content
stringlengths 1
15.9M
|
---|
\section{Introduction} \label{3c454:intro}
Among active galactic nuclei (AGNs), blazars show intense and variable
$\gamma$-ray emission above 100~MeV \citep{Hartman1999:3eg}, with variability timescales
as short as a few days, or a few weeks.
Blazars emit across several decades of energy, from the radio to the TeV
energy band, and their spectral energy distributions (SEDs) typically show two
distinct humps. The first peak occurs in
the IR/Optical band in the Flat Spectrum Radio Quasars (FSRQs) and in
the Low-energy peaked BL Lacs (LBLs), and at UV/X-rays in the
High-energy peaked BL Lacs (HBLs). The second hump peaks at MeV--GeV
and TeV energies in FSRQs/LBLs and in HBLs, respectively.
In the framework of leptonic models, the first peak is commonly
interpreted as synchrotron radiation
from high-energy electrons in a relativistic jet, while the second peak is
interpreted as inverse Compton (IC) scattering of soft seed photons by
the same relativistic electrons.
A recent review of the blazar emission mechanisms and energetics
is given in \cite{Celotti2008:blazar:jet}.
Alternatively, the blazar SED can be explained in the framework of the
hadronic models, where the relativistic protons in the jet are the primary
accelerated particles, emitting $\gamma$-ray radiation by means of photo-pair and
photo-pion production (see \citealt{Mucke2001:hadronic,Mucke2003:hadronic}
for a review on hadronic models).
Since the launch of AGILE{}, the FSRQ \hbox{3C~454.3}{} (PKS~2251$+$158; $z=0.859$) became
one of the most active sources in the $\gamma$-ray sky. Its
very high $\gamma$-ray flux (well above $100 \times 10^{-8}$\,\hbox{photons cm$^{-2}$ s$^{-1}$}{} for
$E>100$~MeV), its flux variability (on a time scale of 1 or 2 days), and
the fact that it was always detected during any AGILE{} pointing, made
it earn the nickname of {\it Crazy Diamond}: \hbox{3C~454.3}{} is now playing
the same role as 3C~279 had for EGRET{} \citep[e.g., ][]
{Hartman2001:3C279:multiwave,Hartman2001:3C279:gammaopt}.
Multiwavelength studies of variable $\gamma$-ray blazars are crucial in order
to understand the physical processes responsible for the emission
along the whole spectrum.
Since the detection of the exceptional 2005 outburst (see
\citealt{Giommi2006:3C454_Swift,Fuhrmann2006:3C454_REM,Pian2006:3C454_Integral}),
several monitoring campaigns were carried out to follow the source multi frequency behavior
\citep{vil06,vil07,rai07,rai08a,rai08b}.
In mid July 2007, 3C~454.3 underwent a new optical brightening,
which triggered observations at all frequencies.
AGILE{} performed a target of opportunity (ToO)
re-pointing towards the source and detected it in a very high $\gamma$-ray state
\citep[][hereafter V08]{Vercellone2008:3C454_ApJ}.
In November and December 2007, AGILE{} detected high $\gamma$-ray activity from
\hbox{3C~454.3}{}, triggering multiwavelength ToO campaigns, whose results
are reported in \citet[][hereafter Paper~I]{Vercellone2008:3c454:ApJ_P1},
in \citet[][hereafter Paper~II]{Donnarumma2009:3c454:subm}, and in
\cite{Anderhub2009:AA:3C454}, respectively.
Paper~I and Paper~II demonstrated that
to fit the simultaneous broad-band SEDs from radio to $\gamma$-ray data,
inverse Compton (IC) scattering
of external photons from the broad line region (BLR) off the relativistic
electrons in the jet was required.
In an earlier work based on the \cite{Vercellone2007:atel1160}
preliminary flux estimate of the July 2007 flare,
\cite{Ghisellini2007:3C454:SED} made a comparison between the
\hbox{3C~454.3}{} SEDs in 2000 (EGRET data), 2005 (optical and X-ray flare),
and 2007 (AGILE{} $\gamma$-ray flare), discussing the role of the bulk Lorentz
factor $\Gamma$ (associated with the emitting source compactness)
during the different epochs.
Moreover, the results of correlation analysis performed in Paper~I
was consistent with no time-lags between
the $\gamma$-ray and the optical flux variations. Such a result was recently
confirmed by \citet{Bonning2008:3C454_apj} who correlated optical, UV,
X-ray and $\gamma$-ray \footnote{The $\gamma$-ray data are taken from the {\it Fermi}{}/LAT
monitored source list light curves available at
\texttt{http://fermi.gsfc.nasa.gov/ssc/data/access/}} data.
In a very recent paper, \cite{Abdo2009:3C454} show the results
of the first three months of {\it Fermi}{} observations of \hbox{3C~454.3}{},
from 2008 July to October.
They present for the first time the signature of a spectral break
above a few GeV, interpreted as a possible break in the energy distribution
of the emitting particles.
In this paper (Paper~III) we present both a re-analysis of the AGILE{}
published data collected during the period July 2007 - December 2007, and
the results of multiwavelength campaigns
on \hbox{3C~454.3}{} during a long-lasting $\gamma$-ray activity period between
2008 May 10 and 2009 January 12.
In particular, we show the results of the AGILE{} campaigns
which took place on May--June 2008 (mj08), July--August 2008 (ja08),
and October 2008 - January 2009 (oj09).
Preliminary $\gamma$-ray results were distributed in
\citet{Donnarumma2008:ATel1545,Vittorini2008:ATel1581,Gasparrini2008:ATel1592,
Pittori2008:ATel1634}, while radio-to-optical data
were published in \citet{Villata2009:3C454:GASP:accep}.
The paper is organized as follows: in Sect.~\ref{3c454:grid} through~\ref{3c454:radio:vlbi}
we present the AGILE{},
{\it Swift}{}, {\it Rossi} X-ray Timing Explorer (RXTE{}),
GLAST-AGILE Support Program within the Whole Earth Blazar Telescope
(GASP-WEBT{}) and radio VLBI data analysis and results;
in Sect.~\ref{3c454:monitoring} we present the simultaneous multiwavelength light-curves.
In Sect.~\ref{3c454:disc} and~\ref{3c454:conc} we discuss the results and draw our conclusions.
Throughout this paper the quoted uncertainties are given at the
$1\sigma$ level, unless otherwise stated, and we adopted a $\Lambda$-CDM cosmology with the
following values for the cosmological parameters: $h = 0.71$, $\Omega_{m}
= 0.27$, and $\Omega_{\Lambda} = 0.73$.
\section{AGILE{} data} \label{3c454:grid}
\subsection{Data Reduction and Analysis} \label{3c454:grid:analysis}
The AGILE satellite \citep{Tavani2008_agile_nima,Tavani2009:Missione},
a mission of the Italian Space Agency (ASI) devoted to high-energy
astrophysics, is currently the only space mission capable of
observing cosmic sources simultaneously in the energy bands
18--60~keV and 30~MeV--50~GeV.
The AGILE scientific instrument combines four
active detectors yielding broad-band coverage from hard X-rays
to gamma-rays:
a Silicon Tracker~\citep[ST;][30~MeV--50~GeV]{Prest2003:agile_st},
a co-aligned coded-mask hard X-ray imager, Super--AGILE
\citep[SA;][18--60~keV]{Feroci2007:agile_sa}, a non-imaging CsI
Mini--Calorimeter~\citep[MCAL;][0.3--100~MeV]{Labanti2009:agile_mcal},
and a segmented Anti-Coincidence System~\citep[ACS;][]{Perotti2006:agile_ac}.
Gamma-ray detection is obtained by the combination of ST,
MCAL and ACS; these three detectors form the AGILE Gamma-Ray Imaging
Detector (GRID).
Level--1 AGILE-GRID data were analyzed using the AGILE Standard Analysis
Pipeline (see~V08 for a detailed description of
the AGILE data reduction). Since \hbox{3C~454.3}{} was observed at high off-axis
angle due to the satellite solar panel constraints,
an ad-hoc $\gamma$-ray analysis was performed. We used $\gamma$-ray events
filtered by means of the \texttt{FM3.119$\_$2a} AGILE{} Filter pipeline.
Counts, exposure, and Galactic background $\gamma$-ray maps were created with
a bin-size of $0.\!\!^{\circ}25 \times 0.\!\!^{\circ}25$\,,
for $E \ge 100$\,MeV. Since the source was observed up to $40^{\circ}$ off-axis,
all the maps were generated including all events collected up to
$60^{\circ}$ off-axis.
We rejected all $\gamma$-ray events whose reconstructed
directions form angles with the satellite-Earth vector smaller
than $85^{\circ}$,
reducing the $\gamma$-ray Earth albedo contamination
by excluding regions within $\sim 15^{\circ}$ from the
Earth limb.
We used the version (\texttt{BUILD-16}) of the Calibration files
(\texttt{I0006}),
which are publicly available at the ASI Science Data Centre
(ASDC) site\footnote{\texttt{http://agile.asdc.asi.it}},
and of the $\gamma$-ray diffuse emission model \citep{Giuliani2004:diff_model}.
We ran the AGILE Source Location task in order
to derive the most accurate location of the source. Then,
we ran the AGILE Maximum Likelihood Analysis (\texttt{ALIKE}) using
a radius of analysis of 10$^{\circ}$,
and the best guess position derived
in the first step. The particular choice of the analysis radius
is dictated to avoid any possible contamination from very off-axis residual
particle events.
\subsection{GRID Results} \label{3c454:grid:results}
Table~\ref{3c454:tab:grid:obs_log} shows the AGILE{}/GRID observation
log during the different time periods. We have re-analysed all the
AGILE{} data already published in V08, Paper~I, and
Paper-II, respectively, according to the procedure described in
Section~\ref{3c454:grid:analysis}. The results are discussed in
Section~\ref{3c454:monitoring}.
\begin{deluxetable}{ccccc}
\tablecolumns{4}
\tabletypesize{\normalsize}
\tablecaption{AGILE{}/GRID observation log.\label{3c454:tab:grid:obs_log}}
\tablewidth{0pt}
\tablehead{
\colhead{Epoch} & \colhead{Start Time} & \colhead{End Time} & \colhead{Exposure} \\
\colhead{} & \colhead{(UTC)} & \colhead{(UTC)} & \colhead{(Ms)} }
\startdata
1 & 2007-07-24 14:30 & 2007-07-30 11:40 & $0.22$ \\
2 & 2007-11-10 12:16 & 2007-12-01 11:38 & $0.64$ \\
3 & 2007-12-01 11:39 & 2007-12-16 12:09 & $0.56$ \\
4 & 2008-05-10 11:00 & 2008-06-09 15:20 & $1.03$ \\
5,6 & 2008-06-15 10:46 & 2008-06-30 11:14 & $0.54$ \\
7 & 2008-07-25 19:57 & 2008-08-14 21:08 & $0.70$ \\
8 & 2008-10-17 12:51 & 2009-01-12 14:30 & $2.86$ \\
\enddata
\end{deluxetable}
In the following paragraphs we report the detailed results of the
still unpublished AGILE{} data.
\subsubsection{May--June 2008} \label{3c454:grid:results:mj08}
The AGILE{} campaign was split into two different periods,
May 10--June 9 (P1) and June 15--30 (P2) because of a ToO re-pointing
towards W~Comae. The total on-source exposure is 1.03 (P1) + 0.54 (P2)~Ms.
\hbox{3C~454.3}{} was detected, during P1 and P2, at a $25.6\sigma$ and $16.3\sigma$
level with an average flux of
$ F_{\rm E>100MeV}^{\rm P1} = (218 \pm 12) \times 10^{-8}$\,\hbox{photons cm$^{-2}$ s$^{-1}$},
and $ F_{\rm E>100MeV}^{\rm P2} = (198 \pm 17) \times 10^{-8}$\,\hbox{photons cm$^{-2}$ s$^{-1}$}\,,
respectively, as derived from the AGILE{} Maximum Likelihood Code analysis.
Figure~\ref{3c454:figure1}, filled circles in panel (a),
shows the $\gamma$-ray light-curve
at 1-day resolution for photons above 100~MeV.
\begin{figure}[ht]
\resizebox{\hsize}{!}{\includegraphics[angle=-90]{figure1.eps}}
\caption{Panel (a): AGILE/GRID $\gamma$-ray light curve at
$\approx 1$-day resolution for E$>$100~MeV in units of
$10^{-8}$\,\hbox{photons cm$^{-2}$ s$^{-1}$}\, during the period May 2008 - January
2009. The downward arrows represent $2\sigma$ upper-limits.
Panel (b): the same as in panel (a), but with a time bin of
three (filled circles and squares), and seven (filled triangles)
days, respectively.
}
\label{3c454:figure1}
\end{figure}
We note that, particularly at the beginning of the campaign,
\hbox{3C~454.3}{} was not always detected on a day-by-day timescale. On
MJD $\sim$~54610 the source began to be detected at a $3\sigma$ level
almost continuously; this clearly indicates the onset of a $\gamma$-ray flaring
activity.
The average $\gamma$-ray flux as well as the daily values were
derived according to the $\gamma$-ray analysis procedure described in
\citet{Mattox1993:1633}.
First, the entire period was analyzed to determine the diffuse
gas parameters and then the source flux density was estimated
independently for each of the eighteen 1-day periods with the diffuse
parameters fixed at the values previously obtained.
Figure~\ref{3c454:figure1}, panel (b), shows the same
AGILE{}/GRID data binned on a time scale of three days. The
light curve clearly shows a strong degree of variability, with
a dynamic range of about four in about two weeks.
Figure~\ref{3c454:figure2}, panel (a), shows the average $\gamma$-ray spectra
extracted over the observing periods P1 and P2.
\begin{figure}[ht]
\resizebox{\hsize}{!}{\includegraphics[angle=0]{figure2.eps}}
\caption{Panel (a): AGILE/GRID average $\gamma$-ray spectrum
for periods P1 and P2.
The blue-dashed and the red-dotted lines represent the best--fit
power law models for P1 and P2, respectively.
Panel (b) and Panel (c) show the average $\gamma$-ray spectra during the
periods July--August 2008 and October 2008 - January 2009,
respectively.
In the three panels only three energy bins were
considered for the spectral fitting: $100 < {\rm E} < 200$~MeV, $200 < {\rm E} < 400$~MeV,
$400 < {\rm E} < 1000$~MeV (see text for details).
[{\it See the electronic edition of the Journal for a color version of
this figure.}]
}
\label{3c454:figure2}
\end{figure}
Each average spectrum was obtained by computing the $\gamma$-ray flux in
five energy bins over each period:
$50 < {\rm E} < 100$~MeV,
$100 < {\rm E} < 200$~MeV, $200 < {\rm E} < 400$~MeV,
$400 < {\rm E} < 1000$~MeV, and $1000 < {\rm E} < 3000$~MeV.
We note that the current instrument response is accurately calibrated
in the energy band 100~MeV--1~GeV, and that the flux
above 1~GeV is underestimated by a factor of about 2. For those
reasons, we fit the data by means of a simple power law model and restricted
our fit to the 100~MeV--1~GeV energy range, obtaining (in units of
${\rm photons\hspace{2mm}cm^{-2} ~s^{-1} ~MeV^{-1}}$):
\begin{equation}\label{3c454:equ:deffluxp1}
F^{\rm P1}(E) =
2.63 \times 10^{-4} \times
\left( \frac{{\rm E}}{1\, {\rm MeV}}\right)^{-(2.05 \pm 0.10)}\,,
\end{equation}
\begin{equation}\label{3c454:equ:deffluxp2}
F^{\rm P2}(E) =
1.58 \times 10^{-4} \times
\left( \frac{{\rm E}}{1\, {\rm MeV}}\right)^{-(1.98 \pm 0.16)}\,.
\end{equation}
The different energy range, and, above all, the different time period,
could explain the different value of the AGILE{} ($2.05 \pm 0.10$,
and $1.98 \pm 0.16$)
and {\it Fermi}{}-LAT ($2.27 \pm 0.03$\,, pre-break, \citealt{Abdo2009:3C454}) $\gamma$-ray photon indices.
\subsubsection{July--August 2008} \label{3c454:grid:results:ja08}
The AGILE{} campaign started immediately after the {\it Fermi}{}/LAT detection
of a very high $\gamma$-ray activity in the period 2008 July 10--21,
\citep{Tosti2008:ATel1628}, which reached, on July 10, a $\gamma$-ray flux
of $ F_{\rm E>100MeV}^{\rm Fermi} = 1200 \times 10^{-8}$\,\hbox{photons cm$^{-2}$ s$^{-1}$}\,
\citep{Abdo2009:3C454}.
The AGILE{} observations covered the period from 2008-07-25 19:57 UT to
2008-08-14 21:08 UT, for a total on-source exposure of about 0.71~Ms.
\hbox{3C~454.3}{} was detected at a $17.5\sigma$
level with an average flux of $ F_{\rm E>100MeV}^{\rm ja08} = (255 \pm 21)
\times 10^{-8}$\,\hbox{photons cm$^{-2}$ s$^{-1}$}, as derived from the AGILE{} Maximum Likelihood
Code analysis.
Figure~\ref{3c454:figure1}, filled squares in panel (a) and in panel (b),
shows the $\gamma$-ray light-curve
at 1-day and at 3-day resolution, respectively, for photons above 100~MeV.
The average $\gamma$-ray flux as well as the daily values were
derived according to the procedure described in \S~\ref{3c454:grid:results:mj08}\,.
Contrary to the May--June data, the July--August light curve does
not show any clear sign of variability.
Figure~\ref{3c454:figure2}, panel (b), shows the average $\gamma$-ray spectrum
derived over the entire observing period.
The average spectrum was obtained by computing the $\gamma$-ray flux in the same
way as in \S~\ref{3c454:grid:results:mj08}.
We fit the data by means of a simple power law model and restricted
our fit to the most reliable energy range (100~MeV--1~GeV):
\begin{equation}\label{3c454:equ:deffluxja08}
F^{\rm ja08}(E) =
3.96 \times 10^{-4} \times
\left( \frac{{\rm E}}{1\, {\rm MeV}}\right)^{-(2.11 \pm 0.14)}\,,
\end{equation}
in units of ${\rm photons\hspace{2mm}cm^{-2} ~s^{-1} ~MeV^{-1}}$.
During this period, which partially overlaps with the {\it Fermi}{} one, the AGILE{}
$\gamma$-ray photon index is, within the statistical errors, in agreement with the
{\it Fermi}{}-LAT result.
\subsubsection{October 2008 - January 2009} \label{3c454:grid:results:oj09}
The AGILE{} observations covered the period from 2008-10-17 12:51 UT to
2009-01-12 14:30 UT, for a total on-source exposure of about 2.86~Ms.
\hbox{3C~454.3}{} was detected at a $17.9\sigma$
level with an average flux of $ F_{\rm E>100MeV}^{\rm oj09} = (77 \pm 5)
\times 10^{-8}$\,\hbox{photons cm$^{-2}$ s$^{-1}$}, as derived from the AGILE{} Maximum Likelihood
Code analysis.
Figure~\ref{3c454:figure1}, filled triangles in panel (a),
shows the $\gamma$-ray light-curve
at 1-day resolution for photons above 100~MeV.
The average $\gamma$-ray flux as well as the daily values were
derived according to the procedure described in \S~\ref{3c454:grid:results:mj08}.
The light curve does not show any clear trend, partly because of a dominant
fraction of upper limits in the data. For this reason, we decided to
rebin the light curve on a time scale of one week. The resulting light
curve is shown in Figure~\ref{3c454:figure1}, filled triangles in panel (b).
On this time scale, a clear trend is present, with the source dimming
as a function of time, with a dynamic range of about a factor of two.
Figure~\ref{3c454:figure2}, panel (c), shows the average $\gamma$-ray spectrum
derived over the entire observing period.
The average spectrum was obtained by computing the $\gamma$-ray flux in the same
way as in \S~\ref{3c454:grid:results:mj08}.
We fit the data by means of a simple power law model and restricted
our fit to the most reliable energy range (100~MeV--1~GeV):
\begin{equation}\label{3c454:equ:deffluxoj09}
F^{\rm oj09}(E) =
2.10 \times 10^{-4} \times
\left( \frac{{\rm E}}{1\, {\rm MeV}}\right)^{-(2.21 \pm 0.13)}\,,
\end{equation}
in units of ${\rm photons\hspace{2mm}cm^{-2} ~s^{-1} ~MeV^{-1}}$
\subsection{Super-AGILE Results} \label{3c454:sa:results}
During the various AGILE{} pointings, \hbox{3C~454.3}{} was located substantially
off-axis in the Super--AGILE field of view (FoV). For this reason, only $3\sigma$ upper limits
can be derived in the 20--60~keV energy band during the AGILE{}/GRID observations.
Table~\ref{3c454:tab:sa} summarizes the Super--AGILE observations and
their results.
\begin{deluxetable*}{cccccc}
\tablecolumns{4}
\tabletypesize{\normalsize}
\tablecaption{Super--AGILE observation results.\label{3c454:tab:sa}}
\tablewidth{0pt}
\tablehead{
\colhead{Start Time} & \colhead{End Time} & \colhead{$\theta_{\rm X}$}
& \colhead{$\theta_{\rm Z}$} & \colhead{Exposure} & \colhead{$F_{\rm 20-60\,keV}$} \\
\colhead{(UTC)} & \colhead{(UTC)} & \colhead{(Deg.)}
& \colhead{(Deg.)} & \colhead{(ks)} & \colhead{(mCrab)}}
\startdata
2008-05-31 10:18 & 2008-06-09 13:38 & $-23.0$ & $+06.0$ & $380$ & $<16$\\
2008-06-15 14:11 & 2008-06-21 12:59 & $-36.0$ & $+08.0$ & $270$ & $<18$\\
2008-07-25 21:39 & 2008-08-02 23:29 & $+03.4$ & $-42.0$ & $345$ & $<18$\\
2008-10-17 18:47 & 2008-10-29 23:12 & $-00.8$ & $-45.0$ & $460$ & $<21$\\
\enddata
\end{deluxetable*}
\section{{\it Swift}{} data} \label{3c454:swift}
\subsection{Data Reduction and Analysis} \label{3c454:swift:analysis}
{\it Swift}{} pointed observations \citep{Gehrels2004:swift} were performed from 2007-07-26
to 2009-01-01. These observations were obtained both by means of several
dedicated ToOs (PI S.\ Vercellone) and by activating {\it Swift}{} Cycle-3
(Obs. ID 00031018 PI A.W.\ Chen) and Cycle-4 Proposals
(Obs. ID 00031216, PI S.\ Vercellone).
A long-lasting monitoring program (P.Is. L.Fuhrmann and S. Vercellone)
covers the period July--October 2008.
\subsubsection{{\it Swift}{}/XRT} \label{3c454:swift:analysis:XRT}
Table~\ref{3c454:tab:xrt:obs_log} summarizes the {\it Swift}{}/XRT observations.
The XRT data were processed with standard procedures
({\tt xrtpipeline} v0.12.1),
adopting the standard filtering and screening criteria, and using FTOOLS
in the {\tt Heasoft} package (v.6.6.1).
\begin{deluxetable*}{
l l l l r }
\tablewidth{0pc}
\tablecaption{{\it Swift}{}/XRT observation log. \label{3c454:tab:xrt:obs_log}}
\tablehead{
\colhead{Sequence} & \colhead{Obs.} & \colhead{Start time (UT)} & \colhead{End time (UT)} & \colhead{Exp.\tablenotemark{a}} \\
\colhead{} & \colhead{Mode} & \colhead{(yyyy-mm-dd hh:mm:ss)} & \colhead{(yyyy-mm-dd hh:mm:ss)} & \colhead{s} \\
\colhead{(1)} & \colhead{(2)} & \colhead{(3)} & \colhead{(4)} & \colhead{(5)}
}
\startdata
00035030013 & PC & 2007-07-26 00:55:44 & 2007-07-26 01:13:58 & 1073 \\
00035030014 & PC & 2007-07-28 07:26:46 & 2007-07-28 10:44:55 & 817 \\
00035030015 & PC & 2007-07-30 10:59:09 & 2007-07-30 14:16:56 & 897 \\
00035030016 & PC & 2007-08-01 11:05:10 & 2007-08-01 11:07:58 & 168 \\
00035030017 & WT & 2007-08-01 09:33:17 & 2007-08-01 13:06:59 & 3903 \\
\enddata
\tablenotetext{1}{NOTE.--Table~\ref{3c454:tab:xrt:obs_log} is published in its entirety in the
electronic edition of the {\it Astrophysical Journal}.
A portion is shown here for guidance regarding its form and
content. }
\tablenotetext{a}{The exposure time is spread over several snapshots during
each observations}
\end{deluxetable*}
The source count rate was variable during the campaigns,
ranging from 0.26 to 1.8 counts\,s$^{-1}$\,. For this reason,
we considered both photon counting (PC) and windowed timing (WT) data,
and further selected XRT event grades 0--12 and 0--2 for the PC and WT
events, respectively \citep{Burrows2005:grades}.
Several {\it Swift}{}/XRT observations showed an average count rate of $>0.5$ counts s$^{-1}$,
therefore in these cases pile-up correction was required for the PC data.
We extracted the source events from an annular
region with an inner radius of 3 pixels (estimated
by means of the PSF fitting technique) and an outer radius of 30 pixels
(1 pixel $\sim2\farcs36$). When the average count rate was $<0.5$
counts s$^{-1}$, we used the full 30-pixel radius region.
We also extracted background events within an
annular region centered on the source with radii of 110 and 116 pixels.
Ancillary response files were generated with {\tt xrtmkarf},
and account for different extraction regions, vignetting and
PSF corrections. We used the spectral redistribution matrices
v011 in the Calibration Database maintained by HEASARC.
{\it Swift}{}/XRT uncertainties are given at
90\% confidence level
for one interesting parameter (i.e., $\Delta \chi^2 =2.71$)
unless otherwise stated.
The {\it Swift}{}/XRT spectra were rebinned in order to have at least
20 counts per energy bin.
We fit the spectra with an absorbed power law model,
(\texttt{wabs*(powerlaw)} in \texttt{XSPEC 11.3.2}).
The Galactic absorption was fixed to the value of
$N_{\rm H}^{\rm Gal} = 1.34 \times 10^{21}$\,cm$^{-2}$,
as obtained by \cite{vil06} by means of a deep {\it Chandra}
observation.
We note the adopted value is consistent with the mean of the
distribution of the $N_{\rm H}$ values obtained by fitting
the spectra with an absorbed power law model and free absorption.
\subsubsection{{\it Swift}{}/UVOT} \label{3c454:swift:analysis:UVOT}
The UVOT data analysis was performed using the {\tt uvotimsum} and
{\tt uvotsource} tasks included in the {\tt FTOOLS} software package
(HEASOFT v6.6.1).
The latter task calculates the magnitudes through aperture photometry within
a circular region and applies specific corrections due to the detector
characteristics.
Source counts were extracted from a circular region with a 5 arcsec radius.
The background was extracted from source-free circular regions in the
source surroundings. The reported magnitudes
are on the UVOT photometric system described in \citet{Poole2008:UVOT}, and are not corrected
for Galactic extinction.
\subsubsection{{\it Swift}{}/BAT} \label{3c454:swift:analysis:BAT}
We analyzed {\it Swift}{}/BAT Survey data in order to study the hard X-ray
emission of \hbox{3C~454.3}{} and to investigate its evolution as a function of time.
We produced a light curve for the source at a 16-d binning
using the procedures described in
\citet[][and references therein; also see
\footnote{\texttt{http://swift.gsfc.nasa.gov/docs/swift/results/transients/Transient\_synopsis.html}}]{Krimm2006_atel_BTM,Krimm2008_HEAD_BTM}.
\subsection{Results} \label{3c454:swift:results}
Figure~\ref{3c454:fig:swift:xrt_uvot_bat}
shows the {\it Swift}{}/XRT fluxes in the 2--10~keV energy range, the {\it Swift}{}/UVOT
observed magnitudes (in the $V$, $B$, $U$, $W1$, $M2$, and $W2$
bands), and the {\it Swift}{}/BAT fluxes in the 15--150~keV energy range
as a function of time for the whole observing period.
\begin{figure}[ht]
\resizebox{\hsize}{!}{\includegraphics[angle=0]{figure3.eps}}
\caption{
{\it Panel (a)}: {\it Swift}{}/UVOT light curves (observed magnitudes) in the $V$
(red triangles), $B$ (green quares), and $U$ (blue circles).
{\it Panel (b)}: {\it Swift}{}/UVOT light curves (observed magnitudes) in the $W1$
(red triangles), $M2$ (green quares), and $W2$ (blue circles).
{\it Panel (c)}: {\it Swift}{}/XRT light curve (observed fluxes) in the 2-10~keV energy
band and in units of $10^{-11}$ erg cm$^{-2}$s$^{-1}$.
{\it Panel (d)}: {\it Swift}{}/BAT light curve in units of mCrab in the
energy band $15-150$\,keV. Downward arrows
mark $3\sigma$ upper limits.
}
\label{3c454:fig:swift:xrt_uvot_bat}
\end{figure}
In order to diminish the statistical uncertainties, we selected observations
with a number of degrees of freedom (dof)~$> 10$.
We note that a common dimming trend is present both in the UV and in
the X-ray energy bands.
As shown in Figure~\ref{3c454:fig:swift:xrt_uvot_bat}, panel (d),
the source has not been always detectable by {\it Swift}{}/BAT
throughout the considered period,
and in several time interval only $3\sigma$ upper limits
can be derived.
Figure~\ref{3c454:fig:swift:hister} shows the {\it Swift}{}/XRT photon index
as a function
of the X-ray flux in the 2--10~keV energy band. Black circles and
red squares represent data acquired in PC and WT mode, respectively.
WT data are not affected by pile-up at the observed count rate
($CR < 3$\,counts\,s$^{-1}$).
\begin{figure}[ht]
\resizebox{\hsize}{!}{\includegraphics[angle=0]{figure4.eps}}
\caption{{\it Swift}{}/XRT photon index as a function of the 2--10~keV
flux. Red squares and black circles mark the {\it Swift}{}/XRT windowed
timing (WT) and photon counting (PC) data, respectively.
}
\label{3c454:fig:swift:hister}
\end{figure}
We investigated the possible presence of a spectral trend in the X-ray data. If we
consider WT data only, a ``harder-when-brighter'' trend seems to be
present. Fitting the data with a constant model, we can exclude this
model at the 99.9993\% level.
When analyzing the PC data only (as well as the sum of the PC and WT
data), this spectral trend vanishes, and a fit with a constant model
still holds. Nevertheless, if we exclude the points at fluxes
$F_{\rm 2-10\,keV} < 2 \times 10^{-11}$\,\hbox{erg cm$^{-2}$ s$^{-1}$} a trend still
holds. These points at low fluxes could correspond to physically
different state of the source than the high fluxes one.
We also note that a deep and prolonged monitoring of \hbox{3C~454.3}{} at mid
and low X-ray states ($F_{\rm 2-10\,keV} \la 10^{-11}$\,\hbox{erg cm$^{-2}$ s$^{-1}$}) will
be crucial to test the possible presence of a spectral trend. Our data-set
contains only four observations (90023002, 90023003, 90023006, and 90023008)
at this flux level, which were acquired during the source low state
in December 2008.
\section{RXTE{} data} \label{3c454:rxte}
\subsection{Data Reduction and Analysis} \label{3c454:rxte:analysis}
The {\it Rossi X-ray Timing Explorer} ({RXTE}) satellite
observed \hbox{3C~454.3}{} in two epochs: from 2007-07-28 to 2007-08-04 and from
2008-05-30 to 2008-06-19. Here we report the analysis of the data
obtained both with the {\it Proportional Counter Array}
\citep[PCA,][]{Jahoda1996SPIE:RXTE},
which is sensitive in the 2--60~keV energy range,
and with the {\it High-Energy X-Ray Timing Experiment }
\citep[HEXTE,][]{Rothschild1998:HEXTE}, which is sensitive in
the 15--250~keV energy range.
RXTE{} data were collected by activating a Cycle 12 ToO proposal
(ID. 93150, PI A.W.~Chen).
The {PCA} is composed of 5 identical
units ({\it Proportional Counter Units}, PCUs), but during our
observations only part of them were used. Since PCU2 was the
only unit always on during our observations and it is the one which
is best calibrated, we report the results obtained from the
PCU2 data only.
The data were processed using the FTOOLS v6.4.1 and screened using
standard filtering criteria.
The net exposure times for the whole data-set in the first and second
epoch were 36.6~ks and 17.4~ks, respectively.
The background light curves and spectra for each observation were
produced using the model appropriate to faint sources.
We restricted our analysis to the 3--20 keV
energy range, in order to minimize the systematic errors due to
background subtraction and calibration of the {\it PCA} instrument.
Figure~\ref{3c454:fig:rxte:lc} shows the 3--20 keV light curve of
the whole RXTE{}/PCA data-set.
\begin{figure}[ht]
\resizebox{\hsize}{!}{\includegraphics[angle=-90]{figure5.eps}}
\caption{Panel (a) and (b) show the RXTE{}/PCA light curve
in the energy band 3-20 keV during the periods
2007 July 29 - August 5 (black points) and 2008 May 30 - June 19
(red triangles), respectively.
[{\it See the electronic edition of the Journal for a color version of
this figure.}]
}
\label{3c454:fig:rxte:lc}
\end{figure}
Strong variability is observed when comparing the count rates of
different observations. Moreover, the average count rate during the second
epoch (panel (b)) is reduced. In order to investigate possible
changes in the spectral shape with time we extracted light curves in
two energy ranges (3--7 keV and 7--20 keV). Their hardness ratio
did not show any significant variation.
A cumulative spectrum for the first and the second epoch was
extracted and simultaneously fitted with a power-law model corrected for
photoelectric absorption (\texttt{wabs*(powerlaw)} model in XSPEC), allowing only
the power-law normalization to assume a different value in the two
spectra.
Figure~\ref{3c454:fig:rxte:spectra} shows the RXTE{}/PCA spectra for both
periods.
A good fit ($\chi^{2}$= 68.3 for 76 degrees of freedom)
was obtained with the following best-fit parameters
(errors are at the 90\% confidence level):
photon index $\Gamma= 1.65\pm$0.02, and a flux
in the 3--20\,keV energy band $F_{\rm 3-20\,keV} = 8.4\times10^{-11} $ erg cm$^{-2}$ s$^{-1}$
and $F_{\rm 3-20\,keV} = 4.5\times10^{-11} $ erg cm$^{-2}$ s$^{-1}$
for the first and
second epoch spectrum, respectively.
\begin{figure}[ht]
\resizebox{\hsize}{!}{\includegraphics[angle=-90]{figure6.eps}}
\caption{RXTE{}/PCA average spectra for both
periods, July 2007 (black points), and May-June 2008 (red triangles),
respectively.
[{\it See the electronic edition of the Journal for a color version of
this figure.}]
}
\label{3c454:fig:rxte:spectra}
\end{figure}
We note that the average RXTE{}/PCA flux during the $\gamma$-ray flare detected
in July 2007 was about a factor of two higher than the
flux detected about 10 months later. Moreover, during both the
July 2007 and the May-June 2008 campaigns, the hard X-ray flux
varied significantly, by about 50\%, on a time scale of about one week.
We also analyzed the data of the HEXTE.
Only the data from cluster
B were analyzed, since the rocking system for background evaluation was
disabled in the other instrument cluster. After a standard
processing,\footnote{\texttt{http://heasarc.gsfc.nasa.gov/docs/xte/recipes/hexte.html}}
we extracted an average spectrum from all the available data, for a
dead-time corrected exposure time of 18~ks. The source was detected up to $\sim 50$~keV and its
spectrum can be fit well ($\chi^{2}$=15.1 for 19 d.o.f.) by a power-law
model with a photon index of $1.6\pm 0.1$, in perfect agreement with the
photon index derived from the PCA spectrum. Also, the normalization of the HEXTE
spectrum is consistent with an high energy extrapolation of the
time-averaged PCA spectrum: the observed flux in the 20--40~keV energy
band is (5$\pm$3)$\times$10$^{-11}$ erg cm$^{-2}$ s$^{-1}$.
This flux (approximately 6~mCrab) is also consistent with the upper limits
obtained by Super--AGILE in the same time periods.
\section{Optical-to-radio data} \label{3c454:optical}
\subsection{GASP-WEBT Data Reduction and Analysis} \label{3c454:webt:analysis}
The Whole Earth Blazar Telescope
(WEBT)\footnote{\texttt{http://www.oato.inaf.it/blazar/webt}, see e.g.
\cite{Villata2004:WEBT:BLLac}.} has been monitoring
\hbox{3C~454.3}{} since the exceptional
2004--2005 outburst \citep{vil06,vil07,rai07,rai08a,rai08b},
throughout the whole period of the AGILE{} observation.
We refer to \cite{rai08a,rai08b} and to \cite{Villata2009:3C454:GASP:accep}
for a detailed presentation and discussion of the radio, mm, near--IR,
optical and {\it Swift}{}/UVOT data.
\begin{figure}[ht]
\resizebox{\hsize}{!}{\includegraphics[angle=-90]{figure7.eps}}
\caption{GASP-WEBT light curve in the $R$ optical band
in the 2007--2008 and 2008--2009 observing seasons.
}
\label{3c454:fig:webt:lc_R}
\end{figure}
Figure~\ref{3c454:fig:webt:lc_R} shows the GASP-WEBT light curve in the
$R$ optical band, displaying several intense flares with a dynamic
range of $\sim 2.4$~mag in about 14 days,
while Figure~\ref{3c454:fig:gamma_gasp} shows
the GASP-WEBT light curves in the near-IR ($J$, $H$, $K$), radio
(5, 8, and 14.5~GHz), and mm (37, 230, and 345~GHz),
respectively\footnote{The radio-to-optical data presented in this
paper are stored in the GASP-WEBT archive; for questions regarding
their availability, please contact
the WEBT President Massimo Villata ({\tt <EMAIL>}).}.
\begin{figure}[ht]
\resizebox{\hsize}{!}{\includegraphics[angle=0]{figure8.eps}}
\caption{{\it Panel (a)}: low-frequency radio data. Red triangles, blue
squares and black circles represent the radio flux at 5, 8, and
14.5 GHz, respectively.
{\it Panel (b)}: high-frequency radio data. Red triangles, blue squares,
and black circles represent the radio flux at 37, 230, and 345 GHz,
respectively.
{\it Panel (c)}: near-IR data. Red triangles, blue squares, and
black circles, represent the $J$, $H$, and $K$ bands, respectively
[{\it See the electronic edition of the Journal for a color version of
this figure.}]
}
\label{3c454:fig:gamma_gasp}
\end{figure}
\section{Radio VLBI data} \label{3c454:radio:vlbi}
\subsection{Radio VLBI Data Reduction and Analysis} \label{3c454:radio:vlbi:analysis}
High resolution radio VLBI data were obtained from the MOJAVE (Monitoring Of
Jets in Active galactic nuclei with VLBA Experiments) project, a long-term
program to monitor radio brightness and polarization variations in jets
associated with active galaxies visible in the northern sky
(\citealt{Lister2009:AJ:Mojave}; see also\footnote{\tt http://www.physics.purdue.edu/MOJAVE}).
The object was observed with the full VLBA at 15\,GHz. We obtained the
calibrated $I$ images and used the \texttt{AIPS} package to derive the position and flux
density of the core and of a few substructures in the jets using
the task \texttt{JMFIT} (Gaussian fit) (see
Figure~\ref{3c454:fig:vlbi:20070809}).
\begin{figure}[ht]
\resizebox{\hsize}{!}{\includegraphics[angle=0]{figure9.eps}}
\caption{VLBI image of 3C~454.3 at 15~GHz on 2007 August 9 (MJD 54321). The peak flux
density is 2.8 Jy\,\,beam$^{-1}$ and contours are traced at $\pm (1, \sqrt(2),
2, ...)\times1.0$ mJy beam $^{-1}$. The cross in the bottom left corner
shows the beam FWHM, which is $1.07 \times 0.52$ mas at $-5.4$ deg.}
\label{3c454:fig:vlbi:20070809}
\end{figure}
Moreover, this source was additionally observed by VLBA at four
epochs during the period of the maximum brightness within the
BK150 VLBA experiment to measure parsec-scale spectra of
$\gamma$-ray bright blazars (Sokolovsky et al., in preparation). We use
15 and 43~GHz results from this program to provide better radio
coverage of the high activity period.
These data are in agreement with MOJAVE results and
give a better statistics in the high active period.
The core is always unresolved by our Gaussian fit and uncertainties
on the flux density are dominated by calibration uncertainties (a few
percent).
In Figure~\ref{3c454:fig:vlbi:components} we show the 3C~454.3 VLBI
radio core flux (panel (a))
at 15 and 43\,GHz, the radio components
flux density at 15 GHz (panel (b)), and the distance of radio components from
the core (panel (c)) as a function of time.
\begin{figure}[ht]
\resizebox{\hsize}{!}{\includegraphics[angle=0]{figure10.eps}}
\caption{Panel (a): radio core flux density at 15~GHz (filled
circles) and at 43~GHz (open squares), respectively. Panel(b): radio components
flux density at 15~GHz. Panel (c): radio components motion at 15~GHz. The vertical dashed
lines represent the start (2007 July 24) and the stop (2009
January 12) of all AGILE{} observations, respectively.
}
\label{3c454:fig:vlbi:components}
\end{figure}
The flux shows a constant increase from 2006 June 15 (MJD 53901) till 2008
October 3 (MJD 54742), followed by a fast decrease towards the last epoch
presented here, 2009 June 25 (MJD 55007).
Jet components show a well defined flux density decrease (component 1) or a
slower flux density decrease which becomes almost constant in the last epochs.
Proper motion is evident, but slowing in time for components 1 and 2; it is almost
absent for component 3.
All data are in agreement with a strong core flux density variability
possibly connected to the $\gamma$-ray activity, while jet components are moving
away and slowly decreasing in flux density, and are not affected by the
recent core activity.
In recent paper, \cite{Kovalev2009:ApJ:Gamma_Radio} indeed find a
connection between the radio and the $\gamma$-ray emission, correlating the
{\it Fermi}{} three month data with the MOJAVE ones, and arguing that the
central region of the blazars being the source of $\gamma$-ray flares.
Nevertheless, a detailed study of the radio structure of \hbox{3C~454.3}{}
is beyond the aims of this paper, therefore the jet properties will be
discussed in depth elsewhere (Lister et al. 2009 in preparation, and Jorstad et al.
2010 in preparation).
For this reason in the following we will concentrate only on the core.
In the last two years this source has also been observed with
the VLBA at 43\,GHz (Jorstad et al. 2010; see also
\footnote{{\tt http://www.bu.edu/blazars/VLBAproject.html}}).
We used the available images to derive the core flux density of the
core at 43\,GHz. Note that, for a better comparison with 15\,GHz VLBI
data, at 43\,GHz we used natural weights and we have not searched
for possible core subcomponents (we refer to Jorstad et al. 2010 for
a detailed study of the radio structure).
The radio core shows an inverted spectrum (self-absorbed), more evident in
the high active regime, followed by a strong flux density decrease. In this
region the radio spectrum is no longer inverted.
\section{Eighteen months of monitoring} \label{3c454:monitoring}
In this section we present a summary of all the observations on \hbox{3C~454.3}{}
in the period between 2007 July 24 and 2009 January 12.
The results of the campaigns performed in
2007 July, November and December were discussed in
\cite{Vercellone2008:3C454_ApJ}, \cite{Vercellone2008:3c454:ApJ_P1}, and
\cite{Donnarumma2009:3c454:subm}, respectively.
\begin{figure}[ht]
\resizebox{\hsize}{!}{\includegraphics[angle=-90]{figure11.eps}}
\caption{AGILE{}/GRID light
curve at $\approx 3$-day resolution for E$>$100~MeV in units of
$10^{-8}$\,\hbox{photons cm$^{-2}$ s$^{-1}$}\,. Different colors correspond to different
observing campaigns, as described in Table~\ref{3c454:tab:grid:obs_log}.
}
\label{3c454:fig:grid:3d:18months}
\end{figure}
Figure~\ref{3c454:fig:grid:3d:18months} shows the AGILE{}/GRID light
curve at $\approx 3$-day resolution for E$>$100~MeV in units of
$10^{-8}$\,\hbox{photons cm$^{-2}$ s$^{-1}$}.
The light curve shows several $\gamma$-ray flares, with a dynamical range of
a factor of 3--4 on a time scale of about ten days. Moreover, a clear
dimming trend in the long-term light curve is present.
Table~\ref{3c454:tab:grid:flux_spec} shows the AGILE{}/GRID fluxes
and spectral indices derived at different epochs.
\begin{deluxetable*}{cccc}
\tablecolumns{4}
\tabletypesize{\normalsize}
\tablecaption{AGILE{}/GRID $\gamma$-ray fluxes and spectral indices above 100~MeV
at different epochs.\label{3c454:tab:grid:flux_spec}}
\tablewidth{0pt}
\tablehead{
\colhead{Start Time} & \colhead{End Time} & \colhead{F$_{\rm E>100\,MeV}$} & \colhead{$\Gamma$}\\
\colhead{(UTC)} & \colhead{(UTC)} & \colhead{($\times 10^{-8}$\,ph\,cm$^{-2}$ s$^{-1}$)} & \colhead{} }
\startdata
2007-07-24 14:30 & 2007-07-30 11:40 & $416.2 \pm 36.0$ & $1.74 \pm 0.16$ \\
2007-11-10 12:16 & 2007-12-01 11:38 & $224.2 \pm 15.3$ & $1.91 \pm 0.14$ \\
2007-12-01 11:39 & 2007-12-16 12:09 & $265.7 \pm 17.5$ & $1.86 \pm 0.12$ \\
2008-05-10 11:00 & 2008-06-09 15:20 & $218.5 \pm 12.2$ & $2.05 \pm 0.10$ \\
2008-06-15 10:46 & 2008-06-30 11:14 & $198.5 \pm 17.1$ & $1.98 \pm 0.16$ \\
2008-07-25 19:57 & 2008-08-14 21:08 & $254.8 \pm 20.6$ & $2.11 \pm 0.14$ \\
2008-10-17 12:51 & 2009-01-12 14:30 & $ 77.0 \pm 5.5$ & $2.21 \pm 0.13$ \\
\enddata
\end{deluxetable*}
\subsection{Multiwavelength light curves} \label{3c454:monitoring:lc}
The AGILE{}/GRID wide field of view allowed for the first
time a long-term monitoring of \hbox{3C~454.3}{} at energies above
100~MeV. Moreover, coordinated and almost simultaneous
GASP-WEBT and {\it Swift}{} observations provided invaluable
information on the flux and spectral behavior from
radio to X-rays.
Figure~\ref{3c454:fig:multi:lc} shows the \hbox{3C~454.3}{} light curves
at different energies over the whole period.
\begin{figure*}[ht]
\resizebox{\hsize}{!}{\includegraphics[angle=0]{figure12.eps}}
\caption{\hbox{3C~454.3}{} light curves at different energies (see Section
\ref{3c454:monitoring:lc} for details) over the whole time period.
}
\label{3c454:fig:multi:lc}
\end{figure*}
The different panels
show, from bottom to top, the AGILE{}/GRID light curve at
$\approx 1$-day resolution for E$>$100~MeV in units of
$10^{-8}$\,\hbox{photons cm$^{-2}$ s$^{-1}$}, the {\it Swift}{}/BAT light curve in the energy range
15--150~keV at $\approx 2$-week resolution, the {\it Swift}{}/XRT (filled circles)
and the RXTE{}/PCA (filled squares) light curves
in the energy range 3--10~keV , the {\it Swift}{}/UVOT light curve
in the UV $W2$ filter, the {\it Swift}{}/UVOT light curve
in the optical $B$ filter, the GASP-WEBT light curve
in the optical $R$ filter, and the VLBI radio core
at 15~GHz (filled circles) and the UMRAO 14.5~GHz (open circles)
light curves, respectively.
We note that RXTE{}/PCA data are systematically higher than
{\it Swift}{}/XRT ones, which is consistent with the 20\% uncertainty
in the relative calibrations of the two instruments in this energy
band, reported by \cite{Kirsch2005:SPIE}.
Figure~\ref{3c454:fig:R:mm:GRID} shows the light curves in the
$R$ band, at 1.3\,mm (230\,GHz), and above 100~MeV. The light curve in
the millimeter wavelength shows a different behavior starting
from the enhanced $\gamma$-ray activity at MJD$\sim 54600$, as will be
discussed in Sect.~\ref{3c454:discu:jet}.
\begin{figure}[ht]
\resizebox{\hsize}{!}{\includegraphics[angle=0]{figure13.eps}}
\caption{Comparison between the light curves in different
bands. Panel (a), (b), and (c) show the light curves in the
optical, millimeter, and $\gamma$-ray energy bands, respectively.
}
\label{3c454:fig:R:mm:GRID}
\end{figure}
Moreover, starting from MJD~54750, the whole jet seems to become less
energetic, with an almost monotonic flux decrease, except for a minor
burst at MJD~54800.
\section{Discussion} \label{3c454:disc}
In the following Sections, we will discuss the
correlations between the flux variations in different energy bands,
the properties of the jet,
and the physical parameters of the emitting source. This latter point
will be addressed by means of complementary approaches, namely the SED
model fitting and discussing the geometrical properties of the jet itself.
\subsection{Correlation analysis} \label{3c454:discu:dcf}
We investigated the correlation between the $\gamma$-ray flux and the optical
flux density in the $R$ band by means of the discrete correlation function
\citep[DCF,][]{ede88,huf92}. This method was developed to study unevenly
sampled data sets and can give an estimate of the accuracy of its results.
Because of the sampling gaps in the light curves, especially at $\gamma$-rays,
we calculated the DCF on four distinct periods: July 2007 (mid 2007),
November--December 2007 (Fall 2007), May--August 2008 (mid 2008), and
October 2007 - January 2009 (Fall 2008).
The upper limits on the $\gamma$-ray fluxes were considered as detections,
with fluxes equal to one half of the limit.
In ``mid 2007'' AGILE was pointed at 3C 454.3 when its optical main peak was
already over; furthermore, we only have 5 $\gamma$-ray points. The low
statistics prevents us to obtain reliable results with the DCF for this
period.
In contrast, the period ``fall 2007'' offers a good opportunity to test the
correlation, since the $\gamma$-ray flux, and even more the optical flux,
exhibited strong variability. Moreover, the period of common monitoring lasted
for more than a month. The corresponding DCF (Figure~\ref{3c454:fig:dcf_go_mc}) shows a
maximum $\rm DCF \sim 0.38$ for a null time lag.
\begin{figure}[ht]
\resizebox{\hsize}{!}{\includegraphics[angle=0]{figure14.eps}}
\caption{Discrete correlation function between the $\gamma$-ray and optical
fluxes during the ``fall 2007'' period. The uncertainty in the
time-lag can be computed according
to the FR/FSS method. The inset shows the
resulting centroid distribution (see text for details).
}
\label{3c454:fig:dcf_go_mc}
\end{figure}
However, the shape of the peak is asymmetric, and if we calculate the
centroid \citep{pet98}, we find that the time lag is $-0.42$\,days,
i.e.\ about 10 hours. This result is in agreement with what was found by
\citet{Donnarumma2009:3c454:subm}
when analyzing the December 2007 observations only, and implies
that the $\gamma$-ray flux variations are delayed by few hours with respect
to the optical ones.
In the period ``mid 2008'' the main optical peak (and also the minor one)
occurred when AGILE was not observing the source.
Finally, we computed the DCF corresponding to the ``fall 2008'' period.
We obtain a broad maximum, indicating a fair correlation
($\rm DCF_{max} \sim 0.66$), but with large errors, peaking at $-2 \, \rm day$
time lag, but with centroid around 0 day. This result is consistent with that
obtained in the ``fall 2007'' period, which appears to be the most robust one.
Hence, for this case we estimated the uncertainty on the time lag by means of
the statistical method known as ``flux randomization / random subset selection''
\citep[FR/RSS][]{pet98,rai03}. We run 2000 FR/RSS realizations and for each
of them calculated the centroid corresponding to the maximum. The resulting
centroid distribution shown in Figure~\ref{3c454:fig:dcf_go_mc} allows us to conclude
that the $\gamma$-optical correlation occurs with a time
lag of $\tau=-0.4^{+0.6}_{-0.8}$, the uncertainty corresponding
to a $1 \sigma$ error for a normal distribution.
This result is consistent with a recent
analysis of the public {\it Fermi}{} data and the optical SMARTS data
by \cite{Bonning2008:3C454_apj}.
\subsection{Variability analysis} \label{3c454:discu:var}
The {\it observed} variance of a light curve for a specific
detector can be written as
\begin{equation}
S^{2} = \frac{1}{N-1}\sum_{i=1}^{N} (x_{i} - \bar{x})^{2},
\end{equation}
where $\bar{x}$ is the average value of the $x_{i}$ measurements.
Moreover, since we deal with different detectors, in order to
take into account the different count rates in different
energy bands, and to compare their variance, we consider
the normalized variance, $S^{2}/\bar{x}^{2}$.
In order to compute the {\it intrinsic} variance of a source
light curve, the measurement errors must be taken into
account, since they contribute an additional term to the
variance.
This approach was treated in detail by \cite{Nandra1997:excvar}
and by \cite{Edelson2002:excvar}, who introduced the term
of ``excess variance'':
\begin{equation}
\sigma_{\rm XS}^{2} = S^{2} - \bar{\sigma^{2}},
\end{equation}
where $\bar{\sigma^{2}}$ is the mean squared error,
\begin{equation}
\bar{\sigma^{2}} = \frac{1}{N} \sum_{i=1}^{N} \sigma_{i}^{2},
\end{equation}
and $\sigma_{i}$ are the measurement uncertainties of a light
curve points $x_{i}$.
Thus, the normalized excess variance,
\begin{equation}
\sigma_{\rm NXS}^{2} = \frac{\sigma_{\rm XS}^{2}}{\bar{x^{2}}},
\end{equation}
can be used to compare variances between different observations.
In order to quantify the flux variability in different
energy bands, we computed the fractional root mean square (rms)
variability amplitude, $F_{\rm var}$, defined as
\begin{equation}
F_{\rm var} = \sqrt{\frac{S^{2} - \bar{\sigma^{2}}}{\bar{x}^{2}} }
\end{equation}
\citep[see also][and references therein]{Vaughan2003:FVAR}.
\begin{figure}[ht]
\resizebox{\hsize}{!}{\includegraphics[angle=0]{figure15.eps}}
\caption{Fractional rms variability amplitude as a function of
frequency.
}
\label{3c454:fig:fvar}
\end{figure}
Figure~\ref{3c454:fig:fvar} shows the fractional rms variability amplitude
at different frequencies.
The optical $R$ band is the one showing the highest degree of
variability. This is partly due to the higher sampling of
the optical data with respect to the the other frequencies.
Nevertheless, a possible trend (higher fractional variability
amplitude at higher frequency) is also present.
This possible trend of the fractional rms variability amplitude
with the logarithm of the frequency was observed in other sources
too \citep[see e.g. PKS~2155$-$304,][]{Zhang2005:PKS2155:fvar},
and it was interpreted as signature of spectral variability.
A more systematic study of the variability properties of
\hbox{3C~454.3}{} will be addressed in a forthcoming paper.
\subsection{Radio VLBI vs $\gamma$-ray Data} \label{3c454:monitoring:vlbi_gamma}
Figure~\ref{3c454:fig:multi:lc} clearly shows a strong enhancement of
the radio core flux starting about on MJD~54500. The
highest flux density is on MJD~54742 at 15~GHz and on MJD~54719 at~43 GHz.
This variability is not well correlated with the variability at
higher frequencies: optical and $\gamma$-ray data data show more different
flares in the period MJD 54400--54800
(see Figure~\ref{3c454:fig:R:mm:GRID} panels (a) and (c), respectively).
Moreover, the radio flux density increase is smooth and longer in time,
while $\gamma$-ray and optical flares are evolving faster.
At 230~GHz the flux density variability mimics the VLBI radio core
properties to MJD~54600, when a large flux density increase is visible,
with a peak at about MJD~54630.
At this frequency the source remains
in an active phase up to MJD~54700 (Figure~\ref{3c454:fig:R:mm:GRID}, panel (b)).
This poses an interesting question as to the nature of such an increase
of the core radio flux.
As reported in \cite{Ghisellini2007:3C454:SED} it is likely that the emitting region is more
compact and has a smaller bulk
Lorentz factor closer to the supermassive black hole.
We can assume that in the region active at 43~GHz, in the quiescent
state, the jet Lorentz factor is $ \Gamma \sim 10$ \citep{Giovannini2001:ApJ:VLBI}.
To obtain the flux density
increase of the core at 43~GHz (from $\sim$ 5 Jy up to 25 Jy)
the Doppler factor has to increase up to $\delta \sim 30$.
Such an increase requires that the source is oriented at a small angle
$\theta$ with
respect to the line of sight, since
a large change in the jet velocity will produce a small increase in the Doppler factor.
A Doppler factor $\delta = 30$ can be obtained if $\theta = 1.5^{\circ}$ and
$\Gamma = 20$, corresponding to a bulk velocity increase from 0.9950 to 0.9987
(note that a larger orientation angle, e.g. $\theta = 3^{\circ}$ with the same
increase in the jet velocity, will produce a small change in the
Doppler factor $\delta$, from 16 to 19).
The presence of one or more new jet components is not
revealed in the high resolution VLBA images, even if the most recent VLBA
images at 43~GHz suggest a jet expansion near to the
radio core starting from MJD $\sim$ 54600.
Because of different properties (multiple bursts at high frequency, a single
peak in the radio band) it is not possible to correlate the radio peak
with a single $\gamma$-ray or optical burst. We can speculate that a multiple source
activity in the optical and $\gamma$-ray bands is integrated in the radio emitting
region in a single event. This event (see
Figure~\ref{3c454:fig:vlbi:components}) has a clear flux
density peak on MJD $\sim$ 54720 and we can assume that 43~GHz is the
self-absorption frequency at that epoch.
This scenario is in agreement with the one discussed by
\cite{Krichbaum2008:ASPC} (see their Figure 3).
According to \cite{Marscher1983:ApJ:SSA}, the self-absorption frequency,
the source size, flux density, and the magnetic field are correlated as follows,
\begin{equation}
B = 3.2 \times 10^{-5} \times \theta^{4}\, \nu_{\rm m}^{5}\, S_{\rm m}^{-2} \,
\delta\, (1+z)^{-1} \,\, {\rm Gauss},
\end{equation}
where $B$ is the magnetic field in Gauss, $\theta$ the angular size in mas
(note that $\theta = 1.8 \times {\rm HPBW}$, where HPBW is the half
power beam width), $\nu_{\rm m}$ is the frequency (in GHz) of the
maximum flux density $S_{\rm m}$ (in Jy), and $\delta$ is the Doppler
factor, respectively.
Moreover, we assume a particular value ($\alpha=0.5$)
of spectral index in the optically thin part of the synchrotron spectrum.
Thus, we can use the radio VLBI data at 43~GHz to constrain the physical
properties in the region where the source will start to be visible at
at this frequency. The angular resolution in the jet direction of VLBA data at
43~GHz is $\sim$ 0.14 mas corresponding (as discussed in \citealt{Marscher1983:ApJ:SSA})
to $\theta \la 0.25$ mas.
Assuming $\delta = 30$, we obtain $B \la 0.5$ Gauss.
It is reasonable to assume that when the source is even smaller, the emission in the
radio band is not visible being self-absorbed, and that the local
magnetic field is $ B \la 0.5$ Gauss when we start to detect the radio emission.
The size of this region should be smaller than 0.25 mas (about 2~pc).
\subsection{Spectral analysis} \label{3c454:discu:sed}
The correlation between the flux level and the spectral slope
in the $\gamma$-ray energy band was extensively studied by means of the
analysis of the EGRET{} data.
\cite{Nandikotkur2007:EGRET:slopes} showed that the behavior of EGRET
blazars is inhomogeneous.
Figure~\ref{3c454:fig:1yr_grid_hyster} shows the AGILE{}/GRID
photon index as a function of the $\gamma$-ray flux at different epochs.
A ``harder-when-brighter'' trend seems to be present in the long time
scale AGILE{} data.
\begin{figure}[ht]
\resizebox{\hsize}{!}{\includegraphics[angle=0]{figure16.eps}}
\caption{AGILE{}/GRID photon index as a function of the $\gamma$-ray flux
above 100~MeV. Number beside each points represents the epochs
listed in Table~\ref{3c454:tab:grid:obs_log}.
}
\label{3c454:fig:1yr_grid_hyster}
\end{figure}
Further long term observations of \hbox{3C~454.3}{} and of other bright $\gamma$-ray
blazars at different flux levels with AGILE{} and {\it Fermi}{} will be
crucial to assess this topic.
Different emission mechanisms can be invoked to explain the $\gamma$-ray
emission. In the leptonic
scenario, the low--frequency peak in the blazar SED is interpreted as synchrotron
radiation from high--energy electrons in the relativistic jet,
while the high--energy peak can be produced by IC on different
kinds of seed photons. In the synchrotron
self--Compton [SSC] model (\citealt{Ghisellini1985:SSC},
\citealt{Bloom1996:SSC}) the seed photons come from the
jet itself. Alternatively, the seed photons can be those of the
accretion disk [external Compton scattering of direct disk
radiation, ECD, \cite{Dermer1992:ECD}] or those of the broad--line
region (BLR) clouds [external Compton scattering from clouds, ECC,
\cite{Sikora1994:ECC}]. The target seed photons can also be those
produced by the dust torus surrounding the nucleus
[external Compton scattering from IR-emitting dust, ERC(IR),
\cite{Sikora2002:ERCIR}].
We fit the SEDs for the different observing periods
by means of a one-zone leptonic model, considering the contributions
from SSC and from external seed photons originating both from the
accretion disk and from the BLR (detailed description of this model
is given in \citealt{Vittorini2009:ApJ:Model}).
Indeed, emission from both of them were detected during faint states
of the source \citep{rai07}.
\begin{figure}[ht]
\resizebox{\hsize}{!}{\includegraphics[angle=0]{figure17.eps}}
\caption{\hbox{3C~454.3}{} SED centered on MJD~54617--54618. Black triangles,
red (blue) squares, red (blue) circles, green circles, and black stars represent radio,
MJD~54617 (54618) {\it Swift}{}/UVOT, MJD~54617 (54618) {\it Swift}{}/XRT, RXTE{}/PCA, and AGILE{}/GRID
data, respectively. UV and X-ray data are de-reddened and corrected for Galactic extintion.
The thin solid, dotted, dashed, dot-dashed, and the triple-dot-dashed,
represent the accretion disk blackbody, the
synchrotron, the SSC, the external Compton on the disk, and
the external Compton on the BLR radiation, respectively. The thick
solid line represent the sum of all the individual components.
}
\label{3c454:fig:sed:mj08}
\end{figure}
The emission along the jet is assumed to be produced in
a spherical blob with comoving radius $R$ by
accelerated electrons characterized by a comoving broken power
law energy density distribution of the form,
\begin{equation}
n_{e}(\gamma)=\frac{K\gamma_{\rm b}^{-1}}
{(\gamma/\gamma_{\rm b})^{\alpha_{\rm l}}+
(\gamma /\gamma_{\rm b})^{{\alpha}_{\rm h}}}\,,
\label{eq:ne_gamma}
\end{equation}
where $\gamma$ is the electron Lorentz factor assumed to vary
between $10<\gamma<10^{4}$, $\alpha_{\rm l}$ and
$\alpha_{\rm h}$ are the pre-- and post--break electron distribution
spectral indices, respectively, and $\gamma_{\rm b}$ is the
break energy Lorentz factor. We assume that the blob
contains an homogeneous and random magnetic field
$B$ and that it
moves with a bulk Lorentz Factor
$\Gamma$ at an angle
$\Theta_{0}$ with respect to the line of sight. The
relativistic Doppler factor is then $\delta = [ \Gamma \,(1 - \beta
\, \cos{\Theta_{0}})]^{-1}$, where $\beta$ is the usual blob bulk
speed in units of the speed of light.
Our modeling of the \hbox{3C~454.3}{} high-energy emission is based
on an inverse Compton model with two main sources of external
target photons:
{(1)} an accretion disk characterized by a blackbody spectrum
peaking in the UV with a bolometric luminosity $L_{\rm d}$
for an IC-scattering blob at a distance $r_{\rm d}=4.6 \times 10^{16}$\,cm
from the central part of the disk;
{(2)} a Broad Line Region with a spectrum peaking in the $V$ band,
placed at a distance from the blob of $r_{\rm BLR}=4 \times 10^{18}$\,cm,
and assumed to reprocess $10\%$ of the irradiating
continuum (\citealt{Tavecchio2008:blr:cloudy,rai07,rai08b}).
\begin{figure}[ht]
\resizebox{\hsize}{!}{\includegraphics[angle=0]{figure18.eps}}
\caption{\hbox{3C~454.3}{} SED during the period MJD~54673--54693.
Black triangles, multicolor squares, circles, and black stars
represent radio, {\it Swift}{}/UVOT, {\it Swift}{}/XRT, and AGILE{}/GRID
data, respectively. UV and X-ray data are de-reddened and corrected for Galactic extintion.
The thin solid, dotted, dashed, dot-dashed, and the triple-dot-dashed,
represent the accretion disk blackbody, the
synchrotron, the SSC, the external Compton on the disk, and
the external Compton on the BLR radiation, respectively. The thick
solid line represent the sum of all the individual components.
}
\label{3c454:fig:sed:ja08}
\end{figure}
These two regions contribute to the ECD and the ECC,
respectively, and it is interesting to test the relative
importance of the two components that can be emitted by the
relativistic jet of \hbox{3C~454.3}{} under different conditions. We summarize here
the main results of our best model for the different time periods.
Table~\ref{3c454:tab:sed:param} shows the best-fit parameters
of our modeling of SEDs corresponding to the following periods
(see Figure~\ref{3c454:figure1}):
(SED1), MJD~54617--54618, when \hbox{3C~454.3}{} entered a phase of
high $\gamma$-ray activity;
(SED2) MJD~54673--54693, when the $\gamma$-ray flux
was almost constant;
(SED3) MJD~54800--54845, when the source flux
was at the minimum level (about $70 \times 10^{-8}$\,\hbox{photons cm$^{-2}$ s$^{-1}$}\,).
In Figures~\ref{3c454:fig:sed:mj08}, \ref{3c454:fig:sed:ja08},
and \ref{3c454:fig:sed:od08},
the thin solid, dotted, dashed, dot-dashed, and the triple-dot-dashed,
represent the accretion disk blackbody, the
synchrotron, the SSC, the external Compton on the disk, and
the external Compton on the BLR radiation, respectively, while the thick
solid line represents the sum of all the individual components.
The insert of Figure~\ref{3c454:fig:sed:od08} shows the
portion of the SED dominated by the contribution of the disk blackbody
radiation, which clearly emerges since the source is a relative low state.
\begin{figure}[ht]
\resizebox{\hsize}{!}{\includegraphics[angle=0]{figure19.eps}}
\caption{\hbox{3C~454.3}{} SED during the period MJD~54800--54845. Black trinagles,
multicolor squares, circles, and black stars represent radio,
{\it Swift}{}/UVOT, {\it Swift}{}/XRT, and AGILE{}/GRID
data, respectively. UV and X-ray data are de-reddened and corrected for Galactic extintion.
The thin solid, dotted, dashed, dot-dashed, and the triple-dot-dashed,
represent the accretion disk blackbody, the
synchrotron, the SSC, the external Compton on the disk, and
the external Compton on the BLR radiation, respectively. The thick
solid line represent the sum of all the individual components.
The insert shows the portion of the SED dominated by the
contribution of the disk blackbody radiation.
}
\label{3c454:fig:sed:od08}
\end{figure}
We find that the three SEDs can be reproduced well by very similar
parameters, the main difference being the shape of the electron
distribution and the break energy Lorentz factor.
We note that the observed minimum variability time scale, of the
order of half a day, is consistent with the
minimum variability time scale ($\sim 10$
hours) allowed by the model fit.
Finally, we computed for the three different SEDs the total power
carried in the jet, $P_{\rm jet}$, defined as
\begin{equation}
P_{\rm jet} = L_{\rm B} + L_{\rm p} + L_{\rm e} + L_{\rm rad}\,\,\,{\rm
erg}\,{\rm s}^{-1},
\label{eq:Pjet}
\end{equation}
where $L_{\rm B}$, $L_{\rm p}$, $L_{\rm e}$, and $L_{\rm rad}$ are
the power carried by the magnetic field, the cold protons, the
relativistic electrons, and the produced radiation, respectively.
We obtain a value of $P_{\rm jet}$ of $3.2 \times 10^{46}$\,erg\,s$^{-1}$,
$3.7 \times 10^{46}$\,erg\,s$^{-1}$, and $2.5 \times 10^{46}$\,erg\,s$^{-1}$
for SED1, SED2 and SED3, respectively.
The total power of the jet is
lower at the end of the AGILE{} observing period, following the general
trend of the multiwavelength light curves.
\begin{deluxetable}{lrrrr}
\tablecolumns{5}
\tabletypesize{\normalsize}
\tablecaption{Input parameters for the model of SED1, SED2, and SED3.
See Section~\ref{3c454:discu:jet} for details. \label{3c454:tab:sed:param}}
\tablewidth{0pt}
\tablehead{
\colhead{Parameter} & \colhead{SED1} & \colhead{SED2} & \colhead{SED3}
& \colhead{Units}}
\startdata
$\alpha_{\rm l}$ & 2.3 & 2.5 & 2.0 & \\
$\alpha_{\rm h}$ & 4.0 & 4.0 & 4.2 & \\
$\gamma_{\rm min}$ & 30 & 30 & 18 & \\
$\gamma_{\rm b}$ & 300 & 280 & 180 & \\
$K$ & 80 & 80 & 100 & cm$^{-3}$ \\
$R$ & 21.5 & 21.5 & 21.5 & 10$^{15}$\,cm\\
$B$ & 2 & 2 & 2 & G\\
$\delta$ & 34 & 34 & 34 & \\
$L_{\rm d}$ & 5 & 5 & 5 & 10$^{46}$\,erg\,s$^{-1}$\\
$r_{\rm d}$ & 0.015 & 0.015 & 0.015 & pc\\
$\Theta_{0}$ & 1.15 & 1.15 & 1.15 & degrees\\
$\Gamma$ & 20 & 20 & 20 & \\
$P_{\rm jet}$ & 3.2 & 3.7 & 2.5 & 10$^{46}$\,erg\,s$^{-1}$\\
\enddata
\end{deluxetable}
\subsection{Jet geometry} \label{3c454:discu:jet}
The light curves in Figure~\ref{3c454:fig:R:mm:GRID} show
a different behavior starting from the end of 2007 among the
different energy bands.
As shown in \cite{Villata2009:3C454:GASP:accep}, a possible interpretation
arises in the framework of a change in orientation of a curved jet,
yielding different alignment configurations within the jet itself.
During 2007, the more pronounced fluxes and variability of the optical
and $\gamma$-ray bands seem to favor the inner portion of the jet as the
more beamed one. On the other hand, the dimming trend in the optical
and in the $\gamma$-ray bands, the higher mm flux emission and its enhanced
variability during 2008, seem to indicate that the more extended
region of the jet became more aligned with respect to the observer
line of sight.
\section{Conclusions} \label{3c454:conc}
The AGILE{} high-energy long-term monitoring of the blazar \hbox{3C~454.3}{} allowed us
to organize a few multiwavelength campaigns, as well as monitoring
programs at lower frequencies, over a time period of about eighteen
months. Thus, we were able to investigate the spectral energy
distributions over several decades in energy, to study the interplay
between the $\gamma$-ray and the optical fluxes, and the physical properties
of the jet producing the non-thermal radiation.
The {\it global} view we obtained after one and a half years of
observations can be summarized as follows:
\begin{enumerate}
\item The $\gamma$-ray emission for energy $E>100$\,MeV is clearly highly
variable, on time scales of the order of one day or even shorter,
with prominent flares reaching, on a day time scale, the order of
magnitude of the Vela pulsar emission, the brightest, persistent
$\gamma$-ray source above 100~MeV.
\item Starting from October 2008, \hbox{3C~454.3}{} entered a prolonged
mid- to low-level $\gamma$-ray phase, lasting several months.
\item Emission in the optical range appears to be weakly correlated with that
at $\gamma$-ray energies above 100 MeV, with a lag (if present) of the $\gamma$-ray
flux with respect to the optical one less than one day.
\item While at almost all frequencies the flux shows a diminishing trend with time,
the 15~GHz radio core flux increases, although no
new jet component seems to be detected.
\item The average $\gamma$-ray spectrum during the different observing
campaigns seems to show an harder-when-brighter trend.
\item Our results support the idea the dominant emission mechanism
in $\gamma$-ray energy band is the inverse Compton scattering of external
photons from the BLR clouds scattering off the relativistic electrons
in the jet.
\item The different behavior of the light curves at different wavelengths could be
interpreted by a changing of the jet geometry between 2007 and 2008.
\end{enumerate}
The simultaneous presence of two $\gamma$-ray satellites, AGILE{} and
{\it Fermi}{}, the extremely prompt response of wide-band satellites such as
{\it Swift}{}, and the long-term monitoring provided from the radio to the
optical by the GASP-WEBT Consortium will assure the chance to
investigate and study the physical properties of several blazars both
at high and low emission states.
\acknowledgements
The AGILE Mission is funded by the Italian Space Agency (ASI) with
scientific and programmatic participation by the Italian Institute
of Astrophysics (INAF) and the Italian Institute of Nuclear
Physics (INFN). This investigation was carried out with partial
support under ASI contract N. {I/089/06/1}. We acknowledge financial
support by the Italian Space Agency through contract
ASI-INAF I/088/06/0 for the Study of High-Energy Astrophysics.
This work is partly
based on data taken and assembled by the
WEBT collaboration and stored in the WEBT archive at the Osservatorio
Astronomico di Torino - INAF\footnote{\texttt{http://www.to.astro.it/blazars/webt/}}.
We thank the {\it Swift}{} and RXTE Teams for making these observations possible,
particularly the duty scientists and science planners.
This research has made use of data from the MOJAVE database
that is maintained by the MOJAVE team.
The 70-cm meniscus observations was partially supported by
Georgian National Science Foundation grant GNSF/ST-08/4-404.
E.K. acknowledges support from the NCS grant No. 96-2811-M-008-033.
K.\ Sokolovsky was supported by the International Max Planck Research School
(IMPRS) for Astronomy and Astrophysics at the universities of Bonn and
Cologne.
We thank L. Fuhrmann for collaborating in the setup of the
AGILE{}/{\it Fermi}{} {\it Swift}{} monitoring campaign in the period
August--October 2008.
We also the Referee for his/her constructive comments.
{\it Facilities:} \facility{AGILE}, \facility{{\it Swift}}, \facility{RXTE},
\facility{WEBT}, \facility{VLBA}, \facility{UMRAO}.
|
\section{Introduction}
Quantum fluctuations of surfaces are important in many physical phenomena. It is
less known these fluctuations can sometimes be expressed in term of a
functional integral over reparametrizations of variables relevant in the
Feynman path integral. In this paper we shall consider the expression
\begin{equation}
W(C)\equiv \int {\cal D}\theta\,{\,\rm e}\,^{-KA[x(\theta)]},
\label{ansatz}
\end{equation}
where $x(\theta)$ is the boundary curve $C$ and where
\begin{equation}
A[x(\theta)]=\frac{1}{8\pi}~\int_0^{2\pi}{\rm d}\sigma \int_0^{2\pi}{\rm d}\sigma' ~
\frac{\left(x(\theta(\sigma))-x(\theta(\sigma'))\right)^2}
{1-\cos(\sigma-\sigma')}.
\label{D}
\end{equation}
The functional (\ref{D}) is known in mathematics as the Douglas integral
\cite{Dou31}, whose minimum with respect to variations of the
reparametrizations $\theta$ gives the minimal area
\begin{equation}
A\left[x(\theta_\star)\right]=S_{\rm min} (C) .
\label{p3}
\end{equation}
Here $\theta_\star(\sigma)$ is the saddle point of the integral
(\ref{ansatz}).
The functional integration (\ref{ansatz})
thus gives the area law to leading order.
The path integral over reparametrizations (\ref{ansatz}) was introduced
in this context in Ref.~\cite{cohen} in connection with an off-shell string
propagator. More recently it was proposed by Polyakov \cite{Pol97} as
an ansatz for the Wilson loop in large $N$ QCD. The leading behavior obviously
gives the leading behavior of the Wilson loop, found in most string models,
where the bulk field $X^\mu(\tau,\sigma)$ satisfies the Dirichlet
boundary condition
\begin{equation}
X^\mu(\tau,\sigma)|_{\rm boundary}=x^\mu (\sigma).
\label{p4}
\end{equation}
To derive Eq.~(\ref{p3}) from the Nambu-Goto action one can follow
Douglas \cite{Dou31}, or the more recent elegant paper by Migdal
\cite{Mig94}.
The functional integral (\ref{ansatz}) can be expanded around the
saddle point $\theta_\star$. With $\theta(\sigma)=\theta_\star(\sigma)+
\beta (\sigma)$ we obtain
the first non-vanishing contribution
\begin{equation}
A_2=\frac{K}{8\pi}~\int_0^{2\pi}{\rm d}\theta\int_0^{2\pi}{\rm d}\theta'\,{\dot x}
(\theta)\cdot{\dot x}(\theta')\frac{\left(\beta(\theta)-\beta(\theta')\right)^2}
{1-\cos (\sigma_\star (\theta)-\sigma_\star (\theta'))}
\label{p5}
\end{equation}
Here $\sigma_\star (\theta)$ is the inverse function of $\theta_\star(\sigma)$.
Some dynamical consequences of Eq.~(\ref{p5}) have been discussed by Rychkov
\cite{Ryc02} and by the authors \cite{MO09}.
However, it is perhaps fair to say
that the physical meaning of the fluctuation integral (\ref{p5}) is not
so clear.
In this paper we shall show that the leading part of the fluctuations of
large loops from
(\ref{p5}) are {\it transverse fluctuations of the minimal surface} embedded
by the curve $x(\theta_\star)$. This we have shown for a rectangle and an
ellipse, but we suspect the result to be more general.
To explain our result we mention that in the
Nambu-Goto action transverse fluctuations add a contribution
\begin{equation}
-\frac{d-2}{2}~{\rm tr ~log}~(-\partial^2)=\frac{(d-2)\pi}{24}~\frac{T}{R}
\label{lut}
\end{equation}
to the area term
for a large $R\times T$ rectangle with $T\gg R$. These quantum fluctuations
(\ref{lut}) are called the L\"uscher term \cite{luscher}. By lattice
Monte Carlo
calculations in three and four dimensions, this term has been found to occur
\cite{monte} in quenched $SU(N)$ for various $N'$s. Thus, for large distances
the two leading terms from the Nambu-Goto action describe QCD quite
well. Therefore, for a rectangular boundary curve the $T/R$ term
can be identified with transverse fluctuations of the {\it minimal surface}.
With these remarks in mind we now give our main result: {\it after
integration over the fluctuations $\beta$ in the functional integral }
\begin{equation}
\int {\cal D}\beta{\,\rm e}\,^{-A_2}
\label{p7}
\end{equation}
{\it we obtain as the leading contribution the L\"uscher term corresponding to
d=}26. In general, we obtain the $d$-dimensional contribution (plus the area
term) from
\begin{equation}
\int {\cal D}\theta {\,\rm e}\,^{-KA[x(\theta)]}~(\det O)^{-(d-26)/48}.
\label{p8}
\end{equation}
Here $O$ is the operator which emerges in the semiclassical expansion of
$A\left[x(\theta)\right]$ to quadratic order as exhibited in Eq.~(\ref{p5}).
This modification of Eq.~(\ref{ansatz}) does
not effect the classical limit leading to the area law, and has the meaning
of a pre-exponential in the semiclassical approximation.
Thus, this equation gives the leading effective QCD string behavior in terms
of functional integration over reparametrizations.
We mention that our considerations may have potential applications
in condensed matter. For example, in the inverse square XY model
one encounters expressions somewhat similar to Eq.~(\ref{p5}), see
for example Ref.~\cite{hof}. We shall,
however, not pursue this track in the present paper.
The plan of this paper is the following: in Sect.~\ref{s:2}
we discuss the general framework for the semiclassical approximation,
and in Sect.~\ref{s:3} we carry out the
functional integral over reparametrizations for a rectangle. A similar
calculation for an ellipse is done in Sect.~\ref{s:4}.
In Sect.~\ref{s:5} we derive the generalization given by Eq.~(\ref{p8})
and in Sect.~\ref{s:6} we make some
conclusions. A number of more technical points
have been discussed in some Appendices.
\section{A semiclassical correction\label{s:2}}
The path integral over reparametrizations~\rf{ansatz} reproduces the
exponential of the minimal area in the classical limit
$K S_{\rm min}\to\infty$.
To calculate the semiclassical correction,
we expand
\begin{equation}
\theta(\sigma)=\theta_*(\sigma)+\beta(\sigma)
\label{expa}
\end{equation}
or
\begin{equation}
\sigma(\theta)=\sigma_*(\theta)+\beta(\theta) \,,
\label{expa1}
\end{equation}
where
\begin{equation}
\beta(0)=\beta(2\pi)=0
\label{bbcb}
\end{equation}
and expand the Douglas integral to
quadratic order in $\beta$ around the classical trajectory $\theta_*(\sigma)$.
This expansion makes sense because typical trajectories in the path
integral over reparametrizations~\rf{ansatz} are smooth as
$K S_{\rm min}\to\infty$ and have Hausdorff dimension one~\cite{BM09}.
Substituting~\rf{expa} into Douglas' integral~\rf{D} and expanding in $\beta$,
we find that the linear term vanishes
because $\theta_*(\sigma)$ is the minimum, while the quadratic part
reads
\begin{equation}
A_2[\beta(\theta)]
=\frac{K}{8\pi}\int_0^{2\pi}{\rm d}\theta_1\int_0^{2\pi}{\rm d}\theta_2\,{\dot x}
(\theta_1)\cdot{\dot x}(\theta_2)\,\frac{\left(\beta(\theta_1)-\beta(\theta_2)
\right)^2}
{1-\cos (\sigma_\star (\theta_1)-\sigma_\star (\theta_2))}\,.
\label{S2}
\end{equation}
The function $\beta(\theta)$ obeys
\begin{equation}
\dot \beta (\theta) \geq -\dot\sigma_*(\theta)
\label{restriction}
\end{equation}
for the derivative of the reparametrizing function to be positive.
This is always satisfied if $\beta$ is small and smooth enough.
In order to calculate the semiclassical correction to the area law, we need
to do the Gaussian integral
\begin{equation}
\int {\cal D}\beta(\theta) {\,\rm e}\,^{-A_2[\beta(\theta)]}
\label{II2}
\end{equation}
with $A_2[\beta(\theta)]$ given by \eq{S2}.
The typical values of $\beta$, which
are essential in the path integral over $\beta$ in \rf{II2},
are $\beta\sim 1/\sqrt{K S_{\rm min}}$, i.e.
small for $\sqrt{K S_{\rm min}} \gg1$.
Hence, the higher terms of an expansion
of $A[\theta_*(\sigma)+\beta]$ in
$\beta$ are suppressed~\cite{Ryc02} at large $\sqrt{K S_{\rm min}} $.
The loop expansion goes in the parameter $1/K S_{\rm min}$ and only one loop
contributes with the given accuracy.
A comment is needed about the measure for the integration over $\beta(\theta)$.
As is explained in Appendix~\ref{appA}, the measure for the integration
over $\sigma(\theta)$ involves a factor
\begin{equation}
\frac{1}{\sigma'}=\frac{1}{\sigma'_*+\beta'}=
\frac{1}{\sigma'_*}-\frac{\beta'}{(\sigma'_*)^2}+
\frac{(\beta')^2}{(\sigma'_*)^3}+
{\cal O}\left((\beta')^3\right).
\end{equation}
Because $\beta\sim 1/\sqrt{K S_{\rm min}}$, the second and third terms
on the right-hand side are not essential to the given order.
Therefore, the measure for the path integration over $\beta(\theta)$
in \eq{II2} is the usual one for smooth functions $\beta(\theta)$, while
this factor will show up to the next order in $ 1/{K S_{\rm min}}$.
\section{Path integral over reparametrizations: rectangle\label{s:3}}
In this section we show how the path integral over
reparametrizations captures the L\"{u}scher term for a rectangle.
The conformal map of the upper half plane onto the interior of
a rectangle is given by the Schwarz--Christoffel mapping
\begin{equation}
\omega =A F\left(\frac{z}{\sqrt{\mu}},\mu\right)
-{\rm i} \frac{AK\left(\sqrt{1-\mu^2}\right)}2,
\end{equation}
where
\begin{equation}
F\left(\frac{z}{\sqrt{\mu}},\mu\right)=\int_0^z
\frac{{\rm d} x}{\sqrt{\mu-x^2}\sqrt{1-\mu x^2}}
\end{equation}
is the incomplete elliptic integral of the first kind.
The two parameters $A$ and $\mu$ are related to the coordinates
of the vertices of the rectangle by
\begin{equation}
A K(\mu)=\frac R2,\qquad A K\left(\sqrt{1-\mu^2}\right)=T,
\label{RandT}
\end{equation}
where $K(\mu)=F(1,\mu)$ is the complete elliptic integral of the first kind.
In deriving \eq{RandT} we used the important identity
\begin{equation}
F(1/\mu,\mu)=K(\mu)+{\rm i} K\left(\sqrt{1-\mu^2}\right).
\end{equation}
Equations~\rf{RandT} relates $\mu$ to the ration of $R/T$ as
\begin{equation}
\frac{2T}{R}=\frac{ K\left(\sqrt{1-\mu^2}\right)}{K(\mu)}.
\label{muvsT/R}
\end{equation}
In the limit $T/R\to\infty$, when $\mu\to0$, this equation simplifies to
\begin{equation}
\pi \frac{T}R=\ln \frac4{\mu}\, .
\label{muto0}
\end{equation}
When $z=s$ runs along the real axis, the variable $\omega$ runs along
the boundary of the rectangle with $s=-1/\sqrt{\mu}, -\sqrt{\mu},
+\sqrt{\mu}, +1/\sqrt{\mu}$ ($\mu<1$)
mapped, respectively, onto the vertices of the rectangle:
$(-R/2,T/2),(-R/2,-T/2),(R/2,-T/2), (R/2,T/2)$. The given choice of the
argument of the mapping preserves the symmetry
$s\to1/s$.
When $z$ has positive imaginary part,
the coordinates
\begin{equation}
X_1(z)=A\, \Re F\left(\frac{z}{\sqrt{\mu}},\mu\right),\qquad
X_2(z)= A\,\Im F\left(\frac{z}{\sqrt{\mu}},\mu\right)-
\frac{AK\left(\sqrt{1-\mu^2}\right)}2
\label{KSmap}
\end{equation}
take their values inside the rectangle. These coordinates are conformal.
For this reason we have
\begin{equation}
x_1(t_*(s))=A \,\Re F\left(\frac{s}{\sqrt{\mu}},\mu\right),\qquad
x_2(t_*(s))= A\,\Im F\left(\frac{s}{\sqrt{\mu}},\mu\right)-
\frac{AK\left(\sqrt{1-\mu^2}\right)}2,
\label{Crecta}
\end{equation}
whose implementation for the function $t_*(s)$ is discussed below.
The boundary contour given by \eq{Crecta} satisfies Douglas'
minimization (see Appendix~\ref{appB}, \eq{Dou1}).
Correspondingly, the Douglas integral
\begin{equation}
\frac 1{4\pi}
\int\nolimits_{-\infty}^{+\infty} {\rm d} s
\int\nolimits_{-\infty}^{+\infty} {\rm d} s' \,\frac{ \left[
x(t_*(s_1))- x(t_*(s_2))\right]^2}{(s-s')^2} =
2 A^2 K(\mu) K\left(\sqrt{1-\mu^2}\right) =R T
\label{71}
\end{equation}
as it should. We have verified these two equations numerically.
A natural parametrization of the boundary of a rectangle is through
$\tau\in S^1$:
\begin{subequations}
\begin{eqnarray}
x_1= \frac T2 \tan \tau \,,~~~ x_2= -\frac T2\qquad&&
- \arctan \frac R T\leq\tau\leq \arctan \frac R T \\*
x_1= \frac R2 \,,~~~ x_2= -\frac R2 \cot \tau \qquad&&
\arctan \frac R T \leq\tau\leq \pi-\arctan \frac R T \\*
x_1= \frac T2 \tan \tau \,,~~~ x_2= \frac T2\qquad&&
\pi- \arctan \frac R T\leq\tau <\pi
\end{eqnarray}
\label{xtau}
\end{subequations}
and analogously for negative $\tau$. Introducing
\begin{equation}
t=\tan\frac{\tau}{2},
\end{equation}
we rewrite \eq{xtau} as
\begin{subequations}
\begin{eqnarray}
x_1= T \frac{t}{1-t^2}\,,~~~ x_2= -\frac T2\qquad&&
- \frac {\sqrt{T^2+R^2}-T}R\leq t\leq \frac {\sqrt{T^2+R^2}-T}R \\*
x_1= \frac R2 \,,~~~ x_2= R \frac{t^2-1}{4t} \qquad&&
\frac {\sqrt{T^2+R^2}-T}R\leq t\leq \frac {\sqrt{T^2+R^2}+T}R \\*
x_1= T \frac{t}{t^2-1}\,,~~~ x_2= \frac T2\qquad&&
\frac {\sqrt{T^2+R^2}+T}R\leq t <+\infty.
\end{eqnarray}
\label{xt}
\end{subequations}
\indent To relate $t$ to $s$, we identify
\begin{subequations}
\begin{eqnarray}
\frac{R}{2K(\mu)} F\left(\frac{ s}{\sqrt{\mu}},\mu\right)
&=& T \frac{t}{1-t^2}\,,\label{I1}\\* \hbox{for}~~
-\sqrt{\mu} \leq s \leq \sqrt{\mu}~~&\hbox{or}&~~
- \frac {\sqrt{T^2+R^2}-T}R\leq t\leq \frac {\sqrt{T^2+R^2}-T}R
\nonumber \\*
\frac{R}{2K(\mu)} \int^{s}_{\sqrt{\mu}}
\frac{{\rm d} x}{\sqrt{x^2-\mu}\sqrt{1-\mu x^2}}- \frac T2&=& R \frac{t^2-1}{4t}
\label{I2}\\*
\hbox{for}~~\sqrt{\mu} \leq s \leq 1/\sqrt{\mu}~~&\hbox{or}&~~
\frac {\sqrt{T^2+R^2}-T}R\leq t\leq \frac {T+\sqrt{T^2+R^2}}R
\,. \nonumber
\end{eqnarray}
\label{st}
\end{subequations}
\hspace*{-1mm}Solving the quadratic equation for $t$ versus $s$,
we obtain the minimizing function $t_*(s)$, which obviously obeys
the boundary condition
\begin{equation}
t(0)=0.
\label{bbc}
\end{equation}
The symmetry $ s\to 1/ s$ plays apparently an important role.
It guarantees that the points $-\infty$, $-1$, $0$, $+1$, $+\infty$
are mapped onto themselves under the reparametrization
$s\to t_*(s)$: $t_*(-\infty)=-\infty$, $t_*(-1)=-1$, $t_*(0)=0$,
$t_*(+1)=+1$, $t_*(+\infty)=+\infty$.
It is convenient to invert \eq{st} using the Jacobi elliptic functions.
Inverting \eq{I1}, we get
\begin{eqnarray}
{s}&=&{\sqrt{\mu}}\,
{\rm sn}\left(\frac{2K(\mu)T}{R}\,\frac{t}{1-t^2},\mu \right),
\label{3}
\\* \hbox{for}~~
-\sqrt{\mu} \leq s \leq \sqrt{\mu}&~~&\hbox{or}~~
- \frac {\sqrt{T^2+R^2}-T}R\leq t\leq \frac {\sqrt{T^2+R^2}-T}R \,.
\nonumber
\end{eqnarray}
The function sn has a nice trigonometric expansion in the parameter (nome)
(Ref.~\cite{GR}, 8.146.1)
\begin{equation}
\exp{\left(-\pi \frac{K(\sqrt{1-\nu^2})}{K(\nu)}\right)}\approx (\mu/4)^2\ll1,
\label{5}
\end{equation}
giving
\begin{equation}
s\approx\sqrt{\mu} \sin \left( \frac{\pi T}{R} \frac{t}{(1-t^2)}\right).
\label{tts1}
\end{equation}
This formula is applicable
for $-R/2T <t<+ R/2T$, when $-\sqrt{\mu}<s<+\sqrt{\mu}$.
We can proceed in the same way with \eq{I2}, whose inverse is
\begin{equation}
s=\sqrt{\mu}\,{\rm sn}\left(K(\mu)+{\rm i} K(\mu)\left(\frac{t^2-1}{2t}+
\frac{T}{R}\right),\mu\right).
\label{7}
\end{equation}
Using the addition formula (\cite{GR}, 8.156.1)
and the reduction of ${\rm sn}(x,0)$ and
${\rm cn}(x,0)$ to $\sin x$ and $\cos x$, we obtain
\begin{equation}
s\approx \sqrt{\mu} \cosh \left[\frac\pi 4 \left(
t-t^{-1}+\frac{2T}R \right) \right].
\label{nsss}
\end{equation}
However, this expansion is useless for large $t \to 2T/R$,
due to the imaginary part of the argument of sn the expansion will
involve hyperbolic functions with arguments that can be large.
Instead we can use the expansion of sn in terms
of inverse sines (\cite{GR}, 8.147.1), where these sines can be large, so
only the first term is relevant:
\begin{equation}
s\approx \frac
1{\displaystyle{\sqrt{\mu}\,\sin \left(\frac{\pi}{2}+{\rm i}\frac{\pi}{4}\left(
\frac{t^2-1}{t}-\frac{2T}{R}\right)\right)}}=
\frac{1}{\displaystyle{\sqrt{\mu} \cosh \left[\frac\pi 4 \left(
t-t^{-1}-\frac{2T}R \right) \right]}}\,.
\label{sss}
\end{equation}
Equation~\rf{nsss} is applicable for $t\to R/2T$ from above and
\eq{sss} is applicable when $t\to 2T/R$.
The quadratic action, describing Gaussian fluctuations around $t_*(s)$,
is like \rf{S2} rewritten for the real-axis parametrization:
\begin{equation}
A_2\left[\beta(t)\right] = \frac K{4\pi} \int{\rm d} t_1 \int {\rm d} t_2
\frac{\dot x(t_1)\cdot \dot x(t_2)}{(s_*(t_1)-s_*(t_2))^2 }
\left[\beta(t_1)-\beta (t_2)\right]^2.
\label{r-aA2}
\end{equation}
The Gaussian approximations is justified for large $K RT$, when
\begin{equation}
\beta(t)\sim \frac1{\sqrt{K R T}}.
\end{equation}
To calculate the path integral over the quantum fluctuations
around the minimizing function $t_*(s)$, we need a mode expansion of
the infinitesimal reparametrizing function $\beta(t)$.
To get rid of the projective symmetry, we keep fixed 3 points, e.g.\
$-1$, $0$, $+1$ or $-\sqrt{\mu}$, $0$, $\sqrt{\mu}$ fixed:
$\beta(-1)=0$, $\beta(0)=0$, $\beta(+1)=0$ or
$\beta(-t_*(-\sqrt{\mu}))=0$, $\beta(0)=0$, $\beta(t_*(\sqrt{\mu}))=0$.
For each segment from $t_i$ to $t_f$, we consider the mode
expansion
\begin{equation}
\beta(t)=\sum_n c_n \sin \left( \pi n \frac{t-t_i}{t_f-t_i} \right),
\end{equation}
obeying the boundary condition $\beta(t_i)=\beta(t_f) =0$.
This set of sines forms a complete basis on the given interval.
Actually we shall need the mode expansion for 4 segments attached to
$t=t_*(\pm \sqrt{\mu})\approx\pm R/2T$ because the large contribution of
$1/\mu$ will appear in $A_2$ only for those.
The appearance of the large factor $1/\mu$ is seen already
from Eqs.~\rf{3}, \rf{7} because $s_*$ in the denominator in \eq{r-aA2} is
proportional to $\sqrt{\mu}$.
We need, however, to show that this $1/\mu$ is multiplied by a factor
$\sim 1$.
Let us analyze the contribution to \rf{r-aA2} that comes
from $-R/2T<t_1,t_2<+ R/2T$, i.e.\ from the bottom side of the rectangle.
Introducing the variable
\begin{equation}
y=\frac{2T}R t \,, \qquad -1< y <1
\end{equation}
and using \eq{tts1}, we write the contribution of this domain to $A_2$ as
\begin{equation}
\frac {K T^2}{4\pi \mu} \int_{-1}^1{\rm d} y_1 \int_{-1}^1 {\rm d} y_2\,
\frac{\left(\beta (y_1)-\beta(y_2)\right)^2}{
\left(\sin(\pi y_1/2)-\sin(\pi y_2/2)\right)^2 } \propto \frac{T}{\mu R} C
\label{AA2}
\end{equation}
with a positive constant $C$.
A similar contribution appears
if $t_1-R/2T\sim t_2-R/2T\sim R^2/(2T)^2$ as is prescribed by \eq{nsss}
for both $t_1>R/2T$ and $t_2>R/2T$. An analogous
contribution (with possibly some powers of $T/R\propto\ln (4/\mu)$) emerges
also when $t_1< R/2T$ and $ t_2> R/2T $
or $t_1>R/2T$ and $ t_2< R/2T $.
Therefore, the path integral over the quantum fluctuations
around the minimizing function $t_*(s)$ gives, using the $\zeta$-function
regularization and \eq{muto0}
(and disregarding the logs to the order in consideration)
\begin{equation}
\prod_{\rm modes} {\sqrt{\mu}} \propto \left(\frac1
{\sqrt[4]{\mu}}\right)^4 \propto \exp{\left(\frac{\pi T}{R} \right)},
\end{equation}
where the 4th power is due to the four sets of modes%
\footnote{This is like
in the computation of the static potential for the Polyakov string in
Ref.~\cite{JM93}.}.
This remarkably reproduces the L\"uscher term~\rf{lut}
in $d=26$!
There was a subtlety in the derivation -- the appearance of a
logarithmic divergence at the corners of the rectangle
if $\dot \beta({\rm corners})\neq0$.
It is seen from \eq{AA2}, where the region near $y_1=y_2=1$ or $y_1=y_2=-1$
(associated with $t_1=t_2\to R/2T$ or $t_1=t_2\to -R/2T$ )
produces the logarithmic divergence
\begin{equation}
\int_{-1+\delta}^{1-\delta}{\rm d} y_1 \int_{-1+\delta}^{1-\delta} {\rm d} y_2\,
\frac{\dot\beta^2 (\pm1)}{
\left(y_1+y_2\mp2\right)^2 } = \dot\beta^2 (\pm1)\ln \frac 1\delta
\label{AAAA2}
\end{equation}
with the upper (lower) sign referring to $y=1$ ($y=-1$).
The coefficient is nonvanishing if $\dot \beta(\pm1)\neq0$.
Analogously, the integral is logarithmically divergent at the corner, when
$t_1,t_2> R/2T$ or $t_1< R/2T$, $t_2> R/2T$
and vise versa.
The logarithmic divergence can be regularized by smoothing
the corners like in Refs.~\cite{luscher,DOP84}.
It is clear from such a regularization that
the contribution of trajectories with $\dot \beta(R/2T)\neq0$
to the path integral over $\beta(t)$ will be suppressed as the
smoothing is removed. Consequently,
this corner divergence does not effect the
result of this section. In the next section we repeat the
consideration for the case of an ellipse, when there are no corners.
If $d\neq 26$, the asymptotic ansatz for the Wilson loops has
to be improved to get the correct factor $(d-2)/24$
in the L\"uscher term~\rf{lut}.
This issue will be described in Sect.~\ref{s:5}.
\section{Path integral over reparametrizations: ellipse\label{s:4}}
In this section we evaluate the path integral over
reparametrizations for an ellipse and obtain a prediction
for the associated L\"{u}scher term.
The necessary formulas are given in Appendix~B of \cite{MO09} and are partially
reproduced in Appendix~B.
We are interested in
the case of a very long ellipse when the ratio $b/a\to0$ and
$\nu\to1$ according to
\begin{equation}
\ln \frac{a+b}{a-b}=\frac{\pi K(\sqrt{1-\nu^2})}{2K(\nu)}\,.
\label{goerlish}
\end{equation}
Using the asymptote
\begin{equation}
K(\nu)\stackrel{\nu\to1}\to \frac 12 \ln \frac 8{(1-\nu)},
\end{equation}
we simplify \eq{goerlish} to
\begin{equation}
\frac ba =\frac {\pi^2}{4\ln \frac 8{(1-\nu)}}.
\end{equation}
For $\nu\to1$ the elliptic function simplifies and we have
\begin{equation}
\theta_*(\sigma)=
\pi \left(\frac{\ln \frac{2\sigma+\sqrt{4\sigma^2+(1-\nu)^2}}{8}}
{\ln \frac8{(1-\nu)}} +1\right)
\label{approxi}
\end{equation}
for $-\pi/2<\sigma<\pi/2$.
Inverting \eq{approxi}, we find
\begin{equation}
\sigma_*(\theta)=\frac{1-\nu}2
\sinh\left(\frac{2\theta}{\pi}\ln\frac8{1-\nu}\right).
\label{inver}
\end{equation}
This is quite similar to Eqs.~\rf{tts1} and \rf{nsss} for a rectangle
with $\sqrt{\mu}$ replaced by $(1-\nu)$.
The calculation of the path integral over reparametrizations at one loop
is quite analogous to that for the rectangle described in the previous section.
We see that the large factor of $(1-\nu)^{-2}$ emerges
in \eq{S2} because $\sigma_*\propto (1-\nu)$ from \eq{inver}.
To evaluate the coefficient, let us consider the domain of small
$\theta_1$ and $\theta_2$:
\begin{equation}
\theta_1,\theta_2 \ll \left(\ln \frac{8}{1-\nu}\right)^{-1}\,,
\end{equation}
which contributes to the integral in \rf{S2}
\begin{equation}
\frac{K}{(1-\nu)^2} \int{\rm d}\theta_1 \int {\rm d}\theta_2
\left(a^2 \theta_1 \theta_2+b^2\right)
\frac{\dot\beta^2 (\theta_1)(\theta_2-\theta_1)^2}{
(\theta_2-\theta_1)^2 \ln^2 \frac{8}{1-\nu}}
\propto \frac{1}{(1-\nu)^2}
\end{equation}
modulo the powers of $b/a \propto \left( \ln \frac8{(1-\nu)} \right)^{-1}$.
The same contribution comes also from
the domain of both $\theta_1$ and $\theta_2$ near $\pi$. We thus have
\begin{equation}
A_2 \propto \frac{1}{(1-\nu)^2}
\end{equation}
for every mode.
Integrating the Gaussian integral for every mode and
using the $\zeta$-function regularization, we get
(disregarding the logs to this order) a pre-factor of the type
\begin{equation}
\prod_{\rm modes} {(1-\nu)}\propto \left(\frac1
{\sqrt{1-\nu}}\right)^4 \propto \exp{\left(\frac{\pi^2}{2}\frac{a}b \right)},
\label{Lue}
\end{equation}
where the product runs over 4 sets of modes, which results in
a factor of 4 in the exponent.
This coincides with the L\"uscher term~\rf{lut} for a
rectangle of the size $R\times T$ in $d=26$ dimensions
provided%
\footnote{It is worth noting that the Wilson loops for a rectangle
and ellipse then coincide~\cite{ABG81} in $d=4$ (the only dimension with a
$T/R$ Coulomb term) to the second order
of perturbation theory.}
\begin{equation}
\frac T R =\frac {\pi a}{2b}.
\label{provided}
\end{equation}
In Appendices \ref{appC} and \ref{appD} we confirm this by
an explicit calculation of the determinant of the Laplace operator
and the conformal anomaly for an outstretched ellipse.
\section{A generalization to arbitrary dimensions\label{s:5}}
The results of two previous sections demonstrate the already
mentioned fact that the ansatz \rf{ansatz} has to be modified in order to
describe the L\"uscher term in $d=4$ dimensions.
A simple modification is based on the form
of the path integral for a rectangle
\begin{equation}
\int {\cal D}\beta(t) {\,\rm e}\,^{-A_2[\beta(t)]}
\stackrel{T\gg R}\propto {\,\rm e}\,^{\pi T/R}
\label{III2}
\end{equation}
with quadratic action $A_2[\beta(t)]$ given by \eq{r-aA2}.
For an arbitrary curve this path integral can be expressed through
the determinant of the corresponding operator, that enters $A_2$,
which we denote as
${O}$:
\begin{equation}
\int {\cal D}\beta(t) {\,\rm e}\,^{-A_2[\beta(t)]} =
\left( \det {O} \right)^{-1/2}.
\label{OO2}
\end{equation}
It is now clear that the following modification of
the ansatz~\rf{ansatz} will provide the correct value of the
L\"uscher term~\rf{lut} in $d$ dimensions:
\begin{equation}
\int {\cal D}t {\,\rm e}\,^{-KA[x(t)]}~(\det O)^{-(d-26)/48}.
\label{pp8}
\end{equation}
This modification of Eq.~(\ref{ansatz}) does
not effect the classical limit, leading to the area law, and has the meaning
of altering a pre-exponential in the semiclassical approximation.
To make the structure of the operator ${O}$ more explicit, it is convenient
to use the expansion \rf{expa} of the direct function
$\theta(\sigma)$ rather than that \rf{expa1} of the inverse function
$\sigma(\theta)$ as above.
