\sth{1}\uth{3}{3} \ess{12}{2}{3} \ddess{23}{3}{3}}{\sess{12}\sess{23}\sess{13}}
+ \frac{3q_{3}\sth{1}\uth{3}{2} }{\sess{12}\sess{13}}
+ \ldots, \elab{l9song3}
\end{eqnarray}
where the dots denote other linearly independent terms. After some rearrangement, using in particular Equations \eqnnum{l9spsqu} and \eqnnum{l9spswap}, we obtain
\begin{eqnarray}
\scfs{2} \left[ \frac{\sgsp{123}}{2 \sess{13}} \right] &=& (3q_{3} -1) \frac{\sth{1}\uth{3}{2} }{\sess{12}\sess{13}}
- (3q_{1} -1) \frac{2\sth{1}\uth{3}{3} \ess{12}{2}{3} \ddess{23}{3}{3}}{\sess{12}\sess{23}\sess{13}} + \ldots.
\end{eqnarray}
We know that $\ensuremath{{q_{0}}} = 1$, so it follows that \eqn{l9scws} is only valid when
\begin{eqnarray}
q_{1} \,\, = \,\, q_{2} &=& q_{3} \,\, = \,\, \frac{1}{3}. \elab{l9qvals}
\end{eqnarray}
Since one can show that \[ \frac{\sgsp{123}}{2 \sess{13}}, \] is invariant under cyclic permutation of 1, 2, and 3, this result can be extended to all coefficients of the form $\sth{p}\uth{q}{1}$. We note that the operators in the algebra are trivially cyclic invariants, given \eqn{l9qvals}. Thus, the coefficient of $\sth{1}\uth{3}{1}$ in $\sws{1} G_{3}$ is the same as the coefficient of $\sth{2}\uth{1}{1}$ and $\sth{3}\uth{2}{1}$, and the result follows. It remains to prove the corresponding result for the coefficient of $\uth{1}{1} \uth{2}{2} \uth{3}{3}$, which has been done, but is very laborious and yields nothing new. Once again, \eqn{l9qvals} must apply for the result to vanish.
From the superconformal algebra, we know that \swk{1}{1} is related to the anticommutator of the special supersymmetry generators, \sws{1} and \swsb{1}, so that Equations \eqnnum{l9scwsb} and \eqnnum{l9scws} together imply \eqn{l9scwk}. Thus,
\begin{eqnarray}
G_3^{0} &\equiv& \frac{\sgsp{123}}{2 \sess{13} } \elab{l9partsol}
\end{eqnarray}
is a solution to all of the $N=1$ Ward identities, given \eqn{l9qvals}.
Recall that we arbitrarily chose the function $f'$ to be $1$ when proving
that the superconformal Ward identities are satisfied.
Of course for a specific set of $R$ charges, and so dilation weights, the three-point function must be unique as a consequence of
the standard results of ordinary conformal field theory. As such, for the charges of \eqn{l9qvals} this result is the unique result. In \sect{uniq}, and
using the results of \sect{partic}, we will
give a superspace proof of this fact and then find the unique three-point chiral Green's function for all possible $R$ charges.
\section{Conditions for Uniqueness of Green's Functions}
\slab{uniq}
A standard argument to establish the uniqueness of Greens functions goes as follows. Consider two Green's functions, $g_{1}$ and $g_{2}$. Their ratio, $r$, will satisfy all the Ward identities with no isotropy transformations and
so will be an invariant. If however, one can prove that there are no invariants then one can establish the uniqueness. If we were to apply this argument for
the case of chiral Green's functions considered in this paper, then we would find that \eqn{l9scwr} implies $r$ is independent of \uth{pq}{1}. From this, it follows that \eqn{l9scwqb} implies $r$ must be independent of \ess{pq}{1}{1} as well, and hence simply a constant. This might, at first sight, appear to suggest that all the chiral Green's functions were unique. Such an argument was suggested in the first version of \bib{l9scfpap8}.
In fact, \eqn{l9scwr} implies that $r$ is of degree zero in \uth{pq}{1}, but the Green's functions are proportional to some power of \uth{pq}{1}
and so are nilpotent. It is of course not correct to divide by nilpotent
quantities. Nonetheless, one might hope
that one could in effect still arrive at the correct result. However, we
note that the above argument does not require $S$-supersymmetry and we have already shown by explicit construction, that the solution to the Ward identities in the absence of \eqnnum{l9scws} is not unique ( even at three points). Therefore the above uniqueness argument must be incorrect.
To see why, we should ask the related question: ``If $g_{1}$ is a Green's function, does there exist another of the form $g_{2} = r(\ess{pq}{1}{1}) g_{1}$ ?'' This is equivalent to studying the ratio $r$, except that no division has occurred. $r$ has to be independent of \uth{pq}{1} or otherwise $g_2$ will not satisfy Ward Identities with the same value $q_0$ as $g_1$, which it must do by hypothesis.
As usual, we deduce from \eqn{l9scwp} that $r$ is a function of the differences, \ess{p,p+1}{1}{1}, and from \eqn{l9scwd}, that $r$ must be of degree zero. Equations \eqnnum{l9scwq}, \eqnnum{l9scwsb} and \eqnnum{l9scwr} are trivially realised on $g_{2}$. From \eqn{l9scwqb}, we deduce
\begin{eqnarray}
\swqb{1} g_{2} &=& g_{1} \, \swqb{1} r \,\,\,\,\,\, = \,\,\,\,\,\, 0.
\end{eqnarray}
This does not imply that $\swqb{1} r = 0$, because $g_{1}$ is a function of Grassmann odd variables. Considering our explicit form of $G_{3}$, we see that it is the square of the spinor \ugsp{123}{1}. If $\swqb{1} r$ was proportional to \ugsp{123}{1} then $g_{2}$ could obey \eqn{l9scwqb}.
However, $\swqb{1} r$ would be non-zero and thus $r$ would be dependent on \ess{pq}{1}{1}, violating the uniqueness argument.
By definition, we take $r$ to be independent of $\theta$ and to be a scalar,
and the dilation Ward identity implies that it is of of degree $0$ in \ess{p, p+1}{}{}. We now show that, when acted on by \swqb{1}, any scalar function of degree zero in \ess{pq}{}{} becomes a function of \dgsp{pqr}{1}.
To see this, note that any scalar function of \ess{p,p+1}{1}{1} can be written in terms of \esspair{p,p+1}{q,q+1}{1}{2}. This is because a scalar is a trace of a product of \ess{p, p+1}{1}{1} and, in order to take the trace of such a term, one must have an even number of \ess{p, p+1}{1}{1}. For example, one can write
\beq \begin{array}{rcl}
\eqal{
\sess{pq} }{=}{ \epsilon^{\grksp{2}\grksp{1}} \esspair{pq}{pq}{1}{2}}
\eqal{
\essdot{pq}{rs} }{=}{ \epsilon^{\grksp{2}\grksp{1}} \esspair{pq}{rs}{1}{2} }
\leqal{
(s_{p}.s_{q}.s_{r}.s_{t}) }{=}{ \epsilon^{\grksp{2}\grksp{1}} \epsilon^{\grksp{4}\grksp{3}} \esspair{p}{q}{1}{3}\esspair{r}{t}{4}{2}
}
\end{array} \eeq
Hence every scalar can be written as the trace of a product of terms of the form \esspair{p,p+1}{q,q+1}{1}{2}.
Furthermore, a scalar expression of degree 0 can always be written in terms of expressions of the form :
\[ \frac{\esspair{p,p+1}{q,q+1}{1}{2}}{\sess{p,p+1}}. \]
Therefore, all we need to do is establish the required result for all expressions of the form
\[ \frac{\esspair{p,p+1}{q,q+1}{1}{2}}{\sess{p,p+1}} \]
and it automatically follows for all scalars by using Leibniz rule for first order linear differential operators.
Direct calculation shows that
\begin{eqnarray}
\swqb{1} \left[ \frac{\esspair{pq}{rs}{1}{2}}{\sess{pq}} \right] &=& - \left( \ugsp{pqr}{2} + \ugsp{qrs}{2} \right) \left(
\frac{\ddess{pq}{1}{1} \ddess{rs}{2}{2}}{\sess{pq}} \right),
\end{eqnarray}
and thus any function of degree zero is a function of terms of the form \dgsp{pqr}{1} when acted on by \swqb{1}. Using the relations given in \app{notapp} we can see that \dgsp{pqr}{1} can always be written in terms of the basis set of functions \dgsp{a, (a+1), (a+2)}{1}. This means that $\swqb{1} k$ can in principle be non-zero and the solution to the Ward identities, excluding \eqn{l9scws} and \eqn{l9scwk}, is not necessarily unique. In particular, at three points $\swqb{1} r$ is simply proportional to \dgsp{123}{1}, and therefore $G_3 \, \swqb{1} r$ vanishes for any $r$ of degree zero, not just constant values.
In order to pursue the question of uniqueness, we must therefore consider the effect of either \swk{1}{1} or \sws{1} on a given Green's function, $g_{1}$. These operators are not independent, as seen from the algebra, and thus we need to consider only one of them. We choose \swk{1}{1} for reasons which will become clear below.
\swk{1}{1} is the sum of a part, \lbswk{1}{1}, which is first order in differential operators and so obeys the Leibniz rule and a multiplicative operator, \mlswk{1}{1}{q}, which contains the isotropy group action. These two parts are given by
\begin{eqnarray}
\lbswk{2}{2} G &=& \sum_{p}^{n} \begin{lbrce} s ^{\alpha \dot \beta } s ^{\beta \dot \alpha } \partial_{\alpha \dot \alpha} +s ^{\alpha \dot \beta } \theta ^{\beta j} \partial_{\alpha j } \end{lbrce} G \\ \mlswk{2}{2}{q} G &=& \sum_{p}^{n} \begin{lbrce} q {(4-N)\over N}s^{\beta \dot \beta} \end{lbrce} G
\end{eqnarray}
Recall that the $p$ index is suppressed on the coordinates and $q$ in the above, and they should properly be written $q_{p}$, etc. We define
\begin{eqnarray}
\remark{\ensuremath{{q_{0}}} &=& \sum_{p=1}^{n} q_{p} \\}
\ensuremath{\mathbf{q} } &=& (q_{1}, q_{2}, q_{3}, \ldots, q_{n})
\end{eqnarray}
and once again we consider
\begin{eqnarray}
g_{2} &=& r(\ess{pq}{1}{1}) \, g_{1},
\end{eqnarray}
where $r$ is degree $0$ in \ess{pq}{1}{1}, as explained above. It follows that
\begin{eqnarray}
\left . \swk{1}{1} g_{2} \right|_{\ensuremath{\mathbf{q} } = \ensuremath{\mathbf{q} }_2} &=& r \, \lbswk{1}{1} g_{1} + g_{1} \, \lbswk{1}{1} r + \mlswk{1}{1}{\ensuremath{\mathbf{q} }_{2}} (r g_{1}),
\end{eqnarray}
for some $\ensuremath{\mathbf{q} }_{2}$, which is by definition the $R$ weight of the Green's function $g_2$.
Defining $\ensuremath{\mathbf{q} }_1$ to be the $R$ weight of the Green's function $g_1$, we have the equation
\begin{eqnarray}
\swk{1}{1} g_{1} &=& \lbswk{1}{1} g_{1} + \mlswk{1}{1}{\ensuremath{\mathbf{q} }_{1}} g_{1} \nonumber \\ &=& 0,
\end{eqnarray}
and thus, writing
\begin{eqnarray}
\ensuremath{\mathbf{q} }_{2} &=& \ensuremath{\mathbf{q} }_{1} + \delta\ensuremath{\mathbf{q} }
\end{eqnarray}
we deduce that, for $g_{2}$ to be a Green's function, we must have
\begin{eqnarray}
\swk{1}{1} g_{2} &=& g_{1} \left( \lbswk{1}{1} r + \mlswk{1}{1}{\delta\ensuremath{\mathbf{q} }} r \right) \nonumber \\ &=& 0. \elab{l9g2isgrn}
\end{eqnarray}
We note that unlike $\swqb{1} r$, $\swk{1}{1} r$ is independent of \uth{pq}{1}
since $r$ is independent of \uth{pq}{1} and so the $\theta $ dependent terms in $K^{\alpha \dot \alpha}$ do not act. It is for this reason that
we chose to investigate the action of \swk{1}{1} in contrast to \sws{1}. As a result, we may demand
\begin{eqnarray}
\lbswk{1}{1} r + \mlswk{1}{1}{\delta\ensuremath{\mathbf{q} }} r &=& 0,
\end{eqnarray}
from \eqn{l9g2isgrn}, irrespective of the nilpotence of $g_1$.
If $\delta\ensuremath{\mathbf{q} } = 0$, this is equivalent to demanding that $r$ be an ordinary conformal invariant in \ess{pq}{1}{1}. If $\delta\ensuremath{\mathbf{q} } \neq 0$, then strictly, the two Green's functions are solutions to {\em different} Ward identities, since $\ensuremath{\mathbf{q} }_1$ and $\ensuremath{\mathbf{q} }_2$ are distinct.
Considering the action of \lbswk{1}{1} on \sess{pq}, we find
\begin{eqnarray}
\lbswk{1}{1} \left( \sess{pq} \right) &=& \sess{pq} \left( \ess{p}{}{} + \ess{q}{}{} \right)^{\grksp{1}\ifdot{\grksp{1}}}, \elab{l9cononssqu}
\end{eqnarray}
which is of the form of \mlswk{1}{1}{\ensuremath{\mathbf{q} }} (\sess{pq}), for a
choice
of \ensuremath{\mathbf{q} }\ whose values are associated with the legs $p$ and $q$. In contrast, when acting on a sum of such terms, we find that
\begin{eqnarray}
\lbswk{1}{1} \left( \sess{pq} + \sess{rs} \right) &=& \sess{pq} \left( \ess{p}{}{} + \ess{q}{}{} \right)^{\grksp{1}\ifdot{\grksp{1}}} + \sess{rs} \left( \ess{r}{}{} + \ess{s}{}{} \right)^{\grksp{1}\ifdot{\grksp{1}}},
\end{eqnarray}
which can never be generated by \mlswk{1}{1}{\ensuremath{\mathbf{q} }} acting on a scalar function. The same is true of any function involving the sum of two or more distinct \sess{pq}. In particular,
\begin{eqnarray}
\essdot{pq}{rt} &=& \sess{pt} + \sess{qr} - \sess{pr} - \sess{qt} \elab{essdotexp}
\end{eqnarray}
implies that
\begin{eqnarray}
\swk{1}{1} \essdot{pq}{rt} &\neq& 0, \elab{l9kpair}
\end{eqnarray}
from which it follows that (see \app{notapp} for notation)
\begin{eqnarray}
\swk{1}{1} \essprd{1}{2}{3}{4} &\neq& 0
\end{eqnarray}
since otherwise, choosing $\ess{3}{}{} = \ess{4}{}{}$, would give a contradiction with \eqn{l9kpair}. This accounts for all scalar expressions and we must therefore construct the function, $r$, from products of \sess{pq}, and ensure that it is of degree zero. For such functions, $\lbswk{1}{1} r$ is of the form $\mlswk{1}{1}{\delta \ensuremath{\mathbf{q} }} r$, and thus we may find a non-vanishing $r$ such that
\begin{eqnarray}
\swk{1}{1} r &=& \lbswk{1}{1} r + \mlswk{1}{1}{\delta\ensuremath{\mathbf{q} }} r \nonumber \\ &=& 0 ,
\end{eqnarray}
for a suitable choice of $\delta\ensuremath{\mathbf{q} }$.
If we restrict our attention to the explicitly known three-point solution, given in Equations \eqnnum{l9partsol} and \eqnnum{l9qvals}, then we see that since no three-point purely conformal invariant exists,
it must be unique once the $R$ charges $\ensuremath{\mathbf{q} }$ of the chiral Greens function are specified. However, there are an infinite number of solutions which have distinct $\ensuremath{\mathbf{q} }$. We can generate a new Green's function from an existing one simply by multiplying by a degree zero function of \sess{pq}, where there is no restriction on $p$ or $q$. At three-points, the new Green's function is a solution to a
set of Ward Identities with the appropriate $R$ charges. Thus, uniqueness survives at the three-point level, but only by virtue of the standard uniqueness of any conformal three-point function. It is is not due to the chirality of the Green's function.
In contrast, one can generate a four-point solution by multiplying together two three-point solutions, with different \ugsp{pqr}{1}. We have
\begin{eqnarray}
G_{4} &=& r(\sess{pq}) \, \frac{\sgsp{123}}{2\sess{13}} \frac{\sgsp{234}}{\sess{24}} ,
\end{eqnarray}
for some choice of $\ensuremath{\mathbf{q} }$. This follows from the three point results for the Leibniz parts of the differential operators, and the remaining terms coming from the isotropy group action can be made to vanish by choosing \ensuremath{\mathbf{q} }\ suitably. In this case, $\ensuremath{{q_{0}}} = 2$. For the particular case of $r=1$, we deduce from our knowledge of the three-point solution that
\begin{eqnarray}
2 q_{1} \, = \, q_{2} &=& q_{3} \, = \, 2 q_{4} \, = \, \frac{2}{3}. \elab{l9qvals4}
\end{eqnarray}
However, as seen from the above discussion, uniqueness depends on the existence of an ordinary conformal $n$-point function, i.e. one which obeys $\lbswk{1}{1} r = 0$. It is well known from ordinary conformal theory that such invariants exist, e.g. the independent cross ratios
\begin{eqnarray}
u \,\, \, \, = \,\,\,\, \frac{\sess{12}\sess{34}}{\sess{13}\sess{24}}, \,\,\,\,\,\,\,\,\, v \,\, \, \, = \,\,\,\, \frac{\sess{23}\sess{14}}{\sess{13}\sess{24}} \elab{uvdefs}.
\end{eqnarray}
The fact that these are conformal invariants is easily seen from \eqn{l9cononssqu}.
If $r$ is a function of $u, v$, then \eqn{l9g2isgrn} is valid when $\delta\ensuremath{\mathbf{q} } = 0$. Thus, for four-points and above, there exist distinct Green's functions, $g_{1}$ and $g_{2}$, which are solutions to precisely the same Ward identities, and thus they are not unique. Given \eqn{l9qvals4}, we may write the corresponding four point solution as
\begin{eqnarray}
G_{4} &=& r(u,v) \, \frac{\sgsp{123}}{2\sess{13}} \frac{\sgsp{234}}{\sess{24}} ,
\end{eqnarray}
where $r(u,v)$ is completely arbitrary.
This result can be traced directly to the nilpotence of
$\ugsp{123}{1}$ and $\ugsp{234}{1}$. In particular, when either of these spinors is raised to the third power they vanish. Thus, one might attempt to find unique solutions, for a given \ensuremath{{q_{0}}}, by restricting the value of \ensuremath{{q_{0}}}\ to $1$, and thus ruling out the possibility of this effect.
Whilst such solutions may well exist, uniqueness is still not guaranteed. In the case where $\ensuremath{{q_{0}}}=1$, the solution may be of the form $\bar{\Pi}^{2}$, where $\bar{\Pi}_{\ifdot{\grksp{1}}}$ is some linear combination of the \dgsp{pqr}{1}. If $\bar{\Pi}_{\ifdot{\grksp{1}}}$ were found to be equal to $\swqb{1} h$, where $h$ was some function of the cross ratios, then we could generate new Green's functions from existing ones by multiplying them by arbitrary functions of $h$.
We shall explore this idea later on in this paper. For the moment, we return to the three point function to find its explicit form for any given \ensuremath{\mathbf{q} }.
\section{The Full $N=1$ Three-Point Solution}
\slab{full}
We know already that \eqn{l9partsol} defines a solution to the Ward Identities in the special case where :
\begin{eqnarray}
\ensuremath{\mathbf{q} } &=& \ensuremath{\mathbf{q} }_1 \,\,\,\,\, = \,\,\,\,\, \left( \frac{1}{3}, \frac{1}{3}, \frac{1}{3} \right)
\end{eqnarray}
We also know from our discussion above that we can generate a new solution to {\em different} Ward Identities, which have $\ensuremath{\mathbf{q} } = \ensuremath{\mathbf{q} }_1 + \delta\ensuremath{\mathbf{q} }$, by multiplying $G_{3}^{0}$ by any degree zero function of \ess{pq}{}{} which obeys
\begin{eqnarray}
\left . \swk{1}{1} r \, \right|_{\ensuremath{\mathbf{q} } = \delta\ensuremath{\mathbf{q} }}&=& 0.
\end{eqnarray}
We have also seen that the only solutions to such an equation must be in the form of a sum of products of \sess{pq}, so that in general
\begin{eqnarray}
r &=& \sum_{a,b} c_{ab} \left( \sess{12}\right)^a \left( \sess{23}\right)^b \left( \sess{13} \right)^{(-a-b)},
\end{eqnarray}
for some constants $c_{ab}$.
Acting on this with \swk{1}{1} and setting the coefficients of the linearly independent terms to zero, we see that for each term individually, we obtain
\begin{eqnarray}
-a + 3 \dq[3] &=& 0, \nonumber \\
a + b + 3 \dq[2] &=& 0, \\
-b + 3 \dq[1] &=& 0. \nonumber
\end{eqnarray}
We can solve this for $a$ and $b$, given $\delta\ensuremath{\mathbf{q} }$, and thus we obtain the general form of the scalar three-point function, as
\begin{eqnarray}
G_{3} &=& \left( \frac{\sess{12}}{\sess{13}} \right)^{3q_3}
\left( \frac{\sess{23}}{\sess{13}} \right)^{3q_1}
\left( \frac{\sess{13}}{\sess{12}\sess{23}} \right)
\sgsp{123} \elab{l9fullsol}
\end{eqnarray}
up to a multiplicative constant. From the above arguments concerning the allowed form of $r$, it also follows that the only way to generalise this to include non-scalar three-point functions is to multiply $G_{3}$ by a constant tensor, such as $\scfeps{1}{2}$, otherwise \swk{1}{1} will not vanish.
\section{The Complete Three-Point Solution in Extended Supersymmetry}
\slab{exten}
\subsection{$N=2$}
We can generalise the discussion of three-point solutions to situations where $N>1$. For example, when $N=2$ we have twice as many spinors, i.e. \dgsp[1]{123}{} and \dgsp[2]{123}{}.
One might suppose that we could construct a new type of solution with $\ensuremath{{q_{0}}} = 1$, but such a solution would have to take the form
\begin{eqnarray}
\left . G_{3}^{(2)} \right|_{\ensuremath{{q_{0}}} = 1} &=& \ugsp[1]{123}{1} f_{\ifdot{\grksp{1}} \ifdot{\grksp{2}}} \ugsp[2]{123}{2},
\end{eqnarray}
for some function, $f_{\ifdot{\grksp{1}} \ifdot{\grksp{2}}}$ of \ess{pq}{}{}.
However, acting on this with \swqb[1]{3}, implies that both the coefficient of \uth[2]{12}{} and of \uth[2]{23}{} in the resulting expression, vanish independently. In particular, \uth[2]{12}{3} implies that
\begin{eqnarray}
\swqb[1]{3} \left( \ugsp[1]{123}{1} f_{\ifdot{\grksp{1}} \ifdot{\grksp{2}}} \duinvess{12}{3}{2} \right) &=& 0.
\end{eqnarray}
In \sect{partic} we showed that this can only be satisfied when $f_{\ifdot{\grksp{1}} \ifdot{\grksp{2}}} = 0$. One can extend this argument to prove that there are no solutions for $\ensuremath{{q_{0}}} < N$ and hence a non-zero $n$-point solution only exists for
\begin{eqnarray}
N \leq \ensuremath{{q_{0}}} \leq (n-2) N \elab{quenone}.
\end{eqnarray}
The upper bound simply follows from the fact that the solution
must be composed from
$(n-2)$ \dgsp[i]{pqr}{}'s, for a given internal symmetry index $i$,
of which there are $N$ types.
As a result, the general $N=2$ three-point function has to be of the form
\begin{eqnarray}
\left . G_{3}^{(2)} \, \right|_{\ensuremath{{q_{0}}} = 2} &=& f(\ess{pq}{}{}) \, \sgsp[1]{123} \, \sgsp[2]{123},
\end{eqnarray}
which is manifestly zero under the action of \swp{1}{1}, \swq[i]{1} and \swsb[i]{1}. However, using our $N=1$ results, the action of \swqb[1]{1} gives
\begin{eqnarray}
\swqb[1]{1} \left . G_{3}^{(2)} \, \right|_{\ensuremath{{q_{0}}} = 2} &=& \frac{\sgsp[1]{123}}{\sess{13}} \, \swqb[1]{1} \left( \sess{13} \, f(\ess{pq}{}{}) \, \sgsp[2]{123} \right)
\end{eqnarray}
and since \swqb[1]{1} can never be zero on a function of \dgsp[2]{123}{1}, at first sight there appears to be no solution.
At this point it helps to consider the action of \ensuremath{D }, which depends on $N$, and hence the $N=2$ solution cannot simply be a product of two $N=1$ solutions, as one might initially expect. Closer inspection reveals that for \ensuremath{D }\ to vanish we require $f$ to be of degree zero in \ess{pq}{}{} and as a result $\sess{13} \, f(\ess{pq}{}{}) \, \sgsp[2]{123}$ has to be degree zero in \ess{pq}{}{} aswell. Expansion of this term in \dth[2]{p q}{1} gives a series of linearly independent terms whose coefficients can all be expressed as combinations of expressions of the form
\[
\frac{\esspair{p,p+1}{q,q+1}{1}{2}}{\sess{p,p+1}}.
\]
We know from \sect{uniq} that such functions must vanish under the action of $\sgsp[1]{123} \, \swqb[1]{1}$ at three-points, and thus it is precisely the same nilpotence which prevented us from obtaining unique solutions, that is responsible for the existence of any three-point solutions at all beyond $N=1$.
Thus, from the $\swqb{1}$ superconformal Ward identity we find no further condition.
As before, we can find the unique form of $f$ by demanding that $\swk{1}{1} G_{3}^{(2)}$ must vanish. It follows from \eqn{l9scwk}, that
\begin{eqnarray}
\!\!\!\!\!\!\!\!\!\! \left . \swk{1}{1} \sgsp[1]{123} \sgsp[2]{123} \right|_{\ensuremath{\mathbf{q} }_1 + \ensuremath{\mathbf{q} }_2} \! \! &=& \! \! \left . \sgsp[1]{123} \left( \swk[2]{1}{1} \sgsp[2]{123} \right) \right|_{\ensuremath{\mathbf{q} }_2} + \left . \left( \swk[1]{1}{1} \sgsp[1]{123} \right) \right|_{\ensuremath{\mathbf{q} }_1} \sgsp[2]{123},
\end{eqnarray}
where,
\begin{eqnarray}
\swk[1]{2}{2} &\equiv& \sum_{p=1}^{n} \begin{lbrce} s ^{\alpha \dot \beta } s ^{\beta \dot \alpha } \partial_{\alpha \dot \alpha} +s ^{\alpha \dot \beta } \theta ^{\beta (1)} \partial_{\alpha (1) } + q{(4-N)\over N}s^{\beta \dot \beta} \end{lbrce} , \, \,\,\,\, \mathrm{etc.}
\end{eqnarray}
The $N=1$ results show that
\begin{eqnarray}
\left . \left( \swk[1]{1}{1} \sgsp[1]{123} \right) \right|_{\ensuremath{\mathbf{q} } = \ensuremath{\mathbf{q} }_1(N)} &=& 0, \,\,\,\,\,\,\,\,\,\,\,\,
\ensuremath{\mathbf{q} }_1(N) \,\,\,\, = \,\,\,\, (0, \frac{N}{4-N}, 0),
\end{eqnarray}
and thus at $N=2$ we find that $\ensuremath{\mathbf{q} }_1(2) = (0,1,0)$, so that $\sgsp[1]{123} \sgsp[2]{123}$ is a solution for
\begin{eqnarray}
\ensuremath{\mathbf{q} } = (0,2,0). \elab{qatfour}
\end{eqnarray}
Once again,
\begin{eqnarray}
f &=& \sum_{a,b} c_{ab} \left( \sess{12}\right)^a \left( \sess{23}\right)^b \left( \sess{13} \right)^{(-a-b)}, \\
\ensuremath{\mathbf{q} } &=& (0,2,0) + \delta\ensuremath{\mathbf{q} },
\end{eqnarray}
from which we deduce that $c_{ab} \neq 0$ only when
\begin{eqnarray}
a &=& \dq[3] \,\,\,\,\, = \,\,\,\,\, q_{3}
\nonumber \\
b &=& \dq[1] \,\,\,\,\, = \,\,\,\,\, q_{1}.
\end{eqnarray}
Therefore, the most general $N=2$ three-point solution is
\begin{eqnarray}
G_{3}^{(2)} &=& \left( \frac{\sess{12}}{\sess{13}} \right)^{q_3}
\left( \frac{\sess{23}}{\sess{13}} \right)^{q_1}
\sgsp[1]{123} \sgsp[2]{123}. \elab{l9fullntwosol}
\end{eqnarray}
\subsection{$N=4$}
In $N=4$ supersymmetry, the $\ensuremath{\mathbf{q} }$ dependence drops out of the Ward Identities because of the factor $(4-N)/ N$ and \ensuremath{R }-symmetry no longer holds. Instead, we can define \ensuremath{{q_{0}}}\ to be half the degree of \uth[i]{pq}{} and the condition given in \eqn{quenone} still holds. In the usual way, we write
\begin{eqnarray}
\left . G_{3}^{(4)} \, \right|_{\ensuremath{{q_{0}}} = 4} &=& f(\ess{pq}{}{}) \, \sgsp[1]{123} \sgsp[2]{123} \sgsp[3]{123} \sgsp[4]{123},
\end{eqnarray}
and \ensuremath{D }-symmetry implies that
\begin{eqnarray}
f &=& (\sess{13})^2 \, r(\ess{pq}{}{}),
\end{eqnarray}
for any $r$ of degree zero in \ess{pq}{}{}.
Written in a different way, the $N=1$ results show that
\begin{eqnarray}
\lbswk[1]{1}{1} \sgsp[1]{123} &=& \ess{2}{1}{1} \sgsp[1]{123},
\end{eqnarray}
and because $\swk{1}{1} = \lbswk{1}{1}$, when $N=4$, we deduce that
\begin{eqnarray}
\left( \swk{1}{1} - 4 \ess{2}{1}{1} \right) \prod_{i=1}^{N} \sgsp[i]{123} &=& 0.
\end{eqnarray}
It follows that the general form of the $N=4$ three-point function is
\begin{eqnarray}
G_{3}^{(4)} &=& \left( \frac{\sess{12}\sess{23}}{\sess{13}} \right)^2 \prod_{i=1}^{N} \sgsp[i]{123}.
\end{eqnarray}
\subsection{General Formula}
A concise summary of all our results at three-points is given by the general formula
\begin{eqnarray}
G_{3}^{(N)} &=& \left( \frac{\sess{12}\sess{23}}{\sess{13}} \right)^{(N-2)} \left( \frac{\sess{12}}{\sess{13}} \right)^{\frac{(4-N)q_3}{N} }
\left( \frac{\sess{23}}{\sess{13}} \right)^{\frac{(4-N)q_1}{N}} \prod_{i=1}^{N} \sgsp[i]{123},
\end{eqnarray}
with $\ensuremath{{q_{0}}} = N$.
This formula is also valid for $N=3$.
\section{The Four-Point Green's Function at $\ensuremath{{q_{0}}} = 1$}
\slab{l9grncalctwo}
The most general four-point function\footnote{In reference \bibnum{pcwfoutnb}, there was some discussion of four-point functions from a different perspective, but the relevant results appear to disagree with those presented here. }, at $\ensuremath{{q_{0}}} = 1$, satisfying Eqs. \eqnnum{l9scwp}, \eqnnum{l9scwq} and \eqnnum{l9scwsb} is,
\begin{eqnarray}
\es{G_{4}} &=&
\frac{f}{\sess{12}} \sgsp{123} +
\frac{g}{\sess{34}} \sgsp{234} +
\frac{2 h}{\sess{23}} \left( \ugsp{123}{} \dgsp{234}{} \right) +
\frac{4 k}{\sess{12}\sess{23}} \left( \ugsp{123}{} \ess{12}{}{} \ess{23}{}{} \dgsp{234}{} \right) + \nonumber \\ & &
\frac{4 l}{\sess{12}\sess{23}} \left( \ugsp{123}{} \ess{23}{}{} \ess{34}{}{} \dgsp{234}{} \right) +
\frac{4 m}{\left(\sess{23}\right)^2} \left( \ugsp{123}{} \ess{12}{}{} \ess{34}{}{} \dgsp{234}{} \right) ,
\end{eqnarray}
where,
\begin{eqnarray}
\left( \ugsp{p}{} \ess{q}{}{} \ess{r}{}{} \dgsp{t}{} \right) &\equiv& \ugsp{p}{1} \udess{q}{3}{1} \duess{r}{3}{2} \dgsp{t}{2}
\end{eqnarray}
and $f$, $g$, $h$, $k$, $l$, $m$ are all arbitrary functions of the six independent 4-point scalars,
$\sess{12}$, $\sess{23}$, $\sess{34}$, \essdot{12}{23}, \essdot{23}{34}, \essdot{12}{34},
from which all other 4-point scalars can be constructed using the relations given in \app{notapp}.
We must now impose the rest of the Ward Identities on \es{G_{4}}, beginning with \ensuremath{D }. This implies that all the arbitrary functions $f$, $g$, $h$, $k$, $l$, $m$ are in fact of degree zero, or in other words are functions of five independent ratios of the scalars given above.
To impose \swk{1}{1} is an enormous calculation to perform by hand, and thus we use a computer algebra package, written in Mathematica, especially for the purposes of this calculation. First of all we operate with \swk{1}{1} on \es{G_{4}}, which is relatively straight forward. We shall not go into the details of how that was done in this paper, but we do discuss some of the details of the simplification of the resulting expression in \app{basapp}.
In particular, this appendix describes how one can define a set of canonical forms in terms of which all other expressions can be written. In this way, the computer can collect like terms and we can then separate out our results into a sum of linearly independent terms. The coefficients of these terms can then be set to zero individually, allowing us to restrict the form of the six arbitrary functions using the resulting equations.
After expanding the expression for $\swk{1}{1} \es{G_{4}}$ in terms of the basis described in \app{basapp}, we can use some of the resulting equations to show that,
\beq \begin{array}{rcl}
m &=& 0 \\
h &=& k + l.
\end{array} \eeq
After some algebra, it follows that we can rewrite the ansatz for \es{G_{4}}\ in a more symmetric way, as
\begin{eqnarray}
\es{
\es{G_{4}} &=&
\frac{f'}{\sess{12}} \sgsp{123} +
\frac{g'}{\sess{34}} \sgsp{234} +
\frac{k'}{\sess{12}} \sgsp{124} +
\frac{l'}{\sess{34}} \sgsp{134},
}
\end{eqnarray}
where $f'$, $g'$, $k'$, $l'$ are degree zero functions of \es{\sess{12}, \sess{23},\sess{34}, \sess{13}, \sess{24}, \sess{14}}\ (see \eqn{essdotexp}). In addition, the rest of the equations imply that we can further restrict the form of these functions so that their arbitrariness comes only from their dependence on the four-point cross ratios $u$ and $v$, defined in \eqn{uvdefs}.
We find
\beq \begin{array}{rcl}
\eqal{ f'(\es{\sess{12}, \sess{23},\sess{34}, \sess{13}, \sess{24}, \sess{14}}) }{=}{
Q_0 \left( \frac{\sess{13}}{\sess{34}} \right)^3 \left( \frac{\sess{24}}{\sess{14}} \right)^2 f''(u,v)}
\eqal{ g'(\es{\sess{12}, \sess{23},\sess{34}, \sess{13}, \sess{24}, \sess{14}}) }{=}{
Q_0 \left( \frac{\sess{13}}{\sess{14}} \right)^2 g''(u,v)}
\eqal{ k'(\es{\sess{12}, \sess{23},\sess{34}, \sess{13}, \sess{24}, \sess{14}}) }{=}{
Q_0 \left( \frac{\sess{13}}{\sess{34}} \right)^3 \left( \frac{\sess{24}}{\sess{14}} \right)^2 k''(u,v)}
\leqal{ l'(\es{\sess{12}, \sess{23},\sess{34}, \sess{13}, \sess{24}, \sess{14}}) }{=}{
Q_0 \left( \frac{\sess{13}}{\sess{14}} \right)^2 l''(u,v)},
\end{array} \eeq
where
\begin{eqnarray}
Q_0 &=& \frac{
\left( \sess{34} \right)^{3\,\left( q_{1} + q_{2} \right)}
\left( \sess{14} \right)^{3\,\left( q_{2} + q_{3} \right)}
}{
\left( \sess{13} \right)^{3\,\left( q_{1} + q_{2} + q_{3} \right)}
\left( \sess{24} \right)^{3\, q_{2}}
}.
\end{eqnarray}
Further restrictions can then be found by imposing \swqb{1} on \es{G_{4}}, in addition to the above restrictions from \es{K}.
The result is the following set of equations for the undetermined functions of $u,v$,
\input{grneqns}
where, for clarity, we have dropped the double primes for the rest of the discussion and
\beq \begin{array}{rcl}
x &=& 1 + u - v \\
y &=& 1 - u - v \\
z &=& 1 - u + v \\
w &=& 1 - 2\,u + {u^2} - 2\,v - 2\,u\,v + {v^2}.
\end{array} \eeq
It is immediately apparent that by specifying four constants, which are the values of $f$, $g$, $k$ and $l$ at some point $u_0, v_0$, we can determine all first derivatives using Equations \eqnnum{qf} - \eqnnum{ql}. By differentiation we can determine all second derivatives in terms of known lower derivatives at $u_0, v_0$ and thus we can determine all higher derivatives at this point by repeating this process indefinitely. Consequently we can construct the solution around $u_0, v_0$ as an infinite Taylor expansion in $u,v$, and thus the solution is uniquely specified by these four constants.
Of course one could instead try to solve the above equations by imposing integrability relations, such as
\begin{eqnarray}
\frac{\partial^{2}}{\partial u \partial v} f &=& \frac{\partial^{2}}{\partial v \partial u} f .
\end{eqnarray}
By calculating $\frac{\partial^{2}}{\partial v \partial u} f $ in two different ways from \eqnnum{qf} and equating these expressions, we ought to find a relationship between $f$, $g$, $k$, $l$ and their first derivatives. Substituting for the first derivatives using Equations \eqnnum{qf} - \eqnnum{ql}, we should get a relationship between $f$, $k$, $g$ and $l$ alone. However, we find in all cases that this is simply the trivial statement that $0=0$. This implies that the equations are integrable in the given form and thus in order to solve them one must specify the values of $f$, $g$, $k$ and $l$ at some initial point, as above. In other words, the integrability relations do not yield further relationships which could be used to reduce the number of constants which have to be specified to obtain a solution.
An alternative approach is to consider Equations \eqnnum{ql}, and rewrite them to make $k$ and $g$ the subjects, as in
\beq \begin{array}{rcl}
\eqal{\!\!\!\!\!\!\! {g\,q_{1}} }{=}{
{{{u\,v\,\left( -1 + u + v \right) \,\partial_{v}l_{2} + 2\,{u^2}\,v\,\partial_{u}l_{2} +
3\,u\,v\,l_{2}\,q_{2} + 3\,f_{2}\,q_{4}}\over {3\,u}}} }
\leqal{\!\!\!\!\!\!\! {k\,q_{3}} }{=}{
{{{2\,{u^2}\,v\,\partial_{v}l_{2} + {u^2}\,\left( -1 + u + v \right) \,\partial_{u}l_{2} +
3\,{u^2}\,l_{2}\,q_{2} + 3\,f_{2}\,q_{4}}
\over 3}}
}
\end{array} \eeq
We then substitute these expressions into Eqs. \eqnnum{qf} to eliminate $k$ and $g$ completely, giving
\beq \begin{array}{rcl}
\eqal{ {\partial_{u}f} }{=}{
{{{{u^2}\,v\,\partial_{v}l +
\left( 3 - 3\,q_{1} - 3\,q_{2} \right)\, f }\over u}} }
\leqal{
{\partial_{v}f} }{=}{
{{{{u^2}\,v\,\partial_{u}l +
\left( -1 + 3\,q_{1} + 3\,q_{4} \right)\, f }\over v}}.
}
\elab{qfprime}
\end{array} \eeq
Note that at no time do we divide by any function of $q_{i}$ as without knowing \ensuremath{\mathbf{q} }, we cannot be sure that such a function is non-zero.
Differentiating the first of Eqs. \eqnnum{qfprime} with respect to $v$ and the second with respect to $u$, we use the integrability condition
\begin{eqnarray}
\frac{\partial^{2}}{\partial u \partial v} f &=& \frac{\partial^{2}}{\partial v \partial u} f
\end{eqnarray}
to combine the two equations. We then use Eqs. \eqnnum{qfprime} to remove the derivatives of $f$ from the resulting expression, giving an equation in $l$ alone,
\begin{eqnarray}
\!\!\!\! v\,\partial_{{v^2}}l - u\,\partial_{{u^2}}l +
\left( 1 - 3q_{1} - 3q_{2} \right) \, \partial_{u}l +
\left( 2 - 3q_{1} - 3q_{4} \right) \, \partial_{v}l &=& 0,
\elab{lone}
\end{eqnarray}
where
\begin{eqnarray}
\partial_{u^n v^m} l &\equiv& \frac{\partial^{n+m}}{\partial u^n \partial v^m}.
\end{eqnarray}
To do this, we only used Equations \eqnnum{ql} and \eqnnum{qf}, and thus we could repeat the process using Equations \eqnnum{ql} and \eqnnum{qg} or Equations \eqnnum{ql} and \eqnnum{qk} to obtain other equations in $l(u,v)$ alone. Of the three second order linear partial differential equations which result, only two are linearly independent. The other can be written as
\beq \begin{array}{rcl}
\left( -1 + v \right)\, v \,\partial_{{v^2}}l + 2\, u v \,\partial_{u v}l +
{u^2}\,\partial_{{u^2}}l + 3\, u \,
\left( q_{1} + 2 q_{2} + q_{3} \right) \,\partial_{u}l + & & \\
3 \, q_{2}\,\left( 2 - 3 q_{4} \right)\, l +
\left( 1 - 3\,q_{2} - 3\,q_{3} + 3\,v\,\left( 1 + q_{2} - q_{4} \right) \right)\, \partial_{v}l &=& 0
\elab{ltwo}
\end{array} \eeq
To investigate these simultaneous equations we proceed as follows.\footnote{We wish to thank Thomas Wolf and Allan Wittkopf for their help with the following argument.}
Once again, we continue to try to impose higher order integrability relations, such as
\begin{eqnarray}
\frac{\partial^{3} l}{\partial u^2 \partial v} &=& \frac{\partial^{3} l}{\partial v \partial u^2} \,\,\,\,\,\, etc,
\end{eqnarray}
by differentiating the above pde's for $l$. We make $\partial_{u^2} l$ the subject of \eqn{lone} and $\partial_{uv} l$ the subject of \eqn{ltwo}. Differentiating \eqn{lone} with respect to $v$ and \eqn{ltwo} with respect to $u$ and eliminating $\partial_{u^2 v} l$, we obtain three differential equations for \[ \partial_{uv} l, \,\,\,\, \partial_{u^2} l, \,\,\,\, \partial_{v^3} l \] (in terms of only $l, \partial_{u} l, \partial_{v} l, \partial_{v^2} l$), from which all other higher derivatives can be obtained. That is to say, we have a system of pde's whose integrability conditions are simply identities which follow as a consequence of these three equations alone and thus impose no further restrictions on $l(u,v)$. This corresponds to the statement above where the complete solution is determined by four independent arbitrary constants.
In this approach one can construct complete solutions given a function $l$ which satisfies the given differential equations. These solutions are not in the form of Taylor expansions and, as an illustration, we have explicitly constructed two distinct solutions for a given \ensuremath{\mathbf{q} }. For example, if we take
\begin{eqnarray}
\ensuremath{\mathbf{q} } &=& (-1/3, 0, q_3, 4/3 - q_3)
\end{eqnarray}
we find that
\beq \begin{array}{rcl}
k &=& \frac{4 - 3q_3}{3 q_3} u^4 v^{(2-3q_3)} \\
g &=& (3q_3 - 4) u^4 v^{(2-3q_3)} \\
f &=& u^4 v^{(2-3q_3)} \\
l &=& \mathrm{constant,}
\end{array} \eeq
is a solution, and so is
\beq \begin{array}{rcl}
k &=& (2 + 2 u - 3 q_3 u - 2 v) (2 - 3q_3) v^{(2-3q_3)} \\
f &=& - \frac{1}{2} (2 - 3q_3) u^2 v^{(2-3q_3)} \\
g &=& \frac{1}{2} (2 - 3q_3) u^2 v^{(2-3q_3)} \\
l &=& v^{(2 - 3q_3)}.
\end{array} \eeq
Clearly these two solutions have the same \ensuremath{\mathbf{q} }, but they differ in the values of the four constants which determine the particular form of the solution. One can construct similar examples for different choices of \ensuremath{\mathbf{q} }.
\section{Conclusions}
In this paper we have found the most general three-point Green's function
for $N= 1,2,3,4$ supersymmetry which is composed of
chiral superfields of a given chirality. We have also shown that although there exist no chiral superconformal invariants, \bibnum{l9scfpap8} the Green's functions of chiral superfields are not uniquely specified above three-points when the the $R$-charge, \ensuremath{{q_{0}}}, is greater than $N$. This result relies crucially on the nilpotent character of such Green's functions. However, for the particular case $\ensuremath{{q_{0}}} = 1, N = 1$ we have shown that the solutions are unique up to the specification of four constants of integration. We have given two equivalent formulations of the solution which should enable one to explicitly construct the solution in any particular case.
The results of our investigations seem to suggest that our findings should generalise to extended supersymmetry, where we expect to find a unique solution at four-points in the case where $\ensuremath{{q_{0}}} = N$. Furthermore, it is tempting to suggest that all higher-point functions may also be uniquely determined in the case where $\ensuremath{{q_{0}}} = N$. At the moment, however, these two statements remain conjectures based upon our explicit results for the $N=1$ four-point solution at $\ensuremath{{q_{0}}} = N$ and for the three-point solutions for $N \geq 1$.
\section{Acknowledgements}
We wish to thank Paul Howe for useful discussions. We also wish to thank Alain Moussiaux at CONVODE and Edgardo S. Cheb-Terrab for advice on the solution of partial differential equations. In particular, we wish to thank Thomas Wolf and Allan Wittkopf for help with their pde solving packages CRACK and RIF, which eventually solved the system of pde's discussed in the paper. AP wishes to thank D.R.T. Jones for discussions and PPARC for a research fellowship.
\section{Note Added}
After this work had been completed a preprint \bibnum{parktwo} appeared on the hep-th archive which also discusses Green's functions in superconformal field theories.
|
\section{Introduction}
\label{sec:intro}
The Mid-Infrared Galaxy Atlas (MIGA) is a mid-infrared (12 and
25~$\mu$m) atlas of part of the Galactic plane ($75^\circ < l <
148^\circ, b = \pm6^\circ$). It was constructed
using IRAS data processed to approximately $0.5'$ resolution using the
HIRES image construction process (\cite{aum90}) including a new point
source ringing suppression algorithm (\cite{cao99}). Parts
of the MIGA along with the far-infrared (60 and 100 $\mu$m)
IRAS Galaxy Atlas (IGA; \cite{cao97}) are being merged with
radio and millimetre data as part of the Canadian Galactic Plane
Survey (\cite{gps98}; CGPS\footnote{Current information on the CGPS
can be found at http://www.ras.ucalgary.ca/CGPS/}), a project to
survey about a quadrant of the Galactic plane at $\leq 1'$
resolution over a wide range of wavelengths (12 $\mu$m -- 190 cm) to
study all of the major components of the interstellar medium (ISM).
The addition of a mid-infrared data set to this data base is important
since one of the goals of the CGPS is to understand the evolution
of dust as it moves through different phases of the ISM. The 25 and
12 micron bands of IRAS have been shown to be good tracers of
the smallest dust particles: the chemically uncharacterized very small grains
(VSG's) and the large carbonaceous molecules, most likely polycyclic
aromatic hydrocarbons (PAH's) (\cite{ona96}), respectively. Therefore, the
MIGA will enable the study of the emission from the smallest
components of interstellar dust at an angular resolution
comparable to that of the complementary data being used to define
different physical environments.
This paper has been designed to be complementary to the paper
describing the IGA (\cite{cao97}). Many of the details of the image
construction algorithms for the MIGA and the IGA are identical and were
described in great detail in a series of papers related to both the
IGA and the parallelization of the HIRES code (\cite{cao97};
\cite{cao96}); thus they will not be repeated here. Rather we
concentrate on pointing out those areas where the MIGA and IGA differ
(most importantly in resolution behaviour and the response to point sources)
and on demonstrating through a series of tests on the data that the
MIGA is a data set of comparable quality to the IGA. This paper, combined
with the papers describing the IGA, gives a complete
guide to the infrared data sets that will be made available to the
general astronomical community as part of the CGPS data releases over
the next few years. The information provided in this paper is relevant
to both MIGA mosaiced images included in the CGPS and stand-alone MIGA images.
In \S~\ref{sec:atlas} we describe the format of the atlas along
with the format of an extension to the IGA (EIGA) that was constructed
specifically for the CGPS. The steps involved in MIGA processing are
outlined in \S~\ref{sec:processing}. Section \ref{sec:character}
describes the characteristics of MIGA images. In
\S~\ref{sec:artifacts} various artifacts of the images are discussed,
including the reduced point source ringing. Finally, sample images are
shown in \S~\ref{sec:images} and future directions for the MIGA
and large-scale HIRES processing are discussed in \S~\ref{sec:sumfut}.
\section{Description of the Atlas --- MIGA and EIGA}
\label{sec:atlas}
The atlas covers a twelve degree wide strip ($b = \pm6^\circ$) of the
Galactic plane from Cygnus to Cassiopeia ($75^\circ < l <
148^\circ$). The higher latitude limit, compared to the
IGA ($b = \pm4.7^\circ$), was required to match the MIGA with the CGPS
coverage which extends from $ -3.56^\circ < b < 5.56^\circ$ in order to
follow the main concentration of HI in this part of the Galaxy. With
this upper bound set, the lower boundary of the atlas was chosen to be
symmetric.
Each MIGA image covers a $1.4^\circ \times 1.4^\circ$ area and consists of
$1^{\mathrm st}$ and $20^{\mathrm th}$ iteration HIRES images along
with ten ancillary maps and tables as listed in Table~\ref{tbl:files}.
See Figures~\ref{fig:allimage_2} and~\ref{fig:allimage_1} for sample images.
The images are in Galactic cartesian (CAR) projection (\cite{gre96})
with a pixel size of $15''$. MIGA images are quite flexible; users
can make seamless mosaics of arbitrary size (see \S~\ref{sec:images})
and rebin and reproject the images as required. We hope to make the
full MIGA available via the web in a similar manner to the existing IGA web
server\footnote{http://irsa.ipac.caltech.edu/applications/IGA/}.
Part of the MIGA, a series of $5.12^\circ \times 5.12^\circ$
mosaics, will be available
through the Canadian Astronomy Data Centre (CADC) as part of the
general public release of the CGPS data. The CGPS mosaics have a
pixel size of $18''$, and so the MIGA and IGA
images were slightly rebinned in creating these mosaics.
The CGPS mosaics will be a very useful way to access both the MIGA and
IGA data, particularly since users will have immediate access to
complementary radio and millimetre data on a uniform grid covering the
same region. Tests done on regular and rebinned MIGA images show that
there is little reduction in the quality of the images due to the
slight rebinning. The general image characteristics described here
for the MIGA, and in Cao et al.\ (1997) for the IGA, apply equally to
the MIGA and IGA in the CGPS mosaic format.
Since the IGA has a high-latitude cutoff of $b = 4.7^\circ$ the
initial CGPS mosaics constructed using the IGA had a blank strip at
high latitudes. In order to match the infrared coverage of the CGPS
with the radio coverage we constructed a high-latitude extension to
the IGA covering $75^\circ < l <148^\circ$
and $ 4.7^\circ < b < 5.56^\circ$. Agreement between the original IGA
images and the EIGA is excellent and the EIGA has been incorporated
into all of the CGPS far-infrared mosaics (see \S~\ref{sec:sumfut}).
We have also constructed a low latitude extension to the IGA from $
-6.0^\circ < b < -4.7^\circ$ to match the IGA and MIGA coverage.
This ability to extend the infrared images of the IGA and the MIGA to
higher/lower Galactic latitudes is an important reason why HIRES and
the IRAS data base, with its almost full-sky coverage, remains an
important tool to study the infrared sky. This will be discussed more
in \S~\ref{sec:sumfut}.
\section{Description of Processing}
\label{sec:processing}
The basic processing of the MIGA (and EIGA) follows the same steps as
discussed in detail in Cao et al.\ (1997). The raw IRAS archive data,
known as CRDD (Calibrated Reconstructed Detector Data) were first sent from
IPAC to the Canadian Institute for Theoretical Astrophysics (CITA) to
allow processing to be done locally. The raw data are
uncompressed and formatted using the programs SnipScan and LAUNDR.
The data are processed in $7^\circ \times 7^\circ$ sections known as
CRDD plates.
Infrared Sky Survey Atlas (ISSA) images corresponding to the plate region are
mosaiced together and used to calibrate the IRAS data using the SmLAUN
program. This step effectively removes the zodiacal emission since
ISSA images have a zodiacal light model subtracted from them. ISSA
mosaicing was done in advance of the other preprocessing steps to
allow the quality of the mosaicing to be checked before use
(see \S~\ref{sec:artifacts}). The data are then reprojected from
equatorial coordinates to Galactic coordinates using the BrkDet
program. At this stage the data are in $1.4^\circ \times 1.4^\circ$
sections spaced every $1^\circ$. All of the preprocessing steps were
done on a Sparc Ultra workstation. For a single CRDD plate
preprocessing took on average about two hours of wall clock time to
complete in addition to the time required to construct and check the
ISSA mosaics. To cover the CGPS region 63 CRDD plates in two bands
had to be processed ($\sim240$ hours).
These data are then HIRES processed to create the final images and ancillary
maps. Unlike the IGA, we used a non-parallel version of the
HIRES code on a SGI Origin 2000 computer at CITA. The
original HIRES code was first modified to run on the SGI architecture and
tests were made comparing IGA release images to 60 and 100 $\mu$m
images constructed at CITA. IGA images were recreated for a field at
$l = 152^\circ, b = -1^\circ$ using the new code and no differences
were seen, beyond that expected for numerical round-off errors:
maximum fractional differences were on the order of $10^{-5}$ and average
fractional differences were on the order of $10^{-8}$.
The construction of the EIGA gave us another chance to
test the new code for compatibility with the IGA production
code. As shown in \S~\ref{sec:sumfut} the match between the IGA images
and the EIGA images produced at CITA is excellent.
Tests showed that processing a single $1.4^\circ \times 1.4^\circ$
region at a given wavelength
took approximately 4.5 minutes of wall clock time. While this is clearly
slower than the processing times reported in Cao et al.\ (1996) for the
parallel-processing machines, it is a vast improvement over the
single-processor times they report. Since we were primarily interested
in covering the CGPS survey region which contains 444 $1.4^\circ \times
1.4^\circ$ areas, and considering other overheads in the production,
this processing speed was adequate. It took 67 hours of wall clock
time to process the entire CGPS region at both 12 and 25 $\mu$m (about a
quarter of the time required to prepare the CRDD data on the Sparc Ultra).
The HIRES code we used also took advantage of a ringing suppression
algorithm that was developed after the IGA was in its production run
(\cite{cao99}). Tests were done before producing the MIGA to compare
the quality of the images with and without the ringing suppression
algorithm in place. Differences between the two images were negligible
away from point sources. In \S~\ref{sec:rings} we discuss the
effect of the ringing suppression algorithm on point sources in more detail.
\section {Characteristics of the Images}
\label{sec:character}
In this section of the paper we describe the resolution, photometric
accuracy, positional accuracy, surface brightness accuracy, and the mosaic
property of the MIGA.
\subsection{Resolution}
\label{sec:resolution}
The basic angular size of the rectangular IRAS detectors is $45''
\times 270''$ at 12 $\mu$m and $45'' \times 276''$ at 25 $\mu$m
(\cite{aum90}). The higher resolution is obtained along the scan
direction of the satellite. As a result of different scan orientations
the position angle of the elongated raw beam varies across the
sky. Improved resolution is possible because each region is covered by
overlapping scans and was usually revisited with a different scan
orientation (see Figure~\ref{fig:allimage_1}).
In order to assess the achieved resolution of the MIGA within each
field simulated beams (PSFs) are constructed using the HIRES IRAS
Simulator mode (\cite{fow94}). Spikes are placed in a regular grid on
a smoothed version of the 20$^{\mathrm th}$ image. As discussed in
Cao et al.\ (1997), the image histogram is used to scale the spikes to an
intensity that represents a point source that is bright enough,
relative to the background emission, that HIRES processing is beneficial.
The image is then scanned with the detector pattern to produce simulated
IRAS data. These data are then regularly HIRES processed to produce
the img\_*bem* maps that show the IRAS beam shape (see
Figure~\ref{fig:allimage_2}). As will be discussed below, it is
important to note that while the results of a 2-D Gaussian fit to
these beams does give a measure of the resolution (*fwhm.txt files),
the actual beam shapes are not 2-D Gaussians even when the beams are
not X-shaped due to large differences in the scan directions
(\cite{ric93}, \cite{mos92}).
To quantify the MIGA resolution we sampled twelve fields scattered
across the atlas region, using the FWHM of 2-D Gaussian fits to
simulated beams as a measure of the achieved resolution.
The results of this test are shown in Table~\ref{tbl:restest} where we
report the mean and standard deviation about the mean for the 49 beams
in each field. The average resolution for the twelve test fields is
$33'' \times 67''$ and $34'' \times 66''$ at 12 and 25 $\mu$m
respectively. This should be compared with the typical
full-resolution co-add (FRESCO; equivalent to 1$^{\mathrm st}$
iteration MIGA and IGA images) resolution of $1' \times 5'$ in both
wavebands or the standardized ISSA resolution of $4' \times 5'$ in all
wavebands. Note that the pixel size of 15$''$ just adequately samples
the beam in the scan direction.
One important thing to notice is the similarity of the resolution at
12 and 25 $\mu$m. Within any given field the resolution may vary
from place to place, but the 12 and 25 $\mu$m resolutions are always
very close in value. It is also important to note that the position
angles of the beams are also very close; in a test of 756 simulated
12 and 25 $\mu$m beams the difference between the position angles was
on average 0.07$^\circ$. In Figure~\ref{fig:rescomp} we show the FWHM
fits to beams on one of the MIGA fields of Table~\ref{tbl:restest}
along with the same data for the corresponding IGA field. MIGA has a
better resolution than the IGA. The resolution between the two
MIGA bands is very well matched at all locations in the image. The
situation is quite different for the 60 and 100 $\mu$m bands of the
IGA where the resolution and the position angle of the beams varies
considerably between the two bands. In a test of 756 simulated 100
and 60 $\mu$m beams the difference between the position angles was on average
7$^\circ$. The reason for this behaviour is that the mid-infrared
detectors on the IRAS focal plane have the same physical size, whereas
the sizes of the far-infrared detectors differ, and that the mid-infrared
detectors are packed more closely together in the focal plane than
the far-infrared detectors so that their scan pattern on the sky is more
similar (see the IRAS Explanatory Supplement [1988] for details on the
IRAS detectors and the layout of the focal plane).
The resolution achieved in HIRES processing is a function of the
number of iterations, the coverage pattern, and the strength of the
point source relative to the background value \emph{during
processing}. The latter two factors are what cause the scatter in
resolution seen at a given wavelength in Figure~\ref{fig:rescomp}. The
best resolution, for a given coverage, is achieved for a high ratio of
point source strength to background which is why a bias level is
applied to the image during HIRES processing to bring the background
level of the image as close to zero as possible (see \S~3.5 of Cao et
al.\ [1997] for more details regarding the calculation of the flux
bias). The bias level that is applied to a given MIGA image is
reported in the image header as a flux in units of W m$^{-2}$. This
value can be converted to an intensity (in Jy sr$^{-1}$) by dividing
the value by the average detector solid angle ($3.2\times10^{-7}$ sr
at 12 $\mu$m and $3.5\times10^{-7}$ sr at 25 $\mu$m [\cite{mos92}])
and by a factor that accounts for the IRAS bandpass shape
($1.348\times10^{-13}$ and $5.16\times10^{-14}$ (Hz) at 12 and 25 $\mu$m
respectively).
In order to examine the variation of resolution with changing point
source to local processing background ratio (PS/BG) for the MIGA we created
simulated beam maps for the $1.4^{\circ} \times 1.4^{\circ}$ field centered at
$l=74^{\circ}$, $b=-6^{\circ}$ using unscaled point sources with
fluxes of 0.05 to 10000 Jy. In this region the average background
level was fairly low, 2.32 MJy sr$^{-1}$ and 6.43 MJy sr$^{-1}$ at
12 $\mu$m and 25 $\mu$m respectively. These values can be converted
to a flux using the average detector solid angles yielding average
background fluxes of 0.74 and 2.2 Jy respectively. To simplify the
test we did not apply a flux bias during processing and so these were
the background values for HIRES processing. The results of this test
are shown in Figure~\ref{fig:psres}. These two plots illustrate that
the resolution of a point source is not dependent on just the strength
of the source but the ratio PS/BG. In the upper plot one sees that for most of
the point sources in this test the resolution achieved at 12 $\mu$m is
slightly better than at 25 $\mu$m, but this is only because of the
lower processing background used at 12 $\mu$m. In the lower panel we
illustrate this directly by plotting the same resolution measurements
for both 12 and 25 $\mu$m against PS/BG. All of the data follow the
expected trend of increasing resolution with increasing point source
strength to background ratio.
In Figure~\ref{fig:psbg} we replot the data from
Figure~\ref{fig:rescomp} now showing the FWHM achieved as a function of
PS/BG. The difference in resolution between the far and
mid-infrared IRAS bands, and the match in resolution between the 12
and 25 $\mu$m bands, holds over a wide range in
PS/BG. Figure~\ref{fig:psbg} also shows that the trend of increasing
resolution with increasing PS/BG is very flat, so that point
sources with a range of PS/BG will still have similar
achieved resolutions at 12 and 25 $\mu$m. Although the beam
simulations are done with a realistic assessment of the processing
background for the actual image, still the resolutions reported in the
*fwhm.txt files are for a particular injected point source and so are
only representative of the resolution in the actual images.
\subsection{Ratio Maps and Cross-beam Simulation}
\label{sec:xbs}
This similarity in resolution between the two bands means that, with
care, high resolution ratio maps can be created directly using the
MIGA images. Care is required because, although the FWHM fits to the
simulated beams are very close, the actual beam shapes are not 2-D
Gaussians and can vary in detail because of the actual PS/BG in the
two images. As an example, Figure~\ref{fig:ratiomaps} shows two ratio
maps of the region around the HII region S151, chosen because of the
unusual irregular cross-shaped beam pattern caused by the significant
difference in the scan angle between the two IRAS coverages of the
region (this effect is more noticeable at high ecliptic latitudes). The first
image was constructed by simply dividing the 12 $\mu$m MIGA image by
the 25 $\mu$m MIGA image.
The other ratio map was constructed using a technique called
cross-band simulation (\cite{fow94}). This technique makes use
of the HIRES IRAS simulator mode. First the simulator scans the 12
$\mu$m HIRES image with the 25 $\mu$m detector pattern to create a
simulated view of the 12 $\mu$m sky. These ``observations'' are then
regularly HIRES processed to create a somewhat lower resolution
version of the 12~$\mu$m image. The same process is followed for the
25 $\mu$m data. The final result is two images at the same resolution
(slightly poorer than the original 25 $\mu$m) with almost identical
beam shapes. The effect of bringing the two wavebands to the same beam
shape is seen in the way the point sources (cross-shaped beam pattern)
in Figure~\ref{fig:ratiomaps} are undistorted in the cross-beam
simulator image while the point sources show some non-physical
structure in the original.
Note that the cross-beam simulation does not compensate for differing
PS/BG that may occur between the two bands. For the most demanding
work, say in examining a particular part of an image with vastly
different PS/BG, flux biases could be chose to better match PS/BG locally.
However, for some purposes even the simple MIGA ratio map may suffice.
The user should closely inspect the simulated beam maps to determine
if the beam shapes are close enough for their purposes. If so, the common
resolution of the MIGA provides a quick and easy way to obtain high resolution
mid-infrared ratio maps. Note that any simple ratio map involving the
IGA will not be as satisfactory, and cross-beam simulation will be
desirable. Therefore we are developing an algorithm to do this using
(M)IGA and its ancillary data, rather than having to return to the raw
IRAS data.
\subsection{Photometry}
\label{sec:photometry}
In order to test the photometric accuracy achieved in the MIGA images
we selected 52 point sources scattered across the range of the atlas.
Sources selected were bright ($>10$ Jy in each band), isolated, and
resolved (as indicated by the point source correlation coefficient in
the IRAS Point Source Catalog (PSC)). While the Maximum-Correlation
Method (MCM; \cite{aum90}) algorithm at the heart of HIRES conserves
flux globally the flux can be redistributed across the image with each
iteration causing changes in the measured flux at a given location.
Photometry was done using a script driving the IPAC Skyview
program\footnote{Skyview is a general purpose FITS image viewing and
analysis tool available at http://www.ipac.caltech.edu/Skyview/}.
Circular apertures of radius $5'$ and $7'$ were drawn around each
point source and a background surface brightness was defined using the
average of twenty points evenly distributed throughout the annulus
between the two circles. Flux values were calculated using the two
apertures and compared. If there was a large discrepancy between the
two values then the background surface brightness was too variable and the
point source was rejected from the sample. Otherwise the average of
the two fluxes was used in the comparison with the PSC. The results
of the photometry are tabulated in Table~\ref{tbl:photometry} and
average values, along with the standard deviation about the average,
are listed in Table~\ref{tbl:avgphoto}.
Both wavelength bands experience the same general trend. There is an
average 6\% positive offset from the PSC at the $1^{\mathrm st}$
iteration and negative 4\% offset after 20 iterations. This
decrease in point-source flux with iteration is not a universal
property of HIRES processing. Cao et al. (1997) report for the IGA,
using a sample of 35 point sources, that at 60 $\mu$m, on average
there are 12\% and 14\% offsets from the PSC after 1 and 20 iterations
respectively, and at 100 $\mu$m the offsets are 1\% and 11\%. In the
IGA case part of the trend is thought to come from a systematic
decrease in the background attributed to increased point-source
ringing, and so one possible cause of the trend observed for the MIGA
might be a systematic increase of the background level with increasing
iterations. However, as shown in Table~\ref{tbl:avgphoto} the
average background value actually decreases very slightly. Also since
we do not see this decrease occurring in every single point source
tested (e.g., PSC~20282+3604 or PSC~23239+5754) it is not a universal
property of the algorithm being used.
Since this behaviour is different than that seen in the IGA, we
compared a subset of 16 point sources processed using the
ringing reduction algorithm and without (as for the IGA). The results
are shown in Table~\ref{tbl:photoringtest} for the 20$^{\mathrm th}$
iteration images. The 1$^{\mathrm st}$ iteration images are
identical in both cases and were remeasured to gain an idea of the
uncertainties involved in the photometric measuring technique being
used. Differences in the measured fluxes were $<0.1\%$; of course a more
sophisticated photometry routine would reduce this
uncertainty further. In general the fluxes measured without the
ringing suppression algorithm in place are higher than the fluxes from
the MIGA images.
Average results for this test are shown in
Table~\ref{tbl:avgphotoringtest}. Without the ringing suppression
algorithm the point source flux tends to increase as the iterations
increase, like the IGA though less dramatically, whereas with the
algorithm in place the point source flux
tends to decrease with increasing iterations as before. However, this behaviour
is not found for every single source; if the point source flux does
increase for a MIGA source, then it tends to increase less than
it does when the ringing suppression algorithm is not used, thus
preserving the sense of the relative behaviour.
While the photometry obtained without the ringing suppression
algorithm in place tends to match the PSC flux values better, we
decided that the possible benefits of having the ringing suppression
algorithm in place outweighed the slightly worse (but still comparing well
to the IGA) photometric performance. Furthermore, as discussed in
\S~\ref{sec:fluxerr} there are other sources of error in flux measurement
that will tend to swamp this uncertainty.
\subsection{Size-Dependent Flux Calibration}
\label{sec:acdc}
One quirk of the IRAS detectors was that their sensitivity was a
function of the dwell-time of a source: the so-called AC/DC effect.
Due to this behaviour two calibrations for IRAS data were developed.
The AC calibration, used for the MIGA and IGA, is suitable for point
sources. The DC calibration is suitable for measuring fluxes from
extended emission $>2^{\circ}$ in extent (\cite{whe94}). For structure
at intermediate scales ($6'$ -- $2^{\circ}$) well defined conversion
factors exist for the mid-infrared IRAS bands and users should consult
Table II.B.1 in Wheelock et al.\ (1994).
In order to convert MIGA data to the DC scale, images need to be
multiplied by 0.78 and 0.82 at 12 and 25 $\mu$m, respectively. Unlike
the 60 and 100 $\mu$m bands, where the correction depends upon the
strength of the source (especially for very bright extended emission),
the AC/DC correction for the mid-infrared IRAS bands appears to be
more stable and does not appear to have any dependence on the source
brightness (IRAS Explanatory Supplement 1988).
\subsection{Flux and Surface Brightness Measurement Uncertainty}
\label{sec:fluxerr}
Section 2.3 of Fich \& Terebey (1996) gives a good general discussion
of the variety of factors involved in estimating uncertainties in IRAS
flux measurements, including instrumental, choice of measurement
technique, and background effects, and is recommended reading for users of both
the MIGA and IGA. It is clear that actual uncertainties in measured point
source flux values from the MIGA will be much higher than the basic
uncertainty of $\stackrel{<}{_\sim}6$\% implied by the photometric
tests against the PSC discussed in \S~\ref{sec:photometry}. Due to
complex background emission, uncertainties in flux measurements can
reach as high as 44\% and 20\% at 12 and 25 $\mu$m, respectively.
Instrumental uncertainties in the mid-infrared IRAS data are smaller than
uncertainties caused by uncertain background estimation and/or
different measurement techniques. As mentioned in \S~\ref{sec:acdc},
the AC/DC effect, although larger than at far-infrared wavelengths, is
well behaved. The absolute calibration between IRAS and DIRBE agrees
to 6\% at 12 $\mu$m and 1\% at 25 $\mu$m (\cite{whe94}), illustrating
how the absolute flux calibration for IRAS is well understood in the
mid-infrared.
Like for the IGA, the ISSA was used as a large scale surface brightness
truth table for the MIGA. In order to check this
calibration we selected five MIGA images from across the atlas region. The
images were reprojected and rebinned to the geometry and pixel size
of the ISSA and compared pixel-by-pixel with the respective ISSA
plate. The AC/DC correction was applied to the ISSA plate before the
comparison was made. The results of the test are tabulated in
Table~\ref{tbl:bright}, and data for one of the test areas are plotted in
Figure~\ref{fig:bright}.
Any
uncertainties inherent to the ISSA can be passed along to the MIGA. Of most
concern is the uncertainty in the zodiacal light model that was
subtracted from the ISSA. Residual zodiacal light removal errors are
3--5\% of the original emission (0.5 MJy sr$^{-1}$ at 12 $\mu$m and
1.0 MJy sr$^{-1}$ at 25 $\mu$m ) for $\beta > 50^\circ$, and are 1.0 MJy
sr$^{-1}$ at 12 $\mu$m and 2.0--2.5 MJy sr$^{-1}$ over scales of
$10^\circ$ for $50^\circ > \beta > 20^\circ$ (\cite{whe94}). Most of
the MIGA coverage is at high ecliptic latitude and so these residual
errors will be minimal.
\subsection{Positional Accuracy}
\label{sec:positions}
We tested the positional accuracy of the MIGA against the PSC using
the same point sources as for the photometry test. Using the IPAC
Skyview program a $5'$ radius circle was drawn around the position of
the point source as given in the PSC (which is a weighted average of the
position at each wavelength), and the flux-weighted centroid
was then measured for the pixels within the circle. This value was taken
as the position of the point source from the MIGA.
The results of this test are shown in Table~\ref{tbl:pos}. At 12
$\mu$m the average measured distance from the PSC position was 7.88$''
\pm 4.0$ ($\pm 1\sigma$ scatter) and at 25 $\mu$m was 6.69$''\pm3.2$. The
positional accuracy of the MIGA is similar to that reported for the
IGA: 7.6$''\pm5.6$ at 60 $\mu$m and 7.1$''\pm4.1$ at 100 $\mu$m. The
position angle of the offset is different for the 12 and 25 $\mu$m
point sources, but no systematic trend was found. The position angle
differences are most likely due to a combination of the differences in
the detailed beam shape and the backgrounds in each band which causes the
flux-weighted centroid of the aperture to shift slightly between each band.
To investigate what effect rebinning the MIGA to produce CGPS mosaics
with $18''$ pixels has on the positional accuracy we repeated the
test on eleven point sources in the W5 region. The results are shown
in Table~\ref{tbl:pos15} and Table~\ref{tbl:pos18}. Both the original
MIGA and the CGPS mosaiced MIGA are in close agreement, with the
original MIGA agreeing only slightly better with the point source
catalog positions. At 12 $\mu$m the average distance from the PSC
positions was 6.89$''\pm2.8$ with 15$''$ pixels and 7.32$''\pm4.0$
with 18$''$ pixels. At 25 $\mu$m the average distance from the PSC
positions was 5.76$''\pm2.3$ and 6.57$''\pm3.5$ with 15$''$ pixels and
18$''$ pixels, respectively.
\subsection{Mosaic Property}
\label{sec:mosaic}
The large angular scale preprocessing of the IRAS data allows large
high-quality mosaics to be constructed from the final HIRES images. In
order to quantify the quality of the mosaics, which tend to be
seamless to the eye, we tested the mosaic property using data from four CRDD
plates. This allowed us to study 134 boundaries between images within
the plates and 19 boundaries across plates. The latter are expected to
be worse because they were preprocessed completely separately.
The tests were done by comparing the pixels along a one pixel
wide strip, one degree long, that is common between adjoining images.
A total of 32294 pixels was examined along boundaries contained
entirely within a single CRDD plate, and 4579 pixels were examined in
the cross-CRDD test.
The pixel ratios ($-1$) and the standard deviations are tabulated in
Tables~\ref{tbl:mosall12} and~\ref{tbl:mosall25} for the 12 and 25
$\mu$m data, respectively. In Table~\ref{tbl:moscomp} the standard deviations
are shown again along with data from the IGA, showing that the mosaic
quality decreases as one moves toward shorter wavelengths. This is
caused by a combination of increasing resolution and more complex
backgrounds making the images less smooth as one moves into the
mid-infrared.
\subsection{Residual Hysteresis}
\label{sec:hysteresis}
IRAS detectors experience a hysteresis effect after passing over a
bright source. While the Galactic plane shadowing effect described by
Cao et al.\ (1997) for the IGA does not affect the 12 and 25 $\mu$m
detectors, hysteresis also can cause bright point sources to have tails
associated with them. Although this can occur at any wavelength, it is
most prominent at 12 and 25 $\mu$m for point sources brighter than
15-20 Jy (\cite{whe94}). As shown in Figure~\ref{fig:pstail}, a single
source can have a number of tails, one for each scan direction.
Clearly this effect can cause difficulties in the interpretation of
structure near bright point sources, and additional care must be taken even
when doing photometry of bright point sources.
\section{Artifacts}
\label{sec:artifacts}
In the following paragraphs we will discuss briefly various processing
artifacts, namely stripes, glitches, coverage artifacts, and
discontinuities, of which users of the MIGA should be aware. Since the
processing of the MIGA closely followed that of the IGA, we refer the
reader to Cao et al.\ (1997) for a detailed discussion. The effect of a new
ringing suppression algorithm is treated in \S~\ref{sec:rings}.
The MIGA used the same destriping technique as in the IGA. Stripes,
which were once the most common artifact in images constructed from
IRAS data, are now almost completely eliminated from HIRES images.
Details on the destriping algorithm and the application to the IGA can
be found in Cao et al.\ (1996) and Cao et al.\ (1997).
A glitch is the term given to a cosmic-ray or trapped energetic
particle hit on the IRAS detectors that shows up in the IRAS data
stream. As with the IGA, the LAUNDR preprocessing
program was used to flag and remove most of the glitches. It is
possible that some glitches did slip through this stage of the
processing although none have been identified so far. For an
example of a glitch artifact in the IGA see Figure 12 of Cao et al.\ (1997)
(glitches in the MIGA [and EIGA] would have the same properties).
Regions of low IRAS detector coverage can
cause spurious structure to appear in HIRES images. Low coverage or
a steep gradient in the coverage can also cause point source positions
to shift. Users of any HIRES product can use the coverage maps
(cvg\_*, see Table~\ref{tbl:files} and Figure~\ref{fig:allimage_1}) to help
determine if observed features could be affected or even caused by
regions of low coverage.
Discontinuities can occasionally exist entirely within a MIGA image or
mosaic as opposed to across image boundaries. These discontinuities
trace their origin to the preprocessing step involving the ISSA images (for
calibration and zodiacal emission removal). ISSA data are required corresponding to
the geometry of the input CRDD plate and so mosaics of ISSA
images were constructed if necessary. Care was taken to minimize the
discontinuities between ISSA images but in some cases a small (on the order of
1 MJy sr$^{-1}$)
discontinuity was unavoidable. This type of discontinuity is usually
not visible in the first iteration image, but is sharpened by the
HIRES algorithm and becomes visible in the twentieth iteration
image. An example of this type of discontinuity is given in Figure 13
of Cao et al.\ (1997).
\subsection{Ringing}
\label{sec:rings}
We were able to apply a ringing suppression algorithm based on Burg
entropy minimization (\cite{cao99}) to the MIGA data. Ringing is still visible
around point sources but the level of the ringing is significantly
reduced. Comparison between images constructed with and without the
algorithm showed little difference in their global properties. For
example, Table~\ref{tbl:bnbbright} shows results from a surface
brightness accuracy test like the tests vs. ISSA presented in
\S~\ref{sec:fluxerr}. In this case we examined five different 10$'$
radius circular apertures at different locations across a 12 $\mu$m
image. Since these tests showed that the images are virtually
indistinguishable from regularly processed HIRES images away from
point sources, the point source photometry tests in
\S~\ref{sec:photometry} were acceptable (see
Tables~\ref{tbl:photoringtest} and~\ref{tbl:avgphotoringtest}), and
the algorithm had been shown to be useful in at least one published
study (\cite{nor97}), the ringing suppression algorithm was adopted
for MIGA processing.
To quantify the beneficial effect of the ringing suppression algorithm on point
sources in MIGA we processed five regions without the ringing
suppression algorithm on (WRS) and selected ten well defined, isolated
point sources. Cuts were then taken across the point sources and the
depth of the ringing on either side of each source was measured
relative to the local background on either side. The change in the
depth of the ringing was then calculated. The effect of the algorithm
is generally to reduce the depth of the rings in every case.
Figure~\ref{fig:psrr} illustrates the effect of the algorithm on three
of the point sources at 12 and 25 $\mu$m. There is also an increase
the peak brightness (not shown). Note the excellent agreement between
the two images as one moves away from the rings.
We found that as the flux of the point source increases the ringing
tends to become more severe, as might be expected. At the same
time the absolute value of the change in the ring depth also increases for
stronger point sources. The result is that the fractional reduction in
the ringing does not exhibit any trend with point source flux. The
results of this calculation, along with data on the point source
fluxes, are summarized in Table~\ref{tbl:psrr}. For the point sources
examined, the average value of ring depth (MIGA/WRS) was
$0.49\pm0.2$ ($\pm1\sigma$ scatter) and $0.45\pm0.2$ at 12 $\mu$m and
25 $\mu$m respectively.
\section {Sample Images}
\label{sec:images}
In this section we display a number of images from MIGA and the CGPS
mosaic version of the MIGA in order to give readers an example of the
data quality. Colour versions of the images along with other samples
are available on the web\footnote{http::/www.cita.utoronto.ca/$\sim$kerton/MIGA.html}.
In Figure~\ref{fig:issacomp} we contrast 12 and 25 $\mu$m images from ISSA
and the corresponding images from MIGA. The improvement in the
data quality is obvious.
In Figure~\ref{fig:w5} we show a $4^\circ \times 3^\circ$ mosaic of
the W5 region at 12 and 25 $\mu$m. This mosaic of twelve MIGA images can
be constructed rapidly since the only operation required on the images
is that they be trimmed before being mosaiced together; no
reprojection step is required.
On larger scales we show in Figure~\ref{fig:mosaic} one of the CGPS
region mosaics. The great usefulness of the CGPS data format is that
radio and millimetre data will be available for the same region at the
same geometry and pixel size, greatly facilitating multiwavelength
analysis of objects.
\section {Summary and Future Directions}
\label{sec:sumfut}
Currently the MIGA covers the CGPS region in longitude and thus
provides a mid-infrared dataset for this multiwavelength survey. We
considered continuing the MIGA processing to encompass more of the
Galactic plane; however, in the future it is expected that
mid-infrared data from the Mid-Course Space Experiment (MSX;
\cite{pri95}) will be made available for the entire Galactic plane
($\pm5^\circ$). Since this data set has a
higher resolution ($18''$) than is possible to achieve using HIRES, we decided
to focus further expansion of the MIGA to higher and lower Galactic
latitudes in certain key areas tied to expansion of the CGPS. The CGPS
is expected to enter a second phase of operation starting in 2000,
where the focus will be on disk-halo interaction and extending the
latitude coverage around Cygnus X and the Cepheus star forming
region. In order to study the disk-halo interaction in our
Galaxy effectively obviously one needs to be able to
explore areas above the Galactic plane beyond $\pm6^\circ$. This
has been clearly demonstrated through investigations of a possible
chimney structure in W4 (\cite{nor96}, \cite{bas99}) and unusual vertical HI
structures (\cite{eng98}) that are analogous to HI ``worms''
(\cite{hei84}).
The MIGA and IGA are flexible enough that new
images can be attached seamlessly to existing images due to their
being based on an all-sky survey and the processing technique. In
Figure~\ref{fig:eiga} we illustrate the addition of the
EIGA images to the original IGA. This is a nice demonstration of the
ease with which both the IGA and MIGA can be extended to higher or
lower latitudes.
Areas can also be mapped as separate regions using the same
type of processing (e.g., the star-forming regions in Taurus and Ophiucus
that are part of the IGA; IGA also includes Orion which was looked at
by MSX).
Processing of regions outside of the CGPS survey area
is currently underway, starting with the Rho-Oph star forming region
and a HI structure at $l=124^\circ$. We intend to make users aware of
the availability of these data via the CITA web pages\footnote{Latest
information is available at http::/www.cita.utoronto.ca/$\sim$kerton/}.
As demonstrated in \S~\ref{sec:xbs} cross-band resolution matching is a
useful technique for the construction of large-scale high resolution
color ratio maps. Unfortunately the means to do this sort of analysis
is not available to the typical user of the MIGA. At the moment the
only option is to request this sort of processing through IPAC or CITA
(facilities with the IRAS data archive and the requisite software). We
are currently working on techniques that will allow users to do cross-beam
simulations using the data that comes as part of the MIGA and IGA.
Once available this technique will greatly improve the utility of both
HIRES data sets.
In summary the MIGA provides users of the CGPS data base with a
mid-infrared data set that, combined with the IGA, will allow users to
study infrared emission from the entire range of dust grain sizes and
thus better study the evolution of dust grains as they move through
different phases of the ISM. Both the MIGA and the IGA can be easily
expanded and built upon to higher and lower Galactic latitudes and
should continue to be useful in the study of disk-halo structures and
star forming regions away from the Galactic plane where higher
resolution infrared data over large angular scales are still not available.
\acknowledgements
We thank Yu Cao for his assistance and suggestions regarding the
processing of the MIGA. Thanks also to Chas Beichman, Ron Beck and
Diane Engler at IPAC for their assistance in obtaining the raw IRAS archive,
and John Fowler for discussions about the HIRES algorithms.
CRK would like to thank the Ontario Graduate Scholarship Program for
support. This research was supported by the Natural Sciences and
Engineering Research Council of Canada.
\clearpage
|
\section[Introduction]{Introduction}
Coupled map lattices (CML) are arrays of low-dim\-en\-sio\-nal dynamical
systems with discrete time, originally introduced in 1984 as
simple models for spatio-temporal complexity \cite{CML:84}.
CMLs have been extensively used in modelling spatio-temporal chaos
in fluids phenomena such as turbulence \cite{Beck:94-Kan:89},
convection \cite{Yanagita:93} and open flows \cite{Willeboordse:95}.
Equally important is the analysis of pattern dynamics, which
has found applications in chemistry \cite{Kapral:94} and
patch population dynamics \cite{Hassell:95-Sole:95}.
One important feature of pattern dynamics is the existence of
travelling fronts, which occur at the pattern boundaries,
and are also seen to emerge from apparently decorrelated
media \cite{Kan:92-93}.
This paper extends the work on the behaviour of a travelling
interface on a lattice developed in
\cite{rcg:thesis,rcg:modloc,rcg:lowdim,Coutinho:98}.
Our main results are:
$i)$ a constructive procedure for the reduction of the
infinitely-dimensional dynamics of a front to one dimension;
$ii)$ a characterization of the behaviour of fronts near the
boundary of parametric stability;
$iii)$ a characterization of the behaviour of fronts near the
continuum limit.
We consider a one-dimensional infinite array of sites.
At the $i$-th site there is a real dynamical variable $x(i)$,
and a local dynamical system ---the {\em local map}.
The latter is given by a real function $f$ which we assume to be
the same at all sites.
The dynamics of the CML is a combination of local dynamics and
coupling, which consists of a weighted sum over some
neighbourhood.
The time-evolution of the $i$-th variable is given by
$$
x_{t+1}(i) = \sum_k \varepsilon_k f(x_t(i+k))
$$
where the range of summation defines the neighbourhood.
The coupling parameters $\varepsilon_k$ are site-independent,
and they satisfy the conservation law $\sum \varepsilon_k=1$,
to prevent unboundedness as time increases to infinity.
The two most common choices for the coupling are
\begin{equation} \label{one-way}
x_{t+1}(i)= (1\!-\!\varepsilon) f(x_t(i)) + \varepsilon f(x_t(i\!-\!1)),
\end{equation}
and
\begin{equation} \label{diffu}
x_{t+1}(i) = (1\!-\!\varepsilon) f(x_t(i))
\displaystyle + {\varepsilon\over 2}\left( f(x_t(i\!-\!1)) + f(x_t(i\!+\!1)) \right)
\end{equation}
which are called {\em one-way\/} and {\em diffusive\/} CML, respectively.
The diffusive CML corresponds to the discrete analogue of the
reaction-diffusion equation with a symmetrical neighbouring interaction.
There is now a single coupling parameter $\varepsilon$ which is constrained
by the inequality $0\leq\varepsilon\leq1$, to ensure that the sign of the
coupling coefficients in (\ref{diffu}) and (\ref{one-way})
({\em i.e.}~$\varepsilon$, $\varepsilon/2$ and $1-\varepsilon$) remains positive.
In this paper we study front propagation in {\em bistable\/} CMLs.
The local mapping $f$ is continuous and has two stable equilibria,
and a {\em front\/} is any monotonic arrangement of the
state variables, linking asymptotically the two equilibria.
We will show how to construct a one-dimensional circle map
describing the motion of the front. Such a mapping
originates from the existence of an invariant function
describing the asymptotic front profile, and of a one-dimensional
manifold supporting the transient motions.
The rotation number of the circle map will then give the velocity
of propagation, resulting in the occurrence of {\em mode-locking},
{\em i.e.}, the parametric stability of the configurations
that correspond to {\em rational\/} velocity.
We will describe the vanishing of this phenomenon in the
continuum limit, as the width of the front becomes infinite.
We shall also be concerned with the evolution of the front
shape near the boundary of parametric stability, where the
continuity of the local map ensures a smooth evolution
of the front shape.
\newpage
Velocity mode-locking is commonplace in nonlinear coupled systems
({\em e.g.}, Frenkel-Kontorova models \cite{Floria:96}, Josephson-junction
arrays \cite{Basler:97}, excitable chemically reactions
\cite{Schreiber:94}, and nonlinear oscillators
\cite{Bressloff:97-Kuske:97}): the present work provides
further support for its genericity, and highlights key
dynamical aspects.
Throughout this paper, the very existence of fronts in the
regimes of interest to us is inferred from extensive
numerical evidence. We are not concerned with existence
proofs here. Fronts have been proved to exist in various
situations, mainly for {\em discontinuous\/} piecewise
affine maps (see \cite{Coutinho:98} and references therein);
in the present context however, continuity is crucial.
Following \cite{rcg:modloc}, we consider a CML whose local map
$f$ is continuous, monotonically increasing and which possesses
exactly two stable fixed points ${x^*_-}$ and ${x^*_+}$.
It then follows that there exists a unique unstable fixed point
$x^*$ such that $x_-^*<x^*<x_+^*$.
The homogeneous fixed states $x(i)=x_\pm^*$,
$\forall\, i \in \Bbb Z$, inherit the stability of the fixed points
$x_\pm^*$ \cite{Gade:93-Zhilin:94b}.
We denote by $I_-=[x_-^*,x^*)$ and $I_+=(x^*,x_+^*]$
the basins of attraction of $x_-^*$ and $x_+^*$, respectively,
while $I=[x^*_-,x^*_+]$.
A {\em minimal mass state\/} is a state satisfying the
monotonicity condition $x(i)\leq x(i+1)$, for all $i$.
It can be shown directly from the system equation that the image
of a minimal mass state has the same property.
A {\em front\/} is a minimal mass state satisfying the
asymptotic condition:\/ $\lim_{i\to\pm\infty}x(i)=x^*_\pm.$
The main properties of a front are its {\em centre of mass\/} $\mu_t$
and its {\em width\/} $\sigma_t^2$, which measure its position and
spread at time $t$, respectively.
They are defined as the mean and variance of the variable $i$
with respect to the time-dependent probability distribution
\begin{equation}\label{prob}
\displaystyle p_t(i)={|\Delta x_t(i)|\over\displaystyle\sum_{i=-\infty}^{\infty}{|\Delta x_t(i)}|},
\end{equation}
where $\Delta x_t(i)=x_t(i+1)-x_t(i)$ is the variation of the local states.
We have
\begin{equation} \label{mu_sigma}
\displaystyle\begin{array}{rcl}
\mu_t & = & \displaystyle\sum_{i=-\infty}^{\infty}{i p_t(i)},\\[3.0ex]
\sigma_t^2 & = & \displaystyle\sum_{i=-\infty}^{\infty}{(i-\mu_t)^2 p_t(i)}.
\end{array}
\end{equation}
A state $X_t=\{x_t(i)\}$ with finite centre of mass and width is said
to be {\em localised}.
In this paper we are interested in fronts of {\em fixed shape},
moving at velocity $v$. They are described by the equation
\begin{equation}\label{eq:Front}
x_t(i)=h(i-vt);
\qquad
v=\lim_{t\to\infty}{\mu_t\over t},
\qquad t, i\in{\Bbb Z}.
\end{equation}
Here the function $h:{\Bbb R} \mapsto [x^*_-,x^*_+]=I$ is to
be determined subject
to the condition that it be monotonic, with
$\lim_{x\to\pm\infty}=\pm x^*_\pm$.
The degree of smoothness of $h$ will depend on the regime being considered.
The object of interest to us is the central part of the front.
Far away from the centre, the lattice is almost homogeneous
({\em i.e.}, $|\Delta x_t(i)|\ll |I|$), and the dynamics is dominated by
the attraction towards the stable points of the local map.
The qualitative evolution of the centre of the front can be
understood as the result of the competition between local
dynamics and coupling (see Figure \ref{pulling.eps}, for the one-way case).
For small $\varepsilon$, the attraction towards the fixed points $x_\pm^*$
overcomes the effect of the coupling, resulting in propagation
failure (zero velocity) \cite{rcg:modloc}.
A sufficiently large coupling will instead cause a site located
within the basin $I_+$ to switch to the basin $I_-$, and move
rapidly towards $x_-^*$. As a consequence, the centre of mass
of the front will move to the right, resulting in propagation.
\oneFIG{pulling.eps}{\FigSize}{\PULLINGCAP}
A similar argument can be applied in the diffusive case.
Now however the coupling is symmetric, and a bias to either
of the stable points will have to be introduced via an
asymmetry in the local map.
For instance, increasing the size of the basin of attraction of
$x_-^*$, will result in propagation from left to right for
an increasing front.
In previous works we have shown that the dynamics of a
{\em finite-size\/} interface in a class of piece-wise linear
one-way CMLs can be reduced to a single one-dimensional map
\cite{rcg:modloc,rcg:lowdim}.
The finiteness of the front depended on the existence of
degenerate superstable fixed points of the local map,
that caused nearby orbits to collapse unto the stable states
in a single iteration.
In this paper we remove such degeneracy, and consider smooth
local maps and infinitely extended fronts (the case of a
discontinuous local map was treated in \cite{Coutinho:98}).
We shall provide evidence that every front evolves towards a
unique asymptotic regime, characterized by a constant velocity
as well as an invariant shape.
Under these assumptions, we then show how the front behaves
at the boundary of the regions of parametric stability
(here the continuity of the local map is essential),
and how the reduction to one-dimensional dynamics can
be achieved.
This paper is organised as follows.
In section \ref{SEC:CONTINUUM} we describe the behaviour of
travelling fronts in the continuum limit, when the density of
interfacial sites is large.
We obtain an ODE describing the shape of the travelling front,
and with it we find new classes of fronts.
In section \ref{SEC:REDUCED} we consider the asymptotic
shape of the front, and we provide extensive evidence that
such a shape is fixed and is described by a continuous function.
This result allows us to derive a procedure for the reduction of
the infinite-dimensional interface dynamics to a one-dimensional
problem described by the {\em auxiliary map.}
In section \ref{SEC:MODLOC} we show that the auxiliary map
is a circle map and we relate its rotation number to the velocity
of the front, from which the mode-locking of the velocity with
respect to the system parameters follows.
Finally, we explain in terms of reduced dynamics the
vanishing effect of mode-locking when the continuum
limit is approached.
\bigskip
\section[The continuum limit]
{The continuum limit{\label{SEC:CONTINUUM}}}
In this section we consider fronts with large widths,
for which the relative density of sites is large, and
the continuum approximation becomes appropriate.
To achieve a front with such features, the attraction
towards $x_\pm^*$ and the repulsion of $x^*$ must be small.
Because $f$ is continuous and monotonic, then $f$ is
necessarily close to the identity, {\em i.e.}
$$
\delta_f=\sup_{x^*_-<x<x^*_+}\,|f(x)-x| \,\ll\, 1.
$$
Choosing functions $f$ such that $\delta_f\rightarrow 0$
is referred to as the {\em continuum limit}.
Inserting equation (\ref{eq:Front}) into the equations of motion
(\ref{one-way}) and (\ref{diffu})
we find that
\begin{equation}\label{functional}\begin{array}{lrcl}
\hbox{\rm a)}& h(z-v)&=&(1-\varepsilon)f(h(z))+ \varepsilon f(h(z-1)),\\[3.0ex]
\hbox{\rm b)}& h(z-v)&=&(1-\varepsilon)f(h(z))\\[1.0ex]
& & &+ \displaystyle{\varepsilon\over 2}\left(f(h(z-1))+f(h(z+1))\right),
\end{array}\end{equation}
for the one-way and diffusive CML, respectively,
where $z=i-vt$.
A function $h$ satisfying the functional equation (\ref{functional})
represents the fixed shape of a front travelling at the velocity $v$.
To solve equation (\ref{functional}) in the continuum limit,
we assume $f$ and $h$ to be twice differentiable, and consider
the Taylor series of $h$ in $z$, up to second order.
The Taylor expansion becomes accurate as the width increases,
since in this case the variation of $h$ over adjacent lattice
sites tends to zero.
We obtain
\begin{equation}\label{taylor}
\begin{array}{rl}
h(z)-f(h(z))& + A\, h'(z)
-\displaystyle\left({{\displaystyle\varepsilon\,f''(h(z))}\over\displaystyle 2}\right) {h'(z)}^2 \\[2.0ex]
&+\displaystyle\left( {\displaystyle v^2-\varepsilon\,f'(h(z))\over\displaystyle 2}\right) h''(z)\,=\,0,
\end{array}
\end{equation}
where $A=\left(\varepsilon\,f'(h(z))- v\right)$ and $A=-v$, for the one-way
and diffusive CML, respectively. In the continuum limit we can
further simplify equation (\ref{taylor}) by considering $f'(x)=1$
and $f''(x)=0$, to obtain
\begin{equation}\label{ODEs}\begin{array}{ll}
\hbox{\rm a)}&
h(z)-f(h(z))+\displaystyle \left({\displaystyle\varepsilon(\varepsilon-1)\over\displaystyle 2}\right) h''(z)=0, \\[4.0ex]
\hbox{\rm b)}&
h(z)-f(h(z)) - v\,h'(z) + \displaystyle
\left({\displaystyle v^2-\varepsilon\over\displaystyle 2}\right) h''(z)=0,
\end{array}\end{equation}
for the one-way and diffusive CML, respectively,
where we set $v=\varepsilon$ in the one-way case since in the continuum limit
$f(x)\rightarrow x$ and thus the rate of information exchange
({\em i.e.}~the velocity) is equal to $\varepsilon$.
For the diffusive case the velocity is not equal to $\varepsilon$, since the
total information exchange comes from the competition between the
left and right neighbours.
Nevertheless, as we shall see, it is possible to give an analytical
approximation to the velocity for the case of an asymmetric cubic
local map.
\oneFIG{pha-spa.eps}{\FigSize}{\PHASPACAP}
Equations (\ref{ODEs}) are similar to those obtained in
\cite{Chow:95-96}, where the travelling front in
a lattice of coupled ODEs, is reduced to a single equation.
The ODEs (\ref{ODEs}) describe the motion of a particle of mass
$m=(v^2-\varepsilon)/2$, subject to the potential $V(x)=\int{(f(x)-x)\,dx}$,
with maxima located at the stable fixed points of the local map
(Figure \ref{pha-spa.eps}).
In the one-way case, the system is conservative.
For numerical experiments, we choose a symmetric local map
$f$ with fixed point $x_\pm^*=\pm1$ and $x^*=0$.
The resulting potential is also symmetric.
There exist two heteroclinic connections, joining $x_-^*$
to $x_+^*$, and $x_+^*$ to $x_-^*$, respectively
(the thick lines in Figure \ref{pha-spa.eps}(a)).
They correspond, respectively, to an increasing and a
decreasing symmetric travelling front for the CML.
In the diffusive case, the differential equation has
the dissipative term $-v\,h'(z)$.
For the local map, we choose $0<x^*=p<1$, which introduces
an asymmetry in the system, and the maxima of the potential
are now unequal:\/ $V(x_-^*)>V(x_+^*)$.
Imposing a heteroclinic connection from $x_-^*$ and $x_+^*$,
constrains the velocity $v$ of the front (see below).
For larger velocities, the separatrix emanating from $x_-^*$
approaches $p$, while for smaller $v$ it escapes to infinity.
Since the presence of friction breaks the time-reversal
symmetry, only one heteroclinic connection is possible, and
the separatrix emanating from $x_+^*$ always approaches
$p$ (the thick lines in Figure \ref{pha-spa.eps}(b)).
\oneFIG{waves.eps}{\FigSize}{\WAVESCAP}
The continuum approximation can be used to construct new
kinds of travelling fronts. For example, the librating
orbits in Figure \ref{pha-spa.eps}(a) (one-way case), correspond
to spatially-periodic travelling fronts that never touch the
stable points (see Figure \ref{waves.eps}(a) $(iii)$).
Such spatially-periodic orbits do not exist in the diffusive case.
Nevertheless, it is possible to construct the travelling front
departing from $x_+^*$ that dissipates down to $p$.
This new solution has a damped oscillatory profile (see Figure
\ref{waves.eps}(b) $(i)$), and it is unstable, because
the fixed point at $p$ is unstable (for the CML).
In the remainder of this section, we briefly examine the case of
a cubic local map, providing the dominant behaviour of a general
bistable local map in the continuum limit.
We use the one-parameter families of cubics
\begin{equation}\label{cubics}
\begin{array}{ll}
\hbox{\rm a)} &
f(x)={\displaystyle x\over \strut \displaystyle 2}(3-\nu-(1-\nu)\, x^2)\, ,
\\[4.0ex]
\hbox{\rm b)} &
f(x)=(1-\nu)\,(p\,x^2-x^3 -p)+(2-\nu)\,x.
\end{array}\end{equation}
for the one-way and diffusive CML, respectively.
Again, $x_\pm^*=\pm 1$ for both cases, while $x^*=0$ in the
one-way case and $x^*=p$ in the diffusive case, where $0<p<1$
controls the asymmetry.
The continuum limit is attained by letting the parameter
$\nu$ approach 1 from below.
Substituting the cubic local maps (\ref{cubics}) in the differential
equations (\ref{ODEs}), one finds expressions for the heteroclinic
connections corresponding to the travelling front solutions:
\begin{equation}\label{wave-sol}
\begin{array}{ll}
\hbox{\rm a)} &
h(z) = \tanh\left(\sqrt{\displaystyle 1-\nu \over\displaystyle 2\varepsilon(1-\varepsilon)}\,z\right), \\[4.0ex]
\hbox{\rm b)} &
h(z) = \tanh\left({\displaystyle(1-\nu)p \over\displaystyle v}\,z\right),
\end{array}\end{equation}
where
\begin{equation}\label{width}
\begin{array}{lll}
\hbox{\rm a)} &
v=\varepsilon &\quad \displaystyle \sigma^2 = {2\pi^2\over 3}{\varepsilon(1-\varepsilon)\over 1-\nu} \\[4.0ex]
\hbox{\rm b)} &
v= p\sqrt{\varepsilon(1-\nu)} & \quad\displaystyle \sigma^2 = {\pi^2\over 3}{\varepsilon \over 1-\nu}.
\end{array}
\end{equation}
for the one-way and diffusive CML, respectively.
In the diffusive case, the expression for the velocity is
derived from imposing an heteroclinic connection, while
the scaling of the width $\sigma^2$ is found from the
solutions (\ref{wave-sol}).
Note that for both models the functional dependence of the width on
the parameter $\nu$ is the same, and it describes the rate at which
the front broadens as the continuum limit is approached.
Moreover, from (\ref{wave-sol}) and (\ref{width}) we have
that in the diffusive case $h$ is independent of $p$.
While in the continuum limit the front is described by a
continuous function $h$ ({\em cf.}~equations (\ref{wave-sol})),
there is no {\em a-priori\/} reason why such a function should continue
to exist away from the limit, due to the discrete nature of the system.
We shall nonetheless give evidence that the dynamics of a
front far from the continuous limit remains one-dimensional.
\section[Reduced dynamics of the travelling front]
{Reduced dynamics of the travelling front{\label{SEC:REDUCED}}}
In this section we provide evidence that every front has a fixed
profile, which can be characterized by an invariant function $h$.
Such a function will then be used to construct a
one-dimensional mapping describing the front evolution
---the {\em auxiliary map}.
If the velocity $v$ of the front is {\em irrational}, then the
collection of points $i-vt$, with $i$ and $t$ integers, form a
set dense on the real line.
Numerical experiments consistently suggest that
in the case of a front, the closure of the set of points
$(i-vt,\,x_t(i))\in{\Bbb R}^2$ forms the graph of a continuous and
monotonic function: \/ $h:{\Bbb Z} \mapsto [x_-^*,\,x_+^*]$, which
is a solution to the functional equation (\ref{functional}).
The results for both CML models are summarised in Figure \ref{sftG.ps},
where we have superposed all translates of the discrete fronts,
after eliminating transient behaviour.
This procedure requires computing $v$ numerically, which was
done using some $10^7$--$10^8$ iterations of the CML.
(In principle, a numerical solution to (\ref{functional}) can
be found using various iterative functional schemes.
However, all the schemes considered were plagued by slow
convergence and are not discussed here.)
\oneFIG{sftG.ps}{\FigSize}{\SFTCAP}
In the case in which $v=p/q$ is {\em rational}, the function $h$
is specified only at a set of $q$ equally spaced points.
It turns out, however, that the definition of $h$ becomes
unequivocal in a prominent parametric regime, corresponding
to the boundary of the so-called mode-locking region or
{\em tongue}. The latter is defined as the collection of
parameters $(\varepsilon,\nu)$ corresponding to a given rational velocity,
where $\nu$ (not necessarily one-dimensional) parametrizes the
family of local maps ---for the one-way CML we typically use
$f(x)=\tanh(x/\nu)$.
We defer the discussion of the origin of such regions to the next section.
Here we consider a sequence of parameters
$(\varepsilon_n,\nu_n)\rightarrow (\varepsilon_*,\nu_*)$,
converging from the outside towards a boundary point
$(\varepsilon_*,\nu_*)$ of the tongue (see Figure \ref{boundary.ps}).
Independently from the path chosen to approach the boundary
point, the front $h$ appears to approach a unique limiting
shape. The limiting shape is a step function with $q$ steps (where $v=p/q$)
for every unit length ---the horizontal length of each step is $1/q$
since there are $q$ equidistant points in every horizontal interval
of unit length for a $v=p/q$ orbit.
In the limit, the front dynamics becomes periodic, with
periodic points corresponding to the midpoint of each step.
This observation suggests that choosing step fronts with the
periodic points at their midpoints ensures continuity of the
front shapes at the resonance tongue boundaries.
\oneFIG{boundary.ps}{\FigSize}{\BOUNDARYCAP}
In the next section, we shall explain this phenomenon in
terms of the dynamics of a one-dimensional map ---the
{\em auxiliary map\/} $\Phi$--- which we now define.
The idea is to describe the evolution of any site in the
front by means of a single site, the {\em central site\/}
$\bar x_t(0)$, defined as the site which is closest to the
unstable point $x^*$.
The position of the central site moves along the lattice with an
average velocity $v(\varepsilon)$, since it follows the centre of the interface.
Following \cite{rcg:modloc,rcg:lowdim}, we define the map $\Phi$\/ as
\begin{equation} \label{eq:AuxiliaryMap}
\bar x_{t+1}(0)=\Phi(\bar x_t(0)).
\end{equation}
If the velocity is {\em irrational}, the domain of definition of
the map is a set of points dense in an interval (see next section),
and the possibility exists of extending $\Phi$ continuously to
the interval.
In Figure \ref{auxmap1G.ps}(a) and (b), we plot the graph of $\Phi$
for a one-way and a diffusive CML, respectively.
The auxiliary map corresponds to the square region depicted
with thick lines,
while the other regions represent delay Poincar\'e maps of
some neighbouring sites.
Indeed for each neighbour $j$ of the central site, there is
a corresponding auxiliary circle map $\Phi_j$, such that
$\bar x_{t+1}(j)=\Phi_j(\bar x_t(j))$, with $\Phi=\Phi_0$
(see below).
\twoFIG{auxmap1G.ps}{auxmap2G.ps}{\FigSize}{\AUXMAPCAP}
If the velocity is {\em rational}, equation (\ref{eq:AuxiliaryMap})
defines $\Phi$ only at a finite set of points, and to extend
the domain of definition, one must make use of equation
(\ref{eq:AuxiliaryMap}) on suitable transients.
We have verified numerically that when a front is perturbed,
the perturbation relaxes quickly onto a one-dimensional
manifold, along which the original front is approached.
The process of randomly disturbing the front amounts to a random
walk path reconstruction of the one-dimensional manifold.
Such one-dimensional transients were found to be independent
of the detail of the perturbation, giving a unequivocal
definition of the auxiliary map also in the rational case.
This is illustrated in Figure \ref{auxmapTongueG.ps}.
Crucially, this construction yields a map that changes
continuously within the tongue, matching the the behaviour
at the boundary of the tongue.
Thus we conjecture that the auxiliary map $\Phi$ depends
continuously on the coupling parameter $\varepsilon$.
In the next section we shall explore some consequences
of the continuity.
\oneFIG{auxmapTongueG.ps}{\FigSize}{\AUXMAPTONGUECAP}
We finally relate the dynamics of the entire front to that
of the central site, governed by $\Phi(x)$.
Let $\bar x_t(j)$ denote the $j$-th neighbouring site of the
central site $\bar x_t(0)$, where $j$ is positive (negative)
for the right (left) neighbours.
The dynamics of $\bar x_t(j)$ can be deduced from that of
$\bar x_t(0)$ and the knowledge of $h$, as follows
\begin{equation}\label{hh-1}
\bar x_t(j)=h\circ \tau_j \circ h^{-1}(\bar x_t(0))
\end{equation}
where $\tau_j$ is the translation by $j$ on ${\Bbb R}$.
Since $\Phi_j(x)$ maps $\bar x_t(j)$ to $\bar x_{t+1}(j)$,
the pair $\left(\bar x_t(j),\,\bar x_{t+1}(j)\right)$ belongs
to the graph of $\Phi_j$.
By applying the operator $h\circ \tau_j \circ h^{-1}$ to the
function $\Phi(\bar x_t(0))$ we obtain:
$$
\begin{array}{rcl}
h\circ \tau_j \circ h^{-1}\Phi(\bar x_t(0))
&=& h\circ \tau_j \circ h^{-1} (\bar x_{t+1}(0))\\[2.0ex]
&=& \bar x_{t+1}(j) \,=\, \Phi_j\left(\bar x_t(j)\right),
\end{array}
$$
where we used equation (\ref{hh-1}) which relates neighbouring sites.
Thus $h\circ\tau_j \circ h^{-1}$ provides a conjugacy between
$\Phi$ and $\Phi_j$ and enables
us to reconstruct the whole interfacial
dynamics from the behaviour of the central site.
\section[Mode-locking of the propagation velocity]
{Mode-locking of the propagation velocity{\label{SEC:MODLOC}}}
In this section we show that the auxiliary map $\Phi$ is a
circle homeomorphism (see Figure \ref{circle.eps}).
The mode-locking of the front velocity will then follow
from the mode-locking of the rotation number of $\Phi$.
Furthermore, the conjectured continuous dependence of $\Phi$
on $\varepsilon$ implies a continuous dependence of the
rotation number on $\varepsilon$, and in particular, $\Phi$ takes
all rotation numbers between any two realised values.
For instance, the front velocity in a one-way CML takes the values
$0$ and $1$ for $\varepsilon=0$ and $1$,
respectively, and thus as the coupling parameter varies,
all velocities $v\in [0,1]$ are realised.
For a diffusive CML only an interval $[0,v_{max}]$ is
attained since the maximum velocity $v_{max}=v(\varepsilon=1)$
does not reach 1 because of the competition between the attractors.
\oneFIG{circle.eps}{\FigSize}{\CIRCLECAP}
Let us consider a continuous and increasing travelling front
$h(i-vt+i_0)$ with positive irrational velocity $0<v<1$.
The largest possible separation between $\bar x_t(0)$ and $x^*$
corresponds to the position of $h$ for which two consecutive
points on the lattice are equally spaced from the unstable point $x^*$.
Suppose that the front shape $h$ is positioned such that for
site $i$, we have $h(i)=x^*$.
We choose $\alpha$ such that
\begin{equation}\label{h_alpha}
h(i-\alpha) = x^* - a \quad {\rm and} \quad
h(i+1-\alpha) = x^* + a
\end{equation}
where $0\leq a \leq\min(|x_+^*-x^*|,|x_-^*-x^*|)$.
By adding the two equations in (\ref{h_alpha}) one obtains
an equation for $\alpha$, and $a$ can then be evaluated.
If the front is at a position where it satisfies the equations
(\ref{h_alpha}) for some $i$, then the $i$-th and $(i+1)$-th
sites are equally spaced from $x^*$, and the dynamics of the site
closest to $x^*$ is contained in the interval $[x^*-a,x^*+a]$.
Any shift of the front will cause either one of the two
sites to be closer to $x^*$ than originally.
We now follow the dynamics of $\bar x_t(0)$ in $[x^*-a,x^*+a]$.
Suppose that at time $\tau$ the $i$-th site is the closest to $x^*$ so
$\bar x_\tau(0) = x_\tau(i)$. We want to know which site will be
closest to $x^*$ at time $\tau+1$. Since we are considering
the case $v>0$ there are two possibilities:
a) the $i$-th site again ($\bar x_{\tau+1}(0)=x_{\tau+1}(i)$)
or b) the $(i+1)$-th site ($\bar x_{\tau+1}(0)=x_{\tau+1}(i+1)$).
Redefining $h_t(i)=h(i-vt+i_0)$, we find two cases
\begin{equation}\label{h1}
h_{\tau+1}^{-1}(\bar x_{\tau+1}(0))=\left\{
\begin{array}{ll}
h_\tau^{-1}(\bar x_\tau(0)) &\quad\mbox{(a)}\\[2.0ex]
h_\tau^{-1}(\bar x_\tau(0))+1 &\quad\mbox{(b)}
\end{array}
\right. .\end{equation}
But, by definition, $h_{\tau+1}(x)=h_{\tau}(x-v)$, so from equation
(\ref{h1}) one obtains
\begin{equation}\label{h2}
\bar x_{\tau+1}(0)=\left\{
\begin{array}{ll}
f_-(\bar x_\tau(0)) &\quad\mbox{(a)}\\[2.0ex]
f_+(\bar x_\tau(0)) &\quad\mbox{(b)}
\end{array}
\right. \end{equation}
where
\begin{equation}\label{h3}
\begin{array}{rcl}
f_-(x) & = & h_\tau\left(h_\tau^{-1}(x)-v\right)\\[2.0ex]
f_+(x) & = & h_\tau\left(h_\tau^{-1}(x)-v+1\right).
\end{array}\end{equation}
The functions $f_-$ and $f_+$ inherit some of the properties of $h$.
In particular, $f_-$ and $f_+$ are continuous and increasing.
In the the interval $[x^*-a,x^*+a]$ we have that $f_-(x)<f_+(x)$,
because $h$ is increasing, so we just evaluate at the following points:
$$
\begin{array}{rcl}
f_-(x^*+a)&=&h_\tau(h_\tau^{-1}(x^*+a)-v)\\[1.0ex]
&=&h_\tau(i+1-\alpha-v)\\[3.0ex]
f_+(x^*-a)&=&h_\tau(h_\tau^{-1}(x^*-a)-v+1)\\[1.0ex]
&=&h_\tau(i-\alpha-v+1)
\end{array}
$$
where we have made use of equations (\ref{h_alpha}).
Thus we have the periodicity condition
\begin{equation}\label{h5}
f_-(x^*+a)=f_+(x^*-a).
\end{equation}
Next we find when $f_-$ and $f_+$ reach the extrema of the interval
$[x^*-a,x^*+a]$.
To this end we determine $c_\pm$ such that $f_\pm(c_\pm)=x^*\pm a$.
So we solve
$$
\begin{array}{rl}
&
\left\{ \begin{array}{rcccl}
f_-(c_-)&=&h_\tau(h_\tau^{-1}(c_-)-v)&=&x^*-a\\[2.0ex]
f_+(c_+)&=&h_\tau(h_\tau^{-1}(c_+)-v+1)&=&x^*+a
\end{array}\right. \\[5.0ex]
\Rightarrow &
\left\{ \begin{array}{rcccl}
h_\tau^{-1}(c_-)-v&=&h_\tau^{-1}(x^*-a)&=&i-\alpha\\[2.0ex]
h_\tau^{-1}(c_+)-v+1&=&h_\tau^{-1}(x^*+a)&=&i+1-\alpha
\end{array}\right. \\[6.0ex]
\end{array}
$$
whence $h_\tau^{-1}(c_-)= h_\tau^{-1}(c_+)$, and since $h$
is monotonic, we have that $c_-= c_+=c.$
Therefore, the map $\Phi$ giving the dynamics of the central
site (\ref{eq:AuxiliaryMap}) is given by
\begin{equation}\label{Phi}
\Phi(x)=\left\{
\begin{array}{lrcccl}
f_+(x) & \mbox{~~if~~} x^*-a &\leq &x&\leq& c \\[2.0ex]
f_-(x) & \mbox{~~if~~} x^*+a &\geq &x& > & c.
\end{array}
\right.
\end{equation}
From the above properties of $f_-$ and $f_+$, it follows that
the auxiliary map $\Phi$ is a homeomorphism of the circle
(see Figure \ref{circle.eps}).
A natural binary symbolic dynamics for $\Phi$ is introduced
by assigning the symbols `0' and `1' whenever the branch
$f_-$ or $f_+$, respectively, is used in (\ref{h1}).
These symbols corresponds to the central site $x(i)$ remaining
unchanged, or being replaced by the new site $x(i+1)$,
respectively.
\oneFIG{intermittencyG.ps}{\FigSize}{\INTERMITTENCYCAP}
Every time a `1' is encountered, the front advances by roughly
one site. So the density of `1's in the sequence gives an approximation
to the velocity, which becomes exact in the limit $t\rightarrow\infty$.
In terms of the circle map, the proportion of `1's in the sequence
corresponds to its rotation number $\rho$:
\begin{equation}
\rho(\varepsilon)=v(\varepsilon)
=\lim_{t\rightarrow\infty}{{\displaystyle 1\over \displaystyle t}\sum_{i=1}^{t}{s_i}},
\end{equation}
where $s_i$ is the $i$-th term in the symbolic sequence.
We have stressed the $\varepsilon$-dependence of $\rho$, since for a fixed
local map, $\Phi$ depends on $\varepsilon$, and so does its rotation number.
Because all sites $\bar x(j)$ belong to the same front, the
site interchanges all occur at the same time, and therefore the
rotation number of any $\Phi_i$ is the same as the one for $\Phi$.
The representation of the motion of a front as a circle map
implies the likelihood of mode-locking for rational velocities,
corresponding to Arnold tongues in parameter space, and it
affords a simple explanation of the various dynamical
phenomena described in the previous sections.
The appearance of a $q$-period tongue as $\varepsilon$ is varied
thorough some critical value $\varepsilon_*$, corresponds to a
fold bifurcation of $\Phi^q$.
Generically, a pair of period-$q$ orbits is created at $\varepsilon=\varepsilon_*$.
Thus the orbits of $\Phi$ will undergo intermittency in the region
of the period-$q$ orbit for $\varepsilon_n$ close to $\varepsilon_*$.
The intermittency will manifest itself in the graph of $\Phi$ as
shown by the darkly shaded areas of the orbit web in Figure
\ref{intermittencyG.ps}.
Moreover, the periodic orbit will form towards the centre of the dark
bands and the corresponding front shape will ``flatten'' at the heights
taken by the periodic points because of the time spent in their
neighbourhood by the orbits of $\Phi$ for $\varepsilon_n\approx \varepsilon_*$.
It then follows that the approximating fronts will form steps for the
periodic front with the periodic points close to their centre points,
and independently from the parametric path chosen to approach the
boundary point (see Figure \ref{boundary.ps}).
\oneFIG{tongues.eps}{\FigSize}{\TONGUESCAP}
In Figure \ref{tongues.eps} we plot
the main mode-locking regions in parameter space
(Arnold tongues), corresponding to $v=p/q$ with small $q$.
Here the local map is given by $f(x)=\tanh(x/\nu)$, while
the parameters vary within the unit square: $(\varepsilon,\nu)\in[0,1]^2$.
We believe that mode-locking is a common phenomenon in front
propagation in CMLs, because the nonlinearity of the local map
induces nonlinearity in the auxiliary map \cite{rcg:modloc,rcg:lowdim},
and mode-locking is generic for such maps.
However, this phenomenon often takes place on very small parametric
scales, since the width of the tongues decreases sharply
with increasing $\nu$ (Figure \ref{tongues.eps}).
This explains why this phenomenon has not been widely reported
(with the notable exception of the large $v=0$ region, corresponding
to the well-known propagation failure in the anti-continuum limit
\cite{Aubry:90-MacKay:95}).
\oneFIG{auxmapContinuumG.ps}{\FigSize}{\AUXMAPCONTINUUMCAP}
In the continuum limit (see Figure \ref{tongues.eps}), the stability
of the attractors $x_\pm^*$ becomes weaker, causing the front to broaden.
In Figure \ref{auxmapContinuumG.ps} we plotted the auxiliary maps $\Phi_i$
corresponding to $\nu=100/101 \simeq 1$ for the one-way CML with
local map $f(x)=\tanh(x/\nu)$. This figure should be compared with
Figure \ref{auxmap1G.ps}, corresponding to a narrower front.
The domain of each $\Phi_i$ is now smaller, since the interval
$I=[x_-^*,x_+^*]$ has to be shared between a larger number of sites.
As a consequence, the nonlinearity of each $\Phi$ is reduced
(note that the auxiliary maps in Figure \ref{auxmapContinuumG.ps}
are almost linear) and with it the size of the tongues.
Thus, the larger the width $\sigma^2$ of the travelling front, the
thinner the mode-locking tongue (see right hand side scale in
Figure \ref{tongues.eps}).
\section*{Acknowledgments}
RCG would like to acknowledge DGAPA-UNAM (M\'exico) for the financial
support during the preparation of this paper. This work was partially
supported by EPSRC GR/K17026.
|
\section{Introduction}
Shear flow has profound effects on complex fluids. It can perturb
equilibrium phase transitions, such as the isotropic-to-nematic (I-N)
liquid crystalline transition in wormlike micelles
\cite{berret94a,berret94b,BPD97}, thermotropic melts
\cite{hess76,olmsted90,olmsted92,mather97}, or rigid-rod suspensions
\cite{see90,olmstedlu97}; the nematic-smectic transition in
thermotropic liquid crystals \cite{safinya91}; and the
isotropic-to-lamellar transition \cite{catesmilner89} in surfactant
systems. Shear can also induce structures, such as the well-known
multi-lamellar vesicles (onions) in surfactant systems
\cite{roux93,diat93,diat95}, that exist only as metastable equilibrium
phases. Another well-known effect is the transition between
orientations of diblock copolymer lamellae in either the steady shear
flow \cite{fredrickson94,goulian95}, or the oscillatory shear flow
\cite{koppi92,winey93,kannan94}, as a function of shear rate or
frequency, and temperature.
\begin{figure}[p]
\par\columnwidth20.5pc
\hsize\columnwidth\global\linewidth\columnwidth
\displaywidth\columnwidth
\epsfxsize=3.5truein
\centerline{\epsfbox[30 12 700 440]{figures/fig_two.eps}}
\caption{ Stress--strain-rate curves for the Doi
model with different excluded volume parameters $u$ (taken from
Fig.~\protect{\ref{fig:constit}} below). The dashed line segments
are unstable (unphysical) steady states. The straight lines indicate
possible coexistence between states $I$ and $II$ under conditions of
common stress (horizontal lines) or strain rate (vertical
line).}
\label{fig:both}
\end{figure}
A related phenomenon is dynamic instability in non-Newtonian fluids
whose theoretical {\sl homogeneous\/} stress--strain-rate constitutive
relations exhibit multi-valued behavior, as in theories of polymer
melts \cite{doiedwards,catesmcleish93} and wormlike micelles
\cite{spenley93,cates90,rehage91}. Such models may describe, for
example, the spurt effect, whereby the flow rate of a fluid in a pipe
changes discontinuously as a function of applied pressure drop
\cite{mcleish86}. A non-monotonic constitutive curve as in
Fig.~\ref{fig:both} typically has a segment (shown as a broken line)
where bulk flow is unstable. If a mean strain rate is imposed which
forces the system to lie on an unstable part of the constitutive
relation, a natural resolution for this instability is to break the
system into two regions, often called \textit{bands}, one on the high
strain rate branch and one on the low strain rate branch, to maintain
the overall applied strain rate. The most important unresolved
question about these banded flows is, what determines the stress at
which the system phase separates into bands? Experiments on many
systems (reviewed in Sec.~IIA), particularly the wormlike micelle
surfactant systems, reveal that there is a well-defined and
reproducible selected stress in a wide class of systems
There have been many suggestions for determining the selected stress.
Some workers have assumed the existence of a non-equilibrium potential
and a variational principle \cite{Zuba96,porte97}. This possibility is
intriguing, although it remains unproven. Early studies postulated a
jump at the top of the stable viscous branch (``top jumping'')
\cite{catesmcleish93,spenley93,schmitt95}, but experiments have shown
that this is not the case \cite{grand97}. Recent studies have solved
the homogeneous flow equations in various geometries using
sophisticated hydrodynamic flow-solvers and found a selected stress
\cite{greco97,espanol96}. However, evidence is growing \cite{history}
that these calculations have history-dependent stress selection (which
is in fact no selection) or introduce gradient terms due to the
discretization of the system. A final method, which we follow here,
has been to incorporate (physically present) non-local contributions
to the stress
\cite{olmsted90,olmsted92,pearson94,spenley96,olmstedlu97,history,jsplanar,goveas99,dhont99},
and examine the equations of motion under steady banded flow
conditions.
Here we extend previous work \cite{olmstedlu97} and calculate phase
diagrams for rigid-rod suspensions in shear flow, solving for the
interfacial profile between phases and using its properties to
determine the coexistence stress. As Fig.~\ref{fig:both} indicates,
phase separation is possible at \textit{either} a specified stress
(horizontal tie lines) \textit{or} a specified strain rate (vertical
tie lines). Only recently has the latter possibility been speculated
upon \cite{schmitt95,olmstedlu97,porte97}, and found experimentally
\cite{Bonn+98}. We explore this possibility explicitly for our model
system, which possesses, in addition to the high and low strain rate
(paranematic and nematic, respectively) branches shown in
Fig.~\ref{fig:both}, a second high strain rate branch in which the
rods stand up in the flow, parallel to the vorticity direction,
instead of lying in the shear plane \cite{bhave93}. We study
coexistence with this so-called `log-rolling' phase and find a rich
non-equilibrium phase diagram.
The summary of this paper is as follows. In Section~II we discuss the
general issues of shear banding and phase separation in flow, and
summarize the primary experimental evidence for this behavior. In
Section~III we present the modified Doi model \cite{doi81,doikuzuu83}
and in Section~IV we briefly discuss our algorithm for calculating the
phase diagram. The general aspects of the interface construction will
be discussed elsewhere \cite{jsplanar}. We present the results for
common stress and strain rate phase separation in Section~V and~VI,
respectively, and discuss some of the implications for metastability
and experiments under controlled stress or controlled strain rate
conditions. We finish in Section~VII with a discussion and summary.
While some of these results have been briefly summarized elsewhere
\cite{faraday}, the current paper is a complete and self-contained
discussion of the problem.
The reader interested in the phenomenology of phase diagrams for
sheared complex fluids rather than liquid crystals may safely skip
Section~III; the rest of the paper is general, and much of the
discussion applies to any system undergoing phase separation in shear
flow. There are, essentially, two steps to calculating phase behavior
in flow. One must derive the dynamical equations of motion for fluid
flow, composition, and the relevant structural order parameter(s),
which is quite difficult. Then, one must understand how to solve them
and interpret the results. While the modified Doi model does not
exhaust all possible phase diagrams (in particular, a shear-thickening
model would be a nice complement), it has many universal features. One
extremely important concept is that density and field variables are
ill-defined in non-equilibrium systems: {\sl either\/} stress {\sl
or\/} strain rate may act as a control parameter analogous to an
equilibrium field variable (\textit{e.g. }pressure, chemical
potential), corresponding to the different orientations of the
interface between coexisting phases. Also, one can gain much
intuition from the underlying stress--strain-rate--composition
\textit{surface}, a fact which we feel has been underappreciated until
now.
\section{Shear Banding}
\subsection{Experimental Evidence}
Shear banding has been confirmed in many systems through direct
optical and NMR visualization, and deduced from rheological
measurements. The best-studied systems are surfactant solutions of
various kinds, including wormlike micelles and onion-lamellar phases.
Rehage and Hoffmann \cite{rehage91} measured a plateau in the
stress--strain-rate relation for wormlike micelles in shear flow.
This behavior has since been seen in a number of wormlike micellar
systems in various flow geometries, by the Montpellier
\cite{berret94a,berret94b,BPD97}, Strasbourg \cite{schmitt94}, Edinburgh
\cite{grand97}, and Massey groups
\cite{Call+96,MairCall96,BritCall97,MairCall97}. Berret \textit{et
al.} \cite{BPD97} visualized shear bands in the plateau region of
the stress--strain-rate curves using optical techniques, providing
proof of banding; and Callaghan \textit{et al.}
\cite{Call+96,MairCall96,BritCall97,MairCall97} used NMR to measure
the velocity profile in various geometries (including Couette,
cone-and-plate, and pipe geometries).
The transition in these cases is to a strongly-aligned, possibly
nematic, phase of wormlike micelles which has a lower viscosity than
the quiescent phase. It is not known how the length distribution
changes in flow, although this is certainly an important aspect of
these `living' systems \cite{Vand94}. Wormlike micellar system can
possess an equilibrium nematic phase, and in some cases the
shear-induced phase is obviously influenced by the proximity of an
underlying nematic phase transition
\cite{berret94a,berret94b,Roux+95,BPD97,Capp+97}. However, many
wormlike micellar systems undergo banding at compositions much more
dilute than that for I-N coexistence, and it is probable that in these
cases flow instability is due to the non-linear rheology of these
systems, which is in many respects similar to that of the Doi-Edwards
model of polymer melts \cite{cates90}. Since there are at lease two
possible effects (a nematic phase transition and flow-instability of
the micellar constitutive relation) apparently leading to
flow-instability, these systems are quite rich. It is tempting to
analyze the extent to which these systems display behavior analogous
to the kinetics of equilibrium phase separation, and groups have
recently begun to study the kinetics of non-equilibrium phase
separation \cite{berret94b,grand97,Berr97}.
Pine and co-workers have recently studied a wormlike surfactant system
at extreme dilutions and found, surprisingly, that for low enough
concentrations (but still above the overlap concentration) shear
induces a viscoelastic phase that they interpret as a gel
\cite{boltenhagen97a,boltenhagen97b,keller97}. The origins and
structure of this gel are currently unknown. In controlled stress
experiments they observe shear banding and a `plateau' for stresses
higher than a certain stress, in which the strain rate
\textit{decreases} as shear induces the gel. Above the stress at which
the gel fills the sample cell the strain rate increases again to
complete a dramatic \textsf{S} curve. For controlled strain rate
experiments the system jumps, at a well defined strain rate, between
the gel and solution phases.
Another well-studied system is the onion lamellar surfactant phase,
originally studied by Roux, Diat, and Nallet
\cite{roux93,diat93,diat95}. These systems display a bewildering
variety of transitions between lamellar, aligned-lamellar, onion, and
onion crystal phases of various symmetries, as functions of applied
shear flow, temperature, and composition. As an example, one
particular system undergoes transitions, with increasing strain rate,
from disordered lamellae to onion, to onion-lamellae coexistence (in
which coexistence is inferred from a plateau in the
stress--strain-rate curves), to well-ordered lamellae \cite{diat93}.
Recently Bonn and co-workers \cite{Bonn+98} found shear-induced
transitions between different gel states of lamellar onion solutions
with shear bands (visualized by inserting tracer particles) oriented
with interface normals in the vorticity direction, indicating phase
separation at common strain rate instead of common stress, as we
clarify below. In this case the averaged stress--strain-rate
constitutive relation followed a sideways \textsf{S} curve under
controlled strain rate conditions.
Mather \textit{et al.} \cite{mather97} have recently studied a
thermotropic polymer liquid crystal using visual and rheological
measurements, and inferred a shear-induced nematic phase transition
and phase separation, the latter which they attribute to
polydispersity.
In summary, shear-banding has been seen in several systems, and in all
cases is associated with some flow-induced change in the fluid
microstructure. Most systems are still poorly-understood
\cite{boltenhagen97a,diat95,Bonn+98} and, given the range of complexity,
it is certain that many qualitatively new phenomena remain to be
discovered.
\subsection{Theoretical Issues}
The crux of the problem from a theoretical point of view may be
appreciated from Fig.~\ref{fig:both}. These stress--strain-rate
curves are somewhat reminiscent of pressure-density ($p-\rho$)
isotherms for a liquid-gas system. Curve segments with negative slope,
$\partial\sigma_{xy}/\partial\dot{\gamma}<0$, are unstable and cannot
describe a physical state of a bulk homogeneous system. Analogously,
isotherms with negative slopes $\partial p/\partial\rho < 0$ have
negative bulk moduli and are unstable. The liquid-gas system resolves
this instability by phase-separating into regions of different
densities (according to the lever rule to maintain the average
density). Similarly, the banded flows seen in the experiments
described above appear to be a non-equilibrium phase separation into
regions of high and low strain-rate, maintaining the applied mean
strain rate.
In previous work \cite{olmsted90,olmsted92,olmstedlu97} we constructed
a `phase diagram' by pursuing an analogy between homogeneous stable
steady states and equilibrium phases. As in equilibrium,
non-equilibrium `phases' may be separated, in field variable space, by
hyper-surfaces representing continuous ({\sl e.g.\/} critical
points/lines) or discontinuous (`first-order') transitions.
Coexistence implies an inhomogeneous state spanning separate branches
of the homogeneous flow curves. Note, however, that there is an
ambiguity in connecting separate branches of the homogeneous flow
curves in Fig.~\ref{fig:both}. The top curve permits coexistence of
states with the same stress and different strain rates, while the
lower curve also allows coexistence of states with the same strain
rate and different stresses \cite{schmitt95,olmstedlu97}!
Fig.~\ref{fig:couette} shows that phase separation at a common
stress occurs such that the interface between bands is parallel to the
vorticity-velocity plane (annular bands, in Couette flow), while phase
separation at a common strain rate occurs with the interface
between bands parallel to the velocity--velocity-gradient plane
(stacked discs, in Couette flow).
{\begin{figure}
\par\columnwidth20.5pc
\hsize\columnwidth\global\linewidth\columnwidth
\displaywidth\columnwidth
\epsfxsize=3.0truein
\centerline{\epsfbox{figures/couette2.eps}}
\caption{Geometries for phase separation at common stress (left) or
strain-rate (right) in a Couette rheometer. For phase separation at
a common stress (left) phases $I$ and $II$ have different strain
rates, while at a common strain rate (right) they have different
stresses. $\hat{\textbf{z}}$ is the vorticity axis,
$\hat{\textbf{x}}$ is the flow direction, and $\hat{\textbf{y}}$ is
the flow gradient axis.}
\label{fig:couette}
\end{figure}}
This highlights a striking contrast between equilibrium and
non-equilibrium systems. In equilibrium the field variables (pressure,
temperature, chemical potential) are uniquely defined and determine
phase coexistence. In sheared fluids, one needs an extra field
variable to determine the extended phase diagram. However, {\sl for a
system with more than one choice of coexisting geometry, the
appropriate field variable may not necessarily be identified a
priori}. The complete answer of how to determine (theoretically)
the dynamic field variable is not known. Of course the nature of the
constitutive relation may help, for example the top curve of the
Fig.~\ref{fig:both} does not allow the strain rate as the field
variable. We will come back to discuss some possible answers to this
interesting problem in subsection \ref{which} (see also
\cite{schmitt95} for other suggestions).
Another important difference from equilibrium systems is evident when,
say assuming the system choose to form shear bands at common stress,
we try to determine at which stress a system shear bands. The
constitutive relations shown in Fig.~\ref{fig:both} are calculated for
homogeneous states, and there is no apparent prescription for
determining the selected banding stress, despite the experimental
evidence for a selected stress. A similar apparent degeneracy occurs
in first order phase transitions in equilibrium statistical mechanics,
but is easily resolved by demanding that the system minimize its total
free energy, or, equivalently, by appealing to the convexity of the
free energy of the equilibrium thermodynamic systems \cite{israel}.
This leads to equality of field variables between two phases and the
common tangent condition ({\sl e.g.\/} the Maxwell equal areas
construction for liquid-gas coexistence \cite{landaustat1}, or the
equal osmotic pressure condition, aided by equal chemical potential,
in rod suspensions \cite{buining93}).
In the shear band problem, an unambiguous resolution of this
degeneracy is to consider the full {\sl inhomogeneous\/} ({\sl i.e.\/}
non-local in space) equations of motion, and determine phase
coexistence by that choice of field variables (appropriately chosen by
hand) for which there exists a {\sl stationary\/} interfacial solution
to the steady-state differential equations of motion
\cite{olmsted90,spenley96,olmstedlu97}. For zero stress this
technique reduces, as it should, to minimization of the free energy.
The importance of inhomogeneous terms in fluid equations of motion has
been noted by several groups, who pointed out that the standard fluid
equations can have ill-defined mathematical solutions \cite{elkareh89}
if such terms are not included. Of course, if the phase diagram
depends sensitively on the form or magnitude of the inhomogeneous
terms one need a detail understanding of the underlying physics. The
use of a stable interface to select among possible coexisting states
was first postulated for non-linear dynamical systems, as far as we
know, by Kramer \cite{kramer81a}, and later by Pomeau \cite{pomeau86};
and first applied (independently) to complex fluids in
Ref.~\cite{olmsted92}. The inclusion of gradient terms in constitutive
relations is rapidly gaining acceptance, as recent preprints by Goveas
(phase separation of model blends of long and short polymers)
\cite{goveas99} and Dhont (introduction of model gradient terms to
resolve stress selection) \cite{dhont99} indicate.
In this work we study a model for rigid-rod suspensions in shear flow.
While there are certainly ongoing experiments on these systems
\cite{mather97}, the primary motivation for this extended work is to
explore the manner in which phase separation and coexistence occurs in
complex fluids in flow. The approximations used in obtaining our
equations are severe (including a decoupling approximation whose
defects are well-known \cite{CLF95}), and we expect qualitative
agreement at best. However, this is the first complete study of which
we are aware of non-equilibrium phase separation of a complex fluid in
flow for a concrete model, and we hope it illuminates the
phenomenology of flow-induced phase transitions.
\section{Methodology}
We seek the equations of motion for a solution of rod-like particles.
The most useful dynamic variables describing the long-wavelength
hydrodynamic degrees of freedom are the volume fraction $\phi({\bf
r})$, the fluid velocity ${\bf v}({\bf r})$, and the nematic order
parameter tensor
\begin{equation}
Q_{\alpha\beta}({\bf r}) = \langle \nu_{\alpha}\nu_{\beta} -
\case13\delta_{\alpha\beta}\rangle,
\end{equation}
where $\boldsymbol{\nu}$ is the rod orientations, $\langle\cdot\rangle$ denotes an average around the point ${\bf r}$. Previous
studies of liquid crystals under shear flow have been either for
thermotropics \cite{olmsted90,olmsted92}, where the issues we
present below associated with composition coupling are not present;
or homogeneous suspensions \cite{see90,bhave93},
where phase coexisting was not considered.
Our work below is based on the model extending that of Doi \cite{doi81,doikuzuu83}.
See, \emph{et al.} \cite{see90} studied the Doi model in shear flow,
but did not attempt to consider phase coexistence. Bhave, \emph{et al}
\cite{bhave93} analyzed this model in more detail, but did not
consider realistic phase separation behavior. We augment this model
with reasonable estimates for translational entropy loss upon phase
separation and for the free energy cost due to spatial
inhomogeneities. Zubarev studied shear-induced phase separation in a
variation of the Doi model in flow based on the equality of
non-equilibrium free energies, calculated from the flow-perturbed
orientational distribution function \cite{Zuba96}. Zubarev only
considered phase separation at a common strain rate, and did not treat
the rheological response (stress) of the system or log-rolling states.
\subsection{Equations of Motion}
The free energy ({\sl e.g.\/} as in Ref.~\cite{buining93}) is given by
\begin{eqnarray}
{\cal F\/}(\phi,\boldsymbol{Q})&& =k_{\scriptscriptstyle B}T \int\!d^3\!r
\left\{\phantom{\bigl\{}\!\!\!{\phi\over v_r}\log\phi +
{\left(1\!-\!\phi\right)\over v_s}\log\left(1\!-\!\phi\right)
\right. \nonumber\\
&&
+ {\phi\over v_r}
\left[ \case12 \left(1\!-\!\case13 u\right)\hbox{\rm Tr}
\,\boldsymbol{Q}^2\!
-\case13 u \hbox{\rm Tr}\,\boldsymbol{Q}^3\!
+\case14 u\left(\hbox{\rm Tr}\,\boldsymbol{Q}^2\right)^2
\right.
\nonumber\\
&&
\left.\left. + \case12 K\left(\nabla_{\alpha}Q_{\beta\lambda}\right)^2 \right]
+ \case12 \frac{g}{v_s} \left(\nabla\phi\right)^2 \right\}.
\label{eq:free}
\end{eqnarray}
Here, ``${\rm Tr}$'' denotes the trace, $v_r$ and $v_s$ are rod and
solvent monomer volumes and
\begin{equation}
u\equiv\nu_2 c d L_0^2,
\end{equation}
is Doi's excluded volume parameter \cite{doi81,doikuzuu83}, where $c$
is the concentration (number/volume) of rods of length $L_0$ and
diameter $d$, and $\nu_2$ is a geometrical prefactor
(Ref.~\cite{doi81} estimated $\nu_2=5\pi/16\simeq0.98$). The
volume fraction $\phi$ is
\begin{equation}
\label{eq:phi}
\phi = c v_r,
\end{equation}
in terms of which
\begin{equation}
u=\phi L \frac{\nu_2}{\alpha},
\end{equation}
where $L=L_0/d$ is the rod aspect ratio and $\alpha$ is an ${\cal
O\/}(1)$ prefactor defined by
\begin{equation}
v_r = \alpha d^2L_0.
\end{equation}
For spherocylinders, $\alpha=\pi [1 - 1/(3L)] / 8$ which reduces, in
the limit $L\rightarrow\infty$, to $\alpha=\pi/8\simeq0.39$. We
use $u$ and $\phi$ interchangeably below as a composition
variable.
In much of what follows we make two further assumptions to reduce
the number of parameters in our model. We fix $v_s$
by assuming
\begin{equation}
v_r = L v_s,
\label{eq:approx1}
\end{equation}
which corresponds to a particular volume of the solvent molecules
relative to that of the rod-like molecules. Further, we assume that
the geometric factor $\nu_2/\alpha$ has the value unity, so that
\begin{equation}
u = \phi L,
\label{eq:approx2}
\end{equation}
which corresponds to a particular shape of the rigid-rod molecules.
These two assumptions specify the detailed shape and volume ratio of
the system we study below. For slightly different systems with $v_r
\ne L v_s$ or $\nu_2/\alpha \ne 1$, our work should still provide an
accurate qualitative picture.
The first two terms of Eq.~(\ref{eq:free}) comprise the entropy of
mixing, and the first three terms in square brackets are from Doi's
expansion of the free energy (derived per solute molecule) in powers
of the nematic order parameter $\boldsymbol{Q}$. These terms were
derived from the Smoluchowski equation for the distribution function
of rod orientations \cite{doi81,doikuzuu83}. We keep the expansion to
fourth order to describe a first order transition and give the correct
qualitative trends.
Assuming Eqs.~(\ref{eq:approx1}-\ref{eq:approx2}),
we calculate the following biphasic coexistence regions,
\begin{eqnarray}
\{u_{I}=2.6925, u_{N}=2.7080\} &&\qquad (L=5.0)\label{eq:INa} \\
\{u_{I}=2.6930, u_{N}=2.7074\} &&\qquad (L=4.7),\label{eq:INb}
\end{eqnarray}
where $u_I$ and $u_N$ are the excluded volume parameters
(compositions) for the coexisting isotropic and nematic phases,
respectively. Note the very weak dependence of the biphasic regime (in
the scaled variable $u=L\phi$) on $L$.
The last two terms in Eq.~(\ref{eq:free}) penalize spatial
inhomogeneities. By adding the single term proportional to $K$ we
have assumed a particular relation for the Frank constants,
($K_1\!=\!K_2\!=\!K, K_3\!=\!0)$ \cite{lubensky70,pgdg}. Although
Odijk has calculated these constants for model liquid crystals (in the
nematic regime) \cite{odijk86}, we will see below that this choice is
probably unimportant for this model. More generally, we expect the
Frank constants to vary as functions of $\boldsymbol{Q(r)}$ in
physical systems, a situation which we have not addressed here. The
final term penalizes composition gradients \cite{gunton}. We are not
aware of any calculations of $g$ for solutions of rod-like particles.
In Eq.~(\ref{eq:free}), we assume an athermal solution with no
explicit interaction energy.
The nematic order parameter obeys the following equation of motion
\cite{doi81,doikuzuu83}:
\begin{equation}
\left(\partial_t + {\rm\bf v}\!\cdot\!\boldsymbol{\nabla}\right) \boldsymbol{Q} =
\boldsymbol{F}(\boldsymbol{\kappa},\boldsymbol{Q}) + \boldsymbol{G}
(\phi, \boldsymbol{Q})
\label{eq:2}
\end{equation}
where $\kappa_{\alpha\beta} = \nabla_{\beta} v_{\alpha}$. In
Eq.~(\ref{eq:2}) the (reactive) ordering term $\boldsymbol{F}$ is given by
\begin{equation}
\boldsymbol{F}(\boldsymbol{\kappa},\boldsymbol{Q})\!=\!\case23\boldsymbol{\kappa}^{s}\!+
\!\boldsymbol{\kappa}\!\cdot\!\boldsymbol{Q}\!+\!
\boldsymbol{Q}\!\cdot\!\boldsymbol{\kappa}^{\scriptscriptstyle T}
\!-\!2(\boldsymbol{Q}\!+\!\case13\boldsymbol{I})\,\hbox{\rm
Tr}(\boldsymbol{Q}\!\cdot\!\boldsymbol{\kappa}),\label{eq:convect}
\end{equation}
where $\boldsymbol{\kappa}^s$ is the symmetric part of
$\boldsymbol{\kappa}$ and $\boldsymbol{I}$ is the identity tensor.
For simplicity, we have chosen the form appropriate for an infinite
aspect ratio (the prefactors differ by ${\cal O\/}(1)$ constants for
finite aspect ratios \cite{doikuzuu83}). The coupling
$\boldsymbol{F}$ to the flow both induces order, and dictates a
preferred orientation. The dissipative portion $\boldsymbol{G}$ is
\begin{equation}
\boldsymbol{G}(\phi,\boldsymbol{Q}) = 6 \frac{ \bar{D}_{\mit
r}}{k_{\scriptscriptstyle B}T}\frac{v_r}{\phi} \boldsymbol{H},
\label{eq:G}
\end{equation}
where
\begin{equation}
\label{eq:Dr}
\bar{D}_r=\frac{\nu_1 D_{r0}}{(1-\case32\,\hbox{\rm Tr}\,\boldsymbol{Q}^2)^2
(c L_0^3)^2},
\end{equation}
is the collective rotational diffusion coefficient and
\begin{equation}
\boldsymbol{H} = -\left[\frac{\delta{\cal F\/}}{\delta \boldsymbol{Q}}
-\case13\,\boldsymbol{I}\,\textrm{Tr}\frac{\delta{\cal F\/}}{\delta
\boldsymbol{Q}}
\right]
\end{equation}
is the molecular field. $D_{\mit r0}$ is the single-rod rotational
diffusion coefficient and $\nu_1$ is an ${\cal O\/}(1)$ geometrical
prefactor, which will be fixed below Eq.~3.30.
The rotational diffusion coefficient is
\begin{equation}
\label{eq:Dr0}
D_{r0} = \frac{k_{\scriptscriptstyle B}T\ln L}{3\pi\eta L_0^3},
\end{equation}
where $\eta$ is the solvent viscosity. The $\boldsymbol{Q}$-dependence
in the denominator of Eq.~(\ref{eq:Dr}) enhances reorientation for
well-ordered systems \cite{doi81}. Our choice for $\bar{D}_r$ is
crude, since it applies to rods in concentrated solution and we use it
in the concentrated and semi-dilute regimes. As with many of our
approximations, this gives us a tractable model system with which to
study the phenomenology of phase separation.
Doi and co-workers derived Eq.~(\ref{eq:2}) for homogeneous systems.
We extend this to inhomogeneous systems by including the gradient
terms implicit in the functional derivative which defines
$\boldsymbol{H}$. Our choice of $\boldsymbol{F}$ is the so-called
quadratic closure approximation to the Smoluchowski equation
\cite{doikuzuu83}. This approximation ensures that the magnitude of
the order parameter remain in the physical range in the limit of
strong ordering, but is known to incorrectly predict phenomena such as
director tumbling and wagging. Many workers have investigated the
subtleties of various closure approximations and the degree to which
they reproduce realistic flow behavior \cite{CLF95}. Since our primary
goal is to explore the method for calculating phase behavior and
outline some of the possibilities for coexistence under flow, we
confine ourselves to this well-studied model.
The fluid velocity obeys \cite{doi81,doikuzuu83,lulu}:
\begin{equation}
\rho\left(\partial_t + {\rm\bf v}\cdot\boldsymbol{\nabla}\right) {\rm\bf v} =
\boldsymbol{\nabla}\!\cdot\!\left[2\eta\boldsymbol{\kappa}^s +
\boldsymbol{\sigma} (\phi, \boldsymbol{\kappa},\boldsymbol{Q}) \right]
+ \frac{\delta\cal F}{\delta\phi}\boldsymbol{\nabla}\!\phi
-\boldsymbol{\nabla} p,
\label{eq:1}
\end{equation}
where $\eta$ is the solvent viscosity, $\rho$ the fluid mass density,
and the pressure $p$ enforces incompressibility,
$\boldsymbol{\nabla}\!\cdot\!{\rm\bf v}=0$. For the low Reynold's
number situations considered here, and for steady shear flow, we will
equate the left-hand side of the equation above to zero.
The constitutive relation for the stess tensor $\boldsymbol{\sigma} (\phi,
\boldsymbol{\kappa},\boldsymbol{Q})$ was derived by Doi and co-workers, and
includes dissipative and elastic parts. Since the elastic stress
dominates \cite{doiedwards}, we keep only this part:
\begin{eqnarray}
\boldsymbol{\sigma}&\simeq& \boldsymbol{\sigma}_{\mit elastic} \nonumber \\
&=& - 3\boldsymbol{H} +
\boldsymbol{H}\!\cdot\!\boldsymbol{Q} - \boldsymbol{Q}\!\cdot\!\boldsymbol{H}
- \boldsymbol{\nabla}Q_{\alpha\beta}\cdot
{\delta{\cal F\/}\over\delta\boldsymbol{\nabla}Q_{\alpha\beta}}.
\label{eq:stress}
\end{eqnarray}
The first term of Eq.~(\ref{eq:stress}) was given by Doi \cite{doi81},
while the last three terms were derived later \cite{olmsted90} and are
equivalent to the elastic stress due to Frank elasticity \cite{pgdg},
generalized to a description in terms of the nematic order parameter
$\boldsymbol{Q}$ rather than the nematic director. Note that the last
three terms vanish for a homogeneous system.
Finally, the composition equation of motion is of the Cahn-Hilliard
form \cite{gunton};
\begin{align}
\left(\partial_t + {\rm\bf v}\cdot\boldsymbol{\nabla}\right) \phi &=
- \boldsymbol{\nabla}\cdot\boldsymbol{J} \nonumber\\
&=\boldsymbol{\nabla}\!\cdot\!\boldsymbol{M}\!\cdot\!\boldsymbol{\nabla} \mu,
\label{eq:3}
\end{align}
where $\boldsymbol{M}$ is the mobility tensor and the chemical potential is
given by
\begin{equation}
\mu = \frac{\delta {\cal F\/}}{\delta\phi}.
\label{eq:mu0}
\end{equation}
The diffusive current is $\boldsymbol{J} =
-\boldsymbol{M}\!\cdot\!\boldsymbol{\nabla}\mu$. The complete
dynamics is thus described by Eqs.~(\ref{eq:2},\ref{eq:1})
and~(\ref{eq:3}).
The dynamical equations of motion for other complex fluids have the
same theoretical structure: equations of motion for the conserved
quantities and the broken-symmetry or flow-induced structural order
parameter (analogous to $\boldsymbol{Q}$), and a constitutive relation
for the stress as a function of composition and order parameter
\cite{lulu}. For a given system and set of equations of motion, the
analysis below is generic. For some local models, internal dynamics
(Eq.~\ref{eq:2}) can be eliminated to give the stress as a history
integral over the strain rate. In polymer melts \cite{doiedwards}, and
in wormlike micelles \cite{cates90} far from a nematic regime, this
leads to non-monotonic stress--strain-rate curves. However, augmenting
these integral theories with non-local terms to calculate interface
profiles is non-trivial.
\subsection{Steady-state conditions}
In this work we study planar shear flow, specified by
\begin{equation}
\frac{\partial v_x({\rm\bf r})}{\partial y} = \dot{\gamma}(\textbf{r}).
\end{equation}
For homogeneous flows ${\rm\bf v}({\rm\bf r}) = \dot{\gamma} y
{\rm\bf\hat{x}}.$
The phase diagram is given by the domains of stable
steady-state solutions to the equations of motion for applied shear
stress or strain-rate, in the phase space spanned by
\begin{align}
({\phi},{\sigma}_{xy}),&&&\text{(common stress)}\\
({\phi},{\dot{\gamma}}),&&&\text{(common strain rate).}
\end{align}
For phase separation at common stress the stress is uniform and the
strain rate partitions between the two phases; while for phase
separation at common strain rate the strain rate is uniform and the
shear stress partitions between the two phases.
The strain rate tensor is given by
\begin{equation}
\boldsymbol{\kappa} = \dot{\gamma}\left(
\begin{array}{ccc}
0 & 1 & 0 \\
0 & 0 & 0 \\
0 & 0 & 0
\end{array} \right).
\end{equation}
Upon rescaling,
\begin{eqnarray}
\widehat{\dot{\gamma}} &=& \frac{\dot{\gamma} L^2}{6D_{\mit r0}\nu_1\nu_2^2} \\
\widehat{\boldsymbol{\sigma}} &=&\frac{\sigma\nu_2 L^3}{3
k_{\scriptscriptstyle B} T},
\end{eqnarray}
the steady-state condition for the order parameter (Eq.~\ref{eq:2}) is
\begin{equation}
0 = \frac{1}{u^2L^2(1-\case32\hbox{\rm Tr}\boldsymbol{Q}^2)^2}
\widehat{\boldsymbol{H}}
+ \widehat{\dot{\gamma}}\,\widehat{\boldsymbol{F}},
\label{eq:Q}
\end{equation}
where $\boldsymbol{F} = \dot{\gamma} \widehat{\boldsymbol{F}}$ and
\begin{equation}
\label{eq:H}
-\widehat{\boldsymbol{H}} = \left(1\!-\!\frac{\scriptstyle u}{\scriptstyle
3}\right)
\boldsymbol{Q} - u\left(\boldsymbol{Q}^2\!-\!
\frac{\boldsymbol{I}}{3}\textrm{Tr}\boldsymbol{Q}\right) +u
\boldsymbol{Q}\textrm{Tr}\boldsymbol{Q}^2 - K\nabla^2\boldsymbol{Q}.
\end{equation}
In steady state planar shear flow the velocity gradients are normal to
the flow direction, so the convective derivative vanishes and
Eq.~(\ref{eq:Q}) specifies the order parameter in a homogeneous flow.
Under these conditions integration of the momentum equation
Eq.~(\ref{eq:1}) gives a constant stress,
\begin{equation}
\boldsymbol{\sigma}_0=\boldsymbol{\sigma} - p\,\boldsymbol{I} + 2 \eta\boldsymbol{\kappa}^s,
\end{equation}
where $\boldsymbol{\sigma}_0$ is the boundary stress. The rescaled
shear stress is
\begin{equation}
\widehat{\sigma}_{xy}^0 = A \widehat{\dot{\gamma}} - u L \left[
\widehat{\boldsymbol{H}}+ K \left(\nabla^2\boldsymbol{Q}\!\cdot\!\boldsymbol{Q} -
\boldsymbol{Q}\!\cdot\!\nabla^2\boldsymbol{Q}\right)\right]_{xy},
\label{eq:sigma}
\end{equation}
where $A=2\nu_1\nu_2^3(\ln L)/(3\pi)$ is a constant of order unity: we
take $A=1$ for the remainder of this work, which corresponds to a
particular choice for $\nu_1$. As with the assumptions of molecular
geometry embodied in $\nu_2$ and $\alpha$
(Eqs.~\ref{eq:approx1}-\ref{eq:approx2}), different values for $A$
should not qualitatively change the nature of our results.
Integrating the steady-state composition equation (\ref{eq:3}) and
using the boundary condition that material cannot enter or leave the
system, we find
\begin{eqnarray}
\mu_0 &=& \mu({\rm\bf r}) \label{eq:mu} \\
\frac{\mu(\textbf{r})}{k_{\scriptscriptstyle B}T} &=&
F_{\mit Doi} + \frac{\partial}{\partial\phi}\left[\phi\ln\phi + L
\left(1-\phi\right)\ln\left(1-\phi\right)\right] \label{eq:mucalc}
\nonumber \\
&& + \phi L \frac{\partial}{\partial u}
F_{\mit Doi} + \case12K \left(\nabla\boldsymbol{Q}\right)^2 - g L
\nabla^2\phi
\label{eq:mu1}
\end{eqnarray}
where Eqs.~(\ref{eq:approx1}-\ref{eq:approx2}) have been used to
specify the molecular geometry, $\mu_0$ is a constant of
integration, and
\begin{equation}
\label{eq:Fdoi}
F_{Doi} = \case12 \left(1\!-\!\case13 u\right)\hbox{\rm Tr}
\,\boldsymbol{Q}^2\!
-\case13 u \hbox{\rm Tr}\,\boldsymbol{Q}^3\!
+\case14 u\left(\hbox{\rm Tr}\,\boldsymbol{Q}^2\right)^2.
\end{equation}
Note that the mobility tensor $\boldsymbol{M}$ plays no role in the
steady-state conditions, or in the resulting phase diagram.
Eqs.~(\ref{eq:Q},\ref{eq:sigma}) and~(\ref{eq:mu}) completely specify
the system in planar shear. Solving these equations will occupy the
remainder of this work. Note that variables $\widehat{\dot{\gamma}},
\widehat{\boldsymbol{\sigma}},$ $\mu/k_{\scriptscriptstyle B} T$ are
all dimensionless quantities.
\section{Calculation of Phase Diagrams}
\subsection{Interface Calculation}
The phase diagram is specified by solving
Eqs.~(\ref{eq:Q},\ref{eq:sigma},\ref{eq:mu}) for given $\mu_0$ and
boundary stress $\sigma_{xy}^0$. Non-equilibrium `phases' are defined
as the stable steady-state space-uniform solutions to these equations.
These {\sl inhomogeneous\/} equations comprise a set of ordinary
differential equations, through the gradients that appear in the
stress and in the functional derivatives that define $\mu$ and
$\boldsymbol{H}$. The only parameters of the theory are the rod aspect
ratio $L$ and the ratio of elastic constants,
\begin{equation}
\lambda=\frac{gL}{K}
\end{equation}
($K$ may be absorbed into the length scale of the system).
{\begin{figure}
\par\columnwidth20.5pc
\hsize\columnwidth\global\linewidth\columnwidth
\displaywidth\columnwidth
\epsfxsize=3.5truein
\centerline{\epsfbox[70 230 670 650]{figures/fig_all.ps}}
\caption{Homogeneous stress $\hat{\sigma}_{xy}$ vs. strain rate
$\widehat{\dot{\gamma}}$ behavior for various excluded volumes,
$L=5.0$ and $\lambda=1.0$. Dotted lines mark unstable branches.
Similarly, curves along which $\partial\mu/\partial\phi<0$ are
linearly unstable.}
\label{fig:constit}
\end{figure}}
We first fix $\phi$ ($u$) and solve the homogeneous algebraic versions
of Eqs.~(\ref{eq:Q}) and~(\ref{eq:sigma}) for $\boldsymbol{Q}$ and
$\dot{\gamma}$ as a function of $\sigma_{xy}^0$\footnote{In the few
cases where the phase diagram in the $\sigma_{xy}\!-\!\mu$ plane has
a transition line parallel to the $\mu$ axis, one must first fix
$\mu_0$, and then determine $\sigma_0$.}. This is done for all
$\phi$. Because ${\cal F\/}(\phi,\boldsymbol{Q})$ describes an I-N
transition, at a given stress, multiple roots exist, with distinct
strain rates and $\boldsymbol{Q}$. Fig.~\ref{fig:constit} shows the
stress strain-rate relations for homogeneous solutions to
Eqs.~(\ref{eq:Q}) and~(\ref{eq:sigma}) for $L=5.0$ and $\lambda=1.0$.
{\begin{figure}
\par\columnwidth20.5pc
\hsize\columnwidth\global\linewidth\columnwidth
\displaywidth\columnwidth
\epsfxsize=3.5truein
\centerline{\epsfbox[0 0 712 370]{figures/fig3D.eps}}
\caption{Stress-strain-composition surface for the curves in
Fig~\protect{\ref{fig:constit}}. The plane is at
$\widehat{\sigma}_{xy}^0=0.05$}
\label{fig:3D}
\end{figure}}
The isotropic branch has a larger viscosity than the nematic branch,
and has an increasing effective viscosity for increasing
concentration, reflecting the contribution $u\widehat{\boldsymbol{H}}$
in Eq.~(\ref{eq:sigma}). Conversely, the nematic branch has a lower
stress at higher concentrations due to the increased nematic order
which permits less-hindered motion.
For a dilute isotropic system (curve \textsf{a}), shear flow
continuosly induces nematic order. A more aligned system has a lower
effective viscosity, so the stress $\sigma(\dot{\gamma})$ increases
slower than linearly (shear-thins) as the magnitude of the order
parameter $\boldsymbol{Q}$ increases. Eventually the system attains,
smoothly, a high strain rate state with a much lower viscosity than in
the limit of zero stress. For more concentrated systems (curves
\textsf{b}, \textsf{c} and \textsf{d}) shear flow induces a transition
to a nematic phase with lower viscosity, and $\sigma(\dot{\gamma})$ is
non-monotonic. There is a region of stresses for which two stable
strain rates exist, on either the nematic or isotropic branches of the
constitutive curve. For compositions inside the biphasic regime
(curve \textsf{e}) both nematic and isotropic branches exist in the
limit of zero stress, with the isotropic branch losing stability at
high enough stress. Finally (not shown) for highly concentrated
systems only nematic branches exist.
As mentioned in Sec.~IIB, we calculate the phase diagram by explicitly
constructing the coexisting interfacial solution \cite{olmsted92}. In
common stress coexistence, for example, the coexisting states have
{\sl different\/} strain rates and, generally, different compositions.
Hence, they connect the high and low strain rate branches of two
different curves in Fig.~\ref{fig:constit}. It is easiest to visualize
this by considering the intersection of a plane at a given stress
$\sigma_{xy}^0$ with the surface $\sigma_{xy}(\dot{\gamma},u)$, as in
Fig.~\ref{fig:3D}.
{\begin{figure}
\par\columnwidth20.5pc
\hsize\columnwidth\global\linewidth\columnwidth
\displaywidth\columnwidth
\epsfxsize=3.25truein
\centerline{\epsfbox[150 220 495 660]{figures/new_mu.ps}}
\caption{(a) Reduced strain rate $\widehat{\dot{\gamma}}(u)$ and
(b) chemical potential $\mu(u)$ for the stress contour in
Fig.~\protect{\ref{fig:3D}} ($\widehat{\sigma}_{xy}=0.05$). The tie
line is calculated using the interface construction.}
\label{fig:strainmu}
\end{figure}}
At a given stress, the strain rate varies with composition as shown in
Fig.~\ref{fig:strainmu}a. At coexistence, the chemical potential
$\mu({\rm\bf r})$ must be constant through the interface, as dictated
by Eq.~(\ref{eq:mu}). The functional form of the non-equilibrium
chemical potential is known from Eq.~(\ref{eq:mu1}), and depends on
the strain rate through the dependence of the nematic order parameter
on the strain rate in steady state. We plot $\mu(u)$ in
Fig.~\ref{fig:strainmu}b. There is a continuum range of $\mu$, which
allow possible coexisting pairs of states. (Recall that $u$ is
proportional to the rod volume fraction $\phi$).
We now impose the interface solvability condition as follows. For a
given stress $\sigma_{xy}^0$, we determine a specific coexistence
chemical potential $\mu_0$, which allows a stable interfacial solution
to Eqs.~(\ref{eq:Q},\ref{eq:sigma},\ref{eq:mu}). In practice, we
eliminate $\dot\gamma({\rm\bf r})$ from Eq.~(\ref{eq:Q}) using
Eq.~(\ref{eq:sigma}), and solve Eqs.~(\ref{eq:Q}) and~(\ref{eq:mu})
for the interfacial profile, with boundary conditions (fixed
$\boldsymbol{Q}$ and $u$) chosen by two points on the low and high
strain rate branches of Fig.~\ref{fig:strainmu}b with the same
$\mu_0$. We adjust $\mu_0$ until a stationary interfacial profile is
found. This solvability criterion give sharp selection on $\mu$, and
in this way determine a tie line on the $\dot{\gamma}\!-\!u$ plane,
Fig.~\ref{fig:strainmu}a. By varying the stress we compute the entire
phase diagram in the $\sigma_{xy}-u$ and $\dot{\gamma}-u$ planes.
For phase separation at a common strain rate the construction is
analogous. One slices a vertical plane through Fig.~\ref{fig:3D} at a
given strain rate, constructs the curve $\mu(u)$ along the
intersection with the surface, and searches for a stationary
interfacial solution.
The interface calculations are carried out by discretizing the system
on a one-dimension mesh and, from smooth initial conditions, evolving
Eqs.~(\ref{eq:Q},\ref{eq:sigma},\ref{eq:mu}) forward using fictitious
dynamics calculated with an implicit Crank-Nicholson scheme. Spatial
variations are only allowed in the direction in which phase separation
occurs, so we replace
\begin{equation}
\label{eq:5}
\nabla \rightarrow \begin{cases}
\dfrac{\partial}{\partial y} & \hbox{common stress} \\[10truept]
\dfrac{\partial}{\partial z} & \hbox{common strain rate,}
\end{cases}
\end{equation}
where $z$ is in the vorticity direction. We fix the values of $u$ and
$\boldsymbol{Q}$ at either side of the interface to lie on the high
and low strain rate branches of Fig.~\ref{fig:strainmu}, begin with
smooth initial conditions, and let the system ``evolve'' towards
steady state. An interface develops between the two phases, and moves
to one boundary or the other. For a given stress, coexistence is
determined by that chemical potential $\mu$ for which a stationary interface
lies in the interior of the system (in the limit of large system size)
\cite{olmsted92}. An analogous construction may be made by
maintaining a fixed mean strain rate on the unstable part of a
homogeneous curve, and then starting up the system and allowing it to
select a stress and chemical potential. In either case the selected
stress is that stress for which a stationary interfacial solution
between the high and low strain rate branches \emph{exists}. Such an
interfacial solution is known in dynamical systems theory as a
heteroclinic orbit \cite{kramer81a,pomeau86,olmstedlu97}, and further
work will investigate this in more detail for simpler model systems
\cite{jsplanar,history,orpdo}.
We restrict the nematic order parameter to
\begin{equation}
\boldsymbol{Q}=\left(
\begin{array}{ccc}
q_1 & q_3 & 0 \\
q_3 & q_2 & 0 \\
0 & 0 & - (q_1 + q_2)
\end{array}
\right),
\end{equation}
since all steady state solutions with non-zero elements $Q_{xz}$ or
$Q_{yz}$ are unstable due to the symmetry of shear flow
\cite{bhave93}. In a similar calculation for thermotropic nematics in
shear flow we have found that this restriction on $\boldsymbol{Q}$
reproduces the same selected stress as that obtained when keeping the
full tensor \cite{olmsted92}.
For planar shear flow and a wide class of equations of motion we have
shown that, if a coexisting solution exists, it occurs at discrete
points in the parameter set\cite{jsplanar}. For example, for a given
stress $\sigma_{xy}^0$, coexistence can occur only at discrete values
for $\mu_0$, that is, along lines in the field variable space spanned
by $\sigma_{xy}\!-\!\mu$. This is analogous to equilibrium systems
where, for example, phase transitions in a simple fluid occur along
lines, rather than within regions, in the pressure-temperature plane.
Note that a one-dimensional calculation does not determine the
stability of the interfacial solution with respect to transverse
undulations (capillary waves), which could be important in,
particularly, the common stress geometry \cite{renardy}.
\subsection{Homogeneous Solutions}
The modified Doi model in the quadratic closure approximation has
three stable solutions in homogeneous planar shear flow. We refer the
reader to Bhave, {\sl et al.\/} for further details \cite{bhave93}.
\begin{itemize}
\item[{\sf I}] \emph{Paranematic:} The paranematic state {\sf I}
induced from a disordered equilibrium phase. The order parameter
$\boldsymbol{Q}$ is small and fairly biaxial, with major axis lying
in the shear plane at an angle of almost $\pi/4$ relative to the
flow direction.
\item[\textsf{N}] \emph{Flow-Aligning Nematic:} The flow-aligning
nematic state is much more strongly-aligned, has slight biaxiality
induced by the flow, and has the major axis of alignment in the flow
plane at an angle of a few degrees relative to the flow direction.
The \textsf{I} and \textsf{N} states have the same symmetry.
\item[\textsf{L}] \emph{Log-Rolling Nematic:} The log-rolling phase is
also a well-aligned and almost uniaxial phase, but with major axis of
alignment in the vorticity $({\rm\bf\hat{z}})$ direction, so the
rods spin about their major axes.
\end{itemize}
{\begin{figure}
\par\columnwidth20.5pc
\hsize\columnwidth\global\linewidth\columnwidth
\displaywidth\columnwidth
\epsfxsize=3.5truein
\centerline{\epsfbox[40 210 590 640]{figures/roots.ps}}
\caption{Constitutive relations for \textsf{I}, \textsf{N}, and
\textsf{L} states, for $L=5.0$ and two values for the excluded
volume parameter.}
\label{fig:log}
\end{figure}}
The {\sf I} phase is stable at lower volume fractions and merges with
the {\sf N} phase at high strain rates. The {\sf L} phase is stable
only at higher volume fractions, and is destabilized at high enough
strain rates. For low strain rates the stress of the {\sf L} state is
lower than that of the {\sf N} state, which is lower than that of the
{\sf I} state (see Fig.~\ref{fig:log}).
Fig.~\ref{fig:stab} shows the regions of stability of the various
states. The loop in Fig.~\ref{fig:stab}a occurs for compositions such
that the constitutive curve $\sigma(\dot{\gamma})$ has the shape of
curve \textsf{b} in Fig.~\ref{fig:constit}. Similar phase-plane plots
were calculated by Bhave \emph{et al.} \cite{bhave93} and See
\emph{et al.} \cite{see90}. They did not consider the mixing entropy
needed to generate a realistic nematic transition, however, and always
generated solutions for a given strain rate instead of a given stress.
(this explains the absence of a loop in their phase-plane plot
$\dot{\gamma}\!-\!\phi$). Their plots (compare Fig.~5 of
Ref.~\cite{bhave93}) correspond to truncating the loop in
Fig.~\ref{fig:stab}a. Fig.~\ref{fig:stab}b has a similar, barely
discernable, loop near the critical point, within which there are no
stable states. This instability is due to the instability of the
composition equation, Eq.~(~\ref{eq:3}). In this region
\begin{equation}
\label{eq:6}
\left.\frac{\partial \mu}{\partial \phi}\right|_{\sigma_{\mit xy}} < 0,
\end{equation}
which is equivalent to a negative diffusion coefficient, and is
analogous to the conventional definition of the spinodal line for
ordinary equilibrium demixing.
{\begin{figure}
\par\columnwidth20.5pc
\hsize\columnwidth\global\linewidth\columnwidth
\displaywidth\columnwidth
\epsfxsize=3.5truein
\centerline{\epsfbox[150 230 500 640]{figures/limits.ps}}
\caption{Regions of stability of paranematic ({\sf I}),
nematic ({\sf N}) and log-rolling ({\sf L}) states in the
strain-rate--composition (a) and stress-composition (b) planes for
$L=5.0$. Note that the loop in (a) contains \emph{no} stable states.
Stability limits are calculated with respect to both order parameter
and composition fluctuations, for a given controlled stress. The
thin loop in (b) encloses a region with no stable states, due to the
instability of the composition equation.}
\label{fig:stab}
\end{figure}}
\section{Common Stress Coexistence}
For common stress coexistence the interface lies in the
velocity-vorticity plane, and inhomogeneities are in the
${\rm\bf\hat{y}}$ direction (see Fig.~\ref{fig:couette}). The stress
balance condition at the interface is
$\boldsymbol{\sigma}\!\cdot\!{\rm\bf\hat{y}}$ uniform. $\sigma_{yy}$
is taken care of by the pressure and $\sigma_{zy}$ vanishes by
symmetry (no flow in the ${\rm\bf\hat{z}}$ direction), leaving
continuity of the shear stress $\sigma_{xy}$ through the interface.
The two coexisting phases $I$ and $II$ have strain rates and
compositions partitioned according to
\begin{eqnarray}
\bar{\phi} &=& \zeta \phi_{\scriptscriptstyle I} + (1-\zeta)
\phi_{\scriptscriptstyle II} \label{eq:lev1}\\
\bar{\dot{\gamma}} &=& \zeta \dot{\gamma}_{\scriptscriptstyle I} + (1-\zeta)
\dot{\gamma}_{\scriptscriptstyle II},\label{eq:lev2}
\end{eqnarray}
where $\bar{\phi}$ and $\bar{\dot{\gamma}}$ are the mean composition
and strain rate and $\zeta$ is the fraction of material in phase $I$.
\subsection{Paranematic--flow aligning coexistence ({\sf I-N})}
\noindent{\textbf{Phase Diagram---}}Fig.~\ref{fig:tie}
shows the tie lines computed on the $(\hat{\sigma}_{xy}\!-\!u)$ and
$(\widehat{\dot{\gamma}}\!-\!u)$ planes according to the procedure
outlined in Section~III. Several features should be noted. Flow
induces nematic behavior in what, in equilibrium, would be an
isotropic phase. The tie lines are horizontal in the
$(\hat{\sigma}_{xy}\!-\!u)$ plane, since phases coexist at a
prescribed stress; and have a positive slope in the
$(\widehat{\dot{\gamma}}\!-\!u)$ plane because the more concentrated
nematic phase flows faster. There is a critical point at sufficiently
strong stress, whose existence is expected since the flow-aligning
nematic and paranematic states have the same symmetry
($\boldsymbol{Q}$ is biaxial) and their major axes in the shear plane.
More interesting is the changing slope of the tie lines. For weak
stresses the equilibrium system is slightly perturbed and the tie
lines are almost horizontal. For high stresses the tie lines become
more vertical and the composition difference between the phases
decreases. The slope of the tie lines determines the shape of the mean
stress--strain-rate relation $\bar{\sigma}_{xy}(\bar{\dot{\gamma}})$
that would be measured in steady state experiments.
{\begin{figure}
\par\columnwidth20.5pc
\hsize\columnwidth\global\linewidth\columnwidth
\displaywidth\columnwidth
\epsfxsize=3.5truein
\centerline{\epsfbox[160 230 485 660]{figures/phase0IN.ps}}
\caption{
Phase diagram in the $(\hat{\sigma}_{xy}\!-\!u)$ (a) and
$(\widehat{\dot{\gamma}}\!-\!u)$ (b) planes for $L=5.0,
\lambda=1.0$, along with the limits of stability of \textsf{I} and
\textsf{N} phases}
\label{fig:tie}
\end{figure}}
\end{multicols}\widetext
{\begin{figure}
\epsfxsize=\displaywidth
\centerline{\epsfbox[100 220 1400 635]{figures/sep_all.ps}}
\caption{Mean stress--strain-rate curves for coexistence at common
stress, for $L=5.0$ and $\lambda=1.0$. The solid lines denotes
\textsf{I} and \textsf{N} branches; the dotted line in each figure
denotes the stable \textsf{N} branch with which the \textsf{I} state
coexists at the low strain rate boundary of the coexistence region,
at a strain rate marked by an open circle $\boldsymbol{\circ}$. The
solid circles $\bullet$ and thick solid line denote the stress that
would be measured in the banded regime. Phase coexistence occurs
between phases of \textit{different} compositions than the mean
compositions ($u=2.555, 2.58, 2.685$). The unstable portion of the
homogeneous flow curve is shown in (a) and (b), but not (c). Note
that the plateaus in the two-phase regions in (a) and (b) rather
obviously do \emph{not} satisfy an equal area construction with the
underlying constitutive curve at the mean composition.}
\label{fig:stress-strainbar0}
\end{figure}}
\begin{multicols}{2} \narrowtext
\noindent{\textbf{Mean Constitutive Relations---}}Consider a
composition in the range where phase-separation occurs. For small
applied stress $\bar{\sigma}_{xy}(\bar{\dot{\gamma}})$ varies smoothly
until the two-phase region is reached. At this stress, a tiny band of
high strain rate strongly-aligned nematic material appears, with
volume fraction determined by the lever rule, Eq.~(\ref{eq:lev1}). The
mean strain rate $\bar{\dot{\gamma}}$ is determined by the lever rule,
Eq.~(\ref{eq:lev2}), and the measured constitutive relation
$\bar{\sigma}_{xy}(\bar{\dot{\gamma}})$ is non-analytic at this point
(see Fig.~\ref{fig:stress-strainbar0}). As the stress is increased
further, the system traverses the two-phase region by jumping from tie
line to tie line. Each successive tie line has a higher stress, a
higher mean strain rate, and a steadily increasing volume fraction of
nematic phase. The compositions of both coexisting phases change
steadily through the two-phase region.
The constitutive relation $\bar{\sigma}_{xy}(\bar{\dot{\gamma}})$
through the two phase region is determined by the spacing and splay of
the tie lines. For mean compositions $\bar{\phi}$ close to the
equilibrium isotropic-nematic transition
(Fig.~\ref{fig:stress-strainbar0}c) the tie lines in the
$(\widehat{\dot{\gamma}}\!-\!u)$ plane are fairly flat, so that the
stress $\sigma_{xy}$ changes significantly through the two-phase
region; and the `plateau' has definite curvature, reflecting the
initial splay of the tie lines. For slightly lower mean compositions
(Fig.~\ref{fig:stress-strainbar0}b) the `plateau' is straighter and
flatter, as can be seen in (Fig.~\ref{fig:stress-strainbar}), because
the lines are more vertical in the $(\widehat{\dot{\gamma}}\!-\!u)$
plane. Finally, for compositions near the critical point the plateau
is flatter still but, more interestingly, phase coexistence occurs in
a region where the stress-strain curve at the mean composition is no
longer non-monotonic (Fig.~\ref{fig:stress-strainbar0}a)! This is
because stability in a two-phase system is also determined by the
stability with respect to composition variations. In fact, the local
chemical potential $\mu(u)$ has negative slope and is unstable on a
segment of this curve. The tie line construction is a graphical
expression of the explanation proposed by Schmitt {\sl et al.\/}
\cite{schmitt95}, who attributed a sloped plateau to
composition-dependence of the stress-strain constitutive relation.
The general relation is given by
\begin{equation}
\frac{\partial\sigma}{\partial\bar{\dot{\gamma}}} = \left[
\frac{\zeta}{\eta_I} + \frac{1-\zeta}{\eta_N} -
m(\sigma)\left\{\frac{1-\zeta}{\dot{\gamma}'_N\eta_N} +
\frac{\zeta}{\dot{\gamma}'_I\eta_I}\right\}
\right]^{-1},
\end{equation}
where $m(\sigma)$ is the slope of tie line with stress value $\sigma$,
the lines $\{\sigma_I(\phi),\sigma_N(\phi), \dot{\gamma}_I(\phi),
\dot{\gamma}_N(\phi)\}$ bound the phase coexistence domains in the
$\sigma\!-\!\phi$ and $\dot{\gamma}\!-\!\phi$ planes;
$\eta_{k}=\partial\sigma_{k}/\partial{\dot{\gamma}}$ is the local
viscosity of the $k$th branch, and
$\dot{\gamma}'_{k}=\partial\dot{\gamma}_{k}/\partial\phi$.
{\begin{figure
\par\columnwidth20.5pc
\hsize\columnwidth\global\linewidth\columnwidth
\displaywidth\columnwidth \epsfxsize=3.5truein \centerline{\epsfbox[60
210 660 620]{figures/sep_all_ovl.ps}}
\caption{$\hat{\sigma}_{xy}$ vs. $\widehat{{\dot{\gamma}}}$
for common stress coexistence for $L=5.0$ and $\lambda=1.0$. The
solid lines connecting the high and low strain rate branches at each
composition denote the composite flow behavior at coexistence.}
\label{fig:stress-strainbar}
\end{figure}}
\noindent{\textbf{Measurements at controlled stress or controlled
strain rate---}}Although these calculations are for phase
separation at a common stress, one may perform experiments at either
controlled stress or strain rate. All three composite curves in
Fig.~\ref{fig:stress-strainbar0} have similar shapes, so we expect the
same qualitative behavior for all compositions. Controlled strain rate
experiments should follow the homogenous flow curves, except for
strain rates in the coexistence regime. Here we expect the steady
state to eventually be the banded state. This should presumably occur
by a nucleation event after some time, for start-up strain rates less
than the \textsf{I} limit shown in Fig.~\ref{fig:tie}a; and occur
immediately for imposed strain rates beyond this stability limit.
Conversely, upon decreasing the strain rate from the nematic phase we
expect nucleated behavior for strain rates larger than the \textsf{N}
limit, and instability for smaller strain rates. In the metastable
regime we expect the flow curve to follow the underlying homogeneous
constitutive curve for the given composition, until the nucleation
event occurs. Interestingly, there is a small region (inside the loop
in Fig.~\ref{fig:tie}a) where the system is unstable when brought, at
controlled strain rate, into this region from either the \textsf{I} or
\textsf{N} states. This corresponds to constitutive curves with the
multi-valued behavior of curve \textsf{b} in Fig.~\ref{fig:constit}.
Controlled stress experiments should exhibit similar behavior.
Consider Fig.~\ref{fig:stress-strainbar0}b. For initial applied
stresses larger than the minimum coexistence stress and less than the
\textsf{I} limit of stability in Fig.~\ref{fig:tie}b, we expect the
system to follow the homogenous flow curve until a nucleation event
occurs. After nucleation the strain rate should increase, until
either the proper plateau strain rate or the high strain rate nematic
state is reached, depending on the magnitude of the stress. For
stresses larger than the limit of stability we expect the system to
become immediately unstable to either a banded flow or a homogeneous
nematic phase, depending on the magnitude of the stress.
\noindent{\textbf{Metastability:~Experiments---}}Experiments on
wormlike micelles \cite{berret94a,berret94b,grand97} have found
constitutive curves analogous to those in, say,
Fig.~\ref{fig:stress-strainbar0}. In these experiments the plateau
appears to be the stable states, while the portion of the constitutive
curve (a `spine') which extends to stresses above the onset of the
stress plateau appears to be a metastable branch on which the system
may remain for a finite period of time under controlled stress or
strain rate conditions. Refs.~\cite{berret94b,grand97,Berr97}
conducted controlled strain-rate experiments and found that the system
follows the composite curves (without `spines' that extend above the
onset of the stress plateau) in Fig.~\ref{fig:stress-strainbar0}, if
care is taken to reach steady state. In these systems the plateaus
were nearly flat, suggesting a very slight dependence of the flow
behavior on composition. For controlled strain rate quenches into
what corresponds to the two-phase region of Fig.~\ref{fig:tie}a, the
system took some time to develop shear bands and phase separate. This
relaxation or `nucleation' time decreased as the mean-strain rate was
increased \cite{grand97}. It is not clear that they reached a limit
of stability (which would be analogous to the \textsf{I} limit in
Fig.~\ref{fig:tie}). The relaxation times were of order $60-600\,{\rm
s}$, depending on temperature, mean composition, and mean strain
rate. We emphasize that these experiments were on micellar solutions,
which probably do not show an isotropic-nematic transition, but still
display the same qualitative stress--strain-rate relationship as curve
\textsf{b} in Fig.~\ref{fig:constit}.
Ref.~\cite{Call+96} revealed different stress plateaus upon
controlling either the strain rate or the shear stress (see Fig.~7 of
Ref.~\cite{Call+96}) in cone-and-plate flow. In controlled stress
experiments the stress plateau occurred at a stress of order $1.5$
times the stress plateau observed under controlled strain rate
conditions. Moreover, the flow curve under controlled stress
conditions exhibited a stress maximum and then a decrease in stress to
an approximate flat plateau. One explanation for the high stress
plateau under controlled stress conditions could be that the `spine'
never nucleated under controlled stress conditions, and the system
smoothly transformed to the high strain rate phase. However, we do
not have an explanation for the decrease and subsequent plateau in
stress under applied strain rate conditions.
In other experiments, controlled stress experiments revealed two kinds
of metastable behavior \cite{grand97}. For
$\sigma_p<\sigma<\sigma_{jump}$, where $\sigma_p$ is the minimum
stress for the onset of banding in controlled strain rate experiments,
the system maintained a strain rate on the `metastable' branch for
indefinite times (measured times were up to $10^4\,\hbox{s}$). For
$\sigma>\sigma_{jump}$ the system accelerated, after of order
$10^3\,\hbox{s}$, and left the rheometer. For these systems it is not
clear whether a stable high shear branch exists. An explanation for
$\sigma_{jump}$ is lacking. Evidently the nucleation processes
governing metastability at controlled stress and controlled strain
rate are different. Clearly we need more experiments and theory about
the nature of nucleation and metastability in controlled stress vs.
controlled strain rate experiments.
\noindent{\textbf{Polydispersity---}}Fig.~\ref{fig:tieL} shows the
effect of rod aspect ratio $L$ on the phase diagram. A smaller rod
aspect ratio couples more weakly to the flow, requiring a slightly
larger strain rate to induce a transition to the nematic phase
(Fig.~\ref{fig:tieL}a). The resulting stress is slightly smaller
because, when the system enters the two phase region the stress is
largely determined by that of the paranematic branch, which decreases
with increasing $L$ (Fig.~\ref{fig:tieL}b). Although the equilibrium
phase boundaries are close (see Eqs.~\ref{eq:INa}-\ref{eq:INb}), the
deviation is amplified considerably by applying flow. This suggests
that flow enhances the natural tendency of length polydispersity to
widen biphasic regimes.
{\begin{figure}
\par\columnwidth20.5pc
\hsize\columnwidth\global\linewidth\columnwidth
\displaywidth\columnwidth
\epsfxsize=3.5truein
\centerline{\epsfbox[140 230 500 650]{figures/L.ps}}
\caption{
Phase diagrams for $L=5.0$ and $L=4.7$ at common stress, for
$\lambda=1.0$.}
\label{fig:tieL}
\end{figure}}
\subsection{Paranematic--Log rolling coexistence ({\sf I-L})}
Fig.~\ref{fig:tieIL} shows the phase diagram calculated for
coexistence between paranematic ({\sf I}) and log rolling ({\sf L})
states. As with \textsf{I-N} coexistence, the zero shear limit
corresponds to the equilibrium biphasic region. However, for non-zero
stress the biphasic region shifts in the direction of higher
concentration. This is reasonable, since the stability limit of the
{\sf L} phase shifts to higher concentrations with increasing stress
(Fig.~\ref{fig:stab}). Note also that, since the {\sf I} and {\sf L}
phases have major axes of alignment in orthogonal directions, there is
no critical point. Instead, the window of phase coexistence ends when
the {\sf I} phase becomes unstable to the {\sf N} phase.
We have also computed phase coexistence between {\sf N} and {\sf L}
phases. This occurs at much higher compositions ($u> u_{\ast} \agt
3.0$) and has a narrow width in composition due to the very slight
difference in viscosities of the two phases. Unfortunately, we cannot
resolve this coexistence regime accurately within the numerical
precision of our calculations and do not present these results here.
The existence of two possible phase diagrams for common stress phase
separation raises an interesting question. Can one observe {\sf I-L}
coexistence? Notice that {\sf I-L} coexistence can only occur for
samples prepared at concentrations at or above that necessary for
equilibrium phase separation. One could prepare a phase separated
isotropic-nematic mixture and, by wall preparation, field alignment,
sedimentation, or other techniques, separate the phases into two
macroscopic domains with the nematic phase in the log-rolling
geometry.
{\begin{figure}
\par\columnwidth20.5pc
\hsize\columnwidth\global\linewidth\columnwidth
\displaywidth\columnwidth
\epsfxsize=3.5truein
\centerline{\epsfbox[140 245 500 645]{figures/phaseIL.ps}}
\caption{
Phase diagram in the $(\hat{\sigma}_{xy}\!-\!u)$ and
$(\widehat{\dot{\gamma}}\!-\!u)$ planes for paranematic--log-rolling
coexistence, for $L=5.0$ and $\lambda=1.0$. The dotted lines are the
limits of stability of the {\sf I} and {\sf L} phases (see
Fig.~\protect{\ref{fig:stab}}).}
\label{fig:tieIL}
\end{figure}}
{\begin{figure}
\par\columnwidth20.5pc
\hsize\columnwidth\global\linewidth\columnwidth
\displaywidth\columnwidth \epsfxsize=3.5truein \centerline{\epsfbox[70
220 740 650]{figures/phase0sig.ps}}
\caption{
Composite phase diagrams for {\sf I-L} and {\sf I-N} coexistence at
common stress for $L=5.0$ and $\lambda=1.0$. We stress that this
represents \emph{two} overlayed phase diagrams, and not a single
phase diagram. For example, there is \emph{no} triple point implied
by the intersection of the \textsf{I-N} and \textsf{I-L} phase
diagrams.}
\label{fig:tieboth}
\end{figure}}
Upon applying shear, the system could then maintain coexistence and
move through the {\sf I-L} two-phase region. However, under
controlled strain rate conditions, the {\sf I} material could decay
into {\sf I-N} coexistence (see Fig.~\ref{fig:tieboth}). The resulting
{\sf I-N} coexistence occurs would quickly destabilize the entire {\sf
I-L} structure. Therefore the three-band structure {\sf N-I-L} will
not be present in this model, and it is probable that \textsf{I-L}
coexistence could only exist under flow as a metastable state. Similar
conclusions may be drawn by examining the phase diagrams in
field-variable space, $\mu-\sigma_{xy}$, as in Fig.~\ref{fig:mu}a. In
this case the chemical potential of the \textsf{I} phase, at
\textsf{I-L} coexistence, is within the \textsf{N} region of the phase
diagram for \textsf{I-N} coexistence, indicating a (possibly
metastable) instability with respect to \textsf{I-N} phase separation.
Moreover, the chemical potentials of the three phases are never the
same, except at rest where the \textsf{L} and \textsf{N} states are
identical apart from the rod orientations.
\section{Common Strain Rate Coexistence}
For coexistence at common strain rate the interface lies in the
velocity--velocity-gradient plane, and inhomogeneities are in the
${\rm\bf\hat{z}}$ direction (see Fig.~\ref{fig:couette}). The stress
balance condition at the interface is
$\boldsymbol{\sigma}\!\cdot\!{\rm\bf\hat{z}}$ uniform. As before,
$\sigma_{zz}$ is taken care of by the pressure while $\sigma_{yz}$ and
$\sigma_{xz}$ are zero by symmetry (and because there are no stable
$q_3$ components in the order parameter tensor). With bands in the
${\rm\bf\hat{z}}$ direction, the strain rate in each band is set by
the relative velocity of the two plates (or cylinders, in a Couette
device), and the shear stresses differ. The mean applied stress
$\bar{\sigma}_{xy}$ is the area average of the stress applied to each
band. The coexisting phases have shear stresses and compositions
partitioned according to
\begin{eqnarray}
\bar{\phi} &=& \zeta \phi_{\scriptscriptstyle
I} + (1-\zeta) \phi_{\scriptscriptstyle II} \label{eq:lev1s}\\
\bar{\sigma}_{xy} &=& \zeta \sigma_{xy}^{\scriptscriptstyle I} + (1-\zeta)
\sigma_{xy}^{\scriptscriptstyle II}, \label{eq:lev2s}
\end{eqnarray}
where $\bar{\sigma}_{xy}$ is the mean shear stress. The interfacial
equations to solve are Eqs.~(\ref{eq:Q}),(\ref{eq:sigma}), and
~(\ref{eq:mu}).
\noindent{\textbf{Phase Diagram---}}Common strain rate
\textsf{I-N} phase coexistence is shown in Fig.~\ref{fig:ties}. In
this case the tie lines are horizontal in the
$(\widehat{\dot{\gamma}}\!-\!u)$ plane. They have a negative slope in
the $(\hat{\sigma}_{xy}\!-\!u)$ plane because the paranematic {\sf I}
phase coexists with a denser and less viscous flow-aligning \textsf{N}
phase. As with phase separation at common stress, there is a (very
small) loop in the limits of stability in the control variable plane
($\dot{\gamma}\!-\!u$) within which there are no stable homogeneous
states. The careful reader will note that the limits of stability at a
given stress (Fig.~\ref{fig:tie}) are different from the limits of
stability at a given strain rate. This is physically correct, and will
be discussed below in Sec.~\ref{sec:spinodals}.
{\begin{figure}
\par\columnwidth20.5pc
\hsize\columnwidth\global\linewidth\columnwidth
\displaywidth\columnwidth
\epsfxsize=3.0truein
\centerline{\epsfbox[170 230 480 650]{figures/phasesIN.ps}}
\caption{Common strain rate phase diagram in the
$(\widehat{\dot{\gamma}}\!-\!u)$ (a) and $(\hat{\sigma}_{xy}\!-\!u)$
(b) planes, for $L=5.0$ and $\lambda=1.0$. Also shown are the limits
of stability of the \textsf{I} and \textsf{N} phases (calculated for
a given imposed strain rate, in contrast to Figures \ref{fig:stab},
\ref{fig:tie}, \ref{fig:tieIL}, and \ref{fig:tieboth}, in which the
stability was calculated for an imposed stress.)}
\label{fig:ties}
\end{figure}}
There is an interesting crossover visible in the
$(\hat{\sigma}_{xy}\!-\!u)$ plane. For higher mean compositions the
fluid has a higher stress in its high strain rate one-phase region
than in its low strain rate one-phase region; that is, respectively
above and below the biphasic region in the Fig.~\ref{fig:ties}a.
Conversely, for low enough compositions $u\alt 2.67$, the stress in
the high strain rate region immediately outside the biphasic regime is
actually {\sl less\/} than the stress just before the system enters
the biphasic region, as can be seen by the crossing of the solid and
dashed phase boundaries in Fig.~\ref{fig:ties}.
This crossover is straightforward to understand. Since phase
separation occurs at a given strain rate, and the stress of the {\sf
N} branch at a given composition and strain rate is always less than
that of the corresponding {\sf I} branch, we expect a decrease in the
stress upon leaving the biphasic regime in cases where the coupling to
composition is less important. We saw in the analysis at common stress
that composition effects are less important (for {\sf I-N}
coexistence) at lower compositions and high strain rates, where the
tie lines are more vertical. We expect this near the critical point
where the two phases become more and more similar. More generally, we
expect this behavior in situations where phase separation occurs at a
common strain rate into a shear-thinning state with only slight
changes in composition. In the more concentrated regime, the
coexistence plateau traverses a wider range of concentrations and
strain rates, and emerges into the pure \textsf{N} phase with a higher
stress (the width in strain rate of the phase coexistence regime is
enough to overcome the shear thinning effect of the nematic phase).
\noindent{\textbf{Mean Constitutive Relations---}}
Figs.~\ref{fig:stressstrainbars0}-\ref{fig:stressstrainbars} show the
mean stress--strain-rate relations. As with common stress phase
separation, the shape of the `plateau' as the strain rate is swept
through the two-phase region is not always flat, and depends on the
splay of the tie lines. At higher concentrations the plateau has a
positive slope while, in accord with the crossover in the
$(\hat{\sigma}_{xy}\!-\!u)$ phase diagram, for lower concentrations
the plateau crosses over to negative slope, which usually signifies a
bulk instability. A simple argument, analogous to that for the
stability of a bulk fluid, supports this. However, we note that a
composite negative slope curve was accessed, and apparently found
stable, by Hu \emph{et al.} \cite{HBP98} under controlled stress
conditions. The negative slope in Fig.~\ref{fig:stressstrainbars0}a
is likely to be inaccessible under controlled stress conditions, and
the instability argument may apply to controlled strain rate
conditions. The general relation for the slope in the composite
region is \cite{schmitt95}
\begin{equation}
\frac{\partial\bar{\sigma}}{\partial{\dot{\gamma}}} =
\eta_I\,\zeta + \eta_N\,(1-\zeta) -
m(\dot{\gamma})\left\{\frac{\eta_N(1-\zeta)}{\sigma'_N} +
\frac{\eta_I\,\zeta}{\sigma'_I}\right\},
\end{equation}
where $m(\dot{\gamma})$ is the slope of the tie line with strain rate
$\dot{\gamma}$ and $\sigma'_{k}=\partial\sigma_{k}/\partial\phi$. In
the limit of no concentration difference ($\delta\phi=0$ or
$m(\dot{\gamma})=\infty$), $\sigma(\dot{\gamma})$ is vertical through
the two phase region.
\end{multicols}\widetext
{\begin{figure}
\displaywidth\columnwidth
\epsfxsize=\displaywidth
\centerline{\epsfbox[80 210 1300 600]{figures/sep_s_all.ps}}
\caption{Mean stress--strain-rate curves for common strain
rate coexistence for $L=5.0$ and $\lambda=1.0$. The solid lines
denote the stable \textsf{I} and \textsf{N} branches; the dotted
line in each figure denotes the stable \textsf{N} branch with which
the \textsf{I} state coexists at the low strain rate boundary of the
coexistence region, at a strain rate marked by an open circle
$\boldsymbol{\circ}$. The solid circles $\bullet$ and thick solid
line denote the stress that would be measured in the banded regime.
The filled circles $\bullet$ and thick solid line denotes the stress
measured under banded conditions. }
\label{fig:stressstrainbars0}
\end{figure}}
\begin{multicols}{2}
\noindent{\textbf{Measurements at controlled stress or controlled
strain rate---}}For controlled strain rate measurements we expect
behavior similar to that for phase separation at common stress. For
start-up experiments with mean strain rates larger than the minimum
strain rate for coexistence at a given composition, we expect the
stress to follow the metastable branch until a nucleation event causes
the stress to decrease to the plateau stress. The exception is a
composition such as that in Fig.~\ref{fig:stressstrainbars0}a, for
which the composite flow curve for \textsf{I-N} coexistence may be
mechanically unstable. Similar results should apply upon decreasing
the strain rate from the shear-induced \textsf{N} phase to below the
\textsf{N} limit. As before, this expectation of a nucleation event is
based on a possibly misguided analogy with equilibrium systems which,
nonetheless, is encouraging given the experiments which see
``nucleation'' type behavior in micelles under flow
\cite{berret94b,grand97,Berr97}.
For controlled stress, the situation is slightly different. For
compositions with mean stress--strain-rate curves of the shape of
Fig.~\ref{fig:stressstrainbars0}c, we expect similar behavior to that
found for common stress phase separation. However, for compositions
that yield curves such as Fig.~\ref{fig:stressstrainbars0}a there is a
window of stresses for which there are \emph{three} possible states:
homogeneous low strain rate and high strain rate branches, and a
banded intermediate branch. We emphasize that we have not determined
the absolute stability of any of these branches. A possibility is
that the system has hysteretic behavior. For example, in start-up
experiments the system would remain on the \textsf{I} branch until a
certain stress, at which point it would nucleate after some time and
transform to either the high strain rate \textsf{N} branch or
coexistence. We cannot tell which state it might go to, from this
analysis, but it seems likely that it would jump straight to the
\textsf{N} branch.\footnote{If the system jumped from the \textsf{I}
branch to the coextence branch, increasing the stress further would
\emph{decrease} the strain rate and return the system to the
\textsf{I} branch. Since it originally nucleated \emph{from} the
\textsf{I} branch, it seems unlikely that the original jump could be
to the coexisting plateau.} The same behavior (in reverse) would be
expected upon reducing the stress from the high strain rate \textsf{N}
phase.
{\begin{figure
\par\columnwidth20.5pc
\hsize\columnwidth\global\linewidth\columnwidth
\displaywidth\columnwidth \epsfxsize=3.5truein \centerline{\epsfbox[60
270 520 570]{figures/sep_s_ovl.ps}}
\caption{$\hat{\sigma}_{xy}$ vs. $\widehat{{\dot{\gamma}}}$
for various compositions, for phase separation at common strain rate
and $L=5$. }
\label{fig:stressstrainbars}
\end{figure}}
Although there have been anecdotal reports of shear banding in the
common strain rate geometry, there have been very few such results
published. Bonn \emph{et al.} \cite{Bonn+98} have recently reported
results for sheared surfactant onion gels, along with visual
confirmation of bands in the common strain rate geometry. In
controlled strain rate experiments they found constitutive curves
analogous to Fig.~\ref{fig:stressstrainbars0}a
or~\ref{fig:stressstrainbars0}b. In controlled stress experiments they
found hysteretic behavior, with the system flipping between high and
low strain rate branches after some delay time, missing the
coexistence `plateau' regime. However, it is not clear that these were
true steady state results.
Stable `negative-slope' behavior was seen in a shear-thickening
systems which phase separates at common stress
\cite{boltenhagen97a,boltenhagen97b}, under controlled stress
conditions. In this case there was a single (mean) strain rate for a
given applied stress, and the measured constitutive relation had an
\textsf{S} shape rather than the sideways \textsf{S} shape of
Fig.~\ref{fig:stressstrainbars0}.
\section{Discussion}\label{sec:discussion}
\subsection{Dependence on gradient terms}
Gradient terms appear in all equations of motion for $K\neq0$ and for
any $g$, so to avoid unphysical equations without gradients (which
cannot resolve interfaces) we must have $\lambda\sim g/K<\infty$. In
the case of $K=0$ and finite $g$ the $\boldsymbol{Q}$ equation of
motion has no explicit gradient terms and hence can, in principle,
support discontinuous solutions. The $\phi$ equation has gradients in
this case, arising from the term $g\left(\nabla\phi\right)^2$ in the
free energy density, Eq.~(\ref{eq:free}), so the system will
eventually reach a state with smooth solutions in both $\phi$ and
$\boldsymbol{Q}$. Conversely, for $g=0$ there are gradient terms in
both the $\boldsymbol{Q}$ and $\phi$ dynamics, with the latter arising
from the term $\phi\left(\nabla\boldsymbol{Q}\right)^2$ in the free
energy density Eq.~(\ref{eq:free}).
Phase boundaries for $\hat{\sigma}_{xy}=0.01, 0.03$ are shown in
Fig.~\ref{fig:acc}. For $\lambda\in(0.0-30.0)$ the phase boundaries
are the same, within the precision of our numerical calculations,
while there is a distinct difference for $\lambda=\infty$. We have
discretized the system on a mesh of 125 points, and the range of
elastic constants is such that the width of the interface is at least
20 mesh points; large enough for smooth behavior and much smaller than
the system size.
{\begin{figure}
\par\columnwidth20.5pc
\hsize\columnwidth\global\linewidth\columnwidth
\displaywidth\columnwidth
\epsfxsize=3.5truein
\centerline{\epsfbox[50 200 750 650]{figures/acc.ps}}
\caption{\textsf{I-N} phase boundaries for common stress
phase separation as a function of $\lambda/L$, for $L=5.0$. The
diamonds $\blacklozenge$ are for $\lambda=\infty (K=0,g=1)$.}
\label{fig:acc}
\end{figure}}
We cannot rule out the possibility that changes in $\lambda$ shift the
phase boundaries by small amounts below our accuracy, which is of
order 0.1\% in $u$, but the apparent independence of the phase
boundaries on $\lambda$ is curious. One might be tempted to generalize
and suggest that, for finite $\lambda$, there exists a selection
criterion which involves only the homogeneous equations of motion,
rather than requiring the inhomogenous terms as in the interface
construction. An interface construction may also be used to determine
equilibrium phase boundaries, in which case a stationary interface is
equivalent to minimizing a free energy and the (relaxational)
dynamical equations derived from a variational principle
\cite{olmsted92}. In the case of a van der Waals fluid this reproduces
the Maxwell construction.
A steady state equation for a single variable $\psi$ with homogeneous
and inhomogeneous terms of the form
\begin{equation}
\label{eq:4}
\sigma_0 = f_{\mit hom}(\psi) + f_{\mit inh}(\partial\psi/\partial y)
\end{equation}
can be integrated to yield a solvability condition for $\sigma_0$,
which is equivalent to the stable interface method. In equilibrium
$f_{\mit inh}$ integrates exactly without an integrating factor, since
it typically arises from a variation of a free energy functional with
respect to $\psi$, and the result $\sigma_0$ (corresponding to the
pressure in the van der Waals fluid) depends only on $f_{\mit hom}$.
Out of equilibrium, integration is not so simple, and the solvability
condition depends, generally, on the form of the gradient terms
\cite{pomeau86,jsplanar}.
In the multivariable case considered here the steady state conditions
for the order parameter and composition are coupled differential
equations which are not integrable in shear flow. This is because of
the terms $\boldsymbol{\kappa}\!\cdot\!\boldsymbol{Q} +
\boldsymbol{Q}\!\cdot\!\boldsymbol{\kappa}^{\scriptstyle T}$ in
Eq.~(\ref{eq:convect}) and
$(\nabla^2\boldsymbol{Q})\!\cdot\!\boldsymbol{Q} -
\boldsymbol{Q}\!\cdot\!(\nabla^2\boldsymbol{Q})$ in
Eq.~(\ref{eq:sigma}). In extensional flow $\boldsymbol{\kappa}$ is
symmetric, so that $\boldsymbol{\kappa}\!\cdot\!\boldsymbol{Q} +
\boldsymbol{Q}\!\cdot\!\boldsymbol{\kappa}^{\scriptstyle T}$
integrates to ${\rm Tr}(Q^2\boldsymbol{\kappa})/2$, while in shear
flow this term can only be integrated by introducing an integral
representation \cite{zwillinger}. Hence a first integral of the steady
state equations cannot be found in shear flow, and it seems unlikely
that a general condition involving only the homogeneous portion of the
steady state equations can determine coexistence. While we appear to
find, for this set of gradient terms, solvability conditions that are
independent of $\lambda$ for $\lambda<\infty$, the relationship of
this to a variational principle remains unknown. We have not
exhausted the possible gradient terms. For example, higher order
gradients in the free energy ($(\nabla^2\boldsymbol{Q})^2$,
\emph{etc.}) would yield higher-order differential equations for the
interfacial profile, and other square gradient terms such as
$Q_{\alpha\beta}\nabla_{\alpha}\nabla_{\beta}\phi$ are possible
\cite{liu93}. Hence, we believe that, for finite $\lambda$,
\textit{the apparent independence of our results on gradient terms
only applies to the particular (simple) family of gradient terms we
have chosen}. The structure of the differential equations describing
the steady states may change abruptly for $\lambda=\infty$, for which
a term is lost in the differential equations, leading to a distinctly
different selection criteria and the shifted phase boundary in
Fig.~\ref{fig:acc}. Unfortunately, this particular set of equations is
too complex for this kind of analysis.
For example, in a study of a
simpler constitutive model, one can demonstrate that the selected
stress depends on the detail form of the gradient terms
\cite{jsplanar}.
Several workers have claimed to find an equal areas construction on
the stress--strain-rate constitutive curve \cite{greco97}. That is,
the ``plateau'' as the system traverses the two-phase region is said
to describe a path such that the areas above and below the plateau,
enclosed by the plateau line and the underlying constitutive curve,
are the same. This is not true here, as can be seen in
Fig.~\ref{fig:stressstrainbars0}.
\subsection{Which phase separation is preferred?}\label{which}
Having calculated \textit{both} common strain rate and common stress
phase separation for the same system, and noticing from
Figs.~\ref{fig:tieboth} and~\ref{fig:ties} that there are compositions
and shear conditions which lie inside the two-phase regions of all
three calculated phase separations, we must address the question of
which phase separation occurs. We have already argued that we expect
\textsf{I-L} phase-separation at common stress to be metastable with
respect to \textsf{I-N} phase separation at common stress. What about
the relative stability of \textsf{I-N} phase separation at either
common stress or common strain rate?
With limited one-dimensonal calculations for systems of different
symmetry (annular bands at common stress and stacked disk-like bands
for common strain rate) it is impossible to calculate the stability of
one interface profile with respect to another. Renardy calculated the
stability of common stress coexistence to capillary fluctuations
\cite{renardy}, which is a start; and such a stability analysis has
been performed, in part, on the layer orientation of smectic systems
in flow \cite{goulian95}. However, some insight can be obtained by
examining the ``phase diagrams'' in the chemical potential--field
variable (either stress or strain rate) planes. The solid lines in
Fig.~\ref{fig:mu} are analogous to lines of phase coexistence in, for
example, the pressure--temperature plane in a simple fluid.
Consider Fig.~\ref{fig:mu}a. Here, $\widehat{\sigma}_{xy}$ and $\mu$
are the proper field variables for phase separation at a common
stress, and the solid lines denote the \textsf{I-L} and \textsf{I-N}
phase boundaries. The dashed line denotes the range of stresses at
coexistence for common strain rate phase separation
(Fig.~\ref{fig:mu}a), for which stress is a generalized density
variable and strain rate the field variable.
Fig.~\ref{fig:mu}a indicates that, for a system undergoing
\textsf{I-N} at a common strain rate, the chemical potential and
stress for the \textsf{I} phase falls within the single phase
\textsf{N} region of the common \emph{stress} phase diagram. Hence, we
expect this \textit{I} phase to be unstable (or metastable) with
respect to phase separation at common stress. Similarly, the control
parameters (chemical potential and stress) for the \textsf{N} phase
coexisting at a common strain rate lie within the single phase
\textsf{I} region for common stress phase separation, which we also
expect to be unstable (or metastable). Conversely, for a system
coexisting at a common stress the \textsf{I} phase lies within the
single phase \textsf{I} region of the common strain rate phase diagram
(Fig.~\ref{fig:mu}b), and similarly for the \textsf{N} phase. This
suggests that phase separation at a common strain rate is unstable (or
metastable) with respect to phase separation at common stress, while
phase separation at a common stress is stable.