Using the identity
\begin{equation}
\beta(t_*(s))=-\frac{1}{\frac{{\rm d} t_*(s)}{{\rm d} s}} \,\beta(s)\,,
\end{equation}
which stems from the definitions~\rf{expa} and \rf{expa1},
we then obtain for the real-axis parametrization:
\begin{equation}
A_2\left[\beta(s)\right] = \frac K{4\pi} \int{\rm d} s_1 \int {\rm d} s_2
\frac{\dot x(t_*(s_1))\cdot \dot x(t_*(s_2))}{(s_1-s_2)^2 }
\left[\beta(s_1)-\beta (s_2)\right]^2,
\label{AAA2}
\end{equation}
which determines the ``momentum'' (with respect to $s$) space operator
\begin{equation}
O(p_1,p_2)= \frac{K}{8\pi} \int {\rm d} q |q|
\Big(
2 \dot x(p_1+q)\cdot \dot x(-p_2-q)-\dot x(p_1-p_2+q) \cdot \dot x(-q)
- \dot x(q)\cdot \dot x(p_1-p_2-q)\Big)
\end{equation}
with
\begin{equation}
\dot x (p) \equiv \int {\rm d} s {\,\rm e}\,^{{\rm i} p s} \dot x(t_*(s))\,.
\end{equation}
We can finally substitute $t_*(s)$ by $t(s)$ in this formula without
changing the semiclassical approximation.
It is worth noting that in contrast to the Laplace operator
of Ref.~\cite{DOP84},
where the L\"uscher term was obtained for the Polyakov string, the present
operator $O$ lives in the boundary, which makes the construction nontrivial.
\section{Conclusions\label{s:6}}
The conclusion is that the reparametrization of the boundary curve involved in
Eq. (\ref{ansatz}) carries information on the transverse fluctuations in 26
dimensions. As is shown in the previous section,
it is possible to generalize this to any dimensions.
Our motivation for the present paper is our previous work on
the QCD/string scattering amplitudes \cite{MO08}, where we used that
the amplitude can be expressed in terms of a Wilson loop through
Feynman path integration. There we only
considered the leading area behavior. However, having developed a
path integral expression for the next term, we hope that the $x^\mu$
integrals can be performed, thereby providing a momentum space
analogue of the L\"uscher term.
We hope this may help to answer a very interesting question as to
how the intercept of the Regge trajectory changes under such
a modification of the ansatz for the Wilson loops.
\begin{acknowledgments}
We are indebted to Andrey Mironov and Niels Obers
for useful discussions.
\end{acknowledgments}
|
\section{Introduction}
These expository notes begin by briefly explaining two constructions of the Jones polynomial (neither of which is the original construction due to Jones \cite{Jones:1985}). The first is via the skien-theoretic construction of the Kauffman bracket \cite{Kauffman:1987}. The second is as a $U_q(\mathfrak{sl}_2)$ quantum group link invariant. This second construction uses a circle of idea developed by a number of authors starting in the late 1980s (see \cite{Turaev} and references therein). We attempt to give some explanation of how quantum group knot invariants work in general, but only fully develop the simplest case. We then discuss how the two constructions are related. Much of the content of these notes can be found in, for instance, \cite[Appendix H]{Ohtsuki}. The main difference here is that we use the non-standard ribbon element from \cite{half_twist}.
The Kauffman bracket is an isotopy invariant of framed links. The functor used in the quantum group construction involves a category where morphisms are tangles of \emph{directed} framed ribbons. In particular, endomorphisms of the trivial representations are \emph{directed} framed links, and the image of such a link is a Laurent polynomial, which is the invariant. In the case we consider, this invariant does not depend on the directing, and, up to an annoying sign, agrees with the Kauffman bracket. Part of the purpose of these notes is to explain the annoying sign, but the real purpose is to describe how the skein relations used in the Kauffman bracket arise naturally in the quantum group construction. To this end we modify the quantum group construction to obtain a functor from a category whose morphisms are tangles of \emph{undirected} framed ribbons. We find it is necessary to use the non-standard ribbon element from \cite{half_twist}. With this change, the annoying minus sign disappears, and the two constructions agree exactly!
The Jones polynomial is an invariant of directed but unframed links, which can be constructed via a simple modification of the Kauffman bracket (explained in Theorem \ref{Jones1} below). We actually compare constructions of invariants of framed but undirected links, so a more accurate subtitle for these notes might be ``two constructions of the Kauffman bracket."
\subsection{Acknowledgements}
These notes are based on a talk I gave in 2008 at the university of Queensland in Brisbane Australia. I would like to thank Murray Elder and Ole Warnaar for organizing that visit. I would also like to thank Noah Snyder for many interesting discussions about knot theory.
\section{The Kauffman bracket construction of the Jones polynomial} \label{Kauffman_bracket}
Up to a change in the variable $q$, the following is the well known construction of the Kauffman bracket \cite{Kauffman:1987}.
\begin{Definition} \label{Kauffman-simplifications}
Let $L$ be a link diagram (i.e. A link drawn as a curve is the plane with crossings). Simplify $L$ using the following relations until the result is a polynomial in $q^{1/2}$ and $q^{-{1/2}}$. That polynomial, denoted by $K(L)$, is the Kauffman bracket of the link diagram.
\begin{enumerate}
\setlength{\unitlength}{0.35cm}
\thicklines
\item \label{ocross_Kauffman}
\begin{picture}(15,2)
\put(1,0){
\begin{picture}(3,3)
\put(0,2){\line(1,-1){1.1}}
\put(2,0){\line(1,-1){1.1}}
\put(0,-1){\line(1,1){3}}
\end{picture}
}
\put(5,0.3){$=$}
\put(9,-1){\line(0,1){3}}
\put(12,-1){\line(0,1){3}}
\put(7,0.3){$q^{1/2}$}
\put(14,0.3){$+$}
\put(16,0.3){$q^{-1/2}$}
\put(20.5,-1){\oval(3,2)[t]}
\put(20.5,2){\oval(3,2)[b]}
\end{picture}
\vspace{0.5cm}
\item \label{circle_Kauffman}
\begin{picture}(10,3)
\put(4,0.4){\circle{2.8}}
\put(7.5,0.3){$=$}
\put(10,0.3){$-q-q^{-1}$}
\end{picture}
\vspace{0.5cm}
\item If two tangle diagrams are disjoint, their Kauffman brackets multiply.
\end{enumerate}
Note that \eqref{ocross_Kauffman} depends on which strand is on top.
\end{Definition}
\begin{Comment}
In order for Definition \ref{Kauffman-simplifications} to make sense, one needs to assume that all crossings are simple crossings. We will always make this assumption about link diagrams. Once we introduce twists, we will also assume that these occur away from crossings. It is clear that, up to isotopy, any link can be drawn with such a diagram (although not in a unique way).
\end{Comment}
The Kauffman bracket is not an isotopy invariant of links, but is instead an isotopy invariant of framed links (that is, links tied out of orientable ``ribbons"), where the framing is ``flat on the page." One can allow twists of the framing to occur in the diagram if one introduces the following
extra relation (here both sides represent a single framed string):
\begin{equation} \label{third-Kauff}
\setlength{\unitlength}{0.42cm}
\begin{picture}(6,2.5)
\put(0,0.9){\surtwist}
\put(0,-1.1){\usrtwist}
\put(2.5,0.2){$=$}
\put(3.5,0.2){$-q^{3/2} $}
\put(6,0){\uid}
\put(6,-2){\uid}
\end{picture}
\end{equation}
\vspace{1cm}
\noindent regardless of directing of the ribbon (note that the direction of the twist, i.e. clockwise versus counter clockwise, does matter).
\begin{Theorem} (see \cite[Theorem 1.10]{Ohtsuki})
The Kauffman bracket as calculated using the above relations is an isotopy invariant of framed links. \qed
\end{Theorem}
We now describe a modification that leads to an invariant of directed but unframed links. The invariant does still depend on more then just the underlying link (i.e. the choice of directing), but now the amount of choice is much smaller. In fact, for knots (i.e. links with a single component), the invariant does not depend on the directing either (see Comment \ref{knots-wd}).
\begin{Definition} \label{def_wbits}
\begin{enumerate}
\item A positive crossing is a crossing of the form
\begin{center}
\setlength{\unitlength}{0.5cm}
\thicklines
\begin{picture}(3,3.7)
\put(0.5,0){\line(1,1){3}}
\put(3,0){\line(-1,1){1}}
\put(1,2){\line(-1,1){1}}
\put(0,0.5){\line(1,1){3}}
\put(3.5,0.5){\line(-1,1){1}}
\put(1.5,2.5){\line(-1,1){1}}
\put(3,3){\vector(1,1){0.8}}
\put(0.5,3){\vector(-1,1){0.8}}
\put(4,0){.}
\end{picture}
\end{center}
That is, a crossing such that, if you approach the crossing along the upper ribbon in the chosen direction and leave along the lower ribbon, you have made a left turn.
\item A negative crossing is a crossing of the form
\begin{center}
\setlength{\unitlength}{0.5cm}
\thicklines
\begin{picture}(3,3.7)
\put(0,0){\line(1,1){1}}
\put(2,2){\line(1,1){1}}
\put(2.5,0){\line(-1,1){3}}
\put(-0.5,0.5){\line(1,1){1}}
\put(1.5,2.5){\line(1,1){1}}
\put(3,0.5){\line(-1,1){3}}
\put(2.5,3){\vector(1,1){0.8}}
\put(0,3){\vector(-1,1){0.8}}
\put(4,0){.}
\end{picture}
\end{center}
That is, a crossing such that, if you approach the crossing along the upper ribbon in the chosen direction, then leave along the lower ribbon, you have made a right turn.
\item A positive full twist is a twist of the form
\begin{center}
\setlength{\unitlength}{0.42cm}
\begin{picture}(2,5)
\put(-0.3,0.9){ \usrtwist }
\put(0.04,3.2){\surtwist}
\put(1.3,4.3){\vector(0,1){0.8}}
\end{picture}
\end{center}
\item A negative full twist is a twist in the opposite direction to a positive full twist.
\item The writhe of a link diagram $L$, denoted by $w(L)$, is the number of positive crossings minus the number of negative crossings plus the number of positive full twists minus the number of negative full twists.
\end{enumerate}
\end{Definition}
\begin{Comment}
Here we have drawn every component with its framing. Sometime we will just draw lines, and use the convention that these stand for ribbons lying flat on the page. This is often referred to as the ``blackboard framing".
\end{Comment}
\begin{Lemma} (see \cite{Kauffman:1987})
The writhe $w(L)$ is an invariant of directed framed links. \qed
\end{Lemma}
The following is one of the more fundamental theorems in knot theory.
\begin{Theorem} (see \cite[Theorem 1.5]{Ohtsuki}) \label{Jones1}
Let $L$ be any link. Then the Jones polynomial,
\begin{equation}
J(L):= (-q^{3/2})^{-w(L)} K(L),
\end{equation}
is independent of the framing. Hence $J(L)$ is an isotopy invariant of directed (but not framed) links. \qed
\end{Theorem}
Theorem \ref{Jones1} is sometimes stated in terms of link diagrams, not framed links. The result for framed links follows by noticing that the positive full twist from Definition \ref{def_wbits} is isotopic to
\begin{center}
\begin{tikzpicture}[scale=0.28, yscale=1]
\draw[line width = 0.06cm] (0,0) .. controls (0,2) and (1,4) .. (2,4) .. controls (3,4) and (3,1) .. (1,3);
\draw[line width = 0.06cm] (0.65, 3.35) .. controls (0.25, 3.75) and (0,5) .. (0,6);
\end{tikzpicture}
\end{center}
with the blackboard framing.
\begin{Comment} \label{knots-wd}
It is straightforward to see that positive full twists are sent to positive full twists if the direction of the ribbon is reversed, and positive crossings are sent to positive crossings if the directions of both ribbons involved are reversed. It follows that the choice of directing only affects the Jones polynomial for links with at least two components.
\end{Comment}
\section{The quantum group construction of the Jones polynomial}
\subsection{The quantum group $U_q(\mathfrak{sl}_2)$ and its representations}
$U_q(\mathfrak{sl}_2)$ is an infinite dimensional algebra related to the Lie-algebra $\mathfrak{sl}_2$ of $2 \times 2 $ matrices with trace zero. It is the algebra over the field of rational functions $\mathbb{C}(q)$ generated by $E,F, K$ and $K^{-1}$, subject to the relations
\begin{equation}
\begin{aligned}
&K K^{-1}=1, \\
&KEK^{-1} = q^2 E, \\
& KFK^{-1} = q^{-2} F, \\
& EF-FE= \frac{K-K^{-1}}{q-q^{-1}}.
\end{aligned}
\end{equation}
For our purposes it is convenient to adjoin a fixed square root $q^{1/2}$ to $\mathbb{C}(q)$.
$U_q(\mathfrak{sl}_2)$ has a finite dimensional representation $V_n$ for each integer $n$ which we now describe. Introduce the ``quantum integers"
\begin{equation}
[n]:= \frac{q^n-q^{-n}}{q-q^{-1}} = q^{n-1}+q^{n-3} + \cdots + q^{-n+1}.
\end{equation}
The representation $V_n$ has $\mathbb{C}(q)$-basis $\{ v_n, v_{n-2}, \cdots, v_{-n+2}, v_{-n} \}$, and the actions of $E,F$ and $K$ are given by
\begin{equation}
\begin{aligned}
E(v_{-n+2j})&=
\begin{cases}
[j+1] v_{-n+2j+2} \quad \text{ if } 0 \leq j < n \\
0 \quad \text{ if } j=n,
\end{cases}
\\
F(v_{n-2j})&=
\begin{cases}
[j+1] v_{n-2j-2} \quad \text{ if } 0 \leq j < n \\
0 \quad \text{ if } j=n,
\end{cases}
\\
K(v_k)&= q^k v_k.
\end{aligned}
\end{equation}
This can be expressed by the following diagram:
\begin{equation}
\setlength{\unitlength}{0.6cm}
\begin{picture}(15,3.5)
\put(1,2){\circle*{0.3}}
\put(3,2){\circle*{0.3}}
\put(5,2){\circle*{0.3}}
\put(11,2){\circle*{0.3}}
\put(13,2){\circle*{0.3}}
\put(15,2){\circle*{0.3}}
\put(7.5,2){$\ldots$}
\put(1.2,2.25){\vector(1,0){1.6}}
\put(3.2,2.25){\vector(1,0){1.6}}
\put(5.2,2.25){\vector(1,0){1.6}}
\put(9.2,2.25){\vector(1,0){1.6}}
\put(11.2,2.25){\vector(1,0){1.6}}
\put(13.2,2.25){\vector(1,0){1.6}}
\put(2.8,1.75){\vector(-1,0){1.6}}
\put(4.8,1.75){\vector(-1,0){1.6}}
\put(6.8,1.75){\vector(-1,0){1.6}}
\put(10.8,1.75){\vector(-1,0){1.6}}
\put(12.8,1.75){\vector(-1,0){1.6}}
\put(14.8,1.75){\vector(-1,0){1.6}}
\put(1.8,2.6){\footnotesize $1$}
\put(3.7,2.6){\footnotesize $[2]$}
\put(5.7,2.6){\footnotesize $[3]$}
\put(9.3,2.6){\footnotesize $[n-2]$}
\put(11.3,2.6){\footnotesize $[n-1]$}
\put(13.7,2.6){\footnotesize $[n]$}
\put(1.7,1){\footnotesize $[n]$}
\put(3.3,1){\footnotesize $[n-1]$}
\put(5.3,1){\footnotesize $[n-2]$}
\put(9.8,1){\footnotesize $[3]$}
\put(11.8,1){\footnotesize $[2]$}
\put(13.8,1){\footnotesize $1$}
\put(0.8,-0.3){$q^n$}
\put(2.8,-0.3){$q^{n-2}$}
\put(4.8,-0.3){$q^{n-4}$}
\put(10.8,-0.3){$q^{-n+4}$}
\put(12.8,-0.3){$q^{-n+2}$}
\put(14.8,-0.3){$q^{-n}$}
\put(-2,2.55){$F:$}
\put(-2,0.95){$E:$}
\put(-2,-0.4){$K:$}
\end{picture}
\end{equation}
\vspace{0.4cm}
There is a tensor product on representations of $U_q(\mathfrak{sl}_2)$, where the action on $a \otimes b \in A \otimes B$ is given by
\begin{equation}
\begin{aligned}
E(a \otimes b) & = Ea \otimes Kb+ a \otimes Eb, \\
F(a \otimes b) & = Fa \otimes b+ K^{-1}a \otimes Fb, \\
K(a \otimes b) & = Ka \otimes Kb.
\end{aligned}
\end{equation}
It turns out that $A \otimes B$ is always isomorphic to $B \otimes A$, and furthermore there is a well known natural system of isomorphisms
\begin{equation}
\sigma_{A,B}^{br}: A \otimes B \rightarrow B \otimes A
\end{equation}
for each pair $A,B$, called the braiding. The braiding has a standard definition, which can be found in, for example \cite{CP} (or Theorem \ref{KR_th} below can also be used as the definition).
Here we only ever apply the braiding to representations isomorphic to the standard 2-dimensional representation of $U_q(\mathfrak{sl}_2)$, so we can use the following:
\begin{Definition} \label{easy-s-def} Let $V$ be the 2 dimensional representation of $U_q(\mathfrak{sl}_2)$. Use the ordered basis $\{ v_1 \otimes v_1, v_{-1} \otimes v_{1}, v_1 \otimes v_{-1}, v_{-1} \otimes v_{-1} \}$ for $V \otimes V$. Then $\sigma_{V,V}^{br}: V \otimes V \rightarrow V \otimes V$ is given by the matrix
\begin{equation*}
\sigma^{br}=
q^{-1/2} \left(
\begin{array}{cccc}
q&0&0&0 \\
0&q-q^{-1}&1&0\\
0&1&0&0 \\
0&0&0&q
\end{array}
\right).
\end{equation*}
To simplify notation, we usually denote $\sigma_{V,V}^{br}$ simply by $\sigma^{br}$.
\end{Definition}
There is a standard action of $U_q(\mathfrak{sl}_2)$ on the dual vector space to $V_n$. This is defined using the ``antipode" $S$, which is the algebra anti-automorphism defined on generators by:
\begin{equation}
\begin{aligned}
&S(E)=-EK^{-1}, \\
&S(F)=-K F, \\
&S(K)=K^{-1}.
\end{aligned}
\end{equation}
For $\hat{v} \in V_n^*$ and $X \in U_q(\mathfrak{sl}_2)$, set $X \cdot \hat v$ to be the element of $V_n^*$ defined by
\begin{equation}
(X \cdot \hat v) (w) := \hat v (S(X) w)
\end{equation}
for all $w \in V_n$. It is straightforward to check that this is in fact an action of $U_q(\mathfrak{sl}_2)$ on $V_n^*$.
It turns out that $V_n$ is always isomorphic to $V_n^*$, which will be important later on.
\begin{Example} \label{an-iso}
An isomorphism between the standard representation of $U_q(\mathfrak{sl}_2)$ and its dual. Let $v_1,v_{-1}$ be the basis for $V$. For $i=\pm 1$, let $\hat v_i$ be the element of $V^*$ defined by
\begin{equation}
\hat v_i (v_j)= \delta_{i,j}.
\end{equation}
Calculating using the above definition, the action of $U_q(\mathfrak{sl}_2)$ on $V^*$ is given by
\setlength{\unitlength}{1cm}
\begin{equation}
\begin{picture}(6,0.8)
\put(0,-1){
\begin{picture}(6,2)
\thicklines
\put(2,1){$\hat v_{-1}$}
\put(5,1){$\hat v_1$,}
\put(2.6,1.2){\vector(1,0){2.3}}
\put(4.9,0.9){\vector(-1,0){2.3}}
\put(0.5,1.5){$F:$}
\put(3.5,1.5){$-q^{-1}$}
\put(0.5,0.35){$E:$}
\put(3.5,0.35){$-q^{}$}
\end{picture}}
\end{picture}
\end{equation}
\vspace{0.7cm}
\noindent Consider the map of vector spaces $f: V \rightarrow V^*$ defined by
\begin{equation}
\begin{cases}
f(v_1)= \hat v_{-1} \\
f(v_{-1}) = -q^{-1} \hat v_1
\end{cases}
\end{equation}
One can easily check that $f$ is in fact an isomorphism of $U_q(\mathfrak{sl}_2)$ representations.
\end{Example}
\begin{Comment}
If one sets $q=1$, the representations $V_n$ described above are exactly the irreducible finite dimensional representations of the usual Lie algebra $\mathfrak{sl}_2$, where one identifies
\begin{gather}
E \leftrightarrow \left( \begin{array}{cc} 0&1 \\ 0&0 \end{array} \right), \quad F \leftrightarrow \left( \begin{array}{cc} 0&0 \\ 1&0 \end{array} \right), \quad \frac{K-K^{-1}}{q-q^{-1}} \leftrightarrow \left( \begin{array}{cc} 1&0 \\ 0&-1 \end{array} \right).
\end{gather}
Of course, one needs to be a bit careful about interpreting the third identification here, since it looks like you divide by $0$. This issue is addressed in \cite[Chapters 9 and 11]{CP}.
For us, this observation will be sufficient justification for thinking of $U_q(\mathfrak{sl}_2)$ as related to ordinary $\mathfrak{sl}_2$.
\end{Comment}
\begin{Comment}
Notice that $K$ acts as the identity on all $V_n$ at $q=1$. $U_q(\mathfrak{sl}_2)$ actually has some other finite dimensional representations where $K$ does not act as the identity at $q=1$. So we have not described the full category of finite dimensional representation of $U_q(\mathfrak{sl}_2)$, but only the so called ``type {\bf 1}" representations. The other representations rarely appear in the theory.
\end{Comment}
\subsection{Ribbon elements and quantum traces}
Much of the following construction can be found in, for example, \cite[Chapter 4]{CP} or \cite{Ohtsuki}. The main difference here is that we work with two ribbon elements throughout. Each satisfies the definition of a ribbon element as in \cite{CP}. Consequently we also have two different quantum traces, and two different co-quantum traces. The non-standard ribbon element $Q_t$ is discussed extensively in \cite{half_twist}.
\begin{Definition} \label{ribbons} The ribbon elements $Q_s$ and $Q_t$ are elements in some completion of $U_q(\mathfrak{sl}_n)$ defined by
\begin{itemize}
\item The standard ribbon element $Q_s$ acts on $V_n$ as multiplication by the scalar $q^{-n^2/2-n}$.
\item The ``non-standard" or ``half-twist" ribbon element $Q_t$ acts on $V_n$ as multiplication by the scalar $(-1)^nq^{-n^2/2-n}$.
\end{itemize}
\end{Definition}
\begin{Comment} One can also think of $Q_s$ or $Q_t$ as a natural system of automorphisms of each finite dimensional type {\bf 1} representations of $U_q(\mathfrak{sl}_2)$. Specifically, $Q_s$ is the system which acts on $V_n$ as multiplication by $q^{-n^2/2-n}$. Similarly, $Q_t$ is the system which acts on $V_n$ as multiplication by $(-1)^nq^{-n^2/2-n}$.
\end{Comment}
\begin{Definition} \label{group-likes} The ``grouplike elements" associated to $Q_s$ and $Q_t$ are elements in some completion of $U_q(\mathfrak{sl}_n)$ defined by
\begin{itemize}
\item $g_s$ acts on $v_{n-2j} \in V_n$ as multiplication by $q^{n-2j}$.
\item $g_t$ acts on $v_{n-2j} \in V_n$ as multiplication by $(-1)^n q^{n-2j}$.
\end{itemize}
\end{Definition}
\begin{Comment}
The group like elements in Definition \ref{group-likes} are related to the ribbon elements in Definition \ref{ribbons} as described in \cite[Chapter 4.2C]{CP}.
\end{Comment}
\begin{Definition} \label{ev-maps} (see
\cite[Section 4.2]{Ohtsuki})
Define the following maps:
\begin{enumerate}
\item $ev$ is the evaluation map $V^* \otimes V \rightarrow F$.
\item $qtr_{Q_s}$ is the standard quantum trace map $V \otimes V^* \rightarrow F$ defined by, \\ for $\phi \in \mbox{End}(V)=V \otimes V^* $, $qtr_{Q_s}(\phi)= trace(\phi \circ g_s)$.
\item $qtr_{Q_t}$ is the ``half-twist" quantum trace map $V \otimes V^* \rightarrow F$ defined by, \\ for $\phi \in \mbox{End}(V)=V \otimes V^* $, $qtr_{Q_t}(\phi)= trace(\phi \circ g_t)$.
\item $coev$ is the coevaluation map $F \rightarrow V \otimes V^*$ defined by $coev(1)= \text{Id}$, where $\text{Id}$ is the identity map in $\mbox{End}(V)= V \otimes V^*$.
\item $coqtr_{Q_s}$ is the standard co-quantum trace map $F \rightarrow V^* \otimes V$ defined by $$coqtr_{Q_s}(1) = (1 \otimes g_s^{-1}) \circ Flip \circ coev (1),$$ where $\mbox{Flip}$ means interchange the two tensor factors.
\item $coqtr_{Q_t}$ is the ``half-twist" co-quantum trace map $F \rightarrow V^* \otimes V$ defined by $$coqtr_{Q_t}(1) = (1 \otimes g_t^{-1}) \circ Flip \circ coev (1).$$
\end{enumerate}
\end{Definition}
\begin{Comment} Although this may not be obvious, the maps in Definition \ref{ev-maps} are all morphisms of $U_q(\mathfrak{sl}_2)$ representations.
\end{Comment}
\begin{Comment}
It is often useful to express the maps from Definition \ref{ev-maps} in coordinates. So, fix $f \in V^*$, $v \in V$, and $\{ e^i \}$, $\{ e_i \}$ be dual basis for $V^*$ and $V$. Then:
\begin{equation}
\begin{aligned}
& ev (f \otimes v) = f(v), \\
& qtr_Q( v \otimes f) = f(gv), \\
& coev(1)= \sum_i e_i \otimes e^i, \\
& coqtr_Q(1) = \sum_i e^i \otimes g^{-1} e_i.
\end{aligned}
\end{equation}
One can choose $Q$ to be either $Q_s$ or $Q_t$, and then one must use the grouplike element $g_s$ or $g_t$ accordingly.
\end{Comment}
\subsection{Two topological categories}
Quantum group knot invariants work by constructing a functor from a certain topological category to the category of representations of the quantum group. We now define the relevant topological category. In fact, we will need two slightly different topological categories.
\setlength{\unitlength}{0.6cm}
\begin{Definition}
$\mathcal{DRIBBON}$ (directed orientable topological ribbons) is the category where:
$\bullet$ An object consists of a finite number of disjoint closed intervals on the real line each directed either up or down. These are considered up to isotopy of the real line. For example:
\vspace{-0.7cm}
\begin{center}
$$\uuv[] \uuv[] \udv[] \udv[] \udv[] \uuv[].$$
\end{center}
`
$\bullet$ A morphism between two objects $A$ and $B$ consists of a ``tangle of orientable, directed ribbons" in $\mathbb{R}^2 \times I$, whose ``loose ends" are exactly $(A, 0, 0) \cup (B,0, 1) \subset \mathbb{R} \times \mathbb{R} \times I $, such that the direction (up or down) of each interval in $A \cup B$ agrees with the direction of the ribbon whose end lies at that interval.
These are considered up to isotopy. For technical details of the definition of ``a ribbon", see \cite{CP}.
$\bullet$ Composition of two morphisms is given by stacking them on top of each other, and then shrinking the vertical axis by a factor of two. For example,
\setlength{\unitlength}{0.5cm}
\begin{center}
\begin{picture}(15,3.8)
\put(-0.22,-0.65){\pplaincrossing}
\put(0,-0.7){\puuv[]}
\put(2,-0.7){\pudv[]}
\put(0,2.7){\pudv[]}
\put(2,2.7){\puuv[]}
\put(4.5,1.5){$\circ$}
\put(5.2,2.5){$\uudcup[]$}
\put(10, 1.5){$=$}
\put(12,1){\halfpplaincrossing}
\put(12,2.7){\pudv[]}
\put(14,2.7){\puuv[]}
\put(12.05,0.8){\halfucup.}
\end{picture}
\end{center}
\end{Definition}
\vspace{0.15cm}
\begin{Definition}
$\mathcal{RIBBON}$ (undirected orientable topological ribbons) is the category obtained from $\mathcal{DRIBBON}$ by forgetting the directings. So an object consists of a finite number of disjoint closed intervals on the real line, a morphism consists of a tangle of undirected ribbons, and composition is still stacking of tangles.
\end{Definition}
\subsection{The functor}
The following is the main ingredient in the quantum group construction of knot invariants. It holds in much greater generality than stated here, which allows for the construction of a great many invariants.
\begin{Theorem} \label{ribbonfunctor} (see \cite[Theorem 5.3.2]{CP}) Let $V$ be the standard 2 dimensional representation of $U_q(\mathfrak{sl}_2)$. For each ribbon element $Q$ (i.e. $Q_s$ or $Q_t$), there is a unique monoidal functor $\mathcal{F}_Q$ from $\mathcal{RIBBON}$ to $U_q(\mathfrak{sl}_2) \text{-rep}$ such that
\begin{enumerate}
\item $\mathcal{F}_Q(\uuv[])=V$ and $\mathcal{F}_Q(\udv[])=V^*$,
\item $\begin{array}{ll} \mathcal{F}_Q \left( \uducap[] \right)= \text{ev}, & \mathcal{F}_Q \left( \uudcap[] \right)= \text{qtr}_Q, \\ \\
\mathcal{F}_Q \left( \uudcup[] \right)= \text{coev}, & \mathcal{F}_Q \left( \uducup[] \right)= \text{coqtr}_Q,
\end{array}$
\vspace{0.3cm}
\setlength{\unitlength}{0.5cm}
\item \label{ft} $\mathcal{F}_Q \left( \begin{picture}(2.2,2.2)
\put(0,0.9){\twist}
\put(0,-1.1){\usltwist}
\end{picture} \right) = Q,$ thought of as an automorphism of either $V$ or $V^*$.
\vspace{0.15in}
\setlength{\unitlength}{0.7cm}
\item $\mathcal{F}_Q \left( \plaincrossing \right) = \sigma^{br}$
\noindent as a morphism from the tensor product of the bottom two objects to the tensor product of the top two objects, regardless of the directions of the ribbons. \qed
\end{enumerate}
\end{Theorem}
\begin{Comment}
If one or both of the ribbons is directed down, one must be cautious using Definition \ref{easy-s-def} to calculate $\sigma^{br}$; one must first choose an explicit isomorphism from $V^*$ to $V$. By naturality, the resulting morphism $\sigma^{br}$ will not depend on this choice. This technicality comes up again in Example \ref{qtr-Ex}, where we deal with it in detail.
\end{Comment}
\begin{Comment}
Notice that $\mathcal{F}_Q$ sends the \emph{negative} full twist to $Q$. This seems like a strange way to set things up, but it is done to match other fairly standard conventions. In some ways it works well; $Q_s$ acts as multiplication by $q$ to a negative exponent, so positive twists correspond to positive exponents.
\end{Comment}
Let $L$ be a directed framed link. Then one can draw $L$ as a composition of the elementary features in Theorem \ref{ribbonfunctor}, and hence find the morphism associated to $L$. This is a morphism from the identity object to itself in the category of $U_q(\mathfrak{sl}_2)$ representations, which is just multiplication by a rational function in $q^{1/2}$ (and this turns out to be a polynomial in $q^{1/2}$ and $q^{-1/2}$). By Theorem \ref{ribbonfunctor}, $\mathcal{F}_Q$ is well defined on tangles up to isotopy. In particular, $\mathcal{F}_Q(L)$, is an isotopy invariant. It is related to the Kauffman bracket as follows:
\begin{Theorem} (see \cite[Theorem 4.19]{Ohtsuki}) \label{with_m} Fix a framed link $L$. Then $\mathcal{F}_{Q_s}(L)$ is independent of the choice of directing of $L$. Furthermore, $\mathcal{F}_{Q_s}(L)= (-1)^{w(L)+\# L} K(L),$ where $w(L)$ is the writhe of $L$ and $\# L$ is the number of components of $L$. \qed \end{Theorem}
\begin{Comment}
Theorem \ref{with_m} is not hard to prove using Corollary \ref{get_rid} below and the observation that $\mathcal{F}_{Q_s}(L)/\mathcal{F}_{Q_t}(L)$ is an isotopy invariant that cannot tell the difference between overcrossings and undercrossings, and hence can only depend on the number of components of $L$ and the writhe of each component mod 2 (see also \cite[Proposition 5.22]{half_twist}). Note that $w(L)$ mod $2$ does not depend on the directing of $L$.
\end{Comment}
The $(-1)^{w(L)+\# L}$ in Theorem \ref{with_m} is the sign referred to in the title of these notes. It is certainly explicitly defined, so in some sense it is not a problem; just an annoyance. Section \ref{matching} develops one way to get rid of this sign by using $Q_t$ in place of $Q_s$, although in some sense this just moves the annoyance into the definition of the ribbon element. The real justification for using $Q_t$ is not so much that it explains the sign, but that it makes Theorem \ref{Fundir} functorial.
\section{Matching the two constructions using the non-standard ribbon element} \label{matching}
We now show how the skein relations used in defining the Kauffman bracket can be explained using the quantum group formulation.
This section is similar to \cite[Appendix H]{Ohtsuki}, although the presentation is simplified by use the non-standard ribbon element $Q_t$ throughout. The idea is to modify the functor from Theorem \ref{ribbonfunctor} to obtain a function from $\mathcal{RIBBON}$ to $U_q(\mathfrak{sl}_2)-\text{rep}$, as opposed to from $\mathcal{DRIBBON}$. One argument for wanting this is that the Kauffman bracket is defined for framed but undirected links, which are morphisms in $\mathcal{RIBBON}$, but not in $\mathcal{DRIBBON}$.
There is only one ``elementary" object in $\mathcal{RIBBON}$ (the single interval), as opposed to two in $\mathcal{DRIBBON}$ (the single interval, but with two possible directions). Our morphism will send this single interval to the two dimensional representation $V$. We must then send each feature in the knot diagram to a morphism between the appropriate tensor powers of $V$. For instance,
\begin{center}
\ucap
\end{center}
\vspace{0.1cm}
should be sent to a morphism from $V \otimes V$ to the trivial object. This is as opposed to the directed case, where such ``caps" are sent to morphisms from $V^* \otimes V$ or $V \otimes V^*$ to the trivial object. To do this, we will use the fact that, in this particular situation, $V$ is isomorphic to $V^*$ (for instance, via the isomorphism from Example \ref{an-iso}). We obtain:
\begin{Theorem} \label{Fundir}
\thicklines
Choose an isomorphism $f: V \rightarrow V^*$. There is a unique functor $\mathcal{F}_{f}: \mathcal{RIBBON} \rightarrow U_q(\mathfrak{sl}_2) \text{-rep}$ such that
\begin{enumerate} \setlength{\unitlength}{0.28cm}
\item $\mathcal{F}_{f}$ takes the object consisting of a single interval to $V$,
\vspace{0.1cm}
\item \label{undir-cap}
$\displaystyle \mathcal{F}_{f} \left(
\begin{picture}(6,2.2)
\put(0,-1){\ucap}
\end{picture}
\right)
=
ev \circ (f \otimes \text{Id}) = qtr_{Q_t} \circ (\text{Id} \otimes f): V \otimes V \rightarrow \mathbb{C}(q)$,
\vspace{0.1cm}
\item \label{undir-cup} $\displaystyle \mathcal{F}_{f} \left(
\begin{picture}(6,2.2)
\put(0,-1.2){\ucup}
\end{picture}
\right)
=
(\text{Id} \otimes f^{-1})\circ coev = (f^{-1} \otimes \text{Id}) \circ coqtr_{Q_t}: \mathbb{C}(q) \rightarrow V \otimes V$,
\vspace{0.2cm}
\setlength{\unitlength}{0.42cm}
\item \label{undir-cross} $\displaystyle \mathcal{F}_{f} \left(
\begin{picture}(5.5,2)
\put(-0.1,-1.1){\plaincrossing}
\end{picture}
\right) =
\sigma^{br},$
\vspace{0.4cm}
\item \label{undir-twist} \setlength{\unitlength}{0.42cm} $\mathcal{F}_{f} \left( \begin{picture}(2.4,2.8)
\put(0,0.9){\twist}
\put(0,-1.1){\usltwist}
\end{picture} \right) = Q_t \quad$ (or, equivalently, multiplication by $-q^{-3/2}$).
\setlength{\unitlength}{0.5cm}
\end{enumerate}
Furthermore, for any link $L$, any choice of directing of $L$, and any choice of $f$, $\mathcal{F}_{f}(L)= \mathcal{F}_{Q_t}(L).$
\end{Theorem}
\begin{Comment}
Theorem \ref{Fundir} implies that, for any link $L$, $\mathcal{F}_{f}(L)$ is independent of the chosen isomorphism $f$. However, the functor $\mathcal{F}_f$ does depend on this choice. For instance, $\mathcal{F}_f$ applied to a cap clearly depends on $f$.
\end{Comment}
\begin{Comment}
Quantum trace and co-quantum trace depend on the ribbon element, and the subscript indicates that we are using the ribbon element $Q_t$. If one tries to use $Q_s$ instead of $Q_t$, then the two expressions on the right sides in Theorem \ref{Fundir} parts \eqref{undir-cap} and \eqref{undir-cup} are off by a minus sign, and the construction does not work. That the two sides of \eqref{undir-cap} and \eqref{undir-cup} agree follows from the fact that $U_q(\mathfrak{sl}_2)-\text{rep}$, along with ``pivotal structure" related to the ribbon element $Q_t$, is unimodal, as defined in \cite{Turaev}. For an explanation of this pivotal structure and a proof that it is unimodal see \cite[Section 5B]{half_twist}. It is also not hard to directly verify that the expressions agree.
\end{Comment}
\begin{proof}[Proof of Theorem \ref{Fundir}]
This proof is a bit informal. You should draw a non-trivial element of $\mathcal{RIBBON}$, then follow what is being said.
First, verify by a direct calculation that the two expressions on the right for parts \eqref{undir-cap} and \eqref{undir-cup} agree. Thus $\mathcal{F}_{f}$ is well defined on framed link diagrams. We will now show that it agrees with $\mathcal{F}_{Q_t}$ calculated with respect to any directing. Since $\mathcal{F}_{Q_t}$ is a functor, this implies that $\mathcal{F}_{f}$ is as well.
Fix a directing of $L$. Insert $ f \circ f^{-1}$ into $\mathcal{F}_{Q_t}(L)$ somewhere along every segment of $L$ that is directed down. This clearly doesn't change the morphism. By the naturality of $\sigma^{br}$,
\begin{equation}
(1 \otimes f) \circ \sigma^{br} = \sigma^{br} \circ (f \otimes 1).
\end{equation}
Use this to pull all the $f$ and $f^{-1}$ through crossings until they are right next to cups and caps. But now you are exactly calculating $\mathcal{F}_{f}(L)$. Hence $\mathcal{F}_{f} = \mathcal{F}_{Q_t}$.
\end{proof}
We are now ready to see how skein relations appear. For the rest of this section we use the blackboard framing (i.e. ribbons lie flat on the page, and are drawn simply as lines). A simple calculation shows that
\begin{equation} \label{123_circle}
\setlength{\unitlength}{0.35cm}
\thicklines
\mathcal{F}_{f} \left(
\begin{picture}(3,1.9)
\put(1.5,0.2){\circle{2.8}}
\end{picture} \right)
=
\text{ multiplication by } -q-q^{-1}.
\end{equation}
\vspace{-0.1cm}
Another direct calculation shows that
\begin{equation} \sigma^{br}= q^{1/2} Id + q^{-1/2} (\text{Id} \otimes f^{-1})\circ coev \circ qtr_{Q_t} \circ (\text{Id} \otimes f): V \otimes V \rightarrow V \otimes V.
\end{equation}
Equivalently,
\thicklines
\begin{equation} \label{kauffman_g}
\setlength{\unitlength}{0.35cm}
\mathcal{F}_{f} \left(
\begin{picture}(3.5,1.9)
\put(0.2,-1.3){\line(1,1){3}}
\put(0.2,1.7){\line(1,-1){1.1}}
\put(2.2,-0.3){\line(1,-1){1.1}}
\end{picture}
\right)
=
q^{1/2}
\mathcal{F}_{f} \left(
\begin{picture}(3.5,1.9)
\put(0.2,-1.3){\line(0,1){3}}
\put(3.2,-1.3){\line(0,1){3}}
\end{picture}
\right)
+
q^{-1/2}
\mathcal{F}_{f}
\left(
\begin{picture}(3.5,1.9)
\put(1.7,-1.3){\oval(3,2)[t]}
\put(1.7,1.7){\oval(3,2)[b]}
\end{picture}
\right).
\end{equation}
\noindent But these are exactly the relations used in Definition \ref{Kauffman-simplifications} to define the Kauffman bracket! Noticing that Equation \eqref{third-Kauff} and Theorem \ref{Fundir}\eqref{undir-twist} are also compatible, this implies that $\mathcal{F}_{f}$ of a framed but undirected link gives the Kauffman bracket. Applying Theorem \ref{Fundir} we see:
\begin{Corollary} \label{get_rid}
Let $L$ be a framed link. Then $\mathcal{F}_{Q_t}(L)$ is independent of the chosen directing, and is equal to the Kauffman bracket $K(L)$. \qed
\end{Corollary}
\begin{Comment}
The non-standard ribbon we use exists in many cases beyond $U_q(\mathfrak{sl}_2)$, and can also help explain the correspondence between various constructions of knot polynomials in those cases.
\end{Comment}
\begin{Example} \label{qtr-Ex}
A way to verify the definition of quantum trace. Recall that $\mathcal{F}_Q$ is supposed to be defined on $\mathcal{DRIBBON}$, and morphisms there are ribbon tangles \emph{up to isotopy}. One can use an isotopy to change a right going cap to the composition of a twist, a crossing, and a left going cap. Since we have only explicitly defined $\sigma^{br}$ acting on $V \otimes V$, not acting on $V \otimes V^*$, we also put in copies of $f$ and $f^{-1}$, where $f$ is the isomorphism from Example \ref{an-iso}. By the naturality of the braiding, this does not affect the morphism. Diagrammatically,
\setlength{\unitlength}{0.5cm}
\thicklines
\begin{center}
\begin{picture}(10,9.1)
\put(6,8){\halfucap}
\put(6,5.2){\otherhalfpplaincrossing}
\put(6.05,8){\line(1,0){1.5}}
\put(6.05,7.3){\line(1,0){1.5}}
\put(7.55,7.3){\line(0,1){0.7}}
\put(6.05,8){\line(0,-1){0.7}}
\put(6.55,7.5){\tiny$f$}
\put(8.05,3.8){\line(1,0){1.5}}
\put(8.05,3.1){\line(1,0){1.5}}
\put(9.55,3.1){\line(0,1){0.7}}
\put(8.05,3.8){\line(0,-1){0.7}}
\put(8.4,3.3){\tiny$f^{-1}$}
\put(6.3,5.8){\line(0,-1){3.5}}
\put(6.33,5.8){\line(0,-1){3.5}}
\put(6.35,5.8){\line(0,-1){3.5}}
\put(7.3,5.8){\line(0,-1){3.5}}
\put(7.33,5.8){\line(0,-1){3.5}}
\put(7.35,5.8){\line(0,-1){3.5}}
\put(8.3,3.1){\line(0,-1){0.7}}
\put(8.33,3.1){\line(0,-1){0.7}}
\put(8.35,3.1){\line(0,-1){0.7}}
\put(9.3,3.1){\line(0,-1){0.7}}
\put(9.33,3.1){\line(0,-1){0.7}}
\put(9.35,3.1){\line(0,-1){0.7}}
\put(8.3,8.1){\line(0,-1){0.8}}
\put(8.33,8.1){\line(0,-1){0.8}}
\put(8.35,8.1){\line(0,-1){0.8}}
\put(9.3,8.1){\line(0,-1){0.8}}
\put(9.33,8.1){\line(0,-1){0.8}}
\put(9.35,8.1){\line(0,-1){0.8}}
\put(8.05,4.75){\halfpsultwist}
\put(8.05,3.78){\halfpusltwist}
\put(7.75,2.1){\udv[],}
\put(5.75,2.1){\uuv[]}
\put(4.6,5){$\simeq$}
\put(0,7.2){\ucap}
\put(0.3,2.3){\line(0,1){5}}
\put(0.33,2.3){\line(0,1){5}}
\put(0.35,2.3){\line(0,1){5}}
\put(1.3,2.3){\line(0,1){5}}
\put(1.33,2.3){\line(0,1){5}}
\put(1.35,2.3){\line(0,1){5}}
\put(2.3,2.3){\line(0,1){5}}
\put(2.33,2.3){\line(0,1){5}}
\put(2.35,2.3){\line(0,1){5}}
\put(3.3,2.3){\line(0,1){5}}
\put(3.33,2.3){\line(0,1){5}}
\put(3.35,2.3){\line(0,1){5}}
\put(-0.25,2.1){\uuv[]}
\put(1.75,2.1){\udv[]}
\end{picture}
\end{center}
\vspace{-1cm}
\noindent where the boxes in the diagram mean ``put in a copy of the isomorphism $f$ when you apply $\mathcal{F}_Q$." Such ``tangles with coupons" are defined precisely in e.g. \cite{CP}. Algebraically, this says
\begin{equation}
\text{qtr}_Q = \text{ev} \circ (f \otimes \text{Id}) \circ \sigma^{br} \circ (\text{Id} \otimes Q^{-1}) \circ (\text{Id} \otimes f^{-1}).
\end{equation}
Since the action of each element on the right side has been explicitly defined, one can now check that the two sides agree on all basis vectors. Note that both sides depend on the choice of ribbon element $Q_s$ or $Q_t$.
\end{Example}
\begin{Comment} For our purposes, we could simply use the calculation in Example \ref{qtr-Ex} to define $\text{qtr}$. However, if we were to more fully develop the theory, the fact that $\text{qtr}$ can be defined as $\text{qtr}_{Q}(\phi)= trace(\phi \circ g)$ for a \emph{grouplike} element $g$ (see Definition \ref{ev-maps}) is important. The reason is that this implies quantum trace is multiplicative on tensor products. See \cite[Remark 1 after Definition 4.2.9]{CP}.
\end{Comment}
\section{Another advantage: the half twist}
We now discuss an invertible element $X$ in a certain completion of $U_q(\mathfrak{sl}_2)$, which is related to the non-standard ribbon element by $Q_t= X^{-2}$. As discussed in \cite{half_twist}, this element has an interesting topological interpretation.
\begin{Definition} \label{defX} $X$ is the operator that acts on $V_n$ by $Xv_{n-2j} = (-1)^{n-j} q^{{n^2/4+n/2}} v_{-n+2j}.$
\end{Definition}
\begin{Comment}
There is actually some choice in how we define $X$: the operator $X'$ that acts on $V_n$ by $X'v_{n-2j} = i^n q^{{n^2/4+n/2}} v_{-n+2j},$ where $i$ is the complex number $i$, also has all the properties discussed below. This type of modification is discussed in \cite[Section 5C]{half_twist} and \cite[Section 8]{Rcommutor}.
\end{Comment}
One can easily check that $X^{-2}=Q_t$. Comparing with Theorem \ref{ribbonfunctor}\eqref{ft}, one may hope that $X$ could be interpreted as an isomorphism, and that the functor $\mathcal{F}_{Q_t}$ could be extended in such a way that
\begin{equation}
\setlength{\unitlength}{0.5cm}
\mathcal{F}_{Q_t} \left(
\begin{picture}(2,1.25)
\put(-0.1,-0.1){\usltwist}
\end{picture}
\right)= X^{-1}.
\end{equation}
Another indication that such an extended functor should exist comes from the following result of Kirillov-Reshetikhin \cite[Theorem 3]{KR:1990} and Levendorskii-Soibelman, \cite[Theorem 1]{LS} (see \cite[Comment 7.3]{Rcommutor} for this exact statement).
\begin{Theorem} \label{KR_th}
$\sigma^{br} = (X^{-1} \otimes X^{-1}) \circ \text{Flip} \circ \Delta(X).$ \qed
\end{Theorem}
\noindent Theorem \ref{KR_th} can be interpreted via the following isotopy:
\setlength{\unitlength}{0.5cm}
\begin{equation}
\begin{picture}(9,2.23)
\put(0,-2){
\begin{picture}(9,5)
\put(0,0.2){\Dtst}
\put(0,3.2){\ptwist}
\put(2,3.2){\ptwist}
\put(4.5,2.5){$\simeq$}
\put(5.75,-0.54){\pplaincrossing}
\put(6,3.2){\uid}
\put(8,3.2){\uid}
\put(9.6,0.3){.}
\end{picture}}
\end{picture}
\end{equation}
\vspace{0.1cm}
\noindent Here $\mbox{Flip} \circ \Delta(X)$ should be interpreted as a morphism corresponding to twisting both ribbons at once by 180 degree, as on the bottom of the left side.
Such an extended functor has been defined precisely in \cite{half_twist}, resulting in a functor from a larger category where ribbons are allowed to twist by 180 degrees, not just by 360 degrees (although Mobius bands are still not allowed). Figure \ref{amorphism} shows an example of a morphism in the resulting category. Notice that elementary objects come in both shaded and unshaded versions.
The construction in \cite{half_twist} can only extend $\mathcal{F}_{Q_t}$, not $\mathcal{F}_{Q_s}$. We feel this gives more evidence that $Q_t$ is natural. One advantage of having such an extended functor is that, since both $\sigma^{br}$ and $Q_t$ are constructed in term of the ``half-twist" $X$, there is in some sense one less elementary feature.
\begin{figure}
$$\mathfig{0.45}{half-ribbons.pdf}$$
\caption{A morphism in the topological category of ribbons with half twists \label{amorphism}}
\end{figure}
|
\section{Introduction and Overview}
The Hubbard model \cite{Hubbard:1963aa}
is a model of spin-half electrons
hopping around on a lattice of atoms
(see \cite{Essler:2005aa} for an introduction).
It has several useful features that make it attractive
for the investigation of aspects of electron transport,
in particular superconductivity.
An unrelated property of its one-dimensional incarnation
is \emph{integrability} which enabled
Lieb and Wu to find the spectrum
by means of Bethe equations \cite{Lieb:1968aa}.
Remarkably, the integrable structure is different from
conventional spin chain models in several respects:
The most striking distinction is, arguably,
that the R-matrix, which was found by Shastry \cite{Shastry:1986bb},
is not of difference form.%
\footnote{Two similar cases have previously been discussed:
These are based on the twisted affine superalgebras
$\alg{gl}(N|N)^{(2\prime)}$ \cite{Nazarov:1999aa}
and $\alg{d}(2,1;e^{2\pi i/3})^{(3)}$ \cite{Leites:1984aa}.
Their Cartan--Killing forms are charged under the twisting automorphism
which leads to unconventional quantum algebras.
Another exceptional case involving the twisted affine superalgebra
$\alg{gl}(2|2)^{(2)}$ is discussed in \protect\secref{sec:rattwist}.}
This implies that the description of the integrable structure through
standard \emph{Yangian} or \emph{quantum affine algebras} \cite{Drinfeld:1985rx,Drinfeld:1986in,Jimbo:1985zk,Jimbo:1985ua}
cannot apply to this case.%
\footnote{The R-matrix must be invariant under the affine shift
which enforces the difference form.}
For a long time the question of the algebraic structure underlying
the Hubbard chain was left at rest.
Recent progress towards this goal came from
a totally unexpected direction:
It turned out that Shastry's R-matrix
is equivalent \cite{Beisert:2006qh} to a
scattering matrix \cite{Beisert:2005tm,Staudacher:2004tk}
found in the context of the AdS/CFT correspondence \cite{Maldacena:1998re}
(see \cite{Beisert:2010jr,Beisert:2004ry,Plefka:2005bk,Arutyunov:2009ga} for reviews
of integrability in AdS/CFT).
This matrix has a \emph{centrally extended} $\alg{psl}(2|2)$ supersymmetry by construction
which includes the two (more or less) manifest $\alg{sl}(2)$ symmetries
of the Hubbard model \cite{Lieb:1989aa,Yang:1989aa}.
Since then, there has been a lot of progress
in the formulation of a quantum symmetry algebra
for Shastry's R-matrix
\cite{Gomez:2006va,Plefka:2006ze,Beisert:2007ds,Matsumoto:2007rh,Spill:2008tp,Torrielli:2008wi,Spill:2008yr,Matsumoto:2009rf}.
In particular, the construction for higher representations
has advanced significantly
\cite{Beisert:2006qh,Chen:2006gp,Arutyunov:2008zt,deLeeuw:2008dp,deLeeuw:2008ye,Arutyunov:2009mi,Arutyunov:2009ce,Arutyunov:2009iq,Arutyunov:2009pw}.
Still, it is fair to say that a satisfactory quantisation
to a \emph{quasi-triangular Hopf algebra}
similar to a Yangian has not yet been achieved.
\smallskip
By quantum-deforming%
\footnote{The q-deformation lifts a rational to a trigonometric R-matrix,
e.g.\ Heisenberg XXX to XXZ.}
the Hubbard chain we hope to get further insights
into the Hopf algebra underlying this special model:
For conventional integrable spin chains based on Lie (super)algebra symmetries,
the quantum deformation lifts the Yangian to a quantum affine algebra.
This has some drawbacks, but also benefits.
One the one hand, the deformation breaks the manifest Lie symmetry
down to its Cartan subalgebra.
On the other hand, one gains a more uniform and symmetric description
of the algebra itself.
It is then possible to return to the undeformed model
and recover the Yangian as a particular limit.
The limit is singular, and it obscures some of the symmetry
of the quantum affine formulation.
An increased internal symmetry will hopefully simplify
the formulation of a quasi-triangular Hopf algebra for the
(quantum-deformed) Hubbard chain.
Another motivation to study the quantum-deformation
is that some of the structures in the centrally extended $\alg{psl}(2|2)$ algebra \cite{Beisert:2005tm}
for Shastry's R-matrix are reminiscent of quantum affine algebras.
The quantum-deformation of the Hubbard Hamiltonian
along with its R-matrix was performed in \cite{Beisert:2008tw}.
It has an additional parameter $q$, and therefore yields a bigger class of models.
It turned out that this class contains a multi-parameter family of
deformations of the Hubbard model proposed earlier
by Alcaraz and Bariev \cite{Alcaraz:1999aa}.
In fact, many of the variants of the Hubbard chain
(see references in \cite{Beisert:2006qh,Beisert:2008tw})
are special cases of this model.
The deformed and undeformed model and R-matrix
have in common a rather complicated structure
which obstructs direct attempts to set up a quasi-triangular Hopf algebra.
\medskip
Fortunately, there is a limit, the \emph{classical limit},
which makes the algebraic structure much more tractable:
The classical framework consists of some Lie algebra $\alg{g}$ along
with an element $r$ of the tensor product $\alg{g}\otimes\alg{g}$
serving as the \emph{classical r-matrix}.
For the quantum algebra $\alg{g}$ is promoted to a
deformation of its universal enveloping algebra
$\grp{U}_q(\alg{g})$
which is substantially bigger than $\alg{g}$ itself.
For r-matrices with spectral parameter,
the Lie algebra $\alg{g}$ is typically of affine Kac--Moody type,
for which an efficient and uniform description exists.
All in all, the manipulations in the classical limit can
usually be performed very explicitly with pen and paper,
much in contradistinction to the quantum case.
The classical limit of Shastry's R-matrix
was derived in \cite{Klose:2006zd,Torrielli:2007mc}.
The underlying Lie algebra with universal classical r-matrix
was found in \cite{Beisert:2007ty}.
This algebra turned out to be a peculiar deformation of the
loop algebra $\alg{gl}(2|2)[u,u^{-1}]$.
Note that the $\alg{gl}(2|2)$ algebra is not simple, it contains central charges
as well as derivations \cite{Moriyama:2007jt,Matsumoto:2007rh,Beisert:2007ty},
and thus it escapes the classification of r-matrices
in \cite{Leites:1984aa}.
The algebra is curious because it is not a loop algebra of some deformed algebra,
the deformation applies to the loop algebra structure itself,
in particular to the derivations and charges.
Yet, surprisingly, the algebra admits a quasi-triangular bialgebra structure.
\medskip
In this paper we will derive the classical r-matrix
for the quantum-deformed Hubbard chain.
This is the trigonometric analog of the
rational r-matrix in \cite{Klose:2006zd,Torrielli:2007mc,Beisert:2007ty}.
We expect that it will be of help in deriving the
full quantum algebra framework for the
(quantum-deformed) Hubbard model.
The paper is organised as follows:
We start with a brief review of the quantum R-matrix in \secref{sec:quantum}.
In the following \secref{sec:classical} we perform the classical limit
and show that it leads to a quasi-triangular Lie bialgebra.
Next we consider its affine extension in \secref{sec:affine}
which provides some more structure to the algebra.
The r-matrix and the algebra have several discrete symmetries
and special points which are discussed in \secref{sec:discrete}.
The last \secref{sec:limits} is devoted to the enumeration of simpler limiting cases
of the r-matrix and the algebra.
Finally, in \secref{sec:concl} we conclude and give an outlook.
\section{Quantum-Deformed S-Matrix}
\label{sec:quantum}
In \cite{Beisert:2008tw} a quantum-deformation of the
centrally extended $\alg{psl}(2|2)$ algebra was defined.
Subsequently, the fundamental R-matrix for this algebra was derived.
In this section we will summarise the results of \cite{Beisert:2008tw}
important to this paper.
\subsection{Serre--Chevalley Presentation}
\begin{figure}\centering
\includegraphics{FigDynkinOXO123.mps}
\caption{Distinguished Dynkin diagram for $\alg{sl}(2|2)$.}
\label{fig:Dynkin}
\end{figure}
We first define the quantum deformation of the extended $\alg{psl}(2|2)$ algebra
in the Serre--Chevalley presentation,
cf.\ \cite{Scheunert:1993aa,Zhang:1999aa} for the case of conventional (affine) $\alg{gl}(2|2)$.
It has 9 Serre--Chevalley generators
$\gen{H}_j,\gen{E}_j,\gen{F}_j$ with $j=1,2,3$.
For the distinguished choice of Dynkin diagram of $\alg{psl}(2|2)$,
see \figref{fig:Dynkin},
the generators $\gen{E}_2,\gen{F}_2$
are fermionic while the remaining 7 are bosonic.
The symmetric Cartan matrix $A_{jk}$ reads%
\footnote{For superalgebras it is sometimes convenient
to flip the signs of some rows/columns to make the matrix symmetric.}
\[
A_{jk}=\matr{rrr}{+2&-1&0\\-1&0&+1\\0&+1&-2}.
\]
\paragraph{Algebra.}
The commutators with symmetrised Cartan elements $\gen{H}_j$
are determined by the Cartan matrix $A_{jk}$
\[
\comm{\gen{H}_j}{\gen{H}_k}=0,
\qquad
\comm{\gen{H}_j}{\gen{E}_k}=+A_{jk}\gen{E}_k,
\qquad
\comm{\gen{H}_j}{\gen{F}_k}=-A_{jk}\gen{F}_k.
\]
The commutators between $\gen{E}_j$ and $\gen{F}_k$
are non-trivial only for $j=k$
\[\label{eq:EFcomm}
\comm{\gen{E}_1}{\gen{F}_1}=
\frac{q^{\gen{H}_1}-q^{-\gen{H}_1}}{q-q^{-1}}\,,
\qquad
\acomm{\gen{E}_2}{\gen{F}_2}=-
\frac{q^{\gen{H}_2}-q^{-\gen{H}_2}}{q-q^{-1}}\,,
\qquad
\comm{\gen{E}_3}{\gen{F}_3}=-
\frac{q^{\gen{H}_3}-q^{-\gen{H}_3}}{q-q^{-1}}\,.
\]
The Serre relations between
alike generators $\gen{E}_j$ or $\gen{F}_j$
read
\< 0\earel{=}
\comm{\gen{E}_1}{\gen{E}_3}
=\comm{\gen{F}_1}{\gen{F}_3} =\gen{E}_2\gen{E}_2
=\gen{F}_2\gen{F}_2
\\
\earel{=}\gen{E}_1\gen{E}_1\gen{E}_2
-(q+q^{-1})\gen{E}_1\gen{E}_2\gen{E}_1
+\gen{E}_2\gen{E}_1\gen{E}_1
=\gen{E}_3\gen{E}_3\gen{E}_2
-(q+q^{-1})\gen{E}_3\gen{E}_2\gen{E}_3
+\gen{E}_2\gen{E}_3\gen{E}_3
\nonumber\\\nonumber
\earel{=}\gen{F}_1\gen{F}_1\gen{F}_2
-(q+q^{-1})\gen{F}_1\gen{F}_2\gen{F}_1
+\gen{F}_2\gen{F}_1\gen{F}_1
=\gen{F}_3\gen{F}_3\gen{F}_2
-(q+q^{-1})\gen{F}_3\gen{F}_2\gen{F}_3
+\gen{F}_2\gen{F}_3\gen{F}_3
.
\>
\paragraph{Central Elements.}
What singles out $\alg{psl}(2|2)$ from the other simple
superalgebras is that it has three non-trivial central extensions
\cite{Nahm:1977tg,Iohara:2001aa}.
Our algebra has two central elements $\gen{C},\gen{D}$,
and they are the key to the peculiar features discussed in this paper.
The standard central element $\gen{C}$ in $\alg{sl}(2|2)$ reads
\[\label{eq:quantumC}
\gen{C}=-\sfrac{1}{2}\gen{H}_1-\gen{H}_2-\sfrac{1}{2}\gen{H}_3.
\]
In addition there are two exceptional central elements $\gen{P}$, $\gen{K}$
which originate from dropping the two Serre relations $\gen{P}=\gen{K}=0$
particular to superalgebras
\cite{Scheunert:1993aa}
\<
\label{eq:quantumPK}
\gen{P}\earel{=} \gen{E}_1\gen{E}_2\gen{E}_3\gen{E}_2
+\gen{E}_2\gen{E}_3\gen{E}_2\gen{E}_1
+\gen{E}_3\gen{E}_2\gen{E}_1\gen{E}_2
+\gen{E}_2\gen{E}_1\gen{E}_2\gen{E}_3
-(q+q^{-1})\gen{E}_2\gen{E}_1\gen{E}_3\gen{E}_2,
\nonumber\\
\gen{K}\earel{=} \gen{F}_1\gen{F}_2\gen{F}_3\gen{F}_2
+\gen{F}_2\gen{F}_3\gen{F}_2\gen{F}_1
+\gen{F}_3\gen{F}_2\gen{F}_1\gen{F}_2
+\gen{F}_2\gen{F}_1\gen{F}_2\gen{F}_3
-(q+q^{-1})\gen{F}_2\gen{F}_1\gen{F}_3\gen{F}_2.
\>
In order to get an interesting quantum algebra structure
the two extra central elements have to be constrained.
We introduce a new central element $\gen{D}$
as well as two global constants $g,\alpha$,
and express $\gen{P},\gen{K}$ through them%
\footnote{Notice the similarity between \eqref{eq:quantumPK,eq:PKident} and \eqref{eq:EFcomm}.}
\[\label{eq:PKident}
\gen{P}=g\alpha(1-q^{2\gen{C}}q^{2\gen{D}}),\qquad
\gen{K}=g\alpha^{-1}(q^{-2\gen{C}}-q^{-2\gen{D}}).
\]
\paragraph{Coalgebra.}
The standard quantum-deformed coproduct applies to all
\emph{bosonic} generators $\gen{E}_j,\gen{F}_j,\gen{H}_j$
(i.e.\ all except $\gen{E}_2$ and $\gen{F}_2$)
\<\label{eq:qcoproduct}
\mathrm{\Delta}(\gen{H}_j)\earel{=} \gen{H}_j\otimes 1+1\otimes \gen{H}_j, \nonumber\\
\mathrm{\Delta}(\gen{E}_j)\earel{=} \gen{E}_j\otimes 1+q^{-\gen{H}_j}\otimes \gen{E}_j, \nonumber\\
\mathrm{\Delta}(\gen{F}_j)\earel{=} \gen{F}_j\otimes q^{\gen{H}_j}+1\otimes \gen{F}_j.
\>
For the two fermionic generators $\gen{E}_2,\gen{F}_2$ an additional braiding with the
generator $\gen{D}$ is introduced
\<
\label{eq:bradedcopr}
\mathrm{\Delta}(\gen{E}_2)\earel{=}
\gen{E}_2\otimes 1+q^{-\gen{H}_2}q^{\gen{D}}\otimes \gen{E}_2,
\nonumber\\
\mathrm{\Delta}(\gen{F}_2)\earel{=}
\gen{F}_2\otimes q^{\gen{H}_2}+q^{-\gen{D}}\otimes \gen{F}_2,
\nonumber\\
\mathrm{\Delta}(\gen{D})\earel{=}\gen{D}\otimes 1+1\otimes\gen{D}.
\>
For convenience we have stated the coproduct of the central charge $\gen{D}$
which actually follows from the other coproducts.
\subsection{Fundamental Representation}
The above algebra has a family of four-dimensional fundamental representations.
Its vector space $\mathbb{V}$ has two bosonic and two fermionic directions.
We assume it to be spanned by the four states
\[
\state{\phi^1},\state{\phi^2}\quad\mbox{and}\quad\state{\psi^1},\state{\psi^2}.
\]
The former two are bosonic and the latter two are fermionic.
\paragraph{Representation.}
The fundamental action of the Chevalley-Serre generators is given by%
\footnote{We have interchanged the states $\state{\psi^1}$ and $\state{\psi^2}$
as compared to \protect\cite{Beisert:2008tw}.}
\[\label{eq:fundquant}
\begin{array}{rclrclrclrcl}
\gen{H}_1\state{\phi^1}\earel{=} -\state{\phi^1},&
\gen{H}_2\state{\phi^1}\earel{=} -(C-\sfrac{1}{2})\state{\phi^1},&
\gen{E}_1\state{\phi^1}\earel{=} q^{+1/2}\state{\phi^2},&
\gen{F}_2\state{\phi^1}\earel{=} c\state{\psi^2}, \\[0.65ex]
\gen{H}_1\state{\phi^2}\earel{=} +\state{\phi^2},&
\gen{H}_2\state{\phi^2}\earel{=} -(C+\sfrac{1}{2})\state{\phi^2},&
\gen{E}_2\state{\phi^2}\earel{=} a\state{\psi^1},&
\gen{F}_1\state{\phi^2}\earel{=} q^{-1/2}\state{\phi^1}, \\[0.65ex]
\gen{H}_3\state{\psi^1}\earel{=} +\state{\psi^1},&
\gen{H}_2\state{\psi^1}\earel{=} -(C+\sfrac{1}{2})\state{\psi^1},&
\gen{E}_3\state{\psi^1}\earel{=} q^{-1/2}\state{\psi^2},&
\gen{F}_2\state{\psi^1}\earel{=} d\state{\phi^2}, \\[0.65ex]
\gen{H}_3\state{\psi^2}\earel{=} -\state{\psi^2},&
\gen{H}_2\state{\psi^2}\earel{=} -(C-\sfrac{1}{2})\state{\psi^2},&
\gen{E}_2\state{\psi^2}\earel{=} b\state{\phi^1},&
\gen{F}_3\state{\psi^2}\earel{=} q^{+1/2}\state{\psi^1}.
\end{array}
\]
The representation parameters $a,b,c,d$
must obey the constraint $(ad-qbc)(ad-q^{-1}bc)=1$.
They can be expressed in terms
of new parameters $\xpm{},\gamma$ as follows%
\footnote{As compared to \cite{Beisert:2008tw}
we have rescaled $\gamma$ by $1/\sqrt{g}$
for later convenience.}
\[\label{eq:xpmparameters}
\begin{array}[b]{rclcrcl}
a\earel{=}\gamma,
&\quad&
b\earel{=}\displaystyle\frac{g\alpha}{\gamma}\,\frac{1}{\xm{}}\,\bigbrk{\xm{}-q^{2C-1}\xp{}},
\\[0.65ex]
c\earel{=}\displaystyle \frac{i\gamma}{\alpha}\,\frac{q^{-C+1/2}}{\xp{}}\,,
&\quad&
d\earel{=}\displaystyle \frac{ig}{\gamma}\,q^{C+1/2}\bigbrk{\xm{} -q^{-2C-1}\xp{}}.
\end{array}
\]
In terms of these parameters the constraint
implies the following quadratic relation between $\xpm{}$
\[\label{eq:xpmrel}
\frac{\xp{}}{q}+\frac{q}{\xp{}}-q\xm{}-\frac{1}{q\xm{}}
+ig(q-q^{-1})\lrbrk{\frac{\xp{}}{q\xm{}}-\frac{q\xm{}}{\xp{}}}=\frac{i}{g}\,.
\]
\paragraph{Central Charges.}
The central charge eigenvalues $D,C$ cannot be written unambiguously using $\xpm{}$,
but the combinations $q^{2D},q^{2C}$ are well-defined
\[\label{eq:q2c}
q^{2D}=\frac{\xp{}}{q\xm{}}\,,
\qquad
q^{2C}=q\,\frac{(q-q^{-1})/\xp{}-ig^{-1}}{(q-q^{-1})/\xm{}-ig^{-1}}
=q^{-1}\frac{(q-q^{-1})\xp{}+ig^{-1}}{(q-q^{-1})\xm{}+ig^{-1}}\,.
\]
The latter two expressions are equivalent upon \eqref{eq:xpmrel}.
Finally, the central charge eigenvalues $P,K$ follow
from \eqref{eq:PKident}
\[\label{eq:PKU}
P=g\alpha\lrbrk{1-q^{2C}q^{2D}},
\qquad
K=
g\alpha^{-1}
\lrbrk{q^{-2C}-q^{-2D}}.
\]
The parameter $\gamma$ adjusts the normalisation
of bosons w.r.t.\ fermions in the representation;
it is unphysical, but there is a preferable choice.
\paragraph{Fundamental R-Matrix.}
The quantum fundamental R-matrix $\mathcal{R}:
\mathbb{V}\otimes \mathbb{V}\to\mathbb{V}\otimes \mathbb{V}$ can be found by demanding
that it satisfies the cocommutativity relation
\[
\widetilde{\copro}(\gen{J})\mathcal{R}=
\mathcal{R}\mathrm{\Delta}(\gen{J})
\]
for all generators $\gen{J}$ of the algebra,
where $\widetilde{\copro}$ is the opposite coproduct.
It turns out to be fully constrained by this
relation up to one overall factor $R^0_{12}$.
The result is lengthy, and it can be found in \cite{Beisert:2008tw};
we refrain from reproducing it here.
\section{Classical Limit}
\label{sec:classical}
The classical limit of quantum-deformed R-matrices typically
consists in sending the deformation parameter $q$ to unity, $q\to 1$.
For the undeformed R-matrix \cite{Beisert:2005tm}, however, the classical limit
involves a large coupling constant, $g\to\infty$,
cf.\ \cite{Klose:2006zd,Torrielli:2007mc}.
Furthermore, the parameters for the fundamental representation
have to scale in a particular fashion such that $\xpm{}$ approach
a common finite value, $\xpm{}\to x$.
We find that a reasonable classical limit consists in setting
\[\label{eq:qclass}
q=1+\frac{h}{2g}+\order{g^{-2}},
\]
with the inverse coupling constant $g^{-1}$ taking the role of the quantum parameter $\hbar$
\[\label{eq:gclass}
g\to\infty,
\]
while $h$ remains a finite deformation parameter even in the classical limit.
For later purposes we shall also introduce $h'$ as the combination
\[
h'=\sqrt{1-h^2}\,.
\]
\subsection{Fundamental Representation}
For the parameters of the fundamental representation we
assume the following classical limit%
\footnote{A simultaneous sign flip of $h'$ and $x$ changes nothing.}
\[\label{eq:xpmclass}
\xpm{}=(h'x-ih)\lrbrk{1\pm\frac{1}{2g}\,\frac{x(hx+ih')}{x^2-1}
+\order{g^{-2}}}.
\]
These obey the constraint \eqref{eq:xpmrel} up to the order given.
The coefficients $a,b,c,d$ in \eqref{eq:xpmparameters}
then take the classical values
\[\label{eq:abcdclass}
a= \gamma,
\qquad
b= -\frac{\alpha (h'x-ih)(hx+ih')}{\gamma h'(x^2-1)}\,,
\qquad
c= \frac{i\gamma}{\alpha (h'x-ih)}\,,
\qquad
d= \frac{x(h'x-ih)}{\gamma h'(x^2-1)}\,.
\]
One can see that $ad-bc=1$ as desired for the classical limit $q\to1$.
The limit of the central charges $D,C,P,K$ in \eqref{eq:q2c,eq:PKU}
then follows as
\[
D=-\sfrac{1}{2} h^{-1}(z+1)q,\qquad
C=\sfrac{1}{2} h^{-1}(z-1)q,\qquad
P=\alpha q,\qquad
K=-\alpha^{-1} z q,
\]
where $z$ and $q$
(the quantum parameter $q$ will not appear in the classical limit
and we can use the letter for a different purpose) are defined by
\[\label{eq:zq
z=\frac{ix}{(h'x-ih)(hx+ih')}\,,\qquad
q=\frac{-(h'x-ih)(hx+ih')}{h'(x^2-1)}\,.
\]
\subsection{Fundamental r-Matrix}
We now take the classical limit
\eqref{eq:qclass,eq:gclass,eq:xpmclass}
on the fundamental R-matrix $\mathcal{R}$ found in \cite{Beisert:2008tw}.
In the strict classical limit it reduces to the unity operator
and the first non-trivial order equals the
fundamental classical r-matrix $r$
\[\label{eq:Rlimit}
(-A_{12}D_{12})^{-1/2}\mathcal{R}=1\otimes 1+\frac{h}{g}\,r+\order{g^{-2}}.
\]
Note that for definiteness we have multiplied the fundamental R-matrix by
a combination of the coefficient functions $A_{12}$ and $D_{12}$ in \cite{Beisert:2008tw}.
This removes the undetermined overall coefficient function $R^0_{12}$,
or it effectively fixes it to a convenient expression.
\begin{table}
\begin{eqnarray*}
r\state{\phi^1\phi^1}\earel{=}
A_{12}\state{\phi^1\phi^1}
\nonumber\\
r\state{\phi^1\phi^2}\earel{=}
\sfrac{1}{2} (A_{12}+B_{12}+1)\state{\phi^2\phi^1}
+\sfrac{1}{2} (A_{12}-B_{12})\state{\phi^1\phi^2}
-\sfrac{1}{2} C_{12} \varepsilon_{\alpha\beta}\state{\psi^\alpha\psi^\beta}
\nonumber\\
r\state{\phi^2\phi^1}\earel{=}
\sfrac{1}{2} (A_{12}-B_{12})\state{\phi^2\phi^1}
+\sfrac{1}{2} (A_{12}+B_{12}-1)\state{\phi^1\phi^2}
+\sfrac{1}{2} C_{12}\varepsilon_{\alpha\beta}\state{\psi^\alpha\psi^\beta}
\nonumber\\
r\state{\phi^2\phi^2}\earel{=}
A_{12}\state{\phi^2\phi^2}
\nonumber\\[10pt]
r\state{\psi^1\psi^1}\earel{=}
-D_{12}\state{\psi^1\psi^1}
\nonumber\\
r\state{\psi^1\psi^2}\earel{=}
-\sfrac{1}{2} (D_{12}+E_{12}+1)\state{\psi^2\psi^1}
-\sfrac{1}{2} (D_{12}-E_{12})\state{\psi^1\psi^2}
+\sfrac{1}{2} F_{12}\varepsilon_{ab}\state{\phi^a\phi^b}
\nonumber\\
r\state{\psi^2\psi^1}\earel{=}
-\sfrac{1}{2} (D_{12}-E_{12})\state{\psi^2\psi^1}
-\sfrac{1}{2} (D_{12}+E_{12}-1)\state{\psi^1\psi^2}
-\sfrac{1}{2} F_{12}\varepsilon_{ab}\state{\phi^a\phi^b}
\nonumber\\
r\state{\psi^2\psi^2}\earel{=}
-D_{12}\state{\psi^2\psi^2}
\nonumber\\[10pt]
r\state{\phi^a\psi^\beta}\earel{=}
G_{12}\state{\phi^a\psi^\beta}
+H_{12}\state{\psi^\beta\phi^a}
\nonumber\\
r\state{\psi^\alpha\phi^b}\earel{=}
K_{12}\state{\phi^b\psi^\alpha}
+L_{12}\state{\psi^\alpha\phi^b}
\nonumber
\end{eqnarray*}
\caption{The fundamental trigonometric r-matrix.}
\label{tab:rmatrix}
\end{table}
\begin{table}
\begin{eqnarray*}
A_{12}
=D_{12}
\earel{=}
\frac{\sfrac{1}{4} z_1+\sfrac{1}{4} z_2+\sfrac{1}{4} z_1 q_1 q_2^{-1}+\sfrac{1}{4} z_2 q_2 q_1^{-1}}{z_1-z_2}
\nonumber\\
\sfrac{1}{2}(A_{12}+B_{12}+1)
=\sfrac{1}{2}(D_{12}+E_{12}+1)
\earel{=} \frac{z_1}{z_1-z_2}
\nonumber\\
\sfrac{1}{2}(A_{12}+B_{12}-1)
=\sfrac{1}{2}(D_{12}+E_{12}-1)
\earel{=} \frac{z_2}{z_1-z_2}
\nonumber\\
\sfrac{1}{2}(A_{12}-B_{12})
=\sfrac{1}{2}(D_{12}-E_{12})
\earel{=}
\frac{-\sfrac{1}{4} z_1-\sfrac{1}{4} z_2+\sfrac{1}{4} z_1 q_1 q_2^{-1}+\sfrac{1}{4} z_2 q_2 q_1^{-1}}{z_1-z_2}
\nonumber\\
C_{12}\earel{=}
\frac{z_1a_1c_2-z_2a_2c_1}{z_1-z_2}
\nonumber\\
F_{12}\earel{=}
\frac{z_1b_1d_2-z_2b_2d_1}{z_1-z_2}
\nonumber\\
G_{12}
=-L_{12}
\earel{=}
\frac{-\sfrac{1}{4} z_1 q_1 q_2^{-1}+\sfrac{1}{4} z_2 q_2 q_1^{-1}}{z_1-z_2}
\nonumber\\
H_{12}\earel{=}
\frac{z_1a_1d_2-z_2b_2c_1}{z_1-z_2}
\nonumber\\
K_{12}\earel{=}
\frac{-z_1b_1c_2+z_2a_2d_1}{z_1-z_2}
\nonumber
\end{eqnarray*}
\caption{The coefficients for the fundamental r-matrix.}
\label{tab:rcoeffs}
\end{table}
The resulting form of the fundamental classical r-matrix is given
in \tabref{tab:rmatrix}.
It is determined by ten coefficient functions $A,B,C,D,E,F,G,H,K,L$.
Their values in the classical limit are given in \tabref{tab:rcoeffs}.
The matrix $r$ inherits the classical Yang--Baxter equation $\cybe{r}{r}=0$
from its quantum counterpart \cite{Beisert:2008tw}, where
\[\label{eq:cybeop}
\cybe{r}{s}:=
\comm{r_{12}}{s_{13}}
+\comm{r_{12}}{s_{23}}
+\comm{r_{13}}{s_{23}}.
\]
Taking a closer look at the coefficients we find four identities
among them: two linear ones
\[\label{eq:rellin}
A-D=-B+E=G+L
\]
and two quadratic identities
\[\label{eq:relquad}
\sfrac{1}{4}(A+B-1)(A+B+1)
=
\sfrac{1}{4}(3A-B)(3D-E)+4GL
=CF+HK.
\]
Note that we cannot in general claim that $A=D$
as suggested by \tabref{tab:rcoeffs}
because it follows only from our above choice
of prefactor $R^0_{12}$ in \eqref{eq:Rlimit}.
In other words, unlike the above four constraints
the latter one is not invariant under the shift
proportional to the identity matrix
\[\label{eq:identityshift}
\delta (A,B,C,D,E,F,G,H,K,L)\sim (+1,-1,0,-1,+1,0,+1,0,0,+1),
\]
which corresponds to changing the overall scattering phase.
Altogether this reduces the $10$ coefficient functions
to merely $6$ independent ones. This equals the number of free parameters:
$x_1$, $x_2$, $\gamma_1$, $\gamma_2$, $h$ and the freedom
to shift by the identity matrix \eqref{eq:identityshift}.
\subsection{Lie Bialgebra}
\label{sec:loops}
In the following we shall derive a framework
for the above r-matrix in terms of a Lie bialgebra.
\paragraph{Fundamental Representation and r-Matrix.}
First we would like to turn the fundamental r-matrix into
a universal one to make it applicable to arbitrary representations.
This is achieved by converting the operations in $r$,
e.g.\ $\state{\phi^1\phi^2}\mapsto\state{\phi^2\phi^1}$,
to representations of symmetry generators, e.g.\ $-\gen{R}^{22}\otimes\gen{R}^{11}$,
acting individually on the two sites.
The operators $\gen{R},\gen{L},\gen{Q},\gen{S},\gen{A},\gen{B}$
are meant to mimic the fundamental representation of $\alg{gl}(2|2)$:
The two sets of $\alg{sl}(2)$ generators $\gen{R}^{ab}=\gen{R}^{ba}$
and $\gen{L}^{\alpha\beta}=\gen{L}^{\beta\alpha}$
act canonically on the two pairs of states $\state{\phi^a},\state{\psi^\alpha}$.
The remaining operators are set up in analogy to \cite{Beisert:2007ty}
to be able to reproduce the coefficients in \tabref{tab:rcoeffs}.
The action of the supercharges $\gen{Q}^{\alpha b}$ and $\gen{S}^{\alpha b}$
is specified through the parameters $a,b,c,d$.
Finally, the action of the derivation $\gen{B}$
and the central charge $\gen{A}$ involves $q$.
Altogether the action reads
\[\label{eq:loopeval}
\begin{array}[b]{rclcrcl}
\gen{R}^{ab}\state{\phi^c}\earel{=} \sfrac{1}{2}\varepsilon^{bc}\state{\phi^a}+\sfrac{1}{2}\varepsilon^{ac}\state{\phi^b},
&&
\gen{L}^{\alpha\beta}\state{\psi^\gamma}\earel{=} \sfrac{1}{2}\varepsilon^{\beta\gamma}\state{\psi^\alpha}-\sfrac{1}{2}\varepsilon^{\alpha\gamma}\state{\psi^\beta},
\\[0.65ex]
\gen{Q}^{\alpha b}\state{\phi^c}\earel{=} a\,\varepsilon^{bc}\state{\psi^\alpha},
&&
\gen{Q}^{\alpha b}\state{\psi^\gamma}\earel{=} -b\,\varepsilon^{\alpha\gamma}\state{\phi^b},
\\[0.65ex]
\gen{S}^{\alpha b}\state{\phi^c}\earel{=} -c\,\varepsilon^{bc}\state{\psi^\alpha},
&&
\gen{S}^{\alpha b}\state{\psi^\gamma}\earel{=} d\,\varepsilon^{\alpha\gamma}\state{\phi^b},
\\[0.65ex]
\gen{A}\state{\phi^a}\earel{=} \sfrac{1}{2} q\,\state{\phi^a},
&&
\gen{A}\state{\psi^\alpha}\earel{=} \sfrac{1}{2} q\,\state{\psi^\alpha},
\\[0.65ex]
\gen{B}\state{\phi^a}\earel{=} -\sfrac{1}{2} q^{-1}\state{\phi^a},
&&
\gen{B}\state{\psi^\alpha}\earel{=} +\sfrac{1}{2} q^{-1}\state{\psi^\alpha}.
\end{array}
\]
The symbol $\varepsilon^{\cdot\cdot}$
is the antisymmetric $2\times 2$ matrix with $\varepsilon^{12}=+1$.
One can make contact with the fundamental
representation of the quantum algebra in \eqref{eq:fundquant}
by means of the following identification
with the Chevalley--Serre generators
\[
\begin{array}[b]{rclcrclcrcl}
\gen{H}_1\earel{=} +2\gen{R}^{12} ,&&
\gen{E}_1\earel{=} -\gen{R}^{22},&&
\gen{F}_1\earel{=} +\gen{R}^{11},
\\[0.65ex]
\gen{H}_2\earel{=}-h^{-1}(z-1)\gen{A}-\sfrac{1}{2}\gen{H}_1-\sfrac{1}{2}\gen{H}_3,&&
\gen{E}_2\earel{=}+\gen{Q}^{11},&&
\gen{F}_2\earel{=}-\gen{S}^{22},
\\[0.65ex]
\gen{H}_3\earel{=} -2\gen{L}^{12} ,&&
\gen{E}_3\earel{=} -\gen{L}^{22},&&
\gen{F}_3\earel{=} +\gen{L}^{11}.
\end{array}\]
We are then led to the following form for the classical r-matrix
from which the various coefficients in \tabref{tab:rcoeffs} are easily reproduced
\<\label{eq:reval}
r\earel{=}
+\frac{\sfrac{1}{2} z_1+\sfrac{1}{2} z_2}{z_1-z_2}\,
2\gen{R}^{12}\otimes \gen{R}^{12}
-\frac{z_1}{z_1-z_2}\,
\gen{R}^{22}\otimes \gen{R}^{11}
-\frac{z_2}{z_1-z_2}\,
\gen{R}^{11}\otimes \gen{R}^{22}
\nl
-\frac{\sfrac{1}{2} z_1+\sfrac{1}{2} z_2}{z_1-z_2}\,
2\gen{L}^{12}\otimes \gen{L}^{12}
+\frac{z_1}{z_1-z_2}\,
\gen{L}^{22}\otimes \gen{L}^{11}
+\frac{z_2}{z_1-z_2}\,
\gen{L}^{11}\otimes \gen{L}^{22}
\nl
-\frac{z_1}{z_1-z_2}\,
\varepsilon_{\alpha\gamma}\varepsilon_{bd}\gen{Q}^{\alpha b}\otimes \gen{S}^{\gamma d}
+\frac{z_2}{z_1-z_2}\,
\varepsilon_{\alpha\gamma}\varepsilon_{bd}\gen{S}^{\alpha b}\otimes \gen{Q}^{\gamma d}
\nl
-\frac{z_1}{z_1-z_2}\,
\gen{A}\otimes \gen{B}
-\frac{z_2}{z_1-z_2}\,
\gen{B}\otimes \gen{A}.
\>
\paragraph{Lie Brackets.}
Next we consider the commutators of the operators in \eqref{eq:loopeval}
to the end that they become the brackets of a Lie algebra
and \eqref{eq:loopeval} define the fundamental representation.
From the way the indices are contracted in \eqref{eq:loopeval},
it is evident that $\gen{R}$ and $\gen{L}$ form two $\alg{sl}(2)$ algebras
and that the generators $\gen{Q}$ and $\gen{S}$
transform in fundamental representations under these
\[\label{eq:RLeval}
\begin{array}[b]{rclcrcl}
\comm{\gen{R}^{ab}}{\gen{R}^{cd}}
\earel{=}
\varepsilon^{bc}\gen{R}^{ad}+\varepsilon^{ad}\gen{R}^{bc},
&\quad&
\comm{\gen{L}^{\alpha\beta}}{\gen{L}^{\gamma\delta}}
\earel{=}
\varepsilon^{\beta\gamma}\gen{L}^{\alpha\delta}+\varepsilon^{\alpha\delta}\gen{L}^{\beta\gamma},
\\[0.65ex]
\comm{\gen{R}^{ab}}{\gen{Q}^{\gamma d}}
\earel{=}
\sfrac{1}{2}\varepsilon^{bd}\gen{Q}^{\gamma a}+\sfrac{1}{2}\varepsilon^{ad}\gen{Q}^{\gamma b},
&\quad&
\comm{\gen{L}^{\alpha\beta}}{\gen{Q}^{\gamma d}}
\earel{=}
\sfrac{1}{2}\varepsilon^{\beta\gamma}\gen{Q}^{\alpha d}+\sfrac{1}{2}\varepsilon^{\alpha\gamma}\gen{Q}^{\beta d},
\\[0.65ex]
\comm{\gen{R}^{ab}}{\gen{S}^{\gamma d}}
\earel{=}
\sfrac{1}{2}\varepsilon^{bd}\gen{S}^{\gamma a}+\sfrac{1}{2}\varepsilon^{ad}\gen{S}^{\gamma b},
&\quad&
\comm{\gen{L}^{\alpha\beta}}{\gen{S}^{\gamma d}}
\earel{=}
\sfrac{1}{2}\varepsilon^{\beta\gamma}\gen{S}^{\alpha d}+\sfrac{1}{2}\varepsilon^{\alpha\gamma}\gen{S}^{\beta d}.
\end{array}
\]
The action of $\gen{Q},\gen{S},\gen{A},\gen{B}$
depends on the parameters $a,b,c,d,q$,
but their commutators
can be written using only $z$ defined in \eqref{eq:zq}
\<\label{eq:DQeval}
\acomm{\gen{Q}^{\alpha b}}{\gen{Q}^{\gamma d}}
\earel{=} 2\alpha \varepsilon^{\alpha\gamma}\varepsilon^{bd} \gen{A},
\nonumber\\
\acomm{\gen{Q}^{\alpha b}}{\gen{S}^{\gamma d}}
\earel{=}
-\varepsilon^{\alpha\gamma}\gen{R}^{bd}
+\varepsilon^{bd}\gen{L}^{\alpha\gamma}
-\varepsilon^{\alpha\gamma}\varepsilon^{bd} h^{-1}(z-1) \gen{A},
\nonumber\\
\acomm{\gen{S}^{\alpha b}}{\gen{S}^{\gamma d}}
\earel{=} -2 \alpha^{-1}\varepsilon^{\alpha\gamma}\varepsilon^{bd}z\gen{A},
\nonumber\\
\comm{\gen{B}}{\gen{Q}^{\alpha b}}
\earel{=}
h^{-1}(z-1)\gen{Q}^{\alpha b}
+2 \alpha \gen{S}^{\alpha b},
\nonumber\\
\comm{\gen{B}}{\gen{S}^{\alpha b}}
\earel{=}
2 \alpha^{-1}z\gen{Q}^{\alpha b}
- h^{-1}(z-1)\gen{S}^{\alpha b}.
\>
Although the above generators and their relations are reminiscent of $\alg{gl}(2|2)$,
they cannot form a Lie algebra as they stand.
The point is that the above commutators depend on the representation parameter $z$,
whereas the structure constants must be universal to the Lie algebra as a whole.
The way out is to consider instead the loop algebra of $\alg{gl}(2|2)[z,z^{-1}]$
in the way proposed in \cite{Beisert:2007ty}:
The variable $z$ can be interpreted as the formal loop variable
and the above action as an evaluation representation.
Then the above commutation relations define Lie brackets
on the loop space $\alg{gl}(2|2)[z,z^{-1}]$.
The algebra is however not $\alg{gl}(2|2)[z,z^{-1}]$
because the above relations are not homogeneous in $z$.
It is rather a non-trivial deformation of $\alg{gl}(2|2)[z,z^{-1}]$.
We observe that several of the coefficients appearing in
\eqref{eq:DQeval} coincide.
It turns out useful to combine these coefficients
as well as $a,b,c,d$ into $2\times 2$ matrices $W$ and $T$, respectively
\[\label{eq:matrices}
T=\matr{cc}{a&-b\\-c&d},
\qquad
W= \matr{cc}{+h^{-1}(z-1)&2\alpha\\2\alpha^{-1} z&-h^{-1}(z-1)}.
\]
We note that $\det T=1$ and $\mathop{\mathrm{Tr}} W=0$.
Introducing a constant matrix $M$ we can write the relations required
to derive \eqref{eq:DQeval} in the compact form
\[\label{eq:Wdef}
TM=q WT,
\qquad
M=\matr{cc}{+1&0\\0&-1}.
\]
The above commutation relations \eqref{eq:DQeval} then read
\<\label{eq:DQmatrix}
\acomm{\gen{Q}^{\alpha b}}{\gen{Q}^{\gamma d}}
\earel{=} \varepsilon^{\alpha\gamma}\varepsilon^{bd} W_{12}(z)\,\gen{A},
\nonumber\\
\acomm{\gen{Q}^{\alpha b}}{\gen{S}^{\gamma d}}
\earel{=}
-\varepsilon^{\alpha\gamma}\gen{R}^{bd}
+\varepsilon^{bd}\gen{L}^{\alpha\gamma}
-\varepsilon^{\alpha\gamma}\varepsilon^{bd}W_{11}(z)\, \gen{A},
\nonumber\\
\acomm{\gen{S}^{\alpha b}}{\gen{S}^{\gamma d}}
\earel{=} -\varepsilon^{\alpha\gamma}\varepsilon^{bd} W_{21}(z)\,\gen{A},
\nonumber\\
\comm{\gen{B}}{\gen{Q}^{\alpha b}}
\earel{=}
W_{11}(z)\,\gen{Q}^{\alpha b}
+W_{12}(z)\,\gen{S}^{\alpha b},
\nonumber\\
\comm{\gen{B}}{\gen{S}^{\alpha b}}
\earel{=}
W_{21}(z)\,\gen{Q}^{\alpha b}
+W_{22}(z)\,\gen{S}^{\alpha b},
\>
suggesting that $\gen{Q}^{\alpha b}$
and $\gen{S}^{\alpha b}$
form a two-component vector on which these matrices can act.
Note that the combinations for the brackets of supercharges
are naturally associated to the symmetric matrix
\[
-WN=\matr{cc}{+W_{12}&-W_{11}\\+W_{22}&-W_{21}}
\quad\mbox{with}\quad
N=\matr{cc}{0&+1\\-1&0}.
\]
\paragraph{Universal r-Matrix.}
The combinations of $z_1$ and $z_2$ appearing in \eqref{eq:reval}
are common for trigonometric classical r-matrices.
We can split all of them into terms proportional to $z_1/(z_1-z_2)$ and $z_2/(z_1-z_2)$
\[\label{eq:rclassfunct}
r_{12}
=\frac{z_1}{z_1-z_2}\,s_{12}+\frac{z_2}{z_1-z_2}\,s_{21}
=s_{12}+\frac{z_2}{z_1-z_2}\,t_{12}.
\]
Here, $s$ and $t$ are following tensor products of generators
\<\label{eq:tr0}
s_{12}\earel{=}
\gen{R}^{12}\otimes \gen{R}^{12}
-\gen{R}^{22}\otimes \gen{R}^{11}
-\gen{L}^{12}\otimes \gen{L}^{12}
+\gen{L}^{22}\otimes \gen{L}^{11}
-\varepsilon_{\alpha\gamma}\varepsilon_{bd}\gen{Q}^{\alpha b}\otimes \gen{S}^{\gamma d}
-\gen{A}\otimes \gen{B},
\nonumber\\
s_{21}\earel{=}
\gen{R}^{12}\otimes \gen{R}^{12}
-\gen{R}^{11}\otimes \gen{R}^{22}
-\gen{L}^{12}\otimes \gen{L}^{12}
+\gen{L}^{11}\otimes \gen{L}^{22}
+\varepsilon_{\alpha\gamma}\varepsilon_{bd}\gen{S}^{\gamma d}\otimes\gen{Q}^{\alpha b}
-\gen{B}\otimes \gen{A},
\nonumber\\
t_{12}\earel{=}
s_{12}+s_{21}
\nonumber\\\earel{=}
-\varepsilon_{ac}\varepsilon_{bd}\gen{R}^{ab}\otimes \gen{R}^{cd}
+\varepsilon_{\alpha\gamma}\varepsilon_{\beta\delta}\gen{L}^{\alpha\beta}\otimes \gen{L}^{\gamma\delta}
\nl
-\varepsilon_{\alpha\gamma}\varepsilon_{bd}\gen{Q}^{\alpha b}\otimes \gen{S}^{\gamma d}
+\varepsilon_{\alpha\gamma}\varepsilon_{bd}\gen{S}^{\alpha b}\otimes \gen{Q}^{\gamma d}
-\gen{A}\otimes \gen{B}
-\gen{B}\otimes \gen{A}.
\>
The term $t$ is (graded) symmetric
and $r$ is (graded) anti-symmetric
\[\label{eq:rantisym}
t_{12}=t_{21}\,,\qquad
r_{12}+r_{21}=0.
\]
We can now consider the classical Yang--Baxter equation $\cybe{r}{r}=0$.
The form of $r$ coincides with the conventional trigonometric r-matrix for
the superalgebra $\alg{gl}(2|2)$ \cite{Leites:1984aa}
for which the CYBE holds indeed.
Therefore the only violations could arise
from the deformations in \eqref{eq:DQmatrix}.
We calculate the terms in $\cybe{r}{r}$ which consist of
one factor of $\gen{A}$ and two supercharges.
These turn out to vanish if the following
three equations hold for
$F_1(z)=zW_{12}(z)$, $F_2(z)=W_{11}(z)$ and $F_3(z)=z^{-1}W_{21}(z)$
\[
\frac{F_{k}(z_1)}{(z_1-z_2)(z_1-z_3)}
-\frac{F_{k}(z_2)}{(z_1-z_2)(z_2-z_3)}
+\frac{F_{k}(z_3)}{(z_1-z_3)(z_2-z_3)}
=0\,.
\]
Setting $z_3=0$ the equation reduces to
\[
\frac{F_k(z_1)-F_k(0)}{z_1}=\frac{F_k(z_2)-F_k(0)}{z_2}\,.
\]
This implies that $F_k(z)$ must be a polynomial of degree 1
which is indeed a solution of the above equation
and which is also true for all matrix elements in \eqref{eq:matrices}.
Therefore the CYBE is fulfilled, and the r-matrix
enhances the loop algebra to a triangular Lie bialgebra.
Note that the above three conditions also guarantee that
the algebra has a positive, a negative and a Cartan subalgebra,
see \eqref{eq:decompose} for more details.
\subsection{Loop Level Form}
\label{sec:levels}
In order to define the loop algebra more rigorously,
we shall provide an alternative presentation in terms of
the generators
at definite levels of the loop algebra
\[
\gen{J}_n\simeq z^n \gen{J}.
\]
This description of the loop algebra is instructive,
and it has in fact a slightly different bialgebra structure.
Nevertheless in the remainder of the paper
we shall mostly employ the functional description introduced above.
\paragraph{Lie Brackets.}
Based on the above operators
$\gen{J}\in\spn{\gen{R},\gen{L},\gen{Q},\gen{S},\gen{A},\gen{B}}=\alg{gl}(2|2)$
we define an algebra spanned by $\gen{J}_n$ for $n\in\mathbb{Z}$.
The vector space of the algebra is the one of $\alg{gl}(2|2)[z,z^{-1}]$,
but the Lie brackets are deformed:
The brackets involving the two sets of $\alg{sl}(2)$ generators $\gen{R}$ and $\gen{L}$
are precisely as in $\alg{gl}(2|2)[z,z^{-1}]$
\[\label{eq:braRL}
\begin{array}[b]{rclcrcl}
\bigcomm{\gen{R}^{ab}_m}{\gen{R}^{cd}_n}
\earel{=}
\varepsilon^{bc}\gen{R}^{ad}_{m+n}+\varepsilon^{ad}\gen{R}^{bc}_{m+n},
&\quad&
\bigcomm{\gen{L}^{\alpha\beta}_m}{\gen{L}^{\gamma\delta}_n}
\earel{=}
\varepsilon^{\beta\gamma}\gen{L}^{\alpha\delta}_{m+n}+\varepsilon^{\alpha\delta}\gen{L}^{\beta\gamma}_{m+n},
\\[0.65ex]
\bigcomm{\gen{R}^{ab}_m}{\gen{Q}^{\gamma d}_n}
\earel{=}
\sfrac{1}{2}\varepsilon^{bd}\gen{Q}^{\gamma a}_{m+n}+\sfrac{1}{2}\varepsilon^{ad}\gen{Q}^{\gamma b}_{m+n},
&\quad&
\bigcomm{\gen{L}^{\alpha\beta}_m}{\gen{Q}^{\gamma d}_n}
\earel{=}
\sfrac{1}{2}\varepsilon^{\beta\gamma}\gen{Q}^{\alpha d}_{m+n}+\sfrac{1}{2}\varepsilon^{\alpha\gamma}\gen{Q}^{\beta d}_{m+n},
\\[0.65ex]
\bigcomm{\gen{R}^{ab}_m}{\gen{S}^{\gamma d}_n}
\earel{=}
\sfrac{1}{2}\varepsilon^{bd}\gen{S}^{\gamma a}_{m+n}+\sfrac{1}{2}\varepsilon^{ad}\gen{S}^{\gamma b}_{m+n},
&\quad&
\bigcomm{\gen{L}^{\alpha\beta}_m}{\gen{S}^{\gamma d}_n}
\earel{=}
\sfrac{1}{2}\varepsilon^{\beta\gamma}\gen{S}^{\alpha d}_{m+n}+\sfrac{1}{2}\varepsilon^{\alpha\gamma}\gen{S}^{\beta d}_{m+n}.
\end{array}
\]
Only the brackets between supercharges $\gen{Q}$, $\gen{S}$
and the derivation $\gen{B}$ are modified.
They follow from the above commutators
for the fundamental representation
\eqref{eq:DQeval}
where the variable $z$ is interpreted as a shift by one level
\<\label{eq:braDQ}
\bigacomm{\gen{Q}^{\alpha b}_m}{\gen{Q}^{\gamma d}_n}
\earel{=} 2\alpha \varepsilon^{\alpha\gamma}\varepsilon^{bd} \gen{A}_{m+n},
\nonumber\\
\bigacomm{\gen{Q}^{\alpha b}_m}{\gen{S}^{\gamma d}_n}
\earel{=}
-\varepsilon^{\alpha\gamma}\gen{R}^{bd}_{m+n}
+\varepsilon^{bd}\gen{L}^{\alpha\gamma}_{m+n}
-\varepsilon^{\alpha\gamma}\varepsilon^{bd} h^{-1}\bigbrk{\gen{A}_{m+n+1}-\gen{A}_{m+n}},
\nonumber\\
\bigacomm{\gen{S}^{\alpha b}_m}{\gen{S}^{\gamma d}_n}
\earel{=} -2 \alpha^{-1} \varepsilon^{\alpha\gamma}\varepsilon^{bd}\gen{A}_{m+n+1},
\nonumber\\
\bigcomm{\gen{B}_m}{\gen{Q}^{\alpha b}_n}
\earel{=}
h^{-1}\bigbrk{\gen{Q}^{\alpha b}_{m+n+1}-\gen{Q}^{\alpha b}_{m+n}}
+2 \alpha \gen{S}^{\alpha b}_{m+n},
\nonumber\\
\bigcomm{\gen{B}_m}{\gen{S}^{\alpha b}_n}
\earel{=}
2 \alpha^{-1}\gen{Q}^{\alpha b}_{m+n+1}
- h^{-1}\bigbrk{\gen{S}^{\alpha b}_{m+n+1}-\gen{S}^{\alpha b}_{m+n}}.
\>
The remaining unspecified Lie brackets are trivial.
Altogether the Jacobi identities are satisfied as
can be confirmed explicitly.
The algebra has a family of four-dimensional evaluation representations
with $\gen{J}_n\simeq z^n \gen{J}$
and the action of $\gen{J}$ specified in \eqref{eq:loopeval}.
\paragraph{Universal r-Matrix.}
The functional r-matrix in \eqref{eq:rclassfunct}
can be cast into the loop level form.
To that end one expands the above function of $z$'s into a geometric series%
\footnote{This formula represents an analytic continuation
of the series, see below for additional distributional contributions.}
\[
\frac{z_2}{z_1-z_2}=
\sum_{k=1}^{\infty} \lrbrk{\frac{z_2}{z_1}}^{k}\,.
\]
The resulting r-matrix then reads explicitly
\<\label{eq:rclass}
r\earel{=}
\gen{R}^{12}_0\otimes \gen{R}^{12}_0
-\gen{R}^{22}_0\otimes \gen{R}^{11}_0
-\gen{L}^{12}_0\otimes \gen{L}^{12}_0
+\gen{L}^{22}_0\otimes \gen{L}^{11}_0
-\varepsilon_{\alpha\gamma}\varepsilon_{bd}\gen{Q}^{\alpha b}_0\otimes \gen{S}^{\gamma d}_0
-\gen{A}_0\otimes \gen{B}_0
\nl
+\sum_{k=1}^\infty \Big[
-\varepsilon_{ac}\varepsilon_{bd}\gen{R}^{ab}_{-k}\otimes \gen{R}^{cd}_{+k}
+\varepsilon_{\alpha\gamma}\varepsilon_{\beta\delta}\gen{L}^{\alpha\beta}_{-k}\otimes \gen{L}^{\gamma\delta}_{+k}
\nl\qquad\qquad
-\varepsilon_{\alpha\gamma}\varepsilon_{bd}\gen{Q}^{\alpha b}_{-k}\otimes \gen{S}^{\gamma d}_{+k}
+\varepsilon_{\alpha\gamma}\varepsilon_{bd}\gen{S}^{\alpha b}_{-k}\otimes \gen{Q}^{\gamma d}_{+k}
\nl\qquad\qquad
-\gen{A}_{-k}\otimes \gen{B}_{+k}
-\gen{B}_{-k}\otimes \gen{A}_{+k}
\Big].
\>
This r-matrix defines a quasi-triangular Lie bialgebra:
First of all the symmetric part of $r$
equals
\[\label{eq:rsyminfsum}
r_{12}+r_{21}=\sum_{k=-\infty}^\infty \lrbrk{\frac{z_2}{z_1}}^k t,
\]
which is an invertible quadratic invariant of the algebra.
It is straight-forward to convince oneself of this fact.
Note that this is slightly different than
in the functional form where $r_{12}+r_{21}=0$, cf.\ \eqref{eq:rantisym}.
Consequently, the resulting algebra is merely \emph{quasi}-triangular.
Furthermore, the classical Yang--Baxter equation $\cybe{r}{r}=0$,
see \eqref{eq:cybeop}, holds.
This is not as easily seen,
in particular it does not follow right away from
the discussion at the end of \secref{sec:loops}
because the expansion into loop levels introduces
slight modifications, cf.\ \eqref{eq:rantisym}
vs.\ \eqref{eq:rsyminfsum}.