{\begin{figure}
\par\columnwidth20.5pc
\hsize\columnwidth\global\linewidth\columnwidth
\displaywidth\columnwidth
\epsfxsize=3.5truein
\centerline{\epsfbox[130 230 500 680]{figures/mu_sig2.ps}}
\caption{Phase diagrams in the chemical potential $\mu$ vs. stress plane
(a) and the $\mu$-strain rate plane (b). The solid lines denote the
phase boundaries for common stress phase coexistence in the
$\mu\!-\!\sigma$ plane (a) and for common strain rate coexistence in
the $\mu\!-\!\dot{\gamma}$ plane (b). The dashed lines denote the
coexisting stresses for common strain rate phase separation within
the common stress phase diagram (a); and vice versa in (b).}
\label{fig:mu}
\end{figure}}
Note that, ultimately, this selection of phase coexistence geometries
follows from the transition being a shear-thinning transition; for a
shear thickening transition the situation could be reversed. In this
case phase coexistence at a common strain rate and a given $\mu$ would
imply a shear-induced phase (analogous to the \textsf{N} phase) with a
higher stress than the \textsf{I} phase. If the phase coexistence
line for common stress (strain rate) lay within a loop corresponding
to the stresses (strain rates) for common strain rate (stress)
coexistence, then common strain rate coexistence would be expected to
be stable, by analogy with the isotropic-nematic shear thinning model.
Obviously this argument is delicate. In a fluid where only one phase
coexistence (either common stress or common strain rate) is supported
by the dynamical equations this argument is moot.
\end{multicols}\widetext
{\begin{figure}
\displaywidth\columnwidth
\epsfxsize=\displaywidth
\centerline{\epsfbox[26 102 571 410]{figures/mu_switch.eps}}
\caption{Phase diagrams in the (a) $\mu\!-\!\dot{\gamma}$ and (b)
$\mu\!-\!\sigma_{xy}$ planes for \textsl{I-N} coexistence (the
\textsl{N} state is stable for higher strain rate or stress,
respectively). The thin vertical solid lines passing through {\tt
B'-C} and {\tt A-B-D} denote phase coexistence at common strain
rate and stress in (a) and (b), respectively. The broken lines
marked $I_\gamma$ and $N_\gamma$ denote the coexisting states at
common strain rate in the $\mu\!-\!\sigma_{xy}$ plane (b); while the
broken lines $I_\sigma$ and $N_\sigma$ denote the coexisting states
at common stress, in the $\mu\!-\!\dot{\gamma}$ plane. (c) is the
mean stress vs. strain rate curve. Shown is a path {\tt A-B-C-D}
taken under the proposition that the system maintains a global
minimum in chemical potential. Point {\tt A} is at coexistence in
the $\mu\!-\!\sigma_{xy}$ plane (b), and hence corresponds to two
points, on lines $I_{\sigma}$ and $N_{\sigma}$, in the
$\mu\!-\!\dot{\gamma}$ plane (a) for the two different strain rates
of the coexisting phases. Similarly, point {\tt C} corresponds to
coexistence at common strain rate in (a), with the coexisting phases
at different stresses lying on lines $I_{\gamma}$ and $N_{\gamma}$
in (b), at the two points {\tt C}. Points {\tt B} and {\tt B'} are
coincident in (c), and correspond to a point switching from phase
separation at common stress ({\tt B}) to phase separation at common
strain rate ({\tt B'}). The path {\tt A-B-C} in (c) may be traced
in (b) by following the upper horizontal arrow until phase
separation at common stress occurs at {\tt A}, then along the
segments {\tt A-B} in (b) or {\tt A-B'} in (a) until phase
separation at common strain rate occurs at {\tt B'}. From this
point until {\tt C} the system phase separates along $I_{\gamma}$
and $N_{\gamma}$, with a mean stress given by the thick diagonal
solid arrow {\tt B-C} in (b) and the thick segment {\tt B'-C} in
(a). The system emerges from the two-phase region at {\tt C} on
$N_{\gamma}$, and continues through {\tt D} on the high strain rate
branch. }
\label{fig:resolve}
\end{figure}}
\begin{multicols}{2}
An alternative possibility is presented in Fig.~\ref{fig:resolve} if
one argues that in steady state, among the possible phases which are
compatible with the interface solvability condition, the chemical
potential reaches its absolute minimum. Consider increasing the
strain rate for a given mean concentration. The thick horizontal
arrows in Fig.~\ref{fig:resolve}a-c denote the $\mu(\sigma_{xy})$ and
$\mu(\dot{\gamma})$ paths for the homogeneous high and low shear rate
states, in the two phase diagrams. In Fig.~\ref{fig:resolve}a the
path is {\tt A-B'-C-D}, in Fig.~\ref{fig:resolve}b the path is {\tt
A-B-C-D}, and in Fig.~\ref{fig:resolve}c the path is {\tt
A-B-C-D}.
This path ensures that the system maintains the minimum chemical
potential for an imposed strain rate. Upon increasing the strain rate
from zero the system remains in the one phase region until {\tt A} is
reached, at which point phase separation at common stress occurs. Note
that {\tt A} spans two points of coexistence in the
$\mu\!-\!\dot{\gamma}$ plane (Fig.~\ref{fig:resolve}a) on the lines
$I_{\sigma}$ and $N_{\sigma}$. Upon further increasing the strain rate
the system continues to phase separate at common stress, following the
segment {\tt A-B} in Fig.~\ref{fig:resolve}b and the two (coexisting)
segments {\tt A-B} in Fig.~\ref{fig:resolve}a. The mean chemical
potential and strain rate follow the diagonal segment {\tt A-B'} in
Fig.~\ref{fig:resolve}a. Upon increasing the \emph{strain rate} above
{\tt B}, the system can continue to maintain its lowest chemical
potential by phase separating with a common strain rate in the two
phases, \emph{i.e.} with shear bands in the vorticity direction.
Hence, the system next follows the path {\tt B'-C} in
Fig.~\ref{fig:resolve}a ($\mu\!-\!\dot{\gamma}$ plane) and the two
coexisting paths {\tt B'-C} in Fig.~\ref{fig:resolve}b
($\mu\!-\!\sigma_{xy}$ plane), with the mean chemical potential and stress
following the diagonal segment {\tt B-C} in Fig.~\ref{fig:resolve}b.
Finally, upon increasing the strain rate above {\tt C} the system
continues along the high strain rate branch. The thin diagonal lines
with arrows, {\tt B'-D} in Fig.~\ref{fig:resolve}a and {\tt B-D} in
Fig.~\ref{fig:resolve}b show the path that would be taken if the
system passed through the two-phase region entirely with a common
stress in the two phases. This scenario follows from minimizing the
chemical potential, and its correctness, of course, should be further
examined by the full time evolution of the original dynamic equations.
It is probable that boundary conditions also play a role. Consider a
Couette device. Typically the walls provide uniform boundary
conditions in the azimuthal direction, while the slight inhomogeneity
of Couette flow induces an asymmetry between the inner and outer
cylinders. The slightly higher stress near the inner wall provides a
preference for the high strain rate nematic phase, and hence might
enhance the stability of common stress phase separation. Similarly,
the intrinsic inhomogeneity (although weaker) in cone-and-plate
rheometry induces a preference for the common stress interfacial
configuration \cite{BritCall97}.
We are also unable to say anything about the number or spacing of
bands. Analogies with equilibrium systems suggest that phase
separation would coarsen until the system formed two bands at
different strain rates (for phase separation at a common stress).
This is the behavior seen in visualizations of flow in Couette,
cone-and-plate, and pipe geometries, where the intrinsic inhomogeneity
provides a ``seed'' for macroscopic phase separation
\cite{berret94b,Call+96,MairCall96,BritCall97,MairCall97,boltenhagen97a}.
Recent visualization of banding in lamellar surfactant systems
\cite{Bonn+98} indicate that phase separation at a common strain rate
can exhibit bands (disc-like bands in Couette flow) whose initial
spacing depends on the applied strain rate and coarsen in time. Unlike
common stress bands, which are expected to (and do) form macroscopic
bands in Couette flow, there is no boundary effect in Couette (aside
from perhaps sedimentation) which would encourage common strain rate
bands to coalesce readily. Normal stresses may play a role in this
process.
\subsection{Stability at prescribed stress or prescribed strain
rate} \label{sec:spinodals}
In calculating the phase diagrams we have calculated the stability of
the fluid under conditions of either fixed strain rate or fixed
stress. These limits of stability, analogous to spinodals in
equilibrium systems, are displayed in Figures~\ref{fig:tie}
and~\ref{fig:ties}. Note that the stability limits and critical points
differ, depending on the control variable (stress $\sigma$ or strain
rate $\dot{\gamma}$). To see why this is, note that schematically the
dynamical equations of motion have the form
\begin{align}
\partial_t \textbf{x} &= \textbf{f}(\textbf{x},\dot{\gamma}) \label{eq:dtx}\\
\sigma &= g(\textbf{x},\dot{\gamma})\label{eq:dtsigma}
\end{align}
where $\textbf{x}$ comprises the dynamical variables (order parameter
$\boldsymbol{Q}$ and composition $\phi$). The second equation relates
the stress to strain rate and dynamical variables at steady state (or
in the zero Reynolds number limit), which implies that the strain rate
$\dot{\gamma}$, for a given stress, is a function $\dot{\gamma} =
\dot{\gamma}(\sigma,\textbf{x})$. Consider fluctuations about a
steady state $\textbf{x}_0$: $\textbf{x} = \delta\textbf{x} +
\textbf{x}_0$. The dynamics for the fluctuation obeys
\begin{align}
\partial_t\,\delta\textbf{x} &=
\left\{\frac{\partial\textbf{f}}{\partial\textbf{x}} +
\frac{\partial\textbf{f}}{\partial\dot{\gamma}}
\frac{\partial{\dot{\gamma}}} {\partial\textbf{x}}
\right\}\cdot\delta\textbf{x}\\
&\equiv \left\{\textsf{M}_{\gamma} +
\delta\textsf{M}_{\sigma}\right\}\cdot\delta\textbf{x}
\end{align}
The limit of stability for common strain rate is calculated using the
fluctuation matrix $\textsf{M}_{\gamma}$, while the limit of stability
for common stress was calculated using $\textsf{M}_{\gamma} +
\delta\textsf{M}_{\sigma}$. These correspond to different stability
criteria.
The question of which spinodal could be observed in an experiment
relies on the accuracy of prescribed stress and prescribed strain rate
rheometers. For a rheometer operating at a prescribed strain rate,
then if $\delta\textbf{x}$ goes unstable through $\textsf{M}_{\gamma}$
(in Eq.~\ref{eq:dtx}), the stress increases due to Eq.~(\ref{eq:dtsigma})
and no attempt is made to control it, leading to instability.
However, consider a rheometer which maintains a prescribed stress. If
the system goes unstable in Eq.~(\ref{eq:dtx}), the bulk stress will
change due to Eq.~(\ref{eq:dtsigma}). A sensitive and fast enough
rheometer will respond by adjusting the strain rate accordingly, to
maintain the imposed stress. Hence, instability would be determined by
the sum $\textsf{M}_{\gamma} + \delta\textsf{M}_{\sigma}$.
Similarly, in equilibrium systems a locus of stability may be defined
by, for example, the diverging of isothermal or adiabatic, or isobaric
or isochoric, response functions (or the vanishing of the appropriate
modulus). For example, the isothermal and adiabatic compressibilities
$K_T$ and $K_S$ differ by a term proportional to the quotient of the
square of the thermal expansion coefficient $\alpha_p$ and the
isobaric heat capacity $c_p$:
\begin{equation}
\label{eq:7}
K_T - K_S = \frac{vT\alpha_p^2}{c_p},
\end{equation}
where $v$ is the specific volume. $K_T^{-1}$ vanishes along the
spinodal line $v_s(T)$, while it is evident that $K_S^{-1}$
(proportional to the sound speed) does not. However, in equilibrium,
the critical point is uniquely defined in phase space, which is
related to the fact that, for example, pressure is a unique
\emph{function} of the volume, and is in fact a state variable.
Conversely, we can see from the shape of the stress--strain rate
curves for the Doi model (\emph{e.g.} Fig.~\ref{fig:constit}c-e),
that the stress can be a multivalued function of strain rate;
\emph{i.e.} it is not a state function. Hence there is no compelling
reason to expect critical points at imposed strain rate to be the same
as critical points at imposed stress. Similarly, the {\em true\/}
spinodal, or locus of stability, is uniquely defined in an equilibrium
system because of the convexity requirement on the entropy
\cite{debenedetti}, and there is no such universal convexity
requirement (barring entropy production, which is minimized only under
restricted conditions, and only locally rather than globally) for
non-equilibrium systems.
\subsection{An analogy with equilibrium systems?}
The liquid crystalline suspension under flow, indeed any system which
undergoes a macroscopic bulk flow-induced phase transition, is
analogous to an equilibrium ternary system comprising species $A$,
$B$, and solvent. In our case, the roles of $A$ and $B$ are played by
the rigid rod composition $\phi$ and either the stress $\sigma$ or
strain rate $\dot{\gamma}$, depending on the nature of the phase
separation. For phase separation at common stress, the phase diagram
in the stress-composition plane $\sigma-\phi$ is analogous to the
$\mu_A-\phi_B$ plane for the equilibrium system, while the
$\dot{\gamma}\!-\!\phi$ plane is analogous to the $\phi_A\!-\phi_B$
plane. In either case, the density variables,
$\left\{\dot{\gamma},\phi\right\}$ in flow and
$\left\{\phi_A,\phi_B\right\}$ in the analogous equilibrium system,
are different in the two coexisting phases. The slope of the
``plateau'' in the $\sigma\!-\dot{\gamma}$ plane, as the system
traverses the two-phase region of the phase diagram, is analogous to a
slope in the $\mu_A\!-\!\phi_B$ plane, the latter indicating that the
chemical potential (or osmotic pressure) of the two phases varies
across the coexistence region.
Can this analogy be extended to the possibility of phase separation at
common stress \emph{or} common strain rate? Certainly, one can
consider a ternary system under conditions of either imposed $\phi_A$
or imposed $\mu_A$, for which one generally expects difference
spinodal lines. That is, the spinodal is determined by the instability
of a matrix in the two-dimensional space spanned by $\phi_A$ and
$\phi_B$, and fixing $\phi_A$ or $\mu_A$ projects this instability onto
different subspaces. Experimental conditions may dictate that the
spinodal line under fixed $\phi_A$ is more likely to be seen by, since
$\phi_A$ is conserved and cannot equilibrate quickly to satisfy an
imposed $\mu_A$. However, we are not aware of any ternary equilibrium
system for which the equilibrium coexistence conditions can differ;
that is, equilibrium is \emph{always} specified by equality of $\mu_A$
and $\mu_B$ in the two phases, and never by equality of $\phi_A$.
\section{Summary}
In this work we have proposed a straightforward phenomenological
extension to the Doi model for a solution of rigid rod particles. We
have added entropic terms, and included inhomogeneous terms in order
to calculate, for the first time, phase separation in shear flow. The
main results of this study are as follows:
\begin{enumerate}
\item Phase separation may occur under conditions of common stress
\emph{or} common strain rate, with different interface orientations
with respect to flow geometry for the two cases.
\item Although both phase separations are possible, the phase diagrams
in the $\mu\!-\!\sigma_{xy}$ and $\mu\!-\!\dot{\gamma}$ planes
(Fig.~\ref{fig:mu}) suggest that phase separation at a common strain
rate is metastable. This can be traced, for this model, to the
shear-thinning character of the transition. For a shear thickening
transition an equivalent argument suggests that (if both are
kinematic possibilities) common stress phase separation is
metastable with respect to strain rate phase separation.
\item The limits of stability (``spinodals'') and critical points for
systems at prescribed stress and prescribed strain rate differ; the
difference of spinodals is similar to equilibrium behavior, while
the difference of critical points is related to the fact that
neither stress or strain rate are always unique state functions.
\item An argument based on minimizing the chemical potential predicts
a complex crossover from common stress to common strain rate phase
separation for controlled strain rate experiments. The veracity of
this assumption is unknown.
\item We have calculated phase coexistence among three phases
(paranematic \textsf{I}, flow-aligning nematic \textsf{N}, and
log-rolling nematic \textsf{L}), where only two phases existed in
equilibrium. We expect \textsf{I-N} phase coexistence to be the
stable configuration (Fig.~\ref{fig:tieboth}), although \textsf{I-L}
phase coexistence could exist as a metastable state with approprate
preparation. We do not expect three phase coexistence for this
model.
\item We have demonstrated how to calculate the mean
stress--strain-rate relationship $\bar{\sigma}(\bar{\dot{\gamma}})$
in the coexistence region. The shape of
$\bar{\sigma}(\bar{\dot{\gamma}})$ is determined by the composition
and strain rates of the coexisting phases \cite{schmitt95}.
\item A phase-separated system can exhibit an apparently unstable
constitutive relation, with negative slope
$\partial\sigma_{xy}/\partial\dot{\gamma}$. Experiments have
accessed such negative slope composite curves under controlled
stress (rather than controlled strain rate) conditions \cite{HBP98}.
\item Our method of solution is general and relies on the existence of
a set of dynamical equations of motion for the structural order
parameter of the particular transition, including the dynamic
response to inhomogeneities.
\item For $\lambda=g/K$ finite the phase boundaries we have found are,
within our accuracy, independent of the relative magnitude of the
gradient terms in our free energy. Although this suggests that, for
the restricted set of inhomogeneities we have incorporated, a
selection criterion exists involving only the homogeneous equations
of motion, this is not true in general for complex fluids in flow
\cite{jsplanar}. For $\lambda=\infty$ the phase boundaries are
slightly shifted.
\item Studies at different aspect ratios suggest that shear flow
enhances polydispersity effects relative to their effect on
equilibrium phase boundaries.
\end{enumerate}
We close by enumerating several open questions. First, systems such as
wormlike micelles probably possess some combination of a perturbed
isotropic-nematic transition and a dynamic instability of the
molecular constitutive relation. It is conceivable that suitable
compositions of these systems could yield a
stress--strain-rate--composition surface (Fig.~\ref{fig:3D}) with
multiple folds. Second, it would be desirable to have a model
shear-thickening system in which to calculate properties of banded
flows, to compare and contrast with the shear-thinning system studied
here and to understand experiments on a wide range of systems,
including clays and surfactant systems. Third, we have not addressed
the number and possible coarsening of bands and band configurations;
and the kinetics of phase separation has hardly been treated
theoretically \cite{berret94b}, with experiments also at an early
stage \cite{berret94b,grand97,Berr97}. Finally, we do not yet know the
conditions which may, if at all, distinguish between common stress or
common strain rate phase coexistence.
\acknowledgements
It is a pleasure to acknowledge helpful conversations and
correspondence with R. Ball, J.-F. Berret, G. Bishko, D. Bonn, M.
Cates, F. Greco, J. Harden, S. Keller, G. Leal, T. McLeish, D. Pine,
G. Porte, O. Radulescu, N. Spenley, L. Walker, and X.-F. Yuan. CYDL
acknowledges funding from St. Catharine's College, Cambridge and the
(Taiwan) National Science Council (NSC 88-2112-M-008-005).
|
\section{Introduction}
The stripe order formed by linearly segregated holes in the oxygen-doped and Sr-doped La$_{2}$NiO$_{4}$ system is studied in detail by a series of works by Tranquada and coworkers \cite{TraO105,TraO125,TraO2_15,Tra97,Tra135,Tra225}. When the spins in this compound form the ordering at low temperatures, a hole stripe separates the antiferromagnetic Ni spin order as an antiphase domain boundary. It was suggested that such characteristic stripe order may persist for a larger hole concentration $n_{h}$ up to $n_{h} \approx \frac{1}{2}$ with keeping the linear relation between the hole concentration $n_{h}$ and the incomensurability $\epsilon$ of the stripe order, {\it i.e.} $\epsilon \sim n_h$\cite{Tra135,Tra225}. According to the resistivity and electron diffraction studies \cite{Che93}, on the other hand, commensurate charge order is speculated for two Sr concentrations $n_h = \frac{1}{3}$ and $\frac{1}{2}$. In fact, very recent experimental studies have established that the $n_h =\frac{1}{3}$ sample exhibits the stripe-type charge order below $T \sim 240$K, and it accompanies with anomalies in optical conductivity and Raman spectra\cite{Kat96,Lee97,raman}.
So far, the hole stripe order in the nickelate is confirmed in the O-doped La$_{2}$NiO$_{4+\delta}$ samples with $\delta = 0.105, 0.125, \frac{2}{15}$ (Refs. \onlinecite{TraO105,TraO125,TraO2_15,Tra97}) and the Sr-doped La$_{2-x}$Sr$_{x}$NiO$_{4}$ samples with $ x =0.135,0.20,0.225$ and $x=\frac{1}{3}$ (Refs. \onlinecite{Tra135,Tra225,Lee97}) by neutron diffraction. For a small hole concentration $n_{h} = x + 2\delta$, the distance between hole stripes is wide enough to accomodate three or more Ni chains in between, and this situation allows for Ni chains to form antiphase antiferromagnetic spin ordering across hole stripes. In contrast, for the samples with larger $n_{h}$ of $\frac{1}{3} \leq n_{h} \leq \frac{1}{2}$, the distance of hole stripes is small, and only one or two Ni chains are accomodated between the hole stripes provided that the hole stripes reside on the Ni sites. Hence it would be interesting to study whether samples with $\frac{1}{3} \leq n_{h} \leq \frac{1}{2}$ can form essentially the same stripe ordering, or they exhibit qualitatively different charge ordering. The information on spin/charge ordering for $\frac{1}{3} \leq n_{h} \leq \frac{1}{2}$ would be also very useful to understand the behavior of the resistivity. In order to elucidate the effects of the higher hole doping to the hole stripe order, we have carried out a neutron diffraction study on La$_{2-x}$Sr$_{x}$NiO$_{4}$ samples with the Sr concentration $x$ for $0.289 \lesssim x \lesssim 0.5$. The preliminary results have been reported elsewhere \cite{yoshi}.
Some of the important findings in the present study are that the incommensurability $\epsilon$ in the Sr-doped nickelate is approximately linear in $n_{h}$ up to $n_{h} \approx \frac{1}{2}$, in sharp contrast with the La$_{2-x}$Sr$_{x}$CuO$_{4}$ system \cite{yam97,tra97b}. A careful examination of the $n_{h}$ dependence of $\epsilon$ further revealed that there is a systematic deviation from an $\epsilon \sim n_{h}$ law around $n_{h}=\frac{1}{3}$, and that such deviation strongly influences transport properties\cite{Kat99}. We also observed that the charge ordering temperature $T_{\rm CO}$ and the spin ordering temperature $T_{\rm N}$ exhibits maxima at $n_{h}=\frac{1}{3}$, and they decrease beyond $n_{h}=\frac{1}{3}$. In addition, the stripe order at low temperatures is of two-dimensional (2D) character, and it consists of a mixture of the $n_{h}=\frac{1}{3}$--type stripe order and the $n_{h}=\frac{1}{2}$--type charge order within the 2D NiO$_2$ planes for $\frac{1}{3} \leq n_{h} \leq \frac{1}{2}$.
Single crystal samples studied in the present study were cut from the same crystals used in the previous measurements of optical and Raman spectra as well as transport properties. They were grown by the floating zone method, and the oxygen off-stoichiometry as well as the hole concetration $n_{h} =x + 2\delta$ were characterized in detail as previously reported \cite{Kat96,raman,Kat99}. All the samples are denoted by the calibrated hole concentration $n_h$ throughout this report.
The neutron scattering experiments were performed using triple axis spectrometers HQR and GPTAS installed at the JRR-3M reactor in JAERI, Tokai, Japan. To optimize the visibility of weak signal from the charge ordering, we chose a combination of horizontal collimators of open-Sample-40$^{\prime}$-Analyzer-open (from monochromator to detector) for the HQR sperctrometer which is installed at the thermal guide tube with a fixed incident neutron momentum of 2.57{\AA}$^{-1}$. The crystals were mounted in an Al can filled with He gas. Following the preceding works, we denote the reciprocal space by the orthorhombic notation, and all the measurements were performed on the $ (h0l) $ scattering plane.
\begin{figure}
\centering \leavevmode
\psfig{file=lsno_prof_diagram.eps,width=\hsize}
\caption{In-plane scan profiles for the spin and charge superlattice peaks observed along $(h01)$ for (a) $n_h = 0.289$, (b) $n_h = 0.332$, and (c) $n_h = 0.39$ samples. }
\label{h-profile}
\end{figure}
In order to characterize the charge and spin ordering in the highly Sr-doped nickelate samples, we first studied the ordering in the NiO$_2$ planes. We found that the samples with $0.289 \lesssim n_h \lesssim 0.5$ show very similar superlattice reflections of the stripe order with those observed in the Sr-doped and O-doped samples with smaller $n_{h}$. Figure \ref{h-profile} shows the profiles of the charge and spin superlattice peaks observed along the $(h01)$ line, on which the superlattice reflections of the spin order were observed at $(2n+1\pm \epsilon, 0, 1)$, while those of the charge order at $(2n\pm 2\epsilon, 0, 1)$ with $n$ integer, respectively. Note that, due to the $\epsilon \sim n_h$ law, an increase of $n_h$ switches the relative positions of the spin and charge superlattice peaks as schematically shown in Fig. \ref{h-profile}(d). At $n_{h}=\frac{1}{3}$, the superlattice peak of the spin order exactly coincides with that of the charge order as seen in Fig. \ref{h-profile}(b), which strongly enhances the stability of the $n_{h}=\frac{1}{3}$ stripe order and gives rise to a distinct anomaly in the resistivity\cite{Che93}.
Reasonably sharp peaks observed in the present samples indicate that the well-developed stripe order is established within the NiO$_2$ planes up to $n_h \lesssim 0.5$. We subsequently examined the stacking of the stripe order perpendicular to the NiO$_2$ planes, by observing the profiles along the $l$ direction (not shown). We found that the scattering profiles are centered at $l=$ interger with stronger intensity at $l=$ odd, being similar to the results observed in the less Sr-doped samples\cite{Tra135,Tra225}. Consequently, the inverse correlation length $\kappa$ of the stripe order along the stacking direction was evaluated by fitting to the formula suggessted by Tranquada {\it et al.} \cite{Tra225},
\begin{equation}
I(l) \sim \frac{1-p^2}{1+p^2-2 p \cos\pi l}
\end{equation}
where $p = e^{-c/2\xi_{l}} = e^{-\frac{c}{2}\kappa_{l}}$. For $\xi_{l}/c \gg 1$, it converges to a conventional Lorentzian form. The fact that the $l$ dependence is well described by Eq. (1) means that the correlation of the stacking of the hole stripes decays exponentially but they are resistered on the lattice at low temperatures even for $\frac{1}{3} \leq n_h \leq \frac{1}{2}$. The profiles observed by $h$ scans are also fitted to Lorentzian, and all the results are summarized in Fig. \ref{x-dep}. The circle symbols denote the in-plane ${\kappa}$, while squares denote ${\kappa}$ perpendicular to the NiO$_2$ planes which is evaluated from $p$.
\begin{figure}
\centering \leavevmode
\psfig{file=lsno_kappa_p.eps,width=0.8\hsize}
\caption{Inverse correlation length of the spin order ${\kappa}_{s}$ (upper panel) and of the charge order ${\kappa}_{c}$ (lower panel) at low temperatures. Triangles denote $p$ defined in Eq. (1). Filled symbols are by the present work, while open symbols from Refs. \protect\onlinecite{Tra135,Tra225}. $\kappa_c$ of $n_h=\frac{1}{3}$ is estimated from the data at 170K.}
\label{x-dep}
\end{figure}
As a function of $n_h$, we can identify three regions for the stripe order. For $n_h \lesssim \frac{1}{4}$, the correlation length is short for both within and perpendicular to the NiO$_2$ planes, and the stripe order is essentially 3 dimensional (3D) short range order (SRO). For $n_h \gtrsim 0.4$, on the other hand, the stripe order is well-developed within the NiO$_2$ planes, but is less correlated between the NiO$_2$ planes, being quasi-2D long range order (LRO). Near $n_h \sim \frac{1}{3}$, $\kappa$ shows a minimum, and the stripe order is quasi-3D LRO, demonstrating the stability of the stripe order at $n_h \sim 1/3$.
The stability of the $\epsilon=\frac{1}{3}$ stripe order is also evident in the $n_h$ dependence of the charge and spin ordering temperatures $T_{\rm CO}$ and $T_{\rm N}$, and they are summarized in the upper panel of Fig. \ref{n_vs_Tc}. In earlier works\cite{Tra135,Tra225,Che93}, $T_{\rm CO}$ and $T_{\rm N}$ were speculated to increase linearly in $n_h$ as indicated by a dashed line for $T_{\rm CO}$. We found, however, that they peak at $n_h =\frac{1}{3}$. We confirmed that the transition temperatures determined in the present work are in excellent accord with the anomalies of the temperature dependence in the resistivity along the $c$ axis ({\bf E} $\parallel c$) \cite{Kat96}. As pointed out earlier \cite{TraO125,Tra135,Tra225}, $T_{\rm CO}$ and $T_{\rm N}$ are different for all the samples we studied with $0.289 \lesssim n_h \lesssim \frac{1}{2}$, indicating that the hole stripe order is established first at $T_{\rm CO}$, and then the antiphase spin order is formed at the lower temperature $T_{\rm N}$ for $n_h \lesssim \frac{1}{2}$.
\begin{figure}
\centering \leavevmode
\psfig{file=lsno_n_vs_Tc_e.eps,width=0.8\hsize}
\caption{Upper panel: $T_{\rm CO}$ (square and diamond symbols) and $T_{\rm N}$ (circles). Lower panel: Hole concentration $n_h$ dependence of the incommensurability $\epsilon$ for $n_h \protect\lesssim \frac{1}{2}$ determined at the lowest temperature studied. Cross symbols indicate the high temperature initial values of $\epsilon$. ($\Diamond$ is taken from Ref. [7], and $\bigcirc, \Box$ from Refs. [1-6].)}
\label{n_vs_Tc}
\end{figure}
Now, we examine the incommensurability $\epsilon$ of the stripe order for $0.289 \lesssim n_h \lesssim \frac{1}{2}$ in detail. Since $\epsilon$ is weakly temperature-dependent, $\epsilon$ of the low temperature limit is plotted at the bottom panel of Fig. \ref{n_vs_Tc} \cite{shift}. We confirmed that $\epsilon$ is approximately linear in $n_h$ up to the limit of the stripe order, $n_h \approx \frac{1}{2}$. This is in strong contrast to the doped cuprate superconductor La$_{2-x}$Sr$_{x}$CuO$_{4}$ and to the O-doped nickelate. For instance, $\epsilon$ saturates at $\epsilon \sim \frac{1}{8}$ beyond the optimum doping in the cuprate\cite{yam97,tra97b}. In the Sr-doped nickelate, on the other hand, the formation of the $\frac{1}{3}$ stripe order is clearly stable, and the region of the stripe order extends to the higher hole concentration.
The analysis of $\epsilon$ further provides important information on the structure of the stripe order. For the O-doped La$_2$NiO$_{4+{\delta}}$ samples, $\epsilon$ is often locked at a rational value given by $\epsilon = (m+n)/(4m+3n)$ \cite{TraO105,TraO125,TraO2_15} because the interstitial oxygen ordering stabilizes the commensurate $\epsilon=\frac{1}{4}$ and $\frac{1}{3}$ stripe orders and introduces the competition between them. In the present study, we found that the low temperature limit of $\epsilon$ can be expressed by the same relation for $n_h <\frac{1}{3}$, but it changes to
\begin{equation}
\epsilon = (n+m^{\prime})/(3n+2m^{\prime}) \ {\rm for}\ n_h \ge\frac{1}{3},
\end{equation}
as tabulated in Table \ref{epsilontable}. To explain the meaning of this formula, the model of the $\epsilon=\frac{1}{3}$ stripe order and the $n_{h}=\frac{1}{2}$ charge order are depicted in Fig. \ref{CO_model}(a) and (b)\cite{comTdep}. The fact that $\epsilon$ is given by Eq. (2) for $\frac{1}{3} \leq n_h \leq \frac{1}{2}$ can be interpreted that the stripe order in this range of hole concentration consists of a combination of the $\frac{3}{2}a$-width unit of the $\epsilon=\frac{1}{3}$ stripe order and the $a$-width unit of the $n_{h}=\frac{1}{2}$ charge order, separated by discommensuration. An example observed in the $n_h=0.425$ sample is depicted in Fig. \ref{CO_model}(c), in which shaded regions correspond to the $\frac{3}{2}a$-width unit. $m$, $n$, and $m^{\prime}$ in Eq. (2) and Table \ref{epsilontable} give the numbers of the $2a$-, $\frac{3}{2}a$-, and $a$-width units in the long-period commensurate unit cell. For $\frac{1}{3} \leq n_h \leq \frac{1}{2}$, an increase of $m^{\prime}$ relative to $n$ in Table \ref{epsilontable} indicates that the fraction of the $n_{h}=\frac{1}{2}$ charge order progressively increases in the stripe ordering pattern. A similar discommensuration pattern of stripe-type ordering was recently reported in a heavily-doped insulating manganite system La$_{1-x}$Ca$_{x}$MnO$_{3}$ for $x > \frac{1}{2}$ \cite{mori}.
\begin{figure}
\centering \leavevmode
\psfig{file=lsno_CO_model.eps,width=0.9\hsize}
\caption{Models of stripe order for $n_h=1/3$ (a), $1/2$ (b), and $n_h=2/5$ (c), respectively. Dashed lines indicate the magnetic unit cell, while the shaded area denotes the $\frac{3}{2}a$-width unit of the $\epsilon = \frac{1}{3}$ stripe order.}
\label{CO_model}
\end{figure}
The high density of hole concentration drastically influences the two ordering temperatures, $T_{\rm CO}$ and $T_{\rm N}$ for $n_h > \frac{1}{3}$. (1) Up to $n_h = \frac{1}{3}$, the spin order is essentially an antiferromagnetic order which is separated by an antiphase domain boundary of hole stripes. For the $n_h = \frac{1}{2}$ ordering pattern, however, the spin order is actually a 2D checkerboard pattern as depicted in Fig. \ref{CO_model}(b), and all the nearest neighbor sites of the spins are occupied by holes. The existence of holes reduces the effective exchange interactions between spins, and lowers $T_{\rm N}$, being consistent with studies of spin dynamics in O-doped nickelates \cite{Nak,TraO97}. (2) As seen in the right column of Table I, the discommensuration of the stripe order progressively increases the fraction of the $\epsilon = \frac{1}{2}$ pattern for $\frac{1}{3} \leq n_h \leq \frac{1}{2}$. In the matrix of the $\epsilon =\frac{1}{3}$ stripe order, intervening $n_h =\frac{1}{2}$ patterns strongly disturb the spin/charge correlation within and between the NiO$_2$ planes as manifested by the $n_h$ depencence of $\kappa$ in Fig. \ref{n_vs_Tc}. These effects in turn cause the suppression of $T_{\rm CO}$ and $T_{\rm N}$ for $n_h > \frac{1}{3}$.
\begin{minipage}{0.85\hsize}
\centering
\begin{table}
\caption{Incommensurability $\epsilon$ observed in the present work. See the text for details.}
\label{epsilontable}
\begin{tabular}{cdcc}
$n_h$ & $\epsilon^{\rm obs}$ & $\frac{m+n}{4m+3n}$ & $\frac{n+m^{\prime}}{3n+2m^{\prime}}$ \\ \hline
0.289 & 0.285 & (1,1) & \\
1/3 & 0.332 & & (1,0) \\
0.398 & 0.365 & & (3,1) \\
0.425 & 0.398 & & (1,1) \\
0.462 & 0.410 & & (2,3) \\
1/2 & $\sim$0.455 & & (1,3) or (1,4)
\end{tabular}
\end{table}
\end{minipage}
Finally, we point out that $\epsilon$ was slightly shifted towards $\epsilon = \frac{1}{3}$ for both sides of $n_h =\frac{1}{3}$ as indicated by a dashed curve as shown in Fig. \ref{n_vs_Tc}. One can see that farther the distance of $n_h$ from $n_h =\frac{1}{3}$, larger the deviation of $\epsilon$ from the $\epsilon \sim n_h$ law. This behavior has an interesting implication in the transport properties \cite{Kat99}. In the stripe model, the hole density in a stripe $n_{\rm st}$ is always $n_{\rm st}=1$ for all $\epsilon$ when the $\epsilon \sim n_h$ law holds. Here, the hole density in a stripe $n_{\rm st}$ is defined as $n_{\rm st} \equiv$ \{number of holes/Ni site\}/\{number of domain walls (DW)/Ni site\} $= n_h / \epsilon$. Because one hole exists per each Ni sites, hole stripes are half filled and they are Mott insulator-like. The deviation of $\epsilon$ from the $\epsilon \sim n_h$ law indicates that the hole density deviates from $1$ for both sides of $n_h = \frac{1}{3}$. For $n_h < \frac{1}{3}$, $n_{\rm st} \lesssim 1$, and the carriers are expected to be electron-like, while for $n_h >\frac{1}{3}$, $n_{\rm st} \gtrsim 1$, and the carriers are hole-like. This consideration is fully consistent with the change of the sign of the Hall coefficient $R_{\rm H}$ at $n_h = \frac{1}{3}$ \cite{Kat99}.
In summary, we have presented that the region of the stripe order extends up to $n_h \sim \frac{1}{2}$ at low temperatures $\sim$10 K in La$_{2-x}$Sr$_{x}$NiO$_4$. Incommensurability $\epsilon$ shows the $\epsilon \sim n_h$ law with systematic deviation around $n_h = \frac{1}{3}$, which controls the nature of carriers in hole stripes. The stripe order consists of combination of the $\epsilon = \frac{1}{3}$ stripe order and the $n_h = \frac{1}{2}$ charge order, and $\epsilon$ is given by $(m+n)/(4m+3n)$ for $n_h < \frac{1}{3}$ or by $(n+m^{\prime})/(3n+2m^{\prime})$ for $n_h > \frac{1}{3}$. The $n_h = \frac{1}{3}$ stripe order is stabilized by the coincidence of the periodicities of the charge, and spin order, and as a result, it forms quasi-3D LRO, while it is 3D SRO for $n_h < \frac{1}{4}$, and quasi-2D LRO for $0.4 \lesssim n_h \lesssim \frac{1}{2}$.
We thank J. M. Tranquada for valuable discussions. This work was supported by a Grant-In-Aid for Scientific Research from the Ministry of Education, Science and Culture, Japan and by the New Energy and Industrial Technology Development Organization (NEDO) of Japan.
|
\section{Introduction}
In certain cosmological schemes, as is well known, the universal constant of
gravitation $G$, changes very slowly with time\cite{r1,r2,r3,r4}. These
include Dirac's large number cosmology and fluctuational cosmology. In these
cases the variation is given by
\begin{equation}
G = \frac{\beta}{T}\label{e1}
\end{equation}
where in fluctuational cosmology referred to (cf. also refs.\cite{r5,r6}), $\beta$
is given in terms o f the constant microphysical parameters by
$$\beta \approx \frac{l \hbar}{m^2}$$
where $l$ is the pion Compton wavelength and $m$ its mass.\\
(It may be pointed out in passing that while Dirac's cosmology has well known
inconsistencies, the latter theory is consistent with observation and predicts
an ever expanding universe as latest observations of distant supernovae do
indeed confirm. In addition, not only are the Large Number coincidences accounted
for, but also Weinberg's mysterious empirical relation between the pion mass and
the Hubble constant is deduced from the theory (cf. references).)\\
In any case, what we now propose to show is, that starting from (\ref{e1}),
we can account for the perhelion precession of the planet Mercury as also for
anomalous acceleration of the planets and more generally for the solar system
bodies and in addition predict anomalous changes in the orbital eccentricities.\\
\section{Solar System Orbits}
We now deduce using (\ref{e1}), the perhelion precession of Mercury. We first
observe that from (\ref{e1}) it follows that
\begin{equation}
G = G_o (1+ \frac{t}{t_o})\label{e2}
\end{equation}
where $G_o$ is the present value of $G$ and $t_o$ is the present age
of the universe and $t$ the time elapsed from the present epoch. Similarly one
could deduce that (cf.ref.\cite{r1}),
\begin{equation}
r = r_o \left(\frac{t_o}{t_o+t}\right)\label{e3}
\end{equation}
We next use Kepler's Third law\cite{r7}:
\begin{equation}
\tau = \frac{2 \pi a^{3/2}}{\sqrt{GM}}\label{e4}
\end{equation}
$\tau$ is the period of revolution, $a$ is the orbit's semi major axis,
and $M$ is the mass
of the sun. Denoting the average angular velocity of the planet by
$$\dot \Theta \equiv \frac{2 \pi}{\tau},$$
it follows from (\ref{e2}), (\ref{e3}) and (\ref{e4}) that
$$\dot \Theta - \dot \Theta_o = - \dot \Theta_0 \frac{t}{t_o},$$
where the subscript $o$ refers to the present epoch, \\
Whence,
\begin{equation}
\omega (t) \equiv \Theta - \Theta_o = - \frac{\pi}{\tau_o t_o} t^2\label{e5}
\end{equation}
Equation (\ref{e5}) gives the average perhelion precession at time '$t$'.
Specializing to the case of Mercury, where $\tau_o = \frac{1}{4}$ year, it
follows from (\ref{e5}) that the average precession per year at time '$t$' is
given by
\begin{equation}
\omega (t) = \frac{4\pi t^2}{t_0}\label{e6}
\end{equation}
Whence, considering $\omega (t)$ for years $t=1,2, \cdots , 100,$ we can
obtain from (\ref{e6}), the usual total perhelion precession per century as,
$$\omega = \sum^{100}_{n=1} \omega (n) \approx 43'' ,$$
if the age of the universe is taken to be $\approx 2 \times 10^{10}$ years.\\
Conversely, if we use the observed value of the precession in (\ref{e6}),
we can get back the above age of the universe.\\
It can be seen from (\ref{e6}), that the precession depends on the epoch.\\
We next demonstrate that orbiting objects will have an anamolous inward
radial acceleration.\\
Using the well known equation for Keplarian orbits (cf.ref.\cite{r7}),
\begin{equation}
\frac{1}{r} = \frac{GMm^2}{l^2} (1 + e cos \Theta)\label{e7}
\end{equation}
\begin{equation}
\dot r^2 = \frac{GM}{rm} - \frac{l^2}{m^2r^2}\label{e8}
\end{equation}
$l$ being the orbital angular momentum constant and $e$ the eccentricity of
the orbit, we can deduce such an extra inward radial acceleration, on
differentiation of (\ref{e8}) and using (\ref{e2}) and (\ref{e3}),
\begin{equation}
a_r = \frac{GM}{2t_o r \dot r}\label{e9}
\end{equation}
It can be easily shown from (\ref{e7}) that
\begin{equation}
\dot r \approx \frac{eGM}{rv}\label{e10}
\end{equation}
For a nearly circular orbit $rv^2 \approx GM$, whence use of (\ref{e10}) in
(\ref{e9}) gives,
\begin{equation}
a_r \approx v/2 t_o e\label{e11}
\end{equation}
For the earth, (\ref{e11}) gives an anomalous inward radial acceleration
$\sim 10^{-9} cm/sec^2,$ which is known to be the case\cite{r8}.\\
We could also deduce a progressive decrease in the eccentricity of orbits. Indeed,
$e$ in (\ref{e7}) is given by
$$e^2 = 1+\frac{2El^2}{G^2m^3M^2} \equiv 1 + \gamma , \gamma < 0.$$
Use of (\ref{e2}) in the above and differenciation, leads to,
$$\dot e = \frac{\gamma}{et_o} \approx - \frac{1}{et_o} \approx - \frac{10^{-10}}
{e} \mbox{per year},$$
if the orbit is nearly circular. (Variation of eccentricity in the usual
theory have been extensively studied (cf.ref.\cite{r9} for a review).\\
We finally consider the anomalous accelerations given in (\ref{e9}) and (\ref{e11})
in the context of space crafts leaving the solar system.\\
If in (\ref{e9}) we use the fact that $\dot r \leq v$ and approximate
$$v \approx \sqrt{\frac{GM}{r}},$$
we get,
$$a_r \geq \frac{1}{et_o} \sqrt{\frac{GM}{r}}$$
For $r \sim 10^{14}cm$, as is the case of Pioneer $10$, this gives,
$a_r \geq 10^{-11}cm/sec^2$\\
Interestingly Anderson et al.,\cite{r10} claims to have observed an anomalous inward acceleration
of $\sim 10^{-9} cm/sec^2$
\newpage
|
\section{Introduction}
How did the first objects form after the Big Bang? In hierarchical
cosmogonies (\eg Turner 1998), the first gravitationally bound systems
may have been stars and small star--forming systems which merge to form
galaxies in large dark matter halos. Arising from the end products of
stellar evolution and mergers, central black holes could grow to become
extremely massive. However, it is not clear how this process would
work at very high redshifts, where little time is available. It has been
suggested that primordial black holes may form well before their host
galaxies (Loeb 1993). In any case, accretion events fueling massive
black holes are thought to manifest themselves as active galactic nuclei
(AGN; \eg Rees 1984). Due to their extreme luminosity, AGN are convenient
beacons for exploring these formative, `Dark Ages' of our Universe.
Extragalactic radio sources have played an important role in identifying
active galaxies at high redshifts. The most distant known {\it galaxies}
have consistently been radio--selected until only very recently. In this
Letter we report the discovery of a radio galaxy at $z = 5.19$. At this
redshift it is the most distant known AGN, surpassing even quasars for the
first time in 36 years. Throughout this paper we use $H_0 = 65 h_{65}
\kmsMpc$, $\Omega_M = 0.3$, and $\Lambda = 0$. For these parameters,
1\arcsec\, subtends 7.0 $h_{65}^{-1}$ kpc at $z = 5.19$ and the Universe
is only 1.08 Gyr old, corresponding to a lookback time 91.1\% of the
age of the Universe.
\section{Source Selection}
The most efficient method to find high--redshift radio galaxies (HzRGs) is to
combine two well--known techniques. The first is to select radio sources with
ultra--steep spectra (USS) at radio wavelengths, i.e.\ very red radio colors (\eg
Chambers, Miley, \& van Breugel 1990). Most powerful radio galaxies have radio
spectral energy distributions which steepen with frequency. Therefore, at fixed
observing frequencies more distant sources exhibit steeper spectra (\eg van
Breugel \etal 1999).
A second selection criterion relies upon the magnitude--redshift
relationship at infrared wavelengths, or $K-z$ Hubble diagram, for
powerful radio galaxies (Figure~\ref{kz}). At low redshifts ($z < 1$),
powerful radio galaxies are uniquely associated with massive galaxies.
The well--behaved $K-z$ diagram suggests that such galaxies can be found
through near--IR identification. This has been confirmed by the discovery
of many $3 < z < 4.4$ radio galaxies which approximately follow the $K-z$
relationship, even to the highest redshifts and despite significant
morphological evolution (van Breugel \etal 1998).
Using several new, large radio surveys we constructed a USS sample
($S_\nu \propto \nu^\alpha; \alpha^{\rm 1.4 GHz}_{\rm 365 MHz} <
-1.30$; De Breuck \etal 1999 [DB99]) which is much larger, more accurate,
and reaches fainter flux density limits than previous such samples.
\tn, with $\alpha^{\rm 1.4 GHz}_{\rm 365 MHz} = -1.63 \pm 0.08$,
is among the steepest sources of our sample. VLA observations
at 4.85 GHz show the source is a slightly resolved $1\farcs2$ double,
with $S_{4.85GHz} = 8.6\pm0.5$ mJy, centered at $\alpha_{\rm J2000} =
09^h24^m19\fs92$, $\delta_{\rm J2000} = -22\arcdeg 01\arcmin 41\farcs5$
(Figure~\ref{kimage}).
\section{Observations}
We obtained $K_s$ images of \tn\ using NIRC (Matthews \&
Soifer 1994) at the Keck~I telescope. We integrated for 32 minutes
on UT 1998 April 18 in photometric conditions with $0\farcs5$ seeing,
and again for 32 minutes on UT 1998 April 19 through light cirrus with
$0\farcs6$ seeing. The observing procedures, calibration, and data reduction
techniques were similar to those described in van Breugel \etal (1998).
The final image comprising 3840~s of on--source integration is shown in
Figure~\ref{kimage}. Using circular apertures of $2\farcs1$ diameter,
encompassing the entire object, we measure $K = 21.15$ for night 1,
and 21.45 for night 2. We estimate that $K = 21.3 \pm 0.3$. If \tn\
is at $z = 5.19$ (\S4), then redshifted {\rm [O~II]}\ at $\lambda = 2.307\mu$m
would be included in the $K_s$ passband and some of the $K$-band flux
might be due to line emission.
We obtained spectra of \tn\ through a 1\farcs5 wide, 3\arcmin\ long
slit using LRIS (Oke \etal 1995) at the Keck~II telescope. The
integration times were 5400~s on UT 1998 December 19 (position angle
0\ifmmode {^{\circ}}\else {$^\circ$}\fi) and 4400~s on UT 1998 December 20 (position angle 180\ifmmode {^{\circ}}\else {$^\circ$}\fi);
both nights were photometric with 0\farcs6 seeing. The observations
used the 150 lines mm$^{-1}$ grating ($\lambda_{\rm blaze} \approx
7500$ \AA; $\Delta\lambda_{\rm FWHM} \approx 17$ \AA), sampling the
wavelength range 4000 \AA\ to 1$\mu$m. Between each 1800~s exposure,
we reacquired offset star A (see Fig.~2), performed 20\arcsec\ spatial
shifts to facilitate removal of fringing in the reddest regions of the
spectra, and blind offset the telescope to return \tn\ within the
slit. We calculated the dispersion using a NeAr lamp spectrum
taken immediately subsequent to the observations (RMS variations of
0.50 \AA), and adjusted the zero point according to telluric emission
lines. Final wavelength calibration is accurate to 1 \AA. The
spectra were flux calibrated using observations of Feige~67 and
Feige~110 obtained on each night and were corrected for foreground
Galactic extinction using a reddening of $E_{B - V} = 0.0168$
determined from the dust maps of Schlegel, Finkbeiner, \& Davis
(1998).
We find a strong, single emission line at $\lambda \sim 7530$
\AA\, which shifts by $\approx 16$ \AA\ between the two nights.
(Figure~\ref{spectrum}; Table~1). The cause of the line offset is
unclear, though it may be related to problems LRIS was experiencing
with slippage in the movable guider at the time of the observations.
The relative brightnesses of other sources on the slit vary between
each 1800~s observation, indicating that despite our precautions of
reacquiring the target after each exposure, guider slippage must have
caused some variations in telescope offsetting. These slight pointing
changes may have caused the slit to sample different regions of
spatially--extended, line--emitting gas. Indeed, \tn\ shows two
separate components at $K$ (Figure~\ref{kimage}), and emission--line
regions of HzRGs are known to be kinematically complex (Chambers,
Miley \& van Breugel 1990; van Ojik \etal 1997).
Line parameters are measured with a Gaussian fit to the emission line
and a flat (in $F_\lambda$) fit to the continuum (Table~1).
Equivalent width values were derived from a Monte Carlo analysis
using the measured line flux and continuum values with errors, subject
to the constraint that both are positive. For UT 1998 Dec.\ 19, when no
continuum was reliably detected, we quote the 90\% confidence
limit, $W^{\rm obs} > 2760$ \AA. For UT 1998 Dec.\ 20, when continuum
was marginally detected, we quote the 90\% confidence interval, $W^{\rm
obs} = 710 - 1550$ \AA.
\section{Redshift Determination}
As discussed by Dey \etal (1998) and Weymann \etal (1998) for two $z >
5$ Ly$\alpha$~-emitting field galaxies, a solitary, faint emission line at
red wavelengths is most likely to be either low-redshift {\rm [O~II]}\ or
high-redshift Ly$\alpha$~. Similar arguments are even more persuasive for
HzRGs because of their strong, rich emission line spectra. For
example, if the line at $\approx$ 7530 \AA\ were [\ion{O}{2}]$\lambda$3727~\ at $z =
1.020$ then composite radio galaxy spectra (McCarthy 1993; Stern \etal
1999a) indicate that the \tn\ spectrum should have shown {\rm CII]}$\lambda$2326~\ at 4699
\AA\, with $\approx 40 - 70$\% the strength of [\ion{O}{2}]$\lambda$3727~, and {\rm {\ion{Mg}{2}$\lambda\lambda$2796,2803~\ at
5653 \AA\, with $\approx 20 - 60$\% the strength of [\ion{O}{2}]$\lambda$3727~. Similar
arguments rule out identifying the emission line with \Ha\ at $z =
0.147$ or {\rm [O~III]}\ at $z = 0.504$, since in these cases even stronger
confirming lines should have been seen.
The large equivalent widths also argue against identifying the emission
line with [\ion{O}{2}]$\lambda$3727~\ at $z = 1.020$, implying $W_{\rm [OII]}^{\rm rest} > 1370$ \AA\ (night
1) and $350 < W_{\rm [OII]}^{\rm rest} < 770$ \AA\ (night 2). Radio galaxy composites
typically have rest--frame [\ion{O}{2}]$\lambda$3727~\ equivalent widths of $\approx 130$
\AA\, (McCarthy 1993; Stern \etal 1999a), though active galaxies with
extreme $W_{\rm [OII]}^{\rm rest}$ are occasionally observed ($W_{\rm [OII]}^{\rm rest} \approx 750$
\AA; Stern \etal 1999b). The equivalent width of \tn\ is more typical
of high-redshift Ly$\alpha$~\ which is often observed with rest frame values
of several $\times$ 100 \AA\ in HzRGs (Table 2). We also note that the
observations from the second night show that Ly$\alpha$~\ is attenuated on the
blue side, presumably due to associated and intervening hydrogen gas,
as is commonly observed in HzRGs (\eg van Ojik \etal\ 1997; Dey 1997)
and normal star-forming galaxies at $z > 5$ (\eg Dey \etal 1998).
Finally, the faint $K$-band magnitude of \tn\ conforms to the
extrapolation of the $K - z$ relation to $z > 5$ (Figure~\ref{kz}).
Identifying the emission line with [\ion{O}{2}]$\lambda$3727~\ would imply a severely
underluminous HzRG (by 3 -- 4 mag). Therefore, the most
plausible identification of the emission line in \tn\ is with Ly$\alpha$~\ at a
(mean) observed wavelength of 7530 \AA\ and $z = 5.19$. Table~1 gives
the dereddened emission--line fluxes.
\section{Discussion}
Among all known $z \mathrel{\spose{\lower 3pt\hbox{$\mathchar"218$} 3.8$ HzRGs, \tn\ is fairly typical in radio
luminosity, equivalent width, and velocity width (Table~2). But this
source has the steepest radio spectrum, consistent with the $\alpha -
z$ relationship for radio galaxies (\eg\ R\"ottgering \etal 1997). \tn\
also has the smallest linear size, perhaps indicating that the source is
relatively young and/or embedded in a denser environment compared to the
other HzRGs, commensurate with its large velocity width (van Ojik\etal
1997) and very high redshift. Together with 8C~1435$+$63, \tn\ appears
underluminous in Ly$\alpha$~, which might be caused by absorption in a relatively
dense cold and dusty medium. Evidence for cold gas and dust in some of
the most distant HzRGs has been found from sub--mm continuum and CO--line
observations of 8C~1435$+$63 and 4C~41.17 (\eg Ivison \etal 1998).
Our observations of \tn\ extend the Hubble $K-z$ diagram for powerful
radio galaxies to $z = 5.19$. Simple stellar evolution models are
shown in Figure~\ref{kz} for comparison with the HzRG. Despite the
enormous $k$--correction effect (from $U_{\rm rest}$ at $z = 5.19$ to
$K_{\rm rest}$ at $z = 0$) and strong morphological evolution (from
radio--aligned to elliptical structures), the $K-z$ diagram remains a
powerful phenomenological tool for finding radio galaxies at extremely
high redshifts. Deviations from the $K-z$ relationship may exist
(Eales \etal 1997; but see McCarthy 1998), and scatter in the $K-z$
values appears to increase with redshift.
The clumpy, radio--aligned $U_{\rm rest}$ morphology resembles that of
other HzRGs (van Breugel \etal 1998; Pentericci \etal 1998). If the
continuum is dominated by star light, as appears to be the case in the
radio--aligned HzRG 4C~41.17 at $z = 3.798$ (Dey \etal 1997), then
$M(U) = -24.4$ for \tn\. Then we can derive a SFR of $\sim$200
M$_\odot$ yr$^{-1}$, assuming a Bruzual \& Charlot (1999) GISSEL
stellar evolution model with metallicity $Z = 0.008$, no extinction,
and a Salpeter IMF. This SFR value is highly uncertain due to the
unknown, but competing, effects of extinction and [\ion{O}{2}]$\lambda$3727~\
emission--line contamination, but is not unreasonable. It is 2.5 times
{\it less} than in 4C~41.17, which has $M(U) = -25.2$ using the same
aperture (Chambers \etal 1990). \tn\ may be a massive, active galaxy
in its formative stage, in which the SFR is boosted by induced star
formation (\eg Dey \etal 1997). For comparison other, `normal' star
forming galaxies at $z > 5$ have 10 -- 30 times lower SFR ($\sim 6 -
20 \msun yr^{-1}$; Dey \etal 1998; Weymann \etal 1998; Spinrad \etal
1998).
Recent $z \sim 3$ and $z \sim 4$ Lyman--break galaxy observations have
suggested a possible divergence of star formation and AGN activity at
high redshift (Steidel \etal 1999), contrary to what was previously
thought (\eg Haehnelt, Natarajan \& Rees 1998). However, if
starbursts and AGN are closely coupled, as suggested to explain the
ultraluminous infrared galaxies (Sanders \& Mirabel 1996), then young
AGN may inhabit especially dusty, obscured galaxy systems. To obtain
a proper census of the AGN population at the very highest redshifts
therefore requires samples which avoid optical photometric selection
and extinction bias, such as our cm--wavelength/$K$-band radio galaxy
sample.
As emphasized by Loeb (1993), if massive black holes form in a
hierarchical fashion together with their host galaxies, this process
must be quick and efficient, as available timescales are short: at $z
= 5.19$ the Universe is only 1 Gyr old. It is unclear how this could
be done, so other models, where primordial massive black holes form
soon after the Big Bang and {\it prior} to the beginning of galaxy
formation, may require additional investigation.
\acknowledgments
We thank G.\ Puniwai, W.\ Wack, R.\ Goodrich and R.\ Campbell for their
expert assistance during our observing runs at the W.M.\ Keck Observatory,
and A.\ Dey, J.R.\ Graham and H.\ Spinrad for useful discussions.
The work by W.v.B., C.D.B. and S.A.S.\ at IGPP/LLNL was performed under
the auspices of the US Department of Energy under contract W-7405-ENG-48.
W.v.B.\ also acknowledges support from NASA grant GO 5940, and D.S. from
IGPP/LLNL grant 98--AP017.
|
\section{Introduction}
One of the recent results leading to better understanding of
quantum entanglement \cite{EPR,Schrodinger} was realizing that there are two
qualitatively different types of entanglement of mixed states of two-component
systems \cite{bound,Pawel}.
Namely,
there is {\it free entanglement} (FE) which can be converted into pure singlet
form by means of local quantum operations and classical communication (LQCC).
Such a process is called distillation \cite{Bennett_pur} and it allows to use
the noisy entanglement for the purposes of quantum communication. However,
there is also {\it bound entanglement} (BE), which cannot be distilled
\cite{bound,Pawel}.
At present the
structure and properties of BE state are being extensively investigated
\cite{aktyw,Popescu99,upb,Terhal,single,Rains}. In particular, a
striking connection between the bound etanglement and nonlocality without
entanglement \cite{nonlocality} has been discovered \cite{upb}.
Also, the bound
entanglement implies
a new approach in entanglement measures: one must, in general leave the
paradigm that a measure of entanglement should vanish only on separable
states. Indeed, at present we know that physically the most relevant
measure of entanglement \cite{huge,Rohrlich,Plenio} which is
distillable entanglement
does not satisfy this condition (it vanishes on the bound entangled states).
The above, more general approach allowed to obtain a new bound on
distillable entanglement
Ref. \cite{Rains}. Due to the connection between entanglement
and positive maps \cite{sep} the investigation of bound entanglement
was also fruitful for pure mathematics. Namely, by use of results
on bound entanglement of Ref. \cite{upb} the first {\it systematic} way
of constructing the so called non-decomposable positive maps was
found in Ref. \cite{Terhal}.
In this paper we would like to investigate the {\it processes} of
interaction with environment, which lead to bound entanglement.
In general, the mixed states emerge from
interaction with environment, which is very hard to be avoided in realistic
situation. Such interaction may completely destroy the initial pure
entanglement, or sometimes there may remain some residual entanglement, free
or bound. We will be interested in the processes for which the residual
entanglemet is the bound one.
To be more
precise, imagine that Alice can send particles to Bob via a quantum channel
$\Lambda$ (representing the interaction with environment). Alice and Bob
are allowed to support the quantum channel by using LQCC operations
and can
enhance the transmission by sending entangled particles down the channel.
The latter means that effectively they have a channel
$\Lambda^{\otimes N}$ for arbitrary $N$.
Now we are interested in such channels that Alice and Bob
(i) cannot send reliably quantum information (equivalently, cannot produce
asymptotically singlet state);
(ii) can produce a BE state. Such channels we will call {\it binding
entanglement channels} (BE channels).
We prove a theorem characterizing such channels, which says that a channel
is BE if and only if sending half of maximally entangled pair through
the channel, one obtains BE state.
It follows that a channel is
BE if there exists a pure entangled state such, that if sent through the
channel it becomes bound entangled. Thus knowing the examples of
BE states, we can
construct the BE channels. We provide a way of constructing BE channel from
any given BE state. Our investigations are based on the
general connections between channels and bipartite states investigated in
\cite{Jamiol,huge,Nielsen,xor,single}.
\section{Binding entanglement channels: characterization}
To begin with, let us introduce some notation. By a channel we mean any
completely positive (CP) trace-preserving map.
A completely positive map $\Lambda:M_m\rightarrow M_n$ will be denoted by
$\Lambda_m^n$ (here $M_n$ denotes the set of $n\times n$ square matrices.
The identity map acting on $M_n$ will be denoted by $I_n$.
Maximally entangled state on the system $M_n\otimes M_n$ of the form
\begin{equation}
P_+^n={1\over \sqrt n}\sum_{i=1}^n |i\rangle|i\rangle
\end{equation}
will be called singlet state. A state acting on
the Hilbert space $C^m\otimes C^n$ will be denoted
by $\varrho_{m,n}$ (or $\sigma_{m,n}$ etc.). Sometimes, if it does not
lead to misunderstanding we will not write the indices explicitly.
Finally, $\varrho_{ikjl}$ denotes matrix element of the state $\varrho$
in product basis
\[
\varrho_{ikjl}\equiv \langle e_i\otimes
f_k|\varrho|e_j\otimes f_l\rangle.
\]
{\bf Definition.} \it We say that a channel $\Lambda$ is \rm binding
entanglement channel \it iff (i) $Q_2(\Lambda)=0$ and (ii) it is possible to
obtain bipartite bound entangled state by means of (possibly multiply) use of
the channel and LQCC operations. \rm
Here $Q_2$ is the quantum capacity of a channel supported by LQCC action (the
subscript $2$ indicates two-way classical communication) \cite{huge}.
Now we will prove a theorem characterizing such channels in terms of
BE bipartite states.
{\bf Theorem.} A channel $\Lambda:M_m\rightarrow M_n$ is binding entanglement
iff the state $(I\otimes \Lambda)P_+^m$ (acting on $C^m\otimes C^n$)
is a BE state.
{\bf Proof.} Let us first prove the sufficiency of the condition. If
$(I\otimes \Lambda_m^n)P_+^m$ is BE state then (ii) is obviously satisfied, so
that one needs to prove that the condition implies also (i). Suppose,
conversely, that $Q_2(\Lambda_m^n)>0$. Then, one can produce asymptotically pure
singlets by use of the channel and LQCC. The first stage of the most
general protocol of producing singlet pairs is sending half of some
state $\sigma_{{k\times N},{m\times N}}$ via
the channel $\Lambda^{\otimes N}$ (denote it by
$\Lambda_{m\times N}^{n\times N}$). The second stage
amounts to distillation of the emerging state
$\varrho_{{k\times N},{m\times N}}=(I_{k\times N}\otimes
\Lambda_{m\times N}^{n\times N}) \sigma$.
Hence, to obtain finally the singlets, the state $\varrho$ must be FE.
We will now show that this implies that $(I_m\otimes \Lambda_m^n) P_+^m$
must be also FE. To see it, note that the state $\sigma$ (as any state)
can be written as
$\sigma=(\Gamma_{m\times N}^{k\times N}\otimes I_{m\times N})P_+^{m\times N}$
(where $\Gamma$ is CP, but not necessarily trace-preserving map).
So we have
\begin{eqnarray}
&&\varrho=(I_{k\times N}\otimes \Lambda_{m\times N}^{n\times N})
(\Gamma_{m\times N}^{k\times N}\otimes I_{m\times N}) P_+^{m\times N}=
\nonumber\\
&&(\Gamma_{m\times N}^{k\times N}\otimes \Lambda_{m\times N}^{n\times N})
P_+^{m\times N}=\nonumber\\
&&(\Gamma_{m\times N}^{k\times N}\otimes I_{n\times N})(I_{m\times N}\otimes
\Lambda_{m\times N}^{n\times N})P_+^{m\times N}
\end{eqnarray}
Now, since $\varrho$ is FE, then also $(I_{m\times N} \otimes
\Lambda_{m\times N}^{n\times N})P_+^{m\times N}$ must be
FE (indeed the action $\Gamma\otimes I$ is LQCC one, hence cannot produce FE
state from a BE one). Now, since $(I_{m\times N}\otimes
\Lambda_{m\times N}^{n\times N})P_+^{m\times N}=
\left( (I_m\otimes \Lambda_m^n)P_+^m\right)^{\otimes N}$ we obtain that
also $(I_m\otimes\Lambda_m^n)P_+^m$ must be FE, which is a contradiction.
Hence, if $(I_m\otimes\Lambda_m^n)P_+^m$ is BE then the condition (i)
is satisfied.
Now, we will show that the condition that
$(I_m\otimes\Lambda_m^n)P_+^m$ is BE is also a necessary one for $\Lambda$
to be BE. Suppose, conversely, that $(I_m\otimes\Lambda_m^n)P_+^m$ is not BE.
Then it can be separable or FE. If its is FE, then one can distill it and
obtain nonzero $Q_2$ so that the condition (i) is violated. If, instead
$(\Lambda_m^n\otimes I_m)P_+^m$ is separable, then we will show that the
condition (ii) is violated. Indeed, if for some state $\sigma_{k,m}$ the state
$\varrho_{m,n}=(I_m\otimes\Lambda_m^n)\sigma_{k,m}$ is BE, then
writing $\sigma$ as $\sigma_{k,m}=(\Gamma_k^m\otimes I_m)P_+^m$ we
obtain, similarly as in the proof of sufficiency, that
$\varrho=(\Gamma_k^m\otimes I_n)(I_m\otimes\Lambda_m^n)P_+^m$.
Then, since $\Gamma\otimes I$ is LQCC, we obtain that
$(I_m\otimes\Lambda_m^n)P_+^m$ cannot be separable (LQCC action cannot make BE
state from separable one). This ends the proof.
From the above characterization of BE channels it follows that
given a channel with $Q_2=0$, if bound entanglement can be created at
all, then it can be created without exchange of classical information between
Alice and Bob but merely by sending half of singlet pair through the channel.
Hence also multiply use of channel is not needed.
\section{Binindg entanglement channels from bound entangled states}
In this section we will provide a procedure of constructing BE channels
from BE states.
As one knows there is an
isomorphism between the set of states $\varrho_{m,n}$ with maximally mixed
reduction $\varrho_A$ and the channels $\Lambda_m^n$.
It is given just by the formula:
\begin{equation}
\varrho_{m,n}=(I_m\otimes \Lambda_m^n)P_+^m
\label{isom1}
\end{equation}
(the maximally mixed reduction is connected with the fact that channels
preserve trace).
In other words, if one has a channel, one can send half of singlets through it
to obtain the state with maximally mixed reduction, and, conversely,
any state of maximally mixed reduction emerges from sending half of singlet
down some channel. Explicitly, the connection between matrix elements of state
and associated channel is the following
\begin{equation}
\langle f_k|\Lambda(|e_i\rangle\langle
e_j)|f_l\rangle\equiv\lambda_{klij}=
\varrho_{ikjl}\equiv \langle e_i\otimes
f_k|\varrho|e_j\otimes f_l\rangle.
\label{isom}
\end{equation}
So we can provide examples of BE channels basing on the known BE states
with maximally mixed reduction. However, one also knows the examples of BE
states with none of reductions maximally mixed \cite{upb}. How to associate
channels with them? As mentioned above, the maximally mixed reduction
is connected with the fact that
the channel acts only on one half of the singlet, so that,
being trace-preserving, it cannot disturb the other one. Since the singlet is
maximally entangled, it has maximally mixed reduction that is inherited by
the final state. Now, if a state with {\it non-maximally}
mixed reductions is concerned, one can imagine it emerges from sending
{\it non-maximally entangled pure state} via a channel. The state must have
the same reduction as the mixed state of interest (as, again, the channel will
not affect that reduction). To recover such a channel from the given state
$\varrho$, we will first transform it into a state $\sigma$ of maximally
mixed reduction by means of LQCC action. Then the channel will be the one
associated with $\sigma$ via the state-channel isomorphism (\ref{isom1}).
Let a BE state $\varrho_{m,n}$ acts on ${\cal H}_A\otimes {\cal H}_B$ and
let ${\cal H}'_A$ be the
support of its reduction $\varrho_A$ with $\dim{\cal H}'_A=k$. Then define
\begin{equation}
\sigma_{r,n}=(r\varrho_A)^{-1/2}\otimes I\,\varrho\,(r\varrho_A)^{-1/2}
\otimes I,
\end{equation}
where $\varrho_A$ was inverted on its support ${\cal H}_A'$
Here we used the fact that the support of any state is equal to
product of the supports of its reductions (see Appendix), so that, in
fact, both
$\varrho$ and
$\sigma$ acts on ${\cal H}'_A\otimes{\cal H}_B$. It is easy to check
that $\sigma_A={I\over r}$. Indeed,
choosing the basis $\{e_i\}\subset{\cal H}_A^i$ to be eigenbasis of
$\varrho_A$ (i.e.
$\varrho_A=\sum_ip_i|e_i\rangle\langle e_i|$) we obtain
\begin{equation}
\sigma_{ikjl}={1\over r\sqrt {p_ip_j}}\varrho_{ikjl},
\end{equation}
hence
\begin{equation}
(\sigma_A)_{ij}=\sum_k\sigma_{ikjk}={1\over r\sqrt {p_ip_j}} p_i\delta_{ij}=
{1\over r} \delta_{ij}
\end{equation}
Now, as the state $\sigma$ was created from $\varrho$ by LQCC action, then
it is BE (the action is called filtering \cite{conc}). Then the
seeked BE channel $\Lambda_A$ corresponding to the given state $\varrho$ is the
one associated with the state $\sigma$ via the formula (\ref{isom})
(the subscript $A$ indicates that we recover the channel by use of the reduction
$\varrho_A$).
Then to obtain explicit form of $\Lambda_A$ one needs to calculate the
map $\Theta$ given by the formula
\begin{equation}
(I_r\otimes \Theta_r^n)P_+^r=\varrho.
\label{map-state}
\end{equation}
Then $\Lambda_A$ is given by
\begin{equation}
\Lambda_A= \Theta\circ \Gamma^T_A
\end{equation}
where
$\Gamma_A(\cdot)={1\over r} \varrho_A^{-1/2}
(\cdot)\varrho_A^{-1/2}$ and $T$ is tranpose in the space of
maps i.e. $(\Theta_B^T)_{klij}=(\Theta_B)_{ijkl}$. If the given map $\Lambda$
is CP (as in our case) and its Stinespring form is
$\Lambda(\cdot)=\sum_iV_i(\cdot)V_i^\dagger$ then the transposed map is
given simply by $\Lambda(\cdot)=\sum_iV^T_i(\cdot){(V_i^T)}^\dagger$.
Thus we obtain that
\begin{equation}
\Lambda_A(\cdot)={1\over
r}\Theta((\varrho_A^T)^{-1/2}(\cdot)(\varrho_A^T)^{-1/2}).
\label{odzysk}
\end{equation}
Note that if only $\varrho_A$ is not maximally mixed then both $\Gamma$ and
$\Theta$ are not trace-preserving. Nevertheless $\Lambda$ is trace-preserving
so that it constitutes a channel.
Of course one can use the other reduction of the state $\varrho$ to obtain a
channel (call it $\Lambda_B$). Then one can get the following formula
\begin{equation}
\Lambda_B(\cdot)={1\over r}
\Theta^T((\varrho_B^T)^{-1/2}(\cdot) (\varrho_B^T)^{-1/2}).
\end{equation}
Now, the state $\varrho$ emerges if (i) Alice send to Bob some pure state $\psi$
of both reductions equal to $\varrho_A$ via the channel $\Lambda_A$, or (ii)
Bob sends to Alice the pure state of reductions $\varrho_B$
through the channel $\Lambda_B$.
\section{Examples}
A simple way of recovering the maps from a given state
via formula (\ref{map-state}) is to use the eigenbasis of the
state. Namely, if
\begin{equation}
\varrho_{m,n}=\sum_ip_i|\psi_i\rangle\langle \psi_i|
\end{equation}
with
\begin{equation}
\psi_i=\sum_{j=1}^m\sum_{k=1}^n c^i_{j,k}|e_i\rangle|f_i\rangle,
\end{equation}
then the associated map $\Theta_m^n$ is given by
\begin{equation}
\Theta(\cdot)=\sum_ip_iV_i(\cdot)V_i^\dagger
\end{equation}
with $\langle e_j|V_i|f_k\rangle=mc^i_{j,k}$. If it is hard to find the
eigenbasis, one can use any decomposition of the BE state into pure ones.
{\it Example 1.}
In the paper Ref. \cite{bound} we modified the St\o{}rmer \cite{Stormer}
matrix to obtain the following family of two-qutrit BE states
\begin{equation}
\sigma_{\alpha}=\frac{2}{7}P_+^3
+\frac{\alpha}{7} \sigma_{+}+
\frac{5-\alpha}{7} \sigma_{-},
\label{target}
\end{equation}
where $3<\alpha\leq4$ and
\begin{eqnarray}
\sigma_{+}={1 \over 3}(|0\rangle|1\rangle \langle 0| \langle 1|
+ |1\rangle|2\rangle \langle 1| \langle 2|+
|2\rangle|0\rangle \langle 2| \langle 0|), \nonumber \\
\sigma_{-}={1 \over 3}(|1\rangle|0\rangle \langle 1| \langle 0|
+ |2\rangle|1\rangle \langle 2| \langle 1|+
|0\rangle|2\rangle \langle 0| \langle 2|).
\label{mix}
\end{eqnarray}
The above state has both reductions maximally mixed, so that we could consider
two channels ($\varrho=(I\otimes \Lambda_1)P_+$ and
$\varrho=(\Lambda_2\otimes I)P_+$). However, due to symmetry of the state, the
two cases give raise to the same family of channels, given by
\begin{eqnarray}
&&\Lambda(\cdot)= {2\over 7} (\cdot)+ {\alpha\over 7} \sum_{k=1}^3 P_{k\oplus 1
k}(\cdot)P_{kk\oplus1}+\nonumber\\
&&{5-\alpha\over 7} \sum_{k=1}^3 P_{k\ominus1
k}(\cdot)P_{kk\ominus1}
\end{eqnarray}
where $P_{ij}=|i\rangle\langle j|$; $\oplus$ and $\ominus$ denote + and $-$
modulo 3 respectively.
{\it Example 2.}
This example will be based on the two-qutrit BE state \cite{Pawel} of the
following form
\begin{equation}
\varrho={1 \over 8a + 1}
\left[ \begin{array}{ccccccccc}
a &0&0&0&a&0&0&0& a \\
0&a&0&0&0&0&0&0&0 \\
0&0&a&0&0&0&0&0&0 \\
0&0&0&a&0&0&0&0&0 \\
a &0&0&0&a&0&0&0& a \\
0&0&0&0&0&a&0&0&0 \\
0&0&0&0&0&0&{1+a \over 2}&0&{\sqrt{1-a^2} \over 2}\\
0&0&0&0&0&0&0&a&0 \\
a &0&0&0&a&0&{\sqrt{1-a^2} \over 2}&0&{1+a \over 2}\\
\end{array}
\right ], \ \ \
\end{equation}
where $0<a<1$. The reduction $\varrho_A$ of the state is given by
\begin{equation}
\varrho_A={1\over 8a+1}\left[\begin{array}{ccc}
3a&0&0\\
0&3a&0\\
0&0&{1+2a}
\end{array}\right]
\end{equation}
hence it is not maximally mixed.
Now, to recover the channel we can apply the formula (\ref{odzysk}).
The map $\Gamma$ is given by
\begin{eqnarray}
&&\Gamma(\cdot)= {a\over 8a+1} \left(3(\cdot) + P_{12} (\cdot) P_{21} +
P_{13} (\cdot) P_{31} \phantom{1\over a}
+ P_{21} (\cdot) P_{12}+\right.\nonumber\\
&& \left. \phantom{1\over a}
P_{23} (\cdot) P_{32}
+ P_{32} (\cdot) P_{23}\right) + {1\over 8a+1} W (\cdot) W^\dagger
\end{eqnarray}
where $W=\sqrt{1+a\over2}P_{31}+
\sqrt{1-a\over 2}P_{33}$.
Since $\varrho_A$ is diagonal we obtain $\Gamma_A^T=\Gamma_A$
Then the final form of the channel $\Lambda_A$ is given by
\begin{eqnarray}
&&\Lambda_A (\cdot)={a\over3} \left(3V(\cdot)V +
{1\over 3a}(P_{12} (\cdot) P_{21} +
P_{32} (\cdot) P_{23} \phantom{1\over a} \right.\nonumber\\
&&+ P_{21} (\cdot) P_{12})+ \left. {1\over 2a+1}(P_{13} (\cdot) P_{31}
+ P_{23} (\cdot) P_{32})\right) + \tilde W (\cdot) \tilde W^\dagger
\end{eqnarray}
where
$V=diag[1/\sqrt{3a}, 1/\sqrt{3a}, 1/\sqrt{2a+1}]$
and $\tilde W=\sqrt{1+a\over6a}P_{31}+\sqrt{1-a\over 2(2a+1)}P_{33}$.
\section{Discussion}
Let us now discuss some possible directions of further
investigation of the binding entnaglement channels. The main goal will be
to find how the BE channels could be useful for quantum communication. The
hint is given by the effect of activation of bound entanglement \cite{aktyw},
where a large amount of BE systems considerably raised the possibilities of
a single FE system. In Ref. \cite{aktyw} we rose a question, whether
the channels associated with the BE states (which, due to theorem, are BE
channels) could exhibit nonadditivity in the following sense. If we have a
channel of some nonzero capacity $Q$, and a BE channel, then by using the
channels jointly, one expects to obtain total capacity greater than $Q$.
Another question arises, if we consider the BE channel as public one (cf.
\cite{aktyw}). This changes the paradigm of entanglement manipulations,
where so far, only classical communication was public. The question is:
what is capacity of some quantum channel of nonzero standard capacity (either
with or without classical comunication) if supplemented with public
BE channel?
It was natural to expect that the capacity of the supported
channel could be strictly greater, especially, because,
as reported in Ref. \cite{upb}, the BE states can have
surprisingly large entanglement of formation ($E_f$). The two-qutrit states
provided in Ref. \cite{upb} have $E_f\simeq 0.2$ of
entanglement of formation while the maximally entangled state of
two-qutrits has $E_f\simeq 1.5$. Now we would like to ask the following
question: may be, {\it the channel supported by public BE channel have
maximal capacity, determined only by the Hilbert space of the sent systems}?
This could be called the effect of cristallization of bound entanglement.
The conjecture is not unreasonable: we have, in fact, infinite amount of
bound entanglement at our disposal.
\begin{appendix}
\vskip5mm
\centerline{\large Appendix}
\vskip2mm
Here we prove the following lemma:
{\bf Lemma.} The support of any state is included in the product of the
supports of the reductions of the state
\begin{equation}
{\rm supp} \varrho\subseteq {\rm supp}
\varrho_A\otimes {\rm supp} \varrho_B
\end{equation}
{\bf Proof.}
Let us first prove the lemma for pure state $|\psi\rangle\langle\psi|$.
Writing the state in the Schmidt decomposition we see that it is a
superposition of the products of the states belonging to supports of
the reductions, so that the thesis of the lemma holds. Now, for
the mixed state
$\varrho=\sum_i|\psi_i\rangle\langle\psi_i|$
we have
\begin{equation}
{\rm supp} \varrho={\rm span} \{\psi_i\}_i
\end{equation}
and
\begin{equation}
{\rm supp} \varrho_A \otimes {\rm supp} \varrho_B=
{\rm span} \{{\rm supp} \varrho^i_A \otimes {\rm supp} \varrho_B^i\}_i,
\end{equation}
where $\varrho_{A,B}^i$ are the reductions of the states $\psi_i$.
Hence we obtain the required inclusion.
\end{appendix}
|
\section{Introduction}
In this Letter we propose a fundamental test for experimentally
discriminating between various classes of theoretical models for
sonoluminescence. It is well known that the optical photons measured
in sonoluminescence are characterized by a broadband spectrum, often
described as approximately thermal with a ``temperature'' of several
tens of thousands of Kelvin~\cite{Physics-Reports}. Whether or not
this ``temperature'' represents an actual thermal ensemble is less
than clear. For instance, according to the ``shock wave approach'' of
Barber, Putterman {\em et al.}, or the ``adiabatic heating
hypothesis'', thermality of the spectrum is due to a high physical
temperature caused by compression of the gases contained in the
bubble. On the other hand, in models based on variants of Schwinger's
``dynamical Casimir approach''~\cite{Schwinger,Laeff,Los-Alamos}, it
is possible to avoid reaching high physical temperatures and yet to
obtain a thermal spectrum (or at least pseudo-thermal characteristics
for the emitted photons) because of the peculiar statistical
properties of the two-photon squeezed-states produced by this class of
mechanism.
We stress that thermal characteristics in single photon measurements
can be associated with {\em at least} two hypotheses: (a) real
physical thermalization of the photon environment; (b) pseudo-thermal
single photon statistics due to tracing over the unobserved member of
a photon pair that is actually produced in a two-mode squeezed state.
We shall call case (a) {\em real thermality}; while case (b) will be
denoted {\em effective thermality}. Of course, case (b) has no
relation with any concept of thermodynamic temperature, though to any
such squeezed state one may assign a (possibly mode-dependent) {\em
effective temperature}.
Our aim is to find a class of measurements able to discriminate
between cases (a) and (b), and to understand the origin of the roughly
thermal spectrum for sonoluminescence in the visible frequency
range. In principle, the thermal character of the experimental
spectrum could disappear at higher frequencies, but for such
frequencies the water medium is opaque, and it is not clear how we
could detect them. (Except through heating effects.) Our key remark
is that it is not necessary to try to measure higher than visible
frequencies in order to get a definitive answer regarding thermality.
It is sufficient, at least in principle, to measure photon pair
correlations in the visible portion of the sonoluminescence spectrum.
Thus regardless of the underlying mechanism, two-photon correlation
measurements are a very useful tool for discriminating between broad
classes of theory and thereby investigating the nature of
sonoluminescence. We note that two-photon correlations have already
been proposed, for the first time in~\cite{Trentalange} and
subsequently in~\cite{HKP,SH}, as an efficient tool for measuring the
shape and the size of the emission region. It was proposed
in~\cite{Trentalange,HKP} that precise Hanbury--Brown--Twiss
interferometry measurements could in principle distinguish between
chaotic (thermal) light emerging from a hot bubble and the possible
production of coherent light via the dynamical Casimir effect.
Unfortunately in the dynamical Casimir effect photons are always
pair-produced from the vacuum in two--mode squeezed states, not in
coherent states. Pair-production via the dynamical Casimir effect
appears to imply that all the photon pairs form two--mode squeezed
states, which are very different from the coherent states analyzed
in~\cite{Trentalange,HKP,SH}.
\section{Real thermal light versus two-mode squeezed states}
The quantum optics mechanism that simulates a thermal spectrum [case
(b)] is based on two-mode squeezed-states defined by
\begin{equation}
|\zeta_{ab}\rangle =\hbox{e}^{- \zeta (a^{\dagger}
b^{\dagger} - b a)} |0_{a},0_{b} \rangle,
\end{equation}
where $\zeta$ is (for our purposes) a real parameter though more
generally it can be chosen to be complex \cite{bk}. In quantum optics
a two-mode squeezed-state is typically associated with a so called
non-degenerate parametric amplifier (one of the two photons is called
``signal'' and the other ``idler'' \cite{bk,bk2,yupo}). Consider the
operator algebra
\begin{equation}
[a,a^{\dagger}]=1=[b,b^{\dagger}],
\qquad
[a,b]=0=[a^{\dagger},b^{\dagger}],
\end{equation}
and the corresponding vacua
\begin{equation}
|0_a \rangle :\ a |0_a \rangle =0,
\qquad
|0_b \rangle :\ b |0_b \rangle =0.
\end{equation}
The two-mode vacuum is the state $| \zeta \rangle \equiv
|0(\zeta) \rangle $ annihilated by the operators
\begin{equation}
A (\zeta)=\cosh (\zeta) \; a -\sinh (\zeta) \; b^{\dagger},
\end{equation}
\begin{equation}
B (\zeta)=\cosh (\zeta) \; b-\sinh (\zeta) \; a^{\dagger}.
\end{equation}
A characteristic of two-mode squeezed states is that if we measure
only one photon and ``trace away'' the second, a thermal density
matrix is obtained \cite{bk,bk2,yupo}. Indeed, if $O_a$ represents an
observable relative to one mode (say mode ``a'') its expectation value
on the squeezed vacuum is given by
\begin{equation}
\langle\zeta_{ab}|O_a|\zeta_{ab}\rangle
=\frac{1}{\cosh^2 (\zeta)} \;
\sum_{n=0}^{\infty}
[\tanh(\zeta)]^{2 n} \; \langle n_{a}|O_a|n_{a} \rangle.
\label{E:termo}
\end{equation}
In particular, if we consider $O_a=N_a$, the number operator in mode
$a$, the above reduces to
\begin{equation}
\langle\zeta_{ab}|N_{a}|\zeta_{ab}\rangle = \sinh^2 (\zeta).
\end{equation}
These formulae have a strong formal analogy with thermofield dynamics
(TFD) \cite{takume,ume}, where a doubling of the physical Hilbert
space of states is invoked in order to be able to rewrite the usual
Gibbs (mixed state) thermal average of an observable as an expectation
value with respect to a temperature-dependent ``vacuum'' state (the
thermofield vacuum, a pure state in the doubled Hilbert
space). In the TFD approach, a trace over the unphysical (fictitious)
states of the fictitious Hilbert space gives rise to thermal averages
for physical observables, completely analogous to the averages
in equation (\ref{E:termo}) {\em except} that we must make the
following identification
\begin{equation}
\tanh(\zeta)= \exp\left(-\frac{1}{2} \frac{\hbar \omega}{k_{B} T}\right),
\end{equation}
where $\omega$ is the mode frequency and $T$ is the temperature. We
note that the above identification implies that the squeezing
parameter $\zeta$ in TFD is $\omega$-dependent in a very special way.
The formal analogy with TFD allows us to conclude that, provided
we measure only one photon mode, the two-mode squeezed-state acts as a
thermofield vacuum and the single-mode expectation values acquire a
pseudo-thermal character corresponding to a ``temperature''
$T_{\mathrm squeezing}$ related with the squeezing parameter $\zeta$
by
\cite{yupo}
\begin{equation}
k_{B}\; T_{\mathrm squeezing} =
\frac{\hbar \; \omega_i}{2 \log(\coth(\zeta))},
\end{equation}
where the index $i=a,b$ indicates the signal mode or the idler mode
respectively; note that ``signal'' and ``idler'' modes can have
different effective temperatures (in general $\omega_{signal} \neq
\omega_{idler}$)~\cite{yupo}.
\section{A toy model and sonoluminescence}
To treat sonoluminescence, we introduce a quantum field theory
characterized by an infinite set of bosonic oscillators (as in bosonic
TFD; not just two oscillators as in the case of ``signal-idler''
systems studied in quantum optics). The simple two-mode squeezed
vacuum is replaced by
\begin{eqnarray}
\label{E:general}
&&|\Omega[\zeta(k,k')]\rangle \equiv
\exp\left[
-\int d^3 k \; d^3k' \; \zeta(k,k') \;
(a_{k} b_{k'}-a^{\dagger}_{k} b^{\dagger}_{k'})\right] \Big|0\Big\rangle,
\end{eqnarray}
where the function $\zeta(k,k')$ is peaked near $k+k'=0$, and becomes
proportional to a delta function in the case of infinite volume
[$\zeta(k,k') \to \zeta(k) \delta(k+k')$] when the photons are emitted
strictly back-to-back~\cite{SL-prl,QED0,QED1,QED2}. To be concrete,
let us refer to the homogeneous dielectric model presented
in~\cite{QED1}. In this limit there is no ``mixing'' and everything
reduces to a sum of two-mode squeezed-states, where each pair of
back-to-back modes is decoupled from the other. The frequency $\omega$
is the same for each photon in the couple, in such a way that we are
sure to get the same ``temperature'' for both. The two-mode squeezed
vacuum then simplifies to
\begin{equation}
|\Omega (\zeta_k)\rangle \equiv
\exp\left[-\int d^3 k\; \zeta_k \;
(a_k a_{-k}-a^{\dagger}_k a^{\dagger}_{-k})\right] \Big|0\Big\rangle.
\end{equation}
The key to the present proposal is that, if photons are pair produced
via the dynamical Casimir effect, then they are actually produced in
some combination of these two-mode
squeezed-states~\cite{SL-prl,QED0,QED1,QED2}. In this case
$T_{\mathrm squeezing}$ is a function of both frequency and squeezing
parameter, and in general only a special ``fine tuning'' would allow
us to get the {\em same} effective temperature for all couples. If we
consider the expectation value on the state $|\Omega (\zeta_k) \rangle
$ of $N_k\equiv a^{\dagger}_k a_k$ we get
\begin{equation}
\langle \Omega(\zeta_k)|N_k| \Omega(\zeta_k)\rangle =\sinh^2 (\zeta_k),
\end{equation}
so we again find a ``thermal'' distribution for each value of $k$
with temperature
\begin{equation}
k_{B}T_k\equiv \frac{\hbar\omega_{k}}{2\;\log(\coth(\zeta_k))}.
\end{equation}
The point is that for $k \neq {\bar{k}}$ we generally get $T_k \neq
T_{\bar{k}}$ {\em unless a fine tuning condition holds}. This
condition is implicitly enforced in the definition of the
thermofield vacuum and it is possible only if we have
\begin{equation}
\coth (\zeta_k)=\hbox{e}^{ \kappa \omega_k},
\label{finet}
\end{equation}
with $\kappa$ some constant, so that the frequency dependence in
$T_k$ is canceled and the same $T_{\mathrm squeezing}$ is obtained for
all couples.
For models of sonoluminescence based on the dynamical Casimir effect
({\em i.e.} squeezing the QED vacuum) we cannot rely on a definition
to provide the fine tuning, but must perform an actual
calculation. Our model~\cite{QED1} is again a useful tool for a
quantitative analysis. We have (omitting indices for notational
simplicity; our Bogolubov transformation is diagonal) the following
relation between the squeezing parameter and the Bogolubov coefficient
$\beta$
\begin{equation}
\langle N \rangle = \sinh^2(\zeta) = |\beta|^2.
\end{equation}
By defining $\tau\equiv \pi\; t_0/ ({\textstyle
n_{\mathrm in}^2+n_{\mathrm out}^2})$, where $t_0$ is the timescale on which the refractive
index changes, one has~\cite{QED1}
\begin{eqnarray}
|\beta(\vec k_1,\vec k_2)|^2
&=&
{
\sinh^2\left(
{\textstyle |n_{\mathrm in}^2 \omega_{\mathrm in} -n_{\mathrm out}^2 \omega_{\mathrm out}|} \; \tau
\right)
\over
\sinh\left(
2\; {\textstyle n_{\mathrm in}^2 } \; \omega_{\mathrm in} \tau
\right) \;
\sinh\left(
2 \; {\textstyle n_{\mathrm out}^2} \; \omega_{\mathrm out} \tau
\right)
} \; {V\over(2\pi)^3} \; \delta^3(\vec k_1 + \vec k_2).
\label{beta-squared}
\end{eqnarray}
In the adiabatic limit (large frequencies) we get a Boltzmann
factor~\cite{QED1}
\begin{equation}
|\beta|^2
\approx
\exp\left(-4 \; \min\{n_{\mathrm in},n_{\mathrm out}\} n_{\mathrm out} \; \omega_{\mathrm out} \; \tau
\right).
\end{equation}
Since $|\beta|$ is small, $\sinh(\zeta)\approx\tanh(\zeta)$, so that
in this adiabatic limit
\begin{equation}
|\tanh(\zeta)|^2 \approx
\exp\left(-4 \; \min\{n_{\mathrm in},n_{\mathrm out}\} n_{\mathrm out} \; \omega_{\mathrm out} \; \tau
\right).
\end{equation}
Therefore
\begin{equation}
k_{B}T_{\mathrm effective} \approx
\frac{\hbar}{8\pi t_{0}}\,
\frac{n_{\mathrm in}^2+n_{\mathrm out}^2}{n_{\mathrm out} \min\{n_{\mathrm in},n_{\mathrm out}\}}.
\end{equation}
Thus for the entire adiabatic region we can assign a {\em single}
frequency-independent effective temperature, which is really a measure
of the speed with which the refractive index changes. Physically, in
sonoluminescence this observation applies only to the high-frequency
tail of the photon spectrum.
In contrast, in the low frequency region, where the bulk of the
photons emitted in sonoluminescence are to be found, the sudden
approximation holds and the spectrum is phase-space-limited (a power
law spectrum), not Planckian~\cite{QED1}. It is nevertheless still
possible to assign a {\em different} effective temperature for each
frequency.
Finite volume effects smear the momentum-space delta function so we no
longer get exactly back-to-back photons. This represents a further
problem because we have to return to the general squeezed vacuum of
equation (\ref{E:general}). It is still true that photons are emitted
in pairs, pairs that are now approximately back-to-back and of
approximately equal frequency. We can again define an effective
temperature for each photon in the couple as in the ``signal-idler''
systems of quantum optics. This effective temperature is no longer the
same for the two photons belonging to the same couple and no ``special
condition'' for getting the same temperature for all the couples
exists. Hence the analysis of these finite volume distortions is not
easy~\cite{QED2}, but the qualitative result that in any dynamic
Casimir effect model of sonoluminescence there should be strong
correlations between approximately back-to-back photons is
robust.
Indeed, if we work with a plane wave approximation for the
electromagnetic eigen-modes (this is essentially a version of the Born
approximation, modified to deal with Bogolubov coefficients instead of
scattering amplitudes) and further modify the infinite-volume model
of~\cite{QED1}, both by permitting a more general temporal profile for
the refractive index, and by cutting off the space integrations at the
surface of the bubble, then the squared Bogolubov coefficient takes
the form
\begin{eqnarray}
|\beta(\vec k_1,\vec k_2)|^2
&=&
F(k_1,k_2; n(t)) \;
\left| S\left(|\vec k_1 + \vec k_2 |\; R\right) \right|^2.
\label{beta-squared2}
\end{eqnarray}
Here $F(k_1,k_2; n(t))$ is some complicated function of the refractive
index temporal profile, which encodes all the dynamics, while
$S\left(|\vec k_1 + \vec k_2 | \;R\right)$ is a purely kinematical
factor arising from the limited spatial integration:
\[
S\left(|\vec k_1 + \vec k_2 | \;R\right) \equiv
\int_{r\leq R} d^3\vec r \; \exp\left[-i( \vec k_1 + \vec k_2 )
\cdot \vec r \right].
\]
Indeed in the infinite volume limit $|S(\vec k_1,\vec k_2)|^2 \to
[V/(2\pi)^3] \; \delta(\vec k_1+\vec k_2)$. It is now a standard
calculation to show that
\[
S\left(|\vec k_1 + \vec k_2 | \;R\right) =
{4\pi\over |\vec k_1 + \vec k_2 |^3}
\left[
\sin(|\vec k_1 + \vec k_2 | \;R) -
(|\vec k_1 + \vec k_2 | \; R) \; \cos(|\vec k_1 + \vec k_2 | \; R)
\right].
\]
So, independent of the temporal profile, kinematics will provide
characteristic angular correlations between the outgoing photons:
this result depends only on the the existence of a vacuum squeezing
effect driven by a time-dependent refractive index (which is what is
needed to make the notion of a Bogolubov coefficient meaningful in
this context).
The plane-wave approximation used to obtain this formula is
valid provided the wavelength of the photons, {\em while they are
still inside the bubble}, are small compared to the dimensions of the
bubble
\begin{equation}
\lambda_{\mathrm{inside}} \ll
R; \qquad \Rightarrow \qquad \omega \gg {c\over n\; R}.
\end{equation}
While there is still considerable disagreement about the physical size
of the bubble when light emission occurs~\cite{QED0}, and almost no
data concerning the value of the refractive index of the bubble
contents at that time, the scenario developed in~\cite{QED1,QED2} is
very promising in this regard. In particular, high frequency photons
are more likely to exhibit the back-to-back effect, and depending on
the values of $R$ and $n$ this could hold for significant portions of
the resulting emission spectrum. Experimentally, one should work at as
high a frequency as possible---at the peak close to the cutoff.
These observations lead us to the following proposal.
\section{Two-photon observables}
Define the observable
\begin{equation}
N_{ab} \equiv N_{a}-N_{b},
\end{equation}
and its variance
\begin{equation}
\Delta (N_{ab})^2=
\Delta N_{a}^{2}+\Delta N_{b}^{2}
-2 \langle N_{a} N_{b}\rangle
+2 \langle N_{a}\rangle \langle N_{b} \rangle.
\end{equation}
These number operators $N_{a},N_{b}$ are intended to be relative to
photons measured, {\em e.g.}, back to back. In the case of true
thermal light we get
\begin{equation}
\Delta N_{a}^{2} = \langle N_{a}\rangle(\langle N_{a} \rangle +1),
\end{equation}
\begin{equation}
\langle N_{a} N_{b}\rangle = \langle N_{a}\rangle \langle N_{b}\rangle,
\end{equation}
so that
\begin{equation}
\Delta (N_{ab})^2_{\mathrm thermal\ light}
=\langle N_{a}\rangle(\langle N_{a}\rangle+1)
+\langle N_{b}\rangle(\langle N_{b}\rangle+1).
\end{equation}
For a two-mode squeezed-state
\begin{equation}
\Delta (N_{ab})^2_{\mathrm two\ mode\ squeezed\ light}=0.
\end{equation}
Due to correlations, $\langle N_{a} N_{b}\rangle \neq \langle
N_{a}\rangle \langle N_{b}\rangle$. Note also, that if you measure
only a single photon in the couple, you get (as expected) a thermal
variance $\Delta N_{a}^{2} = \langle N_{a}\rangle(\langle N_{a}
\rangle +1) $. Therefore a measurement of the covariance $\Delta
(N_{ab})^2$ can be decisive in discriminating if the photons are
really physically thermal or if non classical correlations
between the photons occur \cite{bk2}. If the ``thermality'' in the
sonoluminescence spectrum is of this squeezed-mode type, we will
ultimately desire a much more detailed model of the dynamical
Casimir effect involving an interaction term that produces pairs of
photons in two-mode squeezed-states. Apart from our model~\cite{QED1}
and its finite volume generalization~\cite{QED2}, the Eberlein model
also possesses this property~\cite{Eberlein}. For this type of
squeezed-mode photon pair-production in a linear medium with
spacetime-dependent dielectric permittivity and magnetic permeability
see \cite{bibi}; for nonlinearity effects see~\cite{lomo}.
In summary: The main experimental signature for squeezed-state photons
being pair-produced in sonoluminescence is the presence of strong
spatial correlations between photons emitted back-to-back and having
the same frequency. These correlations could be measured, for example,
by back-to-back symmetrically placed detectors working in coincidence.
Finite-size effects have been shown in~\cite{QED2} to perturb only
slightly this back-to-back character of the emitted photons, in the
sense that back-to-back emission remains largely dominant.
(Additionally it has been verified that the form of the spectrum is
not violently affected.) Of course, a detailed analysis of the many
technical experimental problems (such as e.g. filtering and multi-mode
signals in the detectors) has also to be done (on these topics
see~\cite{Trentalange}), but such technical details are beyond the
scope of the current work.
\section{Discussion}
The main aims of the present Letter are to clarify the nature of the
photons produced in Casimir-based models of sonoluminescence, and to
delineate the available lines of (theoretical as well experimental)
research that should be followed in order to discriminate
Casimir-based models from thermal models, preferably without having to
understand all of the messy technical details of the condensed matter
physics taking place inside the collapsing bubble.
We have shown that ``effective thermality'' can manifest itself at
different levels. What is certainly true is that two-mode squeezed
states will exhibit, at a given fixed three-momentum, occupation
numbers which in that mode follow Bose--Einstein statistics. This can
be called ``thermality at fixed wavenumber''. In contrast, it is
sometimes possible to assign, at least for a reasonably wide range of
wavenumbers, the {\em same} temperature to all modes. This
``thermality across a range of wavenumbers'' gives rise, at least in
this range of wavenumbers, to a spectrum which is approximately
Planckian.
Our sonoluminescence model exhibits Bose--Einstein thermality but not
a truly Planckian spectrum (since the bulk of the photon emission
occurs at frequencies where the sudden approximation holds and a
common temperature for all the momenta is lacking). The spectrum is
generically a power law at low frequencies followed by a
cut-off~\cite{QED1,QED2}. Although precise measurements in the low
frequency tail of the spectrum could also (in principle) allow us to
discriminate class ``a'' models from class ``b'' models, this
possibility has to be considered strongly model-dependent. Furthermore
the spectral data available at the present time is in this regard
relatively crude: spectral analysis by itself does not seem to be an
appropriate tool for discriminating between class ``a" and class ``b"
models.
Despite this limitation we have shown that there is still the
possibility of obtaining a clear discrimination between real and
effective thermality, without relying on the detailed features of the
model, by looking at two-photon correlations. For thermal light one
should find thermal variance for photon pairs. On the other hand,
thermofield--like photons should show zero variance in appropriate
pair correlations. Moreover, our analysis points out that a key point
in discriminating, by means of photon measurements alone, between
classes of models for sonoluminescence is the mechanism of photon
production: Any form of pair-production is associated with two--mode
squeezed states and their strong quantum correlations. In contrast,
any single-photon production mechanism (thermal, partially thermal,
non-thermal) is not. In either case, two-photon correlation
measurements are potentially a very useful tool for looking into the
nature of sonoluminescence.
\acknowledgments
This research was supported by the Italian Ministry of Scientific
Research (DWS, SL, and FB), and by the US Department of Energy
(MV). MV particularly wishes to thank SISSA (Trieste, Italy) and
Victoria University (Te Whare Wananga o te Upoko o te Ika a Maui;
Wellington, New Zealand) for hospitality during various stages of this
research. FB is indebted to A.~Gatti for her very helpful remarks
about photon statistics. SL wishes to thank G.~Barton, G.~Plunien, and
R.~Sch\"utzhold~ for illuminating discussions.
\vskip -0.7 true cm
|
\section{Introduction}
Blue straggler stars (BSS) where first observed in the 1950's
(\cite{sandage53}) in the Galactic globular cluster (GGC) M3. In the
color-magnitude diagram they formed a sparsely populated sequence
extending to higher luminosities than the turn-off point of normal
hydrogen burning main-sequence stars. Superficially they looked like a
population of younger stars, more massive than the turn-off stars, in
an old star cluster. Since there is no other indication of star
formation after the burst which formed the bulk of the cluster stars,
two mechanisms for making BSS are favored. First is the merger of two
stars in a primordial binary system, where ``primordial'' refers to
binaries formed when the cluster formed. Second are collisions in
regions of very high stellar density (\cite{hills76}, \cite{fpfc92},
\cite{bss2pop}, \cite{bss95}, \cite{bailynaraa}, \cite{mh97}). These
collisional-BSS include several classes of objects: direct collisions
producing a more massive star; collisions which harden primordial
binaries until the point of merger; and binaries produced in
collisions which later merge. The dense cores of globular clusters
were obvious targets for observations required to refine our
understanding of BSS. Indeed, more than 20 years ago Hills \& Day (1976)
suggested searching the core of M80 for collisional BSS. However, only
with the advent of the Hubble Space Telescope ({\it HST}) could such
observations be made (\cite{paresce47tuc}, \cite{m15fp}, \cite{m3bss},
\cite{dsm5}, \cite{guham30}).
The BSS population, especially collisional-BSS, can serve as a
diagnostic for the dynamical evolution of GGCs. Because of
gravitational interactions between cluster stars, GGCs evolve
dynamically on time scales generally smaller than their ages. For
example, the first manifestation of dynamical process within a GGC is
that in the inner part of a GGC more massive stars (or binaries)
should settle toward the center. Beyond this, more dramatic dynamical
phases can happen during the cluster's lifetime. Stars with
velocities above the escape velocity continuously evaporate, and
phenomena such as Galactic tidal stripping remove stars from the
outer regions of the cluster and induce substantial changes in the
structure of the cluster itself. As a GGC adjusts to the loss of
stars, the cluster core must contract. Under some circumstances this
process can run away leading to a possibly catastrophic
``core-collapse.'' About 15\% of the GGC population show evidence for
this phenomenon. Binaries are thought to play a fundamental role in
the core collapse: binary-binary collisions could in fact be effective
in halting (or, more probably, delaying) the collapse of the core
avoiding infinite central density.
This time of enhanced binary interactions as the cluster fights off core
collapse could well correspond to a period of unusually large BSS
production. By the end of this phase most of the binaries in the core
will be destroyed by close encounters; the survivors will become highly
hardened (i.e., tightly bound), producing most of the additional
collisional-BSS.
\section{Observations}
To search for BSS (and other blue objects) we have used the {\it Wide
Field Planetary Camera} (WFPC2) of \hst\ to obtain ultraviolet and
visible images of the central region of the high density cluster M80
(NGC~6093). Both the high angular resolution and UV sensitivity of
\hst\ are essential to identify these UV-bright objects among the much
more luminous red giants in the cluster (\cite{m3bss}). The images
were obtained on 5--6 April 1996 (GO-5903, PI: F.R. Ferraro) with the
WFPC2 F160BW (far-UV), F255W (mid-UV), F336W ($U$), and F555W (visible
or $V$) filters. The Planetary Camera (PC, which has the highest
resolution $\sim 0\farcs{046}/{\rm pixel}$) was roughly
centered on the cluster center while the Wide Field (WF) cameras (at
lower resolution $\sim 0\farcs{1}/{\rm pixel}$) sampled the
surrounding outer regions. The BSS identifications are based on
$4\times600\,$s exposures in $U$ and $4\times300\,$s exposures in
F255W. The WFPC2 frames were processed through the standard HST-WFPC
pipeline and photometry was obtained as outlined in our study of BSS
in M3 (\cite{m3bss}). Figure~\ref{image} shows the advantages of using
UV images to search for BSS: in the center of the $V$ image the light
from the bright red giant branch (RGB) stars blends together. In the
UV image the brightest objects are horizontal branch (HB) stars and
BSS; there is little blending even at the center.
\section{Results}
The Ultraviolet Color Magnitude Diagram (CMD) in the
($m_{255},m_{255}-U$) plane for more than 13,000 stars identified in
the HST field of view, is presented in Figure~\ref{uvcmd}. The large
population of BSS defines a narrow nearly-vertical sequence spanning
$\sim 3 $ mag in $m_{255}$. They are clearly separable from the cooler
and fainter Turn-Off and subgiant branch (SGB) stars. However, as
already discussed in previous papers (see \cite{bss95}) one of the
major problem in defining homogeneous samples of BSS is the {\it
operative} definition of the faint edge of the BSS population. This is
true even in UV-CMDs (see for example \cite{m3bss}), since generally
the BSS sequence merges smoothly into the $\rm MS+TO$ region without
showing any gap or discontinuity. In selecting the BSS sample here we
have adopted the same criteria we used in M3, which was recently
observed (\cite{m3bss}) with the same technique and set-up used here.
In order to assure the same BSS limiting absolute magnitude for M80 as
we adopted in M3, we aligned the two $(m_{255},~m_{255}-U)$ CMDs,
using the bright portion of the HB as {\it normalization} region. The
shift in magnitude required to align the two CMDs is $\delta m_{255} =
1.15$. The resulting fainter boundary of the BSS sequence in M80 is
$m_{255}=20.55$. Adopting this figure M80 turns to have a
spectacularly large population of BSS---305 candidates have been found
in the WFPC2 field of view.
Ferraro \etal\ (1997a) split the M3 BSS into {\it bright} and {\it faint}
subsamples. The analogous division in M80 is at
$m_{255}=20.15$. M80 has (1) 129 bright BSS with $m_{255}<20.15$
and (2) 176 faint BSS with $20.15<m_{255}<20.55$.
The BSS region in the CMD is better shown in panel (b) of
Figure~\ref{uvcmd} where the total sample of BSS is plotted as big dots.
Note that the limiting magnitude for the faint BSS is at the ``error
envelope'' of the main sequence region on the CMD. By examining the
adjacent regions of the CMD we estimate that there at most a few MS
stars misidentified as BSS. In addition the faint BSS and bright BSS
have almost identical radial distributions while that of the MS stars is much
less centrally concentrated, similar to that of the RGB+HB stars (see
Fig.~\ref{cumdist} below). This again suggests at most a very minor
contamination of the faint BSS sample.
Table 1 lists the BSS candidates: the first column is the number, then in
columns 2--5 we report the identification number, $m_{255}$ and $U$
magnitudes and the coordinates $(X,~Y)$, respectively. The coordinates are
referred to an arbitrary system and are expressed in {\it ground-based}
pixel units (1 pixel $=0\farcs 35$), after a rotation and translation to
match the complementary ground-based observations (see below).
While not obvious from Figure~\ref{image}, Figure 3 clearly shows that
the BSS (heavy solid line) are far more concentrated towards the
cluster center than either the HB or RGB stars. (The dashed line shows
the combined distribution of the HB+RGB which are individually quite
similar). Half of the BSS population is within $8\arcsec$ from the
cluster center, compared to only $\sim 20\%$ of the HB or RGB in the
same region. The Kolmogorov-Smirnov test applied to the two
distributions shows that the probability of drawing the two
populations from the same distribution is very small, $\sim 10^{-4}$.
This result is consistent with the scenario that BSS are much more
massive population than normal HB, RGB stars. A recent direct
spectroscopic mass measured for a BSS in the core of the GGC 47~Tuc
(\cite{47tucmbss}) also indicates a higher mass
for that star.
Extensive {\it artificial star} tests have been performed to estimate the
degree of completeness of the detected BSS population. The completeness
level is $>80\%$ at the faint edge of the bright sample and $\sim 72\%$ at
the faintest magnitude limit. From these results we estimate that the
{\it true} number of BSS in M80 could be as large as $\sim 400$.
The number of BSS in M80 is huge. The previous record number was in M3
which has a population of $\sim 170$ BSS (about half of the population in
M80) in the WFPC2 field of view (\cite{m3bss}). A quantitative comparison
requires that the BSS number be normalized to account for the size of the
total population. This is done with an appropriate specific frequency:
\begin{displaymath}
{F_{\scriptscriptstyle\rm HB}^{\scriptscriptstyle\rm BSS}} = {{N_{\scriptscriptstyle\rm BSS}} \over
{N_{\scriptscriptstyle\rm HB}}}
\end{displaymath}
\noindent where $N_{\scriptscriptstyle\rm BSS}$ is the number of BSS
and $N_{\scriptscriptstyle\rm HB}$ is the number of HB stars in the
same area. This ratio can be easily computed in the UV-CMDs since the
HB population is quite bright and the sequence well defined. The
specific frequency of BSS in M80 turns to be $\sim 1$. In other
clusters with similar mass, M3, M13 and M92, which have been observed
with similar technique by our group we find substantially lower values
ranging from ${F_{\scriptscriptstyle\rm HB}^{\scriptscriptstyle\rm BSS}} \sim 0.17$ for M13 up to 0.55 and $0.67$ for M92
and M3. Moreover, considering only the field of view of the PC, the
specific frequency of BSS in M80 rises to $\sim 1.7$, i.e., the BSS
are almost twice as abundant as the HB stars.
Several other clusters have recently been surveyed with the WFPC2
covering a region comparable with that of our
observations. The somewhat less massive cluster M30 has a population
of 48 BSS and a specific frequency ${F_{\scriptscriptstyle\rm HB}^{\scriptscriptstyle\rm BSS}} =0.49$ (\cite{guham30}).
While not optimal for BSS searches, the survey of Sosin \etal\ (1997) can
give a rough indication of the central BSS population. The clusters
with the largest BSS population are NGC~6388 and NGC~2808 each with
$\sim 100$ BSS. These clusters are each about a factor 4 more massive
than M80 but still contain only a fraction ($\sim 0.3$) of the BSS
population found in M80. The corresponding specific frequencies of BSS
would be about 0.1 that of M80. Either in terms of number or specific
frequency M80 becomes the Galactic BSS record holder.
\section{Discussion}
One might speculate that the BSS in M80 are produced by an anomalously
large population of primordial binaries. If so, some of these
binaries should be detectable outside the cluster core in the form
primordial-binary-merger BSS, such as those found in the outer region
of M3 (\cite{ph94,bss2pop,m3bss}). However, recent CMDs of the outer
parts of M80 (\cite{bcsspf98,alcainom80}) give no indication for a
large {\it primeval} population comparable to that found in M3. Given
this we turn to the structural characteristics of M80 for an
explanation.
M80 is much more centrally condensed than M3, M92, and M13, a factor
that might promote the production of collisional-BSS. Can that factor
alone account for the BSS population? We suspect not, because the BSS
population in M80 is also large compared with other clusters with high
central density. For example, the central part of 47~Tuc $\log\rho_0
\sim 5.1 \alwaysmath{\,M_\odot}} \def\teff{\alwaysmath{T_{\rm eff}}{\,\rm pc}^{-3}$ compared to $5.4\alwaysmath{\,M_\odot}} \def\teff{\alwaysmath{T_{\rm eff}}{\,\rm pc}^{-3}$ for M80 and in
contrast to $3.5\alwaysmath{\,M_\odot}} \def\teff{\alwaysmath{T_{\rm eff}}{\,\rm pc}^{-3}$ for M3 (\cite{pm93}). Figure 1 of
Sosin \etal\ (1997) shows that 47 Tuc does not have a large population of
BSS---no more than 50 BSS can be counted. Likewise, NGC~2808 and
NGC~6388 have densities of $\log\rho_0 \sim 4.9~{\rm and~} 5.7
\alwaysmath{\,M_\odot}} \def\teff{\alwaysmath{T_{\rm eff}}{\,\rm pc}^{-3}$ respectively and relatively modest BSS populations.
Since high density cannot account for the large number of BSS in M80
perhaps they arise from its dynamical state. M80 has one of the
highest central densities ($\log\rho_0 \sim 5.4 \alwaysmath{\,M_\odot}} \def\teff{\alwaysmath{T_{\rm eff}}{\,\rm pc}^{-3}$) of any
GGC which has shown no previous evidence for having undergone core
collapse (\cite{djor93}). Generally GGCs are considered core-collapsed
or not depending on how well their radial distribution of stars is fit
by King Models (\cite{king66}). These models are characterized by two
parameters, the core radius, $r_c$, and the tidal radius, $r_t$, or,
alternatively, the concentration, $c= \log(r_t/r_c)$. Our data
supplemented with ground-based observations (\cite{bcsspf98}) for
$r>85\arcsec$ provides the best such test to date for M80.
To determine $r_c$ and $c$ we first determined the gravity center
$C_{\rm grav}$ following the procedure of Montegriffo \etal\ (1995). We
computed $C_{\rm grav}$ by simply averaging the $X$ and $Y$ coordinates (in
the local system) of stars lying in the PC camera, and then transforming
them to the absolute system. $C_{\rm grav}$ is located at pixel
$(503\pm5,~418\pm5)$ in our PC image; this corresponds to:
$\alpha_{\rm J2000} = 16^{\rm h}\, 17^{\rm m}\, 02\fs 29,
\delta_{J2000} = -22\arcdeg\, 58\arcmin\, 32\farcs 38$ which is $\sim 4\arcsec$
NW of the center reported in the Djorgovski (1993) compilation. The
$C_{\rm grav}$ is at pixel $(676,~647)$ in the ground-based
coordinate system used in Table~1.
The density
profile with respect to the measured gravity center $C_{\rm grav}$
is shown in
Figure~\ref{kingmods}. It was derived using the standard technique
(\cite{d88}) for all stars with $V<19.5$. A King model with the most
recent values (\cite{trageretal93}), $r_c=9\arcsec$ and $c=1.95$, does
not reproduce the observed density profile for $r<8\arcsec$, however a
King model with a smaller $r_c=6\farcs{5}$ and essentially the same
$c=2.0$ fits the data reasonably well as seen in
Figure~\ref{kingmods}.
Meylan \& Heggie (1997) warn that it can be difficult to differentiate
the dynamical (pre- in- or post-collapse) phase of a GC on the basis
of the shape of the density profile. However, they suggest, as a rule
of thumb, that {\it ``any GC with a concentration $c\sim2.0-2.5$ may
be considered as collapsed or on the verge of collapsing or just
beyond.''} Thus, while the good fit to the King model suggests that
M80 has not yet completed core-collapse, the value of $c$ is
consistent with the suggestion that M80 is on the verge of
collapse. The other piece of information we can bring to bear is the
anomalously large BSS population. Two PCC clusters have been observed
deep enough and with appropriate filters that we have a reasonable
estimate of their central BSS populations. Neither of these, 47~Tuc
(\cite{sosin-ase}) and M30 (\cite{guham30}), has a BSS frequency close
to that of M80. Thus we see that being in a PCC state can not
explain the BSS population of M80.
The most plausible hypothesis at this point is that the BSS arise from
the core collapse process. It is commonly thought that binaries play
an important role on the core collapse (\cite{hut92,mh97}) with the
formation of binaries delaying and eventually halting the
collapse. With its high central density M80 is probably trying very
hard to undergo core collapse but binaries are forming and preventing
this from happening. A large population of collision-BSS should exist
during this time and slightly beyond (until the BSS begin to die off).
This scenario is fully compatible with dynamical evolution times:
following Meylan \& Heggie (1997), without including binary
formation the entire evolution time ($t_{ce}$) of the core is
$t_{ce}\sim 16 t_{rh}(0)$ where $t_{rh}(0)$ is the initial half mass
relaxation time. Using values from Djorgovski (1993), we obtain for
M80 $t_{ce}\sim 4 \times 10^8$, which is 30 times smaller than the
cluster's age.
\section{The Evolved BSS}
With such a large population of BSS we might expect to find a
significant population of evolved BSS (E-BSS). Renzini \& Fusi Pecci
(1988) suggested searching for E-BSS during their core helium burning
phase since they should appear to be redder and brighter than {\it
normal} HB stars. Following this prescription Fusi Pecci et
al. (1992) identified a few E-BSS candidates in several clusters with
predominantly blue HBs where the likelihood of confusing E-BSS stars
with true HB or evolved HB stars was minimized. Because of the small
numbers there always the possibility that some or even most of these
candidate E-BSS were due to field contamination. Near cluster centers
field contamination should be less of a problem. In our \hst\ study of
M3 we identified a sample of E-BSS candidates (see Ferraro et al 1997)
and argued that the radial distribution of E-BSS was similar to that
of the BSS. M80 offers some advantages over M3 in searching for E-BSS:
1) it has a very blue HB so there should be less confusion between red
HB stars and E-BSS; 2) it has a larger number of BSS; 3) we have
identical photometry for M13 which has a very similar BHB to M80
coupled with a much smaller number of BSS---the E-BSS region of the
CMD of M80 should have a substantially larger number of stars than
that of M13. In Figure 3a we show a zoomed $(U,~U-V)$ CMD of the HB
region. The expected location for E-BSS has been indicated as a box;
19 E-BSS (plotted as large filled circles) lie in the box. There are
only 5 E-BSS in the same part of the CMD of M13. $V$ and $U$
magnitudes and position for the E-BSS found in M80 are listed in Table
2.
In the case of M80 it is very unlikely that the E-BSS population is due to
background field contamination. In fact, most (15) of the E-BSS have been
found in the PC field of view, while only 4 E-BSS lie in the most external
WFs. A estimate of the expected field contamination can be computed
adopting the star counts listed by Ratnatunga \& Bahcall (1985). Following
their model, $\sim 0.6$ star per square arcmin is expected in a section of
the CMD which is {\it twice} the size of the region used to isolate the
E-BSS population. (The E-BSS span less than 1 magnitude in $V$, while the
Ratnatunga \& Bahcall (1985) counts are listed for 2 mag-wide bins.) The
expected number of field stars is 0 in the PC field of view and 1.6 stars
in the global field of view of the three WF cameras. For this reason we
can reasonably conclude that the region of the CMD used to select E-BSS
candidate is essentially unaffected by field contamination.
The cumulative radial distribution of the E-BSS stars is shown
(as dotted line) in Figure 3.
The E-BSS cumulative distribution is quite similar to the BSS
distribution and significantly different from that of the HB-RGB. A
Kolmogorov-Smirnov test shows that the probability that the E-BSS and
BSS population has been extracted from the same distribution
is $\sim 67\%$ while the probability that the E-BSS and the
RGB-HB population have the same distribution is
only $\sim 1.6 \%$. This result confirms the expectation that the
E-BSS share the same distribution of the BSS and they are
both a more massive population than the bulk of the
cluster stars. It further strengthens the case that field contamination is
negligible.
Earlier studies (\cite{fpfc92,m3bss}) have suggested that the ratio of
bright BSS (b-BSS) to E-BSS is $N_{\rm b\rm - BSS}/N_{\rm E\rm - BSS} \approx
6.5$. For M80 the number of b-BSS (defined as in \cite{m3bss}) is
$N_{\rm b\rm - BSS} = 129$, and we find $N_{\rm b\rm - BSS}/N_{\rm
E\rm - BSS} = 6.8$ fully consistent with earlier
studies. Because both our BSS and E-BSS samples are so cleanly defined
the ratio of the total number of BSS to E-BSS, $N_{\rm BSS} /N_{\rm
E\rm - BSS} \sim 16$, should be useful in testing lifetimes of
BSS models.
\section{Conclusions}
The emerging scenario for BSS is complex. All GGCs which have been
properly surveyed have some BSS, so BSS must be considered as a normal
component of GGC population. BSS are found in diverse environments and
are probably formed by both merging primordial binaries and stellar
collisions. Some intermediate-low density clusters have only a few BSS
(M13) while similar clusters (M3, M92) have many more. This may arise
from the fact that the initial population of binaries in clusters like
M13 is small. The relatively large population of BSS in the exterior
of M3 (\cite{m3bss}) in contrast to the absence of BSS in the exterior
of M13 (\cite{paltrinm13}) supports the notion of very different
primordial binary populations.
The densest
cluster cores have significant but highly variable BSS
populations (see the discussion in Ferraro, Bellazzini \& Fusi Pecci 1995).
In particular the post-core-collapse clusters 47~Tuc and M30 have
significantly smaller BSS populations than M80.
We suggest that exceptional population in M80 arises because we have
caught a cluster at a critical phase in its dynamical evolution. This
effect could be enhanced by a large fraction of primordial binaries,
but there is no indication for this in the form a large BSS population
in the outer cluster (\cite{bcsspf98,alcainom80}). More
information is needed before a definitive conclusion can be reached. A
search for other indications of a high frequency of stellar
multiplicity in M80, such as a broadening of the main sequence, would
also be very useful. Also, further study of the velocity distribution
would be important to clarify the dynamical state of the cluster
(\cite{mh97}). Core collapse is one of the most spectacular phenomena
in nature. It is important to confirm whether we have caught M80
during the period when the stellar interactions are delaying the
collapse of the core (and producing BSS).
\acknowledgments
This research was partially financed by the Agenzia
Spaziale Italiana (ASI). FRF acknowledges the MURST financial support to
the project {\it Stellar Evolution} and the {\it ESO Visitor Program}
for its hospitality.
RTR \& BD are supported in part by the NASA Long Term Space Astrophysics
grant NAG 5-6403 and STScI/NASA grants GO-6607, 6804.
|
\section{Introduction and Statement of Results}
Let $k$ be a field and $A$ an associative unital $k$-algebra. We
write $\cat{Mod} A$ for the category of left $A$-modules, and
$\cat{D}^{\mrm{b}}(\cat{Mod} A)$ for the bounded derived category.
Let $A^{\circ}$ be the opposite
algebra and $A^{\mrm{e}} := A \otimes_{k} A^{\circ}$
the enveloping algebra, so that
$\cat{Mod} A^{\mrm{e}}$ is the category of $k$-central
$A$-$A$-bimodules.
A {\em two-sided tilting complex} a complex
$T \in \cat{D}^{\mrm{b}}(\cat{Mod} A^{\mrm{e}})$
for which there exists another complex
$T^{\vee} \in \cat{D}^{\mrm{b}}(\cat{Mod} A^{\mrm{e}})$
satisfying
$T^{\vee} \otimes^{\mrm{L}}_{A} T \cong
T \otimes^{\mrm{L}}_{A} T^{\vee} \cong A$.
This notion is due to Rickard \cite{Rd}.
The {\em derived Picard group} of $A$ \tup{(}relative to
$k$\tup{)} is
\[ \operatorname{DPic}_{k}(A) :=
\frac{ \{ \text{two-sided tilting complexes}\
T \in \cat{D}^{\mrm{b}}(\cat{Mod} A^{\mrm{e}}) \} }{
\text{isomorphism} } \]
with identity element $A$, product
$(T_{1}, T_{2}) \mapsto T_{1} \otimes_{A}^{\mrm{L}} T_{2}$
and inverse
$T \mapsto T^{\vee} := \operatorname{R} \operatorname{Hom}_{A}(T, A)$.
See \cite{Ye} for more details.
Since every invertible bimodule is a two-sided tilting complex,
$\operatorname{DPic}_{k}(A)$ contains the (noncommutative) Picard
group $\operatorname{Pic}_{k}(A)$ as a subgroup. It also contains a
central subgroup $\bra{\sigma} \cong \mbb{Z}$, where $\sigma$ is
the class of the two-sided tilting complex $A[1]$. In \cite{Ye}
we showed that when $A$ is either local or commutative one has
$\operatorname{DPic}_{k}(A) = \operatorname{Pic}_{k}(A) \times \bra{\sigma}$.
This was discovered independently by Rouquier-Zimmermann
\cite{Zi}, \cite{RZ}.
On the other hand in the smallest example of a $k$-algebra $A$
that is neither commutative nor local, namely the $2 \times 2$ upper
triangular matrix algebra, this equality fails. These observations
suggest that the group structure of $\operatorname{DPic}_{k}(A)$ should
carry some information about the geometry of the noncommutative
ring $A$.
This prediction is further motivated by another result in
\cite{Ye}, which says that $\operatorname{DPic}_{k}(A)$ classifies the
dualizing complexes over $A$. The geometric significance of
dualizing complexes is well known (cf.\ \cite{RD} and \cite{YZ}).
From a broader perspective, $\operatorname{DPic}_{k}(A)$ is related to the
geometry of noncommutative schemes on the one hand, and to mirror
symmetry and deformations of (commutative) smooth projective
varieties on the other hand. See \cite{BO}, \cite{Ko}, \cite{KR} and
\cite{Or}.
A good starting point for the study of the group $\operatorname{DPic}_{k}(A)$
is to consider {\em finite dimensional} $k$-algebras. The geometric
object associated to a finite dimensional $k$-algebra $A$ is its
quiver $\vec{\Delta}$, as defined by Gabriel (cf.\ \cite{GR} or
\cite{ARS}). It is worthwhile to note that from the point of view of
noncommutative localization theory (cf.\ \cite{MR} Section 4.3)
$\vec{\Delta}$ is the link graph of $A$. More on this in
Remark \ref{rem1.1}.
Some calculations of the groups $\operatorname{DPic}_{k}(A)$ for finite
dimensional algebras have already been done. Let us mention the
work of Rouquier-Zimmermann \cite{RZ} on Brauer tree algebras, and
the work of Lenzing-Meltzer \cite{LM} on canonical algebras.
In this paper we present a systematic study the group
$\operatorname{DPic}_{k}(A)$ when $A$ is a finite dimensional {\em hereditary}
algebra over an {\em algebraically closed} field $k$.
We obtain general results on the structure of $\operatorname{DPic}_{k}(A)$,
as well as explicit calculations. These results carry over to piecewise
hereditary algebras, as well as to certain noncommutative schemes.
The rest of the Introduction is devoted to stating our main results.
The group $\operatorname{Aut}_{k}(A) = \operatorname{Aut}_{\cat{Alg} k}(A)$
of $k$-algebra automorphisms is a linear algebraic group over $k$,
via the inclusion into
$\operatorname{Aut}_{\cat{Mod} k}(A) = \operatorname{GL}(A)$.
This induces a structure of linear algebraic group on the quotient
$\operatorname{Out}_{k}(A)$ of outer automorphisms. We denote by
$\operatorname{Out}^{0}_{k}(A)$ the identity component of
$\operatorname{Out}_{k}(A)$.
Recall that $A$ is a {\em basic} $k$-algebra if
$A / \mfrak{r} \cong k \times \cdots \times k$,
where $\mfrak{r}$ is the Jacobson radical. For a basic algebra one
has $\operatorname{Out}_{k}(A) = \operatorname{Pic}_{k}(A)$.
A hereditary basic algebra $A$ is isomorphic to the {\em path
algebra} $k \vec{\Delta}$ of its quiver.
An algebra $A$ is {\em indecomposable} iff the quiver $\vec{\Delta}$
is connected.
For Morita equivalent $k$-algebras $A$ and $B$ one has
$\operatorname{DPic}_{k}(A) \cong \operatorname{DPic}_{k}(B)$, and the quivers of $A$
and $B$ are isomorphic. According to a result of Brauer (see
\cite{Po} Section 2) one has
$\operatorname{Out}^{0}_{k}(A) \cong \operatorname{Out}^{0}_{k}(B)$.
If $A \cong \prod_{i = 1}^{n} A_{i}$ then
$\operatorname{DPic}_{k}(A) \cong
G \ltimes \prod_{i = 1}^{n} \operatorname{DPic}_{k}(A_{i})$,
where $G \subset S_{n}$ is a permutation group (cf.\ \cite{Ye}
Lemma 2.6). Also
$\vec{\Delta}(A) \cong \coprod \vec{\Delta}(A_{i})$
and
$\operatorname{Out}^{0}_{k}(A) \cong \prod \operatorname{Out}^{0}_{k}(A_{i})$.
Since the main result Theorem \ref{thm0.2} is stated
in terms of $\vec{\Delta}$ and $\operatorname{Out}^{0}_{k}(A)$,
we allow ourselves to assume throughout that $A$ is a basic
indecomposable algebra.
Given a quiver $\vec{Q}$ we denote by $\vec{Q}_{0}$ its vertex set.
For a pair of vertices $x, y \in \vec{Q}_{0}$ we write $d(x, y)$
for the arrow-multiplicity, i.e.\ the number of arrows
$\alpha: x \to y$. Let $\operatorname{Aut}(\vec{Q}_{0})$ be the permutation
group of $\vec{Q}_{0}$, and let
$\operatorname{Aut}(\vec{Q}_{0}; d)$ be the subgroup of arrow-multiplicity
preserving permutations, namely
\[ \operatorname{Aut}(\vec{Q}_{0}; d) = \{ \pi \in \operatorname{Aut}(\vec{Q}_{0})
\mid d(\pi(x), \pi(y)) = d(x, y) \text{ for all } x, y \in
\vec{Q}_{0}\} . \]
Write $\operatorname{Aut}(\vec{Q})$ for the automorphism group of the quiver
$\vec{Q}$. Then $\operatorname{Aut}(\vec{Q}_{0}; d)$ is the image of the
canonical homomorphism
$\operatorname{Aut}(\vec{Q}) \to \operatorname{Aut}(\vec{Q}_{0})$.
The surjection
$\operatorname{Aut}(\vec{Q}) \twoheadrightarrow \operatorname{Aut}(\vec{Q}_{0}; d)$
is split, and it is bijective iff
$\vec{Q}$ has no multiple arrows.
Of particular importance to us is a certain countable quiver
$\vec{\Gamma}^{\mrm{irr}}$.
This is a full subquiver of the Auslander-Reiten quiver
$\vec{\Gamma}(\cat{D}^{\mrm{b}}(\cat{mod} A))$
of $\cat{D}^{\mrm{b}}(\cat{mod} A)$ as defined by Happel
\cite{Ha}. Here $\cat{mod} A$ is the category of finitely generated
$A$-modules. If $A$ has finite representation type (i.e.\
$\vec{\Delta}$ is a Dynkin quiver) then
$\vec{\Gamma}^{\mrm{irr}} \cong \vec{\mbb{Z}} \vec{\Delta}$,
where $\vec{\mbb{Z}} \vec{\Delta}$ is the quiver introduced by
Riedtmann \cite{Rn}. Otherwise
$\vec{\Gamma}^{\mrm{irr}} \cong \mbb{Z} \times \vec{\mbb{Z}}
\vec{\Delta}$. See Definitions \ref{dfn2.1} and \ref{dfn2.2}
for the definition of the quivers
$\vec{\Gamma}^{\mrm{irr}}$ and $\vec{\mbb{Z}} \vec{\Delta}$, and
see Figures \ref{fig2} and \ref{fig3} for illustrations. The group
$\operatorname{DPic}_{k}(A)$ acts on $\vec{\Gamma}^{\mrm{irr}}_{0}$ by
arrow-multiplicity preserving permutations, giving rise to a group
homomorphism
$q: \operatorname{DPic}_{k}(A) \to
\operatorname{Aut}(\vec{\Gamma}^{\mrm{irr}}_{0}; d)$.
Define the bimodule $A^{*} := \operatorname{Hom}_{k}(A, k)$. Then $A^{*}$ is
a two-sided tilting complex, the functor
$M \mapsto A^{*} \otimes^{\mrm{L}}_{A} M \cong
\mrm{R} \operatorname{Hom}_{A}(M, A^{*})$
is the {\em Serre functor} of
$\cat{D}^{\mrm{b}}(\cat{mod} A)$ in the sense of \cite{BK}, and
$M \mapsto A^{*}[-1] \otimes^{\mrm{L}}_{A} M$
is the {\em translation functor} in the sense of \cite{Ha} Section
I.4. We write $\tau \in \operatorname{DPic}_{k}(A)$ for the element
represented by $A^{*}[-1]$. Then $\tau$ is translation of the quiver
$\vec{\Gamma}^{\mrm{irr}}$. Let us denote by
$\operatorname{Aut}(\vec{\Gamma}^{\mrm{irr}}_{0}; d)^{\bra{\tau, \sigma}}$
the subgroup of
$\operatorname{Aut}(\vec{\Gamma}^{\mrm{irr}}_{0}; d)$
consisting of permutations that commute with $\tau$ and $\sigma$.
Here is the main result of the paper.
\begin{thm} \label{thm0.2}
Let $A$ be an indecomposable basic hereditary finite dimensional
algebra over an algebraically closed field $k$, with quiver
$\vec{\Delta}$.
\begin{enumerate}
\item There is an exact sequence of groups
\[ 1 \to \operatorname{Out}^{0}_{k}(A) \to \operatorname{DPic}_{k}(A)
\xrightarrow{q}
\operatorname{Aut}(\vec{\Gamma}^{\mrm{irr}}_{0} ; d)^{\bra{\tau, \sigma}}
\to 1 . \]
This sequence splits.
\item If $A$ has finite representation type then there is an
isomorphism of groups
\[ \operatorname{DPic}_{k}(A) \cong
\operatorname{Aut}(\vec{\mbb{Z}} \vec{\Delta})^{\bra{\tau}} . \]
\item If $A$ has infinite representation type then
there is an isomorphism of groups
\[ \operatorname{DPic}_{k}(A) \cong
\bigl( \operatorname{Aut}((\vec{\mbb{Z}} \vec{\Delta})_{0}; d)^{\bra{\tau}}
\ltimes \operatorname{Out}^{0}_{k}(A) \bigr) \times \mbb{Z} . \]
\end{enumerate}
\end{thm}
The factor $\mbb{Z}$ of $\operatorname{DPic}_{k}(A)$ in part 3 is
generated by $\sigma$. If $\vec{\Delta}$ has no multiple arrows
then so does $\vec{\mbb{Z}} \vec{\Delta}$, and hence
$\operatorname{Aut}((\vec{\mbb{Z}} \vec{\Delta})_{0}; d) =
\operatorname{Aut}(\vec{\mbb{Z}} \vec{\Delta})$.
The proof of Theorem \ref{thm0.2} is in Section 3
where it is stated again as Theorem \ref{thm3.2}.
Recall that a finite dimensional $k$-algebra $B$ is called
{\em piecewise hereditary} of type $\vec{\Delta}$ if
$\cat{D}^{\mrm{b}}(\cat{mod} B) \approx
\cat{D}^{\mrm{b}}(\cat{mod} A)$
where $A = k \vec{\Delta}$ for some finite quiver $\vec{\Delta}$
without oriented cycles. By \cite{Rd} Corollary 3.5 one knows that
$\operatorname{DPic}_{k}(B) \cong \operatorname{DPic}_{k}(A)$. The next corollary
follows.
\begin{cor} \label{cor0.2}
Suppose $B$ is a piecewise hereditary $k$-algebra of type
$\vec{\Delta}$. Then $\operatorname{DPic}_{k}(B)$ is
described by Theorem \tup{\ref{thm0.2}} with $A = k \vec{\Delta}$.
\end{cor}
In Section 4 we work out explicit descriptions of the groups
$\operatorname{Pic}_{k}(A)$ and $\operatorname{DPic}_{k}(A)$ for the Dynkin and affine
quivers, as well as for some wild quivers with multiple arrows.
As an example we present below the explicit description of
$\operatorname{DPic}_{k}(A)$ for a Dynkin quiver of type $A_{n}$
(which corresponds to upper triangular $n \times n$ matrices). The
corollary is extracted from Theorem \ref{thm4.1}.
\begin{cor} \label{cor0.4}
Suppose $\vec{\Delta}$ is a Dynkin quiver of type $A_{n}$ and
$A = k \vec{\Delta}$. Then
$\operatorname{DPic}_{k}(A)$ is an abelian group generated by
$\tau$ and $\sigma$, with one relation
\[ \tau^{n + 1} = \sigma^{-2} . \]
\end{cor}
The relation $\tau^{n + 1} = \sigma^{-2}$ was already discovered
by E. Kreines (cf.\ \cite{Ye} Appendix).
This relation has been known also to Kontsevich, and in
his terminology $\cat{D}^{\mrm{b}}(\cat{mod} A)$
is ``fractionally Calabi-Yau of dimension $\frac{n-1}{n+1}$''
(see \cite{Ko}; note that the Serre functor is $\tau \sigma$).
Suppose $\cat{D}$ is a $k$-linear triangulated category that's
equivalent to a small category. Denote by
$\operatorname{Out}_{k}^{\mrm{tr}}(\cat{D})$ the group of $k$-linear
triangle auto-equivalences of $\cat{D}$ modulo functorial
isomorphisms. For a finite dimensional algebra $A$ one has
$\operatorname{DPic}_{k}(A) \subset
\operatorname{Out}_{k}^{\mrm{tr}}(\cat{D}^{\mrm{b}}(\cat{mod} A)) $,
with equality when $A$ is hereditary (cf.\ Corollary
\ref{cor1.5}).
In \cite{KR} Kontsevich-Rosenberg introduce the noncommutative
projective space $\mbf{NP}^{n}_{k}$, $n \geq 1$. They state that
$\cat{D}^{\mrm{b}}(\cat{Coh} \mbf{NP}^{n}_{k})$
is equivalent to
$\cat{D}^{\mrm{b}}(\cat{mod} k \vec{\Omega}_{n+1})$,
where $\vec{\Omega}_{n+1}$ is the quiver in Figure \ref{fig5},
and $\cat{Coh} \mbf{NP}^{n}_{k}$ is the category of
coherent sheaves.
By Beilinson's results in \cite{Be}, there is an equivalence
$\cat{D}^{\mrm{b}}(\cat{Coh} \mbf{P}^{1}_{k}) \approx
\cat{D}^{\mrm{b}}(\cat{mod} k \vec{\Omega}_{2})$.
Combining Theorem \ref{thm4.3} and Corollary \ref{cor1.5} we
get the next corollary.
\begin{cor}
Let $X$ be either $\mbf{NP}^{n}_{k}$ \tup{(}$n \geq 1$\tup{)} or
$\mbf{P}^{n}_{k}$ \tup{(}$n = 1$\tup{)}. Then
\[ \operatorname{Out}_{k}^{\mrm{tr}}(\cat{D}^{\mrm{b}}(\cat{Coh} X))
\cong
\mbb{Z} \times \bigl( \mbb{Z} \ltimes \operatorname{PGL}_{n+1}(k) \bigr) . \]
\end{cor}
In Section 5 we look at a tree $\Delta$ with $n$
vertices. Every orientation $\omega$ of $\Delta$ gives a quiver
$\vec{\Delta}_{\omega}$. The equivalences between the various
categories
$\cat{D}^{\mrm{b}}(\cat{mod} k \vec{\Delta}_{\omega})$
form the derived Picard groupoid
$\operatorname{DPic}_{k}(\Delta)$. The subgroupoid generated by the
two-sided tilting complexes of \cite{APR} is called the
{\em reflection groupoid} $\operatorname{Ref}(\Delta)$. We show that there
is a surjection $\operatorname{Ref}(\Delta) \twoheadrightarrow W(\Delta)$,
where $W(\Delta) \subset \operatorname{GL}_{n}(\mbb{Z})$ is the Weyl group
as in \cite{BGP}. We also prove that for any orientation $\omega$,
$\operatorname{Ref}(\Delta)(\omega,\omega) = \bra{\tau_{\omega}}$
where $\tau_{\omega} \in \operatorname{DPic}_{k}(A_{\omega})$
is the translation.
\medskip \noindent \textbf{Acknowledgments.}\
We wish to thank A. Bondal, I. Reiten and M. Van den Bergh for
very helpful conversations and correspondences. Thanks to the
referee for suggestions and improvements to the paper.
Some of the work on the paper was done during visits to MIT and
the University of Washington, and we thank the departments of
mathematics at these universities for their hospitality. The
second author was supported by the Weizmann Institute of Science
throughout most of this research.
\section{Conventions and Preliminary Results}
In this section we fix notations and conventions to be used
throughout the paper. This is needed since there are conflicting
conventions in the literature regarding quivers and path algebras.
We also prove two preliminary results.
Throughout the paper $k$ denotes a fixed algebraically closed field.
Our notation for a quiver is $\vec{Q} = (\vec{Q}_{0}, \vec{Q}_{1})$;
$\vec{Q}_{0}$ is the set of vertices, and $\vec{Q}_{1}$ is the set
of arrows. For $x, y \in \vec{Q}_{0}$, $d(x, y)$ denotes the
number of arrows $x \to y$.
In this section the letter $\msf{A}$ denotes a $k$-linear category
that's equivalent to a small full subcategory of itself (this
assumption avoids some set theoretical problems).
Let us write $\operatorname{Aut}_{k}(\msf{A})$ for the class of $k$-linear
auto-equivalences of $\msf{A}$. Then the set
\begin{equation} \label{eqn1.4}
\operatorname{Out}_{k}(\msf{A}) = \frac{\operatorname{Aut}_{k}(\msf{A})}{
\text{functorial isomorphism}}
\end{equation}
is a group.
Suppose $\msf{A}$ is a $k$-linear additive Krull-Schmidt category
(i.e.\ $\operatorname{dim}_{k} \operatorname{Hom}_{\msf{A}}(M, N)$ \linebreak
$< \infty$
and all idempotents split).
We define the quiver $\vec{\Gamma}(\msf{A})$
of $\msf{A}$ as follows: $\vec{\Gamma}_{0}(\msf{A})$ is the
set of isomorphism classes of indecomposable objects of $\msf{A}$.
For two vertices $x, y$ there are $d(x, y)$ arrows
$\alpha : x \to y$, where we choose representatives $M_{x} \in x$,
$M_{y} \in y$,
$\operatorname{Irr}(M_{x}, M_{y}) = \operatorname{rad}(M_{x}, M_{y}) /
\operatorname{rad}^{2}(M_{x}, M_{y})$
is the space of irreducible morphisms
and $d(x, y) := \operatorname{dim}_{k} \operatorname{Irr}(M_{x}, M_{y})$.
See \cite{Rl} Section 2.2 for full details.
If $\msf{A}$ is a $k$-linear category (possibly without direct sums)
we can embed it in the additive category
$\msf{A} \times \mbb{N}$, where a morphism
$(x, m) \to (y, n)$ is an $n \times m$ matrix with entries in
$\msf{A}(x, y) = \operatorname{Hom}_{\msf{A}}(x, y)$.
Of course if $\msf{A}$ is additive then
$\msf{A} \approx \msf{A} \times \mbb{N}$.
If $\msf{A} \times \mbb{N}$ is Krull-Schmidt then we shall write
$\vec{\Gamma}(\msf{A})$ for the quiver
$\vec{\Gamma}(\msf{A} \times \mbb{N})$.
Let $\vec{Q}$ be a quiver. Assume that for every vertex
$x \in \vec{Q}_{0}$ the number of arrows starting or ending at
$x$ is finite, and for every two vertices $x, y \in \vec{Q}_{0}$
there is only a finite number of oriented paths from $x$ to $y$.
Let $k \bra{\vec{Q}}$ be the {\em path category}, whose set of
objects is $\vec{Q}_{0}$, the morphisms are generated by the
identities and the arrows, and the only relations arise from
incomposability of paths. Observe that this differs from the
definition in \cite{Rl}, where the path category corresponds to
$k \bra{\vec{Q}} \times \mbb{N}$ in our notation.
The morphism spaces of $k \bra{\vec{Q}}$ are $\mbb{Z}$-graded,
where the arrows have degree $1$. If $I \subset k \bra{\vec{Q}}$
is any ideal contained in
$\operatorname{rad}^{2}_{k \bra{\vec{Q}}} = \bigoplus _{n \geq 2}
k \bra{\vec{Q}}_{n}$,
and $k \bra{\vec{Q}, I} := k \bra{\vec{Q}} / I$
is the quotient category, then the
additive category $k \bra{\vec{Q}, I} \times \mbb{N}$ is
Krull-Schmidt, and the quiver of $k \bra{\vec{Q}, I}$ is
$\vec{\Gamma}(k \bra{\vec{Q}, I}) = \vec{Q}$.
Let $A$ be a finite dimensional $k$-algebra. In representation
theory there are three equivalent ways to define the quiver
$\vec{\Delta} = \vec{\Delta}(A)$ of $A$. The set
$\vec{\Delta}_{0}$ enumerates either a complete set of primitive
orthogonal idempotents
$\{ e_{x} \}_{x \in \vec{\Delta}_{0}}$, as in \cite{ARS} Section
III.1; or it enumerates the simple $A$-modules
$\{ S_{x} \}_{x \in \vec{\Delta}_{0}}$, as in \cite{Rl} Section
2.1; or it enumerates the indecomposable projective $A$-modules
$\{ P_{x} \}_{x \in \vec{\Delta}_{0}}$, as in \cite{Rl} Section
2.4. The arrow multiplicity is in all cases
\[ d(x, y) =
\operatorname{dim}_{k} e_{x} (\mfrak{r} / \mfrak{r}^{2}) e_{y} =
\operatorname{dim}_{k} \operatorname{Ext}^{1}_{A}(S_{y}, S_{x}) =
\operatorname{dim}_{k} \operatorname{Irr}_{\cat{proj} A}(P_{x}, P_{y}) . \]
Here $\mfrak{r}$ is the Jacobson radical and $\cat{proj} A$ is the
category of finitely generated projective modules, which is
Krull-Schmidt. Observe that the third definition is just
$\vec{\Delta}(A) = \vec{\Gamma}(\cat{proj} A)$.
\begin{rem} \label{rem1.1}
The set $\vec{\Delta}_{0}$ also enumerates the prime spectrum of $A$,
$\operatorname{Spec} A \cong \{ \mfrak{p}_{x} \}_{x \in \vec{\Delta}_{0}}$.
One can show that
$\mfrak{r} / \mfrak{r}^{2} \cong
\bigoplus_{x, y \in \vec{\Delta}_{0}}
(\mfrak{p}_{x} \cap \mfrak{p}_{y}) / \mfrak{p}_{x} \mfrak{p}_{y}$
as $A$-$A$-bimodules. This implies that $d(x, y) > 0$ iff
there is a second layer link
$\mfrak{p}_{x} \rightsquigarrow \mfrak{p}_{y}$
(cf.\ \cite{MR} Section 4.3.7). Thus if we ignore multiple arrows,
the quiver $\vec{\Delta}$ is precisely the link graph of $A$.
\end{rem}
Recall that a {\em translation} $\tau$ is an injective function from
a subset of $\vec{Q}_{0}$, called the set of non-projective vertices,
to $\vec{Q}_{0}$, such that $d(\tau(y), x) = d(x, y)$.
$\vec{Q}$ is a stable translation quiver if it comes with a
translation $\tau$ such that all vertices are non-projective.
A {\em polarization} $\mu$ is an injective function defined on the
set of arrows $\beta: x \to y$ ending in non-projective vertices,
with $\mu(\beta) : \tau(y) \to x$. Cf.\ \cite{Rl} Section 2.2.
\begin{notation} \label{not1.1}
Suppose the quiver $\vec{Q}$ has a translation $\tau$ and a
polarization $\mu$. Given a non-projective vertex $y \in \vec{Q}_{0}$
let $x_{1}, \ldots, x_{m}$ be some labeling, without repetition,
of the set of vertices
$\{ x \mid \text{there is an arrow } x \to y \}$ .
Correspondingly label the arrows
$\beta_{i, j} : x_{i} \to y$
and
$\alpha_{i, j} : \tau(y) \to x_{i}$,
where $i = 1, \ldots, m$;
$j = 1, \ldots, d_{i} = d(x_{i}, y)$; and
$\alpha_{i, j} = \mu(\beta_{i, j})$.
The {\em mesh ending at} $y$ is the subquiver with vertices
$\{ \tau(y), x_{i}, y \}$ and arrows
$\{ \alpha_{i, j}, \beta_{i, j} \}$.
\end{notation}
If $\vec{Q}$ has no multiple arrows then $d_{i} = 1$ and the
picture of the mesh ending at $y$ is shown in Figure \ref{fig0}.
The {\em mesh relation} at $y$ is defined to be
\begin{equation} \label{eqn1.1}
\sum_{i = 1}^{m} \sum_{j = 1}^{d_{i}}
\beta_{i, j} \alpha_{i, j} \in
\operatorname{Hom}_{k \bra{\vec{Q}}} \bigl( \tau(y), y \bigr) .
\end{equation}
It is a homogeneous morphism of degree $2$.
\begin{figure}
\choosegraphics{
\[ \UseTips \begin{xy}
(0,0)*+@{*}="ty"*+!CR{\tau(y)},"ty"
\ar@{->}^{\alpha_{1}} (20,15)*+@{*}="x1"*+!DL{x_{1}},"x1"
\ar@{} (20,7.5)*+@{}*!C{\vdots}
\ar@{->}^{\alpha_{i}} (20,0)*+@{*}="xi"*+!DL{x_{i}},"xi"
\ar@{} (20,-7.5)*+@{}*!C{\vdots}
\ar@{->}^{\alpha_{m}} (20,-15)*+@{*}="xm"*+!UL{x_{m}},"xm"
\ar@{->}^{\beta_{1}} "x1";(40,0)*++@{*}="y"*+!CL{y},"y"
\ar@{->}^{\beta_{i}} "xi";"y"
\ar@{->}^{\beta_{m}} "xm";"y"
\end{xy} \]
}{
\includegraphics[clip]{fig0.eps}
}
\caption{The mesh ending at the vertex $y$ when $d_{i}=1$.}
\label{fig0}
\end{figure}
\begin{dfn} \label{dfn1.1}
Let $I_{\mrm{m}}$ be the mesh ideal in the category
$k \bra{\vec{Q}}$,
i.e.\ the two sided ideal generated by the mesh relations
(\ref{eqn1.1}) where $y$ runs over all non-projective vertices.
The quotient category
\[ k \bra{\vec{Q}, I_{\mrm{m}}} := k \bra{\vec{Q}} / I_{\mrm{m}}
\]
is called the {\em mesh category}.
\end{dfn}
Observe that in \cite{Rn}, \cite{Rl} and \cite{Ha}
the notation for $k \bra{\vec{Q}, I_{\mrm{m}}}$ is
$k(\vec{Q})$.
Now let $\vec{\Delta}$ be a finite quiver without oriented cycles and
$A = k \vec{\Delta}$ the path algebra. Our convention for the
multiplication in $A$ is as follows. If
$x \xrightarrow{\alpha} y$ and $y \xrightarrow{\beta} z$ are
paths in $\vec{\Delta}$, and if
$x \xrightarrow{\gamma} z$ is the concatenated path, then
$\gamma = \alpha \beta$ in $A$.
We note that the composition rule in the path category
$k \bra{\vec{\Delta}}$ is opposite to that in $A$, so that
$\bigoplus_{x, y} \operatorname{Hom}_{k \bra{\vec{\Delta}}}(x, y) =
A^{\circ}$.
For every $x \in \vec{\Delta}_{0}$ let $e_{x} \in A$ be the
corresponding idempotent, and let $P_{x} = A e_{x}$ be the
indecomposable projective $A$-module. So
$\{ P_{x} \}_{x \in \vec{\Delta}_{0}}$
is a set of representatives of the isomorphism classes of
indecomposable projective $A$-modules. Define
$\msf{P} \subset \cat{mod} A$ to be the full subcategory on the
objects $\{ P_{x} \}_{x \in \vec{\Delta}_{0}}$.
Then
$\msf{P} \times \mbb{N} \approx \cat{proj} A$
and
$\vec{\Delta} \cong \vec{\Gamma}(\msf{P}) \cong
\vec{\Gamma}(\cat{proj} A)$.
There is an equivalence of categories
$k \bra{\vec{\Delta}} \xrightarrow{\approx} \msf{P}$
that sends $x \mapsto P_{x}$, and an arrow
$\alpha : x \to y$ goes to the right multiplication
$P_{x} = A e_{x} \xrightarrow{\alpha} P_{y} = A e_{y}$.
We will identify $\msf{P}$ and $k \bra{\vec{\Delta}}$ in this way.
Recall that the automorphism group $\operatorname{Aut}_{k}(A)$ is a linear
algebraic group. Let $H$ be the closed subgroup
\[ H := \{ F \in \operatorname{Aut}_{k}(A) \mid F(e_{x}) = e_{x}
\text{ for all } x \in \vec{\Delta}_{0} \} . \]
\begin{lem} \label{lem1.1}
$H$ is connected.
\end{lem}
\begin{proof}
For each pair $x, y \in \vec{\Delta}_{0}$ the $k$-vector space
$\msf{P}(x,y) := \operatorname{Hom}_{\msf{P}}(x, y) \cong
e_{x} A e_{y}$
is graded. Let $\msf{P}(x,y)_{i}$ be the homogeneous component of
degree $i$, and
\[ Y := \prod_{x, y \in \vec{\Delta}_{0}}
\Bigl( \operatorname{Aut}_{k}
\bigl( \msf{P}(x, y)_{1} \bigr) \times
\operatorname{Hom}_{k} \bigl( \msf{P}(x, y)_{1},
\msf{P}(x, y)_{\geq 2} \bigr) \Bigr) . \]
This is a connected algebraic variety. Since $A$ is generated as
$k$-algebra by the idempotents and the arrows, and the only relations
in $A$ are the monomial relations arising from incomposability of
paths, it follows that any element $F' \in Y$ extends
uniquely to a $k$-algebra automorphism $F$ of $A$ that fixes the
idempotents. Conversely any automorphism $F \in H$ restricts to an
element $F'$ of $Y$. This bijection $Y \to H$ is an isomorphism of
varieties. Hence $H$ is connected.
\end{proof}
The next result is partially proved in \cite{GS} Theorem 4.8 (they
assume $k$ has characteristic $0$).
\begin{prop} \label{prop1.1}
Let $A$ be a basic hereditary finite dimensional algebra over an
algebraically closed field $k$, with quiver $\vec{\Delta}$.
\begin{enumerate}
\item There is a split exact sequence of groups
\[ 1 \to \operatorname{Out}^{0}_{k}(A) \to \operatorname{Pic}_{k}(A) \to
\operatorname{Aut}(\vec{\Delta}_{0};d) \to 1 . \]
\item The group $\operatorname{Out}^{0}_{k}(A)$ is trivial when
$\vec{\Delta}$ is a tree.
\end{enumerate}
\end{prop}
\begin{proof}
1. Since $A$ is basic we have
$\operatorname{Out}_{k}(A) = \operatorname{Pic}_{k}(A)$.
By Morita theory we have
$\operatorname{Pic}_{k}(A) \cong \operatorname{Out}_{k}(\cat{Mod} A)$. Any
auto-equivalence of the category $\msf{P}$ extends to an
auto-equivalence of $\cat{Mod} A$ (using projective resolutions),
and this induces an isomorphism of groups
$\operatorname{Out}_{k}(\msf{P}) \stackrel{\simeq}{\rightarrow} \operatorname{Out}_{k}(\cat{Mod} A)$.
The class of auto-equivalences $\operatorname{Aut}_{k}(\msf{P})$ is actually
a group here. In fact $\operatorname{Aut}_{k}(\msf{P})$ can be identified with
the subgroup of $\operatorname{Aut}_{k}(A)$ consisting of automorphisms that
permute the set of idempotents $\{ e_{x} \} \subset A$.
Define a homomorphism of groups
$q: \operatorname{Out}_{k}(A) \to \operatorname{Aut}(\vec{\Delta}_{0};d)$
by $q(F)(x) = y$ if $F P_{x} \cong P_{y}$. Thus we get a commutative
diagram
\[ \UseTips \xymatrix@M+1ex{
H \ar@{>->}[r]
& \operatorname{Aut}_{k}(\msf{P}) \ar@{>->}[r] \ar@{->>}[d]^{f}
& \operatorname{Aut}_{k}(A) \ar@{->>}[d]^{g} \\
& \operatorname{Out}_{k}(\msf{P}) \ar[r]^{\cong}
& \operatorname{Out}_{k}(A) \ar[r]^{q} &
\operatorname{Aut}(\vec{\Delta}_{0};d) . } \]
For an element $F \in \operatorname{Aut}_{k}(\msf{P})$ we have
$F|_{ \{e_{x}\} } = q f(F)$, and hence
$\operatorname{Ker}(q) = f(H) = g(H)$.
According to Lemma \ref{lem1.1}, $H$ is connected. Because $g$ is a
morphism of varieties we see that $\operatorname{Ker}(q)$ is connected.
But the index of $\operatorname{Ker}(q)$ is finite, so we get
$\operatorname{Ker}(q) = \operatorname{Out}^{0}_{k}(A)$.
In order to split $q$ we choose any splitting of
$\operatorname{Aut}(\vec{\Delta}) \twoheadrightarrow \operatorname{Aut}(\vec{\Delta}_{0};d)$
and compose it with the homomorphism
$\operatorname{Aut}(\vec{\Delta}) \to \operatorname{Aut}_{k}(\msf{P})$.
\medskip \noindent 2.
When $\vec{\Delta}$ is a tree the group $H$ is a torus:
$H \cong \prod_{x, y \in \vec{\Delta}_{0}}
\operatorname{Aut}_{k} \bigl( \msf{P}(x, y)_{1} \bigr)$.
In fact $H$ consists entirely of inner automorphisms that are
conjugations by elements of the form
$\sum \lambda_{x} e_{x}$ with $\lambda_{x} \in k^{\times}$. Thus
$g(H) = 1$.
\end{proof}
The next theorem seems to be known to some experts, but we could not
locate any reference in the literature. Since it is needed in the
paper we have included a short proof. For a left coherent ring $A$
(e.g.\ a hereditary ring) we denote by $\cat{mod} A$ the category
of coherent $A$-modules. In the theorem $k$ could be any field.
\begin{thm} \label{thm1.4}
Suppose $A$ is a hereditary $k$-algebra. Then any $k$-linear
triangle auto-equivalence of $\cat{D}^{\mrm{b}}(\cat{mod} A)$
is standard.
\end{thm}
\begin{proof}
Let $F$ be a $k$-linear triangle auto-equivalence of
$\cat{D}^{\mrm{b}}(\cat{mod} A)$. By \cite{Rd} Corollary 3.5
there exits a two-sided tilting complex $T$ with
$T \cong F A$ in $\cat{D}^{\mrm{b}}(\cat{mod} A)$. Replacing $F$
with $(T^{\vee} \otimes^{\mrm{L}}_{A} -) F$
we may assume that $F A \cong A$. Hence
$F(\cat{mod} A) \subset \cat{mod} A$, and $F|_{\cat{mod} A}$ is an
equivalence. Classical Morita theory says that
$F|_{\cat{mod} A} \cong (P \otimes_{A} -)$ for some invertible
bimodule $P$. So replacing $F$ by $(P^{\vee} \otimes_{A} -) F$ we can
assume that there is an isomorphism
$\phi^{0}: F|_{\cat{mod} A} \cong \mbf{1}_{\cat{mod} A}$.
Now for every object $M \in \cat{D}^{\mrm{b}}(\cat{mod} A)$
we can choose an isomorphism
$M \cong \bigoplus_{i} M_{i}[-i]$ with $M_{i} \in \cat{mod} A$
(cf.\ \cite{Ha} Lemma I.5.2).
Define $\phi_{M}: F M \stackrel{\simeq}{\rightarrow} M$ to be the composition
\[ F M \cong \bigoplus (F M_{i})[-i]
\xrightarrow{\sum \phi^{0}_{M_{i}}[-i]}
\bigoplus_{i} M_{i}[-i] \cong M . \]
According to the proof of \cite{BO} Proposition A.3, for any
morphism $\alpha: M \to N$ one has
$\phi_{N} F(\alpha) = \alpha \phi_{M}$, so
$\phi: F \to \mbf{1}_{\cat{D}^{\mrm{b}}(\cat{mod} A)}$
is an isomorphism of functors.
\end{proof}
\begin{cor} \label{cor1.5}
Suppose $A$ is a hereditary $k$-algebra. Then
\[ \operatorname{DPic}_{k}(A) \cong
\operatorname{Out}_{k}^{\mrm{tr}}(\cat{D}^{\mrm{b}}(\cat{mod} A)) . \]
\end{cor}
\begin{proof}
The group homomorphism
$\operatorname{DPic}_{k}(A) \to
\operatorname{Out}_{k}^{\mrm{tr}}(\cat{D}^{\mrm{b}}(\cat{mod} A))$
is injective, say by \cite{Ye} Proposition 2.2, and it is
surjective by the theorem.
\end{proof}
\section{An Equivalence of Categories}
In this section we prove the technical result Theorem \ref{thm2.3}.
It holds for any finite dimensional hereditary $k$-algebra $A$. In
the special case of finite representation type, Theorem \ref{thm2.3}
is just \cite{Ha} Proposition I.5.6. Our result is the derived
category counterpart of \cite{Rl} Lemma 2.3.3. For notation see
Section 1 above.
We use a few facts about Auslander-Reiten triangles in
$\cat{D}^{\mrm{b}}(\cat{mod} A)$. These facts are well known to
experts in representation theory, but for the benefit of other
readers we have collected them in Theorems \ref{thm2.1} and
\ref{thm2.2}.
Let $\cat{D}$ be a $k$-linear triangulated category, which is
Krull-Schmidt (as additive category).
As in any Krull-Schmidt category, sink and source morphisms
can be defined in $\cat{D}$; cf.\ \cite{Rl} Section 2.2.
In \cite{Ha} Section I.4,
Happel defines Auslander-Reiten triangles in $\cat{D}$,
generalizing the Aus\-lander-Reiten (or almost split) sequences
in an abelian Krull-Schmidt category. A triangle
$M' \xrightarrow{g} M \xrightarrow{f} M'' \to M'[1]$
in $\cat{D}$ is an Auslander-Reiten triangle if $g$ is a source
morphism, or equivalently if $f$ is a sink morphism.
As before, we denote by $M_{x} \in \cat{D}$ an indecomposable object
in the isomorphism class $x \in \vec{\Gamma}(\cat{D})$.
Now let $\vec{\Delta}$ be a finite quiver without oriented
cycles, and $A = k \vec{\Delta}$ the path algebra.
For $M \in \cat{mod} k$ let
$M^{*} := \operatorname{Hom}_{k}(M, k)$.
Define auto-equivalences $\sigma$ and $\tau$ of
$\cat{D}^{\mrm{b}}(\cat{mod} A)$
by $\sigma M := M[1]$
and
$\tau M := \operatorname{R} \operatorname{Hom}_{A}(M, A)^{*}[-1] \cong
A^{*}[-1] \otimes^{\mrm{L}}_{A} M$.
\begin{thm}[Happel, Ringel] \label{thm2.1}
Let $A = k \vec{\Delta}$. Then the following hold.
\begin{enumerate}
\item As an additive $k$-linear category,
$\cat{D}^{\mrm{b}}(\cat{mod} A)$ is a Krull-Schmidt category.
\item The quiver
$\vec{\Gamma} := \vec{\Gamma}(\cat{D}^{\mrm{b}}(\cat{mod} A))$
is a stable translation quiver, and the translation $\tau$
satisfies $M_{\tau(x)} \cong \tau M_{x}$.
\item The Auslander-Reiten triangles in
$\cat{D}^{\mrm{b}}(\cat{mod} A)$ \tup{(}up to isomorphism\tup{)}
correspond bijectively to the meshes in $\vec{\Gamma}$. In the
notation \tup{\ref{not1.1}} with $\vec{Q} = \vec{\Gamma}$ these
triangles are
\[ M_{\tau(y)} \xrightarrow{(g_{i, j})}
\bigoplus_{i = 1}^{m} \bigoplus_{j = 1}^{d_{i}}
M_{x_{i}} \xrightarrow{(f_{i, j})^{\mrm{t}}}
M_{y} \to M_{\tau(y)}[1] . \]
\item A morphism
$(g_{i, j}) : M_{\tau(y)} \to
\bigoplus_{i = 1}^{m} \bigoplus_{j = 1}^{d_{i}} M_{x_{i}}$
is a source morphism iff
for all $i$, $\{ g_{i, j} \}_{j = 1}^{d_{i}}$ is a basis of
$\operatorname{Irr}_{\cat{D}^{\mrm{b}}(\cat{mod} A)}(M_{\tau(y)},
M_{x_{i}})$. Likewise a morphism
${(f_{i, j})^{\mrm{t}}} :
\bigoplus_{i = 1}^{m} \bigoplus_{j = 1}^{d_{i}} M_{x_{i}}
\to M_{y}$ is a sink morphism iff for all $i$,
$\{ f_{i, j} \}_{j = 1}^{d_{i}}$ is basis of
$\operatorname{Irr}_{\cat{D}^{\mrm{b}}(\cat{mod} A)}(M_{x_{i}},
M_{y})$.
\end{enumerate}
\end{thm}
\begin{proof}
1. This is implicit in \cite{Ha} Sections I.4 and I.5. In
particular \cite{Ha} Lemma I.5.2 shows that for any indecomposable
object $M \in \cat{D}^{\mrm{b}}(\cat{mod} A)$ the ring \newline
$\operatorname{End}_{\cat{D}^{\mrm{b}}(\cat{mod} A)}(M)$ is local.
\medskip \noindent
2. See \cite{Ha} Corollary I.4.9.
\medskip \noindent
3. According to \cite{Ha} Theorem I.4.6 and Lemma I.4.8, for each
$y \in \vec{\Gamma}_{0}$ there exists such an Auslander-Reiten
triangle. By \cite{Ha} Proposition I.4.3 these are all the
Auslander-Reiten triangles, up to isomorphism.
\medskip \noindent
4. Since source and sink morphism depend only on the structure of
$k$-linear additive category on $\cat{D}^{\mrm{b}}(\cat{mod} A)$
(cf.\ \cite{Ha} Section I.4.5) we may use \cite{Rl} Lemma 2.2.3.
\end{proof}
The Auslander-Reiten quiver
$\vec{\Gamma}(\cat{D}^{\mrm{b}}(\cat{mod} A))$
contains the quiver $\vec{\Delta}$, as the
full subquiver with vertices corresponding to the indecomposable
projective $A$-modules, under the inclusion
$\cat{mod} A \subset \cat{D}^{\mrm{b}}(\cat{mod} A)$.
\begin{dfn} \label{dfn2.1}
We call a connected component of
$\vec{\Gamma}(\cat{D}^{\mrm{b}}(\cat{mod} A))$
{\em irregular} if it is isomorphic to the connected component
containing $\vec{\Delta}$, and we denote by
$\vec{\Gamma}^{\mrm{irr}}$
the disjoint union of all irregular components of
$\vec{\Gamma}(\cat{D}^{\mrm{b}}(\cat{mod} A))$.
\end{dfn}
The name ``irregular'' is inspired by \cite{ARS} Section VIII.4,
where regular components of $\vec{\Gamma}(\cat{mod} A)$ are
discussed. The quiver $\vec{\Gamma}^{\mrm{irr}}$ will be of special
interest to us. It's structure is explained in Theorem \ref{thm2.2}
below. But first we need to recall the following definition due to
Riedtmann \cite{Rn},
\begin{dfn} \label{dfn2.2}
From the quiver $\vec{\Delta}$ one can construct another quiver,
denoted by $\vec{\mbb{Z}} \vec{\Delta}$. The vertex set of
$\vec{\mbb{Z}} \vec{\Delta}$ is $\mbb{Z} \times \vec{\Delta}_{0}$,
and for every arrow $x \xrightarrow{\alpha} y$ in $\vec{\Delta}$
there are arrows
$(n, x) \xrightarrow{(n, \alpha)} (n, y)$
and
$(n, y) \xrightarrow{(n, \alpha^{*})} (n + 1, x)$
in $\vec{\mbb{Z}} \vec{\Delta}$.
\end{dfn}
The function $\tau (n, x) = (n - 1, x)$
makes $\vec{\mbb{Z}} \vec{\Delta}$ into a stable translation quiver.
Observe that $\tau$ is an automorphism of
the quiver $\vec{\mbb{Z}} \vec{\Delta}$, not just of the vertex set
$(\vec{\mbb{Z}} \vec{\Delta})_{0}$.
$\vec{\mbb{Z}} \vec{\Delta}$ is equipped with a polarization $\mu$,
given by
$\mu(n + 1, \alpha) = (n, \alpha^{*})$
and
$\mu(n, \alpha^{*}) = (n, \alpha)$.
See Figures \ref{fig2} and \ref{fig3} in Section 5 for examples.
We identify $\vec{\Delta}$ with the subquiver
$\{ 0 \} \times \vec{\Delta} \subset \vec{\mbb{Z}}\vec{\Delta}$.
Next let us define a quiver
$\mbb{Z} \times (\vec{\mbb{Z}} \vec{\Delta}) :=
\coprod_{m \in \mbb{Z}} \vec{\mbb{Z}} \vec{\Delta}$;
the connected components are
$\{ m \} \times (\vec{\mbb{Z}} \vec{\Delta})$, $m \in \mbb{Z}$.
Define an automorphism $\sigma$ of
$\mbb{Z} \times (\vec{\mbb{Z}} \vec{\Delta})$
by the action $\sigma(m) = m + 1$ on the first factor.
There is a translation $\tau$ and a polarization $\mu$
of $\mbb{Z} \times (\vec{\mbb{Z}} \vec{\Delta})$
that extend those of
$\vec{\mbb{Z}} \vec{\Delta} \cong \{ 0 \} \times
(\vec{\mbb{Z}} \vec{\Delta})$
and commute with $\sigma$.
The auto-equivalences $\sigma$ and $\tau$ of
$\cat{D}^{\mrm{b}}(\cat{mod} A)$ induce commuting
permutations of $\vec{\Gamma}_{0}$, which we also denote by $\sigma$
and $\tau$ respectively.
\begin{thm}[Happel] \label{thm2.2}
\begin{enumerate}
\item If $A$ has finite representation type then there is a
unique isomorphism of quivers
\[ \rho : \vec{\mbb{Z}} \vec{\Delta} \stackrel{\simeq}{\rightarrow}
\vec{\Gamma}^{\mrm{irr}} \]
which is the identity on $\vec{\Delta}$ and commutes with $\tau$
on vertices. Furthermore
$\vec{\Gamma}^{\mrm{irr}} =
\vec{\Gamma}(\cat{D}^{\mrm{b}}(\cat{mod} A))$.
\item If $A$ has infinite representation type
then there exists an isomorphism of quivers
\[ \rho : \mbb{Z} \times (\vec{\mbb{Z}} \vec{\Delta}) \stackrel{\simeq}{\rightarrow}
\vec{\Gamma}^{\mrm{irr}} \]
which is the identity on $\vec{\Delta}$ and commutes with $\tau$
and $\sigma$ on vertices.
If $\vec{\Delta}$ is a tree then the isomorphism $\rho$
is unique.
\end{enumerate}
\end{thm}
\begin{proof}
This is essentially \cite{Ha} Proposition I.5.5 and Corollary I.5.6.
\end{proof}
Fix once and for all for every vertex
$x \in \vec{\Gamma}^{\mrm{irr}}_{0}$
an indecomposable object
$M_{x} \in \cat{D}^{\mrm{b}}(\cat{mod} A)$
which represents $x$, and such that $M_{x} = P_{x}$ for
$x \in \vec{\Delta}_{0}$.
Define $\msf{B} \subset \cat{D}^{\mrm{b}}(\cat{mod} A)$
to be the full subcategory with objects
$\{ M_{x} \mid x \in (\vec{\mbb{Z}} \vec{\Delta})_{0} \}$.
The additive category $\msf{B} \times \mbb{N}$ is also
Krull-Schmidt, so for $M_{x}, M_{y} \in \msf{B}$ the two
$k$-modules
$\operatorname{Irr}_{\msf{B} \times \mbb{N}}(M_{x}, M_{y})$
and
$\operatorname{Irr}_{\cat{D}^{\mrm{b}}(\cat{mod} A)}(M_{x}, M_{y})$
could conceivably differ (cf.\ \cite{Rl} Section 2.2). But this
is not the case as we see in the lemma below.
\begin{lem} \label{lem2.2}
Suppose $I \subset \mbb{Z}$ is a segment \tup{(}i.e.\
$I = \{i \in \mbb{Z} \mid a \leq i \leq b \}$ with
$a, b \in \mbb{Z} \cup \{\pm \infty\}$\tup{)}. Let
$\msf{B}(I) \subset \cat{D}^{\mrm{b}}(\cat{mod} A)$
be the full subcategory on the
objects
$M_{x}$, $x \in I \times \vec{\Delta}_{0} \subset
\vec{\Gamma}(\cat{D}^{\mrm{b}}(\cat{mod} A))_{0}$.
Then for any
$M_{x}, M_{y} \in \msf{B}(I)$ one has
\[ \operatorname{Irr}_{\msf{B}(I) \times \mbb{N}}(M_{x}, M_{y}) \cong
\operatorname{Irr}_{\cat{D}^{\mrm{b}}(\cat{mod} A)}(M_{x}, M_{y}) . \]
\end{lem}
\begin{proof}
Consider a sink morphism in $\cat{D}^{\mrm{b}}(\cat{mod} A)$
ending in $M_{(n, y)}$,
$(n, y) \in (\vec{\mbb{Z}} \vec{\Delta})_{0}$.
By Theorem \ref{thm2.1}(3) and Theorem \ref{thm2.2}, it is of the
form
${(f_{i, j})^{\mrm{t}}} :
\bigoplus_{i = 1}^{m} \bigoplus_{j = 1}^{d_{i}}
M_{(n - \epsilon_{i}, x_{i})}$ \linebreak
$\to M_{(n, y)}$
with $\epsilon_{i} \in \{ 0, 1 \}$ (cf.\ Notation \ref{not1.1}).
From the definition of a sink morphism we see that this is also a
sink morphism in the category $\msf{B} \times \mbb{N}$.
According to \cite{Rl} Lemma 2.2.3 (dual form), both $k$-modules
\linebreak
$\operatorname{Irr}_{\msf{B} \times \mbb{N}}(M_{(n - \epsilon_{i}, x_{i})},
M_{(n, y)})$
and
$\operatorname{Irr}_{\cat{D}^{\mrm{b}}(\cat{mod} A)}
(M_{(n - \epsilon_{i}, x_{i})}, M_{(n, y)})$
have the morphisms $f_{i, 1}, \ldots, f_{i, d_{i}}$
as basis. And there are no irreducible morphisms $N \to M_{(n, y)}$
for indecomposable objects $N$ not isomorphic to one of the
$M_{(n - \epsilon_{i}, x_{i})}$, in either category. Thus the lemma is
proved for $\msf{B}(I) = \msf{B}$.
Let $x, y \in \vec{\Delta}_{0}$ and $l, n \in \mbb{Z}$. If
$\operatorname{Hom}(M_{(l, x)}, M_{(n, y)}) \neq 0$ then necessarily
$l \leq n$. This is clear for $l = 0$, since $M_{(0, x)}$ is a
projective module, and an easy calculation shows that for $n < 0$,
\[ \mrm{H}^{0}(M_{(n, y)}) \cong
\mrm{H}^{0}(A^{*}[-1] \otimes^{\mrm{L}}_{A} \cdots
\otimes^{\mrm{L}}_{A} A^{*}[-1] \otimes^{\mrm{L}}_{A}
M_{(0, y)}) = 0 . \]
In general we can translate by $\tau^{-l}$.
Now take an arbitrary segment $I$. The paragraph above implies that
for $n, l \in I$ and $i \geq 0$,
$\operatorname{rad}^{i}_{\msf{B}(I) \times \mbb{N}}(M_{(l, x)}, M_{(n, y)}) =
\operatorname{rad}^{i}_{\msf{B} \times \mbb{N}}(M_{(l, x)}, M_{(n, y)})$.
Hence \linebreak
$\operatorname{Irr}_{\msf{B}(I) \times \mbb{N}}(M_{(l, x)}, M_{(n, y)}) =
\operatorname{Irr}_{\msf{B} \times \mbb{N}}(M_{(l, x)}, M_{(n, y)})$.
\end{proof}
Henceforth we shall simply write $\operatorname{Irr}(M_{x}, M_{y})$ when
$x, y \in (\vec{\mbb{Z}} \vec{\Delta})_{0}$.
The lemma implies that the quiver of the category $\msf{B}(I)$ is
the full subquiver
$\vec{I} \vec{\Delta} \subset \vec{\mbb{Z}} \vec{\Delta}$.
Note that for $I = \{ 0 \}$ we get $\msf{B}(I) = \msf{P}$.
Since $\msf{P}$ is canonically equivalent to
$k \bra{\vec{\Delta}}$,
there is a full faithful $k$-linear functor
$G_{0} : k \bra{\vec{\Delta}} \to \msf{B}$
such that $G_{0} x = M_{x} = P_{x}$ for every vertex
$x \in \vec{\Delta}_{0}$,
and
$\{ G_{0}(\alpha_{j}) \}_{j = 1}^{d(x, y)}$
is a basis of $\operatorname{Irr}(M_{x}, M_{y})$
for every pair of vertices $x, y$, where
$\alpha_{1}, \cdots, \alpha_{d(x, y)}$ are the arrows
$\alpha_{j} : x \to y$.
\begin{thm} \label{thm2.3}
Let $\vec{\Delta}$ be a finite quiver without oriented cycles,
$A = k \vec{\Delta}$ its path algebra,
$k \bra{\vec{\mbb{Z}} \vec{\Delta}, I_{\mrm{m}}}$
the mesh category \tup{(}Definitions \tup{\ref{dfn2.2}} and
\tup{\ref{dfn1.1})} and
$\msf{B} \subset \cat{D}^{\mrm{b}}(\cat{mod} A)$
the full subcategory on the objects
$\{ M_{x} \}_{x \in (\vec{\mbb{Z}} \vec{\Delta})_{0}}$.
Then there is a $k$-linear functor
\[ G : k \bra{\vec{\mbb{Z}} \vec{\Delta}, I_{\mrm{m}}}
\to \msf{B} \]
such that
\begin{enumerate}
\rmitem{i} $G x = M_{x}$ for each vertex
$x \in (\vec{\mbb{Z}} \vec{\Delta})_{0}$.
\rmitem{ii} $G|_{k \bra{\vec{\Delta}}} = G_{0}$.
\rmitem{iii} $G$ is full and faithful.
\end{enumerate}
Moreover, the functor $G$ is unique up to isomorphism.
\end{thm}
In other words, there is a unique equivalence $G$ extending $G_{0}$.
\begin{proof}
Let
$\vec{Q}^{+} \subset \vec{\mbb{Z}} \vec{\Delta}$
be the full subquiver with vertex set
$\{ (n, y) \mid n \geq 0 \}$.
Given a vertex $(n, y)$ in $\vec{Q}^{+}$, denote by
$p(n,y)$ the number of its predecessors, i.e.\ the number of
vertices $(m, x)$ such that there is a path
$(m, x) \to \cdots \to (n,y)$ in $\vec{Q}^{+}$.
For any $p \geq 0$ let $\vec{Q}^{+}_{p}$ be the full subquiver
with vertex set
$\{ (n, y) \mid n \geq 0,\ p(n, y) \leq p \}$.
$\vec{Q}^{+}_{p}$ is a translation quiver with polarization, and
$k \bra{\vec{Q}^{+}_{p}, I_{\mrm{m}}} \subset
k \bra{\vec{\mbb{Z}} \vec{\Delta}, I_{\mrm{m}}}$
is a full subcategory.
By recursion on $p$, we will define a functor
$G : k \bra{\vec{Q}^{+}_{p}, I_{\mrm{m}}} \to \msf{B}$
satisfying conditions (i), (ii) and
\begin{enumerate}
\rmitem{iv} Let $x, y$ be a pair of vertices and let
$\alpha_{1}, \cdots, \alpha_{d(x, y)}$ be the arrows
$\alpha_{j} : x \to y$. Then
$\{ G(\alpha_{j}) \}_{j = 1}^{d(x, y)}$
is a basis of $\operatorname{Irr}(M_{x}, M_{y})$.
\end{enumerate}
Take $p \geq 0$.
It suffices to define $G(\alpha)$ for an arrow
$\alpha$ in $\vec{Q}^{+}_{p}$. These arrows fall into
three cases, according to their end vertex $(n, y)$:
\medskip \noindent
(a) $p(n, y) < p$, in which case any arrow $\alpha$ ending in
$(n, y)$ is in
$\vec{Q}^{+}_{p - 1}$, and $G(\alpha)$ is already
defined.
\medskip \noindent
(b) $p(n, y) = p$ and $n = 0$. Any arrow $\alpha$ ending in
$(n, y)$ is in $\vec{\Delta}$, so we define
$G(\alpha) := G_{0}(\alpha)$. By Lemma \ref{lem2.2} condition (iv)
holds.
\medskip \noindent
(c) $p(n, y) = p$ and $n \geq 1$. In this case $(n, y)$ is a
non-projective vertex in $\vec{Q}^{+}_{p}$, and we
consider the mesh ending at $(n, y)$. The vertices with arrows to
$(n, y)$ are $(n - \epsilon_{i}, x_{i})$,
where $i = 1, \ldots, m$; $x_{i} \in \vec{\Delta}_{0}$
and $\epsilon_{i} = 0, 1$
(cf.\ Notation \ref{not1.1}).
Since
$p(n - 1, y) < p(n - \epsilon_{i}, x_{i}) < p$
the arrows $\alpha_{i, j}$ are all in the quiver
$\vec{Q}^{+}_{p - 1}$,
and hence $G(\alpha_{i, j})$ are defined.
According to condition (iv), Lemma \ref{lem2.2} and Theorem
\ref{thm2.1}(4) it follows that there exists an Auslander-Reiten
triangle
\begin{equation} \label{eqn2.7}
M_{(n - 1, y)} \xrightarrow{(G(\alpha_{i, j}))}
\bigoplus_{i = 1}^{m} \bigoplus_{j = 1}^{d_{i}}
M_{(n - \epsilon_{i}, x_{i})}
\xrightarrow{(f_{i, j})^{\mrm{t}}}
M_{(n, y)} \to M_{(n - 1, y)}[1]
\end{equation}
in $\cat{D}^{\mrm{b}}(\cat{mod} A)$. Define
\[ G(\beta_{i, j}) := f_{i, j} : M_{(n - \epsilon_{i}, x_{i})}
\to M_{(n, y)} . \]
Note that the mesh relation
$\sum \beta_{i, j} \alpha_{i, j}$ in $k \bra{\vec{Q}^{+}_{p}}$
is sent by $G$ to
$\sum G(\beta_{i, j}) G(\alpha_{i, j}) = 0$,
so we indeed have a functor
$G : k \bra{\vec{Q}^{+}_{p}, I_{\mrm{m}}} \to \msf{B}$.
Also, by Theorem \ref{thm2.1}(4), for any $i$ the set
$\{ G(\beta_{i, j}) \}_{j = 1}^{d_{i}}$
is a basis of
$\operatorname{Irr}(M_{(n - \epsilon_{i}, x_{i})}, M_{(n, y)})$.
Thus we obtain a functor
$G : k \bra{\vec{Q}^{+}, I_{\mrm{m}}} \to \msf{B}$.
By symmetry we construct a functor
$G : k \bra{\vec{Q}^{-}, I_{\mrm{m}}} \to \msf{B}$
for negative vertices (i.e.\ $n \leq 0$), extending $G_{0}$.
Putting the two together we obtain a functor
$G : k \bra{\vec{\mbb{Z}} \vec{\Delta}, I_{\mrm{m}}} \to \msf{B}$
satisfying conditions (i), (ii) and (iv).
Let us prove $G$ is fully faithful.
For any $n \in \mbb{Z}$ there is a full subquiver
$\vec{\mbb{Z}}_{\geq n} \vec{\Delta} \subset
\vec{\mbb{Z}} \vec{\Delta}$,
on the vertex set
$\{ (i, x) \mid i \geq n \}$.
Correspondingly there are full subcategories
$k \bra{\vec{\mbb{Z}}_{\geq n} \vec{\Delta}, I_{\mrm{m}}} \subset
k \bra{\vec{\mbb{Z}} \vec{\Delta}, I_{\mrm{m}}}$
and
$\msf{B}(\mbb{Z}_{\geq n}) \subset \msf{B}$.
It suffices to prove that
$G : k \bra{\vec{\mbb{Z}}_{\geq n} \vec{\Delta}, I_{\mrm{m}}} \to
\msf{B}(\mbb{Z}_{\geq n})$
is fully faithful. By Lemma \ref{lem2.2} the quiver of
$\msf{B}(\mbb{Z}_{\geq n})$ is $\vec{\mbb{Z}}_{\geq n} \vec{\Delta}$,
which is pre-projective. So we can use
the last two paragraphs in the proof of \cite{Rl} Lemma 2.3.3 almost
verbatim.
Finally we shall prove that $G$ is unique up to isomorphism.
Suppose
$G' : k \bra{\vec{\mbb{Z}} \vec{\Delta}, I_{\mrm{m}}} \to \msf{B}$
is another $k$-linear functor satisfying conditions (i)-(iii).
We will show there is an isomorphism
$\phi : G \stackrel{\simeq}{\rightarrow} G'$ that is the identity on $k \bra{\vec{\Delta}}$.
By recursion on $p$ we shall exhibit an isomorphism
$\phi : G|_{k \bra{\vec{Q}^{+}_{p}, I_{\mrm{m}}}}
\stackrel{\simeq}{\rightarrow} G'|_{k \bra{\vec{Q}^{+}_{p}, I_{\mrm{m}}}}$.
It suffices to consider case (c) above, so let $(n, y)$ be such a
vertex. Then, because
$G'(\alpha_{i, j}) =
\phi_{(n - \epsilon_{i}, x_{i})} G(\alpha_{i, j})
\phi_{(n - 1, y)}^{-1}$,
we have
\[ \sum_{i, j} G'(\beta_{i, j}) \phi_{(n - \epsilon_{i}, x_{i})}
G(\alpha_{i, j}) =
G' \bigl( \sum_{i, j} \beta_{i, j} \alpha_{i, j} \bigr)
\phi_{(n - 1, y)} = 0 . \]
Applying $\operatorname{Hom}(-, M_{(n, y)})$ to the triangle (\ref{eqn2.7})
we obtain a morphism
$a \in$ \linebreak
$\operatorname{End}(M_{(n, y)})$
such that
$G'(\beta_{i, j}) \phi_{(n - \epsilon_{i}, x_{i})} =
a G(\beta_{i, j})$.
Because $G'$ is faithful we see that $a \neq 0$, and since
$\operatorname{End}(M_{(n, y)}) \cong k$
it follows that $a$ is invertible. Set
$\phi_{(n, y)} := a \in \operatorname{Aut}(M_{(n, y)})$.
This yields the desired isomorphism
$\phi : G|_{k \bra{\vec{Q}^{+}_{p}, I_{\mrm{m}}}}
\stackrel{\simeq}{\rightarrow} G'|_{k \bra{\vec{Q}^{+}_{p}, I_{\mrm{m}}}}$.
By symmetry the isomorphism $\phi$ extends to $\vec{Q}^{-}$.
\end{proof}
The uniqueness of $G$ gives the next corollary.
\begin{cor} \label{cor2.9}
Let $F$ be a $k$-linear auto-equivalence of
$k \bra{\vec{\mbb{Z}} \vec{\Delta}, I_{\mrm{m}}}$
fixing all objects, and such that
$F|_{k \bra{\vec{\Delta}}} \cong \mbf{1}_{k \bra{\vec{\Delta}}}$.
Then
$F \cong \mbf{1}_{k \bra{\vec{\mbb{Z}} \vec{\Delta}, I_{\mrm{m}}}}$.
\end{cor}
\begin{rem} \label{rem2.2}
Beware that if $A$ has infinite representation type then
$k \bra{\vec{\Gamma}^{\mrm{irr}}, I_{\mrm{m}}}$ is {\em not}
equivalent to the full subcategory of
$\cat{D}^{\mrm{b}}(\cat{mod} A)$ on the objects
$\{ M_{x} \}_{x \in \vec{\Gamma}^{\mrm{irr}}_{0}}$.
This is because there are nonzero morphisms from the projective
modules (vertices in the component
$\vec{\mbb{Z}} \vec{\Delta}$) to the injective modules
(vertices in $\{ 1 \} \times \vec{\mbb{Z}} \vec{\Delta}$).
\end{rem}
\section{The Representation of $\operatorname{DPic}_{k}(A)$ on the Quiver
$\vec{\Gamma}^{\mrm{irr}}$}
This section contains the proof of the main result of the paper,
Theorem \ref{thm0.2} (restated here as Theorem \ref{thm3.2}). It is
deduced from the more technical Theorem \ref{thm3.1}.
Throughout $k$ is an algebraically closed field, $\vec{\Delta}$ is a
connected finite quiver without oriented cycles, and
$A = k \vec{\Delta}$ is the path algebra. We use the notation of
previous sections.
Recall that $A^{*} = \operatorname{Hom}_{k}(A, k)$ is a tilting complex.
We shall denote by $\tau$ the class of $A^{*}[-1]$
in $\operatorname{DPic}_{k}(A)$, and by $\sigma$ the class of $A[1]$.
We identify an element $T \in \operatorname{DPic}_{k}(A)$ and the induced
auto-equivalence $F = T \otimes^{\mrm{L}}_{A} -$ of
$\cat{D}^{\mrm{b}}(\cat{mod} A)$.
\begin{lem} \label{lem3.8}
$\tau$ and $\sigma$ are in the center of
$\operatorname{DPic}_{k}(A)$.
\end{lem}
\begin{proof}
The fact that $\sigma$ is in the center of $\operatorname{DPic}_{k}(A)$
is trivial. As for $\tau$, this follows immediately from \cite{Rd}
Proposition 5.2 (or by \cite{BO} Proposition 1.3, since
$A^{*} \otimes^{\mrm{L}}_{A} -$ is
the Serre functor of $\cat{D}^{\mrm{b}}(\cat{mod} A)$).
\end{proof}
In Definition \ref{dfn2.1} we introduced the quiver
$\vec{\Gamma}^{\mrm{irr}}$. Recall that for a vertex
$x \in \vec{\Gamma}^{\mrm{irr}}_{0}$,
$M_{x} \in \cat{D}^{\mrm{b}}(\cat{mod} A)$ is the representative
indecomposable object.
\begin{lem} \label{lem3.6}
There is a group homomorphism
\[ q : \operatorname{DPic}_{k}(A) \to
\operatorname{Aut}(\vec{\Gamma}^{\mrm{irr}}_{0}; d)^{\bra{\tau, \sigma}} \]
such that $q(F)(x) = y$ iff $F M_{x} \cong M_{y}$.
\end{lem}
\begin{proof}
Given an auto-equivalence $F$ of $\cat{D}^{\mrm{b}}(\cat{mod} A)$,
the formula $q(F)(x) = y$ iff $F M_{x} \cong M_{y}$ defines a
permutation $q(F)$ of
$\vec{\Gamma}_{0}(\cat{D}^{\mrm{b}}(\cat{mod} A))$
that preserves arrow-multiplicities. Hence it restricts to a
permutation of $\vec{\Gamma}^{\mrm{irr}}_{0}$.
By Lemma \ref{lem3.8}, $q(F)$ commutes with $\tau$ and $\sigma$.
\end{proof}
The group $\operatorname{Out}^{0}_{k}(A)$ was defined to be the identity
component of $\operatorname{Out}_{k}(A)$.
\begin{lem} \label{lem3.2}
$\operatorname{Ker}(q) = \operatorname{Out}^{0}_{k}(A)$.
\end{lem}
\begin{proof}
Let $T \in \operatorname{DPic}_{k}(A)$. By Theorem \ref{thm2.2} we know
that
$\vec{\Gamma}^{\mrm{irr}}_{0} =
\bigcup_{i, j \in \mbb{Z}} \tau^{i} \sigma^{j}(\vec{\Delta}_{0})$.
Hence by Lemma \ref{lem3.8}, $T \in \operatorname{Ker}(q)$ iff $T$ acts
trivially on the set $\vec{\Delta}_{0}$. In particular we see that
$\operatorname{Ker}(q) \subset \operatorname{Pic}_{k}(A)$. Now use Proposition
\ref{prop1.1}.
\end{proof}
\begin{lem} \label{lem3.11}
Suppose $A$ has finite representation type. Then
$\sigma$ is in the center of the group
$\operatorname{Aut}(\vec{\Gamma}^{\mrm{irr}})^{\bra{\tau}}$.
\end{lem}
\begin{proof}
According to \cite{Rn} Section 2, the
group $\operatorname{Aut}(\vec{\mbb{Z}} \vec{\Delta})^{\bra{\tau}}$
is abelian in all cases except $D_{4}$. But a direct calculation
in this case (cf.\ Theorem \ref{thm4.1})
gives $\sigma = \tau^{-3}$.
\end{proof}
Before we can talk about the mesh category
$k \bra{\vec{\Gamma}^{\mrm{irr}}, I_{\mrm{m}}}$
of the quiver
$\vec{\Gamma}^{\mrm{irr}}$, we have to fix a polarization $\mu$ on
it. If the quiver $\vec{\Delta}$ has no multiple arrows then so
does $\vec{\Gamma}^{\mrm{irr}}$ (by Theorem \ref{thm2.2}), and hence
there is a unique polarization on it. If $\vec{\Delta}$ isn't a tree
let us choose an isomorphism
$\rho : \mbb{Z} \times (\vec{\mbb{Z}} \vec{\Delta}) \stackrel{\simeq}{\rightarrow}
\vec{\Gamma}^{\mrm{irr}}$
as in that theorem. This determines a polarization $\mu$ on
$\vec{\Gamma}^{\mrm{irr}}$.
We also get a lifting of the permutation $\sigma$ to an
auto-equivalence of
$k \bra{\vec{\Gamma}^{\mrm{irr}}, I_{\mrm{m}}}$.
\begin{lem} \label{lem3.12}
There are group homomorphisms
\[ p : \operatorname{Out}_{k}(k \bra{\vec{\Gamma}^{\mrm{irr}}, I_{\mrm{m}}})
\to \operatorname{Aut}(\vec{\Gamma}^{\mrm{irr}}_{0}; d)^{\bra{\tau}} \]
and
\[ r : \operatorname{Aut}(\vec{\Gamma}^{\mrm{irr}}_{0}; d)^{\bra{\tau}}
\to
\operatorname{Out}_{k}(k \bra{\vec{\Gamma}^{\mrm{irr}}, I_{\mrm{m}}}) \]
satisfying $p(F)(x) = F x$ for an auto-equivalence $F$ and a
vertex $x$; $p r = 1$; and both $p$ and $r$ commute with $\sigma$.
\end{lem}
\begin{proof}
Since
$\vec{\Gamma}^{\mrm{irr}} \cong
\vec{\Gamma}(k \bra{\vec{\Gamma}^{\mrm{irr}}, I_{\mrm{m}}})$
we get a permutation
$p(F) \in \operatorname{Aut}(\vec{\Gamma}^{\mrm{irr}}_{0}; d)$.
Let's prove that $p(F)$ commutes with $\tau$ in
$\operatorname{Aut}(\vec{\Gamma}^{\mrm{irr}}_{0})$. Consider a
vertex $y \in \vec{\Gamma}^{\mrm{irr}}_{0}$. In the Notation
\ref{not1.1}, there are vertices $x_{i}$ and irreducible morphisms
$\{ F(\alpha_{i, j}) \}_{j=1}^{d_{i}}$ and
$\{ F(\beta_{i, j}) \}_{j=1}^{d_{i}}$
that form bases of
$\operatorname{Irr}_{k \bra{\vec{\Gamma}^{\mrm{irr}}, I_{\mrm{m}}}}
(F \tau y, F x_{i})$
and
$\operatorname{Irr}_{k \bra{\vec{\Gamma}^{\mrm{irr}}, I_{\mrm{m}}}}
(F x_{i}, F y)$
respectively. Since we have
\[ \sum F(\beta_{i, j}) F(\alpha_{i, j}) = 0 \in
\operatorname{rad}_{k \bra{\vec{\Gamma}^{\mrm{irr}}, I_{\mrm{m}}}}^{2}
(F \tau y, F y) /
\operatorname{rad}_{k \bra{\vec{\Gamma}^{\mrm{irr}}, I_{\mrm{m}}}}^{3}
(F \tau y, F y) \]
this must be a multiple of a mesh relation. Hence
$F \tau y = \tau F y$.
Finally to define $r$ we have to split
$\operatorname{Aut}(\vec{\Gamma}^{\mrm{irr}}) \twoheadrightarrow
\operatorname{Aut}(\vec{\Gamma}^{\mrm{irr}}_{0}; d)$
consistently with $\mu$. It suffices to order the set of
arrows $\{ \alpha: x \to y \}$ for every pair of vertices
$x, y \in \vec{\Gamma}^{\mrm{irr}}_{0}$
consistently with $\mu$. We only have to worry about this when
$A$ has infinite representation type. For any
$x, y \in \vec{\Delta}_{0}$ choose some ordering of the set
$\{ \alpha: x \to y \}$. Using $\mu$ and $\sigma$ this ordering can
be transported to all of
$\mbb{Z} \times (\vec{\mbb{Z}} \vec{\Delta})$.
By the isomorphism $\rho$ of Theorem \ref{thm2.2} the ordering is
copied to $\vec{\Gamma}^{\mrm{irr}}$.
\end{proof}
\begin{lem} \label{lem3.5}
There exists a group homomorphism
\[ \tilde{q} : \operatorname{DPic}_{k}(A) \to
\operatorname{Out}_{k}(k \bra{\vec{\Gamma}^{\mrm{irr}}, I_{\mrm{m}}}) \]
such that
$p \tilde{q} = q$.
\end{lem}
\begin{proof}
Choose an equivalence
$G : k \bra{\vec{\mbb{Z}} \vec{\Delta}, I_{\mrm{m}}} \to \msf{B}$
as in Theorem \ref{thm2.3}. If $A$ has infinite representation
type then the isomorphism $\rho$ we have chosen (as in Theorem
\ref{thm2.2}) tells us how to extend $G$ to an equivalence
$G : k \bra{\vec{\Gamma}^{\mrm{irr}}, I_{\mrm{m}}} \to
\coprod_{l \in \mbb{Z}} \msf{B}[l]$
that commutes with $\sigma$ (cf.\ Remark \ref{rem2.2}).
Let $F$ be a triangle auto-equivalence of
$\cat{D}^{\mrm{b}}(\cat{mod} A)$.
Then $F$ induces a permutation $\pi = q(F) $ of the set
$\vec{\Gamma}^{\mrm{irr}}_{0}$
that commutes with $\sigma$. For every vertex
$x \in \vec{\Gamma}^{\mrm{irr}}_{0}$
choose an isomorphism
\[ \phi_{x} : F M_{x} \stackrel{\simeq}{\rightarrow} M_{\pi(x)} \]
in $\cat{D}^{\mrm{b}}(\cat{mod} A)$. Given an arrow
$\alpha : x \to y$ in $\vec{\Gamma}^{\mrm{irr}}$,
define the morphism
$\tilde{q}_{ \{ \phi_{x} \} }(F)(\alpha) : \pi(x) \to \pi(y)$
by the condition that the diagram
\[ \begin{CD}
F M_{x} @> F G(\alpha) >> F M_{y} \\
@V \phi_{x} VV @V \phi_{y} VV \\
M_{\pi(x)} @> G \tilde{q}_{ \{ \phi_{x} \} }(F)(\alpha) >>
M_{\pi(y)}
\end{CD} \]
commutes. Then
$\tilde{q}_{ \{ \phi_{x} \} }(F) \in
\operatorname{Aut}_{k}(k \bra{\vec{\Gamma}^{\mrm{irr}}, I_{\mrm{m}}})$.
If $\{ \phi'_{x} \}$ is another choice of isomorphisms
$\phi'_{x} : F M_{x} \stackrel{\simeq}{\rightarrow} M_{\pi(x)}$
then $\{ \phi'_{x} \phi^{-1}_{x} \}$
is an isomorphism of functors
$\tilde{q}_{ \{ \phi_{x} \} }(F) \to
\tilde{q}_{ \{ \phi'_{x} \} }(F)$,
so the map
$\tilde{q} : \operatorname{DPic}_{k}(A) \to
\operatorname{Out}_{k}(k \bra{\vec{\Gamma}^{\mrm{irr}}, I_{\mrm{m}}})$
is independent of these choices.
It is easy to check that $\tilde{q}$ respects composition of
equivalences.
\end{proof}
\begin{thm} \label{thm3.1}
Let $A$ be an indecomposable basic hereditary finite dimensional
$k$-algebra with quiver $\vec{\Delta}$. Then the homomorphism
$\tilde{q}$ of Lemma \tup{\ref{lem3.5}} induces an isomorphism
of groups
\[ \operatorname{DPic}_{k}(A) \cong
\operatorname{Out}_{k}(k \bra{\vec{\Gamma}^{\mrm{irr}}, I_{\mrm{m}}})
^{\bra{\sigma}} \cong
\begin{cases}
\operatorname{Out}_{k} \bigl( k \bra{\vec{\mbb{Z}} \vec{\Delta}, I_{\mrm{m}}}
\bigr) \quad
\parbox[t]{3cm}{\textup{if } A \textup{ has finite representation
type}} \\[6mm]
\operatorname{Out}_{k} \bigl( k \bra{\vec{\mbb{Z}} \vec{\Delta}, I_{\mrm{m}}}
\bigr) \times \bra{\sigma} \quad
\parbox[t]{2cm}{\textup{otherwise.}}
\end{cases} \]
\end{thm}
\begin{proof}
The proof has three parts.
\medskip \noindent 1. We show that the homomorphism
\[ \tilde{q} : \operatorname{DPic}_{k}(A) \to
\operatorname{Out}_{k}(k \bra{\vec{\Gamma}^{\mrm{irr}}, I_{\mrm{m}}}) \]
of Lemma \ref{lem3.5} is injective.
Let $T$ be a two-sided tilting complex such that
$\tilde{q}(T) \cong
\bsym{1}_{k \bra{\vec{\Gamma}^{\mrm{irr}}, I_{\mrm{m}}}}$.
Then the permutation $q(T)$ fixes the vertices of
$\vec{\Delta} \subset \vec{\Gamma}^{\mrm{irr}}$.
Using the fact that
$A \cong \bigoplus_{x \in \vec{\Delta}_{0}} M_{x}$
we see that
$T \cong A$ in $\cat{D}^{\mrm{b}}(\cat{mod} A)$.
Replacing $T$ with $\operatorname{H}^{0} T$ we may assume
$T$ is a single bimodule.
According to \cite{Ye} Proposition 2.2, we see that $T$ is
actually an invertible bimodule. Since
$k \bra{\vec{\Delta}} \to
k \bra{\vec{\mbb{Z}} \vec{\Delta}, I_{\mrm{m}}}$
is full we get
$\tilde{q}(T)|_{k \bra{\vec{\Delta}}} \cong
\bsym{1}_{k \bra{\vec{\Delta}}}$.
Hence by Morita theory we have $T \cong A$ as bimodules.
\medskip \noindent 2. Assume $A$ has finite representation type,
so that
$\vec{\Gamma}^{\mrm{irr}} \cong \vec{\mbb{Z}} \vec{\Delta}$.
We prove that
\[ \tilde{q} : \operatorname{DPic}_{k}(A) \to
\operatorname{Out}_{k}(k \bra{\vec{\mbb{Z}} \vec{\Delta}, I_{\mrm{m}}}) \]
is surjective.
Consider a $k$-linear auto-equivalence $F$ of
$k \bra{\vec{\mbb{Z}} \vec{\Delta}, I_{\mrm{m}}}$. Let
$\pi := p(F) \in$ \linebreak
$\operatorname{Aut}((\vec{\mbb{Z}} \vec{\Delta})_{0}; d)^{\bra{\tau}}
\cong \operatorname{Aut}(\vec{\mbb{Z}} \vec{\Delta})^{\bra{\tau}}$
as in the proof of Lemma \ref{lem3.5}. According to Lemma
\ref{lem3.11}, $\pi$ commutes with $\sigma$. Define
\[ M := \bigoplus_{x \in \vec{\Delta}_{0}} M_{\pi(0, x)}
\in \cat{D}^{\mrm{b}}(\cat{mod} A) . \]
Then for any $x, y \in \vec{\Delta}_{0}$ and integers $n, i$
the equivalence
$G : k \bra{\vec{\mbb{Z}} \vec{\Delta}, I_{\mrm{m}}} \to \msf{B}$
of Theorem \ref{thm2.3} produces isomorphisms
\[ \begin{aligned}
\operatorname{Hom}_{\cat{D}^{\mrm{b}}(\cat{mod} A)}
(M_{\pi(0, x)}, M_{(n, y)}[i])
& \cong
\operatorname{Hom}_{k \bra{\vec{\mbb{Z}} \vec{\Delta}, I_{\mrm{m}}}}
(\pi(0, x), \sigma^{i}(n, y)) \\
& \cong
\operatorname{Hom}_{k \bra{\vec{\mbb{Z}} \vec{\Delta}, I_{\mrm{m}}}}
((0, x), \sigma^{i} \pi^{-1}(n, y)) \\
& \cong
\operatorname{Hom}_{\cat{D}^{\mrm{b}}(\cat{mod} A)}
(M_{(0, x)}, M_{\pi^{-1}(n, y)}[i]) .
\end{aligned} \]
Therefore
\[ \begin{aligned}
\operatorname{Hom}_{\cat{D}^{\mrm{b}}(\cat{mod} A)}(M, M[i])
& \cong
\bigoplus_{x, y \in \vec{\Delta}_{0}}
\operatorname{Hom}_{\cat{D}^{\mrm{b}}
(\cat{mod} A)}(M_{(0, x)}, M_{(0, y)}[i]) \\
& \cong
\begin{cases}
A^{\circ} & \text{if } i = 0 \\
0 & \text{otherwise} .
\end{cases}
\end{aligned} \]
Also for any $(n, y)$ there is some integer $i$ and
$x \in \vec{\Delta}_{0}$ such that
\[ \operatorname{Hom}_{\cat{D}^{\mrm{b}}(\cat{mod} A)}
(M_{\pi(0, x)}, M_{(n, y)}[i])
\neq 0 . \]
Since any object $N \in \cat{D}^{\mrm{b}}(\cat{mod} A)$ is a direct
sum of indecomposables $M_{(n, y)}$, this implies that
$\operatorname{R} \operatorname{Hom}_{A}(M, N) \neq 0$ if $N \neq 0$. By \cite{Ye}
Theorem 1.8 and the proof of ``(ii) $\Rightarrow$ (i)''
of \cite{Ye} Theorem 1.6 there exists a two-sided tilting complex
$T$ with $T \cong M$ in $\cat{D}(\cat{Mod} A)$
(cf.\ \cite{Rd} Section 3). Replacing
$F$ with $\tilde{q}(T^{\vee}) F$,
where $T^{\vee} := \operatorname{R} \operatorname{Hom}_{A}(T, A)$,
we can assume that $p(F)$ is trivial.
Now that $p(F)$ is trivial, $F$ restricts to an
auto-equivalence of $k \bra{\vec{\Delta}}$, and by Proposition
\ref{prop1.1} we have
$F|_{k \bra{\vec{\Delta}}} \cong \mbf{1} _{k \bra{\vec{\Delta}}}$.
Then Corollary \ref{cor2.9} tells us
$F \cong \mbf{1}_{k \bra{\vec{\mbb{Z}} \vec{\Delta}, I_{\mrm{m}}}}$.
\medskip \noindent 3. Assume $A$ has infinite representation type.
Then the quiver isomorphism $\rho$ of Theorem \ref{thm2.2} induces
a group isomorphism
\[ \operatorname{Out}_{k}(k \bra{\vec{\Gamma}^{\mrm{irr}}, I_{\mrm{m}}})
^{\bra{\sigma}}
\cong
\operatorname{Out}_{k}(k \bra{\vec{\mbb{Z}} \vec{\Delta}, I_{\mrm{m}}})
\times \bra{\sigma} , \]
and $\bra{\sigma} \cong \mbb{Z}$.
We prove that
\[ \tilde{q} : \operatorname{DPic}_{k}(A) \to
\operatorname{Out}_{k}(k \bra{\vec{\mbb{Z}} \vec{\Delta}, I_{\mrm{m}}})
\times \mbb{Z} \]
is surjective.
Take an auto-equivalence $F$ of
$k \bra{\vec{\mbb{Z}} \vec{\Delta}, I_{\mrm{m}}}$, and write
$\pi := p(F) \in$ \newline
$\operatorname{Aut}((\vec{\mbb{Z}} \vec{\Delta})_{0}; d)^{\bra{\tau}}$.
After replacing $F$ with $\tau^{j} F$ for suitable
$j \in \mbb{Z}$, we can assume that
$\pi(0, x) \in \vec{\mbb{Z}}_{\geq 0} \vec{\Delta}$
for all $x \in \vec{\Delta}_{0}$.
Because $\vec{\mbb{Z}}_{\geq 0} \vec{\Delta}$ is the preprojective
component of $\vec{\Gamma}(\cat{mod} A)$ (cf.\ \cite{Rl}), we get
\[ M := \bigoplus_{x \in \vec{\Delta}_{0}} M_{\pi(0, x)}
\in \cat{mod} A . \]
As in part 2 above, $\operatorname{End}_{A}(M) = A^{\circ}$.
Since $M$ is a complete slice, \cite{HR} Theorem 7.2 says that $M$
is a tilting module. So $M$ is a two-sided tilting complex over $A$.
Replacing $F$ by $\tilde{q}(M^{\vee}) F$ we can
assume $p(F)$ is trivial. Let $P$ be an invertible bimodule such that
$\tilde{q}(P)|_{k \bra{\vec{\Delta}}} \cong
F|_{k \bra{\vec{\Delta}}}$.
Replacing $F$ with $\tilde{q}(P^{\vee}) F$
we get
$F|_{k \bra{\vec{\Delta}}} \cong \mbf{1}_{k \bra{\vec{\Delta}}}$.
Then by Corollary \ref{cor2.9} we get
$F \cong \mbf{1}_{\bra{\vec{\mbb{Z}} \vec{\Delta}, I_{\mrm{m}}}}$.
\end{proof}
The next theorem is Theorem \ref{thm0.2} in the Introduction.
\begin{thm} \label{thm3.2}
Let $A$ be an indecomposable basic hereditary finite dimensional
algebra over an algebraically closed field $k$, with quiver
$\vec{\Delta}$.
\begin{enumerate}
\item There is an exact sequence of groups
\[ 1 \to \operatorname{Out}^{0}_{k}(A) \to \operatorname{DPic}_{k}(A)
\xrightarrow{q}
\operatorname{Aut}(\vec{\Gamma}^{\mrm{irr}}_{0} ; d)^{\bra{\tau, \sigma}}
\to 1 . \]
This sequence splits.
\item If $A$ has finite representation type then there is an
isomorphism of groups
\[ \operatorname{DPic}_{k}(A) \cong
\operatorname{Aut}(\vec{\mbb{Z}} \vec{\Delta})^{\bra{\tau}} . \]
\item If $A$ has infinite representation type then
there is an isomorphism of groups
\[ \operatorname{DPic}_{k}(A) \cong
\bigl( \operatorname{Aut}((\vec{\mbb{Z}} \vec{\Delta})_{0}; d)^{\bra{\tau}}
\ltimes \operatorname{Out}^{0}_{k}(A) \bigr) \times \mbb{Z} . \]
\end{enumerate}
\end{thm}
\begin{proof}
1. By Theorem \ref{thm3.1} and Lemma \ref{lem3.12} the homomorphism $q$
is surjective. Lemma \ref{lem3.2} identifies $\operatorname{Ker}(q)$.
\medskip \noindent
2. If $A$ has finite representation type then $\vec{\Delta}$ is a
tree, so $\operatorname{Out}_{k}^{0}(A) = 1$ by Proposition \ref{prop1.1}.
By Theorem \ref{thm2.2} and Lemma \ref{lem3.11} we get
\[ \operatorname{Aut}(\vec{\Gamma}^{\mrm{irr}}_{0}; d)^{\bra{\tau, \sigma}}
\cong \operatorname{Aut}(\vec{\Gamma}^{\mrm{irr}})^{\bra{\tau, \sigma}}
\cong \operatorname{Aut}(\vec{\mbb{Z}} \vec{\Delta})^{\bra{\tau}} . \]
\medskip \noindent
3. If $A$ has infinite representation type then
\[ \operatorname{Aut}(\vec{\Gamma}^{\mrm{irr}}_{0}; d)^{\bra{\tau, \sigma}}
\cong
\operatorname{Aut}((\vec{\mbb{Z}} \vec{\Delta})_{0}; d)^{\bra{\tau}} \times
\bra{\sigma} \]
by Theorem \ref{thm2.2}. We know that $\sigma$ is in the center
of $\operatorname{DPic}_{k}(A)$.
\end{proof}
We end the section with the following problem.
\begin{prob} \label{prob3.1}
The Auslander-Reiten quiver
$\vec{\Gamma}(\cat{D}^{\mrm{b}}(\cat{mod} A))$
is defined for any finite dimensional $k$-algebra $A$ of finite
global dimension. Can the action of $\operatorname{DPic}_{k}(A)$ on
$\vec{\Gamma}(\cat{D}^{\mrm{b}}(\cat{mod} A))$
be used to determine the structure of $\operatorname{DPic}_{k}(A)$ for any
such $A$?
\end{prob}
\section{Explicit Calculations}
In this section we calculate the group structure of
$\operatorname{DPic}_{k}(A)$ for the path algebra $A = k \vec{\Delta}$
for several types of quivers. Throughout $S_{m}$
denotes the permutation group of $\{ 1, \ldots, m \}$.
Suppose $\Delta$ is a tree. Given an orientation $\omega$ of the
edge set $\Delta_{1}$, denote by $\vec{\Delta}_{\omega}$ the
resulting quiver, and by $A_{\omega} := k \vec{\Delta}_{\omega}$.
If $\omega$ and $\omega'$ are two orientations of $\Delta$ then
$\cat{D}^{\mrm{b}}(\cat{mod} A_{\omega}) \approx
\cat{D}^{\mrm{b}}(\cat{mod} A_{\omega'})$.
This equivalence will be discussed in the next section. For now we
just note that the groups
$\operatorname{DPic}_{k}(A_{\omega}) \cong \operatorname{DPic}_{k}(A_{\omega'})$,
so we are allowed to choose any orientation of $\Delta$
when computing these groups. This observation is relevant to
Theorems \ref{thm4.1} and \ref{thm4.2} below.
\begin{thm} \label{thm4.1}
Let $\vec{\Delta}$ be a Dynkin quiver as shown in Figure
\tup{\ref{fig1}}, and let $A := k \vec{\Delta}$ be the path algebra.
Then
$\operatorname{Pic}_{k}(A) \cong \operatorname{Aut}(\vec{\Delta})$
and
$\operatorname{DPic}_{k}(A) \cong
\operatorname{Aut}(\vec{\mbb{Z}} \vec{\Delta})^{\bra{\tau}}$.
The groups $\operatorname{Aut}(\vec{\Delta})$ and
$\operatorname{Aut}(\vec{\mbb{Z}} \vec{\Delta})^{\bra{\tau}}$
are described in Table \tup{\ref{tab1}}.
\end{thm}
\begin{table}
\begin{tabular}{|c|c|c|c|}
\hline
Type &
$\operatorname{Aut}(\vec{\Delta})$ &
$\operatorname{Aut}(\vec{\mbb{Z}} \vec{\Delta})^{\bra{\tau}}$
& Relation
\rule[-1ex]{0ex}{4ex} \\ \hline \hline
$A_{n}$, $n$ even & $1$ &
$\bra{\tau, \sigma}$ $\cong$ $\mbb{Z}$ &
$\tau^{n + 1} = \sigma^{-2}$
\rule[-1ex]{0ex}{4ex} \\ \hline
$A_{n}$, $n$ odd & $1$ &
$\bra{\tau, \sigma} \cong \mbb{Z} \times (\mbb{Z}/2 \mbb{Z})$
& $\tau^{n + 1} = \sigma^{-2}$
\rule[-1ex]{0ex}{4ex} \\ \hline
$D_{4}$ & $S_{3}$ &
$\operatorname{Aut}(\vec{\Delta}) \times \bra{\tau}
\cong S_{3} \times \mbb{Z}$ &
$\tau^{3} = \sigma^{-1}$
\rule[-1ex]{0ex}{4ex} \\ \hline
$D_{n}$, $n \geq 5$ & $S_{2}$ &
$\operatorname{Aut}(\vec{\Delta}) \times \bra{\tau}
\cong S_{2} \times \mbb{Z}$
& $\tau^{n-1} = \theta \sigma^{-1}$, $n$ odd \\
& & & $\tau^{n-1} = \sigma^{-1}$, $n$ even
\rule[-1ex]{0ex}{4ex} \\ \hline
$E_{6}$ & $S_{2}$ &
$\operatorname{Aut}(\vec{\Delta}) \times \bra{\tau}
\cong S_{2} \times \mbb{Z}$ &
$\tau^{6} = \theta \sigma^{-1}$
\rule[-1ex]{0ex}{4ex} \\ \hline
$E_{7}$ & $1$ &
$\bra{\tau} \cong {\mbb Z}$
& $\tau^{9} = \sigma^{-1}$
\rule[-1ex]{0ex}{4ex} \\ \hline
$E_{8}$ & $1$ &
$\bra{\tau} \cong \mbb{Z}$
& $\tau^{15} = \sigma^{-1}$
\rule[-1ex]{0ex}{4ex} \\ \hline
\end{tabular}
\medskip
\caption{The group
$\operatorname{Aut}(\vec{\mbb{Z}} \vec{\Delta})^{\bra{\tau}}$
for a Dynkin quiver. The orientation of $\vec{\Delta}$ is shown
in Figure \ref{fig1}. In types $D_{n}$ and $E_{6}$, $\theta$
is the element of order $2$ in $\operatorname{Aut}(\vec{\Delta})$.}
\label{tab1}
\end{table}
\begin{figure}
\choosegraphics{
\[ \UseTips
\begin{array}{lr}
A_{n} \quad
\begin{xy}
(0,0)*+@{*}="1"*+!U{\scrp{1}},"1"
\ar@{->} (8,0)*+@{*}="2"*+!U{\scrp{2}},"2"
\ar@{->} "2";(16,0)*+@{*}="3"*+!U{\scrp{3}},"3"
\ar@{} "3";(24,0)*+@{}="4","4" |*{\cdots}
\ar@{->} "4";(32,0)*+@{*}="n"*+!U{\scrp{n}},"n"
\end{xy}
& \hspace{15mm} D_{n} \quad
\begin{xy}
(0,0)*+@{*}="1"*+!R{\scrp{1}},"1"
\ar@{->} (0,8)*+@{*}="2"*+!R{\scrp{2}},"2"
\ar@{->} (0,-8)*+@{*}="3"*+!R{\scrp{3}},"3"
\ar@{->} (8,0)*+@{*}="4"*+!U{\scrp{4}},"4"
\ar@{->} "4";(16,0)*+@{*}="5"*+!U{\scrp{5}},"5"
\ar@{} "5";(24,0)*+@{}="6","6" |*{\cdots}
\ar@{->} "6";(32,0)*+@{*}="n"*+!U{\scrp{n}},"n"
\end{xy} \\
E_{6} \quad
\begin{xy}
(0,0)*+@{*}="1"*+!U{\scrp{1}},"1"
\ar@{->} (8,0)*+@{*}="2"*+!U{\scrp{2}},"2"
\ar@{->} "2";(16,0)*+@{*}="3"*+!U{\scrp{3}},"3"
\ar@{<-} "3";(24,0)*+@{*}="5"*+!U{\scrp{5}},"5"
\ar@{<-} "5";(32,0)*+@{*}="6"*+!U{\scrp{6}},"6"
\ar@{<-} "3";(16,8)*+@{*}="4"*+!R{\scrp{4}},"4"
\end{xy}
& \hspace{15mm} E_{7} \quad
\begin{xy}
(0,0)*+@{*}="1"*+!U{\scrp{1}},"1"
\ar@{->} (8,0)*+@{*}="2"*+!U{\scrp{2}},"2"
\ar@{->} "2";(16,0)*+@{*}="3"*+!U{\scrp{3}},"3"
\ar@{<-} "3";(16,8)*+@{*}="4"*+!R{\scrp{4}},"4"
\ar@{<-} "3";(24,0)*+@{*}="5"*+!U{\scrp{5}},"5"
\ar@{<-} "5";(32,0)*+@{*}="6"*+!U{\scrp{6}},"6"
\ar@{<-} "6";(40,0)*+@{*}="7"*+!U{\scrp{7}},"7"
\end{xy} \\[5mm]
& E_{8} \quad
\begin{xy}
(0,0)*+@{*}="1"*+!U{\scrp{1}},"1"
\ar@{->} (8,0)*+@{*}="2"*+!U{\scrp{2}},"2"
\ar@{->} "2";(16,0)*+@{*}="3"*+!U{\scrp{3}},"3"
\ar@{<-} "3";(16,8)*+@{*}="4"*+!R{\scrp{4}},"4"
\ar@{<-} "3";(24,0)*+@{*}="5"*+!U{\scrp{5}},"5"
\ar@{<-} "5";(32,0)*+@{*}="6"*+!U{\scrp{6}},"6"
\ar@{<-} "6";(40,0)*+@{*}="7"*+!U{\scrp{7}},"7"
\ar@{<-} "7";(48,0)*+@{*}="8"*+!U{\scrp{8}},"8"
\end{xy}
\end{array} \]
}{
\includegraphics[clip]{fig1.eps}
}
\caption{Orientations for the Dynkin graphs} \label{fig1}
\end{figure}
\begin{proof}
The isomorphisms are by Theorem \ref{thm0.2} and Proposition
\ref{prop1.1}. The data in the third column of Table \ref{tab1}
was calculated in \cite{Rn} Section 4, except for the shift $\sigma$
which did not appear in that paper. So we have to do a few
calculations involving $\sigma$. Below are the
calculations for types $A_{n}$ and $D_{4}$; the rest are similar
and are left to the reader as an exercise.
\medskip \noindent
Type $A_{n}$: Choose the orientation in Figure \ref{fig1}.
The quiver $\vec{\mbb{Z}} \vec{\Delta}$ looks like Figure \ref{fig2}.
Therefore
$\operatorname{Aut}(\vec{\mbb{Z}} \vec{\Delta})^{\bra{\tau}}
= \bra{\tau, \eta}$
where $\eta(0, 1) = (0, n)$ and
$\eta(0, n) = (n - 1, 1)$.
Now by \cite{Ha} Section I.5.5 and \cite{ARS} Sections VII.1 and
VIII.5, the quiver
$\vec{\Gamma}(\cat{mod} A) \subset \vec{\mbb{Z}} \vec{\Delta}$
is the full subquiver on the vertices in the triangle
$\{ (m, i) \mid m \geq 0,\ m + i \leq n \}$.
The projective vertices are
$(0, i)$ and the injective vertices are
$(n - i, i)$, where $i \in \{ 1, \ldots, n \}$.
We see that $\sigma(0, i) = (i, n + 1 - i)$, and the quiver
$\vec{\Gamma} \bigl( (\cat{mod} A)[1] \bigr) =
\sigma \bigl( \vec{\Gamma}(\cat{mod}A) \bigr)$
is the full subquiver on the vertices in the triangle
$\{ (m, i) \mid m \leq n,\ m + i \geq n + 1 \}$.
Hence $\eta = \tau \sigma$ and
$\operatorname{Aut}(\vec{\mbb{Z}} \vec{\Delta})^{\bra{\tau}}
= \bra{\tau, \sigma}$.
The relation $\tau^{-(n + 1)} = \sigma^{2}$ is easily verified.
\medskip \noindent
Type $D_{4}$: The quiver $\vec{\mbb{Z}} \vec{\Delta}$
is in Figure \ref{fig3}, and
$\vec{\Gamma}(\cat{mod}A) \subset \vec{\mbb{Z}} \vec{\Delta}$
is a full subquiver. From the shape of $\vec{\Delta}$ we know that
$\cat{mod} A$ should have $4$ indecomposable projective modules,
$3$ having length $2$ and one of them simple. From the shape of the
opposite quiver $\vec{\Delta}^{\circ}$ we also know that
$\cat{mod} A$ should have $4$ indecomposable injective modules, $3$
of them simple and one of length $4$.
Counting dimensions using Auslander-Reiten sequences we conclude
that $\vec{\Gamma}(\cat{mod}A)$ is the full subquiver on the
vertices
$\{ 0, 1, 2 \} \times \vec{\Delta}_{0}$. The projective vertices
are $\{ (0, 1), (0, i) \}$,
the injective vertices are $\{ (2, 1), (2, i) \}$, and
the simple vertices are $\{ (0, 1), (2, i) \}$,
where $i \in \{ 2, 3, 4 \}$.
For $i \in \{ 1, 2, 3, 4 \}$ let $P_{i}$, $S_{i}$ and $I_{i}$, be
the projective, simple and injective modules respectively, indexed
such that $P_{i} \twoheadrightarrow S_{i} \rightarrowtail I_{i}$, and with
$P_{i} = M_{(0, i)}$. So $P_{1} = S_{1}$ and
$I_{i} = S_{i}$ for $i \in \{ 2, 3, 4 \}$. By the symmetry of the
quiver it follows that there is a nonzero morphism
$(0, i) \to (2, i)$ in $k \bra{\vec{\mbb{Z}} \vec{\Delta}}$
for $i \in \{ 2, 3, 4 \}$, and hence
$M_{(2, i)} \cong S_{i}$
The rule for connecting $\vec{\Gamma}(\cat{mod}A)$ with
$\vec{\Gamma}(\cat{mod}A[1])$ (see \cite{Ha} Section I.5.5)
implies that
$M_{(3, 1)} \cong M_{(0, 1)}[1] = P_{1}[1]$. Therefore
$M_{(3, i)} \cong P_{i'}[1]$
for $i, i' \in \{ 2, 3, 4 \}$. Now for each such $i$ there is
an Auslander-Reiten triangle
$M_{(2, i)} \to M_{(3, 1)} \to M_{(3, i)} \to M_{(2, i)}[1]$.
When this triangle is turned it gives an exact sequence
$0 \to P_{1} \to P_{i'} \to S_{i} \to 0$, and hence $i' = i$.
The conclusion is that $\sigma(m, i) = (m + 3, i)$ for all
$(m, i) \in (\vec{\mbb{Z}} \vec{\Delta})_{0}$, so
$\sigma = \tau^{-3}$.
\end{proof}
\begin{figure}
\choosegraphics{
\[ \UseTips
\begin{xy}
(0,0)*+@{*}="01"*+!U{\scrp{(0,1)}},"01"
\ar@{->} "01";"01"+(10,10)*+@{*}="02"*++!R{\scrp{(0,2)}},"02"
\ar@{->} "02";"02"+(10,10)*+@{*}="03"*+!D{\scrp{(0,3)}},"03"
\ar@{->} "02";"01"+(20,0)*+@{*}="11"*+!U{\scrp{(1,1)}},"11"
\ar@{->} "03";"02"+(20,0)*+@{*}="12"*++!R{\scrp{(1,2)}},"12"
\ar@{->} "11";"12"
\ar@{->} "12";"12"+(10,10)*+@{*}="13"*+!D{},"13"
\ar@{->} "12";"11"+(20,0)*+@{*}="21"*+!U{\scrp{(2,1)}},"21"
\ar@{->} "13";"12"+(20,0)*+@{*}="22"*++!R{},"22"
\ar@{->} "21";"22"
\ar@{->} "22";"22"+(10,10)*+@{*}="23"*+!D{},"23"
\ar@{->} "22";"21"+(20,0)*+@{*}="31"*++!L{\cdots},"31"
\ar@{->} "23";"22"+(20,0)*+@{*}="32"*++!L{\cdots},"32"
\ar@{->} "31";"32"
\ar@{->} "32";"32"+(10,10)*+@{*}="33"*++!L{\cdots},"33"
\ar@{->} "02"-(20,0)*+@{*}="-12"*++!R{\cdots},"-12";"01"
\ar@{->} "03"-(20,0)*+@{*}="-13"*++!R{\cdots},"-13";"02"
\ar@{->} "01"-(20,0)*+@{*}="-11"*++!R{\cdots},"-11";"-12"
\ar@{->} "-12";"-13"
\end{xy} \]
}{\includegraphics[clip]{fig2.eps}}
\caption{The quiver $\vec{\mbb{Z}} \vec{\Delta}$ for $\vec{\Delta}$
of type $A_{3}$. The vertices in
$\cat{mod} A$ are labeled.}
\label{fig2}
\end{figure}
\begin{figure}
\choosegraphics{
\[ \UseTips
\begin{xy}
(0,0)*+@{*}="01"*+++!U{\scrp{(0,1)}}
*++!R{\cdots},"01"
\ar@{->} "01";"01"+(10,10)*+@{*}="02"*+!D{\scrp{(0,2)}},"02"
\ar@{->} "01";"01"+(10,0)*+@{*}="03"*+!D{\scrp{(0,3)}},"03"
\ar@{->} "01";"01"+(10,-10)*+@{*}="04"*+!U{\scrp{(0,4)}},"04"
\ar@{} (0,0)+(20,0)*+@{*}="11"*+++!U{\scrp{(1,1)}},"11"
\ar@{->} "11";"11"+(10,10)*+@{*}="12"*+!D{\scrp{(1,2)}},"12"
\ar@{->} "11";"11"+(10,0)*+@{*}="13"*+!D{\scrp{(1,3)}},"13"
\ar@{->} "11";"11"+(10,-10)*+@{*}="14"*+!U{\scrp{(1,4)}},"14"
\ar@{->} "02";"11"
\ar@{->} "03";"11"
\ar@{->} "04";"11"
\ar@{} "11"+(20,0)*+@{*}="21"*+++!U{\scrp{(2,1)}},"21"
\ar@{->} "21";"21"+(10,10)*+@{*}="22"*+!D{\scrp{(2,2)}},"22"
\ar@{->} "21";"21"+(10,0)*+@{*}="23"*+!D{\scrp{(2,3)}},"23"
\ar@{->} "21";"21"+(10,-10)*+@{*}="24"*+!U{\scrp{(2,4)}},"24"
\ar@{->} "12";"21"
\ar@{->} "13";"21"
\ar@{->} "14";"21"
%
\ar@{} "21"+(20,0)*+@{*}="31"
\ar@{->} "31";"31"+(10,10)*+@{*}="32"*++!L{\cdots},"32"
\ar@{->} "31";"31"+(10,0)*+@{*}="33"*++!L{\cdots},"33"
\ar@{->} "31";"31"+(10,-10)*+@{*}="34"*++!L{\cdots},"34"
\ar@{->} "22";"31"
\ar@{->} "23";"31"
\ar@{->} "24";"31"
\end{xy} \]
}{ \includegraphics[clip]{fig3.eps}}
\caption{
The quiver $\vec{\mbb{Z}} \vec{\Delta}$ for $\vec{\Delta}$
of type $D_{4}$. The vertices in $\cat{mod} A$ are labeled.}
\label{fig3}
\end{figure}
\begin{thm} \label{thm4.2}
Let $\vec{\Delta}$ be a quiver of type $\tilde{D}_{n}$,
$\tilde{E}_{6}$, $\tilde{E}_{7}$ or $\tilde{E}_{8}$, with the
orientation shown in Figure \tup{\ref{fig4}}. Then
$\operatorname{Pic}_{k}(A) \cong \operatorname{Aut}(\vec{\Delta})$
and
\[ \operatorname{DPic}_{k}(A) \cong \mbb{Z} \times
\operatorname{Aut}(\vec{\mbb{Z}} \vec{\Delta})^{\bra{\tau}} . \]
The structure of the group
$\operatorname{Aut}(\vec{\mbb{Z}} \vec{\Delta})^{\bra{\tau}}$
is given in Table \tup{\ref{tab2}}.
\end{thm}
\begin{proof}
The isomorphisms follow from Theorem \ref{thm0.2} and Proposition
\ref{prop1.1}. The structure of
$\operatorname{Aut}(\vec{\mbb{Z}} \vec{\Delta})^{\bra{\tau}}$
is quite easy to check in all cases. In type $\tilde{D}_{n}$,
$n \geq 5$ odd, the automorphism
$\eta \in \operatorname{Aut}(\vec{\mbb{Z}} \vec{\Delta})^{\bra{\tau}}$
is
\[
\eta(i,j) =
\begin{cases}
(i, n+2 - j) & \text{if } j = 2, n \\
(i - \frac{1-(-1)^j}{2}, n+2 - j) & \text{otherwise} .
\end{cases}
\]
\end{proof}
\begin{figure}
\choosegraphics{
\[ \UseTips
\begin{array}{l}
\tilde{D}_{4} \quad
\begin{xy}
(0,0)*+@{*}="1"*+!R{\scrp{1}},"1"
\ar@{->} (10,6)*+@{*}="2"*+!L{\scrp{2}},"2"
\ar@{->} (10,2)*+@{*}="3"*+!L{\scrp{3}},"3"
\ar@{->} (10,-2)*+@{*}="4"*+!L{\scrp{4}},"4"
\ar@{->} (10,-6)*+@{*}="5"*+!L{\scrp{5}},"5"
\end{xy} \hspace{15mm}
\begin{xy}
(0,0)*+@{*}="1"*+!U{\scrp{1}},"1"
\ar@{->} (8,0)*+@{*}="2"*+!U{\scrp{2}},"2"
\ar@{->} "2";(16,0)*+@{*}="3"*+!U{\scrp{3}},"3"
\ar@{->} "3";(24,0)*+@{*}="4"*+!U{\scrp{4}},"4"
\ar@{->} "4";(32,0)*+@{*}="5"*+!U{\scrp{5}},"5"
\ar@{->} "5";(40,0)*+@{*}="6"*+!U{\scrp{6}},"6"
\ar@{->} "6";(48,0)*+@{*}="7"*+!U{\scrp{7}},"7"
\ar@{->} "7";(56,0)*+@{*}="8"*+!U{\scrp{8}},"8"
\ar@{->} "3";(16,8)*+@{*}="9"*+!R{\scrp{9}},"9"
\ar@{} (8,6)*{\tilde{E}_{8}}
\end{xy} \\
\begin{xy}
(0,0)*+@{*}="3"*+!R{\scrp{3}},"3"
\ar@{->} (0,8)*+@{*}="1"*+!R{\scrp{1}},"1"
\ar@{->} (0,-8)*+@{*}="2"*+!R{\scrp{2}},"2"
\ar@{->} (8,0)*+@{*}="4"*+!U{\scrp{4}};"3"
\ar@{->} "4";(16,0)*+@{}="5","5"
\ar@{} "5";(24,0)*+@{}="6","6" |*{\cdots}
\ar@{->} (32,0)*+@{*}="2m-2";"6"
\ar@{->} "2m-2";(40,0)*+@{*}="2m-1"*+!L{\scrp{2m-1}},"2m-1"
\ar@{->} "2m-1";(40,8)*+@{*}="2m"*+!L{\scrp{2m}},"2m"
\ar@{->} "2m-1";(40,-8)*+@{*}="2m+1"*+!L{\scrp{2m+1}},"2m+1"
\ar@{} (10,6)*{\tilde{D}_{2m}}
\end{xy} \hspace{15mm}
\begin{xy}
(0,0)*+@{*}="1"*+!U{\scrp{1}},"1"
\ar@{->} (8,0)*+@{*}="2"*+!U{\scrp{2}};"1"
\ar@{->} (16,0)*+@{*}="3"*+!U{\scrp{3}};"2"
\ar@{->} (24,0)*+@{*}="4"*+!U{\scrp{4}};"3"
\ar@{->} "4";(32,0)*+@{*}="5"*+!U{\scrp{5}},"5"
\ar@{->} "5";(40,0)*+@{*}="6"*+!U{\scrp{6}},"6"
\ar@{->} "6";(48,0)*+@{*}="7"*+!U{\scrp{7}},"7"
\ar@{->} "4";(24,8)*+@{*}="8"*+!R{\scrp{8}},"8"
\ar@{} (8,6)*{\tilde{E}_{7}}
\end{xy} \\
\begin{xy}
(0,0)*+@{*}="3"*+!R{\scrp{3}},"3"
\ar@{->} (0,8)*+@{*}="1"*+!R{\scrp{1}},"1"
\ar@{->} (0,-8)*+@{*}="2"*+!R{\scrp{2}},"2"
\ar@{->} (8,0)*+@{*}="4"*+!U{\scrp{4}};"3"
\ar@{->} "4";(16,0)*+@{}="5","5"
\ar@{} "5";(24,0)*+@{}="6","6" |*{\cdots}
\ar@{->} "6";(32,0)*+@{*}="2m-1","2m-1"
\ar@{->} (40,0)*+@{*}="2m"*+!L{\scrp{2m}},"2m";"2m-1"
\ar@{->} "2m";(40,8)*+@{*}="2m+1"*+!L{\scrp{2m+1}},"2m+1"
\ar@{->} "2m";(40,-8)*+@{*}="2m+2"*+!L{\scrp{2m+2}},"2m+2"
\ar@{} (10,6)*{\tilde{D}_{2m+1}}
\end{xy} \hspace{15mm}
\begin{xy}
(0,0)*+@{*}="1"*+!U{\scrp{1}},"1"
\ar@{->} (8,0)*+@{*}="2"*+!U{\scrp{2}};"1"
\ar@{->} (16,0)*+@{*}="3"*+!U{\scrp{3}};"2"
\ar@{->} "3";(24,0)*+@{*}="4"*+!U{\scrp{4}},"4"
\ar@{->} "4";(32,0)*+@{*}="5"*+!U{\scrp{5}},"5"
\ar@{->} "3";(16,8)*+@{*}="6"*+!R{\scrp{6}},"6"
\ar@{->} "6";(16,16)*+@{*}="7"*+!R{\scrp{7}},"7"
\ar@{} (8,6)*{\tilde{E}_{6}}
\end{xy}
\end{array} \]
}{\includegraphics[clip]{fig4.eps}}
\caption{Orientations for the affine tree graphs} \label{fig4}
\end{figure}
\begin{table}
\begin{tabular}{|c|c|c|c|}
\hline
Type & $\operatorname{Aut}(\vec{\Delta})$ &
$\operatorname{Aut}(\vec{\mbb{Z}} \vec{\Delta})^{\bra{\tau}}$ &
Relations \rule[-1ex]{0ex}{4ex} \\ \hline \hline
$\tilde{D}_{4}$ & $S_{4}$ &
$\operatorname{Aut}(\vec{\Delta}) \times \bra{\tau} \cong
S_{4} \times \mbb{Z}$ &
\rule[-1ex]{0ex}{4ex} \\ \hline
$\tilde{D}_{n}$, $n \geq 5$ even &
$S_{2} \ltimes S_{2}^{2}$ &
$\operatorname{Aut}(\vec{\Delta}) \times \bra{\tau} \cong
(S_{2} \ltimes S_{2}^{2}) \times \mbb{Z}$ &
\rule[-1ex]{0ex}{4ex} \\ \hline
$\tilde{D}_{n}$, $n \geq 5$ odd &
$S_{2}^{2}$ &
$\operatorname{Aut}(\vec{\Delta}) \times \bra{\eta} \cong
S_{2}^{2} \times \mbb{Z}$ & $\eta^{2} = \tau$
\rule[-1ex]{0ex}{4ex} \\ \hline
$\tilde{E}_{6}$ & $S_{3}$ &
$\operatorname{Aut}(\vec{\Delta}) \times \bra{\tau} \cong
S_{3} \times \mbb{Z}$ &
\rule[-1ex]{0ex}{4ex} \\ \hline
$\tilde{E}_{7}$ & $S_{2}$ &
$\operatorname{Aut}(\vec{\Delta}) \times \bra{\tau} \cong
S_{2} \times \mbb{Z}$ &
\rule[-1ex]{0ex}{4ex} \\ \hline
$\tilde{E}_{8}$ & $1$ &
$\bra{\tau} \cong \mbb{Z}$ &
\rule[-1ex]{0ex}{4ex} \\ \hline
\end{tabular}
\medskip
\caption{The groups
$\operatorname{Aut}(\vec{\mbb{Z}} \vec{\Delta})^{\bra{\tau}}$
for the affine tree quivers shown in Figure \tup{\ref{fig4}}.}
\label{tab2}
\end{table}
\begin{thm} \label{thm4.3}
For any $n \geq 2$ let $\vec{\Omega}_{n}$ be the quiver shown in
Figure \tup{\ref{fig5}}, and let $A := k \vec{\Omega}_{n}$ be the
path algebra. Then
$\operatorname{Pic}_{k}(A) \cong \operatorname{PGL}_{n}(k)$
and
\[ \operatorname{DPic}_{k}(A) \cong \mbb{Z} \times \bigl( \mbb{Z} \ltimes
\operatorname{PGL}_{n}(k) \bigr) . \]
In the semidirect product the action of a generator
$\rho \in \mbb{Z}$ on a matrix $F \in \operatorname{PGL}_{n}(k)$ is
$\rho F \rho^{-1} = (F^{-1})^{\mrm{t}}$.
\end{thm}
\begin{figure}
\choosegraphics{
\[ \UseTips
\vec{\Omega}_{n} \quad
\begin{xy}
(0,0)*+@{*}="1"*+!CR{\scrp{1}},"1"
\ar@/^/@<2.5ex>@{->}|*+{\scrp{\alpha_{1}}}
(20,0)*+@{*}="2"*+!CL{\scrp{2}},"2"
\ar@/^/@<1ex>@{->}|*+{\scrp{\alpha_{2}}} "2"
\ar@{}|{\vdots} "2"
\ar@/_/@<-2ex>@{->}|*+{\scrp{\alpha_{n}}} "2"
\end{xy}
\hspace{25mm}
\vec{T}_{p, q} \quad
\begin{xy}
(0,10)*+@{*}="1"*+!DC{\scrp{1}},"1"
\ar@{->}^{\alpha_{1}} (10,10)*+@{*}="2"*+!DC{\scrp{2}},"2"
\ar@{->}^{\alpha_{2}} "2";(20,10)="3"
\ar@{}|*{\cdots} "3";(30,10)="4"
\ar@{->}^{\alpha_{p - 1}} "4";(40,10)*+@{*}="p"*+!DC{\scrp{p}}
\ar@{->}^{\alpha_{p}}
"p";(40,0)*+@{*}="p+1"*+!UC{\scrp{p+1}},"p+1"
\ar@{->}_{\beta_{q}}
"1";(0,0)*+@{*}="p+q"*+!UC{\scrp{p+q}},"p+q"
\ar@{->}^{\beta_{q-1}}
"p+q";(10,0)*+@{*}="p+q-1"*+!UC{\scrp{p+q-1}},"p+q-1"
\ar@{->}^{\beta_{q-2}} "p+q-1";(20,0)="p+q-2"
\ar@{}|*{\cdots} "p+q-2";(30,0)="p+2"
\ar@{->}^{\beta_{1}} "p+2";"p+1"
\end{xy} \]
}{\includegraphics[clip]{fig5.eps}}
\caption{The quivers $\vec{\Omega}_{n}$ and $\vec{T}_{p, q}$.}
\label{fig5}
\end{figure}
\begin{proof}
As in the proof of Lemma \ref{lem1.1} and Proposition \ref{prop1.1},
the group of auto-equivalences of the path category is
$\operatorname{Aut}_{k}(k \bra{\vec{\Omega}_{n}}) =
\operatorname{Aut}_{k}^{0}(k \bra{\vec{\Omega}_{n}}) \cong
\operatorname{GL}_{n}(k)$.
Hence
$\operatorname{Pic}_{k}(A) \cong
\operatorname{Out}_{k}(k \bra{\vec{\Omega}_{n}}) \cong \operatorname{PGL}_{n}(k)$.
Given $F \in \operatorname{Aut}_{k}(k \bra{\vec{\Omega}_{n}})$
let
$[a_{i, j}] \in \operatorname{GL}_{n}(k)$ be its matrix w.r.t.\ to the basis
$\{ \alpha_{i} \}$, and let
$[b_{i, j}] := ([a_{i, j}]^{-1})^{\mrm{t}}$.
Define an auto-equivalence
$\tilde{F} \in
\operatorname{Aut}_{k}^{0}(k \bra{\vec{\mbb{Z}} \vec{\Omega}_{n}})$
with
$\tilde{F}(m, \alpha_{i}) = \sum_{j} a_{i, j} (m, \alpha_{j})$
and
$\tilde{F}(m, \alpha^{*}_{i}) =
\sum_{j} b_{i, j} (m, \alpha^{*}_{j})$, $m \in \mbb{Z}$.
Then $\tilde{F}$ preserves all mesh relations, and by a linear algebra
argument we see that up to scalars at each vertex, the only
elements of
$\operatorname{Aut}^{0}_{k}(k \bra{\vec{\mbb{Z}} \vec{\Omega}_{n}})$
are of the form $\tilde{F}$.
Let $\rho \in \operatorname{Aut}(\vec{\mbb{Z}} \vec{\Omega}_{n})$ be
$\rho(m, 1) = (m, 2)$ and $\rho(m, 2) = (m + 1, 1)$, with the
obvious action on arrows to make it commute with the polarization
$\mu$. Then
$\operatorname{Out}_{k}(k \bra{\vec{\mbb{Z}} \vec{\Omega}_{n}, I_{\mrm{m}}})$
is generated by $\operatorname{PGL}_{n}(k)$ and $\rho$, so
$\operatorname{Out}_{k}(k \bra{\vec{\mbb{Z}} \vec{\Omega}_{n}, I_{\mrm{m}}})
\cong \mbb{Z} \ltimes \operatorname{PGL}_{n}(k)$.
The formula for $\tilde{F}$ above shows that
$\rho F \rho^{-1} = (F^{-1})^{\mrm{t}}$ for
$F \in \operatorname{PGL}_{n}(k)$.
Finally use Theorem \ref{thm3.1}.
\end{proof}
\begin{rem}
By \cite{Be} and \cite{BO} we see that for $n = 2$ in the theorem
above,
$\operatorname{DPic}_{k}(A) \cong \mbb{Z} \times \mbb{Z} \times
\operatorname{PGL}_{2}(k)$.
The apparent discrepancy is explained by the fact that
$\mbb{Z} \ltimes \operatorname{PGL}_{2}(k) \cong
\mbb{Z} \times \operatorname{PGL}_{2}(k)$
via
$(m, F) \mapsto (m, H^{m} F)$,
where
$H = \left[ \begin{smallmatrix}
0 & -1 \\ 1 & 0
\end{smallmatrix} \right]$.
\end{rem}
For integers $p \geq q \geq 1$ let $\vec{T}_{p, q}$ be the quiver
shown in Figure \ref{fig5}.
Let $\vec{\Delta}$ be a quiver with underlying graph $\tilde{A}_{n}$.
Then $\vec{\Delta}$ can be brought to one
of the quivers $\vec{T}_{p, q}$, $p + q = n+1$, by a sequence of
admissible reflections $s_{x}^{-}$ at source vertices (see Section
6). Therefore
\[ \operatorname{DPic}_{k}(k \vec{\Delta}) \cong
\operatorname{DPic}_{k}(k \vec{T}_{p, q}) . \]
\begin{thm} \label{thm4.4}
Let $A$ be the path algebra $k \vec{T}_{p, q}$.
\begin{enumerate}
\item If $p = q = 1$ then
$\operatorname{Pic}_{k}(A) \cong \operatorname{PGL}_{2}(k)$
and
$\operatorname{DPic}_{k}(A) \cong \mbb{Z} \times
\bigl( \mbb{Z} \ltimes \operatorname{PGL}_{2}(k) \bigr)$.
\item If $p > q = 1$ then
$\operatorname{Pic}_{k}(A) \cong
\left[ \begin{smallmatrix}
k^{\times} & k \\ 0 & 1
\end{smallmatrix} \right]$
and
$\operatorname{DPic}_{k}(A) \cong \mbb{Z} \times
\bigl( \mbb{Z} \ltimes
\left[ \begin{smallmatrix} k^{\times} & k \\ 0 & 1
\end{smallmatrix} \right] \bigr)$.
\item If $p = q > 1$ then
$\operatorname{Pic}_{k}(A) \cong S_{2} \ltimes k^{\times}$
and
$\operatorname{DPic}_{k}(A) \cong \mbb{Z}^{2} \times
\bigl( S_{2} \ltimes k^{\times} \bigr)$.
\item If $p > q > 1$ then
$\operatorname{Pic}_{k}(A) \cong k^{\times}$
and
$\operatorname{DPic}_{k}(A) \cong \mbb{Z}^{2} \times
k^{\times}$.
\end{enumerate}
\end{thm}
\begin{proof}
1. This is because
$\vec{T}_{1, 1} = \vec{\Omega}_{2}$.
\medskip \noindent 2.
Here the group of auto-equivalences of $k \bra{\vec{T}_{p, q}}$
is, in the notation of the proof of Proposition \ref{prop1.1},
$\operatorname{Aut}_{k}(k \bra{\vec{T}_{p, q}}) \cong
(k^{\times})^{p + 1} \times k$,
and the group of isomorphisms is $(k^{\times})^{p}$. Therefore
$\operatorname{Out}_{k}(k \bra{\vec{T}_{p, q}})$
is isomorphic to $k^{\times} \times k$ as varieties, and as matrix
group
$\operatorname{Out}_{k}(k \bra{\vec{T}_{p, q}})
\cong \left[ \begin{smallmatrix} k^{\times} & k \\ 0 & 1
\end{smallmatrix} \right]$.
The auto-equivalence associated to
$\left[ \begin{smallmatrix} a & b \\ 0 & 1
\end{smallmatrix} \right] \in
\left[ \begin{smallmatrix} k^{\times} & k \\ 0 & 1
\end{smallmatrix} \right]$
is $\alpha_{i} \mapsto \alpha_{i}$ and
$\beta_{1} \mapsto a \beta_{1} + b \alpha_{p} \cdots \alpha_{1}$.
The quiver $\vec{\mbb{Z}} \vec{T}_{p, q}$ has no multiple arrows. Let
$\rho$ be the symmetry $\rho(m, i) = (m, i - 1)$ for $i \geq 2$,
and $\rho(m, 1) = (m - 1, p)$. Then $\rho$ generates
$\operatorname{Aut}(\vec{\mbb{Z}} \vec{T}_{p, q})^{\bra{\tau}}$, and we can use
Theorem \ref{thm0.2}. The action of $\rho$ on
$\operatorname{Out}_{k}(k \bra{\vec{T}_{p, q}})$ is
$\rho
\left[ \begin{smallmatrix} a & b \\ 0 & 1 \end{smallmatrix} \right]
\rho^{-1} =
\left[ \begin{smallmatrix} a & -b \\ 0 & 1 \end{smallmatrix}
\right]$.
\medskip \noindent 3.
Here
$\operatorname{Aut}^{0}_{k}(k \bra{\vec{T}_{p, q}}) \cong
(k^{\times})^{2p}$,
and the subgroup of isomorphisms is $(k^{\times})^{2p - 1}$.
The symmetry $\theta \in \operatorname{Aut}(\vec{T}_{p, q})$ of order $2$
acts on $k^{\times}$ by
$\theta a \theta^{-1} = a^{-1}$.
Let $\rho$ be the symmetry
$\rho(m, 1) = (m - 1, p + q)$,
$\rho(m, i) = (m, i - 1)$ if $2 \leq i \leq p + 1$,
and
$\rho(m, i) = (m - 1, i - 1)$ if $p + 2 \leq i \leq p + q$.
Then $\rho$ and $\theta$ commute, and they generate
$\operatorname{Aut}(\vec{\mbb{Z}} \vec{T}_{p, q})^{\bra{\tau}}$. The
action of $\rho$ on $\operatorname{Aut}_{k}(k \bra{\vec{T}_{p, q}})$
is trivial.
\medskip \noindent 4.
Similar to case 3.
\end{proof}
\section{The Reflection Groupoid of a Graph}
In this section we interpret the reflection functors of \cite{BGP}
and the tilting modules of \cite{APR} in the setup of derived
categories.
Let $\Delta$ be a tree with $n$ vertices. Denote by
$\operatorname{Or}(\Delta)$ the set of orientations of the edge set
$\Delta_{1}$. For $\omega \in \operatorname{Or}(\Delta)$ let
$\vec{\Delta}_{\omega}$ be the resulting quiver, and let
$A_{\omega}$ be the path algebra $k \vec{\Delta}_{\omega}$.
Given two orientations $\omega,\omega'$ let
\[ \operatorname{DPic}_{k}(\omega,\omega') :=
\frac{ \{ \text{two-sided tilting complexes }
T \in \cat{D}^{\mrm{b}}(\cat{Mod}(A_{\omega'} \otimes_{k}
A_{\omega}^{\circ})) \} }
{\text{isomorphism}} . \]
The {\em derived Picard groupoid} of $\Delta$ is the groupoid
$\operatorname{DPic}_{k}(\Delta)$ with object set $\operatorname{Or}(\Delta)$ and
morphism sets $\operatorname{DPic}_{k}(\omega,\omega')$. Thus when
$\omega = \omega'$ we recover the derived Picard group
$\operatorname{DPic}_{k}(A_{\omega})$.
For an orientation $\omega$ and a vertex $x$ let
$P_{x, \omega} \in \cat{mod} A_{\omega}$
be the corresponding indecomposable projective module.
Denote by $\tau_{\omega}$ the translation functor of
$\cat{D}^{\mrm{b}}(\cat{mod} A_{\omega})$,
i.e.\ the functor
$\tau_{\omega} = A_{\omega}^{*}[-1] \otimes^{\mrm{L}}_{A_{\omega}}
-$.
Suppose $x \in (\vec{\Delta}_{\omega})_{0}$ is a source. Define
$s^{-}_{x} \omega$ to be the orientation obtained from $\omega$ by
reversing the arrows starting at $x$. Let
\[ T_{x, \omega} := \tau_{\omega}^{-1} P_{x, \omega} \oplus
\bigl( \bigoplus_{y \neq x} P_{y, \omega} \bigr)
\in \cat{mod} A_{\omega} . \]
According to \cite{APR} Section 3,
$T_{x, \omega}$ is a tilting module, with
$\operatorname{End}_{A_{\omega}}(T_{x, \omega})^{\circ} \cong
A_{s^{-}_{x} \omega}$. It is called an {\em APR tilting module}.
One has isomorphisms in $\cat{mod} A_{s^{-}_{x} \omega}$:
\begin{equation} \label{eqn5.1}
\begin{aligned}
\operatorname{Hom}_{A_{\omega}}(T_{x, \omega}, P_{y, \omega}) & \cong
P_{y, s^{-}_{x} \omega} \quad \text{if } y \neq x , \\
\operatorname{Hom}_{A_{\omega}}(T_{x, \omega},
\tau_{\omega}^{-1} P_{x, \omega}) & \cong
P_{x, s^{-}_{x} \omega} .
\end{aligned}
\end{equation}
Under the anti-equivalence between $\cat{mod} A_{\omega}$ and
the category of finite dimensional representations of the quiver
$\vec{\Delta}_{\omega}$, the reflection functor of \cite{BGP} is
sent to
$\operatorname{Hom}_{A_{\omega}}(T_{x, \omega}, -) : \cat{mod} A_{\omega}
\to \cat{mod} A_{s^{-}_{x} \omega}$.
\begin{dfn} \label{dfn5.1}
The {\em reflection groupoid} of $\Delta$ is the subgroupoid
$\operatorname{Ref}(\Delta) \subset$ \linebreak
$\operatorname{DPic}_{k}(\Delta)$
generated by the two-sided tilting complexes
$T_{x, \omega} \in
\cat{D}^{\mrm{b}}(\cat{Mod}
(A_{\omega} \otimes_{k} A_{s^{-}_{x} \omega}^{\circ}))$,
as $\omega$ runs over $\operatorname{Or}(\Delta)$ and $x$ runs over the
sources in $\vec{\Delta}_{\omega}$.
\end{dfn}
Given an orientation $\omega$ the set
$\{ [P_{x, \omega}] \}_{x \in \Delta_{0}}$
is a basis of the Grothendieck group
$\operatorname{K}_{0}(A_{\omega}) =
\operatorname{K}_{0}(\cat{D}^{\mrm{b}}(\cat{mod} A_{\omega}))$.
Let $\mbb{Z}^{\Delta_{0}}$ be the free abelian group with basis
$\{ e_{x} \}_{x \in \Delta_{0}}$. Then
$[P_{x, \omega}] \mapsto e_{x}$ determines a canonical isomorphism
$\operatorname{K}_{0}(A_{\omega}) \stackrel{\simeq}{\rightarrow} \mbb{Z}^{\Delta_{0}}$.
For a two-sided tilting complex
$T \in \operatorname{DPic}_{k}(\omega,\omega')$
let
$\chi_{0}(T) : \operatorname{K}_{0}(A_{\omega}) \stackrel{\simeq}{\rightarrow}
\operatorname{K}_{0}(A_{\omega'})$
be
$\chi_{0}(T)([M]) := [T \otimes^{\mrm{L}}_{A_{\omega}} M]$.
Using the projective bases we get a functor (when we consider a
group as a groupoid with a single object)
\[ \chi_{0} : \operatorname{DPic}_{k}(\Delta) \to
\operatorname{Aut}_{\mbb{Z}}(\mbb{Z}^{\Delta_{0}}) \cong
\operatorname{GL}_{n}(\mbb{Z}) . \]
Recall that for a vertex $x \in \Delta_{0}$ one defines the
reflection
$s_{x} \in \operatorname{Aut}_{\mbb{Z}}(\mbb{Z}^{\Delta_{0}})$
by
\[ \begin{aligned}
s_{x} e_{x} & := -e_{x} + \sum_{ \{x, y\} \in \Delta_{1}} e_{y}
, \\
s_{x} e_{y} & := e_{y} \quad \text{if } y \neq x .
\end{aligned} \]
The {\em Weyl group} of $\Delta$ is the subgroup
$W(\Delta) \subset \operatorname{Aut}_{\mbb{Z}}(\mbb{Z}^{\Delta_{0}})$
generated by the reflections $s_{x}$.
\begin{prop} \label{prop5.1}
Let $x$ be a source in the quiver $\vec{\Delta}_{\omega}$. Then
\[ \chi_{0}(T_{x, \omega}) = s_{x} . \]
\end{prop}
\begin{proof}
There is an Auslander-Reiten sequence
\[ 0 \to P_{x, \omega} \to \bigoplus_{(x \to y) \in
(\vec{\Delta}_{\omega})_{1}} P_{y, \omega} \to
\tau_{\omega}^{-1} P_{x, \omega} \to 0 \]
in $\cat{mod} A_{\omega}$. Applying the functor
$T_{x, \omega}^{\vee} \otimes^{\mrm{L}}_{A_{\omega}} - \cong
\operatorname{R} \operatorname{Hom}_{A_{\omega}}(T_{x, \omega}, -)$
to this sequence, and using formula (\ref{eqn5.1}), we get a
triangle
\[ T_{x, \omega}^{\vee} \otimes^{\mrm{L}}_{A_{\omega}} P_{x, \omega}
\to \bigoplus_{ \{x, y\} \in \Delta_{1}}
P_{y, s_{x}^{-} \omega}
\to P_{x, s_{x}^{-} \omega}
\to (T_{x, \omega}^{\vee} \otimes^{\mrm{L}}_{A_{\omega}}
P_{x, \omega})[1] \]
in $\cat{D}^{\mrm{b}}(\cat{mod} A_{s_{x}^{-} \omega})$.
Hence
\[ [T_{x, \omega}^{\vee}
\otimes^{\mrm{L}}_{A_{\omega}} P_{x, \omega}] =
-[P_{x, s_{x}^{-} \omega}] +
\sum_{ \{x, y\} \in \Delta_{1}} [P_{y, s_{x}^{-} \omega}]
\in \operatorname{K}_{0}(A_{s_{x}^{-} \omega}) . \]
On the other hand for $y \neq x$ we have
$[T_{x, \omega}^{\vee}
\otimes^{\mrm{L}}_{A_{\omega}} P_{y, \omega}] =
[P_{y, s_{x}^{-} \omega}]$.
This proves that
$\chi_{0}(T_{x, \omega}^{\vee}) = s_{x}$;
but $s_{x} = s_{x}^{-1}$.
\end{proof}
An immediate consequence is:
\begin{cor}
$\chi_{0}(\operatorname{Ref}(\Delta)) = W(\Delta)$.
\end{cor}
An ordering $(x_{1}, \ldots, x_{n})$ of $\Delta_{0}$
is called source-admissible for an orientation $\omega$ if
$x_{i}$ is a source in the quiver
$\vec{\Delta}_{s_{x_{i - 1}}^{-} \cdots s_{x_{1}}^{-} \omega}$
for all $1 \leq i \leq n$. Any orientation has source-admissible
orderings of the vertices.
\begin{prop} \label{prop5.2}
Let $(x_{1}, \ldots, x_{n})$ be a source-admissible ordering of
$\Delta_{0}$ for an orientation $\omega$. Write
$\omega_{i} := s_{x_{i}}^{-} \cdots s_{x_{1}}^{-} \omega$,
$A_i := A_{\omega_{i}}$ and
$T_{i} := T_{x_{i}, \omega_{i - 1}}$. Then
\[ T^{\vee}_{n} \otimes^{\mrm{L}}_{A_{n - 1}} \cdots
\otimes^{\mrm{L}}_{A_{2}} T^{\vee}_{2}
\otimes^{\mrm{L}}_{A_{1}} T^{\vee}_{1} \cong
A_{\omega}^{*}[-1] \]
in $\cat{D}^{\mrm{b}}(\cat{Mod} A_{\omega}^{\mrm{e}})$.
\end{prop}
\begin{proof}
For an orientation $\omega$ let
$\vec{\Gamma}^{\mrm{irr}}_{\omega} \subset
\vec{\Gamma}(\cat{D}^{\mrm{b}}(\cat{mod} A_{\omega}))$
be the quiver of definition \ref{dfn2.1}. As usual
$(\vec{\Gamma}^{\mrm{irr}}_{\omega})_{0}$ denotes the set of
vertices of $\vec{\Gamma}^{\mrm{irr}}_{\omega}$.
Let $G(\Delta)$ be the groupoid with object set
$\operatorname{Or}(\Delta)$, and morphism sets
$\operatorname{Iso}\bigl( (\vec{\Gamma}^{\mrm{irr}}_{\omega})_{0},
(\vec{\Gamma}^{\mrm{irr}}_{\omega'})_{0} \bigr)$
for $\omega, \omega' \in \operatorname{Or}(\Delta)$. The groupoid $G(\Delta)$
acts faithfully on the family of sets
$X(\Delta) := \{
(\vec{\Gamma}^{\mrm{irr}}_{\omega})_{0} \}_{\omega \in
\operatorname{Or}(\Delta)}$.
According to Theorem \ref{thm0.2} there is
an injective map of groupoids
$q : \operatorname{DPic}_{k}(\Delta) \rightarrowtail G(\Delta)$.
Let us first assume $\Delta$ is a Dynkin graph. Then there is a
canonical isomorphism of sets
$X(\Delta) \cong \mbb{Z} \times {\Delta}_{0} \times \operatorname{Or}(\Delta)$.
The action of $q(\tau_{\omega})$ on $X(\Delta)$ is
$q(\tau_{\omega})(i, x, \omega) = (i - 1, x, \omega)$. By formula
(\ref{eqn5.1}), the action of $q(T^{\vee}_{x, \omega})$ on
$X(\Delta)$ is
$q(T^{\vee}_{x, \omega})(0, y, \omega) =
(0, y, s_{x}^{-} \omega)$
if $y \neq x$, and
$q(T^{\vee}_{x, \omega})(1, x, \omega)
= (0, x, s_{x}^{-} \omega)$.
Since $q(\tau_{\omega})$ commutes with $q(T^{\vee}_{x, \omega})$
we have
\[ q(T^{\vee}_{n} \otimes^{\mrm{L}}_{A_{n - 1}} \cdots
\otimes^{\mrm{L}}_{A_{1}} T^{\vee}_{1})(i, x, \omega) =
(i - 1, x, \omega) = q(\tau_{\omega})(i, x, \omega) \]
for any $x \in \Delta_{0}$ and $i \in \mbb{Z}$.
If $\Delta$ is not Dynkin then
$X(\Delta) \cong \mbb{Z} \times \mbb{Z} \times
{\Delta}_{0} \times \operatorname{Or}(\Delta)$,
$q(\tau_{\omega})(j, i, x, \omega) = (j, i - 1, x, \omega)$,
etc., and the proof is the same after these modifications.
\end{proof}
\begin{prop} \label{prop5.3}
For any orientation $\omega$,
\[ \operatorname{Ref}(\Delta)(\omega, \omega) = \bra{\tau_{\omega}} . \]
\end{prop}
\begin{proof}
We will only treat the Dynkin case; the general case is proved
similarly with modifications like in the previous proof.
Let $T \in \operatorname{Ref}(\Delta)(\omega, \omega)$. From the proof
above we see that
$q(T)(0, x, \omega) = (i(x), x, \omega)$
for some $i(x) \in \mbb{Z}$. A quiver map
$\pi : \vec{\Delta}_{\omega} \to \vec{\mbb{Z}} \vec{\Delta}_{\omega}$
with $\pi(x) = (i(x), x)$ must have $i(x) = i$ for all $x$,
since $\Delta$ is a tree. Therefore $q(T) =
q(\tau_{\omega}^{-i})$.
\end{proof}
\begin{rem}
The explicit calculations in Section 4 show that the shift
$\sigma = A[1]$ is not in
$\bra{\tau} \subset \operatorname{DPic}_{k}(A)$
for most algebras $A$. Thus
$\operatorname{Ref}(\Delta) \subsetneqq \operatorname{DPic}_{k}(\Delta)$
for most graphs $\Delta$.
\end{rem}
|
\section{Introduction}
\subsection{Representations from quantized symplectic reduction}
Constrained quantization \ci{Dir,MT} is a very useful method that often
allows one to reduce nonlinear problems in mathematical physics to
linear ones. Such a reduction is possible if a given nonlinear
(symplectic) space may be written as the reduced (`physical') phase
space relative to a linear phase space with certain constraints def\/ined
on it. The goal of this paper is to quantize the coadjoint orbits
of a certain inf\/inite-dimensional Lie group, which are highly
nonlinear inf\/inite-dimensional symplectic manifolds, by a mathematically
rigorous version of this method. The group in question (def\/ined
below) has been chosen because it is one of the few
inf\/inite-dimensional Lie groups for which the correspondence between its
irreducible unitary representation s and its coadjoint orbits is known.
Thus it forms an ideal testing ground for the constrained quantization
(as well as for more general constructions in mathematical physics)
of inf\/inite-dimensional phase spaces.
Let $U_0(\H)$ be the Banach Lie group of all unitary operators $U$ on
a separable Hilbert space $\H$ for which $U-{\mathbb I}$ is compact,
equipped with the uniform operator (i.e., norm) topology. The
continuous unitary representation s of $U_0(\H)$ were classif\/ied by Kirillov
\ci{Kir1} and Ol'shanskii \ci{Ols1}. Their classif\/ication
simultaneously applies to the Fr\'{e}chet Lie group $U(\H)$ consisting
of all unitary operators on $\H$, equipped with the strong operator
topology, because all representation s of $U_0(\H)$ are also strongly
continuous, and can therefore be extended to $U(\H)$. Moreover,
$U(\H)$ re-topologized with the uniform topology has the same
irreducible representation s on separable Hilbert spaces as the same group
equipped with the strong topology (whose irreducible representation\ spaces are
automatically separab\-le)~\ci{Pic}. (The representation\ theory of $U(\infty)$
equipped with the inductive limit topology is much more complicated
\ci{Ols4,Boy2} and will not be discussed here.)
A remarkable aspect of the Kirillov-Ol'shanskii classif\/ication is that
all irreducible representation s of $U_0(\H)$ may be thought of as the geometric
quantization of certain of its coadjoint orbits. However, only the
geometric quantization of orbits corresponding to positive eigenvalues
may actually be found in the literature \ci{Boy1}; even this special
case is already fairly involved. It is this quantization that we
venture to redo, and much simplify, by regarding the orbits as
Marsden-Weinstein quotients, and performing a certain constrained
quantization procedure.
Our work was triggered by Montgomery's observation \ci{Mon} (also cf.\
\ci{LMS}) that for f\/inite-dimensional $\H={\mathbb C}} \newcommand{\diri}{\int^{\oplus}^k$ certain coadjoint
orbits of $U(k)$ (namely those characterized by positive eigenvalues)
are Marsden-Weinstein\ quotients of $\H\otimes {\mathbb C}} \newcommand{\diri}{\int^{\oplus}^M$ with respect to $U(M)$, for
suitable $M$ (which depends on the orbit). The left-action of $U(k)$
and the right-action of $U(M)$ on ${\mathbb C}} \newcommand{\diri}{\int^{\oplus}^k\otimes{\mathbb C}} \newcommand{\diri}{\int^{\oplus}^M$ combine to form a
Weinstein dual pair $U(k)\raw {\mathbb C}} \newcommand{\diri}{\int^{\oplus}^k\otimes{\mathbb C}} \newcommand{\diri}{\int^{\oplus}^M\leftarrow U(M)$
\ci{KKS,Ste,Wei83}.
The simplest reduced space thus obtained (viz.\ for $M=1$) is the
projective space $\mathbb P {\mathbb C}} \newcommand{\diri}{\int^{\oplus}^k$; as in the general case, three relevant
symplectic structures, namely its standard form as a K\"{a}hler
manifold, its Lie-Poisson form as a coadjoint orbit, and f\/inally its
Marsden-Weinstein form as a symplectic quotient, all coincide.
We extend Montgomery's result to the situation where the eigenvalues
may be of either sign, and also to the case where $\H$ is
inf\/inite-dimensional. The Weinstein dual pair then~be\-co\-mes
$U_0(\H)\raw \H\otimes{\mathbb C}} \newcommand{\diri}{\int^{\oplus}^M\otimes\overline{{\mathbb C}} \newcommand{\diri}{\int^{\oplus}}^N\leftarrow U(M,N)$, so that
one reduces with respect to the non-com\-pact group $U(M,N)$. Note that
$M$ and $N$ are f\/inite even in the inf\/inite-dimensional case.
The quantization of the `unconstrained system' $U_0(\H)\raw
\H\otimes{\mathbb C}} \newcommand{\diri}{\int^{\oplus}^M\otimes\overline{{\mathbb C}} \newcommand{\diri}{\int^{\oplus}}^N\leftarrow U(M,N)$ is trivially done by
Fock space techniques, yielding a Howe dual pair that quantizes the
clas\-sical Weinstein dual pair in question. To quantize the Marsden-Weinstein\
reduction pro\-cess that led to the classical coadjoint orbits, we
employ a relatively new method~\ci{NPL93,MT}, which is based on the
$C^*$-algebra ic technique of Rief\/fel induction \ci{Rie74,FD,MT}. As explained
in \ci{NPL93,MT}, this method in principle quantizes a symplectic
reduction procedure vastly more general than the Marsden-Weinstein\ one
\ci{MiW,Xu,NPL93,MT}, and improves on more traditional constrained
quantization techniques (such the Dirac or the BRST method) in cases
where the quantized constraints fail to have a joint eigenvalue
zero. In the context of the present paper, this means that for $N=0$,
where one classically reduces with respect to the compact group
$U(M)$, other techniques would apply as well, whereas for $N>0$ these
would break down.
For $N=0$, the induction procedure is easily carried out on the basis
of Weyl's classical results on tensor products and the symmetric group
\ci{Wey,How4}. The case $N>0$, where the coadjoint orbit one
quantizes is characterized by eigenvalues of arbitrary sign, is
considerably more complicated. The quantization of the unconstrained
system $S=\H\otimes{\mathbb C}} \newcommand{\diri}{\int^{\oplus}^M\otimes\overline{{\mathbb C}} \newcommand{\diri}{\int^{\oplus}}^N$ is known explicitly at least
for f\/inite-dimensional $\H={\mathbb C}} \newcommand{\diri}{\int^{\oplus}^k$: it is the $k$-fold tensor product of
the metaplectic (or `oscillator', or `Segal-Shale-Weil') representation\
\ci{Fol}, restricted from $Sp(2(N+M),{\mathbb R}} \newcommand{\bok}{{\mbox{{\bf k}}})$ to its subgroup $U(M,N)$
(see \ci{SW,Ste,BR}).
This tensor product has been decomposed by Kashiwara and Vergne
\ci{KV}, also cf.\ Howe~\ci{How2}. The decomposition of the Hilbert
space quantizing $S={\mathbb C}} \newcommand{\diri}{\int^{\oplus}^k\otimes {\mathbb C}} \newcommand{\diri}{\int^{\oplus}^M\otimes\overline{{\mathbb C}} \newcommand{\diri}{\int^{\oplus}}^N$ under $U(k)$ and
$U(M,N)$ does not ref\/lect the decomposition of $S$ under these group
actions if $k>M+N$ (which is the case of relevance to us, as we are
eventually interested in $k=\infty$), cf.\ \ci{Ada2}. This
fascinating complication implies that for generic coadjoint orbits our
method only works when $\H$ is f\/inite-dimensional.
\subsection{Rief\/fel induction for group actions}
We brief\/ly review how Rief\/fel induction \ci{Rie74,FD,MT} specializes
to the present context. One starts from a strongly Hamiltonian
right-action of a connected Lie group $H$ on a symplectic manifold
$S$, with accompanying equivariant momentum map $J:S\raw \h^*$. We
assume that the reduced space $S^{\mu}\equiv J^{-1}(\O_{\mu})/H$ is a
manifold.
If a Lie group $G$ acts symplectically on $S$ in such a way that its
action commutes with the $H$-action, the reduced space $S^{\mu}$
becomes a symplectic $G$-space in the obvious way; the well-known
`symplectic induction' procedure \ci{KKS,MT} is a special case of this
construction (it is obtained by taking $H\subset G$ and $S=T^*G$).
To quantize the reduced space $S^{\mu}$ and the associated induced
representation\ of $G$, we assume that a quantization of the unconstrained
system as well as of the constraints are given. Hence we suppose we
have f\/irstly found a Hilbert space $\F$, which may be thought of as
the (geometric) quantization of $S$. Secondly, a unitary right-action
(i.e., anti-representation) $U_R(H)$ on $\F$ should be given, which is the
quantization of the symplectic right-action of $H$ on $S$. Thirdly, we
require a unitary representation\ $\plc(H)$ on a Hilbert space $\hlc$, which
`quantizes' the coadjoint action of $H$ on the coadjoint orbit
$\O_{\mu}$ This is only possible if the orbit is `quantizable'; for
$H=U(M)$ there is a bijective correspondence between such orbits and
unitary representation s, and for $U(M,N)$ one obtains at least all unitary
highest weight modules by `quantizing' such orbits \ci{Ada1,Vog}. (In
the latter case the concept of quantization has to be stretched
somewhat to incorporate the derived functor technique to construct
representation s.)
First assuming that $H$ is compact, we construct the induced space
$\huc$ from these data as the subspace of $\F\otimes\hlc$ on which
$U_R^{-1}\otimes\plc$ acts trivially (here $U_R^{-1}$ is the representation\ of~$H$
def\/ined by $U_R^{-1}(h)=U_R(h^{-1})$). If $H$ is only locally compact
(and assumed unimodular for simplicity) with Haar measure $dh$, one
has to f\/ind a dense subspace $L\subset \F$ such that the integral
$\int_H dh\,( (U_R^{-1}\otimes\plc)(h)\Psi,\Phi)\equiv (\Psi,\Phi)_0$ is
f\/inite for all $\Psi,\Phi\in L\otimes\hlc$. This def\/ines a sesquilinear
form $(\cdot,\cdot)_0} \newcommand{\pco}{U_{\rm co}$ on $L\otimes\hlc$ which can be shown to be positive
semi-def\/inite under suitable conditions \ci{NPL93}. The induced space
$\huc$ is then def\/ined as the completion of the quotient of $L\otimes\hlc$
by the null space of $(\cdot,\cdot)_0} \newcommand{\pco}{U_{\rm co}$; its inner product is, of course, given by
the quotient of $(\cdot ,\cdot )_0$. For $H$ compact the integral
exists for all $\Psi,\Phi\in\F$ and $(\Psi,\Phi)_0=(P_0\Psi,P_0\Phi)$,
where $P_0$ is the projector onto the subspace of $\F\otimes\hlc$ carrying
the trivial representation\ of $H$, so that we recover the f\/irst description of
$\huc$.
We now assume that a group $G$ acts on $\F$ through a unitary representation\
$U_L$; it is required that this action commute with $U_R(H)$. The
induced representation\ $U^{\chi}} \newcommand{\hlc}{{\cal H}_{\chi}(G)$ on $\huc$ is now def\/ined as follows. For $H$
compact, $U^{\chi}} \newcommand{\hlc}{{\cal H}_{\chi}$ is simply the restriction of $U_L\otimes{\mathbb I}$ to
$\huc\subset \F\otimes\hlc$; this is well def\/ined because $U_L\otimes{\mathbb
I}$ commutes with $U_R^{-1}\otimes\plc$. In the general case, one has to
assume that $U_L$ leaves $L$ stable; then $U^{\chi}} \newcommand{\hlc}{{\cal H}_{\chi}$ is essentially
def\/ined as the quotient of the action of $U_L\otimes{\mathbb I}$ (on
$L\otimes\hlc$) to $\huc$ as def\/ined above (cf.\ \ci{NPL93} for technical
details pertinent to the general case). The Mackey induction
procedure for group representation s is recovered by assuming that $H\subset
G$, and taking $\F=L^2(G)$, cf.\ \ci{Rie74,FD,MT} for details in the
original setting of Rief\/fel induction, and \ci{NPL93,MT} for the
above setting.
\section{Representations from Rief\/fel induction}
In subsections 2.1 to 2.3 we take $\H$ to be an inf\/inite-dimensional
separable Hilbert space, unless explicitly stated otherwise. All
results (sometimes with self-explanatory modif\/ications) are equally
well valid in the f\/inite-dimensional case, which is considerably
easier to handle; we leave this to the reader. We start with the
simplest case, the def\/ining representation.
\subsection{The quantization of ${\mathbb P}{\cal H}$}
One can realize ${\mathbb P}{\cal H}$ as a Marsden-Weinstein\ quotient with respect to the group
$U(1)$ \ci{AM,MT}. Firstly, $\H$ carries a symplectic form $\omega} \newcommand{\Om}{\Omega$,
expressed in terms of the standard inner product (taken linear in the
f\/irst entry) by $\omega} \newcommand{\Om}{\Omega(\ps,\varphi} \newcommand{\ch}{\chi)=-2\, {\rm Im}\, (\ps,\varphi} \newcommand{\ch}{\chi)$. Secondly,
$U(1)$ (identif\/ied with the unit circle in the complex plane) acts on
$\H$ by $z:\ps\raw z\ps$; this action is symplectic, and yields an
equivariant momentum map \ci{AM} $J:\H\raw {\bf u(1)}^*\equiv {\mathbb R}} \newcommand{\bok}{{\mbox{{\bf k}}}$
given by $J(\ps)=(\ps,\ps)$. Then ${\mathbb P}{\cal H}\simeq J^{-1}(1)/U(1)$.
The quantization of this type of reduced space using Rief\/fel induction
was outlined in the Introduction. We f\/irst need a quantization of the
`unconstrained' system $\H$, which we take to be the symmetric
(bosonic) Fock space $\F=\exp(\H)$ (this is the direct sum of all
symmetrized tensor products $\H^{\otimes n}$ ($n=0,1,\ldots$) of $\H$ with
itself). This quantization is so well-established that we will not
motivate it here; cf.\ \ci{Fol,RS1} for mathematical aspects, and
\ci{Woo} for a derivation in geometric quantization.
The (anti) representation\ $U_R$ of $U(1)$ on $\F$ is obtained by
`quantization' of the right action on $\H$. We choose $U_R$ as the
second quantization of this right action. Labelling this choice
$U_{R,{\rm sq}}$, this yields $U_{R,{\rm sq}}(z) \upharpoonright} \newcommand{\plc}{U_{\chi} \H^{\otimes
n}=z^n{\mathbb I}$. Similarly, the def\/ining representation\ $U_{1}$ of $G=U(\H)$
(the group of all unitary operators on $\H$) on $\H_1=\H$ yields a
symplectic action on $\H$. This is `second' quantized by the representation\
$U_{L,{\rm sq}}$ on $\F$, whose restriction $U_n$ to each subspace
$\H^{\otimes n}\subset\F$ is the symmetrized $n$-fold tensor product of
$U_{1}$ with itself. The representation s $U_{R,{\rm sq}}(U(1))$ and $U_{L,{\rm
sq}}(U(\H))$ obviously commute with each other. Hence $\F$ has a
central decomposition under $U_{L,{\rm sq}}(U(\H))\otimes U^{-1}_{R,{\rm
sq}}(U(1))$, which is explicitly given by
\begin{equation}
\exp(\H)\stackrel{{\rm
sq}}{\simeq} \bigoplus_{n=0}^{\infty} \H_n^{U(\H)}\otimes
\overline{\H}_n^{U(1)}. \ll{dec1}
\end{equation}
Here $\H_n^{U(\H)}$ coincides
with $\H^{\otimes n}$, now regarded as the carrier space of the representation\
$U_n(U(\H))$, which is, in fact, irreducible for all $n$
\ci{Kir1,Ols1} (also cf.\ subsection 3.3 below). Also, ${\H}_n^{U(1)}$
is just ${\mathbb C}} \newcommand{\diri}{\int^{\oplus}$, but regarded as the carrier space of $U_n(U(1))$,
def\/ined by $U_n(z)=z^n$; $\overline{\H}$ stands for the carrier space
of the conjugate representation.
The general context for decompositions of the type (\ref{dec1}) is the
theory of Howe dual pairs \ci{How1,How3}. In the present instance,
this applies to $\H={\mathbb C}} \newcommand{\diri}{\int^{\oplus}^k$, with $U(k)$ and $U(1)$ being the dual pair
in $Sp(2k,{\mathbb R}} \newcommand{\bok}{{\mbox{{\bf k}}})$. (Cf. \ci{Ols4} for the theory of these pairs in the
inf\/inite-dimensional setting.)
The construction of the induced space $\F^1$ is ef\/fortless in this
case. The fact that Marsden-Weinstein\ reduction took place at $J=1$ means that
the orbit of $U(1)$ in question is the point $1\in {\bf
u(1)}^*$. This orbit is quantized by the def\/ining representation\ $U_1$ of
$U(1)$ on $\H_1={\mathbb C}} \newcommand{\diri}{\int^{\oplus}$. By construction, $\F^1$ is the subspace of
$\F\otimes \H_1=\F$ which is invariant under the representation\ $U_R^{-1}\otimes
U_1$. Hence (\ref{dec1}) implies that $\F^1=\H$. The induced representation\
$U^1(U(\H))$ on $\F^1$ is simply the restriction of $U_{L,{\rm
sq}}(U(\H))$ to this space, so that $U^1 \simeq U_{1} $. In other
words, we have recovered the def\/ining representation.
So far, so good, but unfortunately there is a subtlety if one derives
$U_R$ and $U_L$ from geometric quantization. Using the `uncorrected'
formalism (as described, e.g., in Ch.\ 9 of~\ci{Woo}), exploiting the
existence of an invariant positive totally complex polarization, viz.\
the anti-holomorphic one, one f\/inds that $\F$ is realized as the space
of holomorphic functions on $\H$. The quantization $\pi_{\rm qua}$ of the
momentum maps $J_R$ for $U(1)$ and $J_L$ for $U(\H)$ (with respect to
their respective actions on ${\mathbb C}} \newcommand{\diri}{\int^{\oplus}^k$) then reproduces the second
quantizations $U_{R,{\rm sq}}$ and $U_{L,{\rm sq}}$, respectively.
If, however, one is too sophisticated and incorporates the half-form
correction to geometric quantization \ci[Ch.\ 10]{Woo}, one obtains
extra contributions: for $\H={\mathbb C}} \newcommand{\diri}{\int^{\oplus}^k$, $\pi_{\rm qua}(J_R)$ is replaced by
$\pi_{\rm qua}(J_R)+k/2$, whereas $\pi_{\rm qua}(J_L)$ acquires an additional constant
$\mbox{\footnotesize $\frac{1}{2}$}$ (times the unit matrix). These Lie algebra representation s exponentiate
to unitary representation s of double covers $\tilde{U}(k)$ and $\tilde{U}(1)$,
which we denote by $U_{L,{\rm hf}}$ and $U_{R,{\rm hf}}$,
respectively. Under $U_{L,{\rm hf}}(\tilde{U}(k))\otimes U^{-1}_{R,{\rm
hf}}(\tilde{U}(1))$ we then f\/ind the central decomposition
\begin{equation}
\exp(\H)\stackrel{{\rm hf}}{\simeq} \bigoplus_{n=0}^{\infty}
\H_{(n+\mbox{\footnotesize $\frac{1}{2}$},\mbox{\footnotesize $\frac{1}{2}$},\ldots,\mbox{\footnotesize $\frac{1}{2}$})}^{\tilde{U}(k)}\otimes
\overline{\H}_{n+\mbox{\footnotesize $\frac{1}{2}$} k}^{\tilde{U}(1)}. \ll{dec2}
\end{equation}
Here
$\H_{(n+\mbox{\footnotesize $\frac{1}{2}$},\mbox{\footnotesize $\frac{1}{2}$},\ldots,\mbox{\footnotesize $\frac{1}{2}$})}$ carries the representation\ of $\tilde{U}(k)$
with highest weight $ (n+\mbox{\footnotesize $\frac{1}{2}$},\mbox{\footnotesize $\frac{1}{2}$},\ldots,\mbox{\footnotesize $\frac{1}{2}$})$; this is the
tensor product of $\H_n$ and the square-root of the determinant
representation. One observes that the inclusion of half-forms is awkward for
Rief\/fel induction -- we defer a discussion of this point to Chapter 3.
\subsection{The coadjoint orbits of $U_0(\H)$ as reduced spaces}
The Lie algebra $\frak g} \newcommand{\F}{\frak F={\bf u}_0(\H)=i{\mathfrak K}(\H)_{\rm sa}$ of
$G=U_0(\H)$ consists of all skew-adjoint compact operators on $\H$
with the norm topology. The dual $\frak g} \newcommand{\F}{\frak F^*={\bf u}_0(\H)^*$ is the space
of all self-adjoint trace-class operators on $\H$, with topology
induced by the trace norm $\parallel} \newcommand{\h}{{\bf h} \rho} \newcommand{\sg}{\sigma} \newcommand{\ta}{\tau\parallel} \newcommand{\h}{{\bf h}_1={\rm Tr}\, |\rho} \newcommand{\sg}{\sigma} \newcommand{\ta}{\tau|$ (this
coincides with the weak$\mbox{}^*$ topology). The pairing is given by
$\langle} \newcommand{\ra}{\rangle\rho} \newcommand{\sg}{\sigma} \newcommand{\ta}{\tau,X\ra=i\, {\rm Tr}\, \rho} \newcommand{\sg}{\sigma} \newcommand{\ta}{\tau X$.
The coadjoint action of $U_0(\H)$ on ${\mathbf u}_0({\cal H})^*$ is given by
$\pco(U)\rho} \newcommand{\sg}{\sigma} \newcommand{\ta}{\tau=U\rho} \newcommand{\sg}{\sigma} \newcommand{\ta}{\tau U^*$. We are interested in those coadjoint orbits
which are `quantizable' in the sense of geometric quantization, since
their quantization should produce all irreducible representation s of $U_0(\H)$
\ci{Kir1,Kir2}. Each such orbit is labeled by a pair $({\sf m}, {\sf
n})$, where ${\sf m}$ is an ordered $M$-tuple of positive integers
satisfying $m_1\geq m_2\geq\ldots m_M>0$, and $\sf n$ is a similar
$N$-tuple ($M,N<\infty$). The coadjoint orbit ${\cal O}_{\sf m,n}$ consists of all
elements of ${\mathbf u}_0({\cal H})^*$ with eigenvalues
$m_1,m_2,\ldots,m_M,0^{\infty},-n_N,\ldots,-n_1$. The degeneracy of
each numerical eigenvalue $m_i$ (or $-n_j$) is simply the number of
times it occurs in this list. The explicit quantization of the orbits
${\cal O}_{\sf m,n}$ is not discussed in \ci{Kir1,Kir2}; the case where either $\sf
m$ or $\sf n$ is empty is done in \ci{Boy1} using geometric
quantization.
For f\/inite-dimensional $\H$, it was shown by Montgomery \ci{Mon} that
$\O_{{\sf m},0}$ can be written as a Marsden-Weinstein\ reduced space with respect
to the natural right-action of $U(M)$ on $\H\otimes{\mathbb C}} \newcommand{\diri}{\int^{\oplus}^M$. This is a
special instance of the theory of dual pairs. With $\H={\mathbb C}} \newcommand{\diri}{\int^{\oplus}^k$, the
groups $U(\H)$ and $U(M)$ form a Howe dual pair inside the symplectic
group $Sp(2kM,{\mathbb R}} \newcommand{\bok}{{\mbox{{\bf k}}})$ \ci{How1,Ste,How3}, and the momentum maps $J_R$ and
$J_L$ introduced below build a Weinstein dual pair, cf.\
\ci{KKS,Wei83}. General theorems on the connection between coadjoint
orbits of one group and Marsden-Weinstein\ reduced spaces with respect to the other
group in a dual pair are given in \ci{LMS}. We will now generalize the
special case mentioned above to inf\/inite-dimensional $\H$, and general
orbits ${\cal O}_{\sf m,n}$.
We take $S=\H\otimes{\mathbb C}} \newcommand{\diri}{\int^{\oplus}^{M+N}$, which we regard as a Hilbert manifold in
the obvious way. We choose the canonical basis
$\{e_i\}_{i=1,\ldots,M+N}$ in ${\mathbb C}} \newcommand{\diri}{\int^{\oplus}^{M+N}$. The symplectic form $\omega} \newcommand{\Om}{\Omega$ on
$S$ is taken as (we put $\hbar} \newcommand{\cpn}{{\mathbb CP}^n =1$)
\begin{equation}
\omega} \newcommand{\Om}{\Omega(\ps,\varphi} \newcommand{\ch}{\chi)=-2\, {\rm Im}\,\left
( \sum_{i=1}^M(\ps_i,\varphi} \newcommand{\ch}{\chi_i)-\sum_{i=M+1}^{M+N}(\ps_i,\varphi} \newcommand{\ch}{\chi_i)\right) ,
\ll{ommn}
\end{equation}
where we have expanded $\ps=\sum_i\ps_i\otimes e_i$ and
similarly for $\varphi} \newcommand{\ch}{\chi$. It is convenient to introduce an indef\/inite
sesquilinear form on ${\mathbb C}} \newcommand{\diri}{\int^{\oplus}^{M+N}$ by putting $(e_i,e_j)=\pm \dl_{ij}$,
with a plus sign for $i=1,\ldots,M$ and a minus sign for
$i=M+1,\ldots, M+N$. Together with the inner product on $\H$ this
induces an indef\/inite form $(\cdot,\cdot)_S$ on $S$ in the obvious
(tensor product) way. The right-hand side of (\ref{ommn}) then simply
reads $-2\, {\rm Im}\, (\ps,\varphi} \newcommand{\ch}{\chi)_S$. A simple trick shows that $S$ is
strongly symplectic: we can regard $S$ as a Hilbert space
$\H\otimes{\mathbb C}} \newcommand{\diri}{\int^{\oplus}^M\oplus \overline{\H\otimes{\mathbb C}} \newcommand{\diri}{\int^{\oplus}^N}$, with inner product
$(\ps,\varphi} \newcommand{\ch}{\chi)_{\mbox{\footnotesize\rm trick}}=\sum_{i=1}^M(\ps_i,\varphi} \newcommand{\ch}{\chi_i)+
\sum_{i=M+1}^{M+N}(\varphi} \newcommand{\ch}{\chi_i,\ps_i)$. Then $\omega} \newcommand{\Om}{\Omega(\ps,\varphi} \newcommand{\ch}{\chi)=-2\, {\rm
Im}\,(\ps,\varphi} \newcommand{\ch}{\chi)_{\mbox{\footnotesize\rm trick}}$, and the claim follows from
the well-known fact that Hilbert spaces are strongly symplectic~\ci{AM}.
The Lie group $H=U(M,N)$ (which is $U(M)$ or $U(N)$ for $\sf n$ or
$\sf m$ empty) acts on $S$ from the right in the obvious way, i.e., by
$U\raw{\mathbb I}\otimes U^T$. This action is symplectic, with anti-equivariant
momentum map $J_R:S\raw (\h^*)^-$. If we identify $X\in \h$ with a
generator in the def\/ining representation\ of $H$ on ${\mathbb C}} \newcommand{\diri}{\int^{\oplus}^{M+N}$, we obtain (cf.\
\ci[p.\ 501]{KKS})
\begin{equation}
\langle} \newcommand{\ra}{\rangle J_R(\ps),X\ra = i({\mathbb I}\otimes X^T\ps,\ps)_S.
\ll{jr}
\end{equation}
On a suitable Cartan subalgebra $\mathfrak t$ of $\h$, which
we identify as the set of imaginary diagonal operators on ${\mathbb C}} \newcommand{\diri}{\int^{\oplus}^{M+N}$,
with basis $H_j=-iE_{jj}$, this simply reads $\langle} \newcommand{\ra}{\rangle J_R(\ps),H_j\ra =
\pm (\ps_j,\ps_j)$ with a plus sign for $j=1,\ldots,M$ and a minus
sign for $j=M+1,\ldots, M+N$.
We now identify $\wmn$ with an element of $\h^*$ by the pairing $\langle} \newcommand{\ra}{\rangle
\wmn,X\ra=i{\rm Tr}\, D_{\wmn}X$, where $ D_{\wmn}$ is the diagonal
matrix in $M_{M+N}({\mathbb C}} \newcommand{\diri}{\int^{\oplus})$ with entries $m_1,\ldots, m_M,
-n_N,\ldots,-n_1$. This means that $\wmn$ def\/ines a dominant integral
weight on $\mathfrak t$, and vanishes on its complement. The subset
$J_R^{-1}(\wmn)$ of $S$ consists of those vectors $\ps=\sum_i\ps_i\otimes
e_i$ for which $(\ps_i,\ps_i)=m_i$ for $i=1,\ldots, m$, and
$(\ps_{M+j},\ps_{M+j})=n_{N+1-j}$ for $j=1,\ldots, n$, with the
$\ps_k$'s mutually orthogonal. The normalizations come from $J_R$
evaluated on $\mathfrak t$, and the orhtogonality derives from the
constraint that $J_R$ vanish on its complement. {\em Note that the
integrality of the $m_i$ and $n_j$ plays no role in this subsection.}
\begin{lemma}
$J_R^{-1}(\wmn)$ is a submanifold of $S$. \ll{subm}
\end{lemma}
{\em Proof.} According to the theorem on p.\ 550 of \ci{AMP}, we need
to show that $J_R: J_R^{-1}(\wmn)\raw \h^*$ is a submersion, which is
the case if at any point $\ps\in J_R^{-1}(\wmn)\subset S$ the
derivative $(J_R)_*\equiv J_R^{(1)}:T_{\ps}S\raw
T_{J_R(\ps)}\h^*\simeq \h^*$ is surjective and has a complementable
kernel. The former is equivalent to the statement that $\ps$ is a
regular value of the momentum map \ci{AM}. The derivative at $\ps\in
S$ follows from (\ref{jr}) as
\begin{equation} \langle} \newcommand{\ra}{\rangle (J_R^{(1)})_{\ps}(\xi),X\ra= 2
{\rm Re}\, ({\mathbb I}\otimes iX^T\xi,\ps)_S. \ll{derjr}
\end{equation}
This formula shows
that $J_R^{(1)}$ is continuous, so that its kernel is closed. The
complementability of this kernel is then immediate, since $S$ is a
Hilbert manifold. The surjectivity of $J_R^{(1)}$ follows from
(\ref{derjr}) by inspection, but it is more instructive to derive it
from Prop.\ 2.11 (due to Smale) in \ci{Mar}. This states that $\ps$ is
a regular value of the momentum map if\/f the stability group
$H_{\ps}\subseteq H$ of $\ps$ is discrete. Now, as pointed out
earlier, $\ps=\sum_i \ps_i\otimes e_i \in J_R^{-1}(\wmn)$ implies that all
$\ps_i$ are nonzero are orthogonal, so that $H_{\ps}$ is just the
identity. \hfill $\blacksquare$
\medskip
The action of $H$ on $S$ is not proper unless $\sf m$ or $\sf n$ is
empty (in which case $H$ is compact). However:
\begin{lemma} The action of $H$ on $J_R^{-1}(\wmn)$ is proper.
\end{lemma}
{\em Proof.} Let $\ps^{(n)}\raw \ps$ in $S$; equivalently,
$\ps_i^{(n)}\raw \ps_i$ in $\H$ for all $i$. If $\{U^{(n)}\}$ is a
sequence in $H$ and $U^{(n)}\ps^{(n)}$ converges, the fact that for
each $n$ all $\ps_i^{(n)}$ are nonzero and orthogonal implies that
$\{U_{ij}^{(n)}e_j\}$ must converge in ${\mathbb C}} \newcommand{\diri}{\int^{\oplus}^{M+N}$ for each $i$. Since
convergence in the topology on $U(M,N)$ is given by convergence of all
matrix elements in the def\/ining representation, this implies that
$\{U^{(n)}\}$ must converge in $H$. \hfill $\blacksquare$
\medskip
By the standard theory of Marsden-Weinstein\ reduction \ci{Mar74,AM}, these lemmas
imply that the reduced space
\begin{equation}
S^{\wmn}=J_R^{-1}(\wmn)/H_{\wmn}
\ll{sred}
\end{equation}
(where $H_{\wmn}$ is the stability group of
$\wmn\in\h^*$ under the coadjoint action) is a smooth symplectic
manifold. We will proceed to show that it is symplectomorphic to the
coadjoint orbit ${\cal O}_{\sf m,n}\in\frak g} \newcommand{\F}{\frak F^*$, where $G=U_0(\H)$, as explained above.
The required dif\/feomorphism is given by a quotient of the momentum map
$J_L:S\raw \frak g} \newcommand{\F}{\frak F^*$ def\/ined from the natural left-action of $G$ on $S$,
which action is evidently symplectic. Identifying $\frak g} \newcommand{\F}{\frak F$ with the space
of compact skew-adjoint operators $Y$ on $\H$, one easily f\/inds that
this momentum map is given by
\begin{equation} -i\langle} \newcommand{\ra}{\rangle J_L(\ps),Y\ra = (Y\otimes {\mathbb I}
\ps,\ps)_S
=\sum_{i=1}^M(Y\ps_i,\ps_i)-\sum_{i=M+1}^{M+N}(Y\ps_i,\ps_i). \ll{jl}
\end{equation}
Since the left-$G$ action and the right-$H$ action commute, $J_L$
is invariant under $H$ (i.e., $J_L(\ps U)=J_L(\ps)$ for all $U\in H$
and $\ps\in \H$), so that $J_L$ (restricted to $J_R^{-1}(\wmn)$)
quotients to a well-def\/ined map $\tilde{J_L}:S^{\wmn}\raw{\cal O}_{\sf m,n}$. Once we have
shown that $\tilde{J_L}$ is a dif\/feomorphism, it will follow that it is
symplectic, because of the def\/inition of the symplectic structure on
$S^{\wmn}$ and the fact that $J_L$ is equivariant.
Generalizing a standard result in the root and weight theory for
compact Lie groups, see e.g.\ \ci{Kna}, we f\/irst note that the the
stability group of $\wmn\in \h^*$ under the coadjoint action is
$H_{\wmn}=\prod_l U(l)$, where $\sum l=M+N$, and the product is over
the multiplicities within either $\sf m$ or $\sf n$ in $\wmn$; this is
a subgroup of $U(M,N)$ in the obvious block-diagonal form. (For
example, if $\wmn =((2,1,1),(2,2,2))$ the stability group is
$U(1)\times U(2)\times U(3)$.) It then follows from (\ref{jl}) that
$\tilde{J_L}$ is a bijection onto ${\cal O}_{\sf m,n}$.
\begin{prop}
$\tilde{J_L}$ is smooth. \ll{s1}
\end{prop}
{\em Proof.} The manifold structure of ${\cal O}_{\sf m,n}$ is def\/ined by its
embedding in $\frak g} \newcommand{\F}{\frak F^*$, which is a Banach space in the trace-norm
topology (cf.\ the beginning of this section). The smoothness of
$\tilde{J_L}$ then follows from that of $J_L:J_R^{-1}(\wmn)\raw\frak g} \newcommand{\F}{\frak F^*$, since
the Lie group $H$ acts smoothly, freely, and properly on
$J_R^{-1}(\wmn)$.
{\em 1. Continuity of $J_L$.} We prove continuity on all of $S$. As a
map between separable metric spaces ($S$ is separable because $\H$ is
by assumption, and $\frak g} \newcommand{\F}{\frak F^*$ is separable because the f\/inite-rank
operators are dense in it), $J_L$ is continuous if $\ps^{(n)}\raw \ps$
in $S$ implies $J_L(\ps^{(n)})\raw J_L(\ps)$ in $\frak g} \newcommand{\F}{\frak F^*$. The topology
on $\frak g} \newcommand{\F}{\frak F^*$ coincides with the weak$\mbox{}^*$-topology, so the desired
continuity follows from (\ref{jl}), the boundedness of $Y$, and
Cauchy-Schwartz.
{\em 2. Existence and continuity of $J_L^{(1)}$.} The derivative of
$J_L$ at $\ps$ is given by
\begin{equation} \langle} \newcommand{\ra}{\rangle (J_L^{(1)})_{\ps}(\xi),Y\ra= 2
{\rm Re}\, \left(
\sum_{i=1}^M(iY\xi_i,\ps_i)-\sum_{i=M+1}^{M+N}(iY\xi_i,\ps_i)\right).
\ll{derjl}
\end{equation}
By the same reasoning as in the previous item,
$(J_L^{(1)})_{\ps}$ lies in ${\cal L}(S,\frak g} \newcommand{\F}{\frak F^*)$ and is continuous.
The second derivative $J_L^{(2)}:S\times S\raw \frak g} \newcommand{\F}{\frak F^*$ can be read of\/f
from (\ref{derjl}); its existence and continuity are established as
before. Higher derivatives vanish. \hfill $\blacksquare$
\begin{prop}
$\tilde{J_L}^{-1}$ is smooth. \ll{s2}
\end{prop}
{\em Proof.} We pick an arbitrary point $\rho} \newcommand{\sg}{\sigma} \newcommand{\ta}{\tau_0\in{\cal O}_{\sf m,n}$, with stability
group $G_0$. Let $\H=\oplus_l\H_l$ be the decomposition of $\H$ under
which $\rho} \newcommand{\sg}{\sigma} \newcommand{\ta}{\tau_0$ is diagonal (the dimension of each $\H_0$ is the
degeneracy of the corresponding eigenvalue; this dimension is f\/inite
unless the eigenvalue is 0). Then $G_0=\oplus_l U_0(\H_l)$, in
self-evident notation. The Lie algebra $\frak g} \newcommand{\F}{\frak F_0$ of $G_0$ is given by
those operators in $\frak g} \newcommand{\F}{\frak F= i{\mathfrak K}(\H)_{\rm sa}$ which commute with
$\rho} \newcommand{\sg}{\sigma} \newcommand{\ta}{\tau_0$. The manifold ${\cal O}_{\sf m,n}$ is modelled on $\frak g} \newcommand{\F}{\frak F/\frak g} \newcommand{\F}{\frak F_0$. This has the
quotient topology inherited from $\frak g} \newcommand{\F}{\frak F$, i.e., the trace-norm topology
determined by $\parallel} \newcommand{\h}{{\bf h} A\parallel} \newcommand{\h}{{\bf h}_1={\rm Tr}\, |A|$.
We def\/ine a neighbourhood $V_0\subset {\cal O}_{\sf m,n}$ of $\rho} \newcommand{\sg}{\sigma} \newcommand{\ta}{\tau_0$ as
follows. Since $G$ is a Banach-Lie group, by \ci{LT} there exists a
neighbourhoud $V$ of $0\in\frak g} \newcommand{\F}{\frak F$ such that $\exp$ is a dif\/feomorphism on
$V$ into $\frak g} \newcommand{\F}{\frak F$. We put $V_0=\{\pco(\exp(A))\rho} \newcommand{\sg}{\sigma} \newcommand{\ta}{\tau_0| A\in V\}$ (recall
that the coadjoint action is given by $\pco(U)\rho} \newcommand{\sg}{\sigma} \newcommand{\ta}{\tau=U\rho} \newcommand{\sg}{\sigma} \newcommand{\ta}{\tau U^*$). To
def\/ine a chart on $V_0$, we f\/irst show that $\frak g} \newcommand{\F}{\frak F$ (equipped with the
trace-norm topology) admits a splitting $\frak g} \newcommand{\F}{\frak F=\frak g} \newcommand{\F}{\frak F_0\oplus {\bf m}} \newcommand{\ghh}{\Gamma_{\rm hol}({\sf H})_0$. Here
${\bf m}} \newcommand{\ghh}{\Gamma_{\rm hol}({\sf H})_0$ consists of those operators~$A$ in $\frak g} \newcommand{\F}{\frak F$ whose matrix elements
$(A\ps,\varphi} \newcommand{\ch}{\chi)$ vanish if both $\ps$ and $\varphi} \newcommand{\ch}{\chi$ lie in the same space
$\H_l$, for all $l$. It is clear that $\frak g} \newcommand{\F}{\frak F=\frak g} \newcommand{\F}{\frak F_0\oplus {\bf m}} \newcommand{\ghh}{\Gamma_{\rm hol}({\sf H})_0$ as a set,
and it quickly folows that each summand is closed: since $\parallel} \newcommand{\h}{{\bf h} A\parallel} \newcommand{\h}{{\bf h}\leq
\parallel} \newcommand{\h}{{\bf h} A\parallel} \newcommand{\h}{{\bf h}_1$, the uniform topology is weaker than the trace-norm one, so
that closedness in the former implies the corresponding property in
the latter topology. As to the uniform closedness of $g_0$, one has
$\parallel} \newcommand{\h}{{\bf h} [A,\rho} \newcommand{\sg}{\sigma} \newcommand{\ta}{\tau_0]\parallel} \newcommand{\h}{{\bf h}\,\leq 2 \parallel} \newcommand{\h}{{\bf h} A\parallel} \newcommand{\h}{{\bf h}\; \parallel} \newcommand{\h}{{\bf h} \rho} \newcommand{\sg}{\sigma} \newcommand{\ta}{\tau_0\parallel} \newcommand{\h}{{\bf h}$, so that $\frak g} \newcommand{\F}{\frak F_0\ni A_n\raw
A$ implies that $A\in \frak g} \newcommand{\F}{\frak F_0$. On ${\bf m}} \newcommand{\ghh}{\Gamma_{\rm hol}({\sf H})_0$ an even more elementary
inequality does the job. Thus $\frak g} \newcommand{\F}{\frak F/\frak g} \newcommand{\F}{\frak F_0\simeq {\bf m}} \newcommand{\ghh}{\Gamma_{\rm hol}({\sf H})_0$, and we may use
${\bf m}} \newcommand{\ghh}{\Gamma_{\rm hol}({\sf H})_0$ as a modelling space for ${\cal O}_{\sf m,n}$.
We def\/ine a chart on $V_0$ by $\varphi} \newcommand{\ch}{\chi_0:V_0\raw {\bf m}} \newcommand{\ghh}{\Gamma_{\rm hol}({\sf H})_0$, given by
$\varphi} \newcommand{\ch}{\chi_0(\pco(\exp(A))\rho} \newcommand{\sg}{\sigma} \newcommand{\ta}{\tau_0)=A_0$, where $A_0$ is the component of
$A\in\frak g} \newcommand{\F}{\frak F$ in ${\bf m}} \newcommand{\ghh}{\Gamma_{\rm hol}({\sf H})_0$. We would like to model $S^{\wmn}$ on ${\bf m}} \newcommand{\ghh}{\Gamma_{\rm hol}({\sf H})_0$ as
well, but this is not directly possible because it has the wrong
topology. Hence the following detour. Take a $\ps_0\in
J_R^{-1}(\wmn)\subset S$ for which $J_L(\ps_0)=\rho} \newcommand{\sg}{\sigma} \newcommand{\ta}{\tau_0$. Using the fact
that $J_L$ is a bijection, we model $S^{\wmn}=J_R^{-1}(\wmn)/H_{\wmn}$
on the closed linear subspace of $S$ given by $M_0=\{A\otimes{\mathbb I} \ps_0|A\in
{\bf m}} \newcommand{\ghh}{\Gamma_{\rm hol}({\sf H})_0\}$, equipped with the relative topology of $S$. Put
$W_0=\{\exp(A)\otimes{\mathbb I} \ps_0|A\in m_0\}\subset S$. If $pr:J_R^{-1}\raw
J_R^{-1}(\wmn)/H_{\wmn}$ is the canonical projection, we have a chart
on the neighbourhood $pr(W_0)$ of $pr(\ps_0)$ def\/ined by
$\phi} \newcommand{\Ph}{\Phi_0:pr(W_0)\raw M_0$ given by $\phi} \newcommand{\Ph}{\Phi_0(pr(\exp(A)\ps_0))=A\ps_0$.
This procedure respects the manifold structure of $S^{\wmn}$, which by
def\/inition is quotiented from $J_R^{-1}(\wmn)\subset S$.
We now def\/ine
$\mbox{}_0\tilde{J_L}^{-1}=\phi} \newcommand{\Ph}{\Phi_0\circ\tilde{J_L}^{-1}\circ\varphi} \newcommand{\ch}{\chi_0^{-1}$; this is a
map from $\varphi} \newcommand{\ch}{\chi_0 (V_0)\subset {\bf m}} \newcommand{\ghh}{\Gamma_{\rm hol}({\sf H})_0$ to $\phi} \newcommand{\Ph}{\Phi_0\circ pr(W_0)\subset
M_0$. Clearly, $\mbox{}_0\tilde{J_L}^{-1}(A)=A\ps_0$. This immediately
implies that $\mbox{}_0\tilde{J_L}^{-1}$, and therefore $\tilde{J_L}^{-1}$, is
smooth. \hfill $\blacksquare$
To sum up, we have proved
\begin{theorem}
For any separable Hilbert space $\H$, the coadjoint orbit ${\cal O}_{\sf m,n}$ of
the group $U_0(\H)$ (which consists of all trace-class operators on
$\H$ with $M$ specif\/ic positive and $N$ specific negative eigenvalues)
is symplectomorphic to the Marsden-Weinstein\ quotient
$S^{\wmn}=J_R^{-1}(\wmn)/H_{\wmn}$ with respect to $S=\H\otimes{\mathbb C}} \newcommand{\diri}{\int^{\oplus}^{M+N}$
and the natural right-action of $H=U(M,N)$. \ll{omw}
\end{theorem}
\subsection{Representations induced from $U(M)$}
The representation s of $U_0(\H)$ were fully classif\/ied in \ci{Kir1,Ols1,Ols3}
(also cf.\ \ci{Kir2,Ols4,Boy2}). A~re\-mar\-kable fact is that $U_0(\H)$
is a type I group, so that all its factorial representation s are of the form
$U\otimes{\mathbb I}$ on $\H_{U}\otimes\H_{\rm mult}$, where $(U,\H_{U})$ is
irreducible. Each irreducible representation\ cor\-responds to an integral weight $\wmn$ of
the type specif\/ied above, where $M$ and $N$ are arbitrary (but
f\/inite). The carrier space ${\cal H}^{({\sf m,n})}$ is of the form ${\cal H}^{({\sf m,n})}={\cal H}^{{\sf m}}\otimes \overline{{\cal H}}^{{\sf n}}$,
and car\-ries the irreducible representation\ $U^{({\sf m,n})}} \newcommand{\pim}{U^{{\sf m}}=\pim\otimes\overline{U}^{{\sf n}}$. Here ${\cal H}^{{\sf m}}$ is the subspace
of $\otimes^M\H$ obtained by symmetrization according to the Young diagram
whose $k$-th row has length $m_k$, and $\overline{{\cal H}}^{{\sf n}}$ is the conjugate space of
$\H^{{\sf n}}$. The representation\ $\pim$ is the one given by the restriction
of the $M$-fold tensor product of the def\/ining representation\ of $U_0(\H)$ to
${\cal H}^{{\sf m}}$,~etc.
This is almost identical to the theory for f\/inite-dimensional
$\H={\mathbb C}} \newcommand{\diri}{\int^{\oplus}^k$ \ci{Wey,Zel} (which has the obvious restriction that
$M,N\leq k$); the only dif\/ference is that in the inf\/inite-dimensional
case ${\cal H}^{{\sf m}}\otimes\overline{{\cal H}}^{{\sf n}}$ is already irreducible. For $k<\infty$, on the other
hand, one needs to take the so-called Young product \ci{Zel} of ${\cal H}^{{\sf m}}$
and $\overline{{\cal H}}^{{\sf n}}$ rather than the tensor product (this is the irreducible
subspace generated by the tensor product of the highest-weight vectors
in each factor); moreover, the use of conjugate spaces may be avoided
in that case by tensoring with powers of the determinant representation. For
example, ${\mathbb C}} \newcommand{\diri}{\int^{\oplus}^k\otimes \overline{{\mathbb C}} \newcommand{\diri}{\int^{\oplus}}^k$ contains the irreducible subspace
$\sum_{i=1}^k e_i\otimes \overline{e_i}$ which does not lie in the Young
product; for $k=\infty$ this subspace evidently no longer exists. For
$M=0$ or $N=0$ there is no dif\/ference whatsoever.
We will now show how the representation s $(\pim,{\cal H}^{{\sf m}})$ can be obtained by
Rief\/fel induction; the representation s $(\overline{U}^{{\sf n}},\overline{{\cal H}}^{{\sf n}})$ may then be constructed
similarly. This will quantize the coadjoint orbits $\O_{\sf m}\equiv
\O_{({\sf m},\emptyset)} $ and $\O^-_{\sf n}\equiv\O_{(\emptyset,{\sf
n})}$, respectively. We note that $\O^-_{\sf n}$ is $\O_{\sf n}$ with
the sign of the symplectic form changed; this relative minus sign
corresponds to the passage from $\H$ to $\overline{\H}$ upon
quantization.
Our starting point is Theorem \ref{omw}, in which we take
$S=\H\otimes{\mathbb C}} \newcommand{\diri}{\int^{\oplus}^M$, with $H=U(M)$ acting on $S$ from the right and
$G=U_0(\H)$ acting from the left in the natural way; we call these
actions $U_1^T(H)$ and $U_1(G)$, respectively. As explained in the
Introduction, we f\/irst have to quantize $S$ and the group actions
def\/ined on it. We do so by taking the bosonic second quantization, or
symmetric Fock space, $\F=\exp(S)$ over $S$ \ci{RS1,Woo}, cf.\
subsection~2.1. For later use, we equivalently def\/ine this as the
subspace of $\sum_{n=0}^{\infty} \otimes^n S$ on which the natural representation\
of the symmetric group $S_n$ on $\otimes^n S$ acts trivially for all $n$.
As in the $M=1$ case (cf.\ subsection 2.1) we f\/irst investigate the
representation s of $U_0(\H)$ and $U(k)$ on $\F$ obtained by second
quantization, or equivalenty, by geometric quantization without the
half-form modif\/ication. This goes as follows. The groups $H$ and $G$
act on each subspace $\otimes^n S$ by the $n$-fold tensor product of their
respective actions on $S$, and these actions restrict to $\F$. Thus
the actions $U_1^T(H)$ (which we turn into a representation\ by taking the
inverse) and $U_1(G)$ on $S$ are quantized by the unitary representation s
$\Gamma\overline{U}_1(H)$ ($=U_{R,{\rm sq}}^{-1}(H)$ in the notation
of subsection 2.1, and $U_R^{-1}(H)$ in that of the Introduction) and
$\Gamma U_1(G)$ ($=U_{L, {\rm sq}}(G)$), respectively (note that
$U^T_1(h^{-1})=\overline{U}_1(h)$). Here $\Gamma$ is the second
quantization functor \ci{RS1}. This setup, and the associated central
decomposition of $\F$ under these group actions, illustrate Howe's
theory of dual pairs \ci{How1,How2,How3} in an inf\/inite-dimensional
setting, cf.\ \ci{Ols4}.
The fact that the coadjoint orbit $\O_{\sf m}$ of $G$ is
(symplectomorphic to) the Marsden-Weinstein\ quotient of $S$ with respect to ${\sf
m}\in\h^*$, cf.\ Theorem \ref{omw}, should now be ref\/lected, or rather
quantized, by constructing the unitary representation\ $\pim(G)$ (which
according to Kirillov is attached to $\O_{\sf m}$) by Rief\/fel
induction from the representation\ $U_{{\sf m}}} \newcommand{\pll}{U_{{\sf l}}(H)$ attached to the orbit through $\sf
m$ in $H$. Here $U_{{\sf m}}} \newcommand{\pll}{U_{{\sf l}}(U(M))$ is simply the unitary irreducible representation\ given by
the highest weight $\sf m$; it is realized on ${\cal H}_{{\sf m}}$, which is the
subspace of $\otimes^M {\mathbb C}} \newcommand{\diri}{\int^{\oplus}^M$ obtained by symmetrization according to the
Young diagram whose $k$-th row has length~$m_k$.
To f\/ind the carrier space of the induced representation\ $\pim(G)$ we merely
have to identify the subspace of $\F\otimes{\cal H}_{{\sf m}}$ which is invariant under
$\Gamma\overline{U}_1\otimesU_{{\sf m}}} \newcommand{\pll}{U_{{\sf l}}(H)$. This is very easy on the basis of
the following well-known facts \ci{Wey,Zel,How4}:
\begin{enumerate}
\item
The representation s of the symmetric group $S_n$ are self-conjugate; for any
irreducible representation\ $\pll(S_n)$, the tensor product $\pll\otimes\pll$ contains the
identity representation\ once, and $\pll\otimes U_{{\sf l}'}$ does not contain the identity
unless ${\sf l}={\sf l}'$. (Recall that the irreducible representation s of $S_n$ are
labelled by an $n$-tuple of integers ${\sf l}=(l_1,\ldots,l_n)$, where
$l_1\geq l_2\geq \ldots l_n\geq 0$ and $\sum_i l_i=n$.) The
collection of all such $n$-tuples $\sf l$ forms the dual $\hat{S}_n$.
\item
Any unitary irreducible representation\ $\pll(U(M))$ is given by an $M$-tuple ${\sf
l}=(l_1,\ldots,$ $l_M)$ of positive nondecreasing integers (possibly
zero), as in the preceding item, or by the conjugate $\overline{\pll}$
of such a representation. Then $\pll\otimes\overline{\pll}$ contains the identity
representation\ once, but the identity does not occur in any $\pll\otimesU_{{\sf l}'}$, or
in any $\pll\otimes\overline{U_{{\sf l}'}}$ unless in the latter case ${\sf
l}={\sf l}'$.
\item
The def\/ining representation\ of $S_n$ on $\otimes^n {\mathbb C}} \newcommand{\diri}{\int^{\oplus}^M$ commutes with the $n$-fold
tensor product of the conjugate of the def\/ining representation\ of $U(M)$, so
that one has the central decomposition
\begin{equation}
\otimes^n {\mathbb C}} \newcommand{\diri}{\int^{\oplus}^M \simeq \bigoplus_{{\sf l}'\in \hat{S}_n}
{\cal H}_{{\sf l}}^{S^n}\otimes \overline{\H}_{\sf l}^{U(M)}, \ll{cd1}
\end{equation}
where the prime (relevant only when $M<n$) on
the $\oplus$ indicates that the sum is only over those $n$-tuples $\sf
l$ for which $l_{M+1}=0$. Here ${\cal H}_{{\sf l}}^{S^n}$ is the carrier space of
$\pll(S^n)$, and $\overline{\H}_{\sf l}^{U(M)}$ is the carrier space
of the conjugate of the irreducible representation\ of $U(M)$ obtained by making $\sf l$ an
$M$-tuple by adding or removing zeros. (A simliar statement holds
without the conjugation, of course.)
\item
Similarly,
\begin{equation} \otimes^n \H \simeq \bigoplus_{{\sf l}\in \hat{S}_n}
{\cal H}_{{\sf l}}^{S^n}\otimes {\cal H}^{{\sf m}}, \ll{cd2}
\end{equation}
under the appropriate representation s of $S_n$
and $U_0(\H)$, where ${\cal H}^{{\sf m}}$ was introduced at the beginning of this
subsection (for $\H={\mathbb C}} \newcommand{\diri}{\int^{\oplus}^k$ this is equivalent to a classical result in
invariant theory, see e.g. \ci[4.3.3.9]{How4}).
\end{enumerate}
Now consider $\otimes^n (\H\otimes{\mathbb C}} \newcommand{\diri}{\int^{\oplus}^M)\simeq \otimes^n \H\,\otimes\, \otimes^n
{\mathbb C}} \newcommand{\diri}{\int^{\oplus}^M$. This carries the representation\ $U_n^{\H}\otimes U_n^{{\mathbb C}} \newcommand{\diri}{\int^{\oplus}^M}$ of $S_n$, where
$U_n^{\cal K}(S_n)$ is the natural representation\ on $\otimes^n {\cal K}$. Applying
items 4 and 3, and subsequently 1 above, we f\/ind that the subspace
$\otimes^n_s(\H\otimes{\mathbb C}} \newcommand{\diri}{\int^{\oplus}^M)\subset \otimes^n (\H\otimes{\mathbb C}} \newcommand{\diri}{\int^{\oplus}^M)$ which is invariant under
$S_n$ can be decomposed as
\begin{equation}
\bigotimes^n_s(\H\otimes{\mathbb C}} \newcommand{\diri}{\int^{\oplus}^M)\simeq
\bigoplus_{{\sf l}'\in \hat{S}_n} \H^{\sf l}\otimes \overline{\H}_{\sf
l}^{U(M)}, \ll{decohil}
\end{equation}
in the sense that the restriction $\otimes^n_s
(U_1(G)\otimes\overline{U}_1(H))$ of $\Gamma U_1(G)\otimes
\Gamma\overline{U}_1(H)$ (def\/ined on $\F=\exp(\H\otimes{\mathbb C}} \newcommand{\diri}{\int^{\oplus}^M)$) to
$\otimes^n_s(\H\otimes{\mathbb C}} \newcommand{\diri}{\int^{\oplus}^M)\subset \F$ decomposes as
\begin{equation}
\bigotimes^n_s (U_1(G)\otimes\overline{U}_1(H))\simeq
\bigoplus_{{\sf l}'\in \hat{S}_n}
U^{\sf l}(G)\otimes \overline{\pll}(H). \ll{decorep}
\end{equation}
We then apply
item 2 to conclude that the only subspace of $\F\otimes{\cal H}_{{\sf m}}$ which is
invariant under $\Gamma\overline{U}_1\otimesU_{{\sf m}}} \newcommand{\pll}{U_{{\sf l}}(H)$ corresponds to
$n=\sum_{i=1}^M m_i$ (where $m_i$ are the entries of the $M$-tuple
$\sf m$). Moreover, by (\ref{decohil}) this invariant subspace is
exactly ${\cal H}^{{\sf m}}$ as a $U_0(\H)$ module. Hence we have proved
\begin{theorem}
Regard the symmetric Fock space $\F=\exp(\H\otimes{\mathbb C}} \newcommand{\diri}{\int^{\oplus}^M)$ as a left-module
(representation\ space) of $U_0(\H)$ and a right-module of $U(M)$ under the
second quantization of their respective natural actions on
$\H\otimes{\mathbb C}} \newcommand{\diri}{\int^{\oplus}^M$. Applying Rieffel induction to this bimodule, inducing
from the irreducible representation\ $U_{{\sf m}}} \newcommand{\pll}{U_{{\sf l}}(U(M))$ (which corresponds to the highest weight
${\sf m}=(m_1,\ldots,m_M)$), yields the induced space ${\cal H}^{{\sf m}}$ carrying
the irreducible representation\ $\pim(U_0(\H))$. \ll{main}
\end{theorem}
This, then, is the exact quantum counterpart of Theorem \ref{omw},
specialized to ${\sf n}=\emptyset$. As remarked earlier, there exists
an obvious analogue of Theorem \ref{main} for ${\sf m}=\emptyset$, in
which all Hilbert spaces and representation s occurring in the construction are
replaced by their conjugates.
To prepare for the next subsection we will now give a slight
reformulation of the proof. We start with f\/inite-dimensional
$\H={\mathbb C}} \newcommand{\diri}{\int^{\oplus}^k$, with $k>M$. Classical invariant theory \ci{How4} then
provides the decomposition of $\exp(S)$ under $\Gamma U_1(U(k))\otimes
\Gamma \overline{U}_1(U(M))$ as
\begin{equation}
\exp({\mathbb C}} \newcommand{\diri}{\int^{\oplus}^k\otimes{\mathbb C}} \newcommand{\diri}{\int^{\oplus}^M) \stackrel{{\rm
sq}}{\simeq} \bigoplus_{{\sf l}\in D_M} \H_{\sf l}^{U(k)}\otimes
\overline{\H}_{\sf l}^{U(M)}, \ll{howe}
\end{equation}
where the sum is over all
Young diagrams (or tuples) $D_M$ with $M$ rows or less, including the
empty diagram. (Note that it would have been consistent with our
previous notation to write $(\H^{\sf l})^{U(k)}$ for $\H_{\sf
l}^{U(k)}$; both stand for the irreducible representation\ of $U(k)$ def\/ined by
the Young diagram $\sf l$. In what follows, we will reserve the
notation ${\cal H}^{{\sf l}}$ for ${\cal H}_{{\sf l}}(U_0(\H))$, where $\H=l^2$.) Eq.\
(\ref{howe}) is an illustration of the theory of Howe dual pairs
\ci{How1,How2,How3}: it exhibits a multiplicity-free central
decomposition of $\F=\exp(S)$ under the commuting actions of $U(k)$
and $U(M)$ (which form a dual pair in $Sp(2kM,{\mathbb R}} \newcommand{\bok}{{\mbox{{\bf k}}})$, of which $\F$
carries the metaplectic representation).
In order to study the limit $k\raw\infty$ we realize $\exp(\H\otimes{\mathbb C}} \newcommand{\diri}{\int^{\oplus}^M)$
(with $\H=l^2$ now inf\/inite-dimesional) as an (incomplete) inf\/inite
tensor product \ci{vonN} with respect to the vacuum vector
$\Omega\in\exp({\mathbb C}} \newcommand{\diri}{\int^{\oplus}^M)$, that is (recalling $\exp({\mathbb C}} \newcommand{\diri}{\int^{\oplus}^k\otimes{\mathbb C}} \newcommand{\diri}{\int^{\oplus}^M)\simeq
\otimes^k \exp({\mathbb C}} \newcommand{\diri}{\int^{\oplus}^M)$), $\exp(\H\otimes{\mathbb C}} \newcommand{\diri}{\int^{\oplus}^M)\simeq
\otimes_{\Omega}^{\infty}\exp({\mathbb C}} \newcommand{\diri}{\int^{\oplus}^M)$, where the right-hand side is the
Hilbert space closure (with respect to the natural inner product on
tensor products) of the linear span of all vectors of the type
$\ps_1\otimes\ldots \ps_l\otimes\Omega\otimes\Omega\ldots$, $\ps_i\in\exp({\mathbb C}} \newcommand{\diri}{\int^{\oplus}^M)$,
in which only f\/initely many entries dif\/fer from $\Omega$. (The term
`incomplete' refers to the fact that only `tails' close to an inf\/inite
product of $\Om$'s appear.) Thus $\exp({\mathbb C}} \newcommand{\diri}{\int^{\oplus}^k\otimes{\mathbb C}} \newcommand{\diri}{\int^{\oplus}^M)\simeq \otimes^k
\exp({\mathbb C}} \newcommand{\diri}{\int^{\oplus}^M)$ is naturally embedded in $\exp(\H\otimes{\mathbb C}} \newcommand{\diri}{\int^{\oplus}^M)$ by simply
adding an inf\/inite tail of $\Om$'s, and this provides an embedding
$\exp({\mathbb C}} \newcommand{\diri}{\int^{\oplus}^k\otimes{\mathbb C}} \newcommand{\diri}{\int^{\oplus}^M)\subset \exp({\mathbb C}} \newcommand{\diri}{\int^{\oplus}^{k+1}\otimes{\mathbb C}} \newcommand{\diri}{\int^{\oplus}^M)$ as well. Clearly,
$\exp(\H\otimes{\mathbb C}} \newcommand{\diri}{\int^{\oplus}^M)$ coincides with the closure of the inductive limit
$\cup_{k=1}^{\infty} \exp({\mathbb C}} \newcommand{\diri}{\int^{\oplus}^k\otimes{\mathbb C}} \newcommand{\diri}{\int^{\oplus}^M)$ def\/ined by this embedding.
Choosing the natural basis in $\H=l^2$, we obtain an embedding
$U(k)\subset U(k+1)$, with corresponding actions on $\H$; our group
$U_0(\H)$ (realized in its def\/ining representation\ on $\H$) is the norm-closure
of the inductive limit group $\cup_{k=1}^{\infty} U(k)$. Using the
explicit realization of $\H^{\sf l}$ as a Young-symmetrized tensor
product, we similarly obtain embeddings ${\cal H}_{{\sf l}}(U(k))\subset
{\cal H}_{{\sf l}}(U(k+1))$. Thus the inductive limit $\cup_{k=1}^{\infty}
{\cal H}_{{\sf l}}(U(k))$ is well-def\/ined. Using~(\ref{howe}), we then have that
$\exp(\H\otimes{\mathbb C}} \newcommand{\diri}{\int^{\oplus}^M)$ is the closure of $\cup_{k=1}^{\infty} \oplus_{{\sf
l}\in D_M} \H_{\sf l}^{U(k)}\otimes \overline{\H}_{\sf l}^{U(M)}$, which
in turn coincides with the closure of $\oplus_{{\sf l}\in
D_M}\cup_{k=1}^{\infty}\H_{\sf l}^{U(k)}\otimes \overline{\H}_{\sf
l}^{U(M)}$. We now use the fact that the closure of
$\cup_{k=1}^{\infty}\H_{\sf l}^{U(k)}$ is ${\cal H}^{{\sf l}}$ as a representation\ space of
$U_0(\H)$ (this is obvious given the explicit realization of these
spaces, but it is a deep result that an analogous fact holds for all
representation s of $U_0(\H)$ \ci{Ols1,Ols3,Ols4}). This yields the desired
decomposition
\begin{equation}
\exp(\H\otimes{\mathbb C}} \newcommand{\diri}{\int^{\oplus}^M) \stackrel{{\rm sq}}{\simeq}
\bigoplus_{{\sf l}\in D_M} \H_{\sf l} \otimes \overline{\H}_{\sf
l}^{U(M)}, \ll{howeinf}
\end{equation}
under $\Gamma U_1(U_0(\H))\otimes \Gamma
\overline{U}_1(U(M))$. This result was previously derived in \ci{Ols4}
using a technique of holomorphic extension of representation s.
Starting from (\ref{howeinf}), Theorem \ref{main} follows immediately
from item 2 on the list of ingredients of our previous proof.
To end this subsection we register how the half-form correction to
geometric quantization modif\/ies (\ref{howe}), cf.\ subsection 2.1, and
in particular (\ref{dec2}). These corrections are f\/inite only for
$\H={\mathbb C}} \newcommand{\diri}{\int^{\oplus}^k$, $k<\infty$, so we only discuss that case. As for $M=1$,
one f\/inds that the half-form quantizations of the momentum maps
corresponding to the $U(k)$ and $U(M)$ actions on ${\mathbb C}} \newcommand{\diri}{\int^{\oplus}^k\otimes {\mathbb C}} \newcommand{\diri}{\int^{\oplus}^M$ lead
to Lie algebra representation s that can only be exponentiated to representation s
$U_{L,{\rm hf}}$ and $U^{-1}_{R,{\rm hf}}$ of the covering groups
$\tilde{U}(k)$ and $\tilde{U}(M)$ of $U(k)$ and $U(M)$, respectively, on
which the square-root of the determinant is def\/ined. A
straightforward exercise leads to the decomposition
\begin{equation}
\exp({\mathbb C}} \newcommand{\diri}{\int^{\oplus}^k\otimes{\mathbb C}} \newcommand{\diri}{\int^{\oplus}^M) \stackrel{{\rm hf}}{\simeq} \bigoplus_{{\sf l}\in
D_M} \H_{{\sf l}+\mbox{\footnotesize $\frac{1}{2}$} M}^{\tilde{U}(k)}\otimes \overline{\H}_{{\sf l}+\mbox{\footnotesize $\frac{1}{2}$}
k}^{\tilde{U}(M)} \ll{howehf}
\end{equation}
under $U_{L,{\rm hf}}(\tilde{U}(k))\otimes
U^{-1}_{R,{\rm hf}}(\tilde{U}(M))$. Here ${\sf l}+\mbox{\footnotesize $\frac{1}{2}$} M$, regarded as
a highest weight, has components $(l_1+\mbox{\footnotesize $\frac{1}{2}$} M, l_2+\mbox{\footnotesize $\frac{1}{2}$} M,\ldots)$,
and analogously for ${\sf l}+\mbox{\footnotesize $\frac{1}{2}$} k$. Hence $\H_{{\sf l}+\mbox{\footnotesize $\frac{1}{2}$} M}$
carries the tensor product of the representation\ of $\tilde{U}(k)$ characterized
by the Young diagram $\sf l$, and the determinant representation\ to the power
$M/2$, etc. This will be further discussed in subsection
\ref{discussion}.
\subsection{Representations induced from $U(M,N)$}
We are now going to attempt to `quantize' Theorem \ref{omw} for $N\neq
0$. The f\/irst problem is f\/inding a unitary representation\ of $H=U(M,N)$ that
corresponds to the dominant integral weight $\wmn$ on $\mathfrak t$ (or
the corresponding coadjoint orbit in $\h^*$, cf.\ subsection 2.2);
this is the representation\ we should induce from. This problem was solved in
\ci{Ada1}, partly on the basis of the classif\/ication of all unitary
highest-weight modules of $U(M,N)$ \ci{EHW,Jak,Ols2}. In the compact
case, each dominant integral weight corresponds to an irreducible
unitary representation\ with this weight as its highest weight. For $U(M,N)$ on
the other hand, there are two new phenomena. Firstly, there are
further conditions on the dominant integral weight $\wmn$, namely that
all entries of $\sf m$ should be dif\/ferent, and that all entries of
$\sf n$ should be dif\/ferent. Secondly, the representation\ corresponding to
$\wmn$, albeit a highest weight representation, does not in fact have $\wmn$ as
its highest weight. Rather, the highest weight corresponding to $\wmn$
is `renormalized': with $m_1>m_2>\ldots>m_M>0$ and
$n_1>n_2>\ldots>n_N>0$, the highest weight (naively expected to be
$(m_1,\ldots,m_M,-n_N,\ldots,-n_1)$) is in fact
$$
(m_1
+\mbox{\footnotesize $\frac{1}{2}$}(N-M)+\mbox{\footnotesize $\frac{1}{2}$},\ldots,m_i+\mbox{\footnotesize $\frac{1}{2}$}(N-M)+i-\mbox{\footnotesize $\frac{1}{2}$},\ldots, m_M+\mbox{\footnotesize $\frac{1}{2}$}
(N+M)-\mbox{\footnotesize $\frac{1}{2}$},
$$
$$
-(n_N+\mbox{\footnotesize $\frac{1}{2}$}(M+N)-\mbox{\footnotesize $\frac{1}{2}$}),\ldots,-(n_j+\mbox{\footnotesize $\frac{1}{2}$}(M-N)+j-\mbox{\footnotesize $\frac{1}{2}$}),\ldots,
-(n_1+\mbox{\footnotesize $\frac{1}{2}$}(M-N)+\mbox{\footnotesize $\frac{1}{2}$})).
$$
Note that this highest weight is still
dominant; however, it may no longer be integral, so that it def\/ines a
projective representation\ of $U(M,N)$ (single-valued on its double cover
$\tilde{U}(M,N)$). These highest weight representation s belong to the
holomorphic discrete series of $U(M,N)$ \ci{Kna}.
The second problem is the quantization of $S=\H\otimes{\mathbb C}} \newcommand{\diri}{\int^{\oplus}^{M+N}$, with the
corresponding actions of $G=U_0(\H)$ and $H=U(M,N)$. One regards
$U(M,N)$ as a subgroup of $Sp(2(M+N),{\mathbb R}} \newcommand{\bok}{{\mbox{{\bf k}}})$, so that the symplectic
action of the former on ${\mathbb C}} \newcommand{\diri}{\int^{\oplus}^{M+N}$ is the restriction of the action of
the latter \ci{Ste,KKS}. Due to the special way we def\/ined the
$U(M,N)$ action in subsection 2.2 as the inverse of a right-action,
the quantization of this action of $Sp(2(M+N),{\mathbb R}} \newcommand{\bok}{{\mbox{{\bf k}}})$ is then given by
the conjugate of the metaplectic representation\ $U_m$ on $L^2({\mathbb R}} \newcommand{\bok}{{\mbox{{\bf k}}}^{M+N})\equiv
{\cal L}$, cf.\ \ci{KV,SW,Ste}. This def\/ines a representation\ of the inverse
image $\tilde{U}(M,N)$ of $U(M,N)$ in the metaplectic group
$Mp(2(M+N),{\mathbb R}} \newcommand{\bok}{{\mbox{{\bf k}}})$ on $\overline{\cal L}$, which descends to a projective
representation\ of $U(M,N)$, which we denote by $U_{{\rm hf}}(\tilde{U}(M,N))$. As
pointed out in \ci{SW} and \ci{BR} (for $k=1$), this representation\ is
precisely the one obtained from geometric quantization (in a suitable
cohomological variant) if half-forms are taken into account. This
yields a f\/irst candidate for the quantization of the $U(M,N)$ action
on ${\mathbb C}} \newcommand{\diri}{\int^{\oplus}^{M+N}$.
The second possibility is to take the tensor product of the
(restriction of) the metaplectic representation\ of $\tilde{U}(M,N)$ with the
square-root of the determinant, which does def\/ine a unitary representation\
$U_{{\rm sq}}$ of $U(M,N)$ \ci{SW}; see \ci{BR} for a construction of
this representation\ from geometric quantization. It is the representation\ which might be
thought of as being def\/ined by the physicists' second quantization on
$\exp({\mathbb C}} \newcommand{\diri}{\int^{\oplus}^{M+N})$, as in the $U(M)$ case. However, since the action of
$U(M,N)$ on ${\mathbb C}} \newcommand{\diri}{\int^{\oplus}^{M+N}$ is not unitary, this second quantization is
not, in fact, def\/ined. In geometric quantization this lack of
unitarity shows up through the non-existence of a totally complex
invariant polarization on $S$ which is positive. Consequently, one
needs to work with an indef\/inite such polarization \ci{BR}, and this
leads to complications that will eventually cause a shift in the representation
s one would naively expect to occur in the decomposition of the
quantization of $S$.
For f\/inite-dimensional $\H={\mathbb C}} \newcommand{\diri}{\int^{\oplus}^k$ we therefore have a suitable
quantization of $S={\mathbb C}} \newcommand{\diri}{\int^{\oplus}^k\otimes{\mathbb C}} \newcommand{\diri}{\int^{\oplus}^{M+N}$, namely the Hilbert space
$\overline{\cal L}_k\equiv \otimes^k\overline{\cal L}$ (the Fock space realization of this
space is not useful, so we drop the notation $\F$). Moreover, we have
natural unitary representation s $\otimes^k U_{{\rm sq/hf}}$ of $\tilde{U}(M,N)$ on
$\overline{\cal L}_k$, which are quantizations of the symplectic action of $U(M,N)$ on
$S$. Following our notation for $U(M)$, we refer to these representation s as
$U^{-1}_{R,{\rm sq/hf}}$.
In addition, the quantization of the $U(k)$ action on $S$ may be
found (much more easily) from geometric quantization with or without
half-forms. The latter case, in which we call the representation\ $U_{L,{\rm
sq}}(U(k))$, is explicitly given in \ci{KV}. Its half-form variant
$U_{L,{\rm hf}}(U(k))$ dif\/fers from it by the determinant representation\ raised
to the power $(M-N)/2$.
It follows from the theory of Howe dual pairs \ci{How1} that $\overline{\cal L}_k$
decomposes discretely under these representation s. Starting with $U_{L,{\rm
sq}}(U(k))\otimes U^{-1}_{R,{\rm sq}}(U(M,N))$, the explicit decomposition
of $\overline{\cal L}_k$ is given in \ci{KV} as (remember that we have to take the
conjugate of the $U(M,N)$ modules, but not of the $U(k)$ modules used
in \ci{KV}, since our $U(k)$ action is the usual one; also, we use the
conventions of \ci{Ada1} and \ci{How2} for labelling the highest
weight, rather than those of \ci{KV} -- this corresponds to an
interchange of $\sf m$ and $\sf n$)
\begin{equation}
\overline{\cal L}_k \stackrel{{\rm sq}}{\simeq} \bigoplus_{\wmn}
\H_{\wmn}^{U(k)}\otimes \overline{\H}_{({\sf
m}+ k,{\sf n})}^{U(M,N)}, \ll{kave}
\end{equation}
where the sum is over all
pairs $\wmn$ as def\/ined before, with zeros allowed, but neither $\sf
m$ nor $\sf n$ allowed to be empty. $\H_{\wmn}^{U(k)}$ as a representation\
space of $U(k)$ was def\/ined in subsection 2.3, and $\H_{({\sf
m}+k,{\sf n})}^{U(M,N)}$ carries the unitary representation\ of $U(M,N)$ with
highest weight (not subject to further `renormalization')
\[
(m_1+k,\ldots,m_i+k,\ldots,m_M+k,-n_N,\ldots,-n_j,\ldots,-n_1).
\]
The decomposition under $U_{L,{\rm hf}}(U(k))\otimes U^{-1}_{R,{\rm
hf}}(U(M,N))$, on the other hand, reads \ci{How2}
\begin{equation}
\overline{\cal L}_k \stackrel{{\rm hf}}{\simeq} \bigoplus_{\wmn} \H_{({\sf
m}+\mbox{\footnotesize $\frac{1}{2}$}(M-N),{\sf n}-\mbox{\footnotesize $\frac{1}{2}$}(M-N))}^{\tilde{U}(k)}\otimes
\overline{\H}_{({\sf m}+ \mbox{\footnotesize $\frac{1}{2}$} k,{\sf n} +\mbox{\footnotesize $\frac{1}{2}$} k)}^{\tilde{U}(M,N)},
\ll{kave2}
\end{equation}
where the highest weight $({\sf m}+ \mbox{\footnotesize $\frac{1}{2}$} k,{\sf n}
+\mbox{\footnotesize $\frac{1}{2}$} k)$ is explicitly given by
\[
(m_1+k/2,\ldots,m_i+k/2,\ldots,m_M+k/2,-n_N-k/2,
\ldots,n_j-k/2,\ldots,-n_1-k/2),
\]
whereas $\H_{({\sf m}+\mbox{\footnotesize $\frac{1}{2}$}(M-N),{\sf n}-\mbox{\footnotesize $\frac{1}{2}$}(M-N))}$ is the tensor
product of $\H_{\wmn}$, and ${\mathbb C}} \newcommand{\diri}{\int^{\oplus}$, carrying the determinant representation\ of
$U(k)$ to the power $(M-N)/2$, cf.\ \ci{How2}).
Working with (\ref{kave}) for the sake of concreteness, we now wish to
apply Rief\/fel induction from a suitable representation\ of $H=U(M,N)$ to $\overline{\cal L}_k$
in order to extract the copy of $\H_{\wmn}^{U(k)}$ for the value of
$\wmn$ selected by the representation\ we induce from. Firstly, we need a dense
subspace $L\subset \overline{\cal L}_k$ such that the function $x\raw ( U^{-1}_{R,{\rm
sq}}(x)\ps,\varphi} \newcommand{\ch}{\chi)$ is in $L^1(H)$ for all $\ps,\varphi} \newcommand{\ch}{\chi\in L$. This is
easily found: using the decomposition (\ref{kave}), we take $L$ to
consist of vectors having a f\/inite number of components in the
decomposition, each component of which is in the tensor product of
$\H^{U(k)}_{\ldots}$ and the dense subspace of $K$-f\/inite vectors in
the other factor. Since each function of the type $x\raw
(U(x)\ps,\varphi} \newcommand{\ch}{\chi)$, where $U$ is in the discrete series, and $\ps$ and
$\varphi} \newcommand{\ch}{\chi$ are $K$-f\/inite vectors, is in Harish-Chandra's Schwartz
space~\ci{Kna} (which is a subspace of $L^1(H)$), this choice indeed
satisf\/ies the demand. (Based on the explicit realization of $\overline{\cal L}_k$ as
a function space \ci{KV}, a more `geometric' choice of $L$ may also be
found.)
As we are going to induce from holomorphic discrete series representation s of
$U(M,N)$, let us examine the tensor product $\overline{\H}_{({\sf
m}_1,{\sf n}_1)}^{U(M,N)}\otimes \H_{({\sf m}_2,{\sf n}_2)}^{U(M,N)}$.
Recall that $\wmn$ (which here refers to the actual highest weight,
rather than the dominant integral weight that is subject to
renormalization, as sketched above) def\/ines a unitary irreducible
representation\ $U_{\wmn}$ of the maximal compact subgroup $K=U(M)\times U(N)$
with highest weight $(m_1,\ldots,m_M,-n_N,\ldots,-n_1)$. By Theorem 2
in \ci{Rep}, the above tensor product is unita\-ri\-ly equivalent as a
representation\ space of $U(M,N)$ to the representation\ induced (in the usual, Mackey,
sense) from $\overline{U}_{({\sf m}_1,{\sf n}_1)}\otimes U_{({\sf
m}_2,{\sf n}_2)}(K)$. Using the reduction-induction theorem, we can
therefore decompose this induced representation\ as a direct sum over the representation s
induced from the components in the decomposition of
$\overline{U}_{({\sf m}_1,{\sf n}_1)}\otimes U_{({\sf m}_2,{\sf
n}_2)}(K)$.
Let us examine a generic representation\ $U^{\kappa}(H)$ (realized on the
Hilbert space $\H^{\kappa}$ of functions $\ps:G\raw\H_{\kappa}$
satisfying the equivariance condition
$\ps(xk)=U_{\kappa}(k^{-1})\ps(x)$) induced from an irreducible representation\
$U_{\kappa}(K)$. The Rief\/fel induction procedure produces the
semi-def\/inite form $(\cdot ,\cdot )_0$ on $L\otimes\H_{\ch}$ (where, in
this case, $\H_{\ch}=\H_{({\sf m},{\sf n})}^{U(M,N)}$ for certain
$\wmn$). Using (\ref{kave}) and the previous paragraph, we f\/ind that
$L\otimes\H_{\ch}$ is a certain dense subspace of a direct sum with
components of the type $\H_{\wmn}^{U(k)}\otimes\H^{\kappa}$, in which $H$
acts trivially on the f\/irst factor. By our construction of $L$, each
element of $L\otimes\H_{\ch}$ only has components in a f\/inite number of
these Hilbert spaces, so that we can investigate each component
separately. (Had the number of components of elements of $L$ been
inf\/inite, the study of $(\cdot ,\cdot )_0$ would have been more
involved, as this is an unbounded and non-closable quadratic form, so
that $(\sum_i\ps_i,\varphi} \newcommand{\ch}{\chi)_0\neq \sum_i(\ps_i,\varphi} \newcommand{\ch}{\chi)_0$ for inf\/inite
sums.)
Factorizing $\int_H dx= \int_N dn\,\int_K dk$ \ci{Kna}, it follows
from the equivariance condition and the orthogonality relations for
compact groups that in a given component
$\H_{\wmn}^{U(k)}\otimes\H^{\kappa}$ the expression $(\ps,\varphi} \newcommand{\ch}{\chi)_0=\int_H
dx\, ({\mathbb I}\otimes U^{\kappa}(x)\ps,\varphi} \newcommand{\ch}{\chi)$ vanishes unless $U_{\kappa}$
is the identity representation\ $U_{\rm id}$ of $K$. Given a highest weight
representation\ $U_{\ch}(H)$ we Rief\/fel-induce from, there exists a unique pair
$\wmn$ for which $\H_{\wmn}^{U(k)}\otimes\H^{\rm id}$ occurs in the
decomposition of $\overline{\cal L}_k\otimes\H_{\ch}$ as a sum over induced representation s of $H$
in the above sense.
Let $L^{\rm id}$ be the projection of $L\otimes\H_{\ch}$ onto this
$\H_{\wmn}^{U(k)}\otimes\H^{\rm id}$. We def\/ine $\tilde{V}:L^{\rm id}\raw
\H_{\wmn}^{U(k)}$ by linear extension of
$\tilde{V}\ps_1\otimes\ps_2=\ps_1\int_Hdx\, \ps_2(x)$ (where $\ps_1\in
\H_{\wmn}^{U(k)}$ and $\ps_2\in\H^{\rm id}\subset L^2(G)$). The
integral exists by our assumptions on $L$; moreover, the explicit form
of the inner product in $\H^{\rm id}$ (namely $(f,g)=\int_H dx\,
f(x)\overline{g(x)}$, as $K$ is compact) leads to the equality
$(\tilde{V}\ps,\tilde{V}\varphi} \newcommand{\ch}{\chi)=(\ps,\varphi} \newcommand{\ch}{\chi)_0$ (where the inner product on
the left-hand side is the one in $\H_{\wmn}^{U(k)}$). We now extend
$\tilde{V}$ to a map $V$ from $L\otimes\H_{\ch}$ to $\H_{\wmn}^{U(k)}$ by
putting it equal to zero on all spaces involving a factor
$\H^{\kappa}$, where $\kappa\neq {\rm id}$ (and equal to $V$ on
$L^{\rm id}$, of course). Clearly, by this and the preceding
paragraph,
\begin{equation}
(V\ps,V\varphi} \newcommand{\ch}{\chi)=(\ps,\varphi} \newcommand{\ch}{\chi)_0. \ll{V}
\end{equation}
We are now in a
standard situation in the theory of Riefel induction, in which we can
identify the null space of $(\cdot ,\cdot )_0$ with the kernel of $V$,
and the induced space $\H^{\ch}$ (which, we recall, is the completion
of the quotient of $L\otimes\H_{\ch}$ by this null space in the inner
product obtained from this form) with the closure of the image of
$V$. It is clear from our def\/inition of $L$ that the image of $V$
actually coincides with $\H_{\wmn}^{U(k)}$. Also, the def\/inition of
the induced representation\ $U^{\ch}$ of $G=U(k)$ on $\H^{\ch}$ immediately
implies that $U^{\ch}\simeq U_{\wmn}$. Finally, note that (\ref{V})
shows explicitly that $(\cdot,\cdot)_0$ is positive semi-def\/inite, a
fact which was already certif\/ied by Prop.\ 2 in \ci{NPL93}.
Putting these arguments together, we have proved:
\begin{theorem}\ll{old3}
Let $U(k)$ and $U(M,N)$ act on $S={\mathbb C}} \newcommand{\diri}{\int^{\oplus}^k\otimes{\mathbb C}} \newcommand{\diri}{\int^{\oplus}^{M+N}$ (equipped with the
symplectic form (\ref{ommn})) from the left and the right,
respectively, in the natural way, and let $\overline{\cal L}_k$ be the quantization
of $S$, with commuting representation s of $U(k)$ and $U(M,N)$ on $\overline{\cal L}_k$ (which
quantize the above symplectic actions) as given (up to conjugation of
the representation\ of $U(M,N)$) by Kashiwara-Vergne \ci{KV}.
Then Rieffel induction on $\overline{\cal L}_k$ from the holomorphic discrete series
representation\ of $U(M,N)$ with highest weight $({\sf m}+ k,{\sf n})$ (that is,
the highest weight with components
$(m_1+k,\ldots,m_M+k,-n_N,\ldots,-n_1)$) leads to an induced space
${\cal H}_{({\sf m,n})}^{U(k)}$, which as a Rieffel-induced $U(k)$ module carries the
representation\ $U_{\wmn}(U(k))$ (which is the Young product of the representation\ with
Young diagram $\sf m$ and the conjugate of the representation\ with Young
diagram $\sf n$).
Moreover, the induced space is empty if one induces from a highest
weight representation\ of $U(M,N)$ of the form $\wmn$ in which at least one
$m_i$ is smaller than $k$, or is not integral. \ll{dis} \end{theorem}
\section{Discussion}
{The last part of Theorem \ref{old3} is particularly unpleasant for
the quantization theory of constrained system, for it shows that
Theorem \ref{omw} cannot really be `quantized' unless $\sf m$ or $\sf
n$ are empty. For we would naturally induce from the holomorphic
discrete series representation\ of $U(M,N)$ having the `renormalized' highest
weight corresponding to a coadjoint orbit characterized by $\wmn$, as
explained at the beginning of this subsection. But then for $k$ large
enough the induced space will be empty, rather than consisting of
${\cal H}_{({\sf m,n})}^{U(k)}$, as desired. As we have seen, the induction procedure
is only successful if we induce from a representation\ with highest weight
$({\sf m}+k,{\sf n})$, rather than from the ($k$-independent)
renormalized weight we ought to use by f\/irst principles. This is
bizarre, given that the original weight $\wmn$ (or the orbit it
corresponds to) knows nothing about $k$ or $U(k)$. In addition, even
without this problem the induced space will often be empty, because
the `correct' renormalized highest weight one induces from may simply
not occur in the Kashiwara-Vergne decomposition (\ref{kave}) because
of the half-integral nature of its entries (which is a pure `quantum'
phenomenon). (In a rather dif\/ferent setting, the discrepancy for
large $k$ between the `decomposition' of $S$ into pairs of matched
coadjoint orbits for $U(k)$ and $U(M,N)$, and the decomposition of
$\overline{\cal L}_k$ under these groups, must have been noticed by Adams \ci{Ada2},
who points out that there is a good correspondence for $k\leq {\rm
min}\,(M,N)$ only.)
It is peculiar to the non-compact ($N\neq 0$) case that this
dif\/f\/iculty even arises if the half-form correction to quantization is
not applied. For (\ref{kave}) is the non-compact analogue of
(\ref{howe}), and in the latter quantization clearly does commute with
reduction. If we do incorporate half-forms, we obtain (\ref{kave2})
for $U(M,N)$ and (\ref{howehf}) for $U(M)$. In both cases the Rief\/fel
induction process generically (that is, if $M\neq N$) fails to produce
the correct representation\ of $U(k)$, even if one induces from a representation\ whose
highest weight is renormalized (compared to the weight expected from
the orbit correspondence) by the term $k/2$.
Finally, the passage from ${\mathbb C}} \newcommand{\diri}{\int^{\oplus}^k$ to inf\/inite-dimensional Hilbert
spaces is tortuous whenever half-forms are used (the corrections being
inf\/inite for $k=\infty$), and in the non-compact case even without
these. This is partly because of the $k$-dependence of the highest
weights of $U(M,N)$, and partly because $\cal L$ does not contain the
identity representation\ of $U(M,N)$ (recall that in the compact case we used
the carrier space ${\mathbb C}} \newcommand{\diri}{\int^{\oplus}\Omega$ of this representation\ as the f\/ixed `tail' vector
to construct the von Neumann inf\/inite tensor product from).
Clearly, this situation deserves further study. We do not think it is
an artifact of our proposal of using Rief\/fel induction in the
quantization of constrained systems. In fact, this technique comprises
the only method known to us which is precise enough to bring the
embarrassment to light. \ll{discussion} }
|
\section{Introduction}\label{sec:intro}
The possible discovery of an accelerating Universe from observations
of Type Ia supernovae (Perlmutter {\it et al. } 1998, Riess {\it et al. } 1999) has
led to a resurgence of interest in the possibility that the Universe
is dominated by a cosmological constant (for a recent review see
Turner 1999). A number of authors have shown how observations of
distant Type Ia supernovae (SN) can be combined with observations of
CMB anistropies to constrain the cosmological constant and matter
density of the Universe (White 1998, Tegmark {\it et al. } 1998, Lineweaver
1998, Garnavich {\it et al. }, 1998, Efstathiou and Bond 1999, Tegmark 1999,
Efstathiou {\it et al. } 1999). For example, Efstathiou {\it et al. } (1999,
hereafter E99) combine the large SN sample of the Supernova Cosmology
Project (Perlmutter {\it et al. } 1998, hereafter P98; we will refer to these
supernovae as the SCP sample) with a compilation of CMB anisotropy
measurements and find $\Omega_m = 0.25^{+0.18}_{-0.12}$ and
$\Omega_\Lambda = 0.63^{+0.17}_{-0.23}$ ($95 \%$ confidence errors)
for the cosmic densities contributed by matter and a cosmological
constant respectively. These results are consistent with a number of
other measurements, including dynamical measurements of $\Omega_m$,
the large-scale clustering of galaxies and the abundances of rich
clusters of galaxies (Turner 1999, Bridle {\it et al. } 1999, Wang {\it et al. }
1999).
\begin{figure*}
\vskip 3.0 truein
\special{psfile=pgfish1a.ps
hscale=45 vscale=45 angle=-90 hoffset= -10
voffset=240}
\special{psfile=pgfish1b.ps
hscale=45 vscale=45 angle=-90 hoffset= 170
voffset=240}
\caption
{The dashed lines in each panel show $1$, $2$ and $3 \sigma$
likelihood contours in the $\Omega_\Lambda$--$\Omega_m$ plane for the
SCP distant supernova sample as analysed by E99. The solid contours
are derived from the Fisher matrix (equation 4) for the SCP sample
supplemented by 20 SN with a mean redshift of $\langle z \rangle =1$
(Figure 1a) and for twice the SCP sample and $40$ SN with $\langle z
\rangle =1.5$ (Figure 1b). The points show maximum likelihood values
of $\Omega_\Lambda$ and $\Omega_m$ for Monte-Carlo realizations of
these samples, as described in the text.}
\label{figure1}
\end{figure*}
The observational evidence for an accelerating Universe has stimulated
interest in more general models containing a component with an
arbitrary equation of state, $p/\rho = w_Q$ with $w_Q \ge
-1$. Examples include a dynamically evolving scalar field (see {\it
e.g.} Ratra and Peebles 1988 and Caldwell, Dave and Steinhardt 1998,
who have dubbed such a component `quintessence'; we will refer to this
as a `Q' component hereafter) and a frustrated network of topological
defect (Spergel and Pen 1997, Bucher and Spergel 1999). In particular,
Steinhardt , Wang and Zlatev, 1998, have pointed out that for a wide
class of potentials, the evolution of a Q-like scalar field follows
`tracking solutions' , in which the late time evolution is almost
independent of initial conditions.
The purpose of this paper is three-fold. Firstly, to illustrate how
the constraints on $\Omega_m$ and $\Omega_\Lambda$ can be improved by
extending the redshift range of the supernovae samples. At low
redshifts, the magnitude-redshift relation is degenerate for models
with the same value of the decellaration parameter
$q_0$ ($\equiv {1 \over 2}(\Omega_m - 2\Omega_\Lambda)$). This degeneracy
can be broken by observing supernovae at redshifts $\lower.5ex\hbox{\gtsima} 1$ (see,
for example, Goobar and Perlmutter, 1995). Thus, by extending the
redshift range of the current supernovae samples it should be possible
to set tighter limits on $\Omega_m$ and $\Omega_\Lambda$
independently. This is important because there are significant worries
that the SN data may be affected by grey extinction, evolution, or some
other systematic effect. The consistency of SN constraints on
$\Omega_m$ and $\Omega_\Lambda$ with those derived from the CMB
anisotropy measurements would provide an important consistency check
of systematic errors in the SN data and the interpretation of the CMB
data. Secondly, we estimate the accuracy with which a more general
Q-like equation of state can be constrained by high redshift SN and
CMB data. Thirdly, we use the current SN and CMB anisotropy data to
constrain Q-like models in a spatially flat universe and in a universe with
arbitrary spatial curvature.
\section{Analysis of Models with a Cosmological Constant}
\subsection{Constraints from supernovae at $z \sim 1$}
The predicted peak magnitude-redshift relation is given by
\begin{equation}
m^{\rm pred} (z) = {\cal M} + 5{\rm log}{\cal D}_L(z, \Omega_m,
\Omega_\Lambda),
\end{equation}
where ${\cal M}$ is related to the peak absolute
magnitude by ${\cal M} = M - 5 {\rm log}H_0 + 25$.
and ${\cal D}_L = d_L + 5 {\rm log}H_0$
is the Hubble constant-free luminosity distance.
To compute the luminosity distance, we ignore
gravitational lensing and use the standard expression for
a Universe with uniform density (see {\it e.g.} Peebles 1993),
$$
d_L(z, \Omega_m, \Omega_\Lambda) = {c \over H_0} {(1+z) \over \vert
\Omega_k \vert^{1/2}}
{\rm sin}_k \left [ \vert \Omega_k \vert^{1/2} x(z, \Omega_m,
\Omega_\Lambda) \right ],
$$
$$
x(z, \Omega_m, \Omega_\Lambda) =
\int_0^z {dz^\prime
\over
[\Omega_m (1 + z^\prime)^3 + \Omega_k ( 1 + z^\prime)^2 +
\Omega_\Lambda]^{1/2} }
\quad (2)
$$
where $\Omega_k = 1 - \Omega_m - \Omega_\Lambda$ and ${\rm sin}_k =
{\rm sinh}$ if $\Omega_k > 0$ and ${\rm sin_k} = {\rm sin}$ for
$\Omega_k < 0$.
\begin{table}
\label{tab1}
\centerline{\bf Table 1: Fisher Matrix Errors, $\Omega_m$ and $\Omega_\Lambda$.}
\begin{center}
\begin{tabular}{|cccccc|} \hline
\multicolumn{6}{c} {Supernovae Alone} \\
& SCP & \multicolumn{2}{c}{SCP + 20 SN} &
\multicolumn{2}{c}{2$\times$SCP + 40SN} \\
$<z>$ & & $1.0$ & 1.5 & 1.0 & 1.5 \\
$\delta \Omega_m$ & $0.53$ & $0.130$ & $0.081$ & $0.092$ & $0.057$ \\
$\delta \Omega_\Lambda$& $0.71$ & $0.265$ & $0.218$ & $0.19$ & $0.154$\\
$\delta {\cal M}$ & $0.056$ & $0.053$ & $0.049$ & $0.035$ & $0.035$\\\hline
\multicolumn{6}{c} {Supernovae + CMB} \\
& SCP & \multicolumn{2}{c}{SCP + 20 SN} &
\multicolumn{2}{c}{2$\times$P98 + 40SN} \\
$<z>$ & & $1.0$ & 1.5 & 1.0 & 1.5 \\
$\delta \Omega_m$ & $0.073$ & $0.055$ & $0.047$ & $0.039$ & $0.033$ \\
$\delta \Omega_\Lambda$& $0.080$ & $0.060$ & $0.051$ & $0.042$ & $0.036$\\
$\delta {\cal M}$ & $0.046$ & $0.042$ & $0.039$ & $0.030$ & $0.028$\\\hline
\end{tabular}
\end{center}
\end{table}
We assume that we observe $N$ supernovae, with peak magnitude $m_i$,
(corrected for k-term, decline rate-luminosity relation, reddening {\it
etc}), magnitude error $\sigma_i$ and redshift $z_i$, from which we
want to determine a set of parameters $s_k$ by maximising the likelihood
function,
\addtocounter{equation}{1}
\begin{equation}
{\cal L} = \prod_{i=1}^N {1 \over \sqrt{(2 \pi \sigma_i)}}
{\rm exp} \left
( - {(m_i - m_i^{\rm pred})^2 \over 2 \sigma_i^2} \right ).
\end{equation}
In this section we assume that the parameters $s_k$ are $\Omega_m$,
$\Omega_\Lambda$ and ${\cal M}$
(defined in equation 1). An estimate of the covariance matrix,
$C_{ij}$, for these parameters for a given SN data set is given by the
inverse of the Fisher matrix
\begin{equation}
F_{ij} = \sum_k {1 \over \sigma_k^2} {\partial {m_k^{\rm pred}} \over \partial s_i}
{\partial {m_k^{\rm pred}} \over \partial s_j}
\end{equation}
(Kendall and Stewart 1979). The marginalized
error on each parameter (given by $\sqrt C_{ii}$) is listed in Table 1
for several assumed supernova datasets. The column labelled SCP gives
the Fisher matrix errors on $\Omega_m$, $\Omega_\Lambda$ and ${\cal
M}$ derived for sample C ($56$ supernovae) of P98, {\it i.e.}
assuming the magnitude errors, intrinsic magnitude scatter and
redshift distribution of the real sample. The next two columns give
the expected errors for the SCP sample supplemented by 20 supernovae
with a peak magnitude error of $\Delta m = 0.25$ magnitudes and a
Gaussian redshift distribution of dispersion $\Delta z=0.5$ and mean
redshift $\langle z \rangle = 1$ and $1.5$. The upper redshift limit
is close to the maximum for feasible spectroscopic measurements with
$10$ metre-class telescopes (see Goobar and Perlmutter 1995). As these
authors comment, ground based spectroscopy at optical wavelengths
becomes prohibitively expensive for supernovae at higher redshifts
because of the strong K-correction. The last two columns give the
errors for a sample twice as large as the SCP sample supplemented by
40 supernovae with mean redshift of $1.0$ and $1.5$. We adopt a
background cosmology with $\Omega_\Lambda = 0.63$ and $\Omega_m =
0.25$ as indicated by the joint likelihood analysis of the SCP sample
and CMB anisotropies described in E99.
From Table 1 we see that the Fisher matrix analysis of the SCP sample
gives relatively large errors on $\Omega_\Lambda$ and $\Omega_m$, in
agreement with the likelihood analysis presented by P98. However, by
adding $20$ SN at $z \sim 1$, the errors on $\Omega_\Lambda$ and in
particularly $\Omega_m$ are reduced significantly. The last column
shows that an enhanced SCP sample together with $40$ SN at $z \sim
1.5$ (a formidable, but feasible observing programme) can provide a
tight constraint on $\Omega_m$. The parameters $\Omega_\Lambda$ and
$\Omega_m$ are, of course, highly correlated. This is illustrated in
Figure 1 which shows $1$, $2$ and $3\sigma$ error ellipses in the
$\Omega_\Lambda$-$\Omega_m$ plane after marginalizing over $s_3 =
{\cal M}$ assuming a uniform prior distribution. (The components of
the new Fisher matrix after marginalization are given by
$F^\prime_{11} = F_{11} -F^2_{13}/F_{33}$, $F^\prime_{22} = F_{22}
-F^2_{23}/F_{33}$, $F^\prime_{12} = F_{12} -F_{13}F_{23}/F_{33}$.) The
points in the Figure show the results of Monte-Carlo calculations,
where we have simulated the observational samples and determined the
parameters $s_i$ by maximising the likelihood function (equation 2).
By diagonalizing the matrix $F^\prime$ we can find the orthogonal
linear combinations $\Omega_\| = a \Omega_m + b \Omega_\Lambda$ and
$\Omega_\bot = b \Omega_m - a \Omega_m$ defining the major and minor
axes of the likelihood contours shown in Figure 1. The distributions
in these orthogonal directions are shown in Figure 2 and compared with
the distributions determined from the Monte-Carlo simulations. The
Monte-Carlo distributions are very close to Gaussians and show that
the Fisher matrix gives an extremely accurate description of the
errors in the $\Omega_m$--$\Omega_\Lambda$ plane.
\begin{figure}
\vskip 5.4 truein
\special{psfile=pghista.ps
hscale=40 vscale=40 angle=-90 hoffset= -50
voffset=410}
\special{psfile=pghistb.ps
hscale=40 vscale=40 angle=-90 hoffset= -50
voffset=210}
\caption
{Distributions along the major and minor axes of the likelihood
contours shown in Figure 1. The histograms show the distributions derived
from the Monte-Carlo simulations and the dotted lines show Gaussian distributions with variances determined from the Fisher matrix after marginalizing over
the parameter ${\cal M}$.}
\label{figure2}
\end{figure}
Although the errors in $\Omega_m$ and $\Omega_\Lambda$ are
significantly reduced by the addition of high redshift supernovae over
those of the SCP sample, they are still quite large in the parallel
direction $\Omega_\|$. This means that it is difficult to set tight
limits on $\Omega_\Lambda$ from SN measurements alone. The constraints
on the spatial curvature $\Omega_k$ are even weaker. For example, for
the larger sample shown in Figures 1 and 2, the $1\sigma$ error on
$\Omega_k$ is $\delta \Omega_k = 0.19$. This can be reduced by
extending the range to even higher redshifts (see Section 2.2)
or by combining the SN data with cosmic
microwave background anisotropies, as has been done by several authors
(White 1998, Lineweaver 1998, Garnavich {\it et al. } 1998, Tegmark 1999,
E99).
CMB anisotropy measurements, especially with future satellites such
as MAP and Planck, are capable of setting tight constraints on the
locations of the acoustic peaks in the CMB power spectrum. Following
E99, we define an acoustic peak location parameter $\gamma_D(\Omega_m,
\Omega_\Lambda)$ to be the ratio of the peak position in a model with
arbitrary cosmology compared to that in a spatially flat model with
zero cosmological constant. (This parameter depends weakly on the
matter content of the Universe and on the spectral index of the
fluctuations, but we ignore these small dependences in what follows).
CMB measurements are therefore capable of fixing $\gamma_D$, defining
a degeneracy direction in the $\Omega_\Lambda$--$\Omega_m$ plane given
by
\begin{equation}
\Delta \Omega_\Lambda = - {(\partial \gamma_D/\partial \Omega_m)
\over (\partial \gamma_D/\partial \Omega_\Lambda)} \Delta \Omega_m,
\end{equation}
(see Efstathiou and Bond 1998). The results in the lower panel of
Table 1 show the Fisher matrix analysis of the SN samples including
the constraint imposed by equation (5). As is well known,
the combination of SN and CMB measurements can break the degeneracy
between $\Omega_\Lambda$ and $\Omega_m$ and it should be possible to
determine these parameters with an error of less than $0.04$ with an
enlarged supernova sample assuming, of course, that systematic errors
are unimportant.
Although the errors on $\Omega_\Lambda$ from SN measurements alone
converge relatively slowly as the redshift range is increased,
consistency of the cosmological parameter estimates provides a strong
test of systematic errors in the SN data. If we believe that
systematic errors are unimportant, and that our interpretation of the
CMB anisotropies (in terms of adiabatic CDM-like models) are correct,
then current data already constrain $\Omega_m$ and $\Omega_\Lambda$ to
high precision (see Fig 5 of E99). Consistency requires that the
likelihood contours for a high redshift supernova sample converge to
the same answer.
\begin{figure}
\vskip 3.0 truein
\special{psfile=pgNGST.ps
hscale=45 vscale=45 angle=-90 hoffset= -40
voffset=240}
\caption
{Fisher matrix constraints for a sample of SN extending to redshifts
$z>3$ (see text) illustrating that by extending the redshift range one
can determine $\Omega_m$ independently of $\Omega_\Lambda$.}
\label{figure3}
\end{figure}
\subsection{Constraining $\Omega_m$ with NGST}
Observations of very distant supernovae at $z \lower.5ex\hbox{\gtsima} 3$ may be
possible with a Next Generation Space Telescope ({\it e.g.}
Miralda-Escude and Rees 1997, Madau 1998, Livio 1999). We will not
analyse the feasibility of such observations here. Rather, we note
from Figures 1 and 2 that the major axis of the error ellipses in the
$\Omega_\Lambda$-- $\Omega_m$ tilt and become more vertical as the
redshift range of the SN sample is increased. This is because the
magnitude redshift relations for models with very different values
values of $\Omega_\Lambda$ and the same $\Omega_m$ converge at higher
redshifts. The convergence redshift depends on $\Omega_m$ and lies
between $z \approx 2$--$4$ for $\Omega_m$ in the range $0.2$--$1$ (see
Figure 1 of Melnick, Terlevich and Terlevich, 1999).
This is illustrated by Figure 3, which shows the $1$, $2$ and $3
\sigma$ likelihood contours determined from the Fisher matrix for a
sample consisting of twice the SCP sample, $100$ SN with $\langle z
\rangle = 1.5$, $\Delta z = 0.5$, and $40$ SN with $\langle z \rangle
= 3$, $\Delta z = 1$. As expected, these contours are almost vertical
in the $\Omega_m$--$\Omega_\Lambda$ plane. A sample of supernovae (or
some other distance indicator such as HII galaxies, Melnick {\it et al. }
1999) at redshifts $z \sim 3$ can therefore produce a tight constraint
on $\Omega_m$ independently of the value of
$\Omega_\Lambda$.
\section{Constraints on an Arbitrary Equation of State}
In this Section, we analyse the constraints that SN can place on an
arbitrary equation of state. We first consider a constant equation of
state. Models of this type (see Bucher and Spergel) include a
frustrated network of cosmic strings ($p/\rho = -1/3$) and a
frustrated network of domain walls ($p/\rho = -2/3$). A constant
equation of state is also a good approximation to a Q component
obeying tracker solutions. Tracker solutions are discussed in Section
3.2. Constraints on generalised forms of dark matter with
anisotropic stress are discussed by Hu {\it et al. } (1999) and will not be
considered here.
\subsection{Constant equation of state}
\begin{figure*}
\vskip 3.0 truein
\special{psfile=pgfish2a.ps
hscale=45 vscale=45 angle=-90 hoffset= -10
voffset=240}
\special{psfile=pgfish2b.ps
hscale=45 vscale=45 angle=-90 hoffset= 170
voffset=240}
\caption
{As Figure 2, but for an arbitrary constant equation of state in a
spatially
flat Universe.
The dashed lines in each panel show $1$, $2$ and $3 \sigma$ likelihood
contours in the $w_Q$--$\Omega_m$ plane for the SCP distant
supernova sample as analysed in Section 4 (assuming a constant
equation of state). The solid contours are derived from
the Fisher matrix for enhanced samples of high redshift supernovae
and the points show maximum likelihood derived from
Monte-Carlo realizations of these samples.}
\label{figure4}
\end{figure*}
If we include a Q-like component with equation of
state $p/\rho = w_Q$, the expression for the term $x$ in the
luminosity distance (equation 2) is modified to
\begin{eqnarray}
x(z, \Omega_m, \Omega_Q, w_Q) = \qquad \qquad \qquad \qquad \qquad
\qquad \qquad \nonumber\\
\int_0^z {dz^\prime
\over
[\Omega_m (1 + z^\prime)^3 + \Omega_k ( 1 + z^\prime)^2 +
\Omega_Q(1+z^\prime)^{3(1+w_Q)}]^{1/2} }.
\end{eqnarray}
The addition of the parameter $w_Q$ means that it is not possible to
constrain all of the parameters $\Omega_m$, $\Omega_Q$, $w_Q$ to high
accuracy from the supernova data alone (see Section 4.2). Thus,
Garnavich {\it et al. } (1998) analyse the High-z Supernovae Search (HZS)
sample (Riess {\it et al. } 1999) assuming a spatially flat universe and find
that $w_Q < -0.55$ at $95\%$ confidence. A similar analysis of the
SCP sample by Perlmutter, Turner and White (1999) yields $w_Q \lower.5ex\hbox{\ltsima} -0.5$.
Table 2 lists the results of a Fisher matrix analysis for a Q-like
component with a constant $w_Q$. Here we have applied the constraints
$w_Q \ge -1$ and $\Omega_Q \ge 0$. The upper table gives results for
the supernovae magnitude-redshift relation alone assuming a spatially
flat Universe with $\Omega_m = 0.25$ and $w_Q = -1$. The constraints
on $w_Q$ from a sample such as the SCP data are quite poor and
improve relatively slowly as the sample is extended to higher
redshift because of a strong degeneracy between
$w_Q$ and $\Omega_m$ in the magnitude-redshift relation. This is
illustrated in Figure 4, which shows the analogue of Figure
2 for Q-like models. As the supernovae sample is extended to higher
redshift, the likelihood contours narrow but $w_Q$ and $\Omega_m$
remain strongly degenerate.
\begin{table}
\label{tab2}
\centerline{\bf Table 2: Fisher Matrix Errors, $\Omega_m$, $\Omega_Q$ and $w_Q$.}
\begin{center}
\begin{tabular}{|cccccc|} \hline
\multicolumn{6}{c} {Supernovae Alone, $\Omega_k = 0$} \\
& SCP & \multicolumn{2}{c}{SCP + 20 SN} &
\multicolumn{2}{c}{2$\times$SCP + 40SN} \\
$<z>$ & & $1.0$ & 1.5 & 1.0 & 1.5 \\
$\delta \Omega_m$ & $0.14\;\;$ & $0.12\;\;$ & $0.097$ & $0.097$ & $0.073$ \\
$\delta w_Q$ & $0.36\;\;$ & $0.35\;\;$ & $0.32\;\;$ & $0.28\;\;$ & $0.24\;\;$\\
$\delta {\cal M}$ & $0.051$ & $0.051$ & $0.051$ & $0.037$ & $0.036$\\\hline
\multicolumn{6}{c} {Supernovae +CMB, $\Omega_k = 0$} \\
& SCP & \multicolumn{2}{c}{SCP + 20 SN} &
\multicolumn{2}{c}{2$\times$SCP + 40SN} \\
$<z>$ & & $1.0$ & 1.5 & 1.0 & 1.5 \\
$\delta \Omega_m$ & $0.027$ & $0.022$ & $0.0210$ & $0.016$ & $0.015$ \\
$\delta w_Q$ & $0.10\;\;$ & $0.085$ & $0.081$ & $0.061$ & $0.057$\\
$\delta {\cal M}$ & $0.048$ & $0.046$ & $0.045$ & $0.032$ & $0.032$\\\hline
\multicolumn{6}{c} {Supernovae + CMB, $\Omega_k \ne 0$} \\
& SCP & \multicolumn{2}{c}{SCP + 20 SN} &
\multicolumn{2}{c}{2$\times$SCP + 40SN} \\
$<z>$ & & $1.0$ & 1.5 & 1.0 & 1.5 \\
$\delta \Omega_m$ & $0.14\;\;$ & $0.12\;\;$ & $0.095$ & $0.094$ & $0.069$ \\
$\delta \Omega_Q$ & $0.10\;\;$ & $0.083$ & $0.066$ & $0.067$ & $0.048$ \\
$\delta w_Q$ & $0.31\;\;$ & $0.31\;\;$ & $0.27\;\;$ & $0.24\;\;$ & $0.20\;\;$\\
$\delta {\cal M}$ & $0.051$ & $0.051$ & $0.051$ & $0.036$ & $0.036$\\\hline
\end{tabular}
\end{center}
\end{table}
The situation is dramatically improved by the addition of constraints
from CMB anisotropies. The addition of a Q-like component
affects the location of the Doppler peaks (see Caldwell {\it et al. } 1998,
White 1998) and, in analogy with equation (5), an accurate
determination of the CMB power spectrum imposes the constraint
\begin{equation}
\Delta \Omega_Q = - {(\partial \gamma_D/\partial w_Q)
\over (\partial \gamma_D/\partial \Omega_Q)} \Delta w_Q -
{(\partial \gamma_D/\partial \Omega_m)
\over (\partial \gamma_D/\partial \Omega_Q)} \Delta \Omega_m,
\end{equation}
The second panel of Table 2 shows the constraints derived on an
arbitrary equation of state by combining supernovae data with the CMB
constraint of equation (7). For spatially flat models, the
combination of SN and CMB anisotropies constrains $w_Q$ to an
accuracy of better than $0.1$, sufficient to set tight constraints
on the physical parameters of Q-like models (for example, whether
one requires contrived potentials, see Section 4). However, the constraints
on $w_Q$ improve relatively slowly as the SN sample is extended to
higher redshift. Similar conclusions apply if the assumption of a
spatially flat universe is relaxed (see the lower panel of Table 2).
In that case, the parameters $\Omega_m$ and $\Omega_Q$ can be
determined to high precision, but the constraints on $w_Q$ improve
slowly as the SN sample is increased. This implies that it is worth
analysing the constraints on Q-like models with arbitrary spatial
curvature using current SN and CMB data (see Section 4.2).
\begin{figure}
\vskip 5.4 truein
\special{psfile=pgtrack1a.ps
hscale=40 vscale=40 angle=-90 hoffset= -25
voffset=480}
\special{psfile=pgtrack1b.ps
hscale=40 vscale=40 angle=-90 hoffset= -25
voffset=350}
\special{psfile=pgtrack1c.ps
hscale=40 vscale=40 angle=-90 hoffset= -25
voffset=220}
\caption
{The evolution of the equation of state $w_Q$ and its contribution to
the cosmic density parameter $\Omega_Q$ as a function of redshift
derived from solutions to the tracker equation (8) for three
potentials: $V(Q) = M^4({\rm exp}(1/Q) - 1)$ (figures 5a and 5b); $V(Q)
= M^4(M/Q)^2$ (figures 5c and 5d); $V(Q) = M^4(M/Q)^6$ (figures 5e and
5f). The curves in each figure are computed by varying the parameter
$M$, with more negative values of $w_Q$ corresponding to higher values
of $\Omega_Q$.}
\label{figure5}
\end{figure}
\begin{figure}
\vskip 5.4 truein
\special{psfile=pgtrack2a.ps
hscale=40 vscale=40 angle=-90 hoffset= -25
voffset=480}
\special{psfile=pgtrack2b.ps
hscale=40 vscale=40 angle=-90 hoffset= -25
voffset=350}
\special{psfile=pgtrack2c.ps
hscale=40 vscale=40 angle=-90 hoffset= -25
voffset=220}
\caption
{The left hand panels show the tracker solution relations between
$\Omega_Q$ and $w_Q$ at the present day for the three potentials used
in Figure 5. The right hand panels show the derivative $\alpha \equiv
\partial w_Q /\partial {\rm ln}a$ for the tracker solutions as a
function of $w_Q$.}
\label{figure6}
\end{figure}
\subsection{Time varying equation of state: tracker solutions}
In the previous section we have investigated the simplified case of a
constant $w_Q$. If, in fact, the Q-like component arises from a slowly
rolling scalar field evolving in a potential $V(Q)$, the equation of
state of the $Q$ component will vary as a function of time. The
equations of motion of the $Q$ field can be written in the following
compact form (Steinhardt, Wang and Zlatev, 1998)
\begin{eqnarray}
{V^{\prime\prime} V \over (V^\prime)^2} = 1 + {w_B - w_Q \over 2 (1 + w_Q)}
- { 1 + w_B - 2w_Q \over 2 (1 + w_Q)} { \dot x \over 6 + \dot x} \nonumber \\
- {2 \over (1+w_Q)} {\ddot x \over ( 6 + \dot x)^2}, \qquad
x \equiv {(1+w_Q) \over (1 - w_Q)} \qquad
\end{eqnarray}
where primes denote derivatives with respect to $Q$, $\dot x = d {\rm
ln}x/d {\rm ln a}$, $\ddot x = d^2 {\rm ln}x/d {\rm ln^2 a}$ and $a$
is the scale factor of the cosmological model. For a wide class of
potentials, and almost independently of the initial conditions, the
evolution of $Q$ locks on to a tracking solution in which $Q$ and
$w_Q$ vary slowly (see Zlatev, Wang and Steinhardt 1998, Steinhardt
{\it et al. } 1998). Examples of the evolution of $w_Q$ and $\Omega_Q$ at late
times are shown in Figure 5 for three forms of the potential
$V(Q)$. In each case, the evolution of $w_Q$ at $z \lower.5ex\hbox{\ltsima} 4$ is well
approximated by
\begin{equation}
w_Q = w_Q(a_0) + \alpha {\rm ln} (a/a_0)
\end{equation}
where $\alpha$ is a small number determined from the value of $\dot x$
at the present time.
Figure 7 shows the relations between $\Omega_Q$, $w_Q$ and $\alpha$ at
the present time derived from the solutions to equation 4
for the three potentials considered in Figure 5. The minimum value of
$\alpha$ is about $-0.14$, reflecting the fact that $Q$ is evolving
relatively slowly even at late times.
With the approximation of equation (9), the energy density of the $Q$
component evolves according to
\begin{equation}
{\rho_Q(a) \over \rho_Q(a_0)} = \left ( {a \over a_0}
\right)^{-3(1+w_Q(a_0))} {\rm exp} \left( - {3 \over 2} \alpha [ { \rm
ln} (a/a_0) ]^2 \right).
\end{equation}
Note also that with the approximation of equation (9), the
tracker equation (8) becomes an algebraic equation relating
${V^{\prime\prime} V \over (V^\prime)^2}$ to $w_Q$, $\Omega_Q$ and
$\alpha$ ($w_B = w_Q \Omega_Q$ in the matter dominated era).
A small value of $\alpha \sim -0.1$ to $-0.2$ cannot be determined
accurately from SN and CMB observations because it is highly
degenerate with $w_Q$ and $\Omega_m$. As we will show in the next
Section, the introduction of the parameter $\alpha$ provides a
convenient way of testing the sensitivity of constraints on Q-like
models to the time evolution of $w_Q$.
\begin{figure*}
\vskip 3.0 truein
\special{psfile=pglike_sn1.ps
hscale=45 vscale=45 angle=-90 hoffset= -30
voffset=240}
\special{psfile=pglike_sn_cmb1.ps
hscale=45 vscale=45 angle=-90 hoffset= 180
voffset=240}
\caption
{Constraints of $w_Q$ and $\Omega_m$ for spatially flat universes.
Figure 7a shows results for the SCP supernova sample following a
similar analysis to that presented by E99. Figure 7b shows results
for the supernova sample combined with the constraints from the CMB
anisotropy measurements as described in E99. The contours show $1$,
$2$ and $3 \sigma$ likelihood contours.The solid contours are derived
for $\alpha=0$, dotted contours are for $\alpha = -0.1$ and dashed
contours for $\alpha = -0.2$.}
\label{figure7}
\end{figure*}
We note that Huterer and Turner (1998) have recently proposed a
prescription for reconstructing the potential of a $Q$-like component
directly from the magnitude-redshift relation of Type Ia
supernovae. This approach may produce interesting constraints if the
field $Q$ is rapidly evolving at late times. For tracker solutions,
however, the equation of state changes so slowly that it would be
difficult to distinguish the true potential from a perfectly flat one.
\section{Limits on the equation of state
from Type Ia Supernovae and the Cosmic Microwave Background}
\subsection{Spatially flat models}
In this Section, we use current SN and CMB data to constrain the
equation of state of the Universe. The analysis closely follows that
presented in E99. We use the sample of $56$ Type Ia SN of fit C of P98
and adopt the likelihood analysis described by E99 (including a
parametric fit to the luminosity-decline rate correlation), modifying the
expression for luminosity distance to incorporate the parameters of
the Q-like model. The CMB data that we use are plotted in Figure 1 of
E99. We perform a likelihood analysis for these data assuming
scalar adiabatic perturbations, varying the amplitude of the
fluctuation spectrum, the scalar spectral index, the physical
densities of the CDM and baryons $\omega_c = \Omega_c h^2$, $\omega_b
= \Omega_b h^2$ \footnote{$h$ is the Hubble constant in units of $100\;
{\rm km} {\rm s}^{-1} {\rm Mpc}^{-1}$.}, and the Doppler peak location
parameter $\gamma_D$. Modifications to the CMB power spectrum arising
from spatial fluctuations in the Q component are ignored as these are
negligible in the slowly evolving $Q$ models considered here (see
Caldwell {\it et al. } 1998, Huey {\it et al. } 1998). We integrate over the CMB
likelihood assuming uniform prior distributions of the parameters to
compute a marginalized likelihood for $\gamma_D$ as described in E99.
The likelihood functions for the parameters $w_Q$, $\Omega_m$ and
$\Omega_Q$ presented below are constructed from the expression for the
angular diameter distance to the last scattering surface and the
probability distribution of $\gamma_D$.
Figure 7 shows the constraints on $w_Q$ and $\Omega_m$ for spatially
flat universes. The different line types show the constraints for
three different values of the parameter $\alpha$ characterising the
evolution of $w_Q$, $\alpha = 0$ (solid lines), $\alpha = -0.1$
(dotted lines) and $\alpha = -0.2$ (dashed lines). As described in
the previous section, these values span the range found for tracker
solutions for a variety of potentials. These rates of evolution are so
low that they have very little effect on the likelihood contours. The
constraints plotted in Figure 7 are in very good agreement with those
derived by Garnavich {\it et al. } (1998) from an analysis of the HZS sample,
and with the analysis of the SCP sample (Perlmutter, Turner and White
1999) and of the combined HZS and SCP samples (Wang {\it et al. } 1999). The
fact that the constraints are weakly dependent on the size of the SN
sample is a consequence of the strong degeneracy between $w_Q$ and
$\Omega_m$ discussed in Section 3.2.
\begin{figure*}
\vskip 5.5 truein
\special{psfile=pglike_sn2.ps
hscale=60 vscale=60 angle=-90 hoffset= 20
voffset=510}
\special{psfile=pglike_sn_cmb2.ps
hscale=60 vscale=60 angle=-90 hoffset= 20
voffset=310}
\caption
{Analogue of Figure 5, but for quintessence models with arbitrary
spatial curvature. Figures 7a and 7b show marginalized likelihoods
in the $w_Q$--$\Omega_m$ and $\Omega_Q$--$\Omega_m$ planes derived
from Type Ia supernovae. Figures 7c and 7d show the combined
likelihoods for the Type Ia and CMB anisotropies. As in Figure 5,
the solid contours are derived for $\alpha=0$ and
dashed contours for $\alpha = -0.2$.}
\label{figure8}
\end{figure*}
Figure 7b shows the results of combining the SN likelihoods with those
determined from the CMB. The likelihood peaks at $w_Q = -1$, $\Omega_m
= 0.29$. Qualitatively, these results are similar to those of
Perlmutter {\it et al. } (1999); the favoured cosmology has an equation of
state $w_Q = -1$ and $w_Q$ is contrained to be less than $-0.6$ at the
$2 \sigma$ level. However, in detail, the constraints in Figure 7b are
somewhat less stringent than those of Perlmutter {\it et al.},, allowing a
broader range in $\Omega_m$ ($0.15 \lower.5ex\hbox{\ltsima} \Omega_m \lower.5ex\hbox{\ltsima} 0.5$ at the
$2 \sigma$ level). This is because Perlmutter {\it et al. } include
constraints on the power spectrum of galaxy clustering based on the
data compiled by Peacock and Dodds (1994)\footnote{Perlmutter {\it et al. }
(1999) do not combine the SN and CMB likelihoods but analyse the SN
data assuming a spatially flat Universe.}. In our view this is
dangerous because it requires a specific assumption concerning the
distribution of galaxies relative to the mass. Qualitatively, for
nearly scale-invariant adiabatic models, galaxy clustering imposes a
constraint on the parameter combination $\Gamma = \Omega_m h$ of $0.2
\lower.5ex\hbox{\ltsima} \Gamma \lower.5ex\hbox{\ltsima} 0.3$, if galaxies are assumed to trace the mass
fluctuations on large scales (Efstathiou, Bond and White 1992, Maddox,
Efstathiou and Sutherland 1996). Combined with measurements of the
Hubble constant (for which Perlmutter {\it et al. } adopt $h \approx 0.65 \pm
0.05$), galaxy clustering leads to a constraint of $0.25 \lower.5ex\hbox{\ltsima}
\Omega_m \lower.5ex\hbox{\ltsima} 0.5$, partly breaking the degeneracy between $w_Q$ and
$\Omega_m$. The combined SN and CMB analysis in Figure 7b provides
constraints which are nearly as tight, but are much less model
dependent.
The constraints of Figure 7b place strong limits on Q-like models. For
tracking solutions, the constraint $\;w_Q \lower.5ex\hbox{\ltsima}$ $-0.6$ excludes steep
potentials ({\it e.g.} $V(Q) \propto Q^{-\beta}$ with $\beta \lower.5ex\hbox{\gtsima}
2$) and the data clearly favour a standard cosmological term ($w_Q =
-1$). These limits on $w_Q$ are very close to the lower limit
($w_Q \lower.5ex\hbox{\gtsima} -0.7$) allowed for `physically well motivated' tracker
solutions (Steinhardt, Wang and Zlatev, 1998, {\it i.e.}
smooth potentials with simple functional forms). With a slight
improvement of the observations one may be forced to fine-tune the
shape of the potential to construct a viable quintessence model.
The constraints of Figure 7b are somewhat stronger than those of Wang
{\it et al. } (1999), who perform a `concordance analysis' of Q-like models
using a number of observational constraints including those from Type
Ia supernovae and CMB anisotropies. These authors conclude limits of
$-1 \lower.5ex\hbox{\ltsima} w_Q \lower.5ex\hbox{\ltsima} -0.4$. The difference is caused by the different
methods of statistical analysis. The concordance analysis of Wang
{\it et al. } leads, by construction, to more conservative limits than the
maximum likelihood analysis and is more robust to systematic errors in
any particular data set. However, provided systematic errors are
negligible in the CMB and SN datasets, then the constraints of Figure
7b derived by combining likelihoods should be realistic. These small
differences in the upper limits on $w_Q$ are important because they
can place significant restrictions on the physics. As stressed in the
previous paragraph, the upper limit of $w_Q \approx -0.6$ places
strong constraints on tracker models with simple potentials.
\subsection{Models with arbitrary spatial curvature}
Figure 8 shows the results of a likelihood analysis of the SN and CMB
data, but now allowing arbitrary spatial curvature. We show two
projections of the likelihood distributions, marginalizing over
$\Omega_Q$ in Figures 8(a) and 8(c) and over $w_Q$ in Figures 8(b) and
8(c). The constraints, although weaker than those presented in Figure
7, are interesting nevertheless. The combined SN and CMB likelihoods
give a $2\sigma$ upper limit on $w_Q$ of $w_Q \lower.5ex\hbox{\ltsima} -0.4$ (Figure 8d)
and a maximum likelihood solution of $\Omega_m = 0.12$, $\Omega_Q =
0.73$ irrespective of the value of $w_Q$. Evidently, the SN and CMB
data constrain us to a nearly spatially flat Universe dominated either
by a cosmological constant, or a Q-like component with an equation of
state $w_Q \lower.5ex\hbox{\ltsima} -0.4$.
\section{Conclusions}
Observations of distant Type Ia supernovae have provided important
evidence that the Universe may be dominated by a cosmological constant
(P98, Riess {\it et al. } 1999). However, the constraints in the
$\Omega_\Lambda$--$\Omega_m$ plane from current data are degenerate
along a line defined by $\Omega_\Lambda \approx 0.32 + 1.43 \Omega_m$
(Figure 1). This degeneracy can be reduced significantly by extending
the redshift range of the supernovae sample. For example, with $20$
additional supernovae at redshift $z \sim 1.5$ the errors in
$\Omega_m$ and $\Omega_\Lambda$ could be reduced to $\delta \Omega_m
\approx 0.08$ and $\delta \Omega_\Lambda \approx 0.22$. A sample of
supernovae at $z \lower.5ex\hbox{\gtsima} 3$ could provide an accurate estimate of
$\Omega_m$ that is independent of the value of $\Omega_\Lambda$.
The combination of supernovae and CMB anisotropy measurements can
break the degeneracy between $\Omega_\Lambda$ and $\Omega_m$ if the
initial fluctuations are assumed to be adiabatic and characterised by
a smooth fluctuation spectrum. This method applied to recent
supernovae and CMB data suggests a nearly spatially flat universe
dominated by a cosmological term with $\Omega_\Lambda \approx 0.65$
(Lineweaver 1998, Garnavich {\it et al. } 1998, Tegmark 1999, E99). The only
plausible way of avoiding this conclusion is to appeal to some
systematic effect in the supernovae data, for example, grey dust or an
evolutionary effect in the supernovae data such as a metallicity
dependence (see {\it e.g.} P98 for a discussion). The degeneracy
breaking afforded by extending the supernovae data to higher redshift
would provide an important consistency check of such systematic
effects and also on the interpretation of the CMB anisotropy data
The constraints on quintessence-like models with an equation of state
$w_Q = p/\rho$ improve relatively slowly as the supernovae data are
extended to higher redshift. The most promising way of constraining
$w_Q$ seems to be to combine supernovae and CMB measurements. We have
carried out a joint likelihood analysis of CMB anisotropy observations
and the SCP supernovae data. For a spatially flat Universe we derive a
$2\sigma$ upper limit of $w_Q = -0.6$. For universes of arbitrary
spatial curvature, the $2 \sigma$ upper limit is $w_Q = -0.4$. The
combined SN and CMB likelihood peaks at $\Omega_m = 0.12$ and
$\Omega_Q = 0.73$ irrespective of the value of $w_Q$, suggesting that
the Universe is almost spatially flat. The $2 \sigma$ upper limit of
$w_Q = -0.6$ for spatially flat Universes is close to the minimum value
of $w_Q \approx -0.7$ allowed for simple quintessence-models. This
suggests that some fine tuning of the potential may be required to
construct a viable quintessence model.
\vskip 0.2 truein
\noindent
{\bf Acknowledgements.} I thank Richard Ellis, Paul
Steinhardt and Roberto Terlevich
for useful discussions and PPARC for the award of a Senior
Fellowship. I also thank Sarah Bridle, Anthony Lasenby, Mike Hobson
and Graca Rocha for allowing me to use their compilation of CMB
anisotropy data.
|
\section{Introduction}
The problem of obtaining the time involved in the tunneling process
in quantum mechanics is still a controversial issue, despite
considerable efforts in recent years \cite{Reviews}.
In particular, in order to address this issue,
some authors have analyzed the tunneling through time-modulated
potential barriers \cite{Bula82,Bula85,Soko88,Garcia90,Leav91,Leav93,Pimp91,Azbel,AzMa93,LanMar93,Stov93,MarSa95,Wag95}.
One of the pioneer works in this area is the model introduced by
B\"uttiker and Landauer in 1982 \cite{Bula82} in which they consider the
transmission through a time-modulated rectangular barrier, and
introduced a characteristic time for the process.
However, in the above-mentioned papers, there is practically no
mention of the corresponding classical problem;
although the classical limit is straightforward when the potential
barrier does not depend on time,
it is far from trivial when the potential is time modulated.
In this paper I study the classical problem of a rectangular
time-modulated potential barrier, in order to analyze in detail the
traversal time distribution for an ensemble of classical particles.
This classical model was inspired, in part, by the B\"uttiker-Landauer
model mention above.
I will study first the case of a potential barrier located inside
a rigid box \cite{Mateos}. In this case, the classical orbits can
be periodic, quasiperiodic or even chaotic, depending on the parameters and
the initial conditions of the motion. In order to study the dynamics, I
derive first an area-preserving map that allow us to find the orbits for
all times. Then, I study the scattering problem of an ensemble of
particles that interact with an oscillating rectangular potential barrier.
In this case, I will show that the traversal time strongly
depends on the arrival time of the incident particles.
There is a basic difference between these two problems:
(1) In the first case, what we have is the bounded problem of
an oscillating barrier inside a rigid box of finite size. This means
that an incident particle interact with the barrier not once but an
arbitrary number of times, since the particle can cross the barrier
region and then, after bouncing elastically in the box, returns to
the oscillating barrier. Then, the dynamics can
become chaotic, since we have the main ingredients: on one hand,
sensitive dependence on initial condition or arrival times due to
the oscillating barrier, and on the other hand, bounded motion due to
presence of the finite box.
(2) In the second case, we have a scattering problem in which
an incident particle interacts with the barrier only once. Of course, if
this is the case, the problem is straightforward and there is only a
single traversal time. But, if we consider an ensemble of $N$
noninteracting particles with slightly different initial conditions,
say different initial velocities, then we can expect, in general,
$N$ different traversal times that can exhibit a complex distribution
of traversal times.
An approach to the problem of tunnelling times, that is closely related
to the classical trayectories discussed here, is the Bohm
trajectory point of view \cite{Holl}. This approach has been used by
Leavens and Aers \cite{Leav95} to give an unambiguous prescription for
calculating traversal times that are conceptually meaningful within that
interpretation. In particular, Leavens and Aers \cite{Leav91,Leav93}
have treated in detail the case of a time-modulated rectangular barrier,
using Bohm's trajectory interpretation of quantum mechanics \cite{Holl}.
They calculate, among other things, transmission time distributions,
the transmission probability as a function of frequency and Bohm
trajectories.
\section{The model and the map}
Let us study the classical dynamics of a particle in a one-dimensional box,
inside of which there is an oscillating rectangular potential
barrier \cite{Mateos}. This problem consists of a particle moving in one dimension under the action of a time-dependent potential $V(x,t)$. Since
the Hamiltonian of this system is time dependent, the total energy of the particle is not conserved. The Hamiltonian is given by
$H(x,p,t)=p^2/2m + V(x,t)$, where
\begin{equation}
V(x,t)=V_0(x)+V_1(x)f(t).
\end{equation}
The potential $V_0(x)$ goes to infinity when $x<0$ or $x>l+b+L$, is equal to
the constant value $V_0$ when $l\le x\le l+b$, and otherwise is equal to
zero. Thus, what we have is an infinite potential well with a rectangular
potential barrier of width $b$ inside, as shown in Fig. 1a. This potential separates the box in
three regions: region I, $0\le x<l$ of width $l$; region II, where the
rectangular barrier is located, $l\leq x\le l+b$ of width $b$; and region
III, $l+b<x\le l+b+L$ of width $L$.
Clearly, the motion of a particle under the influence of the potential
$V_0(x)$ is regular, that is, we have periodic orbits and the energy is
conserved. However, if we add a time-dependent potential, we can obtain
periodic, quasiperiodic and chaotic orbits, as we will show below. The
potential $V_1(x)$ in eq. (1) is different from zero only inside the
interval $l\le x \le l+b$, where it takes the constant value $V_1$. The
function $f(t)$ in eq. (1) is assumed periodic with period $\tau$, that
is, $f(t+\tau) = f(t)$. In this way, as shown in Fig. 1a, what we have is an
oscillating potential barrier, with an amplitude which oscillates between
$V_0 - V_1$ and $V_0 + V_1$, with frequency $\omega/2\pi$ and period
$\tau=2\pi/\omega$. We will take $V_0>V_1$.
Let us now derive a map that describes the dynamics of a particle under this
potential. The motion is as follows: at the fixed walls at $x=0$ and $x=l+b+L
$, the particle bounces elastically, changing the sign of the velocity but
with the same absolute value. The other two points where there is a change
in the velocity is at the borders of the potential barrier at $x=l$ and
$x=l+b$. The rest of the time the velocity is constant. Thus,
the particle can gain or loose kinetic energy at $x=l$ and $x=l+b$.
The phase space for a typical orbit is depicted in Fig. 1b.
We can analyze the dynamics using a discrete map from the time $t_n$ when
the particle hits the wall at $x=0$, until the next time $t_{n+1}$ when it
hits this wall again. Let us denote by $v_n$ the velocity of the particle
immediately after the $n-th$ kick with the fixed wall at $x=0$, and by
$E_n$ the corresponding total energy. Clearly, $E_n=m{v_n}^2/2$. After
traveling the distance $l$, it arrives at the left side of the barrier
after a time of flight $l/v_n$, where a change in the velocity occurs.
To determine this change let us consider the following: In region I,
the total energy of the
particle is given by $E_n=m{v_n}^2/2$ which is just the kinetic energy,
because in this region the potential energy is zero; when the particle
enters region II, the kinetic energy $E_n^{\prime }$ is changed to
$E_n-V_0-V_1f(t_n+l/v_n)$, that is, the total energy minus the value of the
potential energy at the time of arrival $t_n+l/v_n$. If we denote the new
velocity by $v_n^{\prime }$ (see Fig. 1b), then
$E_n^{\prime }=m{v_n^{\prime }}^2/2$ and
we obtain in this way the change in energy as:
\begin{equation}
E_n^{\prime }=E_n-V_0-V_1f\biggl(t_n+{\frac l{v_n}}\biggr).
\end{equation}
Clearly, if the total energy is less than the potential energy at time
$t_n+l/v_n$, then the particle cannot penetrate region II and simply
reflects elastically and there is only a change in the sign of the
velocity; thus the particle gets trapped in region I and returns to the
wall at $x=0$. After a
time lapse of $2l/v_n$ it will hit again the oscillating barrier and try
again to cross it. If this time the total energy is greater than the
potential energy, then the particle can cross the barrier region;
otherwise, it bounces once more inside region I, and so on.
Now, once the particle overcomes the barrier, it crosses the region II
without changing its velocity $v_n^{\prime }$, even though the barrier is
oscillating in time. When the particle arrives at the right side of the
barrier at $x=l+b$, then another change in the velocity takes place, but
this time the velocity increases in such a way that the kinetic energy $
E_n^{\prime \prime }$ becomes
\begin{equation}
E_n^{\prime \prime }=E_n^{\prime }+V_0+V_1f \biggl(t_n+{\frac l{v_n}}+ {%
\frac b{v_n^{\prime }}}\biggr),
\end{equation}
where $E_n^{\prime \prime }$ is the energy in region III. Clearly, the time
that it takes to arrive at the wall located at $x=l+b+L$ is $%
l/v_n+b/v_n^{\prime }+L/v_n^{\prime \prime }$, where $v_n^{\prime \prime }$
is the velocity in region III (see Fig. 1b). After a time $%
t_n+l/v_n+b/v_n^{\prime}+2L/v_n^{\prime\prime}$, the particle returns to the
right side of the barrier after traveling twice the distance $L$ in region III, and enters once again region II. However, in general, the potential barrier has a different height, given by $V_0+V_1f(t_n+l/v_n+b/v_n^{\prime}+2L/v_n^{\prime\prime})$. Therefore, the new kinetic energy $E_n^{\prime\prime\prime}$ inside region II is now
given by
\begin{equation}
E_n^{\prime\prime\prime} = E_n^{\prime\prime} - V_0 - V_1f\biggl(t_n+{\frac
{l}{v_n}}+{\frac {b}{v_n^{\prime}}} +{\frac { 2L}{v_n^{\prime\prime}}}\biggr),
\end{equation}
Here, once more, there is the possibility that the total energy in region
III is less than the potential energy at time $t_n+l/v_n+b/v_n^{
\prime}+2L/v_n^{\prime\prime}$. In this case, the particle gets trapped in
region III until it can escape by crossing the barrier region.
Finally, after a time $b/|v_n^{\prime\prime\prime}|$, where $
v_n^{\prime\prime\prime}$ is the velocity in region II (see Fig. 1b),
the particle arrives
at the left side of the barrier at $x=l$, where the velocity varies once
more depending on the height of the barrier at time
$t_n+l/v_n+b/v_n^{\prime}+2L/v_n^{\prime\prime}+
b/|v_n^{\prime\prime\prime}|$. We will denote
the velocity in region I, after this time, by $v_{n+1}$, because this is
precisely the velocity after the next hit with the wall at $x=0$. The last
part of this journey is covered in a time span of $l/|v_{n+1}|$; after this,
the particle hits the wall at the origin at time $t_{n+1}$ and start again
its trip to the oscillating barrier, and the whole process starts again.
Therefore we arrive at the following map in terms of energy and time:
\begin{equation}
E_{n+1}=E_n^{\prime \prime \prime }+V_0+V_1f(t_n+T_n)
\end{equation}
and
\begin{equation}
t_{n+1}=t_n+T_n+\sqrt{\frac m{2}} {\frac l{\sqrt{E_{n+1}}}},
\end{equation}
where $T_n$ is given by
\begin{equation}
T_n=\sqrt{\frac m{2}}\biggl({\frac l{\sqrt{E_n}}}+ {\frac b{\sqrt{%
E_n^{\prime }}}}+ {\frac{2L}{\sqrt{E_n^{\prime \prime }}}}+ {\frac b{\sqrt{%
E_n^{\prime \prime \prime}}}}\biggr)
\end{equation}
and $E_n^{\prime }$, $E_n^{\prime \prime }$ and $E_n^{\prime \prime \prime }$
are given by eqs. (2-4), respectively.
Furthermore, it can be shown that, for this
map, the Jacobian is exactly one, that is,
\begin{equation}
J={\frac{\partial (E_{n+1},t_{n+1})}{\partial (E_n,t_n)}}=1.
\end{equation}
This result indicates that this map is an area-preserving one \cite{Lich}.
Let us scale the time using the period $\tau $ of the function $f(t)$. We
define the dimensionless quantities: $\phi _n=(2\pi /\tau )t_n$ and $\Phi
_n=(2\pi /\tau )T_n$. In order to scale the energies we introduce the
dimensionless variables: $e_n=E_n/V_0$,
$e_n^{\prime }=E_n^{\prime }/V_0$,
$e_n^{\prime \prime }=E_n^{\prime \prime }/V_0$ and
$e_n^{\prime \prime \prime }=E_n^{\prime \prime \prime }/V_0$.
With all this definitions we arrive at the following dimensionless map:
\begin{equation}
e_{n+1}=e_n^{\prime \prime \prime }+1+rf\bigl(\phi _n+\Phi _n\bigr),
\end{equation}
and
\begin{equation}
\phi _{n+1}=\phi _n+\Phi _n+{\frac{2\pi M}{\sqrt{e_{n+1}}}},\qquad
(mod 2\pi )
\end{equation}
where $M=l/(w\tau )$, $r=V_1/V_0$ and $w=(2V_0/m)^{1/2}$. Here, $\Phi _n$ is
given by
\begin{equation}
\Phi _n=2\pi M\biggl({\frac 1{\sqrt{e_n}}}+{\frac bl}{\frac 1{\sqrt{%
e_n^{\prime }}}}+{\frac{2L}l}{\frac 1{\sqrt{e_n^{\prime \prime }}}}+{\frac bl%
}{\frac 1{\sqrt{e_n^{\prime \prime \prime }}}}\biggr).
\end{equation}
This map, although more complicated, resembles the structure of the Fermi
Map \cite{Lich}.
\section{Numerical results}
Let us now analyze numerically the map obtained above. First of
all, we notice that we have four dimensionless parameters: the width of the
barrier $b/l$ scaled with the length of region I; the length $L/l$ of region III scaled with $l$; the ratio of the amplitude of oscillation of the
barrier scaled with its height $r=V_1/V_0$; and $M=l/(w\tau )$. The
parameter $M$ is the ratio of the time of flight $l/w$ in region I of
Fig. 1a, with velocity $w$, and the period $\tau$ of oscillation of the barrier. That is, $M$ measures the number of oscillations of the barrier since the particle leaves the wall at $x=0$ until it arrives at the left side of the barrier. On the other hand, we will take the periodic function as:
$f(\phi _n)=\sin(\phi _n)$.
If we fix the barrier position within the one-dimensional box, and choose a
width, then we are fixing the parameters $b/l$ and $L/l$; the remaining two
parameters $M$ and $r$ will control the type of motion. In what follows,
we take the symmetric case, $b/l=1$ and $L/l=1$, which corresponds to the
oscillating barrier centered inside the box, and an oscillating amplitude
of $r=0.5$.
In Fig. 2 we show the energy-phase space $(e_n,\phi _n)$ for $M=4.7$,
using the map given by eqs. (9-11). We plot several orbits that
correspond to different initial conditions. We can clearly see that,
for this system,
we have a phase space with a mixed structure, in which we have
periodic, quasiperiodic and chaotic orbits. Some of the fixed points of
the map can be seen surrounded by elliptic orbits. We notice a fine
structure of smaller islands in the chaotic region, as is usually the
case for other maps \cite{Lich}.
The quantity that we want to analyze in detail is the traversal time in
the barrier region, that is, the time it takes the particle to cross
the region where the barrier is oscillating. We can obtain this quantity
simply as $b/v_n^{\prime }$ or
$b/|v_n^{\prime \prime \prime }|$ (see Fig. 1b).
The structure of this traversal or dwell time depends strongly on the
type of orbit. Clearly, if we have a
periodic orbit, then this time will take only two possible values,
since $v_n^{\prime }$ and $v_n^{\prime \prime \prime }$ does not change
with $n$. On the other hand, if the orbit is quasiperiodic, the velocity
can vary in a full range of values. In this case, the traversal
time can vary only in a limited range. However, when we have a chaotic
orbit, the variation can display a very rich structure \cite{Mateos}.
For the bounded problem, where the oscillating barrier is confined within
a box, we can obtain a chaotic dynamics as shown in Fig. 2. However, if
we remove the walls and leave only the oscillating barrier, we end up with
an open system of the scattering type. In this case, we cannot have
chaotic dynamics, since the particle interacts with the barrier only once.
However, we can study not a single particle, but an ensemble of
noninteracting particles, each of them with different initial conditions.
In Fig. 3 we show a space-time diagram of trayectories for an ensemble of
incident particles. In this case, and for the rest of the figures, we
take $r=0.5$ and $M=77.7$. I use dimensionless distance $x/l$ and
dimensionless time $t$, which is the time scaled with $l/w$. Since
$l=b$, then $l/w$ is the time it takes to cross the barrier region
with a velocity $w=(2V_0/m)^{1/2}$. The barrier is located between
$x/l=1$ and $x/l=2$, and is indicated by horizontal dashed lines in
Fig. 3. We take an ensemble of initial conditions in which the initial
velocity is constant and the initial phase
is uniformily distributed. We see from Fig. 3 that only a subset of
particles in the ensemble can cross the barrier region and that the
traversal time is different for each particle. This is due to the fact that
each particle is influenced differently by the time-modulated barrier,
depending on the arrival time. That is, different arrival times mean
different barrier amplitudes.
The traversal time is defined as the time it takes to cross the region
where the barrier is oscillating, and is given by $b/v_n^{\prime }$.
Since we scale this traversal time with the time $b/w$, the
dimensionless form is given by $1/\sqrt{e_n^{\prime }}$.
For the particles in the ensemble, this time is shown in Fig. 4.
We notice that in many cases the dimensionless
time $t\sim 1$; however, there are some others cases for which $t\gg 1$.
These large peaks occur when the arrival time is such that the
total energy is just above the barrier heigth, and thus the velocity
inside the barrier region is very small and consequently the traversal
time is very large. We can see a strong variation in the traversal
time, that leads to a broad distribution of times. On the other hand,
since the minimum velocity in the barrier region is zero, then there
is no upper bound for the dwell time, and it can acquire very large
values, as seen in Fig. 4.
The traversal time distribution is depicted in Fig. 5. This normalized
distribution has a long-time tail which is a power law.
In Fig. 6 we show the same distribution in a log-log plot that clearly
shows that this is indeed a distribution with a power-law tail
of the form $p(t)\sim t^{-\alpha}$, with $\alpha \simeq 3$. The
straight (dashed) line in this figure has a slope of $-3$.
Another quantity of interest is the transmission coefficient, defined as
the number of particles that cross the barrier region, divided by the
total number of particles in the ensemble. In Fig. 7 we show this
transmission coefficient as a function of $M$. Remember that
$M=l/(w\tau )$ and is, therefore, proportional to the frequency of
oscillation of the barrier. We can see in this figure that the
transmission coefficient vary strongly with $M$, in particular for low
frequencies ($M\sim 1$). On the other hand, for higher frequencies
($M\gg 1$), the transmission coefficient tend towards a constant value.
This last result indicates that for $M\gg 1$, the oscillating potential
barrier acts as an effective potential barrier of average height $V_0$.
Finally, in Fig. 8 we show the average traversal time as a function of
$M$. Again we can see strong fluctuations of this quantity. Since the
distribution of traversal times is a power law with an exponent
$\alpha \simeq 3$, we can expect these large fluctuations; although the
first moment of the distribution is finite in this case, the second or
higher moments can diverge, leading to these large fluctuations, as is
usually the case for L\'evy distributions \cite{Levy}.
\section{Concluding remarks}
In this paper, the dynamics of the classical problem
of an oscillating rectangular potential
barrier is analyzed. When the oscillating barrier is located
within a one-dimensional box, we have a bounded problem and the
corresponding classical dynamics can have a mixed phase space
structure comprising periodic, quasiperiodic and chaotic orbits.
For the scattering problem of a single oscillating barrier,
a distribution of traversal times with a power-law tail is obtained.