The CYBE eventually follows
from the triangular decomposition of $\alg{gl}(2|2)[z,z^{-1}]$
into a positive, negative and Cartan subalgebra
$\alg{g}^+\oplus \alg{g}^- \oplus \alg{g}^0$
with
\<\label{eq:decompose}
\alg{g}^+\earel{=}
\spn{\gen{R}^{11}_0,\gen{L}^{11}_0,\gen{S}_0,\gen{B}_0}
\oplus z\,\alg{gl}(2|2)[z],
\nonumber\\
\alg{g}^0\earel{=}
\spn{\gen{R}^{12}_0,\gen{L}^{12}_0},
\nonumber\\
\alg{g}^-\earel{=}
\spn{\gen{R}^{22}_0,\gen{L}^{22}_0,\gen{Q}_0,\gen{A}_0}
\oplus z^{-1}\,\alg{gl}(2|2)[z^{-1}].
\>
The crucial observation which ensures quasi-triangularity
is that the r-matrix \eqref{eq:rclass}
belongs to the following subspace
(more precisely its compactification)
\[
r\in (\alg{g}^-\otimes \alg{g}^+) \oplus (\alg{g}^0\otimes \alg{g}^0).
\]
This r-matrix takes the form of the classical double of $\alg{g}^+\oplus\alg{g}^0$
divided by the centre generated by a combination of $\alg{g}^0$ and its dual.
Alternatively, one can say that
the decomposition $(\alg{g}^+\oplus \alg{g}^0\oplus\alg{g}^-,\alg{g}^+\oplus \alg{g}^0,\alg{g}^-\oplus \alg{g}^0)$
is a Manin triple up to the double appearance of the
Cartan subalgebra $\alg{g}^0$.
\paragraph{Distributions on the Complex Plane.}
To convert between the above two pictures for loop algebras
one conventionally uses the geometric series
\[\label{eq:geosum}
g(z):=\sum_{n=0}^\infty z^n,
\qquad
g(z)=\frac{1}{1-z}\mbox{ for }|z|<1.
\]
It is convenient to continue the function $g(z)$ analytically
to all $z\neq 1$, but some care is required because
it actually introduces inconsistencies:
Consider the contour integral of $z^k g(z)$
for a circle of radius $r$ around the origin.
One would like to obtain the following result
for the geometric series
(i.e.\ when performing the integral prior to the infinite sum)
\[
\frac{1}{2\pi i}\,\oint_r z^k g(z)\, dz=\delta_{k<0}.
\]
When analytically continuing the series $g(z)$ to all $z\neq 1$ one obtains a
different result
\[
\frac{1}{2\pi i}\,\oint_r \frac{z^k\,dz}{1-z}=
\begin{cases}
+\delta_{k<0}&\mbox{for }r<1,\\
-\delta_{k\geq 0}&\mbox{for }r>1.
\end{cases}\]
The difference between the two integrals equals $-1$ for $r>1$ irrespectively of the value of $k$.
Such a term can be thought of as to originate from a distributional term $\delta_{a,b}(z)$
which is supported on a curve between $a$ and $b$.%
\footnote{The distributions become somewhat more familiar,
see e.g.\ \cite{Faddeev:1987ph},
when the variables and integrations are restricted to the unit circle, $|z|=1$.}
The distribution is defined such that
for each (directed) crossing of the contour through the supporting curve,
the distribution contributes the value of the integrand at $z=0$.
Now the distributional result of the geometric series reads
\[
g(z)=\frac{1}{1-z}+2\pi i\delta_{0,\infty}(z-1),
\]
and the extra term w.r.t.\ \eqref{eq:geosum} is what
reduces a triangular algebra to a \emph{quasi}-triangular one.
Now $g(z)$ has a cut on the positive real axis extending from $z=1$ to $z=\infty$.
Each crossing of the cut from the lower towards the upper half plane
contributes the value of the integrand at $z=1$
\[
\frac{1}{2\pi i}\,\oint_r z^k\, 2\pi i\delta_{0,\infty}(z-1)\,dz=
\begin{cases}
0&\mbox{for }r<1,\\
1&\mbox{for }r>1.
\end{cases}\]
The quadratic invariant requires a geometric series
over both positive and negative powers.
For such series \eqref{eq:rsyminfsum}
the analytic contribution vanishes exactly,
while a distributional contribution remains
\[
\sum_{n=-\infty}^\infty z^n=g(z)+g(1/z)-1=
2\pi i\delta_{0,\infty}(z-1)
+2\pi i\delta_{0,\infty}(1/z-1)
=
2\pi i\delta_{-1,\infty}(z-1).
\]
We have made use of proper transformation rules for this distribution
which are analogous to those for delta functions.
Here the resulting branch cut extends from $z=0$ to $z=\infty$,
and for each crossing it contributes the value of the integrand at $z=1$.
In the remainder of the paper we will only make reference to this
type of distribution, written in the form
\[\label{eq:doublegeometric}
\sum_{n=-\infty}^\infty \lrbrk{\frac{z_1}{z_2}}^n=
2\pi i\delta_{-1,\infty}(z_1/z_2-1)
=:2\pi iz_1\delta(z_1-z_2).
\]
The latter is a convenient abbreviation of the former distribution:
Here the cut extends from $z_1=0$ to $z_1=\infty$
or alternatively from $z_2=\infty$ to $z_2=0$.
\section{Affine Extension}
\label{sec:affine}
A loop algebra can be extended by
one derivation $\gen{D}$ and one central charge $\gen{C}$
to an affine (Kac--Moody) Lie algebra.
Here we show that our deformed loop algebra also admits such
an affine extension.
\subsection{Example}
We shall use the example of $\alg{sl}(2)$ to illustrate the construction
of the affine extension.
The derivation $\gen{D}$ is defined as the following derivative w.r.t.\ $z$
\[\label{eq:derdiff}
\gen{D}=\frac{zd}{dz}\,.
\]
Put differently, $\gen{D}$ generates a scaling transformation of $z$.
Alternatively we can define $\gen{D}$ through its action on
the loop variable $z$ and the base generators $\gen{R}^{ab}$
\[\comm{\gen{D}}{z}=z,
\qquad
\comm{\gen{D}}{\gen{R}^{ab}}=0.
\]
The central charge appears in the brackets as follows
\[
\bigcomm{f(z)\gen{R}^{ab}}{g(z)\gen{R}^{cd}}
=
f(z)g(z)\bigbrk{\varepsilon^{bc}\gen{R}^{ad}+\varepsilon^{ad}\gen{R}^{bc}}
-\sfrac{1}{2}\lrbrk{\varepsilon^{ac}\varepsilon^{bd}+\varepsilon^{ad}\varepsilon^{bc}}
\frac{1}{2\pi i}\oint \bigbrk{f(z)dg(z)}
\,\gen{C}.
\]
Here the contour integral winds once around $z=0$ (or $z=\infty$).
Finally, we can write the quadratic invariant
using the delta distribution in \eqref{eq:doublegeometric}
\[
\hat t=
-2\pi iz_1\delta(z_1-z_2)
\varepsilon_{ac}\varepsilon_{bd}\gen{R}^{ab}\otimes \gen{R}^{cd}
-\gen{D}\otimes\gen{C}-\gen{C}\otimes\gen{D}.
\]
The above construction can be generalised
straight-forwardly to any loop algebra,
but for our deformed loop algebra some more work is needed
because of the non-homogeneous structure of the loop levels in \eqref{eq:DQeval}.
\subsection{Derivation}
Now we have to generalise the brackets with the affine derivation to
all generators of our loop algebra.
First of all, it acts on the loop parameter $z$
as a scaling transformation
\[\label{eq:affderloops}
\comm{\gen{D}}{z}=z
\quad
\mbox{or}
\quad
\gen{D}\simeq \frac{zd}{dz}\,.
\]
The derivations of the two sets of $\alg{sl}(2)$
generators $\gen{R}$ and $\gen{L}$ take the standard form
\[\label{eq:RLaffder}
\comm{\gen{D}}{\gen{R}^{ab}}=0,\qquad
\comm{\gen{D}}{\gen{L}^{\alpha\beta}}=0.
\]
For the remaining generators $\gen{Q},\gen{S},\gen{A},\gen{B}$
we can gain inspiration from the fundamental representation in \eqref{eq:loopeval}.
As compared to the fundamental representation of
the undeformed $\alg{gl}(2|2)$,
the representations of $\gen{Q}$ and $\gen{S}$ (as a 2-vector)
are rotated by the $\grp{SL}(2)$ matrix $T$ in \eqref{eq:matrices}
\cite{Hofman:2006xt}.
Furthermore the action of $\gen{A}$ and $\gen{B}$ is scaled by $q$
w.r.t.\ the undeformed $\alg{gl}(2|2)$.
The parameters $a,b,c,d,q$ depend on $x$ which is related to $z$ via \eqref{eq:zq}.
The derivation $\gen{D}$ transforms $z$ according to \eqref{eq:derdiff},
hence it modifies the matrix $T$.
The brackets of the derivation with the supercharges
must reflect this transformation in order to find a suitable
representation of $\gen{D}$.
We are thus led to the following combinations
\[\label{eq:UVdef}
z\,\frac{dT}{dz}\,T^{-1}
=U+f(z)W,
\qquad
\frac{z}{q}\,\frac{dq}{dz}=V,
\]
with%
\footnote{The conversion to levels of the loop algebra
along the lines of \secref{sec:levels}
is somewhat problematic due to the presence
of poles in $U(z)$ and $V(z)$ at $z\neq0,\infty$.
This issue deserves further investigations.}
\[\label{eq:UVmatrix}
U=
\frac{1}{z+z^{-1}-2+4h^2}\matr{cc}
{-h^2&+h\alpha\\-h\alpha^{-1}&+h^2},
\qquad
V=
-\frac{z-1+2h^2}{z+z^{-1}-2+4h^2}\,.
\]
The precise functional form of $\gamma$
influences the undetermined function $f(z)$.
For $f(z)=0$ we get a reasonably simple final expression
corresponding to the choice
\[
\gamma\simeq \frac{h'x-ih}{h'\sqrt{x^2-1}}\,.
\]
The matrix $U$ now appears as the derivation
of the two-vector of the bare supercharges $\gen{Q}$ and $\gen{S}$.
Altogether the derivations are specified by
\<\label{eq:DQaffder}
\comm{\gen{D}}{\gen{Q}^{\alpha b}}
\earel{=}
U_{11}(z)\,\gen{Q}^{\alpha b}
+U_{12}(z)\,\gen{S}^{\alpha b},
\nonumber\\
\comm{\gen{D}}{\gen{S}^{\alpha b}}
\earel{=}
U_{21}(z)\,\gen{Q}^{\alpha b}
+U_{22}(z)\,\gen{S}^{\alpha b},
\nonumber\\
\comm{\gen{D}}{\gen{A}}
\earel{=} +V(z)\,\gen{A},
\nonumber\\
\comm{\gen{D}}{\gen{B}}
\earel{=} -V(z)\,\gen{B}.
\>
Note that the ambiguity in \eqref{eq:UVdef}
corresponds to shifting $\gen{D}$ by $f(z)\gen{B}$,
cf.\ \eqref{eq:DQmatrix};
nothing is lost by making a specific choice as the above.
The Jacobi identities require
\[
z\,\frac{dW}{dz}=\comm{U}{W}-VW,
\]
which follows by combining \eqref{eq:Wdef} with \eqref{eq:UVdef}.
As an aside, we note that the derivation $\gen{D}$ can be extended to
a Virasoro algebra $\gen{D}_n=z^n\gen{D}$
with a new central charge $\gen{c}$,
but we will not make use of it here.
\subsection{Central Charge}
The central charge appears in the brackets of the two sets of $\alg{sl}(2)$
generators in the standard fashion
\<\label{eq:RLcentral}
\bigcomm{f(z)\gen{R}^{ab}}{g(z)\gen{R}^{cd}}
\earel{=}
f(z)g(z)\bigbrk{\varepsilon^{bc}\gen{R}^{ad}-\varepsilon^{ad}\gen{R}^{bc}}
\nl
-\sfrac{1}{2}\lrbrk{\varepsilon^{ac}\varepsilon^{bd}+\varepsilon^{ad}\varepsilon^{bc}}
\frac{1}{2\pi i}\oint \bigbrk{f(z)dg(z)}\,\gen{C},
\nonumber\\
\bigcomm{f(z)\gen{L}^{\alpha\beta}}{g(z)\gen{L}^{\gamma\delta}}
\earel{=}
f(z)g(z)\bigbrk{\varepsilon^{\beta\gamma}\gen{L}^{\alpha\delta}-\varepsilon^{\alpha\delta}\gen{L}^{\beta\gamma}}
\nl
+\sfrac{1}{2}\lrbrk{\varepsilon^{\alpha\gamma}\varepsilon^{\beta\delta}+\varepsilon^{\alpha\delta}\varepsilon^{\beta\gamma}}
\frac{1}{2\pi i}\oint \bigbrk{f(z)dg(z)}\,\gen{C}.
\>
For the remaining generators $\gen{Q},\gen{S},\gen{A},\gen{B}$
the brackets leading to the central charge have to be adjusted
to the deformations in \eqref{eq:DQaffder}.
There are several ways to derive a central charge
for the above loop algebra. A very convenient method
consists in demanding invariance of the
quadratic invariant, cf.\ \eqref{eq:doublegeometric},
\[
\hat t=
2\pi i z_1\delta(z_1-z_2)t
-\gen{C}\otimes\gen{D}
-\gen{D}\otimes\gen{C}
\]
where $t$ is given in \eqref{eq:tr0}.
The invariance under the loop generators
requires a balancing of two types of terms:
The contributions from brackets with $\gen{D}$
must cancel the contribution from brackets
proportional to the central charge.
One can easily figure out the central charge
contributions complementary to \eqref{eq:DQaffder}
\<\label{eq:DQcentral}
\bigacomm{f(z)\gen{Q}^{\alpha b}}{g(z)\gen{Q}^{\gamma d}}
\earel{=}
\varepsilon^{\alpha\gamma}\varepsilon^{bd} f(z)g(z)W_{12}(z)\, \gen{A}
\nl
+\varepsilon^{\alpha\gamma}\varepsilon^{bd}
\frac{1}{2\pi i}\oint \lrbrk{f(z)g(z)U_{12}(z)\,\frac{dz}{z}}\,\gen{C},
\nonumber\\
\bigacomm{f(z)\gen{Q}^{\alpha b}}{g(z)\gen{S}^{\gamma d}}
\earel{=}
f(z)g(z)
\lrbrk{
-\varepsilon^{\alpha\gamma}\gen{R}^{bd}
+\varepsilon^{bd}\gen{L}^{\alpha\gamma}
-\varepsilon^{\alpha\gamma}\varepsilon^{bd}W_{11}(z)\, \gen{A}
}
\nl
+\varepsilon^{\alpha\gamma}\varepsilon^{bd}
\frac{1}{2\pi i}\oint
\lrbrk{f(z)dg(z)-f(z)g(z)U_{11}(z)\,\frac{dz}{z}}
\,\gen{C},
\nonumber\\
\bigacomm{f(z)\gen{S}^{\alpha b}}{g(z)\gen{S}^{\gamma d}}
\earel{=}
-\varepsilon^{\alpha\gamma}\varepsilon^{bd}f(z)g(z) W_{21}(z)\, \gen{A}
\nl
+\varepsilon^{\alpha\gamma}\varepsilon^{bd}
\frac{1}{2\pi i}\oint
\lrbrk{-f(z)g(z)U_{21}(z)\,\frac{dz}{z}}
\,\gen{C},
\nonumber\\
\bigcomm{f(z)\gen{A}}{g(z)\gen{B}}
\earel{=}
-\frac{1}{2\pi i}\oint
\lrbrk{f(z)dg(z)-V(z) f(z)g(z)\,\frac{dz}{z}}
\,\gen{C}.
\>
Note that the value of the above integrals depends on the
choice of contours. Conventionally one assumes that
the functions $f(z)$ and $g(z)$ are holomorphic
except at $z=0$ and $z=\infty$. In that case the contour can take any path
that winds once around $z=0$ and $z=\infty$.
Here the functions $W(z)$ and $V(z)$ in \eqref{eq:UVmatrix}
introduce two extra poles $z^*_\pm$.
These could be used to define two additional central charges.
It appears that they behave much like $\delta(z-z^*_\pm)\gen{A}$
and therefore there may be no need to enlarge the algebra further.
The issues of how to put the contours and how to define the
affine central charge(s) need further investigations.
\subsection{Affine r-Matrix}
The affine extension of the r-matrix in \eqref{eq:rclassfunct} reads
\[
\hat r=r-\gen{C}\otimes\gen{D}.
\]
In the presence of $\gen{C}$ the additional term is needed
to fulfil the CYBE.
Note that one is free to add an antisymmetric term proportional to
$\gen{C}\otimes\gen{D}-\gen{D}\otimes\gen{C}$
to the above $\hat{r}$
\cite{Reshetikhin:1990ep}.
A curious feature of the r-matrix is that it is not
invariant under the affine derivation $\gen{D}$.
This is because the coefficients $U,V$ in
\eqref{eq:UVmatrix} of the action \eqref{eq:DQaffder}
depend on $z$. Effectively this implies that the
r-matrix is not a function of $z_2/z_1$ alone, but it
depends separately on $z_1$ and $z_2$.
Correspondingly, the cobracket of $\gen{D}$ becomes non-trivial.
A possible benefit of the affine extension is that it may
add further constraints on the r-matrix.
Without the extension it is possible to add to $r$ terms of
the form
\[
z_1^m z_2^n\gen{A}\otimes\gen{A}
\]
because $\gen{A}$ is a central element and hence it cannot be
seen within the CYBE.
One could also view the deformation
as a deformation of $z^m\gen{B}$ by $z^n\gen{A}$
which affects only the coalgebra but not the algebra.
In the presence of the affine extensions such deformations
may no longer be possible because $\gen{A}$ is no longer in the centre;
it has non-trivial brackets with $\gen{D}$ and $\gen{B}$.
Thus it would be interesting to derive constraints
on the permissible deformations of $r$ by $\gen{A}\otimes\gen{A}$.
\subsection{Fundamental Representation}
Let us reconsider the fundamental evaluation representation
\eqref{eq:loopeval}.
The regular loop generators act on a four-dimensional space
spanned by $\state{\phi^a}$ and $\state{\psi^\alpha}$.
Conversely, the derivation $\gen{D}$
acts as a scaling transformation \eqref{eq:derdiff} for the parameter $z$
which is related to the representation parameter $x$ through \eqref{eq:zq}.
Consequently we must promote the states to fields
$\state{\phi^a,x}$ and $\state{\psi^\alpha,x}$ so that $\gen{D}$ can act on them.
The representation of the affine algebra is therefore infinite-dimensional,
and it naturally models a field on a one-dimensional mass shell.
Effectively, the affine derivation corresponds to a (Lorentz) boost
of the mass shell.
In this case the cobrackets for $\gen{D}$ are non-trivial
and therefore Lorentz symmetry must be considered as deformed.
In the picture of fields $\state{\phi^a,x}$ and $\state{\psi^\alpha,x}$,
one can get a clearer understanding of the role of the generator $z^k\gen{B}$
(note that $x$ is related to $z$):
Eq.\ \eqref{eq:loopeval} suggests that it induces a $x$-dependent (i.e.\ gauge)
transformation for the normalisation of bosons w.r.t.\ fermions.
The role of $\gamma$ (which can now depend on $x$) is related:
It serves as a (functional) parameter of the representation,
and it fixes a particular normalisation for it.
This evaluation-type representation clearly has vanishing central charge $\gen{C}\simeq 0$.
However, there surely exist representations with non-vanishing central charge,
such as highest-weight representations.
In the physical context these may correspond to vertex operators.
It would be interesting to investigate charged representations of this algebra.
\section{Discrete Symmetries}
\label{sec:discrete}
Before we continue with particular limiting cases of the classical r-matrix,
we shall discuss some of its discrete symmetries.
These will help us understand the limits better
and also relate some cases to others.
\subsection{Conjugation}
\label{sec:conjugation}
The map \eqref{eq:zq} between $x$ and $z$ is quadratic and thus 2:1.
The underlying reason for this property is that for unitary
superalgebras there are four conjugate fundamental representations.
In the present algebra, however, there is just a one-parameter family of
fundamental representations parametrised through $x$.
For each value of $z$ there are two values of $x$ corresponding to
a pair of representations and its conjugate:
That this is possible in the first place is a special property of $\alg{sl}(2|2)$.
The representation of each $\alg{sl}(2)$ subalgebra is fundamental;
as such it is self-conjugate under transposition and
conjugation by an antisymmetric $2\times 2$ matrix $\varepsilon$,
i.e.\ for a traceless $2\times 2$ matrix $E$ one has
\[
E'=-\varepsilon E^{\scriptscriptstyle\mathrm{T}} \varepsilon^{-1}=E.
\]
However, the representation of the remaining generators
is not self-conjugate under the combined map for a $4\times 4$
supermatrix $E$ written in $2\times2$ blocks
\[\label{eq:stransconj}
E'=-\matr{c|c}{\varepsilon&0\\\hline0&\varepsilon}
E^{\scriptscriptstyle\mathrm{ST}}
\matr{c|c}{\varepsilon^{-1}&0\\\hline0&\varepsilon^{-1}}.
\]
The representation $E$ is parametrised through $x$ and $\gamma$.
The two values of $x$ corresponding to conjugate representations
are related by inversion \cite{Beisert:2008tw}
\[
x'=\frac{1}{x}\,,
\qquad
z'=z.
\]
The parameters \eqref{eq:abcdclass,eq:zq,eq:matrices}
for the fundamental representation map according to
\[
T'=T\matr{cc}{0&-1\\+1&0},\qquad
q'=-q,\qquad
\gamma'=\frac{\alpha}{\gamma}\,\frac{(h'x-ih)(hx+ih')}{h'(x^2-1)}\,.
\]
Note that the matrix multiplying $T$
corresponds to the supertranspose operation
which is $\mathbb{Z}_4$ periodic.
So for each value of $x$ there are two representations
which differ in sign for the odd generators
corresponding to a total of four fundamentals in superalgebras.
See also \secref{sec:uniform} for further comments.
This transformation involves only representations
and thus it can be applied to each of the two sites
of the fundamental r-matrix individually.
Under such a crossing transformation of $x_1,\gamma_1$ the coefficients in
\tabref{tab:rcoeffs} permute as follows
\[
\begin{array}[b]{rclcrcl}
A_{\bar 12}\earel{=} -\sfrac{1}{2} (A_{12}-B_{12}),
&\quad&
D_{\bar 12}\earel{=}-\sfrac{1}{2} (D_{12}-E_{12}),
\\[0.65ex]
\sfrac{1}{2}(A_{\bar 12}-B_{\bar 12})\earel{=} -A_{12},
&\quad&
\sfrac{1}{2}(D_{\bar 12}-E_{\bar 12})\earel{=}-D_{12},
\\[0.65ex]
G_{\bar 12}\earel{=} L_{12},
&\quad&
L_{\bar 12}\earel{=} G_{12},
\\[0.65ex]
H_{\bar 12}\earel{=} -F_{12},
&\quad&
F_{\bar 12}\earel{=} +H_{12},
\\[0.65ex]
K_{\bar 12}\earel{=} -C_{12},
&\quad&
C_{\bar 12}\earel{=} +K_{12}.
\end{array}
\]
The combinations $\sfrac{1}{2} (A_{12}+B_{12}\pm1)$ and $\sfrac{1}{2} (D_{12}+E_{12}\pm1)$
remain invariant.
This transformation is compensated by
the map \eqref{eq:stransconj} on the first site of
the fundamental r-matrix in \tabref{tab:rmatrix}.
The transformation for $x_2,\gamma_2$ is the same as above
except that $C,F,H,K$ transform differently
\[
H_{1\bar 2}= +C_{12},\qquad
C_{1\bar 2}= -H_{12},\qquad
K_{1\bar 2}= +F_{12},\qquad
F_{1\bar 2}= -K_{12}.
\]
Under the combined transformation of both sites
the coefficients are invariant up to the following permutations
\[
C_{\bar 1\bar 2}=F_{12},\qquad
F_{\bar 1\bar 2}=C_{12},\qquad
H_{\bar 1\bar 2}=K_{12},\qquad
K_{\bar 1\bar 2}=H_{12}.
\]
The above transposition map has two fixed points which will be of importance later
\[\label{eq:selfdual}
x^*_\pm=\pm 1\,,
\qquad
z^*_\pm=(ih\pm h')^2\,,
\qquad
q^*=\infty.
\]
These two points will be called \emph{self-dual}.
\subsection{Inversion}
\label{sec:inversion}
Another useful discrete map is the inversion of $z$.
It implies the following transformations of the related parameters
\[
x'= i\,\frac{hx+ih'}{h'x-ih}\,,\qquad
z'= \frac{1}{z}\,.
\]
The parameters of the fundamental representation transform according to
\[
T'= RT,\qquad
q'= zq,\qquad
\gamma'= \frac{\gamma}{h'x-ih}\,,\qquad
R=\matr{cc}{0&i\alpha\\i\alpha^{-1}&0}.
\]
These rules suggest that the following map
\[
\matr{c}{\gen{Q}^{\prime\,\alpha b}\\\gen{S}^{\prime\,\alpha b}}
= R\matr{c}{\gen{Q}^{\alpha b}\\\gen{S}^{\alpha b}},\qquad
\gen{A}'= z\gen{A},\qquad
\gen{B}'= z^{-1}\gen{B}.
\]
together with $z'=1/z$ is an algebra automorphism.
Indeed one can confirm that the algebra
in \secref{sec:classical,sec:affine}
is invariant under the map.
It does not, however, respect the decomposition
in \eqref{eq:decompose} underlying the r-matrix
which is therefore not invariant.
In particular, the subalgebras $\alg{g}^+$ and $\alg{g}^-$
in \eqref{eq:decompose} are interchanged except for the elements
$\gen{R}^{11},\gen{R}^{22},\gen{L}^{11},\gen{L}^{22}$.
To achieve a proper transformation we have to interchange them
using the map $z'=1/z$ with
\[\label{eq:inversionalg}
\begin{array}{rcl}
\gen{R}^{\prime\,ab}\earel{=}\varepsilon_{ac}\varepsilon_{bd}\gen{R}^{cd},
\\[0.65ex]
\gen{L}^{\prime\,\alpha\beta}\earel{=}\varepsilon_{\alpha\gamma}\varepsilon_{\beta\delta}\gen{L}^{\gamma\delta},
\end{array}
\qquad
\matr{c}{\gen{Q}^{\prime\,\alpha b}\\\gen{S}^{\prime\,\alpha b}}
= \varepsilon_{\alpha\gamma}\varepsilon_{bd} R\matr{c}{\gen{Q}^{\gamma d}\\\gen{S}^{\gamma d}},\qquad
\begin{array}{rcl}
\gen{A}'\earel{=} z\gen{A},
\\[0.65ex]
\gen{B}'\earel{=} z^{-1}\gen{B}.
\end{array}
\]
Under the inversion all the coefficients $A_{12},\ldots,L_{12}$ for
the fundamental r-matrix in \tabref{tab:rcoeffs} flip sign.
This yields an overall sign in the r-matrix except for the
elements $\sfrac{1}{2}(A_{12}+B_{12}\pm 1)$ and $\sfrac{1}{2}(D_{12}+E_{12}\pm 1)$
which are permuted. The permutation is compensated by the
transformation in \eqref{eq:inversionalg}.
\subsection{Statistics Flip}
\label{sec:statistics}
The superalgebras of the kind $\alg{psl}(n|n)$
have an exceptional automorphism \cite{Serganova:1985aa,Grantcharov:2004aa}:
It interchanges the two $\alg{sl}(n)$ factors
and thus flips the two gradings in certain representations.
It is responsible for the existence of the two types of
strange superalgebras.
The Lie brackets are invariant under
the exchange of the two $\alg{sl}(2)$ subalgebras
\[
\gen{R}^{\prime\,ab}= \gen{L}^{ab},\qquad
\gen{L}^{\prime\,\alpha\beta}= \gen{R}^{\alpha\beta}.
\]
At the level of the fundamental representation
the exchange is compensated by the map
\[\label{eq:BFmap}
\state{\phi^a}'=\state{\psi^a},\qquad
\state{\psi^\alpha}'=\state{\phi^\alpha}.
\]
Under this map the fundamental r-matrix in \tabref{tab:rmatrix}
flips sign provided that the coefficients
transform according to
\[\begin{array}[b]{rclcrclcrclcrclcrcl}
A'_{12}\earel{=} D_{12},&\quad&
B'_{12}\earel{=} E_{12},&\quad&
G'_{12}\earel{=} -L_{12},&\quad&
C'_{12}\earel{=} F_{12},&\quad&
H'_{12}\earel{=} K_{12},
\\[0.65ex]
D'_{12}\earel{=} A_{12},&\quad&
E'_{12}\earel{=} B_{12},&\quad&
L'_{12}\earel{=} -G_{12},&\quad&
F'_{12}\earel{=} C_{12},&\quad&
K'_{12}\earel{=} H_{12}.
\end{array}
\]
Note that the elements $C,F,H,K$
in \tabref{tab:rmatrix} receive an extra sign
due to the change of statistics of the states
when acting with the bifermionic contributions
\eqref{eq:rclass}.
For the coefficients in \tabref{tab:rcoeffs} this transformation is realised
by mapping the parameter $\gamma$ according to
\[
\gamma'=\frac{\alpha (h'x-ih)(hx+ih')x}{h'\,\gamma (x^2-1)}\,.
\]
The transformation of the coefficients for the fundamental representation
in \eqref{eq:abcdclass,eq:matrices} reads
\[
T'=
RT\matr{cc}{0&1\\1&0}
,
\qquad
q'=q,
\qquad
R=\frac{i}{h'}\matr{cc}{+h&+\alpha z^{-1}\\-\alpha^{-1}z&-h}.
\]
The off-diagonal matrix multiplying $T$ corresponds
to the action \eqref{eq:BFmap}.
The map implies the following transformation for the remaining generators
\[
\matr{c}{\gen{Q}^{\prime\,\alpha b}\\\gen{S}^{\prime\,\alpha b}}=
R
\matr{c}{\gen{Q}^{\alpha b}\\\gen{S}^{\alpha b}},
\qquad
\gen{A}'=\gen{A},
\qquad
\gen{B}'=-\gen{B}.
\]
This transformation respects the algebra in \secref{sec:classical,sec:affine}
and the decomposition \eqref{eq:decompose} while
it flips the sign of the r-matrix in \eqref{eq:rclassfunct}.
\subsection{Duality}
\label{sec:duality}
Further scrutiny suggests that there is a relationship between
r-matrices with global parameters $h$ and $h'$ interchanged.
The quadratic relation $h^2+h'^2=1$ implies various sign ambiguities in the map
which we can lift by choosing a different parameter
\[
h=\sfrac{1}{2}(k+k^{-1})\,,\qquad
h'=-\sfrac{i}{2}(k-k^{-1})\,.
\]
The interchange corresponds to the map $k'=ik$.
The coefficients of the fundamental r-matrix in \tabref{tab:rcoeffs}
turn out to be invariant under the transformation
\[
k'=ik,\qquad
z'=-z,\qquad
x'=x
\,,\qquad
\alpha'=-i\, \frac{k+k^{-1}}{k-k^{-1}}\,\alpha.
\]
The remaining parameters of the fundamental representation transform according to
\[
T'=RT,\qquad
q'=i\,\frac{k-k^{-1}}{k+k^{-1}}\,q,\qquad
\gamma'=\gamma,\qquad
R=\matr{cc}{1&0\\\alpha^{-1}h^{-1}z&1}.
\]
Again the algebra in \secref{sec:classical,sec:affine}
is invariant if one imposes the following map for the generators
\[
\matr{cc}{\gen{Q}^{\prime\,\alpha b}\\\gen{S}^{\prime\,\alpha b}}
=R\matr{cc}{\gen{Q}^{\alpha b}\\\gen{S}^{\alpha b}},
\qquad
\gen{A}'=i\,\frac{k-k^{-1}}{k+k^{-1}}\,\gen{A},
\qquad
\gen{B}'=-i\,\frac{k+k^{-1}}{k-k^{-1}}\,\gen{B}.
\]
Also the decomposition \eqref{eq:decompose} is respected,
and consequently the r-matrix is invariant.
In particular, the coefficients in \tabref{tab:rcoeffs} transform trivially.
\subsection{Reparametrisation}
\label{sec:uniform}
Here we introduce a change of variables
which helps to make some features of the algebra
discussed above somewhat more transparent.
This will be instructive to some extent,
but in the remainder of the paper we shall nevertheless stick
to the old variables.
\paragraph{Reparametrisation.}
We have seen in \secref{sec:conjugation}
that for each value of $z$ there are four
fundamental representations.
They are distinguished by different values of $x$ and $\gamma$.
For instance, for each $z$ the map
\eqref{eq:zq} permits two values for $x$,
and for each $x$ there is a pair of
representations distinguished by different signs for $\gamma$.
In fact one can introduce a new parameter $y$ to distinguish
all four fundamental representations corresponding to a
particular value $z$
\[
x=-\frac{y^2-1}{y^2+1}\,,
\quad
\gamma=\frac{y^2+k^2}{2ky}\,\eta,
\quad
z=-k^2\,\frac{y^4-1}{y^4-k^4}\,,
\quad
q=\frac{-1}{k^2(k-k^{-1})}\,\frac{y^4-k^4}{y^2}\,.
\]
At the same time we shall use the parameter $k$ introduced in
\secref{sec:duality} instead of $h$
\[
h=\sfrac{1}{2}(k+k^{-1}),
\qquad
h'=-\sfrac{i}{2}(k-k^{-1}),
\qquad
\alpha=\sfrac{1}{2}(k-k^{-1})\kappa.
\]
Altogether the following parametrisation yields a slightly more transparent picture.
This can be observed for the coefficients $a,b,c,d$ of the matrix $T$
which now take a very symmetric form
\[
a=\frac{\eta(y^2+k^2)}{2ky}\,,
\quad
b=\frac{-\kappa(y^2-k^2)}{2k\eta y}\,,
\quad
c=\frac{-\eta(y^2+1)}{\kappa (k-k^{-1})y}\,,
\quad
d=\frac{y^2-1}{\eta(k-k^{-1})y}\,.
\]
Notably, all the coefficients in \tabref{tab:rcoeffs}
now factor completely into terms $y^2\pm 1$, $y^2\pm k^2$
and $y_1^2\pm y_2^2$.
\paragraph{Special Points.}
Investigating the above expressions it becomes
clear that the $y$-plane has several special points:
The points $y^\circ_{+1,2,3,4}=\pm 1,\pm i$ map to $z=0$
while $y^\circ_{-1,2,3,4}=\pm k,\pm ik$ map to $z=\infty$.
Finally, the two points $y^*_{\pm}=0,\infty$ map to the self-dual
points $z^*_\pm=-k^{\mp 2}$ or $x^*_\pm=\pm 1$ in \eqref{eq:selfdual}.
The configurations of special points are displayed in \figref{fig:special}.
\begin{figure}\centering
\includegraphics{FigSphereZ.mps}\qquad
\includegraphics{FigSphereX.mps}\qquad
\includegraphics{FigSphereY.mps}%
\caption{The compactified complex plane for $z$, $x$ or $y$, respectively.
The points corresponding to $z^\circ_\pm=0,\infty$ are marked by $\circ$.
The self-dual points corresponding to $z^\ast_\pm$ are marked by $\ast$.
The spheres are divided into one, two or four regions
which are identified by a twist.}
\label{fig:special}
\end{figure}
\paragraph{Eigenbasis.}
A curious feature of the matrix $U$ in \eqref{eq:UVmatrix}
is that the $z$-dependence is in the prefactor only.
Hence the eigenvectors are constants and we
can use them as a new basis for $\gen{Q}$ and $\gen{S}$.
A matrix to perform the similarity transformation to the
eigenvectors is given by
\[
R=\matr{cc}{-2/(k-k^{-1})&\kappa k\\-\kappa^{-1}/(k-k^{-1})&\sfrac{1}{2} k^{-1}}.
\]
The resulting matrix $\tilde T$ containing the coefficients
$\tilde a,\tilde b,\tilde c,\tilde d$ then reads simply
\[
\tilde T=RT=\matr{cc}{\eta y&\kappa \eta^{-1}y\\-\sfrac{1}{2} \kappa^{-1}\eta y^{-1}&\sfrac{1}{2}\eta^{-1}y^{-1}}.
\]
The transformation curiously removes the diagonal terms in the matrix $\tilde W$
\[
\tilde W=RWR^{-1}=\frac{1}{k+k^{-1}}
\matr{cc}{0&-4\kappa (1+k^2z)\\\kappa^{-1}(1+k^{-2}z)&0},
\]
whereas by construction the matrix $\tilde U$ is diagonal
\[
\tilde U=RUR^{-1}=\frac{(k^4-1)z}{4(k^2+z)(1+k^2z)}
\matr{cc}{+1&0\\0&-1}.
\]
The only reason not to perform this similarity transformation once and for all
is that it obscures the linear combinations
of $\tilde{\gen{Q}}$ and $\tilde{\gen{S}}$ which
appear in the contribution \eqref{eq:tr0} to the r-matrix
and in the triangular decomposition \eqref{eq:decompose}.
We will thus stick to the original basis of $\gen{Q}$ and $\gen{S}$.
\paragraph{Embedding.}
The above reparametrisation has led to rational
expressions for the parameters $a,b,c,d,q$ of the fundamental representation.%
\footnote{In fact, also $x$ (but not $z$ itself)
is permissible because $T(x),q(x)$ are rational.}
We can use them to go one step further, and embed our algebra into the
standard algebra $\alg{gl}(2|2)[y,y^{-1}]$
(with $\bar W=M$, $\bar U=\bar V=0$)
in analogy to the transformation in \cite{Beisert:2007ty}
\[
\begin{array}{rcl}
\gen{R}^{ab}\earel{=} \bar{\gen{R}}^{ab},
\\[0.65ex]
\gen{L}^{\alpha\beta}\earel{=} \bar{\gen{L}}^{\alpha\beta},
\end{array}
\quad
\matr{c}{\gen{Q}^{\alpha b}\\\gen{S}^{\alpha b}}=
T(y)\matr{c}{\bar{\gen{Q}}^{\alpha b}\\\bar{\gen{S}}^{\alpha b}},
\quad
\begin{array}{rcl}
\gen{A}\earel{=} q(y)\bar{\gen{A}},
\\[0.65ex]
\gen{B}\earel{=} q(y)^{-1}\bar{\gen{B}},
\end{array}
\]
with $\eta=\sqrt{\kappa}$.
Note that one must allow for pole singularities
at the special points $y^\circ,y^*$.
In this sense, one has to require
that the Riemann surface underlying
the ambient algebra is a sphere with punctures
at all of these points, see \figref{fig:special},
not just at $y=0,\infty$ as for conventional loop algebras.
The reduction to our subalgebra is done
by twisting with the $\mathbb{Z}_4$-periodic automorphism of $\alg{gl}(2|2)$
\footnote{Although the automorphism is $\mathbb{Z}_4$-periodic,
it merely corresponds to a $\mathbb{Z}_2$-periodic \emph{outer} automorphism
of the $\alg{gl}(2|2)$ algebra,
see the discussion below \protect\eqref{eq:aboveZ2}.}
\[
y\to iy,
\quad
\bar{\gen{Q}}\to i\bar{\gen{S}},
\quad
\bar{\gen{S}}\to i\bar{\gen{Q}},
\quad
\bar{\gen{A}}\to -\bar{\gen{A}},
\quad
\bar{\gen{B}}\to -\bar{\gen{B}}.
\]
Furthermore, singularities at the fixed points $y^*=0,\infty$
are restricted to be at most double poles while
there can be poles of arbitrary order at the points $y^\circ$.
As above, this redefinition changes the form of the r-matrix \eqref{eq:rclassfunct}
and the triangular decomposition \eqref{eq:decompose},
and we shall refrain from making use of it subsequently.
It is nevertheless interesting because it
shifts the deformation from the algebra
to the r-matrix,
i.e.\ the conventional affine $\alg{gl}(2|2)$ algebra apparently
admits a non-standard r-matrix.
\paragraph{Discrete Transformations.}
The discrete transformations discussed above also simplify:
Essentially they map the various special points $y^\circ$ and $y^*$
into each other.
The conjugation symmetry discussed in \secref{sec:conjugation}
translates between the four conjugate fundamental representations
for each value of $z$. This is achieved through
\[
y'=iy,\qquad
\eta'=\frac{i\kappa}{\eta}\,.
\]
The inversion symmetry discussed in \secref{sec:inversion}
is invoked by
\[
y\to \frac{k}{y}\,,\qquad
\eta'=i\eta.
\]
The statistics flip symmetry in \secref{sec:statistics} requires
to change $\eta$ according to
\[
\eta'=\frac{\kappa}{\eta}\,\frac{y^2-1}{y^2+1}\,\frac{y^2-k^2}{y^2+k^2}\,.
\]
Finally, there is the duality discussed in \secref{sec:duality}
which relates 4 different values of $k$
\[
k'=ik,\qquad
\eta'=i\,\frac{y^2+k^2}{y^2-k^2}\,\eta,\qquad
\kappa'=-\kappa.
\]
A similar transformation does not change anything in the original
parametrisation
\[
k'=\frac{1}{k}\,,\qquad
y'=\frac{1}{y}\,,\qquad
\kappa'=-\kappa.
\]
Note that the point $k=\sqrt{i}$ is self-dual under a combination
of the above two duality maps.
This map thus becomes an additional symmetry of the $k=\sqrt{i}$ system
\[
y'=\frac{1}{y}\,,\qquad
\eta'=-i\,\frac{y^2+i}{y^2-i}\,\eta,\qquad
x'=-x,\qquad
z'=-z.
\]
It might be worth investigating if the self-dual point $k=\sqrt{i}$
has further interesting properties.
The other self-dual point $k=1$ is
discussed in the following section.
\section{Limits}
\label{sec:limits}
The r-matrix presented in \tabref{tab:rmatrix,tab:rcoeffs}
has a couple of interesting limits which themselves
lead to quasi-triangular Lie algebras.
We shall call the r-matrix of \secref{sec:classical}
the ``full trigonometric r-matrix''.
The limits will modify the attributes of the name accordingly.
\subsection{Full Rational Case}
\label{sec:rational}
The trigonometric r-matrix can be reduced to the
rational r-matrix \cite{Klose:2006zd,Torrielli:2007mc}
obtained in the context of the AdS/CFT duality.
To that end one takes the limit
\[\label{eq:ratlimit}
h=\epsilon\to 0,\qquad
x\sim \epsilon^0,\qquad
z=1+i\epsilon u+\order{\epsilon^2}.
\]
All of the following results are in full agreement with \cite{Beisert:2007ty}
where the structure and the underlying quasi-triangular Lie bialgebra
were obtained.
\paragraph{Fundamental r-Matrix.}
The parameters of the fundamental representation
\eqref{eq:abcdclass,eq:zq} become%
\footnote{For convenience,
one might absorb several factors of $i$ into the
definition of $q,\gen{A},\gen{B},\alpha$.}
\[
a=\gamma,\quad
b=\frac{-i\alpha x}{\gamma(x^2-1)}\,,\quad
c=\frac{i\gamma}{\alpha x}\,,\quad
d=\frac{x^2}{\gamma(x^2-1)}\,,\quad
u=x+\frac{1}{x}\,,\quad
q=\frac{-i x}{x^2-1}\,.
\]
In this limit the r-matrix diverges like $\epsilon^{-1}$ and needs
to be renormalised
\[
\tilde r=i\epsilon r.
\]
Most importantly, the divergence reduces
the structure of the r-matrix in \tabref{tab:rmatrix}
because the constant terms in the
combinations $\sfrac{1}{2} (A+B\pm 1)$ and $\sfrac{1}{2} (D+E\pm 1)$ drop out.
It can then be written in a manifestly $\alg{sl}(2)\oplus\alg{sl}(2)$
invariant fashion known for \emph{rational} r-matrices
\<\label{eq:rrational}
\tilde r\state{\phi^a\phi^b}\earel{=}
\sfrac{1}{2} (\tilde A_{12}+\tilde B_{12})\state{\phi^b\phi^a}
+\sfrac{1}{2} (\tilde A_{12}-\tilde B_{12})\state{\phi^a\phi^b}
+\sfrac{1}{2} \tilde C_{12} \varepsilon^{ab}\varepsilon_{\gamma\delta}
\state{\psi^\gamma\psi^\delta},
\nonumber\\
\tilde r\state{\psi^\alpha\psi^\beta}\earel{=}
-\sfrac{1}{2} (\tilde D_{12}+\tilde E_{12})\state{\psi^\beta\psi^\alpha}
-\sfrac{1}{2} (\tilde D_{12}-\tilde E_{12})\state{\psi^\alpha\psi^\beta}
-\sfrac{1}{2} \tilde F_{12}\varepsilon^{\alpha\beta}\varepsilon_{cd}\state{\phi^c\phi^d},
\nonumber\\
\tilde r\state{\phi^a\psi^\beta}\earel{=}
\tilde G_{12}\state{\phi^a\psi^\beta}
+\tilde H_{12}\state{\psi^\beta\phi^a},
\nonumber\\
\tilde r\state{\psi^\alpha\phi^b}\earel{=}
\tilde K_{12}\state{\phi^b\psi^\alpha}
+\tilde L_{12}\state{\psi^\alpha\phi^b}.
\>
The coefficient functions $\tilde A,\ldots,\tilde L$ are essentially
the same as $A,\ldots,L$ in \tabref{tab:rcoeffs},
but the $z$-dependence reduces according to the limit
\[
\frac{z_1}{z_1-z_2}
\sim
\frac{\sfrac{1}{2} z_1+\sfrac{1}{2} z_2}{z_1-z_2}
\sim
\frac{z_2}{z_1-z_2}
\to
\frac{1}{i\epsilon}\,\frac{1}{u_1-u_2}\,,
\]
where the factor of $1/i\epsilon$ is absorbed into the definition of the
r-matrix $\tilde r$.
The coefficients obey the same linear relations \eqref{eq:rellin}
as in the trigonometric case,
but the constant shift disappears from the
quadratic relations \eqref{eq:relquad}
\[
\sfrac{1}{4}(\tilde A+\tilde B)(\tilde D+\tilde E)
=
\sfrac{1}{4}(3\tilde A-\tilde B)(3\tilde D-\tilde E)+4\tilde G\tilde L
=\tilde C\tilde F+\tilde H\tilde K.
\]
\paragraph{Algebra and Universal r-Matrix.}
The loop algebra derived in \secref{sec:loops}
remains essentially the same.
One difference is that we shall use $u$ as the formal loop variable
instead of $z$. Thus the brackets in \eqref{eq:DQmatrix}
now read
\<\label{eq:DQmatrixrat}
\acomm{\gen{Q}^{\alpha b}}{\gen{Q}^{\gamma d}}
\earel{=} \varepsilon^{\alpha\gamma}\varepsilon^{bd} W_{12}(u)\,\gen{A},
\nonumber\\
\acomm{\gen{Q}^{\alpha b}}{\gen{S}^{\gamma d}}
\earel{=}
-\varepsilon^{\alpha\gamma}\gen{R}^{bd}
+\varepsilon^{bd}\gen{L}^{\alpha\gamma}
-\varepsilon^{\alpha\gamma}\varepsilon^{bd}W_{11}(u)\, \gen{A},
\nonumber\\
\acomm{\gen{S}^{\alpha b}}{\gen{S}^{\gamma d}}
\earel{=} -\varepsilon^{\alpha\gamma}\varepsilon^{bd} W_{21}(u)\,\gen{A},
\nonumber\\
\comm{\gen{B}}{\gen{Q}^{\alpha b}}
\earel{=}
W_{11}(u)\,\gen{Q}^{\alpha b}
+W_{12}(u)\,\gen{S}^{\alpha b},
\nonumber\\
\comm{\gen{B}}{\gen{S}^{\alpha b}}
\earel{=}
W_{21}(u)\,\gen{Q}^{\alpha b}
+W_{22}(u)\,\gen{S}^{\alpha b}.
\>
Furthermore the matrix $W$ in \eqref{eq:matrices}
reduces to%
\footnote{The generators $\gen{B},\gen{A}$ are shifted
by one loop level w.r.t.\ the corresponding ones in
\protect\cite{Beisert:2007ty},
i.e.\ $W$ differs by a factor of $u$.}
\[
W= \matr{cc}{+i u&2\alpha\\2\alpha^{-1}&-i u}.
\]
The limit of the universal trigonometric r-matrix in \eqref{eq:rclassfunct}
reads
\[
\tilde r=\frac{1}{u_1-u_2}\,\tilde t.
\]
When expanded into loop levels using a geometric series (cf.\ \secref{sec:levels})
one finds the analog of \eqref{eq:rclass}
\<
\tilde r\earel{=}
+\sum_{k=0}^\infty \Big[
-\varepsilon_{ac}\varepsilon_{bd}\gen{R}^{ab}_{-1-k}\otimes \gen{R}^{cd}_{+k}
+\varepsilon_{\alpha\gamma}\varepsilon_{\beta\delta}\gen{L}^{\alpha\beta}_{-1-k}\otimes \gen{L}^{\gamma\delta}_{+k}
\nl\qquad\qquad
-\varepsilon_{\alpha\gamma}\varepsilon_{bd}\gen{Q}^{\alpha b}_{-1-k}\otimes \gen{S}^{\gamma d}_{+k}
+\varepsilon_{\alpha\gamma}\varepsilon_{bd}\gen{S}^{\alpha b}_{-1-k}\otimes \gen{Q}^{\gamma d}_{+k}
\nl\qquad\qquad
-\gen{A}_{-1-k}\otimes \gen{B}_{+k}
-\gen{B}_{-1-k}\otimes \gen{A}_{+k}
\Big].
\>
\paragraph{Affine Extension.}
The loop variable $z$ is replaced by $u$ according to \eqref{eq:ratlimit}
\[
z=1+i\epsilon u+\order{h^2}.
\]
After a rescaling
\[
\tilde{\gen{D}}=i\epsilon\gen{D}
\]
the affine derivation \eqref{eq:affderloops}
transforms into a derivative w.r.t.\ $u$
\[
\comm{\tilde{\gen{D}}}{u}=1
\quad
\mbox{or}
\quad
\tilde{\gen{D}}\simeq \frac{d}{du}\,.
\]
The structure of the affine derivations remains the same
as in \eqref{eq:DQaffder}
\<\label{eq:DQaffderrat}
\comm{\tilde{\gen{D}}}{\gen{Q}^{\alpha b}}
\earel{=}
\tilde{U}_{11}(u)\,\gen{Q}^{\alpha b}
+\tilde{U}_{12}(u)\,\gen{S}^{\alpha b},
\nonumber\\
\comm{\tilde{\gen{D}}}{\gen{S}^{\alpha b}}
\earel{=}
\tilde{U}_{21}(u)\,\gen{Q}^{\alpha b}
+\tilde{U}_{22}(u)\,\gen{S}^{\alpha b},
\nonumber\\
\comm{\tilde{\gen{D}}}{\gen{A}}
\earel{=} +\tilde{V}(u)\,\gen{A},
\nonumber\\
\comm{\tilde{\gen{D}}}{\gen{B}}
\earel{=} -\tilde{V}(u)\,\gen{B},
\>
whereas for the central charges in \eqref{eq:DQcentral}
one has to replace $dz/z$ by $du$
\<\label{eq:DQcentralrat}
\bigacomm{f(u)\gen{Q}^{\alpha b}}{g(u)\gen{Q}^{\gamma d}}
\earel{=}
\varepsilon^{\alpha\gamma}\varepsilon^{bd} f(u)g(u)W_{12}(u)\, \gen{A}
\nl
+\varepsilon^{\alpha\gamma}\varepsilon^{bd}
\frac{1}{2\pi i}\oint \bigbrk{f(u)g(u)U_{12}(u)\,du}\,\gen{C},
\nonumber\\
\bigacomm{f(u)\gen{Q}^{\alpha b}}{g(u)\gen{S}^{\gamma d}}
\earel{=}
f(u)g(u)
\lrbrk{
-\varepsilon^{\alpha\gamma}\gen{R}^{bd}
+\varepsilon^{bd}\gen{L}^{\alpha\gamma}
-\varepsilon^{\alpha\gamma}\varepsilon^{bd}W_{11}(u)\, \gen{A}
}
\nl
+\varepsilon^{\alpha\gamma}\varepsilon^{bd}
\frac{1}{2\pi i}\oint
\bigbrk{f(u)dg(u)-f(u)g(u)U_{11}(u)\,du}
\,\gen{C},
\nonumber\\
\bigacomm{f(u)\gen{S}^{\alpha b}}{g(u)\gen{S}^{\gamma d}}
\earel{=}
-\varepsilon^{\alpha\gamma}\varepsilon^{bd}f(u)g(u) W_{21}(u)\, \gen{A}
\nl
+\varepsilon^{\alpha\gamma}\varepsilon^{bd}
\frac{1}{2\pi i}\oint
\bigbrk{-f(u)g(u)U_{21}(u)\,du}
\,\gen{C},
\nonumber\\
\bigcomm{f(u)\gen{A}}{g(u)\gen{B}}
\earel{=}
-\frac{1}{2\pi i}\oint
\bigbrk{f(u)dg(u)-V(u) f(u)g(u)\,du}
\,\gen{C}.
\>
The parameters have to be rescaled w.r.t.\ \eqref{eq:UVmatrix},
now they read, see also \cite{Young:2007wd},
\[
\tilde U=i\epsilon U=
\frac{1}{u^2-4}\matr{cc}
{0&-i\alpha\\+i\alpha^{-1}&0},
\qquad
\tilde V=i\epsilon V=
-\frac{u}{u^2-4}\,.
\]
Note that the non-vanishing of the above parameters
leads to the non-invariance of the r-matrix under $\tilde{\gen{D}}$
and thus to a non-trivial cobracket (see \secref{sec:affine}).
When $\tilde{\gen{D}}$ is interpreted as a two-dimensional (Lorentz) boost,
the corresponding (Lorentz) symmetry would be deformed
along the lines discussed in \cite{Gomez:2007zr,Young:2007wd}.
It appears that the exponentiated affine derivation
$\exp(\sfrac{i}{2} g^{-1}\tilde\gen{D})$
(note that exponentiated generators naturally appear in quantum algebras,
see \secref{sec:quantum})
plays an important role in the quantisation of the algebra:
The generator induces a finite shift of $u$ by an amount
which frequently occurs in the quantum R-matrix,
e.g.\
\[
x^\pm(u)=x(u\pm \sfrac{i}{2} g^{-1})=
\exp(\pm\sfrac{i}{2} g^{-1}\tilde\gen{D})\,x(u).
\]
It would be interesting to pursue the role of the affine derivation further.
\subsection{Conventional Rational Case}
\label{sec:ratconv}
The simplest limit of the fundamental r-matrix
is obtained when the two
parameters $x_k$ approach each other at a generic point $x_0$
\[
x=x_0(1+\epsilon u).
\]
The r-matrix diverges in the limit $\epsilon\to 0$
and one obtains a rational r-matrix $\tilde r$ \eqref{eq:rrational}
\[\label{eq:genericlimit}
\tilde r
=\frac{-hh'\epsilon(x_0^2-1)}{(hx_0+ih')(h'x_0-ih)}\,r.
\]
The new coefficient functions all have the same simple singularity
at $u_1=u_2$
\[
\tilde A_{12}=\tilde B_{12}=\tilde D_{12}=\tilde E_{12}=
\frac{\gamma_2}{\gamma_1}\tilde H_{12}=
\frac{\gamma_1}{\gamma_2}\tilde K_{12}=
\frac{1}{u_1-u_2}\,,
\quad
\tilde C_{12}=\tilde F_{12}=
\tilde G_{12}=\tilde L_{12}=
0.
\]
This r-matrix is the fundamental representation of the
classical rational r-matrix for
the conventional affine $\alg{gl}(2|2)$ algebra.
Interestingly, the parameter $h$ has dropped out
completely from the r-matrix $\tilde r$
and from the associated affine bialgebra.
However this does not mean that the limit is the same for all
$h$ and for all $x_0$. In particular one can see that the prefactor
in \eqref{eq:genericlimit} is singular at certain points $x_0$,
namely $x_0=-ih'/h$, $x_0=ih/h'$ and $x_0=\pm 1$.
The first pair of points corresponds to $z_0=\infty$
and the second pair to the self-dual points $z_0=z^\ast_\pm=(ih\pm h')^2$
discussed around \eqref{eq:selfdual}.
In the following we shall discuss the limits at these points.
\subsection{Conventional Trigonometric Case}
\label{sec:trigconv}
Let us next discuss the point $z_0=\infty$.
The point $z_0=0$ is analogous according to the
discussion in \secref{sec:inversion}, and there is no
need to discuss it separately.
Similarly, we can safely restrict to one of the two corresponding points
$x=ih/h'$ and $x=-ih'/h$, cf.\ \secref{sec:conjugation}.
Here we take the limit
\[
x=\frac{ih}{h'}\lrbrk{1+\frac{\epsilon}{\tilde z}+\order{\epsilon^2}},
\qquad
z=\epsilon^{-1}\tilde z.
\]
At the same time, the parameter $\alpha$
should scale like $\alpha\sim \epsilon^{-1}$.
In this case the r-matrix remains finite
in the limit $\epsilon\to 0$.
Thus the trigonometric structure in \tabref{tab:rmatrix} applies,
and its coefficients in \tabref{tab:rcoeffs} reduce to
\[\label{eq:stdtrigres}
\begin{array}[b]{c}
\displaystyle
A_{12}=B_{12}=D_{12}=E_{12}=
\frac{\sfrac{1}{2} \tilde z_1+\sfrac{1}{2} \tilde z_2}{\tilde z_1-\tilde z_2}\,,
\quad
C_{12}=F_{12}=0,
\\[2ex]
\displaystyle
G_{12}=-L_{12}=
+\frac{1}{4}\,,
\quad
H_{12}=
\frac{\gamma_1}{\gamma_2}\,
\frac{\tilde z_1}{\tilde z_1-\tilde z_2}\,,
\quad
K_{12}=
\frac{\gamma_2}{\gamma_1}\,
\frac{\tilde z_2}{\tilde z_1-\tilde z_2}\,.
\end{array}
\]
These coefficients are precisely the coefficients of the conventional
trigonometric r-matrix for $\alg{gl}(2|2)$.
The underlying algebraic structure is thus the standard affine $\alg{gl}(2|2)$
algebra with trigonometric r-matrix.
\subsection{Twisted Rational Case}
\label{sec:rattwist}
The self-dual points $z^*_\pm$ lead to
a more elaborate limit.
According to \secref{sec:inversion} the two limits
are equivalent and we choose to investigate
\[
z_0=(ih+h')^2,\qquad
x_0=+1.
\]
The limit is defined by
\[
x=1+\epsilon\,\frac{h'-ih}{h'}\,y,
\qquad
z=z_0\lrbrk{1+\frac{ih\epsilon^2 }{h'}\,u},
\qquad
u=y^2,
\qquad
\alpha=\frac{i\epsilon}{h'-ih}\,\tilde\alpha.
\]
Here the r-matrix diverges quadratically
\[
\tilde r=\frac{ih\epsilon^2}{h'}\,r
\]
with $\tilde r$ a rational r-matrix of the form \eqref{eq:rrational}.
The coefficients of this fundamental r-matrix read
\<\label{eq:rtwistrat}
\tilde A_{12}=\tilde D_{12}\earel{=} \frac{1}{4y_1y_2}\,\frac{y_1+y_2}{y_1-y_2}\,,
\nonumber\\
\sfrac{1}{2}(\tilde A_{12}+\tilde B_{12})=
\sfrac{1}{2}(\tilde D_{12}+\tilde E_{12})
\earel{=} \frac{1}{y_1^2-y_2^2}\,,
\nonumber\\
\sfrac{1}{2}(\tilde A_{12}-\tilde B_{12})=
\sfrac{1}{2}(\tilde D_{12}-\tilde E_{12})
\earel{=}
\frac{1}{4y_1y_2}\,\frac{y_1-y_2}{y_1+y_2}\,,
\nonumber\\
\frac{\tilde\alpha}{2\gamma_1\gamma_2}\,\tilde C_{12}=
-\frac{2\gamma_1\gamma_2y_1y_2}{\tilde\alpha}\,\tilde F_{12}\earel{=}
\frac{1}{2}\,\frac{1}{y_1+y_2}\,,
\nonumber\\
\tilde G_{12}=-\tilde L_{12}\earel{=} \frac{1}{4y_1y_2}\,,
\nonumber\\
\frac{\gamma_2 y_2}{\gamma_1}\,\tilde H_{12}=
\frac{\gamma_1 y_1}{\gamma_2}\,\tilde K_{12}\earel{=}
\frac{1}{2}\,\frac{1}{y_1-y_2}\,.
\>
These coefficients along with the rational r-matrix structure in \eqref{eq:rrational}
agree with Eqs.\ (4.2, 4.10) in \cite{Klose:2007rz}
when setting $y=1/2p_-$, $\gamma=\sqrt{\tilde\alpha p_-}$.
This case therefore provides the classical s-matrix
for strings in $AdS_5\times S^5$
in the near flat space limit \cite{Maldacena:2006rv}.
In order to understand the algebra underlying this r-matrix,
we consider the coefficients \eqref{eq:abcdclass,eq:zq} for the fundamental
representation first. It turns out that $a$ and $b$ are finite
while $c$ and $d$ diverge. In combination with $a$ and $b$
one can nevertheless find finite combinations $\tilde c$ and $\tilde d$
\[\label{eq:abcdtwistrat}
a= \gamma,
\quad
b= \frac{\tilde \alpha }{2y\gamma}\,,
\quad
\tilde c=c-\frac{a}{\tilde\alpha\epsilon}=
-\frac{\gamma y}{\tilde \alpha}
\,,
\quad
\tilde d=d-\frac{b}{\tilde\alpha\epsilon}= \frac{1}{2\gamma}
\,,
\quad
\tilde q=\frac{i\epsilon}{x_0}\,q=
\frac{1}{2y}\,.
\]
In the matrix notation \eqref{eq:matrices} this corresponds to
a multiplication by a matrix $R$
\[
\tilde T=RT,\qquad
R=\matr{cc}{1&0\\\tilde\alpha^{-1}\epsilon^{-1}&1}.
\]
This implies that we should consider the following redefined generators
\[
\matr{cc}{\tilde{\gen{Q}}^{\alpha b}\\\tilde{\gen{S}}^{\alpha b}}
=R\matr{cc}{\gen{Q}^{\alpha b}\\\gen{S}^{\alpha b}},
\qquad
\tilde{\gen{A}}=\frac{i\epsilon}{x_0}\,\gen{A},
\qquad
\tilde{\gen{B}}=-\frac{ix_0}{\epsilon}\,\gen{B}.
\]
They have a well-defined algebra in the limit $\epsilon\to 0$,
cf.\ \eqref{eq:DQmatrix,eq:matrices}
with the new matrix
\[
\tilde W=\frac{x_0}{i\epsilon}RWR^{-1}= \matr{cc}{0&2\tilde\alpha\\2\tilde\alpha^{-1}u&0}.
\]
Next we consider the limit of the affine extension of the algebra.
The affine derivation must be rescaled
\[
\tilde{\gen{D}}=\frac{ih\epsilon^2}{h'}\,\gen{D}\simeq \frac{d}{du}\,.
\]
The action on the generators is defined by \eqref{eq:DQaffderrat}
with coefficients $\tilde U,\tilde V$ \eqref{eq:UVmatrix} limiting to
\[
\tilde U=
\frac{ih\epsilon^2}{h'}\,RUR^{-1}=
\frac{1}{4u}\matr{cc}{-1&0\\0&+1},
\qquad
\tilde V=
\frac{ih\epsilon^2}{h'}\,V=
-\frac{1}{2u}\,.
\]
The above algebra is in fact a twisted affine algebra:
This can be observed if we write the Lie brackets
in terms of the generators
\[\label{eq:twistgen}
\bar{\gen{Q}}=u^{+1/4}\tilde{\gen{Q}},\qquad
\bar{\gen{S}}=u^{-1/4}\tilde{\gen{S}},\qquad
\bar{\gen{A}}=u^{+1/2}\tilde{\gen{A}},\qquad
\bar{\gen{B}}=u^{-1/2}\tilde{\gen{B}}.
\]
Now the above algebra is defined by the parameters
\[\label{eq:aboveZ2}
\bar W= \matr{cc}{0&2\tilde\alpha\\2\tilde\alpha^{-1}&0},
\qquad
\bar U=\bar V=0.
\]
I.e.\ the loop levels of the generators add up simply,
and the affine extension acts canonically.
The automorphism defining the above twist has a period of 4.
It corresponds to an \emph{outer} $\mathbb{Z}_2$-automorphism of $\alg{gl}(2|2)$
which acts non-trivially on one of the two $\alg{sl}(2)$ subalgebra.
Note that for the simple superalgebra $\alg{psl}(2|2)$
the corresponding automorphism is inner \cite{Serganova:1985aa,Grantcharov:2004aa},
so the non-triviality of the twist is
only due to the central charge $\gen{A}$ and the derivation $\gen{B}$.
\subsection{Twisted Trigonometric Case}
\label{sec:trigtwist}
We have exhausted all the special values of $z$
for generic values of the global parameter $h$.
As $h$ varies also the self-dual points \eqref{eq:selfdual}
\[
z^*_\pm=\bigbrk{ih\pm h'}^2
\]
move around in the complex plane
while the special values $z=0$ and $z=\infty$ remain fixed.
For particular values of $h$, namely $h=0,\pm1,\infty$,
some of the special values coincide
giving rise to further limits of interest.
We have seen in \secref{sec:duality} that the points $h=\pm 1$
are equivalent to $h=0$, consequently there is no need to discuss them
separately.
Let us first consider the case $h=\infty$.
There both self-dual points approach the other two special values,
$z^*_+=0$, $z^*_-=\infty$.
Now we take the limit
\[
h=\epsilon^{-1},\qquad
z\sim \epsilon^0,\qquad
\epsilon\to 0.
\]
Here the parameters \eqref{eq:abcdclass,eq:zq} of the fundamental representation
\eqref{eq:loopeval} read
\[\label{eq:abcdtwisttrig}
a=\gamma,\quad
b=\frac{\alpha}{2\gamma y}\,,\quad
c=-\frac{\gamma y}{\alpha}\,,\quad
d=\frac{1}{2\gamma}\,,\quad
x=1-\frac{\epsilon}{y}\,,\quad
z=y^2,\quad
q=\frac{1}{2y}\,,
\]
and the fundamental classical r-matrix in \tabref{tab:rmatrix,tab:rcoeffs}
takes the same form using these simplified parameters $a,b,c,d,q$.
In particular we find
\<\label{eq:rtwisttrig}
A_{12}
=D_{12}
\earel{=}
\frac{1}{4}\, \frac{y_1+y_2}{y_1-y_2}\,,
\nonumber\\
\sfrac{1}{2}(A_{12}+B_{12}+1)
=\sfrac{1}{2}(D_{12}+E_{12}+1)
\earel{=} \frac{y_1^2}{y_1^2-y_2^2}\,,
\nonumber\\
\sfrac{1}{2}(A_{12}+B_{12}-1)
=\sfrac{1}{2}(D_{12}+E_{12}-1)
\earel{=} \frac{y_2^2}{y_1^2-y_2^2}\,,
\nonumber\\
\sfrac{1}{2}(A_{12}-B_{12})
=\sfrac{1}{2}(D_{12}-E_{12})
\earel{=}
-\frac{1}{4}\,\frac{y_1-y_2}{y_1+y_2}\,,
\nonumber\\
\frac{\alpha}{2\gamma_1 \gamma_2 y_1y_2}\,
C_{12}=
-\frac{2\gamma_1\gamma_2 }{\alpha}\,
F_{12}\earel{=}
-\frac{1}{2}\,\frac{1}{y_1+y_2}\,,
\nonumber\\
G_{12}
=L_{12}
\earel{=}
0,
\nonumber\\
\frac{\gamma_2}{y_1\gamma_1}\,
H_{12}=
\frac{\gamma_1}{y_2\gamma_2}\,
K_{12}\earel{=}
\frac{1}{2}\,\frac{1}{y_1-y_2}\,.
\>
These coefficients are reminiscent of those
for the twisted rational r-matrix in \eqref{eq:rtwistrat}.
In fact, the representation parameters in \eqref{eq:abcdtwisttrig}
agree precisely with those in \eqref{eq:abcdtwistrat}.
Effectively, it means that the two algebras are equivalent
(up to the affine extensions).
The parameters therefore read
\[
W= \matr{cc}{0&2\alpha\\2\alpha^{-1}z&0},
\qquad
U=\frac{1}{4}\matr{cc}{-1&0\\0&+1},
\qquad
V=-\frac{1}{2}\,.
\]
As explained in \secref{sec:rattwist},
they describe a $\mathbb{Z}_2$-twisted affine $\alg{gl}(2|2)$ algebra.
A quantum R-matrix for this algebra was derived in \cite{Gould:1996aa}.
Therefore one would expect that its classical limit
is related to the trigonometric r-matrix described above.
It is curious to observe that the coefficients
in \eqref{eq:rtwisttrig} match almost exactly
with those found for the scattering matrix derived in
(4.9--4.11) in \cite{Hoare:2009fs}
including the functional form of its prefactor
when equating $y_k=\exp(\theta_k)$.
Nevertheless, there is a crucial difference:
The s-matrix in \cite{Hoare:2009fs} is based on the rational structure \eqref{eq:rrational}
with manifest $\alg{su}(2)\oplus\alg{su}(2)$ symmetry,
while the coefficients are intimately associated
to the the trigonometric structure in \tabref{tab:rmatrix} with
broken $\alg{su}(2)\oplus\alg{su}(2)$.
Effectively $K_2$ in \cite{Hoare:2009fs}
compares to $\sfrac{1}{2} (A_{12}+B_{12})$ rather than
$\sfrac{1}{2} (A_{12}+B_{12}\pm 1)$.
\subsection{Special Trigonometric Case at \texorpdfstring{$h=\infty$}{h=inf}}
\label{sec:triginf}
The value $h=\infty$ considered above is subtle,
and the result depends on the details
of taking the limit $h\to\infty$.
Previously we have assumed that $z$ remains finite,
but there is also the option of scaling $z\to 0$ or $z\to\infty$
in correlation with $h\to\infty$.
Moreover the result generally depends on how fast $z$ converges in
comparison to $h$.
A suitable limit with $z\to 0$
(according to \secref{sec:inversion} this is equivalent to $z\to\infty$)
turns out to be
\[
h=\epsilon^{-1},\qquad
z=-\sfrac{1}{4}\epsilon^2\tilde z,\qquad
x\sim\epsilon^0,\qquad
\alpha=\sfrac{1}{2}\epsilon\tilde\alpha.
\]
This limit is distinguished from the previous one
by the fact that one of the self-dual points \eqref{eq:selfdual}
remains finite while the other approaches infinity
\[
\tilde z^*_-=1,\qquad
\tilde z^*_+\sim 16\epsilon^{-4} .
\]
The parameters $a,b,c,d$ \eqref{eq:abcdclass}
for the supercharges in the fundamental representation remain finite
\[
a=\gamma,\qquad
b=-\frac{\tilde\alpha}{2\gamma}\,\frac{x-1}{x+1}\,,\qquad
c=\frac{2\gamma}{\tilde\alpha}\,\frac{1}{x-1}\,,\qquad
d=\frac{1}{\gamma}\,\frac{x}{x+1}\,,
\]
while the parameters $q,z$ are singular
and must be renormalised
\[
\tilde q=\epsilon q=-\,\frac{x-1}{x+1}\,,\qquad
\tilde z=-4\epsilon^{-2} z=-\frac{4x}{(x-1)^2}\,.
\]
Using these parameters the fundamental
r-matrix takes the same form as in \tabref{tab:rmatrix}.
For the algebra we have to rescale some generators
\[
\tilde{\gen{A}}=\epsilon\gen{A},\qquad
\tilde{\gen{B}}=\epsilon^{-1}\gen{B}.
\]
The parameters $U,V,W$ in \eqref{eq:matrices,eq:UVmatrix}
for the Lie brackets \eqref{eq:DQmatrix,eq:DQaffder,eq:DQcentral}
read in this case
\[
\tilde W=
\epsilon^{-1}W=
\matr{cc}{-1&+\tilde\alpha\\-\tilde\alpha^{-1}\tilde z&+1},
\quad
U=
\frac{1}{4}\,\frac{\tilde z}{\tilde z-1}
\matr{cc}{-1&0\\-2\tilde\alpha^{-1}&+1},
\quad
V=-\frac{1}{2}\,\frac{\tilde z}{\tilde z-1}\,.
\]
\subsection{Special Trigonometric Case at \texorpdfstring{$h=0$}{h=0}}
\label{sec:trig0}
The limit $h\to 0$ was discussed already in \secref{sec:rational};
it yields the full rational r-matrix \cite{Klose:2006zd,Torrielli:2007mc,Beisert:2007ty}.
In this limit it was assumed that $x$ remains finite whereas $z\to 1$.
Likewise one can demand that $z$ remains finite and arbitrary
while $x\to 0$ or $x\to\infty$;
this turns out to yield an inequivalent limit.
Let us consider the case of large $x$
\[
h=\epsilon,\qquad
z\sim \epsilon^0,\qquad
x=-i\epsilon^{-1}\frac{z-1}{z}+\order{\epsilon^0}.
\]
Then the parameters of the fundamental representation \eqref{eq:abcdclass,eq:zq} read
\[
a=\gamma,\qquad
b=c=0,\qquad
d=\frac{1}{\gamma}\,,\qquad
\tilde q=\epsilon^{-1}q=\frac{1}{z-1}\,.
\]
This is almost the fundamental representation of the
standard $\alg{gl}(2|2)$, but the central charge $\tilde q$ behaves differently.
Consequently, the r-matrix coefficients in \tabref{tab:rcoeffs}
take a slightly non-standard form.
The case of $x\to 0$ leads to the conjugate fundamental representation.
Next, let us consider the algebra.
In this case, we should rescale the generators $\gen{A}$ and $\gen{B}$
according to
\[
\tilde{\gen{B}}=\epsilon\gen{B},\qquad
\tilde{\gen{A}}=\epsilon^{-1}\gen{A}
\]
in order to make their action finite.
The algebra now takes the standard form
with the parameters \eqref{eq:matrices,eq:UVmatrix}
\[
\tilde W=
\epsilon W= (z-1)\matr{cc}{+1&0\\0&-1},
\qquad
U=0,
\qquad
V=-\frac{z}{z-1}\,.
\]
All the off-diagonal elements of the matrices are absent
as in the conventional affine $\alg{gl}(2|2)$.
Only the central charge $\gen{A}$ appears with a non-trivial
dependence on the loop variable $z$.
In fact, we can formally make all the algebra relations
like those for affine $\alg{gl}(2|2)$ by redefining
the loop levels of $\tilde{\gen{A}}$ and $\tilde{\gen{B}}$
\[
\bar{\gen{B}}
=
(z-1)^{-1}\tilde{\gen{B}},
\qquad
\bar{\gen{A}}
=
(z-1)\tilde{\gen{A}}.
\]
This leads to the standard affine algebra
with parameters
\[
\bar W= \matr{cc}{+1&0\\0&-1},
\qquad
\bar U=\bar V=0.
\]
The simplification is however at the cost of changing the
universal r-matrix in \eqref{eq:reval}
because the transformation does not respect
the decomposition \eqref{eq:decompose}.
\subsection{Special Rational Case}
\label{sec:ratdef}
There is even a combination of the two different limits at $h\to 0$.
Here $h$ should approach $0$ faster than $z$ approaches $1$.
For example we can define the limit
\[
h=\epsilon^2,\qquad
z=1+i\epsilon u+\order{\epsilon^2},\qquad
x=\frac{u}{\epsilon}+\order{\epsilon^0}.
\]
The parameters of the fundamental representation reduce to
\[
a=\gamma,\qquad
b=c=0,\qquad
d=\frac{1}{\gamma}\,,\qquad
\tilde q=\frac{1}{i\epsilon}\,q=-\frac{1}{u}\,.
\]
The r-matrix diverges and becomes of rational type
\eqref{eq:rrational}
\[\tilde r=i\epsilon r.
\]
The coefficients are almost those of the conventional rational
r-matrix, but there are a few important modifications
\[\begin{array}{c}
\displaystyle
\sfrac{1}{2}(\tilde A_{12}+\tilde B_{12})
=\sfrac{1}{2}(\tilde D_{12}+\tilde E_{12})
=\frac{\gamma_2}{\gamma_1}\,\tilde H_{12}
=\frac{\gamma_1}{\gamma_2}\,\tilde K_{12}
=\frac{1}{u_1-u_2}\,,
\qquad
\tilde C_{12}=\tilde F_{12}= 0,
\\[2ex]\displaystyle
\sfrac{1}{2}(\tilde A_{12}-\tilde B_{12})=\sfrac{1}{2}(\tilde D_{12}-\tilde E_{12})
=\frac{u_1-u_2}{4u_1u_2}\,,
\qquad
\tilde G_{12}=-\tilde L_{12}= \frac{u_1+u_2}{4u_1u_2}\,.
\end{array}
\]
The algebra is specified by the following parameters
\[
\tilde W=i\epsilon W=
u\matr{cc}{-1&0\\0&+1},
\qquad
\tilde U=i\epsilon U=0,
\qquad
\tilde V=i\epsilon V=-\frac{1}{u}\,.
\]
This case may be viewed as the rational analog of the
special trigonometric case at $h=0$ in \secref{sec:trig0}.
\begin{figure}\centering%
\includegraphics[width=\linewidth]{FigLimits.mps}%
\caption{Analytic structure and limits of various r-matrices.
T$(h)$: full trigonometric (\protect\secref{sec:classical});
T$(0)$: special trigonometric at $h=0$ (\protect\secref{sec:trig0});
T$(\infty)$: special trigonometric at $h=\infty$ (\protect\secref{sec:triginf});
T(twist): twisted trigonometric (\protect\secref{sec:trigtwist});
T(conv): conventional trigonometric (\protect\secref{sec:trigconv});
R(full): full rational (\protect\secref{sec:rational});
R(twist): twisted rational (\protect\secref{sec:rattwist});
R(def): special rational (\protect\secref{sec:ratdef});
R(conv): conventional rational (\protect\secref{sec:ratconv}).
Special points $z^\circ_\pm=0,\infty$ and $z^*_\pm$ are marked by
$\circ$ and $\ast$, respectively.
A circle is drawn around coincident special points.
Two cases are connected by an arrow if the second is a particular
limit of the first.}
\label{fig:limits}
\end{figure}
\subsection{Summary}
In this section we have found more than a handful special limits
of the $r$-matrix. What makes these limits special and
how can we be sure that we have not missed an interesting case?
To answer the question we should consider special points in the $z$-plane.
The affine algebra specialises the two points $z^\circ_\pm=0,\infty$.
Furthermore there are two points $z^*_\pm$ which
lead to certain self-duality properties of representations,
see \eqref{eq:selfdual}.
In total there are four special points
\[
z^\circ_\pm=0,\infty,\qquad
z^*_\pm=(ih\pm h')^2.
\]
Above we have constructed limits by zooming into the neighbourhood of
certain points while potentially taking a simultaneous limit for $h$.
There is however a different point of view which makes the various limits
more transparent: By zooming into the neighbourhood of one point we
effectively shift all other special points to the point at infinity.
Hence the various limits correspond to grouping
the special points in different ways.
What is the role of the parameter $h$ in the limits?
Zooming into a neighbourhood can be achieved by M\"obius
transformations of the $z$-plane with coefficients
depending on the limiting procedure.
The transformation maps the special points to different positions,
but there exist one conformal cross-ratio which remains invariant.
Its value $s=(ih+h')^4$ is a function of $h$.
Alternatively one can consider $h=h(s)$ to be a function of the cross-ratio $s$.
This allows us to view the four special points as independent,
and $h=h(z^\circ_\pm,z^\ast_\pm)$ as a function of their distribution
modulo M\"obius transformations.
To understand the various limits, we should group the four special points
in all possible ways. Up to trivial permutations there are nine choices
corresponding to the full trigonometric case with parameter $h$ and
its eight limiting cases considered above,
see \figref{fig:limits}.
Note that the trigonometric cases have two distinct points
$z^\circ_\pm$ while the rational cases have identical points
$z^\circ_+=z^\circ_-$.
Two cases are linked by a limiting procedure
if the special points of the first can be combined
to the special points of the second.
\section{Conclusions and Outlook}
\label{sec:concl}
Classical r-matrices for Lie algebras were classified
in \cite{Belavin:1982aa}.
Three main classes,
distinguished by the distribution of poles in the complex plane,
were identified: rational, trigonometric and elliptic.
The classification is analogous for simple Lie superalgebras \cite{Leites:1984aa}.
In the case of the (non-simple) Lie superalgebra $\alg{gl}(2|2)$
an exceptional r-matrix was identified in \cite{Beisert:2007ty}.
This r-matrix is of rational type, but it is not of difference form.
Its quantisation leads to Shastry's R-matrix for the Hubbard model \cite{Shastry:1986bb}
or equivalently \cite{Beisert:2006qh}
to the S-matrix for the AdS/CFT integrable system \cite{Beisert:2005tm}.
Hence this r-matrix is responsible for the exceptional integrable structure
in these models at the classical level.
\smallskip
In this paper we have developed and investigated
the trigonometric generalisation of the exceptional r-matrix for $\alg{gl}(2|2)$.
The corresponding fundamental quantum R-matrix was derived in \cite{Beisert:2008tw},
and it defines the integrable structure of the Alcaraz--Bariev model \cite{Alcaraz:1999aa}
(type B).
As for the rational case, the underlying Lie algebra is a deformation
of the loop algebra $\alg{gl}(2|2)[z,z^{-1}]$.
The deformation is special in the sense that
the Lie brackets are not homogeneous in the level of the loop algebra.
Nevertheless, the algebra admits solutions to the classical Yang--Baxter equation.
Analogously, it admits a decomposition into positive and negative subalgebras.
An interesting feature of the algebra is
that it has one modulus $h$
whose value has significant impact on the algebra.
One may wonder whether there are other similar cases
of deformed loop algebras or if $\alg{gl}(2|2)$ is truly
exceptional in this regard.
In other words, which is the precise (co)homological
property of $\alg{gl}(2|2)$ or its loop algebra
giving rise to the deformation?
\smallskip
The deformed loop algebra also admits the extension
by a derivation and a central charge to an affine Lie algebra.
This algebra is not of Kac--Moody type,
but its structure is similar in many respects.
The affine derivation serves as a scaling of the loop variable $z$
(or a shift in $u$ in the rational case).
In a physical scattering context, it can be viewed as
a boost operator in analogy to Lorentz boosts in two spacetime dimensions.
Also we must extend the notion of particles to fields,
because the particle momentum does not commute with boosts.
Interestingly, the boost has non-trivial cobrackets,
hence the symmetry should be viewed as deformed or non-commutative \cite{Gomez:2007zr,Young:2007wd}.
Non-invariance of the r-matrix also explains the violation
of difference form for the r-matrix.
Finally, extension of a symmetry often leads to additional restrictions.
Here it would be interesting to see if, e.g.,
the overall prefactor of the r-matrix can be constrained
by the affine extension.
\smallskip
Subsequently, we have investigated discrete transformations
and special points of the r-matrix.
Transformations include conjugation,
inversion of the loop variable, a flip of statistics and
a duality for the global parameter.
Conjugation maps different representations into each other.
In particular, the family of fundamental representations is self-conjugate,
and thus conjugation extends to a crossing symmetry of the r-matrix,
cf.\ \cite{Janik:2006dc}.
Inversion symmetry of the r-matrix can be viewed
as a scattering unitarity condition.
The statistics flip interchanges bosons and fermions in the fundamental representation.
At the level of the algebra it permutes the two $\alg{sl}(2)$ subalgebras.
Last but not least, the duality map relates algebras/r-matrices
with different moduli $h$.
An important insight gained from the discrete transformations
is that next to the special points $z=z^\circ_\pm=0,\infty$,
which exist for any trigonometric r-matrix,
there are two self-dual points $z=z^*_\pm$ whose value depends on $h$.
\smallskip
Finally, several r-matrices with simpler structures
were recovered as limiting cases.
For example, our trigonometric r-matrix reduces
to the exceptional rational r-matrix of \cite{Klose:2006zd,Torrielli:2007mc,Beisert:2007ty}
in a particular limit. The latter can be reduced further to the conventional
rational $\alg{gl}(2|2)$ r-matrix as well as to two other intermediate cases.
In total there is the one-parameter family
of exceptional trigonometric r-matrices and 8 singular cases,
see \figref{fig:limits}.
The trigonometric family has the most sophisticated structure
while the conventional rational r-matrix is the plainest:
All intermediate cases can be obtained from the former and be reduced to the latter.
These include some special cases with $\alg{gl}(2|2)$ structure
discovered earlier in various contexts:
They can be of trigonometric or of rational type,
they are conventional or deformed and
untwisted or $\mathbb{Z}_2$-twisted.
In terms of algebra all cases follow from the one discussed in this paper:
Its structure can be simplified through limits and algebraic contractions
down to the plain $\alg{gl}(2|2)$ affine Kac--Moody algebra.
It would be interesting to find out whether
the trigonometric structure is itself a limiting case
of some exceptional elliptic r-matrix
(note that both $\alg{psl}(2|2)$ and $\alg{osp}(4|2)$ admit
elliptic r-matrices \cite{Leites:1984aa}).
\medskip
With a good part of the classical framework established,
several open questions concerning the
exceptional trigonometric r-matrix remain.
For instance, we would like to promote the Lie bialgebra
to a quantum affine Hopf algebra (cf.\ \cite{Etingof:1995aa}).
Are there any obstacles
due to the non-standard structure of the affine algebra?
So far only the fundamental quantum R-matrix has been established.
However there is little doubt that R-matrices for
higher representations can indeed be constructed
as in the rational case \cite{Beisert:2006qh,Chen:2006gp,Arutyunov:2008zt,deLeeuw:2008dp,deLeeuw:2008ye,Arutyunov:2009mi,Arutyunov:2009ce,Arutyunov:2009iq,Arutyunov:2009pw}.
This would be very suggestive of a universal R-matrix.%
\footnote{Doubts raised in \protect\cite{Arutyunov:2009pw}
apply only to a different type of quantum algebra without derivations $\gen{B}$
and with a minimal set of Serre relations.}
\smallskip
Developing the quantum affine algebra would
establish, as a by-product, the Yangian for the undeformed
Hubbard model or for integrable scattering in AdS/CFT.
One complication in the formulation might reside in the existence
of the tower of derivations $z^n\gen{B}$
for which Drinfeld's first presentation \cite{Drinfeld:1985rx,Drinfeld:1986in}
using Chevalley--Serre generators is not ideally suited.
Instead, Drinfeld's second realisation \cite{Drinfeld:1988aa}
along the lines of \cite{Spill:2008tp} may prove to be more helpful.
\begin{figure}\centering
\includegraphics{FigDynkinOX-O.mps}
\caption{Dynkin diagram for $\alg{d}(2,1;0)$.}
\label{fig:Dynkin2}
\end{figure}
\begin{figure}\centering
\includegraphics{FigDynkinOXO.mps}
\qquad
\includegraphics{FigDynkinXXX.mps}
\qquad
\includegraphics{FigDynkinXOX.mps}
\caption{All Dynkin diagrams for $\alg{sl}(2|2)$.}
\label{fig:Dynkin3}
\end{figure}
Also the choice of Dynkin diagram may play a role:
For instance, the Bethe equations \cite{Lieb:1968aa,Beisert:2005fw}
cannot be formulated (easily)
for the distinguished diagram in \figref{fig:Dynkin}
(leftmost in \figref{fig:Dynkin3}),
but it appears to prefer a structure
reminiscent of the exceptional superalgebra
$\alg{d}(2,1;\alpha)$ with singular parameter $\alpha=0$ in \figref{fig:Dynkin2}.
The latter has a non-symmetrisable Cartan matrix, cf.\ \cite{Hoyt:2007aa}.
It would be interesting to derive the r-matrices
for the various other Dynkin diagrams (see \figref{fig:Dynkin3}),
and to understand how to transform between them,
see also \cite{Khoroshkin:1994aa,Geer:2005aa}.
\paragraph{Acknowledgements.}
The author thanks
B.\ Hoare,
T.\ McLoughlin,
V.\ Schomerus,
V.\ Serganova,
M.\ Staudacher and
A.\ Tseytlin
for interesting discussions.
Useful comments on the manuscript by referees are acknowledged.
The author acknowledges hospitality
by the Galileo Galilei Institute and Durham University
during the workshops ``Non-Perturbative Methods in Strongly Coupled Gauge Theories'' (GGI),
``New Perspectives in String Theory'' (GGI) and ``Gauge and String Amplitudes'' (Durham)
where part of the present work was performed.
|
\section*{Abstract}
In this paper, we give a positive answer to the open question:
Can there exist $4$ limit cycles in quadratic near-integrable polynomial
systems?
It is shown that when a quadratic integrable system has two centers
and is perturbed by quadratic polynomials,
it can generate at least $4$ limit cycles
with $(3,1)$ distribution. The method of Melnikov function is used.
\vspace{0.1in}
\noindent
{\it Keywords}: Hilbert's 16th problem, quadratic near-integrable system,
limit cycle,
\hspace{0.50in}reversible system, Hopf bifurcation, Poincar\'{e} bifurcation,
Melnikov function
\vspace{0.10in}
\noindent
{\it MSC}: 34C07; 34C23
\noindent
\rule{6.5in}{0.012in}
\section{Introduction}\label{intro}
\setcounter{equation}{0}
\renewcommand{\theequation}{1.\arabic{equation}}
The well-known Hilbert's 16th problem is remained unsolved since
Hilbert~\cite{hilbert1900} proposed the 23 mathematical problems at the
Second International Congress of Mathematics in 1990.
Recently, a modern version of
the second part of the 16th problem was formulated by Smale~\cite{smale1998},
chosen as one of the 18 challenging mathematical problems for the 21st
century. To be more specific, consider the following planar system:
\begin{eqnarray}
\frac{dx}{dt} = P_n(x,y), \quad \frac{dy}{dt} = Q_n(x,y),
\label{b1}
\end{eqnarray}
where $\, P_n(x,y) \,$ and $\, Q_n(x,y) \,$ represent $\, n^{\rm th}$-degree
polynomials of $\, x\,$ and $\, y$. The second part of Hilbert's 16th
problem is to find the upper bound $\, H(n) \leq n^q\,$ on the number
of limit cycles that the system can have, where $\, q\,$
is a universal constant, and $H(n)$ is called Hilbert number.
In early 90's of the last century,
Ilyashenko~\cite{Ilyashenko1991}
and \'{E}calle~\cite{Ecalle1992} proved the finiteness theorem pioneered by
Dulac, for given planar polynomial vector fields.
In general the finiteness problem has not been solved even for
quadratic systems.
A recent survey article~\cite{li2003} (and more references therein)
has comprehensively discussed this problem and reported the recent progress.
If the problem is restricted to
the neighborhood of isolated fixed points, then the
question is reduced to studying degenerate Hopf bifurcations, which
give rise to fine focus points.
In the past six decades, many researchers have considered the local
problem and obtained many results
(e.g., see~[6--12]).
In the last 20 years, much progress
on finite cyclicity near a fine focus point or a homoclinic loop has been
achieved. Roughly speaking, the so-called finite cyclicity means that
at most a finite number of limit cycles can exist in some neighborhood
of focus points or homoclinic loop under small perturbations on the
system's parameters.
\pagestyle{myheadings}
\markright{{\footnotesize {\it P. Yu} \& {\it M. Han
\hspace{1.2in} {\it
4 limit cycles in near-quadratic systems}}}}
In this paper, we particularly consider bifurcation of limit cycles
in quadratic systems. Early results can be found in a survey article
by Ye~\cite{Ye1982}. Some recent progress has been reported in
a number of papers (e.g., see~\cite{Roussarie1998,RoussarieSchlomiuk2002}).
For general quadratic system (\ref{b1}) ($n=2$),
in 1952, Bautin~\cite{bautin} proved that there exist $3$ small limit cycles
around a fine focus point or a center. After $30$ years,
until the end of 1970's,
concrete examples were given to show that general quadratic systems can have
$4$ limit cycles~\cite{ChenWang,Shi80}, around two foci
with $(3,1)$ configuration.
Since then, many researchers have paid attention to integrable
quadratic systems, and a number of results have been obtained.
A question was naturally raised:
Can near-integrable quadratic systems have $4$ limit cycles?
A quadratic system is called near-integrable
if it is a perturbation of a quadratic integrable system by quadratic
polynomials. On one hand, it is reasonable to believe that the answer should
be positive since general quadratic systems have at least $4$ limit cycles;
while on the other hand, near-integrable quadratic systems have limitations on
their system parameters and thus it is more difficulty to find $4$ limit cycles
in such systems. In fact, this is still an open problem after
another $30$ years since the finding of $4$ limit cycles in general
quadratic systems.
The study of bifurcation of limit cycles for near-integrable systems is
related to the so called weak Hilbert's 16th problem~\cite{Arnold1977},
which is transformed to finding the maximal number of
isolated zeros of the Abelian integral or Melnikov function:
\begin{equation}
M(h,\delta) = \displaystyle\oint_{H(x,y)=h} Q_n \, dx - P_n \, dy,
\label{b2}
\end{equation}
where $\, H(x,y), \, P_n \,$ and $\, Q_n \,$ are all real polynomials
of $\,x \,$ and $\, y\,$ with $\, {\rm deg} H = n+1$,
and $\, \max\{ {\rm deg} P_n, \, {\rm deg} Q_n \} \le n$.
The weak Hilbert's 16th problem is a very important problem, closely
related to the maximal number of limit cycles of the following
near-Hamiltonian system~\cite{Han2006}:
\begin{equation}
\displaystyle\frac{dx}{dt} = \displaystyle\frac{\partial H(x,y)}{\partial y}
+ \varepsilon \, p_n(x,\,y), \quad
\displaystyle\frac{dy}{dt} =- \, \displaystyle\frac{\partial H(x,y)}{\partial x}
+ \varepsilon \, q_n(x,\,y),
\label{b3}
\end{equation}
where $\, H(x,y)$, $ p_n (x,y) \,$ and $\, q_n(x,y) \,$ are
polynomials of $\, x \,$ and $\, y$, and $\, 0 < \varepsilon \ll 1 \,$
is a small perturbation.
General quadratic systems with one center have been classified by
\.{Z}ol\c{a}dek~\cite{Zoladek1994} using a complex analysis on the
condition of the center, as four systems:
$Q_3^{LV}$ -- the Lotka-Volterra system;
$Q_3^H$ -- Hamiltonian system; $Q_3^R$ -- reversible system; and
$Q_4$ -- codimension-4 system.
In 1994,
Horozov and Iliev~\cite{HI1994} proved that in
quadratic perturbation of
generic quadratic Hamiltonian vector fields with one center and
three saddle points there can appear at most two limit cycles, and this
bound is exact. Later, Gavrilov~\cite{Gavrilov2001}
extended Horozov and Iliev's method to give a fairly complete analysis on
quadratic Hamiltonian systems with quadratic perturbations.
Quadratic Hamiltonian systems, with at most four singularities,
can be classified as three
cases~\cite{Gavrilov2001}: (i) one center and three saddle points;
(ii) one center and one saddle point; and (iii) two centers and
two saddle points.
In~\cite{Gavrilov2001}, Gavrilov showed that like case (i), cases (ii)
and (iii) can also have at most two limit cycles. Therefore,
generic quadratic Hamiltonian systems with quadratic perturbations
can have maximal two limit cycles, and this case has been completely solved.
For the $Q_3^R$ reversible system, there have been many results published.
For example, Dumortier {\it el al.}~\cite{DumortierLiZhang1997}
studied a case of $Q_3^R$ system
with two centers and two unbounded heteroclinic loops,
and presented a complete analysis of quadratic $3$-parameter unfolding.
It was proved that $3$ is the maximal number of limit cycles surrounding a
single focus, and only the $(1,1)$-configuration can occur in case of
simultaneous nests of limit cycles. That is, $3$ is the maximal number
of limit cycles for the system they studied~\cite{DumortierLiZhang1997}.
Later, Peng~\cite{Peng2002} considered a similar case with a homoclinic
loop and showed that $2$ is the maximal number of limit cycles which
can bifurcate from the system.
Around the same time, Yu and Li~\cite{YuLi2002} investigated a similar
case as Peng considered but with a varied parameter in a
certain interval, and obtained the same conclusion as Peng's. Later,
Iliev~{\it et al.}~\cite{IlievLiYu2005} re-investigated the same case
but for the varied parameter in a different interval (which yields two centers)
and got the same conclusion as that of~\cite{DumortierLiZhang1997}, i.e.,
$3$ is the maximal number of limit cycles which can be obtained from
this case.
Recently, Li and Llibre~\cite{LL2006} considered
a different case of $Q_3^R$ system which can exhibit
the configurations of limit cycles: $(0,0)$, $(1,0)$, $(1,1)$ and
$(1,2)$. Again, no $4$ limit cycles were found.
In order to explain why the above authors did not find $4$ limit cycles
from the $Q_3^R$ reversible system,
consider the $Q_3^R$ system with quadratic perturbations,
which can be described by~\cite{DumortierLiZhang1997}
\begin{equation}
\begin{array}{ll}
\dot{x} = -\, y + a\, x^2 + b\, y^2
+ \varepsilon \, (\mu_1 \, x + \mu_2 \, x\,y), \\[1.0ex]
\dot{y} = x\, (1 + c\, y) + \varepsilon\, \mu_3 \, x^2,
\end{array}
\label{b4}
\end{equation}
where $\, a, \, b, \, c \,$ are real parameters, $\, \mu_i, \ i=1,\,2,\,3\,$
are real perturbation parameters, and $\, 0 < \varepsilon \ll 1$.
When $\, \varepsilon = 0$, system (\ref{b4})$_{\varepsilon=0}$ is a reversible
integrable system. It has been noted that in all
the cases considered
in~\cite{DumortierLiZhang1997,Peng2002,YuLi2002,IlievLiYu2005},
the parameters
$\,a \,$ and $\, c \,$ were chosen as $\, a = -\,3, \ c = -\,2$, but with
$\, b=1 \,$ in~\cite{DumortierLiZhang1997};
$\, b=-\,1\,$ in~\cite{Peng2002},
$\, b \in (-\,\infty, -1) \cup (-1,0)\,$ in~\cite{YuLi2002},
and $\, b \in (0,\,2)\,$ in~\cite{IlievLiYu2005}.
In these papers, complete analysis on the perturbation parameters
was carried out with the
aid of Poincar\'{e} transformation and the Picard-Fuchs equation,
but it needed to fix all (or most of) the parameters $\,a, \, b \,$ and $\,c$.
This way it may miss opportunity
to find more limit cycles, such as possible existence of $4$ limit cycles.
As a matter of fact, for the cases considered
in~\cite{YuLi2002,IlievLiYu2005}, a simple scaling on the parameter $b$
($b \ne 0$) can be used to eliminate $b$.
So, suppose the non-perturbed system (\ref{b4})$_{\varepsilon=0}$ has two
free parameters and let us consider the 2-dimensional parameter plane.
Then, all the cases studied in the above mentioned articles are special cases,
represented by just a point or a line segment in the 2-dimensional
parameter plane (see more details in Section 2).
It has been noted that a different method was used in~\cite{LL2006} with
Melnikov function up to second order, but no more
limit cycles were found.
It should be mentioned that Zhang~\cite{PGZhang2002} has proved
that the possible cycle distributions in
general quadratic systems with two foci must be $(0,1)$-distribution
or $(1,i)$-distribution, $\, i=0,\,1,\,2,\,3, \cdots$.
So far, no results have been obtained for $\, i \ge 4$.
This result also rules out the possibility of
$\,(2,2) $-distribution.
It is conjectured that at most $3$ limit cycles
can exist around one focus point.
The problem of bifurcation of $3$ limit cycles near
an isolated homoclinic loop is still open.
In this paper, we turn to a different angle to consider bifurcation of
limit cycles in quadratic near-integrable systems with two centers.
We shall leave more free parameters in the integrable systems, so that
we will have more chances to find more limit cycles.
The basic idea is as follows: we first consider bifurcation of
multiple limit cycles from Hopf singularity, which does not need to
fix any parameters, and use expansion of Melnikov function near
centers to get such limit cycles as many as possible.
This leads to determination of a maximal number of parameters. Then,
for the remaining undetermined parameters, we compute the global
Melnikov function to look for possible large limit cycles.
Indeed, although, due to the complex
integrating factor in the analysis, we are not able to give a complete
analysis for classifying the perturbation unfolding, we do get a positive
answer to the open question of existence of $4$ limit cycles in
quadratic near-integrable systems.
In particular, we will show that perturbing a
reversible, integrable quadratic system
with two centers can have at least $4$ limit cycles, with $(3,1)$
distribution, bifurcating from
the two centers under quadratic perturbations.
The rest of paper is organized as follows.
In Section 2, we give a different classification in real domain
for quadratic systems with one center, and compare it with that
given by \.{Z}ol\c{a}dek~\cite{Zoladek1994}.
Also, we use our classification to present a simple summary on
some of the existing results for the reversible near-integrable system.
Section 3 is devoted to the analysis on bifurcation of small limit cycles
from Hopf singularity. In Section 4, we show how to find large
limit cycles bifurcating from closed orbits to obtain a total of
$4$ limit cycles. Finally, conclusion is drawn in Section 5.
\section{Classification of generic quadratic systems with at least one center}
\setcounter{equation}{0}
\renewcommand{\theequation}{2.\arabic{equation}}
In this section, we give a different classification in real domain
for quadratic systems with a center,
which is consistent with the Hamiltonian systems considered
in~\cite{HI1994,Gavrilov2001}.
We start from the following general quadratic system:
\vspace{0.0in}
\begin{equation}
\begin{array}{ll}
\displaystyle\frac{d z_1}{dt} = c_{100} + c_{110}\, z_1 + c_{101}\, z_2
+ c_{120}\, z_1^2 + c_{111}\, z_1\, z_2 + c_{102}\, z_2^2, \\ [1.5ex]
\displaystyle\frac{d z_2}{dt} = c_{200} + c_{210}\, z_1 + c_{201}\, z_2
+ c_{220}\, z_1^2 + c_{211}\, z_1\, z_2 + c_{202}\, z_2^2,
\end{array}
\label{b5}
\end{equation}
where $\, c_{ijk}$'s are real constant parameters.
It is easy to show that this system has at most four singularities,
or more precisely, it can have $0$, $2$ or $4$ singularities in real domain.
In order for system (\ref{b5}) to have limit cycles, the system must
have some singularity.
In this paper, we assume that system (\ref{b5}) has at least two singularities.
Without loss of
generality, we may assume that one singular point is located at
the origin $(0,0)$, which implies $\, c_{100} = c_{200} = 0$,
and the other at $(p,q)$ ($p^2 + q^2 \ne 0$).
Further assume the origin is a linear center.
Then introducing a series of linear transformations, parameter
rescaling and time rescaling to system (\ref{b5}) yields
the following general quadratic system:
\begin{equation}
\begin{array}{ll}
\displaystyle\frac{dx}{dt} = y + a_1\, x\, y + a_2\, y^2, \\ [1.5ex]
\displaystyle\frac{dy}{dt} = -\, x
+ x^2 + a_3\, x\, y + a_4\, y^2,
\end{array}
\label{b6}
\end{equation}
which has a linear center at the origin $(0,0)$ and another singularity
at $ (1,0)$.
In order to have the origin of system (\ref{b6}) being a center, we
may calculate the focus values of system (\ref{b6}) and find
four cases under which $(0,0)$ is a center, listed in the
following theorem (here we use \.{Z}ol\c{a}dek's
notation in our classification).
\vspace{0.10in}
\noindent
{\bf Theorem 1.1} {\it The origin of (\ref{b6}) is a center if and only
if one of the following conditions is satisfied:
\vspace{0.10in}
\noindent
\underline{$Q_3^R$\,--\,Reversible system}: $\, a_3 = a_2 = 0$, under which
system (\ref{b6}) becomes
\begin{equation}
\begin{array}{ll}
\displaystyle\frac{d x}{dt} = y + a_1\, x\, y, \\[1.0ex]
\displaystyle\frac{d y}{dt} = -\, x + x^2 + a_4 \, y^2 ,
\end{array}
\label{b7}
\end{equation}
with
$$
(1,0) \ \ {being \ a} \ \
\left\{
\begin{array}{ll}
{center} \quad & {if} \ \ a_1 < -\,1, \\ [1.0ex]
{saddle \ point} \quad & {if} \ \ a_1 > -\,1.
\end{array}
\right.
$$
\vspace{0.10in}
\noindent
\underline{$Q_3^H$\,--\,Hamiltonian system}: $\, a_3 = a_1 + 2\, a_4 = 0$,
under which system (\ref{b6}) is reduced to
\begin{equation}
\begin{array}{ll}
\displaystyle\frac{d x}{dt} = y + a_1 \, x\, y + a_2 \, y^2 , \\[1.0ex]
\displaystyle\frac{d y}{dt} = -\,x + x^2 - \displaystyle\frac{1}{2}\, a_1\, y^2 ,
\end{array}
\label{b8}
\end{equation}
with
$$
(1,0) \ \ {being \ a} \ \
\left\{
\begin{array}{ll}
{center} \quad & {if} \ \ a_1 < -\,1, \\ [1.0ex]
{saddle \ point} \quad & {if} \ \ a_1 > -\,1.