This L\'evy-type distribution of times leads to large
fluctuations of the average traversal time as a function of the
frequency of oscillation of the barrier; therefore, it is
difficult to obtain a characteristic time to the process
of crossing the classical oscillating barrier. These large fluctuations
arise due to the sensitive dependence on initial conditions,
typical of the dynamics of chaotic systems. In particular,
for our problem, the quantity that controls the traversal time
is the time of arrival at the barrier. Thus, we obtain a sensitive
dependence on the time of arrival for the classical case.
The possible role for the tunneling time problem, if any,
of the sensitive dependence on the time of
arrival and the difficulty to obtain a characteristic
traversal time in the classical domain, remains to be seen.
\vfill\eject
|
\section{Introduction}
Most galaxies in the nearby universe, including our own Galaxy, belong
to poor groups of galaxies. Despite the ubiquity of the group
environment, we know little about the internal dynamics of groups and
the evolution of group galaxies outside of the Local Group. To learn
whether the diverse and detailed results on the Local Group presented
at this meeting apply elsewhere, we must also survey other poor
groups.
Because poor groups are apparent systems of fewer than five bright
($\leq M^*$) galaxies, past studies have been hampered by small number
statistics. Some of the critical, unanswered questions are (1) whether
most poor groups are real systems instead of chance projections of
galaxies along the line-of-sight, (2) why many poor groups, with their
favorable environments for galaxy-galaxy mergers, have survived until
now, (3) and how galaxies evolve in groups, where the collisional
effects of the intragroup gas and the tidal influences of the global
potential are weaker than in rich clusters. The advent of
multi-object spectroscopy now makes it possible to address such
questions in unprecedented detail. In this talk, I discuss the first
results obtained in collaboration with John Mulchaey (Carnegie
Observatories) from a fiber spectroscopic survey of 12 nearby, poor
groups (Zabludoff \& Mulchaey 1998a; Mulchaey \& Zabludoff 1998;
Zabludoff \& Mulchaey 1998b).
\section{Poor Groups: What are They?}
\subsection{Three Classes of Groups}
Poor groups in the literature have been identified optically and fall
into several classes that can be distinguished by their X-ray
properties and bright galaxy morphologies. Groups with detectable, hot
intra-group gas typically have a giant ($\leq M^* - 1$) elliptical
that is the brightest group galaxy (BGG) and that lies near or on the
peak of the smooth, symmetric X-ray
emission (Mulchaey et al.\ 1996). In contrast,
groups without extended X-ray emission tend to optically resemble the
Local Group, which consists of a few bright, late-type galaxies and
their satellites. If some non-X-ray-detected, late-type-dominated
groups evolve into groups with a central, giant elliptical and a
detectable intra-group medium, then groups in transition may form a
third class of objects. We would expect the cores of such systems to
have signatures of recent dynamical evolution, including interacting
galaxies and X-ray gas that does not coincide with the galaxies. To
construct our sample, we select poor groups from these three classes
that also have complementary X-ray images.
The sample consists of 12 nearby ($1500 < cz < 8000$ km\,s$^{-1}$),
optically-selected groups from the literature for which there are
existing, sometimes serendipitous, pointed ROSAT/PSPC X-ray
observations: NGC\,533, NGC\,5129, NGC\,5946, NGC\,4325, NGC\,741, NGC\,2563,
HCG\,42, HCG\,62, HCG\,90, NGC 664, NCG 491, and NGC\,7582.
Nine of the groups are X-ray-detected. HCG\,90 (Hickson 1982), a
possible transitional object with several interacting galaxies in its
core (Longo et al.\ 1994), is only marginally X-ray-detected
(Ponman et al.\ 1996). It is important to stress that our group
sample is biased with respect to published group catalogs, in which
less than half of the groups are X-ray-detected. The sample
is weighted toward X-ray groups, because of the likelihood that they
are bound systems and the most evolved poor groups.
\subsection{Bound or Superpositions?}
The issue of whether many poor groups, even those identified from
redshift surveys, are bound systems instead of chance superpositions
of galaxies along the line-of-sight has been a puzzle. The existence
of one poor group, our Local Group, is unchallenged. In contrast,
Ramella et al.\ (1989) show that $\sim 30\%$ of groups of three or
four galaxies in the CfA Redshift Survey (Huchra et al.\ 1995) are
probably unbound, geometric projections. The detection of a
significant population of fainter members would be an important first
step in identifying poor groups that are real.
Using the Las Campanas fiber spectrograph designed by Steve Shectman,
we measured the $\sim 100$ brightest galaxies projected within $1.5
\times 1.5$ deg$^2$ of the center of each of the 12 sample groups.
The radial velocity distributions for galaxies in the fields of the
nine X-ray-detected groups each reveal $\sim 20$-50 group members down
to absolute magnitudes of $M_B \sim -14$ to $-16 + 5\log_{10}h$.
The surprisingly large membership, the central concentration of early
type galaxies, the similarity of the BGG's position and orientation to
those of the diffuse X-ray emission, the consistency of the optical
velocity dispersion and the X-ray temperature, and the short crossing
times ($\leq 0.05$ of a Hubble time) of the X-ray groups suggest that
they are bound systems, not geometric superpositions of galaxies, and
that the group cores are close to virialization or virialized.
Because we do not detect diffuse X-ray emission and find only 5-8
members in the three non-X-ray-detected groups, we are unable to
determine their dynamical state. The non-X-ray groups, which consist
of one or two $M^*$ or brighter spirals with several fainter galaxies
that may be satellites, are {\it morphologically} akin to the Local Group
(although our samples are not sufficiently deep to ascertain whether
any group has a dwarf spheroidal population like that of the Local
Group; van den Bergh 1992). If the non-X-ray groups are {\it dynamically}
similar to the Local Group, they are bound (see Zaritsky 1994). Our
current data do not exclude this possibility --- the velocity
dispersions of the non-X-ray groups are consistent with the upper
limits on their X-ray luminosities (Mulchaey \& Zabludoff 1998).
\section{Poor Groups: Why are They?}
If some poor groups are bound systems, then another critical question
is why they exist at all. Poor groups have higher galaxy densities
than the field and lower velocity dispersions than cluster cores,
making them favorable sites for galaxy-galaxy mergers (Barnes 1985).
The likelihood of mergers and the short group crossing times ($\leq
0.05$ of a Hubble time) suggest that most groups should have already
merged into one object. Therefore, either bound groups are collapsing
for the first time like the Local Group (Zaritsky 1994) or only a
small fraction of the group mass is tied to the galaxies, lowering the
rate of galaxy-galaxy interactions relative to a galaxy-dominated
system and allowing the group to survive many crossing times
(Governato et al.\ 1991; Bode et al.\ 1993; Athanassoula et al.\ 1997). To
address this issue by measuring the underlying mass distribution of
poor groups, we use the improved statistics of our spectroscopic
survey.
The velocity dispersion of the combined group sample does
not decrease significantly with radius from the central $\sim
0.1$$h^{-1}$ Mpc to at least $\sim 0.5$$h^{-1}$ Mpc, in contrast to
the more than factor of two decrease that would be observed if the
entire mass were concentrated within 0.1$h^{-1}$ Mpc. The extended
mass is either in the galaxies, in a common halo through which the
galaxies move, or in both the galaxies and a diffuse halo. If all the
mass were tied to the galaxies, most of the mass would be associated
with the bright, central elliptical in those groups in which the BGG
dominates the light. For groups with a few galaxies that have
luminosities comparable to the BGG, the velocity dispersion would be
increased at large radii by subgroups consisting of a massive galaxy
and the subgroup members that are orbiting it. If this picture were
accurate, we would expect the velocity dispersion profiles of groups
with several comparably bright galaxies to be shallower than those in
which the BGG is dominant. However, the combined velocity dispersion
profile of a subsample of two groups (HCG\,42 and NGC\,741) in which the
BGG dominates the light (i.e., the BGG luminosity exceeds the
combined luminosity of the other $M^*$ or brighter galaxies) is
indistinguishable from that of the entire sample. Therefore, we
conclude that most of the group mass lies in a smooth, extended dark
halo.
This result argues that poor groups survive longer than predicted by
models in which all the mass is tied to individual galaxies and may
explain why so many poor groups are observed in lieu of single merger
remnants.
\section{Poor Groups: How do Galaxies Evolve in Them?}
\subsection{Distribution of Early Type Fractions}
The factors that might affect the evolution of galaxies in poor groups
are different from those present in rich clusters.
If clusters evolve hierarchically by accreting
poor groups of galaxies (subclusters), members of an infalling group
have recently experienced the hot, dense cluster environment
for the first time.
Therefore, galaxies in poor groups in the field
are a control sample for understanding the factors that
influence the evolution of their counterparts in subclusters.
For example, we can compare the
morphologies and recent star formation histories of
galaxies in the subclusters of complex clusters like Coma
(Caldwell \& Rose 1997) with those of galaxies in
poor field groups. Differences between the samples would argue
that cluster environment is important in transforming galaxies
at the present epoch.
On the other hand, the lack of such differences
would suggest, as the simplest explanation,
either that star formation and morphology are influenced by
mechanisms present in both field groups and subclusters, such as galaxy-galaxy
encounters, or that the effects of environment on
galaxies are insignificant compared
with conditions at the time of galaxy formation.
Despite the usefulness of group galaxies
as a control sample, their properties, especially at faint magnitudes,
are not well-known.
Past work has included only the four
or five brightest galaxies, which biases samples toward ellipticals,
and has targeted only the central $\leq 0.3$$h^{-1}$ Mpc,
where early types concentrate. Therefore,
to compare the morphologies and star formation histories of group
and cluster members, we must sample both environments to similar
physical radii and absolute magnitude limits.
In the X-ray groups, the early type fraction ($f$)
ranges widely from $\sim 55\%$
(HCG\,62, NGC\,741, and NGC\,533)
to $\sim 25\%$ (e.g., NGC\,2563). The latter value
is characteristic of the field ($\sim 30\%$; Oemler 1992).
We find no early types among the 6-8 galaxies in each of
the three non-X-ray groups.
The early type fractions of $\sim 55\%$ in
NGC\,533, NGC\,741, and HCG\,62 are most surprising, because they are consistent with
those of rich clusters for similar
physical radii and absolute magnitude limits ($\sim 0.55$-0.65;
Whitmore et al.\ 1993).
\subsection{Early Type Fraction vs. Velocity Dispersion}
The correlation between early type fraction and group
velocity dispersion is significant at the $>0.999$ level.
The form of the relation cannot be the same for rich clusters ---
our fit to the group data predicts that the early type fraction
in a $\sigma_r \sim 1000$ km\,s$^{-1}$ cluster is an unphysical $f = 124\%$!
Therefore, the relation must turn up between the poor group and rich
cluster regimes. The group $f - \sigma_r$ relation implies either
that galaxy morphology is set by the local potential size at the
time of galaxy formation (Hickson, Huchra, \& Kindl 1988) and/or that
$\sigma_r$ and $f$ increase as a group
evolves (Diaferio et al.\ 1995).
The early type fractions in our highest velocity dispersion
groups ($\sigma_r \sim 400$ km\,s$^{-1}$) are as high as in rich clusters.
If some early type galaxies are evolved merger remnants,
then the galaxy populations of
higher velocity dispersion groups are more evolved on average.
At some point in the group's evolution, perhaps at a
velocity dispersion near 400 km\,s$^{-1}$, any morphological
evolution resulting from galaxy mergers ceases,
and the fraction of merger remnants in poor groups and rich clusters
is comparable. The implicit upturn in our $f - \sigma_r$ relation
suggests such a saturation point.
Alternatively, the similarity of some group and cluster
early type fractions, and the steepening of
the $f - \sigma_r$ relationship at high $\sigma_r$,
might arise from conditions at the time of galaxy formation. For example,
it is possible that poor groups such as NGC\,533 and rich clusters like
Coma begin as similar mass density perturbations
with correspondingly similar
galaxy populations. In this simple picture, NGC\,533 does not develop a
cluster-size potential, because its field lacks the surrounding
groups that Coma later accretes.
In summary, the cluster-like fraction of early type galaxies
in NGC\,533, NGC\,741, and HCG\,62 indicate that clusters are not the only
environments with copious quantities of E and S0 galaxies.
The simplest explanation is either that fluctuations in
the initial conditions permitted early types to form
in these groups' comparatively low velocity dispersion, low galaxy density
environments or that the galaxies' subsequent evolution
was the product of a mechanism,
such as galaxy-galaxy interactions,
common to both groups in the field and groups that become subclusters.
Although additional environmental
mechanisms may affect the evolution of cluster galaxies, such
cluster-specific processes are not required to explain the current data.
A cluster that evolves hierarchically from subclusters with the properties
of NGC\,533, NGC\,741, and HCG\,62 will have, at least initially, a similar
galaxy population.
\subsection{Star Formation in Early Types}
The star formation histories of galaxies in poor groups provide
additional insight into the environmental factors that may influence
the evolution of galaxies. One approach is to examine the spectra of
the early types for evidence of on-going star formation or of a young
stellar population. We can then compare the fraction of E and S0
group members that have recently formed stars with a sample from rich
clusters with complex structure.
Star formation is on-going or has ended within the last $\sim 2$ Gyr
in eight of the 64 early type group members for which we have
spectra ($12\%$). This fraction is roughly the same for clusters, such as
Coma, with infalling groups ($\sim 15\%$; Caldwell \& Rose 1997).
This result, and the similarity of the early type fractions of some
poor groups to rich clusters, suggests that an environmental mechanism
present in both groups and subclusters may be responsible. For
example, galaxy-galaxy encounters can produce bursts of star formation
(e.g., Londsdale et al.\ 1984; Kennicutt et al.\ 1987;
Sanders {et al.\ 1988) in which the gas is consumed or stripped
away.
Although our observations of poor groups suggest that the effects of
cluster environment are not necessary to produce the early type
fractions and star formation episodes of nearby clusters, we suspect
that the star formation histories of group and subcluster galaxies
will begin to deviate after the subcluster and cluster mix. Proposed
gas removal processes including ram pressure stripping and the tidal
limitation of galaxy halos, which are more
efficient in clusters than in groups, may eventually suppress star
formation in cluster galaxies. Comparative studies of the H{\sc i} content
of field, group, and cluster galaxies will help to resolve this issue.
\acknowledgments
This research was supported by grants from the NSF and NASA.
|
\section{Introduction}
Hickson Compact Groups (hereafter HCGs; Hickson \cite{Hic};
Hickson \cite{Hik}) are small systems of several galaxies (four or more) in an apparent close proximity in the sky.
The debate on their physical reality as bounded systems is still open. A possibility exists that only a part of the sample of HCGs are bound systems and/or that HCG dynamical evolution depends on their environments. Important informations about their real nature could be obtained by studying the rate of merger and interaction between their galaxies. The studies carried out so far agree with the view of a low merging rate inside HCGs with respect to undisrupted systems of galaxies (Zepf et al. \cite{Zep:Whi}, Zepf \cite{Zep}). On the other hand it is not so clear which is the fraction of interacting galaxies in HCGs: photometric and spectroscopic studies (Rubin et al. \cite{Rub}; Mendes de Oliveira et al. \cite{Men:Ama}; Moles et al. \cite{Mol:del}; Mendes de Oliveira et al. \cite{Men:Hic}; Vilchez \& Iglesias Paramo \cite{Vil:Para}; Vilchez \& Iglesias Paramo \cite{Vil:Parb}; Iglesias Paramo \& Vilchez \cite{Par:Vil}) have often given contradictory results.
It is expected that interaction and merger phenomena strongly affect the star formation rate ({\em SFR}) of galaxies.
In particular, interacting galaxies should show an higher star formation rate than field galaxies.
Thus the study of star formation of galaxies in HCGs gives important clues about the interaction and merger phenomena inside them.
Powerful tools to investigate on the star formation activity are the ionization lines emitted by the heated gas surrounding the regions of star formation.
The H$_\alpha$ emission line at 6563 {\AA} can be used as a quantitative and spatial tracer of the rate of massive ($\geq$ 10 M$_\odot$) and therefore recent ($\leq 10^7$ years) star formation (Kennicutt \cite{ken}; Ryder \& Dopita \cite{Ryd:Dop}), unlike the color indexes in the $U,B,V$ filters, that give indications about the past star formation ($> 10^8$ years).
Therefore, by knowing the H$_\alpha$ emission of the HCG galaxies it is possible in principle to carry out important investigations about the present merger and interaction events in these systems.
Up to now, only Rubin et al. (\cite{Rub}) and more recently Vilchez \& Iglesias Paramo (\cite{Vil:Para}) have collected significant samples of H$_\alpha$ data on HCG galaxies.
They published H$_\alpha$ emission-line images respectively for 14 and 16 HCGs. While Vilchez \& Iglesias Paramo (\cite{Vil:Para}) estimate the H$_\alpha$ flux for each of the 63 galaxies of their sample, Rubin et al. do not use flux calibrated and they take into account a sample constituted by disk galaxies only.
H$_\alpha$ data for the galaxies of single groups have been also obtained by Valluri \& Anupama (\cite{Val:Anu}), Mendes de Oliveira et al. (\cite{Men:Pla}) and Plana et al. (\cite{Pla:Men}).
Valluri \& Anupama presented H$_\alpha$ calibrated data for the galaxies of HCG62 and Mendes de Oliveira et al. and Plana et al. reported kinematic observations of H$_\alpha$ emission respectively for four late-type galaxies of HCG16 and for two early-type galaxies and one disk system of HCG90.
With the aim to obtain quantitative informations about the H$_\alpha$ emission of HCGs galaxies we have observed 31 HCG in narrow-band interferometric filters deriving H$_\alpha$ calibrated fluxes for 95 galaxies, 22 out of which are upper limits. In this paper we present the catalogue containing these H$_\alpha$ data.\\
We first describe the sample and the observations in $\S$ 2 and in $\S$ 3. In $\S$ 4 and $\S$ 5 we present the data reduction and calibration procedures used. The Zero Point correction, Galactic and Internal extinction corrections applied to the fluxes are described in $\S$ 6, while $\S$ 7 contains the photometric error derivation.
In $\S$ 8 we present the H$_\alpha$ Catalogue of Galaxies, while in $\S$ 9 we derive the star formation rate for the whole sample. Finally we briefly discuss some of the observed groups in $\S$ 10.
\section{The Sample}
The 100 compact groups catalogue by Hickson (\cite{Hic}) has been revised by Hickson, Kindl \& Auman (\cite{Hic:Kin}) and then by Hickson et al. (\cite{Hic:Men}). By adding a radial velocity criterion Hickson et al. (\cite{Hic:Men}) were able to reject probable non member galaxies. The resulting sample consists of 92 groups each containing three or more "accordant" members, which have radial velocities differing by no more than 1000 km~s$^{-1}$ from the median velocity of the group. Our sample has been drawn from this latest catalog.\\
In this paper, the result of the data reduction and calibration of H$_\alpha$ CCD images are presented for 31 Hickson Compact Groups. The remaining 61 HCGs were not in our sample because the adequate H$_\alpha$ interferometric filters were not available during the observations. This is the only criterion used to select the observed groups.\\
The redshift of observed groups is in the range 0.005$ \leq z \leq$ 0.07 (P. Hickson et al., \cite{Hic:Men}) and their distribution is shown in Figure 1 (the width of each bin is 0.01).
\begin{figure}[h]
\centerline{\psfig{figure=redshift.ps,height=90mm,bbllx=70mm,bblly=60mm,bburx=140mm,bbury=250mm}}
\bigskip
\centerline{{\bf Figure 1:} Redshift distribution of the 31 observed groups.}
\end{figure}
\vskip 0.2truecm
Table 1 lists the observed HCGs as follows:\\
Col.1: Name of the groups according to Hickson's catalogue;\\
Col.2: 1950 right ascension (R.A.) of the centroid of the member galaxies;\\
Col.3: 1950 declination (Dec.) of the centroid of the member galaxies;\\
Col.4: Number of accordant members of the group;\\
Col.5: Velocity dispersion of the group: $\sigma_v$ $(Km~s^{-1})$;\\
Col.6: Mean Redshift of the group.\\
The sample is composed by 134 galaxies, 127 out of which have been observed. 52$\%$ of them are Ellipticals and Lenticulars and the remaining 48$\%$ are Spirals and Irregulars. For each observed galaxy we report in Table 2 the heliocentric radial velocity {\em V} in units of $Km~s^{-1}$, the total magnitude in the photografic band $B_T$, corrected for internal and galactic extinction and the Hubble type, as in Hickson \cite{Hik}. Galaxies are named with the number of HCG plus letter of galaxy itself.\\
The distribution of the total $B_T$ magnitude of the observed galaxies is shown in Figure 2 (the width of each bin is 0.5 magnitude).
\begin{figure}[h]
\centerline{\psfig{figure=magnitudine.ps,height=90mm,bbllx=70mm,bblly=60mm,bburx=140mm,bbury=250mm}}
\bigskip
\centerline{{\bf Figure 2:} Distribution of $B_T$ magnitude of the 127} \centerline{observed galaxies.}
\end{figure}
\begin{table}[h]
{\bf Table 1:} Observed Hickson Compact Groups (Hickson et al., \cite{Hic:Men}): HCG number, right ascension and declination (1950), number of accordant galaxies, velocity dispersion and mean redshift of group.
\begin{flushleft}
\begin{tabular}{cccccc}
\hline
\hline
\noalign{\smallskip}
HCG & R.A. & Dec. & $N^{\circ}$ & $\sigma_v$ & z \\
\ & {\em (1950)} & {\em (1950)} & \ & $(\frac{Km}{s})$ & \ \\
\noalign{\smallskip}
\hline
2 & 0 28 48.97 & 8 10 19.3 & 3 & 54.9 & 0.0144 \\
15 & 2 5 2.95 & 1 54 58 & 6 & 426.6 & 0.0228 \\
33 & 5 7 53.69 & 17 57 51 & 4 & 154.9 & 0.026 \\
34 & 5 19 6.72 & 6 38 5.7 & 4 & 316.2 & 0.0307 \\
35 & 8 41 56.87 & 44 42 16.4 & 6 & 316.2 & 0.0542 \\
37 & 9 10 35.78 & 30 12 58 & 5 & 398.1 & 0.0223 \\
38 & 9 24 58.06 & 12 29 58.4 & 3 & 12.9 & 0.0292 \\
43 & 10 8 39.7 & 0 11 32.7 & 5 & 223.9 & 0.033 \\
45 & 10 15 46.72 & 59 21 27.8 & 3 & 182.0 & 0.0732 \\
46 & 10 19 29.69 & 18 6 39.5 & 4 & 323.6 & 0.027 \\
47 & 10 23 7.57 & 13 59 28.2 & 4 & 42.6 & 0.0317 \\
49 & 10 53 19.24 & 67 26 54.2 & 4 & 33.9 & 0.0332 \\
53 & 11 26 18.96 & 21 2 13.9 & 3 & 81.3 & 0.0206\\
54 & 11 26 38.24 & 20 51 38.4 & 4 & 112.2 & 0.0049 \\
56 & 11 29 53.51 & 53 13 16.5 & 5 & 169.8 & 0.027 \\
59 & 11 45 53.12 & 12 59 51.3 & 4 & 190.5 & 0.0135 \\
66 & 13 36 47.14 & 57 33 45.5 & 4 & 302 & 0.0699 \\
68 & 13 51 29.15 & 40 33 26.9 & 5 & 154.9 & 0.008 \\
69 & 13 53 12.58 & 25 18 44 & 4 & 223.9 & 0.0294 \\
70 & 14 1 54.07 & 33 34 13.7 & 4 & 144.5 & 0.0636 \\
71 & 14 8 45.02 & 25 44 4.1 & 3 & 416.8 & 0.0301 \\
72 & 14 45 36.94 & 19 16 2.6 & 4 & 263 & 0.0421 \\
74 & 15 17 12.89 & 21 4 31.9 & 5 & 316.2 & 0.0399 \\
75 & 15 19 19.7 & 21 21 45.3 & 6 & 295.1 & 0.0416 \\
76 & 15 29 14.96 & 7 29 20.1 & 7 & 245.5 & 0.034 \\
79 & 15 56 59.93 & 20 53 51 & 4 & 130.0 & 0.0145 \\
81 & 16 15 54.25 & 12 54 57.6 & 4 & 177.8 & 0.0499 \\
82 & 16 26 28.03 & 32 56 21 & 4 & 616.6 & 0.0362 \\
83 & 16 33 12.91 & 6 22 9.6 & 5 & 457.1 & 0.0531 \\
92 & 22 33 40.37 & 33 42 12.6 & 4 & 389.0 & 0.0215 \\
96 & 23 25 28.19 & 8 29 55.4 & 4 & 131.8 & 0.0292 \\
\noalign{\smallskip}
\noalign{\smallskip}
\hline
\hline
\end{tabular}
\end{flushleft}
\end{table}
\begin{table*}[h]
{\bf Table 2:} Principal features of the observed galaxies (Hickson \cite{Hik}): Galaxy name, heliocentric velocity of group, total photographic blue magnitude, morphological type of galaxy.
\small
\begin{flushleft}
\begin{tabular}{cccc|cccc|cccc}
\hline
\hline
\noalign{\smallskip}
Galaxy & {\em V} & $B_{T}$ & {\em T} & Galaxy & {\em V} & $B_{T}$ & {\em T} & Galaxy & {\em V} & $B_{T}$\ & {\em T} \\
\ & $(Km~s^{-1})$ & \ & \ & \ & $(Km~s^{-1})$ & \ & \ & \ & $(Km~s^{-1})$ & \ & \ \\
\noalign{\smallskip}
\hline
\noalign{\smallskip}
2a & 4326 & 13.35 & SBd & 56a & 8245 & 15.24 & Sc & 82a & 11177 & 14.14 & E3 \\
2b & 4366 & 14.39 & cI & 56b & 7919 & 14.5 & SB0 & 82b & 10447 & 14.62 & SBa \\
2c & 4235 & 14.15 & SBc & 56c & 8110 & 15.37 & S0 & 82c & 10095 & 14.78 & Im \\
15c & 7222 & 14.37 & E0 & 56d & 8346 & 16.52 & S0 & 82d & 11685 & 15.95 & S0a \\
15d & 6244 & 14.65 & E2 & 56e & 7924 & 16.23 & S0 & 83a & 15560 & 15.99 & E0 \\
15f & 6242 & 15.74 & Sbc & 59a & 4109 & 14.52 & Sa & 83b & 16442 & 16.04 & E2 \\
33a & 7570 & 15.35 & E1 & 59b & 3908 & 15.2 & E0 & 83c & 16520 & 16.7 & Scd \\
33b & 8006 & 15.41 & E4 & 59c & 4347 & 14.4 & Sc & 83d & 15500 & 17.91 & Sd \\
33c & 7823 & 16.4 & Sd & 59d & 3866 & 15.8 & Im & 83e & 15560 & 18.4 & S0 \\
33d & 7767 & 16.73 & E0 & 66a & 20688 & 15.38 & E1 & 92b & 5774 & 13.18 & Sbc \\
34a & 8997 & 14.2 & E2 & 66b & 21472 & 16.5 & S0 & 92c & 6764 & 13.33 & SBa \\
34b & 9620 & 16.56 & Sd & 66c & 20801 & 16.39 & S0 & 92d & 6630 & 13.63 & SB0 \\
34c & 9392 & 16.28 & SBd & 66d & 20850 & 17.45 & E2 & 92e & 6599 & 14.01 & Sa \\
34d & 8817 & 17.57 & S0 & 68a & 2162 & 11.84 & S0 & 96a & 8698 & 13.53 & Sc \\
35a & 15919 & 15.56 & S0 & 68b & 2635 & 12.24 & E2 & 96b & 8616 & 14.49 & E2 \\
35b & 16338 & 15.13 & E1 & 69a & 8856 & 14.94 & Sc & 96c & 8753 & 15.69 & Sa \\
35c & 16357 & 15.69 & E1 & 69b & 8707 & 15.59 & SBb & 96d & 8975 & 16.56 & Im \\
35d & 15798 & 16.81 & Sb & 69c & 8546 & 14.94 & S0 & \ & \ & \ & \ \\
35e & 16773 & 17.05 & S0 & 69d & 9149 & 16.06 & SB0 & \ & \ & \ & \ \\
35f & 16330 & 18.12 & E1 & 70d & 18846 & 15.42 & Sc & \ & \ & \ & \ \\
37a & 6745 & 12.97 & E7 & 70e & 19117 & 15.91 & Sbc & \ & \ & \ & \ \\
37b & 6741 & 14.5 & Sbc & 70f & 19243 & 16.4 & SBb & \ & \ & \ & \ \\
37c & 7357 & 15.57 & S0a & 70g & 19010 & 16.39 & Sa & \ & \ & \ & \ \\
37d & 6207 & 15.87 & Sbdm & 71a & 9320 & 13.75 & SBc & \ & \ & \ & \ \\
37e & 6363 & 16.21 & E0 & 71b & 9335 & 14.9 & Sb & \ & \ & \ & \ \\
38a & 8760 & 15.25 & Sbc & 71c & 8450 & 15.56 & SBc & \ & \ & \ & \ \\
38b & 8739 & 14.76 & SBd & 72a & 12506 & 13.86 & Sa & \ & \ & \ & \ \\
38c & 8770 & 15.39 & Im & 72b & 12356 & 15.48 & S0 & \ & \ & \ & \ \\
43a & 10163 & 15.13 & Sb & 72c & 13062 & 15.47 & E2 & \ & \ & \ & \ \\
43b & 10087 & 15.18 & SBcd & 72d & 12558 & 15.64 & SB0 & \ & \ & \ & \ \\
43c & 9916 & 15.82 & SB0 & 74a & 12255 & 14.06 & E1 & \ & \ & \ & \ \\
43d & 9630 & 16.82 & Sc & 74b & 12110 & 15.07 & E3 & \ & \ & \ & \ \\
43e & 9636 & 17.2 & S0 & 74c & 12266 & 16.1 & S0 & \ & \ & \ & \ \\
45a & 21811 & 15.2 & Sa & 74d & 11681 & 16.32 & E2 & \ & \ & \ & \ \\
45b & 22195 & 17.24 & S0a & 74e & 11489 & 17.8 & S0 & \ & \ & \ & \ \\
45c & 21799 & 17.6 & Sc & 75a & 12538 & 15.2 & E4 & \ & \ & \ & \ \\
46a & 8201 & 16.4 & E3 & 75b & 12228 & 14.9 & Sb & \ & \ & \ & \ \\
46b & 8571 & 16.28 & S0 & 75c & 12292 & 15.93 & S0 & \ & \ & \ & \ \\
46c & 7906 & 16.13 & E1 & 75d & 12334 & 15.82 & Sd & \ & \ & \ & \ \\
46d & 7703 & 16.11 & SB0 & 75e & 12300 & 16.36 & Sa & \ & \ & \ & \ \\
47a & 9581 & 14.61 & SBb & 75f & 13080 & 16.66 & S0 & \ & \ & \ & \ \\
47b & 9487 & 15.67 & E3 & 76a & 10054 & 15.08 & Sa & \ & \ & \ & \ \\
47c & 9529 & 16.63 & Sc & 76b & 10002 & 14.44 & E2 & \ & \ & \ & \ \\
47d & 9471 & 16.2 & Sd & 76c & 10663 & 14.73 & E0 & \ & \ & \ & \ \\
49a & 9939 & 15.87 & Scd & 76d & 10150 & 15.21 & E1 & \ & \ & \ & \ \\
49b & 9930 & 16.3 & Sd & 76e & 10328 & 16.65 & SB0 & \ & \ & \ & \ \\
49c & 9926 & 17.18 & Im & 76f & 10216 & 16.48 & Sc & \ & \ & \ & \ \\
49d & 10010 & 16.99 & E5 & 79a & 4294 & 14.35 & E0 & \ & \ & \ & \ \\
53a & 6261 & 12.91 & SBbc & 79b & 4446 & 13.78 & S0 & \ & \ & \ & \ \\
53b & 6166 & 14.73 & S0 & 79c & 4146 & 14.72 & S0 & \ & \ & \ & \ \\
53c & 6060 & 14.81 & SBs & 79d & 4503 & 15.87 & Sdm & \ & \ & \ & \ \\
54a & 1397 & 13.86 & Sdm & 81a & 14676 & 16.25 & Sc & \ & \ & \ & \ \\
54b & 1412 & 16.08 & Im & 81b & 15150 & 16.51 & S0 & \ & \ & \ & \ \\
54c & 1420 & 16.8 & Im & 81c & 15050 & 17.18 & S0 & \ & \ & \ & \ \\
54d & 1670 & 18.02 & Im & 81d & 14954 & 17.14 & S0a & \ & \ & \ & \ \\
\noalign{\smallskip}
\hline
\hline
\end{tabular}
\end{flushleft}
\end{table*}
\normalsize
\section{Observations}
The CCD images of the HCG sample were obtained during three different observing runs (November 1995, April 1996 and February 1997).
Observations have been carried out at the 2.1 meter telescope (design Ritchey-Chretien) at the National Observatory of Mexico in S. Pedro Martir (SPM). The SPM Cassegrain focus (f/7.5) was coupled with a Tektronix CCD of 1024x1024 pixels, each 24$\mu$m x 24$\mu$m. The telescope scale (13 arcsec/mm) and the pixel dimensions provide a pixel size of 0.3 arcsec/pix with a resulting field of view of 5.12$^\prime\times$5.12$^\prime$. The CCD gain is 4 e$^-$/ADU.\\
During these three runs we observed 31 HCGs.
All images were obtained with seeing conditions in the range 2-2.6 arcsec.
For each HCG two CCD images were taken: the {\em on image}, by using a narrow-band interference filter ({\em H$_\alpha^{on}$ filter}) centered on the wavelength of the H$_\alpha$ line redshifted to the {\em z} of the galaxy
(which isolates the H$_\alpha$ emission-line and underlying
continuum), and the {\em off image}, by using another
interference filter ({\em H$_\alpha^{off}$ filter}) of similar bandwidth but centered on an adjacent region of the spectrum (isolating continuum light only).
Table 3 describes the features of narrow band filters used in this work. In the third column the range of recession velocity that a galaxy should have to give out its H$_\alpha$ line through the interferential filter is shown.\\
\begin{table}[h]
{\bf Table 3:} Features of interferometric filters: central wavelength, FWHM and corresponding velocity interval
\begin{center}
\begin{tabular}{ccr}
\hline
\hline
\noalign{\smallskip}
$\lambda_{central}$ & FWHM & Velocity Interval\\
{\em (\AA)} & {\em (\AA)} & $(Km~s^{-1})$ \\
\noalign{\smallskip}
\hline
\noalign{\smallskip}
6546 & 81 & $-2628 \rightarrow ~ 1074$\\
6564 & 72 & $-1600 \rightarrow ~ 1691$\\
6603 & 80 & $0 \rightarrow ~ 3657$\\
6607 & 89 & $-23 \rightarrow ~ 4045$\\
6641 & 79 & $1760 \rightarrow ~ 5371$\\
6643 & 80 & $1828 \rightarrow ~ 5485$\\
6683 & 80 & $3657 \rightarrow ~ 7314$\\
6690 & 91 & $3725 \rightarrow ~ 7885$\\
6723 & 80 & $5485 \rightarrow ~ 9142$\\
6732 & 74 & $6034 \rightarrow ~ 9416$\\
6742 & 85 & $6240 \rightarrow 10216$\\
6819 & 86 & $9736 \rightarrow 13668$\\
6920 & 88 & $1431 \rightarrow 18330$\\
7027 & 93 & $1908 \rightarrow 23335$\\
\noalign{\smallskip}
\hline
\hline
\end{tabular}
\end{center}
\end{table}
In order to calibrate our data, we have observed some spectrophotometric stars, equally spaced in time during each night, from the list of Massey $\&$ Strobel (\cite{Mas:Str}). Table 4 lists the standards used.
The spectrophotometric standards were observed in the same H$_\alpha$ narrow-band interference filters used to observe Hickson Compact Groups.
\begin{table}[h]
{\bf Table 4:} Standard stars used for calibration
\begin{center}
\begin{tabular}{ccc}
\hline
\hline
\noalign{\smallskip}
Star & $\alpha_{1950}$ & $\delta_{1950}$ \\
\ & ($h$ $m$ $s$) & ($\degr$ $\arcmin$ $\arcsec$) \\ \noalign{\smallskip}
\hline
\noalign{\smallskip}
PG0205+134 & 02 05 21.3 & +13 22 18 \\
Hiltner 600 & 06 42 37.2 & +02 11 25 \\
PG0939+262 & 09 39 58.8 & +26 14 42 \\
Feige34 & 10 36 41.2 & +43 21 50 \\
PG1121+145 & 11 21 39.4 & +14 30 18 \\
Feige66 & 12 34 54.7 & +25 20 31 \\
Feige67 & 12 39 18.9 & +17 47 24 \\
Kopf27 & 17 41 28 & +05 26 04 \\
\noalign{\smallskip}
\hline
\hline
\end{tabular}
\end{center}
\end{table}
The flux from [N$_{II}$] emission lines ($\lambda$=6548 {\AA} and $\lambda$=6584 {\AA}) is included in the on observations. Therefore the flux and luminosity here estimated refer to the sum of H$_\alpha$ and [N$_{II}$] emission lines and not only to H$_\alpha$. Nevertheless through this paper we refer for simplicity to them as f$_{H_\alpha}$ and L$_{H_\alpha}$ respectively.
The aim of the observations was to study the recent star formation rate occurring in HCG galaxies.
Since the $H_\alpha+[N_{II}$] emission a good star formation tracer as well as the H$_\alpha$ line alone (Kennicutt \& Ken, \cite{ken:Ken} ), the presence of [N$_{II}$] does not invalidate our data. Nevertheless, since the H$_\alpha$/[N$_{II}$] ratio is not constant with radius in the largest galaxies, we will refer to the global star formation rate of galaxies, that is to the rate
integrated over all the emitting area of each galaxy.
\vskip 0.2truecm
In Table 5 the journal of the observations is reported as follows:\\
Col.1: Name of the groups;\\
Col.2: Observing date (mm-yy);\\
Col.3: Central wavelength for the {\em H$_\alpha^{on}$ filter} used (\AA);\\
Col.4: Integration time for {\em H$_\alpha^{on}$ filter} exposure (s);\\
Col.5: Central wavelength for the {\em H$_\alpha^{off}$ filter} used (\AA);\\
Col.6: Integration time for {\em H$_\alpha^{off}$ filter} exposure (s).
\begin{table}[h]
{\bf Table 5:} Journal of observations
\begin{center}
\begin{tabular}{cccccc}
\hline
\hline
\noalign{\smallskip}
Group & Date & H$_\alpha^{on}$ & T$_{exp}$ & H$_\alpha^{off}$ & T$_{exp}$ \\
\ & {\em (mm-yy)} & {\em \AA} & {\em (s)} & {\em \AA} & {\em (s)} \\
\noalign{\smallskip}
\hline
\noalign{\smallskip}
HCG2 & Nov.95 & 6643 & 1800 & 6723 & 1800 \\
HCG15 & Nov.95 & 6723 & 1800 & 6643 & 1800 \\
HCG33 & Nov.95 & 6723 & 1800 & 6643 & 1800 \\
HCG34 & Feb.97 & 6732 & 1800 & 6564 & 1800 \\
HCG35 & Feb.97 & 6920 & 1800 & 6690 & 1800 \\
HCG37 & Nov.95 & 6723 & 1800 & 6643 & 1800 \\
HCG38 & Nov.95 & 6723 & 1800 & 6643 & 1800 \\
HCG43 & Feb.97 & 6819 & 1800 & 6607 & 1800 \\
HCG45 & Feb.97 & 7027 & 1800 & 6819 & 1800 \\
HCG46 & Nov.95 & 6723 & 1800 & 6643 & 1800 \\
HCG47 & Apr.96 & 6732 & 1800 & 6564 & 1800 \\
HCG49 & Apr.96 & 6819 & 1800 & 6690 & 1800 \\
HCG53 & Apr.96 & 6690 & 1800 & 6607 & 1800 \\
HCG54 & Nov.95 & 6603 & 1200 & 6683 & 1200 \\
HCG56 & Apr.96 & 6732 & 1800 & 6564 & 1800 \\
HCG59 & Apr.96 & 6690 & 1800 & 6607 & 1800 \\
HCG66 & Apr.96 & 7027 & 1800 & 6920 & 1800 \\
HCG68 & Feb.97 & 6607 & 1200 & 6819 & 1200 \\
HCG69 & Apr.96 & 6732 & 1800 & 6564 & 1800 \\
HCG70 & Apr.96 & 7027 & 1800 & 6920 & 1800 \\
HCG71 & Feb.97 & 6732 & 1800 & 6564 & 1800 \\
HCG72 & Apr.96 & 6819 & 1800 & 6690 & 1800 \\
HCG74 & Feb.97 & 6819 & 1800 & 6607 & 1800 \\
HCG75 & Apr.96 & 6819 & 1800 & 6690 & 1800 \\
HCG76 & Apr.96 & 6819 & 1800 & 6690 & 1800 \\
HCG79 & Apr.96 & 6690 & 1800 & 6607 & 1800 \\
HCG81 & Apr.96 & 6920 & 1800 & 6819 & 1800 \\
HCG82 & Apr.96 & 6819 & 1800 & 6920 & 1800 \\
HCG83 & Apr.96 & 6920 & 1800 & 6819 & 1800 \\
HCG92 & Nov.95 & 6723 & 1800 & 6643 & 1800 \\
HCG96 & Nov.95 & 6723 & 1800 & 6643 & 1800 \\
\noalign{\smallskip}
\hline
\hline
\end{tabular}
\end{center}
\end{table}
\section {Data Reduction}
\subsection{Bias and Flat-Field Corrections}
The science frames have been first bias subtracted.
For each observing run, we have obtained the proper bias by combining several bias frames with a median filter.
Then the images have been corrected for pixel to pixel response variations.
For each night, its own flat field has been constructed by medianing several flat field frames carried out during the night.\\
Two different types of flat fields have been used during the three
observing runs:
the first one has been obtained by medianing twilight sky images and it has been used to reduce the data of 1996.
The other one, used in 1995 and 1997 runs, has been constructed by medianing flat field frames taken on the dome illuminated with twilight sky. No significant differences have been measured by using the two flat fields.\\
These two steps of data reduction are based on the NOAO IRAF package, developed at the Center for Astrophysics.\\
Finally, cosmic rays and bad pixels have been removed from each frame using Munich Image Data Analysis System (MIDAS).\\
\subsection{ H$_\alpha$ Emission-Line Map}
The map of the H$_\alpha$ emission-line flux for each HCG ({\em H$_\alpha$ image}) has been obtained by removing the contribution of the underlying
continuum, that it is by subtracting the H$_\alpha^{off}$
from the H$_\alpha^{on}$ (Pogge, \cite{Pog}).
There are several reasons why the number of
continuum photons per integration time unit passing through the {\em on filter}
can be different from the one through the {\em off filter}.
For example differences between the transmission curves of the two
narrow-band filters, such as different width and/or
transmission peak; or variations of the sky transparency during the night.
This implies that in order to obtain the H$_\alpha$ emission-line flux
image a careful estimation of the underlying continuum to subtract
from the H$_\alpha^{on}$ is required.
In practice, the H$_\alpha^{off}$ have to be rescaled to the continuum of
the H$_\alpha^{on}$ wavelength.
Since stars do not show H$_\alpha$ emission, the number of continuum
photons coming from the stars in each HCG field and passing through the
$on$ and $off$ filters have to be the same.
In-fact, although the $on$ and $off$ filters are sometimes separated by more than 150\AA,
implying that the number of photons coming from the stars is not rigorously the same,
such a difference is negligible.
Thus for each HCG field (and for each couple of filters) we have selected at least three
stars and we have calculated the mean scaling factor $K$
\begin{equation}
K=\frac{1}{N}\sum_{i=1}^{N} \left(\frac{C_{on}}{C_{off}} \right)_i=\left<\left(\frac{C_{on}}{C_{off}} \right) \right>
\end{equation}
where $C_{on}$ and $C_{off}$ are the counts from stars in the $on$ and $off$
image respectively, and N is the number of stars.
Once rescaled, the H$_\alpha^{off}$ have been spatially aligned to the
H$_\alpha^{on}$.
The alignment has been performed by applying the IRAF tasks {\em geomap/geotran}
using the position of at least five stars in the field as
reference coordinates.
Finally, after having additively rescaled the $on$ and $off$ images
to the same median value, we have subtracted the H$_\alpha^{off}$
from the H$_\alpha^{on}$ thus obtaining the image of the H$\alpha$
emission-line flux.
\section{Photometric Calibration}
We have measured instrumental magnitudes of
standard stars by constructing for each star its growth curve
through circular concentric apertures.
The magnitude has been taken at the convergence of the curve.
From the spectral energy distributions of our observed standard stars
(Massey \& Strobel 1988), we have derived their apparent magnitudes at the
effective wavelengths $\lambda_{eff}$ of our filters, through the
usual relation
\begin{equation}
m_{vF}=-2.5 \cdot \log_{10}f_{\lambda_{eff}}(m_{vF})+2.5\log_{10}f_{\lambda_{eff}}(0)
\end{equation}
where F is a generic filter, $f_{\lambda_{eff}}(m_{vF})$ and $f_{\lambda_{eff}}(0)$
are the spectral irradiances in erg cm$^{-2}$ s$^{-1}$ \AA$^{-1}$ within the
$F$ filter having the effective wavelength $\lambda_{eff}(F)$
of a star of magnitude m$_{vF}$ and of a star of $m_{vF}$=0
respectively.
From each star we have derived the zero point $Z_p$ of the photometric
calibration for the different filters and nights.
The standard deviation of the zero point values thus obtained is within 0.05
mag during all but one night.
During this night the scatter is much larger than a factor of four.
Thus with the aim at maintaining the uncertainty on the galaxy photometry
within few hundreds percent, we have considered only those galaxies
observed during photometric nights, i.e. those nights for which $\sigma_{Z_p}\le0.05$ mag.
\section{The H$_\alpha$ Emission of Galaxies}
\subsection{Instrumental H$_\alpha$ Fluxes}
Following the data reduction steps described in section \S 4, we have
obtained 31 emission-line images, one for each HCG of our sample.
We have computed both isophotal and adaptive-aperture H$_\alpha$
fluxes for the HCG galaxies in the 31 fields by using SExtractor
(Bertin et al. \cite{Ber}).
The full analysis of each image is divided in six steps:
sky background estimation, thresholding, deblending, filtering of the detections, photometry and star/galaxy separation.
For each continuum-subtracted H$_\alpha$ image we have used a detection
threshold of one sigma above the background.
The H$_\alpha$ isophotal fluxes have been computed within the region
defined by the detection threshold.
In addition to the isophotal flux we have also considered the corrected
isophotal flux estimated by SExtractor that should take into account
the fraction of flux lost by the isophotal one (Bertin et al. \cite{Ber}).
In addition the adaptive-aperture photometry has also been calculated
(Kron {\cite{Kro:Kro}; Bertin et al. 1996).\\
Out of the 127 accordant observed galaxies belonging to the 31 HCG of our sample, we have been able
to compute isophotal and adaptive-aperture photometry
for 73 and 69 galaxies respectively.
The $1\sigma$ limiting flux, integrated within the mean seeing disk (2.3 arcsec), reached in our observations ranges between $1.43 \cdot 10^{-16}$ and $4.13 \cdot 10^{-17}$ $erg ~ cm^{-2} s^{-1}$.\\
For 22 galaxies, which have not been detected in our {\em H$_\alpha$} images,
we have computed the 3$\sigma$ upper limits above the background:
\begin{equation}
f_{ul}= 3 \cdot rms \cdot \left[\left(\frac{FWHM}{2} \right)^2 \cdot \pi \right]^{\frac{1}{2}}
\end{equation}
being
\[
\begin{array}{lp{0.6\linewidth}}
rms & the sky estimation accuracy $(counts~pix^{-1} s^{-1})$;\\
\left[\left(\frac{FWHM}{2} \right)^2 \cdot \pi \right]^{\frac{1}{2}} & the squareroot of the seeing area in pixels.\\
\end{array}
\]
For the remaining 32 galaxies we have not been able to estimate
the H$_\alpha$ fluxes because of one of the following reasons:
\begin{enumerate}
\item the night was not photometric ($\sigma_{Z_p}>>0.05$ mag);
\item the proper narrow band interference filter was not available;
\item too much imperfections are present on the {\em H$_\alpha$} image probably
due to large variations in seeing conditions between the {\em on} and
{\em off} band exposures, or due to changes in the telescope focus (e.g. because
of substantially different thickness of the filters and/or temperature
variations).
\end{enumerate}
In Table 6 we list the galaxies for which it was not possible
to measure their flux and the corresponding reason (1,2,3).
\subsection{Zero Point Correction}
In order to obtain calibrated fluxes and luminosities for our sample of galaxies,
we have estimated the zero point flux correction, Z$_{flux}$ such that
\begin{equation}
\label{eq:3.13}
f_{H_\alpha}=(f_{on}-f_{off})=Z_{flux} \cdot \left[C_{on}-{C}_{offn} \right]
\end{equation}
where:
\[
\begin{array}{lp{0.7\linewidth}}
f_{on}, f_{off} & are the isophotal or aperture H$_\alpha$ fluxes of galaxy in
the {\em on} and the scaled {\em off filters} respectively;\\
C_{on}, C_{off} & are the counts of the galaxy in the {\em on} and {\em off} band images respectively;\\
\end{array}
\]
It can be proved that the Z$_{flux}$ coefficient of each galaxy is proportional
to Z$_{flux_{on}}$ i.e. the zero point flux correction of the {\em on} band
image.
Knowing the Z$_{flux_{on}}$ in magnitudes ($Z_{p_{on}}$, see \S 5) and the
extinction
coefficient $k_{on}$ of the site relative to each filter, we have derived
the correction factor Z$_{flux_{on}}$ as follows:
\begin{equation}
Z_{flux_{on}}=\Delta\lambda \cdot 10^{-0.4(Z_{p_{on}}-(k_{on} \cdot X_{s})-b)}
\end{equation}
where $X_s$ is the airmass of the standard star and $b$ is
\begin{equation}
b=2.5 \cdot \log_{10}f_{\lambda_{eff}}(0)
\end{equation}
Thus $Z_{flux}$ is given by
\begin{equation}
Z_{flux}=\frac{Z_{flux_{on}}}{10^{[-0.4(k_{on} \cdot X_{s})]}} \cdot 10^{[0.4(k_{on} \cdot X_{g})]}
\end{equation}
where $X_g$ is the airmass of the target galaxy.
This zero point correction was applied to the H$_\alpha$ instrumental
fluxes and to the upper limits estimated for the undetected galaxies.
Fluxes and upper limits have been also corrected so that the H$_\alpha$
emission-line of the galaxy passes exactly in the center of the corresponding
{\em on} filter band, i.e. for the percentage of total flux lost if
the H$_\alpha$ emission line of the galaxy does not pass exactly in
the center of the corresponding {\em on} filter.
The corrected fluxes are reported in Tables 7, 8 and 9.
\subsection{Galactic and Internal Extinction Correction}
The H$_\alpha$ fluxes have been then corrected for the galactic
extinction due to the gas and the dust of our Galaxy.
For each target galaxy we have computed the relative galactic hydrogen
column density N$_h$ ({\em atoms cm$^{-2}$}) as a function of the galaxy
coordinates (R.A. and Dec.).
N$_h$ was obtained interpolating the data available from the
Stark et al. (\cite{Sta:Sta}) data-base.
We computed also an interpolation error defined as the mean of
differences weighted on the distances.
Using the relations:
\begin{equation}
\frac{N_{h}}{A_{B}-A_{V}}= \frac{N_{h}}{E(B-V)}=5.2 \cdot 10^{21}\ atoms\ cm^{-2}\ mag^{-1}
\end{equation}
and:
\begin{equation}
R=\frac{A_{V}}{E(B-V)}= 3.1
\end{equation}
(Ryder \& Dopita \cite{Ryd:Dop}) we have derived the visual extinction
coefficients A$_V$ and A$_B$ ({\em mag}) for each galaxy.
Following Ryder \& Dopita (\cite{Ryd:Dop}), we have obtained the
multiplicative correction $\alpha_G$ to apply to the H$_\alpha$ flux:
\begin{equation}
\alpha_G= 10^{(0.4 \cdot A_{H_{\alpha}})}= 10^{(0.4 \cdot 0.64A_{B})}
\end{equation}
where $A_{H_{\alpha}}$ is the H$_\alpha$ extinction coefficient in magnitudes.
The isophotal fluxes corrected for Galactic Extinction are reported in Table 7.
The minimum and maximum values obtained for $\alpha_G$ are respectively 1.04 and 2.68.\\
The H$_\alpha$ fluxes of spirals have also been corrected for the
Internal Extinction due to the interstellar medium inside the target galaxy
itself.
This correction in the blue band is usually obtained by summing to the galaxy
magnitude the value
\begin{equation}
A_i=c_B \cdot log (r_i)
\end{equation}
(Haynes \& Giovanelli \cite{Hay:Gio})
where:
\[
\begin{array}{lp{0.9\linewidth}}
c_B & is a morphological type dependent correction coefficient in
the $B$ band and \\
r_i & is the intrinsic axial ratio of the galaxy.\\
\end{array}
\]
On the basis of the interstellar extinction curve (e.g. Osterbrook 1974)
we have derived the H$_\alpha$ extinction correction term $c_{H_\alpha}$
using the following transformation:
\begin{equation}
c_B \cdot log(r_i) - c_{H_\alpha} \cdot log(r_i) = -2.5 \cdot log(e_B) + 2.5 \cdot log(e_{H_\alpha})
\end{equation}
where e$_B$ and e$_{H_\alpha}$ are the extinction values at the effective
wavelength respectively of the $B$ and the H$_\alpha$ filters .
Finally we have obtained the flux correction factor $\alpha_i=10^{[0.4 \cdot c_{H_\alpha} \cdot Log(r_i)]}$, to apply
to our spiral galaxies. The minimum and maximum values obtained for $\alpha_i$ are respectively 1.1 and 1.6.\\
On the basis of the fluxes thus obtained we have derived the H$_\alpha$
luminosity L$_{H_{\alpha}}$ of galaxies:
\begin{equation}
L_{H_{\alpha}}=4\pi \cdot f_{H_{\alpha}} \cdot d_{L}^2
\end{equation}
where the luminosity distance $d_L$ is defined as:
\begin{equation}
d_{L}=\frac{c}{H_0 \cdot q_{0}^2} \cdot \left(q_0 \cdot z+(q_0-1) \cdot [-1+(2q_0 \cdot z+1)^{\frac{1}{2}} \right)
\end{equation}
We adopted $H_0=100~km~s^{-1} Mpc^{-1}$ and $q_0=0.5$.
In Table 7 we report the isophotal luminosities of the galaxies ($L_{iso} (1)$) uncorrected for Galactic and Internal Extinction.
\section{Error Estimate}
The uncertainties reported in Tables 7, 8 and 9 regarding the different
flux estimates have been derived as follows:
\scriptsize
\begin{equation}
\sigma_{f_{H_{\alpha}}}= \sqrt{\left(\frac{\partial f_{H_{\alpha}}}{\partial C_{H\alpha}
} \cdot \sigma_{C_{H\alpha}} \right)^2 + \left(\frac{\partial f_{H_{\alpha}}}{\partial Z
_{flux}} \cdot \sigma_{Z_{flux}} \right)^2 + \left(\frac{\partial f_{H_{\alpha}}
}{\partial \alpha_G} \cdot \sigma_{\alpha_G} \right)^2
}
\end{equation}
\normalsize
where:
\[
\begin{array}{lp{0.6\linewidth}}
C_{H\alpha}=C_{on}-C_{off} & is the H$_\alpha$ instrumental
flux; \\
\sigma_{C_{H\alpha}} & is the uncertainty on $C_{H\alpha}$.
It is the standard deviation of the stellar flux residuals measured on the
net H$_\alpha$ images.
Since stars do not show H$_\alpha$ emission we should not
detect any flux at the star positions on the net images.
The detection of net counts could be thus represent simple poisson noise
and/or no perfect continuum subtraction. Therefore the standard deviation
of such measurements gives a good estimation of the pure and not
statistical uncertainties on $C_{H\alpha}$; \\
\sigma_{Z_{flux}} & represents the accuracy on the scale factor
$Z_{flux}$ (\S 6.2) and is given by the difference of the zero points derived by the
two standard stars observed before and after the target HCG;\\
\sigma_{\alpha_G} & is the error about the Galactic
Extinction $\alpha_G$ (\S 6.3) derived by the propagation error formula
to $\alpha_G$.\\
\end{array}
\]
The errors regarding the H$_\alpha$ luminosity (Tables 7 and 9) have been
calculated by applying the error propagation formula.
\section{The H$_\alpha$ Catalogue}
Table 7 lists the H$_\alpha$ isophotal fluxes and luminosities
estimated:
Col.1: Name of the galaxy (Hickson 1982);\\
Col.2: Flux uncorrected for Galactic and Internal Extinction: $f_{iso}$ (1) ($erg~cm^{-2}~s^{-1}$);\\
Col.3: Error about $f_{iso}$ (1): $\sigma_{f_{iso}}(1)$ ($erg~cm^{-2}~s^{-1}$);\\
Col.4: Luminosity uncorrected for Galactic and Internal Extinction: $L_{iso}$ (1) ($erg~s^{-1}$);\\
Col.5: Error about $L_{iso}$ (1): $\sigma_{L_{1}}$ (1) ($erg~s^{-1}$);\\
Col.6: Flux corrected for Galactic Extinction $f_{iso}$ (2) ($erg~cm^{-2}~s^{-1}$) ;\\
Col.7: Error about $f_{iso}$ (2): $\sigma_{f_{iso}}$ (2) ($erg~cm^{-2}~s^{-1}$);\\
Col.8: Flux corrected for Galactic and Internal Extinction for spiral galaxies: $f_{iso}$ (3) ($erg~s^{-1}$);\\
Col.9: Isophotal area at 1$\sigma$ above the background: $A_{iso}$ ($arcsec^2$);\\
Col.10: {\em S/N} ratio computed within the isophotal region defined by the detection threshold.
\vskip 0.5truecm
In Tables 8 and 9 the fluxes are not corrected for Galactic and Internal Extinction. Such corrections can be simply derived from Table 7.
\vskip 0.5truecm
Table 8 lists isophotal corrected and adaptive aperture fluxes:\\
Col.1: Name of the galaxy (Hickson 1982);\\
Col.2: Isophotal flux $f_{isocor}$ ($erg~cm^{-2}~s^{-1}$);\\
Col.3: Error about $f_{isocor}$: $\sigma_{isocor}$ ($erg~cm^{-2}~s^{-1}$);\\
Col.4: Kron flux $f_{Kron}$ ($erg~cm^{-2}~s{-1}$);\\
Col.5: Error about $f_{Kron}$: $\sigma_{Kron}$ ($erg~cm^{-2}~s{-1}$).\\
\vskip 0.5truecm
In Table 9 we report the fluxes and luminosities of upper limits:\\
Col.1: Name of the galaxy (Hickson 1982);\\
Col.2: Flux at 3 $\sigma$ above the background ($f_{3\sigma}$) ($erg~cm^{-2}~s^{-1}$);\\
Col.3: Error about $f_{3\sigma}$: $\sigma$ ($erg~cm^{-2}~s^{-1}$);\\
Col.4: Luminosity at 3 $\sigma$ above the background ($L_{3\sigma}$) ($erg~cm^{-2}~s^{-1}$);\\
Col.5: Error about $L_{3\sigma}$: $\sigma$ ($erg~cm^{-2}~s{-1}$).\\
\vskip 0.5truecm
We have compared the estimated $f_{iso}$, $f_{isocor}$ and $f_{kron}$ for the detected galaxies by using the Kolmogorov-Smirnov test: we found they are drawn from the same parent population. The distributions of the three different fluxes estimated for the detected galaxies are shown in Figure 3.
In the following we make use of $f_{iso}$ in our considerations unless it is differently specified.
\begin{figure}[h]
\centerline{\psfig{figure=flux.ps,height=80mm,bbllx=70mm,bblly=60mm,bburx=140mm,bbury=250mm}}
\bigskip
\bigskip
{{\bf Figure 3:} Distribution of $f_{iso}$ (solid line), $f_{isocor}$ (dashed line) and $f_{kron}$ (dotted line) (corrected for Galactic Extinction) of the 73 detected galaxies. The width of each bin is 0.5 [$10^{-13}$ erg cm$^{-2}$ s$^{-1}$].}
\end{figure}
\vskip 0.5truecm
In Figure 4 the distribution of $H_{\alpha}$ isophotal luminosity $L_{iso}$, corrected for Galactic Extinction and uncorrected for Internal Extinction, of the 73 detected galaxies (shaded histogram) is shown. We over-plot also the distribution of $H_{\alpha}$ upper limits to luminosity for the 22 undetected galaxies (dashed line).
\begin{figure}[h]
\centerline{\psfig{figure=luminosita.ps,height=80mm,bbllx=70mm,bblly=60mm,bburx=140mm,bbury=250mm}}
\bigskip
\bigskip
{{\bf Figure 4:} Distribution of $H_{\alpha}$ luminosity, corrected for Galactic Extinction, for both the 73 detected galaxies (shaded histogram) and for the 22 upper limits (dashed line). The width of each bin is 0.8 [$10^{40}$ erg s$^{-1}$].}
\end{figure}
\vskip 0.5truecm
\begin{table}[h]
{\bf Table 6:} Galaxies without estimated flux (see \S 6.1)
\begin{center}
\begin{tabular}{cccccc}
\hline
\hline
\noalign{\smallskip}
Galaxy & Reason (1,2,3) & \ & \ & Galaxy & Reason (1,2,3) \\
\noalign{\smallskip}
\hline
\noalign{\smallskip}
2c & 1 & \ & \ & 59b & 3 \\
15c & 1 & \ & \ & 59c & 3 \\
15d & 1 & \ & \ & 59d & 3 \\
15f & 1 & \ & \ & 70d & 2 \\
34b & 2 & \ & \ & 79a & 3 \\
43d & 2 & \ & \ & 79b & 3 \\
43e & 2 & \ & \ & 79c & 3 \\
47a & 2 & \ & \ & 79d & 3 \\
47b & 2 & \ & \ & 81a & 3 \\
47c & 2 & \ & \ & 81b & 3 \\
47d & 2 & \ & \ & 81c & 3 \\
54a & 3 & \ & \ & 81d & 3 \\
54b & 3 & \ & \ & 96a & 1 \\
54c & 3 & \ & \ & 96b & 1 \\
54d & 3 & \ & \ & 96c & 1 \\
59a & 3 & \ & \ & 96d & 1 \\
\noalign{\smallskip}
\noalign{\smallskip}
\hline
\hline
\end{tabular}
\end{center}
\end{table}
\begin{table*}[h]
{\bf Table 7:} Isophotal Fluxes and Luminosities
\normalsize
\begin{flushleft}
\begin{tabular}{cccccccccccc}
\hline
\hline
\noalign{\smallskip}
\noalign{\smallskip}
Galaxy & $f_{iso}$ (1) & $\sigma_{f_{iso}}$ (1) & $L_{iso}$ (1) & $\sigma_{L_{iso}}$ (1) & $f_{iso}$ (2) & $\sigma_{f_{iso}}$ (2) & $f_{iso}$ (3) & $A_{iso}$ & {\em S/N}\\
\noalign{\smallskip}
\noalign{\smallskip}
\ & $\left(\frac{erg}{cm^{2}s} \right)$ & $\left(\frac{erg}{cm^{2}s} \right)$ & $\left(\frac{erg}{s} \right)$ & $\left(\frac{erg}{s} \right)$ & $\left(\frac{erg}{cm^{2}s} \right)$ & $\left(\frac{erg}{cm^{2}s} \right)$ & $\left(\frac{erg}{cm^{2}s} \right)$ & $\left(arcsec^2 \right)$ & \ \\
\noalign{\smallskip}
\noalign{\smallskip}
\hline
\noalign{\smallskip}
2a & 6.29E-13 & 3E-14 & 1.39E+41 & 6E+39 & 7.60E-13 & 3E-14 & 9.12E-13 & 1643.6 & 890 \\
2b & 4.94E-13 & 2E-14 & 1.07E+41 & 5E+39 & 5.98E-13 & 3E-14 & \ & 315.8 & 1584 \\
33a & 5.69E-15 & 9E-16 & 3.73E+39 & 6E+38 & 1.52E-14 & 2E-15 & \ & 20.3 & 99 \\
33b & 3.16E-15 & 9E-16 & 2.32E+39 & 6E+38 & 8.47E-15 & 2E-15 & \ & 16.1 & 60 \\
33c & 4.03E-14 & 3E-15 & 2.82E+40 & 2E+39 & 1.08E-13 & 7E-15 & 1.12E-13 & 269.6 & 191 \\
33d & 1.18E-15 & 8E-16 & 8.16E+38 & 6E+38 & 3.17E-15 & 2E-15 & \ & 8.2 & 32 \\
34c & 3.76E-14 & 1E-14 & 3.81E+40 & 1E+40 & 7.57E-14 & 3E-14 & 9.33E-14 & 74.9 & 68 \\
35a & 1.46E-14 & 2E-15 & 4.29E+40 & 6E+39 & 1.66E-14 & 2E-15 & \ & 25.6 & 108 \\
35b & 8.15E-15 & 2E-15 & 2.53E+40 & 6E+39 & 9.28E-15 & 2E-15 & \ & 21.9 & 67 \\
35c & 9.16E-15 & 2E-15 & 2.85E+40 & 6E+39 & 1.04E-14 & 2E-15 & \ & 25.7 & 69 \\
35d & 8.95E-15 & 2E-15 & 2.59E+40 & 5E+39 & 1.02E-14 & 2E-15 & 1.10E-14 & 35.7 & 59 \\
35e & 2.09E-15 & 2E-15 & 6.84E+39 & 6E+39 & 2.38E-15 & 2E-15 & \ & 19.4 & 18 \\
35f & 1.19E-15 & 2E-15 & 3.69E+39 & 6E+39 & 1.36E-15 & 2E-15 & \ & 5.1 & 20 \\
37a & 5.12E-14 & 9E-15 & 2.66E+40 & 5E+39 & 5.65E-14 & 1E-14 & \ & 235.2 & 328 \\
37b & 2.49E-14 & 9E-15 & 1.29E+40 & 5E+39 & 2.74E-14 & 1E-14 & 3.77E-14 & 217.1 & 166 \\
37c & 5.17E-15 & 9E-15 & 3.20E+39 & 6E+39 & 5.71E-15 & 1E-14 & \ & 46.7 & 75 \\
37d & 1.74E-14 & 1E-14 & 7.68E+39 & 5E+39 & 1.93E-14 & 1E-14 & \ & 98.4 & 129 \\
37e & 3.35E-15 & 1E-14 & 1.55E+39 & 5E+39 & 3.69E-15 & 1E-14 & \ & 28.4 & 51 \\
38a & 3.67E-14 & 5E-15 & 3.23E+40 & 4E+39 & 4.31E-14 & 6E-15 & 4.60E-14 & 156.3 & 147 \\
38b & 4.29E-14 & 5E-15 & 3.76E+40 & 5E+39 & 5.03E-14 & 6E-15 & 6.53E-14 & 150.4 & 175 \\
38c & 2.06E-14 & 4E-15 & 1.82E+40 & 3E+39 & 2.41E-14 & 4E-15 & \ & 60.0 & 133 \\
43a & 5.34E-14 & 4E-15 & 6.34E+40 & 4E+39 & 6.19E-14 & 4E-15 & 6.32E-14 & 210.6 & 190 \\
43b & 4.79E-14 & 3E-15 & 5.6E+40 & 4E+39 & 5.55E-14 & 4E-15 & 7.01E-14 & 174.4 & 177 \\
43c & 2.36E-14 & 2E-15 & 2.67E+40 & 2E+39 & 2.74E-14 & 2E-15 & \ & 105.3 & 82 \\
45a & 3.70E-14 & 2E-15 & 2.06E+41 & 1E+40 & 3.85E-14 & 2E-15 & 4.81E-14 & 150.4 & 136 \\
45b & 7.10E-15 & 2E-15 & 4.1E+40 & 1E+40 & 7.39E-15 & 2E-15 & \ & 26.7 & 60 \\
45c & 2.78E-15 & 2E-15 & 1.55E+40 & 1E+40 & 2.89E-15 & 2E-15 & 3.52E-15 & 14.0 & 32 \\
46a & 4.72E-15 & 7E-16 & 3.63E+39 & 5E+38 & 5.42E-15 & 8E-16 & \ & 29.7 & 71 \\
46b & 3.08E-15 & 6E-16 & 2.60E+39 & 5E+38 & 3.54E-15 & 7E-16 & \ & 17.4 & 60 \\
46c & 1.45E-15 & 6E-16 & 1.04E+39 & 4E+38 & 1.66E-15 & 7E-16 & \ & 15.1 & 31 \\
46d & 2.96E-16 & 6E-16 & 2.01E+38 & 4E+38 & 3.41E-16 & 7E-16 & \ & 8.0 & 8.9 \\
49a & 5.05E-14 & 2E-15 & 5.74E+40 & 2E+39 & 5.47E-14 & 2E-15 & 8.64E-14 & 72.8 & 162 \\
49b & 1.18E-13 & 4E-15 & 1.34E+41 & 5E+39 & 1.28E-13 & 5E-15 & 1.75E-13 & 96.3 & 330 \\
49c & 2.35E-14 & 2E-15 & 2.67E+40 & 2E+39 & 2.55E-14 & 2E-15 & \ & 48.9 & 92 \\
49d & 1.21E-14 & 1E-15 & 1.40E+40 & 1E+39 & 1.31E-14 & 1E-15 & \ & 27.0 & 72 \\
53a & 1.44E-13 & 4E-15 & 6.44E+40 & 2E+39 & 1.55E-13 & 4E-15 & 1.55E-13 & 429.4 & 162 \\
53b & 1.48E-14 & 4E-15 & 6.41E+39 & 2E+39 & 1.59E-14 & 4E-15 & \ & 38.4 & 56 \\
53c & 9.37E-14 & 4E-15 & 3.93E+40 & 2E+39 & 1.01E-13 & 4E-15 & 1.30E-13 & 155.3 & 177 \\
\noalign{\smallskip}
\hline
\hline
\end{tabular}
\end{flushleft}
\end{table*}
\newpage
\normalsize
\begin{table*}[h]
{\bf Table 7:} Continued
\normalsize
\begin{flushleft}
\begin{tabular}{cccccccccc}
\hline
\hline
\noalign{\smallskip}
\noalign{\smallskip}
Galaxy & $f_{iso}$ (1) & $\sigma_{f_{iso}}$ (1) & $L_{iso}$ (1) & $\sigma_{L_{iso}}$ (1) & $f_{iso}$ (2) & $\sigma_{f_{iso}}$ (2) & $f_{iso}$ (3) & $A_{iso}$ & {\em S/N}\\
\noalign{\smallskip}
\noalign{\smallskip}
\ & $\left(\frac{erg}{cm^{2}s} \right)$ & $\left(\frac{erg}{cm^{2}s} \right)$ & $\left(\frac{erg}{s} \right)$ & $\left(\frac{erg}{s} \right)$ & $\left(\frac{erg}{cm^{2}s} \right)$ & $\left(\frac{erg}{cm^{2}s} \right)$ & $\left(\frac{erg}{cm^{2}s} \right)$ & $\left(arcsec^2 \right)$ & \ \\
\noalign{\smallskip}
\noalign{\smallskip}
\hline
\noalign{\smallskip}
56a & 2.15E-14 & 1E-15 & 1.68E+40 & 8E+38 & 2.26E-14 & 1E-15 & 2.65E-14 & 131.7 & 87 \\
56b & 1.58E-13 & 3E-15 & 1.13E+41 & 2E+39 & 1.66E-13 & 3E-15 & \ & 150.6 & 601 \\
56d & 2.78E-14 & 1E-15 & 2.22E+40 & 9E+38 & 2.93E-14 & 1E-15 & \ & 68.5 & 156 \\
56e & 7.86E-15 & 9E-16 & 5.65E+39 & 7E+38 & 8.26E-15 & 1E-15 & \ & 39.5 & 58 \\
66b & 5.64E-14 & 4E-15 & 3.04E+41 & 2E+40 & 5.99E-14 & 4E-15 & \ & 30.7 & 579 \\
69a & 8.13E-15 & 2E-15 & 7.32E+39 & 1E+39 & 8.67E-15 & 2E-15 & 1.05E-14 & 56.6 & 45 \\
69b & 4.43E-14 & 4E-15 & 3.86E+40 & 4E+39 & 4.73E-14 & 4E-15 & 6.75E-14 & 50.4 & 193 \\
70e & 2.22E-14 & 9E-16 & 9.46E+40 & 4E+39 & 2.35E-14 & 1E-15 & 3.71E-14 & 43.7 & 60 \\
70g & 1.14E-14 & 7E-16 & 4.87E+40 & 3E+39 & 1.21E-14 & 7E-16 & 1.92E-14 & 39.8 & 42 \\
71a & 1.45E-13 & 7E-15 & 1.45E+41 & 7E+39 & 1.56E-13 & 7E-15 & 2.31E-13 & 571.7 & 342 \\
71b & 8.71E-15 & 8E-16 & 8.72E+39 & 8E+38 & 9.36E-15 & 8E-16 & 1.04E-14 & 97.7 & 50 \\
71c & 2.62E-14 & 1E-15 & 2.14E+40 & 1E+39 & 2.81E-14 & 1E-15 & 3.94E-14 & 228.8 & 162 \\
72a & 8.12E-15 & 2E-15 & 1.47E+40 & 4E+39 & 9.04E-15 & 2E-15 & 1.19E-14 & 34.7 & 43 \\
72b & 3.81E-15 & 2E-15 & 6.72E+39 & 4E+39 & 4.24E-15 & 2E-15 & \ & 18.0 & 28 \\
72c & 2.98E-15 & 2E-15 & 5.87E+39 & 4E+39 & 3.32E-15 & 2E-15 & \ & 10.9 & 28 \\
72d & 3.83E-15 & 2E-15 & 6.97E+39 & 4E+39 & 4.26E-15 & 2E-15 & \ & 13.2 & 33 \\
74a & 2.73E-14 & 2E-15 & 4.73E+40 & 3E+39 & 3.27E-14 & 2E-15 & \ & 61.4 & 169 \\
74b & 6.56E-15 & 7E-16 & 1.11E+40 & 1E+39 & 7.84E-15 & 9E-16 & \ & 40.6 & 52 \\
74c & 5.18E-15 & 7E-16 & 8.98E+39 & 1E+39 & 6.19E-15 & 9E-16 & \ & 22.0 & 53 \\
74d & 5.73E-15 & 7E-16 & 9.01E+39 & 1E+39 & 6.85E-15 & 8E-16 & \ & 23.0 & 63 \\
75a & 2.96E-14 & 1E-15 & 5.37E+40 & 2E+39 & 3.56E-14 & 1E-15 & \ & 137.6 & 86 \\
75b & 6.22E-15 & 6E-16 & 1.07E+40 & 1E+39 & 7.48E-15 & 7E-16 & 8.64E-15 & 21.1 & 46 \\
75c & 9.06E-15 & 7E-16 & 1.58E+40 & 1E+39 & 1.09E-14 & 8E-16 & \ & 54.6 & 42 \\
75d & 2.95E-14 & 1E-15 & 5.19E+40 & 2E+39 & 3.56E-14 & 1E-15 & 4.50E-14 & 96.3 & 148 \\
75f & 1.36E-15 & 4E-16 & 2.69E+39 & 8E+38 & 1.64E-15 & 5E-16 & \ & 44.1 & 10 \\
76a & 1.20E-14 & 2E-15 & 1.39E+40 & 2E+39 & 1.42E-14 & 2E-15 & 1.69E-14 & 23.3 & 61 \\
76b & 1.17E-14 & 2E-15 & 1.35E+40 & 2E+39 & 1.4E-14 & 2E-15 & \ & 55.8 & 38 \\
76c & 1.1E-14 & 2E-15 & 1.44E+40 & 2E+39 & 1.31E-14 & 2E-15 & \ & 57.4 & 33 \\
76d & 1.18E-14 & 2E-15 & 1.4E+40 & 2E+39 & 1.40E-14 & 2E-15 & \ & 54.6 & 33 \\
76f & 5.60E-15 & 2E-15 & 6.72E+39 & 2E+39 & 6.91E-15 & 2E-15 & 9.37E-15 & 13.1 & 36 \\
82b & 4.17E-15 & 1E-15 & 5.23E+39 & 2E+39 & 4.51E-15 & 2E-15 & 4.49E-15 & 32.5 & 23 \\
82c & 3.38E-14 & 2E-15 & 3.96E+40 & 2E+39 & 3.66E-14 & 2E-15 & \ & 83.4 & 109 \\
83b & 1.31E-15 & 1E-15 & 4.12E+39 & 3E+39 & 1.71E-15 & 1E-15 & \ & 21.4 & 8 \\
83c & 3.03E-14 & 3E-15 & 9.59E+40 & 9E+39 & 3.95E-14 & 4E-15 & 4.95E-14 & 69.3 & 104 \\
92c & 5.10E-14 & 3E-15 & 2.67E+40 & 1E+39 & 7.25E-14 & 4E-15 & 1.07E-13 & 120.9 & 164 \\
\noalign{\smallskip}
\hline
\hline
\end{tabular}
\end{flushleft}
\end{table*}
\normalsize
\begin{table*}[h]
{\bf Table 8:} Isophotal Corrected and Kron Fluxes
\small
\begin{flushleft}
\begin{tabular}{ccccc|ccccc}
\hline
\hline
\noalign{\smallskip}
\noalign{\smallskip}
Galaxy & $f_{isocor}$ & $\sigma_{isocor}$ & $f_{Kron}$ & $\sigma_{Kron}$ & Galaxy & $f_{isocor}$ & $\sigma_{isocor}$ & $f_{Kron}$ & $\sigma_{Kron}$\\
\noalign{\smallskip}
\noalign{\smallskip}
\ & $\left(\frac{erg}{cm^2s} \right)$ & $\left(\frac{erg}{cm^2s} \right)$ & $\left(\frac{erg}{cm^2s} \right)$ & $\left(\frac{erg}{cm^2s} \right)$ & \ & $\left(\frac{erg}{cm^2s} \right)$ & $\left(\frac{erg}{cm^2s} \right)$ & $\left(\frac{erg}{cm^2s} \right)$ & $\left(\frac{erg}{cm^2s} \right)$ \\
\noalign{\smallskip}
\noalign{\smallskip}
\hline
\noalign{\smallskip}
2a & 6.64E-13 & 3E-14 & 6.37E-13 & 3E-14 & 74a & 2.87E-14 & 2E-15 & 2.45E-14 & 2E-15 \\
2b & 4.99E-13 & 2E-14 & 4.93E-13 & 2E-14 & 74b & 6.89E-15 & 8E-16 & 6.78E-15 & 8E-16 \\
33a & 6.05E-15 & 9E-16 & 5.38E-15 & 9E-16 & 74c & 6.27E-15 & 8E-16 & 3.99E-15 & 7E-16 \\
33b & 3.56E-15 & 9E-16 & 2.60E-15 & 8E-16 & 74d & 6.6E-15 & 7E-16 & 4.65E-15 & 7E-16 \\
33c & 4.65E-14 & 3E-15 & 4.46E-14 & 3E-15 & 75a & 3.75E-14 & 1E-15 & 1.86E-14 & 9E-16 \\
33d & 1.45E-15 & 8E-16 & 6.50E-16 & 8E-16 & 75b & 7.87E-15 & 6E-16 & 6.96E-15 & 6E-16 \\
34c & 5.54E-14 & 1E-14 & 4.47E-14 & 1E-14 & 75c & 1.24E-14 & 7E-16 & 8.22E-15 & 6E-16 \\
35a & 1.57E-14 & 2E-15 & 1.34E-14 & 2E-15 & 75d & 3.27E-14 & 1E-15 & 3.12E-14 & 1E-15 \\
35b & 8.31E-15 & 2E-15 & 8.75E-15 & 2E-15 & 75f & 2.42E-15 & 4E-16 & \ & \ \\
35c & 9.20E-15 & 2E-15 & 9.01E-15 & 2E-15 & 76a & 1.28E-14 & 2E-15 & 1.25E-14 & 2E-15 \\
35d & 1.03E-14 & 2E-15 & 9.70E-15 & 2E-15 & 76b & 1.09E-14 & 2E-15 & \ & \ \\
35e & 1.93E-15 & 2E-15 & 2.58E-15 & 2E-15 & 76c & 1.01E-14 & 2E-15 & \ & \ \\
35f & 1.73E-15 & 2E-15 & 1.20E-15 & 2E-15 & 76d & 1.14E-14 & 2E-15 & \ & \ \\
37a & 5.45E-14 & 9E-15 & 5.07E-14 & 9E-15 & 76f & 7.08E-15 & 2E-15 & 5.02E-15 & 2E-15 \\
37b & 3.01E-14 & 9E-15 & 2.74E-14 & 9E-15 & 82b & 5.52E-15 & 1E-15 & 8.79E-16 & 1E-15 \\
37c & 6.25E-15 & 9E-15 & 5.36E-15 & 9E-15 & 82c & 3.98E-14 & 2E-15 & 3.47E-14 & 2E-15 \\
37d & 1.98E-14 & 1E-14 & 1.84E-14 & 1E-14 & 83b & 1.46E-15 & 1E-15 & 2.32E-15 & 1E-15 \\
37e & 4.31E-15 & 1E-14 & 3.64E-15 & 1E-14 & 83c & 3.46E-14 & 3E-15 & 3.26E-14 & 3E-15 \\
38a & 4.34E-14 & 5E-15 & 3.89E-14 & 5E-15 & 92c & 5.66E-14 & 3E-15 & 5.28E-14 & 3E-15 \\
38b & 4.87E-14 & 6E-15 & 4.55E-14 & 6E-15 \\
38c & 2.24E-14 & 4E-15 & 2.13E-14 & 4E-15 \\
43a & 6.15E-14 & 4E-15 & 5.34E-14 & 4E-15 \\
43b & 5.42E-14 & 4E-15 & 4.93E-14 & 3E-15 \\
43c & 3.07E-14 & 2E-15 & 2.63E-14 & 2E-15 \\
45a & 4.64E-14 & 2E-15 & 4.95E-14 & 2E-15 \\
45b & 8.54E-15 & 2E-15 & 7.59E-15 & 2E-15 \\
45c & 3.9E-15 & 2E-15 & 5.03E-15 & 2E-15 \\
46a & 5.48E-15 & 7E-16 & 4.76E-15 & 7E-16 \\
46b & 3.47E-15 & 7E-16 & 2.95E-15 & 6E-16 \\
46c & 1.63E-15 & 6E-16 & 1.36E-15 & 6E-16 \\
46d & 3.55E-16 & 6E-16 & 7.21E-17 & 6E-16 \\
49a & 5.47E-14 & 2E-15 & 5.21E-14 & 2E-15 \\
49b & 1.22E-13 & 4E-15 & 1.20E-13 & 4E-15 \\
49c & 2.66E-14 & 2E-15 & 2.47E-14 & 2E-15 \\
49d & 1.25E-14 & 1E-15 & 1.28E-14 & 1E-15 \\
53a & 2.1E-13 & 4E-15 & 1.84E-13 & 4E-15 \\
53b & 1.69E-14 & 4E-15 & 1.57E-14 & 4E-15 \\
53c & 1.05E-13 & 4E-15 & 9.76E-14 & 4E-15 \\
56a & 3.15E-14 & 1E-15 & 2.65E-14 & 1E-15 \\
56b & 1.59E-13 & 3E-15 & 1.58E-13 & 3E-15 \\
56d & 2.98E-14 & 1E-15 & 2.81E-14 & 1E-15 \\
56e & 8.46E-15 & 9E-16 & 8.64E-15 & 9E-16 \\
66b & 5.87E-14 & 4E-15 & 7.98E-15 & 5E-16 \\
69a & 1.06E-14 & 2E-15 & 1.18E-14 & 2E-15 \\
69b & 4.62E-14 & 4E-15 & 4.46E-14 & 4E-15 \\
70e & 2.90E-14 & 1E-15 & 2.43E-14 & 1E-15 \\
70g & 1.3E-14 & 7E-16 & 1.03E-15 & 6E-16 \\
71a & 1.64E-13 & 8E-15 & 1.49E-13 & 7E-15 \\
71b & 6.96E-15 & 7E-16 & 1.12E-14 & 8E-16 \\
71c & 3.22E-14 & 2E-15 & 2.96E-14 & 1E-15 \\
72a & 8.44E-15 & 2E-15 & 8.40E-15 & 2E-15 \\
72b & 3.95E-15 & 2E-15 & 3.94E-15 & 2E-15 \\
72c & 2.92E-15 & 2E-15 & 2.93E-15 & 2E-15 \\
72d & 4.14E-15 & 2E-15 & 4.27E-15 & 2E-15 \\
\noalign{\smallskip}
\hline
\hline
\end{tabular}
\end{flushleft}
\end{table*}
\newpage
\normalsize
\begin{table*}[h]
{\bf Table 9:} Upper Limit to Flux and Luminosity
\begin{flushleft}
\begin{tabular}{ccccc}
\hline
\hline
\noalign{\smallskip}
\noalign{\smallskip}
Galaxy & $f_{3\sigma}$ & $\sigma$ & $L_{3\sigma}$ & $\sigma$ \\
\noalign{\smallskip}
\noalign{\smallskip}
\ & $\left(erg~cm^{-2}s^{-1} \right)$ & $\left(erg~cm^{-2}s^{-1} \right)$ & $\left(erg~s^{-1} \right)$ & $\left(erg~s^{-1} \right)$ \\
\noalign{\smallskip}
\noalign{\smallskip}
\hline
\noalign{\smallskip}
34a & 8.08E-17 & 1E-14 & 7.51E+37 & 9E+39 \\
34d & 7.19E-17 & 8E-15 & 6.41E+37 & 8E+39 \\
56c & 1.29E-16 & 9E-16 & 9.76E+37 & 7E+38 \\
66a & 7.82E-17 & 5E-18 & 3.91E+38 & 2E+37 \\
66c & 7.82E-17 & 5E-18 & 3.96E+38 & 2E+37 \\
66d & 7.82E-17 & 5E-18 & 3.98E+38 & 2E+37 \\
68a & 1.45E-16 & 2E-15 & 7.69E+36 & 8E+37 \\
68b & 1.50E-16 & 2E-15 & 1.18E+37 & 1E+38 \\
69c & 1.86E-16 & 2E-15 & 1.55E+38 & 2E+39 \\
69d & 2.50E-16 & 3E-15 & 2.40E+38 & 3E+39 \\
70f & 2.56E-16 & 6E-16 & 1.10E+39 & 3E+39 \\
74e & 1.19E-16 & 6E-16 & 1.81E+38 & 8E+38 \\
75e & 1.30E-16 & 4E-16 & 2.27E+38 & 7E+38 \\
76e & 2.39E-16 & 2E-15 & 2.94E+38 & 2E+39 \\
82a & 1.72E-16 & 1E-15 & 2.47E+38 & 2E+39 \\
82d & 1.74E-16 & 1E-15 & 2.74E+38 & 2E+39 \\
83a & 2.20E-16 & 2E-15 & 6.18E+38 & 3E+39 \\
83d & 2.06E-16 & 1E-15 & 5.73E+38 & 3E+39 \\
83e & 2.07E-16 & 1E-15 & 5.81E+38 & 3E+39 \\
92b & 2.32E-16 & 3E-15 & 8.84E+37 & 1E+39 \\
92d & 1.73E-16 & 2E-15 & 8.69E+37 & 1E+39 \\
92e & 1.73E-16 & 2E-15 & 8.62E+37 & 1E+39 \\
\noalign{\smallskip}
\hline
\hline
\end{tabular}
\end{flushleft}
\end{table*}
\section {Star formation Rate}
So far this is the largest $H_\alpha$ catalogue of HCG galaxies having $H_\alpha$ calibrated fluxes. Such a sample constitutes a powerful tool to perform quantitative analysis on the recent star formation rate inside HCG galaxies.
We have derived the {\em SFR} for the galaxies of our sample using the results of Kennicutt (1983), which relate the SFR to $H_{\alpha}$ luminosity through the relation:
\begin{equation}
SFR(total)={{L(H_\alpha)}\over{1.12\cdot10^{41} {\rm erg~~s^{-1}}}} M_\odot yr^{-1}
\end{equation}
where a Salpeter initial mass function with an upper mass cutoff of 100
$M_\odot$ has been assumed. {\em SFR} inferred from luminosities for the 73 galaxies detected and for 22 upper limits estimated is shown in Table 10 as follows:\\
Col. 1: Name of detected galaxies;\\
Col. 2: {\em SFR} inferred from isophotal luminosities $L_{iso}$ corrected for Galactic Extinction, $SFR_{iso}$ (2) ;\\
Col. 3: {\em SFR} inferred from isophotal luminosities, corrected for both Galactic and Internal Extinction, $SFR_{iso}$ (3);\\
Col. 5: Name of galaxies for which we have computed the upper limits;\\
Col. 6: {\em SFR} inferred from upper limit to luminosity, corrected for Galactic Extinction, $SFR_{ul}$ (2);\\
Col. 7: {\em SFR} inferred from upper limit to luminosity, corrected for both Galactic and Internal Extinction, $SFR_{ul}$ (3);\\
\begin{table*}[h]
{\bf Table 10:} SFR of galaxies in the sample
\small
\begin{flushleft}
\begin{tabular}{ccccccccc}
\hline
\hline
\noalign{\smallskip}
\noalign{\smallskip}
Galaxy & $SFR_{iso}$ (2) & $SFR_{iso}$ (3) & Galaxy & $SFR_{iso}$ (2) & $SFR_{iso}$ (3) & Upper Limits & $SFR_{ul}$ (2) & $SFR_{ul}$ (3)\\
\noalign{\smallskip}
\noalign{\smallskip}
\ & $M_\odot$ yr$^{-1}$ & $M_\odot$ yr$^{-1}$ & \ & $M_\odot$ yr$^{-1}$ & $M_\odot$ yr$^{-1}$ & \ & $M_\odot$ yr$^{-1}$ & $M_\odot$ yr$^{-1}$\\
\noalign{\smallskip}
\noalign{\smallskip}
\hline
\noalign{\smallskip}
2a & 1.4465 & 1.7363 & 53c & 0.3780 & 0.4860 & 34a & 0.0014 & 0.0014 \\
2b & 1.1584 & 1.1584 & 56a & 0.1572 & 0.1842 & 34d & 0.0012 & 0.0011 \\
33a & 0.0892 & 0.0892 & 56b & 1.0621 & 1.0621 & 56c & 0.0009 & 0.0009 \\
33b & 0.0556 & 0.0556 & 56d & 0.2085 & 0.2085 & 66a & 0.0037 & 0.0037 \\
33c & 0.6754 & 0.7032 & 56e & 0.0530 & 0.0530 & 66c & 0.0038 & 0.0038 \\
33d & 0.0196 & 0.0196 & 66b & 2.8838 & 2.8838 & 66d & 0.0038 & 0.0038 \\
34c & 0.6843 & 0.8435 & 69a & 0.0697 & 0.0844 & 68a & 7E-05 & 7E-05 \\
35a & 0.4360 & 0.4360 & 69b & 0.3669 & 0.5241 & 68b & 0.0001 & 0.0001 \\
35b & 0.2569 & 0.2569 & 70e & 0.8952 & 1.4116 & 69c & 0.0015 & 0.0015 \\
35c & 0.2893 & 0.2893 & 70g & 0.4604 & 0.7295 & 69d & 0.0023 & 0.0023 \\
35d & 0.2633 & 0.2845 & 71a & 1.3874 & 2.0588 & 70f & 0.0105 & 0.0154 \\
35e & 0.0695 & 0.0695 & 71b & 0.0836 & 0.0926 & 74e & 0.0019 & 0.0019 \\
35f & 0.0375 & 0.0375 & 71c & 0.2055 & 0.2877 & 75e & 0.0024 & 0.0033 \\
37a & 0.2623 & 0.2623 & 72a & 0.1457 & 0.1909 & 76e & 0.0031 & 0.0031 \\
37b & 0.1273 & 0.1748 & 72b & 0.0668 & 0.0668 & 82a & 0.0024 & 0.0024 \\
37c & 0.0316 & 0.0316 & 72c & 0.0584 & 0.0584 & 82d & 0.0027 & 0.0027 \\
37d & 0.0757 & 0.0757 & 72d & 0.0693 & 0.0693 & 83a & 0.0072 & 0.0072 \\
37e & 0.0153 & 0.0153 & 74a & 0.5054 & 0.5054 & 83d & 0.0067 & 0.01 \\
38a & 0.3385 & 0.3614 & 74b & 0.1185 & 0.11845 & 83e & 0.0068 & 0.0068 \\
38b & 0.3933 & 0.5111 & 74c & 0.0959 & 0.0959 & 92b & 0.0011 & 0.0013 \\
38c & 0.1901 & 0.1901 & 74d & 0.0962 & 0.0962 & 92d & 0.0011 & 0.0011 \\
43a & 0.6562 & 0.6701 & 75a & 0.5775 & 0.5775 & 92e & 0.0011 & 0.0017 \\
43b & 0.5796 & 0.7323 & 75b & 0.1153 & 0.1331 & \ & \ & \ \\
43c & 0.2765 & 0.2765 & 75c & 0.1698 & 0.1698 & \ & \ & \ \\
45a & 1.9142 & 2.3940 & 75d & 0.5578 & 0.7059 & \ & \ & \ \\
45b & 0.381 & 0.381 & 75f & 0.0289 & 0.0289 & \ & \ & \ \\
45c & 0.1437 & 0.1747 & 76a & 0.1475 & 0.1752 & \ & \ & \ \\
46a & 0.0373 & 0.0373 & 76b & 0.1432 & 0.1432 & \ & \ & \ \\
46b & 0.0267 & 0.0267 & 76c & 0.1524 & 0.1524 & \ & \ & \ \\
46c & 0.0106 & 0.0106 & 76d & 0.1483 & 0.1483 & \ & \ & \ \\
46d & 0.0021 & 0.0021 & 76f & 0.0740 & 0.1004 & \ & \ & \ \\
49a & 0.5545 & 0.8761 & 82b & 0.0505 & 0.0728 & \ & \ & \ \\
49b & 1.2973 & 1.7712 & 82c & 0.383 & 0.383 & \ & \ & \ \\
49c & 0.2576 & 0.2576 & 83b & 0.048 & 0.048 & \ & \ & \ \\
49d & 0.1352 & 0.1352 & 83c & 1.1171 & 1.3999 & \ & \ & \ \\
53a & 0.6193 & 0.6194 & 92c & 0.3387 & 0.5007 & \ & \ & \ \\
53b & 0.06160 & 0.0616 & \ & \ & \ & \ & \ & \ \\
\noalign{\smallskip}
\hline
\hline
\end{tabular}
\end{flushleft}
\end{table*}
\normalsize
In Figure 5 we show the distribution of $SFR_{iso}$ computed taking into account (dotted line) and without taking into account (solid line) Internal Extinction. The two distributions are quite similar, as confirmed also by a Kolmogorov-Smirnov test.
\begin{figure}[h]
\centerline{\psfig{figure=sfr.ps,height=80mm,bbllx=70mm,bblly=60mm,bburx=140mm,bbury=250mm}}
\bigskip
\bigskip
{{\bf Figure 5:} Distribution of SFR for the 73 detected galaxies derived from $L_{iso}$. The solid histogram represents the $SFR_{iso}$ (2) and the dotted histogram represents the $SFR_{iso}$ (3) (see Table 10). The width of each bin is 0.1 [$M_\odot$ yr$^{-1}$].}
\end{figure}
\vskip 0.5truecm
\section {Discussion}
Figures 6 to 11 show the continuum and $H_\alpha$ maps of some of the HCG galaxies of our sample.
A isocontour plot is also shown for the largest galaxies of these groups.
In the $H_\alpha$ images we have removed residuals of field stars for clarity and we display the galaxy flux one sigma above the background.
The scale of the axes are in pixel units and the field of view is 5.12x5.12 arcmin.
East is on the top and North on the left.
For each image the name of HCG is shown in the caption, while the name of the galaxies are reported
in the figure.
In the following we present a brief description of the groups and galaxies in figures 6 to 11.
The values regarding the flux of the lowest isocontours are corrected for Galactic Extinction.
\vskip 0.5truecm
{\bf HCG2-} This group consists of a triplet of galaxies with accordant redshifts (galaxies a, and c) plus a fainter member (galaxy d) which has a higher redshift. In Figure 6 are included only the galaxies a and b, for which we have estimated the $H_\alpha$ fluxes. Galaxy a (late type barred spiral) is brighter in $H_\alpha$ than b (compact irregular), which is also a infrared source. In the $H_\alpha$ map some knots are resolved in the disk of galaxy a. For galaxy b we detect a strong $H_\alpha$ emission in the center and no emission in the outer part, as previously noted by Vilchez \& Iglesias Paramo \cite{Vil:Para}.
The estimated $SFR_{iso}$ for a and b are respectively 1.45 and 1.16 $M_\odot$ yr$^{-1}$.
In the lowest panel of Figure 6 the isocontour plot, showing the shape and
the orientation of the $H_\alpha$ emission, is given.
The lowest contour is at 1$\sigma$ above the background,
corresponding to an $H_\alpha$ emission of $6.27 \cdot 10^{-17}~erg~cm^2s^{-1}arcsec^{-2}$.
The interval among the contours is 3$\sigma$.
The $H_\alpha$ emitting areas have an extension of about
1643.6 and 315.8 $arcsec^2$ for a and b galaxies respectively.
\vskip 0.5truecm
{\bf HCG37-} This is a compact group with five accordant galaxies: a and b are the dominant galaxies of the group. They are radio sources, as galaxy d. The group has a high velocity dispersion (398.1 $km~s^{-1}$) and mass-to-light ratio (123 $M_\odot\L_\odot$), and a short crossing time (0.0054 $Ht_c$). The $H_\alpha$ brightest source of the group is galaxy a. This is a blue elliptical galaxy with a rapidly rotating central disk of ionized gas (Rubin et al. \cite{Rub}). Galaxy b is an edge-on spiral with an intensive $H_\alpha$ emission in the center. This galaxy is also an infrared source.
Galaxies c (SOa), d (SBdm) and e (E0) are all fainter $H_\alpha$ emitters than galaxies a and b.
The $SFR_{iso}$ estimated for a, b, c, d and e galaxies are respectively 0.26, 0.13, 0.03, 0.08 and
0.02 $M_\odot$ yr$^{-1}$.
The lowest panel of Figure 7 presents the isocontours for the largest
galaxies of the group:
the extensions of $H_\alpha$ emission are of about 235.2, 217.1, 46.7, 98.4 and 28.4 $arcsec^2$
for a, b,c, d and e galaxies respectively.
The lowest contour is at 1$\sigma$ above the background,
that is $H_\alpha$ emission higher than $3.64 \cdot 10^{-17}~erg~cm^2s^{-1}arcsec^{-2}$,
while the interval among the contours is 3$\sigma$.
\vskip 0.5truecm
{\bf HCG38-} This group contains the interacting pair Arp 237 (galaxies b and c) with one other galaxy at a similar redshift (galaxy a), plus a fainter high-redshift galaxy (d). Galaxy a is a spiral showing an $H_\alpha$ emission more intense in the center than in the outer disk. The $H_\alpha$ brightest galaxy b (late type barred spiral) is in the interacting pair and it is an infrared source. In galaxy b we reveal a strong $H_\alpha$ emission in the central part of the galaxy and some resolved knots throughout its arm placed in the direction opposite to galaxy c. This last galaxy is of irregular type and it is the $H_\alpha$ dimmest galaxy of the group.
The estimated $SFR_{iso}$ for a, b and c are respectively 0.34, 0.39 and 0.19 $M_\odot$ yr$^{-1}$.
The isocontour plots in Figure 8 show the $H_\alpha$ emission higher than
$7.22 \cdot 10^{-17}~erg~cm^2s^{-1}arcsec^{-2}$ (1$\sigma$ above the background).
The interval among the contours is 1$\sigma$.
The extensions of $H_\alpha$ emission are of about 156.3, 150.4
and 60.0 $arcsec^2$ for a, b and c galaxies respectively.
\vskip 0.5truecm
{\bf HCG46-} This group consists of four early-type galaxies.
The velocity dispersion and mass-to-light ratio of the group is relatively high (respectively 323.6 $km~s^{-1}$ and 478.6 $M_\odot/L_\odot$).
Galaxies b and c appear to be in contact in the continuum image, but not in the $H_\alpha$ map. Two features are in common to all the galaxies of the group: they show a faint $H_\alpha$ emissions that seems confined to the bulge of galaxies.
The estimated $SFR_{iso}$ for a, b, c and d are respectively 0.04, 0.03, 0.01 and 0.002 $M_\odot$ yr$^{-1}$.
The extensions of H$_\alpha$ emission at 1$\sigma$ above the background are reported in Table 7.
\vskip 0.5truecm
{\bf HCG49-} This is a small and very compact group with four accordant galaxies. Its median galaxy separation is only of 12.3 $h^{-1} kpc$ and its velocity dispersion is so low (lower than the uncertainties in the velocity measurements) that no estimate can be made of its mass-to-light ratio (Hickson \cite{Hik}). Galaxies a and b are spiral, while c is an irregular and d an elliptical. Galaxy b is the $H_\alpha$ brightest source of the group, while galaxies a, c and d have comparable $H_\alpha$
emission among them.
The estimated $SFR_{iso}$ for a, b, c and d are respectively 0.56, 1.3, 0.26 and 0.14
$M_\odot$ yr$^{-1}$.
In the lowest panel of Figure 10 the isocontour plot is shown.
The lowest contour is at 1$\sigma$ above the background,
corresponding to an $H_\alpha$ emission of $9.44 \cdot 10^{-17}~erg~cm^2s^{-1}arcsec^{-2}$.
The interval among the contours is 2$\sigma$.
The extensions of H$_\alpha$ emission are of about 72.8, 96.3, 48.9 and 27 $arcsec^2$ for a, b, c
and d galaxies respectively.
\vskip 0.5truecm
{\bf HCG74-} This group contains five early-type accordant galaxies: a, b, d are elliptical and c and e are lenticular. Galaxy a is the dominant one with two very close companions (b and c) and it is also a radio source. All galaxies show an $H_\alpha$ emission confined to their center. We have not revealed e galaxy for which we have estimated the upper limit.
The estimated $SFR_{iso}$ for a, b, c and d detected galaxies are respectively 0.51, 0.12, 0.1 and
0.1 $M_\odot$ yr$^{-1}$.
In the lowest panel of Figure 11 the isocontour plots for a, b and c galaxies
are shown.
The lowest contour is at 1$\sigma$ above the background ($7.52 \cdot 10^{-17}~erg~cm^2s^{-1}arcsec^
{-2}$), while the interval among the contours is 1$\sigma$.
The extensions of H$_\alpha$ emission are 61.38, 40.59, 21.96 and 23.04 $arcsec^2$ for a, b, c and d galaxies respectively.
\vskip 1truecm
\section {Summary}
We have obtained $H_\alpha$ fluxes and luminosities for a sample of 95 galaxies from calibrated observations of 31 HCGs. The sample thus collected comprises 75$\%$ of the accordant galaxies of the observed groups and it represents the largest $H_\alpha$ selected sample of HCG galaxies so far having calibrated fluxes. By the estimated $L_{H_\alpha}$ we have obtained the star formation rate of the sample galaxies.
In a following paper (Severgnini \& Saracco, 1999) we will combine the results obtained from the data presented here
with dynamical, morphological and broad band photometrical data from the literature (Hickson \cite{Hic}, Hickson \cite{Hik}, Rood \& Struble \cite{Rod:Str}) to show that the $H_\alpha$ luminosity of galaxies and hence their current star formation rate are affected by the dynamics of groups in which they reside.\\
Further analysis based on these $H_\alpha$ data will allow us to study the rate of interaction and merger phenomena occurring in HCGs, yielding new insights about the formation and evolution of these systems.
\newpage
\begin{figure*}[h]
\centerline{\psfig{figure=2ab_con.ps,height=75mm,bbllx=70mm,bblly=60mm,bburx=140mm,bbury=250mm}}
\centerline{\psfig{figure=2.ps,height=75mm,bbllx=70mm,bblly=60mm,bburx=140mm,bbury=250mm}}
\vskip - 5truecm
\hskip -0.25truecm
\centerline{\psfig{figure=hcg2.ps,height=170mm,bbllx=70mm,bblly=60mm,bburx=140mm,bbury=250mm}}
\vskip -4truecm
{\bf Figure 6:} Continuum (up), $H_\alpha$ map (middle) and zoomed isocontour
map (down) of galaxies A and B of HCG2.
The lowest contour is at 1$\sigma$ ($6.27 \cdot 10^{-17} erg~cm^2s^{-1}arcsec^{-2}$) above the background.
The interval among the contours is 3$\sigma$.
\end{figure*}
\newpage
\begin{figure*}[h]
\centerline{\psfig{figure=37_con.ps,height=75mm,bbllx=70mm,bblly=60mm,bburx=140mm,bbury=250mm}}
\centerline{\psfig{figure=37.ps,height=75mm,bbllx=70mm,bblly=60mm,bburx=140mm,bbury=250mm}}
\vskip -5.5truecm
\hskip -0.25truecm
\centerline{\psfig{figure=hcg37.ps,height=180mm,bbllx=70mm,bblly=60mm,bburx=140mm,bbury=250mm}}
\vskip -4.5truecm
{\bf Figure 7:} Continuum (up), $H_\alpha$ map (middle) and zoomed isocontour
map (down) of HCG37.
The isocontour plots are given for the largest galaxies only (see Table 7).
The lowest contour is at 1$\sigma$ ($3.64 \cdot 10^{-17}~erg~cm^2s^{-1}arcsec^{-2}$) above the background.
The interval among the contours are 3$\sigma$.
\end{figure*}
\newpage
\begin{figure*}[h]
\centerline{\psfig{figure=38_con.ps,height=75mm,bbllx=70mm,bblly=60mm,bburx=140mm,bbury=250mm}}
\centerline{\psfig{figure=38.ps,height=75mm,bbllx=70mm,bblly=60mm,bburx=140mm,bbury=250mm}}
\vskip -7truecm
\hskip -0.5truecm
\psfig{figure=hcg38bc.ps,height=200mm,bbllx=70mm,bblly=60mm,bburx=140mm,bbury=250mm}
\vskip -22.2truecm
\hskip 6truecm
\psfig{figure=hcg38a.ps,height=240mm,bbllx=70mm,bblly=60mm,bburx=140mm,bbury=250mm}
\vskip -7truecm
{\bf Figure 8:} Continuum (up), $H_\alpha$ map (middle) and zoomed isocontour
map (down) of HCG38.
The lowest contour is at 1$\sigma$ ($7.22 \cdot 10^{-17}~erg~cm^2s^{-1}arcsec^{-2}$) above the background.
The interval is 1$\sigma$.
\end{figure*}
\newpage
\begin{figure*}[h]
\centerline{\psfig{figure=46_con.ps,height=100mm,bbllx=70mm,bblly=60mm,bburx=140mm,bbury=250mm}}
\bigskip
\bigskip
\centerline{\psfig{figure=46.ps,height=100mm,bbllx=70mm,bblly=60mm,bburx=140mm,bbury=250mm}}
\bigskip
\bigskip
{\bf Figure 9:} Continuum (up) and $H_\alpha$ map (down) of HCG46.
\end{figure*}
\newpage
\begin{figure*}[h]
\centerline{\psfig{figure=49_con.ps,height=75mm,bbllx=70mm,bblly=60mm,bburx=140mm,bbury=250mm}}
\centerline{\psfig{figure=49.ps,height=75mm,bbllx=70mm,bblly=60mm,bburx=140mm,bbury=250mm}}
\vskip -5.7truecm
\hskip -0.2truecm
\centerline{\psfig{figure=hcg49.ps,height=180mm,bbllx=70mm,bblly=60mm,bburx=140mm,bbury=250mm}}
\vskip -4.5truecm
{\bf Figure 10:} Continuum (up), $H_\alpha$ map (middle) and zoomed
isocontour map (down) of HCG49.
The lowest contour is at 1$\sigma$ ($9.44 \cdot 10^{-17}~erg~cm^2s^{-1}arcsec^{-2}$) above the background.
The interval among the contours is 2$\sigma$.
\end{figure*}
\newpage
\begin{figure*}[h]
\centerline{\psfig{figure=74_con.ps,height=75mm,bbllx=70mm,bblly=60mm,bburx=140mm,bbury=250mm}}
\centerline{\psfig{figure=74.ps,height=75mm,bbllx=70mm,bblly=60mm,bburx=140mm,bbury=250mm}}
\vskip -5.7truecm
\hskip -0.25truecm
\centerline{\psfig{figure=hcg74.ps,height=180mm,bbllx=70mm,bblly=60mm,bburx=140mm,bbury=250mm}}
\vskip -5truecm
{\bf Figure 11:} Continuum (up), $H_\alpha$ map (middle) and zoomed
isocontour map (down) of HCG74.
The isocontour plots are given for A, B and C
galaxies.
The lowest contour is at 1$\sigma$ ($7.52 \cdot 10^{-17}~erg~cm^2s^{-1}arcsec^{-2}$) above the background.
The interval among the contours is 1$\sigma$.
\end{figure*}
\begin{acknowledgements}
We are grateful to E. Recillas and D. Maccagni which have kindly carried out one of the observing runs. We would like also to thank the S. Pedro Martir observatory staff for the helpful support given during observations. PS would like to acknowledge and to thank B. Catinella and S. Molendi for the useful discussions and suggestions given during this work was in progress.
\end{acknowledgements}
|
\section{Introduction}
\label{intro}
In the period 1995-97, the centre-of-mass energy of the LEP \ee\ collider was
increased in five energy steps from 130 to 183~GeV, so opening a new energy
regime for electroweak cross section and asymmetry measurements. Such
measurements provide a test of the Standard Model (SM) and allow one to
place limits on possible extensions to it.
This paper begins by providing in Section~\ref{def} the definitions of
cross section and asymmetry used here. A brief description of the ALEPH
detector is given in Section~\ref{detector} and details of the luminosity
measurement and data/Monte Carlo samples in Section~\ref{lumi}.
Section~\ref{sec:had} describes the measurement of the \qq\ cross section.
It also explores heavy quark production, providing measurements of
\Rb\ (\Rc) which are here defined as the ratio of the \bb\ (\cc) cross
section to the total \qq\ cross section. Constraints on \qq\
forward-backward asymmetries are obtained using jet charge measurements.
In Section~\ref{sec:leptons}, cross section and asymmetry results are
reported for the three lepton species.
Based on these results, Section~\ref{interpretations} gives limits
on extensions to the SM involving contact interactions,
R-parity violating sneutrinos, leptoquarks, \ZP~bosons and R-parity violating
squarks.
\section{Definition of Cross Section and Asymmetry}
\label{def}
Cross section results for all fermion species are provided for
\begin{enumerate}
\item
the {\it inclusive} process, comprising all events with
$\sqrt{s^{\prime}/s}>0.1$, so including events having hard initial
state radiation (ISR).
\item
the {\it exclusive} process, comprising all events with
$\sqrt{s^{\prime}/s}>0.9$, so excluding radiative events, such
as those in which a return to the Z~resonance occurs.
\end{enumerate}
Here, the variable $s$ is the square of the centre-of-mass
energy. For leptonic final states the variable $s^{\prime}$ is
defined as the square of the mass of the outgoing lepton pair.
For hadronic final states $s^{\prime}$ is defined as the mass squared of
the Z/$\gamma^*$ propagator. This latter choice is necessary, because as
a result of gluon radiation, (which may occur before or after final state
photon radiation), the mass of the outgoing quark pair is not well defined.
Interference effects between ISR and final state radiative (FSR) photons
affect the exclusive cross sections at the level of a few percent
and are not accurately described by existing Monte Carlo (MC) generators.
They are particularly prominent when the outgoing fermions make a small
angle to the incoming \ee\ beams. To reduce uncertainties related to this,
the exclusive cross section and asymmetry results presented here are
defined so as to include only the polar angle region $|{\cos\theta}| < 0.95$,
where $\theta$ is the polar angle of the outgoing fermion. This is
unnecessary for the inclusive cross sections, since they are relatively
insensitive to radiative photons. The inclusive results are therefore
defined to include the full angular acceptance.
When selecting events experimentally, the variable $s^{\prime}_m$
is used, which provides a good approximation to $s^\prime$ when only
one ISR photon is present:
\begin{equation}
s^{\prime}_m\;=\;\frac{\sin\theta_1\;+\;\sin\theta_2\;-\;
|{\sin(\theta_1+\theta_2)}|}{\sin\theta_1\;+\;\sin\theta_2\;+\;
|{\sin(\theta_1+\theta_2)}|}\;\times\;s~.
\end{equation}
Here $\theta_1$ and $\theta_2$ are the angles of the final state
fermions f and $\bar{\mathrm f}$ measured with respect to the direction of the
incoming e$^-$ beam or with respect to the direction of a photon
seen in the apparatus and consistent with ISR. If two or more such photons
are found, the angles are measured with respect to the sum of their
three-momenta. The fermion flight directions
are determined in electron and muon pair events simply from the
directions of the reconstructed tracks, in tau pair events from the jets
reconstructed from the visible tau decay products,
and in hadronic events from the jets formed when forcing the event into
two jets after removing isolated, high energy photons detected in the
apparatus.
If an event contains two or more ISR photons, then unless these photons all
go in the same direction, the variable $s^{\prime}_m$ ceases to be a good
approximation to $s^{\prime}$. Such events can pass the exclusive selection
by being reconstructed with $\sqrt{s^{\prime}_m/s}>0.9$, but nonetheless
have $\sqrt{s^{\prime}/s}<0.9$. They are called `radiative background'.
For dilepton events, the differential cross section is measured
as a function of the angle $\theta^*$, defined by
\begin{equation}
\cos\theta^* = \frac{\sin\frac{1}{2}(\theta_+ - \theta_-)}
{\sin\frac{1}{2}(\theta_+ + \theta_-)}~,
\end{equation}
where $\theta_-$ and $\theta_+$ are the angles of the negatively
and positively charged leptons, respectively, with respect to the incoming
e$^-$ beam. The angle $\theta^*$ corresponds to the scattering angle between
the incoming e$^-$ and the outgoing l$^-$, measured in the \myll\
rest frame, provided that no large-angle ISR photons are present.
The forward-backward asymmetries are determined from the formula
\begin{equation}
\Afb{} = \frac{\sigma_F - \sigma_B}{\sigma_F + \sigma_B}~,
\end{equation}
where $\sigma_F$ and $\sigma_B$ are the cross sections to produce events with
the negatively charged lepton in the forward ($\theta_- < 90^\circ$) and
backward ($\theta_- > 90^\circ$) hemispheres, respectively, defined in the
same limited angular acceptance as given above. Determining the forward-backward
asymmetries from this formula avoids any specific assumption about
the angular dependence of the differential cross section.
\section{The ALEPH Detector}
\label{detector}
The ALEPH detector and performance are fully described in \cite{ALDET}
and \cite{ALPERFORM}. In October 1995, the silicon vertex detector (VDET)
described in these papers was replaced by an improved detector \cite{VDET},
which is used for the analyses presented here. A brief description of the
ALEPH detector follows.
The main tracking detector is a time projection chamber (TPC) lying between
radii of 30 and 180~cm from the beam axis. It provides up to 21
three-dimensional coordinates per track. Inside the TPC is a small drift
chamber (ITC) and within this, the new VDET. The latter has two layers
of silicon, each providing three-dimensional coordinates. All three
tracking detectors contribute coordinates to tracks for polar angles
to the beam axis up to $|{\cos\theta}| < 0.95$. They are immersed in a 1.5~T
axial magnetic field provided by a superconducting solenoid. This allows
the momentum $p$ of charged tracks to be measured with a resolution of
$\sigma(p)/p = 6\times 10^{-4} p_T\oplus 0.005$ (where $p_T$ is the momentum
component perpendicular to the beam axis in \GeVc). The three-dimensional
impact parameter resolution is measured with an accuracy of
$(34 + 70/p)\times (1 + 1.6\cos^4\theta)$~$\mu$m (where $p$ is measured
in \GeVc).
Between the tracking detectors and the solenoid is an electromagnetic
calorimeter (ECAL), which
provides identification of electrons and photons, and measures their
energies $E$ with a resolution of
$\sigma(E)/E = 0.18/\sqrt{E({\mathrm GeV})} + 0.009$. Outside the solenoid,
is a hadron calorimeter (HCAL), which, combined with the ECAL,
measures the energy of hadrons with a resolution of
$\sigma(E)/E = 0.85/\sqrt{E({\mathrm GeV})}$. The HCAL is also used for
muon identification, together with muon chambers lying outside it.
The ECAL and HCAL acceptances extend down to polar angles of 190 and 100 mrad
to the beam axis, respectively.
The luminosity is measured with a lead/proportional-chamber electromagnetic
calorimeter (LCAL) covering the small angle region between 46 and 122 mrad
from the beam axis. A tungsten/silicon electromagnetic calorimeter (SICAL)
covering the angular range from 24 to 58 mrad is used to provide a cross-check.
In general, charged tracks are considered {\it good} for the analyses presented
here if they originate within a cylinder of radius 2~cm and length 10~cm,
centered at the interaction point, and whose axis is parallel to the beam
axis. They must also have at least
four TPC hits, a momentum larger than 0.1~\GeVc\ and a polar angle to the
beam axis satisfying $|{\cos\theta}| < 0.95$.
By relating charged tracks to energy deposits found in the
calorimeters and using photon, electron and muon identification information,
a list of charged and neutral {\it energy flow particles} \cite{ALPERFORM} is
created for each event and used in the following analyses.
\section{Data and Monte Carlo Samples and the Luminosity Measurement}
\label{lumi}
The data used were taken at five centre-of-mass energies
which are given in Table~\ref{tab:lumi}. This table also shows the
integrated luminosity recorded at each energy point, together with its
statistical and systematic uncertainties.
\begin{table}[htbp]
\mycaption{Centre-of-mass energies and integrated luminosities of the
high energy data samples. The two uncertainties quoted on each
integrated luminosity correspond to its statistical and systematic
uncertainty respectively. \label{tab:lumi}}
\begin{center}
\begin{tabular}{|c|c|}
\hline
Energy (GeV) & Luminosity (pb$^{-1}$) \\
\hline
130.2 & $~6.03 \pm 0.03 \pm 0.05$ \\
136.2 & $~6.10 \pm 0.03 \pm 0.05$ \\
161.3 & $11.08 \pm 0.04 \pm 0.07$ \\
172.1 & $10.65 \pm 0.05 \pm 0.06$ \\
182.7 & $56.78 \pm 0.11 \pm 0.29$ \\
\hline
\end{tabular}
\end{center}
\end{table}
For analyses relying heavily on the VDET, such as the study of \bb\ production,
data collected at 130 and 136~GeV in 1995 are discarded,
since the new VDET was not fully installed. However, data
taken at these two energy points in 1997, corresponding to integrated
luminosities of 3.30 and 3.51~$\mathrm{pb}^{-1}$, respectively, are used.
The luminosity is measured using the LCAL calorimeter following the analysis
procedure described in Ref.~\cite{LCAL}, with a slightly reduced
acceptance due to the shadowing of the LCAL detector by the SICAL
below 59~mrad. The systematic uncertainty on the luminosity, which is assumed to be
fully correlated between the different energy points, includes a
theoretical uncertainty, estimated to be 0.25\%~\cite{LUMI_ERR} from the
BHLUMI~\cite{BHLUMI} Monte Carlo generator. The luminosity measurement is
checked by comparison with the luminosity measured independently using the
SICAL detector. The two are consistent within the estimated uncertainties.
Samples of Monte Carlo events were produced as follows. The generator
BHWIDE~v1.01 \cite{BHWIDE} is used for the electron pair channel and
KORALZ~v4.2 \cite{KORALZ} for the muon and tau pair channels.
Simulation of diquark events relies on production of the initial \qq\ system
and accompanying ISR photons with either KORALZ or PYTHIA~v5.7
\cite{PYTHIA}. The simulation of FSR and fragmentation is then
carried out with the program JETSET~v7.4 \cite{PYTHIA}.
The PYTHIA generator is also used for four-fermion
processes such as the Z pair and Ze$^+$e$^-$ channels. The programs
PHOT02~\cite{PHOJET}, HERWIG~v5.9~\cite{HERWIG} and PYTHIA are used to generate
the two-photon events. Finally, backgrounds from W~pair production are studied
using the generators KORALW~v1.21~\cite{KORALW} and EXCALIBUR~\cite{EXCALI}.
\section{Hadronic Final States}
\label{sec:had}
Section~\ref{subhad} describes the measurement of the \qq\ cross section.
Sections~\ref{rb} and \ref{rc} study heavy quark production, providing
measurements of \Rb\ (\Rc) which are here defined as the ratio of the
\bb\ (\cc) cross section to the total \qq\ cross section. The measured
values of \Rb\ (\Rc) are statistically independent of the \qq\ cross
section measurement. Furthermore, by measuring these ratios, rather than
the \bb\ and \cc\ cross sections, one benefits from the cancellation of
some systematic uncertainties.
Section~\ref{jet_charge} places constraints upon \qq\ forward-backward
asymmetries using a jet charge technique applied to b-enriched and b-depleted
event samples.
\
\subsection {The Hadronic Cross Section }
\label{subhad}
The hadronic event selection begins by requiring events to have at least
seven good charged tracks. The energy flow particles are then clustered into
jets using the JADE algorithm~\cite{JADE} with a clustering
parameter $y_{\mathrm cut}$ of 0.008~. Thin, low multiplicity jets
with an electromagnetic energy content
of at least 90\% and an energy of more than 10~GeV are considered to be ISR
photon candidates.
The visible mass \Mvis\ of the event is then measured using charged and
neutral energy flow particles, but excluding these photons and
energy flow particles which make an angle of less than $2^\circ$ to the beam
axis. The distribution of \Mvis\ is shown in Fig.~\ref{evis_183} for
data taken at 183~GeV. It is required to be more than 50~\GeVcc.
The inclusive selection makes the additional requirement that
$\sqrt{s^{\prime}_m/s}>0.1$. There is actually negligible acceptance for
events with $0.1 < \sqrt{s^{\prime}/s} < 0.3$ as a result of the cut on
\Mvis. The inclusive cross sections are therefore extrapolated down to
$\sqrt{s^{\prime}/s} = 0.1$ using the KORALZ generator.
The exclusive selection requires $\sqrt{s^{\prime}_m/s}>0.9$. Here $s^{\prime}_m$ is determined from the
reconstructed jet directions, when the event is clustered into two jets,
after first removing reconstructed ISR photons.
For the exclusive selection, these jets are required to have
$|{\cos\theta}| < 0.95$~.
The $\sqrt{s^{\prime}_{m}/s}$ distributions for centre-of-mass energies of
130, 161, 172 and 183~GeV are displayed in Fig.~\ref{sprim_all}, together
with the expected background.
For the exclusive process, two additional cuts are then applied. Firstly,
\Mvis\ is required to exceed 70\% of the centre-of-mass energy.
This suppresses residual events with a radiative return to the Z. Such events can
have $\sqrt{s^{\prime}_m/s}>0.9$ if they emit two or more ISR photons.
Figure~\ref{mvis_183} shows \Mvis\ for events with
$\sqrt{s^{\prime}_m/s} > 0.9$. The contribution from doubly radiative
events is clearly seen at low masses.
Secondly, when above the \WW\ threshold, (i.e. $\sqrt{s}\geq 161$~GeV), about
80\% of \WW\ background is eliminated by requiring that the thrust of the
event exceeds 0.85~. The thrust distribution is shown in Fig.~\ref{thrust_183}.
\begin{figure}[htbp]
\mbox{\epsfig{file=evis_183.eps,width=0.9\textwidth}}
\mycaption{\label{evis_183}
Visible mass distribution at a centre-of-mass energy of 183~GeV for
events having at least seven tracks. The white histogram shows
the expected signal, whilst the hashed histograms give the expected
contributions of the $\gamma\gamma$ and W~pair backgrounds.
The arrow indicates the cut used in the inclusive selection.}
\end{figure}
\begin{figure}[p]
\mbox{\epsfig{file=sprim_all.eps,width=0.9\textwidth}}
\mycaption{\label{sprim_all}
$\sqrt{s^{\prime}_m/s}$ distribution for hadronic
events at centre-of-mass energies from 130 to 183 GeV.
This is compared to the Monte Carlo expectations shown by the white
histograms. The hashed areas correspond to background
contributions and are dominated by W pair production. The
distribution at 136~GeV is omitted because it closely resembles
that at 130~GeV.}
\end{figure}
\begin{figure}[htbp]
\mbox{\epsfig{file=mvis_183.eps,width=0.9\textwidth}}
\mycaption{\label{mvis_183}
Visible mass distribution at 183~GeV centre-of-mass energy for
hadronic events with $\sqrt{s^{\prime}_m/s}>0.9$. This is compared
with Monte Carlo expectations shown by the white histogram.
The hashed area represents the expected contribution from events
with generated $\sqrt{s^{\prime}/s}<0.9$. The arrow indicates the
cut used in the exclusive selection.}
\end{figure}
\begin{figure}[htbp]
\mbox{\epsfig{file=thrust_183.eps,width=0.9\textwidth}}
\mycaption{\label{thrust_183}
Thrust distribution at 183~GeV centre-of-mass energy for
hadronic events with $\sqrt{s^{\prime}_m/s}>0.9$. The white histogram
shows the expected signal and the hashed one corresponds to the
W pair background. The arrow indicates the cut used in the exclusive
selection.}
\end{figure}
The selection efficiencies are estimated using the KORALZ Monte Carlo
samples. The efficiencies are given for all centre-of-mass
energies in Table~\ref{EFF_ALL}.
\begin{table}[p]
\mycaption{\label{EFF_ALL}
Selection efficiencies and background fractions. For \ee\ production,
numbers are given for (1) $-0.9<\cos\theta^*<0.9$ and
(2) $-0.9<\cos\theta^*<0.7$.}
\begin{center}
\begin{tabular}{| c | c | c || c | c |}
\hline\protect\rule{0ex}{1.4em}
$\sqrt{s^{\prime}/s}$ & $\mathrm E_{cm}$ & Event & Efficiency & Background \\
cut & (GeV) & type & (\%) & (\%) \\
\hline\hline
0.1 & 130 & \qq & $89.6\pm 0.9$ & $~2.4\pm 0.3$ \\
& & $\mu^+\mu^-$ & $78.8\pm 1.6$ & $~0.4\pm 0.1$ \\
& & $\tau^+\tau^-$ & $55.3\pm 1.1$ & $~5.7\pm 2.4$ \\
\cline{2-5}
& 136 & \qq & $89.5\pm 0.7$ & $~1.6\pm 0.2$ \\
& & $\mu^+\mu^-$ & $78.8\pm 1.6$ & $~0.5\pm 0.1$ \\
& & $\tau^+\tau^-$ & $53.9\pm 1.3$ & $~8.9\pm 4.0$ \\
\cline{2-5}
& 161 & \qq & $88.4\pm 0.6$ & $~4.0\pm 0.1$ \\
& & $\mu^+\mu^-$ & $76.5\pm 1.5$ & $~3.3\pm 1.2$ \\
& & $\tau^+\tau^-$ & $44.2\pm 1.4$ & $~8.3\pm 4.0$ \\
\cline{2-5}
& 172 & \qq & $87.3\pm 0.6$ & $10.2\pm 0.1$ \\
& & $\mu^+\mu^-$ & $75.6\pm 1.6$ & $~4.1\pm 1.3$ \\
& & $\tau^+\tau^-$ & $44.8\pm 1.0$ & $11.3\pm 3.7$ \\
\cline{2-5}
& 183 & \qq & $83.9\pm 0.7$ & $17.8\pm 0.2$ \\
& & $\mu^+\mu^-$ & $75.6\pm 1.6$ & $~5.4\pm 1.2$ \\
& & $\tau^+\tau^-$ & $47.4\pm 0.9$ & $12.9\pm 2.6$ \\
\hline\hline
0.9 & 130 & \qq & $92.2\pm 1.0$ & $~9.5\pm 0.3$ \\
& & \ee (1) & $90.0\pm 1.6$ & $10.4\pm 0.9$ \\
& & \ee (2) & $95.5\pm 1.4$ & $11.5\pm 1.0$ \\
& & $\mu^+\mu^-$ & $95.1\pm 1.8$ & $~3.3\pm 0.7$ \\
& & $\tau^+\tau^-$ & $63.9\pm 1.6$ & $11.9\pm 3.2$ \\
\cline{2-5}
& 136 & \qq & $89.2\pm 0.8$ & $~8.8\pm 0.4$ \\
& & \ee (1) & $90.6\pm 1.5$ & $10.8\pm 1.0$ \\
& & \ee (2) & $96.9\pm 1.4$ & $14.1\pm 1.6$ \\
& & $\mu^+\mu^-$ & $95.1\pm 1.8$ & $~3.3\pm 0.7$ \\
& & $\tau^+\tau^-$ & $63.6\pm 1.5$ & $15.9\pm 3.1$ \\
\cline{2-5}
& 161 & \qq & $89.3\pm 0.7$ & $~7.7\pm 0.3$ \\
& & \ee (1) & $89.1\pm 1.5$ & $10.0\pm 0.9$ \\
& & \ee (2) & $94.8\pm 1.3$ & $11.0\pm 1.3$ \\
& & $\mu^+\mu^-$ & $96.0\pm 1.8$ & $~4.5\pm 0.9$ \\
& & $\tau^+\tau^-$ & $62.3\pm 1.7$ & $~9.5\pm 2.7$ \\
\cline{2-5}
& 172 & \qq & $90.2\pm 0.6$ & $~6.4\pm 0.3$ \\
& & \ee (1) & $90.7\pm 1.5$ & $10.7\pm 1.0$ \\
& & \ee (2) & $96.3\pm 1.2$ & $11.9\pm 1.4$ \\
& & $\mu^+\mu^-$ & $96.8\pm 1.8$ & $~6.0\pm 1.3$ \\
& & $\tau^+\tau^-$ & $64.6\pm 1.7$ & $13.9\pm 3.8$ \\
\cline{2-5}
& 183 & \qq & $85.9\pm 0.8$ & $~9.2\pm 0.1$ \\
& & \ee (1) & $87.5\pm 1.4$ & $10.9\pm 0.7$ \\
& & \ee (2) & $95.2\pm 1.2$ & $12.4\pm 0.9$ \\
& & $\mu^+\mu^-$ & $95.9\pm 1.8$ & $~6.5\pm 1.3$ \\
& & $\tau^+\tau^-$ & $67.9\pm 1.8$ & $13.3\pm 3.0$ \\
\hline
\end{tabular}
\end{center}
\end{table}
The uncertainties in the efficiencies arise
from the statistical uncertainties on the Monte Carlo and also from a number
of systematic effects, which are assessed as described below.
\begin{enumerate}
\item Uncertainties related to the simulation of ISR are estimated from
the difference in the efficiencies determined from KORALZ and PYTHIA.
Only the generator KORALZ simulates ISR using YFS exponentiation
\cite{KORALZ}.
\item The overall ECAL energy calibration is determined each year, from
Bhabha events recorded when running at the Z peak. Its uncertainty is
estimated to be $\pm 0.9\%$, based upon a comparison of the detector
response to low energy ($\approx 1$~GeV) electrons in \ggee\ events,
with that to high energy electrons in Bhabha events.
The calibration of the HCAL energy scale uses minimum ionizing
particles. The statistical uncertainty on this calibration is $\pm 2\%$.
These uncertainties on the calorimeter energy scales are considered
to be uncorrelated from year to year.
\item The energy response to 45~GeV jets, in data and Monte Carlo events
produced at the Z resonance, is compared and the difference parametrized as
a function of polar angle to the beam axis. The selection efficiency is
corrected using this parametrization. The systematic uncertainty due
to the energy response is derived from the change in the measured
cross section when this correction is applied.
This uncertainty, which is the largest of the detector
related ones, is considered to be correlated from year to year.
\item The measured polar angles of jets with respect to the beam axis can
suffer from small systematic biases, particularly at low polar angles.
These biases are studied using events produced at the Z~resonance.
Differences between the data and Monte Carlo are taken into account via
their effect on $s^{\prime}_m$. The associated systematic uncertainty
is given by the change in the measured cross section when applying these
corrections. It is considered to be correlated from year to year.
\end{enumerate}
The background fractions remaining after the selection are also given
in Table~\ref{EFF_ALL}. The uncertainty in these backgrounds receives
contributions from the detector calibration and energy response, estimated
as above. Additional uncertainties are studied as follows:
\begin{enumerate}
\item The two-photon background is simulated with the PYTHIA, PHOT02 and
HERWIG generators. It is normalized to the data in the region
$M_{\mathrm vis} < 50$~\GeVcc\ (Fig.~\ref{evis_183}), and the difference
between the expected cross section from PHOT02 and the normalized one
is used to calculate the uncertainty on the process. This background
dominates at centre-of-mass energies of 130 and 136~GeV.
\item Above the W pair production threshold, the \WW\ background is estimated
using KORALW. At 161~GeV the input cross section is taken
from the theoretical expectation, computed with the GENTLE \cite{GENTLE}
program for a W mass of $80.39\pm 0.06$~\GeVcc \cite{WW98}.
At 172 and 183~GeV the cross sections are taken from the average
measurement of the four LEP experiments~\cite{WW98}.
Uncertainties in the W pair cross sections are propagated to
estimate the associated systematic uncertainties on the \qq\ cross sections.
\item Other four-fermion backgrounds such as ZZ and Ze$^+$e$^-$ are
estimated using PYTHIA and EXCALIBUR. They introduce only a $\pm 0.1$\%
systematic uncertainty on the exclusive \qq\ cross section measurement.
\item Events where ISR occurs from the incoming electron and positron remain
an important background for the exclusive selection. Systematic
uncertainties on this background arise from the simulation of ISR
photons and from the energy scale and response of the detector. These
are already taken into account as described above.
\end{enumerate}
The cross section measurements at each centre-of-mass energy are listed in
Table~\ref{CROSSALL}, where they are compared with the SM
predictions. These predictions are discussed in Section~\ref{smpred}.
For the exclusive results, the measurements and predictions
refer to \qq\ final states with $|{\cos\theta}|<0.95$,
where $\theta$ is the polar angle of the quark with respect to the
beam axis. A breakdown of contributions to the systematic uncertainties on
the cross section measurements is given in Table~\ref{HADSYS}.
\begin{table}[p]
\vspace{-1cm}
\mycaption{\label{CROSSALL}
Measured cross sections with statistical and systematic uncertainties for
different channels at centre-of-mass energies from 130 to 183~GeV.
The SM predictions are also given, together with
the number of selected events (before background subtraction).
The exclusive cross sections correspond to the restricted angular
range $|\cos \theta| < 0.95$. For the Bhabha process, results are given
for (1) $-0.9 < \cos\theta^* < 0.9$ and (2) $-0.9 < \cos\theta^* < 0.7$.}
\begin{center}
\begin{tabular}{| c | c | c || r | c | c |}
\hline\protect\rule{0ex}{1.4em}
$\sqrt{s^{\prime}/s}$ & $\mathrm E_{cm}$ & Event & No.~~ & $\sigma$ & SM prediction \\
cut & (GeV) & type & Events & (pb) & (pb) \\
\hline\hline
0.1 & 130 & \qq & 1858 & $335.6~\pm 7.9~\pm 4.5~$ & 327.0~ \\
& & $\mu^+\mu^-$ & 110 & $~22.5~\pm 2.2~\pm 0.5~$ & ~21.9~ \\
& & $\tau^+\tau^-$ & 94 & $~25.9~\pm 2.9~\pm 0.6~$ & ~21.9~ \\
\cline{2-6}
& 136 & \qq & 1558 & $280.8~\pm 7.2~\pm 3.7~$ & 269.8~ \\
& & $\mu^+\mu^-$ & 103 & $~20.4~\pm 2.1~\pm 0.5~$ & ~18.7~ \\
& & $\tau^+\tau^-$ & 67 & $~17.8~\pm 2.5~\pm 0.5~$ & ~18.6~ \\
\cline{2-6}
& 161 & \qq & 1520 & $149.0~\pm 4.0~\pm 1.7~$ & 147.6~ \\
& & $\mu^+\mu^-$ & 107 & $~12.2~\pm 1.2~\pm 0.3~$ & ~11.2~ \\
& & $\tau^+\tau^-$ & 78 & $~14.6~\pm 1.8~\pm 0.3~$ & ~11.2~ \\
\cline{2-6}
& 172 & \qq & 1270 & $122.6~\pm 3.8~\pm 1.2~$ & 122.8~ \\
& & $\mu^+\mu^-$ & 74 & $~~8.8~\pm 1.1~\pm 0.2~$ & ~~9.5~ \\
& & $\tau^+\tau^-$ & 51 & $~~9.5~\pm 1.5~\pm 0.3~$ & ~~9.5~ \\
\cline{2-6}
& 183 & \qq & 6072 & $104.8~\pm 1.6~\pm 0.9~$ & 104.4~ \\
& & $\mu^+\mu^-$ & 406 & $~~8.84\pm 0.47\pm 0.19$ & ~~8.22 \\
& & $\tau^+\tau^-$ & 241 & $~~7.80\pm 0.58\pm 0.17$ & ~~8.22 \\
\hline\hline
0.9 & 130 & \qq & 440 & $~71.6~\pm 3.8~\pm 1.1~$ & ~70.7~ \\
& & \ee (1) & 1186 & $191.3~\pm 6.2~\pm 3.5~$ & 186.7~ \\
& & \ee (2) & 274 & $~41.1~\pm 2.8~\pm 0.9~$ & ~39.0~ \\
& & $\mu^+\mu^-$ & 48 & $~~7.9~\pm 1.2~\pm 0.2~$ & ~~7.0~ \\
& & $\tau^+\tau^-$ & 49 & $~10.9~\pm 1.8~\pm 0.4~$ & ~~7.3~ \\
\cline{2-6}
& 136 & \qq & 351 & $~58.8~\pm 3.5~\pm 0.9~$ & ~57.3~ \\
& & \ee (1) & 1051 & $162.2~\pm 5.6~\pm 3.5~$ & 167.3~ \\
& & \ee (2) & 212 & $~29.5~\pm 2.4~\pm 0.7~$ & ~33.9~ \\
& & $\mu^+\mu^-$ & 43 & $~~6.9~\pm 1.1~\pm 0.2~$ & ~~6.1~ \\
& & $\tau^+\tau^-$ & 27 & $~~5.6~\pm 1.3~\pm 0.2~$ & ~~6.3~ \\
\cline{2-6}
& 161 & \qq & 321 & $~29.94\pm 1.8~\pm 0.4~$ & ~30.7~ \\
& & \ee (1) & 1393 & $119.7~\pm 3.7~\pm 2.3~$ & 119.0~ \\
& & \ee (2) & 302 & $~25.6~\pm 1.7~\pm 0.5~$ & ~24.8~ \\
& & $\mu^+\mu^-$ & 50 & $~~4.49\pm 0.69\pm 0.09$ & ~~3.88 \\
& & $\tau^+\tau^-$ & 44 & $~~5.75\pm 0.96\pm 0.23$ & ~~4.01 \\
\cline{2-6}
& 172 & \qq & 271 & $~26.4~\pm 1.7~\pm 0.4~$ & ~25.1~ \\
& & \ee (1) & 1166 & $107.8~\pm 3.5~\pm 2.1~$ & 102.5~ \\
& & \ee (2) & 268 & $~23.0~\pm 1.6~\pm 0.5~$ & ~21.5~ \\
& & $\mu^+\mu^-$ & 29 & $~~2.64\pm 0.53\pm 0.06$ & ~~3.32 \\
& & $\tau^+\tau^-$ & 26 & $~~3.26\pm 0.74\pm 0.09$ & ~~3.43 \\
\cline{2-6}
& 183 & \qq & 1165 & $~21.71\pm 0.70\pm 0.23$ & ~21.1~ \\
& & \ee (1) & 5063 & $~90.9~\pm 1.4~\pm 1.7~$ & ~90.9~ \\
& & \ee (2) & 1171 & $~18.99\pm 0.63\pm 0.36$ & ~19.11 \\
& & $\mu^+\mu^-$ & 175 & $~~2.98\pm 0.24\pm 0.06$ & ~~2.89 \\
& & $\tau^+\tau^-$ & 129 & $~~2.90\pm 0.29\pm 0.09$ & ~~2.98 \\
\hline
\end{tabular}
\end{center}
\end{table}
\begin{table}[htbp]
\mycaption{\label{HADSYS}{Contributions to the systematic uncertainties on the
\qq\ cross section measurements, for all energies and for both inclusive and
exclusive processes. All quoted values are in percent.}}
\begin{center}
\begin{tabular}{| c | l || l | l | l | l | l |}
\hline\protect\rule{0ex}{1.4em}
$\sqrt{s^{\prime}/s}$ & \multicolumn{1}{c||}{Description} &
\multicolumn{5}{c|}{$\mathrm E_{cm}$ (GeV)} \\
\cline{3-7}
cut & & 130 & 136 & 161 & 172 & 183 \\
\hline\hline
0.1 & MC statistics & 0.3 & 0.3 & 0.3 & 0.3 & 0.1 \\
& ISR simulation & 0.3 & 0.3 & 0.3 & 0.3 & 0.3 \\
& Energy scale & 0.4 & 0.5 & 0.4 & 0.4 & 0.3 \\
& Detector response & 0.6 & 0.5 & 0.5 & 0.5 & 0.6 \\
& \ggqq & 0.3 & 0.2 & 0.05 & 0.05 & 0.04 \\
& \WW & --- & --- & 0.05 & 0.05 & 0.04 \\
& Luminosity & 1.0 & 1.0 & 0.7 & 0.7 & 0.5 \\
\hline
0.9 & MC statistics & 0.4 & 0.4 & 0.4 & 0.4 & 0.2 \\
& ISR simulation & 0.3 & 0.4 & 0.4 & 0.7 & 0.4 \\
& Energy scale & 0.4 & 0.3 & 0.3 & 0.3 & 0.3 \\
& Detector response & 0.9 & 0.9 & 0.7 & 0.7 & 0.7 \\
& \WW & --- & --- & 0.02 & 0.02 & 0.01 \\
& \ZZ & --- & --- & 0.01 & 0.01 & 0.01 \\
& Other four-fermion & --- & --- & 0.03 & 0.03 & 0.03 \\
& Luminosity & 1.0 & 1.0 & 0.7 & 0.7 & 0.5 \\
\hline
\end{tabular}
\end{center}
\end{table}
\subsection{Measurement of the \boldmath$\mathrm b\bar b$ Production Fraction
$R_{\mathrm b}$}
\label{rb}
The ratio \Rb\ of the \bb\ to \qq\ production cross sections is
measured at centre-of-mass energies in the range 130--183~GeV.
Hadronic events are chosen following the exclusive selection described in
Section~\ref{subhad}. The $\rm{b\bar{b}}$ events are separated from the
hadronic ones using the relatively long lifetime of b~hadrons.
From the measured impact parameters of the tracks in an event, the confidence
level that all these tracks originate from the primary vertex is
calculated~\cite{rblep1} and required to be less than a certain cut, which is
chosen to minimize the total uncertainty on $\sigma_{\rm{b}\bar{\rm b}}$.
The estimated selection efficiency and background fraction at each
centre-of-mass energy are given in Table~\ref{rbeff}.
At low centre-of-mass energies about two thirds of the background is due to
radiative events, with the remainder being dominated by \cc\ events.
At higher energies, radiative background, \cc\ events and background from
four-fermion events contribute in roughly equal proportions.
\begin{table}[htbp]
\mycaption{\label{rbeff}
Selection efficiencies and background fractions for the measurement
of \Rb.}
\begin{center}
\begin{tabular}{| c || c | c |}
\hline\protect\rule{0ex}{1.4em}
$\mathrm E_{cm}$ & Efficiency & Background \\
(GeV) & (\%) & (\%) \\
\hline
130 & $40.3\pm 1.0$ & $12.1\pm 1.6$ \\
136 & $40.8\pm 1.0$ & $10.0\pm 1.5$ \\
161 & $40.7\pm 1.5$ & $10.8\pm 1.2$ \\
172 & $38.0\pm 1.4$ & $10.6\pm 1.3$ \\
183 & $37.8\pm 0.9$ & $~9.6\pm 1.0$ \\
\hline
\end{tabular}
\end{center}
\end{table}
The largest contribution to the systematic uncertainty arises from the
b~tagging efficiency. This is evaluated by using the same technique as
above to measure the fraction \Rb\ of \bb\ events in Z data
taken in the same year. The difference between this measurement and the very
precisely known world average measurement of \Rb\ at the Z peak is then taken
as the systematic uncertainty. The b~tagging efficiency is
almost independent of the centre-of-mass energy, because the impact
parameters of tracks from a decaying b~hadron depend little on its energy.
The systematic uncertainties evaluated at the Z peak can therefore be directly
translated to higher centre-of-mass energies.
The small decrease in tagging efficiency with centre-of-mass energy seen in
Table~\ref{rbeff} is due to a decay length cut, which the b~tag uses to reject
tracks from ${\mathrm K}^0$ decay. This introduces no significant additional
uncertainty on the tagging efficiency.
Systematic uncertainties due to the charm and light quark backgrounds are
smaller. If one neglects the dependence of the quark production fractions on
centre-of-mass energy, they are already taken into account by the study
using Z peak data mentioned above. The impact parameter
resolution is monitored using tracks with negative impact parameters.
The number of observed events, together with the measured and predicted
values of \Rb\ are given in Table~\ref{rbcross}. Figure~\ref{bbsum} summarizes
the measurements of \Rb\ as a function of $\sqrt{s}$.
\begin{table}[htbp]
\mycaption{\label{rbcross}
Measured values of \Rb\ with statistical and systematic uncertainties for
\bb\ production with $\sqrt{s^\prime/s} > 0.9$ and
$|{\cos\theta}| < 0.95$, where $\theta$ is the polar angle of the b~quark
with respect to the beam axis.
The SM predictions are also given, together with
the number of selected events (before background subtraction).}
\begin{center}
\begin{tabular}{| c || c | c | c |}
\hline\protect\rule{0ex}{1.4em}
$\mathrm E_{cm}$ & No. & \Rb & SM prediction \\
(GeV) & Events & & \\
\hline
130 & 19 & $0.176\pm 0.044\pm 0.004$ & 0.190 \\
136 & 20 & $0.214\pm 0.050\pm 0.005$ & 0.186 \\
161 & 24 & $0.159\pm 0.034\pm 0.006$ & 0.175 \\
172 & 16 & $0.134\pm 0.036\pm 0.007$ & 0.173 \\
183 & 91 & $0.176\pm 0.019\pm 0.005$ & 0.171 \\
\hline
\end{tabular}
\end{center}
\end{table}
\begin{figure}[htbp]
\mbox{\epsfig{file=rb.eps,width=0.9\textwidth}}
\mycaption{Measured values of the ratio \Rb\ at various centre-of-mass energies,
compared with the SM expectations.
\label{bbsum}}
\end{figure}
\subsection{Measurement of the \boldmath$\mathrm c\bar c$ Production Fraction
$R_{\mathrm c}$}
\label{rc}
The ratio \Rc\ of the \cc\ to \qq\ production cross sections is
measured at a centre-of-mass energy of 183~GeV.
Hadronic events are chosen using the exclusive selection described in
Section~\ref{subhad}, and clustered into two jets using the JADE algorithm.
An additional acceptance cut requiring that both jets
have $|{\cos\theta}| < 0.9$ is applied to ensure that the event is well
contained inside the VDET.
Background from \bb\ events is suppressed by taking advantage of the
relatively long lifetime and high mass of b~hadrons. Three algorithms are used
sequentially. The first rejects events on the basis of track impact parameter
significances \cite{rblep1}, the second using the decay length significance
of reconstructed secondary vertices \cite{Bs_oscillation} and the third
relying on a comparison of the total invariant mass of high impact
parameter significance tracks with the charm hadron mass \cite{rblep1}.
Together the hadronic selection and \bb\ rejection have a 60\% efficiency
for \cc\ events with $\sqrt{s^\prime/s} > 0.9$. The residual fraction of
\bb\ events in the resulting sample is 4\%.
The final \cc\ selection uses a neural network. This was trained to separate
\cc\ jets (giving a neural network output close to one) from light quark
jets (giving an output close to minus one). This network uses twelve variables
per jet. These are listed below, ordered according to decreasing weight.
\begin{itemize}
\setlength{\itemsep}{0mm}
\item The sum of the rapidities with respect to the jet axis of energy flow
particles within $40^\circ$ of this axis.
\item The sphericity of the four most energetic energy flow particles in the
jet, calculated in their rest~frame.
\item The total energy of the four most energetic energy flow objects in the
jet.
\item The number of identified leptons (electrons or muons) in the jet with
momentum larger than 1.5 \GeVc.
\item The transverse momentum squared $p_\perp^2$ with respect to the
jet axis of the $\pi_{\mathrm soft}$ candidate from
${\mathrm D}^*\rightarrow\pi_{\mathrm soft}\mathrm X$, defined as the charged
track in the jet with the smallest value of $p_\perp$ and an energy
between 1 and 4~GeV.
\item The confidence level that all charged tracks in the jet originate
from the primary vertex.
\item The energy of a subjet of mass 2.1~\GeVcc\ built around the leading
energy flow particle in the jet.
\item The momentum of the leading energy flow particle in the jet.
\item The number of energy flow objects in a cone of half-angle
$40^\circ$ around the jet axis.
\item The confidence level that all charged tracks in the jet having a
rapidity with respect to the jet axis exceeding 4.9, originate
from the primary vertex.
\item The decay length significance of a reconstructed secondary vertex.
\item The energy of a reconstructed D~meson (if any). D~meson reconstruction
is attempted in the charged decay channels \mr{K\pi}, \mr{K\pi\pi}
and \mr{K\pi\pi\pi}.
\end{itemize}
The distribution of the sum of the neural network outputs for the two jets
in each event is shown in Fig.~\ref{nnet}. A lower cut is placed on this
to reject light quark events. It was also found that an upper cut served to
reject a tail of remaining \bb\ events. The final selection efficiencies and event
fractions for each flavour are summarized in Table~\ref{flavour}.
\begin{figure}[htbp]
\begin{center}
\mbox{\epsfig{figure=ccbar_sel.eps,width=0.9\textwidth,height=0.7\textwidth}}
\mycaption{Distribution of the sum of the neural network outputs for hadronic
events at 183~GeV centre-of-mass energy after hadronic selection and
\bb\ ~rejection. The lines indicate the chosen cuts.\label{nnet}}
\end{center}
\end{figure}
To evaluate the systematic uncertainty introduced by a given selection
variable, the ratio of distributions of this variable in data to Monte Carlo
is determined. After smoothing to reduce statistical fluctuations, this ratio is
then used to reweight individually \bb, \cc\ and light quark event
flavours in the Monte Carlo. The relative change in the resulting efficiency
for each flavour is taken as its systematic uncertainty. A cross-check is
performed by measuring \Rc\ on the Z data taken in the same year; the
measured value is consistent with the SM prediction. All contributions to
the systematic uncertainty are given in Table~\ref{syst}.
\begin{table}[htbp]
\begin{center}
\begin{minipage}{.45\linewidth}
\mycaption{\label{flavour}
Selection efficiency for each flavour, and its expected fraction in the final
event sample, assuming SM cross sections.}
\begin{center}
\begin{tabular}{|l|r|r|}
\hline
& \multicolumn{1}{c|}{Efficiency (\%)} & \multicolumn{1}{c|}{Fraction (\%)} \\
\hline
\uu & $6.18 \pm 0.25$~ & $14.8 \pm 0.6$~ \\
\dd & $6.66 \pm 0.33$~ & $10.3 \pm 0.5$~ \\
\myss & $6.09 \pm 0.31$~ & $9.7 \pm 0.5$~ \\
\cc & $22.07 \pm 0.48$~ & $51.8 \pm 0.8$~ \\
\bb & $8.61 \pm 0.37$~ & $13.4 \pm 0.6$~ \\
\hline
\end{tabular}
\end{center}
\end{minipage}
\hfil
\begin{minipage}{.50\linewidth}
\mycaption{\label{syst}
Contributions to the systematic uncertainty on the measured value of \Rc.}
\begin{center}
\begin{tabular}{|l|c|}
\hline
\multicolumn{1}{|c|}{Description} & Uncertainty \\
\hline
Luminosity & $\pm 0.001$ \\
Efficiency of \qq\ selection & $\pm 0.001$ \\
Background in \qq\ selection & $\pm 0.001$ \\
Performance of \bb\ rejection tag & $\mbox{}^{+0.012}_{-0.003}$ \\
Performance of neural network & $\mbox{}^{+0.006}_{-0.010}$ \\
MC statistics & $\pm 0.007$ \\
\hline\protect\rule{0ex}{1.4em}
Total & $\mbox{}^{+0.016}_{-0.013}$ \\
\hline
\end{tabular}
\end{center}
\end{minipage}
\end{center}
\end{table}
In the data, 153 events are selected, of which 74.8 are estimated from the
Monte Carlo simulation to be background. The value of \Rc\ at
$\sqrt{s}=183$~GeV, for $\sqrt{s^{\prime}/s}>0.9$ and $|{\cos\theta}| < 0.95$,
where $\theta$ is the polar angle of the c~quark with respect to the
beam axis, is found to be
\begin{equation}
R_{\mathrm c} = 0.276\pm 0.041({\mathrm stat})\
^{+0.016}_{-0.013}({\mathrm syst}) - 0.284\times (R _{\mathrm uds} - 0.584)
- 0.390\times (R_{\mathrm b} - 0.171)~,
\end{equation}
where the dependence on a possible departure of the light quark or \bb\
fractions from their SM expectations is explicitly given. The measured
value of \Rc\ is consistent with its SM expectation of 0.244~.
\subsection{\boldmath Measurement of $A_{\mathrm FB}^{\mathrm q}$ using
Jet Charge}
\label{jet_charge}
Constraints upon the forward-backward asymmetries \Afb{q}\ of \qq\
events at $\sqrt{s}=183$~GeV are obtained using a jet charge technique.
This is performed separately for b-enriched and b-depleted events, which allows
the asymmetry in \bb\ events to be well determined.
The analysis uses the hadronic events with $\sqrt{s'/s}>0.9$, selected as
described in Section~\ref{subhad}. To ensure that the events are well
contained in the tracking chambers, they are also required to satisfy
$|{\cos\theta^*}|<0.9$.
The events are then divided into two samples, according to whether they pass
or fail the b~lifetime tag of Section~\ref{rb} (here used with looser cuts).
This gives a 91\% pure sample of \bb\ events plus a sample of
predominantly light/charm quark events.
After clustering all the events into two jets, the jet charge
$Q_{\mathrm jet}$ of each jet is determined, where
\begin{equation}
Q_{\mathrm jet} =
\left.
{\displaystyle\sum_{i=1}^{\mathrm N_{track}} p_{\parallel i}^{\kappa} Q_i}
\right/
{\displaystyle\sum_{i=1}^{\mathrm N_{track}} p_{\parallel i}^{\kappa}}
\end{equation}
and the sums extend over the charged tracks in the jet. The track momentum
component parallel to the jet and its charge are $p_{\parallel i}$ and $Q_i$,
respectively. The parameter $\kappa$ is set to 0.3, because this minimizes
the uncertainty on the final result.
The mean charge difference between the forward and backward jets
$\Qfb = \langle Q_{\mathrm jet}^{\mathrm F}\rangle -
\langle Q_{\mathrm jet}^{\mathrm B}\rangle$ is then formed. The first row of
Table~\ref{Qfb} shows the observed value of \Qfb\ in the data.
\begin{table}[htbp]
\mycaption{\label{Qfb} Comparison of \Qfb\ in data
with the prediction of Equation~\ref{Qfbeqn} using SM values of
$\sigma_{\mathrm q}$ and \Afb{q}. Also shown is the contribution to this
prediction from signal and background events.}
\begin{center}
\begin{tabular}{|c||c|c|}
\hline
\Qfb & b tagged events & anti-b tagged events \\
\hline
Data & $-0.029\pm 0.018$ & $0.022\pm 0.007$ \\
Expectation in SM & $-0.052$ & ~$0.014$ \\
Contribution from signal & $-0.048$ & ~$0.018$ \\
Contribution from background & $-0.004$ & $-0.003$ \\
\hline
\end{tabular}
\end{center}
\end{table}
The expected value of \Qfb\ depends on the
\qq\ cross sections $\sigma_{\mathrm q}$ and asymmetries \Afb{q}.
Defining $\epsilon_i$ as the selection efficiency for event type $i$,
the prediction is
\begin{equation}
\label{Qfbeqn}
\Qfb = \frac
{\displaystyle \sum_{\mathrm q} \sigma_{\mathrm q} \epsilon_{\mathrm q}
\Afb{q} \delta_{\mathrm q} D_{\mathrm q} +
\sum_{\mathrm x} \sigma_{\mathrm x} \epsilon_{\mathrm x}
{\Qfb}_x
}
{\displaystyle \sum_{\mathrm q} \sigma_{\mathrm q} \epsilon_{\mathrm q} +
\sum_{\mathrm x} \sigma_{\mathrm x} \epsilon_{\mathrm x}
}~,
\end{equation}
where the sums extend over the quark flavours~q and the various background
types~x. The parameters $\delta_{\mathrm q} =
\langle {Q_{\mathrm jet}^{\mathrm q}}\rangle -
\langle {Q_{\mathrm jet}^{\mathrm\bar q}}\rangle$ give the mean charge
separation between the jet containing the quark and that containing the antiquark.
Table~\ref{qsep} shows the parameters $\delta_{\mathrm q}$ for each flavour.
They are obtained from Monte Carlo simulation, but with additional corrections
based upon a precise analysis at the Z peak \cite{Halley}. The parameters
$\delta_q$ vary only slowly with centre-of-mass energy so these corrections
are still applicable at LEP2. The most important correction is that
to $\delta_c$, which amounts to 26\%. This is due to an inadequate simulation
of charm hadron decay modes. However, even this only alters the predicted
value of \Qfb\ by about one half of the statistical error in the data.
Systematic uncertainties on the $\delta_q$ can therefore at present be
neglected.
The parameters $D_{\mathrm q}$ ($\approx 0.98$) give the dilution in
${\Qfb}_{\mathrm q} = \Afb{q} \delta_{\mathrm q} D_{\mathrm q}$ caused by the
small angular dependence of the acceptance efficiency.
\begin{table}[htbp]
\mycaption{\label{qsep} Mean jet charge separations for each quark flavour.}
\begin{center}
\begin{tabular}{|c|c|c|c|c|}
\hline
$\delta_{\mathrm u}$ & $\delta_{\mathrm d}$ & $\delta_{\mathrm s}$ & $\delta_{\mathrm c}$ & $\delta_{\mathrm b}$ \\
\hline
$0.205$ & $-0.130$ & $-0.153$ & $0.155$ & $-0.108$ \\
\hline
\end{tabular}
\end{center}
\end{table}
Assuming SM values of $\sigma_{\mathrm q}$ and \Afb{q}, the predicted values
of \Qfb\ are given in the second row of Table~\ref{Qfb}. They agree with those
measured in the data, both for the b-enriched and the b-depleted samples.
By comparing the measured value of \Qfb\ in each of these two samples with
the predictions of Equation~\ref{Qfbeqn}, two independent constraint equations
are obtained for the allowed values of $\sigma_{\mathrm q}$ and \Afb{q}.
Providing that deviations from the SM are small, it is convenient to
approximate each of these two constraints by a linear equation.
These two linear equations will be used when placing limits on physics beyond
the SM in Section~\ref{interpretations}.
If $\sigma_{\mathrm q}$ and \Afb{q} differ from their SM predictions by
$\Delta\sigma_{\mathrm q}$ and $\Delta\Afb{q}$, respectively, then a Taylor
expansion of Equation~\ref{Qfbeqn} yields
\begin{equation}
\label{Taylor}
\Qfb - {\Qfb}_{\mathrm SM} =
\sum_{\mathrm q} \frac{\partial\Qfb}{\partial\sigma_{\mathrm q}}\Delta\sigma_{\mathrm q} +
\frac{\partial\Qfb}{\partial\Afb{q}}\Delta\Afb{q}~.
\end{equation}
Dividing this equation throughout by the estimated uncertainty on the measured
value of \Qfb\ in the data gives an equation of the form
\begin{equation}
\label{Afbeqn}
\gamma = \sum_{\mathrm q} \alpha_{\mathrm q} \Delta\sigma_{\mathrm q} + \beta_{\mathrm q}\Delta\Afb{q}~.
\end{equation}
Here $\gamma$ is the difference between the measured value of
\Qfb\ and that expected in the SM, divided by
the measurement uncertainty (such that $\gamma$ has an uncertainty
of $\pm 1$). The coefficients $\alpha_{\mathrm q}$ and $\beta_{\mathrm q}$,
defined by this equation, can be expressed in terms of $\epsilon_i$,
$\delta_{\mathrm q}$, $D_{\mathrm q}$ and ${\Qfb}_x$. The values of
$\gamma$, $\alpha_{\mathrm q}$ and $\beta_{\mathrm q}$ for the b-enriched
and b-depleted event samples are given in Table~\ref{Afb}. The uncertainties
in $\alpha_{\mathrm q}$ and $\beta_{\mathrm q}$ can be neglected.
The equation obtained from the b-enriched sample can be interpreted as a
measurement of \Afb{b}, since $\beta_b$ dominates the other coefficients.
It gives
a measured value of $\Afb{b} = 0.33 \pm 0.19$, compared with the SM
prediction of 0.57~.
\begin{table}[htbp]
\mycaption{\label{Afb} Coefficients of the linear constraint
Equation~\ref{Afbeqn} derived for the b-enriched and b-depleted data samples.
The units of $\alpha$ are pb$^{-1}$.}
\begin{center}
\begin{tabular}{|l||r||c|c|c|c|c||c|c|c|c|c|}
\hline
Event Sample & \multicolumn{1}{c||}{$\gamma$} &
$\alpha_u$ & $\alpha_d$ & $\alpha_s$ & $\alpha_c$ & $\alpha_b$ &
$\beta_u$ & $\beta_d$ & $\beta_s$ & $\beta_c$ & $\beta_b$ \\
\hline
$b$ tagged & $1.3\pm 1.0$&
$0.0$ & $ 0.0$ & $ 0.0$ & $0.0$ & $-0.1$ &
$0.0$ & $ 0.0$ & $ 0.0$ & $0.4$ & $-5.4$ \\
anti-$b$ tagged & $1.2\pm 1.0$ &
$0.7$ & $-0.6$ & $-0.7$ & $0.5$ & $-0.2$ &
$6.6$ & $-2.7$ & $-3.4$ & $5.1$ & $-1.1$ \\
\hline
\end{tabular}
\end{center}
\end{table}
\section{Leptonic Final States }
\label{sec:leptons}
\subsection{The \boldmath$\mathrm e^{+}e^{-}$ Channel}
\label{sec:bhabha}
Electron pair events are selected by requiring the presence of two good tracks of
opposite charge with a polar angle to the beam axis of $|{\cos\theta}| < 0.9$.
The sum of the momenta of the two tracks must exceed
30\% of the centre-of-mass energy. The total energy associated with them in
the ECAL must be at least 40\% of the centre-of-mass energy. When calculating
this energy, if one of the tracks passes near a crack in the ECAL,
then the associated HCAL energy is included if it matches the track
extrapolation within 25~mrad. Furthermore, the energy of bremsstrahlung
photons is included when within cones of $20^\circ$ around each
track.
Only the exclusive cross section is measured, so the requirement
$\sqrt{s^\prime_m/s}> 0.9$ is applied.
The main background is due to ISR and this is reduced to a level of 10 -- 12\%,
by requiring that the invariant mass of the \ee\ final state, determined from
the measured track momenta, exceeds
80~\GeVcc. The resulting invariant mass distribution is shown in
Fig.~\ref{dielmass}. The discrepancy between data and Monte Carlo seen in
this figure arises from an inaccurate simulation of the momentum
resolution of tracks at small angles to the beam axis.
As the cut applied on the electron pair invariant mass is very loose, this
does not lead to a significant systematic uncertainty.
The normalization of the radiative background is determined from a fit to
Fig.~\ref{dielmass}, using the expected shapes of the signal and background.
The statistical uncertainty on the fit result leads to the systematic
uncertainty due to radiative background quoted in Table~\ref{ELSYS}.
The selection efficiencies are determined from BHWIDE Monte Carlo samples
at each energy point. They are given, together with the estimated background,
in Table~\ref{EFF_ALL}. The uncertainties on the efficiencies are estimated
by varying the calorimeter energy scales as in Section~\ref{subhad}.
The exclusive cross section is determined in two polar
angle ranges: $-0.9<\cos\theta^*<0.9$ and $-0.9<\cos\theta^*<0.7$. It is
dominated by $t$~channel photon exchange, particularly in the forward region.
The cross section measurements are given in Table~\ref{CROSSALL}.
The contributions to the systematic uncertainties on these measurements are
given in Table~\ref{ELSYS}. Uncertainties from possible bias in the polar
angle measurement of tracks are negligible. They have been assessed
by redetermining the cross section without using tracks having
$|\cos\theta| > 0.9$.
\begin{figure}[htbp]
\begin{center}
\mbox{\epsfig{figure=xm2el_rad183.eps,width=0.9\textwidth}}
\mycaption{Electron pair invariant mass distribution at 183~GeV
centre-of-mass energy. The data, which is shown by points, is compared
with the Monte Carlo expectations for the signal and the radiative background.
\label{dielmass}}
\end{center}
\end{figure}
\begin{table}[htbp]
\mycaption{\label{ELSYS}{Contributions to the systematic uncertainties on the
\ee\ exclusive cross section measurements, for both $\cos\theta^*$
intervals. All quoted values are in percent.}}
\begin{center}
\begin{tabular}{| c | c | l || l | l | l | l | l |}
\hline\protect\rule{0ex}{1.4em}
$\sqrt{s^{\prime}/s}$ & $\cos\theta^*$ range & \multicolumn{1}{c||}{Description} &
\multicolumn{5}{c|}{$\mathrm E_{cm}$ (GeV)} \\
\cline{4-8}
cut & & & 130 & 136 & 161 & 172 & 183 \\
\hline\hline
0.9 & $-0.9<\cos\theta^*<0.9$
& MC statistics & 0.4 & 0.4 & 0.4 & 0.4 & 0.4 \\
& & Energy scale & 1.7 & 1.6 & 1.4 & 1.6 & 1.6 \\
& & \tautau & 0.01 & 0.01 & 0.01 & 0.01 & 0.01 \\
& & Radiative background & 0.9 & 1.0 & 0.9 & 1.0 & 0.9 \\
& & Luminosity & 1.0 & 1.0 & 0.7 & 0.7 & 0.5 \\
\hline
0.9 & $-0.9<\cos\theta^*<0.7$
& MC statistics & 0.6 & 0.5 & 0.6 & 0.6 & 0.6 \\
& & Energy scale & 1.4 & 1.3 & 1.2 & 1.2 & 1.1 \\
& & \tautau & 0.2 & 0.04 & 0.03 & 0.03 & 0.03 \\
& & Radiative background & 1.3 & 1.6 & 1.3 & 1.4 & 1.3 \\
& & Luminosity & 1.0 & 1.0 & 0.7 & 0.7 & 0.5 \\
\hline
\end{tabular}
\end{center}
\end{table}
\subsection{The \boldmath $\mu^+\mu^-$\unboldmath\ Channel}
\label{dimu}
The muon pair selection requires events to contain two good, oppositely charged
tracks with momenta exceeding 6~\GeVc\ and angles to the beam
axis of $|{\cos\theta}|<0.95$. The scalar sum of the momenta of the two
tracks must exceed 60~\GeVc. The total number of good charged tracks in
the events must be no more than eight.
To limit the background from cosmic ray events,
both tracks are required to originate near the primary vertex, and to have at
least four associated ITC hits, which confirms that they were produced within
a few nanoseconds of the LEP beam crossing.
Both tracks must be identified as muons, where muon identification is based
either on the digital hit pattern associated with a track in the HCAL or
on its energy deposition in the calorimeters:
\begin{itemize}
\item
The track should fire at least 10 of the 23 drift tube planes in the HCAL.
It should also fire at least half of the HCAL planes which the track is
expected to cross (taking into account HCAL cracks), and furthermore
should fire 3 or more of the outermost 10 planes of the HCAL.
Alternatively, the track should have at least one associated hit in the
muon chambers.
\item
The energy deposition must be consistent with that of a minimum ionizing
particle,
i.e., the sum of the energies associated with the track in the ECAL and HCAL
should not exceed $60\%$ of the track momentum. Moreover, the sum of these
energies and the track momentum should be smaller than $60\%$ of the
centre-of-mass energy. To control a small misidentification background
related to calorimeter cracks, tracks are also required to have at least
one associated hit in the 10 outermost layers of the HCAL.
\end{itemize}
In the case of the exclusive selection, $\sqrt{s^{\prime}_{m}/s}>$0.9, it is
also required that the muon pair invariant mass exceed 110~\GeVcc, reducing
background from radiative events by about 40\%. The invariant mass
distribution is shown in Fig.~\ref{dimuon_mass}.
\begin{figure}[htbp]
\begin{center}
\mbox{\epsfig{figure=xm2mu183_2.eps,width=0.9\textwidth}}
\mycaption{Distribution of the invariant mass of \mumu\ events at
$\sqrt{s}=183$~GeV, with $\sqrt{s^{\prime}_{m}/s}>$ 0.9. The data points are
compared to the white histogram of Monte Carlo \mumu\ events, normalized to the same
integrated luminosity. The multi-radiative background is shown as the shaded
histogram. The vertical line shows the cut used for the exclusive selection.
\label{dimuon_mass}}
\end{center}
\end{figure}
The efficiency of the kinematic selection cuts is estimated using KORALZ
Monte Carlo events, whilst the muon identification efficiency is measured
using muon pair events in data recorded at the Z~peak in the same year.
This efficiency is typically uncertain by about $\pm 1.6$\% as a result of
the limited number of Z~events.
For the inclusive process, the main background contamination stems
from \ggmm. The systematic uncertainty associated with the normalization
of this background is estimated by comparing data and Monte Carlo in the
region of \mumu\ mass below 50~\GeVcc, which is shown in Fig.~\ref{ggmu}.
\begin{figure}[htbp]
\begin{center}
\mbox{\epsfig{figure=ggmu183dat_2.eps,width=0.9\textwidth}}
\mycaption{Distribution of the invariant mass of \mumu\ events at 183~GeV
centre-of-mass energy in the low mass region where the process \ggmm\ is
dominant. The Monte Carlo prediction is indicated by the shaded histogram.
\label{ggmu}}
\end{center}
\end{figure}
For the exclusive process, the main background comes from radiative events
and is assessed using the mass region below $0.9\sqrt{s}$ of
Fig.~\ref{dimuon_mass}.
Residual cosmic ray contamination is estimated by
relaxing cuts on track impact parameters, and is found to be less than 0.5\%.
Efficiencies and background levels are summarized in Table~\ref{EFF_ALL}, and
the cross section measurements are given in Table~\ref{CROSSALL}.
The contributions to the systematic uncertainties on the cross sections are
given in Table~\ref{MUSYS}.
\begin{table}[htbp]
\mycaption{\label{MUSYS}{Contributions to the systematic uncertainties on the
muon pair cross section measurements, for all energies and for both inclusive and
exclusive processes. All quoted values are in percent.}}
\begin{center}
\begin{tabular}{| c | l || l | l | l | l | l |}
\hline\protect\rule{0ex}{1.4em}
$\sqrt{s^{\prime}/s}$ & \multicolumn{1}{c||}{Description} &
\multicolumn{5}{c|}{$\mathrm E_{cm}$ (GeV)} \\
\cline{3-7}
cut & & 130 & 136 & 161 & 172 & 183 \\
\hline\hline
0.1 & MC statistics & 0.4 & 0.4 & 0.4 & 0.4 & 0.4 \\
& $\mu$ identification & 2.0 & 2.0 & 1.9 & 2.0 & 1.9 \\
& \ggmm & 0.0 & 0.0 & 1.0 & 1.0 & 0.7 \\
& \tautau & 0.04 & 0.05 & 0.1 & 0.1 & 0.1 \\
& \WW & --- & --- & 0.05 & 0.2 & 0.3 \\
& Luminosity & 1.0 & 1.0 & 0.7 & 0.7 & 0.5 \\
\hline
0.9 & MC statistics & 0.4 & 0.4 & 0.4 & 0.3 & 0.4 \\
& $\mu$ identification & 1.9 & 1.9 & 1.8 & 1.8 & 1.9 \\
& \tautau & 0.00 & 0.00 & 0.02 & 0.1 & 0.1 \\
& \WW & --- & --- & 0.0 & 0.03 & 0.2 \\
& Radiative background & 0.5 & 0.5 & 0.6 & 0.8 & 0.6 \\
& Luminosity & 1.0 & 1.0 & 0.7 & 0.7 & 0.5 \\
\hline
\end{tabular}
\end{center}
\end{table}
\subsection{The \boldmath$\tau^+\tau^-$\unboldmath\ Channel}
\label{sec:ditau}
The tau pair selection begins by clustering events into jets,
using the JADE algorithm with the clustering parameter $y_{\mathrm cut}$ equal
to 0.008~.
Tau jet candidates must contain between one and eight charged tracks. Events
with two tau jet candidates are selected, providing that the invariant mass
of the two jets exceeds 25~\GeVcc. This requirement removes a large part of
the $\gamma\gamma$ background.
Following an approach already used for tau identification at
LEP1~\cite{TAULEP}, each tau jet candidate is analysed and classified
as a tau lepton decay into an electron, a muon, charged hadrons or
charged hadrons plus one or more $\pi^{0}$. Both tau jets must be
classified in this way.
Events are required to have at least one tau jet candidate identified
as a decay into a muon or into charged hadrons or charged hadrons plus $\pi^0$.
To suppress background from $\gamma\gamma\rightarrow\mu^+\mu^-$ or \mumu,
events with two jets classified as muonic tau decay are discarded.
For the exclusive selection, W pair events in which both W decay
leptonically represent an important background. However, most of this
background is rejected by requiring that the acoplanarity angle between
the two taus be less than 250~mrad.
Estimated selection efficiencies and background levels are given in
Table~\ref{EFF_ALL}. A more detailed breakdown of the selection efficiencies
for the various tau decay channels is given in Table~\ref{EFF_TAU}.
The main uncertainty on the selection efficiency arises
from the energy scale of the calorimeters and is estimated as in
Section~\ref{subhad}. For the inclusive cross section the dominant background
comes from \ggtt. The systematic uncertainty associated with the normalization
of this background is estimated from a comparison of data and Monte Carlo in
the low visible mass range of selected tau pair events
(15~\GeVcc~$< M_{\mathrm vis} <$~50~\GeVcc).
In the exclusive selection, the dominant background is Bhabha events, which
sometimes pass the selection criteria if they enter cracks in the ECAL
acceptance.
The cross section measurements are listed in Table~\ref{CROSSALL}.
The contributions to the systematic uncertainties are given in
Table~\ref{TAUSYS}.
\begin{table}[htbp]
\mycaption{\label{EFF_TAU} Percentage efficiency of the \tautau\ selection for
the various tau decay channels at a centre-of-mass energy of 183~GeV.}
\begin{center}
\begin{tabular}{| c c c | r | r |}
\hline\protect\rule{0ex}{1.4em}
$\tau_1$ decay & / & $\tau_2$ decay & $\sqrt{{\rm s}^{\prime}{\rm /s}}\,>\,$0.1 \hspace*{3mm} & $\sqrt{{\rm s}^{\prime}{\rm /s}}\,>\,$0.9 \hspace*{3mm} \\
\hline
$\mu\nu_{\mu}\nu_{\tau}$ & / & $\mu\nu_{\mu}\nu_{\tau}$ &
0.8 $\pm$ 0.5 \hspace*{3mm} & 1.1 $\pm$ 1.1 \hspace*{3mm} \\
$\mu\nu_{\mu}\nu_{\tau}$ & / & hadrons$\,\nu_{\tau}$ &
56.3 $\pm$ 1.8 \hspace*{3mm} & 77.1 $\pm$ 2.5 \hspace*{3mm} \\
$\mu\nu_{\mu}\nu_{\tau}$ & / & hadrons$\,\pi^0\,\nu_{\tau}$ &
61.7 $\pm$ 1.3 \hspace*{3mm} & 87.3 $\pm$ 1.5 \hspace*{3mm} \\
$\mu\nu_{\mu}\nu_{\tau}$ & / & e$\nu_{\mathrm e}\nu_{\tau}$ &
49.2 $\pm$ 2.1 \hspace*{3mm} & 74.9 $\pm$ 2.9 \hspace*{3mm} \\
hadrons$\,\nu_{\tau}$ & / & hadrons$\,\nu_{\tau}$ &
48.0 $\pm$ 2.2 \hspace*{3mm} & 67.5 $\pm$ 3.3 \hspace*{3mm} \\
hadrons$\,\nu_{\tau}$ & / & hadrons$\,\pi^0\,\nu_{\tau}$ &
53.6 $\pm$ 1.2 \hspace*{3mm} & 77.9 $\pm$ 1.6 \hspace*{3mm} \\
hadrons$\,\nu_{\tau}$ & / & e$\nu_{\mathrm e}\nu_{\tau}$ &
31.6 $\pm$ 1.7 \hspace*{3mm} & 50.2 $\pm$ 2.9 \hspace*{3mm} \\
hadrons$\,\pi^0\,\nu_{\tau}$ & / & hadrons$\,\pi^0\,\nu_{\tau}$ &
58.7 $\pm$ 1.2 \hspace*{3mm} & 79.9 $\pm$ 1.6 \hspace*{3mm} \\
hadrons$\,\pi^0\,\nu_{\tau}$ & / & e$\nu_{\mathrm e}\nu_{\tau}$ &
35.2 $\pm$ 1.3 \hspace*{3mm} & 48.5 $\pm$ 2.1 \hspace*{3mm} \\
e$\nu_{\mathrm e}\nu_{\tau}$ & / & e$\nu_{\mathrm e}\nu_{\tau}$ &
0.0 \hspace*{15mm} & 0.0 \hspace*{15mm} \\
\hline
\end{tabular}
\end{center}
\end{table}
\begin{table}[htbp]
\mycaption{\label{TAUSYS}{Contributions to the systematic uncertainties on the
ditau cross section measurements, for all energies and for both inclusive and
exclusive processes. All quoted values are in percent.}}
\begin{center}
\begin{tabular}{| c | l || l | l | l | l | l |}
\hline\protect\rule{0ex}{1.4em}
$\sqrt{s^{\prime}/s}$ & \multicolumn{1}{c||}{Description} &
\multicolumn{5}{c|}{$\mathrm E_{cm}$ (GeV)} \\
\cline{3-7}
cut & & 130 & 136 & 161 & 172 & 183 \\
\hline\hline
0.1 & MC statistics & 0.9 & 0.9 & 1.1 & 1.1 & 1.0 \\
& Energy scale & 1.6 & 1.7 & 1.6 & 1.6 & 1.5 \\
& \ggtt & 0.5 & 0.6 & 0.7 & 1.2 & 0.5 \\
& \ggmm & 0.8 & 0.9 & 0.8 & 0.9 & 0.1 \\
& \qq & 0.4 & 0.4 & 0.1 & 0.3 & 0.3 \\
& \ee & 0.7 & 0.7 & 0.5 & 1.0 & 0.6 \\
& \WW & --- & --- & --- & 0.4 & 0.7 \\
& Luminosity & 1.0 & 1.0 & 0.7 & 0.7 & 0.5 \\
\hline
0.9 & MC statistics & 1.4 & 1.3 & 1.4 & 1.2 & 1.2 \\
& Energy scale & 2.0 & 2.0 & 2.2 & 2.1 & 2.3 \\
& \ggtt & 0.5 & 0.8 & 0.6 & 1.0 & 0.4 \\
& \qq & 0.4 & 0.6 & 0.7 & 0.3 & 0.3 \\
& \ee & 1.3 & 0.9 & 1.0 & 0.8 & 0.9 \\
& \mumu & 0.1 & 0.1 & 0.1 & 0.1 & 0.1 \\
& \WW & --- & --- & --- & 0.2 & 0.4 \\
& Radiative background & 1.0 & 1.4 & 2.7 & 0.5 & 0.7 \\
& Luminosity & 1.0 & 1.0 & 0.7 & 0.7 & 0.5 \\
\hline
\end{tabular}
\end{center}
\end{table}
\subsection {Measurement of the Lepton Asymmetries}
\label{sec:lepton_asym}
Figures~\ref{cstel183}, \ref{cstmu183} and \ref{cstau183} show the
observed $\cos \theta^*$ distributions for electron, muon and tau pair events
passing the exclusive selections at a centre-of-mass energy of 183~GeV.
\begin{figure}[htbp]
\begin{center}
\mbox{\epsfig{figure=cstel183.eps,width=0.8\textwidth}}
\mycaption{Distribution of $\cos\theta^*$ in Bhabha events at 183~GeV with
$\sqrt{s^{\prime}_m/s} > 0.9$. The data are represented by points, whilst
the Monte Carlo expectation, normalized to the same integrated luminosity,
is shown by the white histogram. The expected background is indicated by
the shaded histogram.
\label{cstel183}}
\end{center}
\end{figure}
\begin{figure}[htbp]
\begin{center}
\mbox{\epsfig{figure=cstmu183.eps,width=0.8\textwidth}}
\mycaption{Distribution of $\cos\theta^*$ in \mumu\ events at 183~GeV with
$\sqrt{s^{\prime}_m/s} > 0.9$. The data are represented by points, whilst
the Monte Carlo expectation, normalized to the same integrated luminosity,
is shown by the white histogram. The expected background is indicated by
the shaded histogram.
\label{cstmu183}}
\end{center}
\end{figure}
\begin{figure}[htbp]
\begin{center}
\mbox{\epsfig{figure=cstau183.eps,width=0.8\textwidth}}
\mycaption{Distribution of $\cos\theta^*$ in ditau events at 183~GeV with
$\sqrt{s^{\prime}_m/s} > 0.9$. The data are represented by points, whilst
the Monte Carlo expectation, normalized to the same integrated luminosity,
is shown by the white histogram. The expected background is indicated by
the shaded histogram.
\label{cstau183}}
\end{center}
\end{figure}
As discussed in Section~\ref{def}, the dilepton asymmetries are determined
from the fraction of events in which the negatively charged lepton enters
the forward/backward hemispheres with respect to the incoming electron.
They are measured only for the exclusive process and defined in the
range $|{\cos\theta}|<0.95$, where $\theta$ is the polar angle of the
outgoing lepton.
To ensure that the lepton charges are well measured, only events with
$|{\cos\theta^*}|<0.9$ are used. Furthermore, only unambiguously charged
events
are kept, i.e., the product of the charge of the two leptons must be $-1$.
This requirement removes about 0.5\% of the muon pairs and 6\% of the tau
pairs. The remaining charge misidentification level is estimated using
simulated events to be 0.002\% for muon pairs and 0.05\% for tau pairs.
The asymmetries are corrected for backgrounds and acceptance, using the
appropriate Monte Carlo samples. This includes the effect of extrapolating
from $|{\cos\theta^*}|<0.9$ to $|{\cos\theta}|<0.95$.
The measured \mumu\ and \tautau\ asymmetries are shown in Table~\ref{ASYM_ALL}.
For the Bhabha channel, rather than quoting an asymmetry, it is preferred to
give the differential cross section with respect to $\theta^*$, because the
reaction is dominated by $t$~channel photon exchange. This is given in
Table~\ref{Bhab_dsig}.
A major contribution to the systematic uncertainties on the asymmetries is
the background subtraction. The ditau channel is particularly sensitive
to the Bhabha background since this is very peaked in the forward direction.
For the
tau pair asymmetry, this gives a correction of 1.3--3.8\% for centre-of-mass
energies of 130--183~GeV, respectively. The Monte Carlo statistical error on
this correction enters as a part of the systematic uncertainty of the ditau
asymmetry.
Several other sources of systematic uncertainty are considered.
The correction for event acceptance related to the extrapolation in polar
angle introduces a significant uncertainty related to the finite Monte Carlo
statistics. It also leads to a theoretical uncertainty due to ISR/FSR
interference. This is assessed using ZFITTER.
Systematic uncertainties from charge misidentification are negligible.
\begin{table}[htbp]
\mycaption{\label{ASYM_ALL} Lepton forward-backward asymmetries with statistical and systematic
uncertainties, calculated for $\sqrt{s^\prime/s}\,>\,$0.9 in the range $|{\cos\theta}|<0.95$.}
\begin{center}
\begin{tabular}{| c | c || c | c |}
\hline
$\mathrm E_{cm}$ & Lepton & A$_{\mathrm FB}$ & SM prediction \\
(GeV) & Type & & \\
\hline\hline
130 & $\mu^+\mu^-$ & 0.83 $\pm$ 0.08 $\pm$ 0.03 & 0.70 \\
& $\tau^+\tau^-$ & 0.56 $\pm$ 0.12 $\pm$ 0.05 & 0.70 \\
\hline
136 & $\mu^+\mu^-$ & 0.63 $\pm$ 0.12 $\pm$ 0.03 & 0.68 \\
& $\tau^+\tau^-$ & 0.65 $\pm$ 0.15 $\pm$ 0.04 & 0.68 \\
\hline
161 & $\mu^+\mu^-$ & 0.63 $\pm$ 0.11 $\pm$ 0.03 & 0.61 \\
& $\tau^+\tau^-$ & 0.48 $\pm$ 0.14 $\pm$ 0.04 & 0.61 \\
\hline
172 & $\mu^+\mu^-$ & 0.72 $\pm$ 0.13 $\pm$ 0.04 & 0.59 \\
& $\tau^+\tau^-$ & 0.44 $\pm$ 0.20 $\pm$ 0.04 & 0.59 \\
\hline
183 & $\mu^+\mu^-$ & 0.54 $\pm$ 0.06 $\pm$ 0.03 & 0.58 \\
& $\tau^+\tau^-$ & 0.52 $\pm$ 0.08 $\pm$ 0.04 & 0.58 \\
\hline
\end{tabular}
\end{center}
\end{table}
\begin{table}[p]
\vspace{-1.5cm}
\mycaption{Cross sections to produce electron pairs with
$\sqrt{s^\prime/s}>0.9$ and $\cos\theta^*$ in the quoted ranges.
The quoted uncertainties include statistical and systematic components.
\label{Bhab_dsig}}
\begin{center}
\begin{tabular}{|c| @{\mbox{\hspace{0.7cm}}}r r@{\mbox{\hspace{0.7cm}}} ||
r@{$\,\;\pm\;\,$}l@{\mbox{\hspace{0.5cm}}} | r@{\mbox{\hspace{1cm}}} |}
\hline
$\mathrm E_{cm}$ (GeV) &
\multicolumn{2}{c||}{$\cos\theta^*_{\mathrm min},\cos\theta^*_{\mathrm max}$} &
\multicolumn{2}{c|}{$\sigma$ (pb)} & \multicolumn{1}{c|}{SM prediction} \\
\hline
\hline
130 & $-0.9$, & $-0.7$ & 0.19 & 0.34 & 0.37 \\
& $-0.7$, & $-0.5$ & 1.41 & 0.35 & 0.55 \\
& $-0.5$, & $-0.3$ & 1.36 & 0.45 & 1.09 \\
& $-0.3$, & $-0.1$ & 1.23 & 0.48 & 1.19 \\
& $-0.1$, & $ 0.1$ & 2.60 & 0.69 & 2.45 \\
& $ 0.1$, & $ 0.3$ & 3.78 & 0.83 & 3.82 \\
& $ 0.3$, & $ 0.5$ & 8.88 & 1.18 & 7.36 \\
& $ 0.5$, & $ 0.7$ & 21.63 & 2.12 & 22.20 \\
& $ 0.7$, & $ 0.9$ & 149.61 & 6.22 &148.00 \\
\hline
\hline
136 & $-0.9$, & $-0.7$ & 0.73 & 0.20 & 0.22 \\
& $-0.7$, & $-0.5$ & 1.16 & 0.36 & 0.62 \\
& $-0.5$, & $-0.3$ & 0.54 & 0.35 & 0.49 \\
& $-0.3$, & $-0.1$ & 0.52 & 0.41 & 0.89 \\
& $-0.1$, & $ 0.1$ & 1.46 & 0.62 & 2.09 \\
& $ 0.1$, & $ 0.3$ & 2.09 & 0.74 & 2.96 \\
& $ 0.3$, & $ 0.5$ & 6.68 & 1.08 & 6.13 \\
& $ 0.5$, & $ 0.7$ & 16.58 & 1.97 & 20.50 \\
& $ 0.7$, & $ 0.9$ & 132.55 & 5.85 & 133.00 \\
\hline
\hline
161 & $-0.9$, & $-0.7$ & 0.46 & 0.21 & 0.37 \\
& $-0.7$, & $-0.5$ & 0.88 & 0.21 & 0.44 \\
& $-0.5$, & $-0.3$ & 0.55 & 0.28 & 0.79 \\
& $-0.3$, & $-0.1$ & 0.39 & 0.26 & 0.62 \\
& $-0.1$, & $ 0.1$ & 1.24 & 0.40 & 1.43 \\
& $ 0.1$, & $ 0.3$ & 2.37 & 0.47 & 2.07 \\
& $ 0.3$, & $ 0.5$ & 5.35 & 0.73 & 4.95 \\
& $ 0.5$, & $ 0.7$ & 14.38 & 1.27 & 14.10 \\
& $ 0.7$, & $ 0.9$ & 93.76 & 3.76 & 94.20 \\
\hline
\hline
172 & $-0.9$, & $-0.7$ & 0.32 & 0.19 & 0.28 \\
& $-0.7$, & $-0.5$ & 0.88 & 0.19 & 0.34 \\
& $-0.5$, & $-0.3$ & 0.66 & 0.24 & 0.58 \\
& $-0.3$, & $-0.1$ & 0.61 & 0.23 & 0.44 \\
& $-0.1$, & $ 0.1$ & 0.95 & 0.36 & 1.23 \\
& $ 0.1$, & $ 0.3$ & 1.80 & 0.47 & 1.93 \\
& $ 0.3$, & $ 0.5$ & 4.92 & 0.71 & 4.24 \\
& $ 0.5$, & $ 0.7$ & 13.07 & 1.20 & 12.40 \\
& $ 0.7$, & $ 0.9$ & 84.61 & 3.51 & 81.10 \\
\hline
\hline
183 & $-0.9$, & $-0.7$ & 0.24 & 0.07 & 0.21 \\
& $-0.7$, & $-0.5$ & 0.29 & 0.07 & 0.25 \\
& $-0.5$, & $-0.3$ & 0.46 & 0.10 & 0.51 \\
& $-0.3$, & $-0.1$ & 0.71 & 0.12 & 0.64 \\
& $-0.1$, & $ 0.1$ & 0.83 & 0.14 & 0.90 \\
& $ 0.1$, & $ 0.3$ & 1.42 & 0.20 & 1.83 \\
& $ 0.3$, & $ 0.5$ & 3.90 & 0.29 & 3.66 \\
& $ 0.5$, & $ 0.7$ & 12.47 & 0.56 & 11.10 \\
& $ 0.7$, & $ 0.9$ & 71.90 & 1.86 & 71.80 \\
\hline
\end{tabular}
\end{center}
\end{table}
\clearpage
\section{Interpretation in Terms of New Physics}
\label{interpretations}
\subsection{Comparison with Standard Model Predictions}
\label{smpred}
The measured cross sections and asymmetries from Tables~\ref{CROSSALL}
and \ref{ASYM_ALL} are plotted as a function of centre-of-mass energy
in Figs.~\ref{line_lep} and \ref{asym_lep2}, respectively.
The results are compared with SM predictions based on
BHWIDE \cite{BHWIDE} for electron pair production and ZFITTER \cite{ZFITTER}
for all other processes. The ZFITTER predictions
\footnote{
Default ZFITTER flags are used, except for
$\mathrm BOXD=1$, $\mathrm CONV=1$, $\mathrm INTF=1$ and $\mathrm INCL=0$.
The flag $\mathrm FINR=0$ is used for hadronic events and $\mathrm FINR=1$ for
dilepton events.
}
are computed from the input values
$m_{\mathrm Z}=91.1867$~\GeVcc, $m_{\mathrm t}=174.1$~\GeVcc,
$m_{\mathrm H}=127.0$~\GeVcc, $\alpha_{\mathrm em}(M_{\mathrm Z})=1/128.896$
and $\alpha_s(M_{\mathrm Z})=0.120$. The exclusive cross sections and asymmetries are given
in the restricted angular range for the outgoing fermion direction,
$|{\cos\theta}|<0.95$. The inclusive cross sections correspond to the full
angular range.
\begin{figure}[htbp]
\begin{center}
\mbox{\epsfig{figure=line_lep.eps,width=0.9\textwidth}}
\mycaption{Measured cross sections for fermion pair production. The
curves indicate the predictions obtained from BHWIDE for the Bhabha process
and from ZFITTER for the other channels. (Some of the points are shifted
slightly along the horizontal axis to prevent them overlapping).
\label{line_lep}}
\end{center}
\end{figure}
\begin{figure}[htbp]
\begin{center}
\mbox{\epsfig{figure=asym_lep2.eps,width=0.9\textwidth}}
\mycaption{Measured asymmetries for muon and tau pair production.
The curves indicate the predictions obtained from ZFITTER. (Some of the
points are shifted slightly along the horizontal axis to prevent them
overlapping).
\label{asym_lep2}}
\end{center}
\end{figure}
Because of the poor knowledge of the contribution of ISR/FSR interference,
the ZFITTER predictions are assigned a systematic uncertainty
equal to the difference in the predictions with and without interference.
This amounts to 1.5\% for the hadronic cross section and
2\% for the \mumu, \tautau, \bb\ and \cc\ cross sections. These uncertainties would
almost double if one attempted to extrapolate the results to the full angular
range. The Bhabha cross section is assumed to be uncertain by 3\%, based on
a study of different event generators. These uncertainties are taken into
account in the calculation of limits on physics beyond the SM given in the
following sections.
The measured cross sections and asymmetries are consistent with
SM predictions. Similar results have been published by
the OPAL Collaboration~\cite{OPAL}. The L3 Collaboration has published results
covering centre-of-mass energies up to 172~GeV~\cite{L3:cross}.
\subsection{Limits on Four-Fermion Contact Interactions}
\label{contacts}
Comparing the measured exclusive difermion cross sections and
angular distributions with SM predictions allows one to place
limits upon many possible extensions to the SM.
One convenient parametrization of such effects is given by the addition of
four-fermion contact interactions \cite{Eichten:1983} to the known
SM processes. Such contact interactions are characterized by
a scale \lam, interpreted as the mass of a new heavy
particle exchanged between the incoming and outgoing fermion pairs, and a
coupling $g$ giving the strength of the interaction. Contact interactions
are, for example, expected to occur if fermions are composite.
Following the notation of Ref.~\cite{Kroha:1992}, the effective
Lagrangian for the four-fermion contact interaction in the process
$\ee\rightarrow\ff$ is given by
\begin{equation}
\label{lagrangian}
{\cal L}^{CI} = \frac{g^2\eta_{\mathrm sign}}{(1+\delta) \lam^2}
\sum_{i,j=L,R} \eta_{ij}
[\bar{\mathrm e}_i \gamma^\mu {\mathrm e}_i]
[\bar{\mathrm f}_j \gamma_\mu {\mathrm f}_j]~,
\end{equation}
with $\delta = 1$ if $\mathrm f = e$, or 0 otherwise. The fields ${\mathrm e}_{L,R}$
(${\mathrm f}_{L,R}$) are the left- and right-handed chirality
projections of electron (fermion) spinors. The coefficients
$\eta_{ij}$, which take a value between $-1$ and $+1$, indicate the relative
contribution of the different chirality combinations to the Lagrangian.
The sign of $\eta_{\mathrm sign}$ determines whether the contact interaction
interferes constructively or destructively with the SM amplitude.
Several different models are considered in this analysis, corresponding
to the choices of the $\eta_{\mathrm sign}$ and $\eta_{ij}$ given in
Table~\ref{models}.
\begin{table}[hbt]
\mycaption{\label{models} Four-fermion interaction models.}
\begin{center}
\begin{tabular}{|c||c|c|c|c|c|}
\hline
Model & $\eta_{\mathrm sign}$ & $\eta_{\mathrm{LL}}$ &
$\eta_{\mathrm{RR}}$ & $\eta_{\mathrm{LR}}$ & $\eta_{\mathrm{RL}}$ \\
\hline\hline
LL$^{\pm}$ & $\pm 1$ & 1 & 0 & 0 & 0 \\
RR$^{\pm}$ & $\pm 1$ & 0 & 1 & 0 & 0 \\
VV$^{\pm}$ & $\pm 1$ & 1 & 1 & 1 & 1 \\
AA$^{\pm}$ & $\pm 1$ & 1 & 1 & $-1$ & $-1$ \\
LR$^{\pm}$ & $\pm 1$ & 0 & 0 & 1 & 0 \\
RL$^{\pm}$ & $\pm 1$ & 0 & 0 & 0 & 1 \\
LL+RR$^{\pm}$ & $\pm 1$ & 1 & 1 & 0 & 0 \\
LR+RL$^{\pm}$ & $\pm 1$ & 0 & 0 & 1 & 1 \\
\hline
\end{tabular}
\end{center}
\end{table}
In the presence of contact interactions the differential cross section for
$\ee\rightarrow\ff$
as a function of the polar angle $\theta$ of the outgoing fermion
with respect to the $\mathrm e^-$ beam line can be written as
\begin{equation}
\label{xsection}
\frac{{d}\sigma}{{d}\cos\theta} = F_{\mathrm SM}(s,t)
\left[1 + \eps \frac{F_{\mathrm IF}^{\mathrm Born}(s,t)}
{F_{\mathrm SM}^{\mathrm Born}(s,t)}
+ \eps^2 \frac{F_{\mathrm CI}^{\mathrm Born}(s,t)}
{F_{\mathrm SM}^{\mathrm Born}(s,t)} \right]
\end{equation}
with $s$ and $t$ being the Mandelstam variables and
$\eps = g^2\eta_{\mathrm sign}/(4\pi\lam^2)$. The SM cross section
$F_{\mathrm SM}$ is computed as described in Section~\ref{smpred}.
The contributions to the cross section from the SM -- contact interaction
interference and from the pure contact interaction are denoted by
$F_{\mathrm IF}^{\mathrm Born}$ and $F_{\mathrm CI}^{\mathrm Born}$,
respectively. They are calculated in the improved Born approximation. The
Born level formulae can be found in Ref.~\cite{Kroha:1992} and these are
corrected for ISR according to Ref.~\cite{MIZA}.
Because no higher order calculations are available for the contact
interactions, the ratios of these with the improved Born predictions for the
SM cross sections are taken, to allow for a partial cancellation of
higher order effects.
The predictions of Equation~\ref{xsection} are fitted to the data using a
binned maximum likelihood method. For contact interactions affecting the
dilepton channels, the likelihood function ${\cal L}$ is defined by
\begin{equation}
\label{likeli}
{\cal L} = G(\alpha^{\mathrm corr};1)\,\prod_{i}\,G(\alpha^{\mathrm uncorr}_i;1)\,
\prod_{k}\,{\cal P}
\left(\,
N_{ik}^{\mathrm data} , \left[ N_{ik}^{\mathrm pred}(\eps) + \alpha^{\mathrm corr}\Delta n^{\mathrm corr}_{ik}
+ \alpha^{\mathrm uncorr}_i\Delta n^{\mathrm uncorr}_{ik}\right]\,
\right)~.
\end{equation}
The indices $i$ and $k$ run over the centre-of-mass energy points and
angular bins in $\cos\theta^*$, respectively.
The function ${\cal P}$ gives the Poisson probability to observe
$N_{ik}^{\mathrm data}$ events in the data if $N_{ik}^{\mathrm pred}$ are
expected. The systematic uncertainties on the expected number of events which
are (un)correlated between the centre-of-mass energy points are represented
by ($\Delta n^{\mathrm uncorr}$) $\Delta n^{\mathrm corr}$, respectively.
These uncertainties are taken into account using the parameters
$\alpha^{\mathrm corr}$ and $\alpha^{\mathrm uncorr}_i$, which are
constrained using Gaussian distributions $G$ with zero
mean and unit standard deviation.
These parameters are fitted together with the parameter \eps.
The fit range in the angular distribution is chosen to be
$|{\cos\theta^*}| < 0.9$ for the Bhabha channel and
$|{\cos\theta^*}|< 0.95$ for the muon and tau pair channels.
For contact interactions affecting hadronic events, the sum over angular
bins is dropped, and instead two additional terms are added to the likelihood
function to take into account the constraints from the measurement of
the jet charge asymmetries, given by Equation~\ref{Afbeqn} and Table~\ref{Afb}.
The contact interaction can be assumed to couple to all quark flavours
with equal strength. In this case, the jet charge asymmetry measurements
improve the limits for some models by up to 70\%, primarily because of their
sensitivity to the relative cross sections of up- and down-type quarks.
Alternatively, one can assume that contact interactions only affect the \bb\
final state. Such interactions are strongly constrained using the measurement
of \Rb\ from Section~\ref{rb} and the jet charge asymmetry in b-enriched
events of Section~\ref{jet_charge}.
Because of the quadratic dependence of the theoretical cross
sections upon \eps, the likelihood function can have two
maxima.
The 68\% confidence level limits (\eps$^+$ and \eps$^-$) on \eps\
are therefore estimated as follows:
\begin{equation}
\label{uncertainty}
\int_{-\infty}^{\eps^-} {\cal L}(\eps') d\eps' =
\int_{\eps^+}^{\infty} {\cal L}(\eps') d\eps' =
0.16\, \int_{-\infty}^{\infty} {\cal L}(\eps') d\eps'~,
\end{equation}
where for each value of $\eps'$ the parameters $\alpha_c$
and $\alpha_i$ are chosen which maximize the likelihood.
The results for contact interaction affecting leptonic final states are
listed in Table \ref{lepton_results}. Table \ref{quark_results} gives the
results obtained for contact interactions affecting both hadronic events
and all difermion events.
\begin{table}[pht]
\mycaption{\label{lepton_results} Limits on contact interactions coupling to
dilepton final states.
The 68\% confidence level range is given for $\eps$ whilst the 95\% confidence
level limits are given for $\Lambda$.
The results presented for \myll\ assume lepton universality.}
\begin{center}
\begin{tabular}{|c||r|c|c|}
\hline
Model & [$\eps^-$,$\eps^+$] (TeV$^{-2}$)\qquad & $\Lambda^-$ (TeV) & $\Lambda^+$ (TeV) \\
\hline\hline
$\ee\rightarrow\ee$ & & & \\
LL & [$-0.067,+0.021$] & 3.2 & 3.5 \\
RR & [$-0.067,+0.022$] & 3.2 & 3.4 \\
VV & [$-0.017,+0.003$] & 6.4 & 8.0 \\
AA & [$-0.018,+0.019$] & 4.2 & 5.5 \\
LR & [$-0.042,+0.015$] & 4.0 & 4.2 \\
LL+RR & [$-0.038,+0.009$] & 4.2 & 5.0 \\
LR+RL & [$-0.022,+0.006$] & 5.5 & 6.5 \\
\hline
$\ee\rightarrow\mumu$ & & & \\
LL & [$-0.014,+0.040$] & 4.7 & 4.0 \\
RR & [$-0.016,+0.043$] & 4.4 & 3.8 \\
VV & [$-0.005,+0.016$] & 7.7 & 6.3 \\
AA & [$-0.009,+0.015$] & 6.8 & 6.2 \\
LR & [$-0.270,+0.025$] & 1.8 & 3.8 \\
LL+RR & [$-0.007,+0.022$] & 6.6 & 5.4 \\
LR+RL & [$-0.260,+0.019$] & 1.9 & 5.1 \\
\hline
$\ee\rightarrow\tau^+\tau^-$ & & & \\
LL & [$-0.039,+0.032$] & 3.7 & 3.9 \\
RR & [$-0.046,+0.034$] & 3.4 & 3.7 \\
VV & [$-0.012,+0.016$] & 6.2 & 5.9 \\
AA & [$-0.022,+0.013$] & 5.2 & 5.6 \\
LR & [$-0.275,+0.033$] & 1.8 & 3.3 \\
LL+RR & [$-0.020,+0.018$] & 5.2 & 5.2 \\
LR+RL & [$-0.265,+0.025$] & 1.8 & 4.3 \\
\hline
$\ee\rightarrow\myll$ & & & \\
LL & [$-0.014,+0.020$] & 5.5 & 5.3 \\
RR & [$-0.016,+0.021$] & 5.3 & 5.1 \\
VV & [$-0.005,+0.006$] & 9.5 & 9.3 \\
AA & [$-0.007,+0.010$] & 8.0 & 7.5 \\
LR & [$-0.023,+0.019$] & 4.8 & 5.0 \\
LL+RR & [$-0.008,+0.010$] & 7.7 & 7.3 \\
LR+RL & [$-0.011,+0.009$] & 7.1 & 7.2 \\
\hline
\end{tabular}
\end{center}
\end{table}
\begin{table}[htbp]
\mycaption{\label{quark_results} Limits on contact interactions coupling to
hadronic or to all difermion final states.
The 68\% confidence level range is given for $\eps$ whilst the 95\% confidence
level limits are given for $\Lambda$.
The results presented for \ff\ assume that the contact interaction couples
to all the outgoing fermion types equally.}
\begin{center}
\begin{tabular}{|c||r|c|c|}
\hline
Model & [$\eps^-$,$\eps^+$] (TeV$^{-2}$)\qquad & $\Lambda^-$ (TeV) & $\Lambda^+$ (TeV) \\
\hline\hline\protect\rule{0ex}{1.4em}
$\ee\rightarrow\bb$ & & & \\
LL & [$-0.024,+0.013$] & 4.9 & 5.6 \\
RR & [$-0.232,-0.004$] & 1.9 & 3.9 \\
VV & [$-0.029,+0.007$] & 4.6 & 6.5 \\
AA & [$-0.016,+0.009$] & 5.9 & 7.0 \\
LR & [$-0.143,+0.054$] & 2.3 & 3.0 \\
RL & [$-0.028,+0.232$] & 3.6 & 1.9 \\
LL+RR & [$-0.018,+0.009$] & 5.7 & 6.6 \\
LR+RL & [$-0.036,+0.101$] & 3.6 & 2.8 \\
\hline\protect\rule{0ex}{1.4em}
$\ee\rightarrow\qq$ & & & \\
LL & [$-0.008,+0.022$] & 6.2 & 5.4 \\
RR & [$-0.025,+0.036$] & 4.4 & 3.9 \\
VV & [$-0.010,+0.013$] & 7.1 & 6.4 \\
AA & [$-0.004,+0.013$] & 7.9 & 7.2 \\
LR & [$-0.055,+0.079$] & 3.3 & 3.0 \\
RL & [$-0.045,+0.076$] & 4.0 & 2.4 \\
LL+RR & [$-0.007,+0.014$] & 7.4 & 6.7 \\
LR+RL & [$-0.029,+0.099$] & 4.5 & 2.9 \\
\hline\protect\rule{0ex}{1.4em}
$\ee\rightarrow\ff$ & & & \\
LL & [$-0.006,+0.016$] & 7.2 & 6.2 \\
RR & [$-0.013,+0.019$] & 5.8 & 5.4 \\
VV & [$-0.005,+0.005$] & 10.1 & 9.8 \\
AA & [$-0.003,+0.009$] & 9.8 & 8.4 \\
LR & [$-0.024,+0.020$] & 4.8 & 4.9 \\
RL & [$-0.029,+0.006$] & 4.9 & 5.7 \\
LL+RR & [$-0.004,+0.009$] & 9.2 & 8.1 \\
LR+RL & [$-0.014,+0.006$] & 6.8 & 7.6 \\
\hline
\end{tabular}
\end{center}
\end{table}
Although all the physics content is described by the well-defined
parameter \eps, it is conventional to
extract limits on the energy scale \lam, assuming $g^2=4\pi$. The
$95\%$ confidence level limits $\eps_{95}^{\pm}$ are computed according to
\begin{equation}
\int_{0}^{\eps_{95}^{+}} {\cal L}(\eps') d\eps' =
0.95\, \int_{0}^{\infty} {\cal L}(\eps') d\eps'~, \qquad
\int_{\eps_{95}^{-}}^{0} {\cal L}(\eps') d\eps' =
0.95\, \int_{-\infty}^{0} {\cal L}(\eps') d\eps'~,
\end{equation}
which are then used to obtain
\begin{equation}
\label{lambda}
\lam^{\pm} = 1 \left/ \sqrt{|\eps_{95}^{\pm}|}\right.~.
\end{equation}
Limits on the energy scale are listed in Tables \ref{lepton_results} and
\ref{quark_results}. One can drop the assumption $g^2=4\pi$, in which case
these results become limits on $\sqrt{4\pi}\lam/g$.
These results are competitive with previous analyses of contact interactions
already performed at LEP \cite{OPAL,L3:CI:1998}, at the Tevatron
\cite{CDF:CI:1997} and at HERA \cite{H1:CI:1995}. However,
models of $\ee\uu$ and $\ee\dd$
contact interactions which violate parity (LL, RR, LR and RL) are already
severely constrained by atomic physics parity violation experiments,
which quote limits of the order of $15$~TeV~\cite{Deandrea:1997}.
The LEP limits for the fully leptonic couplings or those involving b~quarks
are of particular interest since they are inaccessible
at $\mathrm p\bar p$ or $\mathrm ep$ colliders.