\end{array}
\right.
$$
\vspace{0.10in}
\noindent
\underline{$Q_3^{LV}$\,--\,Lokta-Volterra system}: $\, a_2 =
1+a_4 = 0$, under which system (\ref{b6}) becomes
\begin{equation}
\begin{array}{ll}
\displaystyle\frac{d x}{dt} = y + a_1 \, x\, y , \\[1.0ex]
\displaystyle\frac{d y}{dt} = -\,x + x^2 + a_3 \, x\, y - y^2 ,
\end{array}
\label{b9}
\end{equation}
with
$$
(1,0) \ \ {being \ a} \ \
\left\{
\begin{array}{ll}
{focus} \quad & {if} \ \ a_1 < -\,( 1 + \frac{1}{4}\, a_3^2 \,),
\\ [1.0ex]
{node} \quad & {if} \ \ -\,( 1 + \frac{1}{4}\, a_3^2 \,) < a_1 < -\,1,
\\ [1.0ex]
{saddle \ point} \quad & {if} \ \ a_1 > -\,1.
\end{array}
\right.
$$
\vspace{0.10in}
\noindent
\underline{$Q_4$\,--\,Codimension-4 system}:
\begin{equation}
a_3 - 5\, a_2 = a_1 - (5 + 3\, a_4) = a_4 + 2 (1+a_2^2) = 0,
\label{b10}
\end{equation}
under which system (\ref{b6}) can be rewritten as
\begin{equation}
\begin{array}{ll}
\displaystyle\frac{d x}{dt} = y - (1 + 6\, a_2^2) \, x\, y + a_2 \, y^2 , \\[1.0ex]
\displaystyle\frac{d y}{dt} = -\,x + x^2 + 5\, a_2 \, x\, y - 2\, (1 + a_2^2)\, y^2 ,
\end{array}
\label{b11}
\end{equation}
with $\, (1,0)\,$ being a node for $\, a_2 \ne 0$.
}
\vspace{0.10in}
\noindent
{\bf Remark 1.2.}\
There is one more case found from the above process,
defined by the following conditions:
\begin{equation}
a_3 - 5\, a_2 = a_1 - (5 + 3\, a_4) =
3\,(a_4+2)\,(a_4+1)^2 - (5\,a_4+6)\, a_2^2 = 0.
\label{b12}
\end{equation}
We will show later in this section, when we compare our above real
classification with the complex classification given by
\.{Z}ol\c{a}dek~\cite{Zoladek1994}, that the case defined by
(\ref{b12}) actually belongs to the $\, Q_3^R$-reversible system.
\vspace{0.10in}
\noindent
{\bf Proof.} Necessity is easy to be verified by computing the focus values
of system (\ref{b6}) associated with the origin.
Some focus values will not equal zero if the condition is not satisfied.
For sufficiency, we find an integrating factor for each case when the
condition holds. For the $Q_3^H$\,-\,Hamiltonian system (\ref{b8}),
we know that the
integrating factor is $\,1$, and the Hamiltonian is given by
\begin{equation}
H(x,y) = \displaystyle\frac{1}{2}\, (x^2 + y^2) - \displaystyle\frac{1}{3}\, x^3
+ \displaystyle\frac{1}{2}\, a_1 \, x\, y^2 + \displaystyle\frac{1}{3}\, a_2 \, y^3,
\label{b13}
\end{equation}
which is exactly the same as that given in~\cite{HI1994,Gavrilov2001}.
For the $Q_3^R$\,-\,reversible system (\ref{b7}), the integrating factor is
\begin{equation}
\gamma = |1+a_1 x|^{- \frac{a_1+2 a_4}{a_1}},
\label{b14}
\end{equation}
and the first integral of the system is given by
\begin{equation}
F(x,y) = \displaystyle\frac{1}{2}\,
{\rm sign}(1+a_1 x) \, |1+a_1 x|^{- \frac{2 a_4}{a_1}}
\left[ y^2 + \displaystyle\frac{(1+a_1-a_4)\, (1 + 2\, a_4\, x)}{a_4\, (a_1 - a_4)\,
(a_1 - 2\, a_4)} - \displaystyle\frac{x^2}{a_1 - a_4} \right].
\label{b15}
\end{equation}
For the $Q_3^{LV}$-\,Lokta-Volterra system (\ref{b9}),
we find the integrating factor to be
\begin{equation}
\gamma = | g(x,y) |^{-1}, \quad
{\rm where} \ \
g(x,y) = (1+a_1 x)\, \Big[(x-1)^2 + a_3 \, (x-1) \, y - (1+a_1)\, y^2 \Big],
\label{b16}
\end{equation}
and the first integral of the system is
\begin{equation}
F(x,y) = \!
\left\{ \!\!
\begin{array}{ll}
-\, \displaystyle\frac{{\rm sign} (g(x,y))} {2\, a_1 (1+a_1)}
{\LARGE \Big\{} 2\, \ln |1\!+\!a_1 x| \!+\!
a_1 \ln \left| (1\!+\!a_1) y^2 \!-\! a_3 y (x\!-\!1) \!-\! (x\!-\!1)^2
\right| \\[1.0ex]
\hspace{1.2in}
+\, \frac{ 2\,a_1 \,a_3\,(x-1)}{
\sqrt{[a_3^2 + 4\, (1+a_1)]\,(x-1)^2}}\,
\tanh^{-1} \! \Big[
\frac{ a_3 \,(x-1) - 2 ( 1+a_1) \,y}
{\sqrt{[a_3^2 + 4\, (1+a_1)]\,(x-1)^2}}
\Big] {\LARGE \Big\}}, \\ [2.5ex]
\qquad \qquad \qquad {\rm when} \quad a_3^2 + 4\,(1+a_1) > 0,
\\[2.0ex]
-\, \displaystyle\frac{{\rm sign} (1\!+\!a_1 x)} {2\, a_1 (1+a_1)}
{\LARGE \Big\{} 2\, \ln |1\!+\!a_1 x| \!+\!
a_1 \ln \left[ (1\!+\!a_1) y^2 \!-\! a_3 y (x\!-\!1) \!-\! (x\!-\!1)^2
\right] \\[1.0ex]
\hspace{1.2in}
-\, \frac{ 2\,a_1 \,a_3\,(x-1)}{
\sqrt{[-\,a_3^2 - 4\, (1+a_1)]\,(x-1)^2}}\,
\tan^{-1} \! \Big[
\frac{ a_3 \,(x-1) - 2 ( 1+a_1) \,y}
{\sqrt{[-\,a_3^2 - 4\, (1+a_1)]\,(x-1)^2}}
\Big] {\LARGE \Big\}}, \\ [2.5ex]
\qquad \qquad \qquad {\rm when} \quad a_3^2 + 4\,(1+a_1) < 0.
\end{array}
\right.
\label{b17}
\end{equation}
Finally, for the $Q_4$\,-\,codimension-4 system (\ref{b11}), we have
\begin{equation}
\gamma = | g(x,y) |^{-5/2}, \quad {\rm where} \ \
g(x,y) = 1 - 2\,( 1 + 2\, a_2^2)\,x - 2\, a_2 \, y
+ (1+4\, a_2^2)\, (x + a_2\, y)^2,
\label{b18}
\end{equation}
and the first integral of the system is equal to
\begin{equation}
F(x,y) = \displaystyle\frac{{\rm sign} (g(x,y))}{12\, a_2^6} \,
| g(x,y)|^{-3/2} f(x,y),
\label{b19}
\end{equation}
\vspace{-0.05in}
\noindent
where
\begin{equation}
\begin{array}{rl}
f(x,y) =\!\!& -\,(1+a_2^2)
+ 3\,( x + a_2 \, y + 2\, a_2^2\, x)
\, \Big[ 1+a_2^2 - (1+3\,a_2^2)\, ( x+ a_2\, y) \Big] \nonumber \\ [1.5ex]
\!\!& +\, (1+3\,a_2^2)\, (1+4\,a_2^2)\, (x + a_2\,y)^3.
\end{array}
\label{b20}
\end{equation}
The proof is complete.
\put(10,0.5){\framebox(6,7.5)}
Note that among the four classifications of the integrable system (\ref{b6}),
the first three classified systems (\ref{b7}), (\ref{b8}) and (\ref{b9})
have two free parameters, while the last system (\ref{b11})
has only one free parameter.
\vspace{0.10in}
\noindent
{\bf Remark 1.3.} We now show that our classification in Theorem 1.1 is
equivalent to that given by \.{Z}ol\c{a}dek~\cite{Zoladek1994}.
The general quadratic system considered in~\cite{Zoladek1994} is given
in the complex form:
\begin{equation}
\displaystyle\frac{d z}{dt} = (i + \lambda)\, z + A \, z^2 + B \, z \, \bar{z}
+ C\, \bar{z}^2,
\label{b21}
\end{equation}
where $\, z = x + i\, y$, and $\, A, \ B\,$ and $C\,$ are complex
coefficients. It has been shown in~\cite{Zoladek1994} that
the point $\, z = 0\,$ is a center if and only if one of the
following conditions is fulfilled:
\begin{equation}
\begin{array}{ll}
Q_3^{LV}: & \lambda = B = 0, \\[1.5ex]
Q_3^H: & \lambda = 2\,A + \bar{B} = 0, \\[1.5ex]
Q_3^R: & \lambda = {\rm Im}(AB) = {\rm Im} (\bar{B}^3 C)
= {\rm Im} (A^3 C) = 0, \\[1.5ex]
Q_4: & \lambda = A - 2\,\bar{B} = |C| - |B| = 0.
\end{array}
\label{b22}
\end{equation}
In the following,
we first use real differential equation to give a brief proof
(different from \.{Z}ol\c{a}dek's~\cite{Zoladek1994}),
and then show that our classification is equivalent to \.{Z}ol\c{a}dek's
when system (\ref{b21}) is assumed to have a non-zero singularity.
To prove this, let
$$
A = A_1 + i\, A_2, \quad B = B_1 + i\, B_2, \quad C = C_1 + i\, C_2,
\quad (i^2 = -\,1),
$$
and then rewrite the complex equation (\ref{b21}) in the real form:
\begin{equation}
\begin{array}{ll}
\displaystyle\frac{d x}{dt} = \lambda\,x + y
+ (A_1 + B_1 + C_1)\, x^2
+ 2\, (A_2 - C_2)\, x\, y
- (A_1 - B_1 + C_1)\, y^2 , \\[1.0ex]
\displaystyle\frac{d y}{dt} = -\,x + \lambda\,y
- (A_2 + B_2 + C_2)\, x^2
+ 2\, (A_1 - C_1)\, x\, y
+ (A_2 - B_2 + C_2)\, y^2 ,
\end{array}
\label{b23}
\end{equation}
where $\, y \rightarrow -\, y \,$ has been used.
Letting $\, \lambda = 0 \,$ yields the focus value $\, v_0 = 0$.
Then, it is easy to find the first focus value (or the first Lyapunov constant)
as
\begin{equation}
v_1 = -\,A_1\, B_2 - B_1\, A_2 = -\, {\rm Im}(AB).
\label{b24}
\end{equation}
Letting $\, v_1 = 0 \,$ results in $\, {\rm Im}(AB) = 0$,
which gives
\begin{equation}
B_2 = -\, \displaystyle\frac{B_1\, A_2}{A_1}, \quad
{\rm under \ the \ assumption \ of} \ \ A_1 \ne 0.
\label{b25}
\end{equation}
(The degenerate case $A_1 = 0$ can be similarly analyzed and the
details are omitted here.)
Then, we apply our Maple program (e.g., see~\cite{Yu1998}) to
system (\ref{b23}), with the conditions $\lambda =0$ and (\ref{b25}),
to obtain
$$
v_2 = \displaystyle\frac{-\,f\, (A_1 - 2\, B_1)}{3 A_1^3} , \ \ \
v_3 = \displaystyle\frac{-\,f\, f_3}{216 A_1^5}, \ \ \
v_4 = \displaystyle\frac{-\,f\, f_4}{9720 A_1^7}, \ \ \
v_5 = \displaystyle\frac{-\,f\, f_5}{466560 A_1^9}, \ \ \cdots
$$
where
$$
f = B_1\, (2\,A_1 +B_1)\, ( C_2\, A_1^3 + 3\, C_1\,A_2\, A_1^2 -
3\, A_2^2 \, C_2 \, A_1 - C_1 \, A_2^3),
$$
and $\, f_3, \, f_4$, etc. are polynomials of
$\, A_1,\,A_2,\,C_1,\,C_2\,$ and $\,B_1$.
Letting $\, f =0$, i.e.,
$$
B_1 = 0 \quad {\rm or} \quad
2\,A_1 +B_1 = 0 \quad {\rm or} \quad
C_2 \, A_1^3 + 3 \,C_1 \,A_2 \,A_1^2 -
3 \,A_2^2\, C_2 \,A_1 - C_1 \,A_2^3 = {\rm Im}(A^3 C) = 0
$$
yields $\, v_2 = v_3 = \cdots =0$.
Indeed, $\, B_1 = 0\,$ implies $\, B_2 = 0$ due to the condition (\ref{b25}),
and so $\, B = 0$.
Thus, we obtain $\, \lambda = B = 0$, corresponding to the $Q_3^{LV}\,$ case.
For the condition $\,2\,A_1 +B_1 = 0$, it follows from (\ref{b25}) that
$\, 2\,A_2 - B_2 = 0$, i.e., $\, 2\, A + \bar{B} =0$, which
plus the condition $\,\lambda =0 \,$ gives the $\, Q_3^H \,$ case.
The third condition $\, {\rm Im}(A^3 C) = 0 $, with $\, \lambda = 0 \,$
and $\, {\rm Im}(AB) = 0$, corresponds to the $Q_3^R \,$ case.
Further, it is easy to show that under the condition $\, {\rm Im}(AB) = 0$,
$\, {\rm Im}(A^3 C) = 0 \,$ and $\, {\rm Im}(\bar{B}^3 C) = 0 \,$
are equivalent. Thus, the conditions $\, \lambda = {\rm Im}(AB) =
{\rm Im}(\bar{B}^3 C) = 0 \,$ also applicable for this case.
So for this case, either $\, {\rm Im}(A^3 C) = 0\,$
or $\, {\rm Im}(\bar{B}^3 C) = 0 \,$
is needed, but not both of them.
In the following, we show one more case to join this case, leading to
both the two conditions being needed.
Note that there is one more condition $\, A_1 = 2\, B_1 \,$ which renders
$\, v_2 = 0$. Letting $\, A_1 = 2\, B_1$, and so
$\, A_2 = -\, 2\, B_2 \,$ (see (\ref{b25})), implying that
$\, A - 2\, \bar{B} = 0$.
Under the condition $\, A = 2\, \bar{B} $,
$\, v_1 = v_2 = 0$, and the other focus values become
\begin{eqnarray*}
v_3 &\!\!\!=\!\!\!&
\displaystyle\frac{25}{8} \,
(C_1^2+C_2^2-B_1^2-B_2^2)
( C_2 B_1^3 -3 C_1 B_1^2 B_2 -3 C_2 B_1 B_2^2 + C_1 B_2^3 ), \\[0.5ex]
v_4 &\!\!\!=\!\!\!&
\displaystyle\frac{v_3}{45}\,
\Big[45 B_1^2
+585 B_2^2
+60 (B_1 C_1+B_2 C_2)
-196 (C_1^2+C_2^2)
\Big], \\[0.5ex]
v_5 &\!\!\!=\!\!\!&
\displaystyle\frac{v_3}{6480}\,
\Big[
648 (7 B_1^4 \!+\! 124 B_1^2 B_2^2 \!+\! 1557 B_2^4)
\!-\! 3 (961 B_1^2 C_1^2 \!-\! 7680 B_1 B_2 C_1 C_2 \!+\! 202345 B_2^2 C_2^2)
\ \ \\
&\!\!\! \!\!\!&
\qquad \quad +\, 576 B_1 C_1 (106 B_1^2 + 307 B_2^2)
+ 288 B_2 C_2 (371 B_1^2 + 773 B_2^2)
\\
&\!\!\! \!\!\!&
\qquad \quad -\, 3 (4801 B_1^2 C_2^2 + 206185 B_2^2 C_1^2)
-80688 (C1^2+C2^2) (B_1 C_1 + B_2 C_2)
\\
&\!\!\! \!\!\!&
\qquad \quad +\, 86144 (C1^2+C2^2)^2
\Big], \\[-1.0ex]
&\!\!\! \vdots \!\!\!&
\end{eqnarray*}
Hence, under the conditions $\, \lambda = A - 2\, \bar{B} = 0$,
there are two possibilities such that
$\, v_3 = v_4 = \cdots = 0$. The first possibility is
$$
C_1^2 + C_2^2 - B_1^2 - B_2^2, \quad {\rm i.e.,} \quad
|C| - |B| = 0,
$$
which is one of the conditions given for the $Q_4$ case
(see (\ref{b22})).
The second possibility is given by the condition:
\begin{equation}
C_2 \, B_1^3 -3 \,C_1 \,B_1^2 \,B_2 -3\,C_2\, B_1\, B_2^2 + C_1\, B_2^3 =
{\rm Im} (\bar{B}^3 C) = \textstyle\frac{1}{8}\, {\rm Im} (A^3 C) = 0,
\label{b26}
\end{equation}
due to $\, A = 2 \ \bar{B}$.
Since these conditions can be included in the conditions
$\, \lambda = {\rm Im}(AB) = {\rm Im} (\bar{B}^3 C) = {\rm Im} (A^3 C) = 0$,
this possibility belongs to the $\, Q_3^R \,$ case.
The remaining task is to show that the conditions classified in
(\ref{b22}) are sufficient. This can be done by finding an integrating factor
for each case. For brevity, we only list these integrating factors
below (while the lengthy expressions of the first integrals are omitted):
\begin{equation}
\gamma = \left\{
\begin{array}{ll}
\Big| 1+4 \,(A_2\, x-A_1 \,y) + 4 \,(A_1 C_2+A_2 C_1-2 A_1 A_2) \,x\, y
\\ [1.0ex]
\quad +\, \left[ (A_1+C_1)\,(A_1-3 C_1)+(A_2+C_2)\,(5 A_2-3 C_2)\right] x^2
\\ [1.0ex]
\quad +\, \left[(A_2+C_2)\,(A_2-3 C_2)+(A_1+C_1)\,(5 A_1-3 C_1)\right] y^2
\\ [1.0ex]
\quad +\,
2\, (A_1^2+A_2^2-C_1^2-C_2^2) \left[ (A_2+C_2)\, x^3 - (A_1+C_1)\, y^3
\right. \\ [0.5ex]
\hspace{1.0in} -\left. (A_1-3 C_1)\, x^2 y + (A_2-3 C_2) \,x y^2 \right]
\Big|^{-1}, & {\rm for} \ \ Q_3^{LV}, \\[2.0ex]
1, & {\rm for} \ \ Q_3^{H}, \\[1.5ex]
\left| 1 - 2\, (A_1 - C_1)\, y \right|^{-\, \frac{2\,A_1 + B_1}{A_1-C_1}},
& {\rm for} \ \ Q_3^{R}, \\[2.0ex]
\Big| 1 - 4 ( B_2\, x + B_1\, y)
+ 2 (B_1^2+B_2^2)\, (x^2+y^2) \\ [0.0ex]
\quad +\, 2 (B_1 C_1 + B_2 C_2)\,(x^2-y^2)
+ 4 \,(B_1 C_2 - B_2 C_1)\, x\, y \Big|^{-5/2}, & {\rm for} \ \
Q_4.
\end{array}
\right.
\label{b27}
\end{equation}
For the integrating factors of degenerate cases (e.g.,
$\, A_1 - C_1 = 0$), one can easily find them.
Next, compare the classification listed in (\ref{b22}) with ours
given in Theorem 1.1.
First, consider the $\, Q_3^{LV}\,$ case. Letting
$\, \lambda = B_1 = B_2 = 0\,$ in (\ref{b23}) yields
\begin{equation}
\begin{array}{ll}
\displaystyle\frac{d x}{dt} = y + (A_1 + C_1)\, x^2 + 2\, (A_2 - C_2)\, x\, y
- (A_1 + C_1)\, y^2 , \\[1.0ex]
\displaystyle\frac{d y}{dt} = -\,x - (A_2 + C_2)\, x^2 + 2\, (A_1 - C_1)\, x\, y
+ (A_2 + C_2)\, y^2 .
\end{array}
\label{b28}
\end{equation}
Then, let
\begin{equation}
k = \tan(\theta), \quad {\rm and \ so} \ \
\sin(\theta) = \displaystyle\frac{k}{\sqrt{1+k^2}}, \quad
\cos(\theta) = \displaystyle\frac{1}{\sqrt{1+k^2}},
\label{b29}
\end{equation}
where $\, k\,$ is solved from the following cubic polynomial:
\begin{equation}
P_1(k) = (A_2+C_2)\, k^3 + (A_1 - 3\, C_1)\, k^2 + (A_2 - 3\,C_2)\, k
+ A_1 + C_1 = 0.
\label{b30}
\end{equation}
This cubic polynomial at least has one real solution for $k$, which gives the
slope of the line on which a second fixed point is located.
$\, k = 0 \,$ if $\, A_1 + C_1 =0$, otherwise, $k \ne 0$.
Let $\overline{k}$ be a real root of $P_1(k)$, i.e., $P_1(\overline{k})=0$.
Further, introducing the linear transformation (rotation):
\begin{equation}
x = \cos(\theta)\, u - \sin(\theta)\, v, \quad
y = \sin(\theta)\, u + \cos(\theta)\, v,
\label{b31}
\end{equation}
into (\ref{b28} yields
\begin{equation}
\begin{array}{ll}
\displaystyle\frac{d x}{dt} = y + m_{120}\, x^2 + m_{111}\, x\, y + m_{102}\, y^2, \\[1.0ex]
\displaystyle\frac{d y}{dt} =-\,x + m_{220}\, x^2 + m_{211}\, x\, y + m_{202}\, y^2,
\end{array}
\label{b32}
\end{equation}
where
$$
\begin{array}{ll}
m_{120} = -\, m_{102} = (1+ \overline{k}^2)^{-3/2}\,
P_1(\overline{k})=0,
\\[1.5ex]
m_{220} = -\, m_{202} = (1+ \overline{k}^2)^{-3/2}\, \Big[
(A_1+C_1)\, \overline{k}^3 - (A_2-3\,C_2)\, \overline{k}^2
+ (A_1 - 3\, C_1)\, \overline{k}
-A_2 - C_2 \Big], \\[1.5ex]
m_{111} = -\,2\,(1+ \overline{k}^2)^{-3/2}\, \Big[
(A_1-C_1)\, \overline{k}^3 - (A_2+3\,C_2)\, \overline{k}^2
+ (A_1 + 3\, C_1)\, \overline{k}
-A_2 + C_2 \Big], \\[1.5ex]
m_{211} = 2\,(1+ \overline{k}^2)^{-3/2}\, \Big[
(A_2-C_2)\, \overline{k}^3 + (A_1+3\,C_1)\, \overline{k}^2
+ (A_2 + 3\, C_2)\, \overline{k}
+A_1 - C_1 \Big].
\end{array}
$$
Suppose $\, m_{220} \ne 0$. Then, introducing
$\, \overline{x} = m_{220}\, x, \ \overline{y} = m_{220}\, y \,$ into
(\ref{b32}) results in
\begin{equation}
\begin{array}{ll}
\displaystyle\frac{d \overline{x}}{dt} = \overline{y} + \displaystyle\frac{m_{111}}{m_{220}}\,
\overline{x}\, \overline{y}, \\[2.0ex]
\displaystyle\frac{d \overline{y}}{dt} =-\, \overline{x} + \overline{x}^2
+ \displaystyle\frac{m_{211}}{m_{220}}\, x\, y - \overline{y}^2,
\end{array}
\label{b33}
\end{equation}
which is identical to (\ref{b9}) as long as letting
$\, a_1 = \frac{m_{111}}{m_{220}} \,$
and $\, a_3 = \frac{m_{211}}{m_{220}}$.
This shows that the four parameters
$\, A_1, \, A_2, \, C_1\, $ and $\, C_2 \,$ are not independent. Thus,
alternatively, we may simply take $ \overline{k} = 0 $
(which renders the second singularity of (\ref{b28}) on the $x$-axis),
yielding $\, C_1 = -\,A_1 $. Thus, (\ref{b28}) becomes
$$
\begin{array}{ll}
\displaystyle\frac{d x}{dt} = y + 2\, (A_2 - C_2)\, x\, y , \\[1.0ex]
\displaystyle\frac{d y}{dt} = -\,x - (A_2 + C_2)\, x^2 + 4\, A_1 \, x\, y
+ (A_2 + C_2)\, y^2 .
\end{array}
$$
Suppose $\, A_2 + C_2 \ne 0$. Introducing
$\, \overline{x} = -\,(A_2+C_2)\, x , \
\overline{y} = -\,(A_2+C_2)\, y \, $ into the above equations
we obtain
\begin{equation}
\begin{array}{ll}
\displaystyle\frac{d \overline{x}}{dt} = \overline{y}
- \textstyle\frac{2\,(A_2 - C_2)}{A_2 + C_2} \, \overline{x}\, \overline{y} ,
\\[1.5ex]
\displaystyle\frac{d \overline{y}}{dt} = -\, \overline{x} + \overline{x}^2
- \textstyle\frac{4\, A_1}{A_2 + C_2} \, \overline{x}\, \overline{y}
- \overline{y}^2 ,
\end{array}
\label{b34}
\end{equation}
which is identical to (\ref{b9}) if letting
$\,a_1 = \frac{-\,2\,(A_2 - C_2)}{A_2 + C_2} \,$
and $\, a_3 = \frac{-\,4\, A_1}{A_2 + C_2}$.
In the following, we will use this simple approach for other cases.
For the $\, Q_3^H $ case, substituting $\, \lambda = 0$,
$\,B_1 = -\,2\,A_1 \,$ and
$\, B_2 = 2\, A_2 \,$ into system (\ref{b23}) results in
$$
\begin{array}{ll}
\displaystyle\frac{d x}{dt} = y - (A_1 - C_1)\, x^2 + 2\, (A_2 - C_2)\, x\, y
- (3\,A_1 + C_1)\, y^2 , \\[1.0ex]
\displaystyle\frac{d y}{dt} = -\,x - (3\, A_2 + C_2)\, x^2 + 2\, (A_1 - C_1)\, x\, y
- (A_2 - C_2)\, y^2 .
\end{array}
$$
Further, taking $\, C_1 = A_1\, $
in the above equations gives another singularity on the $x$-axis,
and introducing $\, \overline{x} = -\,(3\,A_2+C_2)\, x , \
\overline{y} = -\,(3\,A_2+C_2)\, y \, $ into the resulting equations
yields
\begin{equation}
\begin{array}{ll}
\displaystyle\frac{d \overline{x}}{dt} = \overline{y}
- \textstyle\frac{2\, (A_2 - C_2)}{3\, A_2 + C_2} \, \overline{x}\, \overline{y}
+ \textstyle\frac{4\, A_1}{3\, A_2 + C_2} \overline{y}^2, \\[1.5ex]
\displaystyle\frac{d \overline{y}}{dt} = -\,\overline{x} + \overline{x}^2
+ \textstyle\frac{A_2 - C_2}{3\, A_2 + C_2} \, \overline{y}^2 ,
\end{array}
\label{b35}
\end{equation}
which is identical to (\ref{b8}) if we set
$\, a_1 = \frac{-\,2\, (A_2 - C_2)}{3\, A_2 + C_2}\,$
and $\, a_2 = \frac{4\, A_1}{3\, A_2 + C_2}$.
For the $Q_3^R$ reversible case, it follows from~\cite{Zoladek1994} that
all the coefficients $A, \, B \,$ and $ C\, $ are real, and thus we obtain
the following real form from the complex system (\ref{b21})
\begin{equation}
\begin{array}{ll}
\displaystyle\frac{d x}{dt} = -\, y + a \, x^2 + b \, y^2 , \\[1.0ex]
\displaystyle\frac{d y}{dt} = x + c\, x\, y,
\end{array}
\label{b36}
\end{equation}
where
$$
a = A_1 + B_1 + C_1, \quad b = B_1 - A_1 - C_1 , \quad
c = 2\, A_1 - 2\, C_1.
$$
Suppose $\,b \ne 0$. Then, introducing
$\, \overline{x} = b\, y, \ \overline{y} = b\, x \,$
into (\ref{b36}) results in
\begin{equation}
\begin{array}{ll}
\displaystyle\frac{d \overline{x}}{dt} = \overline{y} + \displaystyle\frac{c}{b}\,
\overline{x}\, \overline{y}, \\[1.0ex]
\displaystyle\frac{d \overline{y}}{dt} = -\, \overline{x} + \overline{x}^2
+ \displaystyle\frac{a}{b} \, \overline{y}^2,
\end{array}
\label{b37}
\end{equation}
which is identical to (\ref{b7}) if
$$
a_1 = \displaystyle\frac{c}{b} = \displaystyle\frac{2\,(A_1 - C_1)}{B_1 - A_1 -C_1} \quad
{\rm and} \quad
a_4 = \displaystyle\frac{a}{b} = \displaystyle\frac{A_1 + B_1 + C_1}{B_1 - A_1 -C_1}.
$$
For the last $\, Q_4$ case,
under the condition $\, \lambda = A -
2\, \bar{B} = 0 $, by setting
$\, C_1 = -\,3 \, B_1 \, $ (which renders a non-zero singularity
on the $x$-axis) in (\ref{b23}) we obtain
$$
\begin{array}{ll}
\displaystyle\frac{d x}{dt} = y - 2\,(2\,B_2 +C_2) \, x\, y + 2\, B_1 \, y^2,
\\[1.0ex]
\displaystyle\frac{d y}{dt} = -\, x + (B_2-C_2)\, x^2 + 10 \, B_1 \, x\, y
- ( 3\, B_2 - C_2) \, y^2,
\end{array}
$$
Suppose $\, B_2 - C_2 \ne 0$.
Then, introducing $\, \overline{x} = (B_2 - C_2)\, x, \
\overline{y} = (B_2 - C_2)\, y \,$ into the above
equations yields
\begin{equation}
\begin{array}{ll}
\displaystyle\frac{d \overline{x}}{dt} = \overline{y}
- \textstyle\frac{2\,(2\,B_2 +C_2)}{B_2-C_2} \, \overline{x}\, \overline{y}
+ \textstyle\frac{2\, B_1}{B_2-C_2} \, \overline{y}^2,
\\[1.5ex]
\displaystyle\frac{d \overline{y}}{dt} = -\, \overline{x} + \overline{x}^2
+ \textstyle\frac{10 \, B_1}{B_2-C_2} \, \overline{x}\, \overline{y}
- \textstyle\frac{ 3\, B_2 - C_2}{B_2-C_2} \, \overline{y}^2,
\end{array}
\label{b38}
\end{equation}
Comparing the coefficients of the above system (\ref{b38})
with our system (\ref{b6}) results in
\begin{equation}
a_1 = -\,\textstyle\frac{2\,(2\,B_2 +C_2)}{B_2-C_2}, \quad
a_2 = \textstyle\frac{2\, B_1}{B_2-C_2}, \quad
a_3 = \textstyle\frac{10 \, B_1}{B_2-C_2}, \quad
a_4 = -\, \textstyle\frac{ 3\, B_2 - C_2}{B_2-C_2},
\label{b39}
\end{equation}
which in turn implies that
$\, a_3 - 5\, a_2= a_1 - (5+3\,a_4) = 0$,
and
$$
a_4 + 2\, (1+a_2^2) = \textstyle\frac{8\,B_1^2 + C_2^2 - B_2^2}{(B_2-C_2)^2}
= \textstyle\frac{C_1^2 + C_2^2 - B_1^2 - B_2^2}{(B_2-C_2)^2}
= 0, \quad {\rm for} \ \ |C| - |B|=0.
$$
The above conditions are the exact conditions given in (\ref{b10}) for
the $\, Q_4\, $ case.
Finally, we turn to the conditions
given in (\ref{b12}). It follows from (\ref{b39}) that
\begin{equation}
3\,(a_4+2)\,(a_4+1)^2 - (5\,a_4+6)\, a_2^2 =
-\,\textstyle\frac{ 4}{(B_2-C_2)^3} \,
( 3\,B_2^3+3\,B_2^2\,C_2-C_1^2\,B_2-B_1^2\,C_2 ).
\label{b40}
\end{equation}
On the other hand, under the condition $\, C_1 = -\,3\,B_1 $,
the condition (\ref{b26}) for the second possibility becomes
$$
\begin{array}{rl}
C_2 B_1^3 -3 C_1 B_1^2 B_2 -3 C_2 B_1 B_2^2 + C_1 B_2^3
= \!\!&C_2 B_1^3 + C_1^2 B_1 B_2 -3 C_2 B_1 B_2^2 - 3 B_1 B_2^3
\\ [1.5ex]
= \!\!& -\, B_1 \, ( 3\,B_2^3+3\,B_2^2\,C_2-C_1^2\,B_2-B_1^2\,C_2 )
= 0 ,
\end{array}
$$
which implies, by Eq.~(\ref{b40}),
$\, 3\,(a_4+2)\,(a_4+1)^2 - (5\,a_4+6)\, a_2^2 = 0 \,$ for $\, B_1 \ne 0$.
Hence, according to \.{Z}ol\c{a}dek's classification (see (\ref{b22})),
this case should be included in the $Q_3^R \,$ case.
However, one can not prove this by directly using the conditions in (\ref{b12})
as well as that for the $Q_3^R $ case (see Theorem 1.1). One must
trace back to the original system coefficients.
In~\cite{Zoladek1994}, \.{Z}ol\c{a}dek used Bautin's system to verify his
classification. Bautin's system is described by~\cite{bautin}
\begin{equation}
\begin{array}{ll}
\displaystyle\frac{d x}{dt} = \lambda_1 \,x - y
+ \lambda_3 \, x^2
+ (2\, \lambda_2 + \lambda_5)\, x\, y
+ \lambda_6 \, y^2 , \\[1.0ex]
\displaystyle\frac{d y}{dt} = x + \lambda_1 \,y
+ \lambda_2 \, x^2
+ (2\,\lambda_3 + \lambda_4)\, x\, y
- \lambda_2 \, y^2 .
\end{array}
\label{b41}
\end{equation}
It is seen from (\ref{b23}) and (\ref{b41}) that Bautin's system
has only $6$ parameters, while \.{Z}ol\c{a}dek's system
has $7$ (in real domain) parameters. This indicates that
\.{Z}ol\c{a}dek's system has one redundant parameter.
In fact, putting Bautin's system in \.{Z}ol\c{a}dek's complex form gives
the following expressions:
$$
\begin{array}{ll}
\lambda = \lambda_1, \quad
A = \textstyle\frac{1}{4}\,(\lambda_3 + \lambda_4 - \lambda_6 - i\, \lambda_5),
\quad B = -\,\textstyle\frac{1}{2}\, ( \lambda_3 - \lambda_6), \\[1.0ex]
C = \textstyle\frac{1}{4}\, \Big[ - ( 3\, \lambda_3 + \lambda_4 + \lambda_6)
+ i\, (4\,\lambda_2 + \lambda_5) \Big].
\end{array}
$$
Then, applying the formulas given in (\ref{b23}) will immediately
generate the centers conditions obtained by Bautin~\cite{bautin}.
The above expressions clearly show that $\, B_2 = 0$. As a matter of factor,
the integral factor for the system, corresponding to the second possibility,
i.e., when $\lambda = A - 2 \, \bar{B} = {\rm Im} (\bar{B}^3 C ) = 0$, is
given by
$$
\left| 1 + 2 \left[ \textstyle\frac{ C_1 (B_1^2+B_2^2)} {B_1 (B_1^2-3 B_2^2)}
-2 \right] (B_2\, x+ B_1\, y) \right|^{\frac{5 B_1 (B_1^2-3 B_2^2) }
{C_1 (B_1^2+B_2^2)-2 B_1 (B_1^2-3 B_2^2)}}.
$$
For $\, B_2 = 0$, the above expression is reduced to
$$
\Big| 1 - 2\, ( 2\, B_1 - C_1 )\, y \Big|^{ \frac{5 \, B_1}{C_1 - 2\,
B_2}}
= \Big| 1 - 2\, (A_1 - C_1)\, y \Big|^{-\, \frac{2\,A_1 + B_1}{A_1-C_1}}
\quad ({\rm due \ to} \ \, A_1 = 2\, B_1),
$$
which is the integrating factor for the $Q_3^R \,$ system, as shown in
(\ref{b27}).
\vspace{0.10in}
Now we return to system (\ref{b6}).
Among the four classifications,
the Hamiltonian system ($Q_3^H$) has been completely studied
in~\cite{HI1994,Gavrilov2001}: the system can have maximal two limit cycles.
In this paper, we will concentrate on the $Q_3^R$\,-\,reversible case.
Special cases for the reversible system have been investigated by
a number of authors
(e.g., see~\cite{DumortierLiZhang1997,Peng2002,YuLi2002,IlievLiYu2005,LL2006}).
It is easy to see that system (\ref{b7})
is invariant under the mapping $(t,\,y) \to (-t,\, -y)$,
where $\, a_1 \,$ and $\, a_4 \,$ can be considered as perturbation
parameters.
The singular point $(1,0)$ of (\ref{b7}) is a center when $a_1 < -\,1$;
but a saddle point when $\, a_1 > -\,1 $.
$a_1 = -\,1 \,$ gives a degenerate singular point at $(1,0)$.
Further, it is easy to verify that
when $\, (a_1 +1)\,a_4 >0$, there are no more singularity;
while when $\, (a_1 + 1)\, a_4 < 0 $, there exist additional two
saddle points, given by
$$
(x^*\!,y^*) = \Big(- \frac{1}{a_1}, \
\pm \textstyle\frac{ \sqrt{ -\, a_4 \, ( a_1 + 1)}}{a_1 \, a_4} \Big).
$$
$a_4 = 0\,$ is a critical value, yielding the two additional saddle
points at infinity: $ (x^*\!, y^*) = ( -\, \frac{1}{a_1},\, \pm \infty)$.
In summary, the distribution of singularity of the reversible
system (\ref{b7}) has the following possibility
(see Fig.~\ref{fig1}, where 1C+1S stands for one center and one
saddle point, similar meaning applies to 2C, 2C+2S and 1C+3S):
\begin{equation}
\begin{array}{ll}
{\rm Two \ centers \ when} \ \ a_1 < -\,1 \ \, {\rm and} \, \ a_4 < 0; \\[1.0ex]
{\rm Two \ centers \ and \ two \ saddle \ points \ when} \ \
a_1 < -\,1 \, \ {\rm and} \, \ a_4 > 0; \\[1.0ex]
{\rm One \ center \ and \ one \ saddle \ point \ when} \ \
a_1 > -\,1 \, \ {\rm and} \, \ a_4 > 0;&
\\[1.0ex]
{\rm One \ center \ and \ three \ saddle \ points \ when} \ \
a_1 > -\,1 \, \ {\rm and} \, \ a_4 < 0.&
\end{array}
\label{b43}
\end{equation}
In this paper, we pay particular attention to
$\, a_1 < -\,1, \ a_4 < 0$, for which system (\ref{b7}) has only two
singularities at $(0,0)$ and $(1,0)$, both of them are centers.
By adding quadratic perturbations to system (\ref{b7}) we obtain
the following perturbed quadratic system:
\begin{equation}
\begin{array}{rl}
\displaystyle\frac{dx}{dt} = & y \, (1 + a_1\, x)
+ \varepsilon \, P(x,y) \\ [1.0ex]
= & y\, (1 + a_1\, x)
+ \varepsilon \, (a_{10}\, x + a_{01}\, y
+a_{20}\, x^2 + a_{11}\, x\, y + a_{02}\, y^2), \\ [0.5ex]
\displaystyle\frac{dy}{dt}= & -\, x + x^2 + a_4\, y^2 + \varepsilon \, Q(x,y) \\ [1.0ex]
= & -\, x + x^2 + a_4\, y^2
+ \varepsilon \, (b_{10}\, x + b_{01}\, y
+ b_{20}\, x^2 + b_{11}\, x\, y + b_{02}\, y^2),
\end{array}
\label{b44}
\end{equation}
where $\, 0 < \varepsilon \ll 1 $, $a_{ij}$'s and $b_{ij}$'s are perturbation
parameters.
\vspace{0.10in}
\noindent
{\bf Remark 1.4.} The special system considered in~\cite{DumortierLiZhang1997}
is the system (\ref{b4}) with
$$
a = -\,3, \quad c = -\,2, \quad b = 1.
$$
This is equivalent to our system when $\, a_1 = -\,2 \,$ and
$\, a_4 = -\,3$ for which the system has only two centers at $(0,0)$ and
$(1,0)$.
Consider the $a_1$-$a_4$ parameter plane, as shown in Fig.~\ref{fig1}.
It can be seen that the case considered in~\cite{DumortierLiZhang1997} is just a
point, $(a_1, a_4) = (-\,2,\, -\,3) $, in the parameter plane,
marked by a blank circle in the third quadrant
on the line $\, a_4 = \frac{3}{2}\, a_1$ (see Fig.~\ref{fig1}).
The special system studied in~\cite{Peng2002}
is the system (\ref{b4}) with
$$
a = -\,3, \quad c = -\,2, \quad b = -\,1.
$$
This is equivalent to our system when $\, a_1 = 2 \,$ and
$\, a_4 = 3$, for which the system has one center at $(0,0)$
and one saddle point at $(1,0)$.
Thus, this case considered in~\cite{Peng2002} is
again a point, $\, (a_1, a_4) = (2,\, 3) $,
in the $a_1$-$a_4$ parameter plane,
marked by another blank circle in the first quadrant
on the line $\, a_4 = \frac{3}{2}\, a_1$ (see Fig.~\ref{fig1}).
The cases considered in~\cite{YuLi2002,IlievLiYu2005}
correspond to the system (\ref{b4}) with $\, a=-\,3, \ c=-\,2$,
and $\, b \in (-\,\infty, -1) \cup (-1,0)\,$ in~\cite{YuLi2002},
and $\, b \in (0, 2) $ in~\cite{IlievLiYu2005}.
When $\, \varepsilon = 0$ in system (\ref{b4}), one can use the
following transformation:
$$
x = \displaystyle\frac{\tilde{y}}{b}, \quad y = \displaystyle\frac{\tilde{x}}{b},
$$
to transform system (\ref{b4})$_{\varepsilon=0}$ to
\begin{equation}
\begin{array}{ll}
\displaystyle\frac{d \tilde{x}}{dt} =
\tilde{y} \, \Big( 1 + \displaystyle\frac{c}{b} \, \tilde{x} \Big), \\[1.5ex]
\displaystyle\frac{d \tilde{y}}{dt} = -\, \tilde{x} + \tilde{x}^2
+ \displaystyle\frac{a}{b}\, \tilde{y}^2,
\end{array}
\label{b45}
\end{equation}
which is our system (\ref{b7}) with
\begin{equation}
a_1 = \displaystyle\frac{c}{b}, \quad a_4 = \displaystyle\frac{a}{b}.
\label{b46}
\end{equation}
Equation (\ref{b46}) yields
\begin{equation}
a_4 = \displaystyle\frac{a}{c} \ a_1 \qquad (b \ne 0),
\label{b47}
\end{equation}
which represents a line in the $a_1$-$a_4$ parameter
plane, passing through the origin with the slope $\, \frac{a}{c}$.
In particular, the parameter values:
$\, a = -\,3, \ c = -\,2, \
b \in (- \infty, -1) \cup (-1, 0) \cup (0, 2)$, yielding
$a_1 = -\,\frac{2}{b} \,$ and $a_4 = -\, \frac{3}{b}$,
correspond to a part of the line, described by
\begin{equation}
a_4 = \frac{3}{2}\, a_1 \qquad
\forall \, a_1 \in (- \infty, -1) \cup (0, \infty),
\label{b48}
\end{equation}
as shown in Fig.~\ref{fig1},
where the dotted line for $ a_1 \in [-1,0]$ is excluded from the
studies~\cite{YuLi2002,IlievLiYu2005}.
It should be noted that when $\, a=-\,3, \ c=-\,2$,
the point $(0, \frac{1}{b})$ is
a saddle point if and only if
$$
1 + \displaystyle\frac{c}{b} = 1 - \displaystyle\frac{2}{b} > 0 \quad
\Longrightarrow \quad b \in (-\infty, 0) \cup (2, +\infty).
$$
Thus, the case considered in~\cite{YuLi2002} has
one center and one saddle point; while the case studied in~\cite{IlievLiYu2005}
has two centers.
But even these two studies together do not cover the whole line
$\, a_4 = \frac{3}{2}\, a_1$ (the missing part is denoted by a
dotted line segment in Fig.~\ref{fig1}).
Another alternative form for a special case of our system (\ref{b7})
considered in~\cite{Han1997} is described by
\begin{equation}
\begin{array}{ll}
\displaystyle\frac{d \overline{x}}{dt} = \overline{y}\, \Big[ 1 + 2\, ( 1-e)\,
\Big( \overline{x} + \displaystyle\frac{1}{d} \Big) \Big], \\ [1.5ex]
\displaystyle\frac{d \overline{y}}{dt} = \overline{x} + d\, \overline{x}^2 + e\,
\overline{y}^2,
\end{array}
\label{b49}
\end{equation}
where $\, e \,$ and $\, d \ (\ne 0) \,$ are parameters.
This system has a saddle point at the origin and a center at
$\, (\overline{x},\, \overline{y}) = ( -\frac{1}{d},\, 0)$.
Based on the two parameters, seven cases are classified~\cite{Han1997}.
We can apply the following transformation:
$$
\overline{x} = \displaystyle\frac{1}{d} \, ( x -1), \quad
\overline{y} = \displaystyle\frac{1}{d}\, y,
$$
to system (\ref{b13}), yielding
\begin{equation}
\begin{array}{ll}
\displaystyle\frac{d x}{dt} = y\, \Big[ 1 + \displaystyle\frac{2\, ( 1-e)}{d} \, x
\Big], \\ [2.0ex]
\displaystyle\frac{d y}{dt} = -\, x + x^2 + \displaystyle\frac{e}{d} \, y^2 ,
\end{array}
\label{b50}
\end{equation}
which has a center at the origin and a saddle point at $(1,0)$.
Then, setting
\begin{equation}
a_1 = \displaystyle\frac{2\,(1-e)}{d}, \quad a_4 = \displaystyle\frac{e}{d},
\label{b51}
\end{equation}
in system (\ref{b50}) leads to our system (\ref{b7}).
Equation (\ref{b51}) denotes a line, given by
\begin{equation}
a_4 = \displaystyle\frac{e}{2\,(1-e)} \ a_1,
\label{b52}
\end{equation}
in the $a_1$-$a_4$ parameter plane, passing through the origin with the slope
$\, \frac{e}{2\,(1-e)}$. However, it is easy to see that
using our system (\ref{b7}) in analysis is simpler than using
system (\ref{b49}). In fact, all the seven cases classified
in~\cite{Han1997} together denote a region in Fig.~\ref{fig1},
see the shaded area in this figure.
This area covers most of the region, defined by $a_1 > -\,1$.
But the study given in~\cite{Han1997} for the seven cases is restricted
to local analysis on the bifurcation of limit cycles near a homoclinic loop,
except the two lines (see Fig.~\ref{fig1}):
\begin{equation}
a_4 = a_1 \qquad \forall \, a_1 \in (-1, 0) \cup (0, \infty),
\label{b53}
\end{equation}
which corresponds to the parameter value $\, e = \frac{2}{3}$, and
\begin{equation}
a_4 = -\, \displaystyle\frac{1}{2}\, a_1 \qquad \forall \, a_1 \in (0, \infty),
\label{b54}
\end{equation}
which corresponds to $\, e \rightarrow \pm \infty$.
It has been shown~\cite{Han1997} that except the above two lines,
for the parameter values in the shaded area, system (\ref{b49})
can have at most $2$ limit cycles near a homoclinic loop
under quadratic perturbation.
\begin{figure}[!h]
\vspace{-1.45in}
\hspace{-0.50in}
\resizebox{1.2 \textwidth}{!}{\includegraphics{FIG1.eps}}
\vspace{-5.90in}
\caption{Case studies for the $Q_3^R$ reversible system.}
\label{fig1}
\vspace{0.10in}
\end{figure}
Figure~\ref{fig1} shows
the $a_1$-$a_4$ parameter plane associated with the reversible
system (\ref{b7}), where the above mentioned case studies
are indicated on the line $\, a_4 = \frac{3}{2}\, a_1 \,$ as well
as in the shaded area.
More precisely, a complete global analysis given in~\cite{YuLi2002},
which includes the result in~\cite{DumortierLiZhang1997} as a special case,
shows that corresponding to each point on the line segment
$\, a_4 = \frac{3}{2}\, a_1 \ (a_1 > 0)$, the system has one center and
one saddle point, and has maximal $2$ limit cycles. In~\cite{Han1997}
it is shown for each point in the shaded area (except the two line segments
$a_4 = a_1 \ ( a_1 > -1)$ and $ a_4 = -\, \frac{1}{2}\,a_1 \ (a_1 >0)$),
which contains the above line segment, the system has one center and
one (or three) saddle(s), and has maximal $2$ limit cycles, but restricted to
local analysis near one homoclinic loop.
Similarly, a global analysis given in~\cite{IlievLiYu2005},
which contains the result in~\cite{DumortierLiZhang1997} as a special case,
proves that corresponding to each point on the line segment
$\, a_4 = \frac{3}{2}\, a_1 \ (a_1 < -\,1)$, the system has two centers,
and exhibits maximal $3$ limit cycles around one center.
The technique of Poincar\'{e} transformation and Picar-Puchs equation,
used for the above mentioned global analysis on parameter unfolding,
seems not possible to be generalized to consider general situation for
arbitrary points in the $a_1$-$a_4$ parameter plane.
The two particular dash-dotted lines: $a_4 = \frac{1}{3}\,(a_1 - 5) \
\forall \, a_1 \! \in \! (-\infty, \, - 1) \cup (-1, \infty)$,
and $ a_4 = \frac{1}{3} \,(6\,a_1 + 5) \ \forall \, a_1 \! \in \!
(-\infty,\,-1)$, as well as the five dark circles correspond
to our results, presented in the next two sections. In particular,
we will show that there exist $3$ small limit cycles on
the two dash-dotted lines, and at least $4$ limit cycles for the
parameter values marked by the five dark circles.
\vspace{0.10in}
In the following, we will use the perturbed quadratic system (\ref{b44})
for our study on
bifurcation of limit cycles.
Without loss of generality, we may assume (e.g.,
see~\cite{DumortierLiZhang1997})
that $\, a_{01} = a_{20} = a_{11} = a_{02} = b_{10} = b_{20} = b_{02} = 0$.
Thus, system (\ref{b44}) is reduced to
\begin{equation}
\begin{array}{rl}
\displaystyle\frac{dx}{dt} = & y \, (1 + a_1\, x)
+ \varepsilon \, a_{10}\, x , \\ [1.5ex]
\displaystyle\frac{dy}{dt}= & -\, x + x^2 + a_4\, y^2
+ \varepsilon \, ( b_{01}\, y + b_{11}\, x\, y ),
\end{array}
\label{b55}
\end{equation}
where $\, a_1 < -\,1 \,$ and $\, 0 < \varepsilon \ll 1$.
\section{Hopf bifurcation associated with the two centers}
\setcounter{equation}{0}
\renewcommand{\theequation}{3.\arabic{equation}}
In this section, we study Hopf bifurcation of system (\ref{b55})
from two centers $(0,0)$ and $(1,0)$, leading to bifurcation of multiple
limit cycles. The result is summarized in the following theorem.
\vspace{0.1in}
\noindent
{\bf Theorem 2.1.} \ {\it
When $\, a_1 < -\,1$, the quadratic near-integrable system (\ref{b55})
can have small limit
cycles bifurcating from the two centers $(0,0)$ and $(1,0)$ with
distributions: $(3,0)$, $(0,3)$, $(2,0)$, $(0,2)$ and $(1,1)$.
$(2,1)$- or $(1,2)$-distribution does not exist.
}
\vspace{0.10in}
\noindent
{\bf Proof.} \
Consider system (\ref{b55}) for $\, a_1 < -\,1$.
The system (\ref{b55})$_{\varepsilon=0}$ is a reversible integrable system.
In order to compute the Melnikov function near the two
centers $(0,0)$ and $(1,0)$, we need
transform system (\ref{b55})$_{\varepsilon=0}$ to a Hamiltonian system.
The integrating factor $\,\gamma \,$ is given in (\ref{b14}).
Now, introducing $\, d t = \gamma\, d \tau\,$ into (\ref{b55}) yields
the perturbed Hamiltonian system:
\begin{equation}
\begin{array}{rl}
\displaystyle\frac{d x}{d \tau} = & \gamma\, (y + a_1\, x\, y )
+ \varepsilon \, \gamma \, a_{10}\, x , \\[1.5ex]
\displaystyle\frac{d y}{d \tau} = & \gamma \, (-\, x + x^2 + a_4\, y^2 )
+ \varepsilon \, \gamma \, ( b_{01}\, y + b_{11}\, x\, y ),
\end{array}
\label{b56}
\end{equation}
with the Hamiltonian of (\ref{b56})$_{\varepsilon=0}$, given by
\begin{equation}
H(x,y) = \displaystyle\frac{1}{2}\,
{\rm sign}(1+a_1 x) \, |1+a_1 x|^{- \frac{2 a_4}{a_1}}
\left[ y^2 + \displaystyle\frac{(1+a_1-a_4)\, (1 + 2\, a_4\, x)}{a_4\, (a_1 - a_4)\,
(a_1 - 2\, a_4)} - \displaystyle\frac{x^2}{a_1 - a_4} \right],
\label{b57}
\end{equation}
for $\, a_4 \ne 0, \ a_1 \ne a_4 , \ a_1 \ne 2\, a_4$.
The cases $\, a_4 = 0$, $a_1 = a_4 \,$ or $\, \ a_1 = 2\, a_4\, $
will not be considered in this paper.
\begin{figure}[!h]
\vspace{-0.60in}
\hspace{0.00in}
\resizebox{1.10\textwidth}{!}{\includegraphics{FIG2.eps}} \hspace{-0.3in}
\vspace{-6.20in}
\caption{A phase portrait of the reversible system (\ref{b7}) with two
centers for $a_1=-3,\ a_4 = -\, \frac{8}{3}$.}
\label{fig2}
\vspace{0.10in}
\end{figure}
Note that
\begin{equation}
\begin{array}{ll}
h_{00} =
H(0,0) = \displaystyle\frac{1+a_1-a_4}{2\, a_4\, (a_1-a_4)\,(a_1 - 2 a_4)},
& {\rm for} \ \ 1+a_1\,x > 0, \\[2.0ex]
h_{10} = H(1,0) = -\, \displaystyle\frac{(a_1+1)\,(a_4+1)}
{2\, a_4\, (a_1-a_4)\,(a_1 - 2 a_4)}\, (-1-a_1)^{- \frac{2 a_4}{a_1}},
\quad & {\rm for} \ \ 1 + a_1\,x < 0.
\end{array}
\label{b58}
\end{equation}
Since in this paper, we concentrate on the case that system
(\ref{b55})$_{\varepsilon=0}$ has only two centers, we assume
$\, a_1 < -\,1, \ a_4 < 0$. Thus,
$$
\lim_{x \rightarrow -\frac{1}{a_1}^-} H(x,y) = + \,\infty \quad {\rm and}
\quad \lim_{x \rightarrow -\frac{1}{a_1}^+} H(x,y) = -\,\infty.
$$
It is easy to see from system (\ref{b55}) that the trajectories of
(\ref{b55})$_{\varepsilon=0}$ rotate around the center $(0,0)$ in
the clock-wise direction, while rotate around the center $(1,0)$ in
the counter clock-wise direction, as shown in Fig.~\ref{fig2}.
Thus, the values of $h$ in $H(x,y) = h$ are taken from the two intervals:
$\, h \in (h_{00}, \infty) \,$ for $\, 1 + a_1\,x > 0$, and
$\, h \in (-\infty, h_{10}) \,$ for $\, 1 + a_1\,x < 0$.
It should be noted that $\, h_{00} \,$ is not necessarily larger than
$\, h_{10}$. The analyses on the two half-plane in the
$x$-$y$ plane (see Fig.~\ref{fig2}),
divided by the singular line $\, 1+a_1 \,x =0$, are independent.
Next, introduce \vspace{-0.08in}
\begin{equation}
L_{h}: H(x,y) = h \ \left\{
\begin{array}{ll}
h \in (h_{00}, \infty), & {\rm for} \ \ 1+a_1\,x > 0, \\[1.0ex]
h \in (-\infty, h_{10}), \quad & {\rm for} \ \ 1+a_1\,x < 0,
\end{array}
\right.
\label{b59}
\end{equation}
and define the Melnikov function: \vspace{-0.05in}
\begin{equation}
\begin{array}{ll}
M (h, a_{ij}, b_{ij}) =
\displaystyle\oint_{L_{h}} q(x,y, b_{ij}) \, dx - p(x,y, a_{ij}) \, dy,
\end{array}
\label{b60}
\end{equation}
where $\, p(x,y,a_{ij}) = \gamma\, a_{10} \, x \,$
and $\, q(x,y,b_{ij}) = \gamma \, ( b_{01} + b_{11}\, x)\,y $.
Using the results in~\cite{Han2006,Han2000,HC2000}, we can
expand $\, M \,$ near $\, h = h_{00} \,$ and $\, h= h_{10} \,$ as
\begin{equation}
\begin{array}{rl}
M_0(h, a_{ij}, b_{ij}) = & \mu_{00} \, ( h- h_{00})
+ \mu_{01} \, (h-h_{00})^2
+ \mu_{02} \, (h-h_{00})^3
\\ [1.5ex]
& +\, \mu_{03} \, (h-h_{00})^4
+ O((h-h_{00})^5), \quad {\rm for} \ \ 0 < h - h_{00} \ll 1,
\\[1.5ex]
M_1(h, a_{ij}, b_{ij}) = & \mu_{10} \, ( h_{10} -h)
+ \mu_{11} \, (h_{10}-h)^2
+ \mu_{12} \, (h_{10}-h)^3
\\ [1.5ex]
& +\, \mu_{13} \, (h_{10}-h)^4
+ O((h_{10}-h)^5), \quad {\rm for} \ \ 0 < h_{10}-h \ll 1,
\end{array}
\label{b61}
\end{equation}
where the coefficients $\, \mu_{ij}, \ i=0,\,1; \ j=0,\,1,\,2,\, \cdots \,$
can be obtained by using
the Maple programs developed in~\cite{HanYangYu2010} as follows:
\begin{eqnarray*}
\mu_{00} &\!\!\!=\!\!\!& 2\, \pi \, ( a_{10} + b_{01} ), \\[-0.0ex]
\mu_{01} &\!\!\!=\!\!\!&
\displaystyle\frac{\pi}{12} \Big[
( 10 - 13\,a_1 - 14 \, a_4 + 13 \, a_1^2 + 7 \, a_1 \, a_4
- 20 \, a_4^2)\, a_{10} \\ [-0.5ex]
&\!\!\! \!\!\!&
\quad \
+\, ( 10 - a_1 + 10 \, a_4 + a_1^2 - 5 \, a_1 \, a_4 + 4 \, a_4^2)\, b_{01}
+ 12\, (1+a_4)\, b_{11} \Big], \\[-0.0ex]
\mu_{02} &\!\!\!=\!\!\!&
\displaystyle\frac{\pi}{864} \Big[
(1540-980 a_1-280 a_4+861 a_1^2-1512 a_1 a_4-3948 a_4^2-626 a_1^3
+ 1566 a_1^2 a_4 \\ [-1.0ex]
&\!\!\! \!\!\!&
\qquad \ + 1620 a_1 a_4^2 \!-\! 4432 a_4^3 \!+\! 313 a_1^4
\!-\! 1018 a_1^3 a_4 \!-\! 279 a_1^2 a_4^2 \!+\! 3080 a_1 a_4^3
\!-\! 2096 a_4^4) \, a_{10} \qquad \\ [-0.0ex]
&\!\!\! \!\!\!&
\qquad +\, (1540+700 a_1+3080 a_4+21 a_1^2+168 a_1 a_4+2772 a_4^2-2 a_1^3
+ 126 a_1^2 a_4 \\ [-0.0ex]
&\!\!\! \!\!\!&
\qquad \quad \ -\, 828 a_1 a_4^2 +1424 a_4^3 +a_1^4 -58 a_1^3 a_4
+369 a_1^2 a_4^2 - 712 a_1 a_4^3 +400 a_4^4)\, b_{01} \\ [-0.0ex]
&\!\!\! \!\!\!&
\qquad \ + \, 24 \,b_{11} \, (1+a_4)\,
(70+35 a_1+70 a_4+a_1^2-17 a_1 a_4+52 a_4^2) \, b_{11} \Big], \\
\mu_{03} &\!\!\!=\!\!\!&
\displaystyle\frac{\pi}{622080} \Big[
(3403400 \!-\! 300300 a_1 \!+\! 3003000 a_4 \!+\! 690690 a_1^2
\!-\! 4984980 a_1 a_4 \!-\! 7327320 a_4^2 \qquad
\\ [-1.0ex]
&\!\!\! \!\!\!&
\qquad \qquad -\,500885 a_1^3 +3314850 a_1^2 a_4 -4430580 a_1 a_4^2
-17811640 a_4^3 +323121 a_1^4
\\[-0.0ex]
&\!\!\! \!\!\!&
\qquad \qquad
-\, 2444439 a_1^3 a_4 \!+\! 4201218 a_1^2 a_4^2 \!+\! 5794692 a_1 a_4^3
\!-\! 18033936 a_4^4 \!-\! 168603 a_1^5
\\ [-0.0ex]
&\!\!\! \!\!\!&
\qquad \qquad
+\, 1420500 a_1^4 a_4 -3253551 a_1^3 a_4^2 -1296282 a_1^2 a_4^3
+12107904 a_1 a_4^4
\\ [-0.0ex]
&\!\!\! \!\!\!&
\qquad \qquad
-\, 10462368 a_4^5 +56201 a_1^6 -520311 a_1^5 a_4 +1471287 a_1^4 a_4^2
-407053 a_1^3 a_4^3
\\ [-0.0ex]
&\!\!\! \!\!\!&
\qquad \qquad
-\, 4589772 a_1^2 a_4^4
+7149264 a_1 a_4^5
-3159616 a_4^6
)\, a_{10}
\\ [-0.0ex]
&\!\!\! \!\!\!&
\qquad \quad \ \
+\, (3403400 +3303300 a_1 +10210200 a_4 +690690 a_1^2 +5825820 a_1 a_4
\\ [-0.5ex]
&\!\!\! \!\!\!&
\qquad \qquad \ \ \,
+ 14294280 a_4^2 \!+\! 11935 a_1^3 \!+\! 404250 a_1^2 a_4
\!+\! 2721180 a_1 a_4^2 \!+\! 12236840 a_4^3
\\ [-0.0ex]
&\!\!\! \!\!\!&
\qquad \qquad \ \ \,
-699 a_1^4 -11379 a_1^3 a_4 +262458 a_1^2 a_4^2 -1891308 a_1 a_4^3
+6994704 a_4^4
\\ [-0.0ex]
&\!\!\! \!\!\!&
\qquad \qquad \ \ \,
+417 a_1^5
+1380 a_1^4 a_4 -149091 a_1^3 a_4^2 +1121838 a_1^2 a_4^3 -2964576 a_1 a_4^4
\\ [-0.0ex]
&\!\!\! \!\!\!&
\qquad \qquad \ \ \,
+2670432 a_4^5
-139 a_1^6
-291 a_1^5 a_4
+46227 a_1^4 a_4^2
-366193 a_1^3 a_4^3
\\ [-0.0ex]
&\!\!\! \!\!\!&
\qquad \qquad \ \ \,
+1076988 a_1^2 a_4^4
-1335216 a_1 a_4^5
+578624 a_4^6
)\, b_{01}
\\ [-0.0ex]
&\!\!\! \!\!\!&
\qquad \quad \ \
+\, ( 3603600 +3603600 a_1 +10810800 a_4 +790020 a_1^2 + 6597360 a_1 a_4
\\ [-0.0ex]
&\!\!\! \!\!\!&
\qquad \qquad \ \ \,
+15024240 a_4^2 \!+\! 12600 a_1^3 \!+\! 480060 a_1^2 a_4
\!+\! 3764880 a_1 a_4^2 \!+\! 12514320 a_4^3
\\ [-0.0ex]
&\!\!\! \!\!\!&
\qquad \qquad \ \ \,
+180 a_1^4 -10800 a_1^3 a_4 +11340 a_1^2 a_4^2 -618480 a_1 a_4^3
+6566400 a_4^4
\\ [-0.0ex]
&\!\!\! \!\!\!&
\qquad \qquad \ \ \,
+180 a_1^4 a_4
\!-\! 23400 a_1^3 a_4^2
\!+\! 321300 a_1^2 a_4^3
\!-\! 1389600 a_1 a_4^4
\!+\! 1869120 a_4^5
) \, b_{11}
\Big],
\\ [-0.5ex]
&\!\!\! \vdots \!\!\!&
\end{eqnarray*}
and
\begin{eqnarray*}
\mu_{10} &\!\!\!=\!\!\!& 2\, \pi \, (-1-a_1)^{3/2}
\Big[ (1-2\,a_4)\, a_{10} + (1+a_1)\, (b_{01} + b_{11} ),
\\[-0.0ex]
\mu_{11} &\!\!\!=\!\!\!&
\displaystyle\frac{\pi}{12} \ (-1-a_1)^{-\, \frac{2\,(a_1-a_4)}{a_1}} \\
&\!\!\! \!\!\!& \times
\Big[
(10 +33 a_1 -6 a_4 +36 a_1^2 -21 a_1 a_4 -24 a_1^2 a_4
+30 a_1 a_4^2 -8 a_4^3)\, a_{10}
\\ [-0.5ex]
&\!\!\! \!\!\!&
\quad \
+ \, (1+a_1)\, (10 +21 a_1 -10 a_4 + 12 a_1^2 -15 a_1 a_4 +4 a_4^2) \, b_{01}
\\ [-0.0ex]
&\!\!\! \!\!\!&
\quad \
- \, (1+a_1)\,(1+a_4)\, (2 +3\, a_1 -4 \,a_4) \, b_{11}
\Big],
\\[-0.0ex]
\mu_{12} &\!\!\!=\!\!\!&
\displaystyle\frac{\pi}{864} \ (-1-a_1)^{-\, \frac{(5\,a_1-8\,a_4)}{2\,a_1}} \\
&\!\!\! \!\!\!& \times
\Big[ (1540
+7140 a_1
-2800 a_4
+13041 a_1^2
-11592 a_1 a_4
+2212 a_4^2
+11448 a_1^3
\\ [-0.5ex]
&\!\!\! \!\!\!&
\quad \ \
-18072 a_1^2 a_4
\!+\!8628 a_1 a_4^2
\!-\! 1112 a_4^3
\!+\! 752 a_4^4
\!-\! 12024 a_1^3 a_4
\!+\! 12213 a_1^2 a_4^2
\!-\! 5232 a_1 a_4^3 \qquad
\\ [-0.0ex]
&\!\!\! \!\!\!&
\quad \ \
+4320 a_1^4
-1728 a_1^4 a_4
+6192 a_1^3 a_4^2
-7938 a_1^2 a_4^3
+4272 a_1 a_4^4
-800 a_4^5
)\, a_{10}
\\ [-0.0ex]
&\!\!\! \!\!\!&
\quad \
+ \, (1\!+\!a_1)\,( 1540
\!+\! 5460 a_1
\!-\! 3080 a_4
\!+\! 7161 a_1^2
\!-\! 9072 a_1 a_4
\!+\! +2772 a_4^2
\!+\! 4104 a_1^3
\\ [-0.0ex]
&\!\!\! \!\!\!&
\hspace{1.00in}
-\,9030 a_1^2 a_4
\!+\! 6372 a_1 a_4^2
\!-\! 1424 a_4^3
\!+\! 864 a_1^4
\!-\! 3096 a_1^3 a_4
\!+\! 3969 a_1^2 a_4^2
\\ [-0.0ex]
&\!\!\! \!\!\!&
\hspace{1.00in}
-\,2136 a_1 a_4^3
+400 a_4^4
)\, b_{01}
\\ [-0.0ex]
&\!\!\! \!\!\!&
\quad \
- \, (1\!+\!a_1) (1\!+\!a_4) \,
(140
+420 a_1
-420 a_4
+423 a_1^2
-996 a_1 a_4
+576 a_4^2
\\[-0.0ex]
&\!\!\! \!\!\!&
\hspace{1.45in}
+\,144 a_1^3
-633 a_1^2 a_4
+888 a_1 a_4^2
-400 a_4^3
)\, b_{11}
\Big],
\\[-0.0ex]
\mu_{13} &\!\!\!=\!\!\!&
\displaystyle\frac{\pi}{1244160} \ (-1-a_1)^{-\, \frac{(2\,(a_1-3\,a_4)}{a_1}} \\
&\!\!\! \!\!\!& \times
\Big[ (3403400
\!+\! 20720700 a_1
\!-\! 9809800 a_4
\!+\! 53243190 a_1^2
\!-\! 54234180 a_1 a_4
\!+\! 13093080 a_4^2
\\[-1.0ex]
&\!\!\! \!\!\!&
\quad \
+\,74334645 a_1^3
-123735150 a_1^2 a_4
+65571660 a_1 a_4^2
-10776920 a_4^3
+60023916 a_1^4
\\[-0.0ex]
&\!\!\! \!\!\!&
\quad \
-\,147900519 a_1^3 a_4
\!+\! 131934978 a_1^2 a_4^2
\!-\! 49682268 a_1 a_4^3
\!+\! 6439744 a_4^4
\!+\! 27002160 a_1^5
\\[-0.0ex]
&\!\!\! \!\!\!&
\quad \
-\,95460120 a_1^4 a_4
\!+\! 132380865 a_1^3 a_4^2
\!-\! 89408610 a_1^2 a_4^3
\!+\! 29027880 a_1 a_4^4
\!-\! 3527040 a_4^5
\\[-0.0ex]
&\!\!\! \!\!\!&
\quad \
+\,5443200 a_1^6
\!-\! 28946160 a_1^5 a_4
\!+\! 63998532 a_1^4 a_4^2
\!-\! 74879613 a_1^3 a_4^3
\!+\! 48498336 a_1^2 a_4^4
\\[-0.0ex]
&\!\!\! \!\!\!&
\quad \
-\,16296336 a_1 a_4^5
+2181248 a_4^6
-1555200 a_1^6 a_4
+9603360 a_1^5 a_4^2
-24061752 a_1^4 a_4^3
\\[-0.0ex]
&\!\!\! \!\!\!&
\quad \
+\,31232358 a_1^3 a_4^4
-22072536 a_1^2 a_4^5
+8011296 a_1 a_4^6
-1157248 a_4^7
) \, a_{10}
\\ [-0.0ex]
&\!\!\! \!\!\!&
\quad
+ \, (1\!+\!a_1)\,(3403400
\!+\! 17117100 a_1
\!-\! 10210200 a_4
\!+\! 35225190 a_1^2
\!-\! 45225180 a_1 a_4
\\ [-0.0ex]
&\!\!\! \!\!\!&
\hspace{0.90in}
+14294280 a_4^2
+37785825 a_1^3
-79202970 a_1^2 a_4
+54455940 a_1 a_4^2
\\ [-0.0ex]
&\!\!\! \!\!\!&
\hspace{0.90in}
-12236840 a_4^3
+22125636 a_1^4
-68371209 a_1^3 a_4
+77864598 a_1^2 a_4^2
\\ [-0.0ex]
&\!\!\! \!\!\!&
\hspace{0.90in}
-38601828 a_1 a_4^3
+6994704 a_4^4
+6629040 a_1^5
-28984608 a_1^4 a_4
\\ [-0.0ex]
&\!\!\! \!\!\!&
\hspace{0.90in}
+49687587 a_1^3 a_4^2
-41614974 a_1^2 a_4^3
+16953984 a_1 a_4^4
-2670432 a_4^5
\\ [-0.0ex]
&\!\!\! \!\!\!&
\hspace{0.90in}
+777600 a_1^6
-4801680 a_1^5 a_4
+12030876 a_1^4 a_4^2
-15616179 a_1^3 a_4^3
\\ [-0.0ex]
&\!\!\! \!\!\!&
\hspace{0.90in}
+11036268 a_1^2 a_4^4
-4005648 a_1 a_4^5
+578624 a_4^6
) \,b_{01}
\\ [-0.0ex]
&\!\!\! \!\!\!&
\quad
- \, (1\!+\!a_1)(1\!+\!a_4)\,(
200200
\!+\! 900900 a_1
\!-\! 800800 a_4
\!+\! 1600830 a_1^2
\!-\! 3132360 a_1 a_4
\\ [-0.0ex]
&\!\!\! \!\!\!&
\hspace{1.40in}
+1530760 a_4^2
+1397655 a_1^3
-4596480 a_1^2 a_4
+5008500 a_1 a_4^2
\\ [-0.0ex]
&\!\!\! \!\!\!&
\hspace{1.40in}
-1808240 a_4^3
+594864 a_1^4
-3001266 a_1^3 a_4
+5594022 a_1^2 a_4^2
\\ [-0.0ex]
&\!\!\! \!\!\!&
\hspace{1.40in}
-4568112 a_1 a_4^3
+1379936 a_4^4
+97200 a_1^5
-736776 a_1^4 a_4
\\ [-0.0ex]
&\!\!\! \!\!\!&
\hspace{1.40in}
+2162079 a_1^3 a_4^2
\!-\! 3080268 a_1^2 a_4^3
\!+\! 2136528 a_1 a_4^4
\!-\! 578624 a_4^5
)\, b_{11}
\Big],
\\[-0.5ex]
&\!\!\! \vdots \!\!\!&
\end{eqnarray*}
\noindent
{\bf Remark 2.2.} \ The coefficients $\, \mu_{0j}\,$ listed above
are applicable as long as $(0,0)$ is a center,
and the coefficients $\, \mu_{1j}\,$ are applicable
as long as $(1,0)$ is a center, regardless the number
and distribution of the system's singularities.
Therefore, for each point on the whole line $\, a_4 = \frac{1}{3}\,(a_1 - 5)$
(see Fig.~\ref{fig1}),
there always exist $3$ small limit cycles bifurcating
from the center $(0,0)$, no matter whether the system has two centers,
or one center and three saddle points, or one center and one saddle point.
For each point on the line segment $\, a_4 = \frac{1}{3}\,(6\,a_1 + 5)
\ (a_1 < -\,1)$, the system can have $3$ limit cycles bifurcating from
the center $(1,0)$. This indicates that the results given
in~\cite{DumortierLiZhang1997,Peng2002,Han1997} showing that the reversible
near-integrable systems with one center and one saddle point can have
maximal $2$ limit cycles is conservative, since on the part of
the line $\,a_4 = \frac{1}{3}\,(a_1 - 5) \,$ in the first quadrant
($a_1 > 5$) such a system can have at least $3$ limit cycles.
\vspace{0.10in}
First, we consider the maximal number of limit cycles which can
bifurcate from the center $(0,0)$. Setting $\, \mu_{00} = 0\,$ yields
\vspace{-0.10in}
\begin{equation}
b_{01} = -\, a_{10},
\label{b62}
\end{equation}
\vspace{-0.10in}
\noindent
and then we have \vspace{-0.05in}
\begin{equation}
\mu_{01} = \pi \, \Big[ (a_1-1-a_4)(a_1+2 a_4) a_{10} +(1+a_4)\, b_{11} \Big].
\label{b63}
\end{equation}
In order to have $\, \mu_{01} = 0$, we suppose $\, a_4 \ne -\,1 \,$ and
choose
\begin{equation}
b_{11} = -\, \displaystyle\frac{ (a_1-1-a_4)(a_1+2 a_4)} {1+a_4} \ a_{10} .
\label{b64}
\end{equation}
Then, $\, \mu_{02} \,$ and $\, \mu_{03} \,$ are simplified to
\begin{equation}
\begin{array}{ll}
\mu_{02} = \displaystyle\frac{\pi}{3}\, a_1\, (a_1-a_4)\, (a_1+ 2\,a_4)
(a_1 - 3 \,a_4 - 5) \, a_{10}, \\[1.5ex]
\mu_{03} = -\, \displaystyle\frac{\pi}{144}\, a_1 (a_1\!-\!a_4) (a_1\!+\!2 a_4)
( 770
+105 a_1
+1400 a_4
+42 a_1^2
-434 a_1 a_4
\\ [1.5ex]
\hspace{2.30in} +1274 a_4^2
-13 a_1^3
+128 a_1^2 a_4
-415 a_1 a_4^2
+444 a_4^3
)\,a_{10}.
\end{array}
\label{b65}
\end{equation}
There are five choices for $\, \mu_{02} = 0 $. Except the
choice $\, a_1 - 3\, a_4 - 5 = 0$, all other choices lead to
$\, \mu_{0i} = 0, \ i=3,\,4, \cdots $.
Thus, letting
\begin{equation}
a_4 = \displaystyle\frac{1}{3}\, ( a_1 - 5),
\label{b66}
\end{equation}
which implies $\, a_1 \ne 2 \,$ when $\, a_4 \ne -\,1$.
Since we assume $\, a_1 < -\,1$, for this case (i.e., when
the condition (\ref{b66}) holds), $\,a_4 \ne -\,1 \,$ is guaranteed.
Then, we have
\begin{eqnarray*}
\mu_{03}
&\!\!\!=\!\!\!&
- \displaystyle\frac{25 \pi}{162} \, a_1\, (a_1+1)\,(a_1-2)^2 \,
(2\,a_1 + 5)\, a_{10}, \\ [0.0ex]
\mu_{04}
&\!\!\!=\!\!\!&
- \displaystyle\frac{5 \pi}{8748}\, a_1\, (a_1+1)\,(a_1-2)^2 \,
(2\,a_1 + 5)\, (a_1+4)\, (17\,a_1+518) \, a_{10} \\[0.0ex]
&\!\!\! \vdots \!\!\!& \\ [0.0ex]
\mu_{10}
&\!\!\!=\!\!\!&
-\, \displaystyle\frac{10 \pi}{3} \, (-1-a_1)^{-3/2} \,
a_1 \, (2\,a_1 + 5) \, a_{10}, \\[0.0ex]
\mu_{11}
&\!\!\!=\!\!\!&
\displaystyle\frac{25 \pi}{324} \, (-1-a_1)^
{-\,\frac{2\,(2\,a_1+5)}{3\, a_1}} a_1 \, (a_1-2)^2 \,
(2\,a_1 + 5)\, a_{10}, \\[0.0ex]
&\!\!\! \vdots \!\!\!&
\end{eqnarray*}
implying that in addition we need \vspace{-0.10in}
\begin{equation}
(2\,a_1 + 5)\, a_{10} \ne 0.
\label{b67}
\end{equation}
Under the above conditions (\ref{b62}), (\ref{b64}), (\ref{b66}) and
(\ref{b67}), we obtain $\, \mu_{00} = \mu_{01} = \mu_{02} = 0$,
but $\, \mu_{03} \ne 0 , \ \mu_{10} \ne 0$. Hence, at most $3$ small
limit cycles can bifurcate from the center $(0.0)$ with no limit cycles
bifurcating from the center $(1,0)$. Further, giving proper perturbations
to the parameters $a_4 $ (or $a_1$), $b_{11}$ and $b_{01}$, we can obtain
$3$ small limit cycles bifurcating from the origin.
This shows that the conclusion is true for the case of
$(3,0)$-distribution.
Next, consider the $(0,3)$-distribution.
Similarly, letting $\, \mu_{10} = 0\,$ yields
\begin{equation}
b_{01}= -\, b_{11} + \displaystyle\frac{ 2\,a_4-1}{1+ a_1} \, a_{10}.
\label{b68}
\end{equation}
Then, $\, \mu_{11} \, $ becomes
\begin{equation}
\mu_{11} = \pi (-1 - a_1)^{-\, \frac{2\,(a_1-a_4)}{a_1}}
\Big[ ( a_1 + 2\, a_4) ( 2\,a_1 - a_4 + 1)\, a_{10}
- (1+a_1)^2 \, (a_1- a_4 + 1)\, b_{11} \Big].
\label{b69}
\end{equation}
Hence, we set
\begin{equation}
b_{11} = \displaystyle\frac{ ( a_1 + 2\, a_4) ( 2\,a_1 - a_4 + 1)}
{(1+a_1)^2 \, (a_1 - a_4 + 1)} \, a_{10}, \qquad
( a_1 - a_4 + 1 \ne 0 ),
\label{b70}
\end{equation}
to yield $\, \mu_{11} = 0$, and
\begin{equation}
\begin{array}{ll}
\mu_{12} = \displaystyle\frac{\pi}{3}\,
(-1-a_1)^{- \frac{5 a_1 - 8 a_4}{2\,a_1} }
\, a_1\, (a_1-a_4)\, (a_1+ 2\,a_4)
(6\, a_1 - 3 \,a_4 + 5) \, a_{10}, \\[1.5ex]
\mu_{13} = \displaystyle\frac{\pi}{288}\, (-1-a_1)^{- \frac{2 (a_1 - 3 a_4)}{a_1} }
a_1 (a_1\!-\!a_4) (a_1\!+\!2 a_4)
\,( 770
+2205 a_1
-1400 a_4
+2142 a_1^2
\\ [1.5ex]
\hspace{1.20in}
-\,3234 a_1 a_4
+1274 a_4^2
-720 a_1^3
-1962 a_1^2 a_4
+1689 a_1 a_4^2
-444 a_4^3
)\,a_{10}.
\end{array}
\label{b71}
\end{equation}
The only choice for $\, \mu_{12} = 0 \,$ is
$\, 6\, a_1 - 3\, a_4 + 5 = 0$, from which we have
\begin{equation}
a_4 = \displaystyle\frac{1}{3}\, ( 6\, a_1 + 5).
\label{b72}
\end{equation}
This implies that $\, a_1 - a_4 + 1 = -\,(a_1 + \frac{2}{3}) > 0 \,$
for $\, a_1 < -\,1$. Further, we obtain
\begin{eqnarray*}
\mu_{13} &\!\!\!=\!\!\!& -\, \displaystyle\frac{25 \pi}{324} \,
(-1-a_1)^{\frac{10 + 11\, a_1}{a_1} }
a_1\, (3\,a_1+2)^2 \, (3\,a_1 + 5)\, a_{10}, \\ [0.0ex]
\mu_{14} &\!\!\!=\!\!\!& - \displaystyle\frac{5 \pi}{17496}\,
(-1-a_1)^{\frac{80 + 87\, a_1}{6\, a_1} }
a_1\, (3\,a_1+2)^2\, (3\,a_1 + 5)\,
(3\, a_1+4) \, (501\,a_1 + 518)\, a_{10} \\[0.0ex]
&\!\!\! \vdots \!\!\!& \\ [0.0ex]
\mu_{00} &\!\!\!=\!\!\!& \displaystyle\frac{10 \pi}{3\, (1+a_1)^2} \,
a_1 \, (3\,a_1 + 5) \, a_{10}, \\[0.0ex]
\mu_{01} &\!\!\!=\!\!\!& -\, \displaystyle\frac{25 \pi}{324\, (1+a_1)^2}
\, a_1 \, (3\,a_1+5) \,
(3\,a_1 + 2)^2 \, a_{10}, \\[0.0ex]
&\!\!\! \vdots \!\!\!&
\end{eqnarray*}
implying that in addition we require \vspace{-0.05in}
\begin{equation}
(3\,a_1 + 5)\, a_{10} \ne 0.
\label{b73}
\end{equation}
Under the above conditions (\ref{b68}), (\ref{b70}), (\ref{b72}) and
(\ref{b73}), we have $\, \mu_{10} = \mu_{11} = \mu_{12} = 0$,
but $\, \mu_{13} \ne 0 , \ \mu_{00} \ne 0$. Further, by properly
perturbing the parameters $a_4 $ (or $a_1$), $b_{11}$ and $b_{01}$,
we can obtain $3$ small limit cycles bifurcating from the
center $(1,0)$, but no limit cycles from the origin.
This proves the case of $(0,3)$-distribution.
For the case of $(2,0)$-distribution,
it follows from the conditions (\ref{b62})
and (\ref{b64}), and $\, a_4 \ne -\,1\,$ that
$\, \mu_{00} = \mu_{01} =0$, and
$$
\begin{array}{ll}
\mu_{02} = \displaystyle\frac{\pi}{3}\, a_1\, (a_1-a_4)\, (a_1+ 2\,a_4)
(a_1 - 3 \,a_4 - 5) \, a_{10}, \\[1.5ex]
\mu_{10} = -\, \displaystyle\frac{2\,\pi}{(1+a_4)\, (-1-a_1)^{3/2}} \,
a_1 \, (a_1 - a_4)\, ( a_1 + 2\, a_4)\, a_{10}.
\end{array}
$$
Thus, $\, \mu_{02} \ne 0 \,$ implies $\, \mu_{10} \ne 0$,
indicating that the conclusion holds for the case of $(2,0)$-distribution.
if $\, a_4 = -\,1$.
When $\, a_4 = -\,1$, (\ref{b63}) becomes
$$
\mu_{01} = \pi \, a_1 \, ( a_1 - 2) \, a_{10} \ne 0 \qquad
{\rm for} \ \ a_1 < -\,1 \quad {\rm and} \quad a_{10} \ne 0.
$$
Under the conditions $\, b_{01} = -\, a_{10} \,$ and
$\, a_4 = -\,1$, $\, \mu_{10} \,$ and $\, \mu_{11} \,$ becomes
\begin{equation}
\begin{array}{ll}
\mu_{10} = - \, 2\, \pi\, (-1-a_1)^{-3/2} \, \Big[
(a_1 -2 )\, a_{10} - (1+a_1)\, b_{11} \Big], \\[1.5ex]
\mu_{11} = \pi \, (-1-a_1)^{-\, \frac{2+ a_1)}{a_1}} \,
a_1 \, (a_1 -2) \, a_{10} ,
\end{array}
\label{b74}
\end{equation}
which shows that $\, \mu_{11} \ne 0 \,$ for $\, a_1 < -\,1 \,$
and $\, a_{10} \ne 0$. But we can choose
$$
b_{11} = \displaystyle\frac{ a_1 - 2}{ 1+a_1} \, a_{10}
$$
to obtain $\, \mu_{10} =0$. Thus, for this case we have
a $\, (1,1)$-distribution.
Similarly, for the $(0,2)$-distribution, we use the conditions
(\ref{b68}) and (\ref{b70}) to obtain
$$
\begin{array}{ll}
\mu_{12} = \displaystyle\frac{\pi}{3}\,
(-1-a_1)^{- \frac{5 a_1 - 8 a_4}{2\,a_1} }
\, a_1\, (a_1-a_4)\, (a_1+ 2\,a_4)
(6\, a_1 - 3 \,a_4 + 5) \, a_{10}, \\[1.5ex]
\mu_{00} = \displaystyle\frac{ 2 \, \pi}{(1+a_1)^2 \, (a_1 - a_4 + 1)} \,
a_1\, (a_1 - a_4) \, ( a_1 + 2\, a_4) \, a_{10}.
\end{array}
$$
This indicates that $\, \mu_{12} \ne 0 \,$
implies $\, \mu_{00} \ne 0$, and so the conclusion for the
case of $(0,2)$-distribution is also true if $\, a_1 - a_4 + 1 \ne 0$.
When $\, a_1 - a_4 + 1 = 0$, i.e., $a_4 = a_1 + 1 < 0$, (\ref{b69})
is reduced to
$$
\mu_{11} = \pi (-1 - a_1)^{-\, \frac{2\,(a_1-a_4)}{a_1}}
\, a_1 ( 3\, a_1 + 2)\, a_{10} \ne 0 \quad {\rm for} \ \
a_1 < -\,1 \quad {\rm and} \quad a_{10} \ne 0 ,
$$
and $\, \mu_{00} \,$ and $\, \mu_{01} \,$ become
\begin{equation}
\begin{array}{ll}
\mu_{00} = \displaystyle\frac{2\,\pi}{1+a_1} \Big[
(a_1 + 2)\, a_{10} - (1+a_1)\, b_{11} \Big], \\[2.5ex]
\mu_{01} = -\, \displaystyle\frac{\pi}{1+a_1} \, a_1 \, (3\,a_1 + 2)\, a_{10},
\end{array}
\label{b75}
\end{equation}
which clearly shows that $\, \mu_{01} \ne 0 \,$ for
$\, a_1 < -\,1 \,$ and $\, a_{10} \ne 0$. However, we may choose
$$
b_{11} = \displaystyle\frac{ a_1 + 2}{ 1+a_1} \, a_{10}
$$
to obtain $\, \mu_{00} =0$. Thus, for $\, a_1 - a_4 +1 =0$,
we have a $(1,1)$-distribution.
Finally, suppose the condition given in (\ref{b62}) is satisfied, i.e.,
$\, b_{01} = -\, a_{10}$, then substituting this into $\, \mu_{10} \,$
to solve $\, b_{11} $ to obtain
\begin{equation}
b_{11} = \displaystyle\frac{a_1 + 2\, a_4}{1 + a_1}.
\label{b76}
\end{equation}
Then, under the conditions (\ref{b62}) and (\ref{b76}),
we obtain
\begin{equation}
\begin{array}{ll}
\mu_{01} = \displaystyle\frac{\pi}{1+a_1} \, a_1 \, (a_1 - a_4)\, ( a_1 + 2\, a_4) \,
a_{10}, \\[1.5ex]
\mu_{11} = -\, \pi \, (-1-a_1)^{-\, \frac{2 \, (a_1-a_4)}{a_1} }
\, a_1 \, (a_1 - a_4)\, ( a_1 + 2\, a_4) \,
a_{10},
\end{array}
\label{b77}
\end{equation}
which shows that $\, \mu_{01} \ne 0 \,$ implies
$\, \mu_{11} \ne 0$, and thus in general the conclusion is true for
the case of $(1,1)$-distribution.
As we have seen in the above analysis,
if the condition (\ref{b66}), $\, a_4 = \frac{1}{3}\,(a_1 - 5)$,
is not used, then we can only have $2$ limit cycles
bifurcating from the origin, but no limit cycles can occur
from the center $(1,\,0)$. In other words, we can obtain one more limit
cycle, by using the condition $\, a_4 = \frac{1}{3}\,(a_1 - 5)$,
only bifurcating from the center $ (0,0)$.
Similarly, if the condition (\ref{b72}), $a_4 =\frac{1}{3}\,(6 a_1 + 5)$,
is not used, then we can have only $2$ limit cycles
bifurcating from the center $(1,0)$, but no limit cycles can bifurcate
from the origin. Then, condition $a_4 =\frac{1}{3}\,(6 a_1 + 5)$ can be only
used to get one more limit cycle around the center $(1,0)$, rather than
the origin.
Therefore, $(2,1)$- or $(1,2)$-distribution is not possible.
This completes the proof of Theorem 2.1.
\put(10,0.5){\framebox(6,7.5)}
\section{Limit cycles bifurcating from closed orbits}
\setcounter{equation}{0}
\renewcommand{\theequation}{4.\arabic{equation}}
In this section, based on the results of the small limit cycles
obtained in the previous section,
we want to investigate the possibility of existence of large limit cycles
by applying the Melnikov function, defined in (\ref{b60}).
We have the following result.
\vspace{0.1in}
\noindent
{\bf Theorem 4.1.} \ {\it
For the case of bifurcation of small limit cycles from the two centers
$(0,0)$ and $(1,0)$ with $(3,0)$-distribution
(respectively, $(0,3)$-distribution) there exists at least one
large limit cycle near $L_h$ for some
$\, h \in (- \infty, h_{10})$ (respectively for
some $\, h \in (h_{00}, \infty)$).
For the case of limit cycles with $(2,0)$-distribution
(respectively, $(0,2)$-distribution) there exist at least two
large limit cycles, one near $L_{h_1}$ for some $\, h_1 \in (- \infty, h_{10})$
and one near $L_{h_2}$ for some $\, h_2 \in (h_{00}, \infty)$.
The corresponding values of the parameters $\,a_1 \,$ and $\, a_4\,$
for the existence of $4$ limit cycles
can appear at least in some regions in the $a_1$-$a_4$ parameter plane.
}
\vspace{0.1in}
\noindent
{\bf Remark 4.2.} \ Theorem 4.1 gives a positive answer to the open question
of existence of limit cycles in near-integrable quadratic systems: at least
$4$ limit cycles can exist. For the case of $(1,1)$-distribution, so far no more
large limit cycles have been found.
\vspace{0.1in}
\noindent
{\bf Proof.} \ It follows from (\ref{b60}) with
$$
p(x,y,a_{ij}) = |1+a_1\,x|^{- \frac{a_1+2 a_4}{a_1}}
\, a_{10} \, x, \quad
q(x,y,b_{ij}) = |1+a_1\,x|^{- \frac{a_1+2 a_4}{a_1}}
\, ( b_{01} + b_{11}\, x)\, y
$$
that
\begin{eqnarray}
&\!\!\! \!\!\!& M (h,a_1,a_4,a_{10},b_{01},b_{11})
\nonumber \\
&\!\!\!=\!\!\!&
\displaystyle\oint_{L_{h}} q(x,y, b_{ij}) \, dx - p(x,y, a_{ij}) \, dy
\nonumber \\
&\!\!\!=\!\!\!&
\displaystyle\oint_{L_{h}} q(x,y, b_{ij}) \, dx
- \displaystyle\oint_{L_{h}} p(x,y, a_{ij}) \, dy
\nonumber \\
&\!\!\!=\!\!\!&
\displaystyle\oint_{L_{h}} q(x,y, b_{ij}) \, dx
+ \displaystyle\oint_{L_{h}} y\, p_x (x,y, a_{ij}) \, dx \qquad
\nonumber \\
&\!\!\!=\!\!\!&
\displaystyle\oint_{L_{h}} \Big[ q(x,y, b_{ij})
+ y\, p_x (x,y, a_{ij}) \Big] \, dx
\nonumber \\
&\!\!\!=\!\!\!&
\displaystyle\oint_{L_{h}} \left[ |1+a_1\,x|^{- \frac{a_1+2 a_4}{a_1}}\,
( b_{01} + b_{11}\,x)
+ |1+a_1\,x|^{- \frac{a_1+2 a_4}{a_1}}\,
a_{10} \left( 1 \mp \displaystyle\frac{(a_1+2\,a_4)\,x}{ |1+a_1\,x|} \right)
\right] y \, dx \qquad \quad
\nonumber \\
&\!\!\!=\!\!\!&
\displaystyle\oint_{L_{h}} |1+a_1\,x|^{- \frac{a_1+2 a_4}{a_1}}
\left[ ( a_{10} + b_{01} )
+ b_{11}\, x - {\rm sign}(1 + a_1 \,x) \,
a_{10}\, (a_1 + 2\, a_4) \, \displaystyle\frac{x}{|1+a_1\,x|}
\right] y \, dx
\nonumber \\ [1.0ex]
&\!\!\!=\!\!\!&
( a_{10} + b_{01} ) \, I_0(h,a_1,a_4)
+ b_{11} \, I_1(h,a_1,a_4)
+ a_{10}\, I_2(h,a_1,a_4)
\nonumber \\ [1.5ex]
&\!\!\! \equiv \!\!\!&
\left\{ \begin{array}{ll}
M_0(h,a_1,a_4,a_{10}, b_{01}, b_{11}) \quad
{\rm for} \ \ h\in (h_{00},\infty), & {\rm when} \ \
1 + a_1\,x > 0, \\[1.5ex]
M_1(h,a_1,a_4,a_{10}, b_{01}, b_{11}) \quad
{\rm for} \ \ h\in (-\infty, h_{10}), \ \ & {\rm when} \ \
1 + a_1\,x < 0,
\end{array}
\right.
\label{b78}
\end{eqnarray}
where
\begin{eqnarray*}
I_0(h,a_1,a_4) &\!\!\!\!=\!\!\!\!& \displaystyle\oint_{L_h}
|1+a_1\,x|^{- \frac{a_1+2 a_4}{a_1}}\, y \, dx \\[1.0ex]
&\!\!\!\!=\!\!\!\!& \left\{ \begin{array}{lll}
\hspace{0.05in}
2 \displaystyle\int_{x_{\min}}^{x_{\max}}
(1+a_1\,x)^{- \frac{a_1+2 a_4}{a_1}}\, y_+\, dx,
& \forall \ h \in (h_{00},\infty), \!\!& {\rm when} \ \ 1 + a_1 x > 0, \\[2.5ex]
\!\!\! -2 \displaystyle\int_{x_{\min}}^{x_{\max}}
(-1 \!-\! a_1 x)^{- \frac{a_1+2 a_4}{a_1}}\, y_+\, dx,
& \forall \ h \in (-\infty, h_{10}), \!\!&{\rm when} \ \ 1 + a_1 x < 0;
\end{array}
\right.
\\[1.0ex]
I_1(h,a_1,a_4) &\!\!\!\!=\!\!\!\!& \displaystyle\oint_{L_h}
|1+a_1\,x|^{- \frac{a_1+2 a_4}{a_1}}\, x\, y \, dx \\[0.0ex]
&\!\!\!\!=\!\!\!\!& \left\{ \begin{array}{lll}
\hspace{0.05in}
2 \displaystyle\int_{x_{\min}}^{x_{\max}}
(1+a_1\,x)^{- \frac{a_1+2 a_4}{a_1}} x\,y_+ dx,
& \forall \ h \in (h_{00},\infty), \!\!& {\rm when} \ \ 1 \!+\! a_1 x > 0,
\\[2.5ex]
\!\!\! -2 \displaystyle\int_{x_{\min}}^{x_{\max}}
(-1 \!-\! a_1 x)^{- \frac{a_1+2 a_4}{a_1}} x\,y_+ dx,
& \forall \ h \in (-\infty, h_{10}),\!\!& {\rm when} \ \ 1 \!+\! a_1 x < 0;
\end{array}
\right.
\qquad \qquad
\\[1.0ex]
I_2(h,a_1,a_4) &\!\!\!\!=\!\!\!\!& -\, (a_1 + 2\, a_4) \,
{\rm sign}(1+a_1 x) \displaystyle\oint_{L_h}
|1+a_1\,x|^{- \frac{2 (a_1+ a_4)}{a_1}}\, x\, y \, dx \\[1.0ex]
&\!\!\!\!=\!\!\!\!& \left\{ \begin{array}{lll}
\!\!\!
-2 (a_1 \!+\! 2 a_4) \! \displaystyle\int_{x_{\min}}^{x_{\max}}
\! (1\!+\!a_1 x)^{- \frac{2 (a_1+a_4)}{a_1}} x y_+ dx,
& \hspace{-0.05in} \forall \, h \! \in \! (h_{00},\infty), \!\!\!
& \hspace{-0.1in} {\rm when} \ 1 \!+\! a_1 x \!>\! 0,
\\[2.5ex]
\!\!\! 2 (a_1 \!+\! 2 a_4) \!\! \displaystyle\int_{x_{\min}}^{x_{\max}}
\!\! (-1 \!-\! a_1 x)^{- \frac{2(a_1+a_4)}{a_1}} x y_+ dx,
& \!\!\!\! \forall \, h \! \in \! (-\infty, h_{10}),\!\!\!
& {\rm when} \ 1 \!+\! a_1 x \!<\! 0.
\end{array}
\right.
\end{eqnarray*}
Here,
\begin{equation}
y_+ = \left[ \displaystyle\frac{x^2}{a_1 - a_4}
- \displaystyle\frac{(1+a_1-a_4)\,(1+2\,a_4\,x)}{a_4\, (a_1-a_4)\,(a_1 - 2\,a_4)}
+ 2\, h\, {\rm sign}(1 + a_1 \, x) |1+a_1\,x|^{ \frac{2 \, a_4}{a_1}}
\right]^{1/2},
\label{b79}
\end{equation}
and $\, x_{\min} \,$ and $\, x_{\max} \,$ are solved from the
equation, $\, y_+ = 0$,
for $\, h \in (h_{00}, \infty) \,$ when $\, 1 + a_1 \,x > 0$,
and for $\, h \in (-\infty, h_{10})\,$ when $\, 1 + a_1 \,x < 0$.
Since one can not find the closed form of the integrals
$\, I_i(h,a_1,a_4), \ i =0,\,1,\,2$, for
general $\, a_1 \,$ and $\, a_4$, nor the technique of Picard-Fuchs equation
can be applied here, we shall choose some values for
$\, a_1 \,$ and $\, a_4\,$ and then find numerical values of the
integral. We first use the results given in the previous section to determine
$\, b_{01} $, $\, b_{11}$, and $\, a_4$, and then choose proper values for
$\, a_1 \,$ to find more limit cycles.
\begin{figure}[!h]
\vspace{0.05in}
\begin{center}
\hspace{0.00in}
\resizebox{0.70\textwidth}{!}{\includegraphics{fig3a.eps}}
\vspace{-2.80in}
\hspace{-1.20in}
\resizebox{0.33\textwidth}{!}{\includegraphics{fig3aa.eps}}
\vspace{1.20in}
\hspace{0.25in}
$h$
\vspace{-2.0in}
\hspace{-4.6in}
$M_{00}$
\vspace{0.0in}
\hspace{-2.50in}
$h_{00}$
\vspace{1.8in}
\resizebox{0.70\textwidth}{!}{\includegraphics{fig3b.eps}}
\vspace{-0.10in}
\hspace{0.15in}
$h$
\vspace{-1.8in}
\hspace{-4.6in}
$M_{10}$
\vspace{-0.27in}
\hspace{1.05in} $h_1^*$ \hspace{2.35in} $h_{10}$
\vspace{-4.95in} \hspace{3.20in} (a)
\vspace{ 3.40in} \hspace{3.20in} (b)
\vspace{2.8in}
\caption{Functions $M_{00}(h) $ and $M_{10}(h) $
under the conditions
$ \mu_{00} = \mu_{01} = \mu_{02} =0 $, $\, \mu_{03} \ne 0$
and $ \mu_{10} \ne 0$,
for $ a_1 = -\,\frac{30}{7}$ and $ a_4 = \frac{1}{3}(a_1-5)
=- \frac{65}{21}$: (a) $\, M_{00}(h) > 0 $
for $h \in [h_{00}, \, +\infty)$, with
$\, h_0 = -\,\frac{441}{32500} \approx -\,0.01357$;
and (b) $M_{10}(h) $ for $h \in (-\infty, \, h_1]$,
with $\, h_{10} = -\, \frac{33957}{747500}(\frac{23}{7})^{5/9}
\approx -\,0.08797$,
crossing the $h$-axis at $h=h_1^* \in (-\,0.9250363254,\,-\,0.9250363253)$.}
\label{fig3}
\end{center}
\vspace{0.30in}
\end{figure}
\begin{figure}[!h]
\vspace{0.00in}
\begin{center}
\hspace{0.00in}
\resizebox{0.50\textwidth}{!}{\includegraphics{fig4a.eps}}
\vspace{-1.15in}
\resizebox{0.50\textwidth}{!}{\includegraphics{fig4b.eps}}
\vspace{-1.55in}
\hspace{1.93in}${\small \frac{7}{30}}$
\vspace{-6.40in} \hspace{3.70in} (a)
\vspace{ 3.30in} \hspace{3.70in} (b)
\vspace{2.95in}
\caption{Illustration of the existence of $4$ limit cycles
when $\, a_1 = -\,\frac{30}{7}$,
$\, a_4 = \frac{1}{3}(a_1-5) =- \frac{65}{21} - \varepsilon_1$,
and $a_{10}= \frac{1}{2000}$,
$ b_{11} = \frac{230}{21}\,a_{10} - \varepsilon_2$,
$b_{01}=-\,a_{10} - \varepsilon_3$,
where $\, 0 < \varepsilon_3 \ll
\varepsilon_2 \ll \varepsilon_1 \ll \varepsilon $:
(a) An unstable large limit cycle enclosing the center $(1,0)$;
and (b) Zoomed area around the center $(0,0)$ showing the existence of
$3$ small limit cycles.}
\label{fig4}
\end{center}
\vspace{0.20in}
\end{figure}
\vspace{0.05in}
(A) First, consider the $(3,0)$-distribution. For this case, we have
$$
\begin{array}{ll}
a_4 = \displaystyle\frac{1}{3}\, (a_1 - 5), \quad b_{01} = -\,a_{10}, \quad
b_{11} = - \,10\,(1+a_1)\, a_{10}.
\end{array}
$$
Taking $\, a_1 = -\, \frac{30}{7}\,$ yields
$\, a_4 = -\, \frac{65}{21}$, which denotes a point (a blank circle)
on the line $\, a_4 = \frac{1}{3}\, (a_1 - 5)\,$ in the
$a_1$-$a_4$ parameter plane (see Fig.~\ref{fig1}).
Further, we have $\, \ b_{11} = \frac{230}{7}\, a_{10}$, and
$$
\gamma = \Big(1- \frac{30}{7}\,x \Big)^{-\,\frac{22}{9}}
\quad (x \ne \frac{7}{30}).
$$
Then, the Hamiltonian (\ref{b57}) becomes
$$
H(x,y) = \displaystyle\frac{16807\, (16250\,y^2 + 13650\,x^2 +2730\,x -441)}
{32500\,(7-30\,x) \, ( 40353607-172944030\,x)^{4/9}} \quad
{\rm for} \quad x \ne \displaystyle\frac{7}{30},
$$
with
$$
h_{00} = -\,\displaystyle\frac{441}{32500} >
h_{10} = -\,\displaystyle\frac{33957}{747500} \Big(\displaystyle\frac{23}{7} \Big)^{5/9}.
$$
The Melnikov functions $\,M_i(h,a_{10})\,$
can be expressed as
\begin{equation}
M_i(h,a_{10}) = M_{i0}(h)\, a_{10}, \qquad i=0,\,1.
\label{b80}
\end{equation}
Without loss of generality, we may assume
\vspace{-0.05in}
\begin{equation}
a_{10} > 0,
\label{b81}
\end{equation}
\vspace{-0.05in}
\noindent
and thus $\, M_i(h,a_{10}) \,$ and $\, M_{i0}(h)\,$ have the same sign.