\subsection{\boldmath Limits on R-parity Violating Sneutrinos}
\label{snu}
Supersymmetric theories with R-parity violation have terms in
the Lagrangian of the form $\lambda_{ijk} {\mathrm L}_i {\mathrm L}_j
\bar {\mathrm E}_k$, where $\mathrm L$ denotes a lepton doublet superfield
and $\bar{\mathrm E}$ denotes a lepton singlet superfield.
The parameter $\lambda$ is a Yukawa coupling, and $i$, $j$, $k = 1$, 2, 3 are
generation indices. The $\lambda_{ijk}$, which for the purposes of this
analysis are assumed to be real, are non-vanishing only for $i < j$. These
terms allow for single production of sleptons at \ee\ collider experiments.
At LEP, dilepton production cross sections could then differ from their
SM expectations as a result of the exchange of R-parity
violating sneutrinos in the $s$ or $t$~channels \cite{snu_theory}.
Table~\ref{snu_table} shows the most interesting possibilities.
Those involving $s$~channel sneutrino exchange lead to a resonance.
For the results presented here, this resonance is assumed to have a width
of 1~\GeVcc, which can occur if the sneutrino also has R-parity conserving
decay modes~\cite{snu_theory}. For a sneutrino only having the R-parity
violating decay mode into lepton pairs, the width would be much less
than this, leading to slightly better limits.
\begin{table}[htbp]
\mycaption{\label{snu_table}
For each dilepton channel, the table shows the coupling amplitude,
the sneutrino type exchanged, and an indication of whether the exchange
occurs in the $s$ or $t$~channel.}
\begin{center}
\begin{tabular}{|c||c|c|c|}
\hline
$\lambda^2$ & \ee & $\mu^+\mu^-$ & $\tau^+\tau^-$ \\
\hline
$\lambda_{121}^2$ & $\tilde\nu_\mu$ (s,t) & $\tilde\nu_e$ (t) & --- \\
$\lambda_{131}^2$ & $\tilde\nu_\tau$ (s,t) & --- & $\tilde\nu_e$ (t) \\
$\lambda_{121}\lambda_{233}$ & --- & --- & $\tilde\nu_\mu$ (s) \\
$\lambda_{131}\lambda_{232}$ & --- & $\tilde\nu_\tau$ (s) & --- \\
\hline
\end{tabular}
\end{center}
\end{table}
Direct searches for R-parity violating sneutrinos at LEP have led to lower
limits on their masses of 72~\GeVcc\ for $\tilde\nu_e$ and 49~\GeVcc\ for
$\tilde\nu_\mu$ and $\tilde\nu_\tau$ \cite{ALEPH_snu}.
Indirect limits based upon lepton universality and leptonic tau decays imply
that the couplings in Table~\ref{snu_table} must approximately satisfy
$\lambda < 0.1 (m_{\tilde l_R}/200)$~\GeVcc, where $m_{\tilde l_R}$
is the mass of the appropriate right-handed charged slepton \cite{snu_theory}.
These limits can be improved using the dilepton
cross section and asymmetry data presented in this paper.
Limits on the couplings are obtained by comparing the measured dilepton
differential cross sections with respect to the polar angle with the
theoretical cross sections in reference \cite{snu_theory}. The likelihood
function used in the fit and the corrections for ISR, etc. follow the procedure
of Section~\ref{contacts}.
The fit is performed in terms of the parameter $\lambda^2$. Since the
likelihood function can have two minima, limits are again determined by
integrating the likelihood function with respect to $\lambda^2$.
A one-sided limit is used when $\lambda^2$ is positive definite, which occurs
when $\lambda^2 = \lambda_{121}^2$ or $\lambda_{131}^2$, but
not when $\lambda^2 = \lambda_{121}\lambda_{233}$ or
$\lambda_{131}\lambda_{232}$.
Figures~\ref{e_snu}, \ref{mu_snu} and \ref{tau_snu}
show the results for those processes involving sneutrino exchange in the
$s$~channel. Similar results have been obtained by the OPAL
Collaboration~\cite{OPAL} and
for $\tilde\nu_\tau$ by the L3 Collaboration \cite{L3_snu}.
Limits on $\lambda_{121}$ and $\lambda_{131}$ from $t$~channel exchange of
$\tilde\nu_e$ in muon and tau pair production, respectively, give much weaker
limits. These rise from $|\lambda_{1j1}| < 0.5$ at $\tilde\nu_e =$
100~\GeVcc\ to $|\lambda_{1j1}| < 0.9$ at $\tilde\nu_e =$ 300~\GeVcc.
\begin{figure}[htbp]
\begin{center}
\mbox{\epsfig{file=e_snu.eps,height=0.35\textheight}}
\end{center}
\mycaption{\label{e_snu} 95\% confidence level upper limits, obtained from the
Bhabha cross sections, on $|\lambda_{121}|$ versus the assumed
$\tilde\nu_\mu$ mass and on $|\lambda_{131}|$ versus the assumed
$\tilde\nu_\tau$ mass.}
\end{figure}
\begin{figure}[htbp]
\begin{center}
\mbox{\epsfig{file=mu_snu.eps,height=0.35\textheight}}
\end{center}
\mycaption{\label{mu_snu} 95\% confidence level upper limits, obtained from the
\mumu\ cross sections, on $\sqrt{|\lambda_{131}\lambda_{232}|}$
versus the assumed $\tilde\nu_\tau$ mass.}
\end{figure}
\begin{figure}[htbp]
\begin{center}
\mbox{\epsfig{file=tau_snu.eps,height=0.35\textheight}}
\end{center}
\mycaption{\label{tau_snu} 95\% confidence level upper limits, obtained from
the ditau cross sections, on $\sqrt{|\lambda_{121}\lambda_{233}|}$
versus the assumed $\tilde\nu_\mu$ mass.}
\end{figure}
\subsection{Limits on Leptoquarks and R-Parity Violating Squarks}
\label{leptoquarks}
At LEP, the $t$~channel exchange of a leptoquark can modify the \qq\ cross section
and jet charge asymmetry, as described by the Born level equations given in
Ref.~\cite{KAL}. A comparison of the measurements with these equations
allows upper limits to be set on the leptoquark's couplings $g$ as a function
of its mass \MLQ, using the same fit technique and corrections for ISR, etc.
as employed in Section~\ref{contacts}.
Although the leptoquark $t$~channel exchange alters the angular distribution
of the outgoing \qq\ system, this has negligible effect on the \qq\ selection
efficiency.
Limits are obtained for each possible leptoquark species.
The allowed species can be classified according to their spin, weak isospin
$I$ and hypercharge. Scalar and vector leptoquarks are denoted by symbols
${\mathrm S}_I$ and ${\mathrm V}_I$ respectively, and isomultiplets with
different hypercharges are distinguished by a tilde.
An indication ``(L)'' or ``(R)'' after the
name indicates if the leptoquark couples to left- or right-handed leptons.
Where both chirality couplings are possible, limits are set for the two
cases independently, assuming that left- and right-handed couplings are not
present simultaneously. It is also assumed that leptoquarks within a given
isomultiplet are mass degenerate \cite{KAL}.
The \SBL{\half} and \SL{0} leptoquarks are equivalent to up-type anti-squarks and
down-type squarks, respectively, in supersymmetric theories with an R-parity
breaking term $\lambda'_{1jk} {\mathrm L}_1 {\mathrm Q}_j \bar{\mathrm D}_k$
$(j,k=1,2,3)$. Limits in terms of the leptoquark coupling are then
exactly equivalent to limits in terms of $\lambda'_{1jk}$.
Table~\ref{lqtab} gives for each leptoquark type, the 95\% confidence level
lower limits on its mass \MLQ, assuming that it has a coupling
strength equal to the electromagnetic coupling $g = e$. The limits
are given separately, assuming that (i) the leptoquark couples to only
$1^{\mathrm st}$ or only $2^{\mathrm nd}$ generation quarks, or (ii)
to only $3^{\mathrm rd}$ generation quarks. The former limits are derived
using the measured hadronic cross section and jet charge asymmetries, whilst
the latter uses the \bb\ cross section and jet charge asymmetries.
For $g\not= e$, the mass limit scales approximately in proportion to the
coupling if it exceeds about 200~\GeVcc. (This is the contact term limit.)
\begin{table}[htbp]
\mycaption{\label{lqtab}
95\% confidence level lower limits on the leptoquark mass for each species.
Limits are given separately according to the quark generation to which the
leptoquark is assumed to couple. A dash indicates that no limit can be
set and ``N.A.'' denotes leptoquarks coupling only to top quarks and hence
not visible at LEP.}
\begin{center}
\begin{tabular}{|c||c c c c c c c|}
\hline\protect\rule{0ex}{1.4em}
Quark & \multicolumn{7}{c|}{Limit on scalar leptoquark mass (\GeVcc)} \\
\cline{2-8}\protect\rule[-0.6em]{0ex}{1.9em}
Generation & \SL{0} & \SR{0} & \SBR{0} & \SL{1} & \SR{\half} & \SL{\half} & \SBL{\half} \\
\hline\hline
\protect\rule{0ex}{1.4em} $1^{\mathrm st}$ or $2^{\mathrm nd}$
& 200 & --- & 70 & 240 & --- & 20 & --- \\
\protect\rule{0ex}{1.4em} $3^{\mathrm rd}$
& N.A. & N.A. & 180 & 450 & --- & N.A. & 50 \\
\hline
\multicolumn{8}{c}{\null}\\
\hline\protect\rule{0ex}{1.4em}
Quark & \multicolumn{7}{c|}{Limit on vector leptoquark mass (\GeVcc)} \\
\cline{2-8}\protect\rule[-0.6em]{0ex}{1.9em}
Generation & \VR{\half} & \VL{\half} & \VBL{\half} & \VL{0} & \VR{0} & \VBR{0} & \VL{1} \\
\hline\hline
\protect\rule{0ex}{1.4em} $1^{\mathrm st}$ or $2^{\mathrm nd}$
& 150 & 130 & 90 & 340 & 120 & 280 & 470 \\
\protect\rule{0ex}{1.4em} $3^{\mathrm rd}$
& 260 & 160 & N.A. & 400 & 140 & N.A. & 400 \\
\hline
\end{tabular}
\end{center}
\end{table}
Similar limits have been obtained by OPAL and L3~\cite{OPAL, L3:CI:1998}.
Limits from the Tevatron \cite{TEVATRON, D0VEC} depend upon the assumed
branching ratio of the charged leptonic decay mode. However, if this is 100\%,
then the Tevatron excludes leptoquark masses below
$\sim$ 225~\GeVcc. Other experiments can place limits on leptoquarks
which couple to first generation quarks. In particular, low energy data such
as atomic parity violation and rare decays give very stringent limits, usually
as a function of the ratio $g/M_{\mathrm LQ}$. If $g=e$, they imply a
lower limit on the leptoquark mass in the range 430--1500~\GeVcc\
\cite{LQlow}, depending on the leptoquark species. Preliminary results from
HERA~\cite{HERA} exclude scalar leptoquarks with masses below $\approx 250$~\GeVcc\ if
$g = e$.
\subsection{Limits on Extra Z Bosons}
\label{Zprime}
To unify the strong and electroweak interactions, Grand Unification Theories
(GUT) extend the SM gauge group to a group of higher rank, predicting
therefore the presence of at least one extra neutral gauge boson \ZP.
The theories which are considered in this section are the \ES\ \cite{ZPRIME}
and the Left-Right (LR) models \cite{ZPRIME}.
In the \ES\ model, the unification group \ES\ can break into the SM
$SU_{C}(3) \otimes SU_{L}(2) \otimes U_{Y}(1)$ in different ways. Each
symmetry breaking pattern leads to the presence of at least one extra U(1)
symmetry and therefore one extra gauge boson. This is characterized by a
parameter \TES\ ($-\pi/2 \leq$ \TES $\leq \pi/2$) which entirely defines its
couplings to conventional fermions. Contributions from exotic particles
predicted by \ES\ or supersymmetric particles are ignored in
this analysis. Four models derived from \ES\ are studied here, the \ESCHI,
\ESPSI, \ESETA\ and \ESI\ defined by \TES $= 0$, $\pi/2$,
$-\arctan{\sqrt{5/3}}$ and $\arctan{\sqrt{3/5}}$, respectively
\cite{ZPRIME}.
In the LR model, the SM group is extended to $SU_{C}(3) \otimes SU_{L}(2)
\otimes SU_{R}(2) \otimes U_{B-L}(1)$, where B and L are the baryon and lepton
number, respectively. The SM $U_{Y}(1)$ symmetry is recovered by a linear
combination of the generator of $U_{B-L}(1)$ and the third component of
$SU_{R}(2)$. This model, which can arise from symmetry breaking of the GUT
group SO(10), leads to the presence of one extra neutral gauge boson.
Contributions from the extra ${\mathrm W}_{R}$ of $SU(2)_{R}$ are neglected in this
analysis. Couplings to conventional fermions depend only on one parameter
\ALR, where $\sqrt{2/3} \leq \alpha_{LR} \leq
\sqrt{(1-2\sin^2{\theta}_W)/\sin^2{\theta}_W}$, which is function of the
coupling constants $g_{L,R}$ of $SU_{L,R}(2)$ and $\sin^2{\theta}_W$.
Only the Left-Right Symmetric Model (LRS) will be studied here, which
has $g_{L}=g_{R}$, implying that if $\sin^2{\theta}_W=0.23$ then
$\ALR=1.53$~.
Outside the context of GUT, a limit is derived on the mass of the
non-gauge-invariant sequential SM (SSM) \ZP~boson, having the same
couplings as the SM Z, but with a higher mass. Limits are also placed
upon the axial and vector couplings of an arbitrary \ZP\ as a function of
its mass.
In all the models mentioned above, the symmetry eigenstate \ZZEROP\ of the
extra U(1) or $SU_{R}(2)$ can mix with the symmetry eigenstate \ZZERO\ of
$SU_{L}(2) \otimes U_{Y}(1)$ with a mixing angle \TMIX. In such a case, the
Z resonance observed at LEP I must be identified as one of the mass
eigenstates of the \ZZEROP--\ZZERO\ system while the second mass eigenstate
\ZP\ of mass \MZP\ is a free parameter \cite{ZPRIMEMIX, ZEFIT}.
To obtain the 95\% confidence level exclusion limits on the various free
parameters, least squares fits are performed using the set of ALEPH
measurements given in Table~\ref{tab:observ}, taking into account the
correlations between them. The LEP1 measurements are taken from
Ref.~\cite{EWLEP1}, whilst the LEP2 measurements are presented in
Tables~\ref{CROSSALL}, \ref{rbcross} and \ref{ASYM_ALL}.
\begin{table}[htbp]
\mycaption{\label{tab:observ} Set of observables used in the \ZP\ analyses and
the corresponding SM \CHI\ values per degree of freedom.}
\begin{center}
\begin{tabular}{|c||c|c|}
\hline
& Observables & \CHISM/NDF. \\
\hline\hline
LEP1 &
$\sigma^{\mathrm l^+l^-}$, $A^{\mathrm l^+l^-}_{FB}$, $\sigma^{{\mathrm q} \bar{\mathrm q}}$, $l=\mu, \tau$ &
130.9/120 \\
\hline
LEP2 &
$\sigma^{\mathrm l^+l^-}$, $A^{\mathrm l^+l^-}_{FB}$, $\sigma^{{\mathrm q} \bar{\mathrm q}}$, \Rb &
19.4/29 \\
130--183 GeV &
$l=\mu, \tau$, $\sqrt{s'/s} > 0.9$, $|{\cos\theta}| < 0.95$ & \\
\hline
\end{tabular}
\end{center}
\end{table}
Theoretical predictions for difermion cross sections and asymmetries are
obtained from the program ZEFIT 5.0 \cite{ZEFIT}, which is an extension of
ZFITTER 5.0 \cite{ZFITTER} including the one extra neutral gauge boson of
the \ES, LR or SSM models. Theoretical uncertainties on the SM
predictions are taken into account as described in Section~\ref{smpred}.
For all models, the minimum \CHI\ is found to occur when no \ZP\ boson is
present.
For the five models \ESCHI, \ESPSI, \ESETA, \ESI\ and \LRS,
Fig.~\ref{limits2p} shows the 95\% confidence level limits obtained in the
plane of \ZP\ mass versus mixing angle \TMIX. Both parameters are treated
as independent, so these limits correspond to a \CHI\ increase of 5.99~.
The LEP1 data mainly constrain the mixing angle, whilst the LEP2 data
mainly constrain the \ZP\ mass at small mixing angles.
Alternatively, assuming $\TMIX=0$, lower limits on the \ZP\ mass can be
obtained using a one-sided, one-parameter fit ($\Delta\chi^2=2.71$). The
resulting limits are given in Table~\ref{tab:limits1D}, where they are
compared with those from direct \ZP\ searches performed by the CDF
Collaboration \cite{ZPCDF}. This table also gives the mass limit for the
SSM \ZP, which is superior to the limit from CDF.
\begin{table}[htbp]
\mycaption{\label{tab:limits1D} Comparison of 95\% confidence level lower
limits on \MZP\ (\GeVcc) from one parameter electroweak fits (ALEPH) and
direct searches (CDF) for $\TMIX=0$.}
\begin{center}
\begin{tabular}{|c||c|c|}
\hline
Model & ALEPH & CDF direct \\
\hline\hline
\ESCHI\ & 533 & 595 \\
\ESPSI\ & 294 & 590 \\
\ESETA\ & 329 & 620 \\
\ESI\ & 472 & 565 \\
\hline
\LRS\ & 436 & 630 \\
\hline
Sequential SM & 898 & 690 \\
\hline
\end{tabular}
\end{center}
\end{table}
\begin{figure}[p]
\mbox{\epsfig{file= new_col_mix_lim2.eps,width=0.9\textwidth}}
\mycaption{\label{limits2p} 95\% confidence level limits in the \MZP\ vs.
\TMIX\ plane for the \ESCHI, \ESPSI, \ESETA, \ESI\ and \LRS\
models. The shaded regions are excluded.}
\end{figure}
Limits can also be placed on the vector and axial couplings of an arbitrary
\ZP, as a function of its mass. To simplify, such limits will only be given
here for the leptonic couplings (assuming lepton universality) and also only
for the case $\TMIX=0$. Limits are placed on the two couplings simultaneously
($\Delta\chi^2=5.99$). The excluded region is found to be approximately
rectangular in shape and its size is given as a function of \ZP\
mass in Table~\ref{tab:limitsCoupl}.
\begin{table}[htbp]
\mycaption{ \label{tab:limitsCoupl}
95\% confidence level limits on the axial $g_a^\prime$ and vector
$g_v^\prime$ couplings of a \ZP\ boson of mass $m_{\mathrm Z'}$ to a
lepton pair.}
\begin{center}
\begin{tabular}{|c|c|c|}
\hline
$m_{\mathrm Z'}$ (\GeVcc) & $|g_a^\prime|$ & $|g_v^\prime|$ \\
\hline
300 & $\leq 0.36$ & $\leq 0.28$ \\
600 & $\leq 0.81$ & $\leq 0.64$ \\
1000 & $\leq 1.39$ & $\leq 1.11$ \\
\hline
\end{tabular}
\end{center}
\end{table}
\section{Conclusions}
Measurements of the hadronic and leptonic
cross sections and asymmetries at $\sqrt{s}=\mbox{130--183}$~GeV have been
presented.
The ratios of the \bb\ to \qq\ production cross
sections at $\sqrt{s}=130$--183~GeV and of the \cc\ to \qq\ production cross
sections at $\sqrt{s}=183$~GeV have been shown, as well as jet charge
asymmetries.
The results agree with the predictions of the Standard Model and allow
limits to be placed on four-fermion contact interactions, R-parity violating
sneutrinos and squarks, leptoquarks and \ZP\ bosons.
The limits on the energy scale $\Lambda$ of $\ee\ff$ contact interactions
are typically in the range from 2--10~TeV. Those for $\ee\myll$ and $\ee\bb$
interactions are of particular interest, since they are inaccessible at
colliders using proton beams. The new ALEPH limits on R-parity violating
sneutrinos reach masses of a few hundred \GeVcc\ for large values of
their Yukawa couplings.
\section*{Acknowledgements}
We thank our colleagues from the CERN accelerator divisions for the successful
operation of LEP at higher energies. We are indebted to the engineers and
technicians in all our institutions for their contribution to the continuing
good performance of ALEPH. Those of us from non-member states thank
CERN for its hospitality.
|
\section{Introduction}
\indent
Superconducting contacts to semiconductors can be used as a high
resolution spectroscopy tool to understand the mechanism of ohmic
contacts between metals and semiconductors. The subgap conductance of
a normal metal - superconductor (NS) interface is quite sensitive to the
presence of any insulating barriers, varying with the square of the
barrier transmission $T$, rather than proportional to $T$ as in normal
metal contacts. Also, any tunnel barriers spatially separated from the
superconducting contacts give rise to pronounced conductance
resonances. The Blonder-Tinkham-Klapwijk (BTK) formula~\cite{btk}
predicts the differential conductance of different types of NS
contacts~\cite{riedel,chaudhuri} shown in Fig.~\ref{fig:pb01}.
We wish to use the insights from Fig.~\ref{fig:pb01} to better
understand both the superconducting properties and the ohmic contact
mechanism of superconductors and metals to LTG-GaAs and InAs. This
paper compares the electrical characteristics between a composite
In-Nb superconducting contact formed to InAs and to LTG-GaAs. We
observed clear signs of ballistic transport in many of the InAs
samples, but not for the LTG-GaAs samples. However, we did observe
tranmission resonances in the LTG-GaAs samples indicative of a band of
conducting electronic states inside the energy gap of the LTG-GaAs.
\begin{figure}
\centps{pb01.eps}{60}
\caption{Differential conductance for (a) a ballistic NS interface (b)
an NIS Giaever tunnleing contact, and (c) an NINS interface displaying
the McMillan-Rowell resonances. Solid lines on the left indicate the pairing
potential and the grey arrow an insulating (tunnel) barrier.}
\label{fig:pb01}
\end{figure}
Many groups have previously studied NS junctions using GaAs as the
semiconductors~\cite{mwilliams}-\cite{cbagwell}. The main advantages
of GaAs as the semiconductor is the ease with which one can control
the geometry of the electron gas using Schottky gates and its high
electron mobility. The disadvantage of GaAs is that most metals,
including superconductors, form a Schottky contact. The Schottky
barrier eliminates any possibility of ballistic transport through the
NS interface. Low temperature grown (LTG)-GaAs has previously been
investigated because of its ability to make low resistance ohmic
contacts to semiconductor devices.~\cite{melloch} We therefore
reasoned that the tunnel barrier formed at the interface between
LTG-GaAs and a superconductor might be low enough to form a reasonably
high transmission interface.
The energy band diagram of the superconductor - (LTG) GaAs contact,
along with the differential conductance one expects from the BTK
formula, is shown in Fig.~\ref{fig:pb02}. The subgap resonances in
differential conductance, shown on the right of Fig.~\ref{fig:pb02},
are McMillan-Rowell type NINS resonances. Fig.~\ref{fig:pb02} assumes
there is essentially no tunnel barrier between the In-Nb contact and
the LTG-GaAs. That is, the superconductor to LTG-GaAs contact forms a
nearly perfect NS interface. However, there is still a tunnel barrier
which the electons must traverse to enter the GaAs substrate, formed
by the ordinary high-temperature grown GaAs. Therefore, placing a
superconductor on LTG-GaAs forms an NINS junction. If the interface
between the superconductor and LTG-GaAs were not ballistic, one would
simply expect Giaever tunnelling in the differential conductance. Many
such NIS or `super-Schottky' junctions have previously been
experimentally measured in superconductor-GaAs contacts.
\begin{figure}
\centps{pb02.eps}{60}
\caption{Energy band diagram for a
superconductor (In-Nb) to LTG GaAs contact. The band of conducting
states arise from excess As incorporation, traps electrons in the GaAs
between the superconductor and GaAs tunnel barrier. Expected
differential conductance of the sample, including these subgap Andreev
resonances, is shown on the right.}
\label{fig:pb02}
\end{figure}
LTG-GaAs is unique in that it contains a large number of point defects
due to excess As incorporation during growth. The point defects
provide an additional transmissive energy band near the middle of the
semiconductor energy gap, which greatly reduces the barrier between the
metal and the GaAs material~\cite{feenstra}. In addition to the band
of conducting states in the LTG-GaAs, using an LTG-GaAs layer enables
us to achieve effective surface doping 10$^{20}$/cm$^{3}$ rather than
the limit 10$^{18}$/cm$^{3}$ in bulk GaAs.~\cite{woodall} This two
orders of magnitide increase in the surface doping greatly reduces the
Schottky barrier width between the metal and GaAs, permitting the
development of low resistance ohmic contacts to GaAs not attainable
using other methods.
The negative Schottky barrier formed at most metal interfaces with
InAs, on the other hand, indicates that it is possible to make
ballistic NS interfaces to InAs. The surface of InAs accumulates
electrons, forming a natural conduction channel. The surface
accumulation property of InAs is well known, and accounts for the
large number of previous experiments using superconductor-InAs
contacts~\cite{kroemer}-\cite{mhartog}. The energy band diagram of
the superconductor-InAs contact, along with the differential
conductance one expects from the BTK formula, is shown in
Fig.~\ref{fig:pb03}.
\begin{figure}
\centps{pb03.eps}{60}
\caption{Energy band diagram for a superconductor (In-Nb) to InAs
contact. The negative Schottky barrier is shown as a triangular
potential well near the surface. Expected differential conductance of
the sample, including tunneling through the thin Nb and above barrier
resonances, is shown on the right.}
\label{fig:pb03}
\end{figure}
\section{Experimental Results}
\indent
The data below show an interplay between the thin Nb portion of the
superconducting contact and the thicker In superconductor. The Nb
contacts to both InAs and LTG-GaAs semiconductors in this study are
1000 angstroms thick, comparable to the Cooper pair size in the
Nb. Andreev reflections from the superconducting contact Nb alone will
therefore not be perfect, even if the NS interface is ballistic. Only
when the temperature is also lowered below the critical temperature of
In (3.4 K) will there be nearly 100 \% Andreev reflection inside the In
energy gap. Andreev reflection will still be imperfect in the energy
range between the In and Nb gaps. We did not intentionally deposit In
in the growth chamber, using instead the bonding wires to the sample
to form that portion of the superconducting contact.
\subsection{Superconductor to LTG GaAs}
The measured differential conductance from two different
In-Nb/LTG-GaAs samples is shown in
Figs.~\ref{fig:ltg5}-\ref{fig:ltg6}. In both samples we observe
multiple subgap peaks corresponding to the McMillan-Rowell resonances.
The subgap resonances are much clearer in Fig.~\ref{fig:ltg5}, though
they are also present in Fig.~\ref{fig:ltg6}. One can even
distinguish the two different energy gaps of In and Nb by the two
different heights of the conductance resonances in
Fig.~\ref{fig:ltg5}. The larger peaks near zero bias correspond to the
thick In layer, while the weaker peaks above the energy gap of In
correspond to weaker Andreev reflection from the thin Nb
superconductor (in addition to some Andreev reflection outside the In
energy gap).
The McMillan-Rowell resonances in Fig.~\ref{fig:ltg6} are not as well
developed as the ones in Fig.~\ref{fig:ltg5}. Sample 2 may have an
irregular contact geometry, with interface roughness broadening the
Andreev resonances. Sample 2 may also consist of a series of more
closely spaced conductance resonances which are not resolved at the
base temperature of T=1.6K. Both samples we believe are NINS
junctions, with sample 2 being a lower quality (broadened) version of
sample 1. Note that the Nb critical temperature is not 10K in these
samples, due to the compromises necessary to deposit Nb on the
semiconductor structure.
Both LTG - GaAs samples were exposed to air prior to depositing Nb. To
form ballistic Nb - LTG GaAs interfaces we relied on the well known
resistance of LTG GaAs surfaces to oxidation. The appearance of
Andreev resonances in both samples indicates a low degree of surface
oxidation. It is remarkable that these samples show little indication
of surface oxidation, even after exposure to air. The differences
between these two nominally identical samples also shows the
sensitivity of differential conductance spectroscopy using
superconducting contacts. Several additional samples were measured,
giving similar results to those shown in
Figs.~\ref{fig:ltg5}-\ref{fig:ltg6}.
\begin{figure}
\centps{figltg5.eps}{60}
\caption{Clear McMillan-Rowell subgap resonances in LTG-GaAs `Sample
1' confirm the presence of an NINS junction. Therefore only a small
(or no) tunnel barrier is present at the superconductor - LTG-GaAs
interface.}
\label{fig:ltg5}
\end{figure}
\begin{figure}
\centps{figltg6.eps}{60}
\caption{`Sample 2' is a superconductor - LTG-GaAs junction prepared
identially to `Sample 1'. The subgap resonances are weaker and much
broader, with an additional large drop in the differential
conductance near 6.5 meV. Both features suggest an inhomogeneous
contact geometry in this sample 2.}
\label{fig:ltg6}
\end{figure}
\subsection{Superconductor to InAs}
Fig.~\ref{fig:inasbg} shows the differential conductance
characteristics of two nominally identical In-Nb to InAs
junctions. `Sample 3' (top) shows an enhancement of conductance around
zero voltage bias at the base temperature (1.6 K). In the BTK
model~\cite{btk}, such an enhanced conductance near zero bias is
associated with near ballistic transport of Cooper pairs through the
normal metal (InAs) / superconductor (Nb) interface. We see the zero
bias peak develop only as the In becomes superconducting, since the Nb
layer is thin compared to the size of a Cooper pair. `Sample 4'
(bottom) displays Giaever tunneling. One can clearly see the In gap
developing between 5.6 and 1.6 K in `Sample 4'. The Giaever tunneling
peaks due to the Nb remain relatively unaffected as the temperature
varies. The differential conductance of `Sample 4' does not go to zero
inside the gap, since the interface transmission of this tunnel
barrier is of order $T \simeq 0.1$, as opposed to $T = 10^{-6}$ in
typical NIS tunnel junctions.
To avoid the formation of interface oxides before Nb deposition, we
moved the wafer in-situ (under high vacuum) after InAs growth to a Nb
sputtering chamber. We did no addition surface cleaning, such as
striking a plasma, prior to Nb deposition. The results in
Fig.~\ref{fig:inasbg} indicate this procedure is only partially
successful, since there is some variance in interface transmission
from one sample to the next. We measured several additional samples,
with differential conductance results similar to those in
Fig.~\ref{fig:inasbg}.
\begin{figure}
\centps{figinasbg.eps}{60}
\caption{Two identically prepared superconductor - InAs junctions.
`Sample 3' (top) exhibits ballistic transport of Cooper pairs across
the interface to the semiconductor as the In becomes superconducting.
`Sample 4' (bottom) displays a modified Giaever tunneling in which
one can also clearly see the development of the In gap.}
\label{fig:inasbg}
\end{figure}
\subsection{Sample Geometry and Series Resistance}
A few caveats are necessary when attempting to extract detailed
information about the energy gaps of the Nb and In from the measured
data. The actual semiconductor samples are simply two metal Nb pads
deposited on the semiconductor, together with their In bonding wires.
Since the pad separation is 10 microns, the actual sample geometry is
two large NS junctions in series (back to back). The energy gaps one
infers from Figs.~\ref{fig:ltg5}-\ref{fig:inasbg} are larger than
those of In and Nb due to the series resistance of the semiconductor
connecting the two NS junctions. Series resistance is significant in
Figs.~\ref{fig:ltg5}-\ref{fig:inasbg}, since the NS junctions are low
resistance, rather than high resistance (NIS) tunnel junctions. The
actual sample geometry and sample preparation (growth) is described in
detail elsewhere.~\cite{rizkms}
Series resistance stretches the voltage axis (makes the energy gaps
and peak widths appear larger) and compresses the differential
conductance (reduces relative heights of the peaks and
valleys). Measurements of series resistance can be made using a
transmission line structure, but we did not perform such
measurements. We therefore cannot make quantitative comparisons of the
data with a BTK type conductance calculation. We can, however, make
qualitative comparisons of theory and experiment as done in the next
section.
\section{Simulation}
\indent
We simulate the differential conductance $dI/dV$ at zero temperature
using the BTK formula
\begin{equation}
\frac{dI}{dV} = \frac{2e}{h} \left[ 1 - R_e(E) + R_h(E) \right] dE .
\label{fiv}
\end{equation}
Here $R_{e}(E)$ is normal reflection probability and $R_{h}(E)$ is the
Andreev reflection probability. In this paper we wish to model
electron transport through the pairing potential
\begin{eqnarray}
\Delta(x) & = &
\left\{
\matrix{
0 & x < 0 \cr
\Delta_{Nb} & 0<x<W \cr
\Delta_{In} & W<x
}
\right.
\label{pairpot}
\end{eqnarray}
The ordinary electrostatic potential we take as an impulse function
located a distance $L$ away from the Nb, namely
\begin{equation}
V(x) = V_0 \delta(x+L) .
\end{equation}
This combination of pairing and electrostatic potentials forms an of
NINS junction. We can therefore use the reflection amplitudes $r_e$
and $r_h$ calculated in Ref.~\cite{riedel}.
The only difference between the present calculation and that of
Ref.~\cite{riedel} is the form of the pairing potential in
the superconducting contact. We can modify calculation of
Ref.~\cite{riedel} to account for the composite Nb-In contact by the
following scheme: Since the quantity $(v_{0}$/$u_{0}) \exp{-i \phi}$
in Eqs.~(A22)-(A26) of Ref.~\cite{riedel} corresponds to the Andreev
reflection probability of an electron from the NS interface, we simply
replace it by the Andreev reflection probability $r_{a,e}$ from our
new N-S'S interface. The new reflection amplitudes are therefore
\begin{equation}
r_e = \frac{1}{d} \left( \frac{-iZ}{1+iZ} \right)
\left[ 1 - \left( r_{a,e} r_{a,h} \right) e^{2i(k_+ - k_-)L} \right] \; ,
\label{re}
\end{equation}
\begin{equation}
r_h = \frac{1}{d} \left( r_{a,e} \right)
\left( \frac{1}{1+Z^2} \right) e^{i(k_+ - k_-)L} \; ,
\label{rh}
\end{equation}
\begin{equation}
d = 1 - \left( \frac{Z^2}{1+Z^2} \right)
\left( r_{a,e} r_{a,h} \right)
e^{2i(k_+ - k_-)L} \; .
\label{d}
\end{equation}
We then separately calculate the new Andreev reflection probability
$r_{a,e}$ from the composite Nb-In pairing potential step. The
Andreev reflection amplitude of an electron from the pairing potential
in Eq.~(\ref{pairpot}) we find to be
\begin{equation}
e^{i \phi} r_{a,e}
= \frac{v_1}{u_1} +
\left(1 - \frac{v_1^2}{u_1^2} \right) r_{\rm step}
\left[1 + \left( \frac{v_1}{u_1} \right) r_{\rm step}
\right]^{-1} ,
\end{equation}
where
\begin{equation}
r_{\rm step} =
\left( \frac{v_2 u_1 - u_2 v_1}{u_2 u_1 - v_2 v_1} \right)
\exp[ i (k_{e1}-k_{h1} ) W ] .
\label{rstep}
\end{equation}
The Andreev reflection probability for holes we find as $e^{i \phi}
r_{a,e} = e^{-i \phi} r_{a,h}$. The particle current reflection
probabilities are then $R_{e}(E) = |r_e|^2$ and $R_{h}(E) = |r_h|^2$.
Plots of the differential conductance from Eq.~(\ref{fiv}), using the
Andreev reflection probabilities from Eqs.~(\ref{re})-(\ref{rstep}),
are shown in
Figs.~\ref{fig:didv2}-\ref{fig:didv1}. Fig.~\ref{fig:didv2} models the
LTG-GaAs junction, while Fig.~\ref{fig:didv1} simulates the InAs
junction. Solid lines give then conductance when the In is
superconducting, while dashed lines similate a normal In contact. We
have not included thermal broadening in
Figs.~\ref{fig:didv2}-\ref{fig:didv1}.
\begin{figure}
\centps{didv1-2.V2.eps}{60}
\caption{Numerical calculation of differential conductance
corresponding to the In-Nb to LTG-GaAs junction. Strength of the
McMillan-Rowell resonances inside the In gap increase as the In
becomes superconducting. Solid lines give the differential conductance
when the In becomes superconducting.}
\label{fig:didv2}
\end{figure}
Fig.~\ref{fig:didv2} reproduces most of the qualitative features of
the differential conductance taken on the LTG-GaAs semiconductor.
McMillan-Rowell type resonances occur inside the energy gap of both
superconductors, but those inside the In gap become much stronger when
the In goes superconducting. It is interesting that the height of some
resonance peaks outisde the In gap actually decrease (in this
simulation) when the In becomes superconducting. We did not clearly
observe this in the experiment. The calculation also shows weaker
above barrier resonances not observed in experiment. (In
Fig.~\ref{fig:didv2} we have chosen the Nb layer thickness $(W=d_1)$
equal to the coherence length of the In $(\xi_2)$, even though the Nb
is slightly thinner in the actual experiment. We have also arbitrarily
set the spacing between the tunnel barrier to the Nb interface $L =
\xi_2$.)
\begin{figure}
\centps{didv-2.V2.eps}{60}
\caption{Numerical calculation of differential conductance corresponding
to the In-Nb to InAs junction. Effect of the In becoming superconducting
can be seen both in the ballistic junction (top) and tunnel junction
(bottom). Solid lines give the differential conductance
when the In becomes superconducting.}
\label{fig:didv1}
\end{figure}
The simulation in Fig.~\ref{fig:didv1} also confirms the qualitative
features we observed in the differential conductance of the InAs
semiconductor. The ballistic junction (top) corresponds to $Z=0$ in
Fig.~\ref{fig:didv1}, while $Z=1$ corresponds to a tunnel junction
(bottom) with barrier transmission $1/2$. The transmission
coefficient of the junction in its normal state is $T = 1/(1+Z^2)$.
A large peak in the differential conductance near zero bias
appears in the ballistic junction when the In becomes superconducting.
The `envelope' of Andreev reflections also decreases somewhat outside
the energy gap, which we did not observe in experiment, but is consistent
with the simulation in Fig.~\ref{fig:didv2}. The two different energy
gaps of In and Nb are also apparent in the tunnel junction in
Fig.~\ref{fig:didv1} (bottom).
\section{Conclusions}
\indent
We have utilized differential conductance dI/dV versus voltage V in
superconductor-semiconductor contacts as a very sensitive probe for
the energy dependence of current carrying states in the junction. The
superconducting contact is a composite of thin Nb with thick In,
allowing us to probe with two different energy scales near the contact
Fermi level. Since the Nb thickness is less than the Cooper pair size
in Nb, by itself the Nb forms only a partial Andreev mirror.
Junctions between In-Nb and InAs show ballistic transport at the NS
interface, evidenced by the development of a large peak in the
differential conductance near zero bias when the In becomes
superconducting. Junctions between In-Nb and LTG-GaAs show
McMillan-Rowell (NINS) type resosnances. The resonances become
stronger inside the In energy gap when the In becomes superconducting,
since the thick In now makes an effective Andreev mirror. Formation of
such NINS resonances suggests a band of conducting states inside the
energy gap of LTG-GaAs. Interface roughness, series resistance, and
the actual three-dimensional contact geometry broaden and weaken
features in the differential conductance in comparison with an
idealized one-dimensional scattering theory.
\section{Acknowledgements}
\indent
We wish to acknowledge the financial support from the David and Lucile
Packard Foundation and from The MRSEC of the National Science
Foundation under grant No. DMR-9400415. We Thank Supriyo Datta,
Michael McElfresh, and Richard Riedel for many useful discusions.
$^1$ Present Address: Samsung Corporation, Austin, Texas.
$^2$ Present Address: TRW Corporation, Redondo Beach, CA 90278.
$^3$ Present Address: Yale University, Department of Electrical
Engineering, New Haven, CT 06520.
$^3$ Present address: Dept. of Physics, University of North Florida,
Jacksonville, FL 32224.
|
\section{Introduction}
We establish for quaternions an analog of the trace formula obtained by Connes
in \cite{Co} for a commutative local field $K$. This formula has the form
$\Tr(\widetilde{P_\Lambda}P_\Lambda\,U_f) = 2\log(\Lambda)f(1) + W(f) + o(1)$
(for $\Lambda\to\infty$), where $f$ is a test-function on $K^\times$, $U_f$ is
the operator of multiplicative convolution with $f$, $P_\Lambda$ and
$\widetilde{P_\Lambda}$ are cut-off projections (precise definitions will be
given later), all acting on the Hilbert space of square-integrable functions on
$K$. The contant term $W(f)$ was shown by Connes to be exactly the term arising
in the ``Weil's explicit formulae'' \cite{We1} of number theory.
We have shown in this abelian local case (see \cite{Bu1}, \cite{Bu2}, and the
related papers \cite{Bu3} and \cite{Bu4}) that the Weil term $W(f)$ can be
written as $-H(f)(1)$ for a certain dilaton-invariant operator $H$. We study
this operator in the non-commutative context of quaternions and then derive the
(analog of) Connes's asymptotic formula. The proof would go through (with some
simplifications of course) equally well in the abelian case.
We first give some elementary lemmas of independent interest about self-adjoint
operators. We then study in multiplicative terms the additive Fourier Transform,
and this immediately leads to the definition of certain ``Quaternionic Tate
Gamma'' functions and to the analog of Tate's local functional equations
(\cite{Ta}, \cite{We2}). This is of course very much related to the
generalization to $GL(N)$ of Tate's Thesis in the work \cite{GJ} of
Godement-Jacquet (see \cite{Ja1}, \cite{Ja2} for reviews and references to
further works by other authors), where certain ``$\gamma(s,\pi,\psi)$''
functions, local $L$- and $\epsilon$-factors and associated functional equations
are studied. Also relevant is the classic monograph by Stein and Weiss \cite{St}
on harmonic analysis in euclidean spaces. In this paper we will follow a
completely explicit and accordingly elementary approach.
We introduce the ``conductor operator'' $H= \log(|x|) + \log(|y|)$ and show how
it gives an operator theoretic interpretation to the logarithmic derivatives of
the Gamma functions (which are involved in explicit formulae.) It is then a
simple matter to compute Connes's trace, and to obtain the asymptotic formula
$$\Tr(\widetilde{P_\Lambda}P_\Lambda\,U_f) = 2\log(\Lambda)f(1) - H(f)(1) +
o(1)$$ in a form directly involving our operator $H$. Further work leads to a
``Weil-like'' formulation for the constant term $H(f)(1)$, if so desired.
\section{Closed invariant operators}
It is well known that any bounded operator on $L^2(\RR,dx)$ which commutes with
translations is diagonalized by the additive Fourier transform (see for example
the Stein-Weiss monograph \cite{St}.) We need a generalization which applies to
(possibly) unbounded operators on Hilbert spaces of the form $L^2(G,dx)$ where
$G$ is a topological group. Various powerful statements are easily found in the
standard references on Hilbert spaces, usually in the language of spectral
representations of abelian von Neumann algebras. For lack of a reference
precisely suited to the exact formulation we will need, we provide here some
simple lemmas with their proofs.
\begin{definition} Let $L$ be a Hilbert space. A (possibly unbounded) operator
$M$ on $L$ with domain $D$ is said to commute with the bounded operator $A$ if
$$\forall v\in L: v\in D\Rightarrow \Big(\ A(v)\in D\ \hbox{and}\ M(A(v)) =
A(M(v))\ \Big)$$
\end{definition}
\begin{thm}\label{L1}
Let $L$ be a Hilbert space and $G$ a (not necessarily abelian) group of unitary
operators on $L$. Let $\cal A$ be the von Neumann algebra of bounded operators
commuting with $G$. Let $M$ be a (possibly unbounded) operator on $L$, with
dense domain $D$. If the three following conditions are satisfied\\ (1) $\cal
A$ is abelian\\ (2) $(M,D)$ is symmetric\\ (3) $(M,D)$ commutes with the
elements of $G$\\ then $(M,D)$ has a unique self-adjoint extension. This
extension commutes with the operators in $\cal A$.
\end{thm}
\begin{proof} We first replace $(M,D)$ by its double-adjoint so that we can
assume that $(M,D)$ is closed (it is easy to check that conditions (2) and (3)
remain valid). The problem is to show that it is self-adjoint. Let $K$ be the
range of the operator $M + i$. It is a closed subspace of $L$ (as $\|(M +
i)(\varphi)\|^2 = \|M(\varphi)\|^2 + \|\varphi\|^2$, and $M$ is closed). Let $R$
be the bounded operator onto $D$ which is orthogonal projection onto $K$
followed with the inverse of $M + i$. One checks easily that $R$ belongs to
$\cal A$, hence commutes with its adjoint $R^*$ which will also belong to $\cal
A$. Any vector $\psi$ in the kernel of $R$ is then in the kernel of $R^*$ (as
$<R^*\psi|R^*\psi>\ =\ <\psi|R\,R^*\psi>\ =\ 0$). So $\psi$ belongs to the
orthogonal complement to the range of $R$, that is $\psi = 0$ as the range of
$R$ is $D$. So $K = L$ and in the same manner $(M - i)(D) = L$. By the basic
criterion for self-adjointness (see \cite{Re}), $M$ is self-adjoint. Let
$A\in\cal A$. It commutes with the resolvant $R$ hence leaves stable its range
$D$. On $D$ one has $RA(M+i) = AR(M+i) = A = R(M+i)A$ hence $A(M+i) = (M+i)A$
so $A$ commutes with $M$.
\end{proof}
For the remainder of this section we let $G$ be a locally compact, Hausdorff,
topological \emph{abelian} group and $\widehat{G}$ its dual group. We refer to
\cite{Ru} for the basics of harmonic analysis on $G$. In particular we have a
Haar measure $dx$ (unique up to a multiplicative constant) and a Hilbert space
$L = L^2(G, dx)$. We also have a dual Haar measure $dy$ on $\widehat{G}$ such
that the Fourier transform $F(\varphi)(y) = \int \varphi(x) \overline{y(x)} dx$
is an isometry of $L$ onto $L^2(\widehat{G}, dy)$. We sometimes identify the
two Hilbert spaces without making explicit the reference to $F$: so when we
write $f(y)\in L$ we really refer to $F^{-1}(f)\in L$ with $f\in
L^2(\widehat{G}, dy)$. No confusion should arise. We will assume that $dy$ is a
$\sigma$-finite measure so that there exists $\psi \in L$ with the property
$\psi(y)\neq 0\ a.e.$\ .
Let $a(y)$ be a measurable function on $\widehat{G}$, not necessarily
bounded. Let $D_a\subset L$ be the domain of square-integrable (equivalence
classes of measurable) functions $\varphi(x)$ on $G$ such that
$a(y)F(\varphi)(y)$ belongs to $L^2(\widehat{G}, dy)$. And let $M_a$ be the
operator with domain $D_a$ acting according to $\varphi\mapsto M_a(\varphi) =
F^{-1}(a\cdot F(\varphi))$. We write $a=b$ if the two functions $a(y)$ and
$b(y)$ are equal almost everywhere on $\widehat{G}$.
\begin{lem} The operator $(M_a, D_a)$ on $L^2(G,dx)$ commutes
with $G$. Furthermore $D_a$ is dense and $(M_a, D_a)$ is a closed operator. If
$(M_b, D_b)$ extends $(M_a,D_a)$, then in fact $a = b$ and $(M_b, D_b) =
(M_a,D_a)$. The adjoint of $(M_a,D_a)$ is $(M_{\overline{a}},D_{\overline{a}})$
(of course $D_{\overline{a}} = D_a$.)
\end{lem}
\begin{proof} We give the proof for completeness. The commutation with
$G$-translations is clear. Then $D_a$ contains (the inverse
Fourier transform of) $\psi(y)\over\sqrt{1 + |a(y)|^2}$ and all its
translates. Hence if $f$ is orthogonal to $D_a$ then the function
$\overline{f(y)}\psi(y)\over\sqrt{1 + |a(y)|^2}$ on $\widehat{G}$ belongs to
$L^1(\widehat{G},dy)$ and has a vanishing ``inverse Fourier transform'', hence
$f = 0$ (almost everywhere). It is also clear using $\psi(y)\over\sqrt{1 +
|a(y)|^2}$ that if $(M_b, D_b)$ extends $(M_a,D_a)$, then $a = b$. Let us
assume that the sequence $\varphi_j$ is such that $\varphi=\mathop{\rm l{.}i{.}m{.}} \varphi_j$
and $\theta=\mathop{\rm l{.}i{.}m{.}} M_a(\varphi_j)$ both exist. Let us pick a pointwise on
$\widehat{G}$ almost everywhere convergent subsequence
$\varphi_{j_k}(y)$. Using Fatou's lemma we deduce that $\varphi$ belongs to
$D_a$. Using Fatou's lemma again we get the vanishing of $\int_{\widehat{G}}
|\theta(y) - a(y) \varphi(y)|^2\,dy$, and this shows that $(M_a, D_a)$ is a
closed operator. Finally let $f(y)$ be in the domain of the adjoint of
$(M_a,D_a)$. There exists then an element $\theta$ of $L$ such that for any
$\varphi \in D_a$ the equality $$\int f(y)\overline{a(y)\varphi(y)}\,dy = \int
\theta(y)\overline{\varphi(y)}\,dy$$ holds. This implies that the two
following functions of $L^1(\widehat{G}, dy)$:
$${f(y)\overline{a(y)\psi(y)}\over\sqrt{1 + |a(y)|^2}}\hbox{\quad
and\quad}{\theta(y)\overline{\psi(y)}\over\sqrt{1 + |a(y)|^2}}$$ have the same
Fourier transform on $G$, hence are equal almost everywhere. So $f \in
D_{\overline{a}}$ and $(M_a)^*(f) = (M_{\overline{a}})(f)$.
\end{proof}
\begin{thm}\label{T1} Let $(M, D)$ be a \emph{closed operator} on $L^2(G,dx)$
commuting with $G$-translations. Then $(M, D) = (M_a, D_a)$ for a (unique)
multiplicator $a$.
\end{thm}
\begin{note} For a bounded $M$ and $G = \RR$, this is proven in the classical
monograph by Stein and Weiss \cite{St}, as a special case of a more general
statement applying in $L^p$ spaces.
\end{note}
\begin{proof}
Let us first assume that $M$ is bounded. We use the (inverse Fourier transform
of the) function $\psi(y)$ and define $a(y)$ to be
${M(\psi)(y)\over\psi(y)}$. Let us consider the domain $D$ consisting of all
finite linear combinations of translates of $\psi$. It is dense by the argument
using unicity of Fourier transform in $L^1$ we have used previously. Then $(M,
D) \subset (M_a, D_a)$, hence $(M_a, D_a)$ is also an extension of the closure
of $(M, D)$. As $M$ is assumed to be bounded this is $(M, L)$. But this means
that $D_a = L$ and that $M = M_a$ (we then note that necessarily $a$ is
essentially bounded).
The next case is when $M$ is assumed to be self-adjoint. Its resolvents $R_1 =
(M - i)^{-1}$ and $R_2 = (M + i)^{-1}$ are bounded and commute with $G$. Hence
they correspond to multiplicators $r_1(y)$ and $r_2(y)$. The kernel of $R_1$ is
orthogonal to the range of $R_2 = R_1^*$ which is all of $D$, so in fact it is
reduced to $\{0\}$. Hence $r_1(y)$ is almost everywhere non-vanishing. Let $f
\in D$ and $g = M(f)$. As $R_1(M(f) - i\,f) = f$ we get $g(y) = {1 +
i\,r_1(y)\over r_1(y)}\cdot f(y)$ and defining $a(y)$ to be ${1 + i\,r_1(y)\over
r_1(y)}$ we see that $(M_a, D_a)$ is an extension of $(M, D)$. Taking the
adjoints we deduce that $(M, D)$ is an extension of
$(M_{\overline{a}},D_{\overline{a}})$. So all three are equal (and $a$ is
real-valued).
For the general case we use the theorem of polar decomposition (see for example
\cite{Re}). There exists a non-negative self-adjoint operator $|M|$ with the
same domain as $M$ and a partial isometry $U$ such that $M = U|M|$. Further
conditions are satisfied which make $|M|$ and $U$ unique: so they also commute
with $G$. It follows from what was proven previously that $(M, D) \subset (M_a,
D_a)$ for an appropriate $a$ (the product of the multiplicators associated to
the self-adjoint $|M|$ and the bounded $U$). The adjoint $(M^*, D^*)$ also has a
dense domain and commutes with $G$, so in the same manner $(M^*, D^*) \subset
(M_b, D_b)$ for an appropriate $b$. The inclusion
$(M_{\overline{a}},D_{\overline{a}}) \subset (M^*, D^*) \subset (M_b, D_b)$
implies $b = \overline{a}$ and $(M_a, D_a) = (M, D)^{**}$. But the
double-adjoint coincides with the closed operator $(M,D)$.
\end{proof}
Let us mention an immediate corollary:
\begin{corollary} A closed symmetric operator on $L^2(G,dx)$ commuting with
$G$ is self-adjoint, and a symmetric operator which has a dense domain and
commutes with $G$ is essentially self-adjoint.
\end{corollary}
\section{Tate's functional equations}
Our first concern will be to introduce numerous notations. Let $\HH$ be the
space of quaternions with $\RR$-basis $\{1, i, j, k\}$ and table of
multiplication $i^2 = j^2 = k^2 = -1,\ ij = k = -ji,\ jk = i = -kj,\ ki = j =
-ik$. A typical quaternion will be denoted $x = x_0 + x_1 i + x_2 j + x_3 k$,
its conjugate $\overline{x} = x_0 - x_1 i - x_2 j - x_3 k$, its real part
$\mathop{\rm Re}} \newcommand{\Tr}{\mathop{\bf Tr}(x) = x_0$, its (reduced) norm $n(x) = x\overline{x} = \overline{x}x =
x_0^2 + x_1^2 + x_2^2 + x_3^2$.
$\HH$ can also be considered as a left $\CC$-vector space with basis $\{1,
j\}$. We then write $a = x_0 + x_1 i$ and $b = x_2 + x_3 i$. Then $jaj^{-1} =
\overline{a}$, $x = a + bj$, and $n(x) = a\overline{a} + b\overline{b}$. The
action of $\HH$ on itself by right-multiplication sends $x = a + bj$ to the
$2\times2$ complex matrix
$$R_x = \pmatrix{a & -\overline{b} \cr b & \overline{a} \cr}$$ We write $V$ for
the complex vector space of complex-linear forms $\alpha: \HH\to\CC$. The forms
$A: x \mapsto a$ and $B: x \mapsto b$ are a basis of $V$. We have a left action
of $\HH$ on $V$ with $x\in\HH$ acting as $\alpha(y) \mapsto \alpha(yx)$. This
left action represents the quaternion $x$ by the matrix
$$L_x = \pmatrix{a & b \cr -\overline{b} & \overline{a} \cr}$$ Also let $V_N =
\hbox{SYM}^N(V)$, for $N = 0, 1, \dots$ be the $N + 1$-dimensional complex
vector space with basis the monomials $A^j B^{N-j}, 0\leq j\leq N$.
Let $G = \HH^\times$ be the multiplicative group (with typical element $g$) and
$G_0 = \{ g \in G |\ n(g) = 1\}$ its maximal compact subgroup. Through the
assignment $g \mapsto L_g$ an isomorphism $G_0 \sim SU(2)$ is obtained, and the
$V_N$'s give the complete list of (isomorphism classes of) irreducible
representations of $G_0$.
The additive Fourier Transform $\cal F$ is taken with respect to the additive
character $x \mapsto \lambda(x) = e^{-2\pi\,i (x + \overline{x})}$. We note that
$\lambda(xy) = \lambda(yx)$. The choice we make for the normalization of $\cal
F$ is:
$${\cal F}(\varphi)(y) = \widetilde{\varphi}(y) = \int
\varphi(x)\lambda(-xy)\,dx$$ where $dx = 4dx_0dx_1dx_2dx_3$ is the unique
self-dual Haar measure for $\lambda$. With these choices the function $\omega(x)
= e^{-2\pi x\overline{x}}$ is its own Fourier transform.
\begin{definition}
The module $|g|$ of $g\in \HH^\times$ is defined by the equality of additive
Haar measures on $\HH$: $d(gx) = d(xg) = |g|dx$. It is expressed in terms of the
reduced norm by $|g| = n(g)^2$.
\end{definition}
\begin{note}
The multiplicative (left- and right-) Haar measures on $G$ are the multiples of
$dg \over |g|$.
\end{note}
One has a direct product $G = (0, \infty) \times G_0$, $g = r g_0$, $r =
\sqrt{n(g)} = |g|^{1/4}$. We write $d\sigma$ for the Euclidean surface element
on $G_0$ (for the coordinates $x_i$), so that $dx = 4r^3\, drd\sigma$. The rule
for integrating functions of $r$ is $\int g(r) dx = 8\pi^2\,\int_0^\infty g(r)
r^3 dr$ as is checked with $\omega(x)$. So $d\sigma = 2\pi^2\,d^*g_0$ where
$d^*g_0$ is the Haar measure on $G_0$ with total mass $1$.
\begin{definition}
The normalized Haar measure on $G$ is defined to be $d^*g = {1 \over 2
\pi^2}{dg \over |g|} = 4 {dr\over r}\,d^*g_0$. It is chosen so that its
push-forward under the module map $g\mapsto u = |g|\in\RR^{\times+}$ is ${du
\over u} = 4 {dr\over r}$.
\end{definition}
The multiplicative group $G$ acts in various unitary ways on $L^2 := L^2(\HH,
dx)$:\\ \centerline{\hfill$L_1(g) : \varphi(x) \mapsto
|g|^{1/2}\varphi(xg)$\hfill $R_1(g) : \varphi(x) \mapsto
|g|^{1/2}\varphi(gx)$\hfill} and also $L_2(g) = R_1(g^{-1})$ and $R_2(g) =
L_1(g^{-1})$.
\begin{definition}
The \emph{Inversion} $I$ is the unitary operator on $L^2(\HH,dx)$ acting as
$\varphi(x) \mapsto {1\over|x|}\varphi({1\over x})$. The \emph{Gamma operator}
is the composite $\Gamma={\cal F}I$.
\end{definition}
\begin{thm}
The Gamma operator commutes with both left actions $L_1$ and $L_2$ and with both
right actions $R_1$ and $R_2$ of $G$ on $L^2$.
\end{thm}
\begin{proof}
One just checks that ${\cal F}$ intertwines $L_1$ with $L_2$, and also $R_1$
with $R_2$ and that the inversion $I$ also intertwines $L_1$ with $L_2$, and
$R_1$ with $R_2$.
\end{proof}
\begin{definition}
The \emph{basic isometry} is the map $\phi(x) \mapsto f(g) = \sqrt{2\pi^2\,|g|}
\; \phi(g)$ between $L^2(\HH,dx)$ and $L^2(G, d^*g)$.
\end{definition}
\begin{note}
It is convenient to avoid using any notation at all for the basic isometry. So
we still denote by ${\cal F}$ the additive Fourier transform transported to the
multiplicative setting. The inversion $I$ becomes $f(g) \mapsto f(g^{-1})$. The
Gamma operator is still denoted $\Gamma$ when viewed as acting on $L^2(G, d^*g)$.
\end{note}
The spectral decomposition of $L^2((0, \infty),{du\over u})$ is standard Fourier
(or Mellin) theory (alternatively we can apply Theorem {\bf\ref{T1}} here): any bounded
operator $M$ commuting with multiplicative translations is given by a measurable
bounded multiplier $a(\tau)$ in dual space $L^2(\RR, {d\tau \over 2\pi})$:
$$G_1(u) = \mathop{\rm l{.}i{.}m{.}}_{\Lambda \rightarrow \infty} \int_{-\Lambda}^{\Lambda}
\psi(\tau) u^{-i\tau} {d\tau \over 2\pi} \Longrightarrow M(G_1)(u) =
\mathop{\rm l{.}i{.}m{.}}_{\Lambda \rightarrow \infty} \int_{-\Lambda}^{\Lambda} a(\tau)\psi(\tau)
u^{-i\tau} {d\tau \over 2\pi}$$
On the other hand the spectral decomposition of $L^2(G_0,\,d^*g_0)$ is part of
the Peter-Weyl theory: it tells us that $L^2(G_0,\,d^*g_0)$ decomposes under
the $L_1 \times R_1$ action by $G_0 \times G_0$ into a countable direct sum
$\oplus_{N\geq0} W_N$ of finite dimensional irreducible, non-isomorphic,
modules. This is also the isotypical decomposition under either $L_1$ alone or
$R_1$ alone (for which $W_N$ then contains $N+1$ copies of $V_N$.) Using the
standard theory of tensor products of separable Hilbert spaces (see for example
\cite{Re}) we have:
\begin{lem} The isotypical
decomposition of $L^2(G,\,d^*g)$ under the compact group $G_0 \times G_0$ acting
through $L_1 \times R_1$ is
$$L^2(G,\,d^*g) = L^2((0, \infty),{du\over u}) \otimes L^2(G_0,\,d^*g_0) =
\oplus_N L^2((0, \infty),{du\over u})\otimes W_N$$
\end{lem}
\begin{lem}\label{L2}
Let $M$ be a bounded operator on $L^2$ which commutes with both the $L_1$ and
$R_1$ actions of $G$. Then to each integer $N\geq 0$ is associated an
(essentially bounded) multiplicator $a_N(\tau)$ on $\RR$, unique up to equality
almost everywhere, such that
$$\psi\in L^2(\RR, {d\tau \over 2\pi}), \ G_1(u) = \mathop{\rm l{.}i{.}m{.}}_{\Lambda \rightarrow
\infty} \int_{-\Lambda}^{\Lambda} \psi(\tau) u^{-i\tau} {d\tau \over 2\pi}$$
$$\Rightarrow \forall F\in W_N\quad M(FG_1) = FG_2$$
$$\hbox{with }G_2(u) = \mathop{\rm l{.}i{.}m{.}}_{\Lambda \rightarrow \infty}
\int_{-\Lambda}^{\Lambda} a_N(\tau)\psi(\tau) u^{-i\tau} {d\tau \over 2\pi}$$
and where $FG_1$ is the function $g\mapsto F({g\over|g|^{1/4}})G_1(|g|)$ and
$FG_2$ the function $g\mapsto F({g\over|g|^{1/4}})G_2(|g|)$.
\end{lem}
\begin{proof}
Let us take $f \in L^2((0, \infty),{du\over u})$ and consider the linear
operator on $L^2(G_0,d^*g_0)$:
$$F(g_0) \mapsto \left(g_0 \mapsto \int_0^\infty \overline{f(u)}M(f \otimes
F)(g_0\,u^{1/4}){du\over u}\right)$$ It commutes with the action of $G_0 \times
G_0$ hence stabilizes each $W_N$ and is a multiple $a_N^f$ of the identity
there. On the other hand, if we choose $F_1$ and $F_2$ in $W_N$ and consider
$$f \mapsto \left(u \mapsto \int_{G_0} \overline{F_2(g_0)} M(f \otimes
F_1)(g_0\,u^{1/4})\,d^*g_0\right)$$ we obtain a bounded operator $M(F_1,F_2)$ on
$L^2((0, \infty),{du\over u})$ commuting with dilations and such that
$$<f |M(F_1,F_2)(f)> = <F_2 | M_N^f(F_1) > = a_N^f <F_2 | F_1>$$ where the
let-hand bracket is computed in $L^2((0, \infty),{du\over u})$ while the next
two are in $L^2(G_0,\,d^*g_0)$. So $M(F_1,F_2)$ depends on $(F_1,F_2)$ only
through $<F_2 | F_1>$. We then let $a_N(\tau)$ be the spectral multiplier
associated to $M(F, F)$ for an arbitrary $F$ satisfying $<F|F> = 1$.
\end{proof}
\begin{corollary}
The von Neumann algebra $\cal A$ of bounded operators commuting simultaneously
with the left and right actions of the multiplicative quaternions on $L^2(\HH,
dx)$ is abelian.
\end{corollary}
\begin{lem}
A self-adjoint operator $M$ commuting with both left and right actions of $G$
commutes with any operator of the von Neumann algebra $\cal A$. In particular it
commutes with $\Gamma$.
\end{lem}
\begin{proof}
One applies Theorem {\bf\ref{L1}}.
\end{proof}
\begin{definition}\label{D1}
The \emph{quaternionic Tate Gamma functions} are the multiplicators
$\gamma_N(\tau)$ ($N\geq 0$) associated to the unitary operator $\Gamma$.
\end{definition}
\begin{note}
This generalizes the Gamma functions of Tate for $K=\RR$ anf $K=\CC$
(\cite{Ta}). In all cases they are indexed by the characters of the maximal
compact subgroup of the multiplicative group $K^\times$.
\end{note}
\begin{lem}
There is a smooth function in the equivalence class of
$\gamma_N(\tau)$.
\end{lem}
\begin{proof}
If the function $G_1(u)$ on $(0, \infty)$ is chosen smooth with compact support
(so that $\psi(\tau)$ is entire) then, for any $F\in W_N$ the function $FG_1$,
viewed in the additive picture, is smooth on $\HH$, has compact support, and
vanishes identically in a neighborhood of the origin. So its image under the
inversion also belongs to the Schwartz class in the additive picture on
$\HH$. Hence $\Gamma(FG_1)$ can be written as $|g|^{1/2} \phi(g)$ for some
Schwartz function $\phi(x)$ of the additive variable $x$. One checks that this
then implies that $G_2(u)$ is a Schwartz function of the variable $\log(u)$ (we
assume that $F$ does not identically vanish of course), hence that
$\gamma_N(\tau) \psi(\tau)$ is a Schwartz function of $\tau$. The various
allowable $\psi$'s have no common zeros so the conclusion follows.
\end{proof}
\begin{note}
From now on $\gamma_N$ refers to this unique smooth representative. It is
everywhere of modulus $1$ as $\Gamma$ is a unitary operator.
\end{note}
\begin{note}
Any function $F \in W_N$ will now be considered as a function on all of $G =
\HH^\times$ after extending it to be constant along each radial line. It is not
defined at $x=0$ of course.
\end{note}
Let $F\in W_N$. For $\mathop{\rm Re}} \newcommand{\Tr}{\mathop{\bf Tr}(s) > 0$, $F(x) |x|^{s-1}$ is a tempered distribution
on $\HH$, hence has a distribution-theoretic Fourier Transform. At first we only
consider $s = {1\over 2} + i\tau$:
\begin{lem}\label{L4}
As distributions on $\HH$
$${\cal F}(F({1\over x})\,|x|^{-{1\over 2} + i\tau}) = \gamma_N(\tau)
F(x)\,|x|^{-{1\over 2} - i\tau}$$
\end{lem}
\begin{proof}
We have to check the identity:
$$\int F({1\over y})\,|y|^{-{1\over 2} + i\tau}\widetilde{\varphi}(y)\,dy =
\gamma_N(\tau)\cdot\int F(x)\,|x|^{-{1\over 2} - i\tau}\varphi(x)\,dx$$ for all
Schwartz functions $\varphi(x)$ with Fourier Transform
$\widetilde{\varphi}(y)$. Both integrals are analytic in $\tau\in\RR$, hence
both sides are smooth (bounded) functions of $\tau$. It will be enough to prove
the identity after integrating against $\psi(\tau)\, {d\tau \over 2\pi}$ with an
arbitrary Schwartz function $\psi(\tau)$. With the notations of Lemma
{\bf\ref{L2}}, we have to check
$$\int F({1\over y})G_1({1\over y})|y|^{- {1\over 2}}\widetilde{\varphi}(y)\,dy
= \int F(x)\,G_2(x)|x|^{-{1\over 2}}\varphi(x)\,dx$$ But, by Lemma
{\bf\ref{L2}}, and by Definition {\bf\ref{D1}}, $F(x)\,G_2(x)|x|^{-{1\over 2}}$
is just the Fourier Transform in $L^2(\HH, dx)$ of $F({1\over y})G_1({1\over
y})|y|^{- {1\over 2}}$, so this reduces to the $L^2$-identity
$$\int \psi(y)\widetilde{\varphi}(y)\,dy = \int
\widetilde{\psi}(x)\varphi(x)\,dx$$
\end{proof}
\begin{thm}\label{T2}
Let $F\in W_N$. There exists an analytic function $\Gamma_N(s)$ in $0 < \mathop{\rm Re}} \newcommand{\Tr}{\mathop{\bf Tr}(s)
<1$ depending only on $N\in \NN$ and such that the following identity of
tempered distributions on $\HH$ holds for each $s$ in the critical strip
($0<\mathop{\rm Re}} \newcommand{\Tr}{\mathop{\bf Tr}(s)<1$):
$${\cal F}(F({1\over x})\,|x|^{s -1}) = \Gamma_N(s) F(x)\,|x|^{-s}$$
\end{thm}
\begin{proof}
We have to check an identity:
$$\int F({1\over y})\,|y|^{s-1}\widetilde{\varphi}(y)\,dy = \Gamma_N(s)\cdot\int
F(x)\,|x|^{-s}\varphi(x)\,dx$$ for all Schwartz functions $\varphi(x)$ with
Fourier Transform $\widetilde{\varphi}(y)$. Both integrals are analytic in the
strip $0 < \mathop{\rm Re}} \newcommand{\Tr}{\mathop{\bf Tr}(s) <1$, their ratio is thus a meromorphic function, which
depends neither on $F$ nor on $\varphi$ as it equals $\gamma_N(\tau)$ on the
critical line. Furthermore for any given $s$ we can choose $\varphi(x) =
\overline{F(x)} \alpha(|x|)$, with $\alpha$ having very small support around
$|x| = 1$ to see that this ratio is in fact analytic.
\end{proof}
\begin{note}
This is the analog for quaternions of Tate's ``local functional equation''
\cite{Ta}, in the distribution theoretic flavor advocated by Weil \cite{We2}. We
followed a different approach than Tate, as his proof does not go through that
easily in the non-commutative case.
\end{note}
Let $\Gamma(s)$ be Euler's Gamma function ($\int_0^\infty e^{-u}u^s\,{du\over
u}$).
\begin{thm}\label{T3} We have for each $N\in\NN$:
$$\Gamma_N(s) = i^N (2\pi)^{2-4s} {\Gamma(2s + {N\over 2})\over\Gamma(2(1-s) +
{N\over 2})}$$
\end{thm}
\begin{proof}
Let $0\leq j \leq N$ and $\omega_j(x) = \overline{A(x)}^{N-j}\overline{B(x)}^j
e^{-2\pi x\overline{x}} = \overline{a}^{N-j}\,\overline{b}^j\,\omega(x)$. One
checks that $\widetilde{\omega_j}(y) =
(-1)^j\,i^N\,{\alpha}^{N-j}\,\overline{\beta}^j \, \omega(y)$ ($y = \alpha +
\beta j$). We choose as homogeneous function $F_j(x) =
a^{N-j}\,b^j\,|x|^{-N/4}$. For these choices the identity of Theorem
{\bf\ref{T2}} becomes
$$i^N \int (\alpha\overline{\alpha})^{N-j}(\beta\overline{\beta})^{j} e^{-2\pi
y\overline{y}} |y|^{s - 1 - N/4} dy = \Gamma_N(s) \int
(a\overline{a})^{N-j}(b\overline{b})^{j} e^{-2\pi x\overline{x}} |x|^{-s - N/4}
dx$$ Adding a suitable linear combinations of these identities for $0\leq j
\leq N$ gives
$$i^N \int (y\overline{y})^{N}\,e^{-2\pi y\overline{y}} |y|^{s - 1 - N/4} dy =
\Gamma_N(s) \int (x\overline{x})^{N}\,e^{-2\pi x\overline{x}} |x|^{-s - N/4}
dx$$ hence the result after evaluating the integrals in terms of $\Gamma(s)$.
\end{proof}
\section{The central operator $H = \log(|x|) + \log(|y|)$}
\begin{definition}
We let $\Delta \subset {\cal C}^\infty(G)$ be the vector space of finite linear
combinations of functions $f(g) = F(g_0)K(\log(|g|))$ with $F$ in one of the
$W_N$'s (hence smooth) and $K$ a Schwartz function on $\RR$. It is a dense
sub-domain of $L^2$.
\end{definition}
\begin{thm} $\Delta$ is stable under ${\cal F}$.\end{thm}
\begin{proof}
We have to show that $\gamma_N(\tau)$ is a multiplier of the Schwartz class. Let
$h_N(\tau) = -i {\gamma_N^\prime(\tau) \over \gamma_N(\tau)}$. Using Theorem {\bf\ref{T3}}
and the partial fraction expansion of the logarithmic derivative of $\Gamma(s)$
(as in \cite{Bu1} for the real and complex Tate Gamma functions), or
Stirling's formula, or any other means, one finds $h_N(\tau) =
O(\log(1+|\tau|))$, $h_N^{(k)}(\tau) = O(1)$, so that $\gamma_N^{(k)}(\tau) =
O(\log(1+|\tau|)^k)$.
\end{proof}
Let $A$ be the operator on $L^2(\HH,dx)$ of multiplication with $\log(|x|)$. As
it is unbounded, we need a domain and we choose it to be $\Delta$. Of course
$(A, \Delta)$ is essentially self-adjoint. It is unitarily equivalent to the
operator $(B, \Delta)$, $B = {\cal F}A{\cal F}^{-1}$. Clearly:
\begin{lem} The domain $\Delta$ is stable under $A$ and $B$.\end{lem}
\begin{definition}
The \emph{conductor operator} is the operator $H = A + B$:
$$H = \log(|x|) + \log(|y|)$$ This is an unbounded operator defined initially
on the domain $\Delta$.
\end{definition}
\begin{lem} The conductor operator $(H,\Delta)$ commutes with the left and with
the right actions of $G$ and is symmetric.
\end{lem}
This is clear. Applying now Theorem {\bf\ref{L1}} we deduce:
\begin{thm} The conductor operator $(H,\Delta)$ has a unique self-adjoint extension.
\end{thm}
We will simply denote by $H$ and call ``conductor operator'' this self-adjoint
extension.
\begin{thm} The conductor operator $H$ commutes with the inversion $I$.
\end{thm}
\begin{proof} Indeed, it commutes with $\Gamma$ by Theorem {\bf\ref{L1}} and it
commutes with $\cal F$ by construction.
\end{proof}
We now want to give a concrete description of its spectral functions.
\begin{definition} Let for each $N\in\NN$ and $\tau\in\RR$:
$$h_N(\tau) = -i {\gamma_N^\prime(\tau) \over \gamma_N(\tau)}$$
$$k_N(\tau) = - h_N^\prime(\tau)$$
\end{definition}
Explicit computations prove that the functions $h_N$ are left-bounded ($\exists
C\,\forall\tau\,\forall N\ h_N(\tau)\geq -C$) and that the functions $k_N$ are
bounded ($\exists C\,\forall\tau\,\forall N\ |k_N(\tau)|\leq C$.) We need not
reproduce these computations here, which use only the partial fraction expansion
of Euler's gamma function, as similar results are provided in \cite{Bu1} in the
real and complex cases.
Let $f(g) = F(g_0)\phi(|g|)$ be an element of $\Delta$, $F \in W_N \subset
L^2(G_0, dg_0)$, $\phi\in L^2((0,\infty), {du\over u})$, $\phi$ being a Schwartz
function of $\log(u)$ ($u = |g|$). We can also consider $f$ to be given as a
pair $\{F, \psi\}$ with $\psi(\tau) = \int_0^\infty \phi(u) u^{i\tau} {du\over
u}$ being a Schwartz function of $\tau$. Then $A(f)$ is given by the pair $\{F,
D(\psi)\}$ where $D$ is the differential operator ${1\over i}{d\over
d\tau}$. This implies that $\Gamma A\Gamma^{-1} (f)$ corresponds to the pair
$\{F, D(\psi) - h_N\cdot \psi\}$. On the other hand $\Gamma A\Gamma^{-1} = - B$
so $H(f)$ corresponds to the pair $\{F, h_N\cdot \psi\}$. The commutation with
the inversion $I$ translates into $h_N(-\tau) = h_N(\tau)$. Also: $K = i[B, A] =
-i[A, H]$ sends the pair $\{F, \psi\}$ to $\{F, k_N\cdot \psi\}$, hence is
bounded and anti-commutes with the Inversion. We have proved:
\begin{thm}\label{T4}
The operator $\log(|x|) + \log(|y|)$ is self-adjoint, left-bounded, commutes
with the left- and right- dilations, commutes with the Inversion, and its
spectral functions are the functions $h_N(\tau)$. The operator $i\,[\log(|y|),
\log(|x|)]$ is bounded, self-adjoint, commutes with the left- and right-
dilations, anti-commutes with the Inversion and its spectral functions are the
functions $k_N(\tau)$.
\end{thm}
We now conclude this chapter with a study of some elementary
distribution-theoretic properties of $H$. For this we need the analytic
functions of $s$ ($0 < \mathop{\rm Re}} \newcommand{\Tr}{\mathop{\bf Tr}(s) < 1$) indexed by $N\in\NN$:
$$H_N(s) = {d\over ds} \log(\Gamma_N(s))$$ (so that $h_N(\tau) = H_N({1\over 2}
+ i\tau)$).
\begin{lem}
Let $\varphi(x)$ be a Schwartz function on $\HH$. Then $H(\varphi)$ is
continuous on $\HH \backslash\{0\}$, is $O(\log(|x|))$ for $x \rightarrow 0$,
and is $O(1/|x|)$ for $|x| \rightarrow\infty$. Furthermore, for any $F \in W_N$
(constant along radial lines), the following identity holds for $0 < \mathop{\rm Re}} \newcommand{\Tr}{\mathop{\bf Tr}(s) <
1$:
$$\int H(\varphi)(x) F(x) |x|^{-s} dx = H_N(s) \int \varphi(x) F(x) |x|^{-s} dx$$
\end{lem}
\begin{proof}
Assuming the validity of the estimates we see that both sides of the identity
are analytic functions of $s$, so it is enough to prove the identity on the
critical line:
$$\int H(\varphi)(x) F(x) |x|^{-{1\over 2} - i\tau} dx = h_N(\tau) \int
\varphi(x) F(x) |x|^{-{1\over 2} - i\tau} dx$$
As in the proof of Lemma {\bf\ref{L4}}, it
is enough to prove it after integrating against an arbitrary Schwartz function
$\psi(\tau)$. With $G(u) = \int_{-\infty}^\infty \psi(\tau) u^{-i\tau} {d\tau
\over 2\pi}$, and using Theorem {\bf\ref{T4}} this becomes
$$\int H(\varphi)(x) F(x) G(|x|) |x|^{-{1\over 2}} dx = \int \varphi(x) H(FG)(x)
|x|^{-{1\over 2}} dx$$ (on the right-hand-side $H(FG)$ is computed in the
multiplicative picture, on the left-hand-side $H(\varphi)$ is evaluated in the
additive picture). The self-adjointness of $H$ reduces this to
$\overline{H(\overline{\varphi})} = H(\varphi)$, which is a valid identity.
For the proof of the estimates we observe that $B(\varphi)$ is the Fourier
transform of an $L^1$-function hence is continuous, so that we only need to show
that it is $O(1/|x|)$ for $|x| \rightarrow \infty$. For this we use that
$B(\varphi)$ is additive convolution of $-\varphi$ with the distribution $G =
{\cal F}(-\log(|y|))$. The estimate then follows from the formula for $G$ given
in the next lemma.
\end{proof}
\begin{lem}\label{L5} The distribution $G(x) = {\cal F}(-\log(|y|))$ is given as:
$$G(\varphi) = \int_{|x| \leq 1}(\varphi(x) - \varphi(0))\, {dx \over
2\pi^2\,|x|} + \int_{|x| > 1}\varphi(x)\, {dx \over 2\pi^2\,|x|}+\ (4\log(2\pi)
+ 4\gamma_e - 2)\varphi(0)$$
\end{lem}
\begin{proof}
We have used the notation $\gamma_e = - \Gamma^\prime(1)$ for the
Euler-Mascheroni's constant ($=0.577\dots$). Let $\Delta_s$ for $\mathop{\rm Re}} \newcommand{\Tr}{\mathop{\bf Tr}(s)>0$ be
the homogeneous distribution $|x|^{s-1}$ on $\HH$. It is a tempered
distribution. The formula
$$\Delta_s(\varphi) = \int_{|x| \leq 1}(\varphi(x) - \varphi(0))|x|^{s-1}\, dx +
\int_{|x| > 1}\varphi(x)|x|^{s-1}\, dx + {2\pi^2\over s}\varphi(0)$$ defines its
analytic continuation to $\mathop{\rm Re}} \newcommand{\Tr}{\mathop{\bf Tr}(s) > -{1 \over 4}$, with a simple pole at $s =
0$. Using
$${\cal F}(\Delta_s) = \Gamma_0(s) \Delta_{1-s}$$ for $s = 1 - \varepsilon$,
$\varepsilon \rightarrow 0$, and expanding in $\varepsilon$ gives
$$\varphi(0) + \varepsilon G(\varphi) + O(\varepsilon^2) = \Gamma_0(1 -
\varepsilon)\cdot\left\{{2\pi^2\over \varepsilon}\varphi(0) + \int_{|x| \leq
1}(\varphi(x) - \varphi(0))\, {dx\over|x|} + \int_{|x| > 1}\varphi(x)\,
{dx\over|x|} + O(\varepsilon) \right\}$$ As $\Gamma_0(1 - \varepsilon) = {1\over
2\pi^2}(\varepsilon + (4\log(2\pi) + 4\gamma_e - 2) \varepsilon^2 +
O(\varepsilon^3))$ the result follows.
\end{proof}
\section{The trace of Connes for quaternions}
Let $f(g)$ be a smooth function with compact support on $\HH^\times$. Let $U_f$
be the bounded operator $\int f(g) L_2(g)\,d^*g$ on $L^2(\HH, dx)$ of left
multiplicative convolution. So
$$U_f: \varphi(x) \mapsto \int_G f(g){1\over\sqrt{|g|}}\varphi(g^{-1}x)\,d^*g$$
The composition $U_f\,{\cal F}$ of $U_f$ with the Fourier Transform ${\cal F}$
acts as
\begin{eqnarray*}
\varphi(x) &\mapsto& \int_G\int_\HH
f(g){1\over\sqrt{|g|}}\lambda(-g^{-1}xy)\varphi(y)\,dy\,d^*g \\
&=& \int_\HH\int_G
f({1\over
g}){\sqrt{|g|}}\lambda(-gxy)\,d^*g\,\varphi(y)\,dy\\
&=&
{1\over\sqrt{2\pi^2}}\int_{y\in\HH} \left( \int_{Y\in\HH} f({1\over
Y}){1\over\sqrt{2\pi^2|Y|}}\lambda(-Yxy)\,dY\right) \varphi(y)\,dy \\
&=&
{1\over\sqrt{2\pi^2}}\int_\HH {\cal F}(I(f)_a)(xy)\varphi(y)\,dy\\
\end{eqnarray*}
In this last equation $I(f)_a$ is the additive representative
${1\over\sqrt{2\pi^2|Y|}}f({1\over Y})$ of $I(f)$. Finally denoting similarly
with $\Gamma(f)_a$ the additive representative of $\Gamma(f)$ we obtain
$$(U_f{\cal F})(\varphi)(x) = {1\over\sqrt{2\pi^2}}\int_\HH {\Gamma(f)_a}(xy)\varphi(y)\,dy$$
As $f$ has compact support on $\HH^\times$ we note that $I(f)_a$ is smooth with
compact support on $\HH$ and that $\Gamma(f)_a$ belongs to the Schwartz
class. Following Connes (\cite{Co}, for $\RR$ or $\CC$ instead of $\HH$), our
goal is to compute the trace $\Tr(\Lambda)$ of the operator
$\widetilde{P_\Lambda}P_\Lambda\,U_f$, where $\widetilde{P_\Lambda} = {\cal
F}P_\Lambda{\cal F}^{-1}$ and $P_\Lambda$ is the cut-off projection to functions
with support in $|x| \leq \Lambda$. Our reference for trace-class operators will
be \cite{Go}. We recall that if $A$ is trace-class then for any bounded $B$,
$AB$ and $BA$ are trace-class and have the same trace. Also if $K_1$ and $K_2$
are two Hilbert-Schmidt operators given for example as $L^2-$kernels $k_1(x,y)$
and $k_2(x,y)$ on a measure space $(X, dx)$ then $A = K_1^*\, K_2$ is
trace-class and its trace is the Hilbert-Schmidt scalar products of $K_1$ and
$K_2$:
$$\Tr(K_1^*\, K_2) = \int\int \overline{k_1(x,y)}\,k_2(x,y)\ dxdy$$
The operator $P_\Lambda {\cal F}^{-1} P_\Lambda$ is an operator with kernel a
smooth function restricted to a finite box (precisely it is $\lambda(xy),\
|x|,|y|\leq\Lambda$). Such an operator is trace class, as is well-known (one
classical line of reasoning is as follows: taking a smooth function $\rho(x)$
with compact support, identically $1$ on $|x|\leq\Lambda$, and $Q_\rho$ the
multiplication operator with $\rho$, one has $P_\Lambda {\cal F}^{-1} P_\Lambda
= P_\Lambda Q_\rho {\cal F}^{-1} Q_\rho P_\Lambda$, so that it is enough to
prove that $Q_\rho {\cal F}^{-1} Q_\rho$ is trace-class. This operator has a
smooth kernel with compact support, so we can put the system in a box, and
reduce to an operator $K$ with smooth kernel on a torus. Then $K = (1 +
\Delta)^{-n}(1 + \Delta)^{n}K$ with $\Delta$ the positive Laplacian. For $n$
large enough, $(1 + \Delta)^{-n}$ is trace-class, while $(1 + \Delta)^{n}K$ is
at any rate bounded.)\par
So Connes's operator $\widetilde{P_\Lambda}P_\Lambda\,U_f = {\cal F}\cdot
P_\Lambda {\cal F}^{-1} P_\Lambda\cdot U_f$ is indeed trace class and
$$\Tr(\widetilde{P_\Lambda}P_\Lambda\,U_f) = \Tr(P_\Lambda {\cal F}^{-1}
P_\Lambda\cdot U_f{\cal F}) = \Tr(P_\Lambda {\cal F}^{-1} P_\Lambda\cdot
P_\Lambda U_f{\cal F} P_\Lambda)$$ can be computed as a Hilbert-Schmidt scalar
product:
$$\Tr(\widetilde{P_\Lambda}P_\Lambda\,U_f) = {1\over\sqrt{2\pi^2}}\int\int_{|x|,
|y| \leq\Lambda}\lambda(xy)\Gamma(f)_a(xy)\,dxdy$$ using the change of
variable $(x,y) \mapsto (Y= xy, y)$
$$\Tr(\widetilde{P_\Lambda}P_\Lambda\,U_f) = \sqrt{2\pi^2}
\int_{|Y|\leq\Lambda^2} \lambda(Y) \Gamma(f)_a(Y)\left(\int_{{|Y|\over
\Lambda}\leq|y|\leq\Lambda} {dy\over 2\pi^2 |y|}\right)\,dY$$
$$\Tr(\widetilde{P_\Lambda}P_\Lambda\,U_f) = \sqrt{2\pi^2}
\int_{|Y|\leq\Lambda^2} \Big(2\log(\Lambda) - \log(|Y|)\Big)\lambda(Y)
\Gamma(f)_a(Y)\,dY\leqno\bf(C)$$
This integral is an inverse (additive) Fourier transform evaluated at $1$. As
$\Gamma = {\cal F}I$ itself involves a Fourier transform the final result is
just $\sqrt{2\pi^2}M_\Lambda(I(f)_a)(1)$ where $M_\Lambda$ is the self-adjoint
operator $(2\log(\Lambda) - B)_+ = \max(2\log(\Lambda) - B, 0)$. If we recall
that $\sqrt{2\pi^2}$ is involved in the basic isometry from the additive to the
multiplicative picture, we can finally express everything back in the
multiplicative picture:
\begin{thm} The Connes operator $\widetilde{P_\Lambda}P_\Lambda\,U_f$ is a
trace-class operator and satisfies
\begin{eqnarray*}
\Tr(\widetilde{P_\Lambda}P_\Lambda\,U_f) &=& (2\log(\Lambda) - B)_+(I(f))(1)\\
\Tr(\widetilde{P_\Lambda}P_\Lambda\,U_f) &=& 2\log(\Lambda)f(1) - H(f)(1) +
o(1)\\
\end{eqnarray*}
\end{thm}
For the last line we used that $B(I(f))(1) = H(I(f))(1) = H(f)(1)$ as $H =
\log(|x|) + \log(|y|)$ commutes with the Inversion $I$. The error is $o(1)$ for
$\Lambda \rightarrow\infty$ as it is bounded above in absolute value (assuming
$\Lambda > 1$) by
$$\sqrt{2\pi^2} \int_{|Y|\geq\Lambda^2} \log(|Y|)\;
\left|\Gamma(f)_a(Y)\right|\,dY$$ and $\Gamma(f)_a$ is a Schwartz function of
$Y\in\HH$. We note that if needed the Lemma {\bf\ref{L5}} gives to the term
$H(f)(1)$ a form more closely akin to the Weil's explicit formulae of number
theory. We note that Connes's computation in \cite{Co} also goes through an
intermediate stage essentially identical with {\bf (C)} and that the
identification of the constant term with Weil's expression for the explicit
formula of number theory then requires a further discussion. The main result of
\cite{Bu1} and of this paper is thus the direct connection between $H$ and the
logarithmic derivatives of the Tate Gamma functions involved in the explicit
formulae.
{\bf Acknowledgements}
I thank the SSAS (``Soci\'et\'e de secours des amis des sciences'', quai de
Conti, Paris) for its financial support while this work was completed.
\baselineskip=14pt\parskip=12pt
|
\section{INTRODUCTION}
This is the eighth 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 section 2 we consider wavelet approach
for calculation of Arnold--Weinstein curves (closed loops)
in Floer variational approach.
In section 3 we consider
the applications of orbit technique for constructing different types of invariant
wavelet bases in the particular case of affine Galilei group with
the semiproduct structure.
In section 4 we consider applications of very
useful fast wavelet transform (FWT) technique (part 6) to calculations in
KAM theory (symplectic scale of spaces).
This method gives maximally sparse representation of (differential) operator
that allows us to take into account contribution from each level of
resolution.
\section{Floer Approach for Closed Loops}
Now we consider the generalization of wavelet variational
approach to the symplectic invariant calculation of
closed loops in Hamiltonian systems
[9]. As we demonstrated in [3]--[4] we have the parametrization
of our solution by some
reduced algebraical problem but in contrast to the cases from parts 1--4, where
the solution is parametrized by construction based on scalar
refinement equation, in symplectic case we have
parametrization of the solution
by matrix problems -- Quadratic Mirror Filters equations.
Now we consider a different approach.
Let$(M,\omega$) be a compact symplectic manifold of dimension $2n$, $\omega$ is
a closed 2-form (nondegenerate) on $M$ which induces an isomorphism $T^*M\to
TM$. Thus every smooth time-dependent Hamiltonian $H:{\bf R}\times M\to {\bf
R}$ corresponds to a time-dependent Hamiltonian vector field $X_H: {\bf
R}\times M\to TM$ defined by
$
\omega(X_H(t,x),\xi)=-{\rm}d_xH(t,x)\xi
$
for
$\xi\in T_xM$. Let $H$ (and $X_H$) is periodic in time: $H(t+T,x)=H(t,x)$ and
consider corresponding Hamiltonian differential equation on $M$:
$
\dot x(t)=X_H(t,x(t))
$
The solutions $x(t)$ determine a 1-parameter
family of diffeomorphisms
$\psi_t\in {\rm Diff}(M)$ satisfying $\psi_t(x(0))=x(t)$. These diffeomorphisms are
symplectic:
$\omega=\psi_t^*\omega$. Let $L=L_TM $ be the space of contractible loops in $M$
which are represented by smooth curves $\gamma: {\bf R}\to M$ satisfying
$\gamma(t+T)=\gamma(t)$. Then the contractible T-periodic solutions
can be characterized as the critical points of the functional
$S=S_T: L\to {\bf R}$:
\begin{equation}\label{eq:ST}
S_T(\gamma)=-\int_Du^*\omega+\int_0^TH(t,\gamma(t)){\rm d}t,\nonumber
\end{equation}
where $D\subset {\bf C}$ be a closed unit disc and $u: D\to M$ is a smooth
function, which on boundary agrees with $\gamma$, i.e. $u({\rm exp}\{2\pi i
\Theta\})=\gamma(\Theta T)$. Because [$\omega$], the cohomology class of
$\omega$, vanishes then
$S_T(\gamma)$ is independent of choice of $u$.
Tangent space $T_\gamma L$ is the space of vector fields $\xi\in
C^\infty(\gamma^*TM)$ along $\gamma$ satisfying $\xi(t+T)=\xi(t)$.
Then we have for the 1-form ${\rm d}f: TL\to{\bf R}$
\begin{equation}\label{eq:dST}
{\rm d} S_T(\gamma)\xi=\int_0^T(\omega(\dot\gamma,\xi)+{\rm
d}H(t,\gamma)\xi){\rm d}t
\end{equation}
and the critical points of $S$ are contractible loops in $L$ which satisfy the
Hamiltonian equations. Thus the critical points are precisely
the required T-periodic solutions.
To describe the gradient of $S$ we choose $a$ on almost complex structure on $M$
which is compatible with $\omega$. This is an endomorphism $J\in C^\infty({\rm
End}(TM))$ satisfying $J^2=-I$ such that
$
g(\xi,\eta)=\omega(\xi,J(x)\eta),\ \xi,\eta\in T_xM
$
defines a Riemannian metric on M. The Hamiltonian vector field is then
represented by $X_H(t,x)=J(x)\nabla H(t,x)$, where $\nabla$
denotes the gradient w.r.t. the x-variable using the metric.
Moreover the gradient of $S$ w.r.t. the induced metric on $L$ is given by
$
{\rm grad} S(\gamma)=J(\gamma)\dot\gamma+\nabla H(t,\gamma),\
\gamma\in L
$.
Studying the critical points of $S$ is confronted with the well-known
difficulty that the variational integral is neither bounded from below nor from
above. Moreover, at every possible critical point the Hessian of $f$ has an
infinite dimensional positive and an infinite dimensional negative subspaces, so
the standard Morse theory is not applicable.
The additional problem is that the gradient vector field on the loop space $L$:
${\rm d}\gamma/{\rm d}s=-{\rm grad}f(\gamma)$
does not define a well posed Cauchy problem.
But Floer [9] found a way to analyse the space ${\mathcal M}$ of bounded solutions
consisting of the critical points together with their connecting orbits.
He used a combination of variational approach and Gromov's elliptic technique.
A gradient flow line of $f$ is a smooth solution $u: {\bf R}\to M$ of the
partial differential equation
\begin{equation}\label{eq:duds}
\frac{\partial u}{\partial s}+J(u)\frac{\partial u}{\partial
t}+\nabla H(t,u)=0,
\end{equation}
which satisfies $u(s,t+T)=u(s,t)$. The key point is to consider (\ref{eq:duds})
not as the flow on the loop space but as an elliptic boundary value problem. It
should be noted that (\ref{eq:duds}) is a generalization of equation for
Gromov's pseudoholomorphic curves (correspond to the case $\nabla
H=0$ in (\ref{eq:duds})).
Let ${\mathcal M}_T={\mathcal M}_T(H,J)$ the space of bounded solutions of (\ref{eq:duds}), i.e. the
space of smooth functions $u: {\bf C}/ iT{\bf Z}\to M$, which are contractible,
solve equation (\ref{eq:duds}) and have finite energy flow:
\begin{equation}\label{eq:PhiT}
\Phi_T(u)=\frac{1}{2}\int\int_0^T\Big(\arrowvert\frac{\partial u}
{\partial
s}\arrowvert^2+\arrowvert\frac{\partial u}
{\partial t}-X_H(t,u)\arrowvert^2\Big){\rm d}t{\rm
d}s.
\end{equation}
For every $u\in M_T$ there exists a pair $x,y$ of contractible T-periodic
solutions, such that $u$ is a connecting orbit from $y$ to
$x$:
$
\lim_{s\to-\infty}u(s,t)=y(t), \ \lim_{s\to+\infty}=x(t)
$.
Then our approach from preceding parts, which we may apply or on the level of standard
boundary problem or on the level of variational approach and
representation of operators (in our case, $J$ and $\nabla$)
according to part 6(FWT technique) lead
us to wavelet representation of closed loops.
\section{Continuous Wavelet Transform. Bases for Solutions.}
When we take into account the Hamiltonian
or Lagrangian structures from part 7
we need to consider generalized wavelets, which
allow us to consider the corresponding structures instead of
compactly supported wavelet representation from parts 1--4.
We consider an important particular case of constructions
from part 7: affine
relativity group (relativity group
combined with dilations) --- affine Galilei group in n-dimensions. So, we have
combination of Galilei group with independent space and time dilations:
$G_{aff}=G_m\bowtie D_2$,
where $D_2=({\bf R}^{+}_*)^2\simeq {\bf R}^2$, $G_m$ is extended
Galilei group corresponding to mass parameter $m>0$ ($G_{aff}$ is noncentral
extension of $G\bowtie D_2$ by ${\bf R}$, where G is usual Galilei group).
Generic element of $G_{aff}$ is $g=(\Phi,b_0,b;v;R,a_0,a)$, where
$\Phi\in{\bf R}$ is the extension parameter in $G_m$, $b_0\in{\bf R}$,
$b\in{\bf R}^n$ are the time and space translations, $v\in{\bf R}^n$ is the boost
parameter, $R\in SO(n)$ is a rotation and $a_0,a\in{\bf R}^+_*$ are time and
space dilations. The actions of $g$ on space-time is then $x\mapsto
aRx+a_0vt+b$, $t\mapsto a_0t+b_0$, where $x=(x_1,x_2,...,x_n)$.
It should be noted that $D_2$ acts nontrivially on $G_m$.
Space-time wavelets associated to $G_{aff}$ corresponds to unitary irreducible
representation of spin zero. It may be obtained via orbit method. The Hilbert
space is ${\mathcal H}=L^2 ({\bf R}^n\times{\bf R},
{\rm d}k{\rm d}\omega)$, $k=(k_1,...,k_n)$, where ${\bf R}^n\times{\bf R}$ may
be identified with usual Minkowski space and we have for representation:
\begin{equation}
(U(g)\Psi)(k,\omega)=\sqrt{a_0a^n}{\rm exp}i(m\Phi+kb-\omega
b_0)\Psi(k',\omega'),
\end{equation}
with $k'=aR^{-1}(k+mv)$, $\omega'=a_0(\omega-kv-\frac{1}{2}mv^2)$,
$m'=(a^2/a_0) m$.
Mass m is a coordinate in the dual of the Lie algebra and these relations are a
part of coadjoint action of $G_{aff}$. This representation is unitary and
irreducible but not square integrable. So, we need to consider reduction to the
corresponding quotients $X=G/H$. We consider the case in which H=\{phase
changes $\Phi$ and space dilations $a$\}. Then the space $X=G/H$ is parametrized
by points $\bar{x}=(b_0,b;v;R;a_0)$. There is a dense set of vectors
$\eta\in{\mathcal H}$ admissible ${\rm mod}(H,\sigma_\beta)$, where
$\sigma_\beta$ is the
corresponding section.
We have a two-parameter family of functions $\beta$(dilations):
$\beta(\bar{x})=(\mu_0+\lambda_)a_0)^{1/2}$, $\lambda_0, \mu_0\in{\bf R}$.
Then any admissible vector $\eta$ generates a tight frame of Galilean wavelets
\begin{equation}
\eta_{\beta(\bar{x})}(k,\omega)=\sqrt{a_0(\mu_0+\lambda_0a_0)^{n/2}}
{\rm e}^{i(kb-\omega b_0)}\eta(k',\omega'),
\end{equation}
with $k'=(\mu_0+\lambda_0 a)^{1/2}R^{-1}(k+mv)$,
$\omega'=a_0(\omega-kv-mv^2/2)$.
The simplest examples of admissible vectors (corresponding to usual Galilei
case) are Gaussian vector: $\eta(k)\sim{\rm exp}(-k^2/2mu)$ and
binomial vector: $\eta(k)\sim(1+k^2/2mu)^{-\alpha/2}$, $\alpha> 1/2$, where $u$
is a kind of internal energy.
When we impose the relation $a_0=a^2$ then we have the restriction to the
Galilei-Schr\"odinger group $G_s=G_m\bowtie D_s$, where $D_s$ is the
one-dimensional subgroup of $D_2$. $G_s$ is a natural invariance group of
both the Schr\"odinger equation and the heat equation.
The restriction to $G_s$ of the representation (29) splits into the direct
sum of two irreducible ones $U=U_+\oplus U_-$ corresponding to the
decomposition $L^2({\bf R}^n\times{\bf R}, {\rm d}k{\rm d}\omega)=
{\mathcal H}_+\oplus{\mathcal H}_-$, where
$
{\mathcal H}_\pm=L^2(D_{\pm}, {\rm d}k{\rm d}\omega\
=\{ \psi\in L^2({\bf R}^n\times{\bf R},{\rm d}k{\rm d}\omega),\
\psi(k,\omega)=0\ {\textrm for} \ \omega+k^2/2m=0\}.
$
These two subspaces are the analogues of usual Hardy spaces on
${\bf R}$, i.e. the subspaces of (anti)progressive wavelets (see also below,
part III A).
The two representation $U_\pm$ are square integrable modulo the center.
There is a dense set of admissible vectors $\eta$, and each of them generates a
set of $CS$ of Gilmore-Perelomov type.
Typical wavelets of this kind are:\ the Schr\"odinger-Marr wavelet:
$
\eta(x,t)=(i\partial_t+{\triangle}/{2m}){\rm e}^{-(x^2+t^2)/2}
$,
the Schr\"odinger-Cauchy wavelet:
$
\psi(x,t)=(i\partial_t+{\triangle}/{2m})\times
{(t+i)\prod_{j=1}^n(x_j+i)}^{-1}
$.
So, in the same way we can construct different invariant bases with explicit
manifestation of underlying symmetry
for solving Hamiltonian or Lagrangian equations.
\section{SYMPLECTIC HILBERT SCALES VIA WA\-VE\-LETS}
We can solve many important dynamical problems such that KAM
perturbations, spread of energy to higher modes, weak turbulence, growths of
solutions of Hamiltonian equations only if we consider scales of spaces instead
of one functional space. For Hamiltonian system and their perturbations
for which we need take into account underlying symplectic structure we
need to consider symplectic scales of spaces. So, if
$\dot{u}(t)=J\nabla K(u(t))$
is Hamiltonian equation we need wavelet description of symplectic or
quasicomplex structure on the level of functional spaces. It is very
important that according to [12] Hilbert basis is in the same time a
Darboux basis to corresponding symplectic structure.
We need to provide Hilbert scale $\{Z_s\}$ with symplectic structure [12].
All what we need is the following.
$J$ is a linear operator, $J : Z_{\infty}\to Z_\infty$,
$J(Z_\infty)=Z_\infty$, where $Z_\infty =\cap Z_s$.
$J$ determines an isomorphism of scale $\{Z_s\}$ of order $d_J\geq 0$.
The operator $J$ with domain of definition $Z_\infty$ is
antisymmetric in $Z$:
$
<J z_1,z_2>_Z=-<z_1,J z_2>_Z, z_1,z_2 \in $ $ Z_\infty
$.
Then the triple
$\{Z,\{Z_s|s\in R\},\
\alpha=<\bar J dz,dz>\}
$
is symplectic Hilbert scale. So, we may consider
any dynamical Hamiltonian problem on functional level.
As an example, for KdV equation we have
$ Z_s=\{u(x)\in H^s(T^1)|\int^{2\pi}_0 u(x)\mathrm{d} x=0\},\
s\in R,$
$ J=\partial/\partial x,$
is isomorphism of the scale of order one, $\bar J=-(J)^{-1}$ is
isomorphism of order $-1$.
According to [13] general functional spaces and scales of spaces such as
Holder--Zygmund, Triebel--Lizorkin and Sobolev can be characterized
through wavelet coefficients or wavelet transforms. As a rule, the faster
the wavelet coefficients decay, the more the analyzed function is
regular [13]. Most important for us example is the scale of Sobolev spaces.
Let $H_k(R^n)$ is the Hilbert space of all distributions with finite norm
$
\Vert s\Vert^2_{H_k(R^n)}=\int \mathrm{d}\xi(1+\vert\xi\vert^2)^{k/2}\vert
\hat s(\xi)\vert^2.
$
Let us consider wavelet transform
$$
W_g f(b,a)=\int_{R^n}\mathrm{d} x\frac{1}{a^n}\bar g\left(\frac{x-b}{a}\right) f(x),
$$
$ b\in R^n, \quad a>0$,
w.r.t. analyzing wavelet $g$, which is strictly admissible, i.e.
$
C_{g,g}=\int_0^\infty({\mathrm{d} a}/{a})\vert\bar{\hat g(ak)}\vert^2<\infty.
$
Then there is a $c\geq 1$ such that
\begin{eqnarray*}
&&c^{-1}\Vert s \Vert^2_{H_k(R^n)}\leq\int_{H^n}
\frac{\mathrm{d} b\mathrm{d} a}{a}(1+a^{-2\gamma})\vert\times\\
&& W_gs(b,a)\vert^2
\leq
c\|s\|^2_{H_k(R^n)}.
\end{eqnarray*}
This shows that localization of the wavelet coefficients at small
scale is linked to local regularity.
So, we need representation for differential operator ($J$ in our case) in
wavelet basis. We consider it by means of the methods from part 6.
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}
Imaging air \v{C}erenkov telescopes (IACTs) are currently able to detect
{$\gamma$-ray}\
photons of TeV ($10^{12}\,$eV) energy from BL~Lac objects within one hour of
observation and to measure their spectra using a few hours of a good data (for a
review see Weekes et al.~\cite{weekesetal97}). The objects of this class
detected to date are, in order of increasing redshift: Mkn~421 ($z=0.031$)
(Punch et~al.~\cite{punchetal92}, Petry et~al.~\cite{petryetal96}), Mkn~501
($z=0.034$)
(Quinn et al.~\cite{quinnetal96}, Bradbury et al.~\cite{bradburyetal97}),
1ES~2344+514
($z=0.044$) (Catanese et al.~\cite{cataneseetal98}) and PKS~2155-304 ($z=0.117$)
(Chadwick et al.~\cite{chadwicketal98}). Measurements of the spectrum can be
made
over the energy range from $200\,$GeV to about $10\,$TeV using this technique.
Because these photons interact with infra-red radiation to form
electron-positron pairs,
the signal is expected to be attenuated by absorption both within the source
itself and
in the intergalactic medium. Thus, it is possible to use the observations to
study the
intergalactic infrared radiation field (IIRF), given some general, model-dependent
constraints on the
spectrum intrinsic to the source (Stecker, De~Jager,
Salamon~\cite{steckeretal92}).
Determining the IIRF, in turn, allows one to model the evolution of the galaxies
which produce it.
In this {\em Letter} we analyze recent observations of the object Mkn~501
(Konopelko et~al.~\cite{konopelkoetal98b}) which are unique both for the quality of
the
spectra obtained and their energy range (up to $20\,$TeV). These data, taken
during a
period in which the intensity of the source varied strongly,
show a pronounced curvature in the spectrum, being
significantly softer (steeper)
towards higher energy.
We unfold these data using the
upper curve for the spectral energy
distribution of the IIRF given by Malkan \& Stecker (\cite{malkanstecker98},
henceforth MS98), with the corresponding
{$\gamma$-ray}\ opacity as calculated
by Stecker \& De~Jager (\cite{steckerdejager98},
henceforth SD98) and find that the intrinsic
spectrum is flat, ${\rm d} N/{\rm d} E \propto E^{-2}$.
The implications of this result for both the
absorption model and the synchro-self-Compton emission model are discussed.
\section{Measurement of spectrum of Mkn~501}
The BL Lac object Mkn~501 exhibited strong emission in TeV $\gamma$-rays from
March to
October, 1997 (Protheroe et al.~\cite{protheroeetal98}). During this period the
source
was continuously monitored by several ground-based imaging air \v{C}erenkov
telescopes,
including the {\it HEGRA} stereoscopic system of 4 imaging air \v{C}erenkov telescopes
(Aharonian et al.~\cite{aharonianetal97a}), which observed it for a
total
exposure time of $110\,$hours (Aharonian et al.~\cite{aharonianetal99}). The
unprecedented statistics of about 38,000 TeV photons, combined with the good
energy
resolution of $\sim 20$\% over the entire energy range and with detailed studies
of the
detector performance (Konopelko et al.~\cite{konopelkoetal99}), allowed a
determination
of the spectrum in the energy range $500\,$~GeV to $24\,$~TeV (Konopelko et
al.~\cite{konopelkoetal98b}). The Mkn~501 energy spectrum measured by the
{\it HEGRA} collaboration extends well beyond $10\,$TeV, where
uncertainties related to the saturation effect could in principle play a role.
Various data consistency checks were performed in order to avoid these
effects. In addition, simultaneous observations of the Crab Nebula
(the standard-candle TeV source) were undertaken from 1997 September
to 1998 March. Using similar data analysis, the Crab Nebula spectrum
derived from the {\it HEGRA} data was found to be
a pure power law with a differential spectrum index of 2.6 over the
energy range $500\,$GeV -- $23\,$TeV (Konopelko et al.\ \cite{konopelkoetal98a}).
These results are consistent with previous
measurements by the Cangaroo group in the energy range $7$--$50\,$TeV
(Tanimori et al.\ \cite{tanimorietal98}).
The flux of {$\gamma$-rays}, from Mkn~501,
averaged over the entire observation period, was about
three
times that of the Crab Nebula ($3\,$\lq\lq Crab\rq\rq).
Averaged over each day, the {$\gamma$-ray}\
rate
showed strong variations, with a maximum of $10\,$Crab detected on 26/27 June
1997. As
remarked by Aharonian et al.~(\cite{aharonianetal97b}), the hardness ratio of
the
steepening Mkn 501 spectrum appears to be independent of the absolute flux.
The high
{$\gamma$-ray}\ detection rate provided event statistics of a few hundreds within 1
day's
observations ($\sim 3-5$ hours) which suffices to evaluate the energy spectrum
over the
range $1$--$10\,$TeV. The analysis of the spectral shape on a daily basis did
not reveal
any substantial correlation between the {$\gamma$-ray}\ flux and the spectral behavior
(Aharonian
et al.~\cite{aharonianetal99}). This justifies the presentation of a
time-averaged energy
spectrum of Mkn~501 in its active state,
which is shown in Figure 1 over the
energy range from $500\,$GeV to $24\,$TeV. The vertical error bars in this
figure correspond
to statistical errors. Note that the systematic errors at energies below
$1\,$TeV appear to
be quite large, reaching $\sim$50 \% at $500\,$GeV.
Over the entire range, the spectrum shows a gradual softening towards higher energy.
The $19$--$24\,$TeV energy
bin contains a signal with a significance of $3.7\sigma$.
However, the steep energy spectrum and
20~\% energy resolution
do not permit one to exclude the interpretation that these {$\gamma$-rays}\
may have spilled over from the lower energy bins.
Therefore, the energy spectrum is consistent with the hypothesis of
a maximum
energy for the detected {$\gamma$-rays}\ of $\sim 18\,$TeV. The shape of the energy
spectrum is well
described by a power-law with an exponential cutoff. A fit of the
data over the energy region
where the systematic errors are small, i.e., from $1\,$TeV to $24\,$TeV, gives
\begin{eqnarray}
{\rm d} N/{\rm d} E &=& A E^{-\alpha} \exp\left(-E/E_0\right)
{\rm [cm^{-2} s^{-1} TeV^{-1}]}\,. \nonumber\\
A &=&
(9.7\, \pm 0.3\, (\textrm{stat})\, \pm 2.0\, (\textrm{syst}))\cdot 10^{-11}
\nonumber\\
\alpha &=& {-1.9 \, \pm 0.05 \,(\textrm{stat})\, \pm 0.05 \,(\textrm{syst})}
\nonumber\\
E_o &=& 5.7 \,\pm 1.1 (\textrm{stat}) \,\pm 0.6 \,(\textrm{syst}) {\rm TeV}
\nonumber\\
\end{eqnarray}
The logarithmic slope of the energy spectrum (``power-law index'') is $1.8$
in the energy
range $1\,$--$5\,$TeV, and $3.7$ above $5\,$TeV.
Independent measurements of the Mkn~501 TeV energy spectrum by the Whipple
Observatory
(Samuelson et al.~\cite{samuelsonetal98}) are in very good agreement with these
results.
The best fit to the data presented by the Whipple group agrees precisely with a
fit to the
{\it HEGRA} data in the energy range $500\,$GeV -- $10\,$TeV. However the HEGRA group
measured the
spectrum well above $10\,$TeV where it exhibits a further steepening.
\section{Absorption on the diffuse intergalactic infra-red background}
The formulae relevant to absorption calculations involving pair-production are
given and
discussed in Stecker, De Jager \& Salamon~(\cite{steckeretal92}). For {$\gamma$-rays}\
in the TeV
energy range interacting at redshifts $z \ll 1$, the pair-production cross
section
is maximized when the soft photon energy is in the infra-red range:
\begin{eqnarray}
\lambda (E_{\gamma}) \simeq \lambda_{e}{E_{\gamma}\over{2m_{e}c^{2}}} &=&
2.4E_{\gamma,TeV} \; \; \mu{\rm m}
\end{eqnarray}
where $\lambda_{e} = h/(m_{e}c)$ is the Compton wavelength of the electron. For
a $10\,$TeV
{$\gamma$-ray}\, this corresponds to a soft photon in the mid infra-red region of the
spectrum,
having a wavelength around $24\,$$\mu$m.
Pair-production interactions
take place with
photons over a range of wavelengths around the optimal value, as determined by
the energy
dependence of the cross section.
Stecker \& De Jager~(SD98)
have computed
the absorption coefficient of intergalactic
space using a new,
empirically based calculation of the spectral energy distribution (SED) of
intergalactic low
energy photons (MS98).
Assuming that the IIRF is basically in
place by a
redshift $\sim$ 0.3, having been produced primarily at higher redshifts
(Stecker \& De Jager~\cite{steckerdejager97},~\cite{steckerdejager98};
Madau~\cite{madau95}), SD98 limited their calculations to $z<0.3$.
Evolution in stellar emissivity affects the predicted IIRF and
is expected to level off
or decrease at redshifts greater than $\sim 1.5$ (Madau \cite{madau96}).
In this paper, we
assume that evolution continues up to $z=2$, leading to the
higher of the two IIRF used by SD98. This is more consistent with recent
data on IR galaxy evolution, dust absorption, and the lower
limits from IR galaxy counts (Stecker \cite{stecker99}).
To compute the absorption, we adopt the SD98
parametric expressions for $\tau(E,z)$ for $z<0.3$, taking a
Hubble constant of $H_o=65$ km s$^{-1}$Mpc$^{-1}$.
The unfolded {\it HEGRA} data points are also shown in Figure 1,
together with a fit to a power law energy spectrum.
We find
\begin{eqnarray}
{\rm d}N_\gamma/{\rm d}E &=&
1.32\pm0.04 \cdot 10^{-10} \times
\nonumber\\
&&(E/1\,{\rm TeV})^{-2.00\pm0.03}
\nonumber\\
&&{\rm photons\, cm^{-2}\, s^{-1}}
\end{eqnarray}
with $\chi^2$ = 18.6 for 15 degrees of freedom,
giving a high chance probability of 0.2.
In the mid-energy region $1$--$10\,$TeV,
where the measured energy spectrum is
very well-defined,
the data points deviate from the fit by less then 15\%,
which equals the estimated systematic error.
Note that both the statistical and the
systematic uncertainties of the spectrum measurements
increase towards the upper end of the energy range,
where they reach 30\% and 60\%, respectively.
Thus, the data points of the unfolded spectrum are
consistent, within the statistical and systematic errors,
with the simple
power law fit of differential spectrum index 2.0.
Our analysis shows that fitting the unfolded
spectrum using an additional
exponential term, result in a $\chi^2$-value of the same magnitude.
\begin{figure}[h]
\plotone{sp-mrk501.eps}
\caption{\protect\small
The energy spectrum of Mkn~501 as measured by the {\it HEGRA}
IACT array
(open circles) (Konopelko et al.~\protect\cite{konopelkoetal98b}). The combined
power law plus
exponent fit of the
{\it HEGRA}
data is shown by the dotted-dashed curve. The Mkn~501
spectrum
measured by the Whipple group (filled circles) is taken from
Samuelson et al.~(\protect\cite{samuelsonetal98}).
Also shown are the de-absorped {\it HEGRA} points (open circles)
found using the
optical depths to absorption calculated by Stecker \& De
Jager~(\protect\cite{steckerdejager98})
for the ``high'' intergalactic infra-red radiation field.
The power-law fit to these data is shown by the solid line,
and has a photon index of $\protect\alpha=2.00\pm0.03$.}
\end{figure}
\section{Implications for the intrinsic spectrum and the inter-galactic
absorption}
In order to understand the implications of the
absorbed and de-absorbed spectra shown in Figure 1,
it is necessary to adopt a model or scenario for the production of TeV
photons in the source. Of the many suggestions in the
literature, interest has recently centred on those in which a single
population of relativistic
electrons is responsible for both the TeV photons and
for photons in the X-ray region of the spectrum, as in the
synchrotron self-Compton (SSC) and external Compton models
(e.g.,
Bloom \& Marscher \cite{bloommarscher96};
Inoue \& Takahara \cite{inouetakahara96};
Ghisellini \& Madau \cite{ghisellinimadau96};
Dermer, Sturner \& Schlick\-eiser \cite{dermeretal97};
Mastichiadis \& Kirk~\cite{mk97} (henceforth MK97);
Sikora et al.\ \cite{sikoraetal97};
Georganopoulos \& Marscher~\cite{gm98};
Ghisellini et al.\ \cite{ghisellinietal98};
Levinson \cite{levinson98}).
These models are favored because they provide a natural way to
understand the similar variability timescales of the X-ray and
TeV emission.
During the period
of the TeV observations, Mkn~501 was also observed in X-rays
using the BeppoSAX instrument
(Pian et al.~\cite{pianetal98}) and the Rossi X-ray Timing Explorer
(Lamer \& Wagner~\cite{lamerwagner98}). The spectrum in this energy range
varied strongly, with generally a very hard spectral index extending to
much higher energies $\gtrsim100\,$keV than
during less active epochs. There was a close temporal correlation between the
X-ray and TeV fluxes, further
strengthening the case for the origin of X-rays
and TeV in a common electron population.
Extensive studies of the variability properties of the SSC model
have been undertaken (MK97), which show that the
mechanism most likely to be
responsible for the variability shown by the object Mkn~421 is a change in
the maximum energy
(expressed as a Lorentz factor: $\gamma_{\rm max}$)
to which electrons are accelerated. These results were also
applied to Mkn~501 (Mastichiadis \& Kirk \cite{mk99},
henceforth MK99), where,
once again, variability induced by
a change in $\gamma_{\rm max}$ appears to give a reasonable fit
to the X-rays and to the TeV data then available.
For such models, the most important property revealed by the {\it HEGRA}
observations discussed above is the lack of variation in the TeV spectrum,
despite the fact that the correlated variations in the X-rays
show strong spectral variations, consistent with an
increase in $\gamma_{\rm max}$.
Inspection of the model
light curves in the TeV range show that
the softer the spectral slope becomes, the more sensitively it reacts to
changes in $\gamma_{\rm max}$. This is because intrinsic
spectra softer than a photon index of $\alpha=3$ are a direct
result of the electron
cut-off, whereas harder spectra (photon index $\alpha\approx2$)
can be formed
by power-law electrons scattering off
a range of target photon energies.
Thus, on the basis of the homogeneous SSC model, we find that
intensity variations with constant spectral shape imply an intrinsic
spectrum with a photon index of $\alpha\approx2$.
This provides additional evidence that the spectrum of
Mkn~501, which has a slope $\alpha\gtrsim3.7$ above $5\,$TeV,
must be modified by the effects of inter-galactic absorption.
Of the two IIRF considered by SD98, only the
higher provides sufficient
absorption to account for such a strong modification.
We note again that lower limits from galaxy counts in the
mid-IR, as well as other observational data (Dwek et al.~\cite{dweketal98}),
favor the high IIRF (Stecker \cite{stecker99}).
As well as constraining the intergalactic absorption, the
observation of an intrinsic spectrum of $\alpha\approx2$ at $10\,$TeV
requires a higher Doppler
factor than considered by MK97 and MK99. Using the approximate
scaling laws presented in
MK97, we can estimate that a Doppler boosting factor of
$\delta\sim50$ suffices
to produce $\alpha\approx2$ at $10\,$TeV, and have confirmed this by
running full simulations. It is interesting to note that such boosting
factors, although larger
than values measured in sources which display apparent
superluminal motion
(Vermeulen \& Cohen~\cite{vermeulencohen94}), seem to be indicated both by
observations of intra-day variability (Wagner \& Witzel~\cite{wagnerwitzel95})
and of extremely rapid variations in the TeV flux of blazars (e.g.,
Gaidos et al.~\cite{gaidosetal96}).
A completely model independent conclusion is still elusive,
and will remain so until observations
of comparable quality on other blazars of different redshifts are available.
Nevertheless, the two independent arguments we have presented favoring an
intrinsic emission spectrum close to $\alpha=2$ indicate that the effect of
absorption by the intergalactic infra-red background radiation
in the spectrum of Mkn~501 is strong, and
suggest a Doppler boosting factor for this source of $\delta\gtrsim50$.
\acknowledgments{We thank F.~Aharonian, W. Hofmann and H.J.~V\"olk for
stimulating discussions.
A.M. and J.G.K. acknowledge support for this collaboration
by the European Commission under the TMR Programme, contract
ERBFMRX-CT98-0168.
We would like to thank the anonymous referee for helpful comments.}
|
\section{Sinusoidal Modulation}
Elastic X-ray scattering confirmed \cite{kiryu95}
that the distortion in the I phase is incommensurate. The deviation
from dimerization $|q-\pi|$ increases with increasing magnetic field.
This can be illustrated in the unfrustrated XY model which corresponds
via the Jordan-Wigner transformation to free fermions.
The susceptibility towards distortion
becomes maximum at $q=2k_{\rm F}$ since $q=2k_{\rm F}$ allows
to create particle-hole pairs of vanishing energy. A distortion
with $q=2k_{\rm F}= 2\pi m + \pi$ is formed where $m$ is the magnetization.
Since this distortion couples the degenerate
states at $-k_{\rm F}$ and at $k_{\rm F}$ a gap appears there.
Let us now gradually increase the interaction
corresponding to the $S^z_iS^z_{i+1}$ terms at fixed particle number.
For continuity no state can cross the gap such that the picture of a
distortion at $2k_{\rm F}$ remains valid \cite{mulle81,uhrig98a}.
So we have
\begin{equation}
\label{qm-bedg}
|q-\pi| = 2\pi m \ .
\end{equation}
This relation is confirmed numerically \cite{schon98}.
Since experimentally it turned out that higher harmonics of the
distortion are considerably suppressed it is plausible to start with
($q$ given by (\ref{qm-bedg}) \cite{uhrig98a})
\begin{equation}
\label{sinus}
H = J \sum_i[1-\delta\cos(qr_i)] {\bf S}_i {\bf S}_{i+1} \ .
\end{equation}
The distribution of the local magnetizations $m_i=\langle S^z_i\rangle$
is found from NMR experiments \cite{fagot96,horva99}.
With some success, the experimental data were compared to a continuum
theory \cite{fujit84}. This theory, however, is based on a Hartree-Fock
treatment where all Hartree and Fock terms are spatially constant.
In Ref. \cite{uhrig98a} it is shown that this is a too crude approximation
reducing the physics to the one of a XY chain. The antiferromagnetic
correlations found in this way are much smaller than those of an
isotropic XYZ chain. This conclusion is corroborated by several works
\cite{feigu97,forst98,schon98}. But the spin isotropy of
cuprates can hardly be questioned. To account for the smaller amplitudes
it is proposed \cite{uhrig98a,uhrig99b} that experimentally only an
effective magnetization $m_i^{\rm eff}$ is seen which is
an average
\begin{equation}
\label{mittel}
m^{\rm eff}_i = (1-2\gamma)m_i + \gamma(m_{i-1}+m_{i+1}) \ .
\end{equation}
The results for $\gamma=0.2$ agree well with
experiment \cite{uhrig98a}. The microscopic origin of the average
is discussed in Sect.~4.
The reason for strong local magnetizations around the zeros of the
modulation (cf. Fig.~\ref{inkomm}) is found in the localization of a spinon.
Each zero binds exactly one spinon \cite{uhrig99a}. Summing the $m_i$
around a magnetization maximum yields 1/2.
The order of the transition D $\to$ I can be determined by investigating
the ground state energy $E(m)$ as function of the average magnetization
$m$ \cite{schon98}. By means of a Legendre transformation
$\tilde E(h)=E(m) - h m$, one obtains the dependence of ground state
energy $\tilde E(h)$ on the magnetic field $h=g\mu_{\rm B} H$.
It is found that a discontinuous jump for $m\to0$ occurs. This
implies that the transition D $\to$ I for fixed sinusoidal modulation is
of first order. The mean square of $\cos(qr_i)$ jumps discontinuously
from 1 to 1/2 if $q$ deviates infinitesimaly from $\pi$ \cite{schon98}
since the $r_i$ are summed over integer values only.
Experimentally, however, the observed first order jumps are much lower
than those found for fixed sinusoidal modulation
\cite{palme96a,kiryu95,fagot96,ammer97,loren97a}.
\section{Adaptive Modulation}
Since it was stated above that sinusoidal modulation alone does
not account for the weak first order D $\to$ I transition we turn to the
full minimization of the ground state energy of (\ref{hamilton}).
Derivation with respect to $\delta_i$ yields
\begin{equation}
0 = \langle{\bf S}_{i+1}{\bf S}_{i}\rangle -
\langle\langle{\bf S}_{j+1}{\bf S}_{j}\rangle\rangle +K\delta_i\ ,
\label{bondmin}
\end{equation}
where $\langle\langle\cdot \rangle\rangle$ stands for
the expectation value and the average along the chain. The
double-bracketed term accounts for
the constraint of the vanishing average of the $\delta_i$.
The minimization is done iteratively \cite{feigu97,schon98,uhrig99b}.
The generic result is depicted in Fig.~\ref{inkomm}.
\begin{figure}
\vspace*{3mm}
\hfill\psfig{figure=fig-k18-nnneu.eps,width=6.5cm}
\vspace*{-3mm}
\caption[fig2]{Symbols: DMRG-result at $K=18J$, $\alpha=0.35$.
Upper panel: local distortions;
solid line: from
Eq. (\protect\ref{fit1b}) with $\delta= 0.014$, $k_{\rm d}=0.959$,
$\xi_{\rm d} = 10.5$.
Lower panel: local magnetizations; solid line:
Eq. (\protect\ref{fit1a}) with $W=0.21 $, $R=5.0$, $k_{\rm m}=0.992$,
$\xi_{\rm m} = 7.9$.}
\label{inkomm}
\end{figure}
The local magnetizations do not display major differences to the
results for sinusoidal modulation in \cite{uhrig98a} because
the spinon localization is in essence determined only by the slope with
which the modulation vanishes.
Note that the envelope of $m_i$ is proportional to the probability
of finding a spinon at that site \cite{uhrig99a}.
Hence, the magnetic part of the soliton displays localization as for
sinusoidal modulation.
For the distortions the relation (\ref{bondmin}) implies that the
deviations from constantly alternating dimerization are also localized.
The distortion belonging to an isolated soliton is a kink, i.e.
the distortion between two solitons resembles the one in the
D phase. This implies
a crucial advantage over the sinusoidal modulation. For kink-like
solitons the reduction
of the mean square distortion is proportional to the soliton number.
Hence, a low soliton concentration leads only to a small change of the
energy such that $E(m)$ is continuous (in the sense of Lipschitz) on
$m\to0$ \cite{schon98}.
The investigations in Ref. \cite{schon98}
of the model (\ref{hamilton}) yielded a continuous
phase transition even though the magnetization grows very quickly
above the critical field $m\propto -1/\ln(H-H_c)$.
Most of the results of the continuum theories comply also with a
phase transition of second order
\cite{brazo80a,merts81,buzdi83a,fujit84}.
Solely Horovitz mentions the possibility of soliton attraction
in an early work \cite{horov81} implying a first order
transition. Buzdin {\it et al.} expect a first order transition at $T>0$
\cite{buzdi83a}.
In fact, the details of the models matter.
Cross \cite{cross79b} argued already
that an elastic energy {\it with} dispersion $K(q)$ being minimum at
$q=\pi$ leads to a first order transition. The positive curvature
of $K(q)$ around $q=\pi$
suppresses higher harmonics in the distortion. Hence, sinusoidal
modulation is favoured. The concomitant concavity in $E(m)$ at low
magnetization $m$ implies phase separation via the Maxwell construction.
So the phase transition is first order \cite{schon98}.
This finding is in accordance with the conclusion from a phenomenological
Ginzburg-Landau description \cite{bhatt98} that the
D $\to$ I transition is generically of first order.
At the transition the distance between two solitons is
rather large so that the difference between sinusoidal and adaptive modulation
matters most. At higher soliton concentrations the adaptive modulation
becomes more and more sinusoidal but with a concentration dependent amplitude.
The mean square distortion could be determined from the elastic lattice
constants. These experimental results agree well with the
predictions based on the model (\ref{hamilton}), see \cite{loren98}.
The continuum theories applying to the isotropic
Heisenberg chain \cite{nakan80,zang97} provide the following
results (details in Ref. \cite{uhrig99b};
${\rm sn}, {\rm cn}, {\rm dn}$: elliptic Jacobi functions)
\begin{eqnarray}
\label{fit1a}
m_i &=& \frac{W}{2}\left\{ \frac{1}{R}
{\rm dn}\left(\frac{r_i}{k_{\rm m}\xi_{\rm m}},k_{\rm m}\right)
+(-1)^i {\rm cn}\left(\frac{r_i}{k_{\rm m}\xi_{\rm m}},k_{\rm m}\right)\right\}\\
\delta_i &=& (-1)^i \delta\
{\rm sn}\left(\frac{r_i}{k_{\rm d}\xi_{\rm d}},k_{\rm d}\right)
\label{fit1b}
\\
\label{zus1}
\mbox{with}\qquad
\xi&:= &\xi_{\rm m} \, =\, \xi_{\rm d} \quad \Leftrightarrow \quad
k \, :=\, k_{\rm m} \, =\, k_{\rm d}\\ \label{zus2}
1&=& 4 m k_{\rm m/d} K(k_{\rm m/d} )\xi_{\rm m/d} \\
\label{zus3}
1&=& \pi k_{\rm m} \xi_{\rm m}\frac{W}{R}\ .
\end{eqnarray}
The fits in Fig.~\ref{inkomm} are based on Eqs. (\ref{fit1a},\ref{fit1b}).
Identity (\ref{zus1}) is {\it not} complied with, see Sect.~3.
Otherwise no agreement
would be obtained. Relation (\ref{zus2}) is imposed on the fits whereas
Eq. (\ref{zus3}) serves as check. It is fulfilled within 4\%.
The fact that $\xi_{\rm d}/\xi_{\rm m}\approx 1.33$ is considerably
above unity complies nicely with the experimental findings. Elastic
X-ray scattering \cite{kiryu95} found $\xi_{\rm d}=13.6\pm 0.3$ while
the NMR investigations provide $\xi_{\rm m}\approx 10$
\cite{horva99} close to the transition.
There is also another way to introduce
solitons in a spin-Peierls system than the application of magnetic field.
Doping non-magnetic impurities in a spin-Peierls systems
cuts the infinite chains into finite chain segments \cite{khoms96}.
In a number of works
\cite{marti96b} it has been shown that each
impurity frees one spinon which is situated
either before or after the impurity on the chain.
Assuming a fixed dimerization it is easy to see that
the spinon is bound to its generating impurity \cite{uhrig99a}
in accordance with experimental
results \cite{hassa98,els98}.
But in a spin-Peierls system the change of the modulation has
to be taken into account, too. This is done by introducing
\begin{equation}
\label{stoerkopp}
H = J\sum_{i\ge 0}\left[
(1+\delta_i){\bf S}_i{\bf S}_{i+1} +\alpha {\bf S}_i{\bf S}_{i+2}+
\frac{K}{2} \delta_i^2 + f\delta_i(-1)^i\delta_{\rm bulk}
\right]\ ,
\end{equation}
such that the impurity is at site -1. The important amendment
compared to $H$ in Eq. (\ref{hamilton}) is the last term.
If the spinon moves away from the impurity the distortion
pattern is changed between impurity and spinon.
Due to an {\em elastic} interchain interaction (parametrized
by $f$)
a coherent distortion pattern throughout the whole three-dimensional
system is preferred. A deviation from this pattern is
energetically unfavourable. So one is led to include the last
term in Eq. (\ref{stoerkopp}) assuming that the adjacent chains
are dimerized as in the unperturbed D phase. The relation
$K=K_0+f$ ensures the consistency of the distortion
amplitude with its bulk value.
Fig.~\ref{impur} displays the results for various elastic
interchain interactions.
\begin{figure}
\hfill\psfig{figure=fig-stoer.eps,width=6cm}
\caption[fig4]{Local distortions (enhanced) and magnetizations in the
vicinity of an impurity (chain end) at site -1 for $K=18.064$, $\alpha=0.35$
and $f$ as displayed.}
\label{impur}
\end{figure}
The soliton is {\em not} localized at the impurity but at a certain
distance. For lower $f$ values the soliton resembles very much the ones
in the I phase (cf. Fig.~\ref{inkomm}). On increasing $f$ the soliton
is squeezed more and more towards the impurity.
Analogous results can be found by QMC \cite{hanse98}, too.
Sure enough, the soliton is bound to the impurity confirming previous
ideas \cite{khoms96,els98}.
Moreover, the first excitations for the same distortions
and in the same spin sector are found below the singlet-triplet gap at
52\% of $\Delta_{\rm trip}$
for $f=0.01$ and at 64\% for $f=1$. This matters for the
spectroscopic analysis.
\section{Local Renormalization}
Most remarkable is
the difference between the distortive soliton width $\xi_{\rm d}$ and the
magnetic soliton width $\xi_{\rm m}$ in the numerical
results (cf. Fig.~\ref{inkomm}). It amounts to 30\% at $\alpha=0.35$
depending mainly on the frustration for low soliton concentrations
\cite{uhrig99b}.
The challenge is to extend the existing continuum
description to account for this fact. Let us revisit
the semiclassical treatment of the bosonized description of the
spin-Peierls problem \cite{nakan80}. Minimizing the total
energy the variation of the distortion $\delta(x)$ leads to
\begin{equation}
\delta(x) \propto e^{-2\sigma}\cos(2\phi_{\rm class})
\end{equation}
where $\sigma:=\langle \hat\phi^2 \rangle$ denotes the
renormalizing fluctuations
of the local bosonic field $\hat\phi$ about the classical field
$\phi_{\rm class}$.
A soliton corresponds to a solution where $\phi_{\rm class}$ increases
by $\pi$ in a kink-like fashion. If $\sigma$ is assumed to take the
spatially constant value that it has in the ground state \cite{nakan80}
one obtains
\begin{equation}
\delta(x)/\delta = \cos(2\phi_{\rm class}) = \tanh(x/\xi)\ ,
\end{equation}
wherein $\xi$ is given by the ratio $v_{\rm S}/\Delta_{\rm trip}$
of the spin wave velocity and the gap. The alternating component of
the magnetization $a(x)$ is proportional to $\sin(2\phi_{\rm class})$.
Hence, one has $a(x) \propto \sqrt{1-\tanh^2(x/\xi)} = 1/\cosh(x/\xi)$.
But the presence of the soliton induces a deviation $\Delta\sigma$
from its ground state value. Fig.~\ref{locren} displays a generic
result for this deviation.
It is calculated on top of the solution
of Nakano and Fukuyama \cite{nakan80}.
\begin{figure}
\vspace*{-5mm}
\hfill\psfig{figure=fig-sigma.eps,width=6cm}
\vspace*{-5mm}
\caption[fig3]{Local fluctuations as function of $x$. The deviation
$\Delta\sigma$ from the ground state value (solid line) and the
renormalizing factor (dashed) is shown.}
\label{locren}
\end{figure}
The alternating component
\begin{equation}
a(x) \propto \sqrt{1-\exp(4\Delta\sigma)\tanh^2(x/\xi)}
\end{equation}
is thus indeed narrower than before due to the spatial dependence of
the renormalization factor.
The result in Fig.~\ref{locren} is only a first step
since the influence of the altered magnetic behaviour is not included.
Yet it is clear that the renormalization is a local quantity and that this is
the origin of the difference between $\xi_{\rm d}$ and $\xi_{\rm m}$.
\section{Phasons}
In Sect.~1 the discrepancy between
theoretical and experimental amplitudes of the alternating local
magnetizations was solved by an averaging procedure (\ref{mittel}).
Passing to adaptive modulations does not change the amplitudes much
\cite{uhrig99b}. Hence, one still has to find the microscopic origin for the
averaging.
Non-adiabatic effects are in fact responsible. The soliton lattice
oscillates about its equilibrium positions. These
oscillations are best understood in a continuum description which is
well justified if the typical length $\xi$ is noticeably larger than
the lattice constant. In such a continuum description the modulation
$\delta(x)$ can be shifted along the chains without energy cost. This
continuous translational invariance is spontaneously broken by the
soliton lattice. Hence, there are massless Goldstone modes, the
so-called phasons. They are analogous to the phonons of a
crystal lattice except that they do not have three branches
but only with one. While the atoms of a crystal lattice can be shifted
in all three spatial directions the solitons can be shifted only
along the chains.
Albeit the ideal spin-Peierls system is magnetically one-dimensional
the distortions on different chains are elastically coupled. Thus, the
phasons are governed by a 3D, though anisotropic, dispersion.
The dispersion parameters are determined from the anisotropy of the
correlation lengths assuming a Ginzburg-Landau description \cite{bhatt98}.
The corresponding $T^3$ term in the specific heat has been measured
\cite{loren96} and theory and experiment agree astonishingly well
\cite{bhatt98}.
The zero point motion and the excited motion of phasons lead to
the averaging (\ref{mittel}). Let us denote the adiabatic result
for the local magnetizations by
$m_i = a({r}_i)\cos(\pi {r}_i) + u({r}_i)$ where $r_i$
is the component along the chains, $a(r_i)$ the alternating component
and $u(r_i)$ the uniform component of the magnetizations.
A local shift can be implemented by replacing
$\pi {r}_i \to \pi {r}_i+\hat \Theta({\bf r}_i)$ where
$\hat \Theta({\bf r}_i)$ denotes a phase shift operator.
On the long time scales
of a NMR measurement one measures
\begin{equation}
\label{mi-measured}
m_i^{\rm exp} = \langle m_i\rangle
= a({r}_i) \gamma' \cos(\pi {r}_i) + u({r}_i)
\end{equation}
with
$\gamma' := \exp\left(-\frac{1}{2N}\sum_i
\langle\hat \Theta^2({\bf r}_i)\rangle
\right) < 1$.
So there is an amplitude reduction engendered by the local
fluctuations. The factor $\gamma'$ is comparable to a Debye-Waller factor.
Using the values fixed previously \cite{bhatt98} yields
$\gamma' = 0.16 \exp(- (T/T^*)^2/2)$ with $T^*\approx 16.9$K \cite{uhrig99b}.
Eq. (\ref{mittel}) is retrieved by estimating $a(r)$ and $u(r)$ from the
discrete values $m_i$ by
$
a({r}_i) = m_i/2-(m_{i-1}+m_{i+1})/4$
and by
$u({r}_i) = m_i/2+(m_{i-1}+m_{i+1})/4$.
Inserting these formulae into Eq. (\ref{mi-measured}) yields
Eq. (\ref{mittel}) with $\gamma=(1-\gamma')/4$. At $T=0$, $\gamma$
takes the value $0.21$ in accordance with experiment
\cite{uhrig98a,uhrig99b}.
\section{Conclusions}
In this report the modulated phases of spin-Peierls systems were
discussed. Such modulations are induced either by magnetic field
or by impurities. In both ways the singlet pairing in the
D phase is broken and spinons are freed. The lattice
distortion adapts to the spinon by forming a zero to which the spinon
is bound. This new entity constitutes the spin-Peierls soliton.
The order of the transition D $\to$ I phase on increasing field
depends on model details. Imposed sinusoidal modulation leads
to a pronounced first order transition. Allowing the system
to choose an optimum modulation makes the transition continuous
{\em if} the elastic energy is wave vector independent. If the
elastic energy itself pins the modulation to $\pi$ a weak first
order transition is found.
Doping induced solitons are bound to their generating impurity.
The distortion pattern between impurity and soliton is not coherent
with the bulk pattern. This costs energy which acts as a confining
potential. Binding occurs for which experimental evidence exists
\cite{hassa98,els98}.
The difference between magnetic and distortive soliton width
could be traced back to the so far neglected spatial dependence
of the renormalizing local fluctuations. A fully self consistent
analysis is in progress.
The reduction of the alternating magnetic amplitude due to phasons
provides striking evidence for the importance
of the lattice dynamics. The inclusion of non-adiabatic effects
on top of an otherwise adiabatic calculation might still
be unsatisfactory. So other approaches to non-adiabatic behaviour
should be extended to the I phase \cite{uhrig98b}.
I thank C. Berthier, J.P. Boucher, B. B\"uchner, T. Lorenz,
M. Horvati\'c, Th. Nattermann
for fruitful discussions, F. Sch\"onfeld for reliable
numerical work, E. M\"uller-Hartmann for generous support and
G. G\"untherodt's group for intense collaboration.
Financial support of the DFG by the SFB 341 is acknowledged.
|
\section{Introduction} \label{sec:intro}
The conventional framework for particle physics beyond the standard
model (SM) assumes that the fundamental mass scale of nature is the Planck
mass: $\mpl \approx 10^{19}$ GeV. It is then natural to ask: why are the
masses of the
elementary particles so small? Proposed solutions to this
hierarchy problem have a common feature: new non-perturbative gauge
interactions dynamically generate a much lower scale, $M_{dyn}$, from which
electroweak symmetry breaking is generated, and hence all the masses
of the known elementary particles. Schematically, this mass hierarchy is
\begin{equation}
\mpl \rightarrow M_{dyn} \rightarrow M_W \,\,\ldots\,\, m_e.
\label{eq:masshier}
\end{equation}
In supersymmetric theories, $M_{dyn}$ is the scale at which
supersymmetry is broken, and the triggering of electroweak symmetry
breaking may be mediated, for example, by gravitational-scale physics,
or by gauge
interactions at much lower energy scales. Alternatively, $M_{dyn}$ may
be the scale of a new gauge force, technicolor, which forms fermion
condensates that directly break $SU(2) \times U(1)$. Finally, new strong
gauge forces could bind a composite Higgs boson.
Recently an alternative framework has been proposed~\cite{lowmpl} in which
spacetime is enlarged to contain large extra compact spatial dimensions.
At distances smaller than the size of these extra dimensions the
gravitational force varies more rapidly than the inverse square law, so
that the fundamental mass scale of gravity can be made much smaller
than $\mpl$. The conventional mass hierarchy of (\ref{eq:masshier})
is completely avoided if this fundamental mass scale is of order the weak
scale. In this case,
the length scale of the extra dimensions is much larger than the scales
probed experimentally at colliders, and hence this framework requires
that the quarks, leptons and gauge quanta of the SM are
spatially confined to a $3+1$ dimensional sub-space of the enlarged
spacetime.
The physics at the fundamental scale, $\Lambda$, which may well be that of
string theory, will be directly accessible to colliders of
sufficiently high energy; but even at lower energies this physics may
be experimentally probed.
At energies below the fundamental mass scale, physics is described by
an effective Lagrangian, which we take to be the most general set of
$SU(3) \times SU(2) \times U(1)$ invariant operators
involving quark, lepton and Higgs doublet fields of the SM:
\begin{equation}
\CL_{ef\!f} = \CL_{S\!M} + \sum_i {c_i \over \Lambda^p} \CO_i^{4+p}
\label{eq:leff}
\end{equation}
where $\CL_{S\!M}$ is the SM Lagrangian, $i$ runs over all
gauge invariant operators, $\CO_i^{4+p}$, of dimension $4+p$ with $p\geq1$, and
$c_i$ are unknown dimensionless couplings.
In this letter we study consequences of several of the dimension-6
operators. First we derive bounds on the $c_i/\Lambda^2$ from existing
experimental results under very conservative assumptions about
flavor-breaking in the ultraviolet theory. We then re-examine the
precision electroweak bounds on the Higgs boson mass. Analyses within
the standard model find a light Higgs; however, we will show that such
results do not survive the addition of non-renormalizable operators,
even if those operators are suppressed by scales as large as
$11\tev$.
In theories with large extra dimensions there is no good argument for
a light Higgs over a heavy Higgs or a non-linearly realized $SU(2)
\times U(1)$ symmetry, in which case (\ref{eq:leff}) must be replaced by a
chiral Lagrangian.
Finally we examine two operators in particular and their
effects on the discovery of Higgs bosons:
\begin{eqnarray}
\CO_G &=& \phi^\dagger \phi\, G^a_{\mu\nu}G^{a\mu\nu}
\label{eq:hop1} \\
\CO_\gamma&=&\phi^\dagger \phi\, F_{\mu\nu}F^{\mu\nu}
\label{eq:hop2}
\end{eqnarray}
where $G^a_{\mu\nu}$ and $F^{\mu\nu}$
are the QCD and electromagnetic field strengths, and
$\phi$ is the Higgs doublet with Re$\,\phi^0=(v+h)/\sqrt{2}$.
The first operator contributes to Higgs production at hadron colliders via
gluon-gluon fusion, and the second to Higgs decay to $\gamma \gamma$.
There are two reasons why these effects provide a significant
discovery potential for extra dimensions: first, they are competing against a
SM signal which is suppressed by loop factors, and second, the
SM $\Gamma(h\to\gamma\gamma)$ is further suppressed by $e^4
\simeq 10^{-2}$, where $e$ is the electromagnetic coupling
constant.
However we assume that the physics at scale $\Lambda$ which
generates (\ref{eq:hop1})--(\ref{eq:hop2}),
does so in a way that the coefficients are
not suppressed by powers of the SM gauge coupling constants (see also
\cite{bdn}).
Such a behavior is certainly {\em not}\/ expected if the theory at
$\Lambda$ is a 4-dimensional gauge field theory: in that case operators of
the form (\ref{eq:hop1})--(\ref{eq:hop2}) would arise by integrating
out heavy fields, but these fields must couple to $F_{\mu\nu}$ and
$G_{\mu\nu}$ with the usual SM gauge couplings, and further, as shown
in $\cite{einhorn}$, they must be also be loop-suppressed. Thus even
if the gauge theory at $\Lambda$ were strongly-coupled, it seems
unlikely that coefficients of $\CO(1)$ could be generated.
This is very
important --- the effect of the interaction $(e^2/ \Lambda^2)
\phi^\dagger \phi F^2$ on the $h \rightarrow \gamma \gamma$ branching
ratio has been studied, and is small for $\Lambda \geq 1\tev$~\cite{hagiwara}.
Thus observation of the physics we will describe in
Section~\ref{sec:expt} would provide support for
an extra-dimensional theory.
\section{Some Constraints on $\Lambda$}
Are the coefficients $c_{G, \gamma}/ \Lambda^2$ expected to be large
enough for an observable $h \rightarrow \gamma \gamma$ signal? In
general this cannot be excluded, since physics induced by operators
$\CO_i$ will place bounds on
\begin{equation}
\frac{f_i}{\Lambda_i^p} \equiv \frac{c_i}{\Lambda_{\phantom{i}}^p}
\quad\quad\quad (f_i=\pm1)
\end{equation}
not on $\Lambda$. However, it would be unreasonable to expect $c_{G,
\gamma}$ to be orders of magnitude larger than all the other $c_i$.
It is tempting to assume that although the dimensionless coefficients
$c_i$ are unknown, they are all of order unity. However, in this case
operators which violate baryon number constrain $\Lambda \gsim 10^{16}
\gev$, and $CP$ violating operators contributing to $\epsilon_K$
constrain $\Lambda \gsim 10^{5}
\gev$. Thus the framework of large compact extra dimensions, allowing a
fundamental scale close to the weak scale, is clearly excluded unless
the low energy effective theory possesses an approximate flavor
symmetry, in which case one expects
\begin{equation}
c_i = \varepsilon_{Fi} \; c_i'
\label{eq:cprime}
\end{equation}
with $c_i'$ of order unity. The flavor symmetry breaking parameters,
$\varepsilon_{Fi}$, depend on the flavor symmetry group and the pattern
of flavor symmetry breaking. For operators which violate flavor and $CP$
they must be small, while for operators which conserve flavor and $CP$
they may be set to unity.
To allow low values for $\Lambda$, the flavor group should be large,
and its breaking should be kept to a minimum, consistent with the
observed quark and lepton masses and mixings. The maximum flavor
group of the SM is $U(3)^5$. The three
generations of quarks and leptons transform as $q_L = (u_L, d_L) \sim
(3,1,1,1,1)$; $u_R \sim (1,3,1,1,1)$; $d_R \sim (1,1,3,1,1)$;
$\ell_L = (\nu_L, e_L) \sim (1,1,1,3,1)$; $e_R \sim (1,1,1,1,3)$. If
there are only three symmetry breaking parameters, one for each of the
up, down and charged lepton mass matrices, $\varepsilon_u
\sim(3,\bar{3},1,1,1)$; $\varepsilon_d \sim(3,1,\bar{3},1,1)$; $\varepsilon_e
\sim(1,1,1,3,\bar{3})$, then baryon number and
lepton number remain unbroken. (The $\varepsilon_i$ are equal to the
Yukawa couplings up to an $\CO(1)$ factor, $c_i$:
$\lambda_{u,d,e}=c_{u,d,e}\varepsilon_{u,d,e}$.)
However, even after imposing such a
flavor symmetry, there remain operators such as
\begin{equation}
\CO_{qq} = (\bar{q}_L \gamma^\mu \varepsilon_u \varepsilon_u^\dagger q_L)^2
=c_u^4(\bar q_L\gamma^\mu \lambda_u\lambda_u^\dagger q_L)^2
\label{eq:oqq}
\end{equation}
which contribute to $\epsilon_K$ and
constrain $\Lambda \gsim 4.2 \tev\times (\sqrt{c_{qq}}/c_u^2) $. There are
two ways to avoid this bound. First, since the bound depends
quadratically on $c_u$, values slightly larger than 1 will
weaken the bound significantly;
this seems entirely natural to us. Second, one could
postulate that $\varepsilon_{u,d}$ are real and the observed
$\epsilon_K$ has an exotic origin; we view this as disfavored
given that measurements of $V_{ub}/V_{cb}$ and $B-\bar B$ mixing
indicate values of the CKM matrix elements consistent with a standard
model origin of $\epsilon_K$ to better than $30\%$.
For the $h \rightarrow \gamma \gamma$ signal, we are interested in the
operators (\ref{eq:hop1})--(\ref{eq:hop2}),
which conserve $U(3)^5$. Hence,
even if the higher dimension flavor violating operators, such as
(\ref{eq:oqq}), are completely absent, it is important to study
constraints on $\Lambda$ expected from operators which conserve
$U(3)^5$. Such operators include
flavor-conserving four-fermion operators and operators involving the
Higgs doublet and the gauge fields.
There have been many analyses to date which obtain
constraints from these operators, and here we will simply repeat
the results of these analyses, in the notation we are using for
$\Lambda$.
(An analysis similar to ours was recently presented in~\cite{bdn}.)
Among the $CP$-conserving
four-fermion operators, the strongest constraints come from
atomic parity violation. The operator
\begin{equation}
\CO_{\ell q} = (\bar \ell_L\gamma_\mu \ell_L)(\bar q_L\gamma^\mu
q_L)
\label{eq:olq}
\end{equation}
gives a constraint $\Lambda_{lq} >3.0\tev$~\cite{apv} at 95\% CL.
If the operator $(\bar e_R\gamma_\mu e_R)(\bar q_L\gamma^\mu
q_L)$ were generated with the same coefficient, $P$ would be preserved
in atomic systems and the previous limit would vanish.
Although we do not expect $P$ to be a good symmetry of the underlying
theory, a partial cancellation could easily weaken this bound.
Apart from $P$-violation, the best bounds on $\Lambda_{\ell q}$
currently come from OPAL~\cite{eebb},
using the $\ell_1, q_3$ component, and from CDF~\cite{mumuqq}, using
the $\ell_2, q_1$ component. Both find $\Lambda>800\gev$ at
95\% CL.
The bounds on the coefficients of the operators $\CO_{qq, \ell q}$ of
(\ref{eq:oqq})--(\ref{eq:olq}) do not provide strict bounds on the scale
$\Lambda$, because $\Lambda = \Lambda_i \sqrt{c_i}$, and the $c_i$ are
unknown. Nevertheless, if the (flavor-conserving)
$c_i'=c_i$ are expected to be of order unity for
these operators, then $\Lambda \gsim 3 \tev$ is clearly allowed, while a
value of $\Lambda$ as low as $1 \tev$ seems disfavored.
\section{Precision Electroweak Physics and the Higgs Mass Bound}
A second class of constraints arise from precision measurements in the
electroweak gauge sector, namely from the $S$ and $T$ parameters (see,
\eg, \cite{hagiold}). The
strongest of these constraints arise from the operators:
\begin{eqnarray}
\CO_{BW}&=& B^{\mu \nu} (\phi^\dagger \tau^a
W^{a\mu\nu} \phi) \label{obw} \\
\CO_{\Phi}&=& (\phi^\dagger D^\mu\phi)
(D_\mu\phi^\dagger\phi) \label{op1}
\end{eqnarray}
which contribute
\begin{eqnarray}
\Delta S_{new} &=& -\frac{2c_Ws_W}{\alpha}\frac{v^2}{\Lambda_{BW}^2}f_{BW} \\
\Delta T_{new} &=&
-\frac{1}{2\alpha}\frac{v^2}{\Lambda_{\Phi}^2}f_{\Phi}
\end{eqnarray}
where $s_W,c_W$ are the sine and cosine of the weak angle and
$f_{BW}$, $f_\Phi$ are unknown signs.
A global fit to electroweak observables~\cite{erler} yields
$S_{f\!it}=-0.14\pm0.12$ and $T_{f\!it}=-0.22\pm0.15$ assuming
$m_h=100\gev$.~\footnote{The fit in \cite{erler} uses $m_h=600\gev$ and
defines $S=T=0$ in the SM. We rescale to $m_h=100\gev$ using the
parameterization of Ref.~\cite{hagiold} (see
Eqs.~(\ref{dS})--(\ref{dT})). We then treat deviations from
$m_h=100\gev$ as ``new physics.''}
Since each operator contributes only to one of $S$ or $T$, we can find
independent bounds on each. We find that at 95\% CL:
\begin{eqnarray}
\Lambda_{BW}&>& 3.6\tev \label{bwbound}\\
\Lambda_{\Phi}&>& 3.0 \tev.
\end{eqnarray}
We can also extract a bound if
$\Lambda_{BW}=\Lambda_{\Phi}$:
$\Lambda > 4.0\tev$,
allowing the Higgs mass to vary over the range
$100\gev<m_h<800\gev$. We see that the constraints from precision
electroweak physics are very similar in magnitude to those obtained in
the previous section.
How important are these constraints for restricting $\Lambda_\gamma$?
Although the electromagnetic field strength, $F^{\mu \nu}$, is not $SU(2)\times
U(1)$ invariant, the operator $\CO_\gamma$ is generated, after
electroweak symmetry-breaking, from the
invariant operators $\CO_B=
(\phi^\dagger\phi)B_{\mu\nu}B^{\mu\nu}$,
$\CO_W=(\phi^\dagger\phi)W_{\mu\nu}W^{\mu\nu}$ and $\CO_{BW}$ of
Eq.~(\ref{obw}):
\begin{equation}
\frac{f_\gamma}{\Lambda_\gamma^2}=
c_W^2\frac{f_{B}}{\Lambda_{B}^2}+s_W^2 \frac{f_{W}}{\Lambda_{W}^2}+
c_Ws_W\frac{f_{BW}}{\Lambda_{BW}^2}.
\label{foper}
\end{equation}
{\em If}\/ all $f_i$ and $\Lambda_i$ on the right side of Eq.~(\ref{foper})
were equal, then the bound (\ref{bwbound})
on $\Lambda_{BW}$ implies
$\Lambda_\gamma > 3.3\tev$. However, changes in the relative signs
or sizes of each contribution significantly reduces the bound;
thus we have no strong lower bound on the scale $\Lambda_\gamma$ itself.
Likewise we know of no strong constraint on the scale $\Lambda_G$ either.
Finally we wish to address the question of the Higgs mass. It is
well-known that fits to the electroweak data indicate a light Higgs. A
simple fit can be done using only $S$ and $T$ as given above and the
following parameterization of the Higgs contributions from
Ref.~\cite{hagiold}:
\begin{eqnarray}
\Delta S_H&=&0.091 x_H - 0.010 x_H^2 \label{dS}\\
\Delta T_H&=&-0.079 x_H-0.028 x_H^2+0.0026 x_H^3 \label{dT}
\end{eqnarray}
where $x_H=\log(m_h/100\gev)$. Using these forms, one can do a fit demanding
$S_{f\!it}=\Delta S_H+\Delta S_{new}$ and likewise for $T$.
For the SM alone, a 95\% CL upper bound of $255\gev$ has been
obtained~\cite{erler}. However it is clear that from the point of view
of the oblique parameters, shifts in $\Delta S_H$ and $\Delta T_H$ can
be compensated by similar shifts in $\Delta S_{new}$ and $\Delta
T_{new}$. Thus we can derive an effective
``95\% CL bound'' on the Higgs mass as
a function of $\Lambda$ under the requirement that the fit to the
experimentally obtained $S_{f\!it}$ and $T_{f\!it}$ be no worse than
that obtained for $m_h=255\gev$ and $\Lambda\to\infty$. (We do this by
constructing a $\chi^2$ distribution from $S$ and $T$ alone.)
How large can the Higgs mass become with
the inclusion of $\CO_{BW}$ and $\CO_{\Phi}$? The answer is: quite large.
Fitting to $m_h$ as a function of $\Lambda$ and using $S$ and $T$ as
``experimental'' inputs, we find
for particular choices of the signs of the operators (\ie, $f_{BW}=f_\Phi=+1$)
that {\em the precision electroweak bound on the Higgs mass disappears
completely for $4\tev\lsim\Lambda\lsim 11\tev$!} (By ``disappear'' we
mean that the 95\% upper bound on $m_h$ exceeds the unitarity bound of
approximately $800\gev$ and so is meaningless.) Thus, in the context
of gravitational physics at or below $10\tev$, the usual claims that
electroweak physics prefers
a light Higgs do not hold. And even for $\Lambda$ as high as $17\tev$,
the upper limit on the Higgs mass exceeds $500\gev$.
These results are
summarized in Figure~\ref{massfig} where we show the 95\% CL allowed
range for $m_h$ as a function of $\Lambda\equiv\Lambda_{BW}=\Lambda_{\Phi}$.
The hatched region at small $\Lambda$ is ruled out because of its large
contribution to $S$ and $T$, while the region at large $\Lambda$ and
large $m_h$ is ruled out because the new operators contribute too
little to $S$ and $T$ to significantly effect the SM fit to the Higgs
mass. However for intermediate $\Lambda$ (unhatched region)
it is clear that there is
effectively no limit on the Higgs mass thanks to the effects of the
new operators.
\begin{figure}[t]
\centering
\epsfxsize=4truein
\epsffile{higgsmass.eps}
\caption{Precision electroweak limits on the Higgs mass as a function
of the scale of new physics. For this figure, $\Lambda_{BW}$ and
$\Lambda_{\Phi}$ are chosen equal, while the signs $f_{BW}$ and
$f_\Phi$ are chosen to
maximize the allowed region. Hatched regions are disallowed at
95\%, while the dashed line borders the region allowed in the
SM alone.}
\label{massfig}
\end{figure}
(If the physics at $\Lambda$ were weakly-coupled then we would expect
that $c_{BW}\simeq e^2c_Ws_W$; then allowing $c_\Phi\simeq1/4$ would
reproduce Fig.~\ref{massfig}, only with the $\Lambda$ rescaled by
$\sim1/2$. Thus the preference for a light Higgs in the SM is even
removed for a weakly-coupled gauge theory if $\Lambda\sim 2-5\tev$.)
Finally, we note that the one other argument for a light Higgs, namely
triviality, is no longer applicable in these models either. With such
a low ultraviolet cutoff ($\Lambda\sim$ few TeV), the Higgs
self-coupling cannot run to its Landau pole for $m_h\lsim 1\tev$.
\section{Implications for Electroweak Symmetry Breaking}
The mechanism for electroweak symmetry breaking (EWSB) is
unknown. Nevertheless, it is commonly believed that the Higgs boson
exists, and is light. The two indirect indications for this are:
\begin{itemize}
\item The successful prediction of the weak mixing angle from gauge
coupling constant unification. This prediction results in theories
with weak scale supersymmetry which are perturbative to a high scale;
such theories have a light Higgs boson, $m_h \lsim 150 \gev$~\cite{kkw}.
\item The experimental values of the precision electroweak observables
are consistent with the standard model, at 95\% C.L., only if
$m_h \lsim255 \gev$~\cite{erler}.
\end{itemize}
If there are large extra dimensions allowing the fundamental scale,
$\Lambda$, to be in the TeV domain, neither of these points can be used
to argue that the Higgs boson is light. For the first: it has not
been demonstrated that it is
possible to predict the weak mixing angle to the percent level of
accuracy in these theories; furthermore, there is no need for the field theory
below $\Lambda$ to be supersymmetric
since there is no large hierarchy between the weak scale and $\Lambda$.
The argument from fits to the precision electroweak observables
applies only if the standard model is the correct theory up to scales
of at least 10 TeV; it is a very weak bound which is immediately
evaded by large extra dimensions, allowing several scenarios
for EWSB:
\begin{itemize}
\item {\em Light Higgs} ($m_h < 200 \gev$): For $\Lambda \gsim 20 \tev$
some protection mechanism for the Higgs mass would be required; if
this is supersymmetry, the Higgs will be light.
For $ \Lambda \approx 1 - 3 \tev$, if the tree
level Higgs mass happened to vanish, EWSB and a light Higgs boson
could result from 1 loop radiative corrections.
\item {\em Heavy Higgs} ($m_h > 200 \gev$): This could arise for
$\Lambda \approx 1 - 3$ TeV, if the Higgs mass parameter is somewhat less
than $\Lambda$, or alternatively for $\Lambda \approx 3 - 10 \tev$ if
the Higgs mass parameter vanishes at tree level but arises at 1
loop. In both cases a large value for the Higgs self coupling is
needed, and the operators (\ref{obw}) and (\ref{op1}) must mimic
the effects of a light Higgs in the $S$ and $T$ parameters.
\item{\em No Higgs}: Physics at the fundamental scale $\Lambda \approx
1 - 3 \tev$ may itself cause EWSB. An example of this has already
been proposed \cite{ad}. In this case the theory below $\Lambda$
will have $SU(2) \times U(1)$ realized non-linearly, and the chiral
Lagrangian will
have operators analagous to (\ref{obw}) and (\ref{op1}) which mimic
the effects of a light Higgs in the $S$ and $T$ parameters.
\end{itemize}
A light Higgs boson is just one
possibility amongst several for EWSB, and is not
preferred.
We have shown that, in theories with large extra dimensions having
$\CO_{BW,\Phi}$ with $c_{BW,\Phi}$ of order unity, the precision
electroweak data provide a lower bound on the fundamental scale,
$\Lambda_{min}\approx3\tev$. For values of $\Lambda$ in the range
(1--3)$\times\Lambda_{min}$, the signs $f_{BW,\Phi}$ are critical. For
two sign choices, no successful fit can be found for any Higgs
mass. For a third choice, a good fit to the data is found for Higgs
masses all the way up to $m_h=800\gev$. For the final choice, masses
up to $800\gev$ are also obtained, though the fits are less convincing.
Only in the case of very large $\Lambda$ does the data still prefer a
light Higgs, but then the quadratic finetuning of the light Higgs mass
to one part in $m_h^2/\Lambda^2$ is reintroduced.
In view of the bounds on $\Lambda_{min}$ of $3 - 4 \tev$ from each of
$\CO_{qq}$ (\ref{eq:oqq}),$\CO_{\ell q}$ (\ref{eq:olq}), $\CO_{BW}$
(\ref{obw}), and $\CO_{\Phi}$ (\ref{op1}), it may be felt that the
exciting possibility of $\Lambda$ in the $1 - 3 \tev$ range is
unlikely. Why would all the relevant $c_i$ coefficients be small? One
possibility is that the dominant interactions of the new physics at
$\Lambda$ preserve symmetries that are broken by the electroweak gauge
interactions, including $P$, $CP$ and custodial $SU(2)$. If these
symmetries are broken by sub-dominant interactions at $\Lambda$,
then the smallness of the relevant $c_i$ can be naturally explained.
\section{Higgs Production and Decay} \label{sec:expt}
For the case that there is a Higgs boson, either light or heavy,
we now study the effects of $\CO_{G,\gamma}$ of
(\ref{eq:hop1})--(\ref{eq:hop2}) on the
signal for $h \rightarrow \gamma \gamma$ at hadron colliders.
These operators have two immediate consequences.
First, when both Higgs fields are set to their vacuum expectation
values (vev's), the gauge
couplings of QED and QCD are shifted. But these shifts can be reabsorbed
into the definition of the gauge couplings and therefore have no
observable implications. (If one attempts to unify the SM gauge
couplings at some ultraviolet scale, or otherwise define theoretical
relations among them, then these shifts will enter into the relation
between the theoretical couplings and those extracted from
data. However, for all but the lightest $\Lambda$, this
shift is smaller than the experimental uncertainties.)
The second consequence is the possibility of unusual production and decay modes
of the (physical) Higgs bosons. Taking one of the Higgs fields to its vev, one
obtains terms in the effective Lagrangian:
\begin{eqnarray}
\CL_{ef\!f} = \cdots +
f_\gamma \frac{v}{\Lambda_\gamma^2}\,h\,F_{\mu\nu}F^{\mu\nu} +
f_G \frac{v}{\Lambda_G^2}\,h\,G^a_{\mu\nu}G^{a\mu\nu} +\cdots
\label{eq:hleff}
\end{eqnarray}
where $h$ is the physical Higgs boson,
$v=246\gev$ and $f_{\gamma,g}= \pm1$ are unknown signs.
First, $\CO_G$ can contribute to the gluon fusion process $gg\to
h$. It is well-known that the dominant production mode for Higgs
bosons at the Tevatron and the LHC is
through gluon fusion, via a loop of $t$-quarks. Because the process
occurs at one-loop, non-renormalizable operators are more likely to provide a
significant correction to the cross-section. Integrating out the
$t$-quark, the relevant low-energy operator is then (for a recent discussion of
the relevant SM Higgs physics, see~\cite{nlo}):
\begin{equation}
\CL_{G,ef\!f}=\left(-\frac{g\alpha_s}{24\pi
M_W}I_G+f_G\frac{v}{\Lambda_G^2} \right)\, h\,G^a_{\mu\nu}G^{a\mu\nu}
\end{equation}
where $g$ is the SU(2) coupling constant and
$I_G\to1(0)$ for $m_t^2\gg m_h^2$ $(m_t^2\ll m_h^2)$.
For $\Lambda\lsim4.5\tev$, the new
physics will actually dominate the production of Higgs bosons. Note that
the cross-section is maximized for constructive interference, $f_G=-1$,
and minimized for $f_G=+1$.
The operator $\CO_\gamma$ does not contribute to Higgs
production\footnote{However, a large coefficient to $\CO_\gamma$ could turn
the NLC into an $s$-channel Higgs factory when run in $\gamma\gamma$ mode.}.
However it can contribute to the decay of the Higgs into photons:
\begin{equation}
\Gamma(h\to\gamma\gamma)=\frac{|\beta|^2 m_h^3}{4\pi}
\end{equation}
for $\CL=\beta h F_{\mu\nu}F^{\mu\nu}$.
In the SM, this process is dominated by loops of
$W$-bosons and $t$-quarks. Integrating them out yields an effective
operator:
\begin{equation}
\CL_{\gamma,ef\!f}=\left(-\frac{g\alpha}{4\pi M_W}I_\gamma+
f_\gamma\frac{v}{\Lambda_\gamma^2}\right)\, h\,F_{\mu\nu}F^{\mu\nu}
\end{equation}
where $I_\gamma$ varies from roughly $-0.5$ to $-1.3$ as $m_h$ is
varied. Once again, the new physics will dominate the width
for $h\to\gamma\gamma$ given $\Lambda_\gamma\lsim 7\tev$. If $m_h\lsim
150\gev$, its decay width is dominated by final state $b$-quarks; then
$h\to\gamma\gamma$ becomes the dominant decay mode given
$\Lambda_\gamma\lsim 1.5\tev$. However, even for larger
$\Lambda_\gamma$, the branching ratio $h\to\gamma\gamma$ may be more
than sufficient to provide a strong signal. The signal is maximized
for $f_\gamma=+1$ (\ie, constructive interference of the SM and new physics)
and minimized for $f_\gamma=-1$.
(In the context of LEP, Ref.~\cite{eboli} recently examined the effect of
$\CO_\gamma$ and related operators on $e^+e^-\to 3\gamma, qq\gamma\gamma$ and
found sensitivity there to new physics roughly below a scale $\Lambda\lsim
600\gev$.)
Unfortunately, the operator $\CO_G$ can also contribute to the Higgs
decay width via $h\to gg$ which is unobservable among the QCD backgrounds.
In fact, to lowest order,
\begin{equation}
\Gamma(h\to gg)=8\left(\frac{\Lambda_\gamma}{\Lambda_G}\right)^4
\Gamma(h\to\gamma\gamma).
\end{equation}
In the limit in which the new physics is dominating the Higgs decays
and $\Lambda_\gamma\simeq\Lambda_G$, the $h\to gg$ decays suppress the
branching ratio into $h\to\gamma\gamma$ by about a factor of
10. However, once final state $WW/ZZ$ dominate the Higgs width, the decays to
gluons provide no real additional suppression of the
$h\to\gamma\gamma$ branching fraction. Finally we note that the
interference of $\CO_G$ with the SM gives simultaneously
larger (smaller) Higgs cross-sections and larger (smaller) $\Gamma(h\to gg)$.
The sensitivity of any experiment to new physics in the Higgs channel
is then a function of several variables:
$m_h$, $f_\gamma$, $f_G$, $\Lambda_\gamma$ and $\Lambda_G$.
There are four sign choices for $f_\gamma, f_G$; we choose to study
the two cases which maximize/minimize the signal at
current and future colliders. The maximum signal case has $f_\gamma=+1$ and
$f_G=-1$; we checked that over the entire range of interest the
increase in the cross-section implied by $f_G=-1$ more than offset the
corresponding increase in $Br(h\to gg)$. The minimum signal case
has the opposite choice of both signs.
Our analysis then has two parts. First we ignore the $\CO_G$ operator
(\ie, $f_G=0$) and
determine the sensitivity of current and future experiments to new physics
through $\CO_\gamma$ alone. In this case, the production cross-section
is simply that of the SM.
Then in a second analysis we include both $\CO_\gamma$ and $\CO_G$.
As we already noted, the effect of $\CO_G$ is both to enhance the production
but also to diminish the relative branching ratio of
$h\to\gamma\gamma$.
For the purposes of doing the numerical calculations, we have used (in
a greatly modified form) the programs of M.~Spira and
collaborators~\cite{spira}. In all cases, we will work only to leading
order. In the SM it has been found that NLO QCD corrections can change
the cross-sections and decay widths by $\sim60\%$~\cite{nlo}. Naively such
changes appear to correspond only to $\sim10\%$ shifts in $\Lambda$, which are
too small for the physics we are interested in here. However, it is
possible that interference effects and enhanced backgrounds (\ie,
$h\to\gamma\gamma$ in the SM) could produce a larger effect ---
we will not consider that possibility here.
Throughout our analysis we also have to address issues of acceptances and
backgrounds in an approximate manner. In Run I, CDF reported an
efficiency times acceptance approaching 15\% in
inclusive $\gamma\gamma+X$ Higgs searches~\cite{cdf}; we will assume
that this figure prevails at all future facilities. There are also two
major sources of backgrounds for our $\gamma\gamma$ signal: SM
processes which produce or fake $\gamma\gamma$, and the usual SM decay of
$h\to\gamma\gamma$ itself. The latter can be calculated explicitly.
For the former we estimate by fitting to the CDF background
spectrum~\cite{cdf}, appropriately scaled to the luminosity of future Tevatron
runs, or the ATLAS background spectrum~\cite{atlas} appropriately scaled for
LHC runs.
In Figures~\ref{fig1}(a)-(b)
we show the sensitivity to $\Lambda_\gamma$ that can be
obtained at various machines by plotting their $5\sigma$ discovery reaches
(with no $\CO_G$ contribution). The colliders shown are: the
Tevatron with $\sqrt{s}=1.8\tev$ and $100\invpb$ of luminosity (Run I), with
$\sqrt{s}=2\tev$ and $2\invfb$ of luminosity (Run II), with $\sqrt{s}=2\tev$
and $30\invfb$ (a proposed Run III), and the LHC with $\sqrt{s}=14\tev$ and
$10\invfb$ (initial luminosity) and $100\invfb$ (final luminosity)
respectively. (Note that the TeV Run I line falls below the region of
parameter space plotted.)
As one expects, once the $h\to WW,ZZ$ threshold opens up at $\sqrt{s}\simeq
150\gev$, the large $\Gamma(h\to WW,ZZ)$ is sufficient to overwhelm the
photonic width and our experimental sensitivity drops significantly.
Nonetheless, given the possibility of a light Higgs (and the robust
arguments for one in supersymmetric frameworks) experimentalists
should be encouraged
to view $h\to\gamma\gamma$ as a viable and potentially large signal.
\begin{figure}
\centering
\epsfxsize=6truein
\epsffile{limits-5sig-gam.eps}
\caption{$5\sigma$ discovery reaches for $pp,p\bar p\to h\to\gamma\gamma$ in
current and future colliders. Only the $\CO_\gamma$
operator has been included.
In (a), signs are chosen to maximize the signal, while they
are chosen to minimize the signal in (b).}
\label{fig1}
\end{figure}
In terms of extracting a conservative discovery reach for $\Lambda$,
Figure~\ref{fig1}(b)
should be used since it chooses $f_\gamma$ in order to minimize
the signal. We note, for example, that the data from Run I cannot
presently probe (or exclude) $\Lambda$ above $1\tev$,
but that Run II should have
a reach of approximately 1 -- $1.5\tev$ for a light Higgs. However it is
important to realize that for generic $f_\gamma$, the various colliders may
have reaches as high as those shown in Figure~\ref{fig1}(a). Thus, for example,
if the Higgs mass is below the $WW$ threshold, the LHC can possibly find a
signal for $\Lambda$ up to $8\tev$ for a light Higgs! (Unfortunately, that
scale could also be as low as $4\tev$.)
Figures~\ref{fig2}(a)-(b) repeat the same analysis, but now with $\CO_G$
included such that $\Lambda_G=\Lambda_\gamma\equiv\Lambda$.
We view these results
as more realistic compared to those above in which only the
$\CO_\gamma$ operator was kept. We again show the same
set of 5 collider options. Figure~\ref{fig2}(b) is the conservative $5\sigma$
discovery reach, chosen to minimize the $pp,p\bar p\to h\to\gamma\gamma$ rate.
It is interesting
that for a light Higgs, the limits are slightly stronger than those
obtained with $f_G=0$; now even the Tevatron Run I data has the ability to
probe scales above $1\tev$. However the more noticable difference is the
ability to produce larger numbers of heavy Higgs bosons and observe their
$\gamma\gamma$ decays. For example, the LHC is capable of probing scales near
$2\tev$ even for $m_h=1\tev$.
\begin{figure}
\centering
\epsfxsize=6truein
\epsffile{limits-5sig-glu.eps}
\caption{$5\sigma$ discovery reaches for $pp,p\bar p\to h\to\gamma\gamma$ in
current and future colliders. Both $\CO_G$ and $\CO_\gamma$ have been
included, with $\Lambda_\gamma=\Lambda_G\equiv\Lambda$.
In (a), signs are chosen to maximize the signal, while they
are chosen to minimize the signal in (b).}
\label{fig2}
\end{figure}
Figure~\ref{fig2}(b) shows the maximal reach of the various
colliders, with the LHC now extending its sensitivity to $\Lambda$
as high as $10\tev$ for a light Higgs! Finally, we summarize a few of our
results for $m_h=110$, $200$ and $500\gev$ for both exclusion and discovery in
Table~\ref{table1}. All bounds assume $\Lambda_\gamma=\Lambda_G$. For each
choice of the Higgs mass, we have shown a conservative limit on $\Lambda$ which
can be excluded, and a maximum $\Lambda$ below which a signal may be
discovered. Thus for
the exclusion bounds ($2\sigma$) we have taken the interference effects to
minimize the signal; for the maximum discovery reaches ($5\sigma$), we have
chosen the interference effects to maximize the signal.
\begin{table}
\centering
\begin{tabular}{l|cc|cc|cc}
\multicolumn{1}{c}{~} & \multicolumn{6}{c}{$m_h$ (GeV)} \\
\multicolumn{1}{c|}{$\Lambda$ (TeV)} & \multicolumn{2}{c}{110} &
\multicolumn{2}{c}{200} & \multicolumn{2}{c}{500} \\ \hline
Tev Run I & 2.0 & 1.8 & 1.1 & --- & --- & --- \\
Tev Run II & 2.6 & 3.0 & 1.5 & 1.3 & --- & --- \\
Tev Run III & 3.0 & 4.2 & 1.8 & 1.8 & 1.1 & --- \\
LHC ($10\invfb$) & 3.4 & 7.2 & 2.9 & 3.5 & 2.3 & 2.1 \\
LHC ($100\invfb$) & 3.5 & 10.8& 3.2 & 5.8 & 2.9 & 2.9 \\
\end{tabular}
\label{table1}
\caption{Exclusion limits and maximum discovery reaches (in TeV) for various
collider runs for 3 representative Higgs masses. The first column for each
$m_h$ is a conservative $2\sigma$ exclusion reach for each machine; the second
column is the optimistic $5\sigma$ discovery reach. Unfilled columns represent
limits below $1\tev$. We take $\Lambda_G=\Lambda_\gamma$ for the table.}
\end{table}
We have attempted in this analysis to be rather conservative. For one thing,
the $2\sigma$ exclusion limits of the various colliders are often several TeV
higher than the $5\sigma$ discovery limits. Secondly, we have treated the
discovery of the $h\to\gamma\gamma$ signal as simply a counting experiment,
throwing away useful experimental information, for example on the shape of the
diphoton mass spectrum, which would be available experimentally to help extract
the signal from the backgrounds. Lastly, we have not included QCD corrections
to the amplitudes, which we believe could increase the signal (though also
increasing the ``background'' $h\to\gamma\gamma$ signal) by $\sim50\%$.
Therefore we believe that the reaches given here are to be taken as
conservative values, insofar as one should take the scales deduced from naive
power-counting seriously.
\section{Conclusions}
In this paper we have studied two consequences of large extra
dimensions for electroweak symmetry breaking: a relaxation of the precision
electroweak bound on the Higgs boson mass, and an enhanced rate for
$\gamma \gamma$ events at hadron colliders from Higgs decay.
The relaxation of the precision electroweak bound on the Higgs mass
applies when any new physics generates (\ref{obw}) and (\ref{op1}) at
a scale of several TeV. It is well known that
$S$ and $T$ depend only logarithmically on the Higgs boson mass,
but it may not be appreciated that the mass bound can be
evaded completely for a wide range of values of $\Lambda$, extending
as high as 10 TeV. For
example, even a weakened bound of $m_h < 500 \gev$, only applies if
the standard model is
the correct description of nature up to energies of $17\tev$.
We find this implausible, since it implies a fine tuning
in the Higgs mass squared parameter of 1 part in 2000. There is only one
strong argument for a light Higgs boson: the correct successful
prediction of the weak mixing angle at the percent level of accuracy
requires weak scale supersymmetry, and therefore a light Higgs
boson. In theories with large extra dimensions this argument is not
applicable, since the percent level prediction for the weak mixing angle
is lost. Hence, in these theories, there is no preference for a
light Higgs boson, and thus
alternatives with a heavy Higgs or no Higgs should be considered seriously.
If there is a Higgs boson, we have shown that a generic signal of
large extra dimensions is an anomalously large
$\gamma\gamma$ signal at machines capable of producing Higgs bosons.
Expectations from the SM put such a signal out of reach of the
Tevatron. In Figure~\ref{fig2}
we showed the $5 \sigma$ discovery reaches for $h
\rightarrow \gamma \gamma$ at the Tevatron and LHC. At Run II of the
Tevatron collider this signal would be discovered for a light Higgs if
$\Lambda$ is less than 2 (3) TeV for destructive (constructive) interference.
LHC not only increases the discovery potential for a light Higgs boson
mass, up to $10\tev$ for constructive interference, but also has significant
discovery potential up to the largest Higgs masses. This signal
compares favorably with that of graviton production at colliders
\cite{graviton}, especially if the scale which sets the size of the $4+n$
dimensional gravitational coupling is somewhat larger than the
scale $\Lambda$.
\section*{Acknowledgements}
We are grateful to Nima Arkani-Hamed, Michael Chanowitz, Savas Dimopoulos and
Henry Frisch for many
useful conversations. This work was supported in part by the U.S.\
Department of Energy under contract DE--AC03--76SF00098 and by the
National Science Foundation under grant PHY--95--14797.
|
\section{ Introduction }
The ultrarelativistic heavy ion collisions provide a means to
create a new state of matter at high temperature and density
as the quantum chromodynamics (QCD) predicts a phase transition from
normal hadronic matter to quark gluon plasma (QGP), a state of
unconfined quarks and gluons. In recent years a considerable amount
of theoretical and experimental activity is going on in this field.
After the collision of two Lorentz contracted nuclei,
QGP is assumed to be formed and equilibrated in a very small time
of the order of $\sim$ 1 fm.
In the usual description of the evolution of the plasma, it cools by
expanding hydrodynamically till the critical temperature $T_c$ is
reached at which a transition from QGP to normal
hadronic matter takes place.
The temperature remains fixed at $T_c$ until
hadronization gets completed. This hadronization scenario corresponds
to the ideal Maxwell construction. However, in reality hadronization
does not begin at $T=T_c$ due to the large nucleation barrier.
Recently Csernai and Kapusta have proposed a model for nucleation for
the relativistic first order phase transition \cite{CSER}.
Supercooling, through nucleation of hadronic bubbles in QGP has been
studied by several authors \cite{CSER,CSER1,CSER2,DKS} using
the Csernai-Kapusta model of nucleation.
A general outcome of these studies is that the plasma will cool
according to the law $T(\tau)=T_0 (\tau_0/\tau)^{1/3}$ till $T_c$.
The matter continues to cool below $T_c$ until the temperature
goes down to about $\sim$ .8 $T_c$, where bubble formation and
growth becomes sufficient to reheat the system due to the release of
latent heat.
Compared to the idealized Maxwell construction the
supercooling delays the transition and leads to an extra entropy
production as the nucleation process allows dissipation around the
hadronic bubbles. The dynamical prefactor \cite{CSER} in the
nucleation rate includes quark viscosity
coefficients which bring dissipative effect in the medium. In fact,
the nucleation rate is limited by the ability of the dissipative
processes to carry latent heat away from the bubbles's surface.
However, for the dynamical evolution of the plasma the
ideal hydrodynamics is used.
This is not consistent as the viscosity dependent terms in
hydrodynamics would also contribute to the entropy production.
Therefore, in this work we study the supercooling in the viscous
hydrodynamics. The role of viscosity on supercooling and entropy production
has been investigated in detail.
\section{ Csernai-Kapusta model of nucleation }
The nucleation model
computes the probability that a bubble of the
hadronic matter appears in a system, initially in QGP phase near the
critical temperature. Further in the model a small baryon chemical potential
is assumed. The bubble formation is assumed to take place
in a homogeneous QGP phase consisting of $u$ and $d$ quarks and gluons,
ignoring the role of inhomogenities such as strange quarks.
It is further assumed that there is not substantial supercooling.
Langer's theory of nucleation gives the nucleation rate per unit
volume at temperature $T$ as
\begin{eqnarray}\label{basic}
I=\frac{\kappa}{2\pi} \frac{\Omega_0}{V} e^{-\Delta F_*/T},
\end{eqnarray}
where $\Delta F_*$ is the change in the free energy of the system due to
the formation of a critical hadronic droplet, $\Omega_0$ is a
statistical prefactor
which measures the available phase volume and $V$ is the volume of the
system. The dynamical prefactor, $\kappa$ determines the
exponential growth rate of critical droplets which are perturbed from
their equilibrium radius $R_*$.
The coarse-grained effective field theory approximation to
QCD is utilised to obtain $\kappa$ and $\Omega_0$.
The factor $\Omega_0$ is obtained as
\begin{eqnarray}\label{omega}
\frac{\Omega_0}{V} = \frac{2}{3} \left(\frac{\sigma}{3T}\right)^{3/2}
\left(\frac{R_*}{\xi_q}\right)^4,
\end{eqnarray}
where $\sigma$ is the surface free energy and is determined as
50 MeV/fm$^2$ by lattice guage theory simulations without dynamical
quarks. The correlation length $\xi_q$ is estimated as 0.7 fm.
The dynamical prefactor $\kappa$ is obtained as
\begin{eqnarray}\label{kappa}
\kappa = \frac{4 \sigma (4/3 \eta_q + \zeta_q) }
{(\Delta \omega)^2 R_*^3},
\end{eqnarray}
where $\Delta \omega$ is the difference in the enthalpy densities
of the two phases. $\eta$ and $\zeta$ are respectively, the shear and
bulk viscosity coefficients.
Inserting $\Omega_0$ from Eq.~(\ref{omega}) and $\kappa$ from
Eq.~(\ref{kappa}) in Eq.~(\ref{basic}) we get the nucleation rate
per unit volume as
\begin{eqnarray}\label{rate}
I = \frac{4}{\pi} \left( \frac{\sigma}{3T} \right)^{3/2}
\frac{\sigma (4\eta_q/3 + \zeta_q) R_*}{\xi_q^4 (\Delta \omega)^2}
e^{-\Delta F_*/T}.
\end{eqnarray}
The critical radius $R_*$ is given by the Laplace formula as
\begin{eqnarray}\label{crit}
R_*(T) = \frac{2 \sigma}{p_h(T)-p_q(T)},
\end{eqnarray}
where $p_{q/h}$ is the pressure of the quark/hadron
phase at temperature $T$ and the $\Delta F_*$ as
\begin{eqnarray}\label{}
\Delta F_* = \frac{4\pi}{3} \sigma R_*^2.
\end{eqnarray}
From the nucleation rate Eq.~(\ref{rate}), one can calculate the fraction
of volume $h(\tau)$ which has been converted from QCD plasma to hadronic
gas at proper time $\tau$.
If the system cools to $T_c$ at time $\tau_c$, then at some
later time $\tau$ the fraction $h$ of space which has been converted
to hadronic gas is
\begin{eqnarray}\label{frac}
h(\tau) = \int_{\tau_c}^\tau d\tau' I(T(\tau')) [1 - h(\tau')] V(\tau',\tau).
\end{eqnarray}
Here $V(\tau',\tau)$ is the volume of a bubble at time $\tau$ which had
been nucleated at an earlier time $\tau'$; this takes into account
the bubble growth. The factor $1 - h(\tau')$ accounts for the fact that
new bubbles can only be nucleated in the fraction of space not already
occupied by the hadronic gas. The model for bubble growth is
simply taken as \cite{WEIN}
\begin{eqnarray}\label{}
V(\tau',\tau) = \frac{4\pi}{3} \left( R_*(T(\tau')) +
\int_{\tau'}^\tau d\tau'' v(T(\tau'')) \right) ^3,
\end{eqnarray}
where $v(T)$ is the velocity of the bubble growth at temperature $T$.
By definition a critical size bubble is metastable and will not grow
without a perturbation. The growth of bubbles has been studied
numerically with relativistic hydrodynamics by Miller and Pantano
\cite{MILLER}. Their results are consistent with the growth law
\begin{eqnarray}\label{}
v(T) = 3 c [1 - T/T_c]^{3/2}.
\end{eqnarray}
This expression is intended to apply only when
$T > \frac{2}{3} T_c$ so that the growth velocity stays below the speed
of sound of a massless gas, $c/\sqrt {3}$.
At the critical temperature, $R_* \rightarrow \infty$,
$\Delta F_* \rightarrow \infty$ and the rate of
nucleation vanishes. The system must
supercool at least $\sim$ 5 \% to attain a finite rate.
In the evolution of the matter from QGP to hadron phase, the temperature
varies with $\tau$ and its description by scaling hydrodynamics provides
another equation so that $h(\tau)$ and $T(\tau)$ are determined from
these equations at any proper time $\tau$.
\section{ Viscous hydrodynamics }
The scaling viscous hydrodynamics is discussed by
Danielewicz and Gyulassy \cite{DAN} and others. The form of the
dissipative terms depends on the choice of the definition of
what constitutes the local rest frame of the fluid. The
Landau-Lifshitz definition is appropriate for describing systems
with small (or zero) chemical potential. Here we give a simple
derivation of the viscous hydrodynamics. In scaling hydrodynamics
the expansion takes place only along the direction of collision
which we chose as z axis.
The proper time $\tau$ and space-time rapidity $y$ are used in
place of $t$ and $z$ which are defined by
\begin{eqnarray}\label{}
\tau = \sqrt{t^2-z^2}
\end{eqnarray}
and
\begin{eqnarray}\label{}
y = \frac{1}{2} ln \frac{t+z}{t-z}.
\end{eqnarray}
In the rest frame of the fluid we take a volume element as
$\delta V = A_{\bot} \tau \delta y$, where $A_{\bot}$
is the area of the element transverse to z direction.
The expansion takes place with the velocity $v_z = z/\tau$,
in the rest frame of the element.
Due to the viscous effects in longitudinal direction the heat
density per unit time is given by \cite{MIR}
\begin{eqnarray}\label{}
\phi & = & 2 \eta \left( \frac{\partial v_z}{\partial z} \right)^2 +
(\zeta - 2 \eta/3) \left( \frac{\partial v_z}{\partial z} \right)^2, \nonumber \\
{\rm or} \hspace{1in} \phi & = & (4\eta/3 + \xi)/\tau^2.
\end{eqnarray}
After time increment $\Delta \tau$,
the volume expands by $\Delta V = A_{\bot} \Delta \tau \delta y$.
The amount of work done in expansion is $p \Delta V$.
If the energy density at $\tau$ is $\epsilon$ and at $\tau+\Delta \tau$
is $\epsilon + \Delta \epsilon$ then the energy conservation implies
\begin{eqnarray}\label{}
\epsilon \delta V & = & (\epsilon + \Delta \epsilon) (\delta V + \Delta V)
+ p \Delta V - \phi \tau \Delta V, \nonumber \\
{\rm or} \hspace{1in} 0 & \simeq & \Delta \epsilon A_{\bot} \tau \delta y +
(\epsilon + p - \phi \tau)A_{\bot} \Delta \tau \delta y,
\end{eqnarray}
which leads to the scaling Navier-Stokes equation in the limit
$\Delta \tau \rightarrow 0$
\begin{eqnarray}\label{visc}
\frac{d\epsilon}{d\tau} & = & -\frac{\epsilon + p}{\tau} +
\frac{4\eta/3 + \zeta}{\tau^2}.
\end{eqnarray}
Equation~(\ref{visc}) has been
solved earlier \cite{JPG} for the QGP to hadron transition with the Maxwell
construction, i.e., assuming $T=T_c$ for the mixed phase.
To solve Eq.~(\ref{visc}) we require knowledge of the equation of state and
the temperature dependence of $\eta$ and $\zeta$.
In this work we use the bag equation of state for QGP. The energy density,
pressure and entropy densities in pure QGP and hadron phases are taken as
\begin{eqnarray}\label{}
\epsilon_q(T) = 3 a_q T^4 + B,\hspace{.1in} p_q(T) = a_q T^4 - B,
\hspace{.1in} s_q(T) = 4 a_q T^3,
\end{eqnarray}
\begin{eqnarray}\label{}
\epsilon_h(T) = 3 a_h T^4, \hspace{.1in} p_h(T) = a_h T^4,
\hspace{.1in} s_h(T) = 4 a_h T^3.
\end{eqnarray}
Here, $a_q$ and $a_h$ are related to the degrees of freedom
operating in two phases and $B$ is the bag pressure.
For ultrarelativistic gases, the bulk viscosity $\zeta$ is usually much
smaller than the shear viscosity $\eta$ \cite{WEIN}.
Danielewicz and Gyulassy \cite{DAN} give the acceptable range of $\eta$ for
the applicability of the Navier-Stokes equation to the expansion of the
plasma as
\begin{eqnarray}\label{range}
2 T^3 \le \eta \le 3 T^3 (\tau T).
\end{eqnarray}
We define $\mu = (4\eta/3 + \zeta)$ and assume the temperature
dependence of viscosity coefficient as
\begin{eqnarray}\label{}
\mu_q(T) = \mu_{q0} T^3,
\end{eqnarray}
with $q$ being replaced by $h$ for the hadronic phase.
Solving Eq.~(\ref{visc})
for temperature in pure quark phase, we get \cite{JPG}
\begin{eqnarray}\label{}
T = T_0 \left(\frac{\tau_0}{\tau}\right)^{1/3} +
\frac{\mu_{q0}}{8 a_q \tau_0}
\left[\left(\frac{\tau_0}{\tau}\right)^{1/3}-\frac{\tau_0}{\tau}\right]
\end{eqnarray}
for QGP formed at time $\tau_0$ while in pure hadron phase
\begin{eqnarray}\label{}
T = T_h \left(\frac{\tau_h}{\tau}\right)^{1/3} +
\frac{\mu_{h0}}{8 a_h \tau_h}
\left[\left(\frac{\tau_h}{\tau}\right)^{1/3}-\frac{\tau_h}{\tau}\right],
\end{eqnarray}
where $\tau_h$ is the time where hadronization gets completed.
The energy density in the mixed phase at a time $\tau$ can be
written in terms of hadronic fraction $h(\tau)$ as
\begin{eqnarray}\label{}
\epsilon(\tau) & = & \epsilon_q(T) + (\epsilon_h(T)-\epsilon_q(T))
h(\tau), \nonumber \\
& = & 3 [ a_q + (a_h-a_q) h(\tau) ] T^4 + B [1 - h(\tau)],
\end{eqnarray}
while the enthalpy density is
\begin{eqnarray}\label{}
\omega(\tau) = 4 [ a_q + (a_h-a_q) h(\tau) ] T^4.
\end{eqnarray}
The temperature $T$ can be deduced as
\begin{eqnarray}\label{temp}
T(\tau) = \left(\frac{\omega(\tau)}{4 [ a_q + (a_h-a_q) h(\tau) ] }\right)^{1/4}.
\end{eqnarray}
The shear and bulk viscosities are also taken to be the functions of time
according to
\begin{eqnarray}\label{}
\mu(\tau) & = & \mu_q(T) + (\mu_h(T)-\mu_q(T)) h(\tau), \nonumber \\
{\rm or} \hspace{1in} \mu(\tau) & = & [ \mu_{q0} + (\mu_{h0}-\mu_{q0})h(\tau)]
(T(\tau))^3.
\end{eqnarray}
For the Maxwell construction, i.e., for $T=T_c$ in the mixed phase,
solution of Eq.~(\ref{visc}) is
\begin{eqnarray}\label{consh}
h(\tau) = (c-ab)[ Ei(b/\tau) - Ei(b/\tau_c) ] e^{-b/\tau}/\tau
+ (a+1) [1- \tau_c/\tau e^{(b/\tau_c - b/\tau)}].
\end{eqnarray}
Here, $Ei$ is exponential integral and
\begin{eqnarray}\label{}
a = \frac{4}{3} \frac{e_h}{e_q-e_h},
\hspace{.1in} b = \frac{\mu_q-\mu_h}{e_q-e_h},
\hspace{.1in} c = \frac{\mu_h}{e_q-e_h}.
\end{eqnarray}
The constants a, b, and c are evaluated for $T=T_c$.
The solution given by Eq.~(\ref{consh}) does not
account for supercooling.
Due to supercooling, i.e., $T \ne T_c$ in the mixed phase, the temperature
is not constant and depends on $\tau$. Equation~(\ref{visc})
then becomes
\begin{eqnarray}\label{etau}
\frac{d\epsilon}{d\tau} = -\frac{\omega}{\tau}
+ \frac{\mu_{q0} + (\mu_{h0} - \mu_{q0}) h(\tau)}
{[ 4 a_q + 4(a_h-a_q) h(\tau)]^{3/4}}~~~ \frac{\omega^{3/4}}{\tau^2}.
\end{eqnarray}
Equation~(\ref{etau}) and Eq.~(\ref{frac}) are coupled equations,
finally to be solved for $h(\tau)$ and $\epsilon(\tau)$ [ or $T(\tau)$]. Once we
get $h(\tau)$ and $T(\tau)$ we can calculate the entropy density.
Eq.~(\ref{etau}) can also be written in terms of entropy as
\begin{eqnarray}\label{stau}
\frac{ds}{d\tau} = -\frac{s}{\tau} + \left(\frac{p_h(T)-p_q(T)}{T}\right)
~ \frac{dh}{dt}
+ \frac{\mu_{q0} + (\mu_{h0} - \mu_{q0}) h(\tau)}
{[ 4 a_q + 4(a_h-a_q) h(\tau)]^{2/3}}~~~ \frac{s^{2/3}}{\tau^2}.
\end{eqnarray}
Without the second and third terms on the right hand side, this equation
describes the conservation of entropy. The second term is responsible for
the entropy production due to nucleation. For the Maxwell
construction, $p_h-p_q=0$ as $T=T_c$ and this term vanishes.
The third term leads to continuous entropy production due to dissipative
effects.
We discuss the solution of these equations in the next section.
\section{ Results and Discussion}
By solving together Eq.~(\ref{etau}) and Eq.~(\ref{frac}), we have studied the
plasma evolution and calculated $s\tau$---the entropy production as
a function of $\tau$.
The initial conditions are taken at $\tau=\tau_c$ as $T=T_c$,
$h=0$ and $\epsilon(\tau_c)=\epsilon_q(\tau_c)$.
Now, $h(\tau)$ is calculated using Eq.~(\ref{consh}) with step
$\Delta$ in $\tau$. With this value of $h$, Eq.~(\ref{etau}) is solved for
$\epsilon(\tau)$ [or $T(\tau)$]. Then Eq.~(\ref{frac}) is evaluated by the
trapezoidal rule, thereby yielding new value of $h$.
Using new value of $h$, we solve again Eq.~(\ref{etau}) to
improve the value of $\epsilon$. This is repeated till an accuracy of
$\sim 10^{-5}$ is obtained. Then we proceed to the next step $\tau_c+2 \Delta$.
We take
$a_q = 37 \pi^2/90$
and $a_h = 4.6 \pi^2/90$.
The value 4.6 is taken instead of 3 to account for $\rho$,
$\omega$ and $\eta$ mesons apart from pions.
We chose $T_c$ = 160 MeV in this work implying
$B^{1/4}$ = 219 MeV.
For the hadronic matter $\eta_{h0} = 1.5$, $\zeta_{h0} = 1$ while
for the quark matter $\zeta_{q0} = 0$ \cite{DAN,JPG,HOSO}.
The coeficient $\eta_{q0}$ has been varied from 2.5 to 20.
The nucleation involves the surface free energy, correlation length,
and the velocity of bubble growth, which have already been
described in Sec. II.
The volume of a fluid element is $A_{\bot} \tau \delta y$.
As $\tau \delta y$ does not change with $\tau$ in the scaling
hydrodynamics, $s\tau$ provides a measure of total entropy.
So as to understand the role of supercooling
and viscous heat generation, we have
compared various scenarious; ideal hydrodynamics (IHD), IHD with supercooling,
viscous hydrodynamics (VHD) and VHD with supercooling.
We assumed
that at initial time $\tau_i$ = 1 fm/$c$, the temperature $T_i$=268 MeV marked the
beginning of the evolution. This corresponds to energy density $\epsilon$=8.51
Gev/fm$^3$ and entropy $s_i\tau_i$=40.65 fm$^{-2}$. These are
appropriate for the Relativistic Heavy Ion Collider (RHIC) energies.
For further
discussions, it is instructive to examine $\tau$ versus $T$ plot. Figure 1 shows
such a plot for VHD calculations with supercooling included. From a initial
point ($\tau_i,T_i$), one moves on to $(\tau_c,T_c)$ where critical temperature
is reached. However, the nucleation rate is zero at this point and system
does not hadronize. The system keeps cooling to $(\tau_m, T_m)$ where
significant nucleation rate and hadronization is reached. The system cannot cool
further as entropy cannot decrease according to the second law of
thermodynamics.
At this point, the hadron fraction is around 11-18 $\%$ for various cases.
As the temperature increases towards $T_c$, the hadron fraction $h$ increases
and slowly approaches a value of one at the point $(\tau_h, T_c)$.
After this, temperature starts decreasing finally reaching the point $(\tau_f,T_f)$
at the freeze-out temperature $T_f$. This is the general character of $\tau-T$
curve. If one uses the Maxwell construction, one reaches directly from
($\tau_c, T_c)$ to $(\tau_h, T_c)$ without any change in the temperature.
In Figs. 2 and 3, we show the results for
$T$, $h$ and $s\tau$ as a function of $\tau$ for $\eta_{q0}$=2.5 and 5 . For IHD, the critical temperature
$T_c$ is reached earlier than VHD. Similarly with supercooling included,
$T_m$ is
approached earlier in IHD than VHD. The supercooling leads to an abrupt
change in entropy starting from $\tau_m$ onwards, i.e., in the reheating
region.
We find in our calculations that $s\tau$ does not increase much beyond
$\tau \approx $20 fm/$c$.
Therefore, $s_h\tau_h$ represents reasonably well the total entropy
produced. As $s_h$ is evaluated at $T_c$ for all cases, the total entropy
is also a measure of the life time of the system upto the hadronization
stage.
Figure 4 shows $s_h\tau_h$
as a function of $\eta_{q0}$. With VHD, entropy production increases
with viscosity. Supercooling leads almost to a constant shift of the curve for the VHD.
In Fig. 5, the excess entropy production due to supercooling is shown as a
function of viscosity. For all values of $\eta_{q0}$, excess entropy does not
change much. It is almost constant for the VHD calculation.
We also found that the values of $\eta_{q0}<<1$ do not lead
to significant hadronization and the system continues to cool in the QGP
phase. Sufficient value of $\eta_{q0}$ is required for the transition to the
hadron phase. The bubble growth velocity coefficient (3$c$ in our calculation)
also plays an important role in the transition as setting it to zero also
blocks the transition. The viscosity plays very crucial role in the transition
through nucleation. However, once viscosity enters into picture, the viscous
hydrodynamics is to be used for consistency. In the presence of viscosity, the initial
energy density estimates from the experimental rapidity density distribution
would be lower than the ones calculated with the Bjorken formula. Though
supercooling
does not lead to significant increase in entropy production, it is
very much needed. For the description of the phenomena in relativistic nuclear
collisions, the viscous hydrodynamics with supercooling is a more appropriate
framework than the ideal hydrodynamics with or without supercooling. Further
theoretical work is needed to compute the viscosity with narrower bounds for
the analysis of the experimental data.
\section{ Conclusions}
We have studied the entropy production both in
ideal and viscous scaling hydrodynamics with and without supercooling.
Excess entropy produced due to
supercooling in viscous hydrodynamics, weakly depends on viscosity of the plasma phase.
Though, in general entropy produced due to supercooling is
much less than that due to viscous heat generation, the phenomenon
of supercooling provides an important physical mechanism
for the quark-hadron transition to occur. The viscous hydrodynamics with
supercooling also leads to an increase in the lifetime of the plasma.
|
\section{Introduction}
The bcc phase of $^{4}$He has a pronounced quantum nature due to the
relatively open structure of the lattice. Quantum effects are manifested in
strong anharmonicity of some phonon modes and in the large zero-point
kinetic energy of the atoms \cite{glyde}. It is this large kinetic energy
which is thought to help stabilize the bcc phase over the hcp phase. In this
paper we highlight this nature of the bcc phase by proposing a new physical
model for the local atomic motion. For the sake of clarity we have
reproduced here some of the arguments and calculations already given in \cite
{niremil}. We propose that in the bcc $^{4}$He phase the local excitations
of the atoms in their potential wells, result in oscillating local electric
dipoles. The ground-state of these dipoles has the dipoles oscillating in
synchrony, thereby reducing the dipolar interaction energy between them.
Solving a mean-field Hamiltonian describing these dipoles we find that
Bosonic phase fluctuations in the (110) direction reproduce the spectrum of
the T$_{1}(110)$ phonon.
In the following we further explore the nature of the coherent ground-state
of the local-modes in the bcc $^{4}$He. We show that the bcc $^{4}$He is a
unique phase having both Diagonal Long Range Order (DLRO) of the solid
lattice and Off-Diagonal Long Range Order (ODLRO) of the local dipoles.
There is therefore a complex three-component order parameter which describes
the coherently oscillating dipoles in each of the three orthogonal
directions in the lattice. In the ground-state the local dipoles form a
Bose-Einstein condensate in the zero momentum state, and we are able to
estimate the ground-state energy reduction due to this condensation. This
estimate compares favourably with experimental results and consequently we
claim that this condensation energy stabilizes the bcc phase over the hcp
phase. In the hcp phase we expect no coherence or condensation due to the
highly isotropic lattice and the geometric frustration of the hexagonal symmetry. We also comment about the
relation of this work to previous work about the 'super-solid' concept in
quantum solids.
Additionally, we predict a high-energy optical-like mode which has fermionic
statistics. This excitation is confined to the (110) direction and involves
a local 'fliping' of a dipole with respect to the ground-state. This makes
the dipole become anti-symmetric (a $\pi $ phase difference) with respect to
the global phase of the complex order parameter and aquire Fermi-Dirac
statistics. We also give analytic expressions for the scattering intensity
of both the Bose and Fermi excitations along the (110) direction. These
predictions remain to be compared with future experimental data.
\section{Ground-state coherence and Bose excitations.}
The usual treatment of the ground-state and energy of bcc $^{4}$He employs
variational wavefunctions that aim to account for the short-range
correlations between the atoms \cite{glyde}. These correlations arise mainly
due to the hard-core repulsion between the atoms. The atoms have a high
zero-point kinetic energy which is given quite accurately by treating them
as independent particles held in place by the potential of the neighboring
(static) atoms. This type of calculation is the ''particle-in-cell''
approximation which gives surprisingly good agreement with measured
thermodynamic properties of the solid phase \cite{glyde1}. We want to focus
here on the effects of the local motion of the atoms inside this
potential-well on the nature of the ground-state. In this approach we would
like to isolate the lowest energy excited state of the atom inside its
potential well, and treat it as a local excitation of the lattice. This
local excited state consists of a local oscillatory motion of the atom along
a particular direction and produces an oscillating electric dipole,
similarly to that of the usual Van-der Waals interaction. However unlike the
case of the Van-der Waals interaction, in which the dipolar fluctuations are
random, we show that in the bcc solid there are local dipoles which are
correlated and a new ground-state of lower energy is created.
The potential well of an atom in the bcc lattice due to the standard helium
pair-potential $\upsilon (r)$ \cite{glyde}, provided one can take the other
atoms as stationary, can be maped along any direction in the lattice. We
find \cite{niremil} that in the directions normal to the unit cube's faces
(i.e. (100),(010) etc.) the confining potential well is very wide with a
pronounced double-minimum structure (Fig.1). Solving the one-dimensional
Schrodinger's equation for a $^{4}$He atom in this potential, we get a first
excited level with energy $10{\rm K}$, and a wavefunction describing atomic
motion with an amplitude of $\sim 1{\rm \AA }$ (in the (100) direction
(Fig.1)). Based on the above calculation, we shall assume that the atoms
have a local-mode that is highly directional along one of the directions
equivalent to (100). Local atomic motion along the other directions is
assumed to be severly restricted due to the higher excitation energies
(Fig.1). Experimental evidence for the existence of such a ''local mode''
comes from NMR measurements which find an activation energy of 7$\pm 1$K\cite
{schuster,allen}, and we propose to identify this local-mode with
the highly directional motion of the atoms in the normal directions.
Using this identification we can now estimate the size of the local electric
dipole moment that can be created by this local and highly directional
atomic motion. As the atom moves this instantaneous local electric-dipole is
created due to the electronic cloud and the ion being slightly displaced
relative to each other. The electric dipole moment due to mixing of the
lowest $\left| s\right\rangle $ and $\left| p\right\rangle $
electronic-levels of the $^{4}$He atom, is given from perturbation theory as
\begin{eqnarray}
\psi &=&\left| s\right\rangle +\lambda \left| p\right\rangle \Rightarrow
E_{0}\simeq \left\langle \psi \left| E\right| \psi \right\rangle
-\left\langle s\left| E\right| s\right\rangle \simeq \lambda
^{2}\left\langle p\left| E\right| p\right\rangle \nonumber \\
\ &\Rightarrow &\lambda ^{2}\simeq 7/2.46\cdot 10^{4}\simeq 0.00284,\lambda
\simeq 0.0168 \label{lamda}
\end{eqnarray}
where $\left| s\right\rangle $ and $\left| p\right\rangle $ stand for the
ground-state and first excited-state of the $^{4}$He atom, $\lambda $ is the
mixing coefficient and $\left\langle p\left| E\right| p\right\rangle \simeq
2.46\cdot 10^{4}$ K is the excitation energy of the first atomic
excited-state \cite{white}. This small estimated mixing gives the magnitude
of the induced dipole moment as
\begin{equation}
\left| {\bf \mu }\right| =e\left\langle \psi \left| x\right| \psi
\right\rangle \simeq 2e\lambda \left\langle s\left| x\right| p\right\rangle
\simeq e\cdot 0.03{\rm \AA } \label{mu}
\end{equation}
where $\left\langle s\left| x\right| p\right\rangle \simeq 0.9{\rm \AA }$ .