It is noted that for the above chosen parameter values, we have
$$
\mu_{03} = \displaystyle\frac{139150000\, \pi}{453789}\, a_{10} > 0 \quad {\rm and} \quad
\mu_{10} = - \displaystyle\frac{2500 \sqrt{161}\, \pi}{3703} \, a_{10} < 0.
$$
The computation results of $\, M_{00}(h) \,$ for $\, h \in (h_{00}, \infty) \,$
and $\, M_{10}(h) \,$ for $\, h \in (-\infty, h_{10}) \,$ are shown,
respectively, in Figs.~\ref{fig3}(a) and \ref{fig3}(b).
Figure~\ref{fig3}(a) shows that $\, M_{00}(h) > 0 \,$ for
$\, h \in (h_{00}, \infty)$, and its sign agrees with that of
$\, \mu_{03} > 0\,$ for $\, 0 < h-h_{00} \ll 1$, as expected.
It is also noted, as shown in Fig.~\ref{fig3}(b), that
the sign of $\, M_{10}(h)\,$
agrees with that of $\, \mu_{10} < 0 \,$ for
$\, 0 < h_{10} - h \ll 1$. However, unlike the interval
$\, h \in (h_{00}, \infty)$, this interval contains a critical value
$\, h = h_1^* \in (-\,0.9250363254,\,-\,0.9250363253) \,$ at which
$\, M_{10}(h_1^*) = 0 \,$ and the function $\, M_{10}(h) \,$ changes
its sign as $\,h\,$ crosses this critical point.
Thus, for this case, besides the $3$ small limit cycles, there exists
at least one large limit cycle bifurcating from the closed orbit $L_{h_1^*}$
of (\ref{b59}). This large limit cycle is shown in Fig.~\ref{fig4}(a),
which encloses the center $(1,0)$;
and Fig.~\ref{fig4}(b) illustrates the existence of $3$ small limit
cycles around the center $(0,0)$.
\begin{figure}[!h]
\vspace{0.05in}
\begin{center}
\hspace{0.00in}
\resizebox{0.70\textwidth}{!}{\includegraphics{fig5a.eps}}
\vspace{-0.10in}
\hspace{0.15in}
$h$
\vspace{-2.3in}
\hspace{-4.6in}
$M_{00}$
\vspace{-0.30in}
\hspace{0.10in} $h_{00}$ \hspace{2.7in} $h_2^*$
\vspace{2.4in}
\resizebox{0.70\textwidth}{!}{\includegraphics{fig5b.eps}}
\vspace{-2.70in}
\hspace{ 1.30in}
\resizebox{0.33\textwidth}{!}{\includegraphics{fig5bb.eps}}
\vspace{-0.40in}
\hspace{2.95in}
$h_{10}$
\vspace{1.25in}
\hspace{0.08in}
$h$
\vspace{-2.1in}
\hspace{-4.6in}
$M_{10}$
\vspace{-4.75in} \hspace{3.20in} (a)
\vspace{ 3.40in} \hspace{3.20in} (b)
\vspace{2.8in}
\caption{Functions $M_{00}(h) $ and $M_{10}(h) $
under the conditions
$ \mu_{10} = \mu_{11} = \mu_{12} =0 $, $ \mu_{13} \ne 0$
and $\, \mu_{00} \ne 0$,
for $ a_1 = -\,\frac{70}{51}$ and $ a_4 = \frac{1}{3}(6a_1+5)
=- \frac{55}{51}$: (a) $M_{00}(h) $ for $h \in [h_0,\, +\infty)$,
with $\, h_{00} = \frac{7803}{5500} \approx 1.41873$,
crossing the $h$-axis at $h=h_2^* \in (13.3847179116,\,13.3847179117)$;
and (b) $\, M_{10}(h) > 0 \,$
for $h \in (-\infty,\,h_1]$, with
$\, h_{10} = -\,\frac{44217}{104500} ( \frac{19}{51})^{3/7}
\approx -\,0.27714$.}
\label{fig5}
\end{center}
\vspace{0.30in}
\end{figure}
\begin{figure}[!h]
\vspace{0.00in}
\begin{center}
\hspace{0.00in}
\resizebox{0.50\textwidth}{!}{\includegraphics{fig6a.eps}}
\vspace{-1.15in}
\resizebox{0.50\textwidth}{!}{\includegraphics{fig6b.eps}}
\vspace{-1.55in}
\hspace{-1.43in}$\frac{51}{70}$
\vspace{-6.50in} \hspace{3.70in} (a)
\vspace{ 3.30in} \hspace{3.70in} (b)
\vspace{2.95in}
\caption{Illustration of the existence of $4$ limit cycles
when $\, a_1 = -\,\frac{70}{51}$,
$\, a_4 = \frac{1}{3}(6 a_1+5) =- \frac{55}{51} - \varepsilon_1$,
and $a_{10}= 10$,
$ b_{11} = \frac{8670}{361}\,a_{10} - \varepsilon_2$,
$b_{01}=-\,\frac{5611}{361}\,a_{10} + \varepsilon_3$,
where $\, 0 < \varepsilon_3 \ll
\varepsilon_2 \ll \varepsilon_1 \ll \varepsilon $:
(a) An unstable large limit cycle enclosing the center $(0,0)$;
and (b) Zoomed area around the center $(1,0)$ showing the
existence of $3$ small limit cycles.}
\label{fig6}
\end{center}
\vspace{0.20in}
\end{figure}
\vspace{0.05in}
(B) For the case of the $(0,3)$-distribution, we have
$$
\begin{array}{ll}
a_4 = \displaystyle\frac{1}{3}\, (6\,a_1 + 5), \quad
b_{01} = -\, b_{11} + \displaystyle\frac{2\,a_4-1}{1+a_1} \,a_{10}, \quad
b_{11} = \displaystyle\frac{ ( a_1 + 2\, a_4) ( 2\,a_1 - a_4 + 1)}
{(1+a_1)^2 \, (a_1 - a_4 + 1)} \, a_{10}.
\end{array}
$$
By choosing $\, a_1 = -\, \frac{70}{51}$, we have
$\, a_4 = -\, \frac{55}{51}, \ b_{01} = -\, \frac{5611}{361}\, a_{10} \,$ and
$\, b_{11} = \frac{8670}{361}\, a_{10}$.
The point $(a_1, a_4) = (-\, \frac{70}{51}, -\, \frac{55}{51}) \,$
is marked by a blank circle on the line $\, a_4 = \frac{1}{3}\,
(6\, a_1 + 5) \,$ in the $a_1$-$a_4$ parameter plane
(see Fig.~\ref{fig1}). Moreover,
$$
\gamma = \Big(1- \frac{70}{51}\,x \Big)^{-\frac{18}{7}}
\quad (x \ne \frac{51}{70}),
$$
and the Hamiltonian is
$$
H(x,y) = \displaystyle\frac{345025251\, (2750\,y^2+9350\,x^2-16830\,x+7803)}
{5500\,(51-70\,x) \, ( 897410677851- 1231740146070\,x)^{4/7}} \quad
{\rm for} \quad x \ne \displaystyle\frac{51}{70},
$$
with
$$
h_{00} = \displaystyle\frac{7803}{5500} >
h_{10} = -\,\displaystyle\frac{44217}{104500} \Big(\displaystyle\frac{19}{51} \Big)^{3/7}.
$$
For this case, $\mu_{00} \,$ and $\, \mu_{13} \,$ become
$$
\mu_{00} = -\, \displaystyle\frac{10500\, \pi}{361}\, a_{10} < 0 \quad {\rm and} \quad
\mu_{13} = \displaystyle\frac{4561235000 }{565036352721} \,
\Big( \displaystyle\frac{51}{19}\Big)^{2/7} \, \pi \, a_{10} > 0.
$$
The computation results of $\, M_{00}(h) \,$ for $\, h \in (h_{00}, \infty) \,$
and $\, M_{10}(h) \,$ for $\, h \in (-\infty, h_{10}) \,$ are shown
in Figs.~\ref{fig5}(a) and \ref{fig5}(b), respectively.
As shown in Fig.~\ref{fig5}(a), the sign of $\, M_{00}(h)\,$
agrees with that of $\, \mu_{00} < 0 \,$ for
$\, 0 < h - h_{00} \ll 1$, and in addition the function $\, M_{00}(h)\,$
crosses a critical value at $\, h=h_2^* \in (13.3847179116,\,13.3847179117)$,
at which it changes sign.
Figure~\ref{fig5}(b) shows that $\, M_{10}(h) > 0 \,$ for
$\, h \in (-\infty, h_{10})$, and its sign agrees with that of
$\, \mu_{13} > 0\,$ for $\, 0 < h_{10}-h \ll 1$.
Hence, for this case, in addition to the $3$ small limit cycles,
there also exists
at least one large limit cycle bifurcating from the closed orbit $L_{h_2^*}$ of
(\ref{b59}). This large limit cycle is depicted in Fig.~\ref{fig6}(a),
which encloses the center $(0,0)$;
and Fig.~\ref{fig6}(b) illustrates the existence of $3$ small limit
cycles around the center $(1,0)$.
\begin{figure}[!h]
\vspace{0.00in}
\begin{center}
\hspace{0.00in}
\resizebox{0.70\textwidth}{!}{\includegraphics{fig7a.eps}}
\vspace{-2.80in}
\hspace{-0.70in}
\resizebox{0.33\textwidth}{!}{\includegraphics{fig7aa.eps}}
\vspace{1.15in}
\hspace{0.08in}
$h$
\vspace{-1.95in}
\hspace{-4.5in}
$M_{00}$
\vspace{-0.52in}
\hspace{-0.78in} $h_{00}$ \hspace{1.00in} $h_3^*$
\vspace{2.3in}
\resizebox{0.70\textwidth}{!}{\includegraphics{fig7b.eps}}
\vspace{-0.10in}
\hspace{0.15in}
$h$
\vspace{-2.0in}
\hspace{-4.6in}
$M_{10}$
\vspace{-0.28in}
\hspace{0.90in} $h_4^*$ \hspace{1.9in} $h_{10}$
\vspace{-4.75in} \hspace{3.20in} (a)
\vspace{ 3.40in} \hspace{3.20in} (b)
\vspace{2.8in}
\caption{Functions $M_{00}(h) $ and $M_{10}(h) $
under the conditions
$\, \mu_{00} = \mu_{01} = $, $\, \mu_{02} \ne 0$
and $\, \mu_{10} \ne 0$,
for $\, a_1 = -\,4$ and $\, a_4 = - \frac{18}{5}$:
(a) $\, M_{00}(h) \,$
for $h \in [h_{00}, \, +\infty)$, with
$\, h_0 = \frac{25}{384} \approx 0.06510$,
crossing the $h$-axis at $h=h_3^* \in (0.1448192224,\,0.1448192225)$;
and (b) $M_{10}(h) $ for $h \in (-\infty, \, h_1]$,
with $\, h_{10} = -\, \frac{325}{3456}3^{1/5} \approx -\,0.11715$,
crossing the $h$-axis at $h=h_4^* \in (-\,0.5822537644,\,-\,0.5822537643)$.}
\label{fig7}
\end{center}
\vspace{0.30in}
\end{figure}
\begin{figure}[!h]
\vspace{0.00in}
\begin{center}
\hspace{0.00in}
\resizebox{0.50\textwidth}{!}{\includegraphics{fig8.eps}}
\vspace{-1.15in}
\resizebox{0.50\textwidth}{!}{\includegraphics{fig8a.eps}}
\vspace{-1.55in}
\hspace{ 3.58in}${\small \frac{1}{4}}$
\vspace{-6.45in} \hspace{-3.10in} (a)
\vspace{ 3.30in} \hspace{-3.10in} (b)
\vspace{2.95in}
\caption{Illustration of the existence of $4$ limit cycles
when $\, a_1 = -\,4$,
$\, a_4 = -\,\frac{18}{5}$,
and $a_{10}= \frac{1}{100}$,
$\, b_{11} = \frac{392}{65}\, a_{10} - \varepsilon_1$, and
$\, b_{01} = -\, a_{10} - \varepsilon_2$,
where $\, 0 < \varepsilon_2 \ll \varepsilon_1 \ll \varepsilon $:
(a) An unstable large limit cycle enclosing the center $(1,0)$;
and (b) Zoomed area around the center $(0,0)$ showing the
existence of $1$ large limit cycle and $2$ small limit cycles.}
\label{fig8}
\end{center}
\vspace{0.20in}
\end{figure}
\vspace{0.05in}
(C) Now consider the $(2,0)$-distribution. For this case, the condition
$\, a_4 = \frac{1}{3}\, (a_1 - 5) \,$ is not used. We need to
determine the values for both $\,a_1 \,$ and $\,a_4$.
We choose
$$
a_1 = -\, 4, \quad a_4 = -\, \displaystyle\frac{18}{5},
$$
which represents a point in the third quadrant of the
$a_1$-$a_4$ parameter plane (see the dark circle in
Fig.~\ref{fig1} near the line $a_4 = \frac{1}{3} (a_1-5)$). Thus,
$$
\gamma = \Big(1- 4\,x \Big)^{-\frac{14}{5}}
\quad (x \ne \frac{1}{4}).
$$
In addition, we have $\,
b_{01} = -\, a_{10}, \
b_{11} = \frac{392}{65}\, a_{10}$,
and
$$
H(x,y) = \displaystyle\frac{192\,y^2+ 480\,x^2-180\,x+25}
{384\,(1-4\,x)^{9/5}} \quad
{\rm for} \quad x \ne \displaystyle\frac{1}{4},
$$
with
$$
h_{00} = \displaystyle\frac{25}{384} >
h_{10} = -\,\displaystyle\frac{325}{3456} \ 3^{1/5}.
$$
For this case, $\mu_{02} \,$ and $\, \mu_{10} \,$ are reduced to
$$
\mu_{02} = -\, \displaystyle\frac{1344}{125} \, \pi\, a_{10} < 0 \quad {\rm and} \quad
\mu_{10} = -\, \displaystyle\frac{40 \sqrt{3}}{9} \, \pi \, a_{10} < 0.
$$
The computation results of $\, M_{00}(h) \,$ for $\, h \in (h_{00}, \infty) \,$
and $\, M_{10}(h) \,$ for $\, h \in (-\infty, h_{10}) \,$ are shown,
respectively, in Figs.~\ref{fig7}(a) and \ref{fig7}(b).
As shown in Fig.~\ref{fig7}(a), the sign of $\, M_{00}(h)\,$
agrees with that of $\, \mu_{02} < 0 \,$ for
$\, 0 < h - h_{00} \ll 1$. Moreover, the function $\, M_{00}(h)\,$
crosses a critical value at $\, h=h_3^* \in (0.1448192224,\,0.1448192225)\,$
at which it changes sign.
Figure~\ref{fig7}(b) shows $\, M_{10}(h) \,$ for
$\, h \in (-\infty, h_{10})$, whose sign agrees with that of
$\, \mu_{10} < 0\,$ for $\, 0 < h_{10}-h \ll 1$. Also,
$\, M_{10}(h) \,$ crosses a critical value at $\, h= h_4^* \in
(-\,0.5822537644,\,-\,0.5822537643)\,$
at which it changes sign.
Therefore, for this case, besides the two small limit cycles,
there exist at least two large limit cycles bifurcating from the two
different closed orbits $L_{h_3^*}$ and $L_{h_4^*}$ of (\ref{b59}).
One large limit cycle surrounding the center $(1,0)$
is shown in Fig.~\ref{fig8}(a), while another large limit cycle
enclosing the center $(0,0)$ with $2$ small limit
cycles is depicted in Fig.~\ref{fig8}(b).
\begin{figure}[!h]
\vspace{0.00in}
\begin{center}
\hspace{0.00in}
\resizebox{0.70\textwidth}{!}{\includegraphics{fig9a.eps}}
\vspace{-0.10in}
\hspace{0.15in}
$h$
\vspace{-2.3in}
\hspace{-4.6in}
$M_{00}$
\vspace{-0.30in}
\hspace{-0.15in} $h_{00}$ \hspace{2.95in} $h_5^*$
\vspace{2.40in}
\resizebox{0.70\textwidth}{!}{\includegraphics{fig9b.eps}}
\vspace{-2.50in}
\hspace{1.00in}
\resizebox{0.33\textwidth}{!}{\includegraphics{fig9bb.eps}}
\vspace{0.92in}
\hspace{0.22in}
$h$
\vspace{-2.00in}
\hspace{-4.6in}
$M_{10}$
\vspace{-0.40in}
\hspace{1.25in} $h_6^*$ \hspace{0.93in} $h_{10}$
\vspace{-4.65in} \hspace{3.20in} (a)
\vspace{ 3.35in} \hspace{3.20in} (b)
\vspace{2.9in}
\caption{Functions $M_{00}(h) $ and $M_{10}(h) $
under the conditions
$\, \mu_{10} = \mu_{11} = $, $\, \mu_{12} \ne 0$
and $\, \mu_{00} \ne 0$,
for $\, a_1 = -\,\frac{4}{3}$ and $\, a_4 = - \frac{6}{5}$:
(a) $\, M_{00}(h) \,$
for $h \in [h_{00}, \, +\infty)$, with
$\, h_0 = \frac{325}{128} \approx 2.53096$,
crossing the $h$-axis at $h=h_5^* \in (12.6197809949,\,12.6197809950)$;
and (b) $M_{10}(h) $ for $h \in (-\infty, \, h_1]$,
with $\, h_{10} = -\,\frac{75}{128}\ 3^{4/5} \approx -\,1.41107$,
crossing the $h$-axis at $h=h_6^* \in (-\,3.1388150376,\,-\,3.1388150375)$.}
\label{fig9}
\end{center}
\vspace{0.10in}
\end{figure}
\begin{figure}[!h]
\vspace{0.00in}
\begin{center}
\hspace{0.00in}
\resizebox{0.50\textwidth}{!}{\includegraphics{fig10.eps}}
\vspace{-1.15in}
\resizebox{0.50\textwidth}{!}{\includegraphics{fig10a.eps}}
\vspace{-1.55in}
\hspace{-1.88in}${\small \frac{3}{4}}$
\vspace{-6.45in} \hspace{3.70in} (a)
\vspace{ 3.30in} \hspace{3.70in} (b)
\vspace{2.95in}
\caption{Illustration of the existence of $4$ limit cycles
when $\, a_1 = -\,\frac{4}{3}$, $\, a_4 = - \frac{6}{5}$,
and $a_{10}=1$,
$ b_{11} = \frac{1176}{65}\, a_{10} - \varepsilon_1$,
$b_{01}=-\,\frac{513}{65} \, a_{10} + \varepsilon_2$,
where $\, 0 < \varepsilon_2 \ll \varepsilon_1 \ll \varepsilon $:
(a) An unstable large limit cycle enclosing the center $(0,0)$;
and (b) Zoomed area around the center $(1,0)$ showing the
existence of $1$ large limit cycle and $2$ small limit cycles.}
\label{fig10}
\end{center}
\vspace{0.00in}
\end{figure}
\vspace{0.05in}
(D) Finally, consider the $(0,2)$-distribution. For this case, the condition
$\, a_4 = \frac{1}{3}\, (6\,a_1 + 5) \,$ is not used. Taking
$$
a_1 = -\, \displaystyle\frac{4}{3}, \quad a_4 = -\, \displaystyle\frac{6}{5},
$$
yields
$$
\gamma = \Big(1- \displaystyle\frac{4}{3}\,x \Big)^{-\frac{14}{5}}
\quad (x \ne \frac{3}{4}).
$$
The point $(-\frac{4}{3}, -\frac{6}{5})$ is marked by a dark circle
near the line $\, a_4 = \frac{1}{3}\, ( 6\, a_1 + 5)\,$
in the $a_1$-$a_4$ parameter plane (see Fig.~\ref{fig1}).
Further, we have $\,
b_{01} = -\, \frac{513}{65} \, a_{10}, \
b_{11} = \frac{1176}{65}\, a_{10}$,
and
$$
H(x,y) = \displaystyle\frac{243\,(64\,y^2+ 480\,x^2-780\,x+325}
{((3-4\,x)\, (324\,x - 243)^{4/5}} \quad
{\rm for} \quad x \ne \displaystyle\frac{3}{4},
$$
with
$$
h_{00} = \displaystyle\frac{325}{128} >
h_{10} = -\,\displaystyle\frac{75}{128} \ 3^{4/5}.
$$
For this case, $\mu_{00} \,$ and $\, \mu_{12} \,$ are simplified as
$$
\mu_{00} = -\, \displaystyle\frac{896}{65} \, \pi\, a_{10} < 0 \quad {\rm and} \quad
\mu_{12} = -\, \displaystyle\frac{448}{30375} \, 3^{9/10}\, \pi \, a_{10} < 0.
$$
The computation results of $\, M_{00}(h) \,$ for $\, h \in (h_{00}, \infty) \,$
and $\, M_{10}(h) \,$ for $\, h \in (-\infty, h_{10}) \,$ are shown
in Figs.~\ref{fig9}(a) and \ref{fig9}(b), respectively.
As shown in Fig.~\ref{fig9}(a), the sign of $\, M_{00}(h)\,$
agrees with that of $\, \mu_{00} < 0 \,$ for
$\, 0 < h - h_{00} \ll 1$, and the function $\, M_{00}(h)\,$
crosses a critical value at $\, h=h_5^* \in (12.6197809949,\,12.6197809950)\,$
at which it changes sign.
Figure~\ref{fig9}(b) shows $\, M_{10}(h) \,$ for
$\, h \in (-\infty, h_{10})$, whose sign agrees with that of
$\, \mu_{12} < 0\,$ for $\, 0 < h_{10}-h \ll 1$. Moreover,
$\, M_{10}(h) \,$ crosses a critical value at $\, h=h_6^* \in
(-\,3.1388150376,\,-\,3.1388150375)\,$
at which it changes sign.
Therefore, for this case, in addition to the two small limit cycles,
there also exist at least two large limit cycles bifurcating from the two
different closed orbits $L_{h_5^*}$ and $L_{h_6^*}$ of (\ref{b59}).
One large limit cycle surrounding the center $(0,0)$
is shown in Fig.~\ref{fig10}(a), while another large limit cycle
enclosing the center $(1,0)$ with $2$ small limit
cycles is depicted in Fig.~\ref{fig10}(b).
It is noted that all the four sets of values of $a_1$ and $a_4$ chosen above
in (A), (B), (C) and (D) satisfy
\begin{equation}
\displaystyle\frac{a_1 + 2\, a_4}{a_1} = \displaystyle\frac{2\,n}{m}, \quad
{\rm where} \ \, n \ \, {\rm is \ an \ integer \ and} \ \, m \ \,
{\rm is \ an \ odd \ integer},
\label{b82}
\end{equation}
so that a consistent integrating factor (and so a consistent
Hamiltonian function for the whole transformed system) is obtained.
However, this condition is not necessary
since the singular line $\, 1+a_1\,x = 0\,$
divides the phase plane into two parts, and the analysis
does not need the continuity on the singular line.
To demonstrate this, in the following
we present a case for which the condition
(\ref{b82}) is not satisfied. Consider the $(2,0)$-distribution, and choose
$\, a_1 = -\,5\,$ and $\, a_4 = -\,4$. The point $(a_1, a_4) = (-5,\,-4)\,$
is marked by a dark circle in the $a_1$-$a_4$ parameter plane
(see Fig.~\ref{fig1}). Then,
$$
\displaystyle\frac{a_1+2 a_4}{a_1} = \displaystyle\frac{13}{5}, \quad
b_{01}= a_{10}, \quad b_{11} = \displaystyle\frac{26}{3}\, a_{10},
$$
\vspace{-0.05in}
\noindent
and \vspace{0.10in}
$$
H(x,y) = \left\{
\begin{array}{lll}
\displaystyle\frac{ x^2 + y^2}{2 \,( 1 - 5\,x)^{8/5}}, \ \ &
\forall \ h \in (0, \infty), \ \ & {\rm when} \ \ x < \displaystyle\frac{1}{5}, \\[2.0ex]
-\, \displaystyle\frac{ x^2 + y^2}{2 \,( 1 - 5\,x)^{8/5}}, &
\forall \ h \in (-\infty, -\frac{1}{32}\, 2^{4/5}), & {\rm when} \ \
x > \displaystyle\frac{1}{5}.
\end{array}
\right.
$$
For this case, $\, \mu_{02} \,$ and $\, \mu_{10} \,$ become
$$
\mu_{02} = -\, \displaystyle\frac{130}{3}\, \pi \, a_{10} < 0 \quad {\rm and} \quad
\mu_{10} = -\, \displaystyle\frac{65}{12}\, \pi \, a_{10} < 0.
$$
The computation result of $\, M_{00} (h) \,$ shows that
$\, M_{00}(h) < 0 \,$ for $\, 0 < h \ll 1$, agrees with the sign of
$\, \mu_{02}$. Moreover, $\, M_{00}(0.1) = 0.0510077880 > 0$, implying that
there exists $ h = h_7^* \in (0,\, 0.1) \,$ such that $\, M_{00}(h_7^*)=0$,
and so a large limit cycle bifurcates from the closed orbit
$\, L_{h_7^*}$ of (\ref{b59}).
The result of $ \,M_{10}(h) \,$ also shows that $\, M_{10}(h) < 0 \,$
for $\, 0 < -\,\frac{1}{32}\, 2^{4/5} - h \ll 1 $, agreeing with the sign of
$\, \mu_{10}$, and that $\, M_{10}(-\,\frac{1}{32}\, 2^{4/5}-0.8)
= 7.4630743072 > 0$, implying the existence
$\, h = h_8^* \in (-\,\frac{1}{32}\, 2^{4/5}-0.8,\, -\,\frac{1}{32}\, 2^{4/5})
= (-0.8544094102, \, -0.0544094102)\,$ such that
$\, M_{10}(h_8^*)=0$. Thus, there exists another large limit cycle
bifurcating from the closed orbit
$\, L_{h_8^*}$ of (\ref{b59}).
Therefore, this case exhibits $2$ small limit
cycles and $2$ large limit cycles, leading to the existence of
at least $4$ limit cycles.
Summarizing the above results with the continuity of parameters
$\,a_1\,$ and $\, a_4\,$ shows that
at least for some regions in the $a_1$-$a_4$ parameter plane
the reversible near-integrable system (\ref{b55}) can exhibit
at least $4$ limit cycles around the two singular points $(0,0)$ and $(1,0)$
with distribution ether $(3,1)$ or $(1,3)$.
The proof of Theorem 4.1 is finished.
\put(10,0.5){\framebox(6,7.5)}
\section{Conclusion}
In this paper, we have proved that a quadratic non-Hamiltonian integrable
system with two centers can have at least $4$ limit cycles
under quadratic perturbations, with distributions either
$(3,1)$ or $(1,3)$.
This result gives a new record, answering the open problem of the existence of
limit cycles in near-integrable quadratic systems.
It is shown that such systems
can have at least $4$ limit cycles for some regions
in the $2$-dimensional parameter plane, associated with the parameters
of the integrable systems.
Further research is needed on global analysis for all possible parameter
values in the parameter plane.
\section*{Acknowledgments}
This work was supported by
the Natural Sciences and Engineering Research Council of Canada (NSERC)
and the National Natural Science Foundation of
China (NNSFC).
|
\section{Introduction and Summary}
The study of anisotropies in the cosmic microwave background radiation over the past two decades has dramatically improved our understanding of the early universe.
There is now strong evidence that the anisotropies we see today originated from primordial fluctuations generated in the very early universe, and we have learned that these primordial fluctuations have a nearly scale-invariant spectrum. Furthermore, the data still contains no evidence for a deviation from adiabaticity, or Gaussianity~\cite{Komatsu:2010fb}. Even though the case is by no means closed, these properties of the primordial fluctuations certainly support the idea that they originated as quantum fluctuations during inflation, a phase of nearly exponential expansion of the universe~\cite{Guth:1980zm}.
While observations are now good enough to rule out some of the simplest inflationary models involving only a single slowly rolling field with canonical kinetic term, other models in this class are still compatible with all existing data. These models predict an adiabatic spectrum with primordial non-Gaussianities that are too small to be observed, but this is not a generic prediction of inflation. Many models of inflation have been constructed and studied that can lead to an observable departure from Gaussianity. Some popular possibilities are models with multiple fields, non-canonical kinetic terms, light spectator fields and a violation of slow-roll. Beyond an existence proof that observably large non-Gaussianities can be generated, these models provide us with useful theoretical expectations to guide our search.
For a Gaussian signal, all odd $n$-point functions vanish and the higher even $n$-point functions are given in terms of sums of products of the two-point function. The most straightforward way to look for a departure from Gaussianity is then to look for a non-zero three-point function. In Fourier space the three-point function depends on three momenta. Translational invariance of the background geometry ensures that these momenta add up to zero and thus form a triangle. Rotational invariance furthermore dictates that the three-point function can only depend on the three independent scalar products of these momenta.
The information contained in the three-point function can thus be captured by a function of three variables, that can be thought of as two angles and one side of the triangle.
Since the dependence is a priori completely arbitrary, a model independent measurement would be desirable and would provide a precious criterion to discriminate between otherwise indistinguishable models. Unfortunately, progress in this direction is very hard. (For a review see {\it e.g.}~\cite{Liguori:2010hx}). Essentially all phenomenological analyses start from some explicit form of the three-point function guided both by theoretical expectations and by the simplicity of the numerical analysis necessary to compare it with the data. (See, however,~\cite{Fergusson:2009nv}.) Once a ``shape'' has been chosen, only the amplitude of this type of non-Gaussianity remains as a parameter, which is conventionally called $f^\text{shape}$. So far only a handful of scale-invariant shapes have been looked for in the data.
The most recent observational bounds on the magnitude for various shapes from the 7-year WMAP data at 95\% CL are~\cite{Komatsu:2010fb}:
\begin{center}
\begin{tabular}{lc}
local non-Gaussianity & $-10<f^\text{local}<74$\,,\\
equilateral non-Gaussianity& $-214<f^\text{equil}<266$\,,\\
orthogonal non-Gaussianity& $-410<f^\text{ortho}<6$\,.\\
\end{tabular}
\end{center}
Planck data will make it possible to tighten the error bars by about a factor of five, and we may soon find out whether it is necessary to go beyond the simplest models of inflation.
As already briefly mentioned, departures from the slow-roll condition can potentially lead to large non-Gaussianities.
Two conceptually distinct possibilities were first explored by Chen, Easther, and Lim. An otherwise smooth potential might either exhibit a sharp, localized feature~\cite{Chen:2006xjb} (see also~\cite{Bean:2008na}), or it might display a periodic modulation that averages to zero~\cite{Chen:2008wn}.
Chen, Easther and Lim have performed a numerical analysis of both scenarios~\cite{Chen:2006xjb,Chen:2008wn}. They show that a large non-Gaussian signal can be produced without violating the constraints on these models from measurements of the two-point function. They also provide a heuristic estimate of the signal for equilateral configurations for the case of resonant production.
In this work, we analytically compute the scalar primordial bispectrum generated for a modulated potential for arbitrary momentum configurations from first principles. Our work is motivated by a class of models derived from string theory~\cite{McAllister:2008hb, Flauger:2009ab}, but let us stress that these are not the only models in which such oscillations are expected to arise. In large field inflation, the inflaton potential must be flat over a range in field space large compared to $M_{\mathrm{Pl}}$. From the point of view of effective field theory, this seems unnatural unless there is an underlying shift symmetry. Axions are thus natural candidates for the inflaton in large field inflation. It then seems plausible that the potential might receive small periodic contributions from non-perturbative effects. These periodic contributions might be due to instantons in a gauge sector the axion couples to, or, in the context of string theory, they might arise from Euclidean branes or world-sheet instantons. Whether string inspired or not, as soon as we invoke the shift symmetry of axions to explain why the inflaton potential is so flat, we should admit the possibility of small periodic modulations in the potential which may lead to observational consequences. To be specific, the potentials we will consider are of the form
\begin{eqnarray} \label{V}
V(\phi)=V_0(\phi)+\Lambda^4\cos \left(\frac{\phi}{f}\right)\,,
\end{eqnarray}
where $\phi$ is a canonically normalized real scalar field, and $V_0(\phi)$ is assumed to admit slow-roll inflation in the absence of modulations, ${\it i.e.}$ for $\Lambda=0$.
A particular model of this form with $V_0(\phi)=\mu^3\phi$ was obtained from a string theory construction in~\cite{McAllister:2008hb, Flauger:2009ab}, and we will sometimes focus on this special case for concreteness.
The parameters $\Lambda$ and $f$ have dimensions of a mass and are a priori undetermined. However, consistency of a more fundamental description of the system, in our case string theory, will typically limit them to lie in a certain range.
We will assume that the potential is monotonic at least near the values of the scalar field around which the modes we observe in the cosmic microwave background (CMB) exit the horizon and do not consider models in which the inflaton gets trapped. This can be summarized by requiring that the monotonicity parameter
\begin{eqnarray}
b_*\equiv\frac{\Lambda^4}{V_0'(\phi_*)f}<1\,.
\end{eqnarray}
Except for a linear potential, this parameter depends on the value of the field. We will evaluate it at $\phi=\phi_*$, the value of the scalar field at the time when the mode with comoving momentum equal to the pivot scale $k=k_*$ exits the horizon.
To be compatible with WMAP data, the monotonicity parameter must satisfy $b_*\ll1$ for both linear~\cite{Flauger:2009ab} and quadratic $V_0(\phi)$~\cite{Pahud:2008ae}. Other potentials of this form have not been compared to the data, but we expect this to be true for the general case and treat $b_*$ as an expansion parameter.
Our main result is that the three-point function of scalar curvature perturbations to linear order in $b_*$ at some late time $t$, when the modes have exited the horizon, takes the form\footnote{We have set $M_{\mathrm{Pl}}=1$ in this formula and will do so throughout the paper. As usual, it can be re-inserted by dimensional analysis.}
\begin{multline}\label{resu}
\langle\mathcal{R}({\bf k_1},t)\mathcal{R}({\bf k_2},t)\mathcal{R}({\bf k_3},t)\rangle=(2\pi)^7\Delta_\mathcal{R}^4\frac{1}{k_1^2k_2^2k_3^2} \delta^3({\bf k_1}+{\bf k_2}+{\bf k_3})\\\times f^\text{res}\left[\sin\left(\frac{\sqrt{2\epsilon_*}}{f}\ln K/k_*\right)+\frac{f}{\sqrt{2\epsilon_*}} \sum\limits_{i\neq j} \frac{k_i}{ k_j}\cos\left(\frac{\sqrt{2\epsilon_*}}{f}\ln K/k_* \right)+\dots\right]\,.
\end{multline}
with
\begin{equation}\label{fres}
f^\text{res}=\frac{3 b_*\sqrt{2\pi}}{8}\left(\frac{\sqrt{2\epsilon_*}}{f}\right)^{3/2}\,,
\end{equation}
where $\epsilon_*$ denotes the value of the slow-roll parameter derived from the smooth part of the potential $V_0(\phi)$, evaluated at the time the pivot scale $k_*$ exits the horizon. The quantities $\phi_*$ and $\epsilon_*$ are model dependent, but are easy to calculate for any given model. The comoving momentum $K=k_1+k_2+k_3$ is the perimeter of the triangle in momentum space.
The dots in \eqref{resu} stand for terms that can be neglected either because they are suppressed by higher powers in the slow-roll parameters for the smooth part of the potential or by positive powers of $f/\sqrt{2\epsilon_*}$. From \eqref{fres} one can see that large non-Gaussianity requires $f/\sqrt{2\epsilon_*}\ll1$. So this will be the regime of interest in this the paper. The second term is suppressed compared to the first by a factor of $f/\sqrt{2\epsilon_*}$. It is negligible except for squeezed triangles where one of the momenta is much less than the other two. It ensures that the consistency relation of~\cite{Maldacena:2002vr} (see also \cite{Creminelli:2004yq}) holds.
We compare our analytic result with the numerical analysis of~\cite{Hannestad:2009yx} for the linear potential of axion monodromy. We find agreement with their numerical results for the bispectrum at the per cent level.
We compare our results with the numerical analysis of \cite{Chen:2008wn} for the quadratic potential. Identifying their parameter $\tilde{P}^2$ with $9\Delta_\mathcal{R}^4(k_*)/10$, we find agreement with their numerical results as well.
This type of non-Gaussianity, which we refer to as \textit{resonant non-Gaussianity} following the nomenclature in~\cite{Chen:2008wn}, is nearly orthogonal to all commonly studied shapes. The cosine defined in~\cite{Babich:2004gb,Fergusson:2008ra} is less than $10\%$ for the entire range of parameters relevant for axion monodromy inflation.
This is intuitively clear. The shape of resonant non-Gaussianity rapidly oscillates around zero while the other shapes are slowly varying. As the number of oscillations increases, the cosine decreases. The number of oscillations over the scales observed in the cosmic microwave background is approximately given by $\sqrt{2\epsilon_*}/f$. For our string theory example, this implies that an axion decay constant $f=10^{-3} $ leads to about $90$ periods in the cosmic microwave background. The cosine with local, equilateral, and orthogonal shapes is then around $2\%$, and the largest amount of non-Gaussianity for this value of $f$ that is consistent with the observational constraints on the two-point function is $f^\text{res}\approx 80$.
As we will explain, the resonant effect we are studying arises from a term in the interaction Hamiltonian that is higher order in the slow-roll parameters. As a consequence, it is not captured by the simplest version of the effective field theory of inflation~\cite{Cheung:2007st} (see also~\cite{Weinberg:2008hq}). A more detailed discussion is postponed to Section~\ref{s:disc}.
The paper is organized as follows. In Section~\ref{s:power}, we derive the time evolution of the curvature perturbation for a potential~\eqref{V} and obtain the power spectrum. The derivation is independent of the ones presented in \cite{Flauger:2009ab}. It agrees with the results that were found there for the linear potential. In Section~\ref{s:bis}, we calculate the bispectrum and check that it fulfills the consistency relation of~\cite{Maldacena:2002vr} for squeezed triangles. In Section~\ref{s:cosine}, we compute the overlap between the shape of resonant non-Gaussianity and the three most common shapes in the literature that have been compared with the data. We show that it is very small for the range of parameters relevant for our string theory example. Section~\ref{s:disc} contains a discussion of our results. In Appendix \ref{a:gen}, we present the details of the calculation of the inflationary background solutions for the potential \eqref{V}. Appendix~\ref{a:sr} provides the relation between various different definitions for slow-roll parameters commonly used in the literature.
\section{The Mode Functions and the Power Spectrum} \label{s:power}
As we show in Appendix~\ref{a:gen}, the background solution for a scalar field with potential~\eqref{V} can be derived to first order in $b_*$ and to leading order in the slow-roll parameters of $V_0$. In the limit $f\ll\sqrt{2\epsilon_*}$, which, as we shall see, is the regime where large non-Gaussianities are generated, the solution is well approximated by
\begin{eqnarray}
\phi(t)=\phi_0(t)-\frac{3 b_* f^2}{\sqrt{2\epsilon_* }}\sin\left(\frac{\phi_0(t)}{f}\right)\,.
\end{eqnarray}
Here $\phi_*$ is the value of the scalar field when the pivot scale $k=k_*$ exits the horizon,\footnote{For numerical calculations, we will take $k_*=0.002\,\text{Mpc}^{-1}$. The value of $\phi_*$ is model dependent. For the linear potential we use $\phi_*=10.88$.} $\epsilon_*$ is the slow-roll parameter in the absence of modulations evaluated at $\phi_0=\phi_*$, and $\phi_0$ is the solution for the scalar field in the absence of modulations, {\it i.e.} for $b_*=0$.
It can be obtained as a function of time by integrating its equation of motion, but we will not need the result at this time.
Since we will need them later, let us give the expressions for the Hubble slow-roll parameters\footnote{For the convenience of the reader we have collected in other possible definitions of the slow-roll parameters and formulae for the conversion in Appendix \ref{a:sr}.}
\begin{eqnarray}\label{eq:Hsr}
\epsilon\equiv-\frac{\dot{H}}{H^2}\hskip 1cm \text{and}\hskip 1cm\delta=\frac{\ddot{H}}{2\dot{H}H}\,.
\end{eqnarray}
For the potential~\eqref{V}, it is convenient to calculate them in an expansion in the parameter $b_*$
\begin{eqnarray}
\epsilon=\epsilon_0 + \epsilon_1 + {\cal{O}}(b_*^2)\,,\\
\delta=\delta_0 + \delta_1 + {\cal{O}}(b_*^2)\,.
\end{eqnarray}
In the slow-roll approximation for $\phi_0$, and for $f\ll\sqrt{2\epsilon_*}$, one finds
\begin{eqnarray}
&&\epsilon_0=\epsilon_*,\;\;\;\;\delta_0=\epsilon_*-\eta_*\,,\\
&&\epsilon_1=-3b_*f\sqrt{2\epsilon_*}\cos\left(\frac{\phi_0(t)}{f}\right) \,,\\
&&\delta_1=-3b_*\sin\left(\frac{\phi_0(t)}{f}\right)\label{eq:d1}\,,
\end{eqnarray}
where $\eta_*$ and $\epsilon_*$ are the values of the potential slow-roll parameters $\eta_{V_0}\equiv V_0''/V_0$ and $\epsilon_{V_0}\equiv(V_0'/V_0)^2/2$ derived from the smooth part of the potential $V_0$, evaluated at the time the pivot scale $k_*$ exits the horizon. Notice that in this regime both $\epsilon_1\ll1$ and $\delta_1\ll1$ as long as $b_*\ll1$. On the other hand this is not the case for higher slow-roll parameters. For instance, one has
\begin{eqnarray}
\dot\delta_1/H=3b_* \frac{\sqrt{2\epsilon_*}}{f} \cos\left(\frac{\phi_0(t)}{f}\right)\label{eq:dd1}\,,
\end{eqnarray}
which becomes large for small $f/\sqrt{2\epsilon_*}$.
Let us now turn to the spectrum of scalar perturbations. As in~\cite{Flauger:2009ab}, we will choose a slicing such that $\delta\phi({\bf x},t)=0$, and use a spatial diffeomorphism to bring the scalar perturbations in the spatial part of the metric into the form
\begin{eqnarray}
\delta g_{ij}({\bf x},t)=2a(t)^2\mathcal{R}({\bf x},t)\delta_{ij}\,.
\end{eqnarray}
The translational invariance of the background makes it convenient to look for the solution as a superposition of Fourier modes
\begin{eqnarray}\label{eq:rx}
\mathcal{R}({\bf x},t)=\int\frac{d^3{\bf k}}{(2\pi)^{3}}\mathcal{R}({\bf k},t)e^{i{\bf k}\cdot{\bf x}}\,,
\end{eqnarray}
where ${\bf x}$ are the comoving coordinates and ${\bf k}$ denotes the comoving momentum of the mode.
Rotational invariance together with reality (or Hermiticity) of $\mathcal{R}({\bf x},t)$ imply that the Fourier components of the most general solution can be written in the form
\begin{eqnarray}\label{eq:rk}
\mathcal{R}({\bf k},t)=\mathcal{R}_k(t)a({\bf k})+\mathcal{R}^*_k(t) a^\dagger(-{\bf k})\,,
\end{eqnarray}
where $k$ is the magnitude of the comoving momentum ${\bf k}$, $a({\bf k})$ can be thought of as a stochastic parameter in the classical theory or as an annihilation operator in the quantum theory, and $\mathcal{R}_k(t)$ is the mode function. When thought of as creation and annihilation operators $a^\dagger({\bf k}')$ and $a({\bf k})$ satisfy the commutation relation
\begin{eqnarray}
\left[ a({\bf k}), a^\dagger({\bf k}')\right]=(2\pi)^3\delta^3({\bf k}-{\bf k'})\,.
\end{eqnarray}
The time evolution of the mode function $\mathcal{R}_k(t)$ is governed by the Mukhanov-Sasaki equation~\cite{Mukhanov:1985rz,Sasaki:1986hm}. For small $\epsilon$, it can be written in the form~\cite{Weinberg:2008zzc}
\begin{eqnarray} \label{MSx}
\frac{d^2\mathcal{R}_k}{dx^2}-\frac{2(1+2\epsilon+\delta)}{x}\frac{d\mathcal{R}_k}{dx}+\mathcal{R}_k=0\,.
\end{eqnarray}
The initial conditions are such that for $x\gg 1$
\begin{eqnarray}\label{in cond}
\mathcal{R}_k(x)\to-\frac{H}{\sqrt{2k}a\dot\phi}e^{ix}\,,
\end{eqnarray}
where we have used the notation $x\equiv-k\tau$ with conformal time $\tau$ defined as $\tau\equiv \int^t\frac{dt'}{a(t')}$. The Mukhanov-Sasaki equation implies that for $x\ll1$ the mode function $\mathcal{R}_k(x)$ approaches a constant which we denote by $\mathcal{R}_k^{(o)}$, where the superscript ${(o)}$ indicates that the mode is outside the horizon.
It is related to the quantity $\Delta_\mathcal{R}^2(k)$ that is commonly quoted to parameterize the primordial scalar power spectrum by
\begin{eqnarray}\label{eq:psr}
\left|\mathcal{R}_k^{(o)}\right|^2=2\pi^2 \frac{\Delta_\mathcal{R}^2(k)}{k^3}\,.
\end{eqnarray}
In the slow-roll approximation, {\it i.e.}~for $\epsilon\ll 1$, $\delta\ll 1$, and assuming $\dot\delta/H$ is small compared to both $\epsilon$ and $\delta$
\begin{eqnarray}\label{eq:psd}
\Delta_\mathcal{R}^2(k)=\frac{H^2(t_k)}{8\pi^2\epsilon(t_k)} \quad \textrm{(slow-roll approximation)}\,,
\end{eqnarray}
where $t_k$ is the time at which the mode with comoving momentum $k$ exits the horizon.
However, as already pointed out in~\cite{Flauger:2009ab}, the slow-roll approximation breaks down in models with modulated potentials because the magnitude of $\dot\delta/H$ is no longer of quadratic order in $\epsilon$ and $\delta$. Furthermore, the slow-roll parameters are oscillatory functions whose frequency changes in time. When the frequency of this oscillation passes through twice the natural frequency of the mode, which is set by the momentum of the mode, parametric resonance occurs, which is not captured in the slow-roll approximation.
In~\cite{Flauger:2009ab} our main concern was the power spectrum so that only the asymptotic behavior of the mode function was needed but not its detailed behavior as a function of time. The main concern of this work is the calculation of the bispectrum for which the knowledge of the time dependence of the mode functions is important. So let us calculate it. We will neglect the effect of $\epsilon_0$ and $\delta_0$ in equation~\eqref{MSx} for simplicity, but they could be restored without too much extra trouble. Furthermore, we will make use of the fact that the amplitude of $\epsilon_1$ is suppressed compared to that of $\delta_1$ by a factor of $f\sqrt{2\epsilon_*}$ and drop it as well. The Mukhanov-Sasaki equation~\eqref{MSx} then becomes
\begin{eqnarray}
\frac{d^2\mathcal{R}_k}{dx^2}-\frac{2(1+\delta_1(x))}{x}\frac{d\mathcal{R}_k}{dx}+\mathcal{R}_k=0\,.
\end{eqnarray}
As was shown in~\cite{Flauger:2009ab}, for the linear potential parametric resonance occurs around $x_\text{res}=1/(2f\phi_*)$. As we will see, for the general potential it happens at $x_\text{res}=\sqrt{2\epsilon_*}/(2f) $. For $x\gg x_\text{res}$, {\it i.e.}~much before the resonance occurred, we know that the effect of $\delta_1$ is negligible. Therefore the solution is\footnote{To keep the dependence of the mode function on the slow-roll parameters $\epsilon_0$ and $\delta_0$, one should replace $3/2$ by $3/2+2\epsilon_0+\delta_0$ in the discussion below.}
\begin{eqnarray}\label{eq:solunpert}
\mathcal{R}_k(x)=\mathcal{R}_{k,0}^{(o)}i\sqrt{\frac{\pi}{2}}x^{3/2}H_{3/2}^{(1)}(x)\,,
\end{eqnarray}
where $\mathcal{R}_{k,0}^{(o)}$ is the value of $\mathcal{R}_k(x)$ outside the horizon in the absence of modulations and is fixed by the initial condition \eqref{in cond}, and
\begin{eqnarray}
i\sqrt{\frac{\pi}{2}}x^{3/2}H_{3/2}^{(1)}(x)=(1-i x)e^{i x}\,.
\end{eqnarray}
Similarly, for $x\ll x_\text{res}$, {\it i.e.}~long after the resonance has occurred, the background frequency is too high for the mode to keep up with it, and the effect of $\delta_1$ is again negligible. The solution there must take the form
\begin{eqnarray}
\mathcal{R}_k(x)&=&\mathcal{R}_{k,0}^{(o)}\left[c^{(+)}_k i\sqrt{\frac{\pi}{2}}x^{3/2}H_{3/2}^{(1)}(x)-c^{(-)}_k i\sqrt{\frac{\pi}{2}}x^{3/2}H_{3/2}^{(2)}(x)\right]\,.
\end{eqnarray}
A slight generalization of our derivation in~\cite{Flauger:2009ab} implies that at late times $c^{(+)}_k=1+\mathcal{O}(b_*^2)$.
It then seems natural to look for a solution of the form
\begin{eqnarray}\label{eq:solpert}
\mathcal{R}_k(x)=\mathcal{R}_{k,0}^{(o)}\left[i\sqrt{\frac{\pi}{2}}x^{3/2}H_{3/ 2}^{(1)}(x)-c^{(-)}_k(x)i\sqrt{\frac{\pi}{2}}x^{3/2}H_{3/2}^{(2)}(x)\right]\,,
\end{eqnarray}
where $c^{(-)}_k(x)$ vanishes at early times and goes to $c^{(-)}_k$ at late times.
The Mukhanov-Sasaki equation then turns into an equation governing the time evolution of $c^{(-)}_k(x)$. To linear order in $b_*$, this equation is
\begin{eqnarray}\label{tt}
\frac{d}{dx}\left[e^{-2 i x}\left(1-\frac{i}{x}\right)\frac{d}{dx}c^{(-)}_k(x)\right]+e^{-2 i x}\frac{i}{x^2}\frac{d}{dx}c^{(-)}_k(x)=-2i \frac{\delta_1(x)}{x}\,.
\end{eqnarray}
For large $x$, which is where the resonance occurs as long as $f/\sqrt{2\epsilon_*}\ll1$, equation \eqref{tt} can be written in the form
\begin{eqnarray}
\frac{d}{dx}\left[e^{-2 i x}\frac{d}{dx}c^{(-)}_k(x)\right]=-2 i\frac{\delta_1(x)}{x}\,.
\end{eqnarray}
Using the expression for $\delta_1(x)$ given by equation~\eqref{eq:d1} together with\footnote{See Appendix~\ref{a:gen} for a details.}
\begin{equation}
\phi_0(x)=\phi_k+\sqrt{2\epsilon_*}\ln x\;\;\;\;\text{where} \;\;\;\;\phi_k=\phi_*-\sqrt{2\epsilon_*}\ln k/k_*\,,
\end{equation}
where $\phi_k$ is the value of the scalar field when the mode with comoving momentum $k$ exits the horizon,
this can immediately be integrated once to give
\begin{eqnarray}\label{eq:cmprime}
\frac{d}{dx}{c}^{(-)}_k(x)=-6ib\frac{f}{\sqrt{2\epsilon_*}}e^{2 i x}\cos\left(\frac{\phi_k}{f}+\frac{\sqrt{2\epsilon_*}}{f}\ln x\right)\,.
\end{eqnarray}
We will be able to use this result in the next section to argue that the modification of the mode function can be ignored when calculating the bispectrum to leading order in $b_*$.
Let us now integrate this equation once again to get a better idea for what the function ${c}^{(-)}_k(x)$ looks like.
To separate the leading and subleading contributions, it is convenient to write $c^{(-)}_k(x)$ as
\begin{equation}\label{eq:cmeq}
{c}^{(-)}_k(x)=-3ib\frac{f}{\sqrt{2\epsilon_*}}\left[e^{-i\frac{\phi_k}{f}}\int\limits_\infty^x dx' \;e^{2 i x'-i\frac{\sqrt{2\epsilon_*}}{f}\ln x'}+e^{i\frac{\phi_k}{f}}\int\limits_\infty^x dx' \;e^{2 i x'+i\frac{\sqrt{2\epsilon_*}}{f}\ln x'}\right]\,.
\end{equation}
\begin{figure}[h!]
\begin{center}
\includegraphics[width=6.6in]{modefcn}
\caption{The top left shows the dominant contribution to $\text{Re}[{c}^{(-)}_k(x)]$ coming from the first integral in equation~\eqref{eq:cmeq}, while the top right shows the subdominant contribution to $\text{Re}[{c}^{(-)}_k(x)]$, which comes from the second integral in equation~\eqref{eq:cmeq}. The bottom left shows their superposition. The bottom right shows the superposition as black dotted line compared to the numerical solution in orange. As can be seen, the two results are essentially indistinguishable. The dashed line represents the asymptotic value for $x$ going to zero ({\it i.e.} outside the horizon) found by performing the integral in~\eqref{eq:cmeq2} in the stationary phase approximation. All plots are for a linear potential with $f=10^{-3}$, $b=10^{-2}$, $\phi_*=10.88$, and for a value of comoving momentum such that $\phi_k=10.7$. Notice that $\text{Re}[{c}^{(-)}_k(x)]$ changes around $x_\text{res}=1/2f\phi_*\approx 45$ as expected.}
\label{fig:modefcn}
\end{center}
\end{figure}
After some manipulations, these integrals can be recognized as incomplete $\Gamma$-functions or closely related exponential integrals.
The result can be written as
\begin{eqnarray} \label{exact}
c^{(-)}_k(x)&=&-\frac32 b\frac{f}{\sqrt{2\epsilon_*}}\left\lbrace e^{\frac{\sqrt{2\epsilon_*}}{f}\left(\frac{\pi}{2}+i\ln2\right)}e^{-i\frac{\phi_k}{f}}\Gamma\left[1-i\frac{\sqrt{2\epsilon_*}}{f},-2ix\right]+ \right. \\&& \hspace{3cm} \left.+ e^{-\frac{\sqrt{2\epsilon_*}}{f}\left(\frac{\pi}{2}+i\ln2\right)}e^{i\frac{\phi_k}{f}}\Gamma\left[1+i\frac{\sqrt{2\epsilon_*}}{f},-2ix\right] \nonumber\right\rbrace \,.\end{eqnarray}
To gain some intuition, the stationary phase approximation is useful. The phase of the first integrand in \eqref{eq:cmeq} will become stationary near the resonance at $x=x_\text{res}$, while the phase of the second integrand is never stationary (since $x>0$). The first term on the right hand side of equations~\eqref{eq:cmeq} and \eqref{exact} will thus be the dominant contribution, with the subdominant contribution from the second term exponentially decreasing for decreasing $f/\sqrt{2\epsilon_*}$.
Figure~\ref{fig:modefcn} shows the leading and subleading contributions to the real part of ${c}^{(-)}_k(x)$, their superposition, as well as the comparison to the numerical result for the ${c}^{(-)}_k(x)$ from its evolution equation \eqref{tt} for a linear potential. Equation \eqref{tt} is valid for arbitrary $x$ and thus exact at linear order in $b_*$.
As a byproduct, we have given an expression for $c^{(-)}_k\equivc^{(-)}_k(0)$ of the form
\begin{eqnarray}\label{eq:cmeq2}
c^{(-)}_k=3ib\frac{f}{\sqrt{2\epsilon_*}}e^{-i\frac{\phi_k}{f}}\int\limits^\infty_0 dx \;e^{2 i x-i\frac{\sqrt{2\epsilon_*}}{f}\ln x}\,.
\end{eqnarray}
When evaluated exactly as in \eqref{exact} or using the stationary phase approximation, one finds that up to an unimportant phase
\begin{eqnarray} \label{cm}
c^{(-)}_k=3b\sqrt{\frac{\pi}{2}}\left(\frac{f}{\sqrt{2\epsilon_*}}\right)^{1/2}e^{-i\frac{\phi_k}{f}}\,.
\end{eqnarray}
For the linear potential we can set $\sqrt{2\epsilon_*}=1/\phi_*$, and we find that this expression agrees with equation (3.40) in~\cite{Flauger:2009ab} for small $f\phi_*$.\footnote{This is true after dropping the same $k$-independent phase that was dropped there.}
Using equations~\eqref{eq:solpert} and \eqref{cm} as well as the behavior for the Hankel functions for small arguments, one finds that the primordial power spectrum of scalar fluctuations is of the form
\begin{eqnarray}
|\mathcal{R}_k^{(o)}|^2=|\mathcal{R}_{k,0}^{(o)}|^2\left[1+3b_*\left(\frac{ 2\pi f}{\sqrt{2\epsilon_*}}\right)^{1/2}\cos\left(\frac{\phi_k}{f}\right)\right]\,,
\end{eqnarray}
or equivalently
\begin{eqnarray}\label{eq:d2r}
\Delta_\mathcal{R}^2(k)=\Delta_\mathcal{R}^2(k_*)\left(\frac{k}{k_*}\right)^{n_s-1}\left[1+\delta n_s\cos\left(\frac{\phi_k}{f}\right)\right]\,\;\;\;\text{with}\;\;\;\;\delta n_s=3b_*\left( \frac{2\pi f}{\sqrt{2\epsilon_*}}\right)^{1/2}\,,
\end{eqnarray}
where once again $\phi_k$ is the value of the scalar field at which the mode with comoving momentum $k$ exits the horizon, $\epsilon_*$ is the value of $\epsilon_{V_0}$ when the pivot scale $k=k_*$ exits the horizon, and $b_*=\Lambda^4/V_0'(\phi_*)f$. We have restored the dependence on $\epsilon_0$ and $\delta_0$ through the appearance of the scalar spectral index $n_s$, which in the approximation we are using is given by $n_s=1-4\epsilon_0-2\delta_0=1-6\epsilon_*+2\eta_*$.
Everything we have said here about the primordial power spectrum for the scalar modes is valid for small $f/\sqrt{2\epsilon_*}$, which, as we will see, is the regime in which observable non-Gaussianity can be generated. In~\cite{Flauger:2009ab}, the interested reader can find the result for the linear potential for general $f\phi_*$.
We have performed numerical calculations to check these analytic results, and we find good agreement. Most of the numerical calculations were done for the linear potential relevant for axion monodromy inflation, but we have also performed some checks for the case of a quadratic potential as well as $V_0(\phi)=\mu^{10/3}\phi^{2/3}$ motivated by~\cite{Silverstein:2008sg}. Our numerical results for the amplitude of the modulations as well as the frequency agree with our analytic result at the per cent level in all cases.
The discrepancy between our analytic result and the numerical results for the power spectrum in~\cite{Hannestad:2009yx} can be traced to an initial value for $k/aH$ in their numerical calculation that was too small to capture the resonance for small axion decay constants. This issue will be easy to fix.
\section{The Bispectrum}\label{s:bis}
Let us now turn to the calculation of the three-point function. To leading order in perturbation theory, the three-point function in the ``in-in'' formalism~\cite{Schwinger} (see also~\cite{Weinberg:2005vy,Adshead:2009cb} and references therein) is given by
\begin{eqnarray}
\langle\mathcal{R}({\bf k_1},t)\mathcal{R}({\bf k_2},t)\mathcal{R}({\bf k_3},t)\rangle=-i\int^t_{-\infty}\;dt'\langle\left[\mathcal{R}({\bf k_1},t)\mathcal{R}({\bf k_2},t)\mathcal{R}({\bf k_3}, t),H_I(t')\right]\rangle\,,
\end{eqnarray}
where the expectation value is taken in the in-vacuum, and the interaction Hamiltonian $H_I$ was first worked out in \cite{Maldacena:2002vr} (see also~\cite{Chen:2006nt,Seery:2005wm}). The term responsible for the dominant contribution in models with oscillatory potentials in our notation is given by \cite{Chen:2008wn}
\begin{eqnarray}\label{HI}
H_I(t)\supset -\int d^3x\; a^3(t)\epsilon(t)\dot{\delta}(t)\mathcal{R}^2({\bf x},t)\dot{\mathcal{R}}({\bf x}, t)\,.
\end{eqnarray}
Using equations~\eqref{eq:rx} and~\eqref{eq:rk}, one finds that the contribution to the three-point function from the term~\eqref{HI} in the interaction Hamiltonian is given by
\begin{multline}
\langle\mathcal{R}({\bf k_1},t)\mathcal{R}({\bf k_2},t)\mathcal{R}({\bf k_3},t)\rangle=(2\pi)^3i\,\delta^3({\bf k_1}+{\bf k_2}+{\bf k_3})\mathcal{R}_ {k_1}(t)\mathcal{R}_ {k_2}(t)\mathcal{R}_ {k_3}(t)\times\\\int^t_{-\infty}\;dt'\;2a^3(t')\epsilon(t')\dot\delta(t')\left[\mathcal{R}^*_{k_1}(t')\mathcal{R}^*_{k_2}(t')\dot{\mathcal{R}}^*_{k_3}(t')+2\; perm.\right]+c.c.\,
\end{multline}
Observational constraints on the two-point function imply that knowing the result to linear order in $b_*$ will be enough. The oscillatory nature of $\delta$ is what makes this contribution the dominant one. So to linear order in $b_*$, we can replace $\epsilon$ by $\epsilon_0$, $\delta$ by $\delta_1$ and use the unperturbed mode functions for $\mathcal{R}_ {k}(t)$, {\it i.e.} equation~\eqref{eq:solunpert}.
One might be concerned that the derivative of the correction to the mode function becomes large during the resonance and should be kept, but this is not the case. To see this, notice that there is no contribution to the integral after the modes have frozen out. In fact, we will see that the main contribution arises when the modes are still deep inside the horizon. In this limit, {\it i.e.} for large $x$, the ratio of the absolute value of the time derivative of the unperturbed part $\dot{\mathcal{R}}_{k\,,0}(t)$ and the absolute value of the time derivative of the correction $\dot{\mathcal{R}}_{k\,,1}(t)$ becomes
\begin{eqnarray}
\frac{|\dot{\mathcal{R}}_{k\,,1}|}{|\dot{\mathcal{R}}_{k\,,0}|}=\left|c^{(-)}_k(x)+i\frac{d}{dx}c^{(-)}_k(x)\right|\,.
\end{eqnarray}
The results in the last section imply that this is small. Equation~\eqref{exact} tells us that the absolute value of the first term is never significantly larger than $3b_*\sqrt{\pi/2}(f/\sqrt{2\epsilon_*})^{1/2}$, and equation~\eqref{eq:cmprime} reveals that the absolute value of the second term is always less than $6bf/\sqrt{2\epsilon_*}$. As we will see, large non-Gaussianities can only be generated for decay constants satisfying $f\ll\sqrt{2\epsilon_*}$, so that both are small. We conclude that
\begin{eqnarray}
\frac{|\dot{\mathcal{R}}_{k\,,1}|}{|\dot{\mathcal{R}}_{k\,,0}|}\ll1\,.
\end{eqnarray}
In other words, we can use the unperturbed mode functions, and the approximation becomes better for decreasing axion decay constant.
We are interested in the value of the three-point function after horizon exit. In this case, we can replace the factors $\mathcal{R}_ {k_i}(t)$ outside the integral by $\mathcal{R}_ {k_i,0}^{(o)}$, and take the upper limit of the integral to zero. We will drop the dependence of the mode functions on the slow-roll parameters $\epsilon_0$ and $\delta_0$, as well as $\epsilon_1$, and use expression~\eqref{eq:solunpert} for the mode functions inside the integral. This leads to
\begin{eqnarray} \label{gen}
\langle\mathcal{R}({\bf k_1},t)\mathcal{R}({\bf k_2},t)\mathcal{R}({\bf k_3}, t)\rangle&=&(2\pi)^7\Delta_\mathcal{R}^4\frac{1}{k_1^3k_2^3k_3^3} \delta^3({\bf k_1}+{\bf k_2}+{\bf k_3})\\
&&\hskip-3cm\times\int_0^{\infty}dX \frac{\dot \delta_1}{8H} e^{-i X}
\left[-i k_1k_2k_3-\frac1X \sum_{{i\neq j}} k_i^2 k_j +\frac{i}{X^2}K(k_1^2+k_2^2+k_3^2) \right] +c.c\,,\nonumber
\end{eqnarray}
where $H$ and $\dot\delta_1$ should be thought of as functions of $X\equiv-K\tau$, and $K\equiv k_1+k_2+k_3$ is the perimeter of the triangle in momentum space.
When visualizing the results, it is often convenient to introduce a quantity that contains the information about the deviation of the three-point function from scale invariance rather than the three-point function itself. We will use the notation of~\cite{Chen:2008wn} and define\footnote{There is a factor of $9/10$ between our definition of $\mathcal{G}$ and theirs, {\it i.e.} $\mathcal{G}_\text{there}=10\mathcal{G}_\text{here}/9$, which was introduced there presumably to match the WMAP conventions for the local case, where a famous factor of $3/10$ appears. The remaining factor of $3$ arises because the matching is conventionally done in the equilateral limit where three terms become equal.}
\begin{eqnarray}
\langle\mathcal{R}({\bf k_1},t)\mathcal{R}({\bf k_2},t)\mathcal{R}({\bf k_3},t)\rangle&=&(2\pi)^7\Delta_\mathcal{R}^4\frac{1}{k_1^2k_2^2k_3^2} \delta^3({\bf k_1}+{\bf k_2}+{\bf k_3})\frac{\mathcal{G}(k_1,k_2,k_3)}{k_1k_2k_3}\,.
\end{eqnarray}
It can then be seen from equation~\eqref{gen} that for all models whose three-point function receives its dominant contribution from the term in the interaction Hamiltonian~\eqref{HI}, and for which the mode functions are well approximated by the unperturbed ones, one has
\begin{eqnarray}\label{eq:shapegen}
&&\hskip-1cm\frac{\mathcal{G}(k_1,k_2,k_3)}{k_1k_2k_3} =\frac18\int_0^{\infty}dX \frac{\dot \delta_1}{H} e^{-i X}
\left[-i -\frac1X \sum\limits_{i\neq j} \frac{k_i}{ k_j} +\frac{i}{X^2}\frac{K(k_1^2+k_2^2+k_3^2)}{k_1k_2k_3} \right] +c.c\,.
\end{eqnarray}
If the integral receives its main contribution from a small neighborhood around some value of $X=X_\text{res}$, as is the case in our example, we can replace $1/X$ and $1/X^2$ by $1/X_\text{res}$ and $1/X_\text{res}^2$, respectively, and find that the shape of non-Gaussianity is given by
\begin{eqnarray} \label{eq:Gres}
&&\hskip-1cm\frac{\mathcal{G}(k_1,k_2,k_3)}{k_1k_2k_3} =\frac14
\left[\text{Im}\,\mathcal{I}_K -\frac{1}{X_\text{res}} \sum\limits_{i\neq j} \frac{k_i}{ k_j}\text{Re}\,\mathcal{I}_K - \frac{1}{X_\text{res}^2}\frac{K(k_1^2+k_2^2+k_3^2)}{k_1k_2k_3}\text{Im}\,\mathcal{I}_K \right]\,,
\end{eqnarray}
where we have defined the integral
\begin{eqnarray}\label{eq:IKint}
\mathcal{I}_K\equiv\int_0^{\infty}dX \frac{\dot \delta_1}{H} e^{-i X}\,.
\end{eqnarray}
We see that all we need in order to calculate the shape of non-Gaussianities for models of this class is the quantity $\dot\delta_1/H$ as a function of $X$. We have already given this quantity as a function of the scalar field in equation~\eqref{eq:dd1}. It remains to write it as a function of $X$. To leading order in the slow-roll approximation the scalar field is given in terms of $X$ as
\begin{eqnarray}\label{eq:phiofx}
\phi_0(X)=\phi_K+\sqrt{2\epsilon_*}\ln X\;\;\;\;\text{with}\;\;\;\;\phi_K=\phi_*-\sqrt{2\epsilon_*}\ln K/k_*\,,
\end{eqnarray}
where $\phi_K$ is the value of the scalar field at the time the mode with comoving momentum $K$ exits the horizon.\footnote{See Appendix~\ref{a:gen} for details.} This gives
\begin{eqnarray}
\frac{\dot \delta_1}{H}=\frac{3b_*\sqrt{2\epsilon_*}}{f}\cos\left(\frac{\phi_K}{f}+\frac{\sqrt{2\epsilon_*}}{f}\ln X\right)\,.
\end{eqnarray}
The three terms in the integral~\eqref{eq:shapegen} can then be recognized as $\Gamma$-functions, and the integral~\eqref{eq:shapegen} can be done analytically. In the regime $f\ll\sqrt{2\epsilon_*}$, in which the resonance occurs deep inside the horizon, we can also use~\eqref{eq:Gres}.
The integral~\eqref{eq:IKint} then takes the form
\begin{eqnarray}\label{IKgen}
\mathcal{I}_K=\frac{3b_*\sqrt{2\epsilon_*}}{f}\int\limits_0^\infty dX \;e^{- iX}\cos\left(\frac{\phi_K}{f}+\frac{\sqrt{2\epsilon_*}}{f}\ln X\right)\,,
\end{eqnarray}
which can be written in terms of $\Gamma$-functions
\begin{equation}\label{eq:IKGamma}
\mathcal{I}_K=\frac{3ib_* \sqrt{2\epsilon_* }}{2f}\left[e^{\frac{\pi\sqrt{2\epsilon_*}}{2f}}\Gamma\left(1+ i\frac{\sqrt{2\epsilon_*}}{f}\right)e^{i\frac{\phi_K}{f}}+e^{- \frac{\sqrt{2\epsilon_*}\pi}{2f}}\Gamma\left(1-i\frac{\sqrt{2\epsilon_* }}{f}\right)e^{-i\frac{\phi_K}{f}}\right]\,.
\end{equation}
The absolute values of the $\Gamma$-functions in the first and second term are identical so that the first term dominates for small $f/\sqrt{2\epsilon_*}$ because of the exponential factors. This dominant contribution arises from a neighborhood of size $(\sqrt{2\epsilon_*}/f)^{1/2}$ around $X_\text{res}=\sqrt{2\epsilon_*}/f$ where the phase of the integrand in equation~\eqref{IKgen} becomes stationary. Equation~\eqref{eq:Gres} can then be used as long as $f\ll\sqrt{2\epsilon_*}$, which is the regime we are interested in. Either performing the integral directly in the stationary phase approximation or using Stirling's approximation in equation~\eqref{eq:IKGamma}, one finds that up to a $K$-independent phase
\begin{eqnarray}
\mathcal{I}_K=-\frac{3 b_*\sqrt{2\pi}}{2}\left(\frac{ \sqrt{2\epsilon_* }}{f}\right)^{3/2}e^{i\frac{\phi_K}{f}}\,.\end{eqnarray}
Combining this with equation~\eqref{eq:Gres}, we see that the shape of resonant non-Gaussianity is given by
\begin{multline}
\frac{\mathcal{G}(k_1,k_2,k_3)}{k_1k_2k_3}=-\frac{3\sqrt{2\pi}b_*}{8}\left(\frac{\sqrt{2\epsilon_*}}{f}\right)^{3/2}\left[\sin\left(\frac{\phi_K}{f}\right)-\frac{f}{\sqrt{2\epsilon_*}}\sum\limits_{i\neq j} \frac{k_i}{ k_j}\cos\left(\frac{\phi_K}{f}\right)\right.\\\left.-\left(\frac{f}{\sqrt{2\epsilon_*}}\right)^2\frac{K(k_1^2+k_2^2+k_3^2)}{k_1k_2k_3}\sin\left(\frac{\phi_K}{f}\right)\right]\,.
\end{multline}
Other terms in the interaction Hamiltonian also contribute at order $(f/\sqrt{2\epsilon_*})^2$, but these contributions are too small to be phenomenologically interesting and we will drop them. Ignoring a $K$-independent phase, we thus write the final result for the resonant shape as
\begin{eqnarray}\label{eq:Gresfin}
\frac{\mathcal{G}(k_1,k_2,k_3)}{k_1k_2k_3}&=&f^\text{res}\left[\sin\left(\frac{\sqrt{2\epsilon_*}}{f}\ln K/ k_*\right)\right.\\
&&\quad\quad\quad\left.+\frac{ f}{\sqrt{2\epsilon_*}}\sum\limits_{i\neq j} \frac{k_i}{ k_j}\cos\left(\frac{\sqrt{2\epsilon_*}}{f}\ln K/k_*\right)+\dots\right]\,,\nonumber
\end{eqnarray}
or equivalently for the three-point function
\begin{multline}
\langle\mathcal{R}({\bf k_1},t)\mathcal{R}({\bf k_2},t)\mathcal{R}({\bf k_3}, t)\rangle=(2\pi)^7\Delta_\mathcal{R}^4\frac{1}{k_1^2k_2^2k_3^2} \delta^3({\bf k_1}+{\bf k_2}+{\bf k_3})\\\times f^\text{res}\left[\sin\left(\frac{\sqrt{2\epsilon_*}}{f}\ln K/k_*\right)+ \frac{f}{\sqrt{2\epsilon_*}} \sum\limits_{i,j} \frac{k_i}{ k_j}\cos\left(\frac{\sqrt{2\epsilon_*}}{f}\ln K/k_*\right)+\dots\right]\,,
\end{multline}
with
\begin{equation}
f^\text{res}=\frac{3 b_*\sqrt{2\pi}}{8}\left(\frac{\sqrt{2\epsilon_* }}{f}\right)^{3/2}\,.\label{fresgen}
\end{equation}
The dots stand for terms that have been dropped because they are higher order in slow-roll or $f/\sqrt{2\epsilon_*}$.
A few comments on this result are in order. Notice that both the frequency of the oscillation and, when written in terms of the monotonicity parameter $b_*$, the amplitude $f^\text{res}$ depend only on $\sqrt{2\epsilon_*}/f$. If this type of non-Gaussianity were measured, it would thus not be possible to distinguish between different potentials from this measurement alone. However, a measurement of the amplitude of tensor modes would give us a direct measurement of $\epsilon_*$ and hence break the degeneracy.\footnote{For a general discussion of the implications of a measurement of tensor modes for the inflationary theory see~\cite{Baumann:2008aq} and references therein.}
Concerning detectability, notice also that when the axion decay constant becomes too small, the frequency of the oscillations becomes too high to be experimentally resolvable. Let us assume the signal can be resolved if the period is longer than $\Delta \ell\sim1$ around the first Doppler peak, {\it i.e.} near $\ell\sim200$.\footnote{This will then also be true for all larger $\ell$}. The periodicity near a given value of $\ell$ is given by $2\pi \ell f/\sqrt{2\epsilon_*}$. So our condition is $\sqrt{2\epsilon_*}/f\lesssim 2\pi \ell $, and taking $\ell=200$ leads to
\begin{eqnarray}\label{44b}
f^\text{res}\lesssim 4 \times 10^4\, b_*\,.
\end{eqnarray}
In order to go from this to a numerical value for the upper bound on $f^\text{res}$, one needs an upper bound on $b_*$ from comparison of the predicted power spectrum with the data. So far this comparison has only been done for the linear potential \cite{Flauger:2009ab}, which will be the subject of the next subsection, and the quadratic potential \cite{Pahud:2008ae}.\footnote{For another observational constraint on an oscillatory power spectrum due to a deviation from the Bunch-Davies vacuum see~\cite{Okamoto:2003wk}.} However, in \cite{Pahud:2008ae}, only decay constants $f$ ($\beta M_{\mathrm{Pl}}$ in their notation) larger than $5\times 10^{-3}$ were considered, which is too large to give a sizable $f^{\text{res}}$.
It may also be interesting to consider a potential of the form
\begin{eqnarray}
V(\phi)=V_0(\phi)\left[1+\lambda\cos\left(\frac{\phi}{f}\right)\right]\,.
\end{eqnarray}
In the approximation we have been working in, our results can immediately be translated to this type of potential by replacing $b_*f\sqrt{2\epsilon_*}\to\lambda$.
One finds that\footnote{See Appendix~\ref{a:gen} for details.}
\begin{eqnarray}
f^\text{res}=\frac{3\lambda\sqrt{2\pi}}{8f^2} \left(\frac{\sqrt{2\epsilon_*}}{f}\right)^{1/2}\,.
\end{eqnarray}
This type of potential with $V_0(\phi)=\frac12m^2\phi^2$ was studied by Chen, Easther, and Lim in~\cite{Chen:2008wn}. One has
\begin{eqnarray}
\sqrt{2\epsilon_*}=\frac{2}{\phi_*}\,,
\end{eqnarray}
so that our result is
\begin{eqnarray}
f^\text{res}=\frac{3\sqrt{\pi}}{4}\frac{\lambda}{f^{2}\sqrt{f\phi_*}}\,.
\end{eqnarray}
This is smaller than their analytic estimate for the amplitude in the equilateral limit by a factor of $10\sqrt{\pi}/27$, which is in agreement with their statement that their analytic estimate overpredicts the numerical result by about $30\%$.\footnote{For the comparison, recall that there is difference by a factor of $10/9$ between their definition of $f_\text{res}$ and our $f^\text{res}$ from $\mathcal{G}_\text{there}=10\mathcal{G}_\text{here}/9$.}
\subsection{The bispectrum for a linear potential}
We will now consider axion monodromy inflation~\cite{McAllister:2008hb, Flauger:2009ab} in some detail as a special case.
The low energy effective theory describing the system is that of a canonically normalized real scalar field with potential
\begin{eqnarray}\label{eq:V}
V(\phi)=\mu^3\phi+\Lambda^4\cos \left(\frac{\phi}{f}\right)=\mu^3\left[\phi+b f\cos \left(\frac{\phi}{f}\right)\right]\,.
\end{eqnarray}
The parameter $\mu$ has dimensions of a mass and is fixed by COBE normalization to be approximately $\mu\simeq 6\times 10^{-4}$.
For this potential, one finds $\sqrt{2\epsilon_*}=1/\phi_*$, and $b_*=b$ is now independent of $\phi_*$. We thus conclude that the shape of non-Gaussianities in this model is
\begin{equation}\label{eq:Gresfinlin}
\frac{\mathcal{G}(k_1,k_2,k_3)}{k_1k_2k_3}=f^\text{res}\left[\sin\left(\frac{\ln K/k_*}{f\phi_*}\right)+f\phi_*\sum\limits_{i\neq j} \frac{k_i}{ k_j}\cos\left(\frac{\ln K/k_*}{f\phi_*}\right)+\mathcal{O}\left((f\phi_*)^2\right)\right]\,,
\end{equation}
or equivalently for the three-point function
\begin{multline}
\langle\mathcal{R}({\bf k_1},t)\mathcal{R}({\bf k_2},t)\mathcal{R}({\bf k_3},t)\rangle=(2\pi)^7\Delta_\mathcal{R}^4\frac{1}{k_1^2k_2^2k_3^2} \delta^3({\bf k_1}+{\bf k_2}+{\bf k_3})\\\times f^\text{res}\left[\sin\left(\frac{\ln K/k_*}{f\phi_*}\right)+f\phi_* \sum\limits_{i,j} \frac{k_i}{ k_j}\cos\left(\frac{\ln K/k_*}{f\phi_*}\right)+\mathcal{O}\left((f\phi_*)^2\right)\right]\,,
\end{multline}
where
\begin{equation}\label{eq:fres}
f^\text{res}=\frac{3\sqrt{2\pi}b}{8(f\phi_*)^{3/2}}\,.
\end{equation}
Once again, this result is valid at leading order in an expansion in $b$, which is constrained to be much less than one by comparison of the two point function with CMB data \cite{Flauger:2009ab}.\footnote{This is true provided the potential is monotonic, {\it i.e.} $b<1$.} $\phi_K$ is the value of the field at which modes with comoving momentum $K=k_1+k_2+k_3$ exit the horizon, $\phi_*$ is the value of the inflaton field when the pivot scale $k_*$ exits the horizon. Provided this happens around 60 $e$-folds before the end of inflation, one has $\phi_*\sim 11$. Finally the axion decay constant $f$ is a free parameter. A typical value in an explicit string construction is $10^{-4}\lesssim f\lesssim 10^{-1}$. The result is valid provided $f \phi_*\ll 1$. Logarithmic dependences on $k$ which arise from the dependence of the mode functions on $\epsilon_0$ as well as $\delta_0$ were neglected.
As the axion decay constant decreases, the frequency of the oscillations increases linearly and, keeping the parameter $\Lambda$ in the potential~\eqref{eq:V} fixed, the amplitude increases rapidly like $f^{5/2}$. It is a natural question to ask for what values of the axion decay constant an observably large signal can be generated in this model while satisfying the bounds on the power spectrum from the data. As long as $f<10^{-2}$, the bound at 95\% confidence level is summarized approximately by $bf<10^{-4}$~\cite{Flauger:2009ab}.
Combining this with equation~\eqref{eq:fres}, one finds that
\begin{eqnarray}
f^\text{res}\lesssim\frac{3\sqrt{2\pi}}{8f^{5/2}\phi_*^{3/2}}\times 10^{-4}\simeq\left(\frac{6\times 10^{-3}}{f}\right)^{5/2}\,.
\end{eqnarray}
The regime in which the model can simultaneously be consistent with the constraints on the two-point function and generate an observably large three-point function is thus $f\lesssim 6\times 10^{-3}$.
As we mentioned for the general case, requiring that the period of the oscillation should be larger than $\Delta\ell\sim 1$ for $\ell\approx200$ leads to a lower bound on $f$. For the linear case, it is $f\gtrsim 10^{-4}$. The range of axion decay constants for which observably large non-Gaussianities can be generated is thus approximately
\begin{eqnarray}\label{frange}
10^{-4}\lesssim f\lesssim 6\times 10^{-3}\,.
\end{eqnarray}
Notice that in this range $f\phi_*\ll1$ is always satisfied.
The shape of resonant non-Gaussianity for axion monodromy inflation is shown in Figure~\ref{fig:shape} for $b=10^{-2}$, $f\phi_*=2\times10^{-2}$, and fixed $k_1=k_*=0.002\,\text{Mpc}^{-1}$. We chose this value of $f$ because both the leading contribution and the subleading contribution in $f\phi_*$ are clearly visible. Notice that as the value of $k_1$ changes, the phase of the oscillation changes.
\begin{figure}[h]
\begin{center}
\includegraphics[width=5in]{shape}
\caption{This plot shows the shape $\mathcal{G}(k_1,k_2,k_3)/(k_1k_2k_3)$ of resonant non-Gaussianity for the linear potential of axion monodromy inflation with $b=10^{-2}$, $f\phi_*=2\times10^{-2}$ and fixed $k_1=k_*=0.002\,\text{Mpc}^{-1}$. We use the notation $x_2=k_2/k_1$ and $x_3=k_3/k_1$. The triangle inequality implies $x_2+x_3\leq 1$ and the quantity is symmetric under interchange of $x_2$ and $x_3$ so that we show in the plot only the region $1/2 \leq x_2 \leq 1$.}
\label{fig:shape}
\end{center}
\end{figure}
We find that our analytic result for $f^\text{res}$ agrees with the values obtained by numerical integration in~\cite{Hannestad:2009yx} at the per cent level.\footnote{For the comparison, notice that~\cite{Hannestad:2009yx} uses a momentum dependent quantity $\tilde{f}_{NL}$. In the equilateral limit, they extract their quantity $f_A=-\tilde{f}^{(eq)}_{NL}$. This quantity is related to our $f^\text{res}$ according to $f_A=10f^\text{res}/9$.}
\subsection{Consistency relation}
As pointed out in \cite{Maldacena:2002vr} (see also \cite{Creminelli:2004yq}), in the limit in which one of the momenta, say, $k_3$ is much less than the other two, which are then roughly equal, $k_3\ll k_1\approx k_2=k$, the three-point function is related to the two-point function by a consistency relation
\begin{eqnarray} \label{cons}
\lim_{k_3\rightarrow 0}\langle\mathcal{R}({\bf k_1},t)\mathcal{R}({\bf k_2},t)\mathcal{R}({\bf k_3},t)\rangle\simeq-|\mathcal{R}_{k_3}^{(o)}|^2\frac{1}{H(t_k)}\frac{d}{dt_k}\langle\mathcal{R}({\bf k_1},t)\mathcal{R}({\bf k_2},t)\rangle \,,
\end{eqnarray}
where $t_k$ is the time at which $k_1\approx k_2= k$ exit the horizon, and we will take $t$ to be some late time when the modes with comoving momenta $k_1$, $k_2$ and $k_3$ have exited the horizon. In our conventions the two-point function is given by
\begin{eqnarray}
\langle\mathcal{R}({\bf k_1},t)\mathcal{R}({\bf k},t)\rangle=(2\pi)^3\delta({\bf k_1}+{\bf k})|\mathcal{R}_{k}^{(o)}|^2\,.
\end{eqnarray}
Recall that
\begin{eqnarray}
\left|\mathcal{R}_k^{(o)}\right|^2=2\pi^2 \frac{\Delta_\mathcal{R}^2(k)}{k^3}\,,
\end{eqnarray}
and remember that $\Delta_\mathcal{R}^2(k)$ is momentum independent in the scale-invariant limit. It acquires its momentum dependence from the time dependence of the background. Modes with different momenta feel a different background as they exit the horizon. In equation~\eqref{cons}, we can thus make the replacement
\begin{eqnarray}
\frac{1}{H(t_k)}\frac{d}{dt_k}\to\frac{d}{d\ln k}\,,
\end{eqnarray}
with the derivative in $\ln k$ only acting on the momentum dependence in $\Delta_\mathcal{R}^2(k)$. Restoring the ${\bf k_3}$ dependence inside the $\delta$-function, the consistency relation can then be written in the form
\begin{eqnarray}\label{eq:conscond}
\lim_{k_3\rightarrow 0}\langle\mathcal{R}({\bf k_1},t)\mathcal{R}({\bf k_2},t)\mathcal{R}({\bf k_3},t)\rangle\simeq-(2\pi)^3\delta^3({\bf k_1}+{\bf k_2}+{\bf k_3})|\mathcal{R}_{k_3}^{(o)}|^2|\mathcal{R}_{k}^{(o)}|^2 \frac{d\ln\Delta_\mathcal{R}^2(k) }{d\ln k}\,.
\end{eqnarray}
The logarithmic derivative of the amplitude of scalar fluctuations~\eqref{eq:d2r} is given by
\begin{equation}
\frac{d \ln \Delta_\mathcal{R}^2}{d \ln k}\simeq n_s-1+\delta n_s\frac{\sqrt{2\epsilon_*}}{f} \sin \left(\frac{\phi_k}{f}\right)\;\;\;\;\text{with}\;\;\;\;\delta n_s=3b_*\left( \frac{2\pi f}{\sqrt{2\epsilon_*}}\right)^{1/2}\,.
\end{equation}
The consistency condition~\eqref{eq:conscond} together with the power spectrum given in Section 2 implies that the shape in this limit up to a phase in the trigonometric function should take the form
\begin{equation}
\frac{\mathcal{G}(k,k,k_3)}{k^2k_3}=\frac{3\sqrt{2\pi}b_*}{8}\left(\frac{f}{\sqrt{2\epsilon_*}}\right)^{1/2}\frac{2k}{k_3}\cos\left(\frac{\sqrt{2\epsilon_*}}{f}\ln 2k/k_*\right)\,.
\end{equation}
This agrees with our result for the resonant shape~\eqref{eq:Gresfin} after setting $k_1=k_2=k$ and taking the limit $k_3\ll k$.
Notice that in~\cite{Hannestad:2009yx}, the consistency relation \eqref{cons} was used to predict the shape of resonant non-Gaussianities in the squeezed limit. Here we have derived the resonant shape from first principles, and we use the consistency relation as a check of our computation.
\section{Correlations Between Resonant and Other Types of Non-Gaussianities}\label{s:cosine}
The comparison of theoretical models of non-Gaussianity with the data is computationally very challenging. In light of this difficulty, the authors of \cite{Babich:2004gb} have proposed to use a normalized scalar product, or ``cosine'', to assess to which extent two different 3D primordial shapes give rise to a similar 2D signal in the CMB. If two shapes have a large cosine (in absolute value), the observational constraints on one can be exported to the other.
Following this idea, in \cite{Fergusson:2008ra}, an appropriately defined cosine has been used to classify known non-Gaussian models. One class was reserved for models of canonically normalized single-field inflation where some feature is present on the top of an otherwise slow-roll flat potential. The models considered in this work belong to this class.
In this section we show that the shape of resonant non-Gaussianity that we have derived is very different from the shapes of non-Gaussianity that have been constrained by data. In particular we compute the correlation (to be defined soon) between resonant non-Gaussianity and local, equilateral and orthogonal non-Gaussianity and find that it is always less than about $10\%$. The observational constraints on these models are therefore not useful to constrain resonant non-Gaussianity.
\subsection{Scalar product, cosine and shapes}
In this subsection we give the definition of the cosine that we will use in the next subsections to compute the correlation between resonant non-Gaussianity and local, equilateral and orthogonal shapes.
Following \cite{Fergusson:2008ra}, we choose the simplest product that exhibits the same scaling as the optimal CMB estimator. The definition is
\begin{eqnarray} \label{sp}
F(S,S')=\int_{\mathcal{V}} S(k_1,k_2,k_3) S'(k_1,k_2,k_3) \frac{d\mathcal{V}}{K}\,,
\end{eqnarray}
where $d\mathcal{V}=dk_1dk_2dk_3$ and
\begin{eqnarray}
S(k_1,k_2,k_3)\propto \frac{\mathcal{G}(k_1,k_2,k_3)}{k_1k_2k_3}\,.
\end{eqnarray}
The normalization is irrelevant for our purpose because we will only be interested in the normalized scalar product or cosine of two shapes given by
\begin{eqnarray}\label{cosine}
C(S,S')\equiv\frac{F(S,S')}{\sqrt{F(S,S)F(S',S')}}\,.
\end{eqnarray}
Because of the rotational and translational symmetries of the background geometry, the volume of integration $\mathcal{V}$ is three dimensional. The volume three-form $d\mathcal{V}$ and the integration boundaries are conveniently written in the coordinates $\left\lbrace k,\alpha,\beta \right\rbrace$, which are related to $\left\lbrace k_1,k_2,k_3\right\rbrace$ by \cite{Fergusson:2006pr}
\begin{eqnarray}
k\equiv\frac{K}{2}=\frac12\left(k_1+k_2+k_3\right) \,, \quad k_1\equiv k\left(1-\beta\right)\,,\\
k_2\equiv \frac k2\left(1+\alpha+\beta\right)\,,\quad k_3\equiv \frac k2 \left(1-\alpha+\beta\right)\,.
\end{eqnarray}
One virtue of this set of coordinates is that for scale-invariant non-Gaussian models, {\it i.e.} $\mathcal{G}(k,k,k)\propto k^3$ or $S(k,k,k)\propto k^{0}$, the integrals over $dk$ cancel between the numerators and denominators in \eqref{cosine} and we are left with a two dimensional integral. However, resonant non-Gaussianity is not scale invariant (due to the sine and cosine in {\it e.g.} \eqref{Sres}) and having a three-dimensional integration volume is essential to get meaningful results for the cosine \eqref{cosine}.
The integration boundaries in \eqref{sp} require some discussion. In \cite{Fergusson:2008ra}, they were chosen to be $0\leq k\leq \infty$, $-(1-\beta)\leq\alpha\leq(1-\beta)$ and $0\leq \beta\leq1$. Notice that $\alpha$ and $\beta$ parametrize a triangle that is the base of a tetrahedron in the space $\left\lbrace k_1,k_2,k_3\right\rbrace$ with the apex at the origin and the semi-perimeter $k$ parameterizing the height. With the above integration limits, the product \eqref{sp} is typically infinite. Depending on the shapes that are being integrated, there can be an IR divergence for example where $\alpha=\pm 1$ or $\beta=1$, {\it i.e.} at the vertices of the $\left\lbrace \alpha,\beta \right\rbrace$ triangle where one of the $k_i$ vanishes. Local non-Gaussianity \eqref{loc} and resonant non-Gaussianity \eqref{Sres} show this divergence for squeezed configurations. In addition there are generically also UV divergences from the $dk$ integral\footnote{As we said, the UV divergence can be neglected in the case of scale-invariant shapes because the $dk$ integral simplifies in the cosine \eqref{cosine}.}.
Physically it is clear that for a given experiment, {\it e.g.}~observations of the CMB or of large scale structure (LSS), there is a finite range of momenta $\left\lbrace k_{\mathrm{min}},k_{\mathrm{max}}\right\rbrace$ that can be probed. The cosine \eqref{cosine} is useful if it compares two primordial non-Gaussian shapes over the same range of momenta that is probed by a chosen class of experiments. Implementing this is slightly subtle because of the three-dimensional nature of the primordial non-Gaussian shapes as opposed to the two-dimensional nature of the observations (see {\it e.g.} \cite{Babich:2004gb,Fergusson:2008ra} for a discussion). The interesting issue of defining a scalar product suitable for non-scale-invariant shapes is beyond the scope of this work, so we will limit ourselves to specify some $\left\lbrace k_{\mathrm{min}},k_{\mathrm{max}}\right\rbrace$ range of integration for $k_i$ and check that our results do not qualitatively depend on this choice.
Let us now turn to the shapes that we will consider. For resonant non-Gaussianity, we will work with the linear potential derived from the string theoretic construction. It is clear from \eqref{cosine} that the normalization of $S$ is irrelevant. We can thus define
\begin{eqnarray}\label{Sres}
S_\text{res}(k_1,k_2,k_3)\equiv\sin\left(\frac{\ln K}{f\phi_*}\right)+f\phi_*\cos\left(\frac{\ln K}{f\phi_*}\right) \sum_{i\neq j}\frac{k_i}{ k_j}\,.
\end{eqnarray}
We would like to know the correlation of $S_\text{res}$ with other shapes that have already been compared to and constrained by observations. If the cosine were close to one for some of them, we could export their constraints to resonant non-Gaussianity. The best constraints form 7-year WMAP data on local, equilateral, and orthogonal non-Gaussianity at $95\%$ CL are~\cite{Komatsu:2010fb}\footnote{For optimal limits from 5-year WMAP data see~\cite{Smith:2009jr,Senatore:2009gt}.}
\begin{eqnarray}
-10<f^\text{local}<74\,,\quad -214<f^\text{equil}<266\,,\quad -410<f^\text{ortho}<6\,.
\end{eqnarray}
These shapes are defined by
\begin{eqnarray}
S_\text{local}(k_1,k_2,k_3)&\equiv&\frac{k_1^3+k_2^3+k_3^3}{k_1 k_2 k_3}\,,\label{loc}\nonumber\\
S_\text{equil}(k_1,k_2,k_3)&\equiv&\frac{(k_1+k_2-k_3)(k_1+k_3-k_2)(k_3+k_2-k_1)}{k_1k_2k_3}\\\nonumber
&=&\left[-\frac{ k_3^2}{k_1k_2}-\frac{ k_1^2}{k_3 k_2}-\frac{ k_2^2}{k_1k_3}-2+\sum_{i\neq j}\frac{k_i}{ k_j}\right]\,,\\
S_\text{ortho}(k_1,k_2,k_3)& \equiv & 3 S_\text{equil}(k_1,k_2,k_3)-2\,,\nonumber \\
&=&\left[-\frac{3 k_3^2}{k_1k_2}-\frac{3 k_1^2}{k_3 k_2}-\frac{3 k_2^2}{k_1k_3}-8+3\sum_{i\neq j}\frac{k_i}{ k_j}\right]\,.\nonumber
\end{eqnarray}
Notice that these shapes are factorizable approximations to the results of the theoretical calculations (the reader is referred to \cite{Senatore:2009gt} for further details). They are good approximations in the sense that their cosine with the theoretical shapes is very close to one.
\subsection{Numerical results}
We have numerically calculated the cosine \eqref{cosine} between the shape of resonant non-Gaussianity \eqref{Sres} and local, equilateral and orthogonal shapes in \eqref{loc}. We have chosen $k_{\mathrm{min}}=10^{-4} \,\text{Mpc}^{-1}$ and $k_{\mathrm{max}}=10^{-1}\,\text{Mpc}^{-1}$ for the IR and UV cutoff, respectively. In order to implement the IR cutoff in the numerical calculation we have introduced three exponential damping factors $\exp(-k_{\mathrm{min}}/k_i)$ with $i=1,2,3$ in the integral \eqref{sp}. The UV cutoff is taken into account using the integration region $0<k\leq 3k_{\mathrm{max}}/2$. The results of this brute force approach are shown in Figure \ref{fig:num}. We have plotted the value of the three cosines as function of the axion decay constant $f$.
\begin{figure}[ht]
\begin{center}
\includegraphics[width=6.6in]{scatter}
\caption{From the bottom left panel clockwise we plot the cosine between the resonant shape \eqref{Sres}, and orthogonal, local and equilateral shapes \eqref{loc}, evaluated numerically for 50 points equally spaced in $\log f$. We also plot the enveloping profiles from the semi-analytical calculation to show that numerical and semi-analytical approaches give consistent results. In the bottom right panel, numerical results for all shapes are shown together. This makes it graphically clear that resonant non-Gaussianity is essentially orthogonal to the other shapes.\label{fig:num}}
\end{center}
\end{figure}
It is clear from the figure that resonant non-Gaussianity has a correlation smaller than about $10\%$ with the other shapes for any interesting value\footnote{The interesting $f$ range was discussed around \eqref{frange}.} of $f$. The cosines take both positive and negative values and get closer to zero as $f$ decreases. Both these features can be understood analytically and this is the subject of the next subsection.
\subsection{Semi-analytical results}
To better understand the points in Figure \ref{fig:num}, we have calculated the three cosines semi-analytically. We adopted a simplification in implementing the IR cutoff $k_i>k_{\mathrm{min}}$: we took as integration region $-(1-\beta)<\alpha<1-\beta$ and $k_{\mathrm{min}}/k_{\mathrm{max}}\leq \beta\leq 1-k_{\mathrm{min}}/k_{\mathrm{max}}$. This cuts off not only the vertices of the $\left\lbrace\alpha,\beta\right\rbrace$ triangle but also the side of the triangle at $\beta=0$. Given that none of the shapes we are considering has a divergent contribution along a side (as would be {\it e.g.}~the case for flat shapes obtained in the presence of deviations from a Bunch-Davies vacuum \cite{Holman:2007na,Meerburg:2009fi}), the result we get is a good approximation to the one obtained from cutting off only the vertices. As can be seen in Figure \ref{fig:num}, the results of this subsection nicely agree with those of the numerical approach described above, in which only the vertices had been cut off by the exponential damping factors $\exp(-k_{\mathrm{min}}/k_i)$. We have also checked that varying the cutoff in $\beta$ away from $10^{-3}$ by up to an order of magnitude changes the value of the cosines by less than a per cent. The cosines plotted in Figure \ref{fig:ana} as function of $f$ are computed using the above integration boundaries for $\alpha$ and $\beta$ and an integration range $3k_{\mathrm{min}} \leq k \leq 3 k_{\mathrm{max}} /2$ for $k$.
\begin{figure}[ht]
\begin{center}
\includegraphics[width=6.6in]{cosines}
\caption{From the bottom left panel clockwise we plot the cosine between the resonant shape \eqref{Sres}, and orthogonal, local, and equilateral shapes \eqref{loc}, as a function of $f$. In the bottom right panel, showing all the three embedding profiles at the same time, makes it graphically clear that resonant non-Gaussianity has less than about $10\%$ correlation with the other shapes. A linear fit to the enveloping curves is given in \eqref{numfit}.\label{fig:ana}}
\end{center}
\end{figure}
The very rapid oscillations are due to the $k$-independent phase $\phi_*/f$ and do not have a particular physical relevance given that the phase of the oscillations is arbitrary in the present model. On the other hand, the enveloping profiles, which are highlighted in Figures~\ref{fig:num} and \ref{fig:ana}, carry some interesting information. They show that resonant non-Gaussianity has a correlation smaller than about $10\%$ with local, equilateral and orthogonal models. The observational constraints on the latter are hence of little use in constraining resonant non-Gaussianity, the more so the smaller $f$. The enveloping profiles are well described by a linear fit that gives
\begin{eqnarray} \label{numfit}
|C(S_{res},S_\text{local})|&\leq& 1.9 f \phi_*\,,\\
|C(S_{res},S_\text{equil})|&\leq& 2.3 f \phi_*\,\\
|C(S_{res},S_\text{ortho})|&\leq& 1.3 f \phi_*\,.
\end{eqnarray}
This dependence can be understood as follows. Let us consider the cosine $C(S_{res},S)$ for some slowly varying shape $S$. There are three scalar products appearing in the definition \eqref{cosine} of $C(S_{res},S)$. Because of the oscillations of $S_{res}$, the scalar product $F(S_{res},S)$ can get a contribution from at most half a period, because an integration over one or more whole periods gives approximately zero. Notice that for any scalar product the largest contribution comes from the region around $k\sim k_{\mathrm{max}}$ because the overall scaling of \eqref{sp} is $F(S,S')\sim k_{\mathrm{max}}^2$. The periodicity of $S_{res}$ as function of $k$ around $k\simk_{\mathrm{max}}$ is given by $2\pi f \phi_* k_{\mathrm{max}}$. Hence $F(S_{res},S)$ is obtained effectively from integrating over a range of $k$ that is smaller than half a period, \textit{i.e.}~$\pi f \phi_* k_{\mathrm{max}}$. On the other hand, the scalar products $F(S_{res},S_{res})$ and $F(S,S)$ have either squared oscillations or no oscillations at all, respectively. This means that there is now no cancellation due to the oscillations and $F(S_{res},S_{res})$ and $F(S,S)$ get integrated over the whole range of $k$, {\it i.e.} approximately $k_{\mathrm{max}}$. Taking the ratio as in \eqref{cosine}, we see that the absolute value of the cosine $C(S_{res},S)$ can be bounded from above by $\pi f\phi_*$, up to a number smaller than but of order one. Let us say it in other words. If there were no oscillations, $S_{res}$ would be roughly approximated by some linear combination of local and equilateral shapes. Then, always ignoring oscillations, $C(S_{res},S)$ would generically be smaller than but of order one. The presence of oscillations gives the leading effect on the numerator $F(S_{res},S)$ where the range of the $dk$ integral is reduced by a factor $\pi f \phi_*$. Hence, we find again
\begin{eqnarray}
|C(S_{res},S)|&\lesssim \pi f \phi_*\,.
\end{eqnarray}
The result of this heuristic argument nicely agrees with the fit of the semi-analytic computation presented in \eqref{numfit}. Notice that the argument given above applies to the cosine of resonant non-Gaussianity with any slowly varying non-Gaussian shape and not just with those considered here.
\section{Discussion}\label{s:disc}
We have studied the primordial bispectrum of scalar perturbations for models whose potential possesses small modulations.
We do not deny that our work was largely motivated by a class of models derived from string theory that are based on axion monodromy in which such periodic modulations on top of an otherwise flat potential are a generic feature \cite{McAllister:2008hb, Flauger:2009ab,Berg:2009tg}. However, we argue that these are by no means the only models where such oscillations are expected to arise. In large field models of inflation, the inflaton potential is required to be flat over a range in field space much larger than $M_{\mathrm{Pl}}$. From the point of view of effective field theory, a potential that is flat over such a large range seems unnatural unless there is an underlying shift symmetry. This makes axions natural candidates for the inflaton especially in the context of large field inflation. If the inflaton is an axion, it seems plausible that the potential will receive small periodic contributions from non-perturbative effects. These periodic contributions might be due to instantons in a gauge sector the axion couples to, or, in the context of string theory, they might arise from Euclidean branes or world-sheet instantons. String inspired or not, as soon as we use the shift symmetry of axions to explain why the inflaton potential is so flat, we should admit the possibility of small periodic modulations in the potential which may lead to observational consequences.
So if theoretical prejudices have to be employed to isolate a handful of shapes of non-Gaussianity that should be looked for in the data, resonant non-Gaussianity deserves to be one of them. We hope that, even though it is not factorizable, the analytical expression for the shape of resonant non-Gaussianity given in this work will make it possible to obtain observational bounds on this type of non-Gaussianity.
The CMB data is compatible with a small logarithmic running of the power spectrum of primordial fluctuations. This has made it natural to consider phenomenological types of non-Gaussianity that deviate from scale invariance by at most a small logarithmic running. On the other hand, the theoretical considerations expressed above suggest that we should keep in mind the possibility of a scale dependence that reproduces scale invariance only after an adequate average.
Resonant non-Gaussianity typically comes with oscillations in the two-point function which are constrained by observations \cite{Flauger:2009ab}. Remarkably, keeping $b_*$ fixed, the amplitude of modulations in the two- and three-point function scales in opposite directions when varying the frequency. This implies that if oscillations were really imprinted on cosmological perturbations during inflation, they could equally well become observationally accessible in the two-point function, the three-point function, or in both. This disentanglement of two- and three-point functions is a peculiar feature of resonantly produced perturbations.
In~\cite{Cheung:2007st}, an effective field theory for the fluctuations around a quasi de Sitter background was constructed for the case of a single field. As is well known, for a single scalar field coupled to gravity, one can fix a gauge in which the fluctuations in the scalar field vanish, or are eaten up by the metric. As pointed out in~\cite{Cheung:2007st}, this gauge is very much like unitary gauge in a spontaneously broken gauge theory where the Goldstone has become the longitudinal mode of the gauge field. Also very much like in the gauge theory example, the longitudinal mode, or the Goldstone boson dominates the dynamics at high energies so that the non-linear theory of the Goldstone contains all the information about the system at sufficiently high energies. The theory of the Goldstone, even though non-linear, is easier to study and in particular makes relations between different operators transparent that would otherwise be obscure. The effective field theory of inflation as presented in~\cite{Cheung:2007st} or~\cite{Weinberg:2008hq} is the tool of choice if one is interested in high energy corrections, {\it i.e.} higher derivative corrections.
However, this high energy limit is equivalent to a limit in which all slow-roll parameters are taken to zero. Since the effect of resonant non-Gaussianity arises from a term in the interaction Hamiltonian that is higher order in the slow-roll expansion, it should not be surprising that it is not captured by the simplest version of the effective field theory of inflation. These terms could of course be kept~\cite{Cheung:2007sv}, but essentially at the cost of turning the effective field theory of inflation back into the system of a single scalar field coupled to gravity that we have studied here, written in a slightly different notation.
Finally, let us conclude with a couple of interesting directions for future research. We have focused our efforts on the three-point function in this work because it is the obvious observable to look for when looking for a departure from Gaussianity. The four-point function may also be of phenomenological interest in these models, and it can be calculated by the same methods presented here.
Our calculations have shown that the three-point function of primordial curvature perturbations may be large in models with periodically modulated potentials. For a comparison with the data, it still remains to calculate the prediction of the model for the two-dimensional image of the cosmic microwave background. We have seen that our shape is not factorizable. This makes a direct numerical evaluation too time consuming. However, it may be possible to make analytic progress in this direction.
Constraints on resonant non-Gaussianity could also arise from its effect on large scale structures. A very preliminary analysis shows that the effect on the halo bias discussed in \cite{Verde:2009hy} is relatively modest because it comes from the signal in squeezed configuration which is suppressed in our model by the small factor $f/\sqrt{2\epsilon_*}$. It would be interesting to consider other large scale structure observables.
\section*{Acknowledgments}
It is a pleasure to thank Richard Easther, Eiichiro Komatsu, Michele Liguori, Eugene Lim, Liam McAllister, Emiliano Sefusatti, and Gang Xu for many useful comments and discussions. The work of R.F. has been partially supported by the National Science Foundation under Grant No. NSF-PHY-0747868 and the Department of Energy under Grant No. DE-FG02-92ER-40704. The research of E.P. was supported in part by the
National Science Foundation through grant NSF-PHY-0757868.
|
\section{Introduction}
The $\sigma$ and $\kappa$ are controversial particles in hadron
spectroscopy. They were first found in the analysis of $\pi\pi$ and
$\pi K$ scattering data. Because the total phase shifts in the lower
mass region are much less than 180$^{\circ}$ and they do not fit into
ordinary $q \bar q$ meson nonets, they have been the subject of
violent debates. Refs.~(\cite{bes1s} - \cite{s4}) are some recent
analyses that support their existence.
Evidence for $\kappa$ particles comes from the study of production
processes and the re-analysis of $K \pi$ scattering data. Evidence
for the $\kappa$ has been found in $D^+ \to K^- \pi^+
\pi^-$~\cite{s0k}, $J/\psi \to \bar{K}^*(892)^0 K^+
\pi^-$~\cite{d1,bes2k}, and $D^+ \to K^- \pi^+ \mu^+ \nu$~\cite{d2}.
A $K \pi$ s-wave component is found in $D^0 \to K^- K^+
\pi^0$~\cite{d3}, $D^+ \to K^- \pi^+ e^+ \nu_e$~\cite{d3b}, and $\tau
\to K_S \pi^- \nu_{\tau}$~\cite{d4}. But no evidence for the
$\kappa$ in $D^0 \to K^- \pi^+ \pi^0$~\cite{d5} or for the charged
$\kappa$ in $D^0 \to K^- K^+ \pi^0$~\cite{d6} is seen. The $\kappa$
is found in the phenomenological analysis of $K \pi$ scattering phase
shift data~(\cite{s4},~\cite{c1} - \cite{hanqing}). However, some
theorists are not convinced by the evidence~(\cite{c8} - \cite{c11}).
The present status of the $\kappa$ is summarized by the Particle Data
Group PDG~\cite{pdg}.
The $\sigma$ and $\kappa$ particles have been studied with BES data,
where evidence for $\sigma$ and $\kappa$ particles is quite
clear~(\cite{bes1s} - \cite{bes2k}). A neutral $\kappa$ was found in
the decay $J/\psi \to \bar K^*(892)^0 \kappa^0 \to \bar K^*(892)^0 K^+
\pi^-$~\cite{d1,bes2k}. Because of isospin symmetry, if a neutral
$\kappa$ exists, a charged $\kappa$ should exist and could be produced
in $J/\psi \to \bar K^*(892)^{\pm} \kappa^{\mp}$.
In this study, we search for and study the charged $\kappa$ in $J/\psi
\to K^{\pm} K_S \pi^{\mp} \pi^0$. Our analysis is based on 58 million
$J/\psi$ decays collected by BESII at the BEPC (Beijing Electron
Positron Collider). BESII is a large solid-angle magnetic
spectrometer which is described in detail in Ref.~\cite{besd}. The
momentum of charged particles is determined by a 40-layer cylindrical
main drift chamber (MDC). Particle identification is accomplished
using specific ionization ($dE/dx$) measurements in the MDC and
time-of-flight (TOF) information in a barrel-like array of 48
scintillation counters. Outside of the TOF is a barrel shower counter
(BSC) which measures the energy and direction of photons.
\section{Event selection}
In the event selection, candidate tracks are required to have a good
track fit with the point of closest approach of the track to the beam
axis being within the interaction region of 2 cm in $r_{xy}$ and $\pm
20$ cm in $z$ (the beam direction), polar angles $\theta$ satisfying
$|\cos \theta|<0.80$, and transverse momenta $P_t> 50$ MeV/$c$.
Photons are required to be isolated from charged tracks, to come from
the interaction region, and have deposited energy in the BSC greater
than 40 MeV. Events are required to have four good charged
tracks with total charge zero and at least two good photons.
For the $K_S$ reconstruction, we loop over all oppositely
charged pairs of tracks, assuming them to be pions, and fit them to
$K_S \to \pi^+ \pi^-$, which determines vertices and $\pi^+\pi^-$
invariant masses, $M_{\pi\pi}$, for the four possible combinations.
The combination with $M_{\pi\pi}$ closest to $M_{K_S}$ is selected and is
required to satisfy $|M_{\pi^+ \pi^-} - M_{K_S}| < 20$ MeV/$c^2$ and
have its decay vertex in the $xy$-plane satisfy $r_{xy}>0.008$ m.
After $K_S$ selection, the particle type of the remaining two tracks,
that is whether they are $K^+ \pi^-$ or $K^- \pi^+$, is decided by
selecting the combination with smallest $\chi^2_{TOF} +
\chi^2_{DEDX}$.
A four constraint (4C) kinematic fit is applied under the $K^{\pm}
\pi^{\mp} \pi^+ \pi^- \gamma \gamma$ hypothesis, and $\chi^2_{4C} <
20$ is required. Events with a $\gamma \gamma$ invariant mass
satisfying $|M_{\gamma \gamma} - M_{\pi^0}| < 40$ MeV/$c^2$ are fitted
with a 5C kinematic fit to $K^{\pm} \pi^{\mp} \pi^+ \pi^- \pi^0$ with
the two photons constrained to the $\pi^0$ mass, and events with
$\chi^2_{5C} < 50$ are selected. The $\pi^{\mp} \pi^0$ mass
distribution is shown in Fig.~1(a), where the $\rho$ is clearly seen.
Decays with an intermediate $\rho$ are background,
and the requirement $|M_{\pi^{\mp} \pi^0} - M_{\rho}| >
100$ MeV/$c^2$ is applied to remove them. The requirements $| M_{K_S
\pi^0} - 0.897 | > 40$ MeV/$c^2$ and $| M_{K^{\pm} \pi^{\mp}} -
0.897 | > 40$ MeV/$c^2$ are used to remove backgrounds from $J/\psi
\to K^*(892)^0 K^{\pm} \pi^{\mp}$ and $J/\psi \to K^*(892)^0 K_S
\pi^0$. After these requirements, the combined $K^{\pm} \pi^0$ and
$K_S \pi^{\mp}$ mass distribution is shown in Fig.~1(b); the highest
narrow peak is charged $K^*(892)$. The $M_{K^{\pm}\pi^0}$ versus
$M_{K_S \pi^{\mp}}$ scatter plot is shown in Fig.~1(c). There are
two clear bands, a vertical and horizontal band, which
correspond to $J/\psi \to K^*(892)^{\pm} K_S \pi^{\mp} $ and $J/\psi
\to K^*(892)^{\pm} K^{\mp} \pi^0$, respectively. The requirements
$|M_{K^{\pm} \pi^0} - 0.892| < 80$ MeV/$c^2$ and $|M_{K_S \pi^{\pm}} -
0.892| < 80$ MeV/$c^2$ are imposed to select $J/\psi \to K^*(892)^{\pm} K_S
\pi^{\mp} $ and $J/\psi \to K^*(892)^{\pm} K^{\mp} \pi^0$ events,
respectively.
After the above selection, the combined $K_S \pi^{\mp}$ and $K^{\mp}
\pi^0$ invariant mass distribution recoiling against the
$K^*(892)^{\pm}$ is shown in Fig.~1(d). It is the sum of the
four decays listed in Table~1 with about 1000 events for each, as listed
in the table, and a total of 4121 events.
In Fig.~1(d), a clear narrow peak at 892 MeV/$c^2$ and a wider peak at
about 1430 MeV/$c^2$ are seen. In addition, there is a broad low mass
enhancement just above threshold. The spectrum is quite similar to
the spectrum of $K^+ \pi^-$ in the decay $J/\psi \to \bar{K}^*(892)^0
K^+ \pi^-$~\cite{bes2k}. The biggest difference between them is that
the charged $K^*(892)$ peak is much larger than the neutral $K^*(892)$
peak. The $K^*(892) \pi$ invariant mass distribution is shown in
Fig.~1(e), where there are indications of peaks at 1270 MeV/$c^2$ and
1400 MeV/$c^2$. The resulting Dalitz plot is shown in Fig.~1(f). The
two diagonal bands correspond to the low mass enhancement combined
with the 892 MeV/$c^2$ peak and the peak around 1430 MeV/$c^2$ in the
$K \pi$ spectrum.
\begin{table}[htp]
\begin{center}
\doublerulesep 0pt
\renewcommand\arraystretch{1.1}
\begin{tabular}{|c|c|}
\hline
Channel & Events \\
\hline
$J/\psi \to K^*(892)^+ K_S \pi^- \to K^+ \pi^0 K_S \pi^-$ & 1023 \\
\hline
$J/\psi \to K^*(892)^- K_S \pi^+ \to K^- \pi^0 K_S \pi^+$ & 946 \\
\hline
$J/\psi \to K^*(892)^+ K^- \pi^0 \to K_S \pi^+ K^- \pi^0$ & 1055 \\
\hline
$J/\psi \to K^*(892)^- K^+ \pi^0 \to K_S \pi^- K^+ \pi^0$ & 1097 \\
\hline
\end {tabular}
\vspace{0.1in}
\caption {Four signal channels and number of events from each.}
\end{center}
\end{table}
\begin{figure}[htbp]
\begin{flushleft}
{\mbox{\epsfig{file=rho.eps,height=2.6in,width=2.6in}}}
\end{flushleft}
\vspace{-2.6in}
\begin{flushright}
{\mbox{\epsfig{file=kpi.eps,height=2.6in,width=2.6in}}}
\end{flushright}
\begin{flushleft}
{\mbox{\epsfig{file=scat2.eps,height=2.6in,width=2.6in}}}
\end{flushleft}
\vspace{-2.6in}
\begin{flushright}
{\mbox{\epsfig{file=mass.eps,height=2.6in,width=2.6in}}}
\end{flushright}
\begin{flushleft}
{\mbox{\epsfig{file=k892pi.eps,height=2.6in,width=2.6in}}}
\end{flushleft}
\vspace{-2.6in}
\begin{flushright}
{\mbox{\epsfig{file=dalitz2.eps,height=2.6in,width=2.6in}}}
\end{flushright}
\caption[]{ (a) $\pi^{\mp} \pi^0$ mass distribution.
(b) Combined $K^{\pm} \pi^0$ and $K_S \pi^{\mp}$ mass
distribution after final cuts except the $K^*(892)$ cuts.
(c) Scatter plot of $M_{K^{\pm}\pi^0}$ versus $M_{K_S \pi^{\mp}}$.
(d) Invariant mass distribution of $K^{\pm} \pi^0$
and $K_S \pi^{\pm}$ recoiling against
$K^*(892)^{\mp}$. (e) Invariant mass distribution of $K^*(892)
\pi$. (f) Dalitz plot. }
\label{k01}
\end{figure}
\section{Background Studies}
Possible sources of background are studied. First,
sideband backgrounds are studied. The $\pi^+ \pi^-$ mass distribution
is shown in Fig.~2(a), where a clear $K_S$ signal can be seen and the
background level is quite low. The $K\pi$ spectrum from $K_S$
side-band events is shown in Fig.~2(b). (The $K_S$ side-band is
defined by 0.02 GeV/$c^2$ $<|M_{\pi\pi}-0.497|<$ 0.04 GeV/$c^2$.)
There are no clear structures.
The $\gamma \gamma$ spectrum is shown in Fig.~2(c), and the $K\pi$ spectrum of the 284 $\pi^0$ side-band events is shown in
Fig.~2(d). (The $\pi^0$ sideband is defined by 0.04 GeV/$c^2$ $<
|m_{\gamma \gamma}-0.135| < $ 0.08 GeV/$c^2$.)
The structures in Fig.~2(d) are
similar to those in the signal region and come from signal
events in the $\pi^0$ tails. The $K^*(892)$ side-band
background is shown by the dark shaded histogram in Fig.~2(e). (The
$K^*(892)$ side-band is defined by 0.08 GeV/$c^2$ $< | M_{K \pi}
-0.892 |<$ 0.16 GeV/$c^2$.) The clear $K^*(892)$ in the $K \pi$ mass
distribution of $K^*(892)$ side-band events
mainly comes from the cross channel. (There are two bands in the scatter
plot in Fig.~1(c). When we select one band, we will also select some events
from the other band where it crosses the first band. These events correspond
to cross channel background.) Fig.~2(f) shows the $K \pi$ spectrum after
side-band subtraction. The low mass enhancement and $K^*(892)$ peak
survive after side-band subtraction.
Next, we perform Monte Carlo simulation to study the main physics
background processes, including $J/\psi \to \gamma
\eta_c$ $\to \gamma K^* \bar{K}^*$ $\to \gamma K^{\pm} K_S \pi^{\mp}
\pi^0$, $J/\psi \to \gamma \eta_c$ $\to \gamma K^*(892)^{\pm} K^{\mp}
\pi^0$ $\to \gamma K^{\pm} K_S \pi^{\mp} \pi^0$, $J/\psi \to \gamma
\eta_c$ $\to \gamma K^*(892)^{\pm} K_S \pi^{\mp}$ $\to \gamma K^{\pm} K_S
\pi^{\mp} \pi^0$, $J/\psi \to \pi^0 \pi^+ \pi^- \pi^+ \pi^-$, and
$J/\psi \to \pi^0 K^+ K^- \pi^+ \pi^-$. The selection efficiencies
are much lower than $1\%$, and the largest
number of background events contributed is about 6.
Therefore, the physics background is quite low.
\begin{figure}[htbp]
\begin{flushleft}
{\mbox{\epsfig{file=ks_f.eps,height=2.6in,width=2.6in}}}
\end{flushleft}
\vspace{-2.6in}
\begin{flushright}
{\mbox{\epsfig{file=mass_ks-side.eps,height=2.6in,width=2.6in}}}
\end{flushright}
\begin{flushleft}
{\mbox{\epsfig{file=pi0_f.eps,height=2.6in,width=2.6in}}}
\end{flushleft}
\vspace{-2.6in}
\begin{flushright}
{\mbox{\epsfig{file=mass_pi0-side.eps,height=2.6in,width=2.6in}}}
\end{flushright}
\begin{flushleft}
{\mbox{\epsfig{file=mass_side.eps,height=2.6in,width=2.6in}}}
\end{flushleft}
\vspace{-2.6in}
\begin{flushright}
{\mbox{\epsfig{file=mass_sideoff.eps,height=2.6in,width=2.6in}}}
\end{flushright}
\caption[]{(a) $m_{\pi^+\pi^-}$ mass distribution after final event selection
except for the $K_S$ requirement. (b) $K_S$ side-band structure.
(c) $m_{\gamma \gamma}$ mass
distribution after final data selection except the $\pi^0$
requirement. (d) $M_{K \pi}$ distribution from $\pi^0$ side-band
events. (e) $K \pi$ spectrum
recoiling against $K^*(892)$. The dark shaded histogram is from
$K^*$ side-band events. (f) The $K \pi$
spectrum after side-band subtraction. }
\label{k02}
\end{figure}
\section{Partial wave analysis}
A partial wave analysis (PWA), which is based on the
covariant helicity amplitude analysis~(\cite{jacob} - \cite{wu03}),
is performed for the charged $\kappa$.
We add the likelihoods of all four channels together, and find the
minimum of the sum.
This is the same method as
used to fit $J/\psi \to K^*(892)^0 K^+ \pi^-$\cite{bes2k}.
The main difference here is that the decay
$J/\psi \to K^*(892)^{\pm} K^*(892)^{\mp}$ is
included in the fit.
In the PWA analysis, ten resonances, $\kappa$, $K^*_0(1430)$, $IPS$, $K_2^*(1430)$,
$K_2^*(1920)$, $K^*(1410)$ and $K^*(892)$ in the $K\pi$ spectrum,
$K_1(1270)$ and $K_1(1400)$ in the $K^*(892)\pi$ spectrum,
and $b_1(1235)$ in the $K^*(892) K $ spectrum, and two backgrounds,
listed as the last items in Table~2, are considered in the fit.
In Table~2, IPS (Interference Phase Space) refers to a broad
$0^{+}$ structure with a $K \pi$ invariant mass spectrum the same as
phase space,
that interferes with $\kappa$. $K^*$ BG refers to the $K^*(892)$ background coming from the cross
channel. PS (Phase Space) refers to the background with no
interference with resonances and with a shape almost the same as that of phase space.
\begin{table}[htp]
\begin{center}
\doublerulesep 0pt
\renewcommand\arraystretch{1.1}
\begin{tabular}{|c|c|c|c|c|c|}
\hline
\hline
\hline
{\small Resonance } & {\small Spin- } & {\small Decay}
& {\small Mass } &{\small Width }
& Sig.\\
& {\small Parity } & {\small Mode}
& {\small (MeV//$c^2$) } &{\small (MeV/$c^2$) }
& \\
\hline
$\kappa$ (1) & $0^{+}$ & $K\pi$ & $810 \pm 68 ^{+15}_{-24} $
& $536 \pm 87^{+106}_{-47}$ & $> 6 \sigma$ \\
\hline
$\kappa$ (2) & $0^{+}$ & $K\pi$ & $884 \pm 40 ^{+11}_{-22}$
& $478 \pm 77 ^{+71}_{-41}$ & $> 6 \sigma$ \\
\hline
$\kappa$ (3) & $0^{+}$ & $K\pi$ & $ 1165 \pm 58 ^{+120}_{-41}$
& $1349 \pm 500 ^{+472}_{-176}$ & $> 6 \sigma$ \\
\hline
$K^*_0(1430)$ & $0^{+}$ & $K\pi$ & $1400 \pm 86$
& $325 \pm 200$ & $0.6 \sigma$ \\
\hline
IPS & $0^{+}$ & $K\pi$ & --- & --- & $> 6 \sigma$ \\
\hline
$K^*_2(1430)$ & $2^{+}$ & $K\pi$ & $1411 \pm 30$
& $ 111 \pm 46$& $> 6 \sigma$
\\
\hline
$K^*_2(1920)$ & $2^{+}$ & $K\pi$ & $2020 \pm 140$
& $705 \pm 160 $& $> 6 \sigma$
\\
\hline
$K^*(1410)$ & $1^{-}$ & $K\pi$ & $1420 \pm 14$
& $130 \pm 28$& $> 6 \sigma$ \\
\hline
$K^*(892)$ & $1^{-}$ & $K\pi$ & $ 896 \pm 8$
& $57 \pm 12$ &$> 6 \sigma$ \\
\hline
$K_1(1270)$ & $1^{+}$ & $K^*(892) \pi$ & $1254 \pm 14$
& $60 \pm 28$& $> 6 \sigma$ \\
\hline
$K_1(1400)$ & $1^{+}$ & $K^*(892) \pi$ & $1390 \pm 30 $
& $146 \pm 44$& $> 6 \sigma$ \\
\hline
$b_1(1235)$ & $1^{+}$ & $K^*(892) K$ & $1230 \pm 52$
& $142 \pm 38$ &$ 4.5 \sigma$ \\
\hline
$K^*(892)$ BG & $1^{-}$ & $K\pi$ & --- & --- & $> 6 \sigma$ \\
\hline
PS BG & --- & --- & --- & --- & --- \\
\hline
\hline
\end {tabular}
\vspace{0.1in}
\caption {Resonances included in the fit of this channel.
Masses and widths of various resonances are determined by
mass and width scans. $\kappa$ (1), (2) and (3) are results
given by fits using Breit-Wigner functions (1), (2) and (3)
to fit $\kappa$ respectively. IPS refers to the broad $0^+$ structure which interferes with $\kappa$.
$K^*$ BG refers to the $K^*(892)$ background coming from the cross
channel. PS BG refers to the background with
no interference with resonances.}
\end{center}
\end{table}
Three different parameterizations
are used to fit the $\kappa$. They are
\begin{equation}\label{b1}
BW_{\kappa} = \frac{1}{m_{\kappa}^2 -s - i m_{\kappa} \Gamma_{\kappa}},
~~~~~ \Gamma_{\kappa} = {\rm constant} ,
\end{equation}
\begin{equation}\label{b2}
BW_{\kappa} = \frac{1}{m_{\kappa}^2 -s - i \sqrt{s} \Gamma_{\kappa}(s)},
~~~~~ \Gamma_{\kappa}(s) =
\frac{g_{\kappa}^2 \cdot k_{\kappa}}{8 \pi s} ,
\end{equation}
\begin{equation}\label{b3}
BW_{\kappa} = \frac{1}{m_{\kappa}^2 -s - i \sqrt{s} \Gamma_{\kappa}(s)},
~~~~~ \Gamma_{\kappa}(s) = \alpha \cdot k_{\kappa} ,
\end{equation}
where $k_{\kappa}$ is the magnitude of the $K$ momentum
in the $K \pi$, or the $\kappa$,
center of mass system~\cite{hanqing2}, and $\alpha$ is
a constant which will be determined by the fit.
Parameters in the Breit-Wigner function are determined
by mass and width scans. The minima of the scan curves give
the central values of mass and width parameters.
From these, the corresponding Breit-Wigner pole positions can be
directly calculated from equation (1), (2) and (3).
Our final results are listed in Table~3, where
the first errors are statistical, and
the second are systematic.
The mass and width parameters
obtained by different parameterizations are quite different,
but their poles are almost the same, which
is quite similar to what was found in the study of the neutral
$\kappa$. The results for the neutral
$\kappa$~\cite{bes2k} are shown in Table~4 and are consistent with the
charged $\kappa$.
\begin{table}[htp]
\begin{center}
\doublerulesep 0pt
\renewcommand\arraystretch{1.1}
\begin{tabular}{|c|c|c|c|}
\hline
& BW (1) & BW (2) & BW (3) \\
\hline
Mass (MeV/$c^2$) & $810 \pm 68 ^{+15}_{-24} $
& $884 \pm 40 ^{+11}_{-22} $ &
$1165 \pm 58 ^{+120}_{-41} $ \\
\hline
Width (MeV/$c^2$) & $ 536 \pm 87^{+106}_{-47} $
& $478 \pm 77 ^{+71}_{-41} $
& $1349 \pm 500 ^{+472}_{-176} $ \\
\hline
pole (MeV/$c^2$) & $(849 \pm 77 ^{+18}_{-14} )$
& $(849 \pm 51 ^{+14}_{-28} $)
& $(839 \pm 145 ^{+24}_{-7})$ \\
& $-i(256 \pm 40 ^{+46}_{-22} )$
& $-i (288 \pm 101 ^{+64}_{-30} )$
& $-i(297 \pm 51 ^{+50}_{-18})$ \\
\hline
\end {tabular}
\vspace{0.1in}
\caption {Masses, widths and pole positions of the charged
$\kappa$. In the table, the first errors
are statistical, and the second are systematic.
BW (1) means equation (1) is
used to fit the $\kappa$. BW (2) and BW (3)
have similar meanings. }
\end{center}
\end{table}
\begin{table}[htp]
\begin{center}
\doublerulesep 0pt
\renewcommand\arraystretch{1.1}
\begin{tabular}{|c|c|c|c|}
\hline
& BW (1) & BW (2) & BW (3) \\
\hline
Mass (MeV/$c^2$) & $745 \pm 26^{+14}_{-91}$
& $ 874 \pm 25 ^{+12}_{-55}$ & $1140 \pm 39 ^{+47}_{-80}$ \\
\hline
Width (MeV/$c^2$) & $622 \pm 77 ^{+61}_{-78} $
& $518 \pm 65 ^{+27}_{-87}$ & $1370 \pm 156 ^{+406}_{-148}$ \\
\hline
pole (MeV/$c^2$) & $(799 \pm 37 ^{+16}_{-90})$
& $(836 \pm 38 ^{+18}_{-87})$
& $(811 \pm 74 ^{+17}_{-83})$ \\
& $-i(290 \pm 33 ^{+25}_{-38})$
& $-i(329 \pm 66 ^{+28}_{-46})$
& $-i(285 \pm 20 ^{+18}_{-42})$ \\
\hline
\end {tabular}
\vspace{0.1in}
\caption {Masses, widths and pole positions of the neutral
$\kappa$~\cite{bes2k}. BW (1) means equation (1) is
used to fit the $\kappa$. BW (2) and BW (3)
have similar meaning.}
\end{center}
\end{table}
Our final results correspond to
the solution with the minimum least likelihood. Differences among
solutions with similar likelihood values are included as
systematic uncertainties. Also included are
the effect of removing $K_0(1430)$,
IPS, $b_1(1235)$ and $K^*(892) \bar{K}^*(892)$, the result of a
fit with the $K^*(892)$ background level floating, and the result from
a fit using direct side-band subtraction.
The masses and widths of all resonances obtained
by mass and width scans are shown in Table~2.
In the fit, the contribution from $K^*_0(1430)$
is small; its statistical significance is only 0.6$\sigma$.
Because it is expected in this channel, it is included in the
final solution.
The $K \pi $ mass distribution is shown in Fig.~3(a), where points with
error bars are data, and the light shaded histogram is the final fit.
In the figure,
the dark shaded histogram shows the contribution of the charged
$\kappa$. Fig.~3(b) shows the fit for the
$K^*(892) \pi $ spectrum, and Fig.~4 shows the
angular distributions.
\begin{figure}[htbp]
\begin{flushleft}
{\mbox{\epsfig{file=fit-kpi.eps,height=2.6in,width=2.6in}}}
\end{flushleft}
\vspace{-2.6in
\begin{flushright}
{\mbox{\epsfig{file=fit-892pi.eps,height=2.6in,width=2.6in}}}
\end{flushright}
\caption[]{(a) Final fit results for the $K \pi $ spectrum. Points
with error bars
are data,
the light shaded histogram is the final global fit, and the dark
shaded histogram is the contribution of the $\kappa$.
(b) Final fit of $K^*(892) \pi $ spectrum.
Dots with error bars are data, and
the histogram is the final global fit. There are two peaks
in the lower mass region. The lower one is fit by the $K_1(1270)$,
and the higher one is by the $K_1(1400)$. }
\label{k05}
\end{figure}
\begin{figure}[htbp]
\begin{flushleft}
{\mbox{\epsfig{file=fit-angle_t1.eps,height=2.6in,width=2.6in}}}
\end{flushleft}
\vspace{-2.6in}
\begin{flushright}
{\mbox{\epsfig{file=fit-angle_p1.eps,height=2.6in,width=2.6in}}}
\end{flushright}
\begin{flushleft}
{\mbox{\epsfig{file=fit-angle_t2.eps,height=2.6in,width=2.6in}}}
\end{flushleft}
\vspace{-2.6in}
\begin{flushright}
{\mbox{\epsfig{file=fit-angle_p2.eps,height=2.6in,width=2.6in}}}
\end{flushright}
\caption[]{ Final fit result for the angular distributions.
Dots with error bars are data, and the histogram is the final global
fit. The upper left figure shows the final fit of the $\theta_1$
distribution, the upper right shows the final fit of the $\phi_1$
distribution, the lower left shows the final fit of the $\theta_2$
distribution, the lower right shows the final fit of the $\phi_2$
distribution. $\theta_1$ and $\phi_1$ are the polar angle and
azimuthal angle of the $K \pi$ system in the $J/\psi$ center of mass
system. $\theta_2$ and $\phi_2$ are the polar angle and azimuthal
angle of the $K$ meson in the $K \pi$ center of mass system. }
\label{k06}
\end{figure}
\section{Branching ratio measurements}
The decay
$J/\psi \to K^*(892)^{\pm} \kappa^{\mp}$
contributes 655 events. Monte Carlo simulation
of
$J/\psi \to K^*(892)^{\pm} \kappa^{\mp}
\to K^{\pm} K_S \pi^{\mp} \pi^0$
determines an efficiency of 2.33\%,
and the branching ratio of
$J/\psi \to K^*(892)^+ \kappa^-$ or
$J/\psi \to K^*(892)^- \kappa^+$
is
\begin{equation}
BR = \frac{655/4}{2.33\% \times 5.8\times 10^7 \times 1/9 }
= ( 1.09 \pm 0.18 ^{+0.94}_{-0.54} ) \times 10^{-3},
\end{equation}
where, the first error is statistical, the second error
is systematic, the factor $\frac{1}{4}$ is because the events
are the sum of four channels, and $\frac{1}{9}$ is
the isospin factor.
In the previous study of the decay
$J/\psi \to \bar{K}^*(892)^0 K^+ \pi^-$,
the corresponding branching ratio for the neutral $\kappa$
\footnote{The branching ratio of the neutral $\kappa$ was not reported in
Ref.~\cite{bes2k}. In the study of the neutral $\kappa$, the number
of $\kappa$ events was in the range 1891 - 3516, and
the selection efficiency was 14.2\%. Its branching ratio is
$\frac{1891 - 3516}{14.2\% \times 5.8 \times 10^7 \times 4/9} =$
$(0.52 - 0.97) \times 10^{-3}$}
is $(0.52 - 0.97) \times 10^{-3}$.
The two results are consistent
with isospin symmetry.
The systematic error includes
uncertainties from multi-solutions, from different background fit
methods, and from removing some components
from the fit ($K^*_0(1430)$, IPS, $b_1(1235)$,
and $K^*(892)^{\pm} K^*(892)^{\mp}$).
Also interesting is the existence
of the $J/\psi$ electromagnetic decay
$J/\psi \to K^*(892)^{\pm} K^*(892)^{\mp}$.
The peak at 892 MeV/$c^2$ in the $K \pi$ invariant
mass spectrum (see Fig.~1(d)) comes from both
cross channel background and
from $J/\psi \to K^*(892)^{\pm} K^*(892)^{\mp}$.
The number of events from the cross channel
can be calculated approximately from the $K^*(892)$
side-band structure, shown in
Fig.~2(e), where the narrow peak at
892 MeV/$c^2$ is clear. In the final fit, the decay
$J/\psi \to K^*(892)^{\pm} K^*(892)^{\mp}$
contributes 323 events. The Monte Carlo simulation
of the decay
$J/\psi \to K^*(892)^{\pm} K^*(892)^{\mp}
\to K^{\pm} K_S \pi^{\mp} \pi^0$
yields an efficiency of 1.25\%,
and the branching ratio of
$J/\psi \to K^*(892)^+ K^*(892)^-$
is
\begin{equation}
BR = \frac{323/2}{1.25\% \times 5.8\times 10^7 \times 1/9 \times 2}
= ( 1.00 \pm 0.19 ^{+0.11}_{-0.32} ) \times 10^{-3},
\end{equation}
where, the first error is statistical, the second error
is systematic, the factor $\frac{1}{2}$ is because events are counted
twice, $\frac{1}{9}$ is the isospin
factor, and 2 is because the data
comes from two decay channels:
$J/\psi \to K^*(892)^+ K^*(892)^-
\to (K^+ \pi^0) (K_S \pi^-) $
and
$J/\psi \to K^*(892)^- K^*(892)^+
\to (K^- \pi^0) (K_S \pi^+).$
The systematic error includes
uncertainties from multi-solutions, from different background fit
methods, and from removing some components
from fit($K^*_0(1430)$, IPS, and $b_1(1235)$).
\section{Summary}
In conclusion, the charged $\kappa$ is observed
and studied in the decay
$J/\psi \to K^*(892)^{\pm} \kappa^{\mp}
\to K^{\pm} K_S \pi^{\mp} \pi^0$.
The low mass enhancement in the $K \pi$ spectrum cannot be
fit well unless a charged $\kappa$ is added into the solution.
If we use a Breit-Wigner
function of constant width to parameterize the $\kappa$, its
pole locates at
$(849 \pm 77 ^{+18}_{-14}) -i (256 \pm 40 ^{+46}_{-22})$ MeV/$c^2$.
In our analysis, three different $\kappa$ parameterizations
are tried in the fit, and final results are shown in Table~3 and
are consistent with those of the neutral $\kappa$ and are also
in good agreement with those obtained in the analysis of
$K \pi$ scattering phase shifts.
Also, the decay $J/\psi \to K^*(892)^{\pm}
K^*(892)^{\mp}$ is observed for the first time with
the branching ratio
$(1.00 \pm 0.19 ^{+0.11}_{-0.32}) \times 10^{-3}$.
The corresponding decay mode is not observed
in $J/\psi \to \bar{K}^*(892)^0 K^+ \pi^-$.
The decays
$J/\psi \to K^*(892)^{\pm} K^*(892)^{\mp} $
can be produced through $J/\psi$ electromagnetic
decays, while $J/\psi \to K^*(892)^0 \bar{K}^*(892)^0 $
can only be produced through $J/\psi$ hadronic decays, which
would be $SU(3)$ symmetry breaking decays and are suppressed.
\vspace{3mm}
{\bf Acknowledgments}
The BES Collaboration thanks the staff of BEPC and computing
center for their hard efforts. This work is supported in part by
the National Natural Science Foundation of China under contract
Nos. 10491300, 10225524, 10225525, 10425523, 10625524, 10521003, 10821063,
10825524, the Chinese Academy of Sciences under contract No. KJ 95T-03, the
100 Talents Program of CAS under Contract Nos. U-11, U-24, U-25,
and the Knowledge Innovation Project of CAS under contract Nos.
U-602, U-34 (IHEP), the National Natural Science Foundation of
China under contract Nos. 10775077, 10225522 (Tsinghua University), and the
Department of Energy under Contract No. DE-FG02-04ER41291 (U.
Hawaii).
|
\section{\label{sec:introduction}Introduction}
\noindent Colonies of swimming bacteria, \emph{in vitro} mixtures of cytoskeletal filaments and motor proteins, and vibrated granular rods are examples of \emph{active} systems composed of interacting units that consume energy and collectively generate motion and mechanical stresses. Due to their elongated shape, active particles can exhibit orientational order at high concentration and have been likened to ``living liquid crystals"~\cite{Gruler1999}. Their rich collective behavior includes nonequilibrium phase transition and pattern formation on mesoscopic scales. It has been modeled by continuum equations built by modifying the hydrodynamics of liquid crystals to include nonequilibrium terms that account for the activity of the system \cite{TonerRev,SimhaRamaswamySP02,Kruse2004}, or derived from specific microscopic models~\cite{TBLMCM2003,TBLMCMbook}.
A striking property of \emph{confined} active liquid crystals is the instability of the uniform aligned homogeneous state and the onset of spontaneously flowing states, both stationary and oscillatory~\cite{Voituriez06,GiomiMarchettiLiverpool:2008}. This occurs because local orientational order generates active stresses that are in turn balanced by flow, yielding a state that can support local inhomogeneities in the flow velocity and the local alignment, while maintaining a net zero force. Loosely speaking, a confined active liquid crystal ``shears itself'' even in the absence of externally applied forces. It is then not surprising that the rheology of such active liquid crystals in response to an external shear will be very rich.
\begin{figure}[b]
\centering
\includegraphics[width=.9\columnwidth]{img/active-flow3.pdf}
\caption{\label{fig:bacteria}(color online) Schematic example of the flow field surrounding a tensile (left) and contractile (right) swimming microorganism.}
\end{figure}
Phenomenological work by Hatwalne and collaborators~\cite{Hatwalne04} first pointed out that activity lowers the linear bulk viscosity of tensile suspensions, such as most swimming bacteria, while it enhances the viscosity of contractile systems, and that this enhancement may become very large near the isotropic-nematic transition. A semi-microscopic model of contractile suspensions of motor-filaments mixtures confirmed these results and predicted an actual divergence of the viscosity of contractile suspensions at the transition~\cite{TBLMCM06}. Recent numerical studies of active nematic films by Cates \emph{et al}. \cite{CatesEtAl:2008} have confirmed that this result survives when the effect of boundaries is included. In addition, it was found that tensile nematic suspensions can enter a regime of vanishing apparent viscosity in proximity of the isotropic-nematic phase transition. Such a ``superfluid'' window was interpreted by the authors of Ref. \cite{CatesEtAl:2008} as the appearance of bulk shear bands accommodating a range of macroscopic shear-rates at zero stress. Finally, the predicted activity-induced thinning of bacterial suspensions has been demonstrated in recent experiments in \emph{Bacillus subtilis}~\cite{Haines2008,SokolovAranson2009,Haines2009}.
Active particles exert forces on the surrounding fluid, resulting in local tensile or contractile stresses proportional to the amount of orientational order, $\sigma_{ij}^{\alpha}\sim\alpha n_{i}n_{j}$, where $\alpha$ is proportional to the force exerted by the active particles on the fluid and ${\bf n}$ a unit vector denoting the direction of broken orientational symmetry. The sign of $\alpha$ determines whether the flow generated by the active particles is tensile ($\alpha<0$) or contractile ($\alpha>0$). In the case of swimming organisms, the former situation describes ``pushers'', i.e., most bacteria (e.g., E. Coli), while the latter corresponds to ``pullers'' (e.g., Chlamydomonas) (see Fig. \ref{fig:bacteria}).
An important distinction between uniaxial active particles concerns the possibility of forming phases with or without a non-zero macroscopic polarization. Apolar particles are fore-aft symmetric and can form nematic phases in which macroscopic quantities are invariant for ${\bf n}\rightarrow -{\bf n}$. Polar particles can also form phases characterized by a non-zero macroscopic polarization in the direction of a polar director ${\bf p}$ in which they undergo collective motion with mean velocity ${\bf v}\sim \beta\,{\bf p}$, with $\beta$ is the typical self-propulsion velocity. This directed motion occurring in polar suspensions contributes to a non-equilibrium local stress of the form $\sigma_{ij}^{\beta}\sim\beta\,(\partial_{i} p_{j}+\partial_{j} p_{i})$.
Most theoretical work has focused on the rheology of active nematic ($\beta=0$), while the shear response of active polar suspensions is far less explored~\cite{Haines2008,Haines2009}. We find that for a fixed value of $\beta$, the behavior of active suspensions depends on the interplay between the local contractile/tensile stresses, embodied in the parameter $\alpha$, and the flow-aligning behavior of liquid crystalline particles, described by the flow alignment parameter, $\lambda$ \cite{EdwardsYeomans:2009}. Rod-shaped particles typically have $\lambda>0$, spherical particles have $\lambda=0$, while the case $\lambda<0$ describes disk-shaped molecules such as those found in discotic liquid crystals. In passive liquid crystals the magnitude of $\lambda$ controls how the director field responds to a large shear flow away from boundaries. For $|\lambda|>1$ the director tends to align to the flow direction at an angle $\theta_{0}$ such that $\cos 2\theta_{0}=1/\lambda$, while for $|\lambda|<1$ it forms rolls throughout the systems. These regimes are known as ``flow-aligning'' and ``flow-tumbling'' respectively. Understanding of the complex rheology of polar and nematic active suspensions requires exploring the full parameter space, including the important role of boundary conditions. One of the important results of this work is a remarkable exact duality that holds in the regime where the stress-strain relation is linear and shows that tensile ($\alpha<0$) rod-shaped flow-aligning particles ($\lambda>1$) are rheologically equivalent to contractile ($\alpha>0$) discotic flow-tumbling particles ($-1\leq\lambda<0$). Using this result, we present below a unified description of the linear rheology of active suspensions of both polar and apolar particles. Some of the results are summarized in the ``phase diagram" of Fig.~\ref{fig:alpha-lambda}.
\begin{figure}[t]
\centering
\includegraphics[width=.85\columnwidth]{img/alpha-lambda.pdf}
\caption{\label{fig:alpha-lambda} (color online) The figure displays the regions of parameters where spontaneous flow occurs in an unsheared active film on a substrate. The regions of spontaneous flow are bounded by the critical activity $\alpha_{c1}(\beta)$ given in Eq.~\eqref{alphac1} (solid and dashed lines) and are shaded orange, with lighter shades corresponding to increasing values of $\beta$. The same critical activity also separates the regions $|\alpha|<\alpha_{c1}$ where the theoretical stress-strain curves are monotonic and the active suspension is either thinned or thickened by activity at small shear rates, as indicated, from the regions $|\alpha|>\alpha_{c1}$ where the theoretical stress-strain curves are nonmonotonic, with possible ``superfluid'' or hysteretic behavior.}
\end{figure}
This figure shows that the rheological properties of an active film subject to an external shear are closely related to the onset of spontaneous flow in the absence of shear, highlighting the parallel role played in active system by mechanical driving forces, such as a macroscopic strain rate, and internal active driving forces proportional to $\alpha$ and $\beta$.
An unsheared active film exhibits a transition from the homogeneous aligned state to a ``spontaneously flowing'' state, characterized by spatially inhomogeneous velocity and director profiles~\cite{Voituriez06}. The transition occurs at a critical activity $\alpha_{c1}$ in a film bounded by one no-slip substrate and a surface that can freely slide, and at a larger value, $\alpha_{c2}>\alpha_{c1}$, in a film bounded by two no-slip planes. The lines separating regions of different shades in Fig.~\ref{fig:alpha-lambda} are the boundaries $\alpha_{c1}(\beta,\lambda)$ [see Eq.~\eqref{alphac1} below] separating regions of spontaneous flow ($|\alpha|>\alpha_{c1}$) from regions where the homogeneous aligned state is stable ($|\alpha|<\alpha_{c1}$). Interestingly, when the film is subject to an external shear, we find that the flow properties change their {\em qualitative} behaviour at exactly these same critical values of activity. For $\alpha_{c1}<|\alpha|<\alpha_{c2}$, the theoretical stress-strain rate curves obtained from our one dimensional model are nonmonotonic (see Fig.~\ref{fig:stress-strain2}) and the active suspension is strongly non-Newtonian. We suggest a number of different interpretations of the nonmonotonic part of the stress-strain rate curve shown in Fig.~\ref{fig:scenarios}. These include macroscopic ``superfluid-like'' behaviour~\cite{CatesEtAl:2008} with zero effective viscosity, yield-stress behaviour or hysteresis. Finally, for $|\alpha|>\alpha_{c2}$, the theoretical stress-strain curve has a discontinuous jump at zero strain rate, corresponding to a finite ``spontaneous stress'' in the absence of applied shear~\cite{TBLMCM06}.
\begin{figure}[t]
\centering
\includegraphics[width=0.8\columnwidth]{img/sketch.pdf}
\caption{\label{fig:sketch}(color online) Schematic representation of a quasi-one-dimensional film of thickness $L$. In our model the film is sitting on a non-slipping surface and is sheared from the top at constant velocity $v_{0}$. The polar rods form an angle $\theta$ with respect to the infinite direction $x$ of the film. Because of the quasi-one-dimensional geometry, the system is invariant for translations along the $x$ axis.}
\end{figure}
\section{The Model}
Our model of active suspension consists of a two-dimensional film of rod-like particles of length $\ell$ confined to a channel of infinite length along the $x$ axis and finite thickness $L$ along the $y$ axis (see Fig. \ref{fig:sketch}). Because of the chosen geometry, the system is invariant for translations along the $x$ axis. The total density of the suspension, $\rho=Mc+\rho_{\rm solvent}$, with $c$ the concentration of active particles and $M$ their mass, is assumed to be constant, thus $\nabla\cdot{\bf v}=0$, with ${\bf v}$ the flow velocity. We assume that the film is sheared at a constant (macroscopic) rate $\dot{\gamma}$ by keeping the lower plate at $y=0$ fixed, while the upper plate at $y=L$ is moved at constant velocity $v_{0}$. The macroscopic shear-rate is defined then as $\dot\gamma=v_{0}/L=\int_{0}^{L}(dy/L)\,u$, where the rate-of-strain tensor $u_{ij}=(\partial_{i}v_{j}+\partial_{j}v_{i})/2$ has only non-zero components $u_{xy}=u_{yx}=\partial_{y}v_{x}/2\equiv u/2$. Theoretical stress-strain curves are obtained by fixing the macroscopic strain rate $\dot\gamma$ and calculating the resulting stress $\sigma$.
We consider a polarized active suspension and focus only on spatial variations in the direction of the polarization ${\bf P}$. The hydrodynamic equations for an active polar suspension have been formulated by incorporating the active contributions (proportional to the rate of energy consumed by the active units) into the hydrodynamic equations of a passive polar liquid crystalline film.
Some of the active contributions, discussed above, are not allowed by the conditions which define liquid crystal systems at equilibrium and hence are {\em intrinsic} to active systems. Other terms have the same form as those of passive polar liquid crystals and can simply be included by modifying the prefactors of the terms obtained from a passive systems. As such, the {modified} ``passive'' contributions to the equations of motion can be described starting from the non-equilibrium analogue of the Frank free-energy of a suspension of polar particles in a solvent:
\begin{multline*}
F = \int_{\bf r}\,\Big\{
\frac{C}{2}\left(\frac{\delta c}{c_{0}}\right)^{2}+\frac{a_{2}}{2}|{\bf P}|^{2}+\frac{a_{4}}{4}|{\bf P}|^{4}+\frac{K_{1}}{2}(\nabla\cdot{\bf P})^{2}\\[5pt]
+\frac{K_{3}}{2}(\nabla\times{\bf P})^{2}+B_{1}\frac{\delta c}{c_{0}}\,\nabla\cdot{\bf P}+B_{2}|{\bf P}|^{2}\nabla\cdot{\bf P} + \frac{B_{3}}{c_{0}}|{\bf P}|^{2}{\bf P}\cdot\nabla c
\Big\}\,,
\end{multline*}
with $C$ the compressional modulus and $K_{1}$ and $K_{3}$ the splay and bend elastic constant. The parameters $a_i,B_i,K_i,C$ are understood to have both passive and active contributions. In the following we will take $K_{1}=K_{3}=K$. The last three terms in the expression of the free-energy couple concentration and splay and are also present in equilibrium polar suspensions.
The dynamics of the concentration and the polarization are described by
\begin{subequations}
\begin{equation}
\partial_{t}c = -\nabla\cdot\left[c({\bf v}+c \beta_1 {\bf P})+\Gamma'{\bf h}+\Gamma''{\bf f}\right]\,,\label{eq:c1}\\[-10pt]
\end{equation}
\begin{multline}
[\partial_{t}+({\bf v} +c \beta_2{\bf P})\cdot\nabla]P_{i}+\omega_{ij}P_{j} \\= \lambda u_{ij} P_{j}+\Gamma h_{i}+\Gamma'f_{i}\,,\label{eq:p1}
\end{multline}
\end{subequations}
with $\omega_{ij}=(\partial_{i}v_{j}-\partial_{j}v_{i})/2$ the vorticity tensor, ${\bf h}=-\delta F/\delta {\bf P}$ the molecular field and ${\bf f}=-\nabla(\delta F/\delta c)$. The flow velocity satisfies the Navier-Stokes equation~\footnote{We neglect here convective nonlinearities in the Navier-Stokes equations that are unimportant on the long time scales of interest.}:
\begin{equation}
\rho(\partial_t +{\bf v}\cdot\nabla)v_i=\partial_j\sigma_{ij}\;,
\label{NS}
\end{equation}
with $\nabla\cdot{\bf v}=0$ to guarantee incompressibility, and stress tensor given by dissipative, reversible and active contributions, $\sigma_{ij}=2\eta u_{ij}+\sigma_{ij}^r+\sigma_{ij}^\alpha+\sigma_{ij}^\beta$, with
\begin{subequations}
\begin{gather}
\sigma_{ij}^\alpha=\frac{\alpha c^{2}}{\Gamma} \big(P_iP_j+\delta_{ij}\big)\;,\label{sigma-a}\\[5pt]
\sigma_{ij}^\beta=\frac{\beta_3 c^{2}}{\Gamma}\big[\partial_iP_j+\partial_jP_i+\delta_{ij}\bm\nabla\cdot{\bf P}\big]\;,\label{sigma-b}\\[5pt]
\sigma_{ij}^r=-\Pi\delta_{ij}-\frac{\lambda}{2}(P_ih_j+P_jh_i)+\frac{1}{2}(P_ih_j-P_jh_i)\;,\label{sigma-r}
\end{gather}
\end{subequations}
where $\Pi$ is the pressure, $\eta$ the shear viscosity, and
we have assumed an isotropic viscosity tensor.
We now consider a solution deep in the polarized state and neglect fluctuations in the magnitude of the polarization, i.e., assume $|{\bf P}|=\sqrt{-a_{2}/a_{4}}$. For simplicity we also redefine units so that $|{\bf P}| = 1$. The condition ${\bf P} = {\rm constant}$ determines the longitudinal part $h_{\parallel} = {\bf p}\cdot{\bf h}$ of the molecular field that can then be eliminated from the hydrodynamic equations. The details associated with imposing the constancy of the magnitude of the polarization and deriving the hydrodynamic equations solely in terms of the polar director ${\bf p} ={\bf P}/|{\bf P}|$ are given in Appendix \ref{sec:appendix}. With this choice, the hydrodynamic equations for ${\bf p}$ and $c$ can be written in the form
\begin{subequations}
\begin{equation}
\partial_tc+\bm\nabla\cdot c({\bf v}+\beta_1 c {\bf p})=\partial_i\left[D_{ij}\partial_j c+\lambda \gamma' u_{kl}p_{k}p_{l}p_{i}\right]\;,\label{c}\\[-7pt]
\end{equation}
\begin{multline}
[\partial_t+({\bf v}+\beta_2 c {\bf p})\cdot\bm\nabla]p_i+\omega_{ij}p_j\\[5pt]
=\delta_{ij}^T\left[\lambda u_{jk}p_k+\frac{w}{c_{0}}\partial_ic-\frac{\gamma'w}{c_{0}}\partial_{j}\nabla\cdot{\bf p}+\kappa\nabla^{2} p_{j}\right]\;,\label{p}
\end{multline}
\end{subequations}
with $\gamma'=\Gamma'/\Gamma$, $\kappa=\Gamma K$, $w=\Gamma(B_{1}-B_{3})$ and $\delta_{ij}^{T}=\delta_{ij}-p_{i}p_{j}$ the transverse projection operator. $D_{ij}$ is an effective diffusion tensor given by
\begin{equation}\label{eq:diffusion-tensor}
D_{ij} = D_1\delta_{ij}+D_2p_{i}p_{j}\;,
\end{equation}
where $D_1=D-\gamma'w/c_0$ and $D_2=\gamma'w/c_0-D\xi$.
Finally, the reversible part of the stress tensor $\sigma_{ij}^{r}$ becomes:
\begin{align*}
\sigma_{ij}^{r}
&= -\delta_{ij}\Pi + \lambda p_{i}p_{j}p_{k}\left[\frac{w}{c_{0}\Gamma}\,\partial_{k}c+K\nabla^{2}p_{k}\right]\\
&-\frac{\lambda}{2}\left[\frac{w}{c_{0}\Gamma}(p_{i}\partial_{j}c+p_{j}\partial_{i}c)+K(p_{i}\nabla^{2}p_{j}+p_{j}\nabla^{2}p_{i})\right]\\[2pt]
&+\frac{1}{2}\left[\frac{w}{c_{0}\Gamma}(p_{i}\partial_{j}c-p_{j}\partial_{i}c)+K(p_{i}\nabla^{2}p_{j}-p_{j}\nabla^{2}p_{i})\right]\\
&-\lambda\Gamma'\xi\,p_{i}p_{j}(Dp_{k}\partial_{k}c+wp_{k}\partial_{k}\partial_{l}p_{l})
+\frac{\lambda^{2}}{\Gamma} p_{i}p_{j} u_{kl} p_{k}p_{l}\,.
\end{align*}
The equations for an active suspension have been written down phenomenologically and also derived from various semi-microscopic models. The structure of of the equations is generic and applies to a broad class of ``living liquid crystals''. The parameters in the equations are of course system and model specific. In motor/filament mixtures activity arises from clusters of motor proteins crosslinking pairs of filaments. The active couplings are therefore of order $c^2$ in this case~\cite{TBLMCM2003,TBLMCMbook}. In suspensions of swimming microorganisms, activity can be described in terms of the active force $f$ that each swimmer exerts on the surrounding fluid. In this case the active couplings arise even at the single-swimmer level and are of order $c$~\cite{BaskaranMarchetti:2009}. Estimates for the active parameters obtained from semimicroscopic models are summarized in Table~\ref{table1}.
\begin{table}
\centering
\begin{ruledtabular}
\begin{tabular}{crr}
$\quad$ & filaments/motors ($\sim$) & swimmers ($\sim$) \\
\hline
$\beta_1$ & $\tilde{m} u_0\ell^2 \qquad\qquad$ & $ v_{\rm sp}/c \qquad$\\
$\beta_2$ & $-\tilde{m} u_0\ell^2\qquad\qquad$ & $ v_{\rm sp}/c \qquad$\\
$w$ & $\tilde{m} u_0\ell^2 \qquad\qquad$ & $-v_{\rm sp}/c \qquad$\\
$\alpha$ & $\tilde mu_{1} \ell^2\qquad\qquad$ & $ f\ell^3/(\zeta c) \qquad$\\
$\beta_3$ & $\tilde mu_{0} \ell^2\qquad\qquad$ & $ v_{\rm sp}/c \qquad$\\
\end{tabular}
\end{ruledtabular}
\caption{Estimates of active parameters for two types of active suspensions: (i) mixtures of cytoskeletal filaments and cross-linking motor proteins~\cite{TBLMCM2003,TBLMCMbook,defnote}, with $\tilde{m}$ a dimensionless density of crosslinking motor clusters , $u_0$ the speed at which motor proteins walk on filaments, in turn proportional to the rate of ATP consumption, and $|u_1|\sim u_0\ell_m$, with $\ell_m$ the size of a motor cluster; and (ii) swimming microorganisms~\cite{BaskaranMarchetti:2009}, where $v_{\rm sp}\sim (f/\zeta)\,\epsilon$ is the self-propulsion speed of an individual organisms, with $f$ the force that swimmers exert on the fluid, $\epsilon<1$ a dimensionless number determined by the shape of the swimmer and $\zeta\sim 1/\Gamma$ is the longitudinal friction coefficient of a rod-like swimmer of length $\ell$. For both systems the precise values of parameters obtained from each microscopic model differ from the above by numericsl constants of order unity.}
\label{table1}
\end{table}
The equations for an active nematic can be obtained from those of a polar systems by setting $\beta_i=w=0$. In the following we assume $\beta_1=\beta_3=-\beta_2=\beta$, as appropriate for motor filament-systems.
It is convenient to work with dimensionless quantities. Spatial variables are normalized with the length $\ell$ of the rods. Thus $y\rightarrow y/\ell$. Temporal variables are normalized with the time scale of splay and bending fluctuations, thus $t\rightarrow t/\tau$ where $\tau=\ell^{2}/\kappa$. A mass scale is set by $\tau/\Gamma$. All the other quantities are normalized accordingly. In these units the hydrodynamic equations for the rods concentration $\phi=c/c_0$, with $c_0$ the mean density, and the director/polarization angle $\theta$, with ${\bf p}=(\cos\theta,\sin\theta)$, for the geometry of interest are
\begin{subequations}\label{eq:hydrodynamic-equations}
\begin{gather}
\rho(\partial_{t}+v_{y}\partial_{y})v_{x} = \partial_{y}\sigma_{xy}\label{eq:v}\\[7pt]
\partial_t \phi
= \partial_{y}\big\{\beta \phi^{2}\sin\theta+{\cal D}(\theta)\partial_y \phi
+\lambda u\sin\theta\sin2\theta\big\}\,,\label{eq:phi}\\[7pt]
\partial_t{\theta}
= -\beta \phi\sin\theta\partial_y\theta+w\cos\theta\partial_y\phi+{\cal K}(\theta)\partial_y^2\theta \notag\\
+ w\cos\theta\sin\theta (\partial_y\theta)^2 -u(1-\lambda\cos 2\theta)\,, \label{eq:theta}
\end{gather}
\end{subequations}
where ${\cal D}(\theta)=D(1-\xi\sin^2\theta)-w\cos^2\theta$ is a diffusion coefficient, ${\cal K}(\theta)=1-w\cos^2\theta$ describes the energy cost of bend and splay deformations, and $\lambda$ is the flow-alignment parameter. In a steady state the stress tensor $\sigma_{xy} \equiv\sigma$ is constant across the film and it is given by
\begin{align}\label{eq:u}
\sigma &= u\Big[\eta+\lambda^{2}\sin^{2}2\theta\Big]
+\lambda w\sin^2\theta\sin 2\theta(\partial_y\theta)^2\notag\\[3pt]
&\quad+[w-\lambda w_{0}-\lambda(w-w_{0})\cos 2\theta]\cos\theta\partial_y\phi\notag\\[5pt]
&\quad+\alpha\phi^{2}\sin2\theta-2\beta\phi^{2}\sin\theta \partial_y\theta
\,,
\end{align}
with $\eta$ the bare viscosity and $w_{0}$ a constant proportional to the ratio between the translational and orientational diffusion coefficients (i.e. $w_{0}\sim D/K$).
Our goal is to study the relation between the induced shear stress $\sigma$ and the applied shear rate $\dot{\gamma}$ as a function of the two fundamental active parameters $\alpha$ and $\beta$ representing the magnitude of the internal contractile/tensile stress and the velocity scale of directed motion. In order to construct a $\sigma$ vs $\dot{\gamma}$ map, we integrate Eqs. \eqref{eq:hydrodynamic-equations} numerically with boundary conditions $v_{x}(0)=0$ and $v_{x}(L)=v_{0}$, $\theta(0)=\theta(L)=0$ and $j_{y}(0)=j_{y}(L)=0$ which implies $\phi'(0)=\phi'(L)=0$. As initial conditions we choose $\theta(y,0)=0$ and $\phi(y,0)=1$.
In the absence of applied shear, active polar and nematic films exhibit a transition from a quiescent ($v_x=0$) aligned ($\theta=0$) state to a state of spontaneous flow, with both inhomogeneous alignment and velocity profiles. The critical value of activity where the instability occurs depends on boundary conditions. For a film bounded by a no-slip substrate and a surface that can freely slide it is given by \cite{GiomiMarchettiLiverpool:2008}:
\begin{equation}
\label{alphac1}
\alpha_{c1}(\beta,\lambda)
=\left(\frac{\pi}{L}\right)^{2}\frac{\eta(1-w)}{2\phi_{0}^{2}(1-\lambda)}
+\frac{\beta w[\eta+(1-\lambda)^{2}]}{2(1-\lambda)(D-w)}\,,
\end{equation}
and the spontaneously flowing state has $\sigma=0$. For a film bounded by two no-slip surfaces the critical value is $\alpha_{c2}=4\alpha_{c1}$ and the spontaneously flowing state is characterized by a finite value of $\sigma$. The regions of spontaneous flow in the $(\lambda,\alpha)$ plane are displayed in shades of orange in Fig.~\ref{fig:alpha-lambda}. In these regions the film exhibits strongly nonlinear rheology, with nonmonotonic stress-strain curves, as described below.
\begin{figure}[t]
\centering
\includegraphics[width=1.\columnwidth]{img/stress-strain-small-alpha.pdf}
\caption{\label{fig:stress-strain1} Stress ($\sigma$) vs strain ($\dot{\gamma}$) for an active nematic ($\beta=w=0$) suspension for various $\alpha$. Flow-tumbling system with $\lambda=0.1$ are marked by circles and flow-aligning systems with $\lambda=1.9$ by triangles. Other parameters are set $L/\ell=5$, $\eta=1$, $\phi_0=1$, $D=1$ and $\xi=0.3$. The inset shows the comparison with the analytical result given in Eq.~\eqref{eq:eta-eff}.}
\end{figure}
\section{Linear rheology of weakly active systems}
\noindent For $|\alpha|<\alpha_{c1}$, corresponding to the gray regions of Fig.~\ref{fig:alpha-lambda}, the stress strain curves are monotonic and remain linear over a broad range of $\dot\gamma$, as shown in Fig.~\ref{fig:stress-strain1}. Non-Newtonian behavior sets in at smaller values of $\dot\gamma$ with increasing $\alpha$.
As the value of $\alpha$ is increased the slope of the linear portion of the stress-strain curves for $\alpha<\alpha_{c1}$ decreases with increasing $\alpha$, indicating that contractile active stresses lower the effective viscosity of the system. The effective linear viscosity can be calculated analytically by solving Eqs. \eqref{eq:theta} and \eqref{eq:phi} perturbatively in $\sigma$ by expanding the fields $\theta$ and $\phi$ as $\theta = \theta_{0}+\sigma\theta_{1}+\sigma^{2}\theta_{2}\ldots$ and $\phi = \phi_{0}+\sigma\phi_{1}+\sigma^{2}\theta_{2}\ldots$ The quantities $\theta_{0}$ and $\phi_{0}$ represents here the stationary solution of the hydrodynamic equations in absence of shear flow. If the suspension is in an aligned state at $t=0$, when the shear is switched on, then $\theta_{0}=0$ and $\phi_{0} = \text{const}$. We note, however, that this perturbation analysis breaks down in the region $\alpha>\alpha_{c1}$ of spontaneous flow, as in that case both $\theta,\phi$ are spatially varying even at $\sigma=0$. It is straightforward to solve Eqs. \eqref{eq:theta} and \eqref{eq:phi} to first order in $\sigma$. We then obtain the linear apparent viscosity defined as $\eta_{\rm app}=\lim_{\dot\gamma\rightarrow 0}\sigma /\dot{\gamma}$ and given by
\begin{equation}\label{eq:eta-eff}
\eta_{\rm app}=\frac{\eta(1+\zeta)}{\zeta+\tanc\left(\frac{kL}{2}\right)}\,,
\end{equation}
where $\tanc(x) = \tan(x)/x$ and
\begin{subequations}\label{eq:k}
\begin{gather}
\zeta = \frac{\eta w \beta}{(1-\lambda)[\beta w(1-\lambda)-2\alpha(D-w)]}\,,\\[10pt]
k^{2}=\frac{2\alpha\phi_{0}^{2}(1-\lambda)}{\eta(1-w)}-\frac{\beta w\phi_{0}^{2}}{(1-w)(D-w)}\left[1+\frac{(1-\lambda)^2}{\eta}\right]\,,
\end{gather}
\end{subequations}
For passive system $\alpha=\beta=w=0$, and $\eta_{\rm app}=\eta$, as expected. For active nematic, $\beta=w=0$ and the apparent viscosity is simply
\begin{equation}\label{eq:eta-nematic}
\eta_{\rm app} = \frac{\eta}{\tanc\left(\frac{k}{2}\frac{L}{\ell}\right)}\,,
\end{equation}
with $k=\sqrt{2\alpha\phi_{0}^{2}(1-\lambda)/\eta}$. If $\alpha(1-\lambda)<0$, $k$ is imaginary and the $\tan$ function at the denominator of $\eta_{\rm app}$ is replaced by its hyperbolic counterpart. Since $\tanh(x)$ increases more slowly than $x$, the resulting apparent viscosity will increase. If $\alpha(1-\lambda)>0$, $k$ is real and since the $\tan(x)$ function grows more rapidly than $x$ we expect then a rapid decrease in the apparent viscosity as $|\alpha|$ is increased.
This shows that the linear rheology of pullers/contractile systems with $\lambda<1$ are the same as those of pushers/tensile systems with $\lambda>1$. From Eq. \eqref{eq:eta-eff} it is indeed simple to prove that the apparent viscosity $\eta_{\rm app}$ is invariant under the transformation
\begin{equation}\label{eq:eta-invariance}
\eta_{\rm app}(\alpha,\beta,\lambda) = \eta_{\rm app}(-\alpha,\beta,2-\lambda)\,.
\end{equation}
\begin{figure}[b]
\centering
\includegraphics[width=.9\columnwidth]{img/active-flow.pdf}
\caption{\label{fig:active-flow}Schematic example of the flow field surrounding a tensile/flow-aligning (right) and contractile/flow-tumbling active particle. For the choice of the parameters $\alpha$ and $\lambda$ given in Eq. \eqref{eq:eta-invariance} the two flows are identical, leading to an equal apparent viscosity.}
\end{figure}%
Thus flow-aligning pullers with $\lambda=1+\epsilon$ (for $0\le\epsilon<1$) will exhibit the same apparent viscosity of ow-tumbling pushers with $\lambda=1-\epsilon$: $\eta_{\rm app}(-|\alpha|,\beta,1+\epsilon)=\eta_{\rm app}(|\alpha|,\beta,1-\epsilon)$. This duality is displayed in the top frame of Fig.~\ref{fig:eta-apparent} that shows the linear apparent viscosity of active nematic suspensions as a function of $|\alpha|$ for several values of $\lambda$. The solid curves (red online) show that both contractile/flow tumbling suspensions and tensile/flow aligning ones are thinned by activity. The dashed curves (blue online) refer to either contractile/flow aligning suspensions or tensile/flow tumbling ones and show that these systems are thickened by activity.
Bacteria such as E-Coli are pushers ($\alpha<0$) and generally elongated in shape, corresponding to $\lambda>1$. Our results therefore confirm the activity-induced thinning of bacterial suspensions first predicted by Hatwalne et al~\cite{Hatwalne04} and recently observed in ~\cite{SokolovAranson2009}. In contrast, algae like Chamydomonas that propel themselves from the front (and are therefore pullers, with $\alpha>0$). Whether they are thickened or thinned by activity depends intimately on their shape, i.e. on whether they can be described as objects with $\lambda>1$ or $\lambda < 1$. Similarly, motor/filament mixtures are generally contractile ($\alpha>0$) are are expected to be thickened or thinned by activity depending on the effective value of $\lambda$.
\begin{figure}[t]
\centering
\includegraphics[width=.9\columnwidth]{img/eta-app.pdf}
\caption{\label{fig:eta-apparent} (color online) Apparent viscosity $\eta_{\rm app}$ for active nematic (top) and polar (bottom) suspensions. Solid/red lines represent flow-tumbling systems ($\lambda<1$) while dashed/blue lines flow-aligning systems ($\lambda>1$). The corresponding values of $\lambda$ are indicated next to the lines. In the bottom plot $\alpha$ was set to zero. The top frame emphasizes the duality discussed in the text.}
\end{figure}
This duality has a simple interpretation. Active contractile (tensile) particles produce an ingoing (outgoing) flow in the surrounding fluid, but while flow-aligning particles orient at a positive angle with respect to the flow direction, flow-tumbling particles orient at a negative angle under a small applied shear (see Fig. \ref{fig:active-flow}). As a result, the average flow fields produced in the surrounding fluid are identical in the two cases and produce the same resistance to the imposed shear flow. This equivalence holds only for small applied shear stresses. For large shear-rates the configuration of the director field of a flow-tumbling suspension is dramatically different from that of flow-aligning one and the similarity between the two flow-fields no longer holds.
\section{Nonlinear rheology of strongly active systems}
\noindent The linear apparent viscosity given by Eq. \eqref{eq:eta-nematic} vanishes at $\alpha=\alpha_{c1}$, suggesting the onset of a superfluid-like behaviour above this critical value of activity~\cite{CatesEtAl:2008}. For $\alpha>\alpha_{c1}$, { the linearized approximation breaks down} and the stress versus (average) strain rate curve obtained by numerical solution of the equations is nonlinear and nonmonotonic, as shown in Fig.~\ref{fig:stress-strain2}. We emphasize that the flow profiles are always
inhomogeneous with varying velocity gradients and director orientation.
For $\alpha_{c1}<\alpha<\alpha_{c2}$ the theoretical stress versus macroscopic (average) strain rate curve goes through the origin and exhibits a region of negative $d\sigma/d\dot\gamma$, that would in principle be mechanically unstable.
What would be measured in an experiment would, however, depend critically on details of the experimental procedure and the particular apparatus.
To study the steady state rheology there are in general two natural classes of experiments: either (i) one tunes the stress $\sigma$ and measures the resulting strain rate $\dot\gamma$ or (ii) one does a sweep through the values of strain rate $\dot\gamma$ and measures the stress $\sigma$. If the stress-strain rate curve is monotonic, the two procedures are expected to yield the same result. However, this is no longer the case as soon as the response exhibits nonmonotonicity.
An important question, then, is what is the shape of the stress-strain rate curve that would be obtained experimentally for $\alpha>\alpha_{c1}$ in an experiment where one tunes the {\em macroscopic} strain rate $\dot\gamma$. Several scenarios are possible, as shown in Fig.~\ref{fig:scenarios} for a non-monotonic curve with maximum/minimum at $\pm \sigma_{m}$.
\begin{figure}[b]
\centering
\includegraphics[width=1.\columnwidth]{img/scenarios.pdf}
\caption{\label{fig:scenarios}The top left frame display a typical theoretical stress-strain curve of a nematic active suspension in the region $\alpha_{c1}<|\alpha|<\alpha_{c2}$. The theoretical curve is obtained by tuning $\dot\gamma$ and calculating the resulting $\sigma$ and exhibits a region of $d\sigma/d\dot\gamma<0$. The other three frames show three possible experimental stress-strain curves obtained by tuning $\sigma$ and measuring $\dot\gamma$ that could be consistent with the theoretical curve. The top right frame displays the ``superfluid'' scenario suggested in~\cite{CatesEtAl:2008}, with bulk shear bands accommodating different macroscopic shear rates and zero net stress, so that the apparent viscosity of the system is simply zero. The bottom left frame shows a yield-stress like behaviour with a yield stress $\sigma_y=\sigma_m$. The last scenario is described in the bottom right frame and corresponds to a hysteretic stress-strain curve where the suspension can accommodate a range of macroscopic strain rates maintaining a constant total stress $\pm\sigma_0$.}
\end{figure}
(i) One scenario, suggested recently~\cite{CatesEtAl:2008} based on numerical studies in the proximity of the isotropic-nematic phase transition and for small value of the active stress $\alpha$ is the appearance of bulk shear bands accommodating a range of macroscopic shear-rates at zero stress. This would correspond to the bulk stress-strain curve displayed in the top right frame of Fig.~\ref{fig:scenarios} and characterized as ``superfluid'' behavior. In the simplest picture the sheared suspension would separate in bands of constant and opposite strain rates, each with zero stress.
For the systems studied here (deep in the ordered phase, either nematic or polar), we find that the equations of motion provide no mechanism for selecting a particular value of the stress plateau and are unable to find a stable stress-plateau at any value of $|\sigma|< \sigma_m$ (including $\sigma=0$, see Fig. \ref{fig:scenarios}). Furthermore we always find flow profiles with continuously varying gradients of fluid velocity for all values of macroscopic strain-rate $\dot\gamma$ implying that the picture of two bands of constant strain rate would be at best an idealisation.
(ii) An alternative scenario that is observed in other driven systems, such as charge density waves in anisotropic metals~\cite{Maeda1990} and collections of motor proteins~\cite{JulicherProst1995}, is shown in the bottom right frame of Fig.~\ref{fig:scenarios}. In this case the system is expected to exhibit hysteresis, with regions that accommodate coexistence of a range of macroscopic strain rates, corresponding to the constant value $\pm\sigma_0$ of applied stress. In general $\sigma_0$ may coincide with $\sigma_m$ or may be lower, with the system exhibiting ``early swtching". The width of the horizontal hysteretic region of the stress-strain curve decreases with increasing $\alpha$. In this picture the particular steady-state behaviour observed will depend on the initial conditions and particular flow history of each sample.
(iii) Another possibility is that the system shows a yield-stress like behaviour with a yield stress $\pm \sigma_y$ whose sign is determined by the direction of the flow. The value of the yield stress could also be anywhere in the ``unstable'' range of stress: $\sigma_y \le \sigma_m$.
(iv) Finally, there is one more possibility: that he theoretical curve would indeed be reproduced by an experiment which scanned through different values of the macroscopic strain rate. The theoretical curve has been calculated by fixing $\dot\gamma$ and calculating the corresponding value of $\sigma$ under the assumption that there are variations in the director and flow field {\em only in} the gradient direction (i.e. perpendicular to the plates). If this {\em assumption} is valid, every point on this curve does therefore represent a stable state corresponding to this procedure.
\begin{figure}[t!]
\centering
\includegraphics[width=.9\columnwidth]{img/stress-strain.pdf}
\caption{\label{fig:stress-strain2} Stress-strain curves of a nematic suspension ($\beta=w=0$) obtained by numerical solution of the active hydrodynamic equations for several values of $\alpha$. $\alpha_{c1}=0.219$ and $\alpha_{c2}=0.877$ for the parameters chosen in the numerical solution.}
\end{figure}
\begin{figure}[t]
\centering
\includegraphics[width=.9\columnwidth]{img/yieldstress.pdf}
\caption{\label{fig:yieldstress} Yield-stress $\sigma_{c}$ as a function of $\alpha$ for a nematic suspension ($\beta=w=0$) obtained by numerical solution of the active hydrodynamic equations.}
\end{figure}
For $\alpha>\alpha_{c2}$ the stress-strain curve intercepts the $\dot\gamma=0$ axis at a finite value $\sigma_c=\sigma(\dot\gamma=0)$ of the strain rate. The active suspension has a nonzero spontaneous stress even in the absence of applied forces, as indeed observed in the spontaneous flow regime of an active suspension confined between two stationary no-slip planes. In other words, a finite force must be applied to the active suspension to keep it from sliding even at zero mean strain rate. This spontaneous stress $\sigma_c$ is shown as a function of $\alpha$ in Fig.~\ref{fig:yieldstress}. The sign of the stress determines the direction of spontaneous flow.
We now speculate on the possible behavior of the system for each of the scenarios sketched above as $\alpha$ goes through $\alpha_{c2}$. The behavior is shown schematically in Fig.~\ref{fig:scenarios2}.
(i) In the superfluid scenario, the response of the suspension to an applied macroscopic strain rate will show yield stress behavior. The system would smoothly go from the zero-stress plateau to a yield stress which increases from zero at $\alpha_{c2}$.
(ii) In the hysteretic scenario the minimum height of the hysteretic loop becomes $2 \sigma_c$ i.e. $\sigma_c \le \sigma_0 \le \sigma_m$.
(iii) In the yield-stress scenario the system already shows yield stress behaviour which continues for $\alpha > \alpha_{c2}$.
(iv) In the non-monotonic scenario, the non-monotonic stress-strain rate curve shows a jump at $\dot\gamma$ whose magnitude increases from zero at $\alpha_{c2}$.
\begin{figure}[t]
\centering
\includegraphics[width=1.\columnwidth]{img/scenarios2.pdf}
\caption{\protect
Possible scenarios for the transition to the yield stress regime at $\alpha>\alpha_{c2}$. The non-monotonic curve obtained numerically is shown in the top left frame. In the superfluid scenario (top-right) the plateau at $\sigma=0$ divides into two disconnected branches terminating at $\sigma=\pm \sigma_{c}$. In this case the yield stress is expected to grow monotonically from zero. In the yield stress scenario (bottom-left), there is already a non-zero stress at $\dot\gamma=0$ and thus the yield stress simply continues increasing with no qualitative change in the behaviour at $\alpha_{c2}$. In the hysteretic scenario (bottom-right), the loop intersect the positive $\sigma$ axis at $\pm\sigma_0$, with $\sigma_c\leq\sigma_0\leq\sigma_m$.}
\label{fig:scenarios2}
\end{figure}
\section{Discussion and conclusions}
We have studied the rheological behavior of a thin film of polar and apolar active material. For weakly active systems, in the regime of the linear rheology, we have confirmed analytically the prediction of Hatwalne and collaborators~\cite{Hatwalne04} that activity can lower the linear bulk viscosity of tensile suspensions of swimmers as well as enhance the viscosity of contractile systems. We have shown that this result applies also for finite systems, in the presence of boundaries.
An important new result of our work is the role of the {\em shape} of the active particles in controlling the rheological behavior. We find a remarkable exact duality that holds in the regime where the stress-strain rate relation is linear and shows that tensile ($\alpha<0$) rod-shaped flow-aligning particles ($\lambda>1$) are rheologically equivalent to contractile ($\alpha>0$) discotic flow-tumbling particles ($-1\leq\lambda<0$). This means that activity lowers the linear viscosity of both tensile, rod shaped particle and contractile, disc shaped particle suspensions, while it increases the linear viscosity of contractile, rod-shaped particle and tensile, discotic particle suspensions.
For strongly active systems we find that the rheological response is intrinisically nonlinear. The regime of linear rheology at small strain rates vanishes beyond a critical value of activity. In this strongly active regime, we explore a number of possible scenarios for the nonlinear rheology which include a ``superfluid'' phase with vanishing viscosity, hysteresis, yield-stress behavior and non-monotonic behavior. Our one-dimensional analysis does not, however, allow us to determine which of these scenarios is more likely. It is of course possible that allowing for variations of the director and flow field in higher dimensions or allowing for variations in the magnitude of the order parameter would yield a criterion for selecting one of the proposed scenarios.
\begin{acknowledgements}
LG is supported by NSF through the Harvard MRSEC and the Brandeis MRSEC and by the Harvard Kavli Institute for Nanobio Science \& Technology. MCM is supported by NSF grants DMR-075105 and DMR-0806511. TBL acknowledges the support of EPSRC under grant EP/G026440/1. We thank Suzanne Fielding and James Adams for illuminating discussions.
\end{acknowledgements}
|
\section{Introduction}
During winter, the stratospheric circulation at both poles is characterized by a large cyclonic
vortex centered in a region close to the pole (Haynes~\cite{hay2005}).
Extended climatology of the polar vortex have been already conducted, using ECMWF or NCAR-CEP data
set (Karpetchko et al. \cite{kar2005}; Waugh et al. \cite{wa1999a}) with methods based on potential
vorticity to define the vortex.
It is well known that the vortices are much more stronger in mid-winter than in
summer (Waugh et al. \cite{wa1999b}; Harvey et al. \cite{har2002}).
In this study we want to determine the minimum speed of the high altitude wind above
the Antarctica continent.
This might help us in identifying the center of the polar vortex and we could use this information
to qualify the sites for astronomical applications.
\par
Indeed, during a previous study about site characterization for optical turbulence
above the internal Antarctic Plateau using ECMWF analysis, our group
investigated four different sites (Hagelin et al. \cite{hag2008}): Dome A, C and F and South Pole.
One of the conclusions was that in the free atmosphere, above around 10 km, the wind speed increased
monotonically in winter proportionally to the distance to the center of the polar high.
Thus Dome C was the site showing the highest wind speed above 10 km.
We propose in this study to confirm this hypothesis looking at a climatology of the high altitude wind speed
and the corresponding vortex above the Antarctica continent for winter 2005.
Due to the weak variability of the Antarctica vortex proved in other studies, we can deduce that this one-year study
can provide a quantitative estimate with a good accuracy about the position of the minimum wind speed in altitude.
\section{The median high altitude wind speed}
We used the ECMWF analyses from MARS catalog for every day between May 1st and September 30th, 2005,
at 00 UTC.
The analysis employs the 4D-VAR assimilation scheme to assimilate a wide number of observations, including
in-situ conventional data and synthetic data from satellites.
The averaged distance between two horizontal grid points is of the order of $\sim$40~km.
To perform our study, we focused on the wind speed at two different
altitudes, 15~km and 20~km from May 1st and September 30th, every 24 hours at 00 UTC.
We computed the monthly medians of the wind speed and deduced a preferential position for the minimum of wind
speed in altitude for winter 2005.
\par
Fig.\ref{fig:median} shows the monthly median of the wind velocity at two different heights, 15 and 20~km,
computed from May 2005 to September 2005.
\begin{figure}
\begin{center}
\includegraphics[width=0.9\textwidth]{lascaux_2_fig_1.eps}
\end{center}
\caption{Median wind speed at 15~km (column 1) and 20~km (column 2) above the Antarctic continent
(polar stereographic projection), for Winter 2005 (May to September). Units in
m.s$^{-1}$. Isocontours starting from 1 m.s$^{-1}$ with an increment of 1 m.s$^{-1}$.}
{\label{fig:median}}
\end{figure}
\par
The minimum of the median wind is clearly identified at 20~km.
The median polar vortex center (corresponding to the minimum wind speed)
remains in a area between South Pole and Dome A.
Between the four sites investigated by Hagelin \etal~(\cite{hag2008}), Dome C is the farthest away from the
polar vortex center, and by consequence the one with the highest wind speed in altitude.
\section{Conclusion}
This study confirms he conclusion regarding the "position space" of the polar high deduced by
Hagelin \etal~(\cite{hag2008}), and the link between the position of an Antarctic site with respect to this center
and the wind speed in altitude.
Dome C appears to have in winter a wind speed in altitude much higher than other sites like South Pole and
Dome A, closer to the center of the vortex.
Such a quantitative information should be considered by astronomers as a key issue in future plans for
astronomical facilities to be set-up in different sites of the Internal Antarctic Plateau.
\section*{Acknowledgements}
This study has been funded by the Marie Curie Excellence Grant (FOROT) - MEXT-CT-2005-023878.
|
\section{Introduction}
\subsection{Motivation}
Let $\varphi_t^\mathrm{RW}(x)$ be the $t$-step transition
probability for
random walk
on ${\mathbb Z}^d$: $\varphi_0^{\mathrm{RW}}(x)=\delta_{o,x}$ and
\begin{eqnarray}\label{eq:RW}
\varphi_t^{\mathrm{RW}}(x)=(\varphi_{t-1}^
{\mathrm{RW}}*D)(x)\equiv\sum_{y\in{\mathbb Z}^d
}\varphi_{t-1}^{\mathrm{RW}}
(y) D(x-y)\qquad[t\in\mathbb{N}].
\end{eqnarray}
Suppose that the 1-step distribution $D$ is ${\mathbb Z}^d$-symmetric.
How does the
$r$th moment $\sum_x|x|^r\varphi_t^{\mathrm{RW}}(x)$ grow as
$t\to\infty$, where $|\cdot|$ denotes the Euclidean distance?
When $r=2$ and $\sigma^2\equiv\sum_x|x|^2D(x)<\infty$, the answer
is trivial:
$\sum_x|x|^2\varphi_t^{\mathrm{RW}}(x)=\sigma^2t$ since the
variance of the
sum of
independent random variables is the sum of their variances. It is not
so hard
to see that $\sum_x|x|^r\varphi_t^{\mathrm{RW}}(x)=O(t^{r/2})$
as $t\to\infty
$ for other
values of $r>2$, as long as $\sum_x|x|^rD(x)<\infty$. Even so, it
may not be that easy to identify the constant $C\in(0,\infty)$ such that
$(\sum_x|x|^r\varphi_t^{\mathrm{RW}}(x))^{1/r}\sim C\sqrt t$.
Here, and in the
rest of
the paper, ``$f(z)=O(g(z))$'' means that $|f(z)/g(z)|$ is bounded for
all $z$
in some relevant set, while ``$f(z)\sim g(z)$'' means that $f(z)/g(z)$ tends
to 1 in some relevant limit for $z$.
Let $\alpha>0$, $L\in[1,\infty)$ and suppose that
$D(x)\approx|x/L|^{-d-\alpha}$ for large $x$ such that its Fourier transform
$\hat D(k)\equiv\sum_{x\in{\mathbb Z}^d}e^{ik\cdot x}D(x)$ satisfies
\begin{eqnarray}\label{eq:D-asympt}
1-\hat D(k)=v_\alpha|k|^{\alpha\wedge2}\times
\cases{
1+O((L|k|)^\epsilon),&\quad$\alpha\ne2$,
\cr
\log\dfrac1{L|k|}+O(1),&\quad$\alpha=2$
}
\end{eqnarray}
for some $v_\alpha=O(L^{\alpha\wedge2})$ and $\epsilon>0$. If
$\alpha>2$ (or
$D$ is finite-range), then $v_\alpha\equiv\sigma^2/(2d)$. As shown in
Appendix~\hyperref[appendix:D]{A.1}, the long-range Kac potential
\begin{eqnarray}\label{eq:kac}
D(x)=\frac{h(y/L)}{\sum_{y\in{\mathbb Z}^d}h(y/L)}\qquad[x\in
{\mathbb Z}^d],
\end{eqnarray}
defined in terms of a rotation-invariant
function $h$ satisfying
\begin{eqnarray*}
h(x)=\frac{1+O((|x|\vee1)^{-\rho})}{(|x|\vee1)^{d+\alpha
}}\qquad
[x\in{\mathbb R}^d]
\end{eqnarray*}
for some $\rho>\epsilon$, satisfies the above properties. Notice that
$\sum_x|x|^rD(x)=\infty$ for $r\ge\alpha$ and, in particular,
$\sigma^2=\infty$ if $\alpha\le2$. This is of interest in investigating
the asymptotic behavior of $\sum_x|x|^r\varphi_t^\mathrm
{RW}(x)$ for \emph{all}
$r\in(0,\alpha)$ and understanding its $\alpha$-dependence.
In fact, our main interest is in proving sharp asymptotics of the gyration
radius of order $r\in(0,\alpha)$, defined as
\begin{eqnarray*}
\xi_t^{(r)}=\biggl(\frac{\sum_{x\in{\mathbb Z}^d}|x|^r\varphi
_t(x)}{\sum_{x\in{\mathbb Z}^d}
\varphi_t(x)}\biggr)^{1/r},
\end{eqnarray*}
where $\varphi_t(x)\equiv\varphi_t^\mathrm{SAW}(x)$ is the two-point
function for
$t$-step self-avoiding walk whose 1-step distribution is given by $D$, or
$\varphi_t(x)\equiv\varphi_t^\mathrm{OP}(x)$ is the two-point
function for oriented
percolation whose bond-occupation probability for each bond $((u,s),(v,s+1))$
is given by $pD(v-u)$, independently of $s\in{\mathbb Z}_+$, where
$p\ge0$ is the
percolation parameter. More precisely,
\begin{eqnarray}\label{eq:2pt-def}
\varphi_t(x)=
\cases{
\displaystyle\varphi_t^{\mathrm{RW}}(x)\equiv\mathop{\sum_{\omega:o\to x}}_{
(|\omega|=t)}
\prod_{s=1}^tD(\omega_s-\omega_{s-1}),
\cr
\displaystyle\varphi_t^\mathrm{SAW}(x)\equiv\mathop{\sum_{\omega:o\to x}}_
{(|\omega|=t)}
\prod_{s=1}^tD(\omega_s-\omega_{s-1})\prod_{0\le i<j\le t}(1-
\delta_{\omega_i,\omega_j}),
\cr
\varphi_t^\mathrm{OP}(x)\equiv{\mathbb P}_p((o,0)\to
(x,t)),
}
\end{eqnarray}
where $\prod_{0\le i<j\le t}(1-\delta_{\omega_i,\omega_j})$ is the
self-avoiding constraint on $\omega$ and $\{(o,0)\to(x,t)\}$ is the event
that either $(x,t)=(o,0)$ or there is a consecutive sequence of
occupied bonds
from $(o,0)$ to $(x,t)$ in the time-increasing direction. The gyration radius
$\xi_t^{(r)}$ represents a typical end-to-end
distance of a
linear structure of length $t$ or a typical spatial size of a cluster
at time
$t$. It has been expected (and would certainly be true for random walk
in any dimension) that, above the common upper-critical dimension
$d_\mathrm{c}=2(\alpha\wedge2)$ for self-avoiding walk and oriented
percolation,
for every $r\in(0,\alpha)$,
\begin{eqnarray}\label{eq:conjecture}
\xi_t^{(r)}=
\cases{
O\bigl(t^{1/({\alpha\wedge2})}\bigr),&\quad$\alpha\ne2$,
\cr
O\bigl(\sqrt{t\log t}\bigr),&\quad$\alpha=2$.
}
\end{eqnarray}
Heydenreich~\cite{hhh} proved (\ref{eq:conjecture}) for self-avoiding
walk, but
only for small $r<\alpha\wedge2$. Nevertheless, this small-$r$ result is
enough to prove weak convergence of self-avoiding walk to an $\alpha$-stable
process/Brownian motion, depending on the value of~$\alpha$ \cite{hhh}.
As stated below in Theorem~\ref{theorem:main2}, we prove sharp asymptotics
(including the proportionality constant) of
$\sum_x|x_1|^r\varphi_t(x)/\sum_x\varphi_t(x)$ as $t\to\infty$,
where $x_1$
is the first coordinate of $x\equiv(x_1,\dots,x_d)$, and show that
(\ref{eq:conjecture}) holds for all $r\in(0,\alpha)$, solving the
open problems
in \cite{csII,hhh}.
\subsection{Main results}\label{ss:main}
Let $m_{\rm c}\ge1$ be the model-dependent radius of convergence for
the sequence
$\sum_x\varphi_t(x)$. For random walk, $m_{\rm c}=1$ since $\sum
_x\varphi_t^{\mathrm{RW}}(x)$
is always 1. For self-avoiding walk, $m_{\rm c}>1$ due to the
self-avoiding constraint in (\ref{eq:2pt-def}) and, indeed, $m_{\rm
c}=1+O(L^{-d})$ for
$d>d_{\mathrm c}$ and $L\gg1$ \cite{hhs08}. For oriented percolation,
$m_{\rm
c}$ depends on
the percolation parameter $p$ [i.e., $m_{\rm c}=m_{\rm c}(p)$] and was
denoted by $m_p$ in
\cite{csI,csII}. It has been proven \cite{csI} that $m_{\rm c}(p)>1$
for $p<p_{\mathrm c},$
and $m_{\rm c}(p_{\mathrm c})=1$ for $d>d_{\mathrm c}$ and $L\gg1$,
where $p_{\mathrm c}$ is the
critical point
characterized by the divergence of the susceptibility:
$\sum_{t=0}^\infty\sum_{x\in{\mathbb Z}^d}\varphi_t^\mathrm
{OP}(x)\uparrow\infty
$ as
$p\uparrow p_{\mathrm c}$. It has also been\vspace*{1pt} proven \cite{csI} that
$pm_\mathrm{c}=1+O(L^{-d})$
for all $p\le p_{\mathrm c}$.\vspace*{1pt}
Let $C_{\mathrm I}$ and $C_{\mathrm{II}}$ be the constants in \cite
{csI,csII,hhh} such that,
as $t\to\infty$,
\begin{eqnarray}\label{eq:main-I&II}
\sum_{x\in{\mathbb Z}^d}\varphi_t(x)\sim C_{\mathrm I} m_{\rm c}^{-t},
\qquad
\frac{\sum_{x\in{\mathbb Z}^d}e^{ik_t\cdot x}\varphi_t(x)}{\sum
_{x\in{\mathbb Z}^d
}\varphi_t(x)}
\sim e^{-C_{\mathrm{II}}|k|^{\alpha\wedge2}},
\end{eqnarray}
where
\begin{eqnarray}\label{eq:k-scaling}
k_t=k\times
\cases{
(v_\alpha t)^{-1/({\alpha\wedge2})},&\quad$\alpha\ne2$,
\cr
\bigl(v_2 t\log\sqrt t\bigr)^{-1/2},&\quad$\alpha=2$.
}
\end{eqnarray}
Because of this scaling, we have $C_{\mathrm I}^\mathrm
{RW}=C_{\mathrm{II}}^{\mathrm{RW}}=1$ for random
walk. For
self-avoiding walk and critical/subcritical oriented percolation for
$d>2(\alpha\wedge2)$ with $L\gg1$ (depending on the models), it has been
proven that the model-dependent constants $C_{\mathrm I}$ and
$C_{\mathrm{II}}$ are both
$1+O(L^{-d})$ \cite{csI,hhh} and that the $O(L^{-d})$ term in
$C_{\mathrm{II}}$ exhibits
crossover behavior at $\alpha=2$ \cite{csII,hhh}. We will provide precise
expressions for $C_{\mathrm I}$ and $C_{\mathrm{II}}$ at the end of
Section~\ref{ss:outline}.
Our first result is the following asymptotic behavior of the generating
function for the sequence $\sum_x|x_1|^r\varphi_t(x)$.
\begin{theorem}\label{theorem:main1}
Consider the three aforementioned long-range models. For random walk in any
dimension $d$ with any $L$, and for self-avoiding walk and critical/subcritical
oriented percolation for $d>d_{\mathrm c}\equiv2(\alpha\wedge2)$
with $L\gg
1$ (depending
on the models), the following holds for all $r\in(0,\alpha)$: as
$m\uparrow m_{\rm c}$,
\begin{eqnarray}\label{eq:main1}
\qquad\sum_{t=0}^\infty m^t\sum_{x\in{\mathbb Z}^d}|x_1|^r\varphi
_t(x)&=&\frac
{2\sin
({r\pi}/{(\alpha\vee2))}}{(\alpha\wedge2)\sin({r\pi}/\alpha)} \Gamma(r+1)
\frac{C_{\mathrm I}(C_{\mathrm{II}}v_\alpha)^{{r}/{(\alpha
\wedge2)}}}{(1-
{m}/{m_{\rm c}})^{1+
{r}/{(\alpha\wedge2)}}}\nonumber
\\[-8pt]\\[-8pt]
&&{}\times
\cases{
1+O\biggl(\biggl(1-\dfrac{m}{m_{\rm c}}\biggr)^\epsilon\biggr),&\quad$\alpha\ne2$,
\cr
\biggl(\log\dfrac1{\sqrt{1-{m}/{m_{\rm c}}}}
\biggr)^{r/2}+O(1),&\quad$\alpha=2$
}
\nonumber
\end{eqnarray}
for some $\epsilon>0$ when $\alpha\ne2$. The $O(1)$ term for $\alpha=2$
is independent of $m$.
\end{theorem}
It is worth emphasizing that, although $C_{\mathrm I},C_{\mathrm
{II}},m_{\rm c}$ are
model-dependent,
the formula (\ref{eq:main1}) itself is universal. Expanding (\ref
{eq:main1}) in
powers of $m$ and using (\ref{eq:main-I&II}), we obtain the following theorem.
\begin{theorem}\label{theorem:main2}
Under the same condition as in Theorem~\ref{theorem:main1}, as $t\to
\infty$,
\begin{eqnarray}\label{eq:main2}
\frac{\sum_{x\in{\mathbb Z}^d}|x_1|^r\varphi_t(x)}{\sum_{x\in
{\mathbb Z}^d}\varphi
_t(x)}&\sim&
\frac{2\sin({r\pi/(\alpha\vee2)})}{(\alpha\wedge2)\sin
({r\pi}/\alpha)}
\frac{\Gamma(r+1)}{\Gamma({r}/{(\alpha\wedge2)}+1)}\nonumber
\\[-8pt]\\[-8pt]
&&{}\times
\cases{
\bigl(C_{\mathrm{II}}v_\alpha t\bigr)^{{r}/{(\alpha\wedge2)}},&\quad$\alpha\ne2$,
\cr
\bigl(C_{\mathrm{II}}v_2t\log\sqrt{t}\bigr)^{r/2},&\quad$\alpha=2$.
}
\nonumber
\end{eqnarray}
\end{theorem}
We note that $C_{\mathrm{II}}$ is the only model-dependent term in (\ref
{eq:main2}).
As far as we are aware, the sharp asymptotics (\ref{eq:main1}) and (\ref
{eq:main2})
for \emph{all} real $r\in(0,\alpha)$ are new, even for random walk.
Although we focus our attention on the long-range models defined by $D$ that
satisfies (\ref{eq:D-asympt}), our proof also applies to
finite-range models, for which $\alpha$ is considered to be infinity.
Using $|x_1|^r\le|x|^r\le d^{r/2}\sum_{j=1}^d|x_j|^r$ and the
${\mathbb Z}^d$-symmetry
of the models, we are finally able to arrive at the following result.
\begin{corollary}
Under the same condition as in Theorem~\ref{theorem:main1}, (\ref
{eq:conjecture})
holds for all $r\in(0,\alpha)$. In particular, when $r=2<\alpha$,
\begin{eqnarray}\label{eq:mean-square}
\xi_t^{(2)}\mathop{\sim}_{t\to\infty}\sqrt{C_{\mathrm{II}}\sigma^2 t}.
\end{eqnarray}
\end{corollary}
As mentioned earlier, (\ref{eq:conjecture}) has been proven \cite
{hhh} for
self-avoiding walk, but only for small $r<\alpha\wedge2$. The sharp
asymptotics (\ref{eq:mean-square}) has been proven \cite{hs02} for
self-avoiding
walk and critical oriented percolation defined by $D$ that has a finite
$(2+\epsilon)$th moment for some $\epsilon>0$. Our proof is based on a
different method than those used in \cite{hhh,hs02}. It is closer to the
method, explained in the next subsection, used in \cite{ms93} for finite-range
self-avoiding walk and in \cite{ny95} for critical/subcritical finite-range
oriented percolation.
We strongly believe that the same method should work
for lattice trees. Any two points in a lattice tree are connected by a
unique path, so the number of bonds contained in that path can be considered
as time and we can apply the current method to obtain the same
results (with different values for $C_{\mathrm I},C_{\mathrm{II}}$). As
this suggests,
time, or
something equivalent, is important for the current method to work. For
unoriented percolation, for example, it is not so clear what should be
interpreted as time. However, if $D$ is biased in average in one direction,
say, the positive direction of the first coordinate, then $x_1$ can be treated
as time and, after subtracting the effect of the bias, we may obtain the
results even for unoriented percolation.
\subsection{Outline and notation}\label{ss:outline}
In this subsection, we outline the proof of Theorem~\ref
{theorem:main1} and
introduce some notation which is used in the rest of the paper. We also refer
interested readers to an extended version of this subsection in \cite{s09}.
One of the key elements for the proof is to represent the left-hand
side of
(\ref{eq:main1}) in terms of the generating function (i.e., the
Fourier--Laplace
transform)
of the two-point function. We now explain this representation.
Given a function $f_t(x)$, where
$(x,t)\in{\mathbb Z}^d\times{\mathbb Z}_+$, we formally define
\begin{eqnarray*}
\hat f(k,m)=\sum_{t=0}^\infty m^t\sum_{x\in{\mathbb Z}^d}f_t(x)
e^{ik\cdot
x}\qquad[k\in
[-\pi,\pi]^d,~m\ge0].
\end{eqnarray*}
We note that $\hat\varphi(k,m)$ is well defined when $m<m_{\rm c}$
(recall that
$m_{\rm c}\ge1$, as explained at the beginning of Section~\ref{ss:main}).
Let\vspace*{-2pt}
\begin{eqnarray}\label{eq:nablone}
\nabla_{1}^n\hat f(l,m)=\frac{\partial^n\hat f(k,m)}{\partial
k_1^n}\bigg|_{k=l}
\qquad[l\in[-\pi,\pi]^d,~n\in{\mathbb Z}_+].\vspace*{-2pt}
\end{eqnarray}
Then, for $r=2j<\alpha$ ($j\in\mathbb{N}$), we obtain the representation\vspace*{-2pt}
\begin{eqnarray*}
\sum_{t=0}^\infty m^t\sum_{x\in{\mathbb
Z}^d}x_1^{2j}f_t(x)=(-1)^j\nabla_{1}
^{2j}\hat
f(0,m).\vspace*{-2pt}
\end{eqnarray*}
For $r\in(0,\alpha\wedge2)$, we generate the factor $|x_1|^r$ by
using the
constant $K_r\in(0,\infty)$, as follows (see \cite{csII}):\vspace*{-2pt}
\begin{eqnarray}\label{eq:Kr}
K_r\equiv\int_0^\infty\frac{1-\cos v}{v^{1+r}} \,\mathrm{d}v=|x_1|^{-r}
\int_0^\infty\frac{1-\cos(ux_1)}{u^{1+r}} \,\mathrm{d}u.\vspace*{-2pt}
\end{eqnarray}
Suppose, from now on, that $f_t$ is ${\mathbb Z}^d$-symmetric. Then,\vspace*{-2pt}
\begin{eqnarray*}
\sum_{t=0}^\infty m^t\sum_{x\in{\mathbb Z}^d}|x_1|^rf_t(x)&=&\frac
1{K_r}\int
_0^\infty
\frac{\mathrm{d}u}{u^{1+r}}\sum_{t=0}^\infty m^t\sum_{x\in{\mathbb
Z}^d}
\bigl(1-\cos(ux_1)
\bigr)f_t(x)
\\[-2pt]
&=&\frac1{K_r}\int_0^\infty\frac{\mathrm{d}u}{u^{1+r}} \bigl(\hat
f(0,m)-\hat
f(\vec u,m)\bigr),\vadjust{\goodbreak}
\end{eqnarray*}
\noindent
where $\vec u=(u,0,\dots,0)\in{\mathbb R}^d$. Let
\begin{eqnarray}\label{eq:laplace}
\bar\Delta_l\hat f(k,m)
&\equiv&\hat f(k,m)-\frac{\hat f(k+l,m)+\hat f(k-l,m)}2\nonumber
\\[-8pt]\\[-8pt]
&=&\sum_{t=0}^\infty m^t\sum_{x\in{\mathbb Z}^d}\bigl(1-\cos(l\cdot
x)
\bigr)f_t(x)
e^{ik\cdot x}.\nonumber
\end{eqnarray}
We note that $\bar\Delta_l\hat f(k,m)$ is equivalent to
$\frac{-1}2\Delta_l\hat f(k,m)$ in the previous papers (e.g.,
\cite{csI,csII}). In particular,
\begin{eqnarray*}
\bar\Delta_l\hat f(0,m)=\hat f(0,m)-\hat f(l,m).
\end{eqnarray*}
Therefore, for $r\in(0,\alpha\wedge2)$,
\begin{eqnarray*}
\sum_{t=0}^\infty m^t\sum_{x\in\mathbb{Z}^d}|x_1|^rf_t(x)=\frac1{K_r}
\int_0^\infty\frac{\mathrm{d}u}{u^{1+r}} \bar\Delta_{\vec u}\hat f(0,m).
\end{eqnarray*}
For $r=2j+q<\alpha$ [$j\in\mathbb{N}$, $q\in(0,2)$], we combine the above
representations as
\begin{eqnarray*}
\sum_{t=0}^\infty m^t\sum_{x\in{\mathbb
Z}^d}|x_1|^{2j+q}f_t(x)&=&\frac1{K_q}
\int_0^\infty\frac{\mathrm{d}u}{u^{1+q}}\sum_{t=0}^\infty m^t\sum
_{x\in{\mathbb Z}^d}
\bigl(1-\cos(ux_1)\bigr)x_1^{2j}f_t(x)
\\
&=&\frac{(-1)^j}{K_q}\int_0^\infty\frac{\mathrm{d}u}{u^{1+q}}
\bigl(\nabla_{1}^{2j}
\hat f(0,m)-\nabla_{1}^{2j}\hat f(\vec u,m)\bigr)
\\
&=&\frac{(-1)^j}{K_q}\int_0^\infty\frac{\mathrm{d}u}{u^{1+q}}
\bar\Delta_{\vec u}
\nabla_{1}^{2j}\hat f(0,m).
\end{eqnarray*}
From now on, as long as no confusion arises, we will simply omit $m$ and
abbreviate $\hat f(k,m)$ to $\hat f(k)$. Then, the above three
representations are summarized as
\begin{eqnarray}\label{eq:representation}
&&\sum_{t=0}^\infty m^t\sum_{x\in{\mathbb Z}^d}|x_1|^rf_t(x)\nonumber
\\[-8pt]\\[-8pt]
&&\qquad =
\cases{
(-1)^j\nabla_{1}^{2j}\hat f(0)\qquad [r=2j<\alpha,~j\in\mathbb{N}],
\cr
\displaystyle\dfrac{(-1)^j}{K_q}\int_0^\infty\dfrac{\mathrm
{d}u}{u^{1+q}}
\bar\Delta_{\vec u}\nabla_{1}^{2j}\hat f(0)
\cr
\qquad [r=2j+q<\alpha, j
\in{\mathbb Z}_+, q\in(0,2)].
}\nonumber
\end{eqnarray}
Also, we will abbreviate $\hat f(k,m_{\rm c})$ to $\hat f_{\mathrm c}(k)$
whenever it
is well defined. Moreover, we will use the notation
\begin{eqnarray*}
\partial_m\hat f_{\mathrm c}(k)=\frac{\partial\hat f(k,m)}{\partial
m}\bigg|_{m
\uparrow m_{\rm c}}.
\end{eqnarray*}
Another key element for the proof of the main theorem is the lace expansion
(see, e.g., \cite{s06}, Sections~3 and 13),
\begin{eqnarray}\label{eq:lace-exp}
\varphi_t(x)=I_t(x)+\sum_{s=1}^t(J_s*\varphi_{t-s})(x),
\end{eqnarray}
where, for $t\ge0$,
\begin{eqnarray}\label{eq:I-def}
I_t(x)=
\cases{
\delta_{x,o}\delta_{t,0},&\quad\mbox{RW/SAW},
\cr
\pi_t^\mathrm{OP}(x),&\quad$\mathrm{OP}$,
}
\end{eqnarray}
and for $t\ge1$,
\begin{eqnarray}\label{eq:J-def}
J_t(x)=
\cases{
D(x)\delta_{t,1},&\quad\mbox{RW},
\cr
D(x)\delta_{t,1}+\pi_t^\mathrm{SAW}(x),&\quad\mbox{SAW},
\cr
(\pi_{t-1}^\mathrm{OP}*pD)(x),&\quad\mbox{OP}.