The estimation of the mixing $\lambda $ and the dipole-moment $\left| {\bf
\mu }\right| $ serves to set an upper bound on the magnitude of this effect,
since we assumed that the entire excitation energy $E_{0}$ is converted to a
local electric dipole.
It is possible to show that the lowest energy of a correlated dipolar
array
in the bcc lattice preserves the symmetry of the bcc unit cell along one
of
the symmetry axes. In such a case it can be easily shown that there will
be
no contribution to the dipolar interaction energy from oscillating dipole moments
which
are orthogonal, and
the instantaneous dipolar
interaction energy for each of the three orthogonal directions, is given by
\begin{equation}
E_{dipole}=-\left| {\bf \mu }\right| ^{2}\sum_{i\neq 0}\left[ \frac{3\cos
^{2}\left( {\bf \mu }\cdot \left( {\bf r}_{0}-{\bf r}_{i}\right) \right) -1}
\left| {\bf r}_{0}-{\bf r}_{i}\right| ^{3}}\right] \label{edipole}
\end{equation}
where the sum is over all the atoms in the lattice, ${\bf r}_{i}$ being the
instantaneous coordinate of the $i$-th atom. For oscillating dipoles with
random phases, the average instantaneous interaction energy (\ref{edipole})
summed over the lattice would be zero. However, the energy of the dipoles
can be lowered by correlating the phases of the oscillating atoms. Since the
direction of the local dipole shows the instantaneous direction of the
motion or displacement, a state where all the dipoles point in the same
direction is just a uniform motion or translation of the entire lattice. We
therefore have to look for symmetric arrangements with respect to the number
of up/down dipoles, such as is shown in Fig.2. This is the lowest energy
'antiferroelectric' configuration with the periodicity of the bcc unit cell.
We have shown this arrangement for individual dipoles oriented along the
(001) direction, but they are similarly arranged for dipoles along the two
other orthogonal axes. The sum in (\ref{edipole}) for such a configuration
with a unit dipole is given in Fig.2. Thus, the ground state in our picture
has the atoms executing this correlated local oscillation along the three
orthogonal directions.
We therefore have, in addition to the usual (isotropic) Van-der Waals interaction,
highly directional (anisotropic) electric dipoles that become correlated so that they oscillate in synchrony.
This is a state of quantum resonance where the system oscillates between two
equivalent up/down arrangements of the ground-state of the dipoles (Fig.2).
The total interaction between the atoms is now given as the usual
second-order ($\propto 1/r^{6}$) Van-der Waals contribution that is the
result of local-dipoles which have random relative phases, and an additional
long-range (first-order, $\propto 1/r^{3}$) dipolar interaction from the correlated part.
Dipolar interactions that decay as $1/r^{3}$ occur for perfectly correlated
oscillating dipoles, such as a single electric dipole and its image in an
adjacent conducting plate. The coherently oscillating nearest-neighbor
dipoles therefore behave as perfect images of each other (Fig.2), and
oscillate with the same global phase.
The correlated oscillating dipoles do not have an average static dipole
moment, so this is not the case of an antiferroelectric structural phase
transition \cite{landau}. The array shown in Fig.2 is simultaneously
arranged along the other two orthogonal axes. Along each direction the
ground-state is given as a coherent-state of these local dipoles, i.e. has a
well-defined phase and an ill-defined occupation number.
We shall treat the dynamics of the correlated dipolar array as independent
of the other degrees of freedom of the lattice. This assumption needs
justification since there can be phonon modes that will modulate the atomic
motion, thereby coupling with the oscillating dipolar array. The oscillatory
atomic motion induced by the phonons will modulate the relative phases of
the dipoles. Let us look at the ground state of the dipoles, taking for
example dipolar
oscillations oriented along the (001) direction (Fig.2). We now need to
consider only phonons which will modulate the local motion responsible for
the oscillating dipoles in this direction. In the bcc structure, only 3
phonons fulfill this condition: L(001), T(100) and T$_{1}$(110). Let us
calculate the energy of the dipolar array when modulated along these 3
directions. For a modulation along some direction ${\bf k}$ , the dipolar
interaction energy is given by\cite{heller}:
\begin{eqnarray}
X\left( {\bf k}\right) &=&%
-\left| {\bf \mu }\right| ^{2}\sum_{i\neq 0}\left[%
\frac{3\cos ^{2}\left( {\bf \mu }\cdot \left( {\bf r}_{0}-{\bf
r}_{i}\right)%
\right) -1}{\left| {\bf r}_{0}-{\bf r}_{i}\right| ^{3}}\right] \nonumber
\\
&&\ \ \exp \left[
2\pi i{\bf k}\cdot \left( {\bf r}_{0}-{\bf r}_{i}\right) \right]
\label{xk}
\end{eqnarray}
At $k=0$ the interaction matrix $X(k)$ is just the dipolar energy (\ref
{edipole}).
In Fig.3 we plot the value of $X(k)$, the energy of the dipolar array
modulated by the relevant phonons:\ L(001), T(100), and T$_{1}$(110), for
dipole moment $\left| {\bf \mu }\right| =1$. We see that for a modulation by
L(001) and T(100) the periodicity of $X(k)$ is over a full unit-cell, that
is twice the periodicity of these phonons. Since symmetric functions of
periodicities $\pi /a$ and $2\pi /a$ are orthogonal, so are the
eigenfunctions of these particular phonons and dipole-excitations. The
dipole array cannot therefore be excited by any of these two phonon. For the
modulation produced by the T$_{1}$(110) mode, the periodicity of $X(k)$ is
the same as that of the T$_{1}$(110) phonon, which can therefore couple to
the dipole array. We conlude therefore that the coupling of the local modes
to the lattice excitations is limited to a single phonon mode, justifying
our assumption that the local modes can be treated separately to a good
approximation. We shall now calculate the dispersion relation of such an
excitation by a mean-field solution of an effective Hamiltonian. It turns
out that the only phonon mode of the bcc lattice that can couple with the
dipolar array is in fact the natural excitation of the dipolar array in the
(110) direction. Thus, the only elementary (Bose) excitations of the dipole
array would be in the (110) direction, in the form of the T$_{1}$(110)
phonon. The description of this phonon is therefore taken into account by
our treatment of the dynamics of the dipolar array, and will appear as a
solution of the mean-field treatment. This means that our assumption of an
effective decoupling between the dipolar and other degrees of freedom is
justified.
The Hamiltonian treatment of interacting local excitations was developed
originally by Hopfield \cite{hopfield} for the problem of excitons in a
dielectric material. The local excitations are treated as bosons using the
standard Holstein-Primakof procedure, and the effective Hamiltonian
describing their behavior is \cite{anderson}
\begin{eqnarray}
{H_{loc}} &=&{\sum_{k}}(E_{0}+X(k))\left( {{b_{k}}^{\dagger
}}{b_{k}}+{\frac{%
1}{2}}\right) \nonumber \\
&&\ \ +{\sum_{k}}X(k)\left( {{b_{k}}^{\dagger }}{b_{-k}^{\dagger }}%
+b_{k}b_{-k}\right) \label{hloc}
\end{eqnarray}
where ${{b_{k}}^{\dagger },}{b_{k}}$ are Bose creation/anihilation operators
of the local mode, $X(k)$ is given in (\ref{xk}) and $E_0$ is the energy
of exciting a local dipole out of the correlated ground-state.
The Hamiltonian ${H_{loc}}$ (\ref{hloc}) which describes the effective
interaction between localized modes can be diagonalized using the Bogoliubov
transformation ${\beta _{k}}=u(k)b_{k}+v(k)b{^{\dagger }}_{-k}$. The two
functions $u(k)$ and $v(k)$ are given by:
\begin{equation}
{u^{2}}(k)={\frac{1}{2}}\left( \frac{E_{0}{+X(k)}}{{E(k)}}+1\right) ,{v^{2}
(k)={\frac{1}{2}}\left( \frac{E_{0}{+X(k)}}{{E(k)}}-1\right) \label{uv}
\end{equation}
The result of solving by mean-field the effective Hamiltonian for the
correlated dipolar array \cite{niremil}, is a coherent ground-state given by
\cite{huang}
\begin{equation}
\left| \Psi _{0}\right\rangle =\prod_{k}\exp \left( \frac{v_{k}}{u_{k}}{
b_{k}}^{\dagger }}{b_{-k}^{\dagger }}\right) \left| vac\right\rangle
\label{psi0}
\end{equation}
and the energy spectrum is
\begin{equation}
E(k)=\sqrt{E_0\left( E_0+2X(k)\right) } \label{ek}
\end{equation}
In Fig.2 we see that the ground-state arrangement has the dipoles arranged
in alternating planes in the (110) direction. As we have shown the only
naturally occuring Bose excitations of this dipolar field are along this
direction and $X(k)$ is the dipolar interaction matrix element for $k$ in
the (110) direction (Fig.3). In order to calculate the energy spectrum we
now need to fix the size of the coherent dipole moment $\left| {\bf \mu
\right| $. According to our definition of the local mode the energy cost of
flipping the direction of a single dipole out of the ordered ground state is
defined to be $E_{0}$. This is equivalent to having $2\left| X(k=0)\right|
=E_{0}$, which is the condition to have a gapless mode at $k\rightarrow 0$
\ref{ek}). Using this condition, the experimental value of $E_{0}=7$K \cite{schuster,allen} determines the size
of the coherent dipole moment as: $\left| {\bf \mu }\right| \simeq e\cdot 0.01{\rm
\AA }$. This value is indeed smaller than our previous estimation, which
served as an upper bound on the size of the oscillating dipole moment (\ref{mu}).
As we have proposed, the phase modulation in the (110) direction of the
transverse atomic motion in the lattice, with energy $E(k)$ (\ref{ek})
should coincide with the T$_{1}$(110) phonon. In Fig.4 we compare the
experimental values of T$_{1}$(110) taken from neutron scattering data with
the calculated $E(k)$, and we find that the agreement is excellent for all
k $. From (\ref{ek}) and Fig.3 we see that at the edge of the Brillouin zone
the energy $E(k)$ of the phonon should be just the bare energy of the local
mode, $E_{0}$, since $X(\sqrt{2}\pi /a)=0$. We also have that at $k=\sqrt{2
\pi /a$ the dipoles have changed between the two configurations illustrated
in Fig.3, which are the two possible configurations with alternating dipoles
arranged on adjacent planes with the periodicity of the bcc unit cell.
Since the empirical value of $E_{0}$ that we used was taken from NMR data,
the agreement we find with the phonon data taken from inelastic neutron
scattering, emphasizes the self-consistency of our description. We stress
that the value of $E_{0}$ and the lattice vectors are the only empirical
inputs used in the calculation, with the functional behavior completely
given by the lattice structure and the dipolar interactions.
\section{Off-Diagonal-Long-Range-Order and condensation.}
We have found from the mean-field solution at zero temperature that the
ground-state of the bcc phase contains a coherent-state of oscillating
local-dipoles (\ref{psi0}). Since our method predicts the excitation
spectrum of the T$_{1}$(110) phonon with very good accuracy, we expect it to
be valid at the finite temperatures for which the bcc phase exists. We
therefore expect that the basic nature of the bcc phase will be well
described by our results, although the quantitative values may change due to
the finite temperature. The coherent ground-state defines a global phase and
breaks the gauge symmetry of a well-defined occupation number of local
dipoles. In the limit $k\rightarrow 0$ we find that the occupation number of
the local-modes diverges as $1/k$, signaling macroscopic Bose-Einstein
condensation in the zero-momentum state
\begin{equation}
\left\langle n_{k}\right\rangle =v^{2}(k)=\frac{1}{2}\left( \frac{E_{0}{+X(k
}}{E(k)}-1\right) \rightarrow _{k\rightarrow 0}\frac{1}{2}\frac{E_{0}/2}{E(k
}=\frac{E_{0}/2}{2\hbar kc} \label{nk}
\end{equation}
where $c$ is the sound velocity of the T$_{1}$(110) phonon which is the
natural excitation of the dipolar array. This is identical to the result for
a Weakly Interacting Bose Gas (WIBG) problem solved by Bogoliubov \cite{bogo
, where the divergence is related to the occupation of the zero-momentum
state, i.e., the condensate fraction $n_{0}/n$
\begin{equation}
\left\langle n_{k}\right\rangle _{WIBG}=v^{2}(k)\rightarrow _{k\rightarrow 0
\frac{n_{0}}{n}\frac{mc}{2\hbar k} \label{nwibg}
\end{equation}
where in the WIBG case we have $c$ as the $k\rightarrow 0$ sound velocity,
and $\varepsilon _{k}=\hbar ^{2}k^{2}/2m$ is the free particle energy. By
comparing (\ref{nk}) with (\ref{nwibg}) we find that in the bcc case the
role of the condensate-fraction, the WIBG order-parameter, is taken by the
parameter $E_{0}$, which is just $2\left| X(0)\right| $. This can be seen
directly from the form of the ground-state wavefunction (\ref{psi0}) where
the pair-occupation is given by:
\begin{equation}
\left\langle b_{k}^{\dagger }b_{-k}^{\dagger }\right\rangle =2u(k)v(k)=\frac
X(k)}{E(k)} \label{pair}
\end{equation}
Equating the divergent part in (\ref{nk}) and (\ref{nwibg}) we can define an
effective condensate fraction
\begin{equation}
\frac{n_{0}}{n}=\frac{E_{0}/2}{mc^{2}}\simeq \frac{3.5}{10}=35\pm 8\%
\label{nn0}
\end{equation}
where we used for the velocity of sound $c$ the values from our calculation
\ref{ek}) ($\sim 130m/\sec $) and from elastic constants \cite{minki} ($\sim
160m/\sec $). It must be remembered that the mass $m$ in (\ref{nn0}) is not
necessarily the mass of a bare $^{4}$He atom since we are now dealing with
condensation of local dipoles. Comparing with the condensate fraction at
zero temperature in superfluid $^{4}$He \cite{sokol}, which is $\sim 10\%$,
we find that it is lower than the condensate fraction of the local modes in
the bcc phase. We again mention that our result is for T=0 which can be
depleted at finite temperature.
It is clear that it is a non-zero coherent dipole moment $\mu $ that
produces a dipolar interaction matrix $X(k)$ which in turn implies finite
pair occupation (\ref{pair}) and a coherent ground-state. This is just the
condition for the Bose-Einstein condensation of the dipoles in the bcc
ground-state (\ref{nn0}). We therefore have a broken gauge symmetry and a
complex order-parameter in the form of the pair-occupation (\ref{pair}).
This function can be complex since the conditions on $u(k)$ and $v(k)$ allow
for a relative complex phase between them, just as in the WIBG case. A
similar condensation of local dipoles in all three orthogonal axes of local
motion means that there are three independent phases at each lattice site,
since orthogonal dipoles do not interact. The order-parameter in our case
can therefore be described as a vector of three complex functions of
independent magnitude and phase:
\begin{equation}
\Phi ({\bf r})=\left(
\begin{array}{l}
\left| \mu _{x}\right| e^{i\theta _{x}({\bf r})} \\
\left| \mu _{y}\right| e^{i\theta _{y}({\bf r})} \\
\left| \mu _{z}\right| e^{i\theta _{z}({\bf r})}
\end{array}
\right) \label{order}
\end{equation}
If the cubic symmetry is not broken by external stresses, the magnitude
of the coherent dipole moment in the three orthogonal directions should
be the same: $\left| \mu _{x}\right|=\left| \mu _{y}\right|=\left| \mu _{z}\right|$.
In the ground-state the phases are spatially uniform, while the excited
state is described through a periodic phase oscillation, i.e. the T$_{1}
(110) phonon. The order parameter (\ref{order}) is to be contrasted with the
order parameter of superfluid $^{4}$He, which also exhibits ODLRO and which
has a single complex component.
In the hcp phase we do not expect the dipoles to order in a coherent state
since the hexagonal geometry frustrates antiferroelectric-type
configurations. Also the nearly isotropic potential of the hcp lattice
does not allow the highly directional dipole moments as in the bcc case. Indeed there is good agreement between experiments and the
harmonic calculation of the phonons in the hcp phase \cite{minki2}, indicating no strong
quantum corrections, as in the bcc phase.
The bcc $^{4}$He is therefore a unique crystallographic phase having both
Diagonal Long Range Order (DLRO) of the solid lattice and Off-Diagonal Long
Range Order (ODLRO) of the local dipoles. It is not a 'super-solid' \cite
{andreev,leggett,widom,stoof} in that it does not contain both a superfluid and
a solid, but is more similar to the superconductors which have a DLRO of the
atoms in the lattice and ODLRO\ of the superconducting electrons \cite{kohn
. This system is also distinct from the case of Bose-Einstein condensation
of a phonon mode which results in a static deformation of the lattice and a
structural phase-transition \cite{kohn}.
Bose-Einstein condensation of local defects (vacancies) was previously
considered for solid $^{4}$He \cite{andreev,leggett,widom}, and is similar
to our treatment. The main difference is that in our case the physical
picture of the condensed local modes is not a local distortion of the
lattice like a vacancy, and that the condensation is unique to the bcc
phase. The estimate in these works \cite{andreev,leggett,widom} is that in the
ground-state (T=0) the density of vacancies is $\sim 10^{-4}$ per site (at
molar volume of 21cm$^{3}$), and a condensate density of $\sim 10^{-7}$ per
site. In contrast we expect in the bcc phase a sizable fraction (10-30\%) of
condensed local-modes per site (\ref{nn0}).
The fact that the region of existence of the proposed supersolid phase in
the phase diagram of solid $^{4}$He should closely coincide with the region
occupied by the bcc phase, was shown in \cite{stoof}. In this work it was
further shown that in the supersolid there should be a second-sound-like
mode, which is an oscillation in the density of the local-defects. In our
description of the bcc phase this suggests the possibility of an oscillation
in the amplitude of the order-parameter $\Phi ({\bf r})$ (\ref{order}), that
is in the amplitude of the coherently oscillating dipole moment. This is in
contrast to the T$_{1}$(110) phonon mode which is an oscillation in the
phase of the order-parameter. This mode may be produced by modulating the
density of the atoms so that the local excitation energies change and with
them the amplitude of local motion and local electric dipole moment.
Unfortunately we do not expect such a mode to have measurable consequences
which are different from the effects produced by usual longitudinal phonons.
In concluding this section, we would like to mention the recent experiments
on the behaviour of implanted metalic ions (Cs) in solid $^{4}$He \cite
{kanorsky}. These experiments are designed to look for evidence of
time-reversal symmetry breaking which is equivalent to having a static
electric dipole moment. In our description of the bcc phase we do not find a
static but a coherent-dynamic electric dipole moment. We point out that in
these experiments a marked difference between the hcp and bcc phases has
been found. In the bcc phase the electronic-spin relaxation of the Cs atom
is extremely slow and this effect could be a result of the coherence and
long-range order of the dipolar fields. The coherently oscillating $^{4}$He
electrons in the bcc phase will produce a very uniform electromagnetic
interaction with the electronic spin in the Cs atom. By comparison, in the
hcp phase the spin polarization is extremely short lived, indicating a more
random field environment. This result is in accord with our expectation that
the coherent dipoles are unique to the bcc phase. Similar experiments in the
future may allow a probe that will show directly the coherently oscillating
dipoles in the bcc ground-state.
In these experiments \cite{kanorsky} the hyperfine transition in the Cs atom
was also measured. The energy shift of this transition is sensitive to the shape
of the confining cavity of the Cs atom inside $^{4}$He lattice. The width
of the transition is a measure of the fluctuations in this cavity size \cite
{kanorsky2}, and the data show a much smaller spread in the bcc compared
with the hcp phase. Uncorrelated atomic motions of the $^{4}$He atoms will
increase the spread in instantaneous cavity sizes due mainly to breating-like motion of the cavity walls (Fig.5).
This behavior is what we expect
for the hcp case. The correlated atomic motion in the bcc phase should
result in a more constant cavity shape (Fig.5) and a narrow signal, which is indeed measured \cite{kanorsky}.
\section{Ground-state energy and the stability of the bcc phase}
The question of the relative stability of the different crystal structures
in solid He has been a long standing one. The necessity for some
non-Van-der-Waals interactions has been previously proposed to explain the
occurance of fcc over hcp structure in the heavier rare-gas solids \cite
{venables}. The bcc phase is usually found to be more stable than the
close-packed hcp phase due to the large zero-point energy in the He solids
\cite{venables}. The correlations between the dipoles in the ground-state
that we have proposed, lowers the energy of the ground-state of the bcc
phase and further stabilizes it with respect to the hcp phase. The reduction
in ground-state energy acheived by the coherent state of the dipoles along
one of the three orthogonal directions, is
given by \cite{anderson}
\begin{equation}
\Delta E=\sum_{k}\frac{E(k)-(E_{0}+X(k))}{2E_{0}}<0 \label{dele0}
\end{equation}
which is negative since $X(k)<0$ and $E(k)<(E_{0}+X(k))$.
Since the energy reduction integrand (\ref{dele0}) is non-zero only in the
(110) directions it will give a small contribution to the three dimensional
phase-space integration. At zero temperature the summation in (\ref{dele0})
will be confined to one dimensional sections along the (110) direction, so
that the contribution will be zero. At the bcc temperatures ($\sim $1.4K)
the one dimensional chains in the (110) directions are broadened so that the
summation in (\ref{dele0}) is now over finite volume sections of phase
space. We can estimate the maximum width of the conical section in $k$-space
as the momentum which corresponds to a T$_{1}$(110) phonon with energy
k_{B}T$, that is $\sim 0.13$\AA $^{-1}$. The numerical integration of (\ref
{dele0}) over such volume sections gives an energy reduction of $\Delta
E\simeq -2$mK per atom. This result is in agreement with the experimentally
interpolated energy difference between the bcc and hcp phases of solid $^{4}
He \cite{balibar}, which is of the order of a few mK per atom.
This reduction is less than 0.1 percent of the potential and kinetic
energies of the solid, and is therefore very hard to calculate accurately
theoretically \cite{nosanow}. What is shown in the usual calculations is
that the correlations between the motions of the atoms are essential in
lowering the energy of the bcc phase, compared with the hcp phase. Since
part of the correlations in the atomic motion is described by our coherent
dipole model, we expect the condensation energy of the dipoles (\ref{dele0})
to be important in determining the stability of the bcc phase. At finite
temperature the stabilization of the bcc phase compared to the hcp phase is
usually attributed to the lower zero-point energy due to the lower T$_{1}
(110) phonon energy \cite{venables}. This is just the phonon which is
softened by the long-range dipolar interactions that we have described,
indicating again the importance of the coherent dipoles to the stabilization
of the bcc phase of solid $^{4}$He. Our procedure may provide a good
estimate of the small change in energy at the structural phase transition,
by isolating the degree-of-freedom which is most affected by the transition,
i.e. the correlated atomic motion along the directions normal to the unit
cell faces.
The picture we propose is that the dipole condensation mechanism of the bcc
phase competes with the lower potential energy of the hcp phase due to its
higher coordination number. If the hcp phase has a large enough volume
(through thermal expansion or introduction of $^{3}$He impurities) its
potential energy is increased until a critical point is reached where the
bcc phase has a lower total energy due to the dipolar-condensation energy
reduction (\ref{dele0}), which is absent in the hcp phase \cite{niremil}. At
this critical point the structural phase transition occurs.
By comparison, solid $^{3}$He has a stable bcc phase due to the larger
kinetic energy of this lighter isotope. This increased zero-point energy
causes the less dense bcc phase to have a lower ground-state energy than the
hcp phase even at T=0. Since the $^{3}$He atoms are fermions with a spin 1/2
nucleus, the oscillating electric dipoles are not in resonance and can not be
treated as bosons \cite{glyde1}. We therefore do not expect a coherent state
of the atomic motion as in the bcc $^{4}$He, and bcc $^{3}$He is stable due
to its large zero-point kinetic energy alone. On the other hand, at low
enough temperatures where the bcc $^{3}$He becomes an antiferromagnet, there
could be correlations involving both the nuclear spin and electric dipole
degrees of freedom.
\section{Fermionic excitations}
In addition to the fluctuations of the phase of the coherent dipole
ground-state (i.e. T$_{1}$(110) phonons), there can be a localized 'flip' of
a dipole so that it is in anti-phase (phase difference of $\pi $) relative
to the rest of the dipoles, in the ground-state configuration. Such an
excitation is naturally treated as a Fermion since such a flipped dipole is
antisymmetric with respect to the other dipoles, that is with respect to the
global phase $\theta $ (\ref{order}) in one (or more) of the orthogonal
directions of local motion ($x,y,z$). The flipped dipole is no longer a
dipolar image of the nearest-neighboring dipoles but an anti-image, and will
be treated with Fermi-Dirac statistics.
An anti-phase localized-mode (a fermion) is not part of the correlated
ground-state, but nevertheless will feel the effect of the Bose excitations
(T$_{1}$(110) phonons) of the dipolar array as they interact with it. The
effective Hamiltonian describing such a fermion should therefore contain a
term describing the creation and anihilation of pairs of fermions from the
ground-state by a phonon (Boson). This is an off-diagonal term that
describes the fluctuation caused by a T$_{1}$(110) phonon of energy $E(k)$:
it changes a fermion 'particle' into a 'hole' and vice versa. The terms
'particle' and 'hole' are with respect to the ground-state which has
occupation of pairs of localized-modes (i.e. not an 'empty' vacuum).
In addition there should be a term that describes the excitation energy of
the bare fermionic localized-mode, that is $E_{0}$. This is just the energy
to 'flip' a dipole from the ground-state so that all it's interactions with
the other dipoles of the ground-state change sign, i.e. $-2X(0)=E_{0}$.
The many-body effective Hamiltonian that we therefore propose is
\begin{equation}
H_{D}=\sum_{k}E(k)\left( c_{k}^{\dagger }c_{-k}^{\dagger
}+c_{k}c_{-k}\right) -\sum_{k}V_{k}\left( c_{k}^{\dagger
}c_{k}c_{-k}^{\dagger }c_{-k}\right) \label{diracham}
\end{equation}
where $c_{k}^{\dagger },c_{k}$ are the creation and annihilation operators
of the anti-phase (Fermionic) localized-mode. The first term in (\ref
{diracham}) is the 'kinetic' term due to the phonon-roton branch, where the
localized-modes are created/anihilated in pairs. The energy $E(k)$ is the
energy of the T$_{1}$(110) phonon excitation (\ref{ek}). In addition there
is a finite 'potential' energy if there is a finite density of unpaired
fermions, which is $E_{0}$. In the absence of the second term we have just
the Bose ground-state written in terms of fermionic pairs.
We linearize the equations of motion that follow from (\ref{diracham}),
similar to the BCS method \cite{kittel}
\begin{equation}
i\hbar \stackrel{\cdot }{c}_{k}=-E(k)c_{-k}^{\dagger }+\Lambda
_{k}c_{k}\qquad i\hbar \stackrel{\cdot }{c}_{-k}^{\dagger
}=-E(k)c_{k}-\Lambda _{k}^{*}c_{-k}^{\dagger } \label{eqmotion}
\end{equation}
where we used the Fermi anti-commutation relations: $\left\{
c_k,c_k^{\dagger }\right\} =1\qquad \left\{ c_k,c_{-k}^{\dagger }\right\} =0
, and we define
\begin{equation}
\Lambda _k=\Lambda _k^{*}\equiv E_0\equiv \sum_kV_k\left\langle c_k^{\dagger
}c_k\right\rangle ,\sum_k\left\langle c_k^{\dagger }c_k\right\rangle \equiv
1,V_k=E_0 \label{linear}
\end{equation}
From (\ref{linear}) we see that the symbol $E_0$ will now indicate a finite
density of fermions.
The equations of motion have the following eigenvalues:
\begin{equation}
\left|
\begin{array}{cc}
E_{f}(k)-E_{0} & -E(k) \\
-E(k) & E_{f}(k)+E_{0}
\end{array}
\right| =0\Rightarrow E_{f}(k)=\sqrt{E_{0}^{2}+E(k)^{2}} \label{specdir}
\end{equation}
We can now solve the equations using the Bogoliubov-Valatin transformation
for superconductivity \cite{kittel}:
\begin{equation}
c_{k}=u_{k}\alpha _{k}+v_{k}\alpha _{-k}^{\dagger }\qquad c_{-k}^{\dagger
}=-v_{k}\alpha _{k}+u_{k}\alpha _{-k}^{\dagger } \label{bogo}
\end{equation}
with the functions $u_{k},v_{k}$ given by
\begin{equation}
u_{k}^{2}=\frac{1}{2}\left( 1+\frac{E_{0}}{E_{f}(k)}\right) ,v_{k}^{2}=\frac
1}{2}\left( 1-\frac{E_{0}}{E_{f}(k)}\right) \label{uvdir}
\end{equation}
The ground state is
\begin{eqnarray}
\alpha _{k}\left| 0\right\rangle &=&0\Rightarrow \left| 0\right\rangle
=\prod_{k}\alpha _{-k}\alpha _{k}\left| vac\right\rangle \nonumber \\
&=&\prod_{k}\left(
u_{k}+v_{k}c_{k}^{\dagger }c_{-k}^{\dagger }\right) \left|
vac\right\rangle
\label{diracgs}
\end{eqnarray}
In Fig.6 we plot the energy spectrum (\ref{specdir}) compared with the other
phonon modes in the (110) direction \cite{minki}. It is clear that this
optic-like branch should be detectable in the low momentum range where it is
not masked by the signal from the accoustic phonon modes. There is at
present no high resolution neutron-scattering data in this energy and
momentum range, and this prediction can be hopefully checked in future
experiments. This mode could also be observed by Raman scattering, as a peak
at energy $E_{0}$.
\section{Scattering intensity}
A\ neutron scattering inelastically from the solid will create/anihilate an
elementary excitation. An excitation from the effective ground-states of the
T$_{1}$(110) phonon (\ref{psi0}) and of the fermionic mode (\ref{diracgs})
involves an anihilation of a pair of local-modes, leaving an unpaired
local-mode. We therefore expect the experimentaly measured neutron
scattering intensity to be proportional to the density of local-mode
pair-occupation at each wavevector $k$. This gives us the following results
for the two modes:
T$_1$(110) phonon:
\begin{equation}
I\propto \left\langle {{b_k}^{\dagger }}{b_{-k}^{\dagger }}\right\rangle
\frac{E_0}{E(k)}\left| \left( \frac{E(k)}{{E_0}}\right) ^2-1\right|
\label{iphon}
\end{equation}
Fermionic excitation:
\begin{equation}
I\propto \left\langle c_{k}^{\dagger }c_{-k}^{\dagger }\right\rangle =\frac{
}{2}\frac{E(k)}{E_{f}(k)}=\frac{1}{2}\frac{1}{\sqrt{\left( E_{0}/E(k)\right)
^{2}+1}} \label{idirac}
\end{equation}
Both functions (\ref{iphon}),(\ref{idirac}) are plotted with arbitrary scale
in Fig.7. We see that the intensity of the T$_{1}$(110) phonon is such that
at small $k$ it behaves as $1/k$ which is typical for phonons at low $k$ and
was seen experimentally \cite{minki2}, but goes identically to zero at the
edge of the Brillion zone where $E(k)\rightarrow E_{0}$. The fermionic
excitation has an opposite behavior by increasing linearly in intensity with
$k$, until it saturates at the edge of the Brillion zone. The expression for
the intensity of the T$_{1}$(110) phonon is similar to the expression of
the intensity of the phonon-roton excitation spectrum of superfluid $^{4}$He
\cite{nireric}, where it agrees very well with the experimental results.
These predictions for the bcc phase have yet to be checked experimentally.
\section{Conclusion}
In this work we have investigated the nature of the quantum correlations in
the bcc phase of solid $^{4}$He. We identified a three component complex
order parameter and Bose-Einstein condensation in this phase, though not a
'super-solid' \cite{kohn}, i.e. no superfluid component. There can be
further manifestations of the ODLRO of the dipoles in the bcc phase which we
have not explored yet, such as macroscopic topological defects in the
complex order-parameter. The order-parameter or condensate-fraction can also
serve as an extra thermodynamic variable, and this opens the possibilty of
more complicated internal dynamics in the bcc solid, such as the phenomenon
of second sound in superfluid $^{4}$He.
We predict that a local excitation of a dipole out of the coherent
ground-state will behave as a Fermion, and we calculate its energy spectrum.
We find it to behave as an optical-like branch in the (110) direction.
Finally we calculate the scattering intensity as a function of wavevector $k$
for both the Bose (T$_{1}(110)$ phonon) and Fermi (new optical-like branch)
excitations. All these predictions await high-resolution neutron and Raman
scattering experiments to be compared with.
{\bf Acknowledgements}
I thank Emil Polturak for useful discussions and encouragement.
This work was supported by the Israel Science Foundation and by the Technion
VPR fund for the Promotion of Research.
\newpage
\section{Appendix A: Comparison of Bose excitations with Klein-Gordon Hamiltonian.}
We would like to point out that the Hamiltonian describing
the localized dipoles (%
\ref{hloc}) is similar to the Klein-Gordon (KG) Hamiltonian
for a
single spinless boson, written in its first-order form \cite{baymq}:
Klein-Gordon:
\begin{equation}
H_{KG}=\varepsilon _k\left( \sigma _z+i\sigma _y\right) +mc^2\sigma
_z+e\Phi
\widehat{1} \label{hamkg}
\end{equation}
Localized dipoles (\ref{hloc}):
\begin{equation}
H_{loc}=X(k)\left( \sigma _z+i\sigma _y\right) +E_0\sigma _z
\label{hamdip}
\end{equation}
where $\sigma _i$ are the Pauli matrices, $\varepsilon
_k=\widehat{p}^2/2m$, $%
e$ is the electric charge, $\Phi $ is the electrostatic potential, $m$
is
the KG-particle's mass and $c$ is the velocity of light.
We have written the dipolar hamiltonian (\ref{hamdip}) in the basis of a two
component wavefunction
\begin{equation}
\Psi _{loc}=\left(
\begin{array}{c}
c_k^{\dagger } \\
c_{-k}
\end{array}
\right)
\label{dipolewave}
\end{equation}
In this representation we see that exciting a local dipole out of the
ground-state configuration ($c_k^{\dagger }$) has bare energy $E_0$ while
destroying an excited dipole has minus this energy.
There is a freedom of choice weather to define the positive excitation to be a
flipping of an up dipole to a down dipole or vice versa. The sign of the energy of
the dipolar bosons therefore represents this freedom which corresponds to two
equivalent dipolar configurations with a $\pi$ phase difference.
The
two-component
wavefunction of the KG Hamiltonian is:
\begin{eqnarray}
\Psi _{KG}&=&\left(
\begin{array}{c}
\varphi \\
\chi
\end{array}
\right) \nonumber \\
\varphi &=&\frac 12\left( \psi +\frac{i\hbar }{mc^2}\psi
^0\right)
,\chi =\frac 12\left( \psi -\frac{i\hbar }{mc^2}\psi ^0\right)
\label{kleinwave}
\end{eqnarray}
where $\psi $ is the original wavefunction of the second-order KG
equation,
and $\psi ^0=\left( \frac \partial {\partial t}+\frac{ie}\hbar \Phi
\right)
\psi $.
The Hamiltonians (\ref{hamkg},\ref{hamdip}) are the similar
except that the KG density is not normalized to 1 but to $\left\langle
\rho
\right\rangle =E/mc^2$, describing the relativistic increase in the
density
with velocity.
By comparing the two Hamiltonians (\ref{hamdip},\ref{hamkg}) we identify that:
$E_0\leftrightarrow mc^2$, $X(k)\leftrightarrow \varepsilon _k$,
which gives the equivalence of the two hamiltonians.
The peculiarities of the KG equation appear when there is a potential $%
V=e\Phi $ (the Klein paradox for example). The equation for the momentum
of
the KG prticle:
\begin{eqnarray}
2mc^2\varepsilon _k&=&\hbar ^2c^2k^2=\left( E(k)-V\right) ^2-\left(
mc^2\right)^2 \nonumber \\
\Rightarrow k&=&\frac{\sqrt{\left( E(k)-V\right) ^2-\left( mc^2\right)
^2}}{\hbar c} \label{kgk}
\end{eqnarray}
becomes the equation for $X(k)$ in the dipolar case:
\begin{equation}
X(k)=\frac{\left( E(k)-V\right) ^2-\left( E(k) _0\right) ^2}{2E_0}
\label{xkv}
\end{equation}
We see from (\ref{xkv}) that there is a region of energies where the
interaction parameter $X(k)$ is positive and a region where it is
negative.
We saw above that the condensation of the dipoles in the bcc phase is characterized by a negative $X(k)$
which
also gives a gapless excitation
spectrum
at $k\rightarrow 0$. The excitations with $%
E(k)>V+E_0,E(k)<V-E_0$ are therefore not contributing
to
the coherent long-range order.
A fermionic excitation is a local destruction of the coherent order
, and indeed costs at least $E_0$ (for the free case with $V=0$)
to
create (\ref{specdir}).
In the case of the KG equation the sign of the enrgy indicates the charge of
the particle/antiparticle, which have oposite charges.
Charge conjugation therefore interchanges between the two.
What is the meaning of the different signs of the energy
of the dipolar bose excitations in the bcc case ? From our definition of the second-quantized
description of the dipoles, the meaning of the sign of the energy is that the field of
resonating localized-dipoles can have two global configurations shifted
by $%
\pi $ (Fig.2). These two configurations are identical with respect to the magnitude
of the energy
spectrum, but in each the operation of spin flip changes from
up$\rightarrow$down to down$\rightarrow$up. We can therefore identify two ''charges'' for the bcc
to
distinguish between the two shifted phases. Further we find that as in the KG
case the operation of charge conjugation (which reverses the signs of the
dipoles) moves us between the two solutions.
\section{Appendix B: Symmetry breaking of the fermionic excitations.}
We see from (\ref{uvdir}) that when there is no fermion present (i.e. if
we
put $E_{0}=0$ in (\ref{linear})) the ground state has equal numbers of
fermions and holes. The symmetry between particles and holes is broken
by
the free fermion quasiparticle (or quasihole), and the sign of the symmetry-breaking
parameter $%
E_{0}$ determines which of the two kinds is present. The hole/particle
are
with respect to the equilibrium occupation by pairs of fermions in the
ground-state.
A single flipped dipole described as the fermionic excitation, breaks
the
symmetry between the number of up/down dipoles and creates a residual globally
oscillating dipole moment. The parity $P$ symmetry with respect to
reflection along the axis of the global dipole (let us choose to be $z$)
is
broken. The charge $C$ symmetry is also broken since the direction of the global dipole
is
flipped under charge conjugation. The time reversal symmetry $T$ is
unbroken
since the oscillating globel dipole does not define a unique time
direction.
We therefore have that the global $CPT$ symmetry is preserved, as is the
$CP$
and $T$ symmetries individually. The symmetry-breaking parameter in
(\ref
{diracham}) is the sign given to $E_{0}$, which corresponds to choosing
an
up or down dipole to flip. In second quantization langauge this is the
choice between an unpaired particle or hole. The broken symmetry is not
of
the $U(1)$ group such as the $\Phi $ (\ref{order}) order parameter, but
has
a $Z(1)$ discrete symmetry.
We now compare this with the situation of the two-dimensional massive
Dirac
particle \cite{berry}. The Hamiltonian describing a single fermionic
excitation (\ref{diracham}) can be written as
\begin{eqnarray}
i\hbar \frac{\partial }{\partial t}\left(
\begin{array}{c}
c_{k} \\
c_{-k}^{\dagger }
\end{array}
\right) &=&
E(k)\left(
\begin{array}{cc}
0 & 1 \\
1 & 0
\end{array}
\right) \left(
\begin{array}{c}
c_{k} \\
c_{-k}^{\dagger }
\end{array}
\right) \nonumber \\
&+&E_{0}\left(
\begin{array}{cc}
1 & 0 \\
0 & -1
\end{array}
\right) \left(
\begin{array}{c}
c_{k} \\
c_{-k}^{\dagger }
\end{array}
\right)
\label{eqmotionmat}
\end{eqnarray}
while the 2D Dirac particle is described by
\begin{equation}
i\hbar \frac{\partial }{\partial t}\left(
\begin{array}{c}
\varphi \\
\chi
\end{array}
\right) =c(\widehat{{\bf \sigma }}\cdot {\bf p})\left(
\begin{array}{c}
\varphi \\
\chi
\end{array}
\right) +mc^{2}\widehat{\sigma }_{z}\left(
\begin{array}{c}
\varphi \\
\chi
\end{array}
\right) \label{diracmat}
\end{equation}
By assuming momentum ${\bf p}$ in the $\widehat{x}$-direction only we
write
\begin{eqnarray}
i\hbar \frac{\partial }{\partial t}\left(
\begin{array}{c}
\varphi \\
\chi
\end{array}
\right) &=&c\left(
\begin{array}{cc}
0 & -i\hbar \partial _{x} \\
-i\hbar \partial _{x} & 0
\end{array}
\right) \left(
\begin{array}{c}
\varphi \\
\chi
\end{array}
\right) \nonumber \\
&+&mc^{2}\left(
\begin{array}{cc}
1 & 0 \\
0 & -1
\end{array}
\right) \left(
\begin{array}{c}
\varphi \\
\chi
\end{array}
\right)
\label{diracmat2d}
\end{eqnarray}
where $\widehat{{\bf \sigma }}=\left( \widehat{\sigma
}_{x},\widehat{\sigma }%
_{y},\widehat{\sigma }_{z}\right) $, $\widehat{\sigma }_{i}$ the Pauli
matrices, and $\varphi ,\chi $ are the particle/antiparticle scalar
wavefunctions. There is now complete analogy between
(\ref{eqmotionmat})
and (\ref{diracmat2d}). The symmetry-breaking parameter $E_{0}$ is
identical
to the $mc^{2}$ parameter in the 2D Dirac equations-of-motion (\ref
{diracmat2d}). The symmetry that is broken by choosing a non-zero
$mc^{2}$
in two-dimensions is the time-reversal symmetry (TRS) \cite{berry}. As
shown
in (\ref{diracmat}), the time component of the momentum-energy vector in
two
dimensions is taken by the $z$ axis. The broken parity $P$ in the
$z$-axis
for the fermionic excitation of the bcc phase is here replaced by the
TRS
breaking of the heavy two-dimensional Dirac particle.
Similar to the Anderson \cite{andersonbcs} transformation of the BCS
problem
to a magnetic Hamiltonian we can transform (\ref{diracham}) using:
\begin{eqnarray}
n_k &=&c_k^{\dagger }c_k\qquad c_k^{\dagger }c_{-k}^{\dagger }=\sigma
_k^{-}/2\qquad c_kc_{-k}=\sigma _k^{+}/2 \nonumber \\
&\Rightarrow &n_kn_{-k}=\frac 12\left( \sigma _k^z+1\right) \qquad
c_k^{\dagger}c_{-k}^{\dagger }+c_kc_{-k}=\sigma _k^x
\label{magtrans}
\end{eqnarray}
where the $\sigma _k^i$ are Pauli spin-1/2 operators. The basis is such
that
an up-spin in the $\widehat{z}$-direction represents an empty pair,
while a
down-spin represents an occupied pair.
The resulting Hamiltonian is:
\begin{equation}
H_{mag}=\sum \varepsilon _k\sigma _k^x-\frac 12\sum V_k\left( \sigma
_k^z+1\right) \label{hamag}
\end{equation}
This Hamiltonian describes a fictitious magnetic field acting on the
spin $%
\overrightarrow{\sigma }$:
\begin{equation}
\overrightarrow{B}=\varepsilon _k\widehat{x}-\frac 12E_0\widehat{z}
\label{mag}
\end{equation}
where we replaced the potential energy with the constant $V_k=E_0$
(\ref{linear}).
The magnetic field (\ref{mag}) can be compared with the BCS result
\cite{andersonbcs}
\begin{equation}
\overrightarrow{B}_{BCS}=\varepsilon _k\widehat{z}+\frac 12V\sum \left(
\sigma _k^x\widehat{x}+\sigma _k^y\widehat{y}\right) \label{magbcs}
\end{equation}
The alignment of the spins in the ground-state is shown for the two
Hamiltonians in
Fig.8.
In the BCS problem the sign of the symmetry-breaking field in $V$ (\ref
{magbcs}) has to be positive so that it induces ferro-magnetic interaction
between the fictitious spins, otherwise there will not be any rotation of the spins
across the Fermi-energy. Only when the spins rotate do they go through
the
point where the spin is entirely in the $xy$-plane. At this point the
state
has no defined occupation number but a well defined phase, while on both
sides of the domain-wall there is well defined occupation and no phase.
This
superconducting-phase at the fermi energy is just the angle of the fictitious spin in the $xy$-plane,
and
there is broken U(1) symmetry.
In the Dirac case the symmetry-breaking field $E_{0}$ in (\ref{mag}) is
a
constant external field. It can have both signs, which control the
direction
along the $z$-axis that the rotated spin has in the middle of the
domain-wall. This spin describes weather a particle or a hole is
occupied,
while away from the $k=0$ point the spins are in the $xy$-plane, with a
well
defined phase. The symmetry that is broken is therefore the binary Z(1) ($\pm
$) symmetry.
In comparison with the BCS problem we see that in the fermionic excitation in
the bcc the
symmetry-breaking parameter is a finite density of unpaired fermions: $%
\left\langle c_{k}^{\dagger }c_{k}\right\rangle \neq 0$ (\ref{linear}). The
ground-state
without unpaired fermions is a 'vacuum' of pairs of particle-holes in
equal
numbers. In the BCS problem the symmetry-breaking parameter is a finite
pair-density: $\left\langle c_{k}^{\dagger }c_{-k}^{\dagger
}\right\rangle
\neq 0$. The ground-state in the absence of electron pairing is just a
finite density of electrons below the fermi-energy and zero above. In
this
respect the two problems are 'complementary'.
|
\section{Introduction}
The possibility to access new characteristics of hadrons by means of the
deeply virtual Compton scattering \cite{MulRobGeyDitHor94,Ji96,Rad96} and
the hard diffractive hadron electroproduction \cite{Rad96,ColFraStr96}
processes has recently initiated a growing phenomenological interest in
the underlying non-perturbative elements --- the so-called off-forward
parton distributions (OFPD) --- which parametrize hadronic structure in
these reaction making use of the QCD factorization theorems. The main
feature of the processes is a non-zero skewedness, i.e.\ plus component,
$\Delta_+ = \eta$, of the $t$-channel momentum transfer $\Delta$.
One of the central issues which has been addressed in this context is the
description of the scaling violation phenomena in the cross section via the
evolution of the off-forward parton distributions. Since the OFPD is defined
as an expectation value of a non-local string operator, its $Q^2$-dependence
is governed by the renormalization of this operator Fourier transformed to
the momentum fraction space. Inasmuch as the generalized skewed kinematics
can be unambiguously restored \cite{GeyDitHorMulRob88} from the conventional
exclusive one, known as Efremov-Radyushkin-Brodsky-Lepage (ER-BL) region
$\eta = 1$, in what follows we deal formally with renormalization of the
ordinary distribution amplitudes which obey the ER-BL equation
\cite{EfrRad78,BroLep79}
\begin{equation}
\label{ER-BLequation}
\frac{d}{d \ln Q^2} \mbox{\boldmath$\phi$} (x, Q) =
\mbox{\boldmath$V$} \left(x, y | \alpha_s(Q) \right)
\mathop{\otimes}^{\rm e} \mbox{\boldmath$\phi$} (y, Q) .
\end{equation}
Here we have introduced the exclusive convolution
\begin{eqnarray*}
\mathop{\otimes}^{\rm e} \equiv \int_{0}^{1} dy ,
\end{eqnarray*}
to distinguish it from the inclusive one used later. Here
$\mbox{\boldmath$\phi$} = { {^Q\phi} \choose {^G\phi} }$ is the
two-dimensional vector and $\mbox{\boldmath$V$} (x, y | \alpha_s )$
is a $2 \times 2$-matrix of evolution kernels given by a series in
the coupling.
Several methods have been offered so far to solve the off-forward evolution
equation: numerical integration \cite{FraFreGuzStr98}, expansion of OFPD
w.r.t.\ an appropriate basis of polynomials \cite{Beletal97,ManPilWei97},
mapping to the forward case\footnote{This idea has earlier been applied
directly to the kernels in \cite{BelMul98a}.} \cite{Shu99,ShuGolBieMarRys99}
and solution in the configuration space \cite{BalBra89,KivMan99}.
The last three methods are based on the well-known fact that operators
with definite conformal spin do not mix in the one-loop approximation.
Beyond leading order the latter two methods can only be applied in the
formal conformal limit of QCD where the $\beta$-function is set equal
to zero and making use of the conformal subtraction scheme which removes
the special conformal symmetry breaking anomaly appearing in the minimal
subtraction scheme. Thus, we are only left with the former two methods
which allows for a successive improvement of the perturbative
approximations involved.
Up to now only orthogonal polynomial reconstruction method has allowed the
analysis of the scaling violation in the singlet sector in two-loop
approximation since only the anomalous dimensions required in the formalism
were available so far \cite{BelMul98a,Mue94,BelMul98b}. This was sufficient
to get a first insight into the NLO evolution corrections. However, in order
to have an access to the whole kinematical region, especially for small
$x$, $\eta$ and high precision handling of the $x \sim \eta$ domain, one
should look for a more efficient numerical treatment. This can be achieved
with the first method alluded to above. To do the direct numerical
integration of the evolution equation one needs the corresponding evolution
kernels whose Gegenbauer moments define the anomalous dimensions mentioned
earlier. For the time being the former were available at LO order only. The
flavour non-singlet ER-BL kernel ($\eta = 1$) was obtained in NLO by a
cumbersome analytical calculation \cite{Sar84,DitRad84,MikRad85}. As we
have mentioned above the continuation to $\eta \in [0,1]$ is a unique
procedure \cite{GeyDitHorMulRob88}, so that one can obtain in a simple way
the evolution kernels for OFPD. The goal of this paper is to outline a
method that allows one to construct the singlet ER-BL kernels by applying
conformal and supersymmetric constraints where the latter ones arise from
the ${\cal N} = 1$ super Yang-Mills theory \cite{BukFroKurLip85,BelMulSch98}.
In this way we can avoid the direct diagrammatical calculation which would
be very difficult to handle otherwise since no appropriate technology
has been developed yet.
The derivation is based on the fairly well established structure of the
ER-BL kernel in NLO. Up to two-loop order we have
\begin{equation}
\mbox{\boldmath$V$} (x, y | \alpha_s )
= \frac{\alpha_s}{2\pi}
\mbox{\boldmath$V$}^{(0)} (x, y)
+ \left( \frac{\alpha_s}{2\pi} \right)^2
\mbox{\boldmath$V$}^{(1)} (x, y)
+ {\cal O} (\alpha_s^3) ,
\end{equation}
with the purely diagonal LO kernel $\mbox{\boldmath$V$}^{(0)}$ in the
basis of Gegenbauer polynomials and NLO one separated in two parts:
$\mbox{\boldmath$V$}^{(1)} (x, y) = \mbox{\boldmath$V$}^{{\rm D}(1)}
(x, y) + \mbox{\boldmath$V$}^{{\rm ND}(1)} (x, y)$, with the diagonal
part which is entirely determined by the well-known forward DGLAP
splitting functions $\mbox{\boldmath$P$} (z)$ \cite{BelMul98a}
\begin{equation}
\label{PtoVDreduction}
{^{AB} V}^{\rm D} ( x, y )
= \int_{0}^{1} dz\,
\sum_{j = 0}^{\infty} \frac{w (y | \nu)}{N_j(\nu)}
C^{\nu (A)}_j (2x - 1) z^j\, {^{AB}\! P} (z)
C^{\nu (B)}_j (2y - 1),
\end{equation}
where $N_j(\nu)= 2^{ - 4 \nu + 1 }
\frac{ \Gamma^2 (\frac{1}{2}) \Gamma ( 2 \nu + j )}{\Gamma^2 (\nu)
( \nu + j ) j! }$ and $w (y | \nu) = (y \bar y)^{\nu-1/2}$ are the
normalization and weight factors, respectively. The non-diagonal piece
is fixed completely by the conformal constraints \cite{BelMul98b}
\begin{equation}
\label{NDkernel}
\mbox{\boldmath$V$}^{{\rm ND}(1)} (x, y)
= - ( {\cal I} - {\cal D} )\,
\left\{
\mbox{\boldmath$\dot V$} \mathop{\otimes}^{\rm e}
\left(
\mbox{\boldmath$V$}^{(0)} + \frac{\beta_0}{2}\, \hbox{{1}\kern-.25em\hbox{l}}
\right)
+
\left[
\mbox{\boldmath$g$} \mathop{\otimes}^{\rm e}_{,} \mbox{\boldmath$V$}^{(0)}
\right]_-
\right\} (x, y) ,
\end{equation}
in terms of
\begin{equation}
\mbox{\boldmath$V$}^{(0)}
=
\left(
\begin{array}{rr}
C_F\, {^{QQ}V^{(0)}} & 2 T_F N_f\, {^{QG}V^{(0)}} \\
C_F\, {^{GQ}V^{(0)}} & C_A\, {^{GG}V^{(0)}}
\end{array}
\right) ,
\quad
\mbox{\boldmath$g$}
=
\left(
\begin{array}{rr}
C_F\, {^{QQ}g} & 0 \\
C_F\, {^{GQ}g} & C_A\, {^{GG}g}
\end{array}
\right) ,
\end{equation}
the ER-BL kernels at LO and the special conformal symmetry breaking
matrix $\mbox{\boldmath$g$}$. Here $\beta_0 = \frac{4}{3} T_F N_f
- \frac{11}{3} C_A$ is the first expansion coefficient of the QCD
$\beta$-function. In the parity odd sector the dotted kernel,
$\mbox{\boldmath$\dot V$}$, is simply given by a logarithmic
modification of the $\mbox{\boldmath$V$}^{(0)}$. Due to subtleties,
appearing in the parity even case \cite{BelMul98a}, we deal here, for
the sake of simplicity, only with the parity odd and transversity
sectors.
The main problem is thus to restore the diagonal part of the
NLO kernels. Since the use of Eq.\ (\ref{PtoVDreduction}) beyond
LO is extremely complicated in practice, we are forced to look for
other solutions. It turns out that the bulk of contributions in the
ER-BL kernel can be deduced by going to the forward limit making use
of the reduction
\begin{eqnarray}
\label{SingletLimit}
\mbox{\boldmath$P$} (z)
= {\rm LIM}\, \mbox{\boldmath$V$} (x, y)
\equiv \lim_{\tau\to 0} \frac{1}{|\tau|}
\left(
\begin{array}{rr}
{^{QQ} V}
&
\frac{1}{\tau}{^{QG} V}
\\
\frac{\tau}{z} {^{GQ} V}
&
\frac{1}{z}{^{GG} V}
\end{array}
\right)^{\rm ext}
\left( \frac{z}{\tau}, \frac{1}{\tau} \right) .
\end{eqnarray}
Then the difference\footnote{Here $\mbox{\boldmath$V$}^{\rm ND}$
is understood without the $({\cal I} - {\cal D})$-projector.}
\begin{eqnarray*}
\mbox{\boldmath$P$} (z) - \mbox{\boldmath$P$}^{\rm cross-ladder} (z)
- {\rm LIM}\, \mbox{\boldmath$V$}^{\rm ND} (x, y)
\end{eqnarray*}
can be represented in terms of inclusive convolutions of simple splitting
functions and the back transformation to the exclusive kinematics is
trivial. The contributions of the purely diagonal cross-ladder diagrams
$\mbox{\boldmath$V$}^{\rm cross-ladder} (z)$ can be found from the known
$QQ$ sector \cite{Sar84,DitRad84,MikRad85} exploiting the ${\cal N} = 1$
supersymmetric constraints.
The paper is organized as follows. In the next section we analyze the
structure of the known flavour non-singlet ER-BL kernel and state
the benchmarks of the formalism. The structure observed will give
us a guideline for construction of all other kernels: quark chiral
odd sector is considered in Section 3 and parity odd flavour singlet
one is discussed in Section 4. Finally, we give our conclusions and an
outlook .
\section{Structure of ER-BL kernel in non-singlet sector.}
It is very instructive to demonstrate the machinery in the simplest
case of non-singlet sector. Since the explicit two-loop calculation
is available \cite{Sar84,DitRad84,MikRad85} the direct comparison can
be made. The NLO $QQ$-kernel can be decomposed in colour structures
as\footnote{We omit the superscript $QQ$ later in this section.}
\begin{eqnarray}
\label{kernel-NS}
V (x, y | \alpha_s)
&=& \frac{\alpha_s}{2\pi}\, C_F V^{(0)}(x, y) \nonumber\\
&+& \left( \frac{\alpha_s}{2\pi} \right)^2
C_F
\left[
C_F V_F (x, y)
- \frac{\beta_0}{2} V_\beta (x, y)
- \left( C_F - \frac{C_A}{2} \right) V_G (x, y)
\right]_+ \nonumber\\
&+& {\cal O} \left( \alpha_s^3 \right) ,
\end{eqnarray}
with the LO kernel $V^{(0)} (x, y) = \left[ v(x,y)\right]_+$, where
\begin{equation}
v(x,y) = \theta(y - x) f (x, y)
+ \theta(x - y) \overline f (x, y),
\quad\mbox{and}\quad
f (x, y) = \frac{x}{y} \left( 1 + \frac{1}{y - x} \right) .
\end{equation}
The shorthand notations $\bar x = 1 - x$ and $\overline f = f
(\bar x, \bar y)$ are used throughout the paper. The ``+''-prescription
is conventionally defined by
\begin{eqnarray*}
\left[ V (x, y) \right]_+ = V (x, y)
- \delta(x - y) \int_0^1 dz\, V (z, y).
\end{eqnarray*}
Let us now recall a few properties of the kernel that are useful for
the following considerations. Due to absence of the conformal symmetry
breaking counterterms at leading order for the renormalization of the
composite operators with total derivatives, one can use its consequences
to fix the eigenfunctions which turn out to be the Gegenbauer
polynomials $C_j^{3/2} (2x - 1)$ \cite{EfrRad78,BroLep79}. Thus, the
LO kernel is symmetric with respect to the weight function $x \bar x$:
$y \bar y V^{(0)} (x, y) = x \bar x V^{(0)} (y, x)$. Its eigenvalues are
given by the anomalous dimensions appeared in the analysis of deep
inelastic scattering. Thus, it is not surprising that a simple limit
already mentioned in Eq.\ (\ref{SingletLimit}) gives us the DGLAP kernel
\cite{GeyDitHorMulRob88}:
\begin{eqnarray}
\label{LIM-NS}
P(z) = {\rm LIM}\, V (x, y)
\equiv \lim_{\tau \to 0}
\frac{1}{|\tau|}
V^{\rm ext}\left(\frac{z}{\tau},\frac{1}{\tau}\right).
\end{eqnarray}
To perform this limit, we have to extend at first the ER-BL kernel,
originally defined in the domain $0 \leq x, y \leq 1$, to the whole
region $x, y \in ( - \infty, \infty )$ by a unique procedure which is
given in practice by the replacement, e.g.\ at leading order, of the
$\theta$-function by
\begin{eqnarray}
\label{Extension}
\theta (y - x)
\to
\theta \left( 1 - \frac{x}{y} \right)
\theta \left( \frac{x}{y} \right)
\mbox{sign} (y).
\end{eqnarray}
If a kernel is diagonal in the ER-BL representation, we can restore it
from the known DGLAP kernel by the integral transformation
(\ref{PtoVDreduction}). Because of branch cuts appearing in the
convolutions of the NLO terms with the transformation kernel, it is highly
nontrivial to handle the inverse reduction to the exclusive kinematics.
At NLO the kernel (\ref{kernel-NS}) contains besides a pure diagonal part
with respect to the Gegenbauer polynomials also a non-diagonal part
located in $V_F (x, y)$ and $V_\beta (x, y)$. These parts are predicted by
conformal constraints (see $QQ$-entry of Eq.\ (\ref{NDkernel})) and are
fixed by the one-loop special conformal anomaly kernels
\cite{BelMul98a,Mue94,BelMul98b}:
\begin{equation}
\label{def-dV-NS}
\dot v (x, y)
= \theta(y - x) f (x, y) \ln \frac{x}{y}
+ \left\{ x \to \bar x \atop y \to \bar y \right\},
\quad
g (x, y)
= - \theta(y - x)
\frac{ \ln \left( 1 - \frac{x}{y} \right) }{y - x}
+ \left\{ x \to \bar x \atop y \to \bar y \right\} .
\end{equation}
Let us now analyze in detail the contributions to the NLO kernel from
different colour structures. The expressions for $C_F^2$ terms
arise from Feynman diagrams containing quark self-energy insertions
and ladder graphs\footnote{For simplicity we imply the diagrams
in the light-cone gauge \cite{Sar84,DitRad84}.}. In order to subtract
the ultraviolet (UV) divergences in subgraphs it requires the LO
renormalization of the composite operator to which these lines are
attached to. The explicit calculation gives \cite{Sar84,DitRad84,MikRad85}
\begin{eqnarray}
\label{kernel-NS-CFa}
V_F (x, y)
&=& \theta (y - x)
\Bigg\{ \left( \frac{4}{3} - 2 \zeta (2) \right) f
+ 3 \frac{x}{y}
- \left( \frac{3}{2} f - \frac{x}{2 \bar y} \right) \ln \frac{x}{y}
- ( f - \overline f ) \ln \frac{x}{y}
\ln \left( 1 - \frac{x}{y} \right)
\nonumber \\
&+&
\left( f + \frac{x}{2 \bar y} \right) \ln^2 \frac{x}{y} \Bigg\}
- \frac{x}{2\bar y} \ln x \left( 1 + \ln x - 2 \ln \bar x \right)
+ \left\{ {x \to \bar x \atop y \to \bar y } \right\}.
\end{eqnarray}
Making use of the known non-diagonal part (\ref{NDkernel}), $V_F$ can be
represented up to a pure diagonal term, denoted as $D_F(x,y)$, by the
convolution
\begin{equation}
V_F (x, y) =
- \left( \dot{v} \mathop{\otimes}^{\rm e} v
+ g \mathop{\otimes}^{\rm e} v - v \mathop{\otimes}^{\rm e} g \right) (x, y)
+ D_F (x, y).
\end{equation}
To find an appropriate representation of this missing diagonal
element we first take the forward limit. Since the forward limit
of the convolution is\footnote{We remind as well that $[A]_+ \mathop{\otimes} [B]_+
= [C]_+$.}
\begin{equation}
{\rm LIM}\, \left\{ [\dot v]_+ \mathop{\otimes}^{\rm e} [v]_+ \right\} =
\left\{ {\rm LIM}\, [\dot v]_+ \right\}
\mathop{\otimes}^{\rm i} \left\{ {\rm LIM}\, [v]_+ \right\} ,
\end{equation}
where we have introduced the inclusive convolution
\begin{eqnarray*}
P_1 (z) \mathop{\otimes}^{\rm i} P_2 (z)
\equiv \int_0^1 dx \int_0^1 dy \delta( z - xy ) P_1 (x) P_2 (y) ,
\end{eqnarray*}
and the commutator $g \mathop{\otimes} V^{(0)} - V^{(0)} \mathop{\otimes} g$ drops out in the
forward limit, we obtain
\begin{eqnarray}
{\rm LIM}\, V_F (x, y)
= - \dot p \mathop{\otimes}^{\rm i} p
+ {\rm LIM}\, D_F (x,y) ,
\end{eqnarray}
where $\dot p = {\rm LIM}\, \dot v = p(z) \ln z + 1 - z$ and
$p (z) = {\rm LIM}\, v (x, y) = (1 + z^2)/(1 - z)$. The comparison
of ${\rm LIM}\, V_F (x, y)$ with the corresponding part of the DGLAP
kernel \cite{CurFurPet80}
\begin{eqnarray}
P_F (z)
&=& \left\{ \frac{4}{3} - 2 \zeta (2)
- \frac{3}{2} \ln z + \ln^2 z - 2\ln z \ln(1 - z)
\right\} p(z) \nonumber\\
&+& 1 - z + \frac{1 - 3 z}{2} \ln z - \frac{1 + z}{2} \ln^2 z ,
\end{eqnarray}
yields the result in which all double log terms are contained
in the convolution $\dot p \mathop{\otimes} p$ and, therefore, only
single logs survive in $D_F (z) = {\rm LIM} D_F (x, y)$:
\begin{eqnarray}
D_F(z) &=& P_F + \dot p \mathop{\otimes}^{\rm i} p \nonumber\\
&=& - \frac{1}{2} p^a (-z) \ln z
- p^a (z) \left\{ \ln z - 2 \ln (1 - z) - \frac{1}{2} \right\}
- \frac{5}{12} p (z) .
\end{eqnarray}
Here we have introduced for convenience the kernel $p^a (z) = 1 - z$.
The next important point is that the remaining log terms can be
represented as convolutions of $p^a$ and $p$. Thus, we have finally
\begin{eqnarray}
D_F (z)
= \frac{1}{2} p^a \mathop{\otimes}^{\rm i} \left\{ 2\, p + p^a \right\} (z)
+ \frac{1}{12} p(z) + \frac{5}{2} p^a(z) .
\end{eqnarray}
Since $D_F (x, y)$ is by definition diagonal, the extension of $D_F (z)$
towards the ER-BL kinematics is trivial:
\begin{eqnarray}
\label{DF-QQ}
D_F (z) \to D_F (x, y)
= \frac{1}{2} v^a \mathop{\otimes}^{\rm e}
\left( 2\, v + v^a \right) (x, y)
+ \frac{1}{12} v (x, y) + \frac{5}{2} v^a (x, y),
\end{eqnarray}
where a new diagonal element is $v^a (x,y) = \theta(y - x)
\frac{x}{y} + \theta(x - y) \frac{\bar x}{\bar y}$. Evaluating
the convolutions one can establish the equivalence of our prediction
with Eq.\ (\ref{kernel-NS-CFa}).
Next, the Feynman diagrams containing vertex and self-energy corrections
provide $V_\beta$ proportional to $\beta_0$. Its off-diagonal part is
induced by the renormalization of the coupling and is contained in the
dotted kernel (\ref{def-dV-NS})
\begin{equation}
V_\beta (x, y) = \dot v (x, y) + D_\beta (x, y) .
\end{equation}
The remaining diagonal piece, $D_\beta$, is deduced from the known NLO
DGLAP kernel
\cite{CurFurPet80}
\begin{equation}
\label{kernelP-NS-beta}
P_\beta (z) = \frac{5}{3} p (z) + p^a (z) + \dot p (z)
\end{equation}
by going to the forward kinematics and restoring then the missed
contributions from it. Thus,
\begin{equation}
\label{Dbeta-QQ}
D_\beta (x, y) = \frac{5}{3} v (x, y) + v^a (x, y) .
\end{equation}
Indeed, the final result coincides with \cite{Sar84,DitRad84,MikRad85}.
Finally, we come to the contribution which mainly originates from the
crossed ladder diagram proportional to $(C_F - C_A/2)$:
\begin{eqnarray}
\label{kernel-NS-CAa}
V_G (x, y)
= 2 v^a (x, y) + \frac{4}{3} v (x, y)
+ \left( G (x, y)
+ \left\{ x \to \bar x \atop y \to \bar y \right\} \right).
\end{eqnarray}
Since this diagram has no UV divergent subgraph and thus requires no
subtraction, its contribution has to be diagonal w.r.t.\ the Gegenbauer
polynomials. This is obvious for the first two terms appearing in
Eq.\ (\ref{kernel-NS-CAa}). The function\footnote{We have slightly
changed the original definition given in \cite{Sar84,DitRad84} by
$G (x, y) + 2 \theta (y - x) \overline f \ln y \ln \bar x \to G (x, y)$.}
$G(x,y)$ contains in the unusual $\theta (y - \bar x)$-structure
the mixing between quarks and antiquarks
\begin{equation}
\label{kernel-NS-G}
G (x, y) = \theta (y - x) H (x, y)
+ \theta (y - \bar x) \overline H (x, y),
\end{equation}
with
\begin{eqnarray}
H (x, y)
&=& 2 \left[ \overline f
\left( {\rm Li}_2 (\bar x) + \ln y \ln \bar x \right)
- f\, {\rm Li}_2 (\bar y) \right] , \\
\overline H (x, y)
&=& 2 \left[ ( f - \overline f )
\left( {\rm Li}_2 \left( 1 - \frac{x}{y} \right)
+ \frac{1}{2} \ln^2 y \right)
+ f \left( {\rm Li}_2 (\bar y) - {\rm Li}_2 (x)- \ln y \ln x \right)
\right] ,
\end{eqnarray}
where ${\rm Li}_2$ is the dilogarithm. It can be easily checked that
the $G$-contribution (not the terms $H$ and $\overline H$ separately)
is symmetrical w.r.t.\ the weight $x \bar x$. Performing the limit
(\ref{LIM-NS}) we obtain the following correspondence with the
non-singlet DGLAP kernel \cite{MulRobGeyDitHor94}:
\begin{equation}
\label{G-NS}
G(z) \equiv {\rm LIM} G (x, y)
= \theta(z) \theta(1 - z) H (z)
+ \theta(- z) \theta(1 + z) \overline H (z) ,
\end{equation}
where
\begin{eqnarray}
\label{LIM-H-NS}
H (z) &\equiv& {\rm LIM}\, H (x, y)
= p(z) \left( \ln^2 z - 2 \zeta (2) \right) + T (z) , \\
\label{LIM-bH-NS}
\overline H (z) &\equiv& {\rm LIM}\, \overline H (x, y)
= 2 p(z) S_2(- z) + T (-z) .
\end{eqnarray}
Here $S_2 (z) = \int_{z/(1+z)}^{1/(1+z)} \frac{dx}{x}\ln\frac{1 - x}{x}$
and $T (z) = 2 (1 + z) \ln z + 4(1 - z)$. We should emphasize that
this structure of $G$ is the most general, especially, it is the
only contribution that contains Spence functions. In the forward
limit we obtain therefore a typical combinations given in
(\ref{LIM-H-NS}) and (\ref{LIM-bH-NS}), which can be found in all
other channels as well. This observation provides us with a hint for the
construction of all singlet $G$ kernels in the ER-BL representation.
For completeness, we give the corresponding part of the DGLAP kernel
\cite{CurFurPet80}
\begin{eqnarray}
\label{kernelP-NS-CA}
P_G (z) = {\rm LIM}\, V_G (x, y)
= 2 p^a (z) + \frac{4}{3} p(z)+ G(z),
\end{eqnarray}
and $G(z)$ defined above in Eqs.\ (\ref{G-NS})-(\ref{LIM-bH-NS}).
Recapitulating the results obtained in this section, we have observed a
rather simple structure of the non-singlet NLO kernel in the $QQ$-channel.
Up to the diagonal $G$-function, which is in fact the only new element in
the two-loop approximation, we can represent all other terms by a simple
convolution of LO kernels already known. It is not accidental but
a mere consequence of the topology of contributing Feynman graphs
at ${\cal O} (\alpha_s^2)$. Thus, we anticipate the same feature to appear
in all other channels as well.
\section{Quark kernel in chiral odd sector.}
After we have outlined and tested in the preceding section our
formalism, we can apply it to the previously unknown transversity
two-loop ER-BL kernel. We decompose the transversity kernel analogous
to the chiral even case (\ref{kernel-NS}). We also use the same
decomposition for the DGLAP kernels \cite{Vog97}. The leading order
kernel is
\begin{equation}
\label{kernel-tr-0}
V^{(0)T} (x, y)
= \left[ v^b (x, y) \right]_+ - \frac{1}{2} \delta(x - y),
\end{equation}
with
\begin{equation}
v^b (x, y) = \theta(y - x) f^b (x, y)
+ \theta(x - y) \overline f^b (x, y),
\quad\mbox{and}\quad
f^b (x, y) = \frac{x}{y} \frac{1}{y - x}.
\end{equation}
The non-diagonal part has been analyzed in Ref.\ \cite{BelMul98b} and
is completely analogous to the chiral even case discussed above. Thus,
\begin{eqnarray}
V_F^T (x, y)
= - \left\{
\left[ \dot v^b \right]_+ \mathop{\otimes}^{\rm e} V^{(0)T}
+ \left[ g \right]_+ \mathop{\otimes}^{\rm e} V^{(0)T}
- V^{(0)T} \mathop{\otimes}^{\rm e} \left[ g \right]_+
\right\} (x, y) + D_F^T (x, y),
\end{eqnarray}
where $\dot v^b$ is obtained from Eq.\ (\ref{def-dV-NS}) by
replacing $f$ by $f^b$ and the $g$ kernel is the same as in Eq.\
(\ref{def-dV-NS}). Taking the forward limit of $V_F^T (x, y)$
and comparing it with the known result for the DGLAP kernel in
Ref.\ \cite{Vog97}, we find the following trivial representation of
the remaining diagonal part
\begin{eqnarray}
D_F^T (x, y)=
- \frac{2}{3} \left[ v^b (x, y) \right]_+
- \frac{19}{24} \delta(x - y).
\end{eqnarray}
There is essentially no extra work required to find the contribution
proportional to the $\beta_0$-function, since it can be easily traced
from the DGLAP kernel \cite{Vog97} to be
\begin{eqnarray}
\label{kernel-tr-beta}
V_\beta^T (x, y)
= \frac{5}{3} \left[ v^b (x, y) \right]_+
+ \left[ \dot v^b (x, y) \right]_+
- \frac{13}{12} \delta(x - y).
\end{eqnarray}
The case of the $G$ function is easy to handle as well. If we replace
$f$ by $f^b$ in the definition (\ref{kernel-NS-G}), we obtain the
diagonal $G^T (x, y)$ kernel. Taking the forward limit and comparing
it with the DGLAP kernel, we immediately find the remaining
$\delta$-function contribution, so that the whole result reads
\begin{eqnarray}
\label{kernel-tr-CAa}
V^T_G (x, y)
= \left[
G^T (x, y)
+ \left\{ {x \to \bar x \atop y \to \bar y } \right\}
\right]_+
- \frac{19}{6} \delta(x - y).
\end{eqnarray}
This completes the discussion of the quark chiral-odd channel.
\section{Flavour singlet parity odd sector.}
Let us now address the flavour singlet parity odd sector responsible
for the evolution of axial-vector distribution amplitudes. For even parity
there are few subtleties, which will be discussed elsewhere. Here
we would only like to note that in the latter case a direct leading order
calculation provides a result that suffers for the mixed channel from
off-diagonal matrix elements in the unphysical sector. Although the
improved result has been found in Ref.\ \cite{BelMul98a}, it still
remains a difficult task to find an appropriate representation for the
dotted kernels and the two-loop $G$ functions.
Making use of the known non-diagonal part of the ER-BL kernel
(\ref{NDkernel}), the whole NLO result in the axial-vector case reads
\begin{equation}
\label{pred-Sing}
\mbox{\boldmath$V$}^{(1) A}
= - \mbox{\boldmath$\dot V$}^{A} \mathop{\otimes}^{\rm e}
\left(
\mbox{\boldmath$V$}^{(0)A} + \frac{\beta_0}{2}\, \hbox{{1}\kern-.25em\hbox{l}}
\right)
- \mbox{\boldmath$g$} \mathop{\otimes}^{\rm e} \mbox{\boldmath$V$}^{(0)A}
+ \mbox{\boldmath$V$}^{(0)A} \mathop{\otimes}^{\rm e} \mbox{\boldmath$g$}
+ \mbox{\boldmath$D$}^{A} + \mbox{\boldmath$G$}^{A},
\end{equation}
where the kernels $\mbox{\boldmath$D$}^{A} (x, y)$ and
$\mbox{\boldmath$G$}^{A} (x, y)$ are purely diagonal. Here the matrix
of the LO kernels is given in a compact form by
\begin{eqnarray}
\mbox{\boldmath$V$}^{(0)A} (x, y)
=
\left(
\begin{array}{ll}
C_F \left[ {^{QQ} v} (x, y) \right]_+
& - 2 T_F N_f \, {^{QG} v^a} (x, y) \\
C_F\, {^{GQ} v^a} (x, y)
& C_A \left[ {^{GG} v^A} (x, y) \right]_+
- \frac{\beta_0}{2} \delta(x - y)
\end{array}
\right) ,
\end{eqnarray}
where
${^{QQ} v} \equiv {^{QQ} v^a} + {^{QQ} v^b}$ and
${^{GG} v^A} \equiv 2\, {^{GG} v^a} + {^{GG} v^b}$.
The general structure of the functions $v^i$ is
\begin{equation}
{^{AB} v^i}(x, y)
= \theta(y - x) {^{AB}\! f^i}(x, y)
\pm \left\{ {x \to \bar x \atop y \to \bar y } \right\}
\quad
\mbox{for}
\quad
\left\{ {A = B \atop A \not = B } \right. ,
\end{equation}
with
\begin{equation}
\left\{{ {^{AB}\! f^a} \atop {^{AB}\! f^b} }\right\}
= \frac{ x^{\nu(A) - 1/2}}{y^{\nu(B) - 1/2}}
\left\{ { 1 \atop \frac{1}{y - x} } \right\} .
\end{equation}
The index $\nu(A)$ coincides with the index of Gegenbauer polynomials
in the corresponding channel, i.e.\ $\nu(Q) = 3/2$ and $\nu(G) = 5/2$.
The dotted kernels involved in the definition (\ref{pred-Sing}) can
simply be obtained by differentiating LO results w.r.t.\ the index $\nu$
which gives rise to the additional $\ln(x/y)$-multiplier in front of the
former
\begin{equation}
\mbox{\boldmath$\dot V$}^{(0)A} (x, y)
=
\left(
\begin{array}{ll}
C_F \left[ {^{QQ} \dot v} (x, y) \right]_+
& - 2 T_F N_f {^{QG} \dot v}^a (x, y) \\
C_F {^{GQ} \dot v}^a (x, y)
&
C_A \left[ {^{GG} \dot v}^A (x, y) \right]_+
\end{array}
\right) ,
\end{equation}
with the matrix elements
\begin{equation}
{^{AB} \dot v} (x, y) =
\theta(y - x) {^{AB}\! f} (x, y) \ln \frac{x}{y}
\pm \left\{ {x \to \bar x \atop y \to \bar y } \right\} ,
\quad
\mbox{for}
\quad
\left\{ { A = B \atop A \not= B } \right. .
\end{equation}
Note that for $A = B$ the dotted kernels are defined with the
``+''-prescription. The $\mbox{\boldmath$g$}$ function is given by
\begin{eqnarray}
\label{set-g-kernels}
\mbox{\boldmath$g$} (x, y) =
\theta(y - x)
\left(
\begin{array}{cc}
- C_F \left[ \frac{ \ln \left( 1 - \frac{x}{y} \right) }{y - x} \right]_+
& 0 \\
C_F \frac{x}{y}
& - C_A\left[ \frac{ \ln \left( 1 - \frac{x}{y} \right) }{y - x} \right]_+
\end{array}
\right)
\pm
\left\{ x \to \bar x \atop y \to \bar y \right\},
\end{eqnarray}
with ($-$) $+$ sign corresponding to (non-) diagonal elements.
Note, that we have used the property $({\cal I} - {\cal D}) \ln (1 - \frac{x}{y})
= - ({\cal I} - {\cal D}) \frac{x}{y}$ for the element of $GQ$-channel to make
contact with the results of Ref.\ \cite{BelMul98b}.
Next we construct the diagonal $\mbox{\boldmath$G$} (x, y)$ kernel. At
first glance one would naively expect that one can obtain these kernels
by only inserting appropriate ${^{AB}\! f}$ functions in the
definition (\ref{kernel-NS-G}), so that the symmetry properties of the
$f$ functions w.r.t.\ the weight induce then the desired symmetry of
the ${^{AB} G}$ functions. Unfortunately, the symmetry is not sufficient
for the diagonal form of the $\mbox{\boldmath$G$} (x, y)$ kernel. To
ensure the diagonality, we have to add terms containing single logs and
rational functions. Let us define the matrix
\begin{equation}
\label{G-kernel-odd}
\mbox{\boldmath$G$}^A (x, y)
= - \frac{1}{2}
\left(
\begin{array}{cc}
2 C_F \left( C_F - \frac{C_A}{2} \right)
\left[ {^{QQ} G}^A (x, y) \right]_+
&
2 C_A T_F N_f \, {^{QG} G}^A (x, y)
\\
C_F C_A \, {^{GQ} G}^A (x, y)
&
C_A^2 \left[ {^{GG} G}^A (x, y) \right]_+
\end{array}
\right) ,
\end{equation}
with the following general structure
\begin{equation}
{^{AB} G}^A (x, y)
= \theta (y - x)
\left( {^{AB}\! H}^A + \Delta{^{AB}\! H}^A \right) (x, y)
+ \theta (y - \bar x)
\left( {^{AB} \overline H}^A
+ \Delta{^{AB} \overline H}^A \right) (x, y) .
\end{equation}
Here analogous to the non-singlet case we set
\begin{eqnarray}
\label{kernel-S-H}
{^{AB} H}^A (x, y)
\!\!&=&\!\! 2 \left[ \pm {^{AB} \overline f}^A
\left( {\rm Li}_2( \bar x ) + \ln y \ln \bar x \right)
- {^{AB}\! f}^A\, {\rm Li}_2( \bar y ) \right], \\
\label{kernel-S-bH}
{^{AB} \overline{H}}^A (x, y)
\!\!&=&\!\! 2 \left[
\left( {^{AB}\! f}^A \mp {^{AB} \overline f}^A \right)
\left( {\rm Li}_2 \left( 1 - \frac{x}{y} \right)
+ \frac{1}{2} \ln^2 y \right)
+ {^{AB}\! f}^A \left( {\rm Li}_2 ( \bar y )
- {\rm Li}_2 (x) - \ln y \ln x \right) \right],
\nonumber\\
\end{eqnarray}
where the upper (lower) sign corresponds to the $A = B$ ($A \not= B$)
channels. An explicit use of the reduction $P \to V^{\rm D}$ procedure
(\ref{PtoVDreduction}) to restore the $\Delta H$ contributions is rather
involved due to complexity of the integrand function. Rather we have
succeeded to deduce them using different arguments. Since the crossed
ladder diagrams have no UV divergent subgraphs the kernels ${^{AB} G}$
in different channels are related in a scheme independent way by
supersymmetry and conformal covariance of ${\cal N} = 1$ super Yang-Mills
theory \cite{BelMulSch98}. Employing these symmetries we
restore\footnote{The details will be presented elsewhere.} all
necessary terms in a straightforward manner to be
\begin{eqnarray}
\Delta{^{QQ} H}^A (x, y)
&=& \Delta{^{QQ} \overline H}^A (x, y) = 0,
\\
\Delta {^{QG} H}^A (x, y)
&=& 2 \frac{\bar x}{y \bar y} \ln\bar x - 2 \frac{x}{y \bar y} \ln y,
\quad
\Delta{^{QG} \overline H}^A (x, y)
= 2 \frac{x}{y \bar y} \ln x - 2 \frac{\bar x}{y \bar y} \ln y,
\\
\Delta{^{GQ} H}^A (x, y)
&=& 2 \frac{x \bar x}{y} \ln\bar x - 2 \frac{x \bar x}{\bar y} \ln y,
\quad
\Delta{^{GQ} \overline H}^A (x, y)
= - 2 \frac{x \bar x}{y} \ln x + 2 \frac{x \bar x}{\bar y} \ln y,
\\
\Delta{^{GG} H}^A (x, y)
&=& \frac{x^2}{y^2} - \frac{1 + (x - y)^2}{y^2 \bar y^2}
- 2 \frac{x \bar x}{\bar y^2} \ln \frac{x}{y}
+ 2\frac{\bar x (\bar x - x)}{y \bar y} \ln\bar x
- 2 \frac{x(\bar x - x)}{y \bar y} \ln y,
\\
\Delta{^{GG} \overline H}^A (x, y)
&=& 2 \frac{x}{y^2} - \frac{x^2}{\bar y ^2}
+ 2 \frac{1 - x \bar x}{y \bar y^2}
+ 2 \frac{x \bar x}{y^2} \ln\frac{\bar x}{x}
+ 2 \frac{(x + \bar y) \bar x}{y \bar y^2} \ln \frac{x}{y}
- 2 \frac{1 - x \bar x}{y \bar y} \ln x
+ 6 \frac{x \bar x}{y \bar y} \ln y . \nonumber
\end{eqnarray}
Finally, we have to extract the remaining diagonal piece
$\mbox{\boldmath$D$}^A$ of $\mbox{\boldmath$V$}^A$ in the forward
limit (\ref{SingletLimit}) from the known DGLAP kernel
$\mbox{\boldmath$P$}^A$ \cite{MerNeeVog96}. We take into account
the underlying symmetry of the singlet parton distributions to map
the antiparticle contribution, i.e. $z < 0$, into the region $z > 0$.
As expected we find from
\begin{eqnarray}
\mbox{\boldmath$D$}^A (z)
= \mbox{\boldmath$P$}^A(z)
- {\rm LIM}
\left\{
- \mbox{\boldmath$\dot V$} \mathop{\otimes}^{\rm e}
\left( \mbox{\boldmath$V$}^{(0)A} + \frac{\beta_0}{2} \hbox{{1}\kern-.25em\hbox{l}} \right)
- \mbox{\boldmath$g$} \mathop{\otimes}^{\rm e} \mbox{\boldmath$V$}^{(0)A}
+ \mbox{\boldmath$V$}^{(0)A} \mathop{\otimes}^{\rm e} \mbox{\boldmath$g$}
+ \mbox{\boldmath$G$}^{A}
\right\}
\end{eqnarray}
a simple convolution-type representation for ER-BL kernels which can be
immediately deduced from the forward results for singlet $QQ$-channel
\begin{eqnarray}
\label{D-QQ-o}
{^{QQ}\! D}^A
= C_F^2 \left[ D_F \right]_+
- C_F \frac{\beta_0}{2} \left[ D_\beta \right]_+
- C_F \left( C_F - \frac{C_A}{2} \right)
\left[ \frac{4}{3} {^{QQ} v} + 2\, {^{QQ} v}^a \right]_+
- 6\, C_F T_F N_f {^{QQ} v}^a,
\end{eqnarray}
where $D_F$, $D_\beta$ are given by Eqs.\ (\ref{DF-QQ}) and
(\ref{Dbeta-QQ}), respectively. The rest of channels is expressed as
\begin{eqnarray}
{^{QG} D}^A
&=& 3\, C_F T_F N_f
\left\{
{^{QQ} v}^a \mathop{\otimes}^{\rm e} {^{QG} v}^a
- \frac{1}{2} {^{QG} v}^a
\right\} \\
&-& 2\, C_A T_F N_f
\left\{
3\, {^{QQ} v}^a \mathop{\otimes}^{\rm e} {^{QG} v}^a
+ \left[ 1 + 2 \zeta (2) \right] {^{QG} v}^a
\right\}, \nonumber\\
{^{GQ} D}^A
&=& C_F^2
\left\{ \frac{1}{2}
\left[ {^{GG} v}^A \right]_+ \mathop{\otimes}^{\rm e} {^{GQ} v}^a
- \frac{3}{2} {^{GQ} v}^a
\right\}
- C_F \frac{\beta_0}{2}
\left\{
{^{GQ} v}^a \mathop{\otimes}^{\rm e} \left[ {^{QQ} v} \right]_+
- \frac{1}{6} {^{GQ} v}^a
\right\} , \\
&-& C_F C_A
\left\{
\frac{3}{2} \left[ {^{GG} v}^A \right]_+ \mathop{\otimes}^{\rm e} {^{GQ} v}^c
+ \left[
2 \left[ {^{GG} v}^A \right]_+ - \frac{1}{2} {^{GG} v}^a
\right] \mathop{\otimes}^{\rm e} {^{GQ} v}^a
- \left[ \frac{7}{3} - 2 \zeta (2) \right] {^{GQ} v}^a
\right\} , \nonumber\\
{^{GG} D}^A
&=& C_A^2
\left\{
\left[
\left[ {^{GG} v}^A \right]_+ + \frac{1}{2} {^{GG} v}^a
\right] \mathop{\otimes}^{\rm e} {^{GG} v}^a
+ \frac{2}{3} \left[ {^{GG} v}^A \right]_+
- \frac{1}{4} {^{GG} v}^a - 2 \delta(x - y)
\right\} \\
&-& C_A \frac{\beta_0}{2}
\left\{
- \frac{1}{2} {^{GG} v}^a \mathop{\otimes}^{\rm e} {^{GG} v}^a
+ \frac{5}{3} \left[ {^{GG} v}^A \right]_+
+ {^{GG} v}^a + 2 \delta(x - y)
\right\} \nonumber\\
&-& C_F T_F N_f
\left\{
{^{GG} v}^a \mathop{\otimes}^{\rm e} {^{GG} v}^a
- {^{GG} v}^a + \delta(x - y)
\right\} , \nonumber
\end{eqnarray}
where we have introduced a new kernel
\begin{eqnarray}
\label{kernel-c}
{^{GQ} v^c}(x,y)=
\theta(y-x) \frac{x^2}{y}\left(2 \bar{x}y-\bar{y} \right) -
\left\{x\to \bar{x} \atop y\to \bar{y} \right\}.
\end{eqnarray}
These results provide us with the explicit parity odd singlet
evolution kernels.
\section{Conclusions.}
In this paper, we have presented a simple method for construction of
the exclusive evolution kernels in NLO from the knowledge of the
conformal anomalies and the available two-loop splitting functions.
The main task was, of course, the reconstruction of the diagonal part of
the kernel in the basis of Gegenbauer polynomials.
In the course of study we have established convolution-type formulae for
the bulk of contributing two-loop graphs with an exception of cross-ladder
diagrams. The complications which arise in the restoration of the latter
from the known forward kernels has been overcome making use of ${\cal N} = 1$
supersymmetric constraints \cite{BelMulSch98}. The former feature suggests
that by disentangling the topology of corresponding diagrams, it might
allow for an effective and facilitated way of explicit calculation. One
may expect that this property persists for a subset of diagrams at higher
orders and can be used, e.g.\ for diagrammatical derivation of NNLO
splitting functions.
The details of the present formalism together with the flavour singlet
parity even case, where new subtleties appear, will be discussed elsewhere.
\vspace{1cm}
A.B. was supported by the Alexander von Humboldt Foundation.
|
\section{Introduction}
\label{s:intro}
In the evaluation of Feynman integrals, one often needs integrals
and sums related to the dilogarithm and the Riemann zeta function.
This is particularly the case when one considers multi-loop
amplitudes (see \cite{Ritbergen}).
There is actually an intriguing connection between Feynman diagrams,
topology and number theory, which has recently been elucidated by
several authors,
in particular by Broadhurst \cite{Broadhurst}, Kreimer \cite{Kreimer}
and collaborators
(see also Groote, K\"orner and Pivovarov \cite{Groote}).
Many results of this kind have been compiled by Devoto and Duke \cite{D&D}
and by K\"olbig et al \cite{Kolbig}, in addition to those of the standard
tables \cite{Prudnikov}.
In an earlier paper \cite{OgreidOsland} (henceforth referred to as Paper
I\footnote{Often we will
refer to results and identities from our first article on this
subject. Whenever we quote e.g.\ equation (I.13) or
(I.B.2) we are referring to equation (13) or (B.2) in
\cite{OgreidOsland}, respectively.})
we presented results for sums required in the evaluation of Feynman
integrals, related to the Euler series. Several of these series
involve the digamma or psi function.
One such example is the series
\begin{eqnarray}
\sum_{n=1}^\infty\frac{1}{n^2}[\gamma+\psi(n)]\nonumber
\end{eqnarray}
which equals $\zeta(3)$ when summed.
Here, we present further results of this kind, many of which are obtained
using known properties of hypergeometric functions.
The sums of these new series are of the form
\begin{eqnarray}
R_3\zeta(3)+R_2\zeta(2)+R_0\nonumber
\end{eqnarray}
where $R_i$ are all rational numbers.
The starting point of this article will be the well-known result:
\begin{series}\label{series:series1}
\begin{eqnarray}
\sum_{n=1}^\infty\frac{1}{n(n+1)}=1\label{s1}
\end{eqnarray}
\end{series}
This is easily found by recognizing the sum as $\tfrac{1}{2}\ {}_2F_1(1,1;3;1)$
and then using (\ref{2F1unity}).
Using the definition of the Riemann zeta function
along with partial fractioning,
we find the following results as immediate corollaries of
\seriesref{series:series1}:
\begin{series2-6}
\begin{eqnarray}
&&\sum_{n=1}^\infty\frac{1}{n^2(n+1)}=\zeta(2)-1\label{s2}\\
&&\sum_{n=1}^\infty\frac{1}{n(n+1)^2}=-\zeta(2)+2\label{s3}\\
&&\sum_{n=1}^\infty\frac{1}{n^3(n+1)}=\zeta(3)-\zeta(2)+1\label{s4}\\
&&\sum_{n=1}^\infty\frac{1}{n^2(n+1)^2}=2\zeta(2)-3\label{s5}\\
&&\sum_{n=1}^\infty\frac{1}{n(n+1)^3}=-\zeta(3)-\zeta(2)+3\label{s6}
\end{eqnarray}
\end{series2-6}
\addtocounter{series}{5}
The generalization of this type of series is well-known, and is found in
(5.1.24.8) of \cite{Prudnikov1}. These results
will be frequently used throughout the proofs.
\section{One-dimensional series}
\label{s:one-d}
We now turn our attention to some one-dimensional series which bear similarity
to the Euler series as well as to those studied in Paper I.
\begin{series7-15}
\begin{eqnarray}
&&\sum_{n=1}^\infty\frac{1}{n(n+1)}[\gamma+\psi(n)]=1\label{s8}\\
&&\sum_{n=1}^\infty\frac{1}{n(n+1)}[\gamma+\psi(1+n)]=\zeta(2)\label{s7}\\
&&\sum_{n=1}^\infty\frac{1}{n(n+1)}[\gamma+\psi(2+n)]=2\label{s9}\\
&&\sum_{n=1}^\infty\frac{1}{n^2(n+1)}[\gamma+\psi(n)]=\zeta(3)-1\label{s14}\\
&&\sum_{n=1}^\infty\frac{1}{n^2(n+1)}[\gamma+\psi(1+n)]=2\zeta(3)-\zeta(2)
\label{s13}\\
&&\sum_{n=1}^\infty\frac{1}{n^2(n+1)}[\gamma+\psi(2+n)]=2\zeta(3)+\zeta(2)-3
\label{s15} \\
&&\sum_{n=1}^\infty\frac{1}{n(n+1)^2}[\gamma+\psi(n)]=-\zeta(3)-\zeta(2)+3
\label{s11}\\
&&\sum_{n=1}^\infty\frac{1}{n(n+1)^2}[\gamma+\psi(1+n)]=-\zeta(3)+\zeta(2)
\label{s10}\\
&&\sum_{n=1}^\infty\frac{1}{n(n+1)^2}[\gamma+\psi(2+n)]=-2\zeta(3)+3
\label{s12}
\end{eqnarray}
\end{series7-15}
\addtocounter{series}{9}
We prove Series \ref{s7}.
The others follow as corollaries of this result by using the recurrence
relation
(\ref{psirecurrence}), partial fractioning, Series \ref{s2}--\ref{s6},
(I.B.1) and (I.B.2).
\begin{pf*}{Proof of Series \ref{s7}.}
We start by using the integral representation (\ref{psiintegral}) of the
psi function before summing over $n$:
\begin{eqnarray}
&&\sum_{n=1}^\infty\frac{1}{n(n+1)}[\gamma+\psi(1+n)]=
\sum_{n=1}^\infty\frac{1}{n(n+1)}\int_0^1{\rm
d}t\frac{1-t^n}{1-t}\nonumber\\
&&=\int_0^1{\rm d}t\frac{1}{1-t}\left[1-\frac{t}{2}\ {}_2F_1(1,1;3;t)\right]
\nonumber\\
&&=\int_0^1{\rm d}t\frac{1}{1-t}
\left\{1-\frac{1}{t}\left[t+(1-t)\log(1-t)\right]\right\}
=-\int_0^1{\rm d}t\frac{\log(1-t)}{t}=\zeta(2)
\nonumber
\end{eqnarray}
We have used (7.3.2.150) of \cite{Prudnikov} to rewrite ${}_2F_1$. In
the last step we used (3.6.1) of \cite{D&D}.
\qed\end{pf*}
Similar relations can also be found involving the trigamma function (see
Appendix A.2):
\begin{series16-21}
\begin{eqnarray}
&&\sum_{n=1}^\infty\frac{1}{n}\psi^\prime(n)=2\zeta(3)\label{s18}\\
&&\sum_{n=1}^\infty\frac{1}{n}\psi^\prime(1+n)=\zeta(3)\label{s17}\\
&&\sum_{n=1}^\infty\frac{1}{n}\psi^\prime(2+n)=\zeta(3)+\zeta(2)-2\label{s19}\\
&&\sum_{n=1}^\infty\frac{1}{n(n+1)}\psi^\prime(n)=1\label{s21}\\
&&\sum_{n=1}^\infty\frac{1}{n(n+1)}\psi^\prime(1+n)=-\zeta(3)+\zeta(2)
\label{s20}\\
&&\sum_{n=1}^\infty\frac{1}{n(n+1)}\psi^\prime(2+n)=2\zeta(2)-3\label{s22}
\end{eqnarray}
\end{series16-21}
\addtocounter{series}{6}
We prove Series \ref{s17}. The others follow by using the recurrence relation
(\ref{trigammarecurrence}) and partial fractioning, together with
Series \ref{s1}--\ref{s6}.
\begin{pf*}{Proof of Series \ref{s17}.}
We start by using the integral representation
(\ref{trigammaintegral}) of the trigamma function:
\begin{eqnarray}
\sum_{n=1}^\infty\frac{1}{n}\psi^\prime(1+n)
&=&-\sum_{n=1}^\infty\frac{1}{n}\int_0^1{\rm d}t\frac{t^n}{1-t}\log t
=\int_0^1\frac{{\rm d}t}{1-t}\log t\log(1-t)\nonumber\\
&=&\int_0^1\frac{{\rm d}t}{t}\log t\log(1-t)=\zeta(3)
\nonumber
\end{eqnarray}
In the last step we have used (3.6.21) of \cite{D&D}.
\qed\end{pf*}
Next, we consider series which are quadratic or bilinear in psi functions.
\begin{series22-27}
\begin{eqnarray}
&&\sum_{n=1}^\infty\frac{1}{n(n+1)}[\gamma+\psi(n)]^2=\zeta(2)+1\label{s24}\\
&&\sum_{n=1}^\infty\frac{1}{n(n+1)}[\gamma+\psi(n)][\gamma+\psi(1+n)]
=\zeta(3)+\zeta(2)\label{s25}\\
&&\sum_{n=1}^\infty\frac{1}{n(n+1)}[\gamma+\psi(n)][\gamma+\psi(2+n)]
=3\label{s1new}\\
&&\sum_{n=1}^\infty\frac{1}{n(n+1)}[\gamma+\psi(1+n)]^2=3\zeta(3)\label{s23}\\
&&\sum_{n=1}^\infty\frac{1}{n(n+1)}[\gamma+\psi(1+n)][\gamma+\psi(2+n)]
=2\zeta(3)+\zeta(2)\label{s2new}\\
&&\sum_{n=1}^\infty\frac{1}{n(n+1)}[\gamma+\psi(2+n)]^2
=\zeta(2)+3\label{s3new}
\end{eqnarray}
\end{series22-27}
\addtocounter{series}{6}
We prove Series \ref{s23}. The others are immediate corollaries
that follow from using
(\ref{psirecurrence}) along with some of the results derived earlier in this
chapter.
\begin{pf*}{Proof of Series \ref{s23}.}
We start by using the integral representation (\ref{psiintegral}) of the
psi function:
\begin{eqnarray}
&&\sum_{n=1}^\infty\frac{1}{n(n+1)}[\gamma+\psi(1+n)]^2
=\sum_{n=1}^\infty\frac{1}{n(n+1)}
\int_0^1{\rm d}t\frac{1-t^n}{1-t}\int_0^1{\rm d}s\frac{1-s^n}{1-s}
\nonumber\\
&&=\int_0^1\frac{{\rm d}t}{1-t}\int_0^1\frac{{\rm d}s}{1-s}
\left[1-\frac{t}{2}\ {}_2F_1(1,1;3;t)-\frac{s}{2}
\ {}_2F_1(1,1;3;s)\right.\nonumber\\
&&\left.\phantom{=\int_0^1\frac{{\rm d}t}{1-t}\int_0^1\frac{{\rm d}s}{1-s}[}
+\frac{st}{2}\ {}_2F_1(1,1;3;st)\right]\nonumber\\
&&=\int_0^1\frac{{\rm d}t}{1-t}\int_0^1\frac{{\rm d}s}{1-s}
\left\{1-\frac{1}{t}\left[t+(1-t)\log(1-t)\right]\right.\nonumber\\
&&\phantom{=\int_0^1\frac{{\rm d}t}{1-t}\int_0^1\frac{{\rm d}s}{1-s}\{}
-\frac{1}{s}\left[s+(1-s)\log(1-s)\right]\nonumber\\
&&\left.\phantom{=\int_0^1\frac{{\rm d}t}{1-t}\int_0^1\frac{{\rm d}s}{1-s}\{}
+\frac{1}{st}\left[st+(1-st)\log(1-st)\right]\right\}\nonumber
\end{eqnarray}
Here, we have used (7.3.2.150) of \cite{Prudnikov}. We continue to simplify
this expression:
\begin{eqnarray}
&&\int_0^1\frac{{\rm d}t}{1-t}\int_0^1\frac{{\rm d}s}{1-s}
\left[(1-st)\frac{\log(1-st)}{st}\right.\nonumber\\
&&\left.\phantom{\int_0^1\frac{{\rm d}t}{1-t}\int_0^1\frac{{\rm d}s}{1-s}[}
-(1-t)\frac{\log(1-t)}{t}-(1-s)\frac{\log(1-s)}{s}\right]\nonumber\\
&&=\int_0^1\frac{{\rm d}t}{1-t}\left\{
\int_0^1\frac{{\rm d}s}{1-s}
\left[\frac{\log(1-st)}{st}-\frac{\log(1-t)}{t}\right]\right.\nonumber\\
&&\left.\phantom{=\int_0^1\frac{{\rm d}t}{1-t}\{}
+\int_0^1\frac{{\rm d}s}{1-s}\left[\log(1-t)-\log(1-st)\right]
-\int_0^1{\rm d}s\frac{\log(1-s)}{s}\right\}\nonumber\\
&&=\int_0^1\frac{{\rm d}t}{1-t}\left\{
\frac{1}{t}\int_0^1\frac{{\rm d}s}{1-s}\left[\log(1-st)-\log(1-t)\right]
+\frac{1}{t}\int_0^1\frac{{\rm d}s}{s}\log(1-st)
\right.\nonumber\\
&&\left.\phantom{=\int_0^1\frac{{\rm d}t}{1-t}\{}
-\int_0^1\frac{{\rm d}s}{1-s}\left[\log(1-st)-\log(1-t)\right]
+\zeta(2)\right\}\nonumber
\end{eqnarray}
We combine the first and the third of the integrals inside the curly
brackets, whereas the second one is evaluated to give
\begin{eqnarray}
\int_0^1\frac{{\rm d}t}{1-t}\left\{
\frac{1-t}{t}\int_0^1\frac{{\rm d}s}{1-s}\left[\log(1-st)-\log(1-t)\right]
-\frac{1}{t}\mbox{Li}_2(t)+\zeta(2)\right\}.\nonumber
\end{eqnarray}
Next, a change of variables yields:
\begin{eqnarray}
&&\int_0^1\frac{{\rm d}t}{1-t}\left\{
\frac{1-t}{t}\int_0^1\frac{{\rm d}s}{s}\left[\log(1-t+st)-\log(1-t)\right]
-\frac{1}{t}\mbox{Li}_2(t)+\zeta(2)\right\}\nonumber
\end{eqnarray}
We use (3.14.1) and thereafter (2.2.5) of \cite{D&D}:
\begin{eqnarray}
&&\int_0^1\frac{{\rm d}t}{1-t}\left[
-\frac{1-t}{t}\mbox{Li}_2\left(\frac{-t}{1-t}\right)
-\frac{1}{t}\mbox{Li}_2(t)+\zeta(2)\right]\nonumber\\
&&=\int_0^1\frac{{\rm d}t}{1-t}\left\{
\frac{1-t}{t}\left[\mbox{Li}_2(t)+\frac{1}{2}\log^2(1-t)\right]
-\frac{1}{t}\mbox{Li}_2(t)+\zeta(2)\right\}
\nonumber\\
&&=\int_0^1\frac{{\rm d}t}{1-t}\left[\zeta(2)-\mbox{Li}_2(t)\right]
+\frac{1}{2}\int_0^1\frac{{\rm d}t}{t}\log^2(1-t)
=2\zeta(3)+\zeta(3)=3\zeta(3)\nonumber
\end{eqnarray}
In the last step we have used (3.8.9) and (3.6.9) of \cite{D&D}.
\qed\end{pf*}
\section{Two-dimensional series}
\label{s:two-d}
Next, we present results for the sums of several two-dimensional
series. Many of these are proved by using results from the one-dimensional
series of Section 2 and from Paper I.
\begin{series}
\begin{eqnarray}
\sum_{n=1}^\infty\sum_{k=1}^\infty\frac{1}{nk(n+k)}=2\zeta(3)
\end{eqnarray}
\end{series}
\begin{pf*}{Proof.}
\begin{eqnarray}
\sum_{n=1}^\infty\sum_{k=1}^\infty\frac{1}{nk(n+k)}
=\sum_{k=1}^\infty\frac{1}{k^2}[\gamma+\psi(1+k)]=2\zeta(3)\nonumber
\end{eqnarray}
Here, we have used (I.10) and thereafter (I.B.2).
\qed\end{pf*}
\begin{series29-37}
\begin{eqnarray}
&&\sum_{n=0}^\infty\sum_{k=1}^\infty\frac{1}{k(n+k)(1+n+k)}=\zeta(2)
\label{s27}\\
&&\sum_{n=0}^\infty\sum_{k=1}^\infty\frac{1}{k(n+k)^2(1+n+k)}=
2\zeta(3)-\zeta(2)\label{s30}\\
&&\sum_{n=0}^\infty\sum_{k=1}^\infty\frac{1}{k(n+k)(1+n+k)}[\gamma+\psi(k)]
=\zeta(3)\label{s28}\\
&&\sum_{n=0}^\infty\sum_{k=1}^\infty\frac{1}{k(n+k)(1+n+k)}[\gamma+\psi(1+k)]
=2\zeta(3)\label{s29}\\
&&\sum_{n=0}^\infty\sum_{k=1}^\infty\frac{1}{k(n+k)(1+n+k)}[\gamma+\psi(1+n)]
=2\zeta(3)\label{s31}\\
&&\sum_{n=0}^\infty\sum_{k=1}^\infty\frac{1}{k(n+k)(1+n+k)}[\gamma+\psi(n+k)]
=\zeta(3)+\zeta(2)\label{s34}\\
&&\sum_{n=0}^\infty\sum_{k=1}^\infty\frac{1}{k(n+k)(1+n+k)}[\gamma+\psi(1+n+k)]
=3\zeta(3)\label{s33}\\
&&\sum_{n=0}^\infty\sum_{k=1}^\infty\frac{1}{k(n+k)(1+n+k)}[\gamma+\psi(2+n+k)]
=2\zeta(3)+\zeta(2)\label{s32}\\
&&\sum_{n=0}^\infty\sum_{k=1}^\infty\frac{1}{k(n+k)(1+n+k)}
[\gamma+\psi(1+n+2k)]=\tfrac{7}{2}\zeta(3)\label{s35}
\end{eqnarray}
\end{series29-37}
\addtocounter{series}{9}
We need to prove most of these results in different ways. Series \ref{s34}
follows as an immediate corollary of Series \ref{s30} and \ref{s33} after
using (\ref{psirecurrence}).
\begin{pf*}{Proof of Series \ref{s27}, \ref{s28} and \ref{s29}.}
\begin{eqnarray}
&&\sum_{n=0}^\infty\sum_{k=1}^\infty\frac{1}{k(n+k)(1+n+k)}f(k)\nonumber\\
&&=\sum_{k=1}^\infty\frac{1}{k^2(k+1)}\ {}_2F_1(1,k;2+k;1)f(k)
=\sum_{k=1}^\infty\frac{1}{k^2}f(k)\nonumber
\end{eqnarray}
Here, we have used (\ref{2F1unity}) to rewrite ${}_2F_1$.
By replacing $f(k)$ with the
appropriate expression and using the definition of $\zeta(2)$, (I.B.1) or
(I.B.2) the proof is complete.
\qed\end{pf*}
\begin{pf*}{Proof of Series \ref{s30}.}
\begin{eqnarray}
&&\sum_{n=0}^\infty\sum_{k=1}^\infty\frac{1}{k(n+k)^2(1+n+k)}\nonumber\\
&&=\sum_{n=0}^\infty\sum_{k=1}^\infty\frac{1}{k(n+k)^2}
-\sum_{n=0}^\infty\sum_{k=1}^\infty\frac{1}{k(n+k)(1+n+k)}\nonumber\\
&&=\sum_{k=1}^\infty\frac{1}{k}\psi^\prime(k)-\zeta(2)
=2\zeta(3)-\zeta(2)\nonumber
\end{eqnarray}
In the last steps we used (\ref{trigammaidentity}) and Series \ref{s18}.
\qed\end{pf*}
\begin{pf*}{Proof of Series \ref{s31}.}
\begin{eqnarray}
&&\sum_{n=0}^\infty\sum_{k=1}^\infty\frac{1}{k(n+k)(1+n+k)}[\gamma+\psi(1+n)]
\nonumber\\
&&=\sum_{n=0}^\infty\frac{1}{(n+1)(n+2)}\ {}_3F_2(1,1,1+n;2,3+n;1)
[\gamma+\psi(1+n)]
\nonumber\\
&&=\sum_{n=1}^\infty\frac{1}{(n+1)(n+2)}\ {}_3F_2(1,1,1+n;2,3+n;1)
[\gamma+\psi(1+n)]
\nonumber
\end{eqnarray}
Here, we have used the fact that the summand vanishes for $n=0$. Next, we
use (7.4.4.40) of \cite {Prudnikov}. Thereafter we use (6.3.2) of
\cite{Abramowitz} and the recurrence relation for the
psi function:
\begin{eqnarray}
&&\sum_{n=1}^\infty\frac{1}{n(n+1)}[\psi(2+n)-\psi(2)][\gamma+\psi(1+n)]
\nonumber\\
&&=\sum_{n=1}^\infty\frac{1}{n(n+1)}
\left[\gamma+\psi(1+n)+\frac{1}{1+n}-1\right][\gamma+\psi(1+n)]\nonumber\\
&&=\sum_{n=1}^\infty\frac{1}{n(n+1)}[\gamma+\psi(1+n)]^2
+\sum_{n=1}^\infty\frac{1}{n(n+1)^2}[\gamma+\psi(1+n)]\nonumber\\
&&\phantom{=}
-\sum_{n=1}^\infty\frac{1}{n(n+1)}[\gamma+\psi(1+n)]=2\zeta(3)\nonumber
\end{eqnarray}
In the last step we use the results of Series \ref{s7}, \ref{s11} and
\ref{s23}.
\qed\end{pf*}
\begin{pf*}{Proof of Series \ref{s33}.}
We start by using the recurrence relation for the psi
function.
\begin{eqnarray}
&&\sum_{n=0}^\infty\sum_{k=1}^\infty\frac{1}{k(n+k)(1+n+k)}[\gamma+\psi(1+n+k)]
\nonumber\\
&&=\sum_{n=0}^\infty\sum_{k=1}^\infty\frac{1}{k(n+k)(1+n+k)}
\left[\gamma+\psi(2+n+k)-\frac{1}{1+n+k}\right]\nonumber\\
&&=\sum_{n=0}^\infty\sum_{k=1}^\infty
\frac{1}{k(n+k)(1+n+k)}[\gamma+\psi(2+n+k)]\nonumber\\
&&\phantom{=}
-\sum_{n=0}^\infty\sum_{k=1}^\infty\frac{1}{k(n+k)(1+n+k)^2}\nonumber\\
&&=2\zeta(3)+\zeta(2)
-\sum_{n=0}^\infty\sum_{k=1}^\infty\frac{1}{k(n+k)(1+n+k)}
+\sum_{n=0}^\infty\sum_{k=1}^\infty\frac{1}{k(1+n+k)^2}\nonumber\\
&&=2\zeta(3)+\zeta(2)-\zeta(2)
+\sum_{k=1}^\infty\frac{1}{k}\psi^\prime(1+k)
=2\zeta(3)+\zeta(3)=3\zeta(3)\nonumber
\end{eqnarray}
Here, we have used (\ref{trigammaidentity}), Series \ref{s17} and \ref{s27}.
\qed\end{pf*}
\begin{pf*}{Proof of Series \ref{s32}.}
\begin{eqnarray}
&&\sum_{n=0}^\infty\sum_{k=1}^\infty\frac{1}{k(n+k)(1+n+k)}[\gamma+\psi(2+n+k)]
\nonumber\\
&&=\gamma\sum_{n=0}^\infty\sum_{k=1}^\infty\frac{1}{k(n+k)(1+n+k)}
+\left.\frac{{\rm d}}{{\rm d}x}\right|_{x=0}
\sum_{n=0}^\infty\sum_{k=1}^\infty\frac{\Gamma(n+k)}{k\Gamma(2+n+k-x)}
\nonumber\\
&&=\gamma\sum_{k=1}^\infty\frac{1}{k^2(k+1)}\ {}_2F_1(1,k;2+k;1)\nonumber\\
&&\phantom{=}
+\left.\frac{{\rm d}}{{\rm d}x}\right|_{x=0}
\sum_{k=1}^\infty\frac{\Gamma(k)}{k\Gamma(2+k-x)}\ {}_2F_1(1,k;2+k-x;1)
\nonumber\\
&&=\gamma\sum_{k=1}^\infty\frac{1}{k^2}
+\left.\frac{{\rm d}}{{\rm d}x}\right|_{x=0}
\sum_{k=1}^\infty\frac{\Gamma(k)}{k(1-x)\Gamma(1+k-x)}
\nonumber\\
&&=\gamma\sum_{k=1}^\infty\frac{1}{k^2}
+\sum_{k=1}^\infty\frac{1}{k^2}[1+\psi(1+k)]\nonumber\\
&&=\sum_{k=1}^\infty\frac{1}{k^2}[\gamma+\psi(1+k)]
+\sum_{k=1}^\infty\frac{1}{k^2}
=2\zeta(3)+\zeta(2)\nonumber
\end{eqnarray}
Here, we have used (\ref{2F1unity}) to rewrite ${}_2F_1$. In the last step
we have also used (I.B.2).
\qed\end{pf*}
\begin{pf*}{Proof of Series \ref{s35}.}
\begin{eqnarray}
&&\sum_{n=0}^\infty\sum_{k=1}^\infty
\frac{1}{k(n+k)(1+n+k)}[\gamma+\psi(1+n+2k)]\nonumber\\
&&=\sum_{n=0}^\infty\sum_{k=1}^\infty\sum_{j=1}^{n+2k}\frac{1}{jk(n+k)(1+n+k)}
\nonumber\\
&&=\sum_{n=0}^\infty\sum_{k=1}^\infty\sum_{j=1}^{2k}\frac{1}{jk(n+k)(1+n+k)}
+\sum_{n=1}^\infty\sum_{k=1}^\infty\sum_{j=1+2k}^{n+2k}\frac{1}{jk(n+k)(1+n+k)}
\nonumber\\
&&=\sum_{n=0}^\infty\sum_{k=1}^\infty\frac{1}{k(n+k)(1+n+k)}[\gamma+\psi(1+2k)]
\nonumber\\
&&\phantom{=}
+\sum_{n=1}^\infty\sum_{k=1}^\infty\sum_{j=1}^{n}\frac{1}{k(n+k)(1+n+k)(j+2k)}
\nonumber\\
&&=\sum_{k=1}^\infty\frac{1}{k^2(k+1)}\ {}_2F_1(1,k;2+k;1)[\gamma+\psi(1+2k)]
\nonumber\\
&&\phantom{=}
+\sum_{n=0}^\infty\sum_{k=1}^\infty\sum_{j=1}^\infty
\frac{1}{k(n+j+k)(1+n+j+k)(j+2k)}\nonumber\\
&&=\sum_{k=1}^\infty\frac{1}{k^2}[\gamma+\psi(1+2k)]\nonumber\\
&&\phantom{=}
+\sum_{k=1}^\infty\sum_{j=1}^\infty\frac{1}{k(j+k)(1+j+k)(j+2k)}
\ {}_2F_1(1,j+k;2+j+k;1)\nonumber\\
&&=\frac{11}{4}\zeta(3)
+\sum_{k=1}^\infty\sum_{j=1}^\infty\frac{1}{k(j+k)(j+2k)}
=\frac{7}{2}\zeta(3)\nonumber
\end{eqnarray}
Here, we have used (\ref{2F1unity}) to rewrite ${}_2F_1$. We also used (I.13)
and (I.B.4).
\qed\end{pf*}
\begin{series38-47}
\begin{eqnarray}
&&\sum_{n=1}^\infty\sum_{k=1}^\infty\frac{1}{n(k+1)(n+k)}=\zeta(2)
\label{s36}\\
&&\sum_{n=1}^\infty\sum_{k=1}^\infty\frac{1}{n(k+1)(n+k)^2}=
3\zeta(3)-2\zeta(2)\label{s40}\\
&&\sum_{n=1}^\infty\sum_{k=1}^\infty\frac{1}{n(k+1)(n+k)}[\gamma+\psi(k)]
=\zeta(3)+\zeta(2)\label{s37}\\
&&\sum_{n=1}^\infty\sum_{k=1}^\infty\frac{1}{n(k+1)(n+k)}[\gamma+\psi(1+k)]
=3\zeta(3)\label{s38}\\
&&\sum_{n=1}^\infty\sum_{k=1}^\infty\frac{1}{n(k+1)(n+k)}[\gamma+\psi(2+k)]
=2\zeta(3)+\zeta(2)\label{s39}\\
&&\sum_{n=1}^\infty\sum_{k=1}^\infty\frac{1}{n(k+1)(n+k)}[\gamma+\psi(n)]
=2\zeta(3)\label{s41}\\
&&\sum_{n=1}^\infty\sum_{k=1}^\infty\frac{1}{n(k+1)(n+k)}[\gamma+\psi(1+n)]
=2\zeta(2)\label{s42}\\
&&\sum_{n=1}^\infty\sum_{k=1}^\infty\frac{1}{n(k+1)(n+k)}[\gamma+\psi(2+n)]
=2\zeta(3)+\tfrac{1}{2}\label{s43}\\
&&\sum_{n=1}^\infty\sum_{k=1}^\infty\frac{1}{n(k+1)(n+k)}[\gamma+\psi(n+k)]
=\zeta(3)+2\zeta(2)\label{s45}\\
&&\sum_{n=1}^\infty\sum_{k=1}^\infty\frac{1}{n(k+1)(n+k)}[\gamma+\psi(1+n+k)]
=4\zeta(3)\label{s44}
\end{eqnarray}
\end{series38-47}
\addtocounter{series}{10}
We use different proofs for most of these series. Series \ref{s45} follows
as an immediate corollary of Series \ref{s40} and \ref{s44} after using
(\ref{psirecurrence}).
\begin{pf*}{Proof of Series \ref{s36}, \ref{s37}, \ref{s38} and \ref{s39}.}
\begin{eqnarray}
\sum_{n=1}^\infty\sum_{k=1}^\infty\frac{1}{n(k+1)(n+k)}f(k)
&=&\sum_{k=1}^\infty\frac{1}{k(k+1)}[\gamma+\psi(1+k)]f(k) \nonumber
\end{eqnarray}
Here, we have used (I.10). By replacing $f(k)$ by the appropriate expression
and using Series \ref{s7}, \ref{s25}, \ref{s23} or \ref{s2new} in connection
with (\ref{psirecurrence}), the proof is complete.
\qed\end{pf*}
\begin{pf*}{Proof of Series \ref{s40}.}
\begin{eqnarray}
&&\sum_{n=1}^\infty\sum_{k=1}^\infty\frac{1}{n(k+1)(n+k)^2}\nonumber\\
&&=\sum_{k=1}^\infty\frac{1}{(k+1)^3}{}_4F_3(1,1,1+k,1+k;2,2+k,2+k;1)
\nonumber
\end{eqnarray}
We rewrite this using (7.5.3.4) of \cite{Prudnikov}:
\begin{eqnarray}
&&\sum_{k=1}^\infty\frac{1}{(k+1)^3}\left\{
\frac{(1+k)^2}{k^2}[\gamma+\psi(1+k)]-\frac{(1+k)^2}{k}\psi^\prime(1+k)\right\}
\nonumber\\
&=&\sum_{k=1}^\infty\frac{1}{k^2(k+1)}[\gamma+\psi(1+k)]
-\sum_{k=1}^\infty\frac{1}{k(k+1)}\psi^\prime(1+k)
=3\zeta(3)-2\zeta(2)\nonumber
\end{eqnarray}
Here, we have made use of Series \ref{s13} and \ref{s20}.
\qed\end{pf*}
\begin{pf*}{Proof of Series \ref{s41}, \ref{s42} and \ref{s43}.}
\begin{eqnarray}
&&\sum_{n=1}^\infty\sum_{k=1}^\infty\frac{1}{n(k+1)(n+k)}f(n)\nonumber\\
&&=\sum_{k=1}^\infty\frac{1}{(k+1)^2}f(1)
+\sum_{n=2}^\infty\sum_{k=1}^\infty\frac{1}{n(k+1)(n+k)}f(n)
\nonumber\\
&&=\left[\zeta(2)-1\right]f(1)
+\sum_{n=2}^\infty\frac{1}{n(n-1)}[\psi(1+n)-\psi(2)]f(n)\nonumber\\
&&=\left[\zeta(2)-1\right]f(1)
+\sum_{n=1}^\infty\frac{1}{n(n+1)}[\psi(2+n)-\psi(2)]f(n+1)\nonumber
\nonumber\\
&&=\left[\zeta(2)-1\right]f(1)
+\sum_{n=1}^\infty\frac{1}{n(n+1)}\left[\gamma+\psi(2+n)\right]f(n+1)
\nonumber\\
&&\phantom{=}-\sum_{n=1}^\infty\frac{1}{n(n+1)}f(n+1)
\nonumber
\end{eqnarray}
Here, we first used (I.10). Thereafter we used (6.3.2) of \cite{Abramowitz}.
By replacing $f(n)$ with the appropriate expression and using results
derived earlier, the proof is complete.
\qed\end{pf*}
\begin{pf*}{Proof of Series \ref{s44}.}
\begin{eqnarray}
&&\sum_{n=1}^\infty\sum_{k=1}^\infty\frac{1}{n(k+1)(n+k)}[\gamma+\psi(1+n+k)]
\nonumber\\
&&=\gamma\sum_{n=1}^\infty\sum_{k=1}^\infty\frac{1}{n(k+1)(n+k)}+
\left.\frac{{\rm d}}{{\rm d}x}\right|_{x=0}
\sum_{n=1}^\infty\sum_{k=1}^\infty\frac{\Gamma(n+k)}{n(k+1)\Gamma(1+n+k-x)}
\nonumber\\
&&=\gamma\sum_{k=1}^\infty\frac{1}{k(k+1)}[\gamma+\psi(1+k)]\nonumber\\
&&+\left.\frac{{\rm d}}{{\rm d}x}\right|_{x=0}
\sum_{k=1}^\infty
\frac{\Gamma(1+k)}{(k+1)\Gamma(2+k-x)}\ {}_3F_2(1,1,1+k;2,2+k-x;1)
\nonumber
\end{eqnarray}
We have made use of (I.10). Next, we make use of (7.4.4.40) of
\cite{Prudnikov} , and obtain
\begin{eqnarray}
&&\gamma\sum_{k=1}^\infty\frac{1}{k(k+1)}[\gamma+\psi(1+k)]\nonumber\\
&&\phantom{=}
+\left.\frac{{\rm d}}{{\rm d}x}\right|_{x=0}
\sum_{k=1}^\infty\frac{\Gamma(k)}{(k+1)\Gamma(1+k-x)}[\psi(1+k-x)-\psi(1-x)]
\nonumber\\
&&=\gamma\sum_{k=1}^\infty\frac{1}{k(k+1)}[\gamma+\psi(1+k)]\nonumber\\
&&\phantom{=}
+\sum_{k=1}^\infty\frac{1}{k(k+1)}
\left\{\psi(1+k)[\gamma+\psi(1+k)]+\zeta(2)-\psi^\prime(1+k)\right\}
\nonumber\\
&&=\sum_{k=1}^\infty\frac{1}{k(k+1)}[\gamma+\psi(1+k)]^2\nonumber\\
&&\phantom{=}
+\zeta(2)\sum_{k=1}^\infty\frac{1}{k(k+1)}
-\sum_{k=1}^\infty\frac{1}{k(k+1)}\psi^\prime(1+k)=4\zeta(3).
\nonumber
\end{eqnarray}
In the final steps we made use of Series \ref{s1}, \ref{s20}
and \ref{s23}.
\qed\end{pf*}
\begin{series48-52}
\begin{eqnarray}
&&\sum_{n=1}^\infty\sum_{k=0}^\infty\frac{1}{n(k+1)(n+k)}=2\zeta(2)\\
&&\sum_{n=1}^\infty\sum_{k=0}^\infty\frac{1}{n(k+1)(n+k)}[\gamma+\psi(1+k)]
=3\zeta(3)\\
&&\sum_{n=1}^\infty\sum_{k=0}^\infty\frac{1}{n(k+1)(n+k)}[\gamma+\psi(n)]
=3\zeta(3)\\
&&\sum_{n=1}^\infty\sum_{k=0}^\infty\frac{1}{n(k+1)(n+k)}[\gamma+\psi(1+n+k)]
=6\zeta(3)\\
&&\sum_{n=1}^\infty\sum_{k=0}^\infty\frac{1}{n(k+1)(n+k)}[\gamma+\psi(n+k)]
=2\zeta(3)+2\zeta(2)
\end{eqnarray}
\end{series48-52}
\addtocounter{series}{5}
These results all follow immediately from Series \ref{s36}, \ref{s38},
\ref{s41}, \ref{s45}, \ref{s44}, (I.B.1) and (I.B.2).
\begin{series}
\begin{eqnarray}
\sum_{n=1}^\infty\sum_{k=0}^\infty\frac{1}{n(n+1)(1+n+k)^2}
=-\zeta(3)+\zeta(2)
\end{eqnarray}
\end{series}
\begin{pf*}{Proof.}
\begin{eqnarray}
\sum_{n=1}^\infty\sum_{k=0}^\infty\frac{1}{n(n+1)(1+n+k)^2}
&=&\sum_{n=1}^\infty\frac{1}{n(n+1)}\psi^\prime(1+n)=-\zeta(3)+\zeta(2)
\nonumber
\end{eqnarray}
We have used (\ref{trigammaidentity}) and Series \ref{s20} in the last steps.
\qed\end{pf*}
\begin{series}
\begin{eqnarray}
\sum_{n=1}^\infty\sum_{k=1}^\infty\frac{1}{k}
\frac{\Gamma(n)\Gamma(k)}{\Gamma(1+n+k)}
=\zeta(3)
\end{eqnarray}
\end{series}
\begin{pf*}{Proof.}
\begin{eqnarray}
&&\sum_{n=1}^\infty\sum_{k=1}^\infty\frac{1}{k}
\frac{\Gamma(n)\Gamma(k)}{\Gamma(1+n+k)}
=\sum_{k=1}^\infty\frac{1}{k^2(k+1)}\ {}_2F_1(1,1;2+k;1)\nonumber\\
&&=\sum_{k=1}^\infty\frac{1}{k^3}=\zeta(3)\nonumber
\end{eqnarray}
Here, we have used (\ref{2F1unity}) to rewrite ${}_2F_1$.
\qed\end{pf*}
\section{A three-dimensional series}
\label{s:three-d}
\begin{series}
\begin{eqnarray}
\sum_{l=0}^\infty\sum_{n=1}^\infty\sum_{k=1}^\infty\frac{1}{nk(1+l+k)(l+n+k)}
=3\zeta(3)
\end{eqnarray}
\end{series}
\begin{pf*}{Proof.}
\begin{eqnarray}
&&\sum_{n=0}^\infty\sum_{k=1}^\infty\sum_{l=1}^\infty\frac{1}{lk(1+n+k)(n+k+l)}
\nonumber\\
&&=\sum_{n=0}^\infty\sum_{k=1}^\infty\frac{1}{k(n+k)(1+n+k)}
[\gamma+\psi(1+n+k)]=3\zeta(3)\nonumber
\end{eqnarray}
Here, we have used (I.10) and Series~\ref{s33}.
\qed\end{pf*}
|
\section{Introduction}
The interplay between strongly correlated electron systems and disorder is
still an open question. The most efficient method is to start from the weak
disorder in the electronic system and to consider correlations using the
renormalization group (RG) method \cite{1,2,3}. The main result of this
method is prediction of the metal-insulator transition when the symmetry was
broken by the interaction with impurities.
The problem was studied using infinite-U Hubbard model and the t-J model
\cite{4,5,6}. The infinite dimension approach introduced by Metzner and
Vollhardt \cite{7} has been applied by many authors \cite{8,9,10} (for a
complete discussion see Ref. \onlinecite{10}) and using RG method Si and
Kotliar \cite{11} showed that in an extended Hubbard model the disorder can
induce a non-Fermi behavior.
In this paper we will show that the weak disorder in the finite-charge
infinite-U Hubbard model can induced a non-Fermi behavior for a
two-dimensional (2D) electronic system close to the metal-insulator
transition. The paper is organized as follows. In Sect. II we present
the model. The self-energy of the electronic system is calculated in
Sect. III and we show that linear temperature dependence may appear,
which show a typical non-Fermi behavior. The relevance of our results
for the explanation of the experimental results will be discussed in
Sect. IV.
\section{Model}
We consider an electronic system with weak disorder in the finite-charge
infinite-U Hubbard model. The charge susceptibility has been calculated
in \cite{6} as
\begin{equation}
\chi_c(\bq,\omega} \def\G{\Gamma} \def\D{\Delta)=\frac{N(0) N D q^2}
{2\left[Dq^2A_0+D\a q^4/k_F^4-i\omega} \def\G{\Gamma} \def\D{\Delta\right]}
\label{e1}
\end{equation}
where $N(0)$ is the density of state, N is the orbital degeneracy,
$D=v_F^2\tau/2$ is the diffusion coefficient, $A_0=1-2t_0 N(0)$ (with
$t_0$ the base kinetic energy), $\a=3(Q/N)^2(m^8/m)^2/4$ with Q the
total charge and $k_F$ is the Fermi wave vector. For a filling $n_f$
($m/m^*=1-n_f$) close to metal-insulator phase transition $q^4$ term,
given by the quasiparticles interaction, is important and we will show
that is essential in the behavior of the electronic system.
The effect of disorder will be considered as contained in the enhancement
of the charge susceptibility and in order to analyze the effect of it on
the energy of the electronic excitations we take the general form for the
self-energy in one-loop approximation.
\section{Self-energy}
The self-energy of the electrons due to the interaction of electrons with
the charge fluctuations in the presence of disorder has the general form:
\begin{equation}
\S (\bp,\omega} \def\G{\Gamma} \def\D{\Delta)=g^2\int \frac{d^2 q}{(2\pi)^2}\int_{-\infty}^\infty \frac{d\omega} \def\G{\Gamma} \def\D{\Delta'}{2\pi}
\left[\coth{\frac{\omega} \def\G{\Gamma} \def\D{\Delta'}{2T}}-\tanh{\frac{\tilde{\ve}(\bp+\bq)}{2T}}\right]
\frac{Im\chi_c(\bq,\omega} \def\G{\Gamma} \def\D{\Delta')}{\omega} \def\G{\Gamma} \def\D{\Delta+\omega} \def\G{\Gamma} \def\D{\Delta'-\tilde{\ve}(\bp+\bq)+i\a}
\label{e2}
\end{equation}
where $\tilde{\ve}({\bf k}} \def\bq{{\bf q}} \def\bp{{\bf p}} \def\bQ{{\bf Q})=k^2/2m-\mu$. An analytical calculation can be
performed for the two dimensional case. Using the approximation
\begin{equation}
\tilde{\ve}(\bp+\bq)\cong \tilde{\ve}(\bp)+vq\cos(\theta)
\label{e3}
\end{equation}
and the identity
\begin{equation}
\lim_{\a\rightarrow 0}\frac{1}{\omega} \def\G{\Gamma} \def\D{\Delta-\tilde{\ve}+i\a}=\frac{1}{i}\int_o^\infty dt
\exp{\left[i(\omega} \def\G{\Gamma} \def\D{\Delta-\tilde{\ve}+i\a)t\right]}=\frac{1}{i}
\int_0^\infty dt \left[\cos{(\omega} \def\G{\Gamma} \def\D{\Delta-\tilde{\ve})t}+i\sin{(\omega} \def\G{\Gamma} \def\D{\Delta-\tilde{\ve})t}\right]
\label{e4}
\end{equation}
we write Eq. (\ref{e2}) as
\begin{equation}
\S"(\bp,\omega} \def\G{\Gamma} \def\D{\Delta)=-\frac{g^2}{2\pi}\int_0^\infty dt \cos{(\omega} \def\G{\Gamma} \def\D{\Delta-\tilde{\ve})t}
\int_{-\infty}^\infty \frac{d\omega} \def\G{\Gamma} \def\D{\Delta'}{2\pi} Im \chi_c(\bq,\omega} \def\G{\Gamma} \def\D{\Delta')\coth{\frac{\omega} \def\G{\Gamma} \def\D{\Delta'}{2T}}
\int_0^{2\pi}\frac{d\theta}{2\pi}\exp{\left[-ivqt\cos{\theta}\right]}
\label{e5}
\end{equation}
\begin{equation}
\S'(\bp,\omega} \def\G{\Gamma} \def\D{\Delta)=\frac{g^2}{2\pi}\int_0^\infty dt \sin{(\omega} \def\G{\Gamma} \def\D{\Delta-\tilde{\ve})t}
\int_{-\infty}^\infty \frac{d\omega} \def\G{\Gamma} \def\D{\Delta'}{2\pi} Im \chi_c(\bq,\omega} \def\G{\Gamma} \def\D{\Delta')\coth{\frac{\omega} \def\G{\Gamma} \def\D{\Delta'}{2T}}
\int_0^{2\pi}\frac{d\theta}{2\pi}\exp{\left[-ivqt\cos{\theta}\right]}
\label{e6}
\end{equation}
These equations will be transformed if we perform the approximation
$\coth{\omega} \def\G{\Gamma} \def\D{\Delta/2T}\cong 2T/\omega} \def\G{\Gamma} \def\D{\Delta$ and in Eqs. (\ref{e5})-(\ref{e6}) consider
\begin{eqnarray}
S_c&=&\int_{-\infty}^\infty\frac{d\omega} \def\G{\Gamma} \def\D{\Delta'}{2\pi} Im \chi_c(\bq,\omega} \def\G{\Gamma} \def\D{\Delta')
\coth{\frac{\omega} \def\G{\Gamma} \def\D{\Delta'}{2T}}\nonumber\\
&\cong& \frac{T}{\pi}\int_{-\infty}^\infty d\omega} \def\G{\Gamma} \def\D{\Delta' \frac{D q^2 \chi_0}
{(D A_0 q^2+D \a q^4/k_F^2)^2+\omega} \def\G{\Gamma} \def\D{\Delta'^2}=\frac{\chi_0 k_F^2}{\a}\frac{T}{\xi^{-2}+q^2}
\label{e7}
\end{eqnarray}
where
\begin{displaymath}
\chi_0=\frac{N N(0)}{2}
\end{displaymath}
\begin{equation}
\xi^{-2}=\frac{A_0 k_F^2}{\a}
\label{e8}
\end{equation}
Using the exact formulas
\begin{equation}
J_0(z)=\int_0^{2\pi}\frac{d\theta}{2\pi}\exp{\left[-iz\cos(\theta)\right]}
\label{e9}
\end{equation}
\begin{equation}
K_0(kb)=\int_0^\infty dx\frac{x J_0(xb)}{x^2+k^2}
\label{e10}
\end{equation}
where $J_0(x)$ and $K_0(z)$ are the Bassel functions. The imaginary part of
the self-energy given by Eq.(\ref{e5}) has the form
\begin{equation}
\S"(\bp,\omega} \def\G{\Gamma} \def\D{\Delta)=-\frac{g^2T}{2\a\pi}\int_0^\infty dt
\cos{\left[(\omega} \def\G{\Gamma} \def\D{\Delta-\tilde{\ve}(\bp))t\right]}K_0(v_Ft\xi^{-1})
\label{e11}
\end{equation}
where $v_F$ is the Fermi velocity. Performing the integral over t in
Eq. (\ref{e11}) we obtain
\begin{equation}
\S"(\bp,\omega} \def\G{\Gamma} \def\D{\Delta)=-\frac{g^2\chi_0k_F^2}{\a}\frac{\pi}{2}\frac{T}
{\sqrt{(\omega} \def\G{\Gamma} \def\D{\Delta-\tilde{\ve}(\bp))^2+(v_F\xi)^{-2}}}
\label{e12}
\end{equation}
In the approximation $\omega} \def\G{\Gamma} \def\D{\Delta-\tilde{\ve}(\bp)\gg (v_F\xi)^{-1}$
from Eq. (\ref{e12}) we get
\begin{equation}
\S"(\bp,\omega} \def\G{\Gamma} \def\D{\Delta)\cong -\frac{g^2\chi_0T\xi^{-2}}{A_0(\omega} \def\G{\Gamma} \def\D{\Delta-\tilde{\ve}(\bp))}
\label{e13}
\end{equation}
relation which satisfies $\S"\sim 1/\omega} \def\G{\Gamma} \def\D{\Delta$.
From Eq. (\ref{e13}) we can see that due to the coupling with the charge
fluctuations the electronic excitations present a non-Fermi behavior,
obtained also at $T=0$ by Wang et al. \cite{4}, but was considered as a
holon like propagation of the charge fluctuations. The result expressed by
Eq. (\ref{e13}) has been obtained in the approximation $\omega} \def\G{\Gamma} \def\D{\Delta\ll T$ and
following the method proposed by Vilk and Tremblay \cite{12} (See Appendix
D of Ref. \onlinecite{12} for an accurate discussion about the enhancement
of the Fermi behavior in a non-Fermi behavior of electrons interacting with
fluctuations).
An important approximation for this calculation is the existence of an
energy scale for the charge fluctuations in the presence of weak disorder.
More than that this energy scale characterized by a frequency $\omega} \def\G{\Gamma} \def\D{\Delta_0$ has to
satisfy the condition $\omega} \def\G{\Gamma} \def\D{\Delta_0\ll T$, because only the coupling of the
electrons with low energy fluctuations gives a non-Fermi behavior. In this
model we require because of the weak disorder, that $\ve_F \tau=c$ has to be
large. Then we can define $\omega} \def\G{\Gamma} \def\D{\Delta_0\cong\tau^{-1}=c/\ve_F$ and the condition
$\omega} \def\G{\Gamma} \def\D{\Delta_0\ll T$ becomes $ c\ll T E_F$.
\section{Discussion}
We showed that a non-Fermi behavior may appear by the coupling of electrons
in the presence of disorder to the charge fluctuations. Such a mechanism
was also proposed for the coupling of electrons with two-dimensional spin
fluctuations \cite{13,14}. The coupling between electrons and fluctuations
near the quantum critical point has been also proposed \cite{15,16} as the
explanation for the non-Fermi behavior of the electronic system and it seems
to be an appropriate mechanism in the heavy fermion systems.
Our model can be a good explanation for the experimental data obtained by
Boebinger et al. \cite{17} on La$_{2-x}$Sr$_x$CuO$_4$ which present a linear
dependence of the temperature at the insulator-metal crossover. Recently a
similar behavior has been observed for Pr$_{2-x}$Ce$_x$CuO$_4$ (this is an
electron-doped system)\cite{18} at low temperature.
|
\section{Collisional integral}
\label{A}
The appendix is devoted to the explicit calculation of the collisional
integral (\ref{CT}).
Let us introduce the center of mass velocity ${\bf C}$ and the relative
velocity before (${\bf V}$) and after (${\bf V'}$) collision :
\begin{eqnarray}
{\bf U}_1&=&{\bf C}+{\bf V}/2 \nonumber\\
{\bf U}_2&=&{\bf C}-{\bf V}/2 \nonumber\\
{\bf U}_{1^{\prime}}&=&{\bf C}+{\bf V'}/2 \nonumber\\
{\bf U}_{2^{\prime}}&=&{\bf C}-{\bf V'}/2
\label{a1}
\end{eqnarray}
The conservation of kinetic energy during an elastic
collision ensures
\begin{equation}
V^2=V'^2\;,
\label{a2}
\end{equation}
so that the collisional integral can be rewritten in the form
\begin{eqnarray}
\langle\chi_6I_{\rm coll}\rangle=-
(\delta\theta_z-\delta\theta_{\perp})
\frac{3}{128\pi}\frac{m\sigma_0}{N\theta_0^2}\nonumber\\
\int d^3{ r}\;d^3{V}\;d^3{C}\;d^2\Omega\;V
\;f_0(1)\,f_0(2)\,[V_z^2-V_{z^{\prime}}^2]^2\;.
\label{lin2}
\end{eqnarray}
Let us first calculate the angular integral :
\begin{equation}
I_{\Omega}\equiv\int d\Omega\big[ V_z^2-V'^2_z\big]^2
\end{equation}
We introduce a reference frame $\Re$
linked to ${\bf V}$, the vectors of the associated orthonormal basis
beeing $(\hat{\bf a},\hat{\bf b},\hat{\bf c})$.
Without loss of generality we choose
$\hat{\bf a}$ such that ${\bf V}=V\cdot\hat{\bf a}$, and the $z$ axis in
the plane
generated by $(\hat{\bf a},\hat{\bf b})$.
The relative velocity ${\bf V}'$ is characterized in
$\Re$ by two spherical angles
$(\theta',\varphi')$ :
\begin{eqnarray}
{\bf V}'.\hat{\bf a}&=&V\cos\theta' \\
{\bf V}'.\hat{\bf b}&=&V\sin\theta'\cos\varphi' \\
{\bf V}'.\hat{\bf c}&=&V\sin\theta'\sin\varphi'
\end{eqnarray}
Thus
$$
V'_z={\bf V}'\cdot\hat{\bf z}=V_z\cos\theta'+
V\sin\theta'\cos\varphi'(\hat{\bf b}\cdot\hat{\bf z})\;,
$$
where $\hat{\bf z}$ is the unit vector of the $z$ axis and
\begin{eqnarray*}
V'^2_z&=&V^2_z\cos^2\theta'+V^2\sin^2\theta'\cos^2\varphi'(\hat{\bf
b}\cdot
\hat{\bf z})^2\\
&+&2VV_z\cos\theta'\sin\theta'\cos\varphi'(\hat{\bf b}\cdot\hat{\bf
z})\;.
\end{eqnarray*}
With our choice of coordinate, one has
$$\hat{\bf z}\cdot\hat{\bf z}=1=(\hat{\bf a}\cdot\hat{\bf z})^2+
(\hat{\bf b}\cdot\hat{\bf z})^2\;,$$
which implies
$$
V^2(\hat{\bf b}\cdot\hat{\bf z})^2=V^2-V^2_z
$$
and, finally,
\begin{eqnarray*}
V^2_z-V'^2_z&=&V^2_z(1-\cos^2\theta')-\sin^2\theta'\cos^2\varphi'(V^2-V^2_z)\\
&2&VV_z\sin\theta'\cos\theta'\cos\varphi'(\hat{\bf b}\cdot\hat{\bf
z})\;.
\end{eqnarray*}
By integrating the square of the previous expression, one finds
\begin{equation}
I_\Omega=\frac{32\pi}{15}\left(\frac{15}{8}V^4_z+\frac{3}{8}V^4-
\frac{5}{4}V^2_zV^2\right)\;.
\label{Iomega}
\end{equation}
The calculation of
the collisional integral (\ref{lin2}) is now straightforward, and finally
yields the result :
\begin{equation}
\langle\chi_6I_{\rm coll}\rangle=-
(\delta\theta_z-\delta\theta_{\perp})\frac{4}{5m}v_{\rm th}\sigma_0n(0)\;,
\label{collint}
\end{equation}
where $v_{\rm
th}=\sqrt{8\theta_0/\pi
m}$ is the thermal velocity of a particle of the gas.
(\ref{collint}) permits to derive (\ref{tg1},\ref{tg}) with
$\gamma_{\rm coll}$ defined by eq. (\ref{gcoll}).
|
\section{Introduction}
A large number of cosmological probes now suggest that the Universe
is spatially flat with a low mass density (e.g., Perlmutter et al. 1998;
Riess et al. 1998; Lineweaver 1998; Guerra \& Daly 1998; Bahcall \& Fan
1998). In addition to the mass density, gravitational lensing
statistics have allowed limits to be placed on the
cosmological constant. However, current limits on
the cosmological constant from gravitational lensing
arguments are only based on
lensing statistics due to foreground galaxies (e.g., Kochanek 1996;
Falco, Kochanek, Munoz 1998; Cheng \& Krauss 1999; Cooray, Quashnock, Miller
1999; Cooray 1999a; Quast \& Helbig 1999)\footnote{We note that other
techniques, such as the luminosity distance to
Type Ia supernovae at high redshifts (e.g., Perlmutter et al. 1998;
Riess et al. 1998), also allow constraints to be placed on the
cosmological constant}.
An alternative approach is to consider
lensing statistics due to foreground galaxy clusters (e.g.,
Wu \& Hammer 1993; Bartelmann et al. 1998; Cooray 1999b).
It is well known that galaxy cluster evolution
is strongly sensitive to the cosmological mass density of the Universe (e.g.,
Bahcall \& Fan 1998; Viana \& Liddle 1998). Since lensing statistics
are sensitive to the cosmological constant, it is likely that
the number of lensed arcs due to galaxy clusters
can provide strong constraints on both the mass density and the
cosmological constant.
Since the first suggestion that lensed optical arcs can be used as a
cosmological probe (Wu \& Hammer 1993), several studies have addressed
specific issues related to the statistical calculation. These
include the effect of a cosmological constant (Wu \& Mao 1996) and
background source evolution (Hamana \& Futamase 1997).
The numerical works by Bartelmann et al. (1998), using
simulated clusters in three cosmological models, suggested that
current observational statistics on lensed
arcs are consistent with predictions in an open
Universe ($\Omega_\Lambda = 0$) with $\Omega_m \sim 0.3$.
In Cooray (1999b; hereafter C99), we extended the predictions to
general cosmologies and also predicted the existence
of lensed radio and sub-mm towards foreground clusters.
Here, we extend the calculation in C99 by including various
uncertainties in the predicted number of lensed optical sources
to study the possibility of obtaining limits on cosmological
parameters based on the observed number.
In \S~2, we describe our calculation and inputs for the prediction.
In \S~3, we compare the predicted number of lensed arcs to the
observed number and use a reliable lower limit on the observed number to
derive an upper limit on the cosmological mass density of the
Universe. We follow the conventions that the Hubble constant,
$H_0$, is 100\,$h$\ km~s$^{-1}$~Mpc$^{-1}$, the present matter energy
density in units of the closure density is $\Omega_M$, and the normalized
cosmological constant is $\Omega_\Lambda$. Unless otherwise noted,
quoted errors are 1-$\sigma$ statistical errors.
\section{Gravitational Lensing Statistics}
In this section, we briefly describe our calculation and especially
the description of foreground lensing clusters (\S~2.1) and background
sources (\S~2.2). We also introduce a nonsingular isothermal
sphere model to describe galaxy cluster dark matter profile, which is primarily
motivated by recent determinations of the cluster potentials using high
performance numerical inversions of combined strong and weak lensing
data towards a sample of galaxy clusters.
\subsection{Foreground Lenses}
The differential probability that a beam towards a background source
will encounter a foreground lens with a path length of $dz_L$ is:
\begin{equation}
d\tau = n(z_L) a_{\rm lens} \frac{c dt}{dz_L}dz_L,
\end{equation}
where $n(z_L)$ is the number density of foreground lenses at redshift
$z_l$ while $a_{\rm lens}$ is the lensing cross section (e.g.,
Fukugita et al. 1992).
Using the Press-Schechter mass function (Press \& Schechter 1974; PS),
the comoving number density of galaxy clusters, $dn(M,z)$, at
redshift $z$ and mass $(M,M+dM)$, can be written as
\begin{eqnarray}
\frac{dn(M,z)}{dM} = \\ \nonumber
- \sqrt{\frac{2}{\pi}} \frac{\bar{\rho}}{M}\frac{d\sigma(M,z)}{dM}
\frac{\delta_{c}}{\sigma^2(M,z)}
\exp{\left[\frac{-\delta_{c}^2}{2 \sigma^2(M,z)}\right]}\, , \nonumber
\end{eqnarray}
where $\bar{\rho}$ is the comoving background matter density,
$\sigma^2(M,z)$ is the variance of the fluctuation spectrum averaged over
a mass scale $M$, and $\delta_{c}$ is the
linear overdensity of a perturbation which has collapsed and virialized.
Taking an approach similar to the one presented in Viana \& Liddle
(1998), $\sigma(M,z)$ is written as a function of the comoving radius,
$R$, which contains mass $M$ at the current epoch:
\begin{equation}
\sigma(R,z)=\sigma _8(z) \left({ R \over 8 h^{-1}
{\rm Mpc}}\right)^{-\gamma(R)},
\end{equation}
where
\begin{equation}
\gamma(R) = (0.3\Gamma + 0.2)\left[2.92 + \log_{10}
\left({R \over 8 h^{-1} {\rm Mpc}}\right)\right] \, .
\end{equation}
Here, $\Gamma=0.23 \pm 0.05$ (Peacock \& Dodds 1994)
is the CDM shape parameter; our results are insensitive to its
specific value (e.g., Viana \& Liddle 1998).
In order to calculate growth evolution as a function of
redshift in various cosmologies, we write $\sigma_8(z)$ as:
\begin{equation}
\sigma_8(z) = {\sigma_8(0)\over 1+z}\,\,
{g(\Omega_m(z)) \over g(\Omega_m(0))} \, ,
\end{equation}
where, following Carroll, Press \& Turner
(1992), the growth suppression factor is:
\begin{equation}
g(\Omega_m) = {5 \over 2} \Omega_m
\left[ \Omega_m^{4/7} - \Omega_{\Lambda} +
\left(1 + {\Omega_m \over 2} \right)
\left(1 + {\Omega_{\Lambda} \over 70} \right) \right]^{-1} \, .
\end{equation}
The normalization for $\sigma_8$ comes from the local temperature
function (Pen 1998):
\begin{equation}
\sigma_8(0) = \left\{ \begin{array}{cl}
(0.53 \pm 0.05) \, \Omega_m^{-0.46} & {\rm \Omega_\Lambda=0\,} \\
(0.53 \pm 0.05) \,\Omega_m^{-0.53} & {\rm \Omega_m+\Omega_\Lambda=1\, .} \\
\end{array} \right.
\end{equation}
In order to model the cluster lensing potential, we use the
nonsingular singular
isothermal sphere model with the observed velocity
dispersion and an a priori determined value for the core radius of the
dark matter potential of the cluster. The evidence for
a core radius in the dark matter profile of galaxy clusters primarily
comes from the existence of gravitationally lensed arcs in the radial direction
from the cluster center. For simple models for the cluster potential
involving
singular isothermal models, such arcs are located at a distance
equivalent to the core radius of the cluster pontential profile.
Also, recent numerical inversions of galaxy cluster lensing potentials
using Hubble Space Telescope and other ground based high quality images
clearly suggest the presence of a small core
radius (Tyson, Kochanski, Dell'Antonio 1998; Ian Dell'Antonio,
private communication). Thus, it is necessary that we consider a lensing model
which allows for the possible presence of a core radius. Following Hinshaw \&
Krauss (1987), we consider a isothermal sphere model with a core
radius and write the density profile as:
\begin{equation}
\rho = \frac{\sigma_{\rm vel}^2}{2 \pi G(r^2+r_c^2)},
\end{equation}
where $\sigma_{\rm vel}$ is the dark matter velocity dispersion
and $r_c$ is the core radius of the dark matter profile of the
cluster. The conventional singular isothermal sphere (SIS) is
recovered when $r_c$ is zero.
The lensing cross section for the nonsingular isothermal model
is given by:
\begin{equation}
a_{\rm lens} = 16 \pi^3 \left( \frac{\sigma_{\rm vel}}{c} \right)^4
\left( \frac{D_{OL}D_{LS}}{D_{OS}}\right)^2 f(\beta)
\end{equation}
where $D_{OL}$, $D_{OS}$ and $D_{LS}$ are observer to lens,
observer to source and lens to source distances. These distances are
calculated under the filled beam approximation.
In Eq.~9, $f(\beta)$ is a correction factor that takes into account
the nonsingular behavior of the density profile (see, Hinshaw \&
Krauss 1987):
\begin{equation}
f(\beta) = 1 +5\beta-\frac{\beta^2}{2} -
\frac{\sqrt{\beta}(4+\beta)^{3/2}}{2},
\end{equation}
where $\beta$ is the ratio of core radius to critical radius of the
lensing potential, with the latter measured at the redshift of the
cluster:
\begin{equation}
\beta = \frac{r_c c H_0 (1+z_L)}{4 \pi \sigma_{\rm vel}^2} \left(
\frac{D_{OS}}{D_{LS}D_{OS}}\right).
\end{equation}
When the SIS model is considered, $\beta = 0$ and $f(\beta)=1$.
For small core radii, especially for the present case involving
galaxy clusters, one can usually ignore higher order $\beta$
terms associated with $f(\beta)$; we consider, however, the full
formula in deriving cosmological parameters.
Finally, the differential optical depth for the nonsingular isothermal
model is:
\begin{eqnarray}
d\tau = 16 \pi^3 \left(\int_{M_{\rm min}}^{\infty}
\left[\frac{\sigma_{\rm vel}(M')}{c}\right]^4 \frac{dn(M',z_L)}{dM'}\,
f(\beta) dM'\right) \\ \nonumber
\times
(1+z_L)^3 \left( \frac{D_{OL}D_{LS}}{D_{OS}}\right)^2 \frac{cdt}{dz_L}dz_L,
\end{eqnarray}
The total
optical depth to a given background redshift, $z_s$, is given by:
\begin{equation}
\tau(z_s) = \int_{0}^{z_s} \frac{d\tau}{dz_L}dz_L\;.
\end{equation}
In order to calculate the lensing optical depth, we take a two step
approach to relate cluster velocity dispersion to its mass.
We relate velocity dispersion to cluster temperature
using recently updated $\sigma-T$ relation (Wu et al. 1998):
\begin{equation}
\sigma_{\rm vel}(T) = 10^{2.57 \pm 0.03} \left(\frac{T}{{\rm keV}}\right)^{0.56
\pm 0.09}\; {\rm km\; s^{-1}};
\end{equation}
and then to mass using partly theoretical
$M-T$ relation (e.g., Barbosa et al. 1996):
\begin{eqnarray}
T(M,z) = (6.8 \pm 0.5) h^{\frac{2}{3}}\; {\rm keV}\;
\left[\frac{\Omega_m \Delta_c(\Omega_m,z)}{178}\right]^{\frac{1}{3}}
\\ \nonumber
\times \left(\frac{M}{10^{15}h^{-1} M_{\sun}}\right)^{\frac{2}{3}}(1+z).
\end{eqnarray}
We have allowed for an extra
uncertainty in the $M-T$ relation by comparing various
normalizations that have been suggested in the literature.
Also, we note that $\beta$ is dependent on cluster mass through
velocity dispersion (Eq.~11). In addition to velocity dispersion, it is likely
that cluster core radii are also dependent on individual cluster
masses. Even though such variations have been
observationally determined for galaxies
(e.g., Lauer 1985), there is still no observational evidence
for a dependence of galaxy cluster core radii with other physical properties,
such as the X-ray luminosity or temperature. For the purpose of this
calculation, we take a constant value for the core radius based on the
mean value of cluster core radii from numerical
inversions ($\sim$ 35 h$^{-1}$ kpc; Dell'Antonio private
communication).
When deriving cosmological parameters, we vary the exact value
of the core radius to investigate the parameter dependences
on it; As we find later, our limits on
$\Omega_m$ is weakly dependent on the core radius.
\vskip 2mm
\hbox{~}
\centerline{\psfig{file=fig1.ps,width=4.0in,angle=-90}}
\noindent{
\addtolength{\baselineskip}{-3pt}
\vskip 1mm
Fig.~1.\ Optical depth for strong lensing for a background source
at redshift of 1, due to foreground massive clusters. Shown are
the 95\% confidence ranges for both flat ({\it dot-dashed lines}) and open
({\it solid lines}) cosmologies
with and without a cosmological constant. Given the large uncertainty
associated with the optical depth and the small difference between
flat and open cosmological models,
it is unlikely that lensed arc statistics can be used to reliably
place limits on the cosmological constant.
\vskip 3mm
\addtolength{\baselineskip}{3pt}
}
In Fig.~1, as an illustration,
we show the optical depth to lensing due to foreground
clusters with total masses greater than
$7.5 \times 10^{14}\; h^{-1}\; {\rm M_{\sun}}$ for a
background source at a redshift of 1, as a function of $\Omega_m$ for
open and flat cosmologies, and considering a lensing potential
in which $f(\beta)=1$ (SIS model). Shown are the 95\% upper and lower
confidences in each case by considering all possible errors we
have so far considered. The uncertainty in the optical depth is
primarily dominated by the error associated with the normalization of
the PS mass function; since the number density of massive clusters is strongly
sensitive to the exponential term in Eq.~2,
small changes in $\sigma_8$ can produce order of magnitude
changes in the number density. The difference between flat and open
cosmological models is primarily due to the increase in lensing probability
with the addition of $\Omega_\Lambda$. However, this difference
is small, and when errors in observations are also
considered, it is impossible to study the possible
existence of a cosmological constant
using lensed arc statistics. Therefore,
taking a conservative approach, we combine the upper curve valid for flat
cosmologies with the lower curve defined by open models to
combine the 95\% confidence range in the predicted number of lensed sources.
\subsection{Background Sources}
In order to obtain reliable predictions on the number of lensed arcs,
it is important that both the background source evolution
and effects such as ``magnification bias'' (Kochanek 1991) be
included in the calculation. Our description of background sources
comes from the Hubble Deep Field (HDF; Williams et al. 1996). We use the
HDF redshift and magnitude
distribution and the luminosity function from Sawicki et al. (1997).
Such an approach allows us to reliably account for the true
redshift distribution of
background sources, instead of an empirical distribution
or a constant redshift, while also accounting for
intrinsic evolutionary effects which has shown to be
important for lensing predictions (e.g., Hamana \& Futamase 1997)
Using the probability,
$\tau(z,\Omega_m,M_{\rm min})$,
for a source at redshift $z$ to be strongly lensed
and the number of unlensed background sources between rest-frame
luminosity $L$ and $L+dL$ and between redshifts $z$ and $z+dz$,
$\Phi(L,z)dL\,dz$, we can write the number of lensed galaxies,
$d\bar N$, in that luminosity and
redshift interval as (see, also Maoz et al.\ 1992):
\begin{eqnarray}
{d\bar N(L,z)\over dz}=\tau(z,\Omega_m,M_{\rm min}) \\ \nonumber
\times \int\left[\Phi\left({L\over A},z\right)\,
{dL\over A}\right]f(A,L,z)q(A)\,dA \;.
\end{eqnarray}
Here, the integral is over all allowed values of $A$, the amplification
of the brightest lensed image,
$q(A)$ is the probability distribution of amplifications, and
$f(A,L,z)$ is the probability of observing the brightest image
given $A$, $L$, and $z$. Our assumption that the lenses are nonsingular
isothermal spheres implies that the minimum amplification, $A_{\rm
min}$, is a function of $\beta$. In general, the probability
distribution of amplifications can be written as:
\begin{equation}
q(A)\,dA=2 A_{\rm min}^2 A^{-3}\,dA.
\end{equation}
\vskip 2mm
\hbox{~}
\centerline{\psfig{file=fig2.ps,width=4.0in,angle=-90}}
\noindent{
\addtolength{\baselineskip}{-3pt}
\vskip 1mm
Fig.~2.\ Minimum amplification versus
$\beta$, the ratio of core radius to
critical radius at the redshift of the lensing cluster.
\vskip 3mm
\addtolength{\baselineskip}{3pt}
}
In Fig.~2, we show $A_{\rm min}$ as a function of $\beta$, which is
calculated following Cheng \& Krauss (1999).
In practice we use a fitting function that returns $A_{\rm min}$
for a given value of $\beta$, with an accuracy of better than 0.1\%
at all interested values of $\beta$ in the present calculation.
For simplicity, we assume that
$f(A,L,z)$ is a step function, $\Theta[m_{\rm lim},A]$, so that a
lensed image with apparent magnitude brighter than $m_{\rm lim}$
is detected. For a given value of the core radius and the
velocity dispersion, $\beta$ is determined from
Eq.~11. For massive clusters
discussed here with velocity dispersions of the order $\gtrsim$
1000 km s$^{-1}$ and at redshifts $\sim$ 0.2, $\beta \lesssim 0.07$
and $A_{\rm min} \lesssim 5$. Compared to
the SIS model, the addition of a small core radius only produces
slight changes in the lensing probability. As
described in Kochanek (1995), the effect of a core radius
is to increase the magnification bias while increasing the effective
lensing cross section; the overall effect is that the
presence of a core radius is not significantly different from that
of a SIS model.
We assume that the brightness distribution of background galaxies
at any given redshift is described by a Schechter function (Schechter
1976), in which the comoving density of galaxies at redshift $z$ and with
luminosity between $L$ and $L+dL$ is
\begin{equation}
\phi(L,z)\, dL=\phi^*(z)\left[L\over{L^*(z)}\right]^{\alpha(z)}
e^{-L/L^*(z)}\, dL\; ,
\end{equation}
where, as before, both $L$ and $L^*$ are measured in the rest
frame of the galaxy. Following CQM,
we can write the expected number $\bar N$ of lensed sources as
\begin{eqnarray}
\bar N &=\sum_i \tau(z_i,\Omega_m,M_{\rm min})
\int_2^\infty A^{-1-\alpha(z_i)}e^{L_i/L^*(z_i)}
e^{-L_i/AL^*(z_i)} \nonumber \\
&\times \Theta\left[m_{\rm lim},A\right]
{2\over{(A-1)^3}}dA\;,
\end{eqnarray}
where the sum is over each of the background galaxies.
The index $i$ represents
each galaxy; hence, $z_i$, $L_i$, and $m_i$ are, respectively,
the redshift, rest-frame luminosity, and apparent magnitude
of the $i$th galaxy.
Since $L_i$ for individual galaxy is unknown, due to uncertain
K-corrections, following Cooray, Quashnock \& Miller (1999),
we estimate the total average bias
by weighting the integral in Eq.~19 by a normalized
distribution of luminosities $L_i$ drawn from the Schechter function
appropriate for the redshift $z_i$ of galaxy $i$.
We calculated the magnification bias for
individual redshift intervals for which the Schechter function
parameters are available in Table 1 of Sawicki et al.
(1997). In principle, the uncertainties
in the Schechter function parameters at a given redshift can affect
the calculation of the bias, but in practice only the
uncertainty in the power-law slope $\alpha$ has a significant effect.
The effect of varying $\alpha$ on the lensing statistics was discussed
in Cooray, Quashnock \& Miller (1999) for lensing statistics involving
foreground galaxies in the Hubble Deep Field,
which is also valid for the present case
involving galaxy clusters; the general
effect due to uncertainties tabulated in Sawicki et al. (1997) is that the
constraints on cosmological parameters vary by less
than 5\% percent, when $\alpha$ is in general varied by the quoted
1-$\sigma$
uncertainties in Sawicki et al. (1997).
Here, we take a conservative approach
and allow the largest possible bias, so that the expected number is
overestimated by an amount as suggested above. The only effect of this
approach is to
slightly increase our upper limit on $\Omega_m$.
\vskip 2mm
\hbox{~}
\centerline{\psfig{file=fig3.ps,width=4.0in,angle=-90}}
\noindent{
\addtolength{\baselineskip}{-3pt}
\vskip 1mm
Fig.~3.\ Expected number of lensed arcs on the whole sky with
amplifications
greater than 10 and V-band magnitudes brighter than 22 towards
foreground massive clusters. The shaded range shows the 95\%
confidence upper and
lower limits on the expected number of lensed arcs, while the
horizontal lines show the range of current observed numbers. We use the
lower limit on the current observed number to impose an
upper limit on $\Omega_m$.
\vskip 3mm
\addtolength{\baselineskip}{3pt}
}
In Figure~3, we show the expected number of lensed arcs towards
foreground massive clusters with total mass greater than $M_{\rm min}
= 7.5 \times 10^{14}\; h^{-1} {\rm M_{\sun}}$, and assuming a zero
core radius for the lensing potential.
We define an arc as a lensed source which is amplified by a factor
equal to or greater than 10. To make a direct
comparison to both observations and prior predictions, we impose
a limiting V-band magnitude of 22. Our numbers can be directly
compared to previous estimates, especially those of Bartelmann et al.
1998). This study predicted $\sim$ 2400 arcs in an open Universe with
$\Omega_m \sim 0.3$ and $\sim$ 36 arcs in an Einstein-de Sitter Universe.
The number expected in a flat Universe with $\Omega_m \sim 0.3$
and $\Omega_\Lambda \sim 0.7$ was $\sim$ 280. Our estimates for
for an Einstein-de Sitter Universe range from $\sim$ 0.1 to 60
while for $\Omega_m \sim 0.3$ Universe (independent of
$\Omega_\Lambda$) is $\sim$ 50 to 7000 (with the higher end allowed by
the cosmological constant). The primary reason for a lower number of
arcs with $\Omega_\Lambda$ in the study by Bartelmann et al. (1998)
was their assumption that clusters are different in Universes with
a cosmological constant, such that their concentration is lower.
Based on numerical simulations performed by the Virgo Consortium, however,
Thomas et al. (1998) studied a series of clusters in four different
cosmologies, including an open model with $\Omega_m=0.3$ and
a flat model with a cosmological constant of $\Omega_\Lambda=0.7$.
The authors concluded that clusters do not exhibit differences
between open and flat cosmologies with and without
a cosmological constant and that cluster structures
cannot be used to discriminate between the two possibilities.
If Thomas et al. (1998) are correct, then
the inclusion of a cosmological constant is not expected to change cluster mass
profiles to an extent that would affect the gravitational lensing rate.
In any case, such systematic effects are unlikely to
be nearly as large as the current uncertainty in $\sigma_8$ which
dominates the present calculation on the lensing rate.
Ignoring this case, our predictions are
generally consistent with Bartelmann et al. (1998).
As we have demonstrated in Fig.~1, lensed arc statistics are unlikely
to provide useful limits on the cosmological constant.
The same is true for alternatives
to the cosmological constant, such as scalar field and quintessence
models that have recently been introduced
(e.g., Steinhardt et al. 1998).
\section{Constraints on $\Omega_m$}
In order to derive a limit on $\Omega_m$ based on the number of lensed
arcs, we require knowledge on the observed number of such lensing
events. Current surveys of clusters are based on their X-ray
luminosities rather than masses. For an example, the luminosity
cutoff of the followup Einstein Medium Sensitivity Survey (EMSS)
cluster arc survey by Le F\`evre et al. (1994)
is $8 \times 10^{44}\, h^{-2} {\rm ergs\, s^{-1}}$, measured in the
EMSS band of 0.3 to 3.5 keV. Converting this to a total mass
by following through a recently derived $L-T$ relation (Arnaud \&
Evrard 1998) in addition to the above $M-T$ relation suggest a reliable lower
limit on the mass of $7.5 \times 10^{14}\, h^{-1}\, {\rm M_{\sun}}$.
We have taken the lowest limit on mass by considering all
cosmologies -- since the $M-T$ relation and $L$ estimates are
different under varying cosmologies.
The current observed arc statistics (e.g., Le F\'evre et al. 1994; Luppino
et al. 1998), when converted to a whole sky number, suggest
that the number of arcs towards above defined massive clusters
and with amplifications greater than 10 down to a V-band limiting
magnitude of 22 is between 1500 and 2500 (e.g., C99; Bartelmann et
al. 1998). Ignoring the upper value, which is likely to be
unreliable, we use the lower estimate to derive an upper limit on $\Omega_m$.
Since the lower estimate is based on the observed number,
this allows us to put a reliable upper limit on $\Omega_m$.
We also vary this lower limit to study its effects on our constraints.
In order to derive a constraint on $\Omega_m$, we
adopt a Bayesian approach, and take a uniform prior for
$\Omega_m$ between 0 and +1. This is primarily due to the fact that
we do not yet have a precise determination of
$\Omega_m$, and, based on various theoretical arguments,
we do not wish to consider cosmologies in
which either this quantity lies outside the interval [0,1].
Since the prior for $\Omega_m$ is uniform,
the posterior probability density is simply proportional to the
likelihood.
The likelihood ${\cal L}$ --- a function of $\Omega_m$ ---
is the probability of the data, given $\Omega_m$.
The likelihood for $n$ observed arcs (at redshifts
$z_j$) when $\bar N$ is expected is given by (Cooray, Quashnock \&
Miller 1999):
\begin{equation}
\langle {\cal L}(n)\rangle =\prod_{j=0}^n \tau(z_j) \times e^{- \bar N} \times
\left(1 + \sigma_{\tau}^2 \left[ \frac{\bar N^2}{2}-n\bar N +\frac{n(n-1)}{2}\right]\right) \; .
\end{equation}
We have taken into account the uncertainty in predicted lensing rate
by introducing $\sigma_{\tau}$, which is the fractional 1-$\sigma$
error on $\tau$ and then mariginalizing the likelihood
over the variance of it. In order to constrain $\Omega_m$, we also need the
redshifts $z_j$ of the observed arcs. Using published data on
individual lensing clusters, we obtained a median redshift for the lensed
arcs of $\bar z \sim 1.6$. As we find below,
changing this mean redshift to a
reasonably different value does not change our constraints on
the $\Omega_m$ greatly.
When the observed lower limit is
compared with predictions, we find that $\Omega_m \lesssim 0.62$ at
the 95\% confidence when there is no core radius
(SIS). When we include a core radius of 35 h$^{-1}$ kpc, the upper
limit on $\Omega_m$ decreases to 0.56 at the 95\% confidence level; The change
in the upper limit on $\Omega_m$ is only a minor effect.
If the core radius were to be as large as 100 h$^{-1}$ kpc,
then the derived upper limit on $\Omega_m$ can be as low as 0.29 at the
95\% confidence level.
However, such a large core radius
for the cluster dark matter profile is ruled out leaving the
possibility
only for a much smaller core radius of the order 30 to 40 h$^{-1}$ kpc.
When the effective median redshift of lensed
arcs are changed to a lower number as $\sim$ 1, the upper limit
increases to 0.58 from 0.56, while when the redshift is increases to
a value of 3.0 from 1.6, the upper limit on $\Omega_m$ at the 95\%
confidence decreases to 0.52.
When we increase the lower limit from 1500 to 2000, our upper limit on
$\Omega_m$ with a model involving core radius of 35 h$^{-1}$ kpc,
decreases to 0.52 from 0.56. This is primarily due to the fact that
the expected number of lensing events varies by orders of magnitude when
$\Omega_m$ is changed from 1 to 0, with the variation in the
expected number larger at the lower end of $\Omega_m$ values.
For such a small core radius, the limit on $\Omega_m$ is consistent with
current estimates based on other cosmological probes such as
type Ia supernovae and galaxy cluster abundances.
We note that our limit on $\Omega_m$ does not mean that the Universe
is open without a cosmological constant, but rather lensed arc statistics
are not sensitive enough to the cosmological constant
to see its effects above the current uncertainties.
In general, the upper limit on $\Omega_m$ with a cosmological constant
is slightly higher when compared to an open model. However, this
difference is rather small ($\sim$ few percent, see
Fig.~1), and cannot be distinguished using
current observations on cluster number counts and lensed arcs.
\subsection{Uncertainties \& Systematic Effects}
Using a lower limit on the observed number of lensed arcs, we have
derived an upper limit on $\Omega_m$. A major uncertainty is likely to
come when estimating a lower limit on the
observed number of arcs since it is only based on
optical followup observations of EMSS clusters
(e.g., Henry et al. 1992). As a reliable
approach, we have taken the lower limit allowed by the observed number
of lensed arcs towards this sample.
In reality, the true number is likely to be higher but the lower
limit allows us to safely consider upper limits on cosmological
parameters, especially the cosmological mass density.
The present observational number on the number of lensing events is unlikely to
be improved unless large samples of clusters are followed up
at optical wavelengths. Several attempts are currently underway (e.g.,
Luppino et al. 1998), however,
all such surveys our still based on the EMSS sample. It is likely
that the optical followup observations of additional
cluster catalogs, such as the ROSAT Bright Cluster Survey (BCS; Ebeling et
al. 1998), can greatly improve our knowledge on the lensing statistics
due to galaxy clusters allowing better constraints on the cosmological
parameters.
In addition to current low number statistics,
other uncertainties are likely to come
from the conversion of observations, such as cluster X-ray luminosity,
to mass. However, at each step, we have considered various estimates
such that the predicted number of lensed arcs is overestimated;
This approach allows us to consider a reliable limit on $\Omega_m$,
whose upper limit may have been systematically increased by our procedure.
We have also investigated the effect of a core radius on
arc statistics. As found, for luminous optical arcs with
amplifications greater than 10, the effect of a core radius on our
prediction on the number of lensing events is minimal. The upper
limit only varies from 0.62 to 0.56 at the 95\% confidence when
a reasonable core radius of size 35 h$^{-1}$ kpc is introduced.
Increasing the core radius as high as 100 h$^{-1}$ kpc reduces
the upper limit by a factor of $\sim$ 2, however, such a large core
radius is ruled out by current observations of gravitational lensing
of clusters (e.g., the nonexistence of radial arcs at large distances
from the cluster center).
\section{Summary \& Conclusions}
Using a lower limit on the observed number of lensed arcs due to
clusters, we have calculated an upper limit on $\Omega_m$.
Due to large uncertainties in the predicted number of lensed sources,
primarily dominated by the error in $\sigma_8$, we are unable to
place limits on the cosmological constant. However, after considering
possible known errors, and carefully taking account various estimates
such that the upper limit on $\Omega_m$ is not reduced, we
conclude that $\Omega_m \lesssim 0.62$ at the 95\% confidence.
\acknowledgments
I am grateful to Richard Mushotzky for pointing out the possibility
to derive an upper limit on $\Omega_m$, Ian
Dell'Antonio for communicating details of his numerical inversions
and constraints on cluster density profiles.
I also acknowledge useful comments from an anonymous
referee which led to several improvements in the paper and acknowledge
partial support from a McCormick Fellowship at the University of Chicago.
|
\section*{General Introduction}
\label{sec_introduction}
The physics of central collisions is the physics of the Quark Gluon
Plasma. Apart from projects like the search for new physics at very high
rapidities (see the CASTOR subproject at ALICE for a search
for Centauro events at LHC \cite{Angelis99}),
``Non QGP Physics'' may be defined as the physics of
peripheral collisions, which includes the effects of coherent photons
and diffraction effects (Pomeron exchange). It is our aim to show that
one will be able at CMS to address very interesting physics topics in a
rather clean way.
Central collision events are characterized by a very high
multiplicity. On the other hand, the multiplicity in peripheral
collisions is comparatively low. The ions do not interact directly
with each other and move on essentially undisturbed in the beam
direction. The only possible interaction are therefore due to
the long range electromagnetic interaction and diffractive processes.
Due to the coherent action of all the protons
in the nucleus, the electromagnetic field is very strong and the
resulting flux of equivalent photons is large. It is proportional to
$Z^2$, where Z is the nuclear charge. Due to the very short
interaction times the spectrum of these photons extends up to an
energy of about 100GeV in the laboratory system. The coherence
conditions limits the virtuality of the photon to very low values of
$Q^2 < 1/R^2$, where $R=1.2fm A^{1/3}$ is the nuclear size.
Hard diffractive processes in heavy ion collisions have also been studied.
These are interesting processes on their own, but they are
also a possible background to photon-photon and photon-hadron
interactions. The physics potential of such kind of collisions is
discussed in Section \ref{sec_photon} (this is an extension of CMS
note1998/009). It ranges from studies in QCD and strong field QED to
the search for new particles (like a light Higgs particle). This kind
of physics is strongly related to ${\gamma\gamma}$ physics at
$e^+e^-$-colliders with increased luminosity. In view of the strong
interaction background, experimental conditions will be somewhat
different from the ${\gamma\gamma}$ physics at $e^+e^-$-colliders. A
limitation of the heavy ions is that only quasireal but no highly
virtual photons will be available in the A-A collisions.
Another aspect is the study of photon-hadron interactions, extending
the $\gamma$-p interaction studies at HERA/DESY to $\gamma$-A interactions,
also reaching higher invariant masses than those possible at HERA.
At the STAR (Solenoidal Tracker At RHIC) detector at RHIC --- to be
scheduled to begin taking data in 1999 --- a program to study
photon-photon and Pomeron interactions in peripheral collisions
exists \cite{KleinS97a,KleinS97b,KleinS95a,KleinS95b,Nystrand98}.
At RHIC the photon flux will be of the same order of
magnitude, but the spectrum is limited to up to about 3 GeV.
\section{Photon-Photon and Photon-Hadron Physics}
\label{sec_photon}
\subsection{Abstract}
\label{ssec_abstract}
Due to coherence, there are strong electromagnetic
fields of short duration in very peripheral collisions. They give rise to
photon-photon and photon-nucleus collisions with high flux up to an
invariant mass region hitherto unexplored experimentally. After a
general survey photon-photon luminosities in relativistic heavy ion
collisions are discussed. Special care is taken to include the effects
of strong interactions and nuclear size. Then photon-photon physics at
various ${\gamma\gamma}$-invariant mass scales is discussed.
Invariant masses of up to about 100
GeV can be reached at LHC, and in addition the potential for new
physics is available. Photonuclear reactions and other important background
effects, mainly diffractive processes are also discussed.
Lepton-pair production, especially electron-positron pair production
is copious. Due to the strong fields there will be new phenomena,
like multiple $e^+e^-$ pair production.
\subsection{Introduction}
\label{ssec_intro}
The parton model is very useful to study scattering processes at very
high energies. The scattering is described as an incoherent
superposition of the scattering of the various constituents. For
example, nuclei consist of nucleons which in turn consist of quarks
and gluons, photons consist of lepton pairs, electrons consist of
photons, etc.. We note that relativistic nuclei have photons as an
important constituent, especially for low enough virtuality
$Q^2=-q^2>0$ of the photon. This is due to the coherent action of all
the charges in the nucleus. The virtuality of the photon is related
to the size $R$ of the nucleus by
\begin{equation}
Q^2 \mathrel{\mathpalette\vereq<} 1/R^2,
\end{equation}
the condition for coherence. The radius of a nucleus is given
approximately by $R=1.2$~fm~$A^{1/3}$, where $A$ is the nucleon
number. From the kinematics of the process one has
\begin{equation}
Q^2=\frac{\omega^2}{\gamma^2}+q_\perp^2,
\end{equation}
where $\omega$ and $q_\perp$ are energy and transverse momentum of the
quasireal photon. This limits the maximum energy of the quasireal
photon to
\begin{equation}
\omega<\omega_{max} \approx \frac{\gamma}{R},
\label{eq_wmax}
\end{equation}
where $\gamma$ is the Lorentz factor of the projectile
and the perpendicular component of its momentum to
\begin{equation}
q_\perp \mathrel{\mathpalette\vereq<} \frac{1}{R}.
\end{equation}
We define the
ratio $x=\omega/E$, where $E$ denotes
the energy of the nucleus $E= M_N \gamma A$ and $M_N$ is the nucleon mass.
It is therefore smaller than
\begin{equation}
x< x_{max}=\frac{1}{R M_N A} = \frac{\lambda_C(A)}{R},
\end{equation}
where $\lambda_C(A)$ is the Compton wave length of the ion. Here and
also throughout the rest of the paper we use natural units, setting
$\hbar=c=1$.
\begin{figure}[tbh]
\begin{center}
\resizebox{4.5cm}{!}{\includegraphics{PS/nuclx.eps}}
\end{center}
\caption{\it
A fast moving nucleus with charge $Ze$ is surrounded by a strong
electromagnetic field. This can be viewed as a cloud of virtual
photons. These photons can often be considered as real. They are
called equivalent or quasireal photons. The ratio of the photon energy
$\omega$ and the incident ion energy $E$ is denoted by $x=\omega/E$.
Its maximal value is restricted by the coherence condition to
$x<\lambda_C(A)/R\approx 0.175/A^{4/3}$, that is, $x\protect\mathrel{\mathpalette\vereq<}
10^{-3}$ for Ca ions and $x\protect\mathrel{\mathpalette\vereq<} 10^{-4}$ for Pb ions. }
\label{fig_xvar}
\end{figure}
The collisions of $e^+$ and $e^-$ has been the traditional way to
study ${\gamma\gamma}$-collisions. Similarly photon-photon collisions can also be
observed in hadron-hadron collisions. Since the photon number scales
with $Z^2$ ($Z$ being the charge number of the nucleus) such effects
can be particularly large. Of course, the strong interaction of the
two nuclei has to be taken into consideration.
\begin{figure}[tbh]
\begin{center}
\resizebox{5cm}{!}{\includegraphics{PS/aagamgam.eps}}
\end{center}
\label{fig_collision}
\caption{\it
Two fast moving electrically charged objects are an abundant source of
(quasireal) photons. They can collide with each other and with the
other nucleus. For peripheral collisions with impact parameters $b>2R$,
this is useful for photon-photon as well as photon-nucleus
collisions.}
\end{figure}
The equivalent photon flux present in medium and high energy nuclear
collisions is very high. Recent reviews of the present topic
can be found in \cite{BaurHT98,KraussGS97,BaurHT98b,HenckenSTB99}.
This high equivalent photon flux has already
found many useful applications in
nuclear physics \cite{BertulaniB88}, nuclear astrophysics
\cite{BaurR94,BaurR96}, particle physics \cite{Primakoff51} (sometimes
called the ``Primakoff effect''),
as well as, atomic physics \cite{Moshammer97}.
Here our main purpose is to discuss the physics of photon-photon and
photon-hadron (nucleus) collisions in high energy heavy ion
collisions. The ``Relativistic Heavy Ion Collider'' (RHIC),
scheduled to begin taking data in 1999, will have a program to investigate
such collisions experimentally. The equivalent photon spectrum there
extends up to several GeV ($\gamma\approx 100$). Therefore the available
invariant mass range is up to about the mass of the $\eta_c$.
At the recent RHIC/INT (``Institute for Nuclear Theory'') workshop at
the LBNL (Berkeley), the physics of peripheral collisions was discussed
by S. R. Klein and S. J. Brodsky \cite{rhicint99}.
When the ``Large Hadron Collider'' will be scheduled to begin taking data
in 2004/2008, the study of these reactions can be extended to both
higher luminosities but also to much higher invariant masses, hithero
unexplored.
Relativistic heavy ion collisions have been suggested as a general
tool for two photon physics about a decade ago. Yet the study of a
special case, the production of $e^+e^-$ pairs in nucleus-nucleus
collisions, goes back to the work of Landau and Lifschitz in 1934
\cite{LandauL34} (In those days, of course, one thought more about
high energy cosmic ray nuclei than relativistic heavy ion
colliders).
The general possibilities and characteristic features of two-photon
physics in relativistic heavy ion collisions have been discussed in
\cite{BaurB88}. The possibility to produce a Higgs boson via
${\gamma\gamma}$-fusion was suggested in \cite{GrabiakMG89,Papageorgiu89}. In
these papers the effect of strong absorption in heavy ion collisions
was not taken into account. This absorption is a feature, which is
quite different from the two-photon physics at $e^+e^-$ colliders. The
problem of taking strong interactions into account was solved by using
impact parameter space methods in \cite{Baur90d,BaurF90,CahnJ90}. Thus
the calculation of ${\gamma\gamma}$-luminosities in heavy ion collisions is put
on a firm basis and rather definite conclusions were reached by many
groups working in the field, as described, e.g., in
\cite{VidovicGB93,KraussGS97,BaurHT98}. This opens the way
for many interesting applications.
Up to now hadron-hadron collisions have not been used for
two-photon physics. An exception can be found in
\cite{Vannucci80}, where the production of $\mu^+\mu^-$ pairs at the
ISR was studied. The special class of events was selected, where no
hadrons are seen associated with the muon pair in a large solid angle
vertex detector. In this way one makes sure that the hadrons do not
interact strongly with each other, i.e., one is dealing with
peripheral collisions (with impact parameters $b>2R$); the
photon-photon collisions manifest themselves as ``silent events'',
that is, with only a small relatively small multiplicity.
Dimuons with a very low sum of transverse momenta are also considered
as a luminosity monitor for the ATLAS detector at LHC \cite{ShamovT98}.
Experiments are planned at RHIC
\cite{KleinS97a,KleinS97b,KleinS95a,KleinS95b,Nystrand98} and are
discussed at LHC
\cite{HenckenKKS96,Sadovsky93,BaurHTS98}.
We quote J. D. Bjorken \cite{Bjorken99}:
{\it It is an important portion (of the FELIX program at LHC
\cite{Felix97}) to tag on
Weizsaecker Williams photons (via the nonobservation of completely
undissociated forward ions) in ion-ion running, creating a high
luminosity ${\gamma\gamma}$ collider.}
\subsection{From impact-parameter dependent equivalent photon spectra to
{${\gamma\gamma}$-luminosities}}
\label{ssec_lum}
Photon-photon collisions have been studied extensively at $e^+e^-$
colliders. The theoretical framework is reviewed, e.g., in
\cite{BudnevGM75}. The basic graph for the two-photon process in
ion-ion collisions is shown in Fig.~\ref{fig_ggcollision}. Two virtual
(space-like) photons collide to form a final state $f$. In the
equivalent photon approximation (EPA) it is assumed that the square of the
4-momentum of the virtual photons is small, i.e., $q_1^2\approx
q_2^2\approx 0$ and the photons can be treated as quasireal. In this
case the ${\gamma\gamma}$-production is factorized into an elementary cross
section for the process $\gamma+\gamma\rightarrow f$ (with real photons, i.e.,
$q^2=0$) and a ${\gamma\gamma}$-luminosity function. In contrast to the pointlike
elementary electrons (positrons), nuclei are extended, strongly
interacting objects with internal structure. This gives rise to
modifications in the theoretical treatment of two photon processes.
The emission of a photon depends on the (elastic) form factor. Often a
Gaussian form factor or one of a homogeneous charged sphere is used.
The typical behavior of a form factor is
\begin{equation}
f(q^2) \approx
\left\{
\begin{array}{lcl}
Z &\qquad& \mbox{for $|q^2| < \frac{1}{R^2}$}\\
0 &\qquad& \mbox{for $|q^2| \gg \frac{1}{R^2}$}
\end{array}
\right. .
\end{equation}
For low $|q^2|$ all the protons inside the nucleus act coherently,
whereas for $|q^2| \gg 1/R^2$ the form factor is very small, close to
0. For a medium size nucleus with, say, $R=5$ fm, the limiting
$Q^2=-q^2=1/R^2$ is given by $Q^2=(40$MeV$)^2=1.6\times
10^{-3}$~GeV${}^2$. Apart from $e^+e^-$ (and to a certain extent also
$\mu^+\mu^-$) pair production, this scale is much smaller than typical
scales in the two-photon processes. Therefore the virtual photons in
relativistic heavy ion collisions can be treated as quasireal. This is
a limitation as compared to $e^+e^-$ collisions, where the two-photon
processes can also be studied as a function of the corresponding
masses $q_1^2$ and $q_2^2$ of the exchanged photon (``tagged mode'').
\begin{figure}[tbhp]
\begin{center}
\resizebox{3cm}{!}{\includegraphics{PS/feynmgg.eps}}
\end{center}
\caption{\it
The general Feynman diagram of photon-photon processes in heavy ion
collisions: Two (virtual) photons fuse in a charged particle collision
into a final system $f$.
}
\label{fig_ggcollision}
\end{figure}
As was discussed already in the previous section, relativistic
heavy ions interact strongly when the impact parameter is smaller than
the sum of the radii of the two nuclei. In such cases ${\gamma\gamma}$-processes
are still present and are a background that has to be considered in
central collisions. In order to study ``clean'' photon-photon events
however, they have to be eliminated in the calculation of
photon-photon luminosities as the particle production due to the
strong interaction dominates. In the usual treatment of photon-photon
processes in $e^+e^-$ collisions plane waves are used and there is
no direct information on the impact parameter. For heavy ion
collisions on the other hand it is very appropriate to introduce
impact parameter dependent equivalent photon numbers. They have been
widely discussed in the literature, see, e.g.,
\cite{BertulaniB88,JacksonED,WintherA79}.
The equivalent photon spectrum corresponding to a point charge $Z e$,
moving with a velocity $v$ at impact parameter $b$ is given by
\begin{equation}
N(\omega,b) = \frac{Z^2\alpha}{\pi^2} \frac{1}{b^2}
\left(\frac{c}{v}\right)^2 x^2 \left[ K_1^2(x) + \frac{1}{\gamma^2}
K_0^2(x)\right],
\label{eq_nomegab}
\end{equation}
where $K_n(x)$ are the modified Bessel Functions (MacDonald
Functions) and $x=\frac{\omega b}{\gamma v}$.
Then one obtains the probability for a certain electromagnetic process
to occur in terms of the same process generated by an equivalent pulse
of light as
\begin{equation}
P(b) = \int \frac{d\omega}{\omega} N(\omega,b) \sigma_\gamma(\omega).
\end{equation}
Possible modifications of $N(\omega,b)$ due to an extended spherically
symmetric charge distribution are given in \cite{BaurF91}. It should
be noted that Eq.~(\ref{eq_nomegab}) also describes the equivalent
photon spectrum of an extended charge distribution, such as a nucleus,
as long as $b$ is larger than the extension of the object. This is due
to the fact that the electric field of a spherically symmetric system
depends only on the total charge, which is inside it.
As the term $x^2 \left[ K_1^2(x) + 1/\gamma^2 K_0^2(x)\right]$ in
Eq.~(\ref{eq_nomegab}) can be roughly approximated as 1 for $x<1$ and
0 for $x>1$, that is, the equivalent photon number
$N(\omega,b)$ is almost a constant up to a maximum
$\omega_{max}=\gamma/b$ ($x=1$). By integrating
the photon spectrum (Eq.~(\ref{eq_nomegab}))
over $b$ from a minimum value of $R_{min}$ up to infinity (where
essentially only impact parameter up to $b_{max}\approx \gamma/\omega$
contribute, compare with Eq.~(\ref{eq_wmax})),
one can define an equivalent photon number
$n(\omega)$. This integral can be carried out analytically and is
given by \cite{BertulaniB88,JacksonED}
\begin{equation}
n(\omega) = \int d^2b N(\omega,b) = \frac{2}{\pi} Z_1^2 \alpha
\left(\frac{c}{v}\right)^2 \left[ \xi K_0 K_1 - \frac{v^2\xi^2}{2 c^2}
\left(K_1^2 - K_0^2\right)\right] ,
\label{eq_nomegaex}
\end{equation}
where the argument of the modified Bessel functions is
$\xi=\frac{\omega R_{min}}{\gamma v}$.
The cross section for a certain electromagnetic process is then
\begin{equation}
\sigma = \int \frac{d\omega}{\omega} n(\omega) \sigma_{\gamma}(\omega).
\label{eq_sigmac}
\end{equation}
Using the approximation above for the MacDonald functions, we get
an approximated form, which is quite reasonable and is useful for
estimates:
\begin{equation}
n(\omega) \approx \frac{2 Z^2 \alpha}{\pi} \ln
\frac{\gamma}{\omega R_{min}} \qquad \omega<\gamma/R_{min}.
\label{eq_nomegaapprox}
\end{equation}
The photon-photon production cross-section is obtained in a similar
factorized form,
by folding the corresponding equivalent photon spectra of the two
colliding heavy ions \cite{BaurF90,CahnJ90} (for
polarization effects see \cite{BaurF90}, they are neglected here)
\begin{equation}
\sigma_c = \int \frac{d\omega_1}{\omega_1} \int
\frac{d\omega_2}{\omega_2}
F(\omega_1,\omega_2) \sigma_{{\gamma\gamma}}(W_{{\gamma\gamma}}) ,
\label{eq_sigmaAA}
\end{equation}
with
\begin{eqnarray}
F(\omega_1,\omega_2)&=& 2\pi \int_{R_1}^{\infty} b_1 db_1
\int_{R_2}^{\infty} b_2 db_2 \int_0^{2\pi} d\phi \nonumber\\
&&\times N(\omega_1,b_1) N(\omega_2,b_2)
\Theta\left(b_1^2+b_2^2-2 b_1 b_2
\cos\phi-R_{cutoff}^2\right) ,
\label{eq_fw1w2}
\end{eqnarray}
where $W_{{\gamma\gamma}}=\sqrt{4 \omega_1\omega_2}$ is the invariant mass of the
${\gamma\gamma}$-system and $R_{cutoff} = R_1 + R_2$.
(In \cite{Nystrand98} the effect of replacing the simple sharp cutoff
($\Theta$-function) by a more realistic probability of the nucleus to
survive is studied. Apart from the very high end of the spectrum,
modifications are rather small.)
This can also be rewritten in terms of
the invariant mass $W_{{\gamma\gamma}}$ and the rapidity
$Y=1/2 \ln[(P_0+P_z)/(P_0-P_z)]=1/2 \ln(\omega_1/\omega_2)$ as:
\begin{equation}
\sigma_c = \int dW_{{\gamma\gamma}} dY \frac{d^2L}{dW_{{\gamma\gamma}} dY}
\sigma_{{\gamma\gamma}}(W_{{\gamma\gamma}}) ,
\label{eq_sigmaAAMY}
\end{equation}
with
\begin{equation}
\frac{d^2L_{{\gamma\gamma}}}{dW_{{\gamma\gamma}} dY} = \frac{2}{W_{{\gamma\gamma}}}
F\left(\frac{W_{{\gamma\gamma}}}{2} e^Y,\frac{W_{{\gamma\gamma}}}{2} e^{-Y}\right) .
\label{eq_dldwdy}
\end{equation}
Here energy and momentum of the ${\gamma\gamma}$-system in the beam direction are
denoted by $P_0$ and $P_z$. The transverse momentum is of the order of
$P_\perp \le 1/R$ and is neglected here. The transverse momentum
distribution is calculated in \cite{BaurB93}.
In \cite{BaurB93} and \cite{Baur92}
the intuitively plausible formula Eq.~(\ref{eq_fw1w2}) is derived ab
initio, starting from the assumption that the two ions move on a
straight line with impact parameter $b$.
The advantage of heavy nuclei is seen in the coherence factor $Z_1^2
Z_2^2$ contained in the $N(\omega,b)$ in Eq.~(\ref{eq_fw1w2}).
As a function of $Y$, the luminosity $d^2L/dW_{{\gamma\gamma}}{dY}$ for symmetrical
ion collisions has a
Gaussian shape with the maximum at $Y=0$. The width is approximately
given by $\Delta Y = 2 \ln \left[(2\gamma)/(R W_{{\gamma\gamma}})\right]$, see also
Fig.~\ref{fig_y}. Depending on the experimental situation additional cuts
in the allowed $Y$ range are needed.
\begin{figure}[tbh]
\begin{center}
\resizebox{8cm}{!}{\includegraphics{PS/ydep.eps}}
\end{center}
\caption{
The luminosity function $d^2L_{{\gamma\gamma}}/dMdY$ for Pb-Pb collisions with
$\gamma=2950$ as a function of $Y$ for different values of $M$.
}
\label{fig_y}
\end{figure}
Additional effects due to the nuclear structure have been also studied.
For inelastic vertices a photon number $N(\omega,b)$ can also be defined,
see, e.g., \cite{BaurHT98}. Its effect was found to be small. The
dominant correction comes from the electromagnetic excitation of one
of the ions in addition to the photon emission.
We refer to \cite{BaurHT98} for further details.
In Fig.~\ref{fig_lum}
we give a comparison of effective ${\gamma\gamma}$ luminosities, that is the product
of the beam luminosity with the two-photon luminosity
($L_AA \times dL_{{\gamma\gamma}}/dM$) for various collider scenarios.
We use the following collider parameters: LEP200: $E_{el}=100$GeV,
$L=10^{32} cm^{-2} s^{-1}$, NLC/PLC: $E_{el}=500$GeV,
$L=2 \times 10^{33} cm^{-2} s^{-1}$, Pb-Pb heavy-ion mode at LHC:
$\gamma=2950$,
$L=10^{26} cm^{-2} s^{-1}$, Ca-Ca: $\gamma=3750$,
$L=4 \times 10^{30} cm^{-2} s^{-1}$,p-p: $\gamma=7450$,
$L=10^{30} cm^{-2} s^{-1}$.
In the Ca-Ca heavy ion mode, higher
effective luminosities (defined as collider luminosity times
${\gamma\gamma}$-luminosity) can be achieved as, e.g., in the Pb-Pb mode,
since higher AA luminosities can be reached there.
Since the event rates are proportional to the luminosities, and
interesting events are rare (see also below), we think that it is important
to aim at rather high luminosities in the ion-ion runs.
This should be possible,especially for the medium heavy ions like Ca.
For further details see \cite{BrandtEM94,Bruening98,HenckenTB95}.
\begin{figure}[tbh]
\begin{center}
\resizebox{8cm}{!}{\includegraphics{PS/lumcmp.eps}}
\end{center}
\caption{\it Comparison of the effective ${\gamma\gamma}$-Luminosities
($d\tilde L_{{\gamma\gamma}}/dM = L_{AA} \times dL_{{\gamma\gamma}}/dM$) for different ion
species. For comparison the same
quantity is shown for LEP200 and a future NLC/PLC (next linear
collider/photon linear collider), where photons are obtained by laser
backscattering; the results for two different polarizations are shown.}
\label{fig_lum}
\end{figure}
\subsection{$\gamma$-A interactions}
\label{ssec_ga}
There are many interesting phenomena ranging from the excitation of
discrete nuclear states, giant multipole resonances (especially the
giant dipole resonance), quasideuteron absorption, nucleon resonance
excitation to the nucleon continuum.
The interaction of quasireal photons with protons has been studied
extensively at the electron-proton collider HERA (DESY, Hamburg), with
$\sqrt{s} = 300$~GeV ($E_e=27.5$~GeV and $E_p=820$~GeV in the
laboratory system). This is made possible by the large flux of
quasi-real photons from the electron (positron) beam. The obtained $\gamma
p$ center-of-mass energies (up to $W_{\gamma p}\approx200$~GeV) are an
order of magnitude larger than those reached by fixed target
experiments.
Similar and more detailed studies will be possible at the
relativistic heavy ion colliders RHIC and LHC, due to the larger flux
of quasireal photons from one of the colliding nuclei. In the
photon-nucleon subsystem, one can reach invariant masses $W_{\gamma N}$ up
to $W_{\gamma N,max}=\sqrt{4 W_{max} E_N} \approx 0.8 \gamma A^{-1/6}$~GeV.For
Pb at LHC ($\gamma=2950$) one obtains 950~GeV and even higher
values for Ca. Thus one can study
physics quite similar to the one at HERA, with nuclei instead of
protons. Photon-nucleon physics includes many aspects, like the energy
dependence of total cross-sections, diffractive and non-diffractive
processes.
An important subject is the elastic vector meson production $\gamma p
\rightarrow V p$ (with $V=\rho,\omega,\phi,J/\Psi,\dots$). A review of
exclusive neutral vector meson production is given in
\cite{Crittenden97}. The diffractive production of vector mesons
allows one to get insight into the interface between perturbative QCD
and hadronic physics. Elastic processes (i.e., the proton remains in
the ground state) have to be described within nonperturbative (and
therefore phenomenological) models. It was shown in \cite{RyskinRML97}
that diffractive (``elastic'') $J/\Psi$ photoproduction is a probe of
the gluon density at $x\approx \frac{M_{\Psi}^2}{W_{\gamma N}^2}$ (for
quasireal photons). Inelastic $J/\Psi$ photoproduction was also
studied recently at HERA \cite{Breitweg97}.
Going to the hard exclusive photoproduction of heavy mesons on the
other hand, perturbative QCD is applicable. Recent data from HERA on
the photoproduction of $J/\Psi$ mesons have shown a rapid increase of
the total cross section with $W_{\gamma N}$, as predicted by perturbative
QCD. Such studies could be extended to photon-nucleus interactions at
RHIC, thus complementing the HERA studies. Equivalent photon flux
factors are large for the heavy ions due to coherence. On the other
hand, the A-A luminosities are quite low, as compared to HERA. Of
special interest is the coupling of the photon of one nucleus to the
Pomeron-field of the other nucleus. Such studies are envisaged for
RHIC, see \cite{KleinS97a,KleinS97b,KleinS95a,KleinS95b} where also
experimental feasibility studies were performed.
Estimates of the order of magnitude of vector meson production in
photon-nucleon processes at RHIC and LHC are given in \cite{BaurHT98}. In
$AA$ collisions there is incoherent photoproduction on the individual
$A$ nucleons. Shadowing effects will occur in the nuclear environment
and it will be interesting to study these \cite{BauerSYP78}. There is
also the coherent contribution where the nucleus remains in the ground
state. Due to the large momentum transfer, the total (angle
integrated) coherent scattering shows an $A^{4/3}$ dependence. (It
will be interesting to study shadow effects in this case also). This
is in contrast to, e.g., low energy $\nu$A elastic scattering, where
the coherence effect leads to an $A^2$ dependence.
For a general pedagogical discussion of the coherence effects see, e.g.,
\cite{FreedmanST77}.
The coherent exclusive vector meson production at RHIC was studied
recently in \cite{KleinN99}. The increase of the cross section
with $A$ was found there to be between the two extremes ($A^{4/3}$ and
$A^2$) mentioned above. In this context, RHIC and LHC can be considered
as vector meson factories \cite{KleinN99}. In addition there are
inelastic contributions, where the proton (nucleon) is transformed into
some final state $X$ during the interaction (see \cite{Breitweg97}).
At the LHC one can extend these processes to much higher invariant
masses $W$, therefore much smaller values of $x$ will be probed.
Whereas the $J/\Psi$ production at HERA was measured up to invariant
masses of $W\approx 160$~GeV, the energies at the LHC allow for
studies up to $\approx 1$~TeV.
At the LHC \cite{Felix97} hard diffractive vector
meson photoproduction can be investigated especially well in $AA$
collisions. In comparison to previous experiments, the very large
photon luminosity should allow observation of processes with quite
small $\gamma p$ cross sections, such as $\Upsilon$-production. For more
details see \cite{Felix97}.
Photo-induced processes are also of practical importance as they are
a serious source of beam loss as they lead in general to a change of the
charge-to-mass ratio of the nuclei. Especially the cross section for
the excitation of the giant dipole resonance, a collective mode of the
nucleus, is rather large for the heavy systems (of the order of 100b).
The cross section scales approximately with $Z^{10/3}$. The
contribution nucleon resonances (especially the $\Delta$ resonance)
has also been confirmed experimentally in fixed target experiments
with 60 and~200 GeV/A (heavy ions at CERN, ``electromagnetic
spallation'') \cite{BrechtmannH88a,BrechtmannH88b,PriceGW88}. For
details of these aspects, we refer the reader to
\cite{KraussGS97,VidovicGS93,BaltzRW96,BaurB89}, where scaling laws,
as well as detailed calculations for individual cases are given.
\subsection{Photon-Photon Physics at various invariant mass scales}
\label{ssec_ggphysics}
Up to now photon-photon scattering has been mainly studied at $e^+e^-$
colliders. Many reviews \cite{BudnevGM75,KolanoskiZ88,BergerW87}
as well as conference reports
\cite{Amiens80,SanDiego92,Sheffield95,Egmond97,Freiburg99} exist. The
traditional range of invariant masses has been the region of mesons,
ranging from $\pi^0$ ($m_{\pi^0}=135$~MeV) up to about $\eta_c$
($m_{\eta_c}=2980$~MeV). Recently the total ${\gamma\gamma}\rightarrow$~hadron
cross-section has been studied at LEP2 up to an invariant mass range
of about 70~GeV \cite{L3:97}. We are concerned here mainly with the
invariant mass region relevant for LHC (see the
${\gamma\gamma}$-luminosity figures below). Apart from the production of $e^+e^-$
(and $\mu^+\mu^-$) pairs, the photons can always be considered as
quasireal. The cross section for
virtual photons deviates from the one for real photons only for $Q^2$,
which are much larger then the coherence limit $Q^2\mathrel{\mathpalette\vereq<} 1/R^2$
(see also the discussion in \cite{BudnevGM75}). For real photons
general symmetry requirements restrict the possible final states, as
is well known from the Landau-Yang theorem. Especially
it is impossible to produce spin 1 final states. In $e^+e^-$
annihilation only states with $J^{PC}=1^{--}$ can be produced
directly. Two photon collisions give access to most of the $C=+1$
mesons.
In principle $C=-1$ vector mesons can be produced by the fusion of
three (or, less important, five, seven, \dots) equivalent
photons. This cross section scales with $Z^6$. But it is smaller than
the contribution coming from $\gamma$-A collisions, as discussed above,
even for nuclei with large $Z$ (see \cite{BaurHT98}).
The cross section for ${\gamma\gamma}$-production in a heavy ion collision
factorizes into a ${\gamma\gamma}$-luminosity function and a cross-section
$\sigma_{{\gamma\gamma}}(W_{{\gamma\gamma}})$ for the reaction of the (quasi)real photons
${\gamma\gamma} \rightarrow f$, where $f$ is any final state of interest (see
Eq.~(\ref{eq_sigmaAA}). When the final state is a narrow resonance,
the cross-section for its production in two-photon collisions is given
by
\begin{equation}
\sigma_{{\gamma\gamma}\rightarrow R}(M^2) =
8 \pi^2 (2 J_R+1) \Gamma_{{\gamma\gamma}}(R) \delta(M^2-M_R^2)/M_R ,
\label{eq_nres}
\end{equation}
where $J_R$, $M_R$ and $\Gamma_{{\gamma\gamma}}(R)$ are the spin, mass and
two-photon width of the resonance $R$. This makes it easy to calculate
the production cross-section $\sigma_{AA\rightarrow AA+R}$ of a
particle in terms of its basic properties.
In Fig.~\ref{fig_sigmagamma} the function $4\pi^2 dL_{{\gamma\gamma}}/dM /M^2$,
which is universal for a produced resonances, is plotted for various
systems. It can be directly used to calculate the cross-section for
the production of a resonance
$R$ with the formula
\begin{equation}
\sigma_{AA\rightarrow AA+R}(M) = (2 J_R +1) \Gamma_{{\gamma\gamma}} \frac{4 \pi^2
dL_{{\gamma\gamma}}/dM}{M^2} .
\label{eq_aar}
\end{equation}
\begin{figure}[tbhp]
\begin{center}
\resizebox{8cm}{!}{\includegraphics{PS/sigmagamma.eps}}
\end{center}
\caption{\it
The universal function $4\pi^2 dL_{{\gamma\gamma}}/dM_{{\gamma\gamma}} /M_{{\gamma\gamma}}^2$ is
plotted for different ion species at LHC. We use $R=1.2 A^{1/3}$~fm and
$\gamma=2950$, 3750 and 7000 for Pb-Pb, Ca-Ca and $p$-$p$, respectively.
}
\label{fig_sigmagamma}
\end{figure}
We will now give a general
discussion of possible photon-photon physics at relativistic heavy ion
colliders. Invariant masses up to several GeV can be reached at RHIC
and up to about 100 GeV at LHC.
We can divide our discussion into the following two main subsections:
Basic QCD phenomena in ${\gamma\gamma}$-collisions (covering the range of meson,
meson-pair production, etc.) and ${\gamma\gamma}$-collisions as a tool for new
physics, especially at very high invariant masses.
An interesting topic in itself is the $e^+$-$e^-$ pair
production. The fields are strong enough to produce multiple pairs in a
single collisions. A discussion of this subject together with calculations
within the semiclassical approximation can be found in
\cite{Baur90,HenckenTB95a,HenckenTB95b,AlscherHT97}
\subsection{Basic QCD phenomena in ${\gamma\gamma}$-collisions}
\label{ssec_basicqcd}
\subsubsection{Hadron spectroscopy: Light and heavy quark spectroscopy}
One may say that photon-photon collisions provide an independent view
of the meson and baryon spectroscopy. They provide powerful
information on both the flavor and spin/angular momentum internal
structure of the mesons. Much has already been done at
$e^+e^-$ colliders. Light quark spectroscopy is very
well possible at RHIC, benefiting from the high
${\gamma\gamma}$-luminosities. Detailed feasibility studies exist
\cite{KleinS97a,KleinS97b,KleinS95a,KleinS95b}. In these studies, ${\gamma\gamma}$
signals and backgrounds from grazing nuclear and beam gas collisions
were simulated with both the FRITIOF and VENUS Monte Carlo codes. The
narrow $p_\perp$-spectra of the ${\gamma\gamma}$-signals provide a good
discrimination against the background, see also the discussion of a
possible trigger in \ref{ssec_selecting} below. The possibilities to
produce these mesons at the LHC have been discussed in detail in the
FELIX LoI \cite{Felix97}. Also discussed there are how to isolate
${\gamma\gamma}$ events in ion-ion collisions and applications to basic QCD
phenomena like $C=+1$ meson production, vector meson pair production
and total hadronic ${\gamma\gamma}$ cross sections. Rates are given and possible
triggers are discussed. In addition photon-Pomeron and
Pomeron-Pomeron processes are discussed. The general conclusion is
that all these processes are very promising tools for vector meson
spectroscopy.