}
\end{eqnarray}
Recall (\ref{eq:RW}) for random walk. For self-avoiding walk and oriented
percolation, $\pi_t^\mathrm{SAW}(x)$ and $\pi_t^\mathrm
{OP}(x)$ are (alternating
sums of)
the model-dependent lace expansion coefficients (see, e.g., \cite
{s06} for
their precise definitions). By~(\ref{eq:lace-exp}), we obtain
\begin{eqnarray}\label{eq:lace-exp-fourier}
\hat\varphi(k)=\hat I(k)+\hat J(k) \hat\varphi(k).
\end{eqnarray}
From this, we can derive identities for the ``derivatives'' of
$\hat\varphi$ in (\ref{eq:representation}). For example,
\begin{eqnarray}\label{eq:derivative-example}
\bar\Delta_{\vec u}\hat\varphi(0)\equiv\hat\varphi(0)-\hat
\varphi
(\vec u)
&=&\hat I(0)+\hat J(0) \hat\varphi(0)-\bigl(\hat I(\vec u)+\hat
J(\vec u)
\hat\varphi(\vec u)\bigr)\nonumber
\\
&=&\bar\Delta_{\vec u}\hat I(0)+\hat J(0) \hat\varphi(0)-\hat
J(\vec
u) \hat
\varphi(\vec u)\nonumber
\\[-8pt]\\[-8pt]
&=&\bar\Delta_{\vec u}\hat I(0)+\hat\varphi(0) \bar\Delta_{\vec
u}\hat
J(0)+\hat
J(\vec u) \bar\Delta_{\vec u}\hat\varphi(0)\nonumber
\\
&=&\frac1{1-\hat J(\vec u)}\bigl(\bar\Delta_{\vec u}\hat I(0)+\hat
\varphi(0)
\bar\Delta_{\vec u}\hat J(0)\bigr),\nonumber
\end{eqnarray}
where the last line has been obtained by solving the previous equation for
$\bar\Delta_{\vec u}\hat\varphi(0)$. Hence, for $r\in(0,\alpha
\wedge2)$,
\begin{eqnarray}\label{eq:derivative-appl}
\sum_{t=0}^\infty m^t\sum_{x\in{\mathbb Z}^d}|x_1|^r\varphi
_t(x)&=&\frac
{\hat\varphi(0)}
{K_r}\int_0^\infty\frac{\mathrm{d}u}{u^{1+r}} \frac{\bar\Delta
_{\vec
u}\hat J
(0)}{1-\hat J(\vec u)}\nonumber
\\[-8pt]\\[-8pt]
&&{}+\frac1{K_r}\int_0^\infty\frac{\mathrm{d}u}{u^{1+r}}
\frac{\bar\Delta_{\vec u}\hat I(0)}{1-\hat J(\vec u)}.\nonumber
\end{eqnarray}
It is known \cite{csI,hhs08} that as long as $d>d_{\mathrm c}$ (and
$L\gg1$), it
is easier to tame $\hat I$ and $\hat J$, up to $m=m_{\rm c}$, than to tame
$\hat\varphi$. We will thus be able to analyze the integrals on the
right-hand side of (\ref{eq:derivative-appl}) and prove the main theorem.
Before closing this subsection, we provide the following
representations for
the constants $C_{\mathrm I}$ and $C_{\mathrm{II}}$ in (\ref{eq:main1})
in terms of $\hat
I_{\mathrm c}$ and
$\hat J_{\mathrm c}$:
\begin{eqnarray}\label{eq:CICII}
C_{\mathrm I}=\frac{\hat I_{\mathrm c}(0)}{m_{\rm c}\,\partial_m\hat
J_{\mathrm c}(0)},\qquad
C_{\mathrm{II}}=\frac1{m_{\rm c}\,\partial_m\hat J_{\mathrm c}(0)}\lim
_{k\to0}\frac
{\bar\Delta_k\hat J_{\mathrm c}
(0)}{\bar\Delta_k\hat D(0)}.
\end{eqnarray}
In Section~\ref{s:CICII}, we will explain the heuristics for the derivation
of these representations.
\subsection{Organization}
In the remainder of the paper, whenever we consider self-avoiding walk
and oriented
percolation, we assume $d>d_{\mathrm c}$ and $L\gg1$, as well as $p\le
p_{\mathrm c}$
for oriented
percolation.
The paper is organized as follows. In Section~\ref{s:proof}, we
prove Theorem~\ref{theorem:main1} for $r\in(0,\alpha\wedge2)$
(Section~\ref{ss:0<r<alpha,2}), for $r=2j<\alpha$ with $j\in\mathbb{N}$
(Section~\ref{ss:r=2j}) and for $r=2j+q<\alpha$ with $j\in\mathbb
{N}$ and
$q\in(0,2)$
(Section~\ref{ss:r=2j+q}) separately, assuming
Propositions~\ref{proposition:r<alpha&2} and \ref
{proposition:2<r<alpha}. We
prove those key propositions in Section~\ref{s:key-prop}.
We strongly believe that the results for self-avoiding walk and oriented
percolation are the most important and interesting parts of this work.
However, for those who are more interested in random walk,
we make the following suggestion: read up to Section~\ref{s:proof}
for the proof of Theorem~\ref{theorem:main1}, where
Propositions~\ref{proposition:r<alpha&2} and \ref{proposition:2<r<alpha}
are used. However, Proposition~\ref{proposition:r<alpha&2} and a part
[i.e., (\ref{eq:2<r<alpha-IJbds})] of Proposition~\ref{proposition:2<r<alpha}
are trivial for random walk. The remaining part [i.e.,
(\ref{eq:2<r<alpha-varphibd})] of Proposition~\ref{proposition:2<r<alpha}
is the result of Lemma~\ref{lmm:induction1}, which is proved in
Section~\ref{ss:prop2}.
\section{Preliminaries}\label{s:CICII}
In this section, we review in outline the derivation in \cite
{csI,csII,hhh} of
the constants $C_{\mathrm I}$ and $C_{\mathrm{II}}$. During the course
of this, we
summarize the
already known properties of $\hat I$ and $\hat J$ and introduce some
quantities used in the following sections.
We begin by solving (\ref{eq:lace-exp-fourier}) for $\hat\varphi(k)$,
which yields
\begin{eqnarray}\label{eq:lace-exp-solution}
\hat\varphi(k)=\frac{\hat I(k)}{1-\hat J(k)},
\end{eqnarray}
where, by (\ref{eq:I-def}) and (\ref{eq:J-def}),
\begin{eqnarray}\label{eq:Ihat}
\hat I(k)&=&
\cases{
1,&\quad\mbox{RW/SAW},
\cr
\hat\pi^\mathrm{OP}(k),&\quad\mbox{OP},
}
\\\label{eq:Jhat}
\hat J(k)&=&
\cases{
m\hat D(k),&\quad\mbox{RW},
\cr
m\hat D(k)+\hat\pi^\mathrm{SAW}(k),&\quad\mbox{SAW},
\cr
\hat\pi^\mathrm{OP}(k)pm\hat D(k),&\quad\mbox{OP}.
}
\end{eqnarray}
It is known \cite{csI,hhs08} that
\begin{eqnarray}\label{eq:property1}
\hat\pi^\mathrm{SAW}(k)=O(L^{-d}) m^2,\qquad
\hat\pi^\mathrm{OP}(k)-1=O(L^{-d}) (pm)^2,
\end{eqnarray}
where the $O(L^{-d})$ terms are uniform in $k\in[-\pi,\pi]^d$ and
$m\le m_{\rm c}$.
Therefore, $\hat I(k)$ and $\hat J(k)$ are both convergent for all
$k\in[-\pi,\pi]^d$ and $m\le m_{\rm c}$. However, since
$\hat\varphi(0)$ diverges as $m\uparrow m_{\rm c}$, we can
characterize $m_{\rm c}$
by the equation
\begin{eqnarray}\label{eq:critpt}
1=\hat J_{\mathrm c}(0)=
\cases{
m_{\rm c},&\quad\mbox{RW},
\cr
m_{\rm c}+\hat\pi_{\mathrm c}^\mathrm{SAW}(0),&\quad\mbox
{SAW},
\cr
\hat\pi_{\mathrm c}^\mathrm{OP}(0)pm_{\rm c},&\quad\mbox{OP}.
}
\end{eqnarray}
Using this identity, we obtain that, as $m\uparrow m_{\rm c}$ (see
\cite{csI,hhh} for
the precise argument),
\begin{eqnarray*}
\hat\varphi(k)&=&\frac{\hat I(k)}{\hat J_{\mathrm c}(0)-\hat
J_{\mathrm c}
(k)+m_{\rm c}{((\hat
J_{\mathrm c}(k)-\hat J(k))/(m_{\rm c}-m))} (1-{m}/{m_{\rm c}})}
\\
&\sim&\frac{\hat I_{\mathrm c}(k)}{\bar\Delta_k\hat J_{\mathrm
c}(0)+m_{\rm
c}\,\partial_m\hat J_{\mathrm c}(k)
(1-{m/m_{\rm c}})}
\\
&=&\frac{\hat I_{\mathrm c}(k)}{\bar\Delta_k\hat J_{\mathrm
c}(0)+m_{\rm c}\,\partial
_m\hat J_{\mathrm c}(k)}
\sum_{t=0}^\infty\biggl(\frac{m_{\rm c}\,\partial_m\hat J_{\mathrm c}
(k)}{\bar\Delta_k\hat
J_{\mathrm c}(0)+m_{\rm c}\,\partial_m\hat J_{\mathrm c}(k)} \frac
{m}{m_{\rm
c}}\biggr)^t,
\end{eqnarray*}
hence,
\begin{eqnarray}\label{eq:lonkey-markus}
\sum_{x\in{\mathbb Z}^d}\varphi_t(x) e^{ik\cdot x}&\mathop{\sim}\limits_{t\to
\infty}&
\frac{\hat I_{\mathrm c}(k)}{\bar\Delta_k\hat J_{\mathrm
c}(0)+m_{\rm c}\,\partial
_m\hat
J_{\mathrm c}(k)} m_{\rm c}^{-t}\nonumber
\\[-8pt]\\[-8pt]
&&{}\times\biggl(1-\frac{\bar\Delta_k\hat J_{\mathrm
c}(0)}{\bar\Delta
_k\hat J_{\mathrm c}(0)
+m_{\rm c}\,\partial_m\hat J_{\mathrm c}(k)}\biggr)^t.\nonumber
\end{eqnarray}
In particular, at $k=0$,
\begin{eqnarray}\label{eq:lonkey-markus0}
\sum_{x\in{\mathbb Z}^d}\varphi_t(x)\sim\frac{\hat I_{\mathrm
c}(0)}{m_{\rm
c}\,\partial_m
\hat J_{\mathrm c}(0)} m_{\rm c}^{-t},
\end{eqnarray}
which yields the representation for $C_{\mathrm I}$ in (\ref{eq:CICII}).
In the above computation, we have used the fact that the quantities
such\vspace*{1pt} as
$m_{\rm c}\,\partial_m\hat J_{\mathrm c}(0)$ and $\bar\Delta_k\hat
J_{\mathrm c}(0)$
are all convergent
uniformly in $k\in[-\pi,\pi]^d$. To see this, we note that, by
(\ref{eq:Jhat}),
\begin{eqnarray}
\qquad m_{\rm c}\partial_m\hat J_{\mathrm c}(k)&=&
\cases{
m_{\rm c}\hat D(k),&\quad\mbox{RW},
\cr
m_{\rm c}\hat D(k)+m_{\rm c}\,\partial_m\hat\pi_{\mathrm
c}^\mathrm{SAW}(k),&\quad\mbox
{SAW},
\cr
\bigl(\hat\pi_{\mathrm c}^\mathrm{OP}(k)+m_{\rm c}\,\partial
_m\hat\pi_{\mathrm c}^\mathrm{OP}
(k)\bigr)pm_{\rm c}\hat D(k),
&\quad\mbox{OP},
}
\\\label{eq:Jhat-laplace}
\bar\Delta_k\hat J(0)&=&
\cases{
m\bar\Delta_k\hat D(0),&\quad\mbox{RW},
\cr
m\bar\Delta_k\hat D(0)+\bar\Delta_k\hat\pi^\mathrm
{SAW}(0),&\quad\mbox{SAW},
\cr
\bigl(\hat\pi^\mathrm{OP}(0) \bar\Delta_k\hat D(0)+\hat
D(k) \bar\Delta
_k\hat\pi^\mathrm{OP}(0)
\bigr)pm,&\quad\mbox{OP}.
}
\end{eqnarray}
However, it is known that $\pi^\mathrm{SAW}$ and $\pi
^\mathrm{OP}$ both satisfy
\begin{eqnarray}\label{eq:property2}
\qquad \quad|m_{\rm c}\,\partial_m\hat\pi_{\mathrm c}(k)|&\le&\sum_{t=0}^\infty t
m_{\rm
c}^t\sum_{x\in{\mathbb Z}^d}|\pi_t
(x)|\le O(L^{-d}),
\\\label{eq:property3}
|\bar\Delta_k\hat\pi(0)|&\le&\sum_{t=0}^\infty m_{\rm c}^t\sum
_{x\in
{\mathbb Z}^d}\bigl(1-\cos(k
\cdot x)\bigr) |\pi_t(x)|\le O(L^{-d}) \bar\Delta_k\hat
D(0)
\end{eqnarray}
for all $k\in[-\pi,\pi]^d$ and $m\le m_{\rm c}$ for the latter (see
\cite{csII}, Proposition~1, \cite{hhh}, the paragraph below
Theorem~1.2 and
\cite{hhs08}, Proposition~4.1, with an improvement due to monotone
convergence). By these bounds and using (\ref{eq:property1}) and (\ref
{eq:critpt}) and
the fact that $m_{\rm c}^\mathrm{SAW}$ and $pm_{\rm
c}^\mathrm{OP}$ are both
$1+O(L^{-d})$ (see the
beginning of Section~\ref{ss:main}), we conclude that
$m_{\rm c}\,\partial_m\hat J_{\mathrm c}(0)=1+O(L^{-d})$ and
$\bar\Delta_k\hat J_{\mathrm c}(0)=O(\bar\Delta_k\hat D(0))$.
Moreover, it has been proven \cite{csI,csII,hhh} that there exist
$\epsilon=\epsilon(d,\alpha)>0$ and $\delta=\delta(d,\alpha)$,
which is zero
if $\alpha=2$ and $>0$ if $\alpha\ne2$, such that $\pi
^\mathrm{SAW}$ and
$\pi^\mathrm{OP}$ both
satisfy
\begin{eqnarray*}
\sum_{t=0}^\infty t^{1+\epsilon}m_{\rm c}^t\sum_{x\in{\mathbb
Z}^d}|\pi
_t(x)|<\infty,\qquad
\sum_{t=0}^\infty m_{\rm c}^t\sum_{x\in{\mathbb Z}^d}|x|^{\alpha
\wedge
2+\delta}|\pi_t(x)|
<\infty.
\end{eqnarray*}
These bounds imply (see \cite{csI}, equations (6.13) and (6.14),
\cite{csII}, equations (3.3)--(3.4), \cite{hhh}, equations (2.25)--(2.28) and (2.64)--(2.70))
\begin{eqnarray}\label{eq:Jhat-diff}
\frac{\hat J_{\mathrm c}(0)-\hat J(0)}{1-{m}/{m_{\rm c}}}&=&m_{\rm
c}\,\partial_m\hat J_{\mathrm c}(0)
+O\biggl(\biggl(1-\dfrac{m}{m_{\rm c}}\biggr)^\epsilon\biggr),
\\\label{eq:M-source}
\frac{\bar\Delta_k\hat J(0)}{\bar\Delta_k\hat D(0)}&=&M+
\cases{
O(|k|^\delta),&\quad$\alpha\ne2$,
\cr
O\biggl(1/\log\dfrac1{|k|}\biggr),&\quad$\alpha=2$,
}
\end{eqnarray}
where the error terms in (\ref{eq:M-source}), which are zero for
random walk,
are uniform in $m\le m_{\rm c}$ and where $M\equiv M(m)$ is defined as
\begin{eqnarray}\label{eq:M-def}
M=
\cases{
m,&\quad\mbox{RW},
\cr
\displaystyle m+\dfrac{\nabla_{1}^2\hat\pi^\mathrm
{SAW}(0)}{-2v_\alpha}{\mathbh1}_{\{\alpha>2\}},
&\quad\mbox{SAW},
\cr
\displaystyle\biggl(\hat\pi^\mathrm{OP}(0)+\dfrac{\nabla
_{1}^2\hat\pi^\mathrm{OP}
(0)}{-2v_\alpha}
{\mathbh1}_{\{\alpha>2\}}\biggr)pm,&\quad\mbox{OP}.
}
\end{eqnarray}
The crossover terms, which are proportional to ${\mathbh1}_{\{\alpha
>2\}}$,
converge for all $m\le m_{\rm c}$ \cite{csII,hhh}.
By~(\ref{eq:lonkey-markus}) and (\ref{eq:lonkey-markus0}) and (\ref
{eq:M-source}), and
using $\lim_{t\to\infty}t \bar\Delta_{k_t}\hat D(0)=|k|^{\alpha
\wedge2}$,
due to the scaling (\ref{eq:k-scaling}), we obtain that, as $t\to
\infty$,
\begin{eqnarray*}
&&\frac{\sum_{x\in{\mathbb Z}^d}\varphi_t(x) e^{ik_t\cdot x}}{\sum
_{x\in{\mathbb Z}^d
}\varphi_t(x)}
\\
&&\qquad \sim\biggl(1-\frac{\bar\Delta_{k_t}\hat J_{\mathrm c}(0)}{\bar
\Delta
_{k_t}\hat J_{\mathrm c}(0)+m_{\rm c}
\,\partial_m\hat J_{\mathrm c}(k_t)}\biggr)^t\qquad[\because(\ref{eq:lonkey-markus})_{k=k_t}
/(\ref{eq:lonkey-markus0})]
\\
&&\qquad \sim\exp\biggl(-\frac{\bar\Delta_{k_t}\hat J_{\mathrm
c}(0)}{m_{\rm
c}\,\partial_m\hat J_{\mathrm c}(0)}
t\biggr)\qquad[\because 1-\tau\sim e^{-\tau}\mbox{ as }\tau\to
0]
\\
&&\qquad=\exp\biggl(-\frac1{m_{\rm c}\,\partial_m\hat J_{\mathrm c}(0)}
\frac
{\bar\Delta_{k_t}\hat J_{\mathrm c}
(0)}{\bar\Delta_{k_t}\hat D(0)} t \bar\Delta_{k_t}\hat D(0)
\biggr)
\\
&&\qquad\sim\exp\biggl(-\frac{M_{\mathrm c}}{m_{\rm c}\,\partial_m\hat
J_{\mathrm c}(0)}
|k|^{\alpha\wedge2}
\biggr),
\end{eqnarray*}
where $M_{\mathrm c}=M(m_{\rm c})$. This yields the representation for
$C_{\mathrm{II}}$ in
(\ref{eq:CICII}).
\begin{Remark*}
It is natural for some readers to wonder why we do not directly prove
(\ref{eq:main2}) by using the formula (\ref{eq:lonkey-markus}) for
$\sum_x\varphi_t(x) e^{ik\cdot x}$, instead of proving the asymptotics
(\ref{eq:main1}) of its generating function and expanding it in powers
of $m$.
In fact, the first-named author was able to derive an asymptotic
expression for $\sum_x|x_1|^r\varphi_t(x)$ using (\ref
{eq:lonkey-markus}), but
the proportionality constant was in a rather complicated sum form. We then
concluded that using (\ref{eq:lonkey-markus}) would not be an ideal
method for
deriving the simplest possible display of the proportionality constant and
started searching for another method. That turns out to be the use of
the generating
function, as explained in this paper. Later, the first-named author
proved that the aforementioned
sum form is indeed an expansion of the proportionality constant in
(\ref{eq:main2}).
\end{Remark*}
\section{Proof of the main results}\label{s:proof}
\subsection{\texorpdfstring{Proof of Theorem~\protect\ref{theorem:main1} for $r\in(0,\alpha
\wedge2)$}{Proof of Theorem~1.1 for $r\in(0,\alpha
\wedge2)$}}\label{ss:0<r<alpha,2}
In this subsection, we prove Theorem~\ref{theorem:main1} for
$r\in(0,\alpha\wedge2)$. We will discuss the case for $\alpha\ne2$
and that
for $\alpha=2$ simultaneously, until we arrive at the point where we require
separate approaches.
First, we recall (\ref{eq:representation}) and split $\int_0^\infty$ into
$\int_0^U$ and $\int_U^\infty$ for a given $U>0$. Using
(\ref{eq:derivative-example}) for the former integral [as in
(\ref{eq:derivative-appl})] and (\ref{eq:laplace}) for
the latter, we obtain
\begin{eqnarray}\label{eq:small-r:dec1}
&&\sum_{t=0}^\infty m^t\sum_{x\in{\mathbb Z}^d}|x_1|^r\varphi
_t(x)\nonumber
\\
&&\qquad=\frac
{\hat\varphi(0)}
{K_r}\int_0^U\frac{\mathrm{d}u}{u^{1+r}} \frac{\bar\Delta_{\vec
u}\hat J(0)}{1-
\hat J(\vec u)}+\frac1{K_r}\int_0^U\frac{\mathrm{d}u}{u^{1+r}}
\frac{\bar\Delta_{\vec u}\hat I(0)}{1-\hat J(\vec u)}
\\
&&\qquad\quad{}+\frac1{K_r}\int_U^\infty\frac{\mathrm{d}u}{u^{1+r}}\sum
_{t=0}^\infty m^t
\sum_{x\in{\mathbb Z}^d}\bigl(1-\cos(ux_1)\bigr)\varphi_t(x).\nonumber
\end{eqnarray}
We note that, by (\ref{eq:Ihat}) and (\ref{eq:property3}),
$\bar\Delta_k\hat I(0)\equiv0$ for random walk and self-avoiding
walk and
$\bar\Delta_k\hat I(0)=O(\bar\Delta_k\hat D(0))$ uniformly in $m\le
m_{\rm c}$ for
oriented percolation. Since $\bar\Delta_k\hat J(0)$ is also
$O(\bar\Delta_k\hat D(0))$ uniformly in $m\le m_{\rm c}$ [see (\ref
{eq:Jhat-laplace})],
the integrals in the first two terms of (\ref{eq:small-r:dec1}) are of
the same
order and therefore the first term dominates the second term as
$m\uparrow m_{\rm c}$, due to the extra factor $\hat\varphi(0)$,
which exhibits
\begin{eqnarray}\label{eq:chi-asy}
\qquad \hat\varphi(0)&=&\frac{\hat I(0)}{\hat J_{\mathrm c}(0)-\hat
J(0)}=\frac
{\hat I_{\mathrm c}(0)+O
(1-{m}/{m_{\rm c}})}{m_{\rm c}\,\partial_m\hat J_{\mathrm c}(0)
(1-
{m}/{m_{\rm c}})+O((1-{m}/{m_{\rm c}})^{1+\epsilon})}\nonumber
\\[-8pt]\\[-8pt]
&=&\frac{C_{\mathrm I}}{1-{m}/{m_{\rm c}}}+O\biggl(\biggl(1-\frac
{m}{m_{\rm
c}}\biggr)^{-1+\epsilon}\biggr),\nonumber
\end{eqnarray}
where the first equality is due to (\ref{eq:lace-exp-solution}) and
(\ref{eq:critpt}), and the second equality is due to (\ref
{eq:property2}) and
(\ref{eq:Jhat-diff}). These estimates are valid independently of $r$
and thus
used in the later sections as well. By the fact that $0\le1-\cos
(ux_1)\le2$, the last term
in (\ref{eq:small-r:dec1}) obeys
\begin{eqnarray}\label{eq:small-r:dec4}
0&\le&\frac1{K_r}\int_U^\infty\frac{\mathrm{d}u}{u^{1+r}}\sum
_{t=0}^\infty m^t
\sum_{x\in{\mathbb Z}^d}\bigl(1-\cos(ux_1)\bigr)\varphi_t(x)\nonumber
\\[-8pt]\\[-8pt]
&\le&
\frac
{2\hat\varphi(0)}
{K_r}\int_U^\infty\frac{\mathrm{d}u}{u^{1+r}}
=\frac{2\hat\varphi(0)}{K_rr}U^{-r}.\nonumber
\end{eqnarray}
We will choose $U$ to be relatively small so as to make the first term in
(\ref{eq:small-r:dec1}) dominant.
Next, we investigate the integral part of the first term in
(\ref{eq:small-r:dec1}),
\begin{eqnarray}\label{eq:small-r:dec3}
\int_0^U\frac{\mathrm{d}u}{u^{1+r}} \frac{\bar\Delta_{\vec u}\hat
J(0)}{1-\hat J
(\vec u)}=\int_0^U\frac{\mathrm{d}u}{u^{1+r}} \frac{\bar\Delta
_{\vec
u}\hat J(0)}
{\hat J_{\mathrm c}(0)-\hat J(0)+\bar\Delta_{\vec u}\hat J(0)},
\end{eqnarray}
where we have used (\ref{eq:critpt}). By~(\ref{eq:D-asympt}) and
(\ref{eq:M-source}),
we have that, for small $u$,
\begin{eqnarray*}
\bar\Delta_{\vec u}\hat J(0)=\frac{\bar\Delta_{\vec u}\hat
J(0)}{\bar\Delta_{\vec u}
\hat D(0)} \bar\Delta_{\vec u}\hat D(0)=
\cases{
Mv_\alpha u^{\alpha\wedge2}+O(u^{\alpha\wedge2+\epsilon}),
&\quad$\alpha\ne2$,\cr
Mv_2u^2\log\dfrac1u+O(u^2),&\quad$\alpha=2$
}
\end{eqnarray*}
for some $\epsilon>0$, where the error terms are uniform in $m\le
m_{\rm c}$. Let
\begin{eqnarray}\label{eq:mu-def}
\mu=\frac{\hat J_{\mathrm c}(0)-\hat J(0)}{Mv_\alpha}.
\end{eqnarray}
Then,
\begin{eqnarray}\label{eq:small-r:dec5}
&&\frac{\bar\Delta_{\vec u}\hat J(0)}{\hat J_{\mathrm c}(0)-\hat
J(0)+\bar\Delta
_{\vec u}\hat
J(0)}\nonumber
\\[-8pt]\\[-8pt]
&&\qquad=
\cases{
\displaystyle\dfrac{u^{\alpha\wedge2}}{\mu+u^{\alpha\wedge
2}}+\frac
{O(u^{\alpha\wedge2
+\epsilon})}{\mu+u^{\alpha\wedge2}},&\quad$\alpha\ne2$,
\cr
\displaystyle\frac{u^2\log1/u}{\mu+u^2\log1/u}+\frac
{O(u^2)}{\mu+u^2\log1/
u},&\quad$\alpha=2$.
}\nonumber
\end{eqnarray}
We now investigate the integral (\ref{eq:small-r:dec3}) for $\alpha
\ne
2$ and
$\alpha=2$ separately, using (\ref{eq:small-r:dec5}) and the following
proposition.
\begin{proposition}\label{proposition:r<alpha&2}
Under the same conditions as in Theorem~\ref{theorem:main1},
\begin{eqnarray}\label{eq:M-asy}
M&=&M_{\mathrm c}+O\biggl(\biggl(1-\frac{m}{m_{\rm c}}\biggr)^\epsilon
\biggr),
\\\label{eq:mu-asy}
\mu&=&\frac{1-{m}/{m_{\rm c}}}{C_{\mathrm{II}}v_\alpha}+O
\biggl(\biggl(1-\frac
{m}{m_{\rm c}}\biggr)^{1+
\epsilon}\biggr)
\end{eqnarray}
for some $\epsilon>0$, where $M_{\mathrm c}=M(m_{\rm c})$.
\end{proposition}
The proof is deferred to Section~\ref{s:key-prop}. We note that these
estimates are trivial for random walk.
\subsubsection{Proof for $\alpha\ne2$}
We assume that $\epsilon<r$, without loss of generality. By~(\ref
{eq:small-r:dec3})
and (\ref{eq:small-r:dec5}) for $\alpha\ne2$, we have that, for
small $U$,
\begin{eqnarray*}
\int_0^U\frac{\mathrm{d}u}{u^{1+r}} \frac{\bar\Delta_{\vec u}\hat
J(0)}{1-\hat
J(\vec u)}&=&\int_0^U\frac{\mathrm{d}u}u\biggl(\frac{u^{\alpha\wedge
2-r}}{\mu
+u^{\alpha\wedge2}}+\frac{O(u^{\alpha\wedge2-r+\epsilon})}{\mu
+u^{\alpha
\wedge2}}\biggr)
\\
&=&\int_0^\infty\frac{\mathrm{d}u}u \frac{u^{\alpha\wedge2-r}}{\mu
+u^{\alpha
\wedge2}}-\int_U^\infty\frac{\mathrm{d}u}u \frac{u^{\alpha\wedge
2-r}}{\mu
+u^{\alpha\wedge2}}
\\
&&{}+\int_0^U\frac{\mathrm{d}u}u \frac{O(u^{\alpha\wedge
2+\epsilon-r})}
{\mu+u^{\alpha\wedge2}}\bigl({\mathbh1}_{\{\mu\ge u^{\alpha\wedge
2}\}}+{\mathbh1}_{\{\mu<u^{\alpha \wedge2}\}}\bigr)
\\
&=&\int_0^\infty\frac{\mathrm{d}u}u \frac{u^{\alpha\wedge2-r}}{\mu
+u^{\alpha
\wedge2}}+O(U^{-r})+O\bigl(\mu^{-{(r-\epsilon)/(\alpha\wedge2)}}\bigr).
\end{eqnarray*}
Let $U=\mu^{{(1-\epsilon/r)/(\alpha\wedge2)}}$, which is indeed
small as
$m\uparrow m_{\rm c}$, due to Proposition~\ref{proposition:r<alpha&2}.
By the change
of variables $u^{\alpha\wedge2}=\mu z$, we obtain
\begin{eqnarray}\label{eq:standard-cauchy}
&&\int_0^{\mu^{{(1-\epsilon/r)/(\alpha\wedge2)}}}\frac{\mathrm
{d}u}{u^{1+r}}
\frac{\bar\Delta_{\vec u}\hat J(0)}{1-\hat J(\vec u)}\nonumber
\\
&&\qquad =\int
_0^\infty
\frac{\mathrm{d}u}u \frac{u^{\alpha\wedge2-r}}{\mu+u^{\alpha
\wedge2}}
+O\bigl(\mu^{-{(r-\epsilon)/(\alpha\wedge2)}}\bigr)
\\
&&\qquad=\frac{\mu^{-{r/(\alpha\wedge2)}}}{\alpha\wedge2}\int
_0^\infty\frac{\mathrm
{d}z}z \frac{z^{1-{r/(\alpha\wedge2)}}}{1+z}+O\bigl(\mu^{-
{(r-\epsilon)/(\alpha\wedge2)}}\bigr).\nonumber
\end{eqnarray}
However, by the standard Cauchy integral formula, for $\beta\in(0,1)$,
\begin{eqnarray}\label{eq:standard-cauchy1}
\oint_{\gamma_1}\frac{\mathrm{d}z}z \frac{z^{1-\beta}}{1+z}=\oint
_{\gamma_2}
\frac{\mathrm{d}z}z \frac{z^{1-\beta}}{1+z}=2\pi i(-1)^{-\beta
}=2\pi i
e^{-\pi i\beta},
\end{eqnarray}
where, as depicted in Figure~\ref{fig:contour1}, the contour $\gamma_1$
consists of two line segments, an arc of the circle with smaller radius
$\delta\in(0,1)$ and an arc of the circle with larger radius $R\in
(1,\infty)$, and
the contour $\gamma_2$ is the circle centered at $-1$ with radius smaller
than~1. On the other hand, by taking $\delta\to0$ and $R\to\infty$,
we obtain
\begin{eqnarray*}
\mathop{\lim_{R\to\infty}}_{\delta\to0}\oint_{\gamma_1}\frac
{\mathrm{d}z}z
\frac{z^{1-\beta}}{1+z}
=(1-e^{-2\pi i\beta})\int_0^\infty\frac{\mathrm{d}z}z \frac
{z^{1-\beta}}{1+z}.
\end{eqnarray*}
Therefore,
\begin{eqnarray*}
\int_0^\infty\frac{\mathrm{d}z}z \frac{z^{1-\beta}}{1+z}=\frac
{2\pi i e^{-\pi
i\beta}}{1-e^{-2\pi i\beta}}=\frac\pi{\sin(\beta\pi)},
\end{eqnarray*}
which implies that
\begin{eqnarray*}
\int_0^{\mu^{{(1-\epsilon/r)/(\alpha\wedge2)}}}\frac{\mathrm
{d}u}{u^{1+r}}
\frac{\bar\Delta_{\vec u}\hat J(0)}{1-\hat J(\vec u)}&=&\frac\pi
{(\alpha\wedge2)
\sin(r\pi/(\alpha\wedge2))} \mu^{-{r/(\alpha\wedge 2)}}
\\
&&{}+O\bigl(\mu^{-
{(r-\epsilon)/(\alpha\wedge2)}}\bigr).
\end{eqnarray*}
\begin{figure}
\includegraphics{557f01.eps}
\caption{The contours $\gamma_1$ and $\gamma_2$
in the complex plane.}\label{fig:contour1}
\end{figure}
Finally, by substituting (\ref{eq:standard-cauchy}) back into (\ref
{eq:small-r:dec1})
and using (\ref{eq:chi-asy}) and (\ref{eq:mu-asy}), we conclude that there
is an
$\epsilon'\in(0,1)$ such that
\begin{eqnarray}\label{eq:pr4r<2}
\qquad\ \ \sum_{t=0}^\infty m^t\sum_{x\in{\mathbb Z}^d}|x_1|^r\varphi
_t(x)&=&\frac{\pi
K_r^{-1}}
{(\alpha\wedge2)\sin(r\pi/(\alpha\wedge2))} \frac
{C_{\mathrm I}
(C_{\mathrm{II}}v_\alpha)
^{{r/(\alpha\wedge2)}}}{(1-{m/m_{\rm c}})^{1+
{r/(\alpha\wedge2)}}}\nonumber
\\[-8pt]\\[-8pt]
&&{}+O\biggl(\biggl(1-\frac{m}{m_{\rm c}}\biggr)^{-1-{r/(\alpha\wedge
2)}+\epsilon'}\biggr).\nonumber
\end{eqnarray}
However, since (see Appendix~\hyperref[appendix:K]{A.2})
\begin{eqnarray}\label{eq:Kr-evaluation}
\pi K_r^{-1}=2\Gamma(r+1)\sin\frac{r\pi}2,
\end{eqnarray}
this completes the proof of Theorem~\ref{theorem:main1} for
$0<r<\alpha\wedge2$ with $\alpha\ne2$.
\begin{Remark*}
Although the proportionality constant
$(2\sin\frac{r\pi}{\alpha\vee2})/((\alpha\wedge2)\sin\frac
{r\pi}\alpha)$ in
(\ref{eq:main1}) looks slightly different from the constant
$(2\sin\frac{r\pi}2)/((\alpha\wedge2)\sin\frac{r\pi}{\alpha
\wedge2})$ derived
from (\ref{eq:pr4r<2}) and (\ref{eq:Kr-evaluation}), they are equal when
$0<r<\alpha\wedge2$. The reason why we have adopted the former in the
main theorem is due to its applicability to larger values of $r<\alpha
$, which
the latter lacks (e.g., take $r=3<\alpha$).
\end{Remark*}
\subsubsection{Proof for $\alpha=2$}
The proof for $\alpha=2$ is slightly more involved than the above
proof for
$\alpha\ne2$, due to the log corrections in (\ref{eq:small-r:dec5}). By
(\ref{eq:small-r:dec3}) and (\ref{eq:small-r:dec5}) for $\alpha=2$, we
have that,
for small $U$,
\begin{eqnarray*}
\int_0^U\frac{\mathrm{d}u}{u^{1+r}} \frac{\bar\Delta_{\vec u}\hat
J(0)}{1-\hat J
(\vec u)}&=&\int_0^U\frac{\mathrm{d}u}u\biggl(\frac{u^{2-r}\log
1/u}{\mu+u^2
\log1/u}+\frac{O(u^{2-r})}{\mu+u^2\log1/u}\biggr)
\\
&=&\int_0^U\frac{\mathrm{d}u}u \frac{u^{2-r}\log1/u}{\mu
+u^2\log1/u}
+\frac{O(U^{2-r})}\mu,
\end{eqnarray*}
where we have obtained the error term by simply ignoring $u^2\log\frac1u>0$
in the denominator. Let $U=\sqrt\mu$, which is small as $m\uparrow
m_{\rm c}$, as
required, due to Proposition~\ref{proposition:r<alpha&2}. By the
change of
variables $u^2\log\frac1u=\mu z$, we obtain
\begin{eqnarray*}
\int_0^{\sqrt\mu}\frac{\mathrm{d}u}{u^{1+r}} \frac{\bar\Delta
_{\vec
u}\hat J(0)}
{1-\hat J(\vec u)}&=&\int_0^{\sqrt\mu}\frac{\mathrm{d}u}u \frac
{u^{2-r}\log
1/u}{\mu+u^2\log1/u}+O(\mu^{-r/2})
\\
&=&\frac{\mu^{-r/2}}2\int_0^{\log1/{\sqrt\mu}}\frac{\mathrm
{d}z}z \frac{z^{1
-{r}/2}(\log1/{u(z)})^{{r}/2}}{1+z}+O(\mu^{-r/2}).
\end{eqnarray*}
Note that, by taking the logarithm of $u^2\log1/u=\mu z$ and
using the
monotonicity of $(\log\log1/u)/\log1/u$ in $0<u<\sqrt\mu
\ll1$, we have
\begin{eqnarray*}
\log\frac1{u(z)}=\biggl(1+O\biggl(\frac{\log\log1/{\sqrt\mu
}}{\log1/
\mu}\biggr)\biggr)\log\frac1{\sqrt{\mu z}}.
\end{eqnarray*}
Therefore,
\begin{eqnarray*}
&&\int_0^{\sqrt\mu}\frac{\mathrm{d}u}{u^{1+r}} \frac{\bar\Delta
_{\vec u}\hat J(0)}
{1-\hat J(\vec u)}
\\
&&\qquad=\frac{\mu^{-r/2}}2\biggl(1+O\biggl(\frac{\log\log1/{\sqrt
\mu}}{\log1/
\mu}\biggr)\biggr)\int_0^{\log1/{\sqrt\mu}}\frac{\mathrm
{d}z}z \frac{z^{1-
{r}/2}(\log1/{\sqrt{\mu z}})^{{r}/2}}{1+z}
\\
&&{}\qquad\quad +O(\mu^{-r/2}).
\end{eqnarray*}
Suppose that $\log\frac1{\sqrt\mu}\gg1$. Then, by the Cauchy
integral formula (see
Figure~\ref{fig:contour1}),
\begin{eqnarray*}
\oint_{\gamma_1}\frac{\mathrm{d}z}z \frac{z^{1-{r}/2}(\log
1/{\sqrt{\mu
z}})^{{r}/2}}{1+z}&=&\oint_{\gamma_2}\frac{\mathrm{d}z}z \frac
{z^{1-{r}/
2}(\log1/{\sqrt{\mu z}})^{{r}/2}}{1+z}
\\
&=&2\pi i e^{-\pi ir/2}\biggl(\log\frac1{\sqrt\mu}-\frac{\pi
i}2\biggr)^{r/2}
\\
&=&2\pi i e^{-\pi ir/2}\biggl(\log\frac1{\sqrt\mu}\biggr)^{r/2}+O(1),
\end{eqnarray*}
where, as in (\ref{eq:standard-cauchy1}), the contour $\gamma_2$ is
the circle
at $-1$ with radius smaller than 1, while the contour $\gamma_1$
contains an
arc of the circle with radius $\delta\in(0,1)$ and an arc of the
circle with
radius $R\equiv\log\frac1{\sqrt\mu}$. On the other hand, by taking
$\delta\to0$, we obtain
\begin{eqnarray*}
&&\lim_{\delta\to0}\oint_{\gamma_1}\frac{\mathrm{d}z}z \frac
{z^{1-{r}/2}(\log
1/{\sqrt{\mu z}})^{{r}/2}}{1+z}
\\
&&\qquad =(1-e^{-\pi ir})\int_0^{\log1/
{\sqrt\mu}}\frac{\mathrm{d}z}z \frac{z^{1-{r}/2}(\log
1/{\sqrt{\mu
z}})^{{r}/2}}{1+z}+O(1),
\end{eqnarray*}
where the error term is independent of $\mu$. Therefore,
\begin{eqnarray*}
\int_0^{\log1/{\sqrt\mu}}\frac{\mathrm{d}z}z \frac
{z^{1-{r}/2}(\log
1/{\sqrt{\mu z}})^{{r}/2}}{1+z}&=&\frac{2\pi i e^{-\pi ir/2}}{1-
e^{-\pi ir}}\biggl(\log\frac{1}{\sqrt\mu}\biggr)^{r/2}+O(1)
\\
&=&\frac\pi{\sin{(r\pi/2)}}\biggl(\log\frac1{\sqrt\mu}
\biggr)^{r/2}+O(1),
\end{eqnarray*}
which implies
that
\begin{eqnarray}\label{eq:generalized-cauchy}
\qquad \int_0^{\sqrt\mu}\frac{\mathrm{d}u}{u^{1+r}} \frac{\bar\Delta
_{\vec
u}\hat
J(0)}{1-\hat J(\vec u)}=\frac\pi{2\sin({r\pi}/2)} \mu
^{-r/2}\biggl(\log\frac1
{\sqrt\mu}\biggr)^{r/2}+O(\mu^{-r/2}),
\end{eqnarray}
where we have used
\begin{eqnarray*}
O\biggl(\frac{\log\log1/{\sqrt\mu}}{\log1/\mu}
\biggr)\biggl(\log\frac1
{\sqrt\mu}\biggr)^{r/2}=o(1)\qquad[\because r<2].
\end{eqnarray*}
Finally, by substituting (\ref{eq:generalized-cauchy}) back into
(\ref{eq:small-r:dec1}) and using (\ref{eq:small-r:dec4}) with
$U=\sqrt\mu$,
we obtain
\begin{eqnarray*}
\sum_{t=0}^\infty m^t\sum_{x\in{\mathbb Z}^d}|x_1|^r\varphi
_t(x)&=&\bigl(\hat
\varphi(0)
+O(1)\bigr)\frac{\pi K_r^{-1}}{2\sin({r\pi}/2)} \mu^{-r/2}
\biggl(\log\frac1
{\sqrt\mu}\biggr)^{r/2}
\\
&&{}+\hat\varphi(0) O(\mu^{-r/2}).
\end{eqnarray*}
Combining this with (\ref{eq:chi-asy}), (\ref{eq:mu-asy}) and (\ref
{eq:Kr-evaluation})
yields (\ref{eq:main1}) for $\alpha=2$. This completes the proof of
Theorem~\ref{theorem:main1} for $0<r<\alpha=2$.
\subsection{\texorpdfstring{Proof of Theorem~\protect\ref{theorem:main1} for $r=2j<\alpha$
[$j\in\mathbb{N}$]}{Proof of Theorem~1.1 for $r=2j<\alpha$
[$j\in\mathbb{N}$]}}\label{ss:r=2j}
In this subsection, we prove Theorem~\ref{theorem:main1} for positive even
integers $r=2j<\alpha$. First, we recall (\ref{eq:representation})
for $r=2j$:
\begin{eqnarray*}
\sum_{t=0}^\infty m^t\sum_{x\in{\mathbb Z}^d}x_1^{2j}\varphi_t(x)
&=(-1)^j\nabla_{1}^{2j}\hat\varphi(0).
\end{eqnarray*}
Differentiating (\ref{eq:lace-exp-fourier}) and using the ${\mathbb
Z}^d$-symmetry
of the
models [so that
$\nabla_{1}^n\hat J(0)$ and $\nabla_{1}^n\hat\varphi(0)$ are both zero
when $n$ is
odd], we have
\begin{eqnarray*}
\nabla_1^{2j}\hat\varphi(0)=\nabla_{1}^{2j}\hat I(0)+\hat J(0)
\nabla_{1}^{2j}\hat
\varphi(0)+\sum_{l=1}^j\pmatrix{2j\cr 2l}\nabla_{1}^{2l}\hat J(0)
\nabla_{1}^{2(j-l)}
\hat\varphi(0).
\end{eqnarray*}
Solving this equation for $\nabla_1^{2j}\hat\varphi(0)$ and using
(\ref{eq:lace-exp-solution}) for $k=0$, we obtain
\begin{eqnarray}\label{eq:nablone2j-lace}
\nabla_1^{2j}\hat\varphi(0)=\frac{\hat\varphi(0)}{\hat I(0)}
\Biggl(\nabla_{1}^{2j}
\hat I(0)+\sum_{l=1}^j\pmatrix{2j\cr 2l}\nabla_{1}^{2l}\hat J(0) \nabla_{1}
^{2(j-l)}
\hat\varphi(0)\Biggr).
\end{eqnarray}
To identify the dominant term of the right-hand side, we use the following
proposition.
\begin{proposition}\label{proposition:2<r<alpha}
Let $\alpha>2$ and $\langle\alpha\rangle=\max\{j\in\mathbb
{N}\dvtx j<\alpha\}
$ (note that
$\langle\alpha\rangle=\alpha-1$ if $\alpha\ge3$ is an integer).
Under the
same conditions as in Theorem~\ref{theorem:main1},\vspace*{-2pt}
\begin{eqnarray}\label{eq:2<r<alpha-IJbds}
\qquad \quad \left.\begin{array}{r}
\displaystyle\sum_{t=0}^\infty m^t\sum_{x\in{\mathbb Z}^d}|x_1|^\nu
|I_t(x)|\\
\displaystyle\sum_{t=0}^\infty m^t\sum_{x\in{\mathbb Z}^d}|x_1|^\nu|J_t(x)|
\end{array}
\right\}\le
\cases{
O(1),&\quad$0\le\nu\le2$,
\cr
O\biggl(\biggl(1-\dfrac{m}{m_{\rm c}}\biggr)^{1-\nu/2+\epsilon}\biggr),
&\quad$2<\nu<\alpha$
}\hspace*{-15pt}\vspace*{-2pt}
\end{eqnarray}
for some $\epsilon>0$. Moreover,\vspace*{-2pt}
\begin{eqnarray}\label{eq:2<r<alpha-varphibd}
\quad &&|\nabla_{1}^n\hat\varphi(k,me^{i\theta})|\nonumber
\\[-9pt]\\[-9pt]
&&\qquad \le
\cases{
O\biggl(\biggl(1-\dfrac{m}{m_{\rm c}}+|\theta|+|k|^2\biggr)^{-1-{n}/2}
\biggr),
&\quad$n=0,1,2$,
\cr
\dfrac{O((1-{m}/{m_{\rm c}})^{1-{n}/2}
)}{(1-{m}/{m_{\rm c}}
+|\theta|+|k|^2)^2},&\quad$n=3,\dots,\langle\alpha\rangle$,
}\nonumber\vspace*{-2pt}
\end{eqnarray}
where the $O((1-\frac{m}{m_{\rm c}})^{1-{n}/2})$ term is uniform in
$(k,\theta)\in[-\pi,\pi]^{d+1}$.
\end{proposition}
We will use this proposition again in the next subsection to prove
Theorem~\ref{theorem:main1} for the remaining case: $r=2j+q$, where
$j\in\mathbb{N}$
and $q\in(0,2)$. The proof of Proposition~\ref{proposition:2<r<alpha} is
deferred to Section~\ref{s:key-prop}. Note that (\ref{eq:2<r<alpha-IJbds})
is trivial for random walk.
Now we resume the proof of Theorem~\ref{theorem:main1} for $r=2j$.
Notice that\vspace*{-2pt}
\begin{eqnarray}\label{eq:nablone2lJ}
|\nabla_{1}^{2l}\hat J(0)|\le\sum_{t=0}^\infty m^t\sum_{x\in
{\mathbb Z}^d}|x_1|^{2l}
|J_t(x)|\vspace*{-2pt}
\end{eqnarray}
and that a similar bound holds for $I$. By
(\ref{eq:nablone2j-lace})--(\ref{eq:2<r<alpha-varphibd}), we have the
recursion\vspace*{-2pt}
\begin{eqnarray*}
\nabla_{1}^{2j}\hat\varphi(0)&=&\frac{\hat\varphi(0)}{\hat
I(0)}\Biggl(\nabla_{1}^{2
j}\hat I(0)+\pmatrix{2j\cr 2}\nabla_{1}^2\hat J(0) \nabla
_{1}^{2(j-1)}\hat
\varphi(0)
\\[-2pt]
&&{}\hspace*{50pt}+\sum_{l=2}^j\pmatrix{2j\cr 2l}\nabla_{1}^{2l}\hat J(0)
\nabla_{1}^{2(j-l)}
\hat\varphi(0)\Biggr)
\\[-2pt]
&=&\pmatrix{2j\cr 2}\frac{\nabla_{1}^2\hat J(0)}{\hat I(0)} \hat\varphi
(0) \nabla_{1}^{2
(j-1)}\hat\varphi(0)+O\biggl(\biggl(1-\frac{m}{m_{\rm c}}\biggr)^{-1-j+\epsilon
}\biggr),\vspace*{-2pt}
\end{eqnarray*}
where the first term is $O((1-\frac{m}{m_{\rm c}})^{-1-j})$, which is
dominant as
$m\uparrow m_{\rm c}$. Repeated use of this recursion then yields
\begin{eqnarray*}
\nabla_{1}^{2j}\hat\varphi(0)&=&\pmatrix{2j\cr2}\pmatrix{2(j-1)\cr 2}
\biggl(\frac{\nabla_{1}^2
\hat J(0)}{\hat I(0)} \hat\varphi(0)\biggr)^2\nabla
_{1}^{2(j-2)}\hat
\varphi
(0)
\\[-2pt]
&&{}+O\biggl(\biggl(1-\frac{m}{m_{\rm c}}\biggr)^{-1-j+\epsilon}\biggr)
\\
&\vdots&
\\[-2pt]
&=&\prod_{l=2}^j\pmatrix{2l\cr 2}\biggl(\frac{\nabla_{1}^2\hat J(0)}{\hat
I(0)} \hat
\varphi(0)\biggr)^{j-1}\nabla_{1}^2\hat\varphi(0)+O\biggl(\biggl(1-\frac
{m}{m_{\rm c}}\biggr)^{-1-j
+\epsilon}\biggr)
\\[-2pt]
&=&\frac{(2j)!}{2^j}\biggl(\frac{\nabla_{1}^2\hat J(0)}{\hat
I(0)}
\biggr)^j\hat
\varphi(0)^{j+1}+O\biggl(\biggl(1-\frac{m}{m_{\rm c}}\biggr)^{-1-j+\epsilon}\biggr).
\end{eqnarray*}
However, by comparing (\ref{eq:nablone}) and (\ref{eq:laplace}), and using
(\ref{eq:M-asy}),
we have
\begin{eqnarray*}
\nabla_{1}^2\hat J(0)&=&-2v_\alpha\lim_{k\to0}\frac{\bar\Delta
_k\hat
J(0)}{\bar\Delta_k
\hat D(0)}=-2v_\alpha M
\\[-2pt]
&=&-2v_\alpha M_{\mathrm c}+O\biggl(\biggl(1-\frac{m}{m_{\rm c}}\biggr)^\epsilon
\biggr).
\end{eqnarray*}
Recall that $\hat I(0)=\hat I_{\mathrm c}(0)+O(1-\frac{m}{m_{\rm c}})$
[cf.~the numerator
in (\ref{eq:chi-asy})]. Therefore,
\begin{eqnarray}\label{eq:ratio}
\quad \frac{\nabla_{1}^2\hat J(0)}{\hat I(0)}&=&-2v_\alpha\frac{M_{\mathrm
c}}{\hat
I_{\mathrm c}(0)}
+O\biggl(\biggl(1-\frac{m}{m_{\rm c}}\biggr)^\epsilon\biggr)\nonumber
\\[-9pt]\\[-9pt]
&=&-2v_\alpha\frac{C_{\mathrm{II}}}{C_{\mathrm I}}+O\biggl(\biggl(1-\frac
{m}{m_{\rm
c}}\biggr)^\epsilon\biggr)\qquad
[\because(\ref{eq:CICII})\mbox{ and }(\ref{eq:M-source})],\nonumber
\end{eqnarray}
hence
\begin{eqnarray*}
\nabla_{1}^{2j}\hat\varphi(0)&=&\frac{(2j)!}{2^j}\biggl(-2v_\alpha
\frac{C_{\mathrm{II}}}{C_{\mathrm I}}
\biggr)^j\biggl(\frac{C_{\mathrm I}}{1-{m/m_{\rm c}}}
\biggr)^{j+1}+O\biggl(\biggl(1-\frac{m}
{m_{\rm c}}\biggr)^{-1-j+\epsilon}\biggr)
\\[-2pt]
&=&\Gamma(2j+1) \frac{C_{\mathrm I}(-C_{\mathrm{II}}v_\alpha
)^j}{(1-{m/m_{\rm
c}})^{j+1}}+O\biggl(
\biggl(1-\frac{m}{m_{\rm c}}\biggr)^{-1-j+\epsilon}\biggr).
\end{eqnarray*}
This completes the proof of Theorem~\ref{theorem:main1} for positive even
integers $r=2j<\alpha$.
\subsection{\texorpdfstring{Proof of Theorem~\protect\ref{theorem:main1} for $r=2j+q<\alpha$
[$j\in\mathbb{N}$, $q\in(0,2)$]}{Proof of Theorem~1.1 for $r=2j+q<\alpha$
[$j\in\mathbb{N}$, $q\in(0,2)$]}}\label{ss:r=2j+q}
In this subsection, we prove Theorem~\ref{theorem:main1} for the other values
of $r<\alpha$: $r=2j+q$ with $j\in\mathbb{N}$ and $q\in(0,2)$.
First, we recall
(\ref{eq:representation}):
\begin{eqnarray}\label{eq:pr42j+q:1}
\sum_{t=0}^\infty m^t\sum_{x\in{\mathbb Z}^d}|x_1|^r\varphi
_t(x)=\frac{(-1)^j}{K_q}
\int_0^\infty\frac{\mathrm{d}u}{u^{1+q}} \bar\Delta_{\vec u}\nabla_{1}
^{2j}\hat
\varphi(0),
\end{eqnarray}
where, by (\ref{eq:lace-exp-fourier}),
\begin{eqnarray*}
\bar\Delta_{\vec u}\nabla_{1}^{2j}\hat\varphi(0)&=&\nabla
_{1}^{2j}\hat
\varphi(0)
-\nabla_{1}^{2j}\hat\varphi(\vec u)
\\[-2pt]
&=&\nabla_{1}^{2j}\hat I(0)-\nabla_{1}^{2j}\hat I(\vec u)+\sum
_{n=0}^{2j}\pmatrix{2j\cr
n}\bigl(\nabla_{1}^n\hat J(0) \nabla_{1}^{2j-n}\hat\varphi(0)
\\[-2pt]
&&{}\hspace*{150pt}-\nabla_{1}^n\hat J(\vec u) \nabla_{1}^{2j-n}\hat
\varphi
(\vec u)\bigr)
\\
&=&\bar\Delta_{\vec u}\nabla_{1}^{2j}\hat I(0)+\sum_{n=0}^{2j}
\pmatrix{2j\cr n }\bigl(
\nabla_{1}^n\hat J(0)~\bar\Delta_{\vec u}\nabla_{1}^{2j-n}\hat
\varphi (0)
\\
&&{}\hspace*{112pt}+\nabla_{1}^{2j-n}\hat\varphi(\vec u) \bar\Delta_{\vec
u}\nabla_{1}^n
\hat J(0)\bigr).
\end{eqnarray*}
Solving this equation for $\bar\Delta_{\vec u}\nabla_{1}^{2j}\hat
\varphi
(0)$ and
using (\ref{eq:lace-exp-solution}) for $k=0$ and $\nabla_{1}^n\hat J(0)=0$
for odd
$n$, we obtain
\begin{eqnarray*}
\bar\Delta_{\vec u}\nabla_{1}^{2j}\hat\varphi(0)&=&\frac{\hat
\varphi
(0)}{\hat I(0)}
\Biggl(\bar\Delta_{\vec u}\nabla_{1}^{2j}\hat I(0)+\sum
_{l=1}^j\pmatrix{2j\cr 2l}
\nabla_{1}^{2l}\hat J(0) \bar\Delta_{\vec u}\nabla_{1}^{2(j-l)}\hat
\varphi(0)
\\
&&{}\hspace*{86pt}+\sum_{n=0}^{2j}\pmatrix{2j\cr n}\nabla_{1}^{2j-n}\hat\varphi(\vec
u) \bar\Delta_{\vec u}
\nabla_{1}^n\hat J(0)\Biggr).
\end{eqnarray*}
Substituting this back into (\ref{eq:pr42j+q:1}) yields
\begin{eqnarray}\label{eq:pr42j+q:2}
&&\sum_{t=0}^\infty m^t\sum_{x\in{\mathbb Z}^d}|x_1|^r\varphi\nonumber
_t(x)
\\[-8pt]\\[-8pt]
&&\qquad=\frac{\hat
\varphi(0)}
{\hat I(0)}\Biggl(H^{(1)}+\sum_{l=1}^j\pmatrix{2j\cr 2l}H_{2l}^{(2)}+
\sum_{n=0}^{2j}\pmatrix{2j\cr n}H_n^{(3)}\Biggr),\nonumber
\end{eqnarray}
where
\begin{eqnarray}\label{eq:H1-def}
H^{(1)}&=&\frac{(-1)^j}{K_q}\int_0^\infty\frac{\mathrm
{d}u}{u^{1+q}}
\bar\Delta_{\vec u}\nabla_{1}^{2j}\hat I(0)\equiv\sum_{t=0}^\infty m^t
\sum_{x\in{\mathbb Z}^d}|x_1|^rI_t(x),
\\\label{eq:H2-def}
H_{2l}^{(2)}&=&\frac{(-1)^j}{K_q}\int_0^\infty\frac{\mathrm
{d}u}{u^{1+q}}
\nabla_{1}^{2l}\hat J(0) \bar\Delta_{\vec u}\nabla_{1}^{2(j-l)}\hat
\varphi(0)\nonumber
\\[-8pt]\\[-8pt]
&\equiv&(-1)^l\nabla_{1}^{2l}\hat J(0)\sum_{t=0}^\infty m^t\sum
_{x\in
{\mathbb Z}^d}|x_1|^{r
-2l}\varphi_t(x)\nonumber
\end{eqnarray}
and
\begin{eqnarray}\label{eq:H3-def}
H_n^{(3)}&=&\frac{(-1)^j}{K_q}\int_0^\infty\frac{\mathrm
{d}u}{u^{1+q}}
\nabla_{1}^{2j-n}\hat\varphi(\vec u)~\bar\Delta_{\vec u}\nabla
_{1}^n\hat
J(0)\nonumber
\\
&\equiv&\sum_{s,t=0}^\infty m^{t+s}\sum_{x,y\in{\mathbb Z}^d
}x_1^{2j-n}\varphi_t(x)
y_1^nJ_s(y)
\\
&&{}\times\frac1
{K_q}\int_0^\infty\frac{\mathrm{d}u}{u^{1+q}}\times
\cases{
\sin(ux_1) \sin(uy_1),&\quad\mbox{odd }$n$,
\cr
\cos(ux_1) \bigl(1-\cos(uy_1)\bigr),&\quad\mbox{even }$n$.
}
\nonumber
\end{eqnarray}
Next, we isolate error terms from (\ref{eq:pr42j+q:2}) using
Proposition~\ref{proposition:2<r<alpha}. First, by (\ref{eq:2<r<alpha-IJbds}),
we have
\begin{eqnarray}\label{eq:H1-bd}
\bigl|H^{(1)}\bigr|\le\sum_{t=0}^\infty m^t\sum_{x\in{\mathbb Z}^d
}|x_1|^r|I_t(x)|\le O\biggl(\biggl(1
-\frac{m}{m_{\rm c}}\biggr)^{-{r}/2+1+\epsilon}\biggr),
\end{eqnarray}
which gives rise to an error term.
Next, for $H_{2l}^{(2)}$, where $r-2l=2j+q-2l<2j+2-2l<\alpha$, we first
apply Jensen's inequality and then (\ref{eq:2<r<alpha-varphibd}) to obtain
\begin{eqnarray}\label{eq:Jensen}
\quad&&\sum_{t=0}^\infty m^t\sum_{x\in{\mathbb Z}^d}|x_1|^{r-2l}\varphi
_t(x)\nonumber
\\
&&\qquad \le
\Biggl(\frac1
{\hat\varphi(0)}\sum_{t=0}^\infty m^t\sum_{x\in{\mathbb Z}^d
}|x_1|^{2j+2-2l}\varphi_t
(x)\Biggr)^{{(r-2l)/(2j+2-2l)}}\hat\varphi(0)\nonumber
\\
&&\qquad=\biggl(\frac{|\nabla_{1}^{2j+2-2l}\hat\varphi(0)|}{\hat\varphi(0)}
\biggr)^{{(r-2l)/(2j+2-2l)}}\hat\varphi(0)
\\
&&\qquad\le O\biggl(\biggl(1-\frac{m}{m_{\rm c}}\biggr)^{-{(2j+2-2l)}/2}\biggr)^{
{(r-2l)/(2j+2-2l)}}
O\biggl(\biggl(1-\frac{m}{m_{\rm c}}\biggr)^{-1}\biggr)\nonumber
\\
&&\qquad =O\biggl(\biggl(1-\frac{m}{m_{\rm c}}\biggr)^{-1-{(r-2l)}/2}\biggr).\nonumber
\end{eqnarray}
Combining this with (\ref{eq:2<r<alpha-IJbds}) and (\ref
{eq:nablone2lJ}) yields
\begin{eqnarray}\label{eq:H2-bd}
\bigl|H_{2l}^{(2)}\bigr|\le
\cases{
O\biggl(\biggl(1-\dfrac{m}{m_{\rm c}}\biggr)^{-r/2}\biggr),&\quad$l=1$,
\cr
O\biggl(\biggl(1-\dfrac{m}{m_{\rm c}}\biggr)^{-{r}/2+\epsilon}\biggr),
&\quad$l=2,3,\dots,j$.
}
\end{eqnarray}
Finally, for $H_n^{(3)}$ with $n\ge2$ ($H_0^{(3)}$ and
$H_1^{(3)}$
will be investigated in detail later), we use
\begin{eqnarray}\label{eq:crudebd-odd}
\int_0^\infty\frac{\mathrm{d}u}{u^{1+q}} |\sin(ux_1) \sin(uy_1)|
&\le&\int_0^\infty\frac{\mathrm{d}u}{u^{1+q}} (|u^2x_1y_1|\wedge1)
\nonumber
\\[-8pt]\\[-8pt]
&=&O(|x_1y_1|^{q/2}),\nonumber
\\\label{eq:crudebd-even}
\int_0^\infty\frac{\mathrm{d}u}{u^{1+q}} \bigl|\cos(ux_1)
\bigl(1-\cos(uy_1)
\bigr)\bigr|
&\le&\int_0^\infty\frac{\mathrm{d}u}{u^{1+q}} \biggl(\frac{u^2y_1^2}2
\wedge2\biggr)\nonumber
\\[-8pt]\\[-8pt]
&=&O(|y_1|^q),\nonumber
\end{eqnarray}
which are due to the naive bounds $|\sin w|\le|w|\wedge1$, $|\cos
w|\le1$
and $|1-\cos w|\le\frac{w^2}2\wedge2$. By
(\ref{eq:crudebd-odd}) and (\ref{eq:crudebd-even}) and using Jensen's
inequality for
odd $n,$ as in (\ref{eq:Jensen}), we obtain
\begin{eqnarray*}
\bigl|H_n^{(3)}\bigr|
&\le&
\cases{
\biggl(\dfrac{|\nabla_{1}^{2j-n+1}\hat\varphi
(0)|}{\hat\varphi(0)}
\biggr)^{{(2j-n+{q}/2)/(2j-n+1)}}\hat\varphi(0)
\cr
\qquad {}\times \displaystyle\sum
_{s=0}^\infty
m^s\sum_{y\in{\mathbb Z}^d}|y_1|^{n+{q}/2}|J_s(y)|,&\quad\mbox{odd
}$n$,
\cr
\displaystyle|\nabla_{1}^{2j-n}\hat\varphi(0)|\sum_{s=0}^\infty
m^s\sum
_{y\in{\mathbb Z}^d}
|y_1|^{n+q}|J_s(y)|,&\quad\mbox{even }$n$.
}
\end{eqnarray*}
Then, by Proposition~\ref{proposition:2<r<alpha} and using $2j+q=r$,
we obtain
\begin{eqnarray}\label{eq:H3-bd}
\bigl|H_n^{(3)}\bigr|\le O\biggl(\biggl(1-\frac{m}{m_{\rm c}}\biggr)^{-
{r}/2+\epsilon}\biggr)\qquad
[n=2,3,\dots,2j].
\end{eqnarray}
Now, by (\ref{eq:pr42j+q:2}), (\ref{eq:H1-bd}), (\ref{eq:H2-bd}) and
(\ref{eq:H3-bd}), we arrive at
\begin{eqnarray}\label{eq:pr42j+q:3}
\quad \sum_{t=0}^\infty m^t\sum_{x\in{\mathbb Z}^d}|x_1|^r\varphi
_t(x)&=&\frac
{\hat\varphi(0)}
{\hat I(0)}\left(\pmatrix{2j\cr 2}H_2^{(2)}+H_0^{(3)}+\pmatrix
{2j\cr 1}H_1^{(3)}\right)\nonumber
\\[-8pt]\\[-8pt]
&&{}+O\biggl(\biggl(1-\frac{m}{m_{\rm c}}\biggr)^{-1-{r}/2+\epsilon}
\biggr).\nonumber
\end{eqnarray}
Finally, we reorganize the main term of (\ref{eq:pr42j+q:3}) and
complete the
proof of Theorem~\ref{theorem:main1}. First, we note that
\begin{eqnarray*}
\sin(ux_1) \sin(uy_1)&=&\frac{\cos(u(x_1-y_1))-\cos(u(x_1+y_1))}2
\\
&=&\frac{1-\cos(u(x_1+y_1))-(1-\cos(u(x_1-y_1))
)}2,\nonumber
\\
\cos(ux_1) \bigl(1-\cos(uy_1)\bigr)&=&\cos(ux_1)-\frac{\cos(u(x_1+y_1))
+\cos(u(x_1-y_1))}2
\\
&=&\frac{(1-\cos(u(x_1+y_1)))+(1-\cos(u(x_1-y_1))
)}2
\\
&&{}-\bigl(1-\cos(ux_1)\bigr).
\end{eqnarray*}
Then, by (\ref{eq:Kr}), we have the identities
\begin{eqnarray*}
\frac1{K_q}\int_0^\infty\frac{\mathrm{d}u}{u^{1+q}} \sin(ux_1)
\sin(uy_1)&=&
\frac{|x_1+y_1|^q-|x_1-y_1|^q}2,
\\
\frac1{K_q}\int_0^\infty\frac{\mathrm{d}u}{u^{1+q}} \cos(ux_1)
\bigl(1-\cos(u
y_1)\bigr)&=&\frac{|x_1+y_1|^q+|x_1-y_1|^q-2|x_1|^q}2.
\end{eqnarray*}
By these identities and the fact that $r=2j+q$, we obtain
\begin{eqnarray*}
\pmatrix{2j\cr 2}H_2^{(2)}+H_0^{(3)}+\pmatrix{2j\cr 1}H_1^{
(3)}=\sum_{s,t
=0}^\infty m^{t+s}\sum_{x,y\in{\mathbb Z}^d}x_1^{2j-2}\varphi_t(x)
J_s(y)
{\mathcal H}
(x_1,y_1),
\end{eqnarray*}
where
\begin{eqnarray*}
{\mathcal H}(x_1,y_1)&=&\pmatrix{2j\cr 2}|x_1|^qy_1^2+x_1^2\frac
{|x_1+y_1|^q+|x_1-y_1|^q
-2|x_1|^q}2
\\
&&{}+\pmatrix{2j\cr 1}x_1y_1\frac{|x_1+y_1|^q-|x_1-y_1|^q}2.
\end{eqnarray*}
In fact, due to the symmetry
${\mathcal H}(x_1,y_1)={\mathcal H}(x_1,-y_1)={\mathcal
H}(-x_1,y_1)={\mathcal H}(-x_1,\break -y_1)$ for any
$x_1,y_1\in{\mathbb Z}$, the above identity is equivalent to
\begin{eqnarray*}
&&\pmatrix{2j\cr 2}H_2^{(2)}+H_0^{(3)}+\pmatrix{2j\cr 1}H_1^{(3)}
\\
&&\qquad =4\sum_{s,t=0}^\infty m^{t+s}\mathop{\sum_{x,y\in{\mathbb
Z}^d}}_{
(x_1,y_1>0)}
x_1^{2j-2}\varphi_t(x) J_s(y) {\mathcal H}(x_1,y_1).
\end{eqnarray*}
Using the Taylor expansion of
$|x_1\pm y_1|^q\equiv x_1^q(1\pm\frac{y_1}{x_1})^q$ if $x_1>y_1>0$ and
the expansion of $|x_1\pm y_1|^q\equiv y_1^q(1\pm\frac{x_1}{y_1})^q$ if
$y_1>x_1>0$, we have
\begin{eqnarray}\label{eq:cH-identity}
&&{\mathcal H}(x_1,y_1)\nonumber
\\[-9pt]\\[-9pt]
&&\qquad=
\cases{
\displaystyle\left(\pmatrix{2j\cr 2}+\pmatrix{q\cr 2}+\pmatrix{2j\cr 1}q\right)x_1^qy_1^2
+O(x_1^{q-1}y_1^3),\cr \qquad x_1>y_1>0,
\cr
O(y_1^{2+q}),\qquad y_1\ge x_1>0.
}
\nonumber
\end{eqnarray}
Notice that
\begin{eqnarray*}
\pmatrix{2j\cr 2}+\pmatrix{q\cr 2}+\pmatrix{2j\cr 1}q&=&j(2j-1)+\frac{q}2(q-1)+2jq
\\
&=&\biggl(j+\frac{q}2\biggr)(2j+q)-j-\frac{q}2=\frac{r}2r-\frac{r}2
=\pmatrix{r\cr2}.
\end{eqnarray*}
We also notice that, as long as $q\in(0,1]$,
we have
\begin{eqnarray*}
x_1^{q-1}y_1^3=\biggl(\frac{y_1}{x_1}\biggr)^{1-q}y_1^{2+q}\le y_1^{2+q}
\qquad[x_1>y_1>0].
\end{eqnarray*}
Therefore, by Proposition~\ref{proposition:2<r<alpha}, we obtain that,
for $q\equiv r-2j\in(0,1]$,
\begin{eqnarray}\label{eq:pr42j+q:4.5}
&&\pmatrix{2j\cr 2}H_2^{(2)}+H_0^{(3)}+\pmatrix{2j\cr 1}H_1^{(3)}\nonumber
\\[-2pt]
&&\qquad=4\pmatrix{r\cr 2}\sum_{s,t=0}^\infty m^{t+s}\mathop{\sum_{x,y\in
{\mathbb Z}^d
}}_{ (x_1,y_1>
0)}x_1^{2j+q-2}\varphi_t(x) y_1^2J_s(y)\nonumber
\\[-2pt]
&&{}\qquad\quad+\sum_{s,t=0}^\infty m^{t+s}\sum_{x,y\in{\mathbb Z}^d
}x_1^{2j-2}\varphi_t(x)
O(|y_1|^{2+q})J_s(y)
\\[-2pt]
&&\qquad=\pmatrix{r\cr 2}(-\nabla_{1}^2\hat J(0))\sum_{t=0}^\infty
m^t\sum_{x\in{\mathbb Z}^d}
|x_1|^{r-2}\varphi_t(x)\nonumber
\\[-2pt]
&&\qquad{}\quad+\underbrace{|\nabla_{1}^{2j-2}\hat\varphi(0)|\sum
_{s=0}^\infty m^s\sum_{y
\in{\mathbb Z}^d}O(|y_1|^{2+q})J_s(y)}_{O((1-{m/m_{\rm
c}})^{-
{r}/2+\epsilon})}.\nonumber
\end{eqnarray}
For $q\in(1,2)$, we have to deal with the contribution from
$O(x_1^{q-1}y_1^3)$ in (\ref{eq:cH-identity}). However, by Jensen's inequality
and Proposition~\ref{proposition:2<r<alpha},
we have
\begin{eqnarray*}
&&\sum_{s,t=0}^\infty m^{t+s}\mathop{\sum_{x,y\in{\mathbb Z}^d
}}_{(x_1>y_1>0)} x_1^{2j
+q-3}\varphi_t(x) y_1^3|J_s(y)|
\\[-2pt]
&&\qquad \le\biggl(\frac{|\nabla_{1}^{2j-1}\hat\varphi(0)|}{\hat\varphi
(0)}\biggr)^{{(2
j+q-3)/(2j-1)}}\hat\varphi(0)\sum_{s=0}^\infty m^s\sum_{y\in{\mathbb Z}^d
}|y_1|^3|J_s(y)|
\\[-2pt]
&&\qquad \le O\biggl(\biggl(1-\frac{m}{m_{\rm c}}\biggr)^{-{r}/2+\epsilon}\biggr)
\end{eqnarray*}
and thus (\ref{eq:pr42j+q:4.5}) is valid for any $q\in(0,2)$.
Now, by substituting (\ref{eq:pr42j+q:4.5}) back into (\ref
{eq:pr42j+q:3}), we
obtain the recursion
\begin{eqnarray*}
\sum_{t=0}^\infty m^t\sum_{x\in{\mathbb Z}^d}|x_1|^r\varphi
_t(x)&=&\pmatrix{r\cr 2}
\frac{-\nabla_{1}^2\hat J(0)}{\hat I(0)} \hat\varphi(0)\sum
_{t=0}^\infty
m^t\sum_{x\in{\mathbb Z}^d}|x_1|^{r-2}\varphi_t(x)
\\[-2pt]
&&{}+O\biggl(\biggl(1-\frac{m}{m_{\rm c}}\biggr)^{-1-{r}/2+\epsilon}\biggr).
\end{eqnarray*}
Repeatedly using this recursion $j$ times and recalling $r-2j=q$, we
obtain
\begin{eqnarray*}
\sum_{t=0}^\infty m^t\sum_{x\in{\mathbb Z}^d}|x_1|^r\varphi
_t(x)&=&\prod_{i=0}^{j-1}
\pmatrix{r-2i\cr 2}\biggl(\frac{-\nabla_{1}^2\hat J(0)}{\hat I(0)} \hat
\varphi(0) \biggr)^j
\\[-2pt]
&&{}\times\sum_{t=0}^\infty m^t\sum_{x\in{\mathbb Z}^d
}|x_1|^{r-2j}\varphi_t(x)
+O\biggl(\biggl(1-\frac{m}{m_{\rm c}}\biggr)^{-1-{r}/2+\epsilon}\biggr)
\\[-2pt]
&=&\frac{\Gamma(r+1)}{2^j\Gamma(r-2j+1)}\biggl(\frac{-\nabla_{1}
^2\hat J(0)}
{\hat I(0)} \hat\varphi(0)\biggr)^j
\\[-2pt]
&&{}\times\sum_{t=0}^\infty m^t\sum_{x\in{\mathbb
Z}^d}|x_1|^q\varphi_t(x)
+O\biggl(\biggl(1-\frac{m}{m_{\rm c}}\biggr)^{-1-{r}/2+\epsilon}\biggr).
\end{eqnarray*}
Notice that, by (\ref{eq:chi-asy}) and (\ref{eq:ratio}),
\begin{eqnarray*}
\frac{-\nabla_{1}^2\hat J(0)}{\hat I(0)} \hat\varphi(0)=\frac
{2C_{\mathrm{II}}v_\alpha}
{1-{m/m_{\rm c}}}+O\biggl(\biggl(1-\frac{m}{m_{\rm c}}\biggr)^{-1+\epsilon
}\biggr)
\end{eqnarray*}
and that, by (\ref{eq:pr4r<2}) for $\alpha>2$ and (\ref{eq:Kr-evaluation}),
\begin{eqnarray*}
\sum_{t=0}^\infty m^t\sum_{x\in{\mathbb Z}^d}|x_1|^q\varphi
_t(x)=\Gamma
(q+1)
\frac{C_{\mathrm I}(C_{\mathrm{II}}v_\alpha)^{{q}/2}}{(1-
{m/m_{\rm
c}})^{1+{q}/2}}
+O\biggl(\biggl(1-\frac{m}{m_{\rm c}}\biggr)^{-1-{q}/2+\epsilon}\biggr).
\end{eqnarray*}
Therefore, we arrive at
\begin{eqnarray*}
&&\sum_{t=0}^\infty m^t\sum_{x\in{\mathbb Z}^d}|x_1|^r\varphi
_t(x)
\\
&&\qquad =\frac
{\Gamma(r+1)}{2^j
\Gamma(q+1)}\biggl(\frac{2C_{\mathrm{II}}v_\alpha}{1-{m/m_{\rm
c}}}\biggr)^j\Gamma(q+1)
\frac{C_{\mathrm I}(C_{\mathrm{II}}v_\alpha)^{{q}/2}}{(1-
{m/m_{\rm
c}})^{1+{q}/2}}
\\
&&{}\qquad\quad +O\biggl(\biggl(1-\frac{m}{m_{\rm c}}\biggr)^{-1-{r}/2+\epsilon}\biggr)
\\
&&\qquad =\Gamma(r+1) \frac{C_{\mathrm I}(C_{\mathrm{II}}v_\alpha)^{
{r}/2}}{(1-
{m/ m_{\rm c}})^{1+
{r}/2}}+O\biggl(\biggl(1-\frac{m}{m_{\rm c}}\biggr)^{-1-{r}/2+\epsilon
}\biggr).
\end{eqnarray*}
This completes the proof of Theorem~\ref{theorem:main1}.
\subsection{\texorpdfstring{Proof of
Theorem~\protect\ref{theorem:main2}}{Proof of Theorem~1.2}}
It is very easy to identify the main term for $\alpha\ne2$. First, by the
binomial expansion of the main term in (\ref{eq:main1}),
\begin{eqnarray}\label{eq:binom-exp}
\quad&&\biggl(1-\frac{m}{m_{\rm c}}\biggr)^{-1-{r/(\alpha\wedge
2)}}\nonumber
\\
&&\qquad =\sum_{t=0}^\infty
\frac{(-{r/(\alpha\wedge2)}-1)(-{r/(\alpha\wedge
2)}-2)\cdots
(-{r/(\alpha\wedge2)}-t)}{t!}\nonumber
\\
&&{}\hspace*{13pt}\qquad\quad\times \biggl(-\frac{m}{m_{\rm c}}
\biggr)^t
\\
&&\qquad =\sum_{t=0}^\infty\frac{\Gamma({r/(\alpha\wedge
2)}+t+1)}{t! \Gamma
({r/(\alpha\wedge2)}+1)}\biggl(\frac{m}{m_{\rm c}}
\biggr)^t\nonumber
\\
&&\qquad =\frac1{\Gamma({r/(\alpha\wedge2)}+1)}\sum_{t=0}^\infty
\biggl(\frac{m}
{m_{\rm c}}\biggr)^t\frac1{t!}\int_0^\infty x^{t+{r/(\alpha
\wedge2)}} e^{-x}\,
\mathrm{d}x.\nonumber
\end{eqnarray}
Then, by the steepest descent method, we obtain that, for every $\beta
\in{\mathbb R}$,
\begin{eqnarray*}
\int_0^\infty x^{t+\beta} e^{-x} \,\mathrm{d}x\sim\sqrt{2\pi
(t+\beta)}\biggl(
\frac{t+\beta}{e}\biggr)^{t+\beta}\qquad\mbox{as }t\to\infty.
\end{eqnarray*}
Using this for $\beta=0,\frac{r}{\alpha\wedge2}$, we conclude that, as
$t\to\infty$,
\begin{eqnarray*}
\frac1{t!}\int_0^\infty x^{t+{r/(\alpha\wedge2)}} e^{-x}
\,\mathrm{d}x
&\sim&\biggl(\frac{t+{r/(\alpha\wedge2)}}t\biggr)^{t+
1/2}\biggl(\frac{t
+{r/(\alpha\wedge2)}}{e}\biggr)^{{r/(\alpha\wedge2)}}
\\[-2pt]
&\sim&
t^{{r/(\alpha\wedge2)}},
\end{eqnarray*}
which implies that the large-$t$ asymptotic expression for the
coefficient of
$m^t$ in~(\ref{eq:binom-exp}) is
$m_{\rm c}^{-t} t^{{r/(\alpha\wedge2)}}/\Gamma(\frac
{r}{\alpha\wedge2}+1)$,
hence the expression for the constant in (\ref{eq:main2}) for $\alpha
\ne2$.
There are many other ways to derive the above asymptotic expression.
One of
them is to notice that $x^te^{-x}/t!$ in (\ref{eq:binom-exp}) is the
probability
density for the sum of independent mean-one exponential random
variables. Then,
we use Jensen's inequality and apply the law of large numbers if
$\frac{r}{\alpha\wedge2}\le1,$ or exactly compute integer-power
moments for the
exponential random variables if $\frac{r}{\alpha\wedge2}>1$. We omit the
details.
To identify the main term for $\alpha=2$ in (\ref{eq:main2}), as well
as to
obtain the error estimates for all $\alpha>0$, we simply use
\cite{fo90}, Theorems~3A and 4. For convenience, we summarize a slightly
simplified version of these results as follows.
\begin{theorem}[(\cite{fo90}, Theorems~3A and 4)]\label{theorem:fo90}
\textup{(i)} Let
\begin{eqnarray*}
f(z)=(1-z)^{-1-\beta}\biggl(\log\frac1{1-z}\biggr)^\gamma,
\end{eqnarray*}
where $\beta\notin-\mathbb{N}\equiv{\mathbb Z}\setminus{\mathbb
Z}_+$ and $\gamma\notin
{\mathbb Z}_+$ are real or
complex numbers. Then, the coefficient $f_t$ of $f(z)=\sum_tf_tz^t$ satisfies
\begin{eqnarray*}
f_t\sim\frac{t^\beta(\log t)^\gamma}{\Gamma(1+\beta)}\qquad\mbox
{as }
t\to\infty.
\end{eqnarray*}\vspace*{-12pt}
\begin{longlist}[(ii)]
\item[(ii)]
Let $f(z)$ be analytic in $|z|<1$ and
\begin{eqnarray*}
f(z)=O(|1-z|^{-1-\beta})\qquad\mbox{as }z\to1
\end{eqnarray*}
for some real number $\beta>0$. Then, the coefficient $f_t$ of
$f(z)=\sum_tf_tz^t$ satisfies
\begin{eqnarray*}
f_t=O(t^\beta)\qquad\mbox{as }t\to\infty.
\end{eqnarray*}
\end{longlist}
\end{theorem}
The main term for $\alpha=2$ in (\ref{eq:main2}) is obtained by setting
$\beta=\gamma=\frac{r}2$ in Theorem~\ref{theorem:fo90}(i). For the error
estimates, we use Theorem~\ref{theorem:fo90}(ii) with $\beta=\frac
{r}2$ for
$\alpha=2$ and $\beta=\frac{r}{\alpha\wedge2}-\epsilon>0$ for
$\alpha\ne2$.
This completes the proof of Theorem~\ref{theorem:main2}.
\section{Proof of the key propositions}\label{s:key-prop}
In this section, we prove
Propositions~\ref{proposition:r<alpha&2} and \ref{proposition:2<r<alpha},
these being key propositions used in the previous section to prove the main
theorem. In Section~\ref{ss:prop2}, we first prove
Proposition~\ref{proposition:2<r<alpha}. Then, in Section~\ref{ss:prop1},
we use (\ref{eq:2<r<alpha-varphibd}) in Proposition~\ref
{proposition:2<r<alpha}
to show Proposition~\ref{proposition:r<alpha&2} for $\alpha>2$.
\subsection{\texorpdfstring{Proof of
Proposition~\protect\ref{proposition:2<r<alpha}}{Proof of Proposition~3.2}}
\label{ss:prop2}
Below, we prove Proposition~\ref{proposition:2<r<alpha} by using the results
already obtained in \cite{csI,csII,hhh,hhs08} and alternately applying
the following two
lemmas.
\begin{lemma}\label{lmm:induction1}
Let $\alpha>2$, $l\in\{1,2,\dots,\langle\alpha\rangle\}$ and
suppose that
(\ref{eq:2<r<alpha-IJbds}) holds for any $\nu\in\{0,1,\dots,l\vee
2\}$ and
(\ref{eq:2<r<alpha-varphibd}) holds for any $n\in\{0,\dots,l-1\}$.
Then, (\ref{eq:2<r<alpha-varphibd}) holds for $n=l$.
\end{lemma}
\begin{lemma}\label{lmm:induction2}
Let $\alpha>2$ and suppose that (\ref{eq:2<r<alpha-varphibd}) holds
for $n=2l$,
where $l\in\{1,\dots,\langle\frac\alpha2\rangle\}$ (note that
$\alpha-2\le2\langle\frac\alpha2\rangle<\alpha$). Then,
(\ref{eq:2<r<alpha-IJbds}) holds for any $\nu\in(n,n+2]$ if
$n+2<\alpha$,
or for any $\nu\in(n,\alpha)$ if $\alpha\le n+2$.
\end{lemma}
We will prove these lemmas after completing the proof of
Proposition~\ref{proposition:2<r<alpha}. For random walk,
(\ref{eq:2<r<alpha-IJbds}) always holds as mentioned earlier and we
therefore only need Lemma~\ref{lmm:induction1}.
We now begin by proving Proposition~\ref{proposition:2<r<alpha}.
First, we note that (\ref{eq:2<r<alpha-IJbds}) for $\nu\in[0,2]$ and
(\ref{eq:2<r<alpha-varphibd}) for $n=0$ have been proven in the current
setting \mbox{\cite{csI,csII,hhh,hhs08}}; the result in \cite{hhs08} for
self-avoiding walk is only valid at $\theta=0$. However, it is not
hard to extend
the result to nonzero $\theta$ by splitting the denominator in
(\ref{eq:lace-exp-solution}) into $1-\hat J(k,m)$ and
$\hat J(k,m)-\hat J(k,me^{i\theta}),$ and estimating the latter as
$m^t-(me^{i\theta})^t=m^t(1-e^{i\theta})\sum_{s=0}^{t-1}e^{i\theta s}$
[which equals $O(\theta)tm^t$ for $|\theta|\ll1$]. We omit the
details. Then, by
Lemma~\ref{lmm:induction1} with $l=1$, we obtain (\ref{eq:2<r<alpha-varphibd})
for $n=1$. With this conclusion and again using Lemma~\ref{lmm:induction1},
but now with $l=2$, we obtain (\ref{eq:2<r<alpha-varphibd}) for $n=2$.
With this
conclusion and using Lemma~\ref{lmm:induction2}, we further obtain
(\ref{eq:2<r<alpha-IJbds}) for $\nu\in(2,4]$ or $\nu\in(2,\alpha)$,
depending on
whether $\alpha>4$ or $\alpha\le4$. We can repeat this, using
Lemmas~\ref{lmm:induction1} and \ref{lmm:induction2} alternately,
until $n$
reaches $\langle\alpha\rangle$. Let $\tilde l=\langle\frac\alpha
2\rangle$.
We see that
\begin{eqnarray*}
&&
\left.\begin{array}{l}
(\ref{eq:2<r<alpha-IJbds})_{\nu\in[0,2]}\\[5pt]
(\ref{eq:2<r<alpha-varphibd})_{n=0}
\end{array}
\right\}
\stackrel{\mathrm{Lemma~\scriptsize{\ref{lmm:induction1}}}}\Longrightarrow
(\ref{eq:2<r<alpha-varphibd})_{n=1,2}
\stackrel{\mathrm{Lemma~\scriptsize{\ref{lmm:induction2}}}}\Longrightarrow
(\ref{eq:2<r<alpha-IJbds})_{\nu\in(2,4]}
\stackrel{\mathrm{Lemma~\scriptsize{\ref{lmm:induction1}}}}\Longrightarrow
\cdots
\\
&&\hspace*{67pt}\stackrel{\mathrm{Lemma~\scriptsize{\ref
{lmm:induction1}}}}\Longrightarrow
(\ref{eq:2<r<alpha-varphibd})_{n=2\tilde l-1,2\tilde l}
\\
&&\hspace*{67pt}\stackrel{\mathrm{Lemma~\scriptsize{\ref{lmm:induction2}}}}\Longrightarrow
(\ref{eq:2<r<alpha-IJbds})_{\nu\in(2\tilde l,\alpha)}
\bigl(\mathop{\Longrightarrow}\limits^{\mathrm{Lemma~\scriptsize{\ref{lmm:induction1}}}}_{\mathrm{if }\ \alpha>2\tilde l+1}
(\ref{eq:2<r<alpha-varphibd})_{n=2\tilde l+1}\bigr).
\end{eqnarray*}
This completes the proof of Proposition~\ref{proposition:2<r<alpha}.
\begin{pf*}{Proof of Lemma~\ref{lmm:induction1}}
First, by using (\ref{eq:2<r<alpha-IJbds}) for $\nu=2$ and
(\ref{eq:2<r<alpha-varphibd}) for $n=0$,
we prove $|\nabla_{1}\hat I(k)|\le O(|\hat\varphi(k)|^{-1/2})$; the
proof of
$|\nabla_{1}\hat J(k)|\le O(|\hat\varphi(k)|^{-1/2})$ is almost
identical and
thus we omit it. By the ${\mathbb Z}^d$-symmetry of the models
and using\vadjust{\goodbreak} $|\sin(k_1x_1)|\le|k_1x_1|$ and (\ref{eq:2<r<alpha-IJbds})
for $\nu=2$,
we obtain
\begin{eqnarray*}
|\nabla_{1}\hat I(k)|&=&\Biggl|\sum_{t=0}^\infty m^t\sum_{x\in
{\mathbb Z}^d
}x_1\sin(k_1x_1)
I_t(x) e^{i(k_2x_2+\cdots+k_dx_d)}\Biggr|
\\[-2pt]
&\le&|k_1|\sum_{t=0}^\infty m^t\sum_{x\in{\mathbb Z}^d}x_1^2
|I_t(x)|\le O(|k_1|).
\end{eqnarray*}
However, by (\ref{eq:2<r<alpha-varphibd}) for $n=0$, we have
$|\hat\varphi(k)|\le O(|k|^{-2})$, which implies that
$|k_1|\le|k|\le O(|\hat\varphi(k)|^{-1/2})$, as required.
We now use this bound to complete the proof of Lemma~\ref{lmm:induction1}.
First, by
differentiating (\ref{eq:lace-exp-fourier}) and solving the resulting equation
for $\nabla_{1}^l\hat\varphi(k)$, we have that, for $l\in\mathbb{N}$,\vspace*{-2pt}
\begin{eqnarray*}
\nabla_{1}^l\hat\varphi(k)&=&\nabla_{1}^l\hat I(k)+\sum
_{j=0}^l\pmatrix
{l\cr j}\nabla_{1}^j
\hat J(k) \nabla_{1}^{l-j}\hat\varphi(k)
\\[-2pt]
&=&\frac1{1-\hat J(k)}\Biggl(\nabla_{1}^l\hat I(k)+\sum_{j=1}^l\pmatrix
{l\cr j}\nabla_{1}^j
\hat J(k) \nabla_{1}^{l-j}\hat\varphi(k)\Biggr).\vspace*{-2pt}
\end{eqnarray*}
By (\ref{eq:lace-exp-solution}), (\ref{eq:Ihat}) and (\ref
{eq:property1}), we have
$|1-\hat J(k)|^{-1}=O(|\hat\varphi(k)|)$. By (\ref
{eq:2<r<alpha-IJbds}) for
$\nu\ge2$ or using $|\nabla_{1}\hat I(k)|\le O(|\hat\varphi(k)|^{-1/2})$,
we obtain\vspace*{-2pt}
\begin{eqnarray*}
\biggl|\frac{\nabla_{1}^l\hat I(k)}{1-\hat J(k)}\biggr|\le O
(|\hat
\varphi(k)|
)\times
\cases{
|\hat\varphi(k)|^{-1/2},&\quad$l=1$,
\cr
\biggl(1-\dfrac{m}{m_{\rm c}}\biggr)^{1-{l/2}+\epsilon},&\quad$l=2,\dots
,\langle\alpha
\rangle$,
}\vspace*{-2pt}
\end{eqnarray*}
which, by (\ref{eq:2<r<alpha-varphibd}) for $n=0$, is smaller than the
bound in
(\ref{eq:2<r<alpha-varphibd}) for $n=l$, yielding an error term. For $j=1,2$,
we also use (\ref{eq:2<r<alpha-varphibd}) for $n\le l-1$ to obtain\vspace*{-2pt}
\begin{eqnarray*}
&&\biggl|\frac{\nabla_{1}^j\hat J(k)}{1-\hat J(k)} \nabla
_{1}^{l-j}\hat
\varphi(k)
\biggr|
\\[-2pt]
&&\qquad \le O(|\hat\varphi(k)|^{j/2})
\\[-2pt]
&&{}\qquad\quad\times
\cases{
\biggl(1-\dfrac{m}{m_{\rm c}}+|\theta|+|k|^2\biggr)^{-1-{(l-j)}/2},
&\quad$l=j,j+1$,
\cr
\displaystyle\dfrac{(1-{m/m_{\rm c}})^{1-
{(l-j)}/2}}{(1-
{m/m_{\rm c}}+|\theta|
+|k|^2)^2},&\quad$l=j+2,\dots,\langle\alpha\rangle$,
}\vspace*{-2pt}
\end{eqnarray*}
which, again by (\ref{eq:2<r<alpha-varphibd}) for $n=0$, obeys the
required bound
in (\ref{eq:2<r<alpha-varphibd}) for $n=l$. Finally, for $j\ge3$ (hence
for $l\ge3$),\vspace*{-2pt}
\begin{eqnarray*}
&&\biggl|\frac{\nabla_{1}^j\hat J(k)}{1-\hat J(k)} \nabla
_{1}^{l-j}\hat
\varphi(k)
\biggr|
\\[-2pt]
&&\qquad\le\frac{O((1-{m/m_{\rm c}})^{1-{j}/2+\epsilon
})}{1-{m/m_{\rm c}}
+|\theta|+|k|^2}
\\
&&\qquad\quad{}\times
\cases{
\biggl(1-\dfrac{m}{m_{\rm
c}}+|\theta|+|k|^2\biggr)^{-1-{(l-j)}/2}&\quad$[l=j,j+1]$,
\cr
\displaystyle\dfrac{(1-{m/m_{\rm c}})^{1-
{(l-j)}/2}}{(1-
{m/m_{\rm c}}+|\theta|
+|k|^2)^2}&\quad$[l=j+2,\dots,\langle\alpha\rangle]$
}
\\
&&\qquad\le\frac{O((1-{m/m_{\rm c}})^{1-{l}/2+\epsilon
})}{(1-{m/m_{\rm c}}
+|\theta|+|k|^2)^2},
\end{eqnarray*}
which is smaller [by the factor $(1-\frac{m}{m_{\rm c}})^\epsilon$]
than the bound
in (\ref{eq:2<r<alpha-varphibd}), yielding\vspace*{1pt} an error term. This
completes the
proof of Lemma~\ref{lmm:induction1}.
\end{pf*}
\begin{pf*}{Proof of Lemma~\ref{lmm:induction2}}
First, we recall (\ref{eq:I-def}) and (\ref{eq:J-def}). Since\break
$\sum_x|x_1|^\nu D(x)<\infty$ provided that $\nu<\alpha$, (\ref
{eq:2<r<alpha-IJbds})
always holds for random walk. Moreover, for oriented percolation, there
is a
constant $C_\nu<\infty$ such that
\begin{eqnarray*}
\sum_{t=0}^\infty m^t\sum_{x\in{\mathbb Z}^d}|x_1|^\nu
|J_t^\mathrm{OP}(x)|&\le&
p\sum_{t=1}^\infty
m^t\sum_{x,y\in{\mathbb Z}^d}|y_1+x_1-y_1|^\nu|\pi
_{t-1}^\mathrm{OP}(y)| D(x-y)
\\
&\le& C_\nu pm\sum_{t=1}^\infty m^{t-1}\sum_{y\in{\mathbb
Z}^d}(|y_1|^\nu
+1)|\pi_{t-1}^\mathrm{OP}
(y)|,
\end{eqnarray*}
where we have used the fact that, for any $a_1,\dots,a_n\in{\mathbb R}$,
\begin{eqnarray}\label{eq:naive}
\Biggl|\sum_{j=1}^na_j\Biggr|^\nu\le\Bigl(n\max_{1\le j\le
n}|a_j|\Bigr)^\nu
=n^\nu\max_{1\le j\le n}|a_j|^\nu\le n^\nu\sum_{j=1}^n|a_j|^\nu.
\end{eqnarray}
Since $\sum_{s=0}^\infty m^s\sum_{y\in{\mathbb Z}^d}|\pi
_s^\mathrm{OP}(y)|=O(1)$
uniformly in
$m\le m_{\rm c}$ \cite{csI}, it suffices to show that, for
self-avoiding walk and oriented percolation, (\ref
{eq:2<r<alpha-varphibd}) for
$n=2l$, where $l\in\{1,\dots,\langle\frac\alpha2\rangle\}$ implies
that
\begin{eqnarray}\label{eq:2<r<alpha-pibd}
\sum_{t=0}^\infty m^t\sum_{x\in{\mathbb Z}^d}|x_1|^{2l+q}|\pi
_t(x)|\le
O\biggl(\biggl(1-\frac{m}{m_{\rm c}}\biggr)^{1-{(2l+q)}/2+\epsilon}\biggr)
\end{eqnarray}
for any $q\in(0,2]$ if $2l+2<\alpha$, or for any $q\in(0,\alpha
-2l)$ if
$\alpha\le2l+2$.
As we mentioned earlier, $\pi_t(x)$ is an alternating sum of the
lace expansion coefficients. More precisely,
\begin{eqnarray*}
\pi_t(x)=\sum_{N=0}^\infty(-1)^N\pi_t^{(N)}(x),
\end{eqnarray*}
where $\pi_t^{(N)}(x)\ge0$ is the model-dependent $N$th
expansion coefficient (see, e.g., \cite{csI,s06} for the precise
definitions of the expansion coefficients). Due to the subadditivity argument
for self-avoiding walk and by the BK inequality \cite{bk}
for percolation,
it is known that the expansion coefficients satisfy the following diagrammatic
bounds, in which each line corresponds to a 2-point function. For
self-avoiding walk,
\begin{eqnarray}\label{eq:saw-diagbds}
\pi_t^{(0)}(x)&\equiv&0 ,\qquad
\pi_t^{(1)}(x)\le {}\mbox{
\includegraphics{557i01.eps}
},\nonumber
\\[-8pt]\\[-8pt]
\pi_t^{(2)}(x)&\le& {}\mbox{
\includegraphics{557i02.eps}
},\qquad
\pi_t^{(3)}(x)\le {}\mbox{
\includegraphics{557i03.eps}
},\nonumber
\end{eqnarray}
where the bounding diagram for $\pi_t^{(1)}(x)$ is the $t$-step
self-avoiding loop at $x=o$, hence proportional to $\delta_{x,o}$, and the
diagram for $\pi_t^{(2)}(x)$ is the product of three 2-point functions
$\varphi_s^\mathrm{SAW}(x) \varphi_{s'}^\mathrm
{SAW}(x) \varphi_{s''}^\mathrm{SAW}(x)$
summed over
all possible combinations of $s,s',s''\in\mathbb{N}$ satisfying $s+s'+s''=t$,
and so
on. The unlabeled vertices in the diagrams for $\pi_t^{(3)}(x)$ and
the higher order expansion coefficients are summed over ${\mathbb
Z}^d$. For oriented
percolation,
\begin{eqnarray}\label{eq:op-diagbds}
\pi_t^{(0)}(x)\le{}\mbox{
\includegraphics{557i04.eps}
},\qquad
\pi_t^{(1)}(x)\le{}\mbox{
\includegraphics{557i05.eps}
},\qquad
\pi_t^{(2)}(x)\le{}\mbox{
\includegraphics{557i06.eps}
} +
{}\mbox{
\includegraphics{557i07.eps}
},
\end{eqnarray}
where the bounding diagram for $\pi_t^{(0)}(x)$ is $\varphi
_t^\mathrm{OP}(x)^2$
and that for $\pi_t^{(1)}(x)$ is the product of five 2-point functions
concatenated in the depicted way, and so on. The upward direction of the
diagrams is the time-increasing direction and the unlabeled vertices are
summed over space--time ${\mathbb Z}^d\times{\mathbb Z}_+$. For more
details, we refer to
\cite{s07}.
First, we prove (\ref{eq:2<r<alpha-pibd}) for self-avoiding walk. Since
$\pi_t^{(0)}(x)\equiv0$ and $\pi_t^{(1)}(x)\propto\delta_{x,o}$,
it suffices to investigate the contributions from $\pi_t^{
(N)}(x)$ for
$N\ge2$. For $\pi_t^{(2)}(x)$, since
\begin{eqnarray}\label{eq:pi2sawbd}
\pi_t^{(2)}(x)\le\mathop{\sum_{s,s',s''\in\mathbb{N}}}_{(s+s'+s''=t)}
\varphi_s^\mathrm{SAW}(x) \varphi_{s'}^\mathrm
{SAW}(x) \varphi_{s''}^\mathrm{SAW}(x),
\end{eqnarray}
we obtain
\begin{eqnarray*}
\sum_{t=0}^\infty m^t\sum_{x\in{\mathbb Z}^d}|x_1|^{2l+q} \pi_t^{
(2)}(x)&\le&
\biggl(\sum_{x\in{\mathbb Z}^d}|x_1|^q\sum_{s,s'\in\mathbb
{N}}m^{s+s'}\varphi
_s^\mathrm{SAW}(x)
\varphi_{s'}^\mathrm{SAW}(x)\biggr)
\\
&&{}\times\biggl(\sup_{x\in{\mathbb Z}^d}|x_1|^{2l}\sum_{s''\in
\mathbb{N}}m^{s''}
\varphi_{s''}^\mathrm{SAW}(x)\biggr)
\\
&\le& B^{(q)}W^{(2l)},
\end{eqnarray*}
where
\begin{eqnarray*}
B^{(\nu)}&=&\sup_{y\in{\mathbb Z}^d}\sum_{x\in{\mathbb
Z}^d}|x_1|^\nu\sum
_{t\in\mathbb{N}}m^t
\varphi_t^\mathrm{SAW}(x)\sum_{s=0}^\infty m^s\varphi
_s^\mathrm{SAW}(y-x),
\\
W^{(\nu)}&=&\sup_{x\in{\mathbb Z}^d}|x_1|^\nu\sum_{t\in\mathbb
{N}}m^t\varphi
_t^\mathrm{SAW}(x).
\end{eqnarray*}
Similarly to the above and the derivation of \cite{hhh}, formula (2.42), by using
(\ref{eq:naive}) and diagrammatic bounds of the form (\ref{eq:saw-diagbds}),
we can show that
\begin{eqnarray}\label{eq:piNsawbd}
&& \sum_{t=0}^\infty m^t\sum_{x\in{\mathbb Z}^d}|x_1|^{2l+q} \pi_t^{
(N)}(x)\nonumber
\\[-8pt]\\[-8pt]
&&\qquad\le
N^{2l+q+2}\bigl(B^{(0)}\bigr)^{N-2}B^{(q)}W^{(2l)}\qquad[N\ge2].\nonumber
\end{eqnarray}
It is immediate from the definition (\ref{eq:2pt-def}) that
$\varphi_t^\mathrm{SAW}(x)\le\delta_{x,o}\delta
_{t,0}+(D*\varphi
_{t-1}^\mathrm{SAW})(x)$.
By this, we have
\begin{eqnarray}\label{eq:B0sawbd}
B^{(0)}&\le&\sup_{y\in{\mathbb Z}^d}\sum_{x\in{\mathbb Z}^d}\sum
_{t\in\mathbb{N}
}m^t\varphi_t^\mathrm{SAW}(x)
\biggl(\delta_{x,y}+\sum_{s\in\mathbb{N}}m^s(D*\varphi
_{s-1}^\mathrm{SAW}
)(y-x)\biggr)\nonumber
\\
&\le& W^{(0)}+\sup_{y\in{\mathbb Z}^d}\sum_{x\in{\mathbb Z}^d}\sum
_{t\in\mathbb{N}
}m^t(D*\varphi_{t-
1}^\mathrm{SAW})(x)\sum_{s\in\mathbb{N}}m^s(D*\varphi
_{s-1}^\mathrm{SAW})(y-x)
\\
&\le& W^{(0)}+m^2\int_{[-\pi,\pi]^d}\frac{\mathrm{d}^dk}{(2\pi
)^d} \hat
D(k)^2|\hat\varphi^\mathrm{SAW}(k,m)|^2\nonumber
\end{eqnarray}
and
\begin{eqnarray}
W^{(0)}&\le&\sup_{x\in{\mathbb Z}^d}\sum_{t\in\mathbb
{N}}m^t(D*\varphi
_{t-1}^\mathrm{SAW})(x)\nonumber
\\[-2pt]
&\le& m\|D\|_\infty+\sup_{x\in{\mathbb Z}^d}\sum_{t=2}^\infty
m^t(D*D*\varphi
_{t-2}^\mathrm{SAW})
(x)
\\[-2pt]
&\le& m\|D\|_\infty+m^2\int_{[-\pi,\pi]^d}\frac{\mathrm{d}^dk}{(2\pi
)^d} \hat
D(k)^2|\hat\varphi^\mathrm{SAW}(k,m)|.\nonumber
\end{eqnarray}
By (\ref{eq:2<r<alpha-varphibd}) for $n=0$ and $\|D\|_\infty
=O(L^{-d})$, we
can show that $B^{(0)}=O(L^{-d})$ uniformly in $m\le m_{\rm c}$ if $d>4$,
hence the summability of (\ref{eq:piNsawbd}) over $N\ge2$ when $L\gg1$.