In particular the absence of meson production via ${\gamma\gamma}$-fusion is also
of great interest for glueball search. The two-photon width of a
resonance is a probe of the charge of its constituents, so the
magnitude of the two-photon coupling can serve to distinguish quark
dominated resonances from glue-dominated resonances (``glueballs'').
In ${\gamma\gamma}$-collisions, a glueball can only be produced via the
annihilation of a $q\bar q$ pair into a pair of gluons, whereas a
normal $q\bar q$-meson can be produced directly. Therefore we expect
the ratio for the production of a glueball $G$ compared to a normal
$q\bar q$ meson $M$ to be
\begin{equation}
\frac{\sigma({\gamma\gamma} \rightarrow M)}{\sigma({\gamma\gamma} \rightarrow G)}
=
\frac{\Gamma(M \rightarrow {\gamma\gamma})}{\Gamma(G \rightarrow {\gamma\gamma})}
\sim
\frac{1}{\alpha_s^2} ,
\end{equation}
where $\alpha_s$ is the strong interaction coupling constant. On the
other hand glueballs are most easily produced in a glue-rich
environment, for example, in radiative $J/\Psi$ decays, $J/\Psi
\rightarrow \gamma gg$. In this process we expect the ratio of the
cross section to be
\begin{equation}
\frac{\Gamma(J/\Psi \rightarrow \gamma G)}{\Gamma(J/\Psi \rightarrow \gamma
M)} \sim \frac{1}{\alpha_s^2} .
\end{equation}
A useful quantity to describe the gluonic character of a mesonic state
X is therefore the so called ``stickiness'' \cite{Cartwright98}, defined as
\begin{equation}
S_X = \frac{\Gamma(J/\Psi \rightarrow \gamma X)}{\Gamma(X \rightarrow
\gamma \gamma)} .
\end{equation}
One expects the stickiness of all mesons to be comparable, while for
glueballs it should be enhanced by a factor of about $ 1/\alpha_s^4 \sim 20$.
In a recent reference \cite{Godang97} results of the search for $f_J
(2220)$ production in two-photon interactions were presented. There a
very small upper limit for the product of $\Gamma_{{\gamma\gamma}} B_{K_sK_s}$
was given, where $B_{K_s K_s}$ denotes the branching fraction of
its decay into $K_s K_s$. From this it was concluded that this is a
strong evidence that the $f_J(2220)$ is a glueball.
For charmonium production, the two-photon width $\Gamma_{{\gamma\gamma}}$ of
$\eta_c$ (2960 MeV, $J^{PC} = 0^{-+}$) is known from experiment
\cite{PDG98}. But the two-photon widths of $P$-wave charmonium states
have been measured with only modest accuracy. Two photon widths of
$P$-wave charmonium states can be estimated following the PQCD
approach \cite{Bodwin92}. Similar predictions of the bottonia two
photons widths can be found in \cite{Kwong88,Consoli94}. For RHIC the
study of $\eta_c$ is a real challenge \cite{KleinS97b}; the
luminosities are falling and the branching ratios to experimentally
interesting channels are small.
In Table~\ref{tab_ggmeson} (adapted from table~2.6 of \cite{Felix97})
the two-photon production cross-sections
for $c\bar c$ and $b \bar b$ mesons in the rapidity range $|Y|<7$ are
given. Also given are the number of events in a $10^7$ sec run with
the ion luminosities of $4\times 10^{30}$cm${}^{-2}$s${}^{-1}$ for
Ca-Ca and $10^{26}$cm${}^{-2}$s${}^{-1}$ for Pb-Pb. Millions of
$C$-even charmonium states will be produced in coherent two-photon
processes during a standard $10^7$~sec heavy ion run at the LHC. The
detection efficiency of charmonium events has been estimated as 5\%
for the forward-backward FELIX geometry \cite{Felix97}, i.e., one can
expect detection of about $5\times 10^3$ charmonium events in Pb-Pb
and about $10^6$ events in Ca-Ca collisions. This is two to three
orders of magnitude higher than what is expected during five years of
LEP200 operation. Experiments with a well-equipped central detector
like CMS on the other hand should provide a much better
efficiency. Further details, also on experimental cuts, backgrounds and
the possibilities for the study of $C$-even bottonium states are given
in \cite{Felix97}.
\begin{table}[hbt]
\begin{center}
\begin{tabular}{|l|c|c|r|c|c|c|c|}
\hline
State & Mass, & $\Gamma_{{\gamma\gamma}}$ &
\multicolumn{2}{|c}{$\sigma (AA\to AA+X)$} &
\multicolumn{2}{|c|}{rates per $10^7$ sec}\\
\cline{4-5}
& MeV & keV & Pb-Pb & Ca-Ca
& Pb-Pb & Ca-Ca \\
\hline
~~~$\pi_0$ & 134 & $8\times10^{-3}$ & 46 mbarn& 210 $\mu$barn
& $4.6 \times 10^7$ & $8.4 \times 10^9$ \\
~~~$\eta$ & 547 & 0.46 & 20 mbarn& 100 $\mu$barn
& $2 \times 10^7 $ & $4.0 \times 10^9$ \\
~~~$\eta'$ & 958 & 4.2 & 25 mbarn & 130 $\mu$barn
& $2.5 \times 10^7$ & $5.2 \times 10^9$ \\
~~~$f_2(1270)$ & 1275 & 2.4 & 25 mbarn & 133 $\mu$barn
& $2.5 \times 10^7$ & $5.2 \times 10^9$ \\
~~~$a_2(1320)$ & 1318 & 1.0 & 9.2 mbarn& 49 $\mu$barn
& $9.2 \times 10^6$ & $2.0 \times 10^9$ \\
~~~$f_2'(1525)$ & 1525 & 0.1 & 540 $\mu$barn& 2.9 $\mu$barn
& $5.4 \times 10^5$ & $1.2 \times 10^8$ \\
~~~$\eta_c$ & 2981 & 7.5 & 360 $\mu$barn & 2.1 $\mu$barn
& $3.6 \times 10^5$ & $8.4 \times 10^7$ \\
~~~$\chi_{0c}$& 3415& 3.3& 180 $\mu$barn &1.0 $\mu$barn
& $1.8 \times 10^5$ & $4.0 \times 10^7$ \\
~~~$\chi_{2c}$& 3556 & 0.8 & 74 $\mu$barn & 0.44 $\mu$barn
& $7.4 \times 10^4$ & $1.8 \times 10^7$ \\
~~~$\eta_b$ & 9366 & 0.43 &450 nbarn & 3.1 nbarn
& 450 & $1.2 \times 10^4$ \\
~~~$\eta_{0b}$& 9860 & $2.5\times10^{-2}$ & 21 nbarn & 0.15 nbarn
& 21 & 6000 \\
~~~$\eta_{2b}$& 9913 & $6.7\times10^{-3}$ & 28 nbarn & 0.20 nbarn
& 28 & 8000 \\
\hline
\end{tabular}
\end{center}
\caption{\it
Mass, and ${\gamma\gamma}$-widths used to calculate the cross section for meson
production for Pb-Pb and Ca-Ca collisions at CMS.
Masses and widths are taken from \cite{PDG98} and \cite{Felix97}. The
beam luminosities used are $10^{26}$cm${}^{-2}$s${}^{-1}$ for Pb-Pb
and $4\times10^{30}$cm${}^{-2}$s${}^{-1}$ for Ca-Ca.}
\label{tab_ggmeson}
\end{table}
\subsubsection{Vector-meson pair production. Total hadronic
cross-section}
There are various mechanisms to produce hadrons in photon-photon
collisions. Photons can interact as point particles which produce
quark-antiquark pairs (jets), which subsequently hadronize. Often a
quantum fluctuation transforms the photon into a vector meson
($\rho$,$\omega$,$\phi$, \dots) (VMD component) opening up all the
possibilities of hadronic interactions . In hard scattering, the
structure of the photon can be resolved into quarks and
gluons. Leaving a spectator jet, the quarks and gluon contained in the
photon will take part in the interaction. It is of great interest to
study the relative amounts of these components and their properties.
\begin{figure}[tbhp]
\begin{center}
\resizebox{7.5cm}{!}{\includegraphics{PS/Baur16.eps}}
\end{center}
\caption{\it
Diagrams showing the contribution to the ${\gamma\gamma}\rightarrow$hadron
reaction: direct mechanism (a), vector meson dominance (b), single (c)
and double (d) resolved photons.
}
\label{fig_resolved}
\end{figure}
The L3 collaboration recently made a measurement of the total hadron
cross-section for photon-photon collisions in the interval $5 GeV <
W_{{\gamma\gamma}} < 75 GeV$ \cite{L3:97}. It was found that the ${\gamma\gamma}
\rightarrow$hadrons cross-section is consistent with the universal
Regge behavior of total hadronic cross-sections.
The production of vector meson pairs can well be studied at RHIC with
high statistics in the GeV region \cite{KleinS97a}. For the
possibilities at LHC, we refer the reader to \cite{Felix97} and
\cite{BaurHTS98}, where also experimental details and simulations are
described.
\subsection{${\gamma\gamma}$-collisions as a tool for new physics}
\label{ssec_newphysics}
The high flux of photons at relativistic heavy ion colliders offers
possibilities for the search of new physics. This includes the
discovery of the Higgs-boson in the ${\gamma\gamma}$-production channel or new
physics beyond the standard model, like supersymmetry or
compositeness.
Let us mention here the plans to build an $e^+e^-$ linear collider.
Such future linear colliders will be used for $e^+e^-$, $e\gamma$
and ${\gamma\gamma}$-collisions (PLC, photon linear collider).
The photons will be obtained by scattering of
laser photons (of eV energy) on high energy electrons ($\approx$ TeV
region) (see \cite{Telnov95}). Such photons in the TeV energy range
will be monochromatic and polarized. The physics program at such
future machines is discussed in \cite{ginzburg95}, it includes Higgs
boson and gauge boson physics and the discovery of new particles.
While the ${\gamma\gamma}$ invariant masses which will be reached at RHIC will
mainly be useful to explore QCD at lower energies, the ${\gamma\gamma}$ invariant
mass range at LHC --- up to about 100 GeV --- will open up new
possibilities.
A number of calculations have been made for a medium heavy standard
model Higgs \cite{DreesEZ89,MuellerS90,Papageorgiu95,Norbury90}. For
masses $m_H < 2 m_{W^\pm}$ the Higgs bosons decays dominantly into
$b\bar b$. Chances of finding the standard model
Higgs in this case are marginal \cite{BaurHTS98}.
An alternative scenario with a light Higgs boson was, e.g., given in
\cite{ChoudhuryK97} in the framework of the ``general two Higgs
doublet model''. Such a model allows for a very light particle in the
few GeV region. With a mass of 10~GeV, the ${\gamma\gamma}$-width is about 0.1
keV. The authors of
\cite{ChoudhuryK97} proposed to look for such a light neutral Higgs
boson at the proposed low energy ${\gamma\gamma}$-collider. We want to point out
that the LHC Ca-Ca heavy ion mode would also be very suitable for such
a search.
One can also speculate about new particles with strong coupling to the
${\gamma\gamma}$-channel. Large $\Gamma_{{\gamma\gamma}}$-widths will directly lead to large
${\gamma\gamma}$ production cross-sections. We quote the
papers \cite{Renard83,BaurFF84}. Since the ${\gamma\gamma}$-width of a resonance
is mainly proportional to the wave function at the origin, huge values
can be obtained for very tightly bound systems. Composite scalar
bosons at $W_{{\gamma\gamma}}\approx 50$~GeV are expected to have ${\gamma\gamma}$-widths of
several MeV \cite{Renard83,BaurFF84}. The search for such kind of
resonances in the ${\gamma\gamma}$-production channel will be possible at
LHC.
In Refs. \cite{DreesGN94,OhnemusWZ94} ${\gamma\gamma}$-processes at $pp$
colliders (LHC) are studied. It is observed there that non-strongly
interacting supersymmetric particles (sleptons, charginos,
neutralinos, and charged Higgs bosons) are difficult to detect in
hadronic collisions at the LHC. The Drell-Yan and gg-fusion mechanisms
yield low production rates for such particles. Therefore the
possibility of producing such particles in ${\gamma\gamma}$ interactions at
hadron colliders is examined. Since photons can be emitted from
protons which do not break up in the radiation process, clean events
can be generated which should compensate for the small number. In
\cite{DreesGN94} it was pointed out that at the high luminosity of
$L=10^{34}$cm${}^{-2}$s${}^{-1}$ at the LHC($pp$), one expects about
16 minimum bias events per bunch crossing. Even the elastic ${\gamma\gamma}$
events will therefore not be free of hadronic debris. Clean elastic
events will be detectable at luminosities below
$10^{33}$cm${}^{-2}$s${}^{-1}$. This danger of ``overlapping events''
has also to be checked for the heavy ion runs, but it will be much
reduced due to the lower luminosities.
Recent (unpublished) studies done for FELIX and ALICE show that the
chargino pair production can be detectable, if the lighest chargino
would have a mass below 60~GeV$/c^2$. Unfortunately recent chargino
mass limits set by LEP experiments already exclude the existence of
charginos on this mass range. Therefore the observation of
MSSM-particles in ${\gamma\gamma}$-interactions in heavy ion collisions seems to
be hard to achieve.
Similar considerations for new physics were also made in connection
with the planned $eA$ collider at DESY (Hamburg). Again, the coherent
field of a nucleus gives rise to a $Z^2$ factor in the cross-section
for photon-photon processes in $eA$ collisions \cite{KrawczykL95}.
\subsection{Dilepton production}
\label{ssec_leptons}
Electrons (positrons) and to some extent also muons have a special
status, which is due to their small mass. They are therefore produced
more easily than other heavier particles and in the case of $e^+e^-$
pair production also lead to new phenomena, like multiple pair
production. Due to their small mass and therefore large Compton wave
length (compared to the nuclear radius), the equivalent photon
approximation has to be modified when applied to them. For the
muon, with a Compton wavelength of about 2 fm, we expect the standard
equivalent photon approximation to be applicable, with only small
corrections. Both electrons and muons can be produced not only as free
particles but also into an atomic states bound to one of the ions, or
even as a bound state, positronium or muonium.
The special situation of the electron pairs can already be seen from
the formula for
the impact parameter dependent probability in lowest order. Using
the equivalent photon approximation one obtains~\cite{BertulaniB88}
\begin{equation}
P^{(1)}(b) \approx \frac{14}{9 \pi^2} \left(Z \alpha\right)^4
\frac{1}{m_e^2 b^2} \ln^2 \left( \frac{\gamma_{ion} \delta}{2 m_e b}\right) ,
\label{eq_pbapprox}
\end{equation}
where $\delta\approx 0.681$ and $\gamma_{ion}=2\gamma^2-1$ the Lorentz factor
in the target frame, one can see that at RHIC and LHC energies and for
impact parameters of the order of the Compton wave length $b\approx
1/m_e$, this probability exceeds one. Unitarity is restored by
considering the production of multiple pairs
\cite{Baur90,Baur90c,BestGS92,RhoadesBrownW91,HenckenTB95a}. To a good
approximation the multiple pair production can be described by a
Poisson distribution. The impact parameter dependent probability
needed in this Poisson distribution was calculated in lowest order in
\cite{HenckenTB95b,Guclu95}, the total cross section for the one-pair
production in \cite{Bottcher89}, for one and multiple pair production
in \cite{AlscherHT97}. Of course
the total cross section is dominated by the single pair production as
the main contribution to the cross section comes from very large
impact parameters $b$. On the other hand one can see that for impact
parameters $b$ of about $2R$ the number of electron-positron pairs
produced in each ion collision is about 5 (2) for LHC with $Z=82$
(RHIC with $Z=79$). This means that each photon-photon event ---
especially those at a high invariant mass --- which occur
predominantly at impact parameters close to $b \mathrel{\mathpalette\vereq>} 2 R$ --- is
accompanied by the production of several (low-energy) $e^+e^-$ pairs.
\begin{figure}[tbhp]
\begin{center}
\resizebox{8cm}{!}{\includegraphics{PS/epem_p3400.eps}}
\end{center}
\caption{\it
The impact parameter dependent probability to produce $N$
$e^+e^-$-pairs ($N=1,2,3,4$) in one collision is shown for the
LHC ($\gamma=2950$,Pb-Pb). Also shown is the
total probability to produce at least one $e^+e^-$-pair. One sees that at
small impact parameters multiple pair production dominates over
single pair production.
}
\label{fig_pbee1}
\end{figure}
As the total cross section for this process is huge (about 200~kbarn for
Pb-Pb at LHC), one has to take this process into account as a possible
background process. Most of the particles are produced at low
invariant masses (below 10 MeV) and into the very forward direction
(see Fig.~\ref{fig_eee}). High energetic electrons and positrons are
even more concentrated along the beam pipe, most of them therefore are
unobserved. On the other hand, a substantial amount of them is still
left at high energies, e.g., above 1~GeV. These QED pairs therefore
constitute a potential hazard for the detectors, see below in
Sec.~\ref{ssec_selecting}. One the other hand, they can also be useful
as a possible luminosity monitor, as discussed in \cite{Felix97,ShamovT98}.
\begin{figure}[tbh]
\begin{center}
\resizebox{7.5cm}{!}{\includegraphics{PS/eeenergy.eps}}
\resizebox{7.5cm}{!}{\includegraphics{PS/eetheta.eps}}
\end{center}
\caption{
Cross section for the $e^+e^-$ pair production as a function of the
energy (A) of either electron or positron and as a function of the
angle of the electron or positron with the beam axis (B).
Most pairs are produced with energies between 2--5 MeV and in the very
forward or backward direction.
}
\label{fig_eee}
\end{figure}
Differential production probabilities for ${\gamma\gamma}$-dileptons in central
relativistic heavy ion collisions are calculated using the equivalent
photon approximation and an
impact parameter formulation and compared to Drell-Yan and thermal
ones in \cite{Baur92,BaurB93b,Baur92b}. The very low $p_\perp$ values
and the angular distribution of the pairs give a handle for their
discrimination.
Higher order corrections, e.g., Coulomb corrections, have to be taken
into account for certain regions in the phase space.
A classical result for these higher-order
effects can be found in the Bethe-Heitler formula for the process
$Z+\gamma \rightarrow Z + e^+ + e^-$
\begin{equation}
\sigma = \frac{28}{9} Z^2 \alpha r_e^2 \left[ \ln \frac{2 \omega}{m_e} -
\frac{109}{42} - f(Z\alpha) \right] ,
\end{equation}
with the higher-order term given by
\begin{equation}
f(Z\alpha) = (Z\alpha)^2 \sum_{n=1}^{\infty} \frac{1}{n(n^2+(Z\alpha)^2)}
\end{equation}
and $r_e=\alpha/m_e$ is the classical electron radius. As far as total
cross sections are concerned the higher-order contributions tends to a
constant for $\omega \rightarrow \infty$.
A systematic way to take leading terms of higher order
effects into account in $e^+e^-$ pair production is pursued in
\cite{IvanovM97,Ivanov98} using Sudakov variables and the impact-factor
representation. They find a reduction of the single-pair
production cross section of the order of 10\%.
In contrast to this some papers have recently discussed
nonperturbative results using a light-cone approach
\cite{segevW97,EichmannRSW98,BaltzM98}. There it is found that the
single-pair production cross section is identical to the lowest order
result. A calculation of the change of multiple pair production
cross section due to such higher order effects
can be found in \cite{henckenTB99}.
\subsubsection{Equivalent Muons}
Up to now only the production of dileptons was considered, for which
the four-momentum $Q^2$ of the photons was less than about $1/R^2$
(coherent interactions). There is another class of processes, where
one of the interactions is coherent ($Q^2 \le 1/R^2$) and the other
one involves a deep inelastic interaction ($Q^2\gg 1/R^2$), see
Fig.~\ref{fig_dis}. These processes are readily described using the
equivalent electron-- (or muon--, or tau--) approximation, as given,
e.g., in \cite{ChenZ75,BaierFK73}. The equivalent photon can be
considered as containing muons as partons, that is, consisting in part
of an equivalent muon beam. The equivalent muon number is given by
\cite{ChenZ75}
\begin{equation}
f_{\mu/\gamma} ( \omega,x) = \frac{\alpha}{\pi}
\ln\left(\frac{\omega}{m_\mu}\right) \left( x^2 + (1-x)^2 \right),
\end{equation}
where $m_\mu$ denotes the muon mass. The muon energy $E_\mu$ is given by
$E_\mu=x \omega$, where $\omega$ is the energy of the equivalent photon.
This spectrum has to be folded with the equivalent photon spectrum given by
\begin{equation}
f_{\gamma/Z}(u) = \frac{2\alpha}{\pi} \frac{Z^2}{u} \ln\left( \frac{1}{u
m_A R}\right)
\end{equation}
for $u<u_{max}=\frac{1}{R m_A}$.
The deep inelastic lepton-nucleon scattering can now be calculated in
terms of the structure functions $F_1$ and $F_2$ of the nucleon.
The inclusive cross section for the deep-inelastic scattering of the
equivalent muons is therefore given by
\begin{equation}
\frac{d^2\sigma}{dE' d\Omega} = \int dx_1 f_{\mu/Z}(x_1)
\frac{d^2\sigma}{dE' d\Omega} (x_1)
\end{equation}
where $\frac{d^2\sigma}{dE' d\Omega} (x_1)$
can be calculated from the usual invariant variables in
deep inelastic lepton scattering (see, e.g., Eq. 35.2 of \cite{PDG96})
The lepton is scattered to an angle $\theta$ with an energy $E'$. The
equivalent muon spectrum of the heavy ion is obtained as
\begin{equation}
f_{\mu/Z} (x_1) = \int_{x_1}^{u_{max}} du f_{\gamma/Z} (u) f_{\mu/\gamma}(x_1/u).
\end{equation}
This expression can be calculated
analytically and work on this is in progress \cite{BaurHT99}.
\begin{figure}[tbhp]
\begin{center}
\resizebox{5cm}{!}{\includegraphics{PS/inelastic.eps}}
\end{center}
\caption{\it With $Q^2 < 1/R^2$ the photon is emitted coherently from
all ``partons'' inside the ion. For $Q^2 \gg 1/R^2$ the ``partonic''
structure of the ion is resolved.}
\label{fig_dis}
\end{figure}
Such events are characterized by a single muon with an energy $E'$ and
scattering angle $\theta$. The accompanying muon of opposite charge,
as well as the remnants of the struck nucleus, will scatter to small
angles and remain unobserved. The hadrons scattered to large angles
can be observed, with total energy $E_h$ and momentum in the beam
direction of $p_{zh}$. Using the Jacquet-Blondel variable $y_{JB}$ the
energy of the equivalent muon can in principle be reconstructed as
\begin{equation}
E_\mu = \frac{1}{2} \left( E_h - p_{zh} + E' (1-\cos\theta)\right)
\end{equation}
This is quite similar to the situation at HERA, with the difference
that the energy of the lepton beam is continuous, and its energy has
to be reconstructed from the kinematics (How well this can be done in
practice remains to be seen).
\subsubsection{Radiation from $e^+e^-$ pairs}
The bremsstrahlung in peripheral relativistic heavy ion collisions was
found to be small, both for real \cite{BertulaniB88} and virtual
\cite{MeierHTB98} bremsstrahlung photons. This is due to the large
mass of the heavy ions.
Since the cross section for $e^+e^-$ pair production is so large, one
can expect to see sizeable effects from the radiation of these light
mass particles. In the soft photon limit (see, e.g., \cite{Weinberg97}) one
can calculate the cross section for soft
photon emission of the process as
\begin{equation}
Z + Z \rightarrow Z + Z + e^+ + e^- + \gamma
\end{equation}
as
\begin{equation}
d\sigma(k,p_-,p_+) =
- e^2 \left[ \frac{p_-}{p_- k} - \frac{p_+}{p_+ k} \right]^2
\frac{d^3k}{4 \pi^2 \omega} d\sigma_0(p_+,p_-)
\end{equation}
where $d\sigma_0$ denotes the cross section for the $e^+ e^-$ pair
production in heavy ion collisions. An alternative approach is done
by using the equivalent photon approximation (EPA) and
calculating the exact lowest order matrix element for the process
$$
\gamma + \gamma \rightarrow e^+ + e^- + \gamma.
$$
In Fig.~\ref{fig_brems} we show results of calculations for low energy
photons. For this we have used the exact lowest order QED process in the
equivalent photon approximation \cite{HenckenTB99b}.
\begin{figure}[tbhp]
\begin{center}
\resizebox{8cm}{!}{\includegraphics{PS/bremsstrahlung.eps}}
\end{center}
\caption{\it The energy-dependence of bremsstrahlungs-photons from
$e^+ e^-$ pair production is shown for different angles. We show
results for Pb-Pb collisions at LHC.}
\label{fig_brems}
\end{figure}
These low energy photons might constitute a background for the
detectors. Unlike the low energy electrons and positrons, they are of
course not bent away by the magnets. The angular distribution of the
photons also peak at small angles, but again a substantial amount is
still left at larger angles, even at $90^o$. The typical energy of
these low energy photons is of the order of several MeV, i. e., much
smaller than the expected level of the energy equivalent noise in the
CMS ECALs \cite{CMSTechProp}.
\subsubsection{Bound-free Pair Production}
The bound-free pair production, also known as electron-pair production
with capture, is a process, which is also of practical importance in
the collider. It is the process, where a pair is produced but with the
electron not as a free particle, but into an atomic bound state of one
of the nuclei. As this changes the charge state of the nucleus, it is
lost from the beam. Together with the electromagnetic dissociation of
the nuclei (see Sec.~\ref{ssec_ga}) these two processes are the
dominant loss processes for heavy ion colliders.
In \cite{BertulaniB88} an approximate value for this cross section is
given as
\begin{equation}
\sigma_{capt}^K \approx \frac{33\pi}{10} Z_1^2 Z_2^6 \alpha^6 r_e^2
\frac{1}{\exp(2\pi Z_2 \alpha) -1} \left[ \ln\left(\gamma_{ion} \delta / 2\right) -
\frac{5}{3}\right] ,
\label{eq_capture}
\end{equation}
where $\gamma_{ion}=2 \gamma^2 - 1$ is the Lorentz factor of the ion
in the rest frame of the other ion and only capture to the $K$-shell
is included. The cross section for all higher shells is expected to be
of the order of 20\% of this cross section (see Eqs 7.6.23 and 24 of
\cite{BertulaniB88}).
The cross section in Eq.~(\ref{eq_capture}) is of the form
\begin{equation}
\sigma= C \ln \gamma_{ion} + D.
\label{eq_lnAB}
\end{equation}
This form has been found to be a universal one at
sufficient high values of $\gamma$. The constant $C$ and $D$ then only
depend on the type of the target.
The above cross section was found making use of the equivalent photon
approximation (EPA) and also using
approximate wave function for bound state and continuum. More precise
calculations exist
\cite{BaltzRW92,BaltzRW93,BeckerGS87,AsteHT94,AggerS97,RhoadesBrownBS89}
in the literature. Recent calculations within DWBA for high values of
$\gamma$ have shown that the exact first order results do not differ
significantly from EPA results \cite{MeierHHT98,BertulaniB98}.
Parameterizations for $C$ and $D$\cite{BaltzRW93,AsteHT94} for typical
cases are given in Table~\ref{tab_capture}.
\begin{table}[tbhp]
\begin{center}
\begin{tabular}{|c|c|c|c|c|}
\hline
Ion & $C$ & $D$ & $\sigma(\gamma=106)$ & $\sigma(\gamma=2950)$ \\
\hline
Pb & $15.4$barn & $-39.0$barn & 115 barn & 222 barn \\
Au & $12.1$barn & $-30.7$barn & 90 barn & 173 barn \\
Ca & $1.95$mbarn & $-5.19$mbarn & 14 mbarn & 27.8 mbarn\\
O & $4.50\mu$barn & $-12.0\mu$barn & 32 $\mu$barn & 64.3 $\mu$barn \\
\hline
\end{tabular}
\end{center}
\caption{\it
Parameters $C$ and $D$ (see Eq.~(\ref{eq_lnAB}))
as well as total cross sections for the bound-free pair production for
RHIC and LHC. The parameters are taken from
\protect\cite{AsteHT94}.}
\label{tab_capture}
\end{table}
For a long time the effect of higher order and nonperturbative
processes have been under investigation. At lower beam energies, in the
region of few GeV per nucleon, coupled channel calculations have
indicated for a long time, that these give large contributions,
especially at small impact parameters. Newer calculation tend to
predict considerably smaller values, of the order of the first order
result and in a recent article Baltz \cite{Baltz97} finds in the limit
$\gamma\rightarrow \infty$ that contributions from higher orders are even
slightly smaller than the first order results.
The bound-free pair production was measured in two recent experiments
at the SPS, at fixed target $\gamma=168$ \cite{Krause98} and at fixed target
$\gamma \approx 2$ \cite{Belkacem93,Belkacem94}. Both experiments found good
agreement between measurement and calculations.
We note that electron and positron can also form a bound state,
positronium. This is in analogy to the ${\gamma\gamma}$-production of mesons
($q\bar q$ states) discussed in Sec.~\ref{ssec_ggphysics}.
With the known width of the parapositronium $\Gamma((e^+e^-)_{n=1}
{}^1S_0 \rightarrow {\gamma\gamma}) = m c^2 \alpha^5 /2$, the photon-photon
production of this bound state was calculated in \cite{Baur90b}. The
production of orthopositronium, $n=1 {}^3S_1$ was calculated recently
\cite{Ginzburg97}.
As discussed in Sec.~\ref{ssec_ggphysics} the production of
orthopositronium is only suppressed by the factor $(Z\alpha)^2$, which is
not very small. Therefore one expects that both kind of positronium
are produced in similar numbers. Detailed calculation show that the
three-photon process is indeed not much smaller than the two-photon
process \cite{Ginzburg97,Gevorkyan98}.
\subsection{Event rates at CMS}
\label{ssec_evrate}
An overview of the expected event rate for a number of different
photon-photon reactions to either discrete states or continuum states
is given in the following figures. The right hand axes shows both the
number of events per second and per one-year run time (assuming $10^7$
sec per year). We use beam luminosities of
$10^{26}$cm${}^{-2}$s${}^{-1}$ for Pb-Pb and
$4\times10^{30}$cm${}^{-2}$s${}^{-1}$ for Ca-Ca.
The resonances have been calculated using the masses and photon-decay
widths as given in table~\ref{tab_ggmeson}. For the calculation of
the rate for a standard model Higgs boson $H_{SM}$, we use the approach as
discussed in \cite{DreesEZ89}. $H'$ denotes a nonstandard Higgs as
given in the ``general two-Higgs doublet model'' in
\cite{ChoudhuryK97}. As its photon-photon decay width is rather weakly
dependent on its mass in the relevant mass region, we have used a constant
value of 0.1 keV in our calculations.
The total hadronic cross section $\sigma_{{\gamma\gamma}}($hadron$)$ was used in
the form \cite{L3:97}:
\begin{equation}
\sigma_{{\gamma\gamma}}(\mbox{hadron}) = A (s/s_0)^\epsilon + B (s/s_0)^{-\eta}
\end{equation}
with $s_0=1$GeV${}^2$, $\epsilon=0.079$, $\eta=0.4678$, $A=173$ nbarn,
$B=519$ nbarn. For the dilepton and $q\bar q$ production via ${\gamma\gamma}$,
we have used the lowest order QED expression for pointlike fermions.
For the quark masses we use $m_c=1.3$~GeV and $m_b=4.6$~GeV
\cite{DreesKZZ93} (In this reference QCD corrections are also given).
Of course these cross section only correspond to the ``direct
mechanism'' (see Figure~\ref{fig_resolved} above). In addition there
will be also events coming from resolved processes as well as vector
meson dominace \cite{SchulerS95,SchulerS96,SchulerS97}. This explains
the much larger total hadronic cross section compared to the cross
section dye to the ``direct mechanism''. These two-quark processes
will be visible as two-jet events.
\begin{figure}[tbhp]
\begin{center}
\resizebox{10cm}{!}{\includegraphics{PS/overview_CAR.eps}}
\end{center}
\caption{\it Overview of the total cross section and production rates
(both per second and per one year run, assuming 1 year = $10^7$ sec)
of different resonances in Ca-Ca collisions at the CMS. We have used
the parameters as given in the text and in
table~\protect\ref{tab_ggmeson}.}
\label{fig_caexres}
\end{figure}
\begin{figure}[tbhp]
\begin{center}
\resizebox{10cm}{!}{\includegraphics{PS/overview_CAC.eps}}
\end{center}
\caption{\it Overview of the total cross section and production rates
(both per second and per one year run, assuming 1 year = $10^7$ sec)
per GeV for different dilepton and
$q\bar q$ production for Ca-Ca collisions at CMS. Also shown is the total
hadronic cross section. The parameters used are given in the text.}
\label{fig_caexcnt}
\end{figure}
\begin{figure}[tbhp]
\begin{center}
\resizebox{10cm}{!}{\includegraphics{PS/overview_PBR.eps}}
\end{center}
\caption{\it Overview of the total cross section and production rates
(both per second and per one year run, assuming 1 year = $10^7$ sec)
of different resonances in Pb-Pb collisions at the CMS. We have used
the parameters as given in the text and in
table~\protect\ref{tab_ggmeson}.}
\label{fig_pbexres}
\end{figure}
\begin{figure}[tbhp]
\begin{center}
\resizebox{10cm}{!}{\includegraphics{PS/overview_PBC.eps}}
\end{center}
\caption{\it Overview of the total cross section and production rates
(both per second and per one year run, assuming 1 year = $10^7$ sec)
per GeV for different dilepton and
$q\bar q$ production for Pb-Pb collisions at CMS. Also shown is the total
hadronic cross section. The parameters used are given in the text.}
\label{fig_pbexcnt}
\end{figure}
\subsection{Selecting ${\gamma\gamma}$-events}
\label{ssec_selecting}
The ${\gamma\gamma}$-luminosities are rather large, but the ${\gamma\gamma} \rightarrow X$
cross sections are small compared to their hadronic counterparts,
therefore, e.g., the total hadronic production cross section for all
events is still dominated by hadronic events. This makes it necessary
to have an efficient trigger to distinguish photon-photon events from
hadronic ones.
There are some characteristic features that make such a trigger
possible. ${\gamma\gamma}$-events are characterized by the fact that both nuclei
remain intact after the interaction. Therefore a ${\gamma\gamma}$-event will have
the characteristic of a low multiplicity in the central region and no
event in the very forward or backward direction (corresponding to
fragments of the ions). The momentum transfer and energy loss for each
ion are too small for the ion to leave the beam.
It should be noted that in a ${\gamma\gamma}$ interaction with an invariant mass of
several GeV leading to hadronic final states, quite a few particles
will be produced, see , e.g., \cite{L3:97}
A second characteristic is the small transverse momenta of the
produced system due to the coherence condition $q_\perp < 1/R\approx
50$~MeV. If one is able to make a
complete reconstruction of the momenta of all produced particles with
sufficient accuracy, this
can be used as a very good suppression against grazing collisions. As
the strong interaction is short ranged, it has normally a much broader
distribution in the transverse momenta. A calculation using the PHOJET
event generator \cite{EngelRR97} to study processes in central and
grazing collisions by Pomeron-exchange found an average transverse
momentum of about 450 MeV, about a factor of 10 larger than the
${\gamma\gamma}$-events. In a study for the STAR experiment \cite{NystrandK97} it was
also found that triggering for small transverse momenta is an
efficient method to reduce the background coming from grazing
collisions.
Another question that has to be addressed is the importance of
diffractive events, that is, Photon-Pomeron and Pomeron-Pomeron
processes in ion collisions. From experiments at HERA one knows that
the proton has a large probability to survive intact after these
collisions. The theoretical situation unfortunately is not very clear
for these high energies and especially for nuclei as compared to
nucleons. Some calculations within the dual parton model have been
made and were interpreted as an indication that Photon-Pomeron and
Pomeron-Pomeron events are of the same size or even larger than
photon-photon events \cite{EngelRR97}. But these calculations were done
without requiring the condition to have intact nuclei in the final
state. As the nuclei are bound only rather weakly and as mentioned
above the average momentum transfer to the nucleus is of the order of
200 MeV, it is very likely that the nucleus will break up in such a
collision. First estimates based on this model indicate, that this
leads to a substantial suppression of diffractive events, favoring
again the photon-photon events.
The cross section ratio of photon-photon to Pomeron-Pomeron processes
depends on the ion species. Roughly it scales with $Z^4/A^{1/3}$, see
\cite{Felix97}. Thus for heavy ions, like Pb, we may expect dominance
of the photon-photon processes whereas, say in $pp$-collisions, the
Pomeron-Pomeron processes will definitely dominate in coherent
collisions.
Nevertheless diffractive events are of interest in ion collisions
too. As one is triggering again on an intact nucleus, one expects that
the coherent Pomeron emission from the whole nucleus will lead to a
total transverse momentum of the produced system similar to the
${\gamma\gamma}$-events. Therefore one expects that part of the events are coming
from diffractive processes. It is of interest to study how these could
be further distinguished from the photon-photon events.
Another class of background events are additional electromagnetic
processes. One of the dominating events here is the electromagnetic
excitation of the ions due to an additional single-photon exchange. As
mentioned above this is one of the dominant beam-loss processes for
Pb-Pb collisions. The probability to excite at least one of the ions
for Pb-Pb collisions is about 65\% and about 2\% for Ca-Ca for an impact
parameter of $2R$. Especially at large invariant masses, ${\gamma\gamma}$-events
occur at impact parameter close to $2R$, therefore in the case of Pb-Pb
collisions one has to expect that most of them are accompanied by the
excitation or dissociation of one of the ions \cite{HenckenTB95,BaltzS98}. Most
of the excitation lead into the giant dipole resonance (GDR), which
has almost all of the dipole strength. As it decays predominantly via
the emission of a neutron, this leads to a relativistic neutron with
an energy of about 3 TeV in the forward direction. Similarly all other
low energy breakup reactions in the rest frame of one of the ions are
boosted to high energy particles in the laboratory. In order to
increase the ${\gamma\gamma}$-luminosity it would be interesting to include these
events also in the ${\gamma\gamma}$-trigger. On the other hand one has
to make sure, that this does not obscure the interpretation of these
events as photon-photon events.
Another background process is the production of electron-positron
pairs, see Sec.~\ref{ssec_leptons}. Due to their small mass, they are
produced rather copiously. They are predominantly produced at low
invariant masses and energies and in the forward and backward
direction. Figure~\ref{fig_eedifT} shows cross section as a function
of energy and angle for different experimental cuts.
On the other hand, as the total cross section for
this process is enormous ($\approx$ 230 kbarn for Pb-Pb collisions, 800
barn for Ca-Ca collisions), a significant cross section remains even at
high energies in the forward direction. This has to be taken into
account when designing forward detectors. Table~\ref{tab_ee} shows the
cross section for $e^+e^-$ production where the energy of both
particles is above a certain threshold value.
\begin{figure}[tbhp]
\begin{center}
\resizebox{7.5cm}{!}{\includegraphics{PS/ginesE.eps}}
\resizebox{7.5cm}{!}{\includegraphics{PS/ginesT.eps}}
\end{center}
\caption{\it
The single differential cross section for a number of experimental
constraints. (a) for different angular-ranges as a function of energy,
(b) for different energies as a function of the angle
with the beam axis $\theta$.
}
\label{fig_eedifT}
\end{figure}
\begin{table}[htb]
\caption{
Cross sections of $e^+e^-$ pair production when {\em both} electron
and positron have an energy above a threshold value.}
\begin{center}
\begin{tabular}{c|rr}
\hline
$E_{thr}$ (GeV) & $\sigma$(Pb-Pb) & $\sigma$(Ca-Ca)\\
\hline
0.25& 3.5 kbarn & 12 barn \\
0.50& 1.5 kbarn & 5.5 barn \\
1.0 & 0.5 kbarn & 1.8 barn \\
2.5 & 0.08 kbarn & 0.3 barn \\
5.0 & 0.03 kbarn & 0.1 barn \\
\hline
\end{tabular}
\end{center}
\label{tab_ee}
\end{table}
\subsection{Conclusion}
\label{ssec_conclusion}
In this contribution to the CMS heavy ion chapter the basic properties
of peripheral hadron-hadron collisions are described. Electromagnetic
processes, that is, photon-photon and photon-hadron collisions, are an
interesting option, complementing the program for central collisions.
It is the study of events, with relatively small multiplicities and a
small background. These are good conditions to search for new physics.
The method of equivalent photons is a well established tool to
describe these kinds of reactions. Reliable results of quasireal
photon fluxes and ${\gamma\gamma}$-luminosities are available. Unlike electrons
and positrons heavy ions and protons are particles with an internal
structure. Effects arising from this structure are well under
control and minor uncertainties coming from the exclusion of central
collisions and triggering can be eliminated by using a luminosity
monitor from muon-- or electron--pairs. A trigger for peripheral
collisions is essential in order to select photon-photon events. Such
a trigger seems to be possible based on the survival of the nuclei
after the collision and the use of the small transverse momenta of the
produced system. A problem, which is difficult to judge quantitatively
at the moment, is the influence of strong interactions in grazing
collisions, i.e., effects arising from the nuclear stratosphere and
Pomeron interactions.
The high photon fluxes open up possibilities for photon-photon as well
as photon-nucleus interaction studies up to energies hitherto
unexplored at the forthcoming colliders RHIC and LHC. Interesting
physics can be explored at the high invariant ${\gamma\gamma}$-masses, where
detecting new particles could be within range. Also very interesting
studies within the standard model, i.e., mainly QCD studies will be
possible. This ranges from the study of the total ${\gamma\gamma}$-cross section
into hadronic final states up to invariant masses of about 100~GeV to
the spectroscopy of light and heavy mesons. The production via
photon-photon fusion complements the production from single photons in
$e^+$--$e^-$ collider and also in hadronic collisions via other
partonic processes.
Peripheral collisions using Photon-Pomeron and Pomeron-Pomeron
collisions, that is, diffractive processes are an additional
application. They use essentially the same triggering conditions and
therefore one should be able to record them at the same time as
photon-photon events.
|
\section{Introduction}
Various mechanical problems can be elegantly approached by the Hamiltonian
formalism, which not only found well-established ground in classical
theories\cite{one}, but also provided much physical insight in the early
development of quantum theories\cite{two,three}. It is curious though that
the concept of canonical transformations, which plays a fundamental role in
the Hamiltonian formulation of classical mechanics, has not attracted as
much attention in the corresponding formulation of quantum mechanics. A
relatively small quantity of literature is available as of now on
this subject [4--10].
The main reason for this is probably that canonical variables in quantum
mechanics are not c-numbers but noncommuting operators, manipulation of which
is considerably involved. In spite of this difficulty, the great success of
canonical transformations in classical mechanics makes it desirable to
investigate the possibility of application of the concept of canonical
transformations in quantum mechanics at least to the extent allowed in view
of the analogy with the classical case.
The usefulness of the classical canonical transformations is most visible in
the Hamilton-Jacobi theory where one seeks a generating function that makes
the transformed Hamiltonian become identically zero\cite{one}. A quantum
analog of the Hamilton-Jacobi theory has previously been considered by
Leacock and Padgett\cite{eight} with particular emphasis on the quantum
Hamilton's characteristic function and applied to the definition of the
quantum action variable and the determination of the bound-state energy
levels\cite{one2}. However, the {\em dynamical} aspect of the quantum
Hamilton-Jacobi theory appears to remain untouched. In the present study,
we concentrate on this aspect of the problem, and derive the time-dependent
quantum Hamilton-Jacobi equation following closely the procedure that lead
to the classical Hamilton-Jacobi equation.
The analogy between the classical and quantum Hamilton-Jacobi theories can
be best exploited by employing the idea of the quantum generating function
that was first introduced by Jordan\cite{four} and Dirac\cite{five}, and
recently reconsidered by Lee and l'Yi\cite{ten}. The ``well-ordered''
operator counterpart of the quantum generating function is used in
constructing our quantum Hamilton-Jacobi equation, which resembles in form
the classical Hamilton-Jacobi equation. By means of well-ordering, a unique
operator is associated with a given c-number function, thereby the ambiguity
in the ordering problem is removed. We identify the quantum generating
function accompanying the quantum Hamilton-Jacobi theory as the quantum
Hamilton's principal function, and apply this theory to find the dynamical
solutions of quantum problems.
The prevailing conventional belief that physical observables should be
Hermitian operators invokes in our discussion the unitary transformation
that transforms one Hermitian operator to another. This along with the
fact that the unitary transformation preserves the fundamental quantum
condition for the new canonical variables $[\hat{Q},\hat{P}]=i\hbar$ if
the old canonical variables satisfy $[\hat{q},\hat{p}]=i\hbar$ provides a
good reason why we call the unitary transformation the quantum canonical
transformation. This definition of the quantum canonical transformation is
analogous to the classical statement that the classical canonical
transformation keeps the Poisson brackets invariant, i.e.,
$[Q,P]_{PB}=[q,p]_{PB}=1$. In our current discussion of the quantum
canonical transformation we will consider exclusively the case of the
unitary transformation.
The paper is organized as follows. In Sec.\ II the quantum canonical
transformation using the idea of the quantum generating function is briefly
reviewed, and the transformation relation between the new Hamiltonian and
the old Hamiltonian expressed in terms of the quantum generating function is
derived. From this relation, and by analogy with the classical case, we
arrive at the quantum Hamilton-Jacobi equation in Sec.\ III. It will be found
that the unitary transformation of the special type
$\hat{U}(t)=\hat{T}(t)\hat{A}$ where $\hat{T}(t)$ is the time-evolution
operator and $\hat{A}$ is an arbitrary time-independent unitary operator
satisfies the quantum Hamilton-Jacobi equation.
Sec.\ IV is devoted to the discussion of
the quantum phase-space distribution function under canonical transformations.
The differences between our approach and that of Ref.\cite{one1} are described.
Boundary conditions and simple applications of the theory are given
in Sec.\ V, where to perceive the main idea easily most of the discussion
is developed with the simple case $\hat{A}=\hat{I}$, the unit operator, while
keeping in mind that the present formalism is not restricted to this case.
Finally, Sec.\ VI presents concluding remarks.
\section{Quantum Canonical Transformations}
Let us begin our discussion by reviewing the theory of the quantum canonical
transformations\cite{five,ten}. A quantum generating function that is
analogous to a classical generating function is defined in terms of the
matrix elements of a unitary operator as follows\cite{five},
\begin{equation}
e^{iF_1(q_1,Q_2,t)/\hbar}\equiv \langle q_1|Q_2\rangle _t
=\langle q_1|\hat{U}(t)|q_2\rangle ,
\end{equation}
where the unitary operator $\hat{U}(t)$ transforms an eigenvector of
$\hat{q}$ into an eigenvector of $\hat{Q}=\hat{U}\hat{q}\hat{U}^\dagger$,
i.e., $|Q_1\rangle _t=\hat{U}(t)|q_1\rangle$ (and $|P_1
\rangle _t=\hat{U}(t)|p_1\rangle$).\footnote{
An eigenvalue $X_1$ and an eigenvector $|X_1\rangle$ of an
operator $\hat{X}$ are defined by the equation, $\hat{X}|X_1\rangle
=X_1|X_1\rangle$ ($X=q$, $p$, $Q$, and $P$).
Different subindices are used to distinguish different eigenvalues or
eigenvectors, e.g., $X_2$, $|X_2\rangle$, etc.; The subscript $t$ on a
ket $|\rangle _t$ (bra ${_t\langle}|$) expresses time dependence of the
ket $|\rangle _t$ (bra ${_t\langle}|$).}
Different types of the quantum
generating function can be defined similarly\cite{ten}, i.e.,
$e^{iF_2(q_1,P_2,t)/\hbar}\equiv \langle q_1|P_2\rangle _t
=\langle q_1|\hat{U}(t)|p_2\rangle$,
$e^{iF_3(p_1,Q_2,t)/\hbar}\equiv \langle p_1|Q_2\rangle _t
=\langle p_1|\hat{U}(t)|q_2\rangle$, and
$e^{iF_4(p_1,P_2,t)/\hbar}\equiv \langle p_1|P_2\rangle _t
=\langle p_1|\hat{U}(t)|p_2\rangle$.
The quantum canonical transformation, or the unitary transformation,
corresponds to a change of representation or equivalently to a rotation
of axes in the Hilbert space. The unitary transformation guarantees that
the fundamental quantum condition $[\hat{Q},\hat{P}]
=[\hat{q},\hat{p}]=i\hbar$ holds, the new canonical variables
$(\hat{Q},\hat{P})$ are Hermitian operators, and the eigenvectors of
$\hat{Q}$ or $\hat{P}$ form a complete basis. One should keep in mind
that the eigenvalue $Q_1$ has the same numerical value as the eigenvalue
$q_1$ because the unitary transformation preserves the eigenvalue spectrum
of an operator\cite{three}. In cases where it is convenient, one is free to
interchange $q_1$ $ (p_1)$ with $Q_1$ $(P_1)$.
Transformation relations between $(\hat{q},\hat{p})$ and $(\hat{Q},\hat{P})$
can be expressed in terms of the ``well-ordered'' generating operator
$\bar{F}_1(\hat{q},\hat{Q},t)$\cite{five} that is an operator counterpart
of the quantum generating function $F_1(q_1,Q_2,t)$ as follows:\footnote{
A well-ordered operator $\bar{G}(\hat{X},\hat{Y})$ is
developed from a c-number function $G(X_1,Y_2)$ such that
$\langle X_1|\bar{G}(\hat{X},\hat{Y})|Y_2
\rangle = G(X_1,Y_2)\langle X_1|Y_2\rangle $\cite{five}. For example,
if $G(X_1,Y_2)=X_1Y_2+Y_2^2X_1^3$, then $\bar{G}(\hat{X},\hat{Y})=
\hat{X}\hat{Y}+\hat{X}^3\hat{Y}^2$.}
\begin{equation}
\hat{p}=\frac{\partial \bar{F}_1(\hat{q},\hat{Q},t)}{\partial \hat{q}},
\hspace{1cm}
\hat{P}=-\frac{\partial \bar{F}_1(\hat{q},\hat{Q},t)}{\partial \hat{Q}}.
\end{equation}
Similar expressions for other types of the generating operators can be
immediately inferred by analogy with the classical relations. For a later
reference, we present the relations for $\bar{F}_2(\hat{q},\hat{P},t)$ below,
\begin{equation}
\hat{p}=\frac{\partial \bar{F}_2(\hat{q},\hat{P},t)}{\partial \hat{q}},
\hspace{1cm}
\hat{Q}=\frac{\partial \bar{F}_2(\hat{q},\hat{P},t)}{\partial \hat{P}}.
\end{equation}
It is interesting to note that, whereas the four types of the generating
functions in classical mechanics are related with each other through the
Legendre transformations\cite{one}, the relations between the quantum
generating functions of different types can be expressed by means of the
Fourier transformations. For example, the transition from
$F_1(q_1,Q_2,t)$ to $F_2(q_1,P_2,t)$ can be accomplished by
\begin{eqnarray}
e^{iF_2(q_1,P_2,t)/\hbar}&=&\int dQ_2 \langle q_1|Q_2\rangle _t\hspace{0.7mm}
{_t\langle} Q_2|P_2\rangle _t, \nonumber \\
&=&\frac{1}{\sqrt{2\pi \hbar}} \int dQ_2 e^{iF_1(q_1,Q_2,t)/\hbar}
e^{iP_2Q_2/\hbar}.
\end{eqnarray}
The usefulness of the concept of the quantum generating function can be
revealed, for example, by considering the unitary transformation
$\hat{U} =e^{ig(\hat{q})/\hbar}$ where $g$ is an arbitrary real function.
From the definition of the quantum generating function, we have
\begin{eqnarray}
e^{iF_2(q_1,P_2)/\hbar}&=&\langle q_1|e^{ig(\hat{q})/\hbar}|p_2 \rangle ,
\nonumber \\
&=& \frac{1}{\sqrt{2\pi \hbar}}e^{\frac{i}{\hbar}[g(q_1)+q_1P_2]}.
\end{eqnarray}
The well-ordered generating operator is then given by
\begin{equation}
\bar{F}_2(\hat{q},\hat{P})=g(\hat{q})+\hat{q}\hat{P}+i\frac{\hbar}{2}
\ln 2\pi \hbar ,
\end{equation}
and Eq.\ (3) yields the transformation relations
\begin{eqnarray}
\hat{Q}&=& \hat{q}, \\
\hat{P}&=&\hat{p}-\frac{\partial g(\hat{q})}{\partial \hat{q}}.
\end{eqnarray}
This shows that, in some cases, an introduction of the quantum generating
function can provide an effective method of finding the transformation
relations between ($\hat{q},\hat{p}$) and ($\hat{Q},\hat{P}$) without
recourse to the equations $\hat{Q}=\hat{U}\hat{q}\hat{U}^\dagger$ and
$\hat{P}=\hat{U}\hat{p}\hat{U}^\dagger$.
Now we consider the dynamical equations governing the time-evolution of
quantum systems. The time-dependent Schr\"{o}dinger equation for the system
with the Hamiltonian $H(\hat{q},\hat{p},t)$ is given in terms of a
time-dependent ket $|\psi \rangle _t$ by
\begin{equation}
i\hbar \frac{\partial}{\partial t}|\psi \rangle _t
=H(\hat{q},\hat{p},t)|\psi \rangle _t.
\end{equation}
In $Q$-representation the time-dependent Schr\"{o}dinger equation takes
the form
\begin{equation}
i\hbar \frac{\partial}{\partial t}\psi ^Q(Q_1,t)=K\left( Q_1,
-i\hbar \frac{\partial}{\partial Q_1},t\right) \psi ^Q(Q_1,t),
\end{equation}
where $\psi ^Q(Q_1,t)={_t\langle}Q_1|\psi \rangle _t$, and
\begin{equation}
K(\hat{Q},\hat{P},t)=H(\hat{q},\hat{p},t)+i\hbar \hat{U}
\frac{\partial \hat{U}^\dagger}{\partial t}.
\end{equation}
The second term on the right hand side of Eq.\ (11) arises from the fact
that we allow the time dependence of the unitary operator $\hat{U}(t)$,
which indicates that, even though we adopt here the Schr\"{o}dinger picture
where the time dependence associated with the dynamical evolution of a
system is attributed solely to the ket $|\psi \rangle _t$, $\hat{Q}$ and
$|Q_1\rangle _t$ may depend on time also. In terms of the generating
operator $\bar{F}_1(\hat{q},\hat{Q},t)$, Eq.\ (11) can be written as
\begin{equation}
K(\hat{Q},\hat{P},t)=H(\hat{q},\hat{p},t)+\frac{\partial \bar{F} _1
(\hat{q},\hat{Q},t)}{\partial t}.
\end{equation}
The equivalence of Eqs.\ (11) and (12) can be proved as shown in Appendix A.
It is important to note that $K(\hat{Q},\hat{P},t)$ plays the role of the
transformed Hamiltonian governing the time-evolution of the system in
$Q$-representation. The analogy with the classical theory is remarkable.
\section{Quantum Hamilton-Jacobi Theory}
We are now ready to proceed to formulate the quantum Hamilton-Jacobi theory.
One can immediately notice that, if $K(\hat{Q},\hat{P},t)$ of Eq.\ (12)
vanishes, the time-dependent Schr\"{o}dinger equation in $Q$-representation
yields a simple solution, $\psi ^Q=$\ const. This observation along with
Eq.\ (2) naturally leads us to the following quantum Hamilton-Jabobi equation,
\begin{equation}
H\left( \hat{q},\frac{\partial \bar{S}_1(\hat{q},\hat{Q},t)}{\partial
\hat{q}},t\right) +\frac{\partial \bar{S}_1(\hat{q},\hat{Q},t)}{\partial t} =0,
\end{equation}
where, following the classical notational convention, we denote the
generating operator that is analogous to the classical Hamilton's principal
function by $\bar{S}_1(\hat{q},\hat{Q},t)$. Eq.\ (13) bears a close formal
resemblance to the classical Hamilton-Jacobi equation. It, however, differs
from the classical equation in that it is an operator partial differential
equation. The procedure of solving dynamical problems is completed if we
express the wave function in the original $q$-representation as
\begin{eqnarray}
\psi ^q(q_1,t)&=&\int \langle q_1|Q_2 \rangle _t\hspace{0.7mm}
{_t\langle}Q_2| \psi \rangle _t dQ_2, \nonumber \\
&=&\int e^{iS_1(q_1,Q_2,t)/\hbar}\psi ^Q(Q_2)dQ_2,
\end{eqnarray}
where $S_1(q_1,Q_2,t)$ is the c-number counterpart of
$\bar{S}_1(\hat{q},\hat{Q},t)$, and is obtained by replacing the well-ordered
$\hat{q}$ and $\hat{Q}$ in $\bar{S}_1$, respectively, with $q_1$ and $Q_2$.
In Eq.\ (14), $t$ is dropped from $\psi ^Q$, since
${_t\langle}Q_2| \psi \rangle _t =$ const.
As is the case for the classical Hamilton-Jacobi equation, the mission of
solving dynamical problems is assigned to the quantum Hamilton-Jacobi
equation.
Even though we arrive at the correct form of the quantum Hamilton-Jacobi
equation, it seems at first sight quite difficult to attain solutions of it
due to its unfamiliar appearance as an operator partial differential equation.
Thus it seems desirable to search a corresponding c-number form of the
quantum Hamilton-Jacobi equation. For this task, we note that, if the unitary
operator $\hat{U}(t)$ is assumed to be separable into
$\hat{U}(t)=\hat{T}(t)\hat{A}$, where $\hat{T}(t)$ is the time-evolution
operator
and $\hat{A}$ is an arbitrary time-independent unitary operator, then
$\psi ^Q(Q_1,t) ={_t\langle}Q_1|\psi \rangle _t=\langle q_1|\hat{A}^\dagger
\hat{T}^\dagger (t)\hat{T}(t)|\psi (t=0) \rangle =\langle q_1|\hat{A}^\dagger
|\psi (t=0)\rangle =$\ const. This means that the left hand side of Eq.\ (10)
becomes zero, i.e., the canonical transformation mediated by a separable
unitary operator is exactly the one that we seek. Assuming
$\hat{U}(t)=\hat{T}(t)\hat{A}$, we rewrite Eq.\ (1) as
\begin{equation}
e^{iS_1(q_1,Q_2,t)/\hbar} =\langle q_1|\hat{T}(t)\hat{A}|q_2\rangle .
\end{equation}
Differentiating this equation with respect to time, we obtain
\begin{eqnarray}
\frac{i}{\hbar}\frac{\partial S_1}{\partial t}e^{iS_1/\hbar} &=&
\langle q_1 |\frac{\partial \hat{T}}{\partial t}\hat{A}|q_2\rangle
=\frac{1}{i\hbar}\langle q_1|\hat{H}\hat{T}\hat{A}|q_2\rangle , \nonumber \\
&=&\frac{1}{i\hbar}H\left( q_1,-i\hbar \frac{\partial}{\partial q_1},t
\right) \langle q_1|\hat{T}\hat{A}|q_2\rangle ,\nonumber \\
&=&\frac{1}{i\hbar}H\left( q_1,-i\hbar \frac{\partial}{\partial q_1},t
\right) e^{iS_1/\hbar}.
\end{eqnarray}
Eq.\ (16) leads immediately to the desired c-number form of the quantum
Hamilton-Jacobi equation
\begin{equation}
\left[ H\left( q_1,-i\hbar \frac{\partial}{\partial q_1},t\right)
+\frac{\partial
S_1(q_1,Q_2,t)}{\partial t}\right] e^{iS_1(q_1,Q_2,t)/\hbar}=0.
\end{equation}
Substitution of $S_2(q_1,P_2,t)$ for $S_1(q_1,Q_2,t)$ generates another
c-number form of the quantum Hamilton-Jacobi equation. The equations for
the cases of $S_3(p_1,Q_2,t)$ and $S_4(p_1,P_2,t)$ can be derived through
a similar process.
Consider a one-dimensional nonrelativistic quantum system whose
Hamiltonian is given by
\begin{equation}
H(\hat{q},\hat{p},t)=\frac{\hat{p}^2}{2}+V(\hat{q},t).
\end{equation}
The c-number form of the quantum Hamilton-Jacobi equation (17) for this
problem becomes
\begin{equation}
\frac{1}{2}\left(\frac{\partial S_1}{\partial q_1}\right)^2
-i\frac{\hbar}{2}\frac{\partial ^2S_1}{\partial q_1^2}+V(q_1,t)
+\frac{\partial S_1}{\partial t}=0.
\end{equation}
We can see clearly that, in the limit $\hbar \rightarrow 0$, the above
equation reduces to the classical Hamilton-Jacobi equation. The second term
of Eq.\ (19) represents the quantum effect. We note that it has been known
from the early days that substitution of
$\psi (q,t)=e^{iS(q,t)/\hbar}$ into the Schr\"{o}dinger equation gives rise
to the same Hamilton-Jacobi equation for $S(q,t)$,\footnote{For a stationary
state of a system whose Hamiltonian does not depend explicitly on time, one
may put $S(q,t) = W(q) -Et$ and obtain a differential equation for $W(q)$.
To find a solution to the resulting equation, one may then use the expansion
of $W$ in powers of $\hbar$. This approach has been extensively considered
in connection with the well-known WKB approximation. In the present paper,
the formalism is developed for general nonstationary states (of systems
that can possibly have time-dependent Hamiltonians).}
where $S(q,t)$ is interpreted merely as the complex-valued phase of the
wave function (see, for example, Ref.\cite{schiff}).
The present approach
more clearly shows the strong analogy between the classical and quantum
Hamilton-Jacobi theories emphasizing that the quantum Hamilton's principal
function $S_1$ which is related with the wave function via Eq.\ (14) plays
the role of the quantum counterpart of the classical generating function.
Moreover, as discussed later in Sec.\ V, $e^{iS_1/\hbar}$ defined in Eq.\ (14)
can be interpreted as a propagator under a certain choice of $\hat{A}$.
It may be viewed that the Hamilton-Jacobi equation in the form of Eq.\ (19)
is no more tractable analytically than the Schr\"{o}dinger equation
for general potential problems. Nevertheless, it would be possible at least to
obtain an approximate solution of it using a perturbative method as follows.
Since the solution of Eq.\ (19) is given by the classical Hamilton's
principal function in the limit $\hbar \rightarrow 0$, we can expand the
general solution in powers of $\hbar$:
\begin{equation}
S_1=S_1^{(0)}+\hbar S_1^{(1)}+\hbar ^2S_1^{(2)}+\cdots ,
\end{equation}
where $S_1^{(0)}$ is the classical Hamilton's principal function.
Substituting Eq.\ (20) into Eq.\ (19) and collecting coefficients of the
same orders in $\hbar$, we can obtain
\begin{equation}
\frac{1}{2}\left(\frac{\partial S_1^{(0)}}{\partial q_1}\right)^2+V(q_1,t)
+\frac{\partial S_1^{(0)}}{\partial t}=0,
\end{equation}
and
\begin{equation}
\frac{1}{2}\sum_{k=0}^{n}\frac{\partial S_1^{(k)}}{\partial q_1}
\frac{\partial S_1^{(n-k)}}{\partial q_1}
-\frac{i}{2}\frac{\partial ^2S_1^{(n-1)}}{\partial q_1^2}
+\frac{\partial S_1^{(n)}}{\partial t}=0, \hspace{0.5cm} n\geq 1.
\end{equation}
Given the solution $S_1^{(0)}$ of the classical Hamilton-Jacobi equation (21),
we solve Eq.\ (22) to find $S_1^{(1)}$. $S_1^{(2)}$ can be determined
subsequently from the knowledge of $S_1^{(0)}$ and $S_1^{(1)}$, and so forth.
We note that Eq.\ (22) is linear in $S_1^{(n)}$ and first-order differential
in $q_1$ for $S_1^{(n)}$. Thus, from a practical viewpoint, Eqs.\ (21) and
(22) could
be more advantageous to deal with than Eq.\ (19) as long as
the classical Hamilton's
principal function that is the solution of Eq.\ (21) is readily available.
The present formalism provides an encouraging point that the well-ordered
operator counterpart of the quantum Hamilton's principal function gives also
the solutions of the Heisenberg equations through Eq.\ (2). If we consider
the case $\hat{U}(t)=\hat{T}(t)$, we can obtain in the Heisenberg picture the
relations $(\hat{q}_H,\hat{p}_H)\equiv
(\hat{T}^\dagger \hat{q}_S\hat{T}, \hat{T}^\dagger \hat{p}_S\hat{T})$
and $(\hat{Q}_H,\hat{P}_H)\equiv (\hat{T}^\dagger \hat{Q}_S\hat{T},
\hat{T}^\dagger \hat{P}_S\hat{T})= (\hat{T}^\dagger \hat{T}\hat{q}_S
\hat{T}^\dagger \hat{T},\hat{T}^\dagger \hat{T}\hat{p}_S\hat{T}^\dagger
\hat{T})=(\hat{q}_S,\hat{p}_S)$, where we attached the subscript $_S$ and
$_H$ to operators to explicitly denote, respectively, the Schr\"{o}dinger and
the Heisenberg pictures. Thus, when expressed in the Heisenberg
picture Eq.\ (2) turns into
\begin{equation}
\hat{p}_H=\frac{\partial \bar{S}_1(\hat{q}_H,\hat{q}_S,t)}{\partial \hat{q}_H},
\hspace{1cm}
\hat{p}_S=-\frac{\partial \bar{S}_1(\hat{q}_H,\hat{q}_S,t)}{\partial \hat{q}_S},
\end{equation}
and from these transformation relations we can obtain $\hat{q}_H$ and
$\hat{p}_H$ as functions of time and the initial operators $\hat{q}_S$ and
$\hat{p}_S$. Obviously, $\hat{q}_H(\hat{q}_S,\hat{p}_S,t)$ and
$\hat{p}_H(\hat{q}_S,\hat{p}_S,t)$ obtained
in this way evolve according to the Heisenberg equations.
\section{Quantum Phase-Space distribution functions and canonical
transformations}
Since our theory of the quantum canonical transformations is formulated with
the canonical position $\hat{q}$ and momentum $\hat{p}$ variables on an equal
footing, it would be relevant to consider the phase-space picture of quantum
mechanics, exploiting the distribution functions in relation to the present
theory.
\subsection{Distribution functions}
For a given density operator $\hat{\rho}$, a general way of defining quantum
distribution functions proposed by Cohen\cite{one5} is that
\begin{equation}
F^f(q_1,p_1,t)=\frac{1}{2\pi ^2\hbar}\int \int \int dxdydq_2 \langle
q_2+y|\hat{\rho} |q_2-y\rangle f(x,2y/\hbar )e^{ix(q_2-q_1)}
e^{-i2yp_1/\hbar}.
\end{equation}
Various choices of $f(x,2y/\hbar)$ lead to a wide class of quantum
distribution functions\cite{one6}. To mention only a few, the choice $f=1$
produces the well-known Wigner distribution function \cite{one7}, while the
choice $f(x,2y/\hbar)= e^{-\hbar x^2/4m\alpha -m\alpha y^2/\hbar}$ yields
the Husimi distribution function that recently has found its application in
nonlinear dynamical problems\cite{one8}. The transformed distribution
function is defined in ($Q_1,P_1$) phase space likewise by
\begin{equation}
G^f(Q_1,P_1,t)=\frac{1}{2\pi ^2\hbar}\int \int \int dXdYdQ_2\hspace{0.7mm}
{_t\langle} Q_2+Y|\hat{\rho} |Q_2-Y\rangle _tf(X,2Y/\hbar )e^{iX(Q_2-Q_1)}
e^{-i2YP_1/\hbar}.
\end{equation}
Our main objective here is to find a relation between the old and the
transformed distribution functions. After a straightforward algebra, which
is displayed in Appendix B, it turns out that the transformation relation
between the two distribution functions can be expressed as
\begin{equation}
G^f(Q_1,P_1,t)=\int \int dq_2dp_2\kappa (Q_1,P_1,q_2,p_2,t)
F^f(q_2,p_2,t),
\end{equation}
where the kernel $\kappa$ is given by
\begin{eqnarray}
\kappa (Q_1,P_1,q_2,p_2,t)=\frac{1}{2\pi ^3\hbar} \int \int \int
\int \int \int dXdYdQ_2dxdyd\alpha \frac{f(X,2Y/\hbar )}
{f(x,2y/\hbar )} \nonumber \\
\times e^{\frac{i}{\hbar}[F_1(q_2+\alpha -y,Q_2-Y,t)-F_1^*(q_2+\alpha +y,
Q_2+Y,t)]} e^{i[X(Q_2-Q_1)-\alpha x]} e^{\frac{2i}{\hbar}
[yp_2-YP_1]}.
\end{eqnarray}
This expression for the kernel can be further simplified if integrations
in Eq.\ (27) can be performed with a specific choice of the function $f$.
For instance, the simple choice $f=1$ provides the following kernel for the
Wigner distribution function,
\begin{eqnarray}
\kappa (Q_1,P_1,q_2,p_2,t)=\frac{2}{\pi \hbar} \int \int dYdy
e^{\frac{i}{\hbar} [F_1(q_2-y,Q_1-Y,t)-F_1^*(q_2+y,Q_1+Y,t)]}
e^{\frac{2i}{\hbar}[yp_2-YP_1]}.
\end{eqnarray}
This equation was first derived by Garcia-Calder\'on and Moshinsky\cite{one9}
without employing the idea of the quantum generating function.
Curtright {\it et al.}\cite{one10} also obtained an equivalent expression
in their recent discussion of the time-independent Wigner
distribution functions.
We wish to point out that the quantum canonical transformation described here
is basically different from that considered earlier by Kim and
Wigner\cite{one1}.
While the present approach deals with the transformation between operators
($\hat{q},\hat{p}$) and ($\hat{Q}, \hat{P}$), their approach is about the
transformation between c-numbers ($q,p$) and ($Q,P$). For the transformation
$Q=Q(q,p,t)$ and $P=P(q,p,t)$, their approach yields for the kernel the
expression
\begin{equation}
\kappa (Q_1,P_1,q_2,p_2,t)=\delta [Q_1-Q(q_2,p_2,t)]\delta
[P_1-P(q_2,p_2,t)],
\end{equation}
where $Q(q,p,t)$ and $P(q,p,t)$ satisfy the classical Poisson brackets
relation, $[Q,P]_{PB}=[q,p]_{PB}=1$. The kernels of Eq.\ (28) and Eq.\ (29)
coincide with each other for the special case of a linear canonical
transformation, as was shown by Garcia-Calder\'on and Moshinsky\cite{one9}.
Specifically, for the case of the Wigner distribution function, they showed
that the linear transformation for operators, $\hat{Q}=a\hat{q}+b\hat{p}$ and
$\hat{P}=c\hat{q}+d\hat{p}$, and that for c-number variables, $Q=aq+bp$ and
$P=cq+dp$, yield the same kernel
$\kappa (Q_1,P_1,q_2,p_2)=\delta [Q_1-(aq_2+bp_2)]\delta [P_1-(cq_2+dp_2)]$.
In general cases, however, Eq.\ (27) and Eq.\ (29) give rise to different
kernels. As an example, let us consider the unitary transformation
$\hat{U}=e^{ig(\hat{q})/\hbar}$ considered in Sec.\ II. The first-type
quantum generating function has the form
$e^{iF_1(q_1,Q_2)/\hbar}=e^{ig(q_1)/\hbar}
\delta (q_1-Q_2)$. This nonlinear canonical transformation yields for the
Wigner distribution function the kernel
\begin{equation}
\kappa (Q_1,P_1,q_2,p_2)= \frac{\delta (Q_1-q_2)}{\pi \hbar} \int
dy e^{\frac{i}{\hbar}
[g(q_2-y)-g(q_2+y)]} e^{\frac{2i}{\hbar}(p_2-P_1)y}.
\end{equation}
It is apparent that the integral of the above equation cannot generally be
reduced to the $\delta$-function of Eq.\ (29) except for some trivial cases,
e.g., $g=$const, $g=q$, and $g=q^2$. Distribution functions other than the
Wigner distribution function do not usually allow the simple expression for
the kernel in the form of Eq.\ (29), even if one considers a linear canonical
transformation.
\subsection{Dynamics}
In this subsection we describe how the quantum Hamilton-Jacobi theory can
lead to dynamical solutions in the phase-space picture of quantum mechanics.
For this task, we first consider the time evolution of the transformed
distribution function in ($Q_1,P_1$) phase space. Differentiating Eq.\ (25)
with respect to time, we can get
\begin{eqnarray}
\frac{\partial G^f}{\partial t}=\frac{1}{2\pi ^2\hbar}
\int \int \int dXdYdQ_2 \left[ \left( \frac{\partial}
{\partial t}{_t\langle} Q_2+Y| \right) \hat{\rho} |Q_2-Y\rangle _t+
{_t\langle}Q_2+Y|\frac{\partial \hat{\rho}}{\partial t}|Q_2-Y\rangle _t
\right. \nonumber \\
\left. +{_t\langle}Q_2+Y|\hat{\rho} \left( \frac{\partial}{\partial t}
|Q_2-Y\rangle _t\right) \right] f(X,2Y/\hbar)e^{iX(Q_2-Q_1)}
e^{-i2YP_1/\hbar}.
\end{eqnarray}
We now substitute into Eq.\ (31) the time evolution equations
\begin{eqnarray}
\frac{\partial \hat{\rho}}{\partial t}&=&-\frac{i}{\hbar}
[\hat{H},\hat{\rho}], \\
\frac{\partial}{\partial t}{_t\langle}Q_2+Y|&=&{_t\langle}
Q_2+Y|\hat{U}\frac{\partial \hat{U}^\dagger}{\partial t}, \\
\frac{\partial}{\partial t}|Q_2-Y\rangle _t &=&\frac{\partial \hat{U}}
{\partial t}\hat{U}^\dagger |Q_2-Y\rangle _t=-\hat{U}\frac{\partial
\hat{U}^\dagger}{\partial t}|Q_2-Y\rangle _t,
\end{eqnarray}
where $\hat{H}=H(\hat{q},\hat{p},t)$ is the Hamiltonian governing the
dynamics of the system, and obtain
\begin{eqnarray}
\frac{\partial G^f}{\partial t}=\frac{1}{2\pi ^2\hbar}
\int \int \int dXdYdQ_2\hspace{0.7mm} {_t\langle}Q_2+Y|\left( -\frac{i}{\hbar}
[\hat{K},\hat{\rho}]\right) |Q_2-Y\rangle _tf(X,2Y/\hbar)e^{iX(Q_2-Q_1)}
e^{-i2YP_1/\hbar},
\end{eqnarray}
where $\hat{K}=K(\hat{Q},\hat{P},t)$ is just the transformed Hamiltonian
already defined in Eq.\ (11). Eq.\ (35) should be compared with the
following equation that governs the time evolution of the distribution
function in ($q_1,p_1$) phase space,
\begin{eqnarray}
\frac{\partial F^f}{\partial t}=\frac{1}{2\pi ^2\hbar}
\int \int \int dxdydq_2 \langle q_2+y|\left( -\frac{i}{\hbar}
[\hat{H},\hat{\rho}]\right) |q_2-y\rangle f(x,2y/\hbar)e^{ix(q_2-q_1)}
e^{-i2yp_1/\hbar}.
\end{eqnarray}
We can easily see that, through the quantum canonical transformation, the
role played by $\hat{H}$ is turned over to $\hat{K}$.
Just as the wave function has a trivial solution in the representation
where the transformed Hamiltonian $K(\hat{Q},\hat{P},t)$ vanishes, so does
the distribution function in the corresponding phase space, as can be seen
from Eq.\ (35). With the trivial solution $G^f=$\ const., we go back to the
original space via the inverse of the transformation equation (26) to obtain
$F^f(q_1,p_1,t)$. For example, for the case of the Wigner distribution
function the transformation can be accomplished by
\begin{equation}
F^W(q_1,p_1,t)=\int \int dQ_2dP_2\tilde{\kappa} (q_1,p_1,Q_2,P_2,t)
G^W(Q_2,P_2),
\end{equation}
where $\tilde{\kappa}$ is given in terms of the quantum principal function by
\begin{eqnarray}
\tilde{\kappa} =\frac{2}{\pi \hbar} \int \int dydY e^{\frac{i}{\hbar}
[S_1(q_1+y,Q_2+Y,t)-S_1^*(q_1-y,Q_2-Y,t)]}
e^{\frac{2i}{\hbar}[YP_2-yp_1]}.
\end{eqnarray}
Thus, once the quantum Hamilton-Jacobi equation is solved and the quantum
principal function $S_1$ is obtained, the dynamics of the distribution
function, as well as that of the wave function, can be determined.
\section{Boundary conditions and Applications}
Up to this point the whole theory has been developed for the case
$\hat{U}(t)=\hat{T}(t)\hat{A}$ with $\hat{A}$ taken to be arbitrary
unless otherwise mentioned.
To see how the quantum Hamilton-Jacobi theory is used to achieve the
dynamical solutions of quantum problems, it would be
sufficient, though, to consider the case of $\hat{A}=\hat{I}$, the unit
operator. This case was considered by Dirac in
connection with his action principle (see Sec.\ 32 of Ref.\cite{three}). He
showed that $S_1$ defined by Eq.\ (15) equals the classical action function
in the limit $\hbar \rightarrow 0$. It should be mentioned that this
particular case allows the quantum generating functions to attain the
property that $e^{iS_1/\hbar}$ is the propagator in position space and
$e^{iS_4/\hbar}$ the propagator in momentum space. We will henceforth work
on the case $\hat{U}(t)=\hat{T}(t)$. The general case
$\hat{U}(t)=\hat{T}(t)\hat{A}$ will be briefly treated in Appendix C.
Before applying the theory it is necessary to provide some remarks
concerning the
quantum Hamilton-Jacobi equation (17) and its solution. First, if
the Hamiltonian depends only on either $\hat{q}$ or $\hat{p}$, we do not
need to solve Eq.\ (17). Instead, since the unitary operator has the simple
form $\hat{U}=\hat{T}=e^{-iH(\hat{q})t/\hbar}$ or $e^{-iH(\hat{p})t/\hbar}$,
we can obtain $S_1$ directly from Eq.\ (15) by calculating the matrix
elements of $\hat{U}$. For example, for a free particle,
$\hat{U}=e^{-i\hat{p}^2t/2\hbar}$, it is convenient to calculate
$e^{iS_2/\hbar}=\langle q_1|e^{-i\hat{p}^2t/2\hbar}|p_2\rangle$, and we get
$S_2(q_1,P_2,t)=-\frac{P_2^2t}{2}+q_1P_2+i\frac{\hbar}{2} \ln 2\pi \hbar$.
Second, in order to solve Eq.\ (17), we need to impose proper boundary
conditions on $S_1$. Since here we are dealing with unitary transformations,
we immediately get from the definition of $S_1$ the condition
\begin{equation}
\int dQ_3 e^{i[S_1(q_1,Q_3,t)-S_1^*(q_2,Q_3,t)]/\hbar} =\delta (q_1-q_2),
\end{equation}
which follows from the calculation of the matrix elements of
$\hat{U}(t)\hat{U}^{\dagger}(t)=\hat{I}$.
This unitary condition ensures that the well-ordered operator
counterpart of $S_1$ yields Hermitian operators for $\hat{Q}$ and $\hat{P}$
from Eq.\ (2). Mathematically, Eq.\ (17) can have
several solutions, and there is an arbitrariness in the choice of the
new position variable, because any function of the constant of integration
of Eq.\ (17) can be a candidate for the new position variable. Not all the
possible solutions correspond to the unitary transformations, and from the
possible solutions we choose only those which satisfy Eq.\ (39) and thus give
Hermitian position and momentum operators that are observables.
These solutions correspond to
the unitary transformations of the type $\hat{U}(t)=\hat{T}(t)\hat{A}$.
Further, from these solutions we single out the one that corresponds to the
case $\hat{A}=\hat{I}$ by imposing the condition
$e^{iS_1(q_1,Q_2,t=0)/\hbar}=\delta (q_1-Q_2)$ as an initial
condition. The appropriate form for $S_2$ corresponding to this condition is
that $e^{iS_2(q_1,P_2,t=0)/\hbar}=\frac{1}{\sqrt{2\pi \hbar}}
e^{iq_1P_2/\hbar}$. In the limit $\hbar \rightarrow 0$, $S_2$ in this
equation reduces to the correct classical generating function for the
identity transformation, $S_2=q_1P_2$. In solving the Hamilton-Jacobi
equation perturbatively using Eqs.\ (21) and (22), in order to consistently
satisfy the initial condition, we start with the classical Hamilton's
principal function $S_1^{(0)}$ that gives at initial
time the relations $q_1=Q_2$ and $p_1=P_2$ from the classical c-number
counterpart of Eq.\ (2). An arbitrary additive constant $c$ to the solution
of Eq.\ (17) that always appears in the form $S_1+c$ when we deal with a
partial differential equation such as Eq.\ (19) which contains only partial
derivatives of $S_1$\cite{one} can also be fixed by the initial condition.
Depending
whether boundary conditions can readily be expressed in a simple form, one
type of the quantum generating function may be favored over another. The
existence and uniqueness of the independent solution of Eq.\ (17) satisfying
the above conditions can be guaranteed from the consideration of the equation
$e^{iS_1(q_1,Q_2,t)/\hbar}=\langle q_1|\hat{T}(t)|q_2\rangle$, in which
$S_1$ is just given by the matrix elements of $T(t)$. It is clear that
these matrix elements exist and are uniquely defined.
As illustrations of the application of the theory, we consider the following
two simple systems.
{\sl Example 1. A particle under a constant force.}
As a first example, let us consider a particle moving under a constant force
of magnitude $a$, for which the Hamiltonian is $\hat{H}=\hat{p}^2/2-a\hat{q}$.
We start with the following classical principal function that is
the solution of Eq. (21),
\begin{equation}
S_1^{(0)}=\frac{(q_1-Q_2)^2}{2t} +\frac{at(q_1+Q_2)}{2}
-\frac{a^2t^3}{24}.
\end{equation}
Substituting $S_1^{(0)}$ into Eq.\ (22) and solving the resulting equation,
we find that the first order term in $\hbar$ has the general solution
\begin{equation}
S_1^{(1)}=\frac{i}{2}\ln t +f\left( \frac{q_1-Q_2}{t} -\frac{a}{2}t\right) ,
\end{equation}
where $f$ is an arbitrary differentiable function. To satisfy the proper
boundary condition $e^{iS_1(q_1,Q_2,t=0)/\hbar} =\delta (q_1-Q_2)$,
$f$ and all higher order terms of $S_1$ are chosen to be zero, and the overall
additive constant to be $c=\hbar \frac{i}{2}\ln i2\pi \hbar$.
By well-ordering terms, we get the generating operator
\begin{equation}
\bar{S_1}(\hat{q},\hat{Q},t)=\frac{\hat{q}^2-2\hat{q}\hat{Q}
+\hat{Q}^2}{2t}+\frac{at}{2}(\hat{q}+\hat{Q})
-\frac{a^2t^3}{24} +\hbar\frac{i}{2}\ln i2\pi \hbar t.
\end{equation}
We can easily check that the operator form of the quantum Hamilton-Jacobi
equation (13) is satisfied by the above generating operator.
From Eq.\ (14) we obtain the wave function
\begin{equation}
\psi ^q(q_1,t)=\int \frac{1}{\sqrt{i2\pi \hbar t}}
e^{\frac{i}{2\hbar t}
[(q_1-Q_2)^2 +at^2(q_1+Q_2)-a^2t^4/12]}\psi ^Q(Q_2)dQ_2.
\end{equation}
Because $\psi ^Q(Q_2)$ is constant in time, we can express it in terms of the
initial wave function. For the present case in which we use the first-type
quantum generating function $S_1$ and $\hat{A}=\hat{I}$, we have simply
$\psi ^q(q_2=Q_2,t=0)=\psi ^Q(Q_2)$. We note that Eq.\ (43) is in exact
agreement with the result of Feynman's path-integral approach\cite{two0}.
For the time evolution of the distribution function, we find from Eq.\ (38)
the following kernel for the Wigner distribution function,
\begin{eqnarray}
\tilde{\kappa} (q_1,p_1,Q_2,P_2,t)&=&\frac{1}{\pi ^2\hbar ^2 t}\int \int
dYdy e^{-\frac{2i}{\hbar}\left(Q_2-q_1+p_1t-\frac{at^2}{2}\right)
\frac{y}{t}} e^{\frac{2i}{\hbar}\left( P_2-\frac{q_1-Q_2-at^2/2}{t}
\right) Y}, \nonumber \\
&=&\delta(Q_2-q_1+p_1t-at^2/2) \delta \left(P_2-\frac{q_1-Q_2-at^2/2}
{t}\right) .
\end{eqnarray}
Substituting Eq.\ (44) into Eq. (37), we obtain
\begin{eqnarray}
F^W(q_1,p_1,t)=F^W(q_1-p_1t+at^2/2,p_1-at,0),
\end{eqnarray}
where use has been made of the relation $F^W(q_1,p_1,t=0)=G^W(q_1,p_1)$.
As has been mentioned, the present Hamilton-Jacobi theory also
provides the solutions of the Heisenberg equations via the transformation
relations between the two sets of canonical operators.
From Eqs.\ (2) and (42) we can obtain
\begin{eqnarray}
\hat{q}_S&=&\hat{Q}_S(t)+\hat{P}_S(t)t+\frac{a}{2}t^2, \\
\hat{p}_S&=&\hat{P}_S(t) +at.
\end{eqnarray}
In the Heisenberg picture, the above equations become
\begin{eqnarray}
\hat{q}_H(t)&=&\hat{q}_S+\hat{p}_St+\frac{a}{2}t^2, \\
\hat{p}_H(t)&=&\hat{p}_S +at,
\end{eqnarray}
which are the solutions of the Heisenberg equations.
By setting $a=0$, we can obtain the free particle solution.
{\sl Example 2. The harmonic oscillator}
For the harmonic oscillator whose Hamiltonian is given
by $\hat{H}=\hat{p}^2/2+\hat{q}^2/2$, the classical Hamilton-Jacobi
equation (21) can be solved to give the classical principal function
\begin{equation}
S_1^{(0)}=\frac{1}{2}(q_1^2+Q_2^2)\cot t-q_1Q_2\csc t.
\end{equation}
With the boundary condition $e^{iS_1(q_1,Q_2,t=0)/\hbar}=\delta (q_1-Q_2)$,
Eq.\ (22) can be solved to give
\begin{equation}
S_1^{(1)}=\frac{i}{2}\ln \sin t,
\end{equation}
and $S_1^{(2)}=\cdots =0$. The additive constant has the form
$c=\hbar \frac{i}{2}\ln i2\pi \hbar$. The well-ordered generating operator
is then written as
\begin{equation}
\bar{S_1}(\hat{q},\hat{Q},t)=\frac{1}{2}(\hat{q}^2+\hat{Q}^2)\cot t
-\hat{q}\hat{Q}\csc t +\hbar \frac{i}{2} \ln i2\pi \hbar \sin t.
\end{equation}
The wave function takes the form
\begin{equation}
\psi ^q(q_1,t)=\int \frac{1}{\sqrt{i2\pi \hbar \sin t}}
e^{\frac{i}{2\hbar \sin t}
[(q_1^2+Q_2^2)\cos t -2q_1Q_2]}\psi ^q(Q_2,0)dQ_2,
\end{equation}
and the kernel and the distribution function are given respectively by
\begin{equation}
\tilde{\kappa} (q_1,p_1,Q_2,P_2,t)=\delta (Q_2-q_1\cos t +p_1\sin t )
\delta (P_2+Q_2\cos t -q_1\csc t),
\end{equation}
and
\begin{equation}
F^W(q_1,p_1,t)=F^W(q_1\cos t-p_1\sin t,q_1\sin t+p_1\cos t,0).
\end{equation}
This equation shows that the Wigner distribution function for the harmonic
oscillator rotates clockwise in phase space.
The quantum Hamilton-Jacobi equation for other types of generating operators
can be solved by a similar technique. For instance, we can obtain the
following solution for the second-type generating operator,
\begin{equation}
\bar{S_2}(\hat{q},\hat{P},t)=-\frac{1}{2}(\hat{q}^2+\hat{P}^2)\tan t
+\hat{q}\hat{P}\sec t+\hbar \frac{i}{2} \ln 2\pi \hbar \cos t.
\end{equation}
The solutions of the Heisenberg equations can be obtained from Eqs.\ (2) and
(52) (or Eqs. (3) and (56)). In the Heisenberg picture we have
\begin{eqnarray}
\hat{q}_H(t)&=&\hat{q}_S\cos t+\hat{p}_S\sin t, \\
\hat{p}_H(t)&=&-\hat{q}_S\sin t+\hat{p}_S\cos t.
\end{eqnarray}
It should be mentioned that, even though we restricted our discussion
in this section only to the case $\hat{A}=\hat{I}$ by imposing
the special initial condition, it is very probable that another choice of
$\hat{A}$ satisfying the quantum Hamilton-Jacobi equation happens to be more
readily obtainable. In that case, the initial condition that
is derived from $e^{iS_1(q_1,Q_2,0)/\hbar}=\langle q_1|\hat{A}|q_2\rangle$ is
of course different from that described above. As an example, for the
harmonic oscillator, we presented a different solution for $S_1$ in Appendix C
where the unitary operator $\hat{A}$ corresponds to the transformation that
interchanges the position and momentum operators.
\section{Concluding remarks}
We wish to give some final remarks concerning the quantum Hamilton-Jacobi
theory. In this approach, the quantum Hamilton-Jacobi equation takes the
place of the time-dependent Schr\"{o}dinger equation for solving dynamical
problems, and the quantum Hamilton's principal function $S_1$ that is the
solution of the former equation gives the solution of the latter equation
through Eq.\ (14). As mentioned in Sec.\ V, $e^{iS_1/\hbar}$ becomes the
propagator in position space for the case $\hat{A}=\hat{I}$. To find the
propagator, Feynman's path-integral approach divides the time difference
between a given initial state and a final state into infinitesimal time
intervals, and then lets the quantum generating function for the
infinitesimal transformation equal the classical action function plus
a proper additive constant that vanishes in the limit $\hbar \rightarrow
0$, and finally takes the sum of the infinitesimal
transformations. On the other hand, the present approach seeks the
quantum generating function that
directly transforms the initial state to the final state.
The present formalism gives also the solutions of the Heisenberg equations
through the transformation relations which in the Heisenberg picture can
be expressed as Eq.\ (23).
In conclusion, it is clear that the present approach, which has its origin
in Dirac's canonical transformation theory, helps better comprehend the
interrelations among the existing different formulations of quantum mechanics.
Finally, one more remark may be worth making as to the extent to which the
quantum Hamilton-Jacobi theory can stretch the range of its validity.
Even though our work here deals with the
unitary transformation to ensure that the new operators become Hermitian,
and hence observables, the main idea presented in this paper could be
extended so as to include the non-unitary transformation that deals with
non-Hermitian operators. The theory would then have the form of the
quantum Hamilton-Jacobi equation, but it would be associated with
different types of
transformations, such as $\hat{U}(t)=\hat{T}(t)\hat{B}$ where $\hat{U}(t)$ and
$\hat{B}$ are not unitary. However, it may then be necessary to pay particular
attention
and care to the completeness of the eigenstates of the new operators
$\hat{Q}$ and $\hat{P}$, for the property is crucial to several relations
derived and has been used implicitly throughout the paper.
|
\section{Introduction}
\indent
The reduction of the absorption strength of high energy real photons on nuclei
is known as shadowing effect.
This effect is generally described considering
the real photon as a superposition of a bare photon and of a hadronic fluctuation
with the same quantum numbers $(J^{PC}=1^{--})$.
Within this model the shadowing is produced by the coherent multiple scattering
of the hadronic intermediate state on different nucleons
inside the nucleus.
The amount of the shadowing mainly depends
on macroscopic nuclear parameters like the mass number A and the radius $r_A$,
and on properties of the hadronic fluctuation like the
coherence length $\lambda_h$ and the interaction cross section $\sigma_{hN}$ with the nucleon.
In earliest simple models \cite{BR69}, the hadronic component of the photon is
given by the low-lying
vector mesons $\rho, \omega$ and $\phi$. These Vector Meson Dominance (VMD) models
qualitatively reproduce the photonuclear absorption cross section
behavior in the several GeV domain \cite{HEY71}.
Generalized Vector Meson Dominance (GVMD) models \cite{DI76},
which include higher
mass vector mesons and non diagonal terms, better explain higher energy
real photon absorption and virtual photon absorption in deep inelastic electron scattering.
On the contrary, at low real photon energies
most of the calculations fail to reproduce the experimental results \cite{MIR97}.
Two recent VMD calculations,
that describe the vector-meson mass distributions with $\delta$-functions \cite{PI95,BO96}
and consider an energy independent vector-meson nucleon cross section $\sigma_{VN}$ \cite{PI95},
do not predict the nuclear
damping of the photoabsorption strength observed below 2 GeV,
as shown in Fig.~\ref{Fig1} for the carbon case.
In addition they also underestimate the experimental shadowing effect
between 2 and 3 GeV.
The result of a GVMD calculation \cite{ENG97}, in which the energy behavior of
$\sigma_{VN}$ cross section is taken into account, is also given in Fig.~\ref{Fig1}. It clearly
shows a better agreement with the experimental shadowing ratio, but it is not able to reproduce
the absolute value of the total photonuclear cross section.
The shadowing phenomena, also observed in deep inelastic lepton nucleus scattering,
is also studied within a VMD model
in which the photon hadronic spectral function is derived from the empirical cross sections
of the {\it $e^+e^-\rightarrow$ hadrons} processes \cite{PI90}.
Besides the vector meson mass spectra,
this model also includes the low energy $\pi^+\pi^-$ non-resonant production,
and the high energy quark-antiquark continuum.
The importance of the hadronic spectral function
in the description of the process is shown
in Fig.~\ref{Fig2} where the coherence length $\lambda_V= 2k / m_V^2$
of vector mesons of lower mass $m_V$ are given as a function of the photon energy $k$.
The shadowing effect starts to manifest at an energy for which $\lambda_V$
is bigger than the typical intranucleon distance
($d_{NN}\sim 1.8$ fm) so that scattering on at least two nucleons is possible.
Then the strength of the effect starts to saturate at an energy for which $\lambda_V$
is bigger than the nuclear size ($\sim 2r_A$).
Clearly a low energy shadowing can be only induced by the lowest mass hadronic
components of the photon spectral function and by their possible modifications
in the nuclear medium.
Both the reduction of the vector-meson mass \cite{BRO91,HAT92}
and the modification of the $\rho$-meson spectral function \cite{RAP97,KLI97} can
decrease the photon energy at which the coherence length starts to exceed the intranucleon distance
thus producing an earlier onset of the shadowing effect.
In this paper a model is derived to describe the photonucleon and the photonuclear
total absorption cross sections above the nucleon resonance region ($k \geq$ 1.65 GeV).
In particular the experimental hadronic spectral function, vector-meson nucleon
cross sections and effective $\rho$-coupling constant are taken into account.
A possible modification of the hadronic spectral function
inside the nuclear medium is also considered.
\begin{figure}[t]
\vspace{13cm}
\leavevmode
\special{psfile=vmd_fig1.eps vscale=110 hscale=100 voffset=-450 hoffset=-30 angle=0}
\caption{a) Total photonuclear cross section and b)
ratio to the photonucleon cross section for carbon.
Different symbols refer to different experiments.
Also shown in a) is the total cross section on hydrogen (thin solid line).
Dashed \protect\cite{PI95},
dot-dashed \protect\cite{BO96} and dotted \protect\cite{ENG97} lines are two VMD
and one GVMD predictions.
}
\label{Fig1}
\end{figure}
\begin{figure}[t]
\vspace{8cm}
\leavevmode
\special{psfile=vmd_fig2.eps vscale=100 hscale=100 voffset=-500 hoffset=-40 angle=0}
\caption{Coherence length of the hadronic spectral function as a function of the photon energy.
The average intranucleon distance $d_{NN}$, the carbon $2r_{C}$ and lead
$2r_{Pb}$ nuclear diameters are also shown.}
\label{Fig2}
\end{figure}
\section{Model}
In the description of the photohadronic absorption process,
the physical photon is considered as a superposition of a bare photon
and a hadronic component made up of a quark-antiquark state ($q\overline{q}$).
The photonucleon cross section $\sigma_{\gamma N}$ is decomposed in
a term $\sigma_{\gamma N}^{dir}$ due to the direct coupling of the bare photon
with the nucleon and an hadronic term $\sigma_{\gamma N}^{had}$.
At small total center of mass energy the hadronic components of the absorbed
photon are mainly formed by strongly correlated $q\overline{q}$ pairs, while at higher energy
$q\overline{q}$ pairs from the so-called continuum are also important.
\subsection{Photoabsorption on the nucleon}
\indent
The hadronic contribution to the photoabsorption cross section on the proton
is expressed by a
spectral relation of the form \cite{PI95,ENG97}:
\begin{equation}
\sigma_{\gamma p}^{had}(k) = 4 \pi \alpha_{em} \int^{s_u}_{s_0} \frac{d\mu^2}{\mu^2}
\Pi (\mu^2) \sigma_{hp} (\mu^2\, , k) \, \,,
\label{eq1}
\end{equation}
being $\Pi (\mu^2)$ the spectrum of the hadronic fluctuation of mass $\mu$
and $\sigma_{hp}$ the effective hadron-proton cross section.
The integration limits are the two pion production threshold $s_0 \equiv (2\, \, m_\pi)^2$
and ${s_u}={(\sqrt{s}-m_p)^2}$ with $s$ the total center of mass energy
and $m_p$ the proton mass.
The hadronic spectral function of the photon $\Pi (\mu^2)$ is
related to the measured cross section of the
$e^+ e^- \rightarrow$ $hadrons$ process by
\begin{equation}
\Pi(s) = \frac{1}{12\pi^2} \frac{\sigma_{e^+e^- \rightarrow hadrons}(s)}
{\sigma_{e^+e^- \rightarrow \mu^+ \mu^-}(s)} \, \,.
\label{eq2}
\end{equation}
At low energy ($s \leq {s_{1}} = m_{\phi}^2 \approx 1$ GeV$^2$), $\Pi(s)$ is dominated by the resonance contribution $\Pi^R(s)$
due to the sum of
the low-mass vector meson spectral function $G_V (s)$.
At higher energy ($s > s_{1}$), besides the narrow charmonium and upsilon resonances, the spectral function is dominated by
the contribution $\Pi^C(s)$ of the continuum quark-antiquark fluctuations. Then, the total spectral function $\Pi (s)$ can be written as
\begin{eqnarray}
\Pi(s) = \Pi^R(s) + \Pi^C(s) =
\sum_{V=\rho , \omega , \phi , J/\psi , \psi '} G_V(s) + \Pi^C(s).
\label{eq61}
\end{eqnarray}
Substituting Eq.~(\ref{eq61}) in Eq.~(\ref{eq1}), the $\sigma_{\gamma p}^{had}$ is written in terms of the resonance
and the continuum contributions:
\begin{eqnarray}
\sigma_{\gamma p}^{had}(k) & = & \sigma_{\gamma p}^R (k)+ \sigma_{\gamma p}^C (k) = \nonumber \\
& = & 4\pi \alpha_{em} \sum_V \int^{s_u}_{s_0} \frac{d \mu^2}{\mu^2} G_V (\mu^2) \sigma_{Vp}
(k) +\nonumber \\
& + & \, 4\pi \alpha_{em} \int^{s_u}_{s_1} \frac{d \mu^2}{\mu^2} \Pi^C (\mu^2) \sigma_{qp}
(\mu^2,k) \, \,,
\label{eq5}
\end{eqnarray}
where $\sigma_{Vp}$ and $\sigma_{qp}$ are the interaction cross sections of the vector mesons
and of the continuum
quark-antiquark pairs respectively.
In this work, $G_V (s)$ are derived directly from Eq.~(\ref{eq2}) by taking into account
the experimental resonance widths \cite{MIR98}:
\begin{eqnarray}
G_V (s) & = &\frac{1}{\pi} (\frac{m_V}{g_V})^2 \frac{B_V m_V \Gamma_V(s)}{(s- m^2_V)^2 + (m_V \Gamma _V(s))^2} \,\,,
\label{eq9}
\end{eqnarray}
where $g_V$ are the $\gamma V$ coupling constants,
$\Gamma _V(s)$ are the total hadronic widths of the resonances and $B_V$ are the branching
ratios for the decay $V \rightarrow e^+ e^-$ \cite{PDG}.
The continuum contribution is written as:
\begin{equation}
\Pi^C(s) = \frac{1}{12 \pi ^2} \Sigma_f \, \, \, 3 q^2_f
\label{eq4}
\end{equation}
and the sum is extended over all quark flavors $f$ of fractional charge $q_f$
which are energetically accessible.
In Fig.~\ref{Fig3} the resonance and the continuum contributions to the spectral function
are shown.
\begin{figure}[t]
\vspace{8cm}
\leavevmode
\special{psfile=vmd_fig3.eps vscale=100 hscale=100 voffset=-520 hoffset=-40 angle=0}
\caption{Continuum (dashed line) and resonance (solid line) contributions to the
hadronic spectral function used in the model.}
\label{Fig3}
\end{figure}
In order to evaluate the resonance contribution in Eq.(~\ref{eq5}),
experimental vector-meson proton cross sections $\sigma_{Vp}(k)$ are considered.
In particular, the $\sigma_{\rho p}(k)$ is derived from photoproduction data
on hydrogen \cite{ABB68}.
The $\rho$-meson photoproduction cross section is related to the elastic scattering
of transversely polarized vector meson on nucleons by the VMD relationship
and, through the optical theorem, to the total cross
section $\sigma_{\rho p}$ :
\begin{equation}
\frac{d\sigma}{dt}(\gamma p \rightarrow \rho p)\mid _{t=0}=\frac{\alpha_{em}}{64\pi}\frac{4\pi}{{g_{\rho}}^2}
(1+{\eta_{\rho}}^2)(\frac{q_{\rho}}{q_{\gamma}})^2{\sigma_{{\rho}p}}^2
\end{equation}
where $\eta_{\rho}$ is the ratio of the real to imaginary forward-scattering amplitude,
and $q_{\rho}$ and $q_{\gamma}$ are the center of mass momenta of the $\rho p$ and $\gamma p$ systems at the
same invariant collision energy $\sqrt{s}$ \cite{KON98}.
The values of the $\eta_{\rho}$ and of the effective coupling constant
$\frac{4\pi}{{g_{\rho}}^2}$
are from Ref. \cite{PAU98}, where the effective $\rho$-coupling constant
is reproduced by GVMD with physical coupling
and the non diagonal $\rho p \rightarrow \rho^{'} p$ term.
The $\sigma_{\rho p}$ is shown in Fig.~\ref{Fig4}. As it is seen
$\sigma_{\rho p}$ is higher at low energies; its energy behaviour is
parameterized as
\begin{equation}
\sigma_{\rho p}(k) = p_{1} + \frac{p_{2}}{\sqrt{k}} \, \,,
\label{eq11}
\end{equation}
where $p{_1}=$18 mb and $p{_2}=$27 mb GeV$^{1/2}$.
The cross sections of the higher-mass vector mesons are fixed to
$\sigma_{\omega p}(k)=\sigma_{\rho p}(k)$,
$\sigma_{\phi p} = 12$~mb,
$\sigma_{J/\psi p} = 2.2$~mb and $\sigma_{\psi' p} = 1.3$~mb~\cite{PI95}.
\begin{figure}[ht]
\vspace{8cm}
\leavevmode
\special{psfile=vmd_fig4.eps vscale=95 hscale=100 voffset=-480 hoffset=-40 angle=0}
\caption{ Fit (solid curve) to the ${\rho}$-meson interaction cross section
for the proton ${\sigma_{{\rho}p}}$
derived from Refs.\protect\cite{ABB68} (open circles).
Dashed curve is the continuum interaction cross section ${\sigma_{qp}}$
derived from Eq.(\protect\ref{eq10}).}
\label{Fig4}
\end{figure}
The $\sigma_{qp}$ is determined by the transverse size of the
$q\overline{q}$-fluctuations \cite{PI95}:
\begin{eqnarray}
\sigma_{qp} (\mu^2,k)\!\!\!&=&\!\!\! \int^1_0 \sigma_{qp} (\mu^2,k,\alpha) d \alpha = \nonumber\\
\!\!\!&=&\!\!\! (\!q_{1}\!+\!\frac{q_{2}}{\sqrt{k}}\!)\!\left[\!\frac{8}{\mu^2} ln \!
(\frac{\!1\!+\!x\!}{\!1\!-\!x\!})\! +\! R^2_c (\!1\!-\!x\!)\! \right]
\label{eq10}
\end{eqnarray}
where the integration is performed
over the fraction $\alpha$ of the light-cone momentum carried by the quark \cite{MIR98}.
Here $q_{1}$ and $q_{2}$ are free parameters, $x=\sqrt{1-(\frac{2}{\mu R_c})^2}$ where
$R_c$ is the maximum
transverse size of the $q\overline{q}$-fluctuations.
The continuum contribution $\sigma_{\gamma p}^C$
is derived by fitting to the Eq.(\ref{eq5}) the proton photoabsorption
cross section data \cite{PDG} at photon energy higher than 5 GeV,
where the direct contribution is assumed to be negligible.
The direct contribution $\sigma_{\gamma p}^{dir}$ is calculated as:
\begin{equation}
\sigma_{\gamma p}^{dir}=\sigma_{\gamma p}-\sigma_{\gamma p}^{had},
\label{eq10c}
\end{equation}
where $\sigma_{\gamma p}(k)$ is parameterized as
$\sigma_{\gamma p}=67.7s^{0.08}+129s^{-0.45}$ \cite{DON92}.
In Fig.~\ref{Fig5} the result of the calculation for $\sigma_{\gamma p}$ in the
energy range 1.65 GeV $< k <$ 30 GeV are presented together with the experimental data.
The resonance $\sigma_{\gamma p}^R$ and
the continuum $\sigma_{\gamma p}^C$ contributions
to the total cross section $\sigma_{\gamma p}$ are also given.
The $\rho$-meson accounts for about 85$\%$ of the resonance contribution,
the $\omega$-meson for the 9$\%$ , the $\phi$-meson for the 4$\%$.
The small bump in the calculation that occurs at $k\sim$8 GeV
is due to the opening of charm channels which account for about 1$\%$.
The ${\sigma_{\gamma p}}^{dir}$ contribution to the total cross section
is also shown in Fig.~\ref{Fig5}.
The hadronic and the direct contributions to the total cross section for the neutron case
have been also derived from the deuteron photoabsorption data~\cite{PDG} by using a procedure
similar to the one described for the proton.
This allows to evaluate the isospin weighted nucleon cross sections
($\sigma_{\gamma N}, \sigma_{\gamma N}^{had},
\sigma_{\gamma N}^R, \sigma_{\gamma N}^C, \sigma_{hN}, \sigma_{VN}$ and $\sigma_{qN}$)
for each nucleus.
\begin{figure}[ht]
\vspace{8cm}
\leavevmode
\special{psfile=vmd_fig5.eps vscale=95 hscale=100 voffset=-480 hoffset=-40 angle=0}
\caption{ Predictions of the model (thick solid line) for the photoabsorption cross section
on the proton.
Dotted curves are the hadronic contributions due to resonance (R) and to the continuum (C).
Dashed curves are the individual $\rho$, $\omega$ and $\phi$-mesons contributions.
The contribution from direct processes (dir) is shown as a thin solid curve.
}
\label{Fig5}
\end{figure}
\subsection{Photoabsorption on nuclei}
\indent
The nuclear photoabsorption cross section $\sigma_{\gamma A}$ is written as:
\begin{equation}
\sigma_{\gamma A}(k) = \sigma_{\gamma A}^{dir}(k)+\sigma_{\gamma A}^{had} (k).
\label{eq11b}
\end{equation}
The direct term $\sigma_{\gamma A}^{dir}(k)$ is equal to the
incoherent sum of the corresponding terms on proton and neutron:
\begin{equation}
\sigma_{\gamma A}^{dir}(k)=Z \sigma_{\gamma p}^{dir}(k)+N \sigma_{\gamma n}^{dir}(k).
\label{eq11c}
\end{equation}
The hadronic term is derived by substituting
in Eq.~(\ref{eq1}) the
hadron-proton cross section $\sigma_{h p}$ with the
hadron-nucleus cross section $\sigma_{h A}$:
\begin{equation}
\sigma_{\gamma A}^{had} (k) = 4 \pi \alpha_{em} \int^{s_u}_{s_0} \frac{d\mu^2}{\mu^2}
\Pi (\mu^2) \sigma_{hA} (\mu^2\, , k) \, \,.
\label{eq12}
\end{equation}
Inside the nucleus the intermediate hadronic system undergoes a coherent scattering on bound nucleons.
The interference between
multiple scattering amplitudes reduces the hadron-nucleus cross section $\sigma_{h A}$ compared
to $ A \sigma_{h N}$ thus leading to shadowing.
This process is described by the Glauber-Gribov multiple scattering formalism
\cite{GRI70}.
Considering the scattering on one up to five nucleons, $\sigma_{hA}$ is given by:
\begin{eqnarray}
\lefteqn{\sigma_{hA} (\mu^2\, , k) = A \sigma_{hN} \left [1 - \, a_{2} \, (A - 1) \,
\frac{\sigma_{hN}}{\pi\overline{r^2} }
\, F_2(\epsilon) + \right .}\nonumber \\
&& + \left . a_{3} \, (A - 1) \, (A - 2) \, [\frac{\sigma_{hN}}
{\pi\overline{r^2}}]^2 \,
F_3(\epsilon) - \right .\nonumber \\
&&- \left . a_{4} \, (A - 1) \, (A - 2) \, (A - 3) \, [\frac{\sigma_{hN}}
{\pi\overline{r^2}}]^3 \,
F_4(\epsilon) + \right .\nonumber \\
&&\! +\! \left . a_{5}\! \, (A\! - \!1)\! (A\! - \!2\!) \! (A\! - \!3)\! (A\! -\! 4)
[\frac{\sigma_{hN}} {\pi\overline{r^2}}]^4
F_5(\epsilon) \right ] \! \! ,
\label{eq14}
\end{eqnarray}
where $a_n$ are numerical coefficients which are strongly decreasing with $n$,
$F_n(\epsilon)$ are functions of
$\epsilon (\mu^2\, , k)=\sqrt{\overline{r^2}}/~{\lambda_{h}}$ which
depend on the nuclear density distribution.
The quantity $\overline{r^2}$
is the rms electron-scattering
radius given in Ref.\cite{DEJ74}.
When $\lambda_{h}\!\!\ll\!\!\sqrt{\overline{r^2}}$, $F_n(\epsilon)$
approximately vanish and $\sigma_{h A} = A \sigma_{h N}$.
Otherwise there is shadowing and the shadowing cross section reduction
${\Delta \sigma}(k)$=$\sigma_{\gamma A}^{had}(k) - A \sigma_{\gamma N}^{had}(k)$
is given by
\begin{equation}
{\Delta \sigma} (k) = {\Delta \sigma}^C(k) + {\Delta \sigma}^R(k)
\label{eq16a}
\end{equation}
\noindent with
\begin{eqnarray}
\lefteqn{{\Delta \sigma}^C\,(k)\,= 4\,\,\pi\alpha_{em} A (A - 1)\times} \nonumber \\
\lefteqn{\times \left[ \frac{ a_{2} }{\pi \overline{r^2}} \,
\int^{s_u} _{s_1} \,\frac{d \mu^2}{\mu^2 }\, \Pi^C (\mu^2) \, \Sigma_{qN}^{(2)} (\mu^2,k) \, F_2 (\epsilon)-
\right.} \nonumber \\
&&\!\!\!\!\!\!\!\!\!\!\!\!\!\!a_{3}\frac{(A - 2)}{ [\pi \overline{r^2}]^2}
\, \int^{s_u} _{s_1} \, \frac{d \mu^2}{\mu^2 } \Pi^C (\mu^2)
\Sigma_{qN}^{(3)} (\mu^2,k) \, F_3 (\epsilon)+ \nonumber\\
&&\!\!\!\!\!\!\!\!\!\!\!\!\!\!a_{4}\frac{(A-2)(A-3)}{[\pi \overline{r^2}]^3}\,
\int^{s_u} _{s_1} \! \frac{d \mu^2}{\mu^2 }\! \Pi^C (\mu^2)
\Sigma_{qN}^{(4)} (\mu^2,k) \, F_4 (\epsilon)- \nonumber\\
&&\!\!\!\!\!\!\!\!\!\!\!\!\!\!\left.a_{5}\frac{(A\! -\! 2)\!(A\!-\!3)\!(A\!-\!4)\!}{\!
[\pi\! \overline{r^2}]^4}\!\int^{s_u} _{s_1} \!
\frac{d \mu^2}{\mu^2 }\! \Pi^C\! (\mu^2) \Sigma_{qN}^{(5)}\! (\mu^2,k) \! F_5\! (\epsilon)\!\right]
\nonumber\\
\label{eq16b}
\end{eqnarray}
\noindent and
\begin{eqnarray}
\lefteqn{{\Delta \sigma}^R(k) =4 \pi \alpha_{em} A (A - 1)\times\,} \nonumber \\
\lefteqn{ \times \sum_V \left[\frac{ a_{2}}{ \pi \overline{r^2}}
\int^{s_u}_{s_0} \frac{d \mu^2}{\mu^2 } G_V (\mu^2) \Sigma_{VN}^{(2)}(k) F_2 (\epsilon)-\right.} \nonumber \\
&&\!\!\!\!\!\!\!\!\!\!\!\!\!\! a_{3} \frac{(A - 2)}{ [\pi \overline{r^2}]^2} \int^{s_u}_{s_0} \frac{d \mu^2}{\mu^2 }
G_V (\mu^2) \Sigma_{VN}^{(3)}(k) F_3 (\epsilon)+ \nonumber \\
&&\!\!\!\!\!\!\!\!\!\!\!\!\!\!a_{4} \frac{(A - 2)(A-3)}{ [\pi \overline{r^2}]^3} \int^{s_u}_{s_0} \frac{d \mu^2}{\mu^2 }
G_V (\mu^2) \Sigma_{VN}^{(4)}(k) F_4 (\epsilon)- \nonumber \\
&&\!\!\!\!\!\!\!\!\!\!\!\!\!a_{5}\!\left. \frac{(A\!-\!2)\!(A\!-\!3)\!(A\!-\!4)}{ [\pi \overline{r^2}]^4}\!
\int^{s_u}_{s_0} \!
\frac{d \mu^2}{\mu^2 }\! G_V (\mu^2)\! \Sigma_{VN}^{(5)}(k)\! F_5 (\epsilon)\! \right] \nonumber\\
\label{eq16c}
\end{eqnarray}
where
\begin{equation}
\Sigma_{qN}^{(i)} (\mu^2,k) = \int^1_0 d \alpha \left[\!\sigma_{qN}(\mu^2, \alpha,k)\!\right]^i \, \,,
\label{eq17}
\end{equation}
and for each nucleus
\begin{equation}
\left[\sigma_{qN} (\mu^2,\alpha,k)\right] ^i =\left[\!\frac{Z \sigma_{qp}(\mu^2,\alpha,k) +
N \sigma _{qn}(\mu^2,\alpha,k)}{A}\!\right]^i \,
\label{eq18a}
\end{equation}
\begin{equation}
\Sigma_{VN}^{(i)}(k) = [\sigma_{VN}(k)]^i =\left[ \!\frac{Z \sigma_{Vp}(k) + N \sigma _{Vn}(k)}{A}\!\right] ^i \, \,.
\label{eq18}
\end{equation}
Two different parameterizations of the nuclear density are used in the evaluation of the functions $F_n$, specifically
a Gaussian and a uniform density distributions for light and heavy nuclei, respectively.
In both cases the experimental
average nuclear density and $\overline{r^2}$ values
are well reproduced \cite{MIR98}.
Being each term in Eq.~(\ref{eq14}) proportional to $A^{\frac{n+2}{3}}$,
the third, fourth, and fifth terms
give a non negligible
contribution only for the heavy nuclei.
Then for the light nuclei the first and second terms in Eq.~(\ref{eq14}) are only considered.
The results of the calculation are shown in Figs.~\ref{Fig6} and ~\ref{Fig7} for five nuclei.
The comparison with the data is performed for both the photonuclear cross section and
the ratio between photonuclear and photonucleon cross sections.
As it is seen the results are in slightly better agreement with the data with respect to previous
models \cite{PI95, BO96, ENG97}. However the calculation still shows
a stronger energy dependence than data.
In particular, at low energy it overestimates the experimental result thus suggesting
the need of further mechanisms for the description of the process.
\section{Medium effects on the hadronic spectral function}
\indent
The calculation described in the previous section is based on the assumption that the spectral
function of the hadronic fluctuation of the photon does not
change inside the nuclear medium.
In order to improve the phenomenological description of the low energy
photonuclear data, the effect of the
possible hadronic mass modification in the nuclear
medium is now considered.
The ${\rho}$-meson mass modification in the nuclear medium is predicted by several theoretical
approaches which consider effective chiral Lagrangians,
in-medium scaling properties based on QCD sum rules,
quark bag models combined with quantum hadrodynamics (for a recent review see Ref. \cite{CAS99}).
Many of these theories predict a mass reduction ${\delta} m_{\rho}$ proportional to the
average nuclear density and
amounting up to about 100-200 MeV for the nuclear matter density.
The decrease of the ${\rho}$-meson mass
inside the nucleus increases the coherence length $\lambda_{\rho} = 2k / m_{\rho}^2$
and thus decreases the energy threshold for the shadowing.
Other theories predict a broadening or a complete distortion of the in-medium $\rho$-meson
mass distribution.
Considering a possible $\rho$-meson mass shift
in nuclei, a fit to the photonuclear absorption data is performed
by using the previously described calculation.
In the spectral function $\Pi(\mu^2)$ of Eq.~(\ref{eq12}), the $\rho$-meson mass
$m_{\rho}$ is replaced by $m_{\rho}$ + $\delta m_{\rho}$, with
$\delta m_{\rho}$ free parameter.
In order to reduce the number of free parameters in the fitting procedure,
no mass modifications of other vector-mesons are considered
since their contributions are small.
The fits are shown in Figs.~\ref{Fig6} and ~\ref{Fig7}; the relevant $\chi^{2}$
improves by about a factor of two with respect to the calculations with $\delta m_{\rho}$=0.
It is worth to mention that also a distortion of the $\rho$-meson mass distribution, which
enhances the low mass hadronic spectral function, will result in a better agreement with
the experimental data.
The values of the $\delta m_{\rho}$ obtained from the fits are given in Table~\ref{Table5}:
they range from $-$63 MeV to $-$163 MeV and
the shift in carbon is more than a factor of two bigger than in lead.
The values of the $\delta m_{\rho}$ obtained for the lightest nuclei are
significantly larger than most of the theoretical expectations, while are
in qualitative agreement with
a recent measurement performed via the
${^3}He({\gamma},{\pi}{^+}{\pi}{^-})X$ reaction
\cite{LOL98,HUB98} which suggests a $\sim$ 160 MeV reduction of the ${\rho}$-mass
in $^3$He.
This reduction is so large that
cannot be explained
by the mean field picture of nuclear matter ~\cite{SAI97}.
In this latter reference, unlike all other calculations which consider infinite nuclear matter,
the experimental charge density distributions are used,
resulting in a shift in
$^4$He about a factor of two bigger than in $^{12}$C due to the higher $^4$He core
density.
Also a recent calculation that accounts for the local density distributions in
$^3$He, shows a substantial changes in the $\rho$-meson mass \cite{BHA99}.
In this respect the large shift observed in the fit of the light nuclei photoabsorption
data could be ascribed to the high core density, while
for the heavier nuclei the mass-shifts agree with the theoretical predictions
which account for the mean nuclear field alone.
Moreover, the higher local density distributions
for the lighter nuclei can reduce
the local intranucleon distance $d_{NN}$, thus accounting for an
earlier onset of the shadowing effect on these nuclei.
\begin{table}[h]
\caption{$\rho$-meson mass shifts $\delta m_{\rho}$ extracted from the photoabsorption data
fits. The errors
indicate the statistical and the systematic uncertainties.}
\begin{center}
\begin{tabular}{lc@{\hspace{1cm}}lc} \hline \hline
Nucleus & $\delta m_{\rho}[MeV]$ & Nucleus & $\delta m_{\rho}[MeV]$ \\
\hline
C & -163 $\pm 14$ $\pm 50$ & Sn & -115 $\pm 17$ $\pm 53$ \\
Al & -133 $\pm 11$ $\pm 40$ & Pb & -63 $\pm 20$ $\pm 62$ \\
Cu & -104 $\pm 14$ $\pm 57$ & & \\
\hline \hline
\end{tabular}
\label{Table5}
\end{center}
\end{table}
\section{Conclusions}
Total photoabsorption cross sections for nucleon and nuclei are calculated in the
energy range 1.65-30 GeV.
The process is described taking into account both the direct and the hadronic fluctuation
interactions of the photon. The latter is computed with a hadronic spectral function
which includes the effective $\rho$-coupling constant, the finite width of vector-meson resonances and the
quark-antiquark continuum.
Realistic and energy dependent interaction cross section for the $\rho$-meson is derived
from photoproduction data.
The shadowing effect is evaluated in the framework of a Glauber-Gribov
multiple scattering theory up to the 5$^{th}$ order.
The low energy onset of the shadowing effect is interpreted as a possible
signature of a modification of the hadronic spectral function in
the nuclear medium. In particular, a decrease of the $\rho$-meson mass in nuclei
is suggested for a better description of the experimental data.
This reduction is larger for the light nuclei and cannot be accounted for by mean
field consideration alone.
\section{Acknowledgments}
We would like to express our gratitude to
K. Saito and A. Sibirtsev for useful discussions, and to
A. Bhattacharyya for providing us with results prior of publication.
|
\section*{Introduction}
In a recent Physical Review Letter~\cite{BOROS},
Boros {\em et al.}
proposed a model in which
a substantial charge symmetry violation (CSV) for parton
distributions in the nucleon accounts for the experimental
discrepancy between neutrino
(CCFR)~\cite{CCFR} and muon (NMC)~\cite{NMC} nucleon structure
function data at low $x$.
Charge symmetry (sometimes also referred to as isospin symmetry)
is a symmetry which interchanges protons
and neutrons, thus simultaneously interchanging up and
down quarks, which implies the equivalence between the up (down) quark
distribution in the proton and the down (up) quarks in the neutron.
Currently, all fits to Parton Distribution
Functions (PDFs) are preformed under the assumption of
charge symmetry between neutrons and protons.
Boros {\em et al.} have proposed~\cite{BOROS} that charge symmetry
is broken such that
the $d$ sea quark distribution
in the nucleon is larger than the $u$ sea quark distribution
for $x<0.1$, which also results in a violation of flavor symmetry.
Their paper notes
that structure functions extracted in neutrino deep
inelastic scattering experiments are dominated by
the higher statistics data taken with neutrino (versus antineutrino)
beams. They note that neutrino-induced charged current interactions
couple to $d$ quarks and not to $u$ quarks, while the muon coupling to
the 2/3 charged $u$ quark is much larger than the coupling to
the 1/3 charge $d$ quark. Therefore,
if the $d$ sea quark distribution
is significantly larger than the $u$ sea quark distribution
in the nucleon, there would be a significant difference between
the nucleon structure functions as measured in neutrino and
muon scattering experiments. However, both neutrino and muon
scattering data have been taken on approximately isoscalar targets,
such as iron or deuterium.
Isoscalar targets have an equal number of neutrons and
protons.
A larger number of $d$ sea quarks than $u$ sea quarks
in an isoscalar target
implies a violation of charge symmetry. Therefore,
Boros {\em et al.} proposed that a large charge symmetry
violation of the sea quarks in the nucleon might explain
the observed discrepancy $(10\sim15 \%)$ between neutrino and muon structure
function data.
Boros {\em et al.} define the following charge
symmetry violations in the nucleon sea.
\begin{eqnarray}
\delta \ubar (x) & = \ubar^p(x) - \overline{d}^n(x), \\
\delta \overline{d} (x) & = \overline{d}^p(x) - \ubar^n(x),
\label{eq:delta}
\end{eqnarray}
where $\ubar^p(x)$ and $\overline{d}^p(x)$
are the distribution of the
$u$ and $d$ sea anti-quarks in the proton, respectively.
Similarly $\ubar^n(x)$ and $\overline{d}^n(x)$
are the distribution of the
$u$ and $d$ sea anti-quarks in the neutron, respectively.
The distributions for the quarks and antiquarks in
the sea is assumed to be the same.
The relations for CSV in the sea quark distributions
are analogous to equations (1) and (2) for the sea anti-quarks.
Charge symmetry in the valence
quarks is assumed to be conserved, since there is good
agreement between the neutrino and muon scattering
data for $x>0.1$.
Within this model,
Boros {\em et al.} extract a large CSV
from the difference in structure functions
as measured in neutrino and muon scattering
experiments.
Theoretically, such a large charge
symmetry violation (of order of 25\% to 50\%)
is very unexpected.
Therefore, the article has generated
a significant amount of interest
both within and outside
the high energy physics community~\cite{science}.
If the proposed model is valid,
all parametrizations of PDFs would have to be modified. In addition,
physics analyses which rely on the knowledge of PDFs (e.g. the
extraction of the electro-weak mixing angle from the ratio
of neutral current and charged current cross sections) would be significantly
affected.
In this communication we show that the CSV models proposed
by Boros {\em et al.} are
ruled out by the $W$ charge asymmetry measurements
made by the CDF experiment at the Fermilab
Tevatron collider~\cite{CDF}. These $W$ data provide a
very strong constraint on the ratio of
$d$ and $u$ quark momentum distributions
in the proton over the $x$ range of $0.006$ to $0.34$.
Figure~\ref{fig:DELTA} shows the quantity
$x\Delta (x) = x[\delta \overline{d}(x) - \delta \ubar(x)]/2$ required
to explain the difference between neutrino and muon data,
as given in Fig. 3 of Boros {\em et al.}~\cite{BOROS}.
The average $Q^2$ of these data is about 4 (GeV$/c$)$^2$.
The dashed line is the strange sea quark distribution [$xs(x)$]
in the nucleon as measured by the CCFR collaboration using
dimuon events produced in neutrino nucleon interactions.
Boros {\em et al.} state that
the magnitude of implied charge symmetry violation is somewhere between
the full magnitude of the strange sea and half the magnitude
of the strange sea. Since the strange sea itself has been
measured to be about half of the average of the $d$ and $u$
sea, this implies a charge symmetry violation of
order 25\% (at $x=0.05$) and 50\% (at $x=0.01$).
However, as can be seen in Fig.~\ref{fig:DELTA}, the
shape of the strange sea does not provide a good
parametrization of the charge symmetry violation, therefore,
we have parametrized $x\Delta(x) $ (at $Q^2 =
4$ (GeV$/c$)$^2$ ) as follows. For $x > 0.1$,
$x\Delta(x) = 0.$ For $x < 0.01$,
$x\Delta(x) =0.15$, and for
$0.01< x < 0.1$,
$x\Delta(x) = .15[log(x)-log(.1)]/[log(.01)-log(.1)]$.
This parametrization is shown as the solid line in
Fig.~\ref{fig:DELTA}.
The dot-dashed line shows the value of our parametrization
when evolved to $Q^2 = M_W^2$.
Boros {\em et al.}
suggest that it is theoretically
expected that
$\Delta(x) = \delta \overline{d} (x) = - \delta \ubar (x)$,
which means that the sum of $u$ and $d$ sea distributions
for protons and neutrons is the same. Within the assumption that
$\Delta(x) = \delta \overline{d} (x) = - \delta \ubar (x)$, we use
two models to parametrize the range of allowed changes in PDFs
to introduce the proposed charge symmetry violations.
\begin{figure}[t]
\centerline{\psfig{figure=csv_delta_v2.ps,width=3.0in,height=3.0in}}
\caption{Charge symmetry violating distribution,
$x\Delta (x)$ = $x (\delta \overline{d} (x) - \delta \ubar (x))/2$ required
to explain the difference between neutrino and muon data,
as given in Fig. 3 of Boros {\it et al}.
The dashed line is the strange sea quark distribution
in the nucleon [$xs(x)$] as measured by the CCFR collaboration using
dimuon events produced in neutrino nucleon interactions.
The solid line is our parametrization at $Q^2 =
4$ (GeV$/c$)$^2$ as described in the text. The dot-dashed line
is our parametrization when evolved to $Q^2 = M_W^2$. }
\label{fig:DELTA}
\end{figure}
In Model 1, it
is assumed that the standard PDF parametrizations
are dominated by
neutrino data and therefore represent the average
of $d$ and $u$ sea quark distributions. Therefore,
half of the CSV
is introduced into the $u$ sea quark distribution and
half of the effect is introduced into the $d$ sea
quark distribution such that the average of $d$ and $u$ sea
quark distributions is unchanged.
\begin{eqnarray}
\ubar^p(CSV) & = \ubar^p-\Delta(x)/2, \\
\overline{d}^p(CSV) & = \overline{d}^p+\Delta(x)/2, \\
\ubar^n(CSV) & = \ubar^n-\Delta(x)/2, \\
\overline{d}^n(CSV) & = \overline{d}^n+\Delta(x)/2.
\label{eq:model1}
\end{eqnarray}
In Model 2,
it is assumed that standard PDFs are dominated
by muon scattering data, and therefore are good
representation of the 2/3 charge $u$ quark distribution.
In this model, the entire effect is introduced into the
$d$ sea quark distribution as follows;
\begin{eqnarray}
\ubar^p(CSV) & = & \ubar^p, \\
\overline{d}^p(CSV) & = & \overline{d}^p+\Delta(x),\\
\ubar^n(CSV) & = & \ubar^n,\\
\overline{d}^n(CSV) & = & \overline{d}^n+\Delta(x).
\label{eq:model2}
\end{eqnarray}
Model 2 would change the total quark sea.
In order to have a precise test for the CSV effect, all PDFs
have to be refitted based on the above two models. However, the ratio of $d$
and $u$ distribution will be almost
the same whether we refit the PDFs or not.
The $d/u$ ratio which has been extracted
from $F_2^n/F_2^p$ measurements (assuming charge symmetry)
is in fact the quantity $u^n/u^p$ which does
not have any sensitivity to the proposed CSV effect.
In order to test for CSV effects, measurements of
$d^p/u^p$ or $d^n/u^n$ are required.
Therefore, the CDF measurements of the $W$ charge asymmetry
in $p\overline{p}$ collisions provide a unique test
of CSV effects, because of the direct sensitivity of these
data to the $d/u$ ratio in the proton (note that the $d$ and
$u$ quark distributions at small $x$ are dominated by
the quark-antiquark sea).
We now proceed to show that these implementations
of CSV in the nucleon sea are
ruled out by the CDF $W$ charge asymmetry measurements
at the Tevatron.
At Tevatron energies, $W^+$ ($W^-$) bosons are produced in
$p\overline{p}$ collisions primarily by the annihilation of
$u$ ($d$) quarks in the proton and $\overline{d}$ ($\overline{u}$)
quarks from the antiproton. Because $u$ quarks carry on average
more momentum than $d$ quarks~\cite{CTEQ4M},
the $W^+$ bosons tend to follow the direction
of the incoming proton and the $W^-$ bosons' that of the antiproton.
The charge asymmetry in the production of $W$ bosons as a function
of rapidity ($y_W$) is therefore
related to the difference in
the $u$ and $d$ quark distributions, and
is roughly proportional~\cite{ELB}~\cite{ADM}
to the ratio of the difference and the sum of the
quantities $d(x_1)/u(x_1)$ and $d(x_1)/u(x_2)$, where
$x_1$ and $x_2$ are the fractions of the proton momentum carried
by the $u$ and $d$ quarks, respectively. (Note that
the quark distributions in the proton are equal to the
antiquark distributions in the antiproton). At large rapidity,
$x_1$ is larger than 0.1, which is a region
where CSV does not exist. On the other hand $x_2$
is in general less than 0.1, and a 25\% to 50\%
CSV effect would imply
a very large effect on the $W$ asymmetry. Since the $W$ charge
asymmetry is sensitive to the $d/u$ ratio, it does not
matter if the CSV effect at small $x$ is present in either $d$
or $u$ sea quark. All of these models would result in
a similar change in the $W$ asymmetry.
Experimentally, the $W$ rapidity is not determined
because of the unknown longitudinal momentum of the neutrino from the $W$
decay. What is actually measured
by the CDF collaboration
is the lepton charge asymmetry which is a convolution of
the $W$ production charge asymmetry and the well known asymmetry from the
$V$-$A$ $W$ decay.
The two asymmetries are in opposite directions
and tend to cancel at large values of rapidity.
However, since the
$V$-$A$ asymmetry is well understood, the lepton asymmetry is still sensitive
to the parton distributions.
The lepton charge asymmetry is defined as:
\begin{equation}
A(y_l)=\frac{d\sigma^+/dy_l-d\sigma^-/dy_l}
{d\sigma^+/dy_l+d\sigma^-/dy_l},
\end{equation}
where $d\sigma^+$ ($d\sigma^-$) is the cross section for
$W^+$~($W^-$) decay leptons as a function of lepton rapidity, with positive
rapidity being defined in the proton beam direction.
The CDF data~\cite{CDF} shown in
Fig.~\ref{fig:WASYM} span the broad range of lepton
rapidity ($0.0<|y_l|<2.2$),
and provide information about the $d/u$ ratio in the proton
over the wide $x$ range ($0.006<x<0.34$).
Therefore, the CDF $W$ asymmetry data
would provide a strong tool to test the CSV models over a broad range
of $x$, and not just in part of the range proposed in the Boros {\em et al.}
model.
\begin{figure}[t]
\centerline{\psfig{figure=csv_wasym_50_v2.eps,width=3.0in,height=3.0in}}
\caption{ The CDF $W$ Asymmetry data. The solid line is the
prediction from the standard CTEQ4M PDF(CSV=0).
The dashed-dotted line is the CTEQ4M PDF modified
for larger $d$ quark distribution at large $x$
as proposed by Yang and Bodek(CSV=0). The dashed and dotted
lines are predictions from the CTEQ4M PDF modified
to include the Boros {\em et al.} charge symmetry violation
in the quark sea as described in the text. All theoretical predictions
are calculated in NLO QCD using the DYRAD program. }
\label{fig:WASYM}
\end{figure}
Also shown in Fig.~\ref{fig:WASYM} (solid line)
are the predictions for the $W$ asymmetry from QCD
calculated to Next-to-Leading-Order (NLO)
using the program DYRAD~\cite{Dyrad}, with
the CTEQ4M PDF~\cite{CTEQ4M} parametrization
for the $d$ and $u$ quark distributions
in the proton ( we have used CTEQ4
because it is the PDF set that has been used by Boros {\em et al.}
in their paper ).
As pointed out by Yang and Bodek~\cite{YANG},
the small difference between the data and the prediction of the
CTEQ4M PDF at high rapidity is because
the $d$ quark distribution is somewhat underestimated
at high $x$ in the standard PDF parametrizations.
The predictions of the CTEQ4M PDF with the proposed
modifications by Yang and Bodek are shown as the dashed-dotted
line in the figure.
The two dotted lines in
Fig.~\ref{fig:WASYM} show
the predicted $W$ asymmetry for the
CTEQ4M PDF with the proposed Boros {\em et al.} charge symmetry
violation in the sea for Model 1 and
Model 2, respectively. The
CDF $W$ data clearly rule out these models.
Most striking in this analysis is the broad range of lepton
rapidity over which this disagreement occurs with the
CSV models. This is suggestive that models of this class
would be ruled out over a broad range of $x$, and not just
in part of the range proposed in the Boros {\em et al.}
model.
In the direct measurement of the $W$ mass at the Tevatron,
the CDF $W$ asymmetry data have been used to limit the
error on $M_W$ from PDFs to about 15 MeV. This has
been done by calculating
the deviation between the error weighted
average measured asymmetry over the
rapidity range of the data, and the predictions from various
PDFs. This measured average asymmetry for
the data is $0.087\pm0.003$. The predicted average asymmetries
(weighted by the same errors as the data) are 0.094, 0.125, and
0.141 for the CTEQ4M PDF, and for Model 1 and Model 2, respectively.
If we only accept PDFs which are within two standard deviations
of the CTEQ4M PDF, the $W$ asymmetry data rule out CSV effects
at the level of more than 10 standard deviations for the two models
with CSV effects.
Another precision measurement which is sensitive to CSV effects is the
measurement of neutral-current scattering in neutrino-nucleon collisions.
Just as the magnitude of the couplings to $u$ and $d$ quarks differ in
neutral-current $\mu$--$q$ scattering at NMC, the couplings
to $u$ and $d$ quarks also differ in
neutral-current $\nu$--$q$ scattering. In this case, the left-handed and
right-handed couplings of the neutral current to quarks are given by
$g_L=I_3-Q\sin^2\theta_W$ and $g_R=-Q\sin^2\theta_W$, where $Q$ is the quark
charge and $I_3$ is the third component of the weak isospin in the quark
doublet, $+1/2$ for $u$-type quarks and $-1/2$ for $d$-type quarks.
Therefore the CSV-inspired enhancement in the
$d$ quark distributions will
change the the cross-section for neutral-current scattering, even
for an isoscalar target. Because these cross-section
measurements are used to extract electroweak parameters, a CSV effect could
then affect the precision measurements of $\sin^2\theta_W$.
The most precise measurements of neutral-current neutrino-quark scattering
come from the CCFR~\cite{CCFR-NC} and NuTeV~\cite{NuTeV-NC} experiments. As
noted above, CCFR had a beam of mixed neutrinos and anti-neutrinos, dominated
by neutrinos. The NuTeV experiment
uses separate neutrino and anti-neutrino beams in its
measurements to allow separation of neutral-current neutrino and
anti-neutrino interactions. The NuTeV and CCFR experiments
measure combinations of the
ratios, $R^\nu$ and $R^\nub$, (above a fixed hadron energy
$\nu$
threshold of $20$~GeV
or $30$~GeV in NuTeV and CCFR, respectively), where
\begin{equation}
R^{\nu(\nub)}\equiv
\frac{\sigma_{\nu(\nub)}^{\rm NC}}{\sigma_{\nu(\nub)}^{\rm CC}}=
\frac{1}{2}-\sin^2\theta_W+\frac{5}{9}(1+r^{\pm 1} )\sin^4\theta_W,
\end{equation}
and $r\equiv\sigma_\nub^{\rm CC}/\sigma_\nu^{\rm CC}\sim0.4$. The NuTeV
experiment has
extracted $\sin^2\theta_W$ using the combination $R^\nu-rR^\nub$ which is
insensitive to the effects of sea quarks, and thus not changed by CSV effects
in sea, as in the Boros {\em et al.\,} model. However, the CCFR measurement
with a mixed beam
is equivalent to $R^{\nu}+0.13R^{\nub}$ in which the sea quark
contributions do not cancel.
Within the framework of Model 1, the modified PDFs leave the charged
current neutrino data unchanged, but affect the level of
the neutral current cross section.
The effect of the Model 1 implementation of the Boros {\em et al.}
model on the CCFR result has been calculated using the CTEQ4L
PDF~\cite{CTEQ4M} in the cross-section model. The CCFR experiment
extracts a
$\sin^2\theta_W$ which is equivalent to
$M_W=80.35\pm0.21$ GeV~\cite{CCFR-NC}, which can be compared to the current
average of all direct $M_W$ measurements, $80.39\pm0.06$ GeV. Model 1 would
increase the CCFR measured $M_W$ by $0.26$~GeV.
Since the CDF $W$ asymmetry data rule out a CSV effect at the
the level of 1/5 of the magnitude of Model 1, the error from possible CSV
effects in PDFs is less than 50 MeV. This illustrates
the value of the CDF $W$ asymmetry
data in limiting the systematic error from PDF uncertainties
not only in the direct measurement of the $W$ mass in hadron colliders,
but also in the indirect measurement of the $W$ mass in neutrino
experiments.
In conclusion, the CDF $W$ asymmetry data rule out the Boros {\em et al.}
model for charge
symmetry violation in parton distributions~\cite{note} as the
source of the difference between neutrino (CCFR) and
muon (NMC) deep inelastic scattering data. Sources such as
a possible difference in nuclear effects between neutrino
and muon scattering, or a possible underestimate of the
strange quark sea in the nucleon have been ruled out~\cite{BOROS}.
The experimental systematic errors between the two experiments,
and improved theoretical analyses of massive charm
production in both neutrino and muon scattering are both
presently being investigated~\cite{private}
as possible reasons for this discrepancy.
|
\section{Introduction}
The ultimate goal in the study of collisions of heavy ions at the
highest available energies is the production and characterization of an
extended volume of deconfined quarks and gluons, the quark gluon plasma
(QGP)\cite{qm97}.
Due to the high multiplicities in central Pb+Pb collisions at 158~GeV per Nucleon,
recorded in the NA49 large acceptance spectrometer,
a statistically significant determination
of momentum space distributions and particle ratios can be performed
for single events, allowing for a study of
event-by-event fluctuations\cite{sto95,rol97}. In this paper we will
focus on fluctuations in the average transverse momentum of individually measured
charged particles from event to event.\\
One expects that the fluctuation patterns
are altered in the vicinity of the QCD phase transition \cite{sto95,mrow93}.
This conjecture is supported by recent calculations in an effective
model of the strong interaction \cite{step98,stephanov99}, which suggest
that near a tri-critical point in the QCD phase diagram the event-by-event fluctuation
pattern in transverse momentum should change significantly. \\
A precise measurement of event-by-event fluctuations allows for a
test of the hypothesis of thermal equilibrium \cite{gazd92}
and the extraction of thermodynamical properties of the system in a model comparison.
Model studies\cite{random} have shown that
non-equilibrium models of nuclear collisions based purely on
initial state scattering can be tested by measurements of
transverse momentum fluctuations. Model calculations on transverse
momentum fluctuations have been performed in many of the commonly
used microscopic models of nuclear collisions \cite{bleicher98,liu98,cap99},
in particular focussing on the question of how the fluctuations
change when going from nucleon-nucleon to nucleus-nucleus collisions.\\
It has also been suggested that for a thermodynamical picture of the
strongly interacting system formed in the collision, the strength of
fluctuations is directly related to fundamental properties of the
system like the specific heat \cite{stod95,shur98a}
and matter compressibility \cite{mrow98a}. A detailed discussion
of transverse momentum fluctuations in a resonance gas model can
be found in \cite{stephanov99}.\\
One of the most intensely discussed topics related to fluctuations
at the QCD phase transition is the formation of so-called disoriented
chiral condensates (DCCs) \cite{raja93,anselm91} as a consequence of
the transient restoration of chiral symmetry, which may lead to a production of
pions with much larger fluctuations of the charged-to-neutral pion ratio than
expected from Poisson-statistics.
The sensitivity of our measurement to these
fluctuations is discussed.\\
NA49 is currently pursuing two different,
but complementary approaches to the characterization of the
single events. In the approach presented here
we characterize the event by global observables
like the mean transverse momentum of individually detected charged particles in the event,
averaging over a large interval in momentum space.
Global quantities in general also include contributions
from particle correlations occuring at smaller scales, i.e.
smaller intervals in momentum space.
NA49 is also studying a system of differential measures of
event morphology which aim at a multiscale characterization
of the correlation content of single events, which should eventually
provide a decomposition of the global fluctuations as a
function of scale \cite{train98b}.
\section{Experimental Setup and Data selection}
The setup of the NA49 experiment is described in \cite{nimpaper}.
We used a data set of central Pb+Pb collisions that where
selected by a trigger on the energy deposited in
the NA49 forward calorimeter. The trigger accepted only the 5\% most
central events, corresponding to an impact parameter range of $b < 3.5$~fm.
The event vertex was reconstructed using information from beam position
detectors and the fit of the measured particle trajectories. Only events
uniquely reconstructed at the known target position were used.
The NA49 large acceptance hadron spectrometer allows the detection of
more than 1000 individual charged particles for a single central Pb+Pb
collision.\\
In this analysis particles were selected that had a measured track length of
more than 2~m in one of the two Main Time Projection Chambers (MTPC) outside
the magnectic field and were also observed in at least
one of the Vertex TPC's inside the superconducting magnets.
We studied particles in a region of $0.005 < p_T < 1.5$~GeV/c
and rapidity $4 < y_{\pi} < 5.5$. A cut on the extrapolated impact parameter
of the particle track at the primary vertex
was used to reduce the contribution
of non-vertex particles originating from weak decays and secondary
interactions. We estimate that about 60~\% of such particles are
rejected by the vertex cuts.
From a full simulation of our apparatus using a GEANT \cite{geant} based
Monte-Carlo code and a parametrization of the detector response
we obtained an average reconstruction efficiency of 90\%. The
average resolution in transverse momentum for the particles used here
is around 3~MeV/c, dominated by multiple Coulomb scattering.
The two-track resolution was determined using both the
simulation and a mixed-event technique. For particles selected
by the track cuts the pair
detection efficiency drops from around 80\% at an average distance in
the Main TPC of $d = 2.5$~cm to around 20\% at an average distance
of $d = 1.5$~cm.
In table~1 the most important parameters of the inclusive
and event-by-event distributions of accepted particles are summarized.
Throughout this paper brackets ($\langle x \rangle$) will
denote averages over events and bars ($\overline{x}$) will denote inclusive averages over
all (accepted) particles and all events.
\begin{table}[t]
\caption{Measured parameters of the inclusive and event-by-event accepted particle distributions.
Errors are statistical only.}
\begin{tabular}{|l|l|}
No. of events & 98426 \\
\tableline
$\langle N\rangle$ & $ 270.13 \pm 0.07 $ \\
\tableline
$(\langle N^2\rangle - \langle N\rangle^2)^\frac{1}{2}$ &$23.29 \pm 0.05 $ \\
\tableline
$\overline{p_T}$ & $376.75 \pm 0.06$~MeV/c \\
\tableline
$(\overline{p_T^2} - \overline{p_T}^2)^\frac{1}{2}$ & $282.2 \pm 0.1$~MeV/c \\
\end{tabular}
\end{table}
\section{Analysis and results}
For each of the events we characterize the observed particle distribution in the
acceptance region by calculating the mean of the transverse momentum
distribution of the $N$ accepted particles in the event,
\begin{equation}
M(p_T) = 1/N \cdot \sum_{i=1}^{N} p_{Ti}.
\end{equation}
The resulting distribution of $M(p_T)$ is shown in fig.~1. \\
The distribution of $M(p_T)$ has approximately Gaussian shape.
No significant excess of 'anomalous' events outside the main distribution
is observed.
For the variance of
the $M(p_T)$ distribution we get
\begin{eqnarray*}
V(M(p_T))/\overline{p_T} = 4.65 \pm 0.01\%.
\end{eqnarray*}
The biggest contribution to the observed variance is expected to come from
finite-number statistics. The main task in the remainder of the paper will
be to extract possible non-statistical contributions
on top of the trivial statistical variation from event to event.
A first impression of the possible size of non-statistical contributions can
be obtained by a comparison to the same distribution calculated for
so-called mixed events
(solid line in fig.~1). The mixed events were constructed by combining
particles drawn randomly from different events while reproducing
the multiplicity distribution of the real events.
\noindent Only one track
of any original event was used in a given mixed event and no further selection
was made regarding the impact parameter or multiplicity of the
original events.
By construction the mixed events have the same single-particle distributions as the real
events, but no internal correlations. The variance of the mixed event $M(p_T)$
distribution is therefore determined by finite number statistics, giving
\begin{eqnarray*}
(\overline{p_T^2} - \overline{p_T}^2)^\frac{1}{2}/
( \overline{p_T} \cdot \sqrt{\langle N \rangle}) = 4.6\%.
\end{eqnarray*}
The mixed event distribution resembles, very closely, the single event distribution,
thus suggesting that large amplitude non-statistical fluctuations are small and/or
rare.\\
To further quantify and study the deviation of the $M(p_T)$ distribution
a number of methods has been discussed recently \cite{kadija92,bialas99,alberico99,belkacem99}.
In this analysis we follow the approach suggested in \cite{gazd92}.
We define for every particle $i$
\begin{equation}
z_i = p_{Ti} - \overline{p_T} .
\end{equation}
For every event we calculate
\begin{equation}
Z = \sum_{i=1}^{N} z_i .
\end{equation}
With this definitions
we use the following measure to quantify the degree of fluctuation
in mean transverse momentum from event to event:
\begin{equation}
\Phi_{p_T} = \sqrt{\frac{\langle Z^2 \rangle}{\langle N \rangle}} - \sqrt{\overline{z^2}}.
\end{equation}
One limiting case for this fluctuation measurement is particle emission according
to a parent distribution that
remains unchanged for all events, i.e.\
every single event is just a
random sample of finite multiplicity taken from the same parent
distribution. $\Phi_{p_T}$ was defined such that for this case a
value of zero is assumed. This value also corresponds to the fluctuations
for an ideal gas of classical particles\cite{mrow98b}. \\
\begin{figure}[t]
\centerline{\epsfig{file=figure1.eps,height=7.5cm }}
\caption{Event-by-event distribution of the mean transverse
momentum $M(p_T)$ of accepted particles in the event (points). For
comparison, the solid line shows the $M(p_T)$ distribution
for mixed events.}
\label{1}
\end{figure}
Our goal is to detect or exclude fluctuations that are
compatible with changes, event-to-event, in the
parent distribution in transverse momentum.
Such changes would in general lead
to values of $\Phi_{p_T} > 0$. $\Phi_{p_T}$, as we will demonstrate
later, is also sensitive to internal correlations or anti-correlations
of particles within single events, which result in $\Phi_{p_T} > 0$
or $\Phi_{p_T} < 0$, respectively.\\
In our data set we measure a value of
\begin{eqnarray*}
\Phi_{p_T} = 0.6 \pm 1.0~\mbox{MeV/c},
\end{eqnarray*}
compatible with zero.
The error was estimated by calculating $\Phi_{p_T}$ separately
on independent subsamples of approximately 10000 events each.
Before discussing the implications of the small value measured
for $\Phi_{p_T}$ for the existence of collective non-statistical
fluctuations, we first turn to investigating the sensitivity of
this measure to correlations at small scales, which are known to
be present in these events.
Previous studies \cite{hbtpaper} have quantitatively analyzed two-particle
correlations in relative momentum, with quantum statistics and
Coulomb final state interactions giving the strongest contributions.
We have developed an analysis procedure to identify the contribution
of these known effects, folded with the NA49 experimental response,
to the observed value of $\Phi_{p_T}$.
This comparison is done as follows.
\begin{enumerate}
\item From the original events we construct a sample
of mixed events with the same overall multiplicity
distribution as the real events.
Without further modifications, the mixed events give a
value of $\Phi_{p_T} = -0.2 \pm 0.4$~MeV/c, again consistent with zero.
\item In the second step we model
the contributions from particle pair
correlations at small relative momenta ('small scales').
The effect of Bose-Einstein or Fermi-Dirac
statistics on transverse momentum fluctuations has been discussed in
\cite{mrow98b}. \\
The effects of quantum statistics and final-state (Coulomb) interactions
partially cancel each other and are further diminished by particles emitted
from long-lived resonances, particles originating from weak decays
and by including combinations of non-identical particles.
A detailed evaluation of two-particle correlations in NA49 is
given in \cite{hbtpaper}.
To include the sum of all these effects in the mixed events
we use a procedure that was described in \cite{kad92}. In this procedure
the momenta of particles are altered pairwise to introduce the desired
form of the two-particle correlation function, making sure
that in the mixed events on the average the two-particle correlations as a function
of relative momentum
closely match those
observed in the data, only corrected for the two-track resolution.
The analysis of the modified mixed events provides an estimate of the
minimal event-by-event $M(p_T)$ fluctuations that we expect
as a consequence of the observed average two-particle correlations.
The contribution from the two-particle correlation function
alone is $\Delta \Phi_{p_T} = 5 \pm 1.5$~MeV/c.
\item In our data set we observe a slight but statistically significant
correlation between
the multiplicity of the event $N$ and the average transverse momentum
$M(p_T)$. The correlation is characterised by a linear correlation
coefficient of
\begin{eqnarray*}
\frac{\langle (M(p_T) - \overline{p_T}) \cdot (N - \langle N ) \rangle}
{V(M(p_T)) \cdot V(N)} = -0.03 \pm 0.01,
\end{eqnarray*} where $V(N)$ is the variance of the multiplicty distribution. We therefore
observe a slight decrease of $M(p_T)$
with increasing multiplicity within the central collision data set.
Introducing this correlation in the
mixed events gives a negligible contribution to the width of the
$M(p_T)$ distribution ($\Delta \Phi_{p_T} \ll 1$~MeV/c).
\item We then apply an experimental filter on each of the modified mixed events
that simulates the influence of the two-track resolution and momentum
resolution of the NA49 apparatus. While the contribution from
momentum resolution is found to be negligible for the range of
fluctuations considered here, the two-track resolution results in an
effective anti-correlation between particles in momentum space
and gives a contribution of $\Delta \Phi_{p_T} = -4 \pm 0.5$~MeV/c.
\end{enumerate}
Combining all effects, we find that the observed value $\Phi_{p_T}$ is
compatible with independent particle production:
Including both the effects of two-particle correlations in momentum
space and the experimental two-track resolution leads to a
cancellation resulting in a very small net contribution.
Any additional contributions beyond those mentioned above either have
to be small or cancel with sufficient accuracy to be compatible
with $\Phi_{p_T} = 0.6$~MeV/c. \\
Here it is worthwhile to note that the effects of two-particle
momentum correlations and two-track resolution, as included
in our simulations, both are strongly multiplicity-dependent
and become negligible for multiplicities comparable to those
observed in p+p collisions. This suggests that the physical origin of
transverse momentum fluctuations as measured using $\Phi_{p_T}$
changes when comparing the value of 0.6~MeV/c for Pb+Pb to the
preliminary NA49 measurement of $5\pm 1$~MeV/c for p+p collisions
\cite{roland98}.
To further study the sensitivity of our
measurement we have introduced explicit non-statistical fluctuations
in the mixed event sample and studied the response in $\Phi_{p_T}$
as a function of the parameters controlling the strength of these
fluctuations. By comparing the value obtained for
$\Phi_{p_T}$ observed in such models with that in the data, we
can determine the sensitivity of our measurement to various
kinds of fluctuations and eventually derive limits on the
amplitude or frequency of occurence for fluctuations in specific
models.\\
In the first model we examine the sensitivity to non-statistical
fluctuations introduced by scaling the transverse momentum for
all tracks in a given event by a constant factor $x$. The resulting
change in the $p_T$ parent distribution from event to event resembles
that of an event by event change in the inverse slope parameter of
an exponential transverse momentum distribution.
We obtain a random number $x$ for each event distributed
according to
\begin{equation}
P(x) = \frac{1}{\sqrt(2 \pi \sigma_{fluc}^2)} \exp(-(x-1)^2/(2 \sigma_{fluc}^2))
\end{equation}
and multiply the transverse momentum of each particle in the event by
the same factor $x$. Here the amplitude of the event by event fluctuation
is controlled by the parameter $\sigma_{fluc}$.
Adding these fluctuations to the mixed events and neglecting
two-particle correlations and two-track resolution, $\Phi_{p_T}$ increases
proportionally to $\sigma_{fluc}^2$.\\
When adding all effects, we find that for this model a fluctuation strength
$\sigma_{fluc} = 1.2$\% corresponds to $\Phi_{p_T} = 7$~MeV/c. Given
the definition of $\Phi_{p_T}$ in Eq.~4 and net contributions to $\Phi_{p_T}$ as
observed and simulated in steps
1 to 4 above we can establish an upper limit on $\sigma_{fluc} < 1.2\%$ at
90\% confidence level.\\
While a precise limit for specific models can only be set using a simulation
procedure as outlined above, we typically find that for various types
of non-statistical fluctuations our measurement is sensitive when the
fluctuations lead to an effective non-statistical variation in the mean
transverse momentum from event to event of about 1\%.
We can also use the simulation procedure to establish limits on fluctuations of amplitude
$\sigma_{fluc}$ that occur only in a fraction $F$ of all events.
The resulting exclusion plot is shown in fig.~2,
where the relative frequency $F$ of events exhibiting
fluctuations of amplitude $\sigma_{fluc}$ is plotted versus
$\sigma_{fluc}$.
We see that for $F = 1$ fluctuations of a relative
amplitude of $\sigma_{fluc} > 1.2\%$ are ruled out at 90\% confidence
level, whereas fluctuations occuring in 1\% of the events
can only be ruled out for $\sigma_{fluc} > 10\%$.
\begin{figure}[htp]
\centerline{\epsfig{file=figure2.eps,height=7.5cm}}
\caption{Limit on the amplitude of fluctuations in the
$p_T$ parent distribution
as a function of the frequency of events showing the fluctuation.}
\label{3}
\end{figure}
It is important to note that the measurement of fluctuations in
$M(p_T)$ is also
relevant for models of processes that lead to
non-statistical fluctuations \em localized \em in transverse
momentum. The most widely discussed example of such
a process is the formation of
disoriented chiral condensates (DCCs)\cite{raja93,anselm91}, which has been
postulated as a consequence of the possible restoration
of chiral symmetry in non-equilibrium scenarios for heavy-ion collisions.
DCC models predict the formation of domains
that eventually emit pions where the ratio $f$ of neutral
to all pions varies as
\begin{equation}
P(f) = \frac{1}{2 \sqrt{f}}.
\label{dcc}
\end{equation}
The models also suggest that pions emitted from DCC domains
will be preferentially produced at low transverse
momenta\cite{raja93}.
This provides for a translation of the number-fluctuations
predicted by the DCC models into $p_T$ fluctuations accessible
to our experiment. A limit on DCC production has already been
set by the WA98 collaboration \cite{agga98}, based on a study of
fluctuations in the relative multiplicties of charged and neutral
particles near mid-rapidity.\\
For comparison purposes we used the same DCC model as in \cite{agga98},
where the DCC production is characterized by the probability $F$
to form a single DCC domain in an event and the fraction $\xi$
of pions coming from the DCC. We make the additional assumption
that the DCC pions are produced with $p_T < p_T^{max} = m_{\pi}$.
The ratio of neutral to charged
pions was chosen randomly according to equ.~\ref{dcc}.
The isospin fluctuations of pion production from DCCs then
lead to multiplicity fluctuations of charged pions at low
transverse momenta and therefore to non-statistical fluctuations in $M(p_T)$.
For DCCs occuring in every event ($F = 1$) the fluctuations observed in the
data rule out DCC sizes
of $\xi > 3.5$~\%, which is about a factor of 5 smaller than
the previous limit set in \cite{agga98}. This limit could be further
improved by restricting the analysis to the region of small
transverse momenta.\\
\section{Summary}
In summary, event-by-event fluctuations in the average transverse
momentum of charged particles in the forward hemisphere
of central Pb+Pb collisions have been measured.
The distribution of average transverse momentum
per event $M(p_T)$ has an approximately Gaussian shape, with
no excess of 'anomalous' events falling out of the distribution.
The fluctuation strength in the data is characterized by
a value of $\Phi_{p_T} = 0.6 \pm 1$~MeV/c.
Using a procedure based on mixed events
we find that the fluctuations in $M(p_T)$ from event to event
are compatible with
independent particle production modified by the
known two-particle correlations due to quantum statistics and final state
interactions and taking into account the response
of the NA49 apparatus, without requiring further variations in the
transverse momentum parent distribution from event to
event.\\
For a model of non-statistical fluctuations in average $p_T$ we
use a detailed simulation procedure to determine an upper limit on
the strength of fluctuations occuring in every event of
$\sigma_{fluc} < 1.2\%$ at 90\% confidence level.\\
We also demonstrate that high precision measurements of charged particle
transverse momentum fluctuations provide a sensitive test for models
predicting the formation of disoriented chiral condensates in heavy-ion collisions.
Finally, we have provided the first measurement to compare to predictions of
event-by-event transverse momentum fluctuations in thermodynamical descriptions of the strongly
interacting system produced in Pb+Pb collisions at the SPS.\\
\input{NA49acknow}
|
\section{Introduction }
\label{sec-introd}
\andy{intro}
The usual, {\em static} version of the quantum Zeno effect (QZE) consists in
hindering (and eventually halting) the time evolution of a quantum system by
repeatedly checking if it has decayed \cite{von}.
In a few words, this is due to the fact that in time $dt$, by the Schr\"odinger equation,
the phase of a state $\psi(t)$ changes by $\hbox{O}(dt)$ while the
absolute value of its scalar product with the initial state changes by
$\hbox{O}(dt^2)$.
The {\it dynamic\/} quantum Zeno effect exploits the above features and
forces the evolution through an arbitrary trajectory by a series of repeated
measurements \cite{von1,AA87}: Let there be a family of states
$\phi_k$, $k=0,1,\ldots, N$, such that $\phi_0=\psi(0)$, and such that
successive states differ little from one another (i.e.,
$|\langle\phi_{k+1} | \phi_k \rangle|$ is nearly 1). Now let $\delta T =
T/N$ and at $T_k=k\delta T$ project the evolving wave function on
$\phi_k$. Then for sufficiently large $N$, $\psi(T) \approx \phi_{_N}$.
[The static QZE is the special case $\phi_k=\phi_0 (=\psi(0)) \ \forall \
k$.]
In the following we will show how guiding a
system through a closed loop in its state space (projective Hilbert
space) leads to a geometrical phase
\cite{AA87,Panchar,BerryQuantal}. We will first summarize some
results valid for neutron spin \cite{continous,Berry} and then
consider the case of photon polarization \cite{BerryKlein}.
\section{Neutron spin}
\label{sec-neutron}
\andy{neutron}
Assume first that there is {\em no} Hamiltonian acting on the system: the
neutron crosses a region where no magnetic
field is present. It starts with spin up along the $z$-axis and is
projected on the family of states
\andy{projfamily}
\begin{equation}}\newcommand{\eeq}{\end{equation}
\phi_k \equiv \exp(-i\theta_k\mbox{\boldmath $\sigma$}\cdot\mbox{\boldmath $n$})\coltwovector10 \qquad
\hbox{with~} \theta_k \equiv \frac{ak}N \;,
\qquad k=0,\ldots,N \ ,
\label{eq:projfamily}
\eeq
where $\mbox{\boldmath $\sigma$}$ is the vector of the Pauli matrices and $\mbox{\boldmath $n$}} \def\bmA{\mbox{\boldmath $A$} =
(n_x,n_y,n_z)$ a unit vector (independent of $k$).
The neutron evolves for a time $T$ with projections at times $T_k =
k\delta T$ ($k=1,\dots,N$ and $\delta T=T/N$). The final state is
$\left[\phi_0 = \coltwovector10\right]$
\andy{finstate}
\begin{eqnarray}}\newcommand{\earr}{\end{eqnarray}
\ket{\psi(T)}
&=& |\phi_N\rangle \langle \phi_N|
\phi_{N-1}\rangle \cdots \langle \phi_2|
\phi_1\rangle \langle \phi_1| \phi_0\rangle \nonumber \\
&=& \cos^N \left(\frac{a}{N} \right)
\left(1 + i n_z \tan \frac {a}{N} \right)^N
|\phi_N\rangle \nonumber \\
&\stackrel{N\rightarrow \infty}{\longrightarrow} &
\exp (ia n_z) \exp (-ia\mbox{\boldmath $\sigma$}\cdot\mbox{\boldmath $n$}) | \phi_0\rangle .
\label{eq:finstate}
\earr
If
$a=\pi$, \andy{finstatepi}
\begin{equation}}\newcommand{\eeq}{\end{equation}
\psi(T)
= \exp [-i \pi (1-\cos \Theta)] \phi_0
= \exp (-i \Omega/2 ) \coltwovector10 ,
\label{eq:finstatepi}
\eeq
where $\cos\Theta \equiv n_z$ and $\Omega$ is the solid angle
subtended by the curve traced by the
spin during its evolution. The factor $ \exp (-i\Omega/2)$ is a Berry
phase and it is due only to measurements (the Hamiltonian is zero).
Notice that, as discussed by Pati and Lawande \cite{Pati}, no Berry phase appears
in the usual quantum Zeno context,
namely when $\phi_k \propto \phi_0 \ \forall \ k$, because in that case
$a=0$ in (\ref{eq:finstate}).
We now look at the process (\ref{eq:finstate}) for $N$
finite. The spin goes back to its initial state after describing a
regular polygon on the Poincar\'e sphere, as in Figure 1a.
\begin{figure}
\centerline{\epsfig{file=olospin.eps, height=5cm}}
\caption{Fig. 1. a) Spin evolution due to $N=5$ measurements.
b) Spin evolution with very frequent measurements and non-zero Hamiltonian.}
\end{figure}
After $N (<\infty)$ projections the final state is \cite{Berry}
\begin{equation}}\newcommand{\eeq}{\end{equation}
\psi(T)=\rho_N \exp(-i\beta_N)\phi_0,
\eeq
where
\andy{rhoN, betaN}
\begin{eqnarray}}\newcommand{\earr}{\end{eqnarray}
\rho_N = \left(\cos^2\frac{\pi}{N}+n^2_z
\sin^2\frac{\pi}{N}\right)^{\frac{N}{2}}, \qquad
\beta_N = \pi-N\arctan\left(\cos\Theta\tan\frac{\pi}{N}\right).
\label{eq:betaN}
\earr
The quantity $\rho_N$ accounts for the probability loss ($N$ is
finite and there is no QZE). It is easy to check that in the
``continuous measurement" limit (QZE) we recover the result
(\ref{eq:finstatepi}).
The relation between the solid angle and the geometrical phase is valid
also with a finite number of polarizers $N$. Indeed, it is
straightforward to show that the solid angle subtended by a regular $N$-sided polygon
(Figure 1a) is
\begin{equation}}\newcommand{\eeq}{\end{equation}
\Omega_{N}
=2\pi-2N\arctan\left(\cos\Theta\tan\frac{\pi}{N}\right)=2\beta_N.
\eeq
This result is of course in agreement with other analyses \cite{SM}
based on the Pancharatnam connection \cite{Panchar}.
Let us now consider the effect of a non-zero Hamiltonian
(neutron spin in a magnetic field)
\andy{Hamadd}
\begin{equation}}\newcommand{\eeq}{\end{equation}
H=\mu \mbox{\boldmath $\sigma$}\cdot\mbox{\boldmath $b$} ,
\label{eq:Hamadd}
\eeq
where $\bmb = (b_x,b_y,b_z)$ is a unit vector, in general different from
$\mbox{\boldmath $n$}} \def\bmA{\mbox{\boldmath $A$}$. See Figure 1b.
If the system starts with spin up it has the
following ``undisturbed" evolution
\andy{undisturb}
\begin{equation}}\newcommand{\eeq}{\end{equation}
\psi(t) = \exp(-i\mu t\mbox{\boldmath $\sigma$}\cdot\mbox{\boldmath $b$})\phi_0 .
\label{eq:undisturb}
\eeq
Now let the system evolve for a time $T$ with projections at times
$T_k=k\delta T$ ($k=1,\dots,N$ and $\delta T=T/N$) and Hamiltonian
evolution in between. It is not difficult to show that, in the
continuum limit ($N\to\infty$), the final state reads \cite{Berry}:
\andy{finpsi}
\begin{eqnarray}}\newcommand{\earr}{\end{eqnarray}
\psi(T)
=\exp\left(-i \int_0^T \langle \psi(t) | H | \psi(t) \rangle dt\right)
\exp\left(i a n_z - ia\mbox{\boldmath $\sigma$}\cdot\mbox{\boldmath $n$} \right) \phi_0.
\label{eq:finpsi}
\earr
The first factor in (\ref{eq:finpsi}) is obviously the dynamical
phase and the remaining phase, when the spin goes back to its
initial state, is the geometrical phase: when $a=\pi$
\andy{fun}
\begin{equation}}\newcommand{\eeq}{\end{equation}
\psi(T) = \exp \left( - i\Omega/2 \right)
\exp\left(- i\mu T (\bmb \cdot \mbox{\boldmath $n$}} \def\bmA{\mbox{\boldmath $A$}) n_z \right) \coltwovector10 ,
\label{eq:fun}
\eeq
where $\Omega$ is the solid angle subtended by the curve traced out by
the spin, as in (\ref{eq:finstatepi}), and $\mu T (\bmb \cdot \mbox{\boldmath $n$}} \def\bmA{\mbox{\boldmath $A$})
n_z$ is the dynamical phase.
\section{Photon polarization}
\label{sec-photon}
\andy{photon}
The experiments described in Section \ref{sec-neutron} involve
neutrons, or in general {\it massive} spin-$1/2$ particles. This
allowed us to neglect the neutron momentum in our analysis: the
axis of the neutron polarizer can have an arbitrary direction with
respect to the neutron momentum. Equations very similar to those of
Section \ref{sec-neutron} were obtained by Berry and Klein in their
beautiful work on polarized light \cite{BerryKlein}. However, when
one deals with photons, the additional constraint of transversality
makes things more complicated: the photon polarization must be
perpendicular to momentum and for this reason, in Ref.\
\cite{BerryKlein}, the projection on states of non-linear
polarization is achieved by making use of ``retarders" \cite{SiMu}.
It is interesting, in this context, to discuss a geometric
configuration proposed by A.G.\ Klein \cite{Klein1}. A photon is
sent into a polygonal cylinder made up of $N$ perfect plane mirrors
and emerges after $N$ reflections: see Figure 2 ($N=6$).
\begin{figure}
\centerline{\epsfig{file=mirror.eps, height=5cm}}
\caption{Fig. 2. (a) Klein's arrangement: a photon is sent into a polygonal
cylinder, bouncing off $N=6$ perfect mirrors and emerging from the
opposite side. (b) Evolution of the polarization. The points 1-6
represent the tip of the polarization vector and the shadow area
yields the Berry phase. }
\end{figure}
Let us consider a photon with momentum parallel to $\hat{\bm
k}=(\sin\theta\cos\phi,\sin\theta\sin\phi,\cos\theta)$ and
polarization $\ket{p}$, which is reflected by an ideal mirror with
normal unit vector $\bm n$. After reflection, the photon helicity
is reversed and the momentum becomes
\begin{equation}}\newcommand{\eeq}{\end{equation}
\hat{\bm k}'=-R_{\bmsub n}(\pi)\hat{\bm k},
\eeq
where $R_{\bmsub n}(\phi)$ represents a rotation by an angle $\phi$
around direction $\bm n$. Hence, after a reflection, the
polarization state becomes (ideal mirror, infinite conductivity)
\andy{reflection}
\begin{equation}}\newcommand{\eeq}{\end{equation}\label{eq:reflection}
\ket{p'}=M(\bm n)\ket{p},\quad\mbox{with}\quad
M(\bm n)=\exp\left(-i\bm n\cdot\bm J\pi\right),
\eeq
where $\bm J=(J_x,J_y,J_z)$ are the generators of rotations in {\bf
R}$^3$. Consider now a polygonal cylinder of ideal mirrors, whose
normal vectors are
\andy{mirror}
\begin{equation}}\newcommand{\eeq}{\end{equation}\label{eq:mirror}
\bm n_1=
\left(-1,0,0\right),
\quad \bm n_\ell= R_z[(\ell-1)\alpha]\bm n_1
=[-\cos(\ell-1)\alpha,-\sin(\ell-1)\alpha,0],
\eeq
where $\alpha=2\pi/N$ for a regular $N$-sided polygon [see
Figure~2(a)]. The operator representing the action of the $\ell$-th
mirror is $M_\ell=M(\bm n_\ell)$. By using (\ref{eq:reflection})
and (\ref{eq:mirror}) one gets
\begin{equation}}\newcommand{\eeq}{\end{equation}
M_\ell
=-\exp\left(-i\alpha(\ell-1) J_z\right)
M_1 \exp\left(i\alpha(\ell-1) J_z\right),\quad M_1=\exp(i\pi J_x).
\eeq
Let $\ket{p_0}$
be the initial photon polarization; after $N$
reflections, the photon emerges with final polarization
\andy{finpol}
\begin{equation}}\newcommand{\eeq}{\end{equation}\label{eq:finpol}
\ket{p_N}=M_N\cdots M_2 M_1 \ket{p_0}
=\exp \left( -i N \alpha J_z \right)\widetilde M^N\ket{p_0},
\eeq
where
\begin{equation}}\newcommand{\eeq}{\end{equation}
\widetilde M=\exp\left(i\alpha J_z\right) M_1
= \exp\left(i\alpha J_z\right) \exp(i\pi J_x).
\eeq
By using $[J_i, J_j]=i\varepsilon_{ijk}J_k$, we obtain
\andy{nrot}
\begin{equation}}\newcommand{\eeq}{\end{equation}\label{eq:nrot}
\widetilde M=\exp\left(-i\pi\tilde{\bm n}\cdot\bm J\right),
\quad\mbox{with}\quad
\tilde{\bm n}=\left(-\cos\frac{\alpha}{2},\sin\frac{\alpha}{2},0 \right),
\eeq
so that $\widetilde M$ is a $\pi$ rotation around
$\tilde{\bm n}$. Remembering that $N\alpha=2\pi$ and using
(\ref{eq:nrot}), the final polarization (\ref{eq:finpol}) reads
\begin{equation}}\newcommand{\eeq}{\end{equation}
\ket{p_N}=\exp(-i 2\pi J_z)\exp\left(-i N\pi
\tilde{\bm n}\cdot\bm J \right)\ket{p_0}
= \exp\left(-i N\pi
\tilde{\bm n}\cdot\bm J \right)\ket{p_0} .
\eeq
For odd $N=2m+1$, one gets
\begin{equation}}\newcommand{\eeq}{\end{equation}
\ket{p_N}
=\exp\left(-i (2m+1)\pi \tilde{\bm n}\cdot\bm J\right)\ket{p_0}
=\exp\left(-i \pi
\tilde{\bm n}\cdot\bm J\right)\ket{p_0}
\eeq
and the final polarization is not in the same direction as the
initial one. On the other hand, for even $N=2m$, the polarization
vector does describe a closed loop, but one obtains
\begin{equation}}\newcommand{\eeq}{\end{equation}
\ket{p_N}=\exp(-i 2m \pi \tilde{\bm n}\cdot\bm J)\ket{p_0}
=\ket{p_0}
\eeq
and the photon acquires no geometrical phase. The reason for this
result is shown in Figure~2(b): for even $N$, the polarization
vector {\it always} describes a solid angle $\Omega=2\pi$: the
photon always acquires a Berry phase $\beta=\Omega=2\pi$, with no
physical effects.
The experiment just described is not equivalent to the one analyzed
in the previous section. Indeed, the dynamics of reflections is
always unitary. The difficulty in obtaining a geometrical phase is
due to the condition of transversality of the electromagnetic
field. To encompass this situation one needs (at least three)
mirrors whose normal vectors do not lie in the same plane, as shown
in \cite{Kitano}.
\medskip
\noindent {\bf Acknowledgments:}
We thank A.G.\ Klein and L.S.\ Schulman for many useful discussions
and M.V.\ Berry and A.K.\ Pati for interesting comments on Ref.\
\cite{Berry}. The mirror geometry investigated in
Section~\ref{sec-photon} was proposed by A.G.\ Klein.
|
\section{Introduction}
\label{sec:Intro}
\quad
Further precise measurement of
the anomalous magnetic dipole moment of the muon,
conventionally denoted as $a_\mu \equiv \frac{1}{2}(g-2)_\mu$,
is now underway
at Brookhaven National Laboratory (BNL).
The perspective for the goal of this experiment is
\cite{future-exp}
\begin{equation}
\Delta a_\mu({\rm expt}) =
4.0 \times 10^{-10}\, .
\label{eq:future-precision}
\end{equation}
The recent report for its test-running
course at BNL \cite{future-exp,BNL-NV}
combined with the previous one at CERN \cite{CERN}
gives to muon $g-2$
\begin{equation}
a_\mu(\textrm{expt}) = 11659~235~(73) \times 10^{-10}\, ,
\label{eq:present_value}
\end{equation}
where the numerals in the parenthesis represent
the uncertainty in the final few digits.
Thus the precision (\ref{eq:future-precision})
amounts to the determination of its value by one further digit.
\\
\quad
Our primary interest
is what we can learn when invoking
such a improvement in muon $g-2$ experiment.
At present the standard model predicts
\begin{equation}
a_\mu(\textrm{SM}) =
11659~160.5~(6.5) \times 10^{-10} ,
\end{equation}
which includes the up-dated estimate on
the leading hadronic vacuum polarization contribution
\cite{Davier},
(See \cite{Eidelman} due to analysis without recourse to $\tau$ decay data),
and $\mathcal{O}(\alpha^3)$ contribution
\cite{Krause}.
The designed precision (\ref{eq:future-precision})
is put forward to find out the existence
of the $W$, $Z$ boson effect to muon $g-2$
\cite{weak-2-loop}
\begin{equation}
a_\mu(\textrm{weak}) =
151~(4) \times 10^{-11} .
\label{eq:weak_effect}
\end{equation}
The electron $g-2$ does not receive observable effects
from $Z$ and $W$ bosons
and are entirely saturated by QED effect.
This fact enables us to find out the validity of
calculation scheme of quantum field theory,
including the perturbative renormalization procedure.
Rather electron $g-2$ provides
the most accurate determination of the fine structure constant
$\alpha$ \cite{Kinoshita} at present.
From (\ref{eq:weak_effect}) and (\ref{eq:future-precision}),
the muon $g-2$ is expected not only to obtain
a conceivable evidence for
such a structure about the electroweak interactions
as involved in standard model,
but also has the potential
to indicate the existence of much richer ingredient
associated with some theoretical problem. \\
\quad
When we incline to use muon $g-2$
as a probe of new physics,
the theme of the talk assigned to me as well,
it would be important to recall
the motivation or the merits of each model.
Thus, here, I will focus on two concrete models
considered to approach to ''hierarchy problem'',
although it is somewhat a conceptual viewpoint.
Here ''hierarchy problem'' stands for
the following question in some narrow sense;
what is assuring the stability of the electroweak scale,
represented by $W$ boson mass, $M_W$,
against quantum fluctuation associated with
high momentum modes below some cutoff scale.
The cutoff scale here is the scale
at which the gauge interactions appearing in standard model would
become subject to some kind of modification.
It may be the GUT scale $M_G$,
at which the standard model gauge symmetry group
is merged into a larger symmetry group.
Or it may be the scale at which the gauge boson is resolved
into more fundamental structure, for instance, into string. \\
\quad
Here we take up two models considered with such a motivation.
One is TeV scale gravity discussed in the next section,
and the other supersymmetry in Sec. \ref{sec:SUSY}.
Sec. \ref{sec:discussion} concludes
with remark on several facets
for muon $g-2$ to probe new physics practically.
\section{TeV scale gravity}
\label{sec:gravity}
\quad
The laboratory experiments check the structure of
gravity has been met up to the order of millimeters.
With this in mind let us turn our attention
to the scheme introduced a couple of years ago
\cite{Dimopoulos}. \\
\quad
Let us imagine that our world is confined in a three-brane,
the extended object with three-spatial directions,
flowing in a higher dimensional space
($(n+4)$-dimension in total).
These $n$-directions are compactified to an $n$-dimensional torus
with the same length $2\pi R$, for simplicity.
Then there are an infinite tower of Kaluza-Klein states
from the four-dimensional space-time point of view.
They are the states with non-zero momentum in the extra directions.
In four-dimensional world
such a momentum appears as the mass,
the scale of which is characterized by inverse of $R$. \\
\quad
The behavior of static potential between two point-like sources
prepared with separation $r$
illustrates an important aspect of the present setting
\cite{Arkani-Hamed} .
When $r$ is large compared to $R$,
the potential behaves as $\sim 1/r$,
nothing but the form of the usual Newtonian one.
The Planck scale $M_{\rm PL} \simeq 10^{19}$ characterizes
its strength.
However, once the separation $r$ reaches below $R$,
the probability that Kaluza-Klein states get excited
cannot become neglected any more.
Thus counting those states
which propagate essentially like massless states
between two sources
leads $r^{-(n+1)}$ as
$r$-dependence of the potential at short distance,
characteristic of $(n+3)$-spatial dimension.
This only reflects the fact
that the local structure less than the compactification scale
is that of $(n+4)$-dimensional space-time.
A noteworthy point is
that the strength of force is
then characterized by another Planck scale,
$M_*$;
\begin{equation}
M_* =
\left(
\frac{M_{\rm PL}^2}{R^n}
\right)^{(n+2)}\, .
\end{equation}
This is
the strength of the gravitational interaction in the bulk theory,
the fundamental Planck scale.
\\
\quad
Now we take $R$ equal to 1 mm
\footnote{
Energy scale is related to the length scale $L$ through
\begin{equation}
E = 0.197\,{\rm eV} \times \frac{1\,{\rm mm}}{L}\, .
\nonumber
\end{equation}
},
which corresponds to the length scale one-order less than
the current reach of the experiment on gravity.
With two extra compactified directions
$M_*$ = 1 TeV
\footnote{
It was commented by A. Kataev in this workshop
that consideration on the effect
to the life-time of the red giant stars
appears to reject the possibility, $M_*$ = 1 TeV,
which he heard at the seminar by Arkani-Hamed
\cite{Arkani-Hamed}.}.
\\
\quad
If the three-brane were further of Dirichlet-brane type,
on which the open string can attach,
our standard model gauge bosons
would become the tangential component of
the ground state of such an open string.
Then the gauge bosons
would get resolve into strings higher than 1 TeV
as long as the string coupling constant is of order unity.
Thus there is no hierarchy problem ab initio.
\\
\quad
At present we are lacking the precise
formulation of theories and
the detail knowledge on its dynamical aspects
(especially on the compactification mechanism).
However the dimension counting argument
with symmetry consideration
has been an enough tool
when we estimate the order of magnitude
about some effect in low energy phenomena
unless the effective theory description breaks down,
although we rather expect to superstring theory
that many miracles beyond this assumption occur.
\\
\quad
The argument begins with the chiral symmetry for muon.
When muon was massless, the magnetic dipole coupling would be absent.
Thus the magnitude of magnetic moment will
be proportional to muon mass $m_\mu$
and the corresponding operator appears
in the effective Lagrangian:
\begin{equation}
\mathcal{L} = e\,m_\mu \,A\,\bar{\mu}\,
\sigma^{\lambda\rho} F_{\lambda\rho}\, \mu.
\label{eq:muon_dipole}
\end{equation}
with the the extra effect $A$ due to the new structure
characterized by mass scale $M_*$.
Since the mass dimension of $A$ turns out to be minus one
in four space-time dimension,
the dimension counting now gives
\begin{equation}
A = c \times \frac{1}{M_*},
\label{eq:A_const}
\end{equation}
with the numerical constant $c$ of order unity.
Insertion of (\ref{eq:A_const}) into
the effective interaction (\ref{eq:muon_dipole})
with a slight rearrangement yields
\begin{equation}
\mathcal{L} =
\frac{e}{4 m_\mu}\,
\left[
4c\,
\left(
\frac{m_\mu}{M_*}
\right)^2
\right]\,
\bar{\mu} i\sigma^{\lambda\rho} F_{\lambda\rho}\,\mu.
\end{equation}
The quantity found in the square bracket of the above expression
corresponds to the additional contribution
to $a_\mu$ due to the presence of TeV scale gravity.
Thus the additional effect to $a_\mu$
can be read off as
\begin{equation}
\delta a_\mu
= 4c\,
\left( \frac{m_\mu}{M_*} \right)^2
\times 10^{10},
\end{equation}
which becomes for $M_*$ = 1 TeV
\begin{equation}
\delta a_\mu \sim
\left( 4c \times 100 \right) \times 10^{-10}.
\end{equation}
Thus a crude estimate shows that
the effect from TeV scale gravity is in the marginal
situation to be detected even
with the current accuracy (\ref{eq:present_value}).
The future accuracy (\ref{eq:future-precision})
is quite adequate to detect the existence
of new aspect of gravity characterized by TeV scale
\footnote{
During the preparation of the talk,
I have noticed that more concrete demonstration
has been performed in a similar context
\cite{Graesser}
}.
\section{Supersymmetric Model}
\label{sec:SUSY}
\quad
Now we will take our attention to the search of supersymmetry
with the use of muon $g-2$,
which has been discussed
as a machinary to check the various adovocated models
or from some generic standpoint
\cite{Kosower,Chattopadhyay,Polonsky,Gabrielli}.
The detail formula and so on in this section
will also be found in a separate literature
\cite{CHH}.
\\
\quad
In the narrow sense
of the ''hierarchy problem'' defined in the introductory remark,
supersymmetry makes it possible
to extend the standard model gauge group into
the larger gauge group in grand unified theory
(GUT).
This is realized in the form of
the cancellation of the dangerous quantum corrections
within each supermultiplet.
The bosonic partner of muon,
for instance, has not been observed yet
so that it must be much heavier.
Smuon can be let heavier by giving it a lifting-up mass, $m_S$.
This procedure does not spoil quantum stability
as far as $m_S$ is not so apart from the electroweak scale.
The scale $m_S$ works
as the ultraviolet cutoff scale for the phenomena
accessed by low momentum probe
while it works as the infrared cutoff
from the supersymmetric high-energy side.
It has been recognized that
unification of the coupling constants of three gauge interactions,
$\textrm{SU(2)}_{\rm L}$, $\textrm{U(1)}_{\rm Y}$
and color $\textrm{SU(3)}_{\rm C}$,
is achieved by assuming the particle content
of the minimal supersymmetric extension of standard model
above such $m_S$.
For the future purposes it deserves to recall here
that the knowledge from precise
measurement around $Z$ pole plays an indispensable role
to establish this fact.
\\
\quad
Now we examine the effect on muon $g-2$
induced from supersymmetric theories.
They come essentially from two diagrams.
One is the chargino-sneutrino loop,
where the charginos are the admixture of $\textrm{SU(2)}$
gaugino $\tilde{w}^-$ and the charged Higgsino.
The other is the neutralino-smuon loop,
where the neutralinos are the admixture of
their neutral counterparts.
The other contributions,
such as the charged Higgs loop one,
are so small that they are irrelevant
even for future study.
\\
\quad
\begin{figure}[htb]
\includegraphics[origin=Bc, angle=-90, scale=0.6]{g-2_50_200.eps}
\caption{$\mu$ dependence of $(a_\mu)_{\rm SUSY}$
for $\tan\beta = 50$ and
$m_{\tilde{\mu}_L} = m_{\tilde{\mu}_R} = 200$ GeV.}
\label{fig:g-2_50_200}
\end{figure}
Fig. \ref{fig:g-2_50_200}
is intended to demonstrate
what magnitude of those effects to muon $g-2$
is expected.
The figure shows
the supersymmetric contribution to muon $g-2$
as a function of $\mu$ parameter
(the supersymmetric mass common to two Higgs supermultiplets),
for relatively large $\tan\beta$
(the ratio of the vacuum expectation values of two Higgs doublets),
the supersymmetry breaking slepton mass set equal to 200 GeV,
and for three choices of supersymmetry breaking ${\rm SU(2)_L}$
gaugino mass $M_2$ with ${\rm U(1)_Y}$ gaugino mass given
here through the GUT relation
\begin{equation}
M_1 = \frac{5}{3} \tan^2\theta_W M_2\, .
\end{equation}
As the weak effect is $15 \times 10^{-10}$,
the supersymmetric effect can become substantial.
Actually the muon $g-2$ even with the current accuracy
excludes the region of negative sign of $\mu$
for this set of the other parameters.
\\
\quad
Fig. \ref{fig:g-2_50_200} is drawn without paying any attention
to the other constraints on supersymmetric models already present.
The direct search of superpartners of the known species
puts the lowest bound (93 GeV) to the lightest chargino mass,
and the bound (78 GeV) to the mass of each lighter slepton
\cite{Mihara}.
In fact the chargino mass bound requires that
the absolute magnitude of $\mu$ parameter be greater than
about 100 GeV
for the gaugino mass in our interest,
while the slepton mass bound demands $\left| \mu \right|$
less than about 400 GeV.
With those regions excluded
Fig. \ref{fig:cnt_50_200} shows
the contours each of which has equal magnitude of muon $g-2$
on the $\mu$-$M_2$ plane.
Since the future accuracy is much smaller than
the interval between the neighboring contours
in Fig. \ref{fig:cnt_50_200},
we have a great chance to observe a signal
coming from the existence of supersymmetry through muon $g-2$.
\\
\begin{figure}[thb]
\begin{center}
\includegraphics[origin=Bc, scale=0.9]{cont_50_200.eps}
\caption{Contours with equal $(a_\mu)_{\rm SUSY}$
in $\mu$(horizontal direction)-$M_2$ plane
for $\tan\beta = 50$ and
$m_{\bar{\mu}_L} = m_{\bar{\mu}_R} = 200$ GeV.
The two islands are the regions escaping from
any other constraints.
The contours are drawn with the interval of
$50 \times 10^{-10}$
from $-200 \times 10^{-10}$ to $200 \times 10^{-10}$
for $(a_\mu)_{\rm SUSY}$.
Darker face corresponds to
smaller $(a_\mu)_{\rm SUSY}$.}
\label{fig:cnt_50_200}
\end{center}
\end{figure}
\quad
Now we are tempted to grasp the specific feature of muon $g-2$
in search of supersymmetric theory.
It will turn out that muon $g-2$ seems to have a peculiar property
which is not shared by any other observables.
\\
\quad
In the most honest region of parameter space
$\left| \mu \right| \ge (2 \sim 3)\, M_2$,
chargino-sneutrino loop contribution dominates over
neutralino-smuon loop contribution.
The current basis analysis
helps one to catch up with the qualitative dependence on
the various parameters.
As long as $\tan\beta \ge 3$,
the chirality flips due to the vacuum expectation value
$ \left< H_U \right> $
of the Higgs field giving mass to the up-type quarks,
turning $\tilde{w}^-$ to the charged component of $\tilde{H}_U$,
which transformed to the charged $\tilde{H}_D$ due to $\mu$-term.
Picking $\sin\beta$ from $ \left< H_U \right> $
and $1/\cos\beta$ from a yukawa-type coupling involving muon,
the dominant contribution in the present situation
becomes
\begin{equation}
(a_\mu)_{\rm SUSY} \propto +\mu \tan\beta \, ,
\label{eq:a_susy_1}
\end{equation}
although the overall sign needs a detail computation.
From this expression we can read off such properties that
\begin{enumerate}
\item[(a)]
The effect to muon $g-2$
is greatly enhanced for large $\tan\beta$ \cite{Kosower}.
In fact, when $\tan\beta$ is small,
the overall magnitude of SUSY effect is drastically
reduced as shown in Fig. \ref{fig:g-2_3_100}.
Thus in this case the current experiment
could not put any restriction on its existence.
But the future accuracy in muon $g-2$
is quite sufficient to explore it
\footnote{
The renomalization group analysis shows that
QED correction tends to decrease those new effect
about 6\%, and this fact should be recalled
at the critical stage of confronting
with the experimental data \cite{Degrassi}.
}.
\item[(b)]
The sign of this contribution is govern by the sign of $\mu$.
\end{enumerate}
\begin{figure}[htb]
\includegraphics[origin=Bc, angle=-90, scale=0.6]{g-2_3_100.eps}
\caption{$\mu$ dependence of $(a_\mu)_{\rm SUSY}$
for $\tan\beta = 3$ and
$m_{\tilde{\mu}_L} = m_{\tilde{\mu}_R} = 100$ GeV.}
\label{fig:g-2_3_100}
\end{figure}
It is interesting to remind that
a large $\tan\beta$ is a natural consequence
of the gauge mediated SUSY breaking scenario,
and elaborate analysis on muon $g-2$
has been performed in this line \cite{Gabrielli}.
Or it is a necessary ingredient
for unification of all yukawa coupling constants
of the third generation.
\\
\quad
As observed at both ends of $\mu$-direction
in Fig. \ref{fig:g-2_3_100},
supersymmetric effect to muon $g-2$ does not decouple
even if we can let the absolute magnitude of $\mu$ large
while those other parameters remain fixed.
Note that small $\tan\beta$ case allows
relatively large absolute magnitude of $\mu$ parameter
without conflicting with slepton mass bound,
as the mixing between left- and right-sleptons are
proportional to a combination $\mu \tan\beta$.
\\
\quad
Such a phenomenon can be understood from the following observation.
When $\left| \mu \right|$ is large
the chargino-sneutrino effect decouples,
but the neutralino-smuon effect increases.
Let us consider a diagram in which
the chiral flip occurs due to the mixing between the left
and right-handed smuons in the current eigen-basis.
As the Higgino does not propagate,
suppression factor due to the inverse power of $\mu$ is now absent.
Thus $(a_\mu)_{\rm SUSY}$ becomes proportional to $-\mu\tan\beta$.
(The sign is also opposite to the chargino-sneutrino effect
(\ref{eq:a_susy_1}).)
This is the reason for such a behavior
in this large $\left| \mu \right|$ region.
\\
\quad
From those observations,
muon $g-2$ seems to play the major role
to find the sign and the magnitude of $\mu$ term.
As far as I know, such a property sensible to $\mu$
is not shared by any other observables.
Recall the following two facts, that is,
\begin{enumerate}
\item[(a)] $\mu$ is a supersymmetric parameter
which is associated with common mass of Higgs supermultiplet.
Thus this is a parameter independent of
the supersymmetry breaking parameters by nature.
\item[(b)] Although supersymmetry
assures the quantum stability of electroweak scale,
supersymmetry does not set $\mu$ to this region
at tree level automatically.
This is the most annoying matter called as ``$\mu$ problem.
This problem stands out especially in the context of GUT.
\end{enumerate}
Thus, once supersymmetry is established
also by the other experiments,
the determination of $\mu$ parameter through muon $g-2$
may develop further theoretical access
to the origin of $\mu$, the origin of electroweak scale.
\\
\quad
Before addressing to the future testing possibility
in the small $\tan\beta$ regime,
we remind the additional constraints
implied from precision measurement at $Z$ pole.
As was mentioned,
the result of this precision measurement
has given an indispensable information
to argue grand unification.
It also has killed the naive technicolor models.
Thus we should discuss the effect on muon $g-2$
on the region of the parameter space
consistent with those measurements.
\\
\quad
They are summarized by four parameters.
Three of these parametrizes the ``oblique'' corrections
from new physics,
with respect to ``reference'' standard model;
here we take the one specified by
\begin{equation}
m_t = 175\ {\rm GeV}, \quad
M_H = 100\ {\rm GeV}\, .
\end{equation}
The last one is associated
with the modification of coupling of bottom quark to $Z$ boson.
This is neglected here
by assuming that the squarks are so heavy enough
that their effects decouple.
Since it has been recognized that
the SUSY effect to $W$ boson mass and coupling of $\tau$ to $Z$
is not relevant within the current accuracy\cite{Cho},
we concentrate on $S$ and $T$ parameters.
\begin{figure}[htb]
\begin{center}
\includegraphics[origin=Bc, scale=0.6]{ST_100.eps}
\caption{Constraint on supersymmetric theory
from $S$-$T$ parameters for $\mu > 0$.
Slepton brings $S$ parameter to decrease
(Each line follows the response to the change
of $m_{\tilde{l}_L}$.)
while chargino and neutralino tend to
increase it
(Each line shows the response to the change of $\mu$
parameter.).
Therefore they
partly cancels when added together
(the line with square dots
for $m_{\tilde{l}_L}$ = 100 GeV,
with triangle dots for $m_{\tilde{l}_L}$ = 200 GeV
for $\tan\beta$ = 3).}
\end{center}
\label{fig:ST_100}
\end{figure}
\\
\quad
Fig. \ref{fig:ST_100}
\footnote{
The author thanks G. C. Cho for drawing this figure
several times.
} shows
a constraint implied from $S$ and $T$ parameters
\footnote{
Both axes are essentially $S$, $T$ themselves here.
}.
The reference standard model is at the origin on this plane
located in the contour of 90 \% confidence level.
The slepton contribution brings $S$ parameter to negative,
while the chargino and neutralino ones to positive.
A set of two lines in the left-hand side pursues
the response
of the slepton effects
to the change of SUSY breaking slepton mass
for two values of $\tan\beta$
(solid line for $\tan\beta = 3$,
dashed line for $\tan\beta = 50$.)
The one on the right-hand side follows
the response of the chargino effects
against the change of $\mu$ parameter.
The solid line with the square (triangular) marks
represents the locus followed by
the sum of the these two contributions
for the slepton mass 100 GeV (200 GeV) and $\tan\beta$ equal to 3
when $\mu$ is changed to about 500 GeV .
Thus such a parameter set with $\tan\beta $ equal to 3
is allowed at 95 \% confidence level.
But in the case of $\tan\beta$ = 50
it is rather difficult to take slepton mass equal
to 100 GeV.
Once the slepton mass is taken larger,
for instance, at 200 GeV,
there is no restriction from this analysis.
\\
\quad
Now for $\tan\beta = 3$ the contour with equal $(a_\mu)_{\rm SUSY}$
in the $\mu$-$M_2$ plane
is drawn in Fig. \ref{fig:cnt_3_100}.
\begin{figure}[htb]
\begin{center}
\includegraphics[origin=Bc, scale=0.9]{cont_3_100.eps}
\caption{Similar contours with equal $(a_\mu)_{\rm SUSY}$
in $\mu$-$M_2$ plane
for $\tan\beta = 3$ and
$m_{\bar{\mu}_L} = m_{\bar{\mu}_R} = 100$ GeV.
The contours are drawn with the interval of
$5\times 10^{-10}$
between $-25 \times 10^{-10}$ and $10 \times 10^{-10}$
for $(a_\mu)_{\rm SUSY}$.}
\label{fig:cnt_3_100}
\end{center}
\end{figure}
With the future accuracy, which amounts to the interval
between the neighboring contours in that figure,
we can extract SUSY effect and may obtain
precise information on the model.
\section{Discussion and Summary}
\label{sec:discussion}
\quad
Now let us turn back to the theoretical uncertainty.
As was mentioned by several talks in this workshop,
besides QED contribution,
$a_\mu({\rm SM})$
is also dominated by the leading order QCD contribution
which arises through the hadronic vacuum polarization.
Its improvement is now awaiting
for the precise knowledge about the low energy
hadron production cross section
planned to be accumulated
at Novosivirsk, Frascatti and Beijing.
\\
\quad
The hadronic light-by-light scattering effect \cite{HK},
which requires purely theoretical evaluation,
may become an obstacle.
Thus the reduction of its error also needs
further challenge.
\\
\quad
To summarize we discussed
the effect to muon $g-2$ from two candidates of models
each of which accesses to ``hierarchy problem''.
We found that the potential signatures are expected
from the existence of both two candidates
by future measurement of muon $g-2$
even on account of the precision measurement
at $Z$ pole.
But this program cannot be accomplished without
improvement in measurement of the hadron production cross section
in low energy domain.
\\
{\bf Acknowledgement} \\
\vspace{0.4cm} \\
\quad
The author thanks S. Eidelman
for hospitality at Novosivirsk.
He also thanks to N. Sakai for occasion visiting at TIT
after this March,
and to G. C. Cho and K. Hagiwara
for the various comments on the preparatory content of talk.
|
\section{Introduction}
Mira and semiregular variables (SRV's) are pulsating
low and intermediate mass red giants located on the asymptotic
giant branch (AGB). The importance of these variables is highlighted by
the fact that they are primary sources for the enrichment of
interstellar medium via mass loss.
The observed pulsational behaviour may lead to a better understanding
of inner physical processes having crucial effects on
stellar evolution.
The classification scheme according to the General Catalog of
Variable Stars (GCVS) is based only on the
amplitude and regularity of the visual variation. SRV's have amplitudes
smaller than
2.5 mag in V, while typical periods range from 25 to hundreds of days.
Their basic properties (classification, temperature, luminosity, space
distribution, important spectral features) were studied in general by
Kerschbaum \& Hron (1992), Jura \& Kleinmann (1992), Kerschbaum \&
Hron (1994), Lebzelter et al. (1995),
Kerschbaum \& Hron (1996), Kerschbaum et al. (1996), Hron et al. (1997).
Although they are usually treated separately from Mira-type variables,
there has been increasing evidence of a closer relationship between
the two types of variables. Kerschbaum \& Hron (1992, 1994) claimed
that some semiregulars are more closely related to Miras than the
pure classification suggests. Szatm\'ary et al. (1996) found V Boo to
have dramatically decreasing amplitude over decades of time mimicing
evolution from the Mira to the semiregular state. A similar phenomenon
was found by Bedding et al. (1998) for R Dor, which implies that
certain groups of semiregulars may belong to a subset of Mira variables.
Bedding \& Zijlstra (1998) reached similar conclusion based on HIPPARCOS
period-luminosity relations for Mira and semiregular variables.
The mode of pulsation in SRV's raised many questions during
the last decades. A detailed review is given by Percy \& Polano (1998),
who showed that the presence of higher overtone pulsation is suggested by
the
observations (up to the third and fourth overtone). Wood et al. (1998)
presented 5 different period-luminosity sequences for the LMC
red variables based on the MACHO photometric database,
concluding similarly to Percy \& Polano (1998) that even third and
fourth overtones could be the dominant excited modes.
Bedding et al. (1998) claim that the observed mode switching
in R Dor occurs between the first and the third overtone.
All of these studies support the idea that fundamental plus
first overtone pulsation in SRV's is an oversimplified
assumption and the complex light variations may be due to
many simultaneously excited modes.
As has been mentioned above, the typical time scale of SRV's can be
hundreds of days, and consequently there are very few high-quality
photometric observations (photographic or photoelectric) in the literature.
Although micro-lensing projects (MACHO, EROS, OGLE) yielded many
theoretical constraints on stellar pulsation interpretations
(see Welch 1998 for a review), the majority of SRV's need much
longer (a few decades, at least) continuous time-series of observations.
First results
concerning red variables in the LMC have already appeared
from the MACHO group (Cook et al. 1997, Minitti et al. 1998,
Alves et al. 1998, Wood et al. 1998), but periodicities in
SRV's in our own Galaxy deserve further study.
Fortunately, a large fraction of bright SRV's have been observed
visually by amateur astronomers all around the world. There exist
50--70 years long data series which are perfectly usable for
studying periodicities in the light curves (see e.g.
Percy et al. 1993, Mattei et al. 1998, Andronov 1998).
The main aims of this study are to present a detailed light curve
analysis for 93 SRV's based on long-term visual observations
and to demonstrate the
general trends and the most interesting phenomena we found
in the analysed sample.
The paper is organised as follows. Observations are discussed and
tested
in Sect.\ 2., while Sect.\ 3. deals with the results of
period analysis, especially with multiperiodicity as a
consequence of multimode pulsation.
Interesting special cases (triple periodicity, long-term amplitude
decrease and amplitude modulation) are briefly summarized in Sect.\ 4.
\begin{table*}
\small
\begin{center}
\caption{The list of programme stars. Variability types, periods and
spectral types are taken from the GCVS.}
\begin{tabular}{lllrllllrl}
\hline
GCVS & IRAS & Type & Period & Sp. type & GCVS & IRAS & Type & Period & Sp. type \\
\hline
{\bf O-rich} & & & & & Y UMa & 12380+5607 & SRb & 168 & M7II-III:\\
RU And & -- & SRa & 238 & M5e-M6e& Z UMa & 11538+5808 & SRb & 196 & M5IIIe\\
RV And & 02078+4842 & SRa & 171 & M4e& RY UMa & 12180+6135 & SRb & 310 & M2-M3IIIe\\
V Aqr & 20443+0215 & SRa & 244 & M6e& ST UMa & 11251+4527 & SRb & 110 & M4-M5III\\
S Aql & 20093+1528 & SRa & 146 & M3e-M5.5e& R UMi & 16306+7223 & SRb & 326 & M7IIIe\\
GY Aql & 19474$-$0744 & SR & 204 & M6III:e-M8& V UMi & 13377+7433 & SRb & 72 & M5IIIab\\
T Ari & 02455+1718 & SRa & 317 & M6e-M8e& SW Vir & 13114$-$0232 & SRb & 150 & M7III\\
RS Aur & -- & SRa & 170 & M4e-M6e& RU Vul & -- & SRa & 174 & M3e-M4e\\
U Boo & 14520+1753 & SRb & 201 & M4e& & & & \\
V Boo & 14277+3904 & SRa & 258 & M6e& {\bf C-rich} & & &\\
RV Boo & 14371+3245 & SRb & 137 & M5e-M7e& ST And & 23362+3529 & SRa & 328 & C4,3e-C6,4e\\
RS Cam & 08439+7908 & SRb & 89 & M4III& VX And & 00172+4425 & SRa & 369 & C4,5\\
RR Cam & 05294+7225 & SRa & 124 & M6 & AQ And & 00248+3518 & SR & 346 & C5,4\\
RY Cam & 04261+6420 & SRb & 136 & M3III& V Aql & 19017$-$0545 & SRb & 353 & C5,4-C6,4\\
RT Cnc & 08555+1102 & SRb & 60 & M5III& S Aur & 05238+3406 & SR & 590 & C4-5\\
V CVn & 13172+4547 & SRa & 192 & M4e-M6eIIIa:& UU Aur & 06331+3829 & SRb & 234 & C5,3-C7,4\\
SV Cas & 23365+5159 & SRa & 265 & M6,5& S Cam & 05356+6846 & SRa & 327 & C7,3e\\
AA Cas & 01163+5604 & Lb & & M6III& U Cam & 03374+6229 & SRb & & C3,9-C6,4e\\
SS Cep & 03415+8010 & SRb & 90 & M5III& ST Cam & 04459+6804 & SRb & 300 & C5,4\\
DM Cep & 22073+7231 & Lb & & M4& T Cnc & 08538+2002 & SRb & 482 & C3,8-C5,5\\
RS CrB & 15566+3609 & SRa & 322 & M7 & X Cnc & 08525+1725 & SRb & 195 & C5,4\\
W Cyg & 21341+4508 & SRb & 131 & M4e-M6eIII& Y CVn & 12427+4542 & SRb & 157 & C5,4J\\
RU Cyg & 21389+5405 & SRa & 233 & M6e-M8e& RT Cap & 20141$-$2128 & SRb & 393 & C6,4\\
RZ Cyg & 20502+4709 & SRa & 276 & M7,0-M8,2ea& WZ Cas & 23587+6004 & SRb & 186 & C9,2\\
TZ Cyg & 19147+5004 & Lb & & M6& RS Cyg & 20115+3834 & SRa & 417 & C8,2e\\
AB Cyg & -- & SRb & 520 & M4IIIe& RV Cyg & 21412+3747 & SRb & 263 & C6,4e\\
AF Cyg & 19287+4602 & SRb & 93 & M5e-M7& TT Cyg & 19390+3229 & SRb & 118 & C5,4e\\
U Del & 20431+1754 & SRb & 110 & M5II-III& AW Cyg & 19272+4556 & SRb & 340 & C4,5\\
CT Del & 20270+0943 & Lb & & M7& V460 Cyg & 21399+3516 & SRb & 180 & C6,4\\
CZ Del & 20312+0920 & SRb & 123 & M5& RY Dra & 12544+6615 & SRb: & 200 & C4,5\\
EU Del & 20356+1805 & SRb & 60 & M6,4III& UX Dra & 19233+7627 & SRa: & 168 & C7,3\\
S Dra & 16418+5459 & SRb & 136 & M7& RR Her & 16028+5038 & SRb & 240 & C5,7e-C8,1e\\
TX Dra & 16342+6034 & SRb & 78 & M4e-M5& U Hya & 10350$-$1307 & SRb & 450 & C6,5\\
AH Dra & 16473+5753 & SRb & 158 & M7& V Hya & 10491$-$2059 & SRa & 531 & C6,3e-C7,5e\\
SW Gem & 06564+2606 & SRa & 680 & M5III& W Ori & 05028+0106 & SRb & 212 & C5,4\\
X Her & 16011+4722 & SRb & 95 & M6e& Y Per & 03242+4400 & M & 249 & C4,3e\\
ST Her & 15492+4837 & SRb & 148 & M6-7IIIas& SY Per & 04127+5030 & SRa & 474 & C6,4e\\
UW Her & 17126+3625 & SRb & 104 & M5e& S Sct & 18476$-$0758 & SRb & 148 & C6,4\\
g Her & 16269+4159 & SRb & 89 & M6III& Y Tau & 05426+2046 & SRb & 242 & C6.5,4e\\
RT Hya & 08272$-$0609 & SRb & 290 & M6e-M8e& SS Vir & 12226+0102 & SRa & 364 & C6,3e\\
RY Leo & 10015+1413 & SRb & 155 & M2e& & & & \\
U LMi & 09516+3619 & SRa & 272 & M6e& {\bf Uncertain} & & & & \\
RX Lep & 05090$-$1155 & SRb & 60 & M6,2III& T Cen & 13388$-$3320 & SRa & 90 & K0:e-M4II:e\\
SV Lyn & 08003+3629 & SRb & 70 & M5III& AI Cyg & 20297+3221 & SRb & 197 & M6-M7\\
X Mon & 06548$-$0859 & SRa & 156 & M1eIII-M6ep& GY Cyg & -- & SRb & 300 & M7p\\
BQ Ori & 05540+2250 & SR & 110 & M5IIIe-M8III& V930 Cyg & 19371+3021 & Lb & & \\
UZ Per & 03170+3150 & SRb & 927 & M5II-III& RS Gem & 06584+3035 & SRb & 140 & M3-M8\\
$\tau^4$ Ser & 15341+1515 & SRb & 100 & M6IIb-IIIa& RX UMa & 09100+6728 & SRb & 195 & M5\\
W Tau & 04250+1555 & SRb & 265 & M4-M6.5& & & & &\\
V UMa & 09047+5118 & SRb & 208 & M5-M6& & & & &\\
\hline
\end{tabular}
\end{center}
\end{table*}
\section{Observations}
The bulk of the analysed data were taken from three
international databases of visual observations.
These belong to the Association Francaise des Observateurs d'Etoiles
Variables (AFOEV\footnote{\tt ftp://cdsarc.u-strasbg.fr/pub/afoev}),
the Variable Star Observers' League in Japan
(VSOLJ\footnote{\tt http://www.kusastro.kyoto-u.ac.jp/vsnet/gcvs})
and the Hungarian Astronomical Association -- Variable Star Section
(HAA/VSS\footnote{\tt http://www.mcse.hu/vcssz/data}).
The compiled visual estimates are stored at publicly
available web-sites as Julian Date + magnitude files.
A smaller fraction of data originated from the American
Association of Variable Star Observers (AAVSO International
Database which includes HAA/VSS, and part of AFOEV and
VSOLJ observations).
\begin{figure}
\begin{center}
\leavevmode
\psfig{figure=8380f1.eps,width=\linewidth}
\caption{A comparison of the visual light curves for Y UMa from
AFOEV and VSOLJ.
The differences do not exceed the uncertainty of the individual points
which is about $\pm0.3$ mag. The top panel shows the original data, while
the averaged curves are plotted in the bottom panel.}
\end{center}
\label{f1}
\end{figure}
The main selection criterion in choosing the sample was the length
and the continuity of the light curves. In order to reach high resolution
in the frequency domain, we usually kept only those stars with at least 10 years
of continuous data. This is equivalent to a frequency resolution
($\sim$time$^{-1}$) of {$2.7\cdot10^{-4}$} cycles/day. In most cases
the length of the analysed data is about 50 years, and occasionally it is
70--80 years. The final sample containing 93 semiregular variables
is summarized in Table 1, with the main information taken from the GCVS.
Y~Per (classified as a Mira star in the GCVS) was included because of its
recently observed semiregular nature (see Sect.\ 4.2.). A few
Lb-type variables are also included, as their classification is a
quite uncertain issue; recent studies of Kerschbaum et al. (1996)
and Kerschbaum \& Olofsson (1998) pointed out the close similarity
of selected Lb's and SRV's based on the infrared and mass-loss properties.
There are two steps in the data handling that precede
before the period analysis: (1) data averaging using 10-day bins, and
(2) merging of observations of different origin. We performed a few
simple tests
to decide whether merging should precede or follow the averaging. We plotted
the different original light curves together for the best observed
stars and found that the systematic differences did not exceed the level
of the scatter in the data. One example can be seen in Fig.\ 1, where we
have plotted
the French and Japanese data for Y~UMa (type SRb). The error of
an individual point is estimated to be about $\pm0.3$ mag.
The two curves are very similar,
which suggests that the comparison sequences define
a well determined system of visual magnitudes. A similar conclusion was drawn
for the majority of the stars in our sample, so we simply merged the
available data before calculating the averaged curves.
We could possibly reach somewhat better precision by introducing personal
corrections for the most active observers, but as other tests have
shown, the length of data is much more important in determining
periodicities -- which is our main goal -- than is the accuracy of
the individual measurements (see below). A thorough review of the
homogeneity of visual photometry is given by Sterken \& Manfroid (1992).
The averaging procedure consisted of taking 10-day bins and calculating
the mean value from the individual points. Since the typical time scale
of the period
in our sample of semiregular variables is about one hundred days,
this binning
procedure does not smooth out significant detail in the light variations.
A rough estimate
of the resulting improvement in
precision is as follows. As we mentioned earlier, the error of an
individual observation is about $\pm0.3$ mag. For a given 10-day bin with
10 points within it, the standard error of the mean value will
be $0.3/\sqrt{10}\approx0.1$ mag. The amount of data and their
distribution in our sample in most cases permit such precision
to be realized. Extremely deviant points differing from the mean value by
more than $\sim 3 \sigma$ were rejected in the original data
during a close visual inspection of all light curves.
\subsection{Tests of the quality and usability}
\begin{figure*}
\begin{center}
\leavevmode
\psfig{figure=8380f2.eps,width=14cm}
\caption{Comparison between the photoelectric and visual observations
for RY~Dra.}
\end{center}
\label{f2}
\end{figure*}
\begin{figure*}
\begin{center}
\leavevmode
\psfig{figure=8380f3.eps,width=14cm}
\caption{Artifical data with different S/N ratios and their corresponding
DFT spectra. It can be seen that length is much more important that
the quality of data. Note that top panels show only subsets.}
\end{center}
\label{f3}
\end{figure*}
Three tests were made to check the reliability and usability
of the resulting
mean light curves. The first was comparison with available
simultaneous photoelectric V-measurements. Although the spectral
response function of the human eye differs from the Johnson
V filter, there were and continiue to be several attempts to
find calibrations of transformations.
Zissell (1998) modified the zeropoint of the conversion formula proposed
by Stanton (1981) and gave the following relation:
\begin{center}
$$m_{vis.}~=~ V~+~0.182~(B-V)~-~0.032. \eqno (1)$$
\end{center}
\noindent According to the available
photoelectric observations of semiregular variables, in most cases their
$B-V$ colour changes with much smaller amplitude than V brightness does,
thus it is a straightforward simplifying assumption that there is
a constant shift between the photoelectric V and visual light curves.
This is, of course, true only at a level of about 0.1 mag, which is
in the range of the scatter of the visual data.
A direct comparison is shown in Fig.\ 2, where we compare visual data
for RY~Dra with simultaneous photoelectric V measurements carried
out at Grinnel College. The top curve is the
photoelectric one, while the bottom curve is the corresponding
10-day mean of visual data. The middle curve is a noise filtered
version of the lower curve, where noise filtering was done by
a simple Gaussian smoothing with 8 days FWHM. Note that while
the visual curves were shifted in a vertical direction for clarity,
the distance between the smoothed visual and the photoelectric curve
is the real difference caused by the colour effects.
The observed average shift of 0.60 mag is in good agreement
with the predicted 0.57 mag by Eq.\ 1 ($\langle B-V \rangle\approx~3.3$
for RY~Dra).
The agreement between
the visual and photoelectric curves are very good, even the smallest
humps and bumps, of 0.1 mag are clearly visible in the
visual data. A similar conclusion can be drawn using Hipparcos Tycho V
data (ESA 1997\footnote{\tt http://astro.estec.esa.nl/Hipparcos}):
visual observations define light curves that are very similar
to the photoelectric ones. We have to note, that the Gaussian data
smoothing was applied only here because of its illustrative power,
our main analyses were based on the 10-day binned light curves only.
Another test was performed as a numerical simulation in order
to study the effect of the length of the data set versus the
signal-to-noise ratio (S/N).
We generated artificial time-series by adding three monoperiodic
signals with very similar periods and amplitudes to those
observed in real variables (e.g. A$_0=0.7$ mag, P$_0=1000$ days,
A$_1=0.3$ mag, P$_1=140$ days, A$_2=0.5$ mag, P$_2=77$ days).
Additional white noise was added to get artificial S/N values of 100
and 1, respectively. The ``observed'' time ranged from JD 2435000 to 2451000.
We calculated the Discrete Fourier Transforms (DFT) for both datasets
and the results are shown in Fig.\ 3.
It is obvious that the period (and amplitude) determination is
almost completely independent of the S/N ratio if the time-series
is long enough.
This can be explained by the fact that the applied noise
is independent of the current brightness and consequently
these two quantities are also independent in the frequency domain.
Real observations
come from many different observers who made their estimates
independently, therefore the observational noise is uncorrelated.
We have extensively
explored this question and our conclusion is that the analysed time-series
fulfill all requirements for accurate period analysis. This result is similar
to that of Szatm\'ary \& Vink\'o (1992). We have to
note that the independence of noise and observations can be
assumed only for bright semiregulars with amplitudes
that are not too large.
For Mira variables, which can become quite faint at minimum light,
the data obtained by observers using small telescopes will
have more scatter near the minima in the light curves.
Following the referee's note on using the averaged data, we have
performed a third test addressed to the effects of the binning. The most
important effect is the decrease of amplitude due to the binning, while
the resulting frequencies may differ a bit, too.
We explored this question by analysing the unbinned, 5-day and 10-day mean
light curves.
\begin{figure}
\begin{center}
\leavevmode
\psfig{figure=8380f4.eps,width=\linewidth}
\caption{The effects of binning for RY~UMa.
{\it Top panel}: comparison of the different Fourier-spectra
calculated from the original data and 5-day, 10-day bins.
{\it Bottom panel}: comparison of the four-component fitted curves
with the observations (small dots). A vertical shift
of $\pm$1 mag was applied for clarity.}
\end{center}
\label{f4}
\end{figure}
\begin{table}
\begin{center}
\caption{The frequencies and amplitudes of four principal peaks in the
Fourier-spectra plotted in Fig.\ 4.}
\begin{tabular} {llll}
\hline
& unbinned & 5-day & 10-day\\
\hline
f$_0$ & 0.000104 & 0.000100 & 0.000100\\
A$_0$ & 0.119 & 0.146 & 0.152\\
f$_1$ & 0.000272 & 0.000252 & 0.000248\\
A$_1$ & 0.178 & 0.170 & 0.165\\
f$_2$ & 0.003276 & 0.003276 & 0.003271\\
A$_2$ & 0.145 & 0.153 & 0.156\\
f$_3$ & 0.003500 & 0.003480 & 0.003480\\
A$_3$ & 0.119 & 0.108 & 0.113\\
\hline
\end{tabular}
\end{center}
\end{table}
The results of this test are briefly summarized in Fig.\ 4 with those
of obtained for RY~UMa (type SRb). We plotted three different Fourier-spectra
calculated from the binned and original, unbinned data.
The principal peaks have slightly different frequencies and amplitudes
(Table 2),
but the four-component fits (bottom panel in Fig.\ 4) do not differ
significantly. This suggests that the differences are mainly due to the
uncertainty of the whole analysis caused by the noisy data and not
particularly due to the
averaging. We conclude that the averaging procedure does not
introduce significant alias structures, if the light curves are
densely covered by many independent observations. We obtained similar
results even for stars with periods of about 100 days (e.g. TX~Dra) suggesting
that data binning does not affect too seriously the calculated periods.
The amplitudes have, of course, larger uncertainties, but as they
may have cycle-to-cycle changes, this aspect is beyond our present scope.
\section{Period analysis}
We calculated Discrete Fourier Transforms (DFT) of the merged and
averaged time-series. The frequency ranged from 0 to 0.025 cycles/day,
while the frequency step was chosen as $4\cdot10^{-6}$ c/d.
The code used was Period98 (Sperl 1998\footnote{\tt
http://dsn.astro.univie.ac.at/period98}).
A few sample power spectra are presented in Figs.\ 5-6 and
Sect.\ 4.
\begin{figure*}
\begin{center}
\leavevmode
\psfig{figure=8380f5.eps,width=\textwidth}
\caption{Sample spectra of low-amplitude variables
with one- (DM~Cep and EU~Del), two- (TT~Cyg) and three-component
(UW~Her) fits. The averaged light curves were calculated with 10-day bins.}
\end{center}
\label{f5}
\end{figure*}
\begin{figure*}
\begin{center}
\leavevmode
\psfig{figure=8380f6.eps,width=\textwidth}
\caption{Sample spectra of medium amplitude variables
with one- (S~Cam), two- (U~LMi and V~Hya) and three-component
(RX~UMa) fits. The averaged light curves were calculated with
10-day bins.}
\end{center}
\label{f6}
\end{figure*}
\begin{table}
\begin{center}
\caption{Monoperiodic variables. $\langle m \rangle$ is the
mean visual brightness, $\Delta T$ is the length of the time-series.
The numbers in brackets denote
the estimated uncertainty based on the width of the peaks in the
Fourier spectra. $A$ means the semi-amplitude.}
\begin{tabular} {lrrlll}
\hline
Star & $\langle m \rangle$ & $\Delta T$ & P (GCVS) & P & $A$\\
\hline
RV And & 10.0 & 22300 & 171 & 165(5) & 0.30\\
S Aql & 10.5 & 25800 & 146 & 143(3) & 0.98\\
GY Aql & 12.3 & 3000 & 204 & 464(4) & 2.35\\
T Ari & 9.5 & 33000 & 317 & 320(3) & 0.91\\
S Aur & 11.2 & 18800 & 590 & 596(6) & 0.61\\
U Boo & 11.2 & 23300 & 201 & 204(3) & 0.62\\
RV Boo & 8.5 & 7000 & 137 & 144(2) & 0.09\\
S Cam & 9.3 & 28000 & 327 & 327(1) & 0.81\\
RY Cam & 8.4 & 8500 & 136 & 134(1) & 0.16\\
T Cnc & 9.0 & 12400 & 482 & 488(4) & 0.34\\
RT Cap & 7.5 & 24000 & 393 & 400(4) & 0.31\\
T Cen & 7.0 & 12400 & 90 & 91(1) & 0.62\\
DM Cep & 7.9 & 6500 & -- & 367(3) & 0.12\\
RS CrB & 7.5 & 12400 & 322 & 331(1) & 0.19\\
AI Cyg & 9.0 & 3000 & 197 & 146(2) & 0.18\\
GY Cyg & 10.6 & 8000 & 300 & 143(1) & 0.13\\
V460 Cyg & 6.5 & 6400 & 180 & 160(10) & 0.08\\
V930 Cyg & 12.5 & 2000 & -- & 247(3) & 0.72\\
EU Del & 6.2 & 11000 & 60 & 62(1) & 0.08\\
SW Gem & 8.8 & 11800 & 680 & 700(10) & 0.10\\
RR Her & 9.0 & 27800 & 240 & 250(10) & 0.54\\
RT Hya & 8.1 & 12400 & 290 & 255(3) & 0.20\\
U Hya & 5.3 & 27800 & 450 & 791(5) & 0.06\\
X Mon & 8.4 & 26600 & 156 & 148(7) & 0.59\\
SY Per & 10.7 & 5000 & 474 & 477(9) & 0.89\\
UZ Per & 8.7 & 4000 & 927 & 850(10) & 0.25\\
W Tau & 10.4 & 24800 & 265 & 243(3) & 0.27\\
V UMa & 10.4 & 12400 & 208 & 198(2) & 0.19\\
SS Vir & 8.3 & 25500 & 364 & 361(1) & 0.81\\
\hline
\end{tabular}
\end{center}
There are some episodes in the light variation of
RV~And, S~Aql and U~Boo caused by possible mode
switching (Cadmus et al. 1991).
\end{table}
We did not try to extract as many periods as possible from the
power spectra because the excited frequencies in semiregular variables
are not stable over time (see, e.g., Mattei et al. 1998).
The DFT may contain many misleading peaks
because of the cycle-to-cycle variations. These changes make
impossible to fit simple sums of sines. Our approach was to accept only
the most dominant periods which were
tested by whitening, cleaning and alias filtering. One-year
alias peaks occur for many stars, while in some cases cross production
terms are present as well (e.g. $f_0, f_1, f_0\pm f_1$). We did
an iterative period determination allowed by Period98 (Sperl 1998), in which
we checked the consistency of the fitted harmonics and the light
curve itself after every step. The frequency identification was a
quite difficult task in certain stars (e.g. TT~Cyg, TZ~Cyg and other
low-amplitude variables), mainly because of the long-term
changes in the mean brightness. They may cause fairly high false
peaks in the low-frequency region which have to be subtracted and
neglected in searching for pulsational periods. This involves
some additional uncertainty which can be hardly avoided.
The instabilities of the excited frequencies may cause multiple
peaks scattering around local average values in the Fourier spectra,
therefore,
in some cases (TT~Cyg, X~Her, TZ~Cyg, SW~Gem, U~Hya, S~Sct, SW~Vir)
we could only estimate the periods and amplitudes
with help of a paralel comparison of the multiple structures of the DFT
and the observed cyclic changes in the light curves. The determined
``periods'' in the quoted stars should be rather considered as
characteristic times of variations instead of real periods.
The resulting periodicities can be summarized as follows.
Among the 93 semiregulars, we have found 29 purely
monoperiodic stars, 56 stars with unambiguous multiperiodic
behaviour (44 bi- and 12 triperiodic), and 8 stars which
turned out to be rather irregular (meaning
that we did not find any peak higher than the calculated noise
level for those variables). We present the
main observational properties (average brightness, length of analysed
data in days) as well as the calculated
periods and their amplitudes in Tables 3--5. The period uncertainty
was estimated
from the width of the peaks in the spectra at 90\% of maximum.
One star, RS Cam, apparently has a fourth period ($P_3=81$ days,
$A_3=0.1$ mag.) as well.
The amplitude values have more uncertainty than do the periods
because of the instability of the periods. This issue was studied
by wavelet analysis (see Sect.\ 4 for examples), which is a
useful tool for studying temporal variations in the frequency content
(see, e.g., Bedding et al. 1998, Barth\'es \& Mattei 1997,
Szatm\'ary et al. 1996, Foster 1996, G\'al \& Szatm\'ary 1995a,
Szatm\'ary et al. 1994,
Koll\'ath \& Szeidl 1993, Szatm\'ary \& Vink\'o 1992). Therefore,
the amplitude values listed in Tables 3--5 only serve to indicate
the approximate relative strengths of the corresponding periods.
\begin{table*}
\begin{center}
\caption{Biperiodic variables. The symbols are the same as in Table 3.
Subscripts ``0'' and ``1'' simply correspond to the longer and shorter
periods. Periods being in good agreement with the values listed in the GCVS
are typesetted separately.}
\begin{tabular} {lrrlllll}
\hline
Star & $\langle m \rangle$ & $\Delta T$ & P (GCVS) & $P_0$ & $A_0$ & $P_1$ & $A_1$\\
\hline
ST And & 10.0 & 5300 & {\bf 328} & {\bf 338(2)} & 1.15 & 181(1) & 0.20\\
VX And & 8.5 & 5800 & {\bf 369} & 904(5) & 0.16 & {\bf 375(2)} & 0.27\\
AQ And & 8.6 & 8200 & {\bf 346} & {\bf 346(1)} & 0.15 & 169(1) & 0.18\\
V Aqr & 8.7 & 22800 & {\bf 244} & 689(5) & 0.25 & {\bf 241(2)} & 0.35\\
V Aql & 7.4 & 28300 & 353 & 400(50) & 0.10 & 215(1) & 0.11\\
RS Aur & 10.0 & 22300 & {\bf 170} & {\bf 173(1)} & 0.29 & 168(1) & 0.23\\
UU Aur & 5.7 & 21800 & {\bf 234} & 441(2) & 0.15 & {\bf 235(2)} & 0.09\\
V Boo & 8.7 & 27800 & {\bf 258} & {\bf 257(1)} & 0.86 & 137(1) & 0.19\\
RR Cam & 10.5 & 12400 & {\bf 124} & 223(2) & 0.09 & {\bf 124(1)} & 0.10\\
V CVn & 7.5 & 26400 & {\bf 192} & {\bf 194(1)} & 0.42 & 186(1) & 0.13\\
SV Cas & 8.6 & 25000 & {\bf 265} & 460(4) & 0.48 & {\bf 262(2)} & 0.32\\
WZ Cas & 7.2 & 11000 & {\bf 186} & 373(1) & 0.16 & {\bf 187(1)} & 0.09\\
SS Cep & 7.3 & 27000 & {\bf 90} & 340(10) & 0.07 & {\bf 100(5)} & 0.05\\
W Cyg & 6.2 & 33400 & {\bf 131} & 240(5) & 0.05 & {\bf 130(5)} & 0.14\\
RS Cyg & 8.0 & 27800 & {\bf 417} & {\bf 422(4)} & 0.48 & 211(2) & 0.20\\
RU Cyg & 8.5 & 26800 & {\bf 233} & 441(1) & 0.16 & {\bf 234(1)} & 0.96\\
RZ Cyg & 11.8 & 28800 & {\bf 276} & 537(2) & 0.63 & {\bf 271(4)} & 0.86\\
TT Cyg & 8.0 & 25500 & 118 & 390(10) & 0.03 & 188(5) & 0.03\\
TZ Cyg & 10.8 & 9500 & -- & 138(1) & 0.06 & 79(1) & 0.05\\
AB Cyg & 7.9 & 25400 & {\bf 520} & {\bf 513(2)} & 0.17 & 429(2) & 0.07\\
AW Cyg & 8.9 & 22300 & 340 & 3700(50) & 0.10 & 387(3) & 0.10\\
U Del & 7.0 & 29800 & 110 & 1146(10) & 0.21 & 580(5) & 0.05\\
S Dra & 8.9 & 26600 & 136 & 311(1) & 0.12 & 172(2) & 0.12\\
RY Dra & 7.0 & 8800 & 200 & 1150(20) & 0.20 & 300(10) & 0.10\\
UX Dra & 6.7 & 8000 & {\bf 168} & 317(2) & 0.08 & {\bf 176(1)} & 0.10\\
AH Dra & 7.8 & 8600 & 158 & 189(1) & 0.25 & 107(1) & 0.12\\
RS Gem & 10.6 & 12400 & {\bf 140} & 271(1) & 0.25 & {\bf 148(1)} & 0.22\\
X Her & 6.7 & 33500 & {\bf 95} & 178(5) & 0.05 & {\bf 102(5)} & 0.03\\
ST Her & 7.9 & 25000 & {\bf 148} & 263(2) & 0.08 & {\bf 149(1)} & 0.08\\
g Her & 5.1 & 9000 & {\bf 89} & 887(5) & 0.20 & {\bf 90(1)} & 0.07\\
V Hya & 9.1 & 32600 & {\bf 531} & 6400(50) & 1.22 & {\bf 531(3)} & 0.66\\
RY Leo & 10.2 & 22000 & {\bf 155} & {\bf 160(1)} & 0.40 & 145(1) & 0.28\\
U LMi & 11.8 & 10000 & {\bf 272} & {\bf 272(2)} & 0.45 & 144(1) & 0.16\\
W Ori & 6.5 & 33300 & {\bf 212} & 2390(20) & 0.15 & {\bf 208(1)} & 0.08\\
BQ Ori & 7.9 & 26000 & 110 & 240(6) & 0.14 & 127(2) & 0.10\\
Y Per & 9.4 & 24000 & {\bf 249} & {\bf 245(1)} & 0.36 & 127(1) & 0.16\\
S Sct & 7.3 & 22400 & {\bf 148} & 269(5) & 0.08 & {\bf 149(2)} & 0.10\\
$\tau^4$ Ser & 6.7 & 5800 & 100 & 1240(10) & 0.10 & 111(1) & 0.09\\
Y Tau & 7.4 & 27000 & {\bf 242} & 461(2) & 0.18 & {\bf 242(1)} & 0.18\\
Z UMa & 7.8 & 31000 & {\bf 196} & {\bf 195(1)} & 0.33 & 100(1) & 0.05\\
RY UMa & 7.3 & 9800 & {\bf 310} & {\bf 305(1)} & 0.16 & 287(1) & 0.11\\
ST UMa & 6.8 & 26000 & 110 & 5300(20) & 0.11 & 615(6) & 0.09\\
R UMi & 9.7 & 28000 & {\bf 326} & {\bf 325(1)} & 0.42 & 170(1) & 0.11\\
RU Vul & 9.3 & 19500 & 174 & 369(2) & 0.13 & 136(1) & 0.11\\
\hline
\end{tabular}
\end{center}
\end{table*}
\begin{table*}
\begin{center}
\caption{Triply periodic variables.}
\begin{tabular} {lrrlllllll}
\hline
Star & $\langle m \rangle$ & $\Delta T$ & P (GCVS) & $P_0$ & $A_0$ & $P_1$ & $A_1$ & $P_2$ & $A_2$\\
\hline
U Cam & 8.2 & 26800 & -- & 2800(100) & 0.13 & 400(30) & 0.09 & 220(5) & 0.09\\
RS Cam & 8.7 & 25000 & {\bf 89} & 966(10) & 0.17 & 160(1) & 0.15 & {\bf 90(1)} & 0.12\\
ST Cam & 7.3 & 28000 & 300 & 1580(10) & 0.10 & 372(3) & 0.12 & 202(2) & 0.08\\
X Cnc & 6.7 & 25800 & {\bf 195} & 1870(10) & 0.08 & 350(3) & 0.08 & {\bf 193(1)} & 0.09 \\
Y CVn & 5.7 & 28500 & {\bf 157} & 3000(100) & 0.08 & 273(3) & 0.06 & {\bf 160(2)} & 0.05\\
AF Cyg & 7.2 & 26600 & {\bf 93} & 921(10) & 0.08 & 163(1) & 0.11 & {\bf 93(1)} & 0.11\\
TX Dra & 7.6 & 26800 & {\bf 78} & 706(2) & 0.10 & 137(1) & 0.06 & {\bf 77(3)} & 0.07\\
UW Her & 8.1 & 8600 & {\bf 104} & 1000(10) & 0.09 & 172(1) & 0.08 & {\bf 107(1)} & 0.09\\
Y UMa & 8.6 & 32000 & {\bf 168} & 324(1) & 0.16 & 315(1) & 0.09 & {\bf 164(2)} & 0.06\\
RX UMa & 10.6 & 33200 & {\bf 195} & {\bf 201(1)} & 0.37 & 189(1) & 0.26 & 98(0.5) & 0.16\\
V UMi & 8.1 & 29000 & {\bf 72} & 737(10) & 0.06 & 126(2) & 0.04 & {\bf 73(0.5)} & 0.06\\
SW Vir & 7.6 & 8500 & {\bf 150} & 1700(50) & 0.15 & 164(1) & 0.13 & {\bf 154(1)} & 0.20\\
\hline
\end{tabular}
\end{center}
\end{table*}
\subsection{Discussion of the multiperiodic nature}
Mattei et al. (1998, hereafter M98) found 30 semiregular variables
with two periods.
Our periods for the 16 common stars
are in very good agreement. This is also true for the two triply
periodic variables, V~UMi and TX~Dra. Since the semiregulars
have quite noisy light curves due to the intrinsic short-timescale
structure, it is worth comparing the independently
determined periodicities by plotting period ratios
against our periods (Fig.\ 7). The significantly deviant points are
those of V~UMi, Y~CVn and S~Dra. This can be explained by the
instability of the periods and the different length of the analysed
data in M98.
The time span of the dataset studied here is more than twice
that of M98. Because the period and amplitude may be changing
over time, the results obtained with datasets covering different
time spans will be different. Thus, we conclude that
the applied period determination and alias filtering
give consistent results with the earlier independent study.
\begin{figure}
\begin{center}
\leavevmode
\psfig{figure=8380f7.eps,width=\linewidth}
\caption{Intercomparison of 24 periods for 16 stars in common with
Mattei et al. (1998). The differences do not exceed the amount of
instrinsic instability of the periods (up to a few percent cycle-to-cycle
changes).}
\end{center}
\label{f7}
\end{figure}
In order to examine the general distribution of the periods, we
made pairs of periods in 56 multiperiodic variables. This was done
in triply periodic stars
by sorting the periods and choosing the two neighbouring values.
Shorter periods against the longer ones are plotted in Fig.\ 8.
Three sequences are clearly present while two others are suggested.
All of them are marked by dashed lines that were
drawn by fitting a least-squares linear trend to the most
populated sequence and shifting that line to match the other sequences.
It is very interesting how parallel these ridges of data points are.
One would expect
such a separation between stars pulsating in different modes (i.e., for
the same ``longer'' period value different ``shorter'' periods
correspond), assuming that the periodicities are due to pulsation.
\begin{figure}
\begin{center}
\leavevmode
\psfig{figure=8380f8.eps,width=\linewidth}
\caption{The ``shorter'' periods vs. the ``longer'' periods of pairs
of periods for the multiply periodic variables. Three sequences are obvious
and two other ones are possible.}
\end{center}
\label{f8}
\end{figure}
Since pulsation theory usually uses period ratios to predict the modes,
we plotted the Petersen diagram (period ratios vs. periods) in Fig.\ 9.
The most populated region contains stars
with period ratios between 1.80 and 2.00, while the other
sequences are those of with ratios of around 10--12, 6.0, 3.60--3.90
and 1.02--1.10. Mattei et al. (1998) pointed
out that in their sample 63\% (19 of 30) of the multiperiodic stars
have period ratios between 1.80 and 2.00.
Our larger sample supports this statistic,
as 39 of 56 stars (70\%) have period ratios of around $1.90\pm0.15$. The
dashed lines in Fig.\ 9 were fitted by the same procedure as in Fig.\ 8.
\begin{figure}
\begin{center}
\leavevmode
\psfig{figure=8380f9.eps,width=\linewidth}
\caption{The Petersen diagram for all multiperiodic semiregulars
with triply-periodic variables shown individually.}
\end{center}
\label{f9}
\end{figure}
The other period ratios are interesting, too. The lowest sequence is that
of stars with period ratios somewhat higher than 1. This means two closely
separated periods, assuming that the close peaks are not due to
small changes of one period. We have checked the original light
curves and found very clear examples for beating (e.g. RU~And, RX~UMa).
On the other hand, some stars should be considered as monoperiodic
variables with slightly and randomly changing period. Unfortunately,
it is very difficult to distinguish between these possibilities.
The upper sequence in Fig.\ 9 is populated by stars with period ratios
of around
10. This is a well-known period ratio for semiregulars
(e.g. Houk 1963, Wood 1976, Percy \& Polano 1998). The intermediate
ratios were not
discussed in the earlier papers, although they are present in their
observational analyses to a smaller extent (see Fig.\ 6 in M98).
The assumption of a few simultaneously excited modes can be tested
by the triply-periodic
variables. If their period ratios fall on the sequences defined
by the doubly periodic stars that would support the assumption.
Fig.\ 9 shows those stars separately. The
main sequences are evidently well covered by the triply-periodic
semiregulars only, too.
Most recently, Bedding et al. (1998) have studied the mode switching
in R~Dor ($P_{long}/P_{short}=1.81$) concluding that
it probably pulsates in the
first and third overtones. Furthermore, they suggest that all stars
with similar period ratios pulsate in these modes. Wood et al. (1998)
presented multiple structures in a diagram similar to Fig.\ 8 for
the LMC red variables, and also suggested higher modes than fundamental
and first overtone.
Although Figs.\ 8--9 supports the idea that the segregation is a consequence
of the presence of many modes of pulsation, other possible explanations
could not be excluded. Older models by Fox \& Wood (1982) predict high
period ratios (6--10) for masses as high as 6--8 $M_{\odot}$, while
the periods of fundamental and first overtone radial modes
have ratios of about 2 in many theoretical models
(e.g. Ostlie \& Cox 1986, Fox \& Wood 1982). On the other hand,
quasi-periodic cycles might be caused by physical mechanisms other
than pulsation (e.g. duplicity, distorted stellar shapes,
rotation -- Barnbaum et al. 1995, dust-shell dynamics --
H\"ofner et al. 1995).
Nevertheless, we can claim that: {\it i)}
a significant percentage
of semiregular stars show multiperiodic behaviour; {\it ii)} there is
supporting evidence provided by the triply periodic variables
that the segregation in Figs.\ 8--9 is due to different modes of pulsation.
We have tried to find correlations among the periods, the period
ratios, and several main physical properties, such as the infrared
JHKL'M colours (Kerschbaum \& Hron 1994), galactic latitude, and
mass-loss rates (Loup et al. 1993). No correlation was
found among these parameters. We have also tried to find
a distinction between the C-rich and O-rich variables, but the
photometric parameters studied did not allow to determine such
a discrimination. Nevertheless, we plotted the period distribution
of the two types of stars in Fig.\ 10. This diagram is strongly
biased by the effects of the sample selection, as noted by the
referee: long-period O-rich stars would have on the average
larger amplitudes (due to the O-rich opacity sources, such as
VO, TiO), and would be classified as Miras and consequently not enter the
sample. Visual C-rich stars with small amplitudes have on
the average higher luminosities and therefore longer periods.
The simple Gaussian fits marked by the solid and
dashed lines were used to estimate the maximum and the spread
of the distributions (186 and 295 days for O-rich and C-rich
stars, respectively; the FWHM is 0.44$\pm$0.07 dex for both fits).
\begin{figure}
\begin{center}
\leavevmode
\psfig{figure=8380f10.eps,width=\linewidth}
\caption{The period distribution of C-rich and O-rich semiregulars.}
\end{center}
\label{f10}
\end{figure}
\section{Special cases}
This section deals with examples illustrating our analysis procedure and the
application of wavelet analysis. The stars mentioned below as well all
other stars will
be investigated in more details in a subsequent paper. Here we
briefly outline only what we found especially interesting. The examples
cover triple periodicity (TX~Dra and V~UMi), amplitude
modulation (RY~UMa) and long-term amplitude decrease
(V~Boo, RU~Cyg and Y~Per).
\subsection{TX Draconis \& V Ursae Minoris}
\begin{figure}
\begin{center}
\leavevmode
\caption{Fourier spectrum and wavelet contour map for TX Dra. The small
insert shows the window function (the frequency range is
$-$0.015--0.015 c/d). There are a few occasions (indicated
by horizontal arrows) when the 77 days mode (vertical arrow) dominated
the spectrum.}
\end{center}
\label{f11}
\end{figure}
\begin{figure}
\begin{center}
\leavevmode
\caption{The same as in Fig.\ 10 for V UMi. The close similarity is evident.}
\end{center}
\label{f12}
\end{figure}
The clearest examples of triple periodicity are TX~Dra and
V~UMi. In many respects they are twins in their pulsational
characteristics. The dominant modes of TX Dra and V UMi have periods
of 77 days and 73 days, respectively.
The other periods
are also very similar: 706 and 137 days for TX~Dra versus
737 and 126 days for V~UMi. This result is in perfect
agreement with that of Mattei et al. (1998), except for the longest
period in V~UMi. This can be explained by the instability of this
period which caused a double peak around 750 days in the Fourier
spectrum with slightly differing amplitudes. These peaks correspond
exactly to our 737$\pm$10 days and M98's 773 days periods. The data
distribution is not the same in the two analysed data sets which affected
the calculated amplitudes.
We studied the stability of the frequency content by wavelet
analysis (e.g. Szatm\'ary et al. 1996, Foster 1996, Szatm\'ary et al. 1994).
The resulting three-dimensional wavelet contour maps and the
corresponding Fourier spectra are shown in Figs.\ 11-12.
The most unstable period in TX~Dra is the shortest one (77 days). While the
other two modes seem to be quite stable over thousands of days,
the short period component sometimes switches on and off. That is
why there are many peaks in the power spectrum scattering around the
average value of 77 days.
Observers should note that the dominant mode is, as of February 1999,
this rapid one, and based on earlier behaviour, we expect
that it will switch off around 1999--2000, thereby offering
a very good opportunity to observe mode switching in real-time!
V~UMi is generally very similar to TX~Dra, even in the instability
of the excited modes. Obviously the pulsation in these stars
is not a smooth and repetitive process. Mild chaos is probably
present, too, as suggested by, e.g., Mattei et al. (1998). The extended
atmosphere with strong inner convection creates a very complex environment
where slight changes in the actual parameters have very serious
effects on the resulting pulsational properties.
The change between pulsational modes has been detected in the case of some
53~Per stars (Smith 1978), in a rapidly oscillating Ap star
(Kreidl et al. 1991), and in F supergiant (Fernie 1983).
Mode switching in red semiregular stars was reported by Cadmus et al. (1991),
G\'al \& Szatm\'ary (1995b), and Percy \& Desjardins (1996).
Further cases of mode switching and models for this phenomenon are discussed
by Bedding et al. (1998).
\subsection{RY Ursae Majoris}
\begin{figure}
\begin{center}
\leavevmode
\psfig{figure=8380f13.eps,width=\linewidth}
\caption{RY UMa, the best example of amplitude modulation.
The cycle length of the modulation is about 4000 days, which
equals about 13 $P_{puls.}$.}
\end{center}
\label{f13}
\end{figure}
Amplitude modulation in pulsating variables is mainly associated
with RR~Lyrae variables showing the Blazhko-effect (e.g. Kov\'acs 1995,
Szeidl 1988, Moskalik 1986), with $\delta$ Scuti-type stars with
very complex light variations
(e.g. Mantegazza et al. 1996, Breger 1993), one
known classical Cepheid, V473~Lyr, which has strong amplitude
modulation (Van Hoolst \& Waelkens 1995), and some
Mira and semiregular variables (Mattei 1993, Mattei et al. 1998,
Mattei \& Foster 1999a, b, Barth\'es \& Mattei 1997).
\begin{figure*}
\begin{center}
\leavevmode
\caption{Compressed light curves, power spectra and wavelet maps
for V Boo and
RU Cyg. These stars probably evolve from the Mira state to the semiregular
state.}
\end{center}
\label{f14}
\end{figure*}
In our sample, one of the best examples of semiregular variables
with amplitude modulation is RY~UMa, which is classified as
an SRb star. Its light curve reveals a clear amplitude
modulation which is strongly supported by the results of wavelet
analysis (Fig.\ 13). Although the Fourier spectrum suggests two closely
separated periods (see Table 4), accepting them would be misleading, as
the wavelet map shows slight frequency changes of the principal peak
accompanied with amplitude modulation. Therefore, that close
peak is an artifact caused by the instability of the principal
one. The underlying physical mechanism in unknown: there are several
possibilities such as the rotation, magnetic activity change or
duplicity effects. Unfortunately the presently available observations
do not allow finding a reliable model for this modulation. Nevertheless,
it is very interesting that the characteristic time of the
amplitude modulation is about 4000 days being around a typical
value of theoretically calculated rate of rotation of red giant
stars (as estimated using rotational velocities studied by
Schrijver \& Pols 1993).
\subsection{V~Bootis, RU~Cygni \& Y~Persei}
V~Boo is the prototype of SRa variables suffering from long-term amplitude
decrease (Szatm\'ary et al. 1996). Recently Bedding et al. (1998)
presented a very similar phenomenon for R~Dor which was classified in GCVS
as an SRb star. The proposed explanation for the amplitude decrease in R~Dor
is that the star is evolving from the Mira state to the semiregular state.
We found another two examples of amplitude decrease in
RU~Cyg (SRa) and Y~Per (Mira), which are consequently the third and
fourth candidates for that interesting evolutionary status.
Unfortunately, these substantial changes of the lightcurves are only
hints of probable change of the variability type, further spectroscopic
or near-infrared photometric observations would be desirable.
V~Boo and RU~Cyg are compared in Fig.\ 14, where long-term
light curves are plotted together with
the corresponding power spectra and wavelet maps. The similarity is quite
conspicuous.
Y~Per differs from V~Boo and RU~Cyg in a very important aspect.
While the dominant frequencies of V~Boo and RU~Cyg did not change
significantly, Y~Per seems to be a clear example of a transition
from Mira to SRb. This is presented in Fig.\ 15, where the
compressed light curve is plotted with two power spectra
corresponding to two data subsets (before and after JD 47000).
The earlier monoperiodicity (P=253 day) was replaced by
a biperiodicity (P$_{0}$=245 day, P$_{1}$=127 day). Apparently
the Mira star Y~Per was transformed to a typical doubly-periodic SRb star.
The most surprising result is the abruptness
of the mode switching.
\section{Summary}
\begin{figure}
\begin{center}
\leavevmode
\psfig{figure=8380f15.eps,width=\linewidth}
\caption{Y~Per: compressed light curve and power spectra of two subsets
separated by the dashed line.}
\end{center}
\label{f15}
\end{figure}
Based on the light curve analyses presented in the previous sections
we have obtained the following results:
1. We have analysed long-term visual observations of 93 red semiregular
variables in order to determine their dominant periods. Direct comparison
with photoelectric measurements demonstrated the
usefulness of the low quality visual data. The most important requirement
of precise frequency determination is to have as long a time-series
as possible, as has been illustrated by frequency analysis of
artificial data.
2. We have found 29 monoperiodic and 56 multiperiodic stars
(44 with two and 12 with three significant and essentially stable periods).
8 variables do not show any unambiguous periodicity. The distribution
of periods and period ratios in multiperiodic variables suggests
the existence of up to five different groups among the variables studied,
which is most probably due to different modes of pulsation.
3. We have highlighted a few interesting special cases:
\begin{itemize}
\item{} TX~Dra and V~UMi are very similar triply-periodic variables
with nearly equal periods. While the longer two periods are
stable over decades of time, the high-frequency mode switches on and off
from time to time. We predict another mode change of
TX~Dra soon, most probably in 1999-2000.
\item{} RY~UMa is one of the best documented examples for amplitude
modulation in SRV's. Unfortunately the visual data alone are not enough
to determine the underlying physical process.
\item{} We have discussed three stars (V~Boo, RU~Cyg and Y~Per)
that show a gradual decrease in amplitude. All of them seem to evolve from
mira-like to semiregular type (as does R~Dor, according to
Bedding et al. 1998), which suggests that Miras and
SRV's may be much more closely related than was thought earlier.
Y~Per differs from the other stars, because, in addition to the amplitude
decrease, a new mode has appeared with a quite high amplitude.
The observed period ratio (1.93) is very typical in the majority
of doubly periodic SRV's.
\end{itemize}
\begin{acknowledgements}
We sincerely thank variable star observers of AFOEV, VSOLJ, HAA/VSS and
AAVSO whose dedicated observations over many decades made this
study possible. The referee (Dr. F. Kerschbaum) has greatly
improved the paper with his notes and suggestions.
This research was supported by Hungarian OTKA Grants \#F022249,
\#T022259 and Szeged Observatory Foundation. The NASA ADS Abstract
Service was used to access data and references.
\end{acknowledgements}
|
\section{Introduction}
The art of controlling Josephson current transport through mesoscopic
superconducting junctions poses many challenges for theory and
experiment from both fundamental and applied points of
view. \cite{BagwellSupM} Control of Josephson current requires
multi-terminal devices - superconducting transistors. One example is
the Josephson field effect transistor (JOFET), \cite{Houten,Akazaki} where
control of the Josephson {\em equilibrium} current is imposed via an
electrostatic gate. Another solution is to connect the normal region
to a normal voltage biased reservoir. Recent progress in fabrication
of superconducting junctions has brought forward a number of
interesting multiterminal structures, e.g. 2DEG,
junctions\cite{Akazaki,Takayanagi,Braginski} metallic junctions,
\cite{Morpurgo1,Baselmans} and high-Tc junctions. \cite{Lombardi}
Injection of electrons and holes allows nonequilibrium quasiparticle
distributions to be maintained in the N-region, making it possible to
control the {\em nonequilibrium} Josephson current. The problem of
nonequilibrium current injection in ballistic junctions is of
particular interest: the Josephson current is transported through
bands of Andreev levels.\cite{Andreev} This provides means for
achieving a dramatic variation of the Josephson current.
The purpose of this paper is to provide a broad description of
Josephson current transport through ballistic SNS junctions under
conditions of nonequilibrium in the normal region due to contact with
a voltage biased normal reservoir. Connection of the normal part of an
SNS junction to a normal electron reservoir gives rise to broadening
of the Andreev bound levels. Van Wees et al. \cite{Vanwees} were the
first ones to consider this broadening in perfect SNS junctions and
to describe essential aspects of the variation of the Josephson
current with voltage, in terms of nonequilibrium population of Andreev
levels. Moreover, Wendin and Shumeiko \cite{Wendin,Wendin2} have
predicted that nonequilibrium filling of Andreev levels may reveal
very large Josesphson currents with different directions, and that
pumping between levels could reverse the direction of the Josephson
current. This has been investigated in detail by Bagwell and coworkers
\cite{Bagwell} and by Samuelsson et al.
\cite{Samuelsson}.
Recently, Morpurgo et al. \cite{Morpurgo97b} observed Andreev levels
by using the injection lead as a spectroscopic probe. Suppression of
the Josephson current due to injection has been demonstrated in both
ballistic \cite{Braginski} and diffusive SNS junctions
\cite{Morpurgo1}. The physical mechanism of the effect in diffusive
junctions is essentially the same as in ballistic
junctions.\cite{Volkov95} Very recently, Baselmans
et. al. \cite{Baselmans} were able also to reverse the direction of
the Josephson current.
A decisive step beyond the work of van Wees et al. \cite{Vanwees} was
taken by Samuelsson et al. \cite{Samuelsson}, who showed that an
essential aspect is the ability of the scatterer at the injection
point to shift the phases of the quasiparticles. In such a case, the
connection to the injection lead also affects the form of the wave
function of the Andreev resonances, and therefore affects Josephson
currents flowing through the resonances. As a result, modification of
the Josephson current under injection does not reduce to the effect of
non-equilibrium population. This is particularly dramatic for long
junctions, where the equilibrium Josephson current is exponentially
small at finite temperature. \cite{Kulik}
In contrast, this {\em anomalous} nonequilibrium Josephson current
does not depend on the length of the junction (long-range Josephson
effect). This means that, in principle, a dissipationless current of
the order of the equilibrium Josephson current of a short junction can
be restored under conditions of filling up all the Andreev levels in
the gap. The effect is most pronounced in junctions with a small
number of transport modes. This opens up the possibility for a new
kind of Josephson transistor where the supercurrent is turned on when
the gate voltage is switched from $eV=0$ to $eV=\Delta$.\cite{Patent}
The complete picture of the nonequilibrium current also includes the
current injected into the junction. This {\em injection} current is
dependent on the properties of the Andreev levels in the junction. It
therefore provides information on the nonequilibrium Josephson
current. We found that it is closely related to the anomalous current
and has similar properties. The injection current has also in itself
been at the focus of great interest in recent literature.
\cite{Lambertrew}
The paper is organized as follows. In section $II$ we present a
general discussion of the currents in a 3-terminal SNS device. In
section $III$ we describe our model based on the stationary BdG
equation. We derive all currents in the case of a three terminal
junction without barriers at the NS-interface in section $IV$. In
section $V$ we discuss the equilibrium and nonequilibrium Josephson
currents, both in a short and long junctions. The effect of barriers
at the NS-interfaces is discussed in section $VI$ and the injection
current and the conductance are analyzed in section $VII$. In section
$VIII$ we discuss the four terminal junction and how it differs from
the three terminal one. Finally, in section $IX$ we present our
conclusions.
\section{Nonequilibrium Josephson currents}
We will consider two junction configurations: 3- and 4-terminal (see
Fig. \ref{juncfig}). The normal part of the junction is inserted between
two superconducting electrodes. The superconducting electrodes are
connected with each other to form a loop and the magnetic flux threading the
loop allows us to control the phase difference $\phi=\phi_R-\phi_L$ across
the junction.
\begin{figure}
\centerline{\psfig{figure=fig3.eps,height=3.5cm}}
\caption{A schematic picture of the three terminal SNS junction setup
under consideration, with a normal reservoir attached to the normal
part of the junction. The normal reservoir is connected to the
superconducting loop (grounded) via a voltage source biased at
$V$. The right figure shows a close-up of the junction area with the
arrows showing the direction of the current flow in the junction}
\label{juncfig}
\end{figure}
We consider a junction in the ballistic limit, i.e when the length
$L=L_2+L_3$ of the normal part of the junction is shorter then both
the elastic and inelastic scattering lengths, $L\ll l_e,l_i$. The
3-terminal configuration is an elementary structure which gives all
necessary information for understanding also the properties of the
4-terminal junction, to be discussed below. We use a simplified
description of the connection point, modeling it by a scattering
matrix $S$ that connects ingoing and outgoing wave function
amplitudes \cite{Buttiker}
\begin{equation}
\Psi_{out} = S \Psi_{in},
\end{equation}
with
\begin{equation}
S=\left( \begin{array}{lll}
\sqrt{1-2\epsilon} & \sqrt\epsilon& \sqrt\epsilon\\
\sqrt\epsilon & r & d\\
\sqrt\epsilon & d & r\\
\end{array} \right),
\label{scatmat}
\end{equation}
where $r$ and $d$ are reflection and transmission amplitudes for
scattering between lead 2 to lead 3 and $\sqrt{\epsilon}$ is the
scattering amplitude from the injection lead 1 to lead 2 or 3. In a
multichannel treatment, $r,d$ and $\epsilon$ become matrices
describing the scattering between the channels. In this paper we
however choose to consider a single-mode structure.
In the junction presented in Fig. \ref{juncfig}, the current $I_1$
injected into the junction from the normal reservoir splits at the
connection point. At the NS-interfaces, the normal current is
converted into a supercurrent. The supercurrent flows around the loop
and is drained at a point connected to the normal reservoir via a
voltage source biased at voltage $V$. There are two major questions
about the currents: (i) what is the current $I_1$ in injection
electrode 1 as function of the applied voltage, and (ii) how is the
current split between the arms 2 and 3. The first problem has been
discussed earlier, \cite{Nakano,Kadigrobov} the picture is the
following: due to Andreev quantization the problem is equivalent to a
resonant transmission problem. For weak coupling to the normal
reservoir, $\epsilon \ll 1$, the probability of an incoming electron
to be reflected is large unless its energy coincides with an Andreev
level. In such a case, the electron is back scattered as a hole which
produces a current density peak. The current as a function of applied
voltage between the normal reservoir and the junction (IVC) thus
increases stepwise, typical for resonant transport, with position and
height of the steps depending on the phase difference between the
superconductors.
The current distribution among the left and right arms of the junction
is also phase dependent. However, there is a less trivial aspect of
the problem related to the Josephson current in the loop. There is no
possibility to distinguish the Josephson current which flows along the
loop (as the result of an applied phase difference) from the split
injection current {\em except} in the limit of weak coupling to the
external reservoir. In the limit $\epsilon \ll 1$ the injection
current $(\sim\epsilon)$ vanishes while the Josephson current remains
finite. This allows us to separate the problem of the Josephson
current under injection from the problem of splitting of the injection
current.
The scattering states carrying the current can qualitatively be
described as electrons or holes entering the SNS junction from the
injection lead 1, being split at the connection point, scattered back
and forth in the junction by Andreev reflections at the NS-interfaces
and normal reflections at the connection point, and then finally
leaving the junction, having effectively transported current from one
superconductor to the other. When the lifetime of the Andreev
resonances is smaller than the inelastic scattering time in the
junction, the quasiparticle distribution in the normal region is
determined by the Fermi distribution function of the normal reservoir,
and the current in the leads $j=2$ or $3$ from injected quasiparticles
can be written
\begin{equation}
I_j=\int_{-\infty}^{\infty} dE (i^e_j n^e +i^h_jn^h)
\end{equation}
where $i_j^{e(h)}$ is the current density for injected electrons
(holes) and $n^{e(h)}=n_F(E \pm eV)$ are the Fermi distribution
functions in the normal reservoir, with
$n_F=[1+\mbox{exp}(E/kT)]^{-1}$. This current can conveniently be rewritten
\begin{equation}
I_j=\int_{-\infty}^{\infty}
dE\left[\frac{i^+_j}{2}(n^e+n^h)+\frac{i^-_j}{2}(n^e-n^h)\right]=I^+_j+I^-_j
\label{curreq}
\end{equation}
where $i^+=i^e+i^h$ and $i^-=i^e-i^h$. Quasiparticles are also
injected from the superconductors for energies above the
superconducting gap. Since the superconductors are grounded ($V=0$),
the current from the superconductors is an equilibrium current. This
current plus the current $I^+=I_2^+= I_3^+$ injected from the normal
reservoir in absence of applied voltage, is the total {\em equilibrium}
current. Applying a bias voltage $(V\ne0)$, $I^+$ becomes the {\em
nonequilibrium} current due to population of the empty Andreev levels,
giving rise to current jumps when the injection energy $eV$ equals the
Andreev level energies (see Fig. \ref{spectro}). This makes it
possible to probe the energy of the Andreev levels.
\cite{Vanwees,Samuelsson,Bagwell}
\begin{figure}
\centerline{\psfig{figure=andspektro.eps,height=6.5cm}}
\caption{The current voltage characteristics (IVC) for $I^+$ (upper) and
$I^-$ (lower) for a junction with seven Andreev levels for $0<E<\Delta$.
The currents jump every time the voltage $eV$ is equal to the energy of an
Andreev level, typical for resonant transport}
\label{spectro}
\end{figure}
The $I^-$ part of the current is entirely nonequilibrium current. It partly
consists of the injection current; however, there is also a component which
does not vanish in the limit of weak coupling to the reservoir: we call
this the {\em anomalous Josephson current}. \cite{Samuelsson} This current
results from a different form of the Andreev resonance wave functions in
the {\em open} junction compared with the wave functions of true Andreev
{\em bound} states.
The origin of the anomalous current can qualitatively be described by
considering the lowest order quasiparticle classical paths which contribute
to the resonances in transparent junctions ($R\ll1$) with perfect NS
interfaces.
\begin{figure}
\centerline{\psfig{figure=pccd2.eps,height=6.5cm}}
\caption{The charge current density for two resonant Andreev levels
for injected electrons $i^e$ (dotted) and holes $i^h$ (dashed), their
sum $i^+$ (solid) and difference $i^{-}$ (dash-dotted). Note that the
difference current $i^-$ has the same sign for both resonances. Inset:
Two lowest order paths for an injected electron (solid) or a hole
(dashed) at a resonance. The grey ellipse denotes the effective
scatterer due to the three lead connection. The difference of the
currents due to these processes is proportional to
$\mbox{Im}(rd^*)\sin\phi$, the first order term of the anomalous
current }
\label{ccd2}
\end{figure}
Consider a resonant state where the most of the electrons move to the
left and the holes to the right, only a fraction of them travellling
in the opposite direction due to normal scattering at the connection
point. An injected electron gives rise to a leftgoing electron in lead
2 with the amplitude $1+e^{i\phi_R}d^*e^{-i\phi_L}r$ with
$\phi=\phi_R-\phi_L$ (not taking electron and hole dephasing and the
energy dependent phase picked up when Andreev reflecting into account)
thus giving a contribution to the current of the order
$1+RD+\mbox{Re}(rd^*e^{i\phi})$ (see inset in
Fig. \ref{ccd2}). Correspondingly, an injected hole gives rise to a
rightgoing hole in lead 3 with amplitude $1+e^{-i\phi_L} de^{i\phi_R}
r^*$ and a contribution to the current of order
$1+RD+\mbox{Re}(rd^*e^{-i\phi})$ (see right figure in inset in
Fig. \ref{ccd2}). The difference current $i^-$ thus contains a part
proportional to
$\mbox{Re}[rd^*(e^{i\phi}-e^{-i\phi})]=2\mbox{Im}(rd^*)\sin(\phi)$,
which is the leading term in the anomalous current. At a resonant
state where the particles move in the opposite direction, i.e the
electrons to the right and the holes to the left, we find from the
same arguments that the anomalous current is again proportional to
$2\mbox{Im}(rd^*)\sin(\phi)$, {\em with the same sign}. The anomalous
current thus flows in the same direction for all resonances, in
contrast to the equilibrium Josephsson current which changes sign from
one level to the next. The IVC for $I^-$ is thus a staircase, as shown
in Fig. \ref{spectro}, saturating at $eV>\Delta$ due to the absence of
sharp resonances for energies above the superconducting gap. This has
a dramatic effect on the long range properties of the Josephson
current.
For a long junction $(L\gg \xi_0=\hbar v_F/\Delta)$, the IVC in Fig.
\ref{spectro} becomes dense, since there is a large number $\sim
L/\xi_0$ of Andreev levels in the junction. The spacing between the
Andreev levels is $\sim \hbar v_F/L$, so at temperatures exceeding
the interlevel distance, the current $I^+$ is averaged to zero while
$I^-$ is reduced to a smooth ramp function. We thus get a current
$I^-$ that increases linearly with voltage up to $eV=\Delta$ and
saturates at a level of the order of the equilibrium Josephson current
of a short junction, $I\sim e\Delta/\hbar$. This current is
independent of the length of the junction, since there is a large
number of levels $\sim L$ each carrying a current $\sim 1/L$.
\section{Calculation of the current}
\subsection{General formulation}
We consider a three-terminal junction with asymmetric current
injection ($L_2\neq L_3$) and perfect transmission at the NS
interfaces. The junction can be described by the stationary 1-D
Bogoliubov-de Gennes (BdG) equation\cite{Degennes}
\begin{equation}
\label{bdgeq}
\left[ \begin{array}{cc}
H_{0}&\Delta\\
\Delta^*&-H_{0}
\end{array} \right] \Psi=E \Psi \hspace{1cm}
H_{0}=-\frac{\hbar^{2}}{2m}\frac{d^2}{dx^2}-E_{F}
\end{equation}
which gives $E$ as a departure from $E_{F}$. We apply the approximation
\cite{Likharev} with $\Delta (x)$ constant in the
superconductors and zero in the normal region.
\begin{equation}
\Delta(x)= \left\{ \begin{array}{ll}
\Delta e^{i\phi_L} & x<-L_2 \\
0 & -L_2 <x<L_3 \\
\Delta e^{i\phi_R} & x>L_3
\end{array}
\right. ,
\end{equation}
where the phase difference between the superconductors is
$\phi=\phi_R-\phi_L$. We can then make an ansatz with plane waves in
the different regions of the junction. For positive energies $E>0$ we
put in the normal regions $j=1,2,3$,
\begin{eqnarray}
\Psi_{j}&=&c^{+,e}_{j}\left[\begin{array}{c} 1\\ 0 \end{array}\right] e^{i
k^ex}+c^{h,-}_{j}\left[\begin{array}{c} 0 \\ 1 \end{array}\right]
e^{-ik^hx} \nonumber \\
&& +c^{e,-}_{j}\left[\begin{array}{c} 1\\ 0 \end{array}\right]
e^{-ik^ex}+c^{h,+}_{j}\left[\begin{array}{c} 0\\ 1
\end{array}\right] e^{ik^hx}
\end{eqnarray}
and in the superconductors $j=L,R$
\begin{eqnarray}
\Psi_{j}&=&d^{e,+}_{j}\left[\begin{array}{c} ue^{i\phi_j}\\ v
\end{array}\right] e^{i q^ex}+d^{h,-}_{j}\left[\begin{array}{c}
ve^{i\phi_j}\\ u \end{array}\right] e^{-iq^hx} \nonumber \\
&& +d^{e,-}_{j}\left[ \begin{array}{c} ue^{i\phi_j}\\ v
\end{array}\right] e^{-iq^ex}+d^{h,+}_{j}\left[\begin{array}{c}
ve^{i\phi_j}\\ u \end{array}\right] e^{iq^ex}.
\end{eqnarray}
The coherence factors $u$ and $v$ are defined as
\begin{equation}
u(+),v(-)= \left\{ \begin{array}{ll}
\sqrt{\frac{1}{2}(1\pm\xi/E)} & E>\Delta \\
\sqrt{\frac{1}{2}(E\pm\xi)/\Delta} & E<\Delta
\end{array}
\right.
\end{equation}
where $\xi=\sqrt{E^2-\Delta^2}$ for $E>\Delta$ and
$\xi=i\sqrt{\Delta^2-E^2}$ for $E<\Delta$. The wavevectors are
$q^{e,h}=\sqrt{2m/\hbar^2}\sqrt{E_{F}\pm \xi}$ in the superconductors
and and $k^{e,h}=\sqrt{2m/\hbar^2}\sqrt{E_{F}\pm E}$ in the normal
regions. The wavefunctions are matched at the NS-interfaces and at the
injection point. The three-terminal injection point is modeled by the
scattering matrix\cite{Buttiker,Nakano} given by
Eq. (\ref{scatmat}). The scattering amplitudes $\epsilon$ ($0\leq
\epsilon\leq 0.5$), $d$ and $r$ obey the relations
$Re(rd^\ast)=-\epsilon/2$ and $D+R=1-\epsilon$ ($D=|d|^2, R=|r|^2$)
due to the unitarity of the scattering matrix. Moreover,
$\mbox{Im}(rd^*)=\sigma\sqrt{RD-\epsilon^2/4}$, with $\sigma=\pm 1$
dependent on the phase of the scatterer. For simplicity the coupling
parameter $\epsilon$ is chosen real and positive. The scattering
amplitudes are assumed to be energy independent, which gives the
scattering matrix for hole wavefunction amplitudes $S_h=S_e^*$.
Assuming $\Delta \ll E_F$ we make the approximation $q^{e}=q^{h}=k_{F}$
in the superconductors and $k^{e}=k^{h}=k_{F}$ in the normal region
except in exponentials where we put $k^{e,h}=k_{F}\pm E/(\hbar
v_F)$. At energies $E<\Delta$, only electrons and holes from the normal
reservoirs are injected in the junction. For $E>\Delta$ quasiparticles
from the superconductors are also injected. The current density in the
three normal regions, which is what is needed to calculate all
currents in the junction, are calculated using the quantum mechanical
formula \cite{BTK}
\begin{equation}
i_j(E)=\frac{e}{h}(|c_j^{+,e}|^2-|c_j^{-,e}|^2-|c_j^{+,h}|^2+|c_j^{-,h}|^2).
\label{btkcurr}
\end{equation}
We now define energy dependent phases
$\theta_{2,3}=\gamma-\beta_{2,3}$ in each of the leads $2$ and $3$,
consisting of the phase $\gamma=\arccos(E/\Delta)$ picked up by the
electrons and holes when Andreev reflecting, and the dephasing
$\beta_{2,3}=(k^e-k^h)L_{2,3}=2EL_{2,3}/(\hbar v_F)$ of the electrons
and holes while propagating ballistically through the normal
region. Furthermore, it is convenient to separate out the specific
features of asymmetry by introducing sum phases
2$\theta=\theta_2+\theta_3$, $\beta = \beta_2+\beta_3$, and the
difference phases $\chi=\theta_2-\theta_3$, defining essential phase
parameters characterizing the junction,
\begin{eqnarray}
\theta = \gamma-\beta/2 = \arccos(E/\Delta) - EL/(\hbar v_F) \\
\label{thetadef}
\chi = \beta_3-\beta_2 = 2El/(\hbar v_F)
\label{chidef}
\end{eqnarray}
where $L=L_2+L_3$ and $l= L_3-L_2$
The current densities of the scattering states in leads $2$ and $3$ from
electrons $i_{2,3}^e$ and holes $i_{2,3}^h$ are then given by
\begin{eqnarray}
i_2^{e,h}&=&-{e\over h}{\epsilon\over Z} \left\{2D\sin\phi\sin2\theta
\right. \nonumber \\ && \left. \pm \left[\sigma
2\sqrt{RD-\epsilon^2/4}\sin\phi(\cos\chi-\cos2\theta)
\right. \right. \nonumber \\ && \left. \left. +\epsilon
\left[1-\cos(2\gamma-\beta_2) + \cos\phi (\cos\chi-\cos2\theta)
\right]\right]\right\}
\label{ieh2}
\end{eqnarray}
\begin{eqnarray}
i_3^{e,h}=&&-{e\over h}{\epsilon\over Z}\left\{2D\sin\phi\sin2\theta
\right. \nonumber \\ && \left. \pm
\left[2\sigma\sqrt{RD-\epsilon^2/4}\sin\phi(\cos\chi-\cos2\theta)
\right. \right. \nonumber \\ && \left. \left. -\epsilon
\left[1-\cos(2\gamma-\beta_3) + \cos\phi (\cos\chi-\cos2\theta)
\right]\right]\right\}
\label{ieh3}
\end{eqnarray}
where
\begin{equation}
Z=[(1-\epsilon)\cos2\theta -R\cos\chi -D\cos\phi]^2+\epsilon^2\sin^22\theta
\label{Z}
\end{equation}
From Eqs. (\ref{ieh2}) and (\ref{ieh3}) it follows that the sum of the
electron and hole current densities, $i^+=i^e+i^h$, are equal in leads 2 and 3,
giving the sum current density
\begin{equation}
i^+=i_3^+=i_2^+=-{4e\over h}{\epsilon\over
Z}\left\{D\sin\phi\sin2\theta\right\}.
\label{sumcurr}
\end{equation}
The difference current densities $i^-=i^e-i^h$ in leads 2 and 3 are
not equal, however. We therefore define the anomalous current density
$i_a$ as that part of the difference current density which survives in
the limit $\epsilon\rightarrow 0$,
\begin{equation}
i_a=-\sigma \frac{4e}{h}\frac{\epsilon}{Z}\left\{\sqrt{RD-\epsilon^2/4}
\sin\phi(\cos\chi-\cos2\theta) \right\}
\label{avdiffcurr}
\end{equation}
The injection current density $i_{inj}=i_3^- -i_2^-$ is given by,
\begin{equation}
i_{inj}=\frac{4e}{h}\frac{\epsilon^2}{Z}\left\{\sin^2\chi + (\cos\chi
+\cos\phi) (\cos\chi-\cos2\theta) \right\}
\label{injdiffcurr}
\end{equation}
and splits asymmetrically between the two horizontal arms $2$ and $3$,
\begin{equation}
i_{inj2,3}=\pm \frac{2e}{h}\frac{\epsilon^2}{Z}\left\{1-\cos(2\theta-\beta_{2,3}).
+ \cos\phi(\cos\chi-\cos2\theta) \right\}
\end{equation}
From the relations $i^{+}(E)=-i^{+}(-E)$ and $i^{-}(E)=i^{-}(-E)$ one
can calculate the current densities for all energies inside the gap
$|E|<\Delta$. The continuum current density, for energies outside the
gap $|E|>\Delta$, is calculated in the same way. However, since the
Andreev reflection probability decays very rapidly outside the gap,
the Andreev resonances become very broad and contribute much less to
the current. Only the quasiparticles injected from the superconductors
contribute significantly to the current, as will be discussed
below. The full formulas for the continuum current density for a
symmetric junction $l=0$ is presented in Appendix A.
\subsection{Weak coupling limit}
Throughout the paper we will mainly discuss the situation when the normal reservoir is weakly coupled to the normal part of the junction, $\epsilon \ll 1$. In this limit the Andreev resonances are very sharp and the current densities are calculated by evaluating the expression $\epsilon/Z$ appearing in the Eqs. (\ref{sumcurr})-(\ref{injdiffcurr}), in the limit $\epsilon \rightarrow 0$. This is done in detail in Appendix B, and gives [see Eq. (\ref{epszeta})]
\begin{equation}
\lim_{\epsilon\rightarrow 0} \frac{\epsilon}{Z}=\sum_{n,\pm} \frac{\pi}{D|\sin \phi \sin 2\theta|}\left|\frac{dE}{d\phi}\right| \delta(E-E_n^{\pm}).
\label{epszero}
\end{equation}
where $E_n^{\pm}$ are the energies of the bound Andreev states. To
calculate the current density, information about the bound state
energies as well as the derivative of the energy with respect to phase
difference is thus needed. The bound state energies are given by the
zeros of the denominator $Z$ [Eq. (\ref{Z})] at $\epsilon=0$, namely
\cite{Bagwell4}
\begin{equation}
\cos2\theta = R\cos\chi + D\cos\phi.
\label{bsteq}
\end{equation}
The energy of the Andreev levels as a function of phase difference
$\phi$ is plotted in Fig. \ref{asandr}. In the figure it is shown that
the Andreev levels appear in pairs, labeled by $n$, with an upper
($+$) and a lower ($-$) level (referring to $E>0$). The index $n$ is
zero for the pair of levels with positive energy closest to $E_F$. In
the case of one single bound state, the level is labeled by $E_0^-$.
\begin{figure}
\centerline{\psfig{figure=andrspek.eps,height=6.5cm}}
\caption{Andreev bound state energies as a function of phase difference
$\phi$ for different lengths $L=0$ (left), $L\sim \xi_0$ (middle) and
$L\gg\xi_0$ (right) of the junction with $D=0.7$. Solid lines are for a symmetric
junction $l=0$, dashed for an asymmetric one. A gap opens up in the
spectrum at $\phi=0$ due to the asymmetry.}
\label{asandr}
\end{figure}
The derivative of the bound state energy with respect to phase is
obtained by differentiating Eq. (\ref{bsteq}), giving
\begin{eqnarray}
&&\frac{dE_n^{\pm}}{d\phi}=
\frac{D\sin\phi}{2\sin2\theta} \nonumber \\
&& \times \left( \frac{1}{\sqrt{\Delta^2-(E_n^{\pm})^2}}
+\frac{L}{\hbar v_F } + \frac{l}{\hbar v_F} R
\frac{\sin\chi}{\sin2\theta}\right)^{-1}.
\label{asymdedfi}
\end{eqnarray}
The expression for the sum current density is given by inserting Eqs. (\ref{epszero})-(\ref{asymdedfi}) into Eq. (\ref{sumcurr}), giving
\begin{equation}
i^+=\frac{2e}{\hbar}\sum_{n,\pm}\frac{dE}{d\phi}\delta(E-E_n^{\pm}),
\label{sumepsgotozero}
\end{equation}
where the relation $\mbox{sgn}[(dE/d\phi) \sin\phi \sin 2\theta]=-1$
[see Eq. (\ref{signrel})] has been taken into account. The expression
(\ref{sumepsgotozero}) coincides with the equation for the Andreev
bound state current \cite{Vitaly1} derived directly from the BdG
equation. From the alternating slopes of the energy-phase relation
$E(\phi)$, plotted in Fig. \ref{asandr}, it is clear that the sum
current density ($\sim dE/d\phi$) changes sign between two subsequent
Andreev resonances (see Fig. \ref{ccd2}).
The anomalous current density $i_a$ is given directly by inserting Eq. (\ref{epszero}) into (\ref{avdiffcurr}), namely
\begin{eqnarray}
i_a&=&-\sigma {2e\over\hbar} \mbox{sgn}(\sin\phi)\sqrt{RD} \nonumber \\
&& \times \sum_{n,\pm} \frac{\cos\chi -\cos\phi}{|\sin2\theta|} \left|{dE_n^{\pm} \over d\phi}\right| \delta(E-E_n^{\pm}).
\label{ianzero}
\end{eqnarray}
For a symmetric junction $l=0, \cos\chi=1$, the anomalous current
density does not change sign as a function of energy, opposite to the
sum current density (see Fig. \ref{ccd2}). For finite asymmetry, the
anomalous current might change sign. However, this does not lead to
strong suppression of the total anomalous current, as will be shown
below in section VB.
The injection current $i_{inj}=i_3^- -i_2^-$ is proportional to
$\epsilon ^2$ and thus goes to zero for $\epsilon \ll 1$. We
approximate the injection current in the weak coupling limit by the
first order term in $\epsilon$, given by inserting the expression for
$\epsilon/Z$ in the zero coupling limit into Eq. (\ref{epszero})
\begin{eqnarray}
i_{inj}&&= \epsilon{8e\over \hbar} \sum_{n,\pm}
\frac{\sin^2\chi + D(\cos\chi -\cos\phi)^2}{|\sin2\theta|}
\nonumber \\
&& \times \left |\frac{dE_n^{\pm}}{d\phi}\right| \delta(E-E_n^{\pm}).
\label{injcurrdens}
\end{eqnarray}
The injection current density is closely related to the anomalous
current density $i_a$, in the sense that the injection current density
is positive for all energies and values of the phase difference
$\phi$.
\subsection{Structure of the nonequilibrium current}
Including the continuum contribution from the superconductors
(Appendix A) in Eq. (\ref{curreq}), we can finally write down the
structure of the total current in each lead:
\begin{eqnarray}
I_j = \int_{-\infty}^{\infty}dE
\left[\frac{i_j^+}{2}(n^e+n^h)+\frac{i_j^-}{2}(n^e-n^h)+i^s n_F \right],
\label{totcurr}
\end{eqnarray}
where $i^s$ is the current density from the quasiparticles injected
from the superconductors. The equilibrium current $(V=0)$ flowing in
leads $2$ and $3$ is given by
\begin{equation}
I_{eq}=\int dE \left[ i^+ + i^s \right] n_F
\label{ieqdef}
\end{equation}
while in lead $1$ it is zero. Subtracting the equilibrium current from
the total current we get the {\em nonequilibrium} current in the horizontal
leads $2$ and $3$. We divide the nonequilibrium current into the
the {\em regular current} $I_r$ associated with the nonequilibrium
population of the existing resonant states,
\begin{equation}
I_r=\int dE \left[\frac{i^+}{2}(n^e+n^h-2n_F) \right],
\label{irdef}
\end{equation}
the {\em anomalous current} $I_a$ associated with the essential modification
of the Andreev states due to the open normal lead,
\begin{equation}
I_a=\int dE \left[\frac{i_a}{2}(n^e-n^h) \right],
\label{iadef}
\end{equation}
and the injected current $I_1$
\begin{equation}
I_1= I_{inj}=\int dE \left[\frac{i_{inj}}{2}(n^e-n^h) \right].
\label{i1def}
\end{equation}
With these definitions, the total currents in leads 2 and 3 may be
written as
\begin{eqnarray}
I_2=I_{eq}+I_r+I_a-I_{inj,2}, \\
I_3=I_{eq}+I_r+I_a+I_{inj,3}, \nonumber
\end{eqnarray}
where $I_{inj}= I_{inj,2}+I_{inj,3}$. As discussed in Section II, the
separation of the anomalous current is arbitrary, and has physical
meaning only in the weak coupling limit when $I_{inj} \rightarrow 0$.
In the weak coupling limit, the integrals in Eqs.
(\ref{irdef})-(\ref{iadef}) become sums over resonant states
\begin{equation}
I_r=\frac{e}{\hbar}\sum_{n,\pm}\frac{dE_n^{\pm}}{d\phi}\left[n^e(E_n^{\pm})+
n^h(E_n^{\pm})-2n_F(E_n^{\pm})\right],
\label{irbound}
\end{equation}
\begin{eqnarray}
I_a&=&-\sigma {e\over\hbar} \mbox{sgn}(\sin\phi)\sqrt{RD} \nonumber \\
&& \times \sum_{n,\pm} \frac{\cos\chi -\cos\phi}{|\sin2\theta|} \left|\frac{dE_n^{\pm}}{d\phi}\right|\left[n^e(E_n^{\pm})-n^h(E_n^{\pm})\right].
\label{iabound}
\end{eqnarray}
The equilibrium current for energies $|E|<\Delta$ is given by inserting Eq. (\ref{sumepsgotozero}) into (\ref{ieqdef}),
\begin{equation}
I_{eq}^{b}=\frac{2e}{\hbar}\sum_{n,\pm}\frac{dE_n^{\pm}}{d\phi}n_F(E_n^{\pm}),
\label{sumboundstate}
\end{equation}
For energies above the gap, the equilibrium current results from quasiparticles injected from the superconductors only, since this current is the only continuum current being finite in the weak coupling limit (see Appendix A).
\section{Josephson current of a short junction}
For a short junction $L=l=0$, there is exactly one resonance for
positive energies $0<E<\Delta$. For no coupling to the normal
reservoir $\epsilon=0$, this resonant Andreev state is converted into
a bound Andreev state, with the dispersion relation $E_0^-=\Delta
\sqrt{1-D\sin^2(\phi/2)}$. The equilibrium current of a short junction
is thus given by the well known
\cite{Furusaki} relation
\begin{equation}
I_{eq}=\frac{e\Delta}{\hbar}\frac{D \sin \phi}{2\sqrt{1-D\sin^2(\phi/2)}}
\tanh(E_0^-/2kT).
\label{shortieq}
\end{equation}
The continuum current is zero, which can be seen by putting $L=0$
$(\beta=0)$ in the equations for the continuum current in Appendix
A. At zero temperature and zero applied bias, only the level with
negative energy $-E_0^-$ is populated. For an applied a voltage bias
$V>0$, the electron (hole) population is shifted upwards (downwards)
in energy. When the voltage $eV=E_0^-$, the energy of the resonant
level, the level becomes populated and there is an abrupt jump of the
current. The regular part of the current, $I_r$, jumps an amount
$\delta I_r=-I_{eq}$, thus cancelling the equilibrium Josephson
current. This has recently been observed in experiments.
\cite{Braginski,Morpurgo1}
The anomalous current jumps by the amount
\begin{figure}
\centerline{\psfig{figure=pshortivc.eps,height=6.5cm}}
\caption{The currents $I_{eq}+I_r$ (dash-dotted), $I_a$ (dashed) and the
total Josephson current $I_{eq}+I_r+I_a$ (solid) in the horizontal leads
$2$ and $3$ as a function of voltage $V$ at $T=0$ for a short
junction $L=0$ with $D=0.8, \phi=3\pi/4, \epsilon=0.01$ and $\sigma=-1$.
The total current is $I_{eq}$ for $eV<E_0^-$ and $\delta I_a$ for
$eV>E_0^-$.}
\label{shortIVC}
\end{figure}
The effect of finite temperature in a zero length junction is merely to
smear the steps in the IVC.
In the symmetric case ($l=0)$ it is interesting to extend the
discussion to a longer junction with two resonant levels (see
Fig. \ref{asandr}), since the current distribution between the levels
becomes nontrivial. \cite{Wendin,Bagwell} In the limit $D\ll1$, both
resonances have energies close to the gap edge, $E_0^{\pm} \approx
\Delta$, and with the additional approximation $\beta/2>\sqrt{D}$ we
obtain the expression for the derivative of energy with respect to
phase [see Eq. (\ref{asymdedfi})]
\begin{equation}
\frac{dE_0^{\pm}}{d\phi}=\pm\frac{\Delta\sqrt{D}}{4}\frac{L}{\xi_0}\frac{\sin(\phi)}{|\sin(\phi/2)|\sqrt{1-D\sin^2(\phi/2)}}.
\end{equation}
The equilibrium bound state current becomes proportional to $I_{eq}^b \sim
dE_0^+/d\phi+dE_0^-/d\phi\sim D$ (taking terms of order $D$ into account),
but for the currents of the individual levels $\sim \sqrt{D}$. The resonant
levels thus carry opposite ``giant'' currents which almost cancel in
equilibrium. For $L>0$, we also have to take the continuum contribution
into account. In has been shown \cite{Wendin} that the
continuum contribution to the equilibrium current is
$I_{eq}^c=-1/2I_{eq}^b$, thus giving the total equilibrium current
$I_{eq}=1/2I_{eq}^b$.
At zero temperature, when a voltage equal to the lowest lying level
$eV=E_0^-$ is applied, the regular and anomalous current jumps
\begin{equation}
\delta
I_r=\frac{e\Delta}{\hbar}\frac{L}{\xi_0}\frac{\sqrt{D}\sin(\phi)}{2|\sin(\phi/2)
|\sqrt{1-D\sin^2(\phi/2)}},
\end{equation}
\begin{equation}
\delta
I_a=\sigma \frac{e \Delta}{\hbar}\frac{L}{\xi_0}
\frac{\sqrt{RD}\sin\phi}{\sqrt{1-D\sin^2(\phi/2)}}.
\end{equation}
Both jumps are proportional to $\sqrt{D}$, and the magnitude of the total
current at $E_0^-<eV<E_0^+$ is then much larger than the equilibrium
current.
When the bias voltage is further increased to $eV=E_0^+$ there is a second
current jump: the regular current jumps in the {\em opposite} direction and
becomes equal to the small negative bound state equilibrium current
$I_r=-I_{eq}^b$. The anomalous current, however, again jumps $\delta I_a$
in the {\em same} direction. For voltages $eV>E_0^+$ the total current in
the junction is thus $I_{eq}^c+2\delta I_a$. The full formulas for all the
individual currents including temperature dependence is given in Appendix C.
\section{Josephson current of a long junction}
We now discuss the Josephson current in a long ($L\gg\xi_0$) symmetric
($l=0$) junction, and we treat the effects of asymmetry below. In a
long junction there are many $(N=[L/(\xi_0 \pi)])$ pairs of
resonances, as seen in Fig. \ref{asandr}. The width of each resonance
is $\Gamma=\epsilon \hbar v_F/(2L)$. This width must not be too small
if the quasiparticles are to be able to enter and leave the junction
without being scattered inelastically ($\Gamma > \hbar
v_F/l_i$). This gives an upper limit for the length $L < l_i
\epsilon$.
The derivative of energy with respect to phase $dE/d\phi$ in Eq.
(\ref{asymdedfi}), which determines the current in Eqs. (\ref{irbound})-
(\ref{sumboundstate}), can be simplified in a long junction
$L\gg\xi_0$,
\begin{equation}
\frac{dE^{\pm}}{d\phi}=\pm\frac{\hbar v_F}{L}
\frac{\sqrt{D}\sin(\phi)}{4|\sin(\phi/2)|\sqrt{1-D \sin^2(\phi/2)}}.
\label{longdedfi}
\end{equation}
This expression holds everywhere except close to the gap edge,
$\Delta-E_n \sim (\hbar v_F/L)(\xi_0/L)$, a distance much smaller than
the energy distance $\pi\hbar v_F/L$ between the pairs of
levels. Therefore, equation (\ref{longdedfi}) can be used for
calculation of the currents of all levels except the last pair of
levels closest to the energy gap. The current from this last pair of
levels must always be treated on a separate footing.
According to Eq. (\ref{longdedfi}), each of the Andreev levels carry
a current of the order of $1/L$. Furthermore, as follows from the
exact Eq. (\ref{asymdedfi}), each pair of levels carries a small
net current, $dE_{n}^+/d\phi-dE_{n}^-/d\phi$, of the order of
$(1/L)^3$. The sum of the currents from all bound states is thus
determined by the current ($\sim 1/L$) from the last pair of levels.
\subsection{Equilibrium current}
For the equilibrium current of a long junction, the contribution from
the Andreev bound states at $|E|<\Delta$ and from the continuum at
$|E|>\Delta$ are of the same magnitude (see
Fig. \ref{boundcontcurr}). The continuum current (see Appendix A) is
given by
\begin{eqnarray}
&&I_{eq}^{c}=\frac{e}{h}\left(\int_{-\infty}^{-\Delta}+\int^{\infty}_{\Delta}\right)
dE n_F\nonumber \\
&&\times \frac{4D\sin\phi\sin\beta \sinh 2\gamma_c
}{(\cos\beta\cosh2\gamma_c-R-D\cos\phi)^2+(\sin\beta \sinh 2
\gamma_c)^2},
\label{contcurr2}
\end{eqnarray}
with $\gamma_c=\mbox{arccosh}(E/\Delta)$. Following the method by
Ishii \cite{Ishii} and Svidzinsky et al. \cite{Kulik}, one can
rewrite this integral as a sum over the residues,
\begin{equation}
I_{eq}^{c}=-I_{eq}^{b}+4kT\pi i\sum_{p=0}^{\infty}i(E_p),
\label{contcurr3}
\end{equation}
where the first term results from the poles of the current density
$i(E)$ in Eq. (\ref{contcurr2}) and the second term from the poles of
the distribution function $n_F(E)$, given by $E_p=i2kT\pi(1/2+p)$. The
first term in (\ref{contcurr3}) is the current carried by the bound
states with negative sign. The total equilibrium current $I_{eq}$ is
then just given by the second term. For zero temperature this sum
over poles turns into an integral and the total equilibrium current is
then given by
\begin{equation}
I_{eq}(T=0)=\frac{e}{\hbar}\frac{\hbar
v_F}{L}\frac{\sqrt{D}\sin(\phi)\arccos(R+D\cos\phi)}{2\pi
|\sin(\phi/2)|\sqrt{1-D \sin^2(\phi/2)}}.
\label{longzerotemp}
\end{equation}
At high temperatures $\hbar v_F/L\ll kT \ll \Delta$, only the the first term
$(p=0)$ in the sum $(\ref{contcurr3})$ needs to be included, and the
equilibrium current becomes
\begin{equation}
I_{eq}(kT \gg \hbar v_F/L)=\frac{e \Delta}{\hbar}\pi
\left(\frac{4kT}{\Delta}\right)^2 D \sin \phi e^{-2\pi L/\xi_T},
\label{longfinitetemp}
\end{equation}
where $\xi_T=\hbar v_F/kT$. The equilibrium current for a
long junction at finite temperature is thus exponentially small.
\cite{Kulik}. Expressions (\ref{longzerotemp}) and
(\ref{longfinitetemp}) extend earlier results \cite{Kulik,Ishii}
to the case of arbitrary transparency $D$ of the junction.
\begin{figure}
\centerline{\psfig{figure=bccurr.eps,height=6.5cm}}
\caption{The equilibrium bound state (a) and continuum (b) currents
and their sum (dashed) as a function of length $L$ for finite
$kT=0.2\Delta$, $D=0.8$, $\phi=3\pi/4$ and $\epsilon=0.01$. There is a
cusp in both the bound state and continuum currents when a new bound
state forms out of the continuum. The total equilibrium current,
however, dies monotonically with increased length. Inset: The
equilibrium bound state and continuum currents as a function of
temperature for a long junction $L=15\xi_0$ with $D=0.8$, $\phi=3
\pi/4$ and $\epsilon=0.01$. The bound state current (a) decreases from
$I_{eq}^b(T=0)=i^*$ to $-I_{eq}^c$, when the temperature is increased
from zero to $kT \ll \hbar v_F/L$. The continuum current (b) is
unaffected in this temperature regime.}
\label{boundcontcurr}
\end{figure}
We are also interested in analyzing the separate behavior of the
bound state current, because this current is revealed in
nonequilibrium, as will be discussed in detail below. Therefore, using
relation (\ref{longdedfi}) we can write Eq. (\ref{sumboundstate}) on
the form
\begin{eqnarray}
I_{eq}^{b}&=&\frac{e}{\hbar}\frac{\hbar
v_F}{L}\frac{\sqrt{D}\sin(\phi)}{2|\sin(\phi/2)|\sqrt{1-D\sin^2(\phi/2)}}
\nonumber \\
&&\times \sum_{n=0}^{N-1} \left[\tanh(E_n^-/2kT)-\tanh(E_n^+/2kT)\right]
\nonumber \\
&& +i^{*}\tanh(\Delta/2kT).
\label{longeq}
\end{eqnarray}
The term $i^*$ results from the last pair of levels at
$E\approx\Delta$, and is of the order $1/L$, as discussed above. At
$T=0$, the sum in Eq. (\ref{longeq}) is zero, and we thus find that
$i^{*}=I_{eq}^b(T=0)$. When the temperature is increased, the sum in
Eq. (\ref{longeq}) starts to contribute with negative sign and the
bound state current is decreased. The continuum current (and also
$i^*$), however, is independent of temperature for $kT\ll\Delta$,
since it is an integral over states with $|E|>\Delta$ [see
Eq. (\ref{contcurr2})]. At $kT \gg \hbar v_F/L$ the {\em total}
equilibrium current is exponentially small (see
Eq. (\ref{longfinitetemp}) and has thus decreased an amount
$I_{eq}(T=0)-I_{eq}(kT\gg\hbar v_F/L) \approx I_{eq}(T=0)$. This is
thus solely due to decrease of the bound state current, as shown in
the inset in Fig. \ref{boundcontcurr}.
\subsection{Regular current}
The regular current can be written, inserting relation
(\ref{longdedfi}) into Eq. (\ref{irbound}), on the form
\begin{eqnarray}
I_r&=&\frac{e}{\hbar}\frac{\hbar
v_F}{2L}\frac{\sqrt{D}\sin(\phi)}{2|\sin(\phi/2)|\sqrt{1-D\sin^2(\phi/2)}}
\nonumber \\
&& \times \sum_{n=0}^{N-1}[g(E_n^-)-g(E_n^+)]+\frac{i^*}{2}g(\Delta)
\label{ireglong}
\end{eqnarray}
where $g(E)=\tanh[(E+eV)/2kT]+\tanh[(E-eV)/2kT]-2\tanh(E/2kT)$. The regular
current $I_r$ jumps up or down every time $eV=E_n^{\pm}$ [see Fig.
\ref{ireg}]. Each current jump has the magnitude
\begin{equation}
\delta I_r=\frac{e}{\hbar}\frac{\hbar
v_F}{L}\frac{\sqrt{D}\sin(\phi)}{2|\sin(\phi/2)|\sqrt{1-D\sin^2(\phi/2)}}
\end{equation}
at zero temperature. At voltages $eV>\Delta$, the regular current is
the sum of all states in the range $0<E<\Delta$, and is equal to
the negative bound state current $-i^*$.
\begin{figure}
\centerline{\psfig{figure=ireg.eps,height=6.5cm}}
\caption{The equilibrium current $I_{eq}$ plus the regular current $I_r$ vs
voltage. $L=10 \xi_o$, $\phi=\pi/2$, $D=0.8$, $\epsilon=0.05$. Solid
line - $T=0$, dashed-dotted - $kT=0.04\Delta$, dashed -
$kT=0.07\Delta$. The regular current jumps alternating by $\pm \delta I_r$
every time the voltage is equal to the energy of an Andreev
resonance. For $kT \gg \hbar v_F/L$ and $eV>\Delta$ the current
$I_{eq}+I_r=I_{eq}^c$}
\label{ireg}
\end{figure}
It is interesting to study the sum $I_{eq}+I_r$, plotted in
Fig. \ref{ireg}, at temperatures $kT\gg \hbar v_F/L$. In this
temperature regime the equilibrium current is exponentially small and
also the regular current steps in the IVC in Fig. \ref{ireg} are
suppressed. For a voltage $eV\sim \Delta$, the last level, carrying the
major part ($i^*$) of the bound state current, is populated and the
current $I_{eq}+I_r$ jumps to $I_{eq}^c$, the value of the continuum
current, since all bound states are populated. This current $I^c_{eq}$
is of the order of $1/L$ and the current $I_r+I_{eq}$ is {\it
increased } from zero to be $\sim 1/L$ when increasing
the voltage from $V=0$ to $eV \sim \Delta$.
\subsection{Anomalous current}
The anomalous current is given by inserting Eq.
(\ref{longdedfi}) into Eq. (\ref{iabound}),
\begin{equation}
I_a=-\sigma\frac{e}{\hbar}\frac{\hbar
v_F}{4L}\frac{\sqrt{RD}\sin\phi}{1-D\sin^2(\phi)}\sum_{n=0}^{N-1}[h(E_n^+)+h(E
_n^-)],
\label{longian}
\end{equation}
where $h(E)=\tanh[(E-eV)/2kT]-\tanh[(E+eV)/2kT]$. We have neglected
the current from the last level close to $E=\Delta$, because the
currents of all levels add up and the current from the last level is
negligible. The IVC at zero temperature looks like a staircase, as
shown in Fig. \ref{panvolt}.
\begin{figure}
\centerline{\psfig{figure=anvolt.eps,height=6.5cm}}
\caption{The anomalous current as a function of voltage $V$ for (a)
$\phi=\pi/4$ and (b) $\phi=3\pi/4$ for $L=10 \xi_0$, $D=0.8$,
$\epsilon=0.05$ and $\sigma=-1$. Temperature $T=0$ (solid) and
$T=0.1\Delta$ (dashed line). The current steps with magnitude $\delta
I_a$ are smeared to a straight line for $kT \gg \hbar v_F/L$. Upper inset: The
critical anomalous current at $eV=\Delta$ as a function of
transparency $D$ for coupling constant $\epsilon=0.1$. Due to finite
coupling $\epsilon$, the critical current always goes to zero for
$R=0$. Lower inset: The anomalous current $I_a(eV=\Delta,kT \gg \hbar v_F/L)$
as a function of phase difference $\phi$ for different transparencies
$D=0.1, 0.5$ and $0.9$. The highest amplitude corresponds to the
highest transparency and vice versa.}
\label{panvolt}
\end{figure}
The magnitude of the current step at zero temperature is given by
\begin{equation}
\delta I_a=\frac{e}{\hbar}\frac{\hbar
v_F}{2L}\frac{\sqrt{RD}\sin\phi}{1-D\sin^2(\phi/2)}.
\end{equation}
At temperatures larger then the interlevel distance, $kT \gg \hbar
v_F/L$, the staircase IVC is smeared out to a straight slope, as shown
in Fig. \ref{panvolt}. The exact position of each level becomes
irrelevant and we can write the sum over bound states in
(\ref{longian}) as an integral, noting that the expression
$dE/dn=\pi\hbar v_F/L$ holds for all levels in the sum
(\ref{longian}),
\begin{eqnarray}
&&\sum_{n=0}^{N}[h(E_n^+)+h(E_n^-)] \nonumber \\
&&\approx \frac{2L}{\pi \hbar
v_F}\int_{0}^{\Delta}dE[\tanh(E+eV)-\tanh(E-eV)] \nonumber \\
&&=\frac{4L}{\hbar v_F \pi}f(V,T)
\label{sumint}
\end{eqnarray}
where
\begin{equation}
f(V,T)=kT\ln\left(\frac{\cosh(\Delta+eV)/kT}{\cosh(\Delta-eV)/kT}\right),
\label{fdef}
\end{equation}
and the anomalous current takes the simple form
\begin{equation}
I_a=-\sigma\frac{e}{\hbar}\frac{\sqrt{RD}\sin\phi}{\pi[1-D\sin^2(\phi/2)]}f(V,T)
.
\label{iasimple}
\end{equation}
In the limit $\hbar v_F/L\ll kT\ll \Delta$, $f(V,T)=\min(eV,\Delta)$:
the anomalous current thus scales linearly with applied voltage up to
$\Delta$. It follows from Eq. (\ref{iasimple}) that $I_a$ is
independent of the length of the junction, being the sum of $N\sim L$
levels which each carries a current $I_n\sim 1/L$. This gives that the
anomalous current roughly is equal to the total equilibrium current of
the short junction. The critical anomalous current is plotted with
respect to transparency in the inset in Fig.~\ref{panvolt}. In the
limit $D\ll 1$ it is given by
$(I_a)_c=(e/\hbar)(\sqrt{D}/\pi)f(V,T)$. It is proportional to the
first power of $\Delta$ for $T$ close to $T_c$, therefore surviving up
to $kT\approx\Delta$. The anomalous current-phase relation (see inset
in Fig. \ref{panvolt}) is $2\pi$ periodic and resembles that of the
equilibrium Josephson current. The direction of the anomalous current
is however proportional to $\sigma$, i.e dependent on the phase of the
scatterer at the connection point, which is not the case for the
equilibrium Josephson current.
To get the complete picture of the Josephson current in a long
junction, $I=I_{eq}+I_r+I_a$ is plotted as a function of bias voltage
for different temperatures in Fig. \ref{totalcurr}.
\begin{figure}
\centerline{\psfig{figure=longivc.eps,height=6.5cm}}
\caption{The total current $I=I_{eq}+I_r+I_a$ as a function of
voltage. At zero temperature we have $I_r+I_{eq}$ (dash-dotted), $I_a$
(dashed) and the total current $I_r+I_{eq}+I_a$ (solid). The total
current for temperatures $kT \gg \hbar v_F/L$ is plotted
(dotted). Junction parameters are $D=0.8, \phi=3\pi/4, L=20 \xi_0$,
$\epsilon=0.01$ and $\sigma=-1$}
\label{totalcurr}
\end{figure}
The zero temperature total Josephson current oscillates strongly
around a constant slope as a function of voltage, showing steps
whenever the voltage passes an Andreev level. The step structures are
washed out for temperatures $kT \gg \hbar v_F/L$, and in this limit
the total current roughly coincides with the anomalous current, given
by Eq. (\ref{iasimple}).
\subsection{Asymmetric junction}
The effect of asymmetry is most pronounced in the long limit when the
asymmetry is much larger than the coherence length but much smaller
than the total length of the junction, $L \gg l \gg \xi_0$. In this
limit, the derivative of energy with respect to phase $dE/d\phi$ in
Eq. \ref{asymdedfi} reduces to the expression of a symmetric long
junction (\ref{longdedfi}), since $|\sin 2\theta|>R|\sin \chi|$ (see
Appendix B). The equilibrium current $I_{eq}$ and the regular current
$I_r$ are not substantially changed in comparison to the symmetric
case. In contrast, the anomalous current is modified in a nontrivial
way, taking the form
\begin{eqnarray}
I_a&=&-\sigma {e\over \hbar} \frac{\hbar v_F}{L} \sqrt{R}D^{3/2}
\sin\phi \nonumber \\ &&\times
\sum_{n,\pm}^N\frac{\cos\chi-\cos\phi}{1-(D\cos\phi+R\cos\chi)
^2}(n^e-n^h),
\label{ianascurr}
\end{eqnarray}
obtained by inserting Eq. (\ref{longdedfi}) into (\ref{iabound}).
For $T=0$ the step structure in the IVC is modified due to the change
of Andreev levels as a result of the asymmetry (see
Fig.~\ref{asandr}). Already for small asymmetry $l\sim\xi_0$ the
anomalous current might change dramatically. Depending on the phase
difference of the junction, the IVC is renormalized and changes sign
for $-\pi/2<\phi<\pi/2$.
\begin{figure}
\centerline{\psfig{figure=anasvolt.eps,height=5.5cm}}
\caption{The asymmetric anomalous current $I_{a}$ vs voltage for
different asymmetries (a) $l=0\xi_0$, (b) $l=2\xi_0$ and (c)
$l=40\xi_0$ for $kT\gg \hbar v_F/L$, $D=0.8$, $\epsilon=0.05$,
$L=60\xi_0$, $\phi=\pi/4$ and $\sigma=-1$. The IVC is changed
dramatically already for as small asymmetry $l\sim \xi_0$, if the
phase difference $-\pi/2<\phi<\pi/2$}
\label{ianas}
\end{figure}
When the temperature is increased beyond the interlevel distance
$kT\gg \hbar v_F/L$, the step structure becomes smeared and we get a
periodic modulation of the IVC on the scale of $eV\sim\hbar
v_F/l$. This modulation arises from the factor $\cos \chi$.
When the temperature is further increased to $kT\gg \hbar v_F/l$ this
periodic structure is smeared out and the IVC once again becomes a
straight line, but with renormalized slope. In this high temperature limit
the amplitude of the terms in the sum in Eq. (\ref{ianascurr}) oscillates
with a period $\hbar v_F/l$. During this period, the filling factors
$n$ can be taken to be constant, and we can sum over one period to get the
average value. Performing this summation in the continuum limit, we get
\begin{eqnarray}
&&\sum_{one~period}\frac{\cos\chi-\cos\phi}{1-(D\cos\phi+R\cos
\chi)^2} \nonumber \\ &&\approx \frac{\hbar
v_F}{2l}\int_0^{2\pi}\frac{\cos(\chi)-\cos(\phi)}{1-(D\cos\phi+R
\cos\chi)^2} d\chi \nonumber \\
&&=\frac{L}{l}\frac{1}{8R\sqrt{D}}\left(
\frac{|\sin(\phi/2)|}{\sqrt{1-D\cos^2(\phi/2)}}-\frac{|\cos(\phi/2)|}{\sqrt{1-D\sin^2(\phi/2)}}\right). \nonumber \\
\label{ianassum}
\end{eqnarray}
This quantity is energy independent and we can then sum over the filling
factors
following the procedure from the symmetric case (\ref{sumint})
\begin{equation}
\sum_{averaged~periods}(n^e-n^h)\approx\frac{4l}{\hbar v_F\pi}f(V,T).
\end{equation}
The anomalous current then becomes
\begin{eqnarray}
&&I_a=-\sigma 2 {e\over \hbar} \frac{D}{\pi \sqrt{R}} \sin\phi
\nonumber \\
&&\times
\left(\frac{|\sin(\phi/2)|}{\sqrt{1-D\cos^2(\phi/2)}}-\frac{|\cos(\phi/2)|}{\sqrt{1-D\sin^2(\phi/2)}}\right)f(V,T), \nonumber \\
\label{renormanasym}
\end{eqnarray}
which is independent of both the length $L$ and the asymmetry $l$. We
also find that the renormalized anomalous current becomes
$\pi$-periodic. This can qualitatively be explained by the fact that
the $2\pi$-periodic part of the anomalous current density is very
sensitive to asymmetry, oscillating fast with energy on the scale of
$\hbar v_F/l$, becoming washed out during summation over bound states
at high temperatures $kT\gg\hbar v_F/l$. The $\pi$ periodic part of
the current does not have this sensitivity and is the only part of the
anomalous current that survives. The asymmetric anomalous
current-phase relation is shown in Fig. \ref{ianasph}.
\begin{figure}
\centerline{\psfig{figure=anasfi.eps,height=6.5cm}}
\caption{The asymmetric anomalous current $I_a$ at $eV=\Delta$ and $kT\gg
\hbar v_F/L$ as a function of phase difference $\phi$ for different
transparencies $D=0.1, 0.5$ and $0.9$. Inset: The critical anomalous
current $(I_a)_c$, for $eV=\Delta$ and $kT\gg \hbar v_F/L$, as a function of
transparency $D$ for coupling constant $\epsilon=0.1$.}
\label{ianasph}
\end{figure}
The $\pi$-periodicity and the zeros at $\phi=n\pi/2$ give the
condition that the slope of the IVC must change sign due to asymmetry
in the range $-\pi/2<\phi<\pi/2$, as shown in Fig. \ref{ianas}. The
critical asymmetric anomalous current as a function of transparency
$D$ is shown in the inset in in Fig. \ref{ianasph}. The critical
asymmetric anomalous current as a function of transparency $D$ is
shown in the inset in in Fig. \ref{ianasph}. The behavior is very
similar to the critical anomalous current in the symmetric case, the
main difference being that the amplitude is reduced by roughly a
factor of two.
\section{Interface barriers}
In any realistic experimental situation, normal reflection at the
NS-interface, modeled by a barrier with reflection amplitude $r_b$,
must be taken into account. \cite{Chrestin} The
general expression, considering both the interface barriers and the
midpoint scatterer, becomes analytically intractable. We can however
analyze the case where the midpoint scatterer is absent ($R=0$) to get
an understanding of the effect of NS-barriers on the junction
properties, and then treat the general case with injection and
midpoint scatterer numerically.
In the absence of the superconducting leads (a NININ-junction), the
two barriers give rise to normal Breit-Wigner resonances for the
electrons and holes. Understating the properties of these resonances
turns out to be crucial for describing the behavior of Andreev levels
and current transport. The energies of the electron and hole
resonances are calculated straightforwardly
\begin{eqnarray}
E^e_n=-2E_F\left[1-\frac{\pi(n-\nu)}{k_FL}\right], \nonumber \\
E^h_m=2E_F\left[1-\frac{\pi(m-\nu)}{k_FL}\right],
\end{eqnarray}
where $r_b=\sqrt{R_b}e^{i\nu\pi}$ and $n(m)$ are integers denoting the
index of the electron (hole) resonances. The intersection between
electron and hole resonances ($E^e_n=E^h_m$) is given by
$L_{n+m}=\lambda_F/4(m+n-2\nu)$ with the Fermi wavelength
$\lambda_F=2\pi/k_F$. These normal resonances are plotted in
Fig. \ref{bwandrres}.
\begin{figure}
\centerline{\psfig{figure=barrand.eps,height=6.5cm}}
\caption{The Andreev levels (solid) and the normal electron and hole
resonances (dotted) as a function of length $L$ of the junction with
four Andreev levels in (a) weak resonance limit $R_b\ll 1$ (b) strong
resonance limit $R_b\sim 1$. The lengths of two subsequent
intersections of normal resonances $L_{n+m}$ and $L_{n+m-1}$ are shown
with arrows}
\label{bwandrres}
\end{figure}
For the junction with superconducting leads, one can in the same way
as before calculate the equation for the bound Andreev states
($|E|<\Delta$), with
the result \cite{Wendin2}
\begin{eqnarray}
&&D_b^2\cos\phi+2R_b\cos\beta-\cos(2\gamma-\beta)-R_b^2\cos(2\gamma+\beta)
\nonumber \\
&&-4R_b\sin^2\gamma\cos(\beta_0)=0
\end{eqnarray}
where we have defined $\beta_0=\pm 2E^{e,h}/(\hbar v_F/L)$ and $+(-)$ denotes
hole(electron) resonance energies. One can draw some qualitative conclusions
on how the Andreev levels are related to the normal resonances by looking
at Fig. \ref{bwandrres}. In the limit of high barrier transparency
$R_b\ll 1$, the Andreev levels are
weakly modified by the barriers. In the opposite limit $R_b\sim 1$, the
Andreev levels get pinned at the normal resonances, but there are no
level crossings at the points where the normal electron and hole
resonances intersect.
We find the same interlevel distance $\hbar v_F \pi/L$ in the junction
with the superconducting leads (SINIS junction) and normal leads
(NININ-junction). The main difference is that the normal resonance
move very quickly through the junction when the length $L$ increases,
while the Andreev levels oscillate up and down.
Considering Andreev state energies close to the Fermi level,
$E\ll \Delta$, one can derive a simplified dispersion relation
\cite{Kadigrobov2}
\begin{equation}
\sin^2(\beta/2)=\frac{D_b^2\cos^2(\phi/2)+4R_b\sin^2(\beta_0/2)}{(1+R_b)^2}.
\label{barrel}
\end{equation}
Using this relation we can study the bound state current in different
length limits.
In the short limit, $L\ll \xi_0$ there are two cases to be considered. For
nearly transparent barriers $D_b\sim 1$, and thus broad resonances
$\Gamma=D_b
\hbar v_F/L\gg \Delta$, one can neglect dephasing (putting $\beta=0$) and
just get the total transparency of the junction
$D=D_b^2/(D_b^2+4R_b\sin^2(\beta_0/2))$ to be put into the standard zero
length junction equilibrium current formula. In the strong barrier case
$D_b\ll 1$ the resonances are sharp $\Gamma\ll \Delta$ and one can not
neglect the dephasing. Assuming that the resonance is close to Fermi energy
$E^{e,h}\ll\Delta$, we can put $\beta\ll 1$ in (\ref{barrel}) and
obtain\cite{Wendin2,Beenakker4}
\begin{equation}
E=\pm \sqrt{\Gamma^2\cos^2(\phi/2)+(E^{e,h})^2}.
\end{equation}
When the resonance is exactly at the Fermi energy $E^{e,h}=0$, the
Josephson current is given by
\begin{equation}
I=\frac{e\Gamma}{\hbar}\sin(\phi/2)\tanh\left(\frac{\Gamma\cos(\phi/2)}{2kT}
\right).
\end{equation}
The critical current at low temperatures ($kT\ll \Gamma$) is thus smaller
than the critical current of a short, clean junction by a factor
$\Gamma/\Delta$.
For a long junction $L\gg\xi_0$ we can calculate the derivative of energy
with respect to phase,
\begin{equation}
\frac{dE}{d\phi}=\pm\frac{\hbar
v_F}{2L}\frac{D_b^2\sin\phi}{\sqrt{(1+R_b)^4-[D_b\cos\phi-4R_b\cos(\beta_0)]^2}}
,
\label{dedfibarr}
\end{equation}
In the weak barrier limit $R_b\ll 1$, this just causes oscillations
with length around the clean junction ($R_b=0$) result. In the strong
barrier limit $R_b\sim 1$, one can distinguish two limits: When the
length of the junction is far away from the length $L_{n+m}$ where the
electron and hole resonances intersect, the junction is out of
resonance. The second term in Eq. (\ref{dedfibarr}) is negligible and
the current from the individual levels thus becomes
\begin{equation}
I=\pm\frac{e v_F}{4L} D_b^2\sin\phi.
\end{equation}
It is proportional to $D_b^2$ and thus strongly suppressed. In the
opposite limit, when the length of the junction
$L=L_{n+m}=\lambda_F(m+n-2\nu)/4$, the junction is in resonance. When
$n+m$ is even we get the current carried by each level
\begin{equation}
I=\pm\frac{e v_F}{L}\frac{D_b\sin\phi}{4|\cos(\phi/2)|}
\label{ires1}
\end{equation}
and when $n+m$ is odd we get
\begin{equation}
I=\pm\frac{e v_F}{L}\frac{D_b\sin\phi}{4|\sin(\phi/2)|}.
\label{ires2}
\end{equation}
We see that the current is proportional to $D_b$, just as expected for
the junction in resonance. The current carried is thus of the order of
the single barrier junction current. An interesting feature is that
the current is dependent on the parity of the sum of the electron and
hole resonance indices $n+m$. When the third lead is connected to the
junction, the scattering at the connection point just splits the Breit
Wigner resonances, and the qualitative picture for the bound states
derived without the third lead connected survives.
To calculate the total equilibrium, regular or anomalous current, the
currents carried by all individual levels have to be summed up. In the
weak barrier limit $R_b\ll 1$ we just find that all properties
calculated above for the symmetric junction without barriers hold,
with a small length dependent modulation $\sim R_b$ with a period
$\delta L\sim \lambda_F$. In the strong barrier limit $R_b\sim 1$, the
result will depend on whether the junction is in or out of resonance.
\begin{figure}
\centerline{\psfig{figure=barreq.eps,height=6.5cm}}
\caption{Short segment of the equilibrium bound state current as a
function of length $L$, to illustrate the resonant behavior. The
junction is long $L\gg \xi_0$ with $R_b=0.9$, $\epsilon=0.01$ and
$\phi=\pi/2$.}
\label{paseq}
\end{figure}
Fig. \ref{paseq} shows the resonant behavior of the equilibrium
current as a function of length. The current has a peak around lengths
$L=\lambda_F/4(m+n+2\nu)$. The phase dependence of the current at the
resonant peaks is well described by the expressions for the single
level currents (\ref{ires1}) and (\ref{ires2}).
The anomalous current is also strongly length dependent and when the
junction is in resonance we have an anomalous current $I_a\sim \sigma
D_b \sqrt{RD}$ while when we are out of resonance $I_a\sim \sigma
D_b^2 \sqrt{RD}$. It turns out that there is an anomalous current even
without scattering at the connection point, but it oscillates around
zero as a function of length with the period $\sim \lambda_F$.
\section{Injection current and conductance}
Although the nonequilibrium Josephson current is at the focus of this
paper, the injection current that flows between the normal reservoir
and the SNS-junction is also of great interest: it determines the
conductance of the circuit. The conductance of NS-structures has been
studied intensely in recent years \cite{Lambertrew} and is an
interesting quantity in itself. It can also be used to determine the
direction of the Josephson current in the junction \cite{Vanwees} or
to detect a large Josephson current in the superconducting loop which
changes the applied external flux vs phase dependence, thus modifying
the phase dependence of the conductance. \cite{Phaserel}
\subsection{Injection current}
We start by discussing the symmetric junction $l=0$, and comment on
the modifications due to asymmetry below. The current injected in
lead $1$ for energies $|E|<\Delta$ can be calculated to lowest order
in $\epsilon$ by inserting Eq. (\ref{injcurrdens}) into
Eq. (\ref{i1def})
\begin{eqnarray}
I_1&=&\epsilon\frac{e}{\hbar}\frac{|\cos(\phi/2)|}{\sqrt{D}\sqrt{1-D\sin^2(\phi/
2)}}
\nonumber \\ &&\times
\sum_{n,\pm}\left|\frac{dE_n^{\pm}}{d\phi}\right|\left[n^e(E_n^{\pm})-n^h(E_n^{\pm})\right].
\label{iinj}
\end{eqnarray}
This current is proportional to $\epsilon$ (unlike the Josephson
current discussed above), which is also true for the continuum
contribution. For $|E|>\Delta$, the injection current density in
Eq. $(\ref{i1def})$, $i_{inj} = i_1^-=i_3^--i_2^-$, is obtained from
Eq. (\ref{contdiffcurr}). This current density is roughly described by
the normal current density $4\epsilon e/h$, with oscillations around
this value due to the resonances in the scattering states (see
Fig. \ref{injection}). These oscillations are strongest around $E\sim
\Delta$ and decrease with increasing energy.
The injection current is proportional to the modulus $|dE/d\phi|$,
just like the anomalous current, as discussed above. The IVC thus has
the shape of a staircase, as shown in Fig. \ref{injection}.
\begin{figure}
\centerline{\psfig{figure=i1volt.eps,height=6.5cm}}
\caption{The injection current in lead 1 as a function of voltage for
(a) $\phi=\pi/4$, (b) $\phi=3\pi/4$ and (c) $\phi=\pi$. Zero $T$
(solid lines) and $T=0.05\Delta$ (dashed line) with $D=0.8$,
$L=10\xi_0$ and $\epsilon=0.05$. For $eV>\Delta$ the IVC approaches
the value of a normal junction}
\label{injection}
\end{figure}
We can derive expressions for the injection current in different
length limits. A complete set of formulas are given in Appendix A. Here we
only present the heights of the current steps in the IVC:s in some
representative cases.
In the limit of zero length of the junction ($L=0$) there is only one
Andreev level (for $0<E<\Delta$) and the magnitude of the step is
\begin{equation}
\delta
I_1=\epsilon\frac{e}{\hbar}\frac{\sqrt{D}}{2}\frac{|\sin(\phi/2)|\cos^2(\phi/2)}
{1-D\sin^2(\phi/2)}
\end{equation}
while for two levels close to $\Delta$ we get ($L\sim \xi_0$)
\begin{equation}
\delta I_1=\epsilon
\frac{e\Delta}{\hbar}\frac{L}{\xi_0}\frac{\cos^2(\phi/2)}{\sqrt{1-D\sin^2(\phi/2)}}.
\end{equation}
In the long junction limit ($L \gg \xi_0$) the current is given by
putting Eq. (\ref{longdedfi}) into (\ref{iinj}),
\begin{equation}
I_1=\epsilon\frac{e}{\hbar}\frac{\hbar
v_F}{2L}\frac{\cos^2(\phi/2)}{1-D\sin^2(\phi/2)}\sum_{n=0}^{N}[h(E_n^-)+h(E_n^+)
].
\label{iinjlong}
\end{equation}
The current step at zero temperature is
\begin{equation}
\delta I_1=\epsilon\frac{e}{\hbar}\frac{\hbar
v_F}{L}\frac{\cos^2(\phi/2)}{1-D\sin^2(\phi/2)}.
\end{equation}
For high temperatures $kT \gg \hbar v_F/L$, the sum (\ref{iinjlong}) can be
converted to an integral, just as for the anomalous current
(\ref{sumint}), which gives the current for $eV<\Delta$
\begin{equation}
I_1=\epsilon\frac{4e}{h}\frac{\cos^2(\phi/2)}{1-D\sin^2(\phi/2)} f(V,T).
\label{iinjhight}
\end{equation}
where $f(V,T)$ is given by Eq. (\ref{fdef}). The $IVC$ thus becomes
linear for $eV<\Delta$, with the slope independent of length and
temperature, as is seen in Fig. \ref{injection}. All information about
individual Andreev levels is washed out.
The effect of asymmetry on the injection current is drastic in the
limit of a long junction with large asymmetry $L\gg l\gg \xi_0$, just
as for the asymmetric anomalous current. This shows the strong
relationship between the two currents. The injection current is given
by inserting Eq. (\ref{injcurrdens}) into (\ref{i1def})
\begin{equation}
I_1=\epsilon\frac{e}{\hbar}\frac{\hbar v_F}{L}\sum_{n=0}^{N}\frac{D
\sin^2(\phi)+R\sin^2\chi}{1-(D\cos\phi+R\cos\chi)^2}(n^e-n^h).
\end{equation}
Averaging over periods and summing up the filling factors, just as in
the case of the anomalous current [see Eq. (\ref{ianassum})], the
injection current becomes
\begin{eqnarray}
&&I_1=\epsilon\frac{e}{\hbar}\frac{8}{R\pi} \left[1-\sqrt{D} \right.
\nonumber \\ && \left. \times
\left(\frac{|\sin(\phi/2)|^3}{\sqrt{1-D\cos^2(\phi/2)}}+\frac{|\cos(\phi/2)|^3}{\sqrt{1-D\sin^2(\phi/2)}}\right)\right] f(V,T). \nonumber \\
\label{iinjasym}
\end{eqnarray}
The injection current in this limit does not depend on either the
length $L$ or the asymmetry $l$. It is $\pi$-periodic,
$I_1(\phi+\pi)=I_1(\phi)$, for the same reasons as discussed above for
the asymmetric anomalous current. The implications of the $\pi$
periodicity for the conductance is discussed below.
\subsection{Conductance}
For the conductance, we discuss the whole range of coupling parameters
$\epsilon$, not only the weak coupling limit. The conductance is defined
\begin{eqnarray}
&&G(V,\phi)= \frac{dI_1}{dV}
\frac{d}{dV}(n^e-n^h)
\nonumber \\
&&= \frac{e}{kT} \int_{-\infty}^{\infty}\frac{i_1^-}{2}
[\cosh^{-2}\frac{(E+eV)}{2kT}+ \cosh^{-2}\frac{(E-eV)}{2kT}].
\end{eqnarray}
At zero temperature $T=0$ the conductance for a symmetric
junction $l=0$ can be written \cite{Nakano} for $eV<\Delta$,
noting that $i_1^-(E)=i_1^-(-E)$,
\begin{equation}
G(V,\phi)=\epsilon^2 \frac{4e^2}{h}\frac{4
\cos^2(\phi/2)\sin^2\theta}{[(1-\epsilon)\cos2\theta-R-D\cos\phi]^2+
\epsilon^2\sin^2 2\theta}.
\label{conduct}
\end{equation}
At a voltage fulfilling the relation $1-\cos 2
\theta=D(1-\cos\phi)/\sqrt{1-2\epsilon}$, which is exactly at an
Andreev resonance, and a phase difference $\phi=0$ mod $2\pi$, the
conductance is $G=4e^2/h$, which is the conductance for a perfect
NS-interface. This holds for any transparency $D$, length $L$ and
coupling constant $\epsilon$.
For $L=0$ we get $\cos\theta = eV/\Delta$ to be inserted into
(\ref{conduct}). This gives rise to a peak in the conductance at the
voltage $eV\approx E_0^-$, the energy of the single Andreev
resonance. In the long limit $(L\gg \xi_0)$ for $E \ll \Delta$ we get
$\theta=\pi/2-eV/E$ which results in conductance oscillations as a
function of applied voltage
\cite{Morpurgo97b,Volkzait,Dimoulas,Lesovik}, with a period $\pi \hbar
v_F/L$, the distance between the pairs of Andreev levels. This is made
clear by taking the derivative of current with respect to voltage in
Fig. \ref{injection}.
The conductance vs phase relation $G(V,\phi)$ is also altered when
voltage is applied. For voltages around $eV/(\hbar v_F/L) \approx n
\pi$, i.e at an energy between the pairs of Andreev resonances, the
conductance has a maximum around $\phi\approx\pi$ and a local minimum
at $\phi=0$ [apart from the absolute minimum at $\phi=\pi$ due to the
factor $\cos^2(\phi/2)$ in Eq. (\ref{conduct})]. For $eV/(\hbar v_F/L)
\approx\pi/2+n\pi$, i.e at an energy between the two Andreev
resonances in the pair, the maximum shifts to $\phi=0$ and the minimum
to $\phi=\pi$ [see Fig. (\ref{condflucv})]. This behavior has
recently been observed in both ballistic \cite{Morpurgo97b} and
quasiballistic \cite{Dimoulas} junctions. It has also been predicted
for diffusive junctions. \cite{Volkzait,Leadbeater} This voltage
dependence of the conductance holds for $kT \ll \hbar v_F/L$.
\begin{figure}
\centerline{\psfig{figure=condflucv.eps,height=6.5cm}}
\caption{The conductance as a function of $\phi$ for different voltages
(a) $eV=0$, (b) $eV=0.075\Delta$ and c) $eV=0.15\Delta$ at temperature
$kT=0.01\Delta$. $D=0.9$, $\epsilon=0.05$ and $L=20\xi_0$.}
\label{condflucv}
\end{figure}
For the additional condition zero voltage $V=0$, the expression
(\ref{conduct}) reduces to
\begin{equation}
G(0,\phi)= \epsilon^2 \frac{2e^2}{h}\frac{2\cos^2(\phi/2)}{R+D\cos^2(\phi/2)}.
\end{equation}
For $D\leq R$ the maximum conductance has a universal magnitude
$G_{max}(0)=G_N\epsilon/(1-\epsilon)^2$ where $G_N=\epsilon 4e^2/h$ is
the conductance of the junction in the normal state.
When the coupling is weak, $\epsilon \ll 1$, the maximum conductance
is $G_{max} \sim \epsilon G_N$, i.e much smaller than the normal
conductance. In this limit the Andreev resonances are sharp and there
are no available Andreev states at $E_F$, because the scattering at
the three lead connection point opens up a gap in the Andreev spectrum
at $\phi=\pi$ (see Fig. \ref{asandr}). The conductance is thus
strongly suppressed. The conductance rapidly increases with voltage
and temperature and has a maximum at $eV,kT\sim \hbar v_F/L$. This
happens because electrons (holes) with energies $E>0$ ($E<0$) tunnel
into the first resonant Andreev state at finite energy. This gives
rise to a finite energy conductance peak \cite{Kastalsky} (see
Fig. \ref{condtemp}). The maximum conductance $G_{max}$ is plotted as
a function of temperature in Fig. \ref{condtemp} (note that the
minimum conductance always is zero).
\begin{figure}
\centerline{\psfig{figure=condt.eps,height=6.5cm}}
\caption{The maximum conductance $G_{max}/G_N$ as a function of $T$
for zero voltage. $L=10\xi_0$, $D=0.9$ and $\epsilon=0.05$. At $T=0$
the maximum conductance is $G_{max} \ll G_N$. It increases with
increasing $T$, reaches a maximum around $kT\sim \hbar v_F/L$, drops
again for $kT>\hbar v_F/L$ but saturates at a constant value
$G_{max}=G_N$ for $kT \gg \hbar v_F/L$. Inset: The conductance as a function
of $\phi$ for different temperatures $kT=0, 0.01\Delta, 0.025\Delta,
0.05\Delta$ and $0.1\Delta$ at zero voltage $V=0$. Increasing
temperature from bottom to top at $\phi=0$. $D=0.9$, $\epsilon=0.05$
and $L=20\xi_0$.}
\label{condtemp}
\end{figure}
The conductance as a function of phase difference $\phi$ behaves
differently for different temperatures $T$, as is shown in the inset
in Fig. \ref{condtemp}. For $kT\gg \hbar v_F/L$, a limit only accessible for
the long junction $L\gg\xi_0$, the maxima of the conductance around
$\phi\approx\pi$ are shifted to a maximum at $\phi=0$ mod $2\pi$. This
holds independent of voltage applied. A similar effect in a
quasiballistic system has been reported by Dimoulas et al.\cite{Dimoulas}
In the weak coupling limit $\epsilon \ll 1$ we can use expression
(\ref{iinjhight}) to get the conductance in the long limit for $\hbar
v_F/L \ll kT\ll \Delta$ and $eV<\Delta$
\begin{equation}
G(\phi)=\epsilon \frac{4e^2}{h}\frac{\cos^2(\phi/2)}{1-D\sin^2(\phi/2)}
\label{condsimp}
\end{equation}
which is independent of voltage, temperature and length of the
junction and has a maximum $G_{max}=G_N$ at $\phi=0$ mod $2\pi$.
In the same limit we get the conductance in the asymmetric junction
from expression (\ref{iinjasym})
\begin{eqnarray}
&&G(\phi)=\epsilon\frac{e}{\hbar}\frac{8}{R\pi} \left[1-\sqrt{D}
\right. \nonumber \\
&& \left. \times
\left(\frac{|\sin(\phi/2)|^3}{\sqrt{1-D\cos^2(\phi/2)}}+\frac{|\cos(\phi/2)|^3}{\sqrt{1-D\sin^2(\phi/2)}}\right)\right].
\label{condas}
\end{eqnarray}
It is $\pi$-periodic and has a maximum for $\phi=\pi(n+1/2)$ and a
minimum for $\phi=\pi n$. This $\pi$-periodicity can be qualitatively
explained by considering the lowest order paths giving rise to the
conductance.
\begin{figure}
\centerline{\psfig{figure=pathanas.eps,height=4cm}}
\caption{The paths in the asymmetric junction giving the first order terms
of the conductance. Electrons are drawn with solid lines, holes with
dashed. The grey ellipse denotes the effective scatterer due to the three
lead connection. The upper paths give rise to a $2\pi$ periodic component
of the current, suppressed at finite temperature $kT\gg \hbar v_F/L$. The lower
paths, time reversed, give rise to a $\pi$ periodic component of the
current, not suppressed by temperature.}
\label{asym}
\end{figure}
The upper paths in Fig. \ref{asym}, corresponding to an injected
electron and giving rise to an outgoing hole, produce a current of the
order $i_1\sim
|e^{i(\phi_L+\beta_2)}+e^{i(\phi_R+\beta_3)}|^2=2+2\cos(\phi+\chi)$. This
part of the current is $2\pi$ periodic in the phase, rapidly
oscillating in energy with a period $\hbar v_F/l$. It is washed out
when summing up the levels in a long junction at finite
temperature. The lower paths in Fig. \ref{asym}, corresponding to an
injected electron giving rise to an outgoing electron, produce a
current of the order $i_1 \sim
|d^*e^{i(\phi+\beta)}+d^*e^{i(-\phi+\beta)}|^2=D(2+2\cos2\phi)$. This
part of the current is $\pi$ periodic in phase and not sensitive to
asymmetry.
The discussion about the periodicity of the conductance oscillations
with respect to phase goes back to the early eighties. A
$\pi$-periodic contribution to the weak localization correction to the
conductance in a similar system was predicted by Spivak et
al.\cite{Spivak} and discussed further by Altshuler et
al. \cite{Altshuler} It has been shown in numerical simulations for a
structure similar to ours that the full conductance, i.e not only the
weak localization contribution, might become $\pi$-periodic at finite
temperatures. \cite{Hui} A large $\pi$-periodic conductance
oscillation with phase was also observed in diffusive samples
\cite{Petrashov}. Whether the explanation to the crossover from $2\pi$
to $\pi$ periodicity with increased temperature discussed above can
account for these observations remains to be investigated.
\section{Four terminal junction}
One problem with the three terminal junction is, as discussed above,
that it is not possible to separate the injection current from the
Josephson current in a clear way for arbitrary coupling $\epsilon$. In
a four terminal
junction\cite{Bagwell,Volkov95,Kadigrobov2,Petrashov4,Nazstoof} this
is possible under certain conditions, which makes it interesting to
discuss this configuration separately.
We consider two different types of junction configurations (see
Fig. \ref{fourterm}). The upper junction is a straightforward
extension of the three terminal device pictured in
Fig. \ref{juncfig}. Two normal reservoirs are connected to the normal part of the junction. The reservoirs are then connected to
the grounded superconducting loop via voltage sources biased at
$V_1$ and $V_4$ respectively.
\begin{figure}
\centerline{\psfig{figure=fig4.eps,height=6.5cm}}
\caption{Two different setups of the four terminal junction. In (a) the
normal reservoirs are biased independently at $V_1$ and $V_4$ with
respect to the superconducting loop (grounded), in (b) only the potential
difference between the normal reservoirs, $V$, is determined. In the
right figure, a close up of the junction area is shown, with the
direction of the currents showed with arrows}
\label{fourterm}
\end{figure}
The current injected from a normal reservoir is split at the
connection point. One part of the current flows through the junction
directly into the other reservoir and the other part of the current is
divided between the leads $2$ and $3$. In the general case, the
currents in lead $1$ and $4$ are not equal $I_1 \neq I_4$. However, by
adjusting the potentials $V_1$ and $V_4$, the currents in the vertical
lead can be put equal and from current conservation at the connection
point it follows $I_2=I_3$ and we have a clear separation between the
injection current $I_1=I_4$ and the Josephson current circulating in
the superconducting loop $I_2=I_3$.
In the lower junction, this separation follows directly from current
conservation at the connection point, since $I_2=I_3$. In this
junction a bias $V$ is applied between the normal reservoirs, which
are only connected to the superconducting loop via the four lead
connection point. We define the potential of the superconducting loop
in this junction to be zero and the potentials of the normal
reservoirs to $V_1$ and $V_4$ resepectively, just as for the upper
junction. Our biasing arrangement then gives $V=V_1-V_4$. The
condition of current conservation, $I_1(V_1,V_4)=I_4(V_1,V_4)$, gives
a second condition on $V_1$ and $V_4$. With these definitions we can
calculate the current in both junctions in the same way.
The cross-shaped connection point is modelled by the scattering matrix
\begin{equation}
S=\left( \begin{array}{llll}
r_{\perp} & \sqrt\epsilon & \sqrt\epsilon & d_{\perp}\\
\sqrt\epsilon & r & d & \sqrt\epsilon\\
\sqrt\epsilon & d & r & \sqrt\epsilon \\
d_{\perp} & \sqrt\epsilon & \sqrt\epsilon & r_{\perp}\\
\end{array} \right)
\end{equation}
where the $\epsilon$ describes the coupling of the SNS junction to the
vertical normal lead ($0\leq \epsilon\leq 0.25$). The horisontal scattering
amplitudes now obey the relations $Re(rd^*)=-\epsilon$ and
$D+R=1-2\epsilon$. The same holds for the vertical scattering
amplitudes $r_{\perp}$ and $d_{\perp}$.
The current densities $i_{j}^{e(h),1(4)}$, with the upper index $1$ or
$4$ denoting the lead from which the quasiparticles are injected, are
calculated in the same way as in the case of the three terminal
junction. Due to the symmetry of the scattering matrix, quasiparticles
injected from leads $1$ or $4$ give rise to the same current density
in leads $2$ and $3$, i.e $i_{2(3)}^{e(h),1}=i_{2(3)}^{e(h),4}$.
The expressions for the sum and anomalous current densities become
very similar to the three terminal expressions [see
Eq. (\ref{sumcurr}) and (\ref{avdiffcurr})], i.e one just changes
$\epsilon \rightarrow 2\epsilon$ [also changing $Z=Z(\epsilon
\rightarrow 2\epsilon)$] and divides by two, noting that
$i^{+,1}=i^{+,4}$ and $i_a^{1}=i_a^{4}$. Neither the vertical
transparency $D_{\perp}$ nor the reflectivity $R_{\perp}$ thus appear
explicitly in these expressions. The factor one half simply reflects
that there are {\it two} normal leads connected to the normal part of
the junction. In the limit of weak coupling $\epsilon \ll 1$, the sum
of the current densities from both normal reservoirs is equal to the
current density from the single normal reservoir in the three terminal
junction, $i^{+,1}+i^{+,4}=i^+$ and $i_a^1+i_a=i_a$ (simply reflecting
that one cannot create more Josephson current by adding more normal
leads).
In this weak coupling limit the current in the horisontal lead
$I=I_2=I_3$ is given by
\begin{equation}
I=I_{eq}+\frac{1}{2}[I_r(V_1)+I_r(V_4)]+\frac{1}{2}[I_a(V_1)+I_a(V_4)]
\end{equation}
with $I_{eq},I_r$ and $I_a$ the same as in the three terminal case,
given by Eq. (\ref{irbound})-(\ref{sumboundstate}). Noting the
relations $I_r(-V)=I_r(V)$ and $I_a(-V)=-I_a(V)$, we see that (i) for
bias $V_1+V_4=0$ the anomalous current is zero, and (ii) for
$V_1-V_4=0$ the regular current is zero. We can thus control the
regular and anomalous currents in the upper junction in
Fig. \ref{fourterm} independently by adjusting either the potential
difference $V_1-V_4$ or the sum $V_1+V_4$ between the normal
reservoirs, keeping the other quantity constant.
\subsection{Injection current and conductance}
The injection currents $I_1$ and $I_4$ in the four terminal device is
qualitatively different from the injection current in the three
terminal device, since in the four terminal junction the injected
quasiparticles from one normal reservoir can travel directly through
the junction to the other normal reservoir.
The current in leads $j=1,4$ can be written [see Eq. (\ref{totcurr})]
\begin{eqnarray}
I_j&=&\int_{-\infty}^{\infty}\ dE \left[\frac{i_j^{+,1}}{2}(n^{e,1}+n^{h,1})+
\frac{i_j^{-,1}}{2}(n^{e,1}-n^{h,1})+ \right. \nonumber \\
&& \left.
+\frac{i_j^{+,4}}{2}(n^{e,4}+n^{h,4})+\frac{i_j^{-,4}}{2}(n^{e,4}-n^{h,4})
\right]
\label{fourinj}
\end{eqnarray}
The symmetry of the junction gives that $i^{-,1}_4=-i^{-,4}_1$ and
$i^{-,4}_4=-i^{-,1}_1$. This leads to that for $V_1=-V_4=V/2$ the
currents in the vertical leads are equal, i.e $I_{inj}=I_1=I_4$ and
thus no injection current flows into the superconductors. In this case
Eq. (\ref{fourinj}) reduces to
\begin{equation}
I_{inj}=\int_{-\infty}^{\infty} dE \left[ \frac{i_1^{-,1}-i_1^{-,4}}{2}(n^{e,1}-n^{h,1})\right].
\end{equation}
Considering the simplest case with a symmetric junction $l=0$ without
barriers at the NS-interfaces. The injection current density is given
by straightforward calculation
\begin{eqnarray}
&&i_1^{-,1}-i_1^{-,4}=4\frac{e}{h}\left\{D_{\perp}+\epsilon+\frac{\epsilon}{Z_f
}\left[\left([1-2\epsilon]\cos2\theta-D\cos\phi \right. \right. \right.
\nonumber \\
&&\left. \left. \left. -R
\right)[R_{\perp}-D_{\perp}-2\mbox{Re}[d(d_{\perp}-r_{\perp})]\sin^2(\phi/2)]+\left(\sin^22\theta\nonumber \right. \right. \right. \\
&&\left. \left. \left. -1+2\epsilon+(D\cos\phi+R)\cos2\theta
\right)(R_{\perp}-D_{\perp})\right]\right\}.
\end{eqnarray}
Some general comments can be made about the conductance
$G=dI_{inj}/dV$. When the vertical and horisontal leads are decoupled
($\epsilon \rightarrow 0$), the conductance reduces to
$G=(e^2/h)D_{\perp}$, the conductance of the normal vertical
channel. For finite coupling, an additional term is added to the
conductance $\delta G\sim\epsilon$. This additional term $\delta G$ is
dependent on the phase difference $\phi$, but it is also, unlike for
the three terminal junction, dependent on the scattering amplitudes
$r,d,r_{\perp}$ and $d_{\perp}$, i.e not only the scattering
probabilities $R,D,R_{\perp}$ and $D_{\perp}$. This becomes clear when
we note that we can rewrite the expression
$\mbox{Re}[d(d_{\perp}-r_{\perp})]=1/2[R_{\perp}-D_{\perp}+([R_{\perp}-D_{\perp}][R-D]-4\mbox{Im}[r_{\perp}d_{\perp}^*]\mbox{Im}[rd^*])/(1-4\epsilon)]$,
i.e dependent on $\sigma \sigma_{\perp}$, just like the anomalous
current. This contribution can be explained qualitatively by
interference between quasiparticle paths where one path describes
scattering in the vertical direction and the other one in the
horisontal direction, thereafter leaving the junction.
In the case of zero temperature and voltage, the conductance becomes
\begin{eqnarray}
G&=&\frac{2e^2}{h}\left[D_{\perp}\right. \nonumber \\ && \left.
+\epsilon\left(1+\frac{D_{\perp}-R_{\perp}+\sin^2(\phi/2)\mbox{Re}[d(d_{\perp}-r
_{\perp})]}{R+D\cos^2(\phi/2)}\right)\right]
\end{eqnarray}
which is independent of the length $L$ of the junction. The
conductance at zero phase difference given by
$G(\phi=0)=(2e^2/h)[D_{\perp}/(1-2\epsilon)]$. From this
value, the conductance then increases or decreases, depending on the
phases of the scattering amplitudes, monotonically with
$\phi\rightarrow\pi$, as is seen in Fig. \ref{fourtermcond}.
\begin{figure}
\centerline{\psfig{figure=fourcond.eps,height=6.5cm}}
\caption{The zero voltage, zero temperature four terminal conductance
$G$ as a function of phase difference $\phi$ for different horisontal
transparency $D=0.2$ (dashed), $D=0.5$ (solid) and $D=0.8$
(dotted). The product $\sigma \sigma_{\perp}=+1$ for increase of
conductance due to finite phase difference, $\sigma \sigma_{\perp}=-1$
for decrease. The vertical transparency is $D_{\perp}=0.5$
and $\epsilon=0.05$.}
\label{fourtermcond}
\end{figure}
For finite energies, the injected quasiparticles tunnel into the
Andreev resonances and there is a finite bias anomaly of the
conductance. However, there is not always a peak in the
conductance. This is made clear by considering the conductance at
finite voltage, zero temperature and zero phase difference (to avoid
dependence on $\sigma \sigma_{\perp}$), given by
\begin{equation}
G=\frac{2e^2}{h} \left(D_{\perp}+\epsilon\left[1+\frac{(R_{\perp}-D_{\perp})[\cos^2\theta-(1-2\epsilon)]}{(1-2\epsilon)^2\sin^2 \theta+4\epsilon^2\cos^2\theta} \right]\right).
\end{equation}
In the weak coupling limit, $\epsilon \ll 1$, for
$R_{\perp}>D_{\perp}$, there is an increase in the conductance for
finite energies, just as for the three terminal device. For
$R_{\perp}<D_{\perp}$, i.e for highly transmissive junctions in the
vertical direction, however, there is a {\em decrease} in conductance
for finite energies. The presence of Andreev resonances thus decreases
the probability of a quasiparticle to be transmitted through the
junction.\cite{Morpurgo97b}
Other properties of the conductance are similar to the three terminal
junction, also taking the scatting phases into account (via $\sigma$
and $\sigma_{\perp}$). The conductance is periodic in voltage with a
period $\pi \hbar v_F/L$, the distance between the pairs of Andreev
resonances, for temperatures lower than this inter pair distance $kT\ll
\hbar v_F/L$. For such low temperatures the phase dependence of the
conductance is also dependent on the voltage. For high temperatures
$kT \gg \hbar v_F/L$, all the features of the individual Andreev
levels are washed out and the conductance becomes independent on
junction length and applied voltage $eV<\Delta$. For $eV>\Delta$, the
conductance is equal to the normal conductance of the junction
$G_N=(2e^2/h)(D_{\perp}+2\epsilon)$.
\section{Conclusions}
We have analyzed the equilibrium and nonequilibrium Josephson currents
and conductance in a ballistic, multiterminal, single mode SNS
junction. The nonequilibrium is created by means of quasiparticle
injection from a normal reservoir connected to the normal part of the
junction. By applying a voltage $V$ to the normal reservoir, up to the
superconducting gap $\Delta$, the equilibrium current of a short
junction $L\ll\xi_0$ can be suppressed. When the junction is longer
$L\geq \xi_0$, the direction of the Josephson current changes sign
as a function of applied voltage. For a junction longer then the
thermal length $L \gg \xi_T$, the {\em equilibrium} Josephson current
is exponentially small. The {\em nonequilibrium} Josephson current in
this regime is dominated by the {\em anomalous current}, arising from
the modification of the current carrying Andreev states due to
coupling to the normal reservoir. This anomalous current scales
linearly with applied voltage and saturates at a magnitude of the
order of the equilibrium current carried by a short junction, $I\sim
e\Delta/\hbar$.
The conductance oscillates as a function of the phase difference
$\phi$ between the superconductors, with a period of $2\pi$ in a
symmetric junction. The position of the conductance minima, $\phi=0$
or $\pi$, is dependent both on applied voltage and temperature. The
conductance exhibits a finite bias anomaly, at $eV \sim \hbar v_F/L$,
the position of the first current carrying Andreev level.
Asymmetric injection gives rise to oscillations of all currents on the
scale of $eV\sim \hbar v_F/l$ where $l$ is the length difference
between the two leads. At temperatures above this energy, these
oscillations are smeared and a we get renormalized anomalous and
injection currents that are $\pi$-periodic.
Introduction of barriers at the NS-interfaces give a strong length
dependence of all currents, governed by the Breit Wigner resonances between
the normal barriers. There are resonant current peaks at lengths where
the normal electron and hole resonances cross.
Connecting a second normal reservoir to the normal part of the
junction allows a clear separation between the injection current,
flowing between the two normal reservoirs, and the Josephson current,
flowing between the superconductors.
\acknowledgements
This work has been supported by research grants from NFR, TFR, NUTEK
(Sweden) and by a NEDO International Joint Research Grant (Japan).
|
\section{Introduction}
An investigation of possibility that dimension of our world is more then
four is not new. Nearly all papers on this direction are done with the
framework of standard Kaluza-Klein models where extra dimensions are curled
up to an unobservable size (see for example review \cite{OW}). Besides of
obvious achievements this approach encounters some problems such as: Why
four dimensions are extended and others are curled; How to choose the
signature of multidimensional space; Physical meaning of extra components of
Einstein's equations is unclear; There exists the problem of stability.
An alternative proposal that the extra dimensions are extended and the
matter is trapped in 4-dimensional submanyfold was advanced in papers
\cite{RS,V,S,BK}. This approach has properties similar to four dimensions - all
dimensions are extended and equal at the beginning and the signature has the
form (+,-,-, ... ,-).
Models of this kind also do not contradict to present time experiments
\cite{OW}. Multidimensionality in these models was used to solve several
problems, such as, cosmological constant, dark matter, non-locality or
hierarchy problems \cite{RS,V,S,KH,G}.
For the simplicity, we investigate here only the case of five dimensions.
The general procedure immediately generalizes to arbitrary dimensionality.
Using extended dimensions approach we want to consider our Universe as a
3-shell expanding in 5-dimensional space-time \cite{G}. This model supported
by at least two observed facts. First is the isotropic runaway of galaxies,
which only for obviousness usually explained as an expansion of a 3-sphere
in five dimensions. Second is the existence of a preferred frame in the
Universe where the relict radiation is isotropic. In the framework of the
closed-Universe model without boundaries this can also be explained if
Universe is a bubble and the mean velocity of the background radiation is
zero with respect to its center in the fifth dimension.
\section{Stability Condition}
\setcounter{equation}{0}
In models of large extra dimensions we need the mechanism of confining a
matter inside of the 4-dimensional submanyfold which must be sufficiently
narrow along the extra dimensions and flat along four others. It is natural
to think that such a splitting of 5-dimensional space with trapping of a
matter into four dimensions is the result of existence of the special
solution of multidimensional Einstein equations
\begin{equation} \label{1.1}
^5R_{AB}-\frac{1}{2}g_{AB}~^5R = 6\pi ^2GT_{AB}~~.
\end{equation}
Here $G$ is 5-dimensional gravitational constant and $A,B,...=0,1,2,3,5$.
We need stabile macroscopic solution, so it is natural to consider only
classical gravitational, electromagnetic and scalar fields which can form
extended solutions. For the beginning let us consider only gravitational and
electromagnetic fields with the Lagrangian
\begin{equation} \label{1.3}
L= - \sqrt{g}\left[ \frac{1}{12\pi ^2G}~^5R +
\frac{1}{4}F_{AB}F^{AB}\right] ~~.
\end{equation}
Generallisation for any Yang-Mills field is obvios.
To obtain stabile splitting of multidimensional space momentum toward the
extra - fifth dimensions must be zero
\begin{equation} \label{1.4}
P_5=\int T_5^5dtdV+\int T_5^\alpha dS_\alpha =0
\end{equation}
(Greek indices $\alpha ,\beta ...=0,1,2,3$ numerate coordinates in four
dimensions). Other dimensions can expand and our world can be expanding
bubble.
For our choice of gravitational Lagrangian the energy-momentum tensor of
gravitation and electromagnetic fields in five dimensions $T_{AB}$ has the
form of so named Lorentz energy-momentum complex
\begin{equation} \label{1.5}
T_A^B=t_A^B+\tau_A^B=\frac{1}{12\pi^2G\sqrt{g}}\partial_CX_A^{BC}~~,
\end{equation}
where
\begin{equation} \label{1.6}
X_A^{BC}=-X_A^{CB}=\sqrt{g}[g^{BD}g^{CE}(\partial _Dg_{AE}-
\partial _Eg_{AD})]~~.
\end{equation}
In (\ref{1.5})
\begin{eqnarray}
t_{A}^{B} = \frac{1}{12\pi^2G}(g^{BD}\partial_{A}\Gamma^{E}_{DE} -
g^{ED}\partial_{A}\Gamma^{B}_{DE} + \delta^{B}_{A}~^5R)~~, \nonumber \\
\tau_{A}^{B} = - F^{BC}F_{AC} + \frac{1}{4}\delta^{B}_{A} F^{DE}F_{DE}
\label{1.7}
\end{eqnarray}
are energy-momentum tensor of gravitational and electromagnetic fields
respectively.
To obey the stability condition (\ref{1.4}) for the solutions we must have
\begin{equation} \label{1.8}
T_5^\alpha =T_5^5=0~~.
\end{equation}
Using (\ref{1.6}) from (\ref{1.8}) we obtain
\begin{equation} \label{1.9}
\partial_\beta g_{5A}=\partial_5g_{5\alpha }=0~~.
\end{equation}
Simple solution of (\ref{1.9}) is
\begin{equation} \label{1.10}
g_{5\alpha }=0~~,~~~g_{55}=const=-1~~.
\end{equation}
In general $g_{55}$ can be any function of $x^5$, but this function in all
formulae will appear as a coefficient and would not influence on the
dynamic of the theory.
From the other hand from first condition of (\ref{1.8}) using explicit form
(\ref{1.7}) we obtain
\begin{equation} \label{1.11}
\partial_5\Gamma_{\beta \gamma }^\alpha =0~~,~~~F_{5\alpha }=0
\end{equation}
and equation of electromagnetic field has standard Maxwell 4-dimensional
form
\begin{equation} \label{1.12}
D_\nu F^{\mu \nu }=0~~.
\end{equation}
Here we want to notice that in model of Visser \cite{V} electromagnetic
field don't obey the condition (\ref{1.11}) ($F_{5\alpha }$ not equal to
zero) and therefore his solution is unstable.
Finally from (\ref{1.10}) and first of (\ref{1.11}) we obtain the metric
tensor corresponding to stable splitting of multidimensional space-time
\begin{equation} \label{1.13}
g_{\alpha \beta }=\lambda ^2(x^5)\eta_{\alpha \beta }~~,
~~~g_{55}=-1~~,~~~g_{5\beta }=0~~,
\end{equation}
where $\eta_{\alpha \beta }$ is the 4-dimensional metric tensor and
$\lambda ^2(x^5)$ in the meanwhile is the arbitrary function of fifth
coordinate. This solution which we received from stability conditions
exactly coincide with the anzats of Rubakov-Shaposhnikov \cite{RS}.
Splitting (\ref{1.13}) was made in the frame of the 4-dimensional wall. If
we consider our Universe as an expanding bubble in the frame of the bubble
center (\ref{1.13}) needs Lorentz transformation. Bubble expansion $\lambda
(x^5)\rightarrow \lambda (x^5-vt)$, where $v$ is velocity of the wall, means
conformal transformation of 4-dimensional metric $\eta_{\alpha \beta }$. To
keep stability, the theory must be invariant under conformal transformations
in the submanyfold. This condition fixes dimension of our world. Indeed
using formulae (\ref{1.12}) and (\ref{1.13}) Lagrangian of electromagnetic
field in any dimensions can be written in the form
\begin{equation} \label{1.14}
L=\sqrt{\lambda ^{2n}\eta }\frac {1}{4\lambda ^4}\eta^{\alpha
\gamma }\eta^{\beta \delta }F_{\alpha \beta }F_{\gamma \delta }~~,
\end{equation}
where $n$ is dimension of submanyfold of trapping. Only in case of four
dimensions ($n=4)$ we have conformal invariance and stabile splitting is
possible. Thus only 3-dimensional expanding bubbles can be survived for a
long time and our Universe can be one of them.
\section{Trapping}
\setcounter{equation}{0}
In previous section it was shown that in Gausian normal coordinates the
5-dimensional metric of our Universe can be written in the form
\begin{equation} \label{2.1}
ds^2=-(dx^5)^2+\lambda ^2(x^5) \eta_{\alpha \beta }dx^\alpha dx^\beta ~~.
\end{equation}
In this coordinates components of Christoffel symbols with two or three
indices $5$ are equal to zero, while with the one index $5$ forms the tensor
of extrinsic curvature \cite{MWT}
\begin{eqnarray}
\nonumber K_{\alpha\beta} = \Gamma^{5}_{\alpha\beta } =
\frac{1}{2}\partial_{5}g_{\alpha\beta } =
\lambda\lambda^{^{\prime}}\eta_{\alpha\beta }~~, \\
\label{2.2}K^{\alpha\beta} = - \frac{1}{2}\partial_{5}g^{\alpha\beta } ~~,
\end{eqnarray}
where prime denotes derivative with respect of the coordinate $x^5$. Also we
want to represent some useful formulae
\begin{eqnarray} \label{2.3}
g^{\alpha\gamma}K_{\gamma\beta} = \Gamma^{\alpha}_{5\beta } =
\lambda\lambda^{^{\prime}}\delta^{\alpha}_{\beta }~~, \nonumber \\
K = g^{\alpha\beta}K_{\alpha\beta} = g_{\alpha\beta}K^{\alpha\beta} =
4\lambda^{^{\prime}}/\lambda~~, \\
\partial_5 K= g^{\alpha\beta}\partial_5 K_{\alpha\beta} -
2K^{\alpha\beta}K_{\alpha\beta}~~. \nonumber
\end{eqnarray}
Any vector and tensor naturally is split-up into its components orthogonal
and tangential to the shell. Using decomposition of the curvature tensor
\begin{eqnarray} \label{2.4}
^5R_{\alpha\beta} = R_{\alpha\beta } + \partial_5 K_{\alpha\beta } -
2K_{\alpha}^{\gamma} K_{\gamma\beta } + K K_{\alpha\beta }~~, \nonumber \\
^5R_{55} = - \partial_5 K - K^{\alpha\beta}K_{\alpha\beta}~~, \\
^5R = R + K^{\alpha\beta}K_{\alpha\beta} + K^2 + 2\partial_5 K \nonumber
\end{eqnarray}
one can find decomposition of Einstein's equations
\begin{eqnarray} \label{2.5}
R_{\alpha\beta } - \frac{1}{2}\eta_{\alpha\beta}R + \partial_5 K_{\alpha\beta }
- 2K_{\alpha}^{\gamma} K_{\gamma\beta } + K K_{\alpha\beta } - \frac{
\lambda^2}{2}\eta_{\alpha\beta }( K^{\gamma\delta}K_{\gamma\delta} + K^2 +
2\partial_5 K) = \nonumber \\
= \frac{6\pi ^2G}{\lambda^2}(- F^{\delta}_{\alpha}F_{\delta\beta} +
\frac{1}{4}\eta_{\alpha \beta } F^{\gamma\delta}F_{\gamma\delta} )~~ , \\
\frac{1}{2\lambda^2}R + \frac{1}{2}(K^2 - K^{\gamma\delta}K_{\gamma\delta})
= - \frac{3\pi ^2G }{2\lambda^4} F^{\alpha\beta }F_{\alpha\beta }~~ .
\nonumber
\end{eqnarray}
Using last formula of (\ref{2.3}) we noticed that since
\begin{equation}
t_{5}^{5} = \frac{1}{12\pi ^2G}(- \partial_{5}K -
g^{\alpha\beta}\partial_{5}K_{\alpha\beta} + ~^5R)~~,
\label{2.51}
\end{equation}
fifteenth Einstein's equation (last equation of (\ref{2.5})) is nothing else
than stability condition (\ref{1.8}) - $T_5^5=0$. So extra
component of Einstein's equations, whose physical meaning is unclear in
standard Kaluza-Klein models, in our approach coincide with the condition of
stability.
Using the explicit form of extrinsic curvature tensor (\ref{2.3}) from the
system of Einstein's equations (\ref{2.5}) we find
\begin{equation}
-R=12\lambda \lambda ^{"} ~~. \label{2.6}
\end{equation}
Then from second equation of (\ref{2.5}) for function $\lambda (x^{5})$ we
have
\begin{equation}
\lambda \lambda ^{"}- \lambda ^{^{\prime }2}-\frac{\pi ^{2}G}{4}F^{\gamma
\delta }F_{\gamma \delta } =0~~. \label{2.7}
\end{equation}
This equation gives trapping solution
\begin{equation} \label{2.8}
\lambda =cosh(Ex^5)~~,
\end{equation}
where
\begin{equation} \label{2.9}
E=\sqrt{\frac{\pi ^2G}4F^{\gamma \delta }F_{\gamma \delta }}~~.
\end{equation}
Width of our world $\Delta \sim 1/E$ depended on gravitational constant and
density of electromagnetic field.
To see how gravitational trapping works let us consider simple example of
the real scalar field $\phi $ in the background metric (\ref{1.13})
with $\lambda $ expressed by (\ref{2.8}). If we put
\begin{equation} \label{2.10}
\phi =\lambda u(x^\alpha )
\end{equation}
to the 5-dimensional Klein-Gordon equation
\begin{equation} \label{2.11}
\frac 1{\sqrt{g}}\partial _A(\sqrt{g}g^{AB}\partial _B\phi )+m^2\phi =0~~,
\end{equation}
for $u$ we find
\begin{equation} \label{2.12}
\eta^{\alpha \beta }\partial _\alpha \partial _\beta u+
\frac{m^2\lambda ^4-1}{\lambda ^2}u=0~~.
\end{equation}
According to (\ref{2.8}) - $\lambda =cosh(Ex^5)$ and we see that ''mass'' of
the scalar field in this equation has its minimum in the 4-dimensional
submanyfold and growth rapidly far from the wall. So field $\phi $ is in the
potential well.
\section{Scalar fields}
\setcounter{equation}{0}
Now let us consider more complicate model adding to (\ref{1.3}) the
Lagrangian of complex scalar fields
\begin{equation} \label{3.1}
L_\psi =\sqrt{g}[D_A\bar \psi D^A\psi -\xi |\psi |^2~^5R-U(|\psi |)]~~,
\end{equation}
where $U(|\psi |)$ in the meanwhile is any function of $\bar \psi \psi $.
Variation of this Lagrangian by metric tensor gives energy-momentum tensor
of scalar fields
\begin{equation} \label{3.2}
T_{AB}=t_{AB}+2\xi (~^5R_{AB}-\frac 12g_{AB}~^5R+D_AD_B-g_{AB}D_CD^C)|\psi |^2~~,
\end{equation}
where
\begin{equation} \label{3.3}
t_{AB}=D_A\bar \psi D_B\psi +D_B\bar \psi D_A\psi -g_{AB}[D_C\bar \psi
D^C\psi -U(|\psi |)]~~.
\end{equation}
Stability condition for the shell (\ref{1.8}) - $T_{5\alpha }=0$, now except
of conditions $~^5R_{5\alpha }=F_{5\alpha }=0$ gives the new condition for
scalar fields
\begin{equation} \label{3.4}
D_5\psi =(\partial _5-iA_5)\psi =0~~.
\end{equation}
Maxwell equations now take the standard 4-dimensional form with the source
\begin{equation} \label{3.5}
D_\mu F^{\mu\nu }=j^\nu =\bar \psi \partial ^\nu \psi -\psi \partial ^\nu
\bar \psi ~~,
\end{equation}
and the equation for the scalar field is
\begin{equation} \label{3.6}
(\eta^{\mu \nu }\partial _\mu \partial _\nu +\xi ~^5R)\psi +
\lambda ^2\frac{\partial U(|\psi |)}{\partial \bar \psi }~=0~~.
\end{equation}
Using formulae (\ref{2.3}) for the extrinsic curvature tensor $K_{\mu \nu }$,
splitting of Einstein's equations (\ref{2.5}) now has the form
\begin{eqnarray}
\left( \frac{1}{6\pi ^2G } + 2\xi|\psi|^2\right)(R_{\alpha\beta } -
\frac{1}{2}\eta_{\alpha \beta }R) +
\frac{1}{2\pi ^2G } \eta_{\alpha \beta } \frac{\lambda^{"}}{\lambda} = \nonumber \\
= \frac{1}{\lambda^2}(- F^{\delta}_{\alpha}F_{\delta\beta} +
\frac{1}{4} \eta_{\alpha\beta} F^{\gamma\delta }F_{\gamma\delta} ) +
D_{\alpha}\bar{\psi}D_{\beta}\psi + D_{\beta}\bar{\psi}D_{\alpha}\psi - \nonumber \\
\label{3.7}- \eta_{\alpha\beta}[\eta^{\mu\nu}D_{\mu}\bar{\psi}D_{\nu}\psi -
\lambda^2U(|\psi|)] + 2\xi(D_{\alpha}D_{\beta} -
\eta_{\alpha\beta}D_{\mu}D^{\mu}) |\psi|^2 ~~, \\
\left( \frac{1}{6\pi ^2G } + 2\xi|\psi|^2\right)\frac{1}{2\lambda^2}R +
\frac{1}{\pi ^2G }\frac{(\lambda^{^{\prime}})^2}{\lambda^2} = \nonumber \\
\label{3.8}= - \frac{1}{4\lambda^4} F^{\alpha\beta }F_{\alpha\beta } +
\frac{1}{\lambda^2}\eta ^{\mu\nu}D_{\mu}\bar{\psi}D_{\nu}\psi - U(|\psi|) -
\frac{2\xi}{\lambda^2}D_{\mu}D^{\mu}|\psi|^2 ~~.
\end{eqnarray}
Substituting of (\ref{3.6}) and (\ref{3.8}) in trace of equation (\ref{3.7})
one can find that solution of the system (\ref{3.6}) - (\ref{3.8}) is
\begin{eqnarray} \label{3.9}
6\xi = 1~~, \nonumber \\
R = -12\lambda \lambda^{"} ~~, \\
\lambda = cosh(Ex^5)~~ . \nonumber
\end{eqnarray}
We see that constant of coupling gravitational and scalar fields $\xi $
fixed on the value corresponding to conformal invariance of scalar field
equation in four dimensions. So again we received 4-dimensionality from the
condition of stability towards the extra dimensions. Only conformal
invariant form of function $U(|\psi |)$ in four dimensions, when $\psi
=\lambda (x^5)/u(x^\nu )$, is
\begin{equation} \label{3.10}
U(|\psi |)=\mu|\psi |^4/2~~,
\end{equation}
where $\mu$ is coupling constant. Finally equation of massless scalar field
in five dimensions (\ref{3.6}) has the form
\begin{equation} \label{3.11}
(\eta ^{\mu \nu }D_\mu D_\nu +2E^2+\mu |\psi |^2)\psi =0~~.
\end{equation}
We see that because of coupling with gravitational field in four dimensions
scalar field has "mass" $E^2$ expressed with gravitational constant and
density of electromagnetic field by (\ref{2.9}).
|
\section{Introduction.}
\vspace{1cm}
It is well known that dynamical symmetry breaking (DSB) presents an attractive alternative to the Higgs mechanism
of electroweak symmetry breaking in Standard Model and in a natural way solves the so called hierarchy problem
connected with
the quadratic divergence of the mass of the Higgs boson (for a general introduction to DSB see \cite{book}).
Morever, DSB allows at least in principle
to deduce all relevant parameters of symmetry breaking. However, usually DSB requires strong coupling
($\alpha_c \ge 1$)
that essentially restricts the choice of models that can be used and it makes also
quantative studying of DSB a difficult problem. Physically, it is easy to understand why
$\alpha_c \ge 1$. For the state with the condensate of fermion-antifermion pairs to have lower energy than
the trivial
vacuum, it is necessary that energy of the corresponding
fermion-antifermion bound state be negative. Then, for example, in QED, in view of the uncertainty principle,
it imples $\alpha_c \ge 1$. Therefore, it is very interesting to consider situations when DSB takes place
in the regime of weak coupling ($\alpha_c \approx 0$).
Two examples of DSB in the regime of weak coupling are known. The first is symmetry breaking in the presence
of the Fermi surface (i.e. chemical potential is nonzero).
In this case, as well known from the Bardeen--Cooper--Schrieffer theory of superconductivity
\cite{BCS}, a bound state forms for any (however small) attraction between fermions. The effective field theory
description of this phenomenon
based on the renormalization group was developed in \cite{Pol}, where it was shown that renormalization group
scaling takes place only in the direction perpendicular to the Fermi surface, therefore, from the viewpoint of
renormalization
group scaling the effective dimension of spacetime is 1 + 1. Since in two-dimensional
spacetime a bound state forms even in the case of arbitrary small attraction, we obtain that $\alpha_c = 0$ in this case.
(Note that this is one of the key ideas of QCD color superconductivity at finite density in the regime of weak coupling,
which has been actively studied in recent years \cite{Alf, Rap}).
The other example of DSB in the regime of weak coupling is DSB in external constant magnetic field. This phenomenon was
discovered in \cite{GMSh1,GMSh2}, where it was shown that chiral symmetry is dynamically broken
in the Nambu---Jona-Lasinio (NJL) model \cite{NJL} and QED
in external constant magnetic field $B$ for an arbitrary weak interaction,
i.e. the critical coupling constant is zero in
this case\footnote{The case of the discrete three-dimensional NJL model was considered in \cite{Kri} and the effect of
enhancement of the chiral condensate in supercritical ($g > g_c > 0$) phase of the four-dimensional NJL model
was studied in \cite{Kle}.}.
The essence of this strong magnetic catalysis \cite{GMSh1, GMSh2} is that electrons are
effectively (1 + 1)-dimensional when their energy is much less than the Landau gap $\sqrt{|eB|}$. The lowest Landau level
plays here the role similar to that of the Fermi surface in the BCS theory of superconductivity, leading to dimensional
reduction in dynamics of fermion pairing.
Recently another example of DSB in the regime of weak coupling was discovered. By using the NJL-type models it was
shown \cite{BK, IMO} that critical coupling constant is zero in spaces with constant negative curvature
(for an excellent review of DSB in curved
spacetime see \cite{Ina}), i.e. chiral symmetry is always broken for any $g > 0$ (note that the fact of impossibility
of keeping chiral invariance for free
massless fermions in spaces with negative constant curvature was also noted in \cite{Wil}).
The physical explanation of this very interesting fact is lacking. The authors of these works
calculated the effective potential for an order parameter and then showed that it has a nontrivial minimum for any $g > 0$.
To find an explanation of this result in more physical terms was one of the main motivations of the present work.
\vspace{1cm}
\section{The model.}
\vspace{1cm}
For our aims it is enough to consider the NJL model
in curved spacetime
\begin{equation}
S = \int\! \sqrt{-g} d^{4}x \left[\sum_{k=1}^{N}
\bar{\psi_k}i\gamma^{\mu}\nabla_{\mu}\psi_k
+ \frac{G}{2N} \left( (\sum_{k=1}^{N}\bar{\psi_k}\psi_k )^{2} -
(\sum_{k=1}^{N}\bar{\psi_k}\gamma_5 \psi_k)^{2} \right)
\right],
\end{equation}
where N is the number of flavors,
$g = \mbox{det} (g_{\mu\nu})$ the determinant of metric, $\nabla_{\mu} = \partial_{\mu} + i \omega^{ab}_{\mu}
\sigma_{ab}$ the covariant derivative with spin connection $\omega^{ab}_{\mu}$, and $\gamma^{\mu}$ matrices in curved
spacetime are expressed through the Dirac $\gamma^a$ matrices in flat spacetime with the help of vierbeins
$\gamma^{\mu} = V^{\mu}_{a}\gamma^{a}$.
The action (1) is invariant with respect to chiral transformations
$\psi \rightarrow e^{i\gamma_5\beta}\psi$.
For practical calculations in four-fermion theories
it is convenient to use the so called auxiliary
field method \cite{Hub, Stra}, where
Lagrangian (1) is represented in the equivalent form
\begin{equation}
L= \sum_{k=1}^{N}\left(i\bar{\psi_k} \gamma^\mu \nabla_\mu\psi_k + \bar{\psi_k}(\sigma + i\gamma_5\pi)\psi_k\right)
- \frac{N}{2G}\sigma^2,
\end{equation}
where $\sigma$ and $\pi$ are auxiliary fields. If we integrate over $\sigma$ and $\pi$, we obtain the initial
action (1). If the field $\sigma$ acquires a nonzero vacuum expectation value, then obviously fermions acquire
mass and chiral symmetry is broken.
To find the effective action for the fields $\sigma$ and $\pi$, we integrate over the fermion fields. We obtain
(without loss of generality one can set $\pi = 0$ because it is always possible to restore the dependence on $\pi$
by requiring chiral symmetry of the effective action)
\begin{eqnarray}
\Gamma(\sigma_c) = -N\int \sqrt{-g}d^{4}x
\frac{\sigma_c^{2}}{2G}
-i\mbox{Ln Det}(i\gamma^{\mu}\nabla_{\mu}-\sigma_{c}),
\end{eqnarray}
where $\sigma_{c}(x)=<0|\sigma|0>$.
The effective potential $V(\sigma_c)$ (we set $\sigma_c(x)$ = const) is given by the expression
\begin{equation}
V(\sigma_c)=-\frac{\Gamma(\sigma_c)}{\int\sqrt{-g}d^{4}x}.
\end{equation}
Furthermore,
\begin{eqnarray}
\mbox{Ln Det}(i\gamma^{\mu}\nabla_{\mu}-\sigma_c) = \mbox{TrLn}(i\gamma^{\mu}\nabla_{\mu}-\sigma_c) = \nonumber \\
\frac{1}{2} \mbox{TrLn}(i\gamma^{\mu}\nabla_{\mu}-\sigma_c) +
\frac{1}{2} \mbox{TrLn}(\gamma_5(i\gamma^{\mu}\nabla_{\mu}-\sigma_c)\gamma_5) = \nonumber \\
\frac{1}{2} \mbox{TrLn}\left((i\gamma^{\mu}\nabla_{\mu}-\sigma_c)(-i\gamma^{\nu}\nabla_{\nu}-\sigma_c) \right).
\end{eqnarray}
By using the Schwinger proper time method \cite{Sch}, we get the effective potential (we also perform the Wick rotation)
\begin{equation}
V(\sigma_c) = N\,\left(\frac{\sigma_c^{2}}{2G} + \frac{1}{2} \int_{\frac{1}{\Lambda^2}}^{\infty}\frac{ds}{s}\, \mbox{tr}
\,<x|e^{-sH}|x>\right),
\end{equation}
where $H = -(\gamma_E^{\mu}\nabla_{\mu})^2 + \sigma_c^2$ ($\gamma_E^{\mu}$ are Euclidean $\gamma$-matrices).
Consequently, the gap equation ($\frac{dV}{d\sigma_c}|_{\sigma_c=m} = 0$) is
\begin{equation}
1\,\, = \,\,G\int_{\frac{1}{\Lambda^2}}^{\infty}ds \, \mbox{tr} <x|e^{-sH}|x>.
\end{equation}
Thus, we need to find the diagonal heat kernel $\mbox{tr} <x|e^{-sH}|x>$ in spaces with
constant negative curvature. Before doing it we
first describe what these spaces are (for a very good introduction see \cite{Bal}).
The D-dimensional Riemannian space of constant negative curvature $H^D$
(hyperbolic space) can be described as a hyperboloid
\begin{equation}
- x_0^2 + x_1^2 + x_2^2 + ... + x_D^2 = - a ^{2}
\end{equation}
embedded in (D+1)-dimensional Minkowski space with metric $ds^2=-dx_0^2 + dx_1^2 + ... + dx_D^2$.
It is easy to show that the Minskowski metric becomes positive definite on the surface given by Eq.(8)
(this is the reason why we have
chosen the Minkowski metric in the form (--, +, +,..., +)). Obviously
by construction hyperbolic space has the group of
isometry SO(1, D) and is a homogeneous space because any two points on $H^D$ can be connected by some isometry
(all points of this space are equivalent).
By using the parametrization
\begin{eqnarray}
x_0 = a\cosh \sigma, \,\,\,
x_1 = a \sinh \sigma \cos \theta_1, \nonumber \\
x_2 = a \sinh \sigma \sin \theta_1 \cos \theta_2, \,\,\,
x_3 = a \sinh \sigma \sin \theta_1 \sin \theta_2 ,\,.\,.\,.\,,
\end{eqnarray}
the line element $ds^2=-dx_0^2 + dx_1^2 + ... + dx_D^2$ becomes
\begin{equation}
ds^{2}=a^{2}(d\sigma^{2}+\sinh^{2}\sigma d\Omega_{D-1}),
\end{equation}
where $d\Omega_{D-1}$ is the metric on unit (D-1)-dimensional sphere and the curvature is equal to
\begin{equation}
R= - \frac{D(D-1)}{a^2}.
\end{equation}
Recall that for Euclidean space the linear element in spherical coordinates is
\begin{equation}
ds^{2}=(dr^{2}+r^{2}d\Omega_{D-1}).
\end{equation}
By comparing Eq.(10) and Eq.(12), we see that the difference between flat and hyperbolic space is that the volume of
sphere in hyperbolic space grows with radius $r$ as $a^{2}\sinh^{2}\frac{r}{a}$ instead $r^{2}$ as in flat space.
In the present work we actually consider the (D+1)-dimensional ultrastatic spacetime $R \times H^{D}$, where the components
of metric are time independent and the conditions $g_{00}=1$ and $g_{0i} = 0$ are true
in an appropriate system of coordinates (thus, time coordinate describes evolution of fields on $H^{D}$).
\vspace{1cm}
\section{The effective reduction.}
\vspace{0.8cm}
\subsection{Heat kernels.}
\vspace{1cm}
Since the metric on $R \times H^{D}$ is time independent, the heat kernel in Eq.(6) trivially factorizes and we are left
with the problem of calculation of the heat kernel \\ $\mbox{tr} \, <x|e^{-sH}|x>$ on the hyperbolic space $H^{D}$.
As we mentioned in Introduction it was shown that $g_{c}=0$ in spaces with constant negative
curvature. The authors of these works \cite{BK, IMO} calculated heat kernel in the closed form either using
the Schwinger method for
calculation of $<x|e^{-sH}|x>$ \cite{Sch} or expressing it through the spinor Green function, which was obtained as a
solution of the corresponding differential equation in \cite{Campo}. Heat kernel is in a certain sense
an integral characteristic. To
reveal the underlying dynamics which gives $g_{c}=0$, we need more detailed information about the system. For this we
calculate heat kernel by summing over the eigenfunctions of the Dirac operator on $H^D$ that allows us to investigate
what dynamics is responsible for $g_{c} = 0$ in spaces with negative curvature\footnote{We would like to thank
V.P. Gusynin for suggesting this approach.}.
To calculate heat kernel in the form of sum over eigenfunctions, we use the method and the results of \cite{Byts}.
To illustrate the method, we first calculate the heat kernel for scalar field $h_{scalar} =
<x|e^{-\frac{s}{a^2}A}|x>$,
where $A = -\Delta + m^2a^2$ and $-\Delta$ is
the Laplace--Beltrami operator on $H^D$ (we will
use this heat kernel when we discuss the role of Goldstone bosons), which is given by
\begin{equation}
\Delta=\frac{\partial^{2}}{\partial \sigma^{2}}+(D-1)
\coth\sigma\frac{\partial}{\partial\sigma}+(\sinh\sigma)^{-2}\Delta_{S^{D-1}}
\:,
\end{equation}
where the last term denotes the Laplace--Beltrami operator
on the unit sphere $S^{D-1}$. If the eigenfunctions of the operator (13) are known, then
we can insert their complete set in the martix element for the heat kernel. Then the heat kernel is represented
in the form of sum over eigenfunctions
\begin{equation}
h_{scalar} = <x|e^{-\frac{s}{a^2}(-\Delta + m^2a^2)}|x> = \sum_{\lambda}
e^{-\frac{s}{a^2}(\lambda + m^2a^2)}|\phi_{\lambda}(x)|^2,
\end{equation}
where $\phi_{\lambda}$ are eigenfunctions
($-\Delta\phi_{\lambda}=\lambda\phi_{\lambda}$). It is obvious from Eq.(13) that the equation for eigenfunctions
admits the separation of variables,
therefore, we seek them in the form
$\phi=f_{\lambda}(\sigma)Y_{lm}$, where $Y_{lm}$ are the spherical
harmonics on $S^{D-1}$
\begin{eqnarray}
\Delta_{S^{D-1}}Y_{lm}=-l(l+D-2)Y_{lm} \:.\nonumber
\end{eqnarray}
Thus, the radial wave functions satisfy the ordinary differential
equation
\begin{equation} f_{\lambda}^{''}+(D-1)\coth\sigma
f_{\lambda}^{\prime}
+\left[\lambda-\frac{l(l+D-2)^{2}}{\sinh^{2}\sigma}\right]f_{\lambda}=0
\:.\nonumber\end{equation}
The only bounded solutions of Eq.(15) are
\begin{equation}
f_{\lambda}(\sigma)= C \: \Gamma\left(\frac{D}{2}\right)
\left(\frac{\sinh\sigma}2\right)^{1-D/2} P_{-1/2+ir}^{\mu}(\cosh\sigma)
\:,
\end{equation}
where $P_{\nu}^{\mu}(z)$ are the associated
Legendre functions of the first kind
\cite{Grad},
$r=(\lambda-\rho_{D}^{2})^{1/2}$ is used as a label for the continuum
spectrum, $\rho_{D} = \frac{D-1}{2}$, and $C$ is
the normalization constant. The asymptotic behaviour of $P_{\nu}^{\mu}(z)$ for $\mid z\mid \gg 1$ is
\begin{equation}
P_{\nu}^{\mu}(z)\approx
\frac{2^{\nu}\Gamma(\nu+1/2)}{\pi^{1/2}\Gamma(\nu-\mu+1)}
z^{\nu}+\frac{\Gamma(-\nu-1/2)}{2^{\nu+1}\pi^{1/2}
\Gamma(-\nu-\mu)}z^{-\nu-1} \:,\nonumber
\end{equation}
from which we
obtain the asymptotic behavior of the eigenfunctions
\begin{equation}
f_{\lambda}(\sigma)\simeq C\frac{2^D\Gamma(D/2)\Gamma(ir)}{4\pi^{1/2}\Gamma(\rho_D+ir)}
e^{-\rho_D\sigma+ir\sigma}+h.c. \:.
\end{equation}
The radial functions are bounded at infinity provided
the parameter $r$ is real, which is equivalent to the condition
$\lambda\geq\rho_{D}^{2}$. Thus, the spectrum of the Laplace--Beltrami operator has a
gap which is determined by the curvature and depends on $D$.
Since $H^D$ is a homogeneous space (i.e. all points are equivalent), the heat kernel
does not depend on $x$ and one can use any point to calculate the heat kernel. In the spherical coordinates it is very
convenient to use the origin because as follows from the
explicit solutions (Eq.(16)) only modes with {\it l} $\: = 0$ are not equal to zero at this point.
We normalized eigenfunctions so that $f_{\lambda} (0) = C$ at $x=0$. The
invariant measure defining the scalar product between eigenfunctions is
\begin{equation}
(f_{\lambda},f_{\lambda'})=\Omega_{D-1}\int_{0}^{\infty}
f_{\lambda}^{*}f_{\lambda'}(\sinh\sigma)^{D-1}d\sigma\:,
\end{equation}
where $\Omega_{D-1}$ is the volume of the (D-1)-dimensional sphere and the factor $\sinh^{D-1}\sigma$ follows
from the square root of the determinant of metric. The normalization constant is determined from
the usual condition of normalization of eigenfunctions of continuous spectrum
\begin{equation}
(\phi_{\lambda},\phi_{\lambda^{\prime}})
=\delta(\lambda-\lambda^{\prime}).
\end{equation}
The easiest way to calculate this scalar product and determine the normalization constant for eigenfunctions
(16) is to use the
fact that the scalar product of two eigenfunctions
is expressed through the derivative of their Wronskian $W[\cdot,\cdot]$ at an arbitrary point. For us it is
most convenient to calculate the Wronskian at infinity because we know the asymptotic behavior of eigenfunctions there.
Thus, we obtain
\begin{equation}
(f_{\lambda},f_{\lambda^{\prime}})=
\frac{2^{D-1}\pi^{\frac{D}{2}}\Gamma(\frac{N}{2})}
{(r')^2 - r^2} \lim_{\sigma\rightarrow \infty}(1-\cosh^{2} \sigma)
W[P_{-\frac{1}{2} - ir}^{\mu}(\cosh \sigma),P_{-\frac{1}{2} + ir^{\prime}}^{\mu}(\cosh \sigma)]\:,
\end{equation}
where the limit is taken in the sense of distributions.
Thus, we find
\begin{equation}
|C(r)|^2=\frac{2}{(4\pi)^{D/2}\Gamma(D/2)}
\frac{\mid\Gamma(ir+l+\rho_D)\mid^{2}}{\mid \Gamma(ir)\mid^{2}} \:
\end{equation}
for the square of the normalization constant.
Since we normalized eigenfunctions as
$f_{\lambda}(0)=C$, the heat kernel is given by
\begin{equation}
h_{scalar} = \frac{1}{a^D}\int_{0}^{\infty}e^{-\frac{s}{a^2} (r^2 + m^2a^2)}|C^2(r)| dr,
\end{equation}
where we have made the change of variables $r=\sqrt{\lambda - \rho^2_D}$.
Note that $C(r)dr$ is the measure of eigenfunctions. It defines the number
of states per unit volume in the range $dr$.
We now consider the heat kernel for the Dirac operator on $R \times H^D$. Before doing it we first discuss
what we mean by chiral symmetry in spacetimes of arbitrary dimension (we consider again spacetimes whose metric
has Minkowski spacetime signature).
In Section 2, we described the chirally invariant NJL model in four-dimensional spacetime.
As well known chiral symmetry is connected with properties of representations of the Clifford
algebra (for a good description of spinors in n-dimensional spacetime see, e.g., \cite{Sohn}).
The Clifford algebra for n-dimensional spacetime of even dimension has only one
complex, irreducible representation in the $2^{n/2}$-dimensional
spinor space. These spinors are reducible with respect to the even
subalgebra (generated by products of an even number of Dirac matrices)
and split in a pair of
$2^{n/2-1}$-component irreducible Weyl spinors ($\gamma_{n+1} = \gamma_0 ... \gamma_{n-1}$ is
an analog of the $\gamma_5$ matrix in n-dimensional spacetime and $\frac{1 \pm \gamma_{n+1}}{2}$ are the corresponding
chiral projectors). In odd-dimensional spacetimes, there are two different representations of the Clifford algebra (they
differ by the sign of the $\gamma$-matrices) and chiral symmetry is not defined because $\gamma_{n+1}$ is proportional
to the unity. In order to define chiral symmetry in odd-dimensional spacetimes, it is the usual practice to assume that
fermion fields are in a reducible representation of the Clifford algebra so that we can define an analog of chiral
symmetry (for an explicit example in
(2 + 1)-dimensional spacetime see, e.g., \cite{App}). In what follows
we understand chiral symmetry in odd-dimensional spacetimes in this sense.
In the scalar case it is easy to
factorize the
part of heat kernel, which contains time derivatives. It is a little bit more elaborated for spinors because
the time derivative is multiplied by the $\gamma_0$-matrix.
By using (see Eq.(5))
\begin{eqnarray}
(i\gamma^{\mu}\nabla_{\mu}+m)(i\gamma^{\nu}\nabla_{\nu}-m) = (i\nabla_0 + i\vec{\gamma}\gamma_0\vec{\nabla}
+m\gamma_0)\gamma_0\gamma_0(i\nabla_0 + i\gamma_0\vec{\gamma}\vec{\nabla}-m\gamma_0) = \nonumber \\
(i\nabla_0 - i\vec{\alpha}\vec{\nabla}+m\gamma_0)(i\nabla_0 + i\vec{\alpha}\vec{\nabla}-m\gamma_0) =
(i\nabla_0)^2 - (-i\vec{\alpha}\vec{\nabla}+m\gamma_0)^2,
\end{eqnarray}
where $\vec{\alpha} = \gamma_0 \vec{\gamma}$, we get rid of the $\gamma_0$-matrix.
It is no wonder why such an operator
for the heat
kernel for spatial coordinates appears after the separation
of time derivatives. Indeed, the Dirac equation can be written in
the form of a Schr\"{o}dinger equation
\begin{eqnarray}
i\frac{\partial \psi}{\partial t} = H \psi \nonumber
\end{eqnarray}
with the Hamiltonian $H = -i\vec{\alpha}\vec{\nabla} + \beta m$,
where $\vec{\alpha}=\gamma_0\vec{\gamma}$ and $\beta = \gamma_0$.
Therefore, we immediately recognize our operator $(-i\vec{\alpha}\vec{\nabla}+m\gamma_0)^2$ as the square
of the Hamiltonian,
which is obviously a positive definite operator. Thus, the gap equation (see Eq.(7)) on $R \times H^D$ is
\begin{equation}
1\,\, = \,\,G\int_{\frac{1}{\Lambda^2}}^{\infty} ds \,\mbox{tr}
<t,x|e^{-s(-(\nabla_0)^2 + (-i\vec{\alpha}\vec{\nabla}+m\gamma_0)^2)}|t,x>,
\end{equation}
where we again performed the Wick rotation.
The contribution of time derivatives to the heat kernel is easy to be found.
Therefore, we are left with the problem of calculation of heat
kernel on $H^D$ for the operator $(-i\vec{\alpha}\vec{\nabla}+m\gamma_0)^2$.
To calculate the heat kernel for the operator $(-i\vec{\alpha}\vec{\nabla}+m\gamma_0)^2$ we use as in the scalar case
expansion in eigenfunctions.
The equation for eigenfunctions is
\begin{eqnarray}
(-i\vec{\alpha}\vec{\nabla} + m\gamma_0)\psi_{\lambda}=\lambda\psi_{\lambda} \nonumber\:.
\end{eqnarray}
The covariant derivative of a spinor field on $H^{D}$ can be
decomposed in a radial part plus the covariant derivative along the
unit $S^{(D-1)}$-sphere. Furthermore,
making the decomposition of $2^{\frac{D+1}{2}}$-dimensional representation in a Dirac-like
representation of $\gamma$-matrices, the equation for eigenfunctions takes the form of a
coupled system (for more details see \cite{Byts})
\begin{equation}
i\gamma_{1}\left(\partial_{\sigma}+\rho_D
\coth\sigma\right)\psi_{1}+\frac{1}{\sinh\sigma}i\not\!\nabla_s\psi_{1}
=-a(\lambda+m)\psi_{2} \:,
\end{equation}
\begin{equation}
i\gamma_{1}\left(\partial_{\sigma}+\rho_D
\coth\sigma\right)\psi_{2}+\frac{1}{\sinh\sigma}i\not\!\nabla_s\psi_{2}
=-a(\lambda-m)\psi_{1} \:,
\end{equation} where
$\psi_{1,2}$ are the $2^{\frac{D+1}{2}-1}$-components Weyl spinors and $i\not\!\nabla_s$ is the
Dirac operator on $S^{D-1}$. The spinors $\psi_{1,2}$ transform
irreducibly under $SO(D)$ so that we can put
$\psi_{1,2}=f_{1,2}(\sigma)\chi_{1,2}$, where $\chi_{1,2}$ are spinors
on $S^{D-1}$.
The eigenvalues of the Dirac operator on $S^{D-1}$ are known to be
$\kappa=\pm(l+\rho_D)$, $l=0,1,2,...$ \cite{Cand}.
The solutions of Eqs.(26) and (27) are given in terms
of hypergeometric functions as follows:
\begin{eqnarray}
f_{1}^{+}(\sigma)&=&A\frac{ia(\lambda+m)}{l+N/2}\left(\frac{\lambda-m}{4 \lambda}\right)^{\frac{1}{2}}
(1+z)^{\frac{l}{2}}(z-1)^{\frac{l+1}{2}}F\left(\alpha,\alpha^*;
l+\rho_N+\frac{3}{2};\frac{1-z}{2}\right)\:,\nonumber\\
f_{2}^{+}(\sigma)&=&A\left(\frac{\lambda-m}{4 \lambda}\right)^{\frac{1}{2}}(1+z)^{\frac{l+1}{2}}(z-1)^{\frac{l}{2}}
F\left(\alpha,\alpha^*;l+\rho_N+
\frac{1}{2};\frac{1-z}{2}\right)\:,\nonumber\\
f^{-}_{1}(\sigma)&=&A\left(\frac{\lambda+m}{4 \lambda}\right)^{\frac{1}{2}}(1+z)^{\frac{l+1}{2}}(z-1)^{\frac{l}{2}}
F\left(\alpha,\alpha^*;l+\rho_N+
\frac{1}{2};\frac{1-z}{2}\right)\:, \\
f^{-}_{2}(\sigma)&=&A\frac{ia(\lambda-m)}{l+N/2}\left(\frac{\lambda+m}{4 \lambda}\right)^{\frac{1}{2}}
(1+z)^{\frac{l}{2}}(z-1)^{\frac{l+1}{2}}F\left(\alpha,\alpha^*;
l+\rho_N+\frac{3}{2};\frac{1-z}{2}\right) \:, \nonumber
\end{eqnarray}
where $f_{1,2}^{\pm}(r)$ are the
solutions with $\kappa =\pm(l+\rho_D)$, $z=\cosh \sigma$, $\alpha=l+D/2+ir$,
$r=a\sqrt{\lambda^2 - m^2}$, and $A$ is the normalization constant.
As in the scalar case the normalization constant $A$ is determined from the usual $\delta$-function
condition of normalization of eigenfunctions of continuous spectrum
\begin{equation}
|A(r)|^2=\frac{\Gamma(\frac{D}{2})}{\pi^{\frac{D}{2}+1}2^{D+1+2l}}
\frac{|\Gamma(D/2+l+ir)|^2|\Gamma(ir)|^2}{|\Gamma(2ir)|^2}\:.
\end{equation}
Note that the
solutions remain bounded at infinity if $r$ is real. Hence, the
spectrum of the Dirac operator on $H^D$ is $|\lambda|\geq m$.
Thus, unlike the scalar case, there is no gap for fermions on $H^D$ (this fact is very important
for what follows). Nevertheless,
the solutions are exponentially vanishing at infinity.
Having determined the normalization constant, we immediately get the heat kernel for spinors on $H^D$
(again as in the scalar case only modes with $l\,=\,0$ are not equal to zero at the origin)
\begin{equation}
h_{H^D} \, = \,\frac{2^{[\frac{D+1}{2}]}}{a^D}\int_0^\infty
e^{-\frac{s}{a^2}(r^2+m^2a^2)}|A(r)|^2dr\,,
\end{equation}
where $[\frac{D+1}{2}]$ denotes the
integer part of $\frac{D+1}{2}$, which results from the trace over the spinor indices.
(Note that this heat kernel calculated by 'brute force' through summation over eigenfunctions coincides
with the heat kernel calculated in \cite{Campo}, which is expressed through the spinor Green function obtained as a
solution of the corresponding differential equation).
\vspace{1cm}
\subsection{Analysis.}
\vspace{1cm}
To interprete the heat kernel obtained, we remind the results of the corresponding calculations in flat spacetime.
The gap equation in flat spacetime is
\begin{equation}
1\,\, = \,\,G\int_{\frac{1}{\Lambda^2}}^{\infty} ds \,\, h_{flat},
\end{equation}
where $h_{flat} \, = \, tr<x|e^{-s(-(\gamma_E^{\mu}\nabla_{\mu})^2 + m^2)}|x>$.
The eigenfunctions of the Dirac operator in flat spacetime are just plane waves, therefore, the corresponding
heat kernel is
\begin{equation}
h_{flat} \, = \,2^{\frac{n}{2}}\int \frac{d^n k}{(2\pi)^n}
e^{-s(k^2+m^2)}.
\end{equation}
By integrating over angular variables,
we obtain
\begin{equation}
h_{flat}\, = \, 2^{\frac{n}{2}}\int_0^{\infty}
\frac{2 \, dk k^{n-1}}{(4 \pi)^{\frac{n}{2}} \Gamma(\frac{n}{2})} e^{-s(k^2+m^2)} =
\frac{2^{\frac{n}{2}}e^{-s m^2}}{(4 \pi s)^{\frac{n}{2}}}
\end{equation}
Thus, we see that the function $k^{n-1}$ defines
a measure in space of eigenfunctions and for $m = 0$ determines the asymptotic behavior of the heat kernel at large s.
Obviously, every new dimension gives an additional factor
$s^{-\frac{1}{2}}$ to the asymptotic behavior of
the heat kernel. Furthermore, we see from the gap equation (31) that in two-dimensional spacetime $G \to 0$ if $m \to 0$
because the integral over $s$ is divergent on the upper limit in this case. Thus, the critical value
of coupling constant is zero.
We now return to the heat kernel on $R \times H^D$. It is
\begin{equation}
h_{R \times H^D} \, = \, \frac{2^{[\frac{D+1}{2}]}}{a^D}\int_0^\infty
\frac{e^{-\frac{s}{a^2}(r^2+m^2a^2)}}{(4 \pi s)^{\frac{1}{2}}}|A(r)|^2dr\,,
\end{equation}
where the factor $(4 \pi s)^{\frac{1}{2}}$ is the contribution to the heat kernel from time coordinate
and the rest is the heat kernel on $H^D$ (see Eq.(30)).
Let us explicitly calculate the measure $|A(r)|^2$, which is given by Eq.(29) with $l\, = \, 0$.
By using the formulas \cite{Grad}
\begin{eqnarray}
|\Gamma(iy)|^2 = \frac{\pi}{y\sinh(\pi y)}, \nonumber \\
|\Gamma(\frac{1}{2} + iy)|^2 = \frac{\pi}{\cosh(\pi y)},
\end{eqnarray}
we get
\begin{equation}
|A(r)|^2 = \frac{r \coth(\pi r) \: \prod_{j=1}^{\frac{D-2}{2}} \: (r^2 + j^2)}
{\pi^{\frac{D}{2}}2^{D-1} \Gamma(\frac{D}{2})}
\end{equation}
for even D and
\begin{equation}
|A(r)|^2 = \frac{\prod_{j=\frac{1}{2}}^{\frac{D-2}{2}} \: (r^2 + j^2)}
{\pi^{\frac{D}{2}}2^{D-1} \Gamma(\frac{D}{2})}
\end{equation}
for odd D.
The asymptotic behavior of the heat kernel for large s in the case of critical coupling
constant ($m = 0$) is determined by the behavior of the integrand at small $r$.
As follows from Eqs. (36) and (37), for small $r$, the measure $|A(r)|^2$ tends to a constant for any D.
Consequently, we obtain that the leading term of the heat kernel (34) is $h_{R \times H^D} \sim \frac{1}{s}$.
Thus, the fermion dynamics on $R \times H^D$ in the infrared region corresponds to the dynamics of
(1 + 1)-dimensional theory.
(Note that in the opposite limit of small s (large energies that corresponds to large $r$)
the measure $|A(r)|^2$ tends to
$\frac{2r^{D-1}}{(4 \pi)^{\frac{D}{2}} \Gamma(\frac{D}{2})}$.
Therefore, the leading term of the heat kernel on $R \times H^D$ at $s \to 0$ is
$\frac{1}{(4 \pi s)^{\frac{D+1}{2}}}$ that corresponds to the behavior of (D+1)-dimensional theory
as expected). Consequently, we can say that the effective reduction of the dimension of spacetime
$1 + D \to 1 + 1$ takes place in the infrared region for fermion fields
for any 1 + D. This explains why $g_c = 0$ in spaces with constant negative curvature (it immediately follows
from the gap equation if $h \sim \frac{1}{s}$ for large $s$). For
completeness we present the corresponding results of the effective reduction for the case of the NJL model in
four-dimensional spacetime in external magnetic field \cite{GMSh1, GMSh2}.
The heat kernel for the Dirac operator in constant magnetic field is
\begin{eqnarray}
h_{magnetic}\, = \, \frac{e^{-s m^2} eB \cot (eBs)}{16 \pi^2 s}\,,
\end{eqnarray}
where $e$ is the charge of the electron and $B$ is magnetic field.
The heat kernel (38) evidently also corresponds to
(1 + 1)-dimensional theory in the infrared because $\coth(eBs)$ tends to 1 for large s.
\vspace{1cm}
\subsection{Nambu--Goldstone bosons.}
\vspace{1cm}
In the preceding subsection we found that the dynamics of fermions in the infrared is effectively (1 + 1)-dimensional.
Potentially, it may present a problem for dynamical symmetry breaking of a continuous symmetry because,
according to the Coleman--Mermin--Wagner theorem \cite{Cole}, spontaneous symmetry breaking of a continuous symmetry
is not possible in 1 + 1 due to strong infrared divergences connected with massless Nambu--Goldstone bosons
(the existence of
this potential problem in theories with the effective reduction of dimension of spacetime was indicated in
\cite{GMSh1, GMSh2}).
For example, in the NJL model the following diagram of the next-to-leading (in $\frac{1}{N}$) correction to vacuum energy
is infrared divergent in 1 + 1:
\begin{figure}[htb]
\begin{center}
\begin{picture}(200,100)(0,0)
\CArc(100,50)(50,0,360)
\DashLine(50,50)(150, 50){5}
\end{picture}
\end{center}
\end{figure}
Fig. 1. The next-to-leading order correction to vacuum energy in the NJL model. The fermion propagators
are denoted by solid lines. A dashed line denotes the propagators of $\sigma$ and $\pi$ in the leading order in $1/N$.
\vspace{1cm}
If the effective reduction of dimension of spacetime in the infrared region took place for scalars, then we
would have a problem connected with infrared divergent radiative corrections due to massless Nambu--Goldstone bosons.
For the case of the
effective reduction in external magnetic
field, Gusynin, Miransky, and Shovkovy \cite{GMSh1, GMSh2} presented an elegant solution of this potential problem.
They indicated
that since in the case of chiral symmetry breaking the condensate $<0|\bar{\psi}\psi|0>$ is neutral and the
Nambu--Goldsone bosons are
neutral particles, the effective dimensional reduction (which for fermions reflects the fact that the motion
of charged particles is restricted in the directions perpendicular to the magnetic field) does not affect the dynamics
of the center of mass of neutral excitations.
Therefore, as they showed by explicit calculations the propagators of Nambu--Goldstone bosons have
(3 + 1)-dimensional form in the infrared region. Evidently such a solution
cannot be used in the case of gravitational field because gravity is universal and
all particles including Nambu--Goldstone bosons directly interact
with gravitational field. Therefore, we should seek another solution. For this end we consider
the propagator of massless
scalar field. This propagator can be expressed through the nondiagonal heat kernel of the Laplace--Beltrami operator.
Time dependence is trivially factorized and we are left with the problem of
calculating heat kernel on $H^D$. In Subsection 3.1 we have calculated the diagonal
heat kernel $h_{scalar} = <x|e^{-\frac{s}{a^2}(-\Delta)}|x>$ (see Eq.(23)).
The nondiagonal heat kernel was calculated in
\cite{Cam}
\begin{equation}
<x|e^{-\frac{s}{a^2}(-\Delta)}|y> = \frac{1}{a^D} \int_0^{\infty} \phi_r (\tau) |C(r)|^2
e^{-\frac{s}{a^2}(r^2 + \rho_D^2)} dr,
\end{equation}
where $\phi_r (\tau) = F(ir + \rho_D, -ir + \rho_D, \frac{D}{2}; - \sinh^2 \frac{\tau}{2a})$ ($\tau$ is the geodesic
distance between points $x$ and $y$) and
$|C(r)|^2$ is given by Eq.(22).
Thus, we obtain the propagator for scalar massless particles
\begin{equation}
G(t-t^{\prime}, \tau) = \frac{1}{a^D}\int^{\infty}_{1/\Lambda^2}
ds \frac{e^{-\frac{t-t^{\prime}}{4s}}}{(4\pi s)^{\frac{1}{2}}}
\int_0^{\infty} \phi_r (\tau) |C(r)|^2
e^{-\frac{s}{a^2}(r^2 + \rho_D^2)} dr\,,
\end{equation}
where we have performed the Wick rotation in time coordinate.
Obviously, since there is a gap in the spectrum of the Laplace--Beltrami operator, there are not any problems with infrared
behavior of massless scalar particles.
Indeed, in the proper time method infrared divergences are connected with the
divergence of the integral over $s$ on the upper
limit of integration. Since there is a gap in the spectrum, the integrand has the factor
$e^{-\frac{s}{a^2}\rho_D^2}$. Therefore,
infrared divergences are absent.
Thus, we conclude that the effective reduction of dimension of spacetime in
the infrared region for fermion fields does not contradict the Mermin--Wagner--Coleman theorem.
Note that our calculations show that there are not gapless bosonic modes. Consequently, there are not gapless
Nambu--Goldstone bosons in this model. However, this does not contradict the Goldstone theorem: this theorem
has been proved only for
Minkowski space. The problem of a possible extending the theorem to the case of curved spacetime will be considered
elsewhere.
\vspace{1cm}
\section{Conclusion.}
\vspace{1cm}
In the present paper we studied chiral symmetry breaking in the NJL model
in spaces with constant negative curvature. We showed that
zero value of critical coupling constant $g_c = 0$ is connected with the effective reduction of dimension of spacetime
$1 + D \to 1 + 1$ for fermions in the infrared region.
Note that this effective reduction has a universal
character
in the sense that the initial theory reduces in the infrared region to two-dimensional one in the fermion sector
for any dimension $1 + D$. In this respect
this is similar to the effective reduction in the presence of the Fermi
surface when the net fermion charge is not equal to zero.
By analysing the scalar propagator,
we showed that such an effective reduction is absent in the scalar sector, therefore, the
effective reduction of the dimension of spacetime for fermions and symmetry breaking are consistent
and there is not a contradiction with
the Coleman--Mermin--Wagner theorem, which states that spontaneous symmetry breaking is not possible
in 1 + 1.
Finally let us mention that the hyperbolic space $H^D$ is an Euclidean analog of anti-de Sitter space
(the Wick rotation of AdS gives
$H^D$). Recently the dynamics of quantum fields on AdS has received a lot of attention in view of the conjectured CFT/AdS
correspondence \cite{Mal}. Therefore, it is a natural problem to study what the dynamics we discuss here means
in the context of this correspondence. The results of this study will be published elsewhere.
The author thanks V.P. Gusynin for the suggestion to use expansion of the heat kernel in eigenfunctions of the
Dirac operator and for the
critique of an earlier version of this paper. The author thanks V.A. Miransky for useful remarks and suggestions and
also acknowledges helpful discussions with
A.S. Belyaev, M. Nowakowski, L.C.B. Crispino, I.L. Shapiro, and Yu.V. Shtanov.
I am grateful to I.L. Shapiro for bringing my attention to \cite{Byts}.
This work was supported in part by FAPESP grant No. 98/06452-9.
|
\section*{Motivation}
Understanding the mechanism of the glass transition is one of the most challenging tasks of modern condensed matter
physics. Despite the ample experimental and theoretical work on this subject there is still no universally accepted
view thoroughly describing all physical aspects. Among the experimental approaches used to investigate this phenomenon,
dielectric spectroscopy is a well established method probing the rotational dynamics of dipolar molecules and the
translational dynamics of charged particles.
Our group has access to an extraordinary wide range of frequencies from 10$^{-6}$ to 10$^{14}$\,Hz almost covering 20
decades continuously. The various experimental setups include a home-made time-domain spectrometer (10\,$\mu$Hz $\leq
\nu \leq 1$\,kHz), commercially available autobalance bridges ($20$ Hz $\leq \nu \leq 20$ MHz), and radio-frequency and
microwave setups using coaxial reflection and transmission techniques ($1$ MHz $\leq \nu \leq 30$ GHz). At $40$ GHz
$\leq \nu \leq 1.2$ THz a quasi-optical submillimeter-wave spectrometer is employed measuring the complex transmission
coefficient. Higher frequencies (450\,GHz $\leq \nu \leq 10$\,THz) are investigated with a commercially available
Fourier-transform infrared spec\-tro\-meter.
We applied these techniques to investigate the glass transition in the molecular glass formers glycerol ($T_g =185$\,K,
hydrogen-bonded network) \cite{Lunkigly,Lunkiorl,Lunkikyo,Lunkibos,Sch98} and propylene carbonate (PC) ($T_g=160$\,K,
van-der-Waals glass) \cite{Lunkigly,Lunkiorl,Lunkikyo,Lunkibos,Sch99} and in the ionic melts
[Ca(NO$_3$)$_2$]$_{0.4}$[KNO$_3$]$_{0.6}$ (CKN) \cite{Lunkigly,Lunkiorl,Lunkikyo,Lunkibos,Lun97} and
[Ca(NO$_3$)$_2$]$_{0.4}$--[RbNO$_3$]$_{0.6}$ (CRN) (for both: $T_g =333$\,K). The frequency dependent complex
dielectric permittivity $\varepsilon{}^*=\varepsilon{}'-i\varepsilon{}''$ obtained from the experiments allows for an
investigation of a variety of dynamic processes known to be present in glass-forming materials including
$\alpha$-process, fast $\beta$-processes, and microscopic response (boson peak). We are able to trace the dynamics of
the $\alpha$-relaxation from almost total arrest near the glass temperature $T_g$ up to temperatures in the liquid
state where it starts to merge with the microscopic response.
Of special interest is the high-frequency r\'egime in the GHz -- THz range, which allows for an investigation of the
possible fast processes, that have been predicted by various theoretical approaches to be inherent to supercooled
liquids. We provide a detailed analysis of these processes using the predictions of the mode coupling theory (MCT) of
the glass transition. \cite{Goe92} This theory developed in the last 20 years describes the glass transition as a
dynamic phase transition at $T_c > T_g$. It is still controversially discussed and until recently could not be tested
properly with dielectric spectra because of the limited frequency range available in typical dielectric experiments.
The measurements of $\varepsilon{}''(\nu)$ in the far-infrared region at $\nu\approx1$ THz give access to the r\'egime
of the boson peak, known already from neutron and light scattering experiments with which our data will be compared.
\section*{Results and Analysis}
Figure~\ref{abb:fig1} shows the high frequency part of the dielectric loss spectra for glycerol
\cite{Lunkigly,Lunkiorl,Lunkikyo,Lunkibos,Sch98}, PC \cite{Lunkigly,Lunkiorl,Lunkikyo,Lunkibos,Sch99} and CKN
\cite{Lunkigly,Lunkiorl,Lunkikyo,Lunkibos,Lun97}. We observe a temperature dependent $\alpha$-peak, a shallow minimum
region and the boson-peak (not in CKN). According to the so-called idealized MCT \cite{Goe92}, above $T_{c}$, the
minimum region can be approximated by the interpolation formula:
\begin{equation} \varepsilon{}''(\nu)=\frac{\varepsilon{}''_{\min }}{(a+b)}\left[a\left( \frac{\nu}{\nu_{\min }}\right)^{-b}+b\left(\frac{\nu}{\nu_{\min }}\right)^{a}\right]\label{eq:mct}\end{equation}
$\nu _{\min}$ and $\varepsilon{}''_{\min}$ denote the position the minimum. The exponents $a$ and $b$ are temperature
independent and are constrained by the exponent parameter $\lambda =\Gamma ^{2}(1-a)/\Gamma (1-2a)=\Gamma
^{2}(1+b)/\Gamma (1+2b)$ where $ \Gamma $ denotes the Gamma function. This formula restricts the exponent $a$ to values
below $0.4$, i.e.~a significantly sublinear increase of $\varepsilon{}''(\nu )$ at frequencies above $\nu_{\rm min}$ is
predicted. The critical temperature $T_c$ should manifest itself in the temperature dependence of the
$\varepsilon{}''(\nu )$-minimum. For $T>T_c$ MCT predicts the following relations: $\nu _{\rm min}\sim
(T-T_c)^{1/(2a)}$ and $\varepsilon _{\rm min}''\sim (T-T_c)^{\frac{1}{2}}$. Additionally, MCT predicts critical
behaviour also for the $\alpha$-relaxation: the peak frequency follows $\nu_{max} \sim (T-T_c)^\gamma$ with
$\gamma=1/(2a)+1/(2b)$. In the case of the ionic conductors we chose the imaginary part of the dielectric modulus
$M''(\nu)={\rm Im}\{1/\varepsilon{}^*(\nu)\}$ to determine $\nu_{max}$ for the $\alpha$-relaxation since the maxima in
$\varepsilon{}''(\nu)$ are hidden by large conductivity contributions. This procedure is commonly used for the
evaluation of dielectric data on ionic conductors \cite{Moyni} and justified by the finding that the results for
$M''(\nu)$ follow closely those obtained by mechanical experiments.\cite{Pimenov}
\begin{figure}[tbp]
\centering
\begin{minipage}[t]{0.48\textwidth}
\centering
\includegraphics[clip,width=7.7cm]{fig1.eps}
\sf \caption{High frequency $\varepsilon{}''(\nu)$ spectra of glycerol, PC and CKN for various temperatures. The lines are fits according to the predictions of the idealized MCT \protect\ref{eq:mct} (for parameters see text).}
\rm \label{abb:fig1}
\end{minipage}\hfill
\begin{minipage}[t]{0.48\textwidth}
\centering
\includegraphics[clip,width=7.8cm]{fig2.eps}
\sf \caption{High frequency $\varepsilon{}''(\nu)$ spectra of glycerol, PC and CKN for various temperatures. The spectra with symbols are the dielectric data of $\varepsilon{}''(\nu)$, the lines are $\chi ''(\nu)$ as calculated from light and neutron scattering data.}
\rm \label{abb:fig2}
\end{minipage}
\end{figure}
The solid lines in figure~\ref{abb:fig1} are fits of the minimum region of $\varepsilon{}''(\nu)$ with
equation~\ref{eq:mct}. For glycerol reasonable fits are obtained for $\nu$\,\,${\scriptstyle\lesssim}$\,\,$\nu_{\rm
min}$ with $\lambda =0.705$, $a=0.325$, $b=0.63$. \cite{Lunkigly,Lunkibos} These values agree reasonably with the
results from other techniques.\cite{Wut94} At high frequencies the fits are limited by an additional steeper increase.
It may be argued that these deviations are due to vibrational contributions (the so-called boson-peak) which are not
included in the idealized version of MCT. A different behaviour is observed in the PC spectra: vibrational
contributions seem to be of less importance and we find a consistent description of the $\varepsilon{}''(\nu)$-minima
at $T\geq 193$ K using $\lambda =0.76$ which implies $a=0.29$ and $b=0.5$. The obtained value of the exponent parameter
$\lambda $ is consistent with the results from various other measurement techniques \cite{Du,Ohl,Berg}. The strong
boson-peak in glycerol is in accord with the findings of Sokolov {\it et al.}\,\cite{Sok} that the amplitude ratio of
boson peak and fast process is largest for 'strong' glass formers \cite{Angell}, glycerol being much stronger than PC.
The high frequency CKN spectra above 379\,K can be fitted very well using equation~\ref{eq:mct}, yielding $\lambda
=0.76$, $a=0.3$, $b=0.54$.~\cite{Lun97} The parameters are in good agreement to those obtained from light scattering
experiments.~\cite{Li}. For CRN (spectra not shown) we obtained the following set of parameters: $\lambda =0.91$,
$a=0.2$, $b=0.35$.~\cite{Lun97}
The critical temperature $T_c$ should manifest itself in the temperature dependence of the
$\varepsilon{}''(\nu)$-minimum and the $\alpha$-peak (see above). Figure~\ref{abb:fig3} presents minimum amplitude and
frequency and the $\alpha$-peak frequency in representations that lead to straight lines, extrapolating to $T_c$, if
the predicted critical behaviour is obeyed. For all materials the sets of parameters can be described consistently with
$T_c\approx 262$\,K for glycerol, 187\,K for PC, 375\,K for CKN and 365\,K for CRN. These values lie in the same region
as the $T_c$'s obtained from other techniques.~\cite{Wut94,Du,Ohl,Berg,Li,Knaak} The deviations of the data from the
predicted critical behaviour seen near $T_c$ can be ascribed to hopping processes considered in extended versions of
MCT.~\cite{Goe92} Especially for glycerol some uncertainties for $T_c$ follow as a consequence of the choice from which
temperature range the extrapolation is made.
\begin{figure}[tbp]
\centering
\begin{minipage}[t]{0.7\textwidth}
\centering
\includegraphics[angle=-90,clip,width=12cm]{fig3.eps}
\end{minipage}\quad
\begin{minipage}[t]{0.25\textwidth}
\centering
\sf \caption{Plots of the critical dependences for the position of the minimum for glycerol and PC[(a): height
$\varepsilon{}''_{\min}$, (b): frequency $\nu_{\min}$] and (c) the $\alpha$\,-peak; (d), (e) and (f) are the same plots
for CKN and CRN. The lines are drawn according to MCT as linear extrapolations to the corresponding $T_c$.}
\label{abb:fig3} \rm
\end{minipage}
\end{figure}
Finally we want to compare our spectra to those obtained from light and neutron scattering experiments. In
figure~\ref{abb:fig2} the dielectric spectra are plotted as symbols, the dotted lines are the imaginary part of the
susceptibility $\chi_{ls}''(\nu)$ calculated from light scattering and the solid lines are $\chi_{ns}''(\nu)$ from the
neutron scattering data taken from the literature~\cite{Wut94,Du,Ohl,Li,Knaak}. The scattering spectra, given in
arbitrary units, are scaled to match the $\varepsilon{}''(\nu)$ data in the boson-peak r\'egime. Comparing the three
materials in figure~\ref{abb:fig3} we observe the following universal characteristics: (1) The ratio of the structural
processes ($\alpha$-relaxation) and the boson-peak is largest in the dielectric and smallest in the neutron scattering
data. This is also found in molecular dynamics simulations of ortho-terphenyl~\cite{Wahn} and of a system of rigid
diatomic molecules.~\cite{Schsim} (2) The frequency of the $\alpha$-peak (where observable) is higher in the light
scattering data compared to the dielectric data. (3) The minimum in $\varepsilon{}''(\nu)$ only coincides with the
minima of the scattering data in CKN; for both glycerol and PC there are differences in the position of $\nu_{\min}$.
The differences in the spectra are caused by the different probes, each method couples to. A possible explanation of
the underlying microscopic processes was given considering the different dependencies of the probes on orientational
fluctuations.~\cite{Leb} Additionally, the MCT was recently generalized to molecular liquids with orientational degrees
of freedom~\cite{Schkug} thereby providing an explanation for the different ratios of $\alpha$- and boson-peak
amplitude for the different probes. In addition, very recently MCT was successfully applied to simultaneously describe
both the present dielectric and light scattering data~\cite{Li} of PC by means of a schematic model.~\cite{voigtmann}
\section*{Acknowledgements}
Our gratitude is directed to A. Maiazza for preparing the ionic conducting materials and A. Pimenov, Yu. Goncharov, B.
Gorshunov and M. Dressel for help with installing the submillimeter-wave spectrometer and performing some of the
measurements. This work was supported by the Deutsche Forschungsgemeinschaft, Grant-No. LO264/8-1 and the BMBF,
contract-No. 13N6917.
|
Subsets and Splits