Moreover,
by (\ref{eq:2<r<alpha-varphibd}) for $n=2l$,
\begin{eqnarray}\label{eq:W2lsawbd}
W^{(2l)}&\le&\int_{[-\pi,\pi]^d}\frac{\mathrm{d}^dk}{(2\pi)^d}
|\nabla_{1}^{2l}
\hat\varphi^\mathrm{SAW}(k,m)|\nonumber
\\[-2pt]
&\le&
O\biggl(\biggl(1-\frac
{m}{m_{\rm c}}\biggr)^{1-l}
\biggr)\int_{[-\pi,\pi]^d}\frac{\mathrm{d}^dk}{|k|^4}
\\[-2pt]
&\stackrel{d>4}=& O\biggl(\biggl(1-\frac{m}{m_{\rm
c}}\biggr)^{1-l}\biggr).\nonumber
\end{eqnarray}
Therefore,
\begin{eqnarray*}
\sum_{N=2}^\infty\sum_{t=0}^\infty m^t\sum_{x\in{\mathbb
Z}^d}|x_1|^{2l+q}
\pi_t^{
(N)}(x)\le O\biggl(\biggl(1-\frac{m}{m_{\rm c}}\biggr)^{1-l}\biggr)B^{(q)}.
\end{eqnarray*}
To complete the proof of (\ref{eq:2<r<alpha-pibd}) for self-avoiding
walk, it
suffices to show that there is an $\epsilon>0$ such that
$B^{(q)}=O((1-\frac{m}{m_{\rm c}})^{-{q}/2+\epsilon})$. For $q=2$,
we use (\ref{eq:2<r<alpha-varphibd}) for $n=2$ and take an arbitrary
$\epsilon\in(0,1\wedge\frac{d-4}2)$ to obtain
\begin{eqnarray*}
B^{(2)}&\le&\int_{[-\pi,\pi]^d}\frac{\mathrm{d}^dk}{(2\pi)^d}
|\hat
\varphi^\mathrm{SAW}(k,m) \nabla_{1}^2\hat\varphi
^\mathrm{SAW}(k,m)|
\\
&\le& O\biggl(\biggl(1-\frac{m}{m_{\rm c}}\biggr)^{-1+\epsilon}\biggr)\int_{[-\pi
,\pi]^d}
\frac{\mathrm{d}^dk}{|k|^{2(2+\epsilon)}}\le O\biggl(\biggl(1-\frac
{m}{m_{\rm c}}\biggr)^{-1
+\epsilon}\biggr).
\end{eqnarray*}
For $q\in(0,2)$, we first note that
\begin{eqnarray}\label{eq:Bqprebd}
\quad B^{(q)}\le\frac1{K_q}\int_0^\infty\frac{\mathrm
{d}u}{u^{1+q}}\int_{[-\pi,
\pi]^d}\frac{\mathrm{d}^dk}{(2\pi)^d} |\hat\varphi^\mathrm
{SAW}(k,m)
\bar\Delta_{\vec
u}\hat\varphi^\mathrm{SAW}(k,m)|.
\end{eqnarray}
It is known that, by \cite{hhs08}, Proposition~2.6, with an improvement
due to the same argument as in \cite{csII}, Proposition~2.1,
\begin{eqnarray*}
&&|\bar\Delta_{\vec u}\hat\varphi^\mathrm{SAW}(k,m)|
\\
&&\qquad \le\sum_{(j,j')=(0,\pm1),(1,-1)}\frac{O(1-\hat D(\vec
u))
}{1-{m/m_{\rm c}}
+1-\hat D(k+j\vec u)}
\\
&&\hspace*{98pt}{}\quad{}\times \frac{1}{1-{m/m_{\rm c}}+1-\hat D(k+j'\vec
u)}
\end{eqnarray*}
holds in the current setting, where the $O(1-\hat D(\vec u))$ term is
uniform in $k\in[-\pi,\pi]^d$ and $m\le m_{\rm c}$. Substituting
this, and
(\ref{eq:2<r<alpha-varphibd}) for $n=0,$ into (\ref{eq:Bqprebd}), and
using the
translation invariance and the ${\mathbb Z}^d$-symmetry of $D$ and the Schwarz
inequality (see \cite{csII}, formulas (4.27)--(4.29)), we end up with
\begin{eqnarray*}
B^{(q)}
&\le&\int_0^\infty\frac{\mathrm{d}u}{u^{1+q}}\int_{[-\pi,\pi]^d}\frac{\mathrm{d}^dk}{(2\pi)^d}
\frac{O(1-\hat D(\vec u))}{(1-
{m/m_{\rm c}}
+1-\hat D(k))^2}
\\
&&{}\hspace*{81pt}\times\frac{1}{1-{m}/{m_{\rm c}}+1-\hat D(k-\vec
u)}
\\
&\le&\int_0^\infty\mathrm{d}u \frac{1-\hat D(\vec u)}{u^{1+q}}\int
_{[-\pi,
\pi]^d}\frac{\mathrm{d}^dk}{(2\pi)^d} \frac{O((1-{m/m_{\rm
c}})^{-{q}/2
+\epsilon})}{(1-\hat D(k))^{2-{q}/2+\epsilon}(1-\hat D(k-\vec u))}
\end{eqnarray*}
for any $\epsilon\in(0,\frac{q}2)$. However, by following the proof of
\cite{csII}, formula (4.30), we can show that
\begin{eqnarray*}
\int_{[-\pi,\pi]^d}\frac{\mathrm{d}^dk}{(2\pi)^d} \frac1{(1-\hat D(k))^{2
-{q}/2+\epsilon}(1-\hat D(k-\vec u))}\le O
\bigl(u^{(d-6+q-2\epsilon)
\wedge0}\bigr),
\end{eqnarray*}
hence
\begin{eqnarray}\label{eq:Bqbd}
\quad B^{(q)}&\le& O\biggl(\biggl(1-\frac{m}{m_{\rm c}}\biggr)^{-{q}/2+\epsilon
}\biggr)\biggl(
\int_0^1\frac{\mathrm{d}u}u u^{(d-4-2\epsilon)\wedge(2-q)}+\int
_1^\infty
\frac{\mathrm{d}u}{u^{1+q}}\biggr)\nonumber
\\[-9pt]\\[-9pt]
&=&O\biggl(\biggl(1-\frac{m}{m_{\rm c}}\biggr)^{-{q}/2+\epsilon}
\biggr)\nonumber
\end{eqnarray}
if $\epsilon<\frac{d-4}2$. This completes the proof of (\ref
{eq:2<r<alpha-pibd})
for self-avoiding walk.
For oriented percolation, similarly to the proof of \cite{csII},
Lemma~3, by
using (\ref{eq:naive}) and diagrammatic bounds of the form (\ref
{eq:op-diagbds}),
we can show that, for $N\ge0$,
\begin{eqnarray}\label{eq:piNopbd}
\qquad &&\sum_{t=0}^\infty m^t\sum_{x\in{\mathbb Z}^d}|x_1|^{2l+q} \pi_t^{
(N)}(x)\nonumber
\\[-2pt]
&&\qquad\le(N+1)^{2l+q}\bigl(T^{(0)}\bigr)^{N-2}\bigl(\bigl(N\bigl(1+T^{(0)}\bigr)
+T^{(0)}\bigr) T^{(0)} V^{(q)}
\\[-2pt]
&&\hspace*{102pt}\qquad\quad{}+N\bigl((N-1)\bigl(1+T^{(0)}\bigr)+3T^{(0)}\bigr)T^{
(q)}
V^{(0)}\bigr),\nonumber
\end{eqnarray}
where
\begin{eqnarray*}
V^{(\nu)}
&=&\sup_{(x,t)\in{\mathbb Z}^{d+1}}\sum_{(y,s)\in{\mathbb Z}^{d+1}}|y_1|^{2l}
(mD*\varphi_s^\mathrm{OP})(y) m^s |y_1-x_1|^\nu
\\[-2pt]
&&{}\hspace*{81pt}\times (D*\varphi
_{s-t}^\mathrm{OP}
)(y-x),
\\[-2pt]
T^{(\nu)}
&=&\sup_{(x,t)\in{\mathbb Z}^{d+1}}\sum_{(y,s),(y',s')\in{\mathbb Z}^{d+1}}
(mD*\varphi_s^\mathrm{OP})(y) m^s |y_1-x_1|^\nu
\\
&&{}\hspace*{105pt}\times(D*\varphi
_{s'-t}^\mathrm{OP}
)(y'-x)
\\[-2pt]
&&{}\hspace*{105pt}\times\bigl(\varphi_{s-s'}^\mathrm
{OP}(y-y')+\varphi_{s'-s}^\mathrm{OP}
(y'-y)\bigr).
\end{eqnarray*}
Notice that
\begin{eqnarray}\label{eq:T0bd}
\qquad\quad T^{(0)}\le2m\int_{[-\pi,\pi]^d}\frac{\mathrm{d}^dk}{(2\pi
)^d} \hat D(k)^2
\int_{-\pi}^\pi\frac{\mathrm{d}\theta}{2\pi} |\hat\varphi
^\mathrm{OP}
(k,me^{i\theta})|
|\hat\varphi^\mathrm{OP}(k,e^{i\theta})|^2.
\end{eqnarray}
Using (\ref{eq:2<r<alpha-varphibd}) for $n=0$ and $\|D\|_\infty
=O(L^{-d})$, we
can show that $T^{(0)}=O(L^{-d})$ uniformly in $m\le m_{\rm c}$ if
$d>4$ and
$L\gg1$, hence the summability of (\ref{eq:piNopbd}) over $N\ge0$. Moreover,
by (\ref{eq:2<r<alpha-varphibd}) for $n=0,2l$ and using $|\hat
D(k)|\le1$,
we have
\begin{eqnarray}\label{eq:V0bd}
V^{(0)}
&\le&2^{2l}m\biggl(\int_{[-\pi,\pi]^d}
\frac{\mathrm{d}^dk}{(2\pi)^d}\int_{-\pi}^\pi\frac{\mathrm{d}\theta
}{2\pi}
|\nabla_{1}^{2l}\hat\varphi^\mathrm{OP}(k,me^{i\theta
})||\hat\varphi^\mathrm{OP}(k,e^{i
\theta})|\nonumber
\\[-2pt]
&&{}\hspace*{26pt}+\sum_{x\in{\mathbb Z}^d}|x_1|^{2l}D(x)\int_{[-\pi,\pi
]^d}\frac
{\mathrm{d}^dk}
{(2\pi)^d}\int_{-\pi}^\pi\frac{\mathrm{d}\theta}{2\pi} |\hat
\varphi^\mathrm{OP}
(k,me^{i\theta})|
\\[-2pt]
&&\hspace*{182pt}{}\hspace*{26pt}\times |\hat\varphi^\mathrm{OP}(k,e^{i\theta
})|\biggr)\nonumber
\\[-2pt]
&\stackrel{d>4}=&O\biggl(\biggl(1-\frac{m}{m_{\rm c}}\biggr)^{1-l}\biggr).\nonumber
\end{eqnarray}
To complete the proof of (\ref{eq:2<r<alpha-pibd}), it thus suffices to
show that
there is an $\epsilon>0$ such that
\begin{eqnarray*}
T^{(q)}=O\biggl(\biggl(1-\frac{m}{m_{\rm c}}\biggr)^{-{q}/2+\epsilon
}\biggr),\qquad
V^{(q)}=O\biggl(\biggl(1-\frac{m}{m_{\rm c}}\biggr)^{1-l-{q}/2+\epsilon
}\biggr).
\end{eqnarray*}
Here, we only explain the proof of the bound on $V^{(2)}$; the
bound on
$T^{(2)}$ can be proven quite similarly and the bounds on $T^{(q)}$
and $V^{(q)}$ for $q\in(0,2)$ can be proven by following a
similar line
of argument from (\ref{eq:Bqprebd}) through to (\ref{eq:Bqbd}). To prove
the bound
on $V^{(2)}$, we first note that
\begin{eqnarray}\label{eq:V2prebd}
\quad V^{(2)}&\le&2^{2l+2}m\biggl(\int_{[-\pi,\pi]^d}\frac{\mathrm
{d}^dk}{(2\pi)^d}
\int_{-\pi}^\pi\frac{\mathrm{d}\theta}{2\pi} |\nabla
_{1}^{2l}\hat
\varphi^\mathrm{OP}(k,
me^{i\theta})||\nabla_{1}^2\hat\varphi^\mathrm
{OP}(k,e^{i\theta})|\nonumber
\\
&&{}\hspace*{37pt}+\sigma^2\int_{[-\pi,\pi]^d}\frac{\mathrm{d}^dk}{(2\pi
)^d}\int_{-
\pi}^\pi\frac{\mathrm{d}\theta}{2\pi} |\nabla_{1}^{2l}\hat
\varphi
^\mathrm{OP}(k,me^{i
\theta})||\hat\varphi^\mathrm{OP}(k,e^{i\theta})|\nonumber
\\
&&{}\hspace*{37pt}+\sum_{x\in{\mathbb Z}^d}x_1^{2l}D(x)\int_{[-\pi,\pi
]^d}\frac{\mathrm{d}^dk}
{(2\pi)^d}\int_{-\pi}^\pi\frac{\mathrm{d}\theta}{2\pi} |\hat
\varphi^\mathrm{OP}(k,
me^{i\theta})|\nonumber
\\[-8pt]\\[-8pt]
&&{}\hspace*{197pt}\times |\nabla_{1}^2\hat\varphi^\mathrm
{OP}(k,e^{i\theta})|\nonumber
\\
&&{}\hspace*{37pt}+\sigma^2\sum_{x\in{\mathbb Z}^d}x_1^{2l}D(x)\int_{[-\pi
,\pi]^d}
\frac{\mathrm{d}^dk}{(2\pi)^d}\int_{-\pi}^\pi\frac{\mathrm{d}\theta
}{2\pi}
|\hat\varphi^\mathrm{OP}(k,me^{i\theta})|\nonumber
\\
&&{}\hspace*{37pt}\hspace*{186pt}\times |\hat\varphi
^\mathrm{OP}(k,e^{i\theta
})|\biggr).\nonumber
\end{eqnarray}
It is immediate from (\ref{eq:2<r<alpha-varphibd}) for $n=0,2$ that the
last two
lines are both $O(1)$ for $d>4$. Moreover, by (\ref{eq:V0bd}), the
second line
is $O((1-\frac{m}{m_{\rm c}})^{1-l})$ for $d>4$. For the first line,
we use the
following bounds due to (\ref{eq:2<r<alpha-varphibd}) for $n=2,2l$:
for any
$\epsilon\in(0,1)$,
\begin{eqnarray*}
|\nabla_{1}^{2l}\hat\varphi^\mathrm{OP}(k,me^{i\theta})|\le
\frac{O
((1-{m}/
{m_{\rm c}})^{-l+\epsilon})}{(|\theta|+|k|^2)^{1+\epsilon}},\qquad
|\nabla_{1}^2\hat\varphi^\mathrm{OP}(k,e^{i\theta})|\le O(|k|^{-4}),
\end{eqnarray*}
where the $O((1-\frac{m}{m_{\rm c}})^{-l+\epsilon})$ term is uniform in
$(k,\theta)\in[-\pi,\pi]^{d+1}$ and the $O(|k|^{-4})$ term is
uniform in
$\theta\in[-\pi,\pi]$. We then obtain that
\begin{eqnarray*}
&&\mbox{the first line of (\ref{eq:V2prebd})}
\\
&&\qquad \le\int_{[-\pi,\pi]^d}
\frac{\mathrm{d}^dk}{|k|^4}\int_{-\pi}^\pi\frac{\mathrm{d}\theta
}{2\pi}
\frac{O((1-{m/m_{\rm c}})^{-l+\epsilon})}{(|\theta
|+|k|^2)^{1
+\epsilon}}
\\
&&\qquad\le O\biggl(\biggl(1-\frac{m}{m_{\rm c}}\biggr)^{-l+\epsilon}\biggr)\int_{[-\pi
,\pi]^d}
\frac{\mathrm{d}^dk}{|k|^{4+2\epsilon}}=O\biggl(\biggl(1-\frac{m}{m_{\rm c}}\biggr)^{-l+
\epsilon}\biggr)
\end{eqnarray*}
if $\epsilon<\frac{d-4}2$. This completes the proof of (\ref
{eq:2<r<alpha-pibd})
for oriented percolation. This completes the proof of Lemma~\ref
{lmm:induction2}.
\end{pf*}
\subsection{\texorpdfstring{Proof of
Proposition~\protect\ref{proposition:r<alpha&2}}{Proof of Proposition~3.1}}
\label{ss:prop1}
First, we note that (\ref{eq:M-asy}) implies (\ref{eq:mu-asy}). To see
this, we
first substitute (\ref{eq:Jhat-diff}) and (\ref{eq:M-asy}) into (\ref
{eq:mu-def}) and
then use (\ref{eq:CICII}) [see (\ref{eq:M-source})] to obtain
\begin{eqnarray*}
\mu&=&\frac{m_{\rm c}\,\partial_m\hat J_{m_{\rm c}}(0) (1-
{m/m_{\rm c}})+O((1-{m/m_{\rm c}})^{1+\epsilon})}{M_{\mathrm c}v_\alpha+O
((1-
{m/m_{\rm c}})^\epsilon
)}
\\
&=&\frac{1-{m/m_{\rm c}}}{C_{\mathrm{II}}v_\alpha}+O
\biggl(\biggl(1-\frac{m}
{m_{\rm c}}\biggr)^{1+\epsilon}\biggr).
\end{eqnarray*}
Therefore, to complete the proof of Proposition~\ref{proposition:r<alpha&2},
it suffices to show (\ref{eq:M-asy}).
It is easier to prove (\ref{eq:M-asy}) for $\alpha\le2$. In this case,
$M$ in
(\ref{eq:M-def}) is reduced to
\begin{eqnarray*}
M=
\cases{
m,&\quad\mbox{RW/SAW},
\cr
\hat\pi^\mathrm{OP}(0)pm,&\quad\mbox{OP}.
}
\end{eqnarray*}
Therefore, (\ref{eq:M-asy}) is trivial for random walk and
self-avoiding walk.
For oriented percolation, we use (\ref{eq:critpt}) and (\ref{eq:property2})
to obtain
\begin{eqnarray*}
M_{\mathrm c}-M&=&\hat\pi_{\mathrm c}^\mathrm{OP}(0)p(m_{\rm
c}-m)+\bigl(\hat\pi_{\mathrm c}^\mathrm{OP}
(0)-\hat\pi^\mathrm{OP}(0)\bigr)
pm
\\
&=&1-\frac{m}{m_{\rm c}}+O(L^{-d}) \biggl(1-\frac{m}{m_{\rm c}}\biggr),
\end{eqnarray*}
where the $O(L^{-d})$ term is uniform in $m\le m_{\rm c}$.
This implies (\ref{eq:M-asy}).
It remains to prove (\ref{eq:M-asy}) for $\alpha>2$. In fact, we only need
investigate the crossover terms in (\ref{eq:M-def}) that are
proportional to
${\mathbh1}_{\{\alpha>2\}}$ and show that
\begin{eqnarray}\label{eq:desiredbd}
|\nabla_{1}^2\hat\pi_{\mathrm c}(0)-\nabla_{1}^2\hat\pi(0)|\le
O
\biggl(\biggl(1-\frac{m}
{m_{\rm c}}\biggr)^\epsilon\biggr)
\end{eqnarray}
since the above proof for $\alpha\le2$ directly applies to the noncrossover
terms. Notice that, for $\epsilon\in(0,1)$,
\begin{eqnarray*}
0&\le& m_{\rm c}^t-m^t\le m_{\rm c}^t\biggl(1-\biggl(\frac{m}{m_{\rm
c}}\biggr)^t\biggr)^{1-\epsilon}\biggl(
\frac{1-({m/m_{\rm c}})^t}{1-{m/m_{\rm c}}}
\biggr)^\epsilon\biggl(1-\frac{m}{m_{\rm c}}\biggr)^\epsilon
\\
&\le& m_{\rm c}^t t^\epsilon\biggl(1-\frac{m}{m_{\rm
c}}\biggr)^\epsilon
\end{eqnarray*}
so that
\begin{eqnarray*}
|\nabla_{1}^2\hat\pi_{\mathrm c}(0)-\nabla_{1}^2\hat\pi(0)|&\le&
\sum_{t\in
\mathbb{N}}(m_{\rm c}^t-m^t)
\sum_{x\in{\mathbb Z}^d}x_1^2 |\pi_t(x)|
\\
&\le&\biggl(1-\frac{m}{m_{\rm c}}\biggr)^\epsilon\sum_{t\in\mathbb
{N}}t^\epsilon
m_{\rm c}^t\sum_{x\in{\mathbb Z}^d}
x_1^2 |\pi_t(x)|.
\end{eqnarray*}
Moreover, since
\begin{eqnarray*}
t^\epsilon=\frac{t}{t^{1-\epsilon}}=\frac{t}{\Gamma(1-\epsilon
)}\int_0^\infty
\ell^{-\epsilon}e^{-\ell t}\, \mathrm{d}\ell,
\end{eqnarray*}
we have
\begin{eqnarray}\label{eq:desiredprebd}
&&|\nabla_{1}^2\hat\pi_{\mathrm c}(0)-\nabla_{1}^2\hat\pi(0)|\nonumber
\\[-8pt]\\[-8pt]
&&\qquad \le
\frac
{(1-{m/m_{\rm c}})
^\epsilon}{\Gamma(1-\epsilon)}\int_0^\infty\frac{\mathrm{d}\ell
}{\ell^\epsilon}
\sum_{t\in\mathbb{N}}t (m_{\rm c}e^{-\ell})^t\sum_{x\in{\mathbb
Z}^d}x_1^2|\pi_t(x)|.\nonumber
\end{eqnarray}
To show (\ref{eq:desiredbd}), it thus suffices to prove that the above
integral with respect to $\ell$ is $O(1)$ for sufficiently small
$\epsilon$.
First, we consider self-avoiding walk. By the diagrammatic bound on
$\pi_t^{(2)}(x)$ in (\ref{eq:saw-diagbds}) [see (\ref
{eq:pi2sawbd})], we
readily obtain
\begin{eqnarray*}
\sum_{t\in\mathbb{N}}t m^t\sum_{x\in{\mathbb Z}^d}x_1^2 \pi
_t^{(2)}(x)&\le&
\sum_{s,s',
s''\in\mathbb{N}}(s+s'+s'') m^{s+s'+s''}
\\
&&{}\qquad\hspace*{11pt}\times\sum_{x\in{\mathbb Z}^d}x_1^2 \varphi
_s^\mathrm{SAW}(x)
\varphi_{s'}^\mathrm{SAW}(x) \varphi_{s''}^\mathrm
{SAW}(x)
\\
&\le&3W^{(2)}\sum_{x\in{\mathbb Z}^d}\sum_{s,s'\in\mathbb{N}}s
m^{s+s'}\varphi_s^\mathrm{SAW}(x)
\varphi_{s'}^\mathrm{SAW}(x)
\\
&\le&3B'W^{(2)},
\end{eqnarray*}
where
\begin{eqnarray}\label{eq:B'-def}
B'\equiv B'(m)=\sup_{y\in{\mathbb Z}^d}\sum_{x\in{\mathbb Z}^d}\sum
_{t\in\mathbb{N}}t
m^t\varphi_t^\mathrm{SAW}(x)
\sum_{s=0}^\infty m^s\varphi_s^\mathrm{SAW}(y-x).
\end{eqnarray}
Similarly to the above and the derivation of (\ref{eq:piNsawbd}), we
can show
that, by (\ref{eq:saw-diagbds}) and~(\ref{eq:naive}),
\begin{eqnarray*}
\sum_{t\in\mathbb{N}}t m^t\sum_{x\in{\mathbb Z}^d}x_1^2 \pi
_t^{(N)}(x)\le N^4
\bigl(B^{(0)}\bigr)^{N-2}B'W^{(2)}\qquad[N\ge2].
\end{eqnarray*}
Since $B^{(0)}=O(L^{-d})$ and $W^{(2)}=O(1)$ uniformly in
$m\le m_{\rm c}$ if
$d>4$ [see formulas (\ref{eq:B0sawbd})--(\ref{eq:W2lsawbd})], we obtain that, for
$L\gg1$,
\begin{eqnarray}\label{eq:desiredprebd-saw1}
&&\int_0^\infty\frac{\mathrm{d}\ell}{\ell^\epsilon}\sum_{t\in
\mathbb{N}
}t (m_{\rm c}e^{-
\ell})^t\sum_{x\in{\mathbb Z}^d}x_1^2|\pi_t(x)|\nonumber
\\[-8pt]\\[-8pt]
&&\qquad \le\underbrace{\sum_{N=2}^\infty O(N^4)
O(L^{-d})^{N-2}}_{O(1)}\int_0^\infty
\frac{\mathrm{d}\ell}{\ell^\epsilon} B'(m_{\rm c}e^{-\ell
}).\nonumber
\end{eqnarray}
We now show that the integral of $B'(m_{\rm c}e^{-\ell})/\ell
^\epsilon$ is
uniformly bounded\vspace*{1pt} if $\epsilon<\frac{d-4}2$. First, we replace
$t\varphi_t^\mathrm{SAW}(x)$ in (\ref{eq:B'-def}) by the
following bound due to
subadditivity:
\begin{eqnarray*}
t\varphi_t^\mathrm{SAW}(x)=\sum_{s=1}^t\varphi
_t^\mathrm{SAW}(x)\le\sum
_{s=1}^t(\varphi_{s
-1}^\mathrm{SAW}*D*\varphi_{t-s}^\mathrm{SAW})(x).
\end{eqnarray*}
Then, by using $|\hat D(k)|\le1$ and (\ref{eq:2<r<alpha-varphibd})
for $n=0$,
we obtain
\begin{eqnarray}\label{eq:B'bd}
B'(m_{\rm c}e^{-\ell})&\le& m_{\rm c}e^{-\ell}\int_{[-\pi,\pi
]^d}\frac{\mathrm{d}^dk}{(2
\pi)^d} |\hat\varphi^\mathrm{SAW}(k,m_{\rm c}e^{-\ell})|^3\nonumber
\\[-8pt]\\[-8pt]
&\le& O(1)\int_{[-\pi,\pi]^d}\frac{\mathrm{d}^dk}{|k|^4} \frac
{e^{-\ell}}{1-
e^{-\ell}+|k|^2},\nonumber
\end{eqnarray}
where the $O(1)$ term is independent of $\ell$. However, for
$\epsilon\in(0,1)$,
\begin{eqnarray*}
\int_0^\infty\frac{\mathrm{d}\ell}{\ell^\epsilon} \frac{e^{-\ell
}}{1-e^{-\ell}
+|k|^2}&\le&\frac1{1-e^{-1}}\biggl(\int_0^1\frac{\mathrm{d}\ell
}{\ell^\epsilon}
\frac1{\ell+|k|^2}+\int_1^\infty\frac{\mathrm{d}\ell}{\ell
^\epsilon} e^{-\ell}\biggr)
\\
&\le&\frac1{1-e^{-1}}\biggl(\int_0^{|k|^2}\frac{\mathrm{d}\ell}{\ell
^\epsilon}
\frac1{|k|^2}+\int_{|k|^2}^1\frac{\mathrm{d}\ell}{\ell^{1+\epsilon
}}+1\biggr)
\\
&=&O(|k|^{-2\epsilon}).
\end{eqnarray*}
Therefore, if $\epsilon<\frac{d-4}2$, then we obtain
\begin{eqnarray}\label{eq:desiredprebd-saw2}
\int_0^\infty\frac{\mathrm{d}\ell}{\ell^\epsilon} B'(m_{\rm
c}e^{-\ell})\le O(1)
\int_{[-\pi,\pi]^d}\frac{\mathrm{d}^dk}{|k|^{4+2\epsilon}}=O(1).
\end{eqnarray}
Combining (\ref{eq:desiredprebd}), (\ref{eq:desiredprebd-saw1}) and
(\ref{eq:desiredprebd-saw2}), we complete the proof of (\ref
{eq:desiredbd}) for
self-avoiding walk.
For oriented percolation, similarly to the derivation of (\ref{eq:piNopbd}),
we can show that, for $N\ge0$,
\begin{eqnarray*}
&&\sum_{t\in\mathbb{N}}t m^t\sum_{x\in{\mathbb Z}^d}x_1^2 \pi
_t^{(N)}(x)
\\
&&\qquad\le
(N+1)^2\bigl(
T^{(0)}\bigr)^{N-2}\bigl(\bigl(N\bigl(1+T^{(0)}\bigr)+T^{(0)}
\bigr)
T^{(0)} V'
\\
&&{}\hspace*{88pt}\qquad\quad+N\bigl((N-1)\bigl(1+T^{(0)}\bigr)+3T^{(0)}\bigr)T' V^{(0)}\bigr),
\end{eqnarray*}
where
\begin{eqnarray*}
V'\equiv V'(m)
&=&\sup_{(x,t)\in{\mathbb Z}^{d+1}}\sum_{(y,s)\in{\mathbb
Z}^{d+1}}|y_1|^2(mD*
\varphi_s^\mathrm{OP})(y) m^s |s-t+1|\nonumber
\\
&&\hspace*{81pt}{}\times (D*\varphi
_{s-t}^\mathrm{OP})(y-x),\nonumber
\\
T'\equiv T'(m)
&=&\sup_{(x,t)\in{\mathbb Z}^{d+1}}\sum_{(y,s),(y',s')\in{\mathbb
Z}^{d+1}}(mD*
\varphi_s^\mathrm{OP})(y) m^s |s'-t+1|\nonumber
\\[-8pt]\\[-8pt]
&&\hspace*{105pt}{}\times (D*\varphi
_{s'-t}^\mathrm{OP})(y'-x)\nonumber
\\
&&\hspace*{105pt}{}\times\bigl(\varphi_{s-s'}^\mathrm
{OP}(y-y')+\varphi_{s'-s}^\mathrm{OP}\nonumber
(y'-y)\bigr).
\end{eqnarray*}
Since $T^{(0)}=O(L^{-d})$ and $V^{(0)}|_{l=1}=O(1)$ uniformly in
$m\le m_{\rm c}$ if $d>4$ and $p\le p_{\mathrm c}$ [see formulas (\ref
{eq:T0bd}) and (\ref{eq:V0bd})], we obtain
that, for $L\gg1$,
\begin{eqnarray}\label{eq:desiredprebd-op1}
&&\int_0^\infty\frac{\mathrm{d}\ell}{\ell^\epsilon}\sum_{t\in
\mathbb{N}
}t (m_{\rm c}
e^{-\ell})^t\sum_{x\in{\mathbb Z}^d}x_1^2|\pi_t(x)|\nonumber
\\[-8pt]\\[-8pt]
&&\qquad \le O(1)\int_0^\infty\frac{\mathrm{d}\ell}{\ell^\epsilon}
\bigl(V'(m_{\rm c}
e^{-\ell})+T'(m_{\rm c}e^{-\ell})\bigr).\nonumber
\end{eqnarray}
However, by the Markov property,
\begin{eqnarray*}
(t+1)(D*\varphi_t^\mathrm{OP})(x)=\sum_{s=0}^t(D*\varphi
_t^\mathrm{OP})(x)\le\sum_{s=0}^t
(\varphi_s^\mathrm{OP}*D*\varphi_{t-s}^\mathrm{OP})(x).
\end{eqnarray*}
Applying this bound to the definitions of $V'$ and $T'$ and then using
$|\hat D(k)|\le1$ and (\ref{eq:2<r<alpha-varphibd}) for $n=0,2$, we obtain
\begin{eqnarray*}
\left.\begin{array}{c}
V'(m_{\rm c}e^{-\ell}) \\[2pt] T'(m_{\rm c}e^{-\ell})
\end{array}
\right\}\le O(1)\int_{[-\pi,\pi]^d}\frac{\mathrm{d}^dk}{|k|^4}
\frac{e^{-\ell}}
{1-e^{-\ell}+|k|^2}.
\end{eqnarray*}
Recalling (\ref{eq:B'bd}) and (\ref{eq:desiredprebd-saw2}), we conclude
that (\ref{eq:desiredprebd-op1})
is uniformly bounded. This completes the proof of (\ref{eq:desiredbd}) for
oriented percolation. We have thus completed the proof of
Proposition~\ref{proposition:r<alpha&2}.
\begin{appendix}
\section*{Appendix}
\subsection{Asymptotics of $1-\hat D(k)$ for small $k$}\label{appendix:D}
In this appendix, we will use the following notation for convenience:
\begin{eqnarray*}
|\hspace*{-1,5pt}\|x\|\hspace*{-1,5pt}|_\ell=|x|\vee\ell\qquad[\ell>0].
\end{eqnarray*}
\renewcommand{\thetheorem}{A.\arabic{theorem}}
\setcounter{theorem}{0}
\begin{lemma}\label{lemma:D}
Let $\alpha,\rho>0$ and
\begin{eqnarray*}
h(x)=\frac{1+O(|\hspace*{-1,5pt}\|x|\hspace*{-1,5pt}\|_1^{-\rho})}{|\hspace*{-1,5pt}\|x|\hspace*{-1,5pt}\|_1^{d+\alpha}}\qquad
[x\in{\mathbb R}^d].
\end{eqnarray*}
Suppose that $h$ is a rotation-invariant function. Then, there exist
$\epsilon>0$ and $v_\alpha=O(L^{\alpha\wedge2})$ such that, for $|k|<1/L$,
the 1-step distribution $D$ in (\ref{eq:kac}) satisfies
\renewcommand{\theequation}{A.\arabic{equation}}
\setcounter{equation}{0}
\begin{eqnarray}\label{eq:1-hatDasy}
1-\hat D(k)=v_\alpha|k|^{\alpha\wedge2}\times
\cases{
1+O((L|k|)^\epsilon),&\quad$\alpha\ne2$,
\cr
\log\dfrac1{L|k|}+O(1),&\quad$\alpha=2$.
}
\end{eqnarray}
\end{lemma}
\begin{pf}
The case for $\alpha>2$ is easy. By the Taylor expansion of
$1-\cos(k\cdot x)$ and using the ${\mathbb Z}^d$-symmetry of $D$,
\begin{eqnarray*}
1-\hat D(k)=\sum_{x\in{\mathbb Z}^d}\bigl(1-\cos(k\cdot x)\bigr)
D(x)=\frac
{|k|^2}{2d}
\sum_{x\in{\mathbb Z}^d}|x|^2D(x)+O((L|k|)^{2+\epsilon})
\end{eqnarray*}
holds provided that $0<\epsilon<2\wedge(\alpha-2)$. This proves
(\ref{eq:1-hatDasy})
with $v_\alpha\equiv\sigma^2/(2d)=O(L^2)$.
It remains to prove (\ref{eq:1-hatDasy}) for $\alpha\le2$. First, we
note that,
by definition,
\begin{eqnarray*}
D(x)=\frac{c_h}{L^d} h(x/L)\qquad[x\in{\mathbb Z}^d],
\end{eqnarray*}
where
\begin{eqnarray*}
c_h=\biggl(\frac1{L^d}\sum_{y\in{\mathbb Z}^d/L}h(y)
\biggr)^{-1}=\int_{{\mathbb R}^d
}h(y)\, \mathrm{d}^d
y+O(L^{-1}).
\end{eqnarray*}
Taking the Fourier transform yields
\begin{eqnarray*}
1-\hat D(k)&=&\frac{c_h}{L^d}\sum_{x\in{\mathbb Z}^d}\bigl(1-\cos
(k\cdot
x)\bigr) h\biggl(
\frac{x}L\biggr)
\\
&=&\frac{c_h}{(L|k|)^d}\biggl(|k|^d\sum_{y\in|k|{\mathbb Z}^d}
\bigl(1-\cos
(e_k\cdot y)\bigr)
h\biggl(\frac{y}{L|k|}\biggr)\biggr),
\end{eqnarray*}
where $e_k=k/|k|$. By the Riemann sum approximation for small $k$ and the
rotational invariance of $h$, we obtain
\begin{eqnarray*}
1-\hat D(k)&=&\frac{c_h(1+O(|k|))}{(L|k|)^d}\int_{|y|\ge|k|}
\bigl(1-\cos(e_k
\cdot y)\bigr) h\biggl(\frac{y}{L|k|}\biggr)\, \mathrm{d}^dy
\\
&=&\frac{c_h(1+O(|k|))}{(L|k|)^d}\int_{|y|\ge|k|}(1-\cos y_1) h
\biggl(\frac{y}
{L|k|}\biggr)\, \mathrm{d}^dy
\\
&=&c_h(L|k|)^\alpha\bigl(1+O(|k|)\bigr)
\\
&&{}\times\int_{|y|\ge|k|}(1-\cos y_1)\biggl(\frac1{|\hspace*{-1,5pt}\|y|\hspace*{-1,5pt}\|_{L|k|}^{d+
\alpha}}+\frac{O((L|k|)^\rho)}{|\hspace*{-1,5pt}\|y|\hspace*{-1,5pt}\|_{L|k|}^{d+\alpha+\rho
}}\biggr)\,
\mathrm{d}^dy.
\end{eqnarray*}
This is the starting point of the analysis for $\alpha\le2$.
For $\alpha<2$, we note that
\begin{eqnarray*}
\int_{|y|\ge|k|}\frac{1-\cos y_1}{|\hspace*{-1,5pt}\|y|\hspace*{-1,5pt}\|_{L|k|}^{d+\alpha}}
\,\mathrm{d}^dy
&=&\int_{|y|\ge L|k|}\frac{1-\cos y_1}{|y|^{d+\alpha}}\, \mathrm{d}^dy
\\
&&{}+\underbrace{\int_{|k|\le|y|<L|k|}\frac{1-\cos
y_1}{(L|k|)^{d+\alpha}}
\,\mathrm{d}^dy}_{O((L|k|)^{2-\alpha})}
\\
&=&\int_{{\mathbb R}^d}\frac{1-\cos y_1}{|y|^{d+\alpha}} \,\mathrm{d}^dy-
\underbrace{\int_{|y|<L|k|}\frac{1-\cos y_1}{|y|^{d+\alpha}}\, \mathrm{d}^d
y}_{O((L|k|)^{2-\alpha})}
\\
&&{}+O((L|k|)^{2-\alpha}),
\end{eqnarray*}
where we have used $L|k|<1$ to estimate the error terms. Moreover,\vspace*{-2pt}
\begin{eqnarray*}
\int_{|y|\ge|k|}\frac{1-\cos y_1}{|\hspace*{-1,5pt}\|y|\hspace*{-1,5pt}\|_{L|k|}^{d+\alpha+\rho
}} \,\mathrm{d}^dy
&=&\underbrace{\int_{|y|\ge1}\frac{1-\cos y_1}{|y|^{d+\alpha+\rho
}}\, \mathrm{d}^d
y}_{O(1)}
+\int_{L|k|\le|y|<1}\frac{1-\cos y_1}{|y|^{d+\alpha+\rho}}
\,\mathrm{d}^dy
\\[-2pt]
&&{}+\underbrace{\int_{|k|\le|y|<L|k|}\frac{1-\cos
y_1}{(L|k|)^{d+\alpha+
\rho}} \,\mathrm{d}^dy}_{O((L|k|)^{2-\alpha-\rho})},\vspace*{-2pt}
\end{eqnarray*}
where\vspace*{-2pt}
\begin{eqnarray*}
\int_{L|k|\le|y|<1}\frac{1-\cos y_1}{|y|^{d+\alpha+\rho}}\, \mathrm{d}^dy=
\cases{
O(1),&\quad$\rho<2-\alpha$,\cr
O\biggl(\log\dfrac1{L|k|}\biggr),&\quad$\rho=2-\alpha$,\cr
O((L|k|)^{2-\alpha-\rho}),&\quad$\rho>2-\alpha$.
}\vspace*{-2pt}
\end{eqnarray*}
This proves (\ref{eq:1-hatDasy}) with $0<\epsilon<1\wedge(2-\alpha
)\wedge\rho$ and\vspace*{-2pt}
\begin{eqnarray*}
v_\alpha=c_hL^\alpha\int_{{\mathbb R}^d}\frac{1-\cos
y_1}{|y|^{d+\alpha}}
\,\mathrm{d}^dy.\vspace*{-2pt}
\end{eqnarray*}
For $\alpha=2$, we note that\vspace*{-2pt}
\begin{eqnarray*}
\int_{|y|\ge|k|}\frac{1-\cos y_1}{|\hspace*{-1,5pt}\|y|\hspace*{-1,5pt}\|_{L|k|}^{d+2}} \,\mathrm{d}^dy
&=&\underbrace{\int_{|y|\ge1}\frac{1-\cos y_1}{|y|^{d+2}}\, \mathrm
{d}^dy}_{O(1)}
+\int_{L|k|\le|y|<1}\frac{1-\cos y_1}{|y|^{d+2}} \,\mathrm{d}^dy
\\[-2pt]
&&{}+\underbrace{\int_{|k|\le|y|<L|k|}\frac{1-\cos
y_1}{(L|k|)^{d+2}}
\,\mathrm{d}^dy}_{O(1)}.\vspace*{-2pt}
\end{eqnarray*}
By the Taylor expansion of $1-\cos y_1$ and using $|y|^2=\sum_{j=1}^dy_j^2$,
we obtain\vspace*{-2pt}
\begin{eqnarray*}
\int_{L|k|\le|y|<1}\frac{1-\cos y_1}{|y|^{d+2}} \,\mathrm{d}^dy
&=&\frac12\int_{L|k|\le|y|<1}\frac{y_1^2}{|y|^{d+2}}\, \mathrm
{d}^dy+O(1
)\\[-2pt]
&=&\frac1{2d}\int_{L|k|\le|y|<1}\frac1{|y|^d}\, \mathrm{d}^dy+O(1)
\\[-2pt]
&=&\frac{\omega_d}
{2d}\log\frac1{L|k|}+O(1),\vspace*{-2pt}
\end{eqnarray*}
where $\omega_d\equiv2\pi^{d/2}/\Gamma(d/2)$ is the surface area of
the unit
$d$-sphere. Moreover,\vspace*{-2pt}
\begin{eqnarray*}
\int_{|y|\ge|k|}\frac{1-\cos y_1}{|\hspace*{-1,5pt}\|y|\hspace*{-1,5pt}\|_{L|k|}^{d+2+\rho}}
\,\mathrm{d}^dy
&=&\underbrace{\int_{|y|\ge L|k|}\frac{1-\cos y_1}{|y|^{d+2+\rho}}
\,\mathrm{d}^d
y}_{O((L|k|)^{-\rho})}
\\[-2pt]
&&{}+\underbrace{\int_{|k|\le|y|<L|k|}\frac{1-\cos
y_1}{(L|k|)^{d+2+\rho}}
\,\mathrm{d}^dy}_{O((L|k|)^{-\rho})}.\vspace*{-2pt}
\end{eqnarray*}
This proves (\ref{eq:1-hatDasy}) with $v_2=c_hL^2\omega_d/(2d)$.
\end{pf}
\subsection{Identity for the constant $K_r$}\label{appendix:K}
\begin{lemma}\label{lemma:K}
For $r\in(0,2)$,
\begin{eqnarray}\label{eq:lemma:K}
K_r\equiv\int_0^\infty\frac{1-\cos v}{v^{1+r}}\, \mathrm{d}v=\frac
\pi{2\Gamma(r+1)
\sin({r\pi}/2)}.
\end{eqnarray}
\end{lemma}
\begin{pf}
Below, we prove (\ref{eq:lemma:K}) only for $r\in(0,1]$. Since the
definition of
$K_r$ and the rightmost expression in (\ref{eq:lemma:K}) are both
analytic in
$r\in{\mathbb C}$ with \mbox{$0<\Re(r)<2$}, we can extend (\ref
{eq:lemma:K}) to $r\in(1,2)$
using analytic continuation.
First, we rewrite $K_r$ as
\begin{eqnarray}\label{eq:Kr-rewr}
K_r&=&\int_0^\infty\frac{\mathrm{d}u}{u^{1+r}}\int_0^u\sin v\, \mathrm{d}v
=\frac1r\int_0^\infty\frac{\sin v}{v^r}\,\mathrm{d}v\nonumber
\\[-8pt]\\[-8pt]
&=&\mathop{\lim_{R\to\infty}}_{\delta\to0}\frac1{2ir}\int
_\delta^R
\frac{e^{iv}-e^{-iv}}{v^r}\,\mathrm{d}v.\nonumber
\end{eqnarray}
For $a>0$, we let
\begin{eqnarray*}
\gamma_a^\pm&=&\biggl\{z=ae^{\pm i\theta}\dvtx \theta\mbox{ increases from 0
to }
\dfrac\pi2\biggr\},
\\
\eta^\pm&=&\{z=\pm iv\dvtx v\mbox{ increases from $\delta$ to }R\}.
\end{eqnarray*}
Then, by the Cauchy integral formula,
\begin{eqnarray*}
\int_\delta^R\frac{e^{iv}}{v^r}\,\mathrm{d}v&=&\int_{\gamma_\delta
^+}\frac{e^{iz}}
{z^r}\,\mathrm{d}z+\int_{\eta^+}\frac{e^{iz}}{z^r}\,\mathrm{d}z-\int
_{\gamma_R^+}
\frac{e^{iz}}{z^r}\,\mathrm{d}z
\\
&=&i\int_0^{\pi/2}\frac{e^{i\delta e^{i\theta}}}{(\delta
e^{i\theta})^{r-1}}
\,\mathrm{d}\theta+i^{1-r}\int_\delta^R\frac{e^{-v}}{v^r}\,\mathrm{d}v-
\underbrace{i\int_0^{\pi/2}\frac{e^{iRe^{i\theta
}}}{(Re^{i\theta})^{r-1}}\,
\mathrm{d}\theta}_{O(R^{-r})}.
\end{eqnarray*}
Similarly,
\begin{eqnarray*}
\int_\delta^R\frac{e^{-iv}}{v^r}\,\mathrm{d}v&=&\int_{\gamma_\delta^-}
\frac{e^{-iz}}{z^r}~\mathrm{d}z+\int_{\eta^-}\frac
{e^{-iz}}{z^r}\,\mathrm{d}z
-\int_{\gamma_R^-}\frac{e^{-iz}}{z^r}\,\mathrm{d}z
\\
&=&-i\int_0^{\pi/2}\frac{e^{-i\delta e^{-i\theta}}}{(\delta e^{-i
\theta})^{r-1}}\,\mathrm{d}\theta+(-i)^{1-r}\int_\delta^R\frac{e^{-v}}{v^r}\,
\mathrm{d}v+O(R^{-r}).
\end{eqnarray*}
Substituting these expressions back into (\ref{eq:Kr-rewr}) yields
\begin{eqnarray}\label{eq:Kr-rewr2}
K_r&=&\mathop{\lim_{R\to\infty}}_{ \delta\to0}\biggl(\frac{\delta
^{1-r}}{2r}
\int_0^{\pi/2}\biggl(\frac{e^{i\delta e^{i\theta}}}{e^{i\theta(r-1)}}+
\frac{e^{-i\delta e^{-i\theta}}}{e^{-i\theta(r-1)}}\biggr)\,\mathrm
{d}\theta\nonumber
\\[-9pt]\\[-9pt]
&&{}\hspace*{64pt}+i^{-r}\frac{1+(-1)^{-r}}{2r}\int_\delta^R\frac{e^{-v}}{v^r}\,\mathrm{d}v
\biggr).\nonumber
\end{eqnarray}
If $r=1$, then the second term is absent due to the cancelation $1+(-1)=0$.
By dominated convergence, we obtain
\begin{eqnarray}\label{eq:K1}
K_1=\lim_{\delta\to0}\frac12\int_0^{\pi/2}(e^{i\delta
e^{i\theta}}
+e^{-i\delta e^{-i\theta}})\,\mathrm{d}\theta=\int_0^{\pi/
2}\,\mathrm{d}
\theta=\frac\pi2.
\end{eqnarray}
If $r\in(0,1)$, on the other hand, the first term in (\ref
{eq:Kr-rewr2}) is
$O(\delta^{1-r})$ and therefore goes to zero as $\delta\to0$. Since
$(-1)^{-r}=(-1)^r=i^{2r}$ and $i^r+i^{-r}=2\cos\frac{r\pi}2$, we obtain
\begin{eqnarray*}
K_r=\frac{\cos({r\pi}/2)}r\int_0^\infty\frac
{e^{-v}}{v^r}\,\mathrm{d}v
=\frac{\cos({r\pi}/2)}r\Gamma(1-r).
\end{eqnarray*}
Using the well-known relations $\Gamma(1-r) \Gamma(r)=\pi/\sin
(r\pi)$ and
$r\Gamma(r)=\Gamma(r+1)$, we finally arrive at
\begin{eqnarray*}
K_r=\frac{\cos({r\pi}/2)}{r\Gamma(r)}~\frac\pi{\sin(r\pi)}
=\frac\pi{2\Gamma(r+1)\sin({r\pi}/2)}.
\end{eqnarray*}
This is also valid for $r=1$, due to (\ref{eq:K1}). This completes
the proof of Lem-\break ma~\ref{lemma:K}.
\end{pf}
\end{appendix}
\section*{Acknowledgments}
The first-named author is grateful to the staff of L-Station at
Hokkaido University for their
support and hospitality during a visit (July 23--August 5, 2008). The
second-named author is
grateful to Tai-Ping Liu and the Institute of Mathematics at Academia Sinica
in Taiwan for their support and hospitality during a visit (November
5--14, 2009). We would like to thank the referee for many valuable comments
regarding an earlier version of this paper.
|
\section{Introduction}
\setcounter{equation}{0}
The control of the Schr\"odinger equation has received a lot of
attention in the last decades. (See e.g.
\cite{Zua-2} for an excellent review of the contributions up to 2003).
Significant progresses have been made
for the linear Schr\"odinger equation on its controllability and
stabilizability properties (see \cite{jaffard,KL,lebeau,liu,mach-2,
mach-3,miller,RTTT} for control issues, and \cite{BP,CG,CCG,mor,YY} for Carleman estimates and their applications to inverse problems).
For the control of the so-called {\em bilinear}
Schr\"odinger equation, in which the bilinear term is linear in both
the control and the state function, see e.g.
\cite{BMS,CLP,BCG,Beauchard,BKP,MRT,ILT-2,BC,BS} and the references therein.
By contrast, the study of the nonlinear Schr\"odinger equation is still at
its early stage. Recently, Illner, Lange and Teismann
\cite{ILT-1,ILT-2} considered the internal controllability
of the nonlinear Schr\"odinger equation posed on a finite interval
with periodic boundary conditions:
\begin{equation}
\label{dim1}
iu_t+u_{xx}+f(u)=ia(x)h(x,t).
\end{equation}
In (\ref{dim1}), $a$ denotes a smooth real function which is strictly
supported in $\mathbb T$, the one-dimensional torus. They showed that the
system \eqref{dim1}
is locally exactly controllable in the space $H^1(\mathbb T )$.
Their approach was based on the well-known Hilbert Uniqueness
Method (HUM) and Schauder's fixed point theorem.
Later, Lange and Teismann \cite{LT-1} considered internal control for the
nonlinear Schr\"odinger equation (\ref{dim1}) posed on a finite interval with
the homogeneous Dirichlet boundary conditions
\begin{equation}
\label{dirichlet}
u(0,t)=u(\pi ,t)=0
\end{equation}
and
established local exact controllability of the system
(\ref{dim1})-(\ref{dirichlet}) in the space $H^1_0 (0, \pi)$
around a special ground state of the
system. Their approach was mainly based upon HUM and the implicit
function theorem. Dehman, G\'erard and Lebeau
\cite{DGL} studied the internal control and stabilization of a class
of defocusing nonlinear Schr\"odinger equations
posed on a two-dimensional compact Riemannian manifold $M$
without boundary
$$
iu_t+\Delta u + f(u)=ia(x)h(x,t).
$$
They demonstrated, in particular, that the system is (semiglobally)
exactly controllable and stabilizable in the space $H^1(M)$
assuming that the Geometric Control Condition and
some unique continuation condition are satisfied.
Recently, the authors proved in \cite{RZ2007b} that
the cubic Schr\"odinger equation on the torus $\mathbb T$ with a localized
control
\begin{equation}
\label{cubic}
iu_t + u_{xx} + \lambda |u|^2 u = ia(x)h(x,t),\quad x\in \mathbb T,
\end{equation}
is locally exactly controllable in $H^s(\mathbb T )$ for all
$s\ge 0$ (hence, in $L^2(\mathbb T )$). Inspired by the work of Russell-Zhang
in \cite{rz-1}, the method of proof combined the momentum approach and
Bourgain analysis. In the same paper, the local stabilization
by the feedback law $h=a(x) u(x,t)$ was established by applying
the contraction mapping theorem in some Bourgain space.
Finally, similar results were obtained with
Dirichlet (resp. Neumann) homogeneous boundary conditions thanks to
an extension argument.
More recently, Laurent has shown in \cite{laurent} that the system
(\ref{cubic}) is semiglobally exactly controllable and stabilizable.
The same result has also been derived by Laurent in \cite{laurent2} for
certain manifolds of dimension 3, including $\mathbb T ^3$, $S^3$,
and $S^2\times S^1$.
The propagation of compactness and regularity proved in
\cite{laurent,laurent2} plays a crucial role in the derivation of the
stabilization results in these papers.
See also \cite{LRZ} for another application of these ideas to
the semiglobal stabilization of the periodic Korteweg-de Vries equation.
In addition, the authors considered in \cite{RZ2008}
the following nonlinear Schr\"odinger equation
$$
iu_t + \Delta u + \lambda |u|^2 u =0
$$
posed on a bounded domain $\Omega $ in $\mathbb R ^n$ with either
the Dirichlet boundary conditions or the Neumann boundary conditions.
They showed that if
\[ s>\frac{n}{2}, \]
or
\[ 0\leq s< \frac{n}{2} \mbox{ with } 1\leq n < 2+2s,\]
or
\[ s=0,1 \mbox{ with } n=2,\]
then the systems with control inputs acting on the whole boundary of
$\Omega$ are locally exactly controllable in the classical Sobolev
space $H^s (\Omega) $ around any smooth solution of the
Schr\"odinger equation.
The aim of this paper is to extend the results of \cite{RZ2007b} to
any dimension. More precisely, we shall assume that the spatial variable
lives in the rectangle
$$
\Omega =(0,l_1)\times \cdots \times (0,l_n).
$$
We shall investigate the control properties of the semilinear
Schr\"odinger equation
\begin{equation}
iu_t + \Delta u +\lambda |u|^{\alpha }u = i a(x)h(x,t),
\end{equation}
where $\lambda \in \mathbb R$ and $\alpha\in 2\mathbb N^*$,
by combining new linear controllability results in the spaces
$H^s(\Omega )$ with Bourgain analysis. Let us briefly review
the results proved in this paper.
The internal controllability of the linear Schr\"odinger equation
on $\T ^n$
\begin{equation}
\label{LS}
iu_t + \Delta u = i a(x)h(x,t), \quad x\in \T ^n,\ t\in (0,T)
\end{equation}
is established in $H^s(\T ^n)$ for any $s\ge 0$ and
any function $a\not\equiv 0$. (Note that the Geometric Control Condition
is not required.) It is derived from a well-known
result in $L^2(\T ^n)$, due to Jaffard \cite{jaffard} when $n=2$
and Komornik \cite{komornik} for any $n\ge 2$, by an argument allowing
to shift the (state and control) space from $L^2(\T ^n )$ to $H^s(\T ^n )$.
In particular,
the exact controllability in $H^s(\T ^n )$ will require a control input
$h\in L^2(0,T;H^s(\T ^n))$. Similar results with Dirichlet or Neumann
homogeneous boundary conditions are deduced by
using the extension argument from \cite{RZ2007b}.
The boundary controllability of the linear Schr\"odinger equation is considered
both with Dirichlet control
\begin{equation}
\label{dirichletN}
u=1_{\Gamma _0}h(x,t)
\end{equation}
and with Neumann control
\begin{equation}
\label{NeumannN}
\frac{\partial u}{\partial \nu}= 1_{\Gamma _0}h(x,t).
\end{equation}
In \eqref{dirichletN} and in \eqref{NeumannN}, $\Gamma_0$ denotes an open
set in $\partial \Omega$. For the Dirichlet control, we shall prove that
in {\em any} dimension $n\ge 2$ the exact controllability holds in
$H^{-1}(\Omega )$ whenever $\Gamma_0$ is a neighborhood of a vertex of
$\Omega$. The observability inequality for this
(arbitrarily small) control region is actually
derived from the corresponding observability
inequality for internal control by multiplier techniques.
For the Neumann control, the exact controllability
in $L^2(\Omega )$ is obtained
in any dimension when $\Gamma _0$ is a side. Finally, the
results with Dirichlet (resp Neumann) boundary control
are extended to any Sobolev space $H^s(\Omega)$ with $s<1/2$
(resp. $s<1$)
by considering control inputs more regular in time, namely
$h\in H^{\frac{s+1}{2}}(0,T;L^2(\partial \Omega ))$
(resp. $h\in H^{\frac{s}{2}}(0,T;L^2(\partial \Omega ))$).
The extension of the above exact controllability results to the semilinear
Schr\"odinger equation
\begin{equation}
\label{NLS}
iu_t+\Delta u +\lambda |u|^{\alpha} u = i a(x)h(x,t)
\end{equation}
is performed on the basis of Bourgain analysis. The needed linear
and multilinear estimates are combined with a fixed-point argument to produce
local exact controllability results. Sharp results (for the support
of the control input) are given for the internal control. Boundary
controllability results are derived from those established for the
linear equation with the aid of estimates in Bourgain spaces of solutions
of boundary-value problems with boundary terms given by HUM.
Finally, the local exponential stabilization with an internal feedback law
is proved by following the same approach as in \cite{RZ2007b}.
The paper is organized as follows. The controllability results
for the linear Schr\"odinger equation are collected in Section 2.
Section 3 is devoted to the controllability
of the semilinear equations. Section 4 deals with the
internal stabilization issue.
Multilinear estimates for nonlinearities
of the form $u^{\alpha _1}{\overline u}^{\alpha _2}$ are established in
Appendix.
\section{Linear systems}
\subsection{Internal control}
We first consider the linear open loop control system for the Schr\"odinger
equation posed on $\T ^n:=(-\pi ,\pi )^n$ with periodic boundary conditions:
\begin{equation}
\label{I1}
iu_t+\Delta u =iGh:=ia(x)h(x,t),\ \ u(x,0)=u_0(x),
\end{equation}
where $a\in C^\infty (\T ^n)$ is a given smooth real-valued function
and $h=h(x,t)$ is the control input.
We denote by $H^s(\mathbb T ^n)$ the Sobolev space of the functions $u$ defined
on the torus $\T ^n$ (i.e. defined on $\mathbb R ^n$ and periodic of period
$2\pi$ with respect to each variable $x_i$)
for which the $H^s$ norm
$$
||u||_s=||(1-\Delta )^{s/2} u||_{L^2(\T ^n)}
$$
is finite.
We first establish an internal observability inequality for the solution
$v(t)=W(t)v_0$ of
\begin{equation}
\label{A0}
\left\{
\begin{array}{l}
iv_t +\Delta v=0 \qquad (x,t)\in \T ^n\times \mathbb R ,\\
v(0)=v_0.
\end{array}
\right.
\end{equation}
\begin{prop}
\label{prop1}
{\bf (Observability inequality in $H^{-s}(\mathbb T ^n)$)}
Let $a\in C^\infty (\T ^n)$ with $a\ne 0$ and $T>0$.
Then for any $s\ge 0$ there exists a constant $c>0$ such that
for any solution $v$ of
\eqref{A0} with $v_0\in H^{-s}(\mathbb T ^n)$, it holds
\begin{equation}
\label{A1}
||v_0||^2_{-s} \le c \int_0^T||av(t)||^2_{-s}dt.
\end{equation}
\end{prop}
\noindent{\em Proof.} We proceed in several steps.\\
\noindent
{\em Step 1.} Assume that $s=0$, and let
$$
\omega =\{ x\in (-\pi ,\pi)^n;\ |a(x)|>||a||_{L^\infty(\T ^n)} /2 \}.
$$
Then, by \cite[Lemma 8.9]{KL},
there exists some positive constant $c$ such that for any
square-summable sequence
$(c_k)_ {k\in \mathbb Z ^n \setminus \{ 0 \}}$ we have
\begin{equation}
\label{A2}
\sum_{k\ne 0}|c_k|^2
\le c\int_0^T\!\!\!\int_\omega \left|\sum_{k\ne 0}
c_k e^{ i (k\cdot x -|k|^2t)}\right|^2 dxdt.
\end{equation}
The result is still valid when the set of indices is changed into
$\mathbb Z ^n$ by \cite[Proposition 8.4]{KL}. This yields \eqref{A1} when $s=0$. \\
\noindent
{\em Step 2.} We prove the weaker inequality
\begin{equation}
||v_0||^2_{-s} \le c
\left( \int_0^T || av(t)||^2_{-s}dt + ||v_0||^2_{-s-1}
\right)
\label{A20}
\end{equation}
by contradiction.
If \eqref{A20} is false, then there exists a sequence
$\{ v_j\}$ of solutions of
\eqref{A0} in $C([0,T];H^{-s}(\T ^n ))$ such that
\begin{equation}
1=||v_j(0)||^2_{-s} \ge j
\left( \int_0^T || av_j(t)||^2_{-s}dt + ||v_j(0)||^2_{-s-1}
\right) .
\label{A3}
\end{equation}
Since $v_j$ is bounded in $L^\infty ([0,T];H^{-s}(\T ^n ))$ and $(v_j)_t$ is
bounded in $L^\infty ([0,T];H^{-s-2}(\T ^n ))$ by \eqref{A0}, we infer
from Aubin's lemma that, for a subsequence again
denoted by $\{ v_j\}$, we have for $j\to \infty$
$$
\left\{
\begin{array}{ll}
v_j\to v \qquad &\text{ in } L^\infty ([0,T];H^{-s}(\T ^n ))
\quad \text{weak }*\\
v_j\to v \qquad &\text{ in } C([0,T];H^r(\T ^n ))\quad \forall r<-s
\end{array}
\right.
$$
where $v\in C_w([0,T];H^{-s}(\T ^n ))$ is a solution of \eqref{A0}.
In particular, $v_j(0)\to v(0)$ in $H^r(\T ^n )$ for any $r<-s$. Since
$v_j(0)\to 0$ in $H^{-s-1}(\T ^n )$ by \eqref{A3}, we conclude that $v\equiv 0$.
Let $w_j=(1-\Delta )^{-s/2}v_j$. Then $w_j\in L^\infty ([0,T];L^2(\T ^n))$ and
$$
\left\{
\begin{array}{ll}
w_j\to 0 \qquad &\text{ in } L^\infty ([0,T];L^2(\T ^n))\quad \text{weak }* \\
w_j\to 0 \qquad &\text{ in } C([0,T];H^r (\T ^n ))\quad \forall r<0.
\end{array}
\right.
$$
Let us split $aw_j$ into
$$
aw_j=(1-\Delta )^{-s/2}(av_j) -(1-\Delta )^{-s/2}[a,(1-\Delta )^{s/2}]w_j.
$$
As the pseudodifferential operator $[a,(1-\Delta )^{s/2}]$ maps continuously
$H^r(\T ^n )$ into $H^{r-s+1}(\T ^n )$, we have that
\begin{equation}
\label{A4}
(1-\Delta )^{-s/2} [a,(1-\Delta )^{s/2} ]w_j \to 0 \qquad
\text{ in } C([0,T];H^r(\T ^n ))\ \text{ for any } r<1.
\end{equation}
Therefore, using \eqref{A3} and \eqref{A4}, we obtain that
$$
aw_j \to 0 \text{ in } L^2([0,T]; L^2(\T ^n)).
$$
Clearly, $w_j$ satisfies also the linear Schr\"odinger equation \eqref{A0},
so we infer from the observability inequality \eqref{A1}
established for $s=0$ that
$$
w_j(0)\to 0 \text{ in } L^2(\T ^n).
$$
It follows that $v_j(0)=(1-\Delta )^{s/2} w_j(0)\to 0$ in $H^{-s}(\T ^n )$,
contradicting the fact that $||v_j(0)||_{-s}=1$ for all $j$.\\
\noindent{\em Step 3.} We prove \eqref{A1} by contradiction.
If \eqref{A1} is false, there exists a sequence $\{ v_j\}$ of solutions
of \eqref{A0} in $C([0,T];H^{-s}(\T ^n ))$ such that
\begin{equation}
\label{A5}
1=||v_j(0)||^2_{-s}\ge j\int_0^T||av_j(t)||^2_{-s}dt \qquad
\forall j\ge 0.
\end{equation}
Extracting a subsequence if needed, we may assume that
\begin{eqnarray}
v_j\to v \qquad &&\text{ in } L^\infty ([0,T];H^{-s}(\T ^n )) \quad \text{weak}*
\label{A6}\\
v_j\to v \qquad &&\text{ in } C([0,T];H^r(\T ^n ))\quad \forall r<-s
\label{A7}
\end{eqnarray}
for some solution $v\in C_w([0,T];H^{-s}(\T ^n ))$ of \eqref{A0}, where
$C_w([0,T];H^{-s}(\mathbb T ^n)$ denotes the space of weakly sequentially
continuous functions from $[0,T]$ to $H^{-s}(\mathbb T ^n)$
(see \cite[Lemme 8.1]{LM}).
Clearly, $a v_j\to av$ in $L^\infty ([0,T];H^{-s}(\T ^n ))$ weak $*$ which,
combined to \eqref{A5}, yields $a v\equiv 0$.
An application of Holmgren theorem (see e.g. \cite[Theorem 8.6.5]{hormander})
gives $v\equiv 0$. On the other hand,
\eqref{A7} gives $v_j(0)\to 0$ in $H^{-s-1}(\T ^n )$.
It then follows from \eqref{A20} that $v_j(0)\to 0$ in $H^{-s}(\T ^n )$,
and this contradicts \eqref{A5}.
{\hfill$\quad$\qd\\}
Applying HUM \cite{L} with $L^2(\mathbb T ^n)$ as pivot space, we infer from
Proposition \ref{prop1} the following internal controllability
of the linear Schr\"odinger equation in $H^s(\mathbb T ^n)$.
\begin{thm}
\label{thm1}
Let $T>0$ and $s\ge 0$ be given. Then for any
$(u_0,u_1)\in H^s(\mathbb T ^n )\times H^s(\mathbb T ^n)$ there exists
a control $h\in L^2([0,T];H^s(\mathbb T ^n))$ such that the system \eqref{I1}
admits a unique solution
$u\in C([0,T]; H^s (\mathbb T ^n))$ satisfying $u(T)=u_1$.
Moreover, we can define a bounded operator
$$\Phi :H^s(\mathbb T ^n)\times H^s(\mathbb T ^n)\to L^2([0,T];H^s(\mathbb T ^n))
$$
such that for any $(u_0,u_1)\in H^s(\mathbb T ^n)\times H^s(\mathbb T ^n)$ it holds
\begin{equation}
\label{oper}
W(T)u_0+\int_0^T W(T-\tau )
( G(\Phi (u_0,u_1)))(\cdot , \tau )d\tau =u_1.
\end{equation}
\end{thm}
The (small) control region is represented in Figure \ref{tore1}.
Trapped rays are drawn to mean that the wave equation fails to be
controllable with such control regions.
\begin{figure}[hbtp]
\begin{center}
\epsfig{file=tore1,width=7cm}
\caption{Internal control of the Schr\"odinger equation.}
\label{tore1}
\end{center}
\end{figure}
\subsection{Boundary control}
In this section $\Omega =(0,\pi )^n$, and $\Gamma _0$
denotes an open set in $\partial \Omega$.
\noindent
\subsubsection{Dirichlet boundary control}
We first adopt the following definition.
\begin{defi}
\label{def1}
The open set $\Gamma _0\subset\partial \Omega$ is called a {\em Dirichlet
control domain} if given any $u_0,\ u_1\in H^{-1}(\Omega)$ and
any time $T>0$, one may find a control $h\in L^2(0,T;L^2(\Gamma _0))$
such that the solution $u=u(x,t)$ of
\begin{equation}
\label{B0}
\left\{
\begin{array}{ll}
iu_t+\Delta u = 0\qquad &\text{ in } \Omega\times (0,T)\\
u=1_{\Gamma _0}h(x,t)\qquad &\text{ on } \partial\Omega \times (0,T) \\
u(0)=u_0&
\end{array}
\right.
\end{equation}
satisfies $u(T)=u_1$.
\end{defi}
The following result provides Dirichlet control domains which are arbitrary
small in {\em any} dimension $n\ge 2$. Note that the wave equation
fails to be controllable with such control domains.
\begin{figure}[hbtp]
\begin{center}
\epsfig{file=tore2, width=7cm}
\caption{Boundary control of the Schr\"odinger equation.}
\label{tore2}
\end{center}
\end{figure}
\begin{thm}
\label{thm2}
Let $\Omega =(0,\pi )^n$, and
let $\Gamma _0\subset \partial \Omega$ be any open set
containing a vertex of $\partial \Omega$. Then $\Gamma _0$ is a Dirichlet
control domain.
\end{thm}
By Dolecki-Russell test of controllability (or HUM),
Theorem \ref{thm2} is a direct consequence of the following
boundary observability result for the system
\begin{equation}
\label{B1}
\left\{
\begin{array}{ll}
iv_t+\Delta v = 0\qquad &\text{ in } \Omega\times (0,T)\\
v=0\qquad &\text{ on } \partial\Omega \times (0,T) \\
v(0)=v_0.&
\end{array}
\right.
\end{equation}
\begin{prop}
\label{prop2}
Assume that the (open) control region $\Gamma _0\subset \partial \Omega$
contains a vertex of $\partial \Omega$. Then for every $T>0$, there exists a
constant $c>0$ such that
\begin{equation}
\label{B7}
||\nabla v_0||^2_{L^2(\Omega )}
\le c\int_0^T\!\!\! \int_{\Gamma _0}
\left\vert\frac{\partial v}{\partial \nu}
\right\vert ^2d\sigma dt
\end{equation}
for any solution $v$ of \eqref{B1} with $v_0\in H^1_0(\Omega )$.
\end{prop}
\noindent{\em Proof.} We proceed in several steps.\\
{\em Step 1.} First, we prove an observability inequality in $H^1_0(\Omega)$
with an internal observation in an arbitrary subdomain of $\Omega$.
\begin{lem}
\label{lem1}
Let $\omega \subset \Omega$ be an arbitrary nonempty open set.
Then there exists a constant $c>0$ such that
\begin{equation}
\label{B6}
||\nabla v_0||^2_{L^2(\Omega )}
\le c\int_0^T\!\!\! \int_{\omega}
\left\vert\nabla v (x,t)\right\vert ^2 dxdt
\end{equation}
for every solution $v$ of \eqref{B1} with
$v_0\in H^1_0(\Omega )$.
\end{lem}
\noindent
{\em Proof of Lemma \ref{lem1}.} Extend $v$ to $(-\pi,\pi)^n\times (0,T)$ in
such a way that $v$ is an odd function of $x_i$ for each $i=1,...,n$, and
extend the initial state $v_0$ in a similar way. Then $v$ solves
\eqref{A0}. Writing $v_0=\sum_{k\in \mathbb Z ^n}c_ke^{ik\cdot x}$, we have that
$$
\nabla v(x,t)=\sum_{k\in \mathbb Z ^n} ic_k e^{i(k\cdot x-|k|^2t)}k.
$$
It follows then from \eqref{A2} that
\begin{eqnarray*}
||\nabla v_0||^2_{L^2(\T ^n )}
&=& \sum_{j=1}^n \sum_{k\in \mathbb Z ^n} |k_j|^2|c_k|^2 \\
&\le& c\sum_{j=1}^n \int_0^T\!\!\!\int_\omega
\big\vert \sum_{k\in \mathbb Z ^n} c_ke^{i(k\cdot x - |k|^2t)}k_j\big\vert ^2dxdt \\
&\le& c \int_0^T\!\!\!\int_\omega |\nabla v|^2 dxdt.
\end{eqnarray*}
The lemma is proved.{\hfill$\quad$\qd\\}
\noindent
{\em Step 2.} We use the multiplier method to reduce the boundary
observation inequality to an internal observation inequality. Without
loss of generality, we may assume that $\Gamma_0$ is a (small)
neighborhood of the vertex $M=(\pi, ..., \pi)$ defined as
$$
\Gamma _0=\{ x\in \partial \Omega;\
x_1 + \cdots + x_n > n \pi -{\varepsilon} \},
$$
where ${\varepsilon}$ is a (possibly small) positive number.
The following lemma is needed.
\begin{lem}
\label{lem2}
There exists a nonnegative function $\theta \in C^3(\mathbb R ^n)$
which is null on $\{ x\in\mathbb R ^n;\ x_1\le 0\}$ and strictly convex on
$(0,+\infty )^n\cap B_1(0)$.
\end{lem}
\noindent
{\em Proof of Lemma \ref{lem2}.}
Set $y^+=\max (y,0)$ for all $y\in \mathbb R$. Let
$$
\theta (x_1,...,x_n)=(x_1^+)^4
\big( 1+\delta \sum_{j=2}^n(x_j^+)^4 \big)
$$
where $\delta >0$ is a small number whose value will be specified later.
Clearly, $\theta$ is a nonnegative function
of class $C^3$ on $\mathbb R ^n$, which vanishes on the set $\{ x_1\le 0\}$.
To prove that $\theta$ is strictly convex on $(0,+\infty)^n
\cap B_1(0)$, it is sufficient
to check that the Hessian matrix
\begin{equation}
\label{hessian}
H(x)=\left( \frac{\partial ^2\theta }{\partial x_i\partial x_j} (x)\right)
\end{equation}
is positive definite for every $x\in (0,+\infty )^n\cap B_1(0)$.
Simple computations give that for any $\xi\in\mathbb R ^n$,
$$
\xi ^TH(x)\xi =12 x_1^2 (1+\delta \sum_{j=2}^n x_j^4)\xi _1^2
+12\delta x_1^4 \sum_{j=2}^n x_j^2 \xi _j^2
+32\delta x_1^3\xi _1\sum_{j=2}^n x_j^3\xi _j.
$$
From Young inequality, we obtain that
$$
32 |x_1^3x_j^3\xi _1\xi _j|\le 26 x_1^2 x_j^4 \xi _1 ^2
+10x_1^4 x_j^2 \xi _j^2,
$$
therefore
\begin{equation}
\xi ^TH(x)\xi \ge (12-26(n-1)\delta ) x_1^2 \xi_1 ^2
+2\delta x_1^4 \sum_{j=2}^n x_j^2 \xi _j^2 \ge c|\xi |^2
\end{equation}
if $x\in (0,+\infty)^n\cap B_1(0)$ and $\delta <(6/13)(n-1)^{-1}$. {\hfill$\quad$\qd\\}
At this position, we need an identity from \cite{mach-2}.
\begin{lem}\cite[Lemma 2.2]{mach-2}
\label{lem3}
For any $q\in H^2(\Omega ,\mathbb R ^n)$ and any solution $v$ of
\eqref{B1} issued from $v_0\in H^1_0(\Omega )$, it holds
\begin{eqnarray}
&&\frac{1}{2}\int_0^T\!\!\!\int_{\partial \Omega}(q\cdot \nu)
\left\vert \frac{\partial v}{\partial \nu}\right\vert ^2 d\sigma dt
= \frac{1}{2}\text{\rm Im }
\int_{\Omega}(v q\cdot \nabla \bar v)dx\vert _0^T
\nonumber\\
&& +\frac{1}{2}\text{\rm Re }
\int_0^T\!\!\!\int_{\Omega}(v\nabla (\text{div }q)\cdot
\nabla \bar v )dxdt
+\text{\rm Re } \int_0^T\!\!\!\int _{\Omega}\sum_{j,k=1}^n
\frac{\partial q_k }{\partial x_j}\,
\frac{\partial \bar v}{\partial x_k}\,
\frac{\partial v }{\partial x_j}\, dxdt. \label{A10}
\end{eqnarray}
\end{lem}
Let
$$
\omega =\{ x\in \Omega;\ x_1+\cdots + x_n > n\pi -{\varepsilon} \}.
$$
We readily infer from Lemma \ref{lem2} that there exists a
convex function $\theta \in C^3(\overline{\Omega})$ which is
strictly convex on $\omega$ and null on $\overline{\Omega \setminus \omega}$.
Using \eqref{A10} with $q=\nabla \theta$ we obtain
\begin{equation}
\label{A11}
\int_0^T\!\!\!\int_\omega
\nabla {\bar v}(x)^TH(x)\nabla v (x)\, dxdt
\le c\int_0^T\!\!\!\int _{\Gamma _0} \left\vert
\frac{\partial v}{\partial \nu }\right\vert ^2d\sigma dt +
C_\delta \int_{\Omega} |v_0|^2 dx +\delta \int_\Omega
|\nabla v_0|^2 dx,
\end{equation}
where $\delta >0$ is a small number and $H(x)$ denotes the Hessian matrix
given in \eqref{hessian}.
In \eqref{A11}, we used the fact that both quantities
$||v(t)||_{L^2(\Omega )}$ and $||\nabla v(t)||_{L^2(\Omega )}$ are conserved.
Using Lemma \ref{lem1} and the fact that the Hessian matrix
$H(x)=(\partial ^2\theta/\partial x_i\partial x_j)(x)$
is positive definite on $\omega$, we obtain
\begin{equation}
\label{A40}
||\nabla v_0||^2_{L^2(\Omega )}
\le c\int_0^T\!\!\!\int_{\Gamma _0}
\left\vert \frac{\partial v}{\partial \nu}\right\vert ^2
d\sigma dt + C_\delta \int_\Omega |v_0|^2dx.
\end{equation}
for a convenient choice of $\delta$. The proof of the estimate
\begin{equation}
\label{A41}
||v_0||^2_{L^2(\Omega)} \le c\int_0^T\!\!\! \int_{\Gamma _0}
\left\vert \frac{\partial v}{\partial \nu}\right\vert ^2
d\sigma dt \end{equation}
is classical (see e.g. \cite[pp. 27-28]{mach-2}). Then
\eqref{B7} follows from \eqref{A40}-\eqref{A41}. This completes
the proof of Proposition \ref{prop2} and of Theorem \ref{thm2}.
{\hfill$\quad$\qd\\}
\begin{rem}
\begin{enumerate}
\item Theorem \ref{thm2} is stated for a square $\Omega =(0,\pi )^n$,
but it is valid (with the same proof) for any rectangle
$\Omega =(0,l_1)\times \cdots \times (0,l_n)$.
\item Using a frequential criterion and number theoretic arguments,
Ramdani et al. \cite{RTTT} proved that when $n=2$,
$\Gamma _0\subset \partial \Omega$ is a Dirichlet control
domain if and only if $\Gamma _0$
has both a horizontal and a vertical components. It is however
unclear whether the approach in \cite{RTTT} can yield
a similar result for $n\ge 3$.
\item Using Theorem \ref{thm1} on a rectangle $\tilde \Omega
=(-1 ,\pi )\times (0,\pi )^{n-1}$ with a control input
supported in $\tilde \Omega \setminus \Omega$, and next taking the
restriction to $\Omega$, we infer that the linear Schr\"odinger
equation is
controllable in $L^2(\Omega )$ with a control supported on a side.
(This fact can also be deduced from the Carleman inequalities
established in \cite{mor}.) This suggests that the condition for a domain
to be a Dirichlet control domain is less restrictive when the state
space is smoothed.
\end{enumerate}
\end{rem}
We now aim to extend Theorem \ref{thm2} to a control result in a space
$H^s(\Omega)$, with $s\ge -1 $. We define
$H^s_D(\Omega )=D(A_D^{\frac{s}{2}})$, where
$A_D$ is the Dirichlet Laplacian; i.e., $A_Du=-\Delta u$ with domain
$D(A_D)=H^2(\Omega)\cap H^1_0(\Omega ) \subset L^2(\Omega )$.
We first need to replace the characteristic function $1_{\Gamma _0}$ by a
smooth controller function $g\in L^\infty (\partial \Omega )$. We adopt the following
\begin{defi}
Let $g\in L^\infty (\partial \Omega )$. We say that $g$ is a
{\em smooth Dirichlet controller} if
\begin{enumerate}
\item[(i)] there exists a constant $C>0$ such that
\begin{equation}
\label{AA1}
||\nabla v_0||^2_{L^2(\Omega )} \le
C\int_0^T\!\!\!\int_{\partial \Omega } g(x)
\left\vert \frac{\partial v}{\partial \nu}\right\vert ^2 d\sigma dt
\end{equation}
for any solution $v$ of \eqref{B1} emanating from $v_0\in H^1_0(\Omega )$
at $t=0$;
\item for any face $F$ of $\partial \Omega $, $g_F=g_{\vert _F}\in C^\infty (F)$
and for all $k\ge 0$
\begin{equation}
\label{AA10}
\frac{\partial ^{2k+1} g_F}{\partial \nu ^{2k+1}} =0 \quad \text{ on }
\partial F.
\end{equation}
\end{enumerate}
\end{defi}
Note that for any nonempty open set $\Gamma _0\subset \partial \Omega$ one
can construct a smooth Dirichlet controller $g$ supported in $\Gamma _0$.
Consider for example a small neighborhood $\Gamma _0 = [0,\varepsilon ]^n\cap
\partial \Omega $ of $0$ in $\partial \Omega$. A smooth Dirichlet controller
$g$ supported in $\Gamma _0$ is given by
$$
g(x_1,...,x_n)=\prod_{i=1}^n \rho (x_i)
$$
where $\rho \in C^\infty (\mathbb R )$ fulfills
$$
\rho (s) =
\left\{
\begin{array}{ll}
1 & \text{\rm if } s\le \frac{\varepsilon}{4},\\
0 & \text{\rm if } s\ge \frac{\varepsilon}{2}.
\end{array}
\right.
$$
Note also that $g\in C^0(\partial \Omega )$ and that the set
$\{ x\in \partial \Omega;\ g(x)>0 \}$
is an open neighborhood of $0$ in $\partial \Omega$.
Let $g$ be a smooth Dirichlet controller, and let $S$ denote the bounded
operator $H^1_0 (\Omega )\to H^{-1}(\Omega )$ defined by $Sv_0=u(T)$, where
$u=u(x,t)$ solves
\begin{equation}
\label{AA2}
\left\{
\begin{array}{ll}
iu_t+\Delta u = 0\qquad &\text{ in } \Omega\times (0,T)\\
u=g(x)h(x,t)\qquad &\text{ on } \partial\Omega \times (0,T) \\
u(0)=0&
\end{array}
\right.
\end{equation}
with $h(x,t)=(\partial v/\partial \nu ) (x,t)$, $v=W_D(t)v_0$ denoting
the solution of
\begin{equation}
\label{AA2bis}
\left\{
\begin{array}{ll}
iv_t+\Delta v = 0\qquad &\text{ in } \Omega\times (0,T)\\
v=0\qquad &\text{ on } \partial\Omega \times (0,T) \\
v(0)=v_0.
\end{array}
\right.
\end{equation}
Applying HUM, we infer from the observability inequality \eqref{AA1}
that $S$ is invertible. We shall prove that a similar result holds in more
regular spaces.
\begin{thm}
\label{thm3}
Pick any number $s\in [-1,\frac{1}{2} )$. Then $S$ is an isomorphism from
$H_D^{s+2}(\Omega )$ onto $H^s_D(\Omega )$. More precisely, for any
$T>0$ and any $u_T \in H^s_D(\Omega )$, if we set
$h(x,t)=(\partial v/\partial \nu ) (x,t)$ where $v$ denotes the solution of \eqref{AA2bis} with $v_0=S^{-1}u_T$, then $v_0\in H^{s+2}_D(\Omega )$,
$h\in H^{\frac{s+1}{2}}(0,T;L^2(\partial \Omega ))$, and
the solution $u$ of \eqref{AA2} satisfies $u\in C([0,T]; H^s_D(\Omega ))$ and
$u(T)=u_T$.
\end{thm}
\noindent{\em Proof.}
{\em Step 1.} Let us first check that $S^{-1}$ is a bounded operator from
$H^s_D(\Omega )$ into $H^{s+2}_D(\Omega )$ for $s\in [-1,\frac{1}{2})$.
The result is already known for $s=-1$. Assume first that
$ -1 < s < 0 $, and pick any $u_T \in H^s_D(\Omega )$ decomposed as
$$
u_T(x) =\sum_{p\in (\mathbb N ^*)^n} u_{T,p} \sin(p_1 x_1)\cdots \sin(p_nx_n),
$$
with $\sum_{p\in (\mathbb N ^*)^n} |p|^{2s}|u_{T,p}|^2<\infty$.
Let $v_0 = S^{-1}(u_T)\in H^1_D(\Omega )$
decomposed as
\begin{equation}
\label{XY1}
v_0(x) =\sum_{p\in (\mathbb N ^*)^n}
v_p \sin(p_1x_1)\cdots \sin(p_nx_n),
\end{equation}
and let $v$ denote the solution of \eqref{AA2bis}. The control
given by HUM driving \eqref{AA2} from $0$ to $u_T$ reads
\begin{equation}
\label{h1992}
h(x,t):=\partial v/\partial \nu
=\sum_{p\in (\mathbb N ^*)^n}
v_p e^{-i|p|^2t}
\frac{\partial}{\partial \nu}
(\sin(p_1x_1)\cdots \sin(p_nx_n)).
\end{equation}
Let us write the solution $u=u(x,t)$ of \eqref{AA2} in the form
\begin{equation}
\label{XY2}
u(x,t) =\sum_{p\in (\mathbb N ^*)^n} u_p(t)
\sin(p_1x_1)\cdots \sin(p_nx_n).
\end{equation}
The moments $\{ u_p(t)\} _{p\in (\mathbb N ^*)^n}$ can be computed from
the control input $h$ by using duality. Scaling in \eqref{AA2}
by $\overline{w}$, where $w=W_D(t)w_0$ is a smooth solution, we obtain
$$
i\int_\Omega u(x,t)\overline{w(x,t)}\, dx =
\int_0^t \int_{\partial \Omega} g(x) h(x,\tilde t)
\overline{\frac{\partial w}{\partial \nu}}\, d\sigma (x) d\tilde t.
$$
Pick any $q\in (\mathbb N ^*)^n$ and choose
$w_0(x)=\sin ( q_1x_1)\cdots \sin(q_nx_n)$.
We obtain from \eqref{h1992} that
\begin{eqnarray}
(\frac{\pi}{2})^n ie^{i|q|^2t} u_q(t)
&=& \int_0^t \int_{\partial \Omega} g(x)
h (x,\tilde t) e^{i|q|^2\tilde t }
\frac{\partial}{\partial\nu}(\sin (q_1 x_1)\cdots \sin (q_n x_n))
d\sigma (x)d\tilde t \nonumber \\
&=& \sum_{p \in (\mathbb N ^*)^n} v_p
(\int_0^t e^{i(|q|^2-|p|^2)\tilde t } d\tilde t) \nonumber \\
&&\!\!\!\!\!\! \times\!\!\!
\int_{\partial \Omega} g(x)
\frac{\partial}{\partial\nu}(\sin (p_1 x_1)\cdots \sin (p_n x_n))
\frac{\partial}{\partial\nu}(\sin (q_1 x_1)\cdots \sin (q_n x_n)) d\sigma (x).
\label{XY3}
\end{eqnarray}
It follows that for $t=T$
\begin{equation}
\label{solution}
S(v_0)=u_T=u(T)=\sum_{q\in (\mathbb N ^*)^n}
\left( \sum_{p\in (\mathbb N ^*)^n} a_{q,p} v_p \right)
\sin (q_1 x_1)\cdots \sin (q_n x_n)
\end{equation}
with
\begin{equation}
\label{AA3}
a_{q,p}=
-(\frac{2}{\pi}) ^{n} \frac{e^{-i|p|^2T} - e^{-i|q|^2T}}{|q|^2-|p|^2}
\int_{\partial \Omega} g(x)
\frac{\partial}{\partial\nu}(\sin (p_1 x_1)\cdots \sin (p_n x_n))
\frac{\partial}{\partial\nu}(\sin (q_1 x_1)\cdots \sin (q_n x_n)) d\sigma (x).
\end{equation}
In \eqref{AA3}, we used the convention that
\begin{equation}
\label{XY4}
\frac{e^{-i|p|^2t} - e^{-i|q|^2t}}{|q|^2-|p|^2}
=it e^{-i|q|^2t} \qquad \text{ for }\ |p|=|q|.
\end{equation}
Introduce the operator $D^\sigma$ defined by
$$
D^\sigma
\left(
\sum_{p\in (\mathbb N ^*)^n} c_p\sin (p_1x_1)\cdots \sin (p_nx_n)
\right)
=\sum_{ p \in (\mathbb N ^*)^n} |p|^\sigma c_p\sin (p_1x_1)\cdots \sin (p_nx_n).
$$
In what follows, $\sum _p$ and $\sum _q$ will stand for
$\sum _{p\in (\mathbb N ^*)^n}$ and $\sum _{q\in (\mathbb N ^*)^n}$, respectively.
We aim to prove that $v_0\in H^{s+2}_D (\Omega )$ for
$u_T \in H^s_D ( \Omega )$. For $v_0$ given by \eqref{XY1}, let
$$
||v_0||_s^2=\sum_p |p|^{2s} |v_p|^2.
$$
$C$ denoting a constant varying from line to line, we have that
\begin{eqnarray}
||v_0||_{s+2}
&\le& ||D^{s+1} v_0||_1 \nonumber\\
&\le & C||S(D^{s+1}v_0)||_{-1} \nonumber\\
&\le & C \left( ||D^{s+1}(Sv_0)||_{-1} + ||[S,D^{s+1}]v_0||_{-1}\right)
\nonumber\\
&\le & C\left( ||u_T||_s + || [S,D^{s+1} ] v_0 ||_{-1} \right) .
\label{ABCD1}
\end{eqnarray}
Clearly
$$
[S,D^{s+1}] v_0 = \sum_q \left( \sum _p a_{q,p}
(|p|^{s+1} - |q|^{s+1}) v_p\right) \sin (q_1x_1)\cdots \sin (q_nx_n),
$$
hence
$$
||[S,D^{s+1}] v_0 ||_{-1}^2 =
\sum _q |q|^{-2} [ \sum_p a_{q,p} (|p|^{s+1} - |q|^{s+1}) v_p ] ^2.
$$
Writing $\partial \Omega =\cup_{ 0\le l < 2^n -1 } F_l$,
where the $F_l$'s denote the faces of $\Omega$, the integral term
in \eqref{AA3} may be written $\sum_{0\le l < 2^n - 1 } I_{F_l}$, with
$$
I_{F_l}:=
\int_{F_l} g(x)
\frac{\partial}{\partial\nu}(\sin (p_1 x_1)\cdots \sin (p_n x_n))
\frac{\partial}{\partial\nu}(\sin (q_1 x_1)\cdots \sin (q_n x_n)) d\sigma (x).
$$
Let us estimate $I_{F_l}$ for $F_0:=\{ x\in \partial \Omega ; x_n=0\} =
[0,\pi ]^{n-1}\times \{ 0 \} $. Then
\begin{eqnarray*}
|I_{F_0}| &=& p_n q_n
\left\vert
\int_{[0,\pi ]^{n-1}} g(x_1,...,x_{n-1},0)
[\prod_{j=1}^{n-1}\sin (p_jx_j)\sin (q_jx_j)]
dx_1\cdots dx_{n-1}
\right\vert\\
&=& p_n q_n
\left\vert
\int_{[0,\pi ]^{n-1}} g(x_1,...,x_{n-1},0)
[\prod _{j=1}^{n-1}\frac{1}{2}
(\cos (p_j-q_j)x_j - \cos (p_j+q_j)x_j )
] dx_1\cdots dx_{n-1} \right\vert.
\end{eqnarray*}
Using \eqref{AA10} and integrations by parts, we see that for every
$k\in \mathbb N$, we have for some constant $C_k>0$
\begin{equation}
\label{XYZ}
|I_{F_0} | \le C_k p_nq_n \prod _{j=1}^{n-1} \langle p_j-q_j\rangle ^{-k}.
\end{equation}
The corresponding contribution in $||[S, D^{s+1}]v_0||_{-1}^2$
is therefore estimated by
$$A_{F_0}=\sum _q |q|^{-2}
\left(
\sum_p p_n q_n (\prod _{j=1}^{n-1} \langle p_j- q_j\rangle ^{-k})
\langle |q|^2 - |p|^2 \rangle ^{-1} ||p|^{s+1}-|q|^{s+1}| |v_p|
\right) ^2.$$
Since
$$
\frac{\vert |p|^{s+1} -|q|^{s+1}\vert }{\langle |q|^2 - |p|^2 \rangle}
\le C \frac{\left\vert |p| - |q| \right\vert (|p|^s +|q|^s)}
{\langle |q|^2 -|p|^2 \rangle}
\le C \frac{|p|^s + |q|^s}{ |p| + |q| }
$$
we have by Cauchy-Schwarz
\begin{eqnarray}
A_{F_0} &\le&
C\sum_q
[\sum _p p_n (\prod _{j=1}^{n-1} \langle p_j-q_j\rangle ^{-k})
\frac{|p|^s + |q|^s}{|p|+ |q|} |v_p| ]^2\nonumber \\
&\le& C\sum_q
\left( \sum_p \frac{|p|^{2s} + |q|^{2s}}{(|p|+ |q|)^2}\prod_{j=1}^{n-1}
\langle p_j-q_j\rangle ^{-k} \right) \cdot
\left( \sum _p p_n^2 |v_p|^2 \prod _{j=1}^{n-1}
\langle p_j - q_j \rangle ^{-k} \right) \label{AA20}
\end{eqnarray}
Pick any $k>1$. Then, as $s<0$, if we choose $k>1$
$$
\sum_{q_n}\sum_p \frac{|p|^{2s} + |q|^{2s}}{(|p|+ |q|)^2}
\prod_{j=1}^{n-1} \langle p_j-q_j\rangle ^{-k}
\le
\sum_{q_n}\sum_{p_n} \frac{p_n^{2s} + q_n^{2s}}{(p_n+q_n)^2}
\sum_{p_1,\ldots ,p_{n-1}} \prod_{j=1}^{n-1}\langle p_j-q_j\rangle ^{-k}<\infty.
$$
Therefore
\begin{eqnarray*}
A_{F_0} &\le& C\sum_{q_1,\ldots ,q_{n-1}} \sum_p p_n^2 |v_p|^2
\prod_{j=1}^{n-1} \langle p_j-q_j\rangle ^{-k} \\
&\le & C\sum_p |p|^2 |v_p|^2 \sum_{q_1,\ldots ,q_{n-1}}
\prod_{j=1}^{n-1} \langle p_j-q_j\rangle ^{-k} \\
&\le & C\sum_{p} |p|^2 |v_p|^2.
\end{eqnarray*}
The estimate for another face $F_l$ is similar. We conclude that
$$
||[S,D^{s+1}]v_0||^2_{-1} \le C ||v_0||_1^2 \le C ||u_T||_{-1}^2
$$
hence, with \eqref{ABCD1}, $v_0\in H_D ^{s+2}(\Omega )$.
Let us now assume that $u_T \in H_D ^s (\Omega )$ with $0\le s<\frac{1}{2}$.
The proof is carried out as above when $-1<s<0$, except for the estimate of
$A_{F_0}$ in \eqref{AA20}. We know from the lines above that
$v_0\in H_D ^\sigma (\Omega )$ for any $\sigma <2$. Then, by Cauchy-Schwarz
inequality,
\begin{eqnarray}
A_{F_0}
&\le& C\sum_q \left(\sum _p p_n \prod_{j=1}^{n-1}
\langle p_j-q_j\rangle ^{-k} \frac{|p|^s+|q|^s}{|p| + |q|} |v_p| \right)^2
\label{orion1}\\
&\le& C\sum_q
\left(\sum _p \frac{|p|^{2s}+|q|^{2s}}{(|p| + |q|)^2}
|p|^{-1} \prod_{j=1}^{n-1}\langle p_j-q_j\rangle ^{-k} \right)
\left(\sum _p p_n^2 |p||v_p|^2
\prod_{j=1}^{n-1}
\langle p_j-q_j\rangle ^{-k} \right) . \nonumber
\end{eqnarray}
Note that
$$
\sum_{q_n}\left( \sum_p \frac{|p|^{2s}+|q|^{2s}}{(|p|+|q|)^2} |p|^{-1}
\prod _{j=1}^{n-1}\langle p_j-q_j\rangle ^{-k}\right)
\le C(S_1+S_2+S_3)
$$
where
\begin{eqnarray*}
S_1&=& \sum_{q_n}\left( \sum_p
\frac{ |p|^{2s-1} }{ (|p|+|q|)^2} \prod_{j=1}^{n-1}
\langle p_j-q_j\rangle ^{-k}\right) \\
S_2&=& \sum_{q_n}\left( \sum_p
\frac{ q_n^{2s}|p|^{-1} }{ (|p|+|q|)^2} \prod_{j=1}^{n-1}
\langle p_j-q_j\rangle ^{-k}\right) \\
S_3&=& \sum_{q_n}\left( \sum_p
\frac{ |q'|^{2s}|p|^{-1} }{ (|p|+|q|)^2} \prod_{j=1}^{n-1}
\langle p_j-q_j\rangle ^{-k}\right) \qquad \text{ where }\ q=(q',q_n).
\end{eqnarray*}
Since $2s-1<0$,
$$
S_1\le \sum_{q_n}
\left(
\sum_p \frac{p_n^{2s-1}}{(p_n+q_n)^2}
\prod_{j=1}^{n-1}\langle p_j-q_j\rangle ^{-k}
\right) \le const<\infty .
$$
Also,
\begin{eqnarray*}
S_2 &\le & \sum_{q_n}\left(
\sum_p \frac{q_n^{2s}p_n^{-1}}{(p_n+q_n)^2} \prod_{j=1}^{n-1}
\langle p_j-q_j\rangle ^{-k} \right) \\
&\le& C\sum_{p_n}\sum_{q_n} \frac{q_n^{2s} p_n^{-1} }{ (p_n+q_n)^2} \\
&\le& C\sum_{p_n}\left( \frac{1}{p_n(p_n+1)} + p_n^{2s-3} +
\int_1^\infty
\frac{x^{2s}}{p_n(p_n+x)^2} dx \right) \\
&\le&
C\left( 1+\sum_{p_n\ge 1} p_n^{2s-2} \int_0^{+\infty} \frac{y^{2s}}{(1+y)^2}dy \right)\\
&\le& const<\infty.
\end{eqnarray*}
Finally,
$$
S_3 \le |q'|^{2s} \sum_{q_n}\sum_p \frac{p_n^{-1}}{(p_n+q_n)^2}
\prod_{j=1}^{n-1} \langle p_j-q_j\rangle ^{-k}
\le C |q'| ^{2s}.
$$
It follows that
$$
A_{F_0} \le C \sum_{q_1,\ldots ,q_{n-1}} \sum_p p_n^2 |p| |v_p|^2
|q'| ^{2s} \prod _{j=1}^{n-1} \langle p_j-q_j\rangle ^{-k}.
$$
Note that
$$
\sum_{q_1,\ldots ,q_{n-1}} |q'| ^{2s}
\prod_{j=1}^{n-1} \langle p_j - q_j \rangle ^{-k}
\le C |p'| ^{2s}
$$
since, for $k>2s+1$,
$$
\sum_{q_j} q_j^{2s} \langle p_j-q_j \rangle ^{-k}
\le Cp_j^{2s}.
$$
(Split the sum into one for $q_j\le 2p_j$, and another one for $q_j>2p_j$.)
Therefore, since $0\le s<1/2$,
\begin{equation}
\label{orion2}
A_{F_0} \le C \sum_p |p|^{3+2s} |v_p|^2 =
||v_0||^2_{s+\frac{3}{2}} \le C ||u_T||^2_{s-\frac{1}{2}}.
\end{equation}
Thus, we have proved that $S^{-1}$ is bounded from $H^s_D(\Omega )$ into $H^{s+2}_D(\Omega )$ for $-1\le s<\frac{1}{2}$. Note that, for $v_0\in
H^{s+2}_D(\Omega )$,
$h\in H^{\frac{s+1}{2}}(\mathbb T ; L^2(\partial \Omega ))$
by \eqref{h1992}. \\
{\em Step 2.} Since $S$ is an isomorphism from $H^{1}_D(\Omega )$ onto
$H^{-1}_D(\Omega )$, it remains to prove that $S$ maps $H^{s+2}_D(\Omega )$ into
$H^s_D(\Omega )$. The proof of Theorem \ref{thm3}
will thus be complete with the following result.
\begin{prop}
\label{regularite}
Let $s\in [-1,\frac{1}{2})$ and $T>0$. For any $v_0\in H^{s+2}_D(\Omega )$,
let $u=\Gamma v_0$ denote the solution of \eqref{AA2} associated with
$h=\partial v/\partial \nu$, where $v(t)=W_D(t) v_0$. Then $\Gamma $ is a bounded
operator from $H^{s+2}_D(\Omega )$ into $C([0,T];H^s_D(\Omega ))$.
\end{prop}
\noindent
{\em Proof of Proposition \ref{regularite}.}
It is well known that for any $h\in L^2(0,T;L^2(\partial \Omega ))$,
there exists a unique solution $u\in C([0,T];H^{-1}(\Omega ))$
in the transposition sense of \eqref{AA2} (see e.g. \cite{mach-2}).
The result is therefore true for $s=-1$. Let us now assume that $s\in (-1,1/2)$.
From Step 1, we know that $u$ is given by
\begin{equation}
\label{XY5}
u(t)
= -\left( \frac{2}{\pi}\right) ^n
\sum_{q\in (\mathbb N ^*)^n}
\left(
\sum_{p\in (\mathbb N ^*)^n}
v_p\frac{ e^{-i|p|^2 t} - e^{-i|q|^2 t}}{|q|^2-|p|^2}
I(g,p,q)\right)
\sin(q_1x_1)\cdots \sin (q_nx_n)
\end{equation}
where
\begin{equation}
\label{XY6}
I(g,p,q)=\int_{\partial \Omega} g(x)\frac{\partial}{\partial \nu}
(\sin (p_1x_1)\cdots \sin (p_nx_n))\frac{\partial}{\partial \nu}
(\sin (q_1x_1)\cdots \sin (q_nx_n)) d\sigma (x).
\end{equation}
Again $I(g,p,q)=\sum_{0\le l<2^n-1}I_{F_l}$, where the $F_l$'s denote the faces
of $\Omega$ and $I_{F_l}$ is given in \eqref{XY3}.
We have that
\begin{eqnarray*}
||\Gamma v_0||_{L^\infty (0,T;H^s_D (\Omega ))}
&= & ||D^{s+1}(\Gamma v_0)||_{L^\infty (0,T;H^{-1}_D(\Omega ))} \\
&\le&
||\Gamma (D^{s+1}v_0)||_{L^\infty (0,T;H^{-1}_D(\Omega ))}
+||[\Gamma ,D^{s+1}]v_0||_{L^\infty (0,T;H^{-1}_D(\Omega ))}
\cdot
\end{eqnarray*}
Since
$$
||\Gamma (D^{s+1} v_0)||_{L^\infty (0,T;H^{-1}_D(\Omega ))}
\le C ||D^{s+1}v_0||_1 \le C ||v_0||_{s+2},
$$
it remains to estimate the commutator $[\Gamma , D^{s+1}]v_0$.
Clearly
\begin{equation}
([\Gamma ,D^{s+1}]v_0)(t)
= -\left( \frac{2}{\pi}\right) ^n
\sum_{q} \left(
\sum_{p;|p|\ne |q|} v_p \frac{|p|^{s+1}-|q|^{s+1}}{|q|^2-|p|^2}
(e^{-i|p|^2 t} - e^{-i|q|^2 t})
I(g,p,q)\right)
\prod_{j=1}^n\sin(q_jx_j).
\end{equation}
The contribution in $||([\Gamma , D^{s+1}]v_0)(t)||_{-1}^2$
due to $F_0=\{ x\in\partial \Omega; \ x_n=0\}$ is estimated with
\eqref{XYZ} by
\begin{eqnarray*}
B_{F_0} &\le& C\sum_q |q|^{-2}\left( \sum_{p,|p|\ne |q|}
|v_p| \frac{|p|^s+|q|^s}{|p|+|q|}|I_{F_0}|\right) ^2\\
&\le& C\sum_q
\left(\sum_{p;|p|\ne |q|} |v_p| \frac{|p|^s + |q|^s}{|p|+|q|} p_n \prod_{j=1}^{n-1}
\langle p_j-q_j\rangle ^{-k} \right) ^2.
\end{eqnarray*}
Therefore, using the estimation of the r.h.s. of \eqref{orion1} in
\eqref{orion2}, we conclude that for $s<1/2$
$$
B_{F_0} \le C||v_0||^2_{s+\frac{3}{2}},
$$
the constant $C$ being uniform in $t\in [0,T]$. Therefore
$$
||[\Gamma, D^{s+1}]v_0||_{L^\infty (0,T;H^{-1}_D(\Omega ))}
\le C ||v_0||_{s+2}\cdot
$$
Thus, we have proved that
\begin{equation}
\label{PQR}
||u||_{L^\infty (0,T;H^s_D(\Omega ))}
\le C ||v_0||_{H^{s+2}_D(\Omega )}\cdot
\end{equation}
Since $u\in C([0,T];H^{-1}_D(\Omega ))$, we conclude that
$u\in C_w ([0,T];H^s_D(\Omega ))$. If we pick $\tilde s\in (s,1/2)$
and $\tilde v_0\in H^{\tilde s +2}_D(\Omega )$, the corresponding solution
$\tilde u$ belongs to $C_w ([0,T];H_D^{\tilde s}(\Omega ))$, hence to
$C([0,T];H^s_D(\Omega ))$, the embedding
$H^{\tilde s}_D (\Omega ) \subset H^s_D (\Omega )$ being compact.
It follows from \eqref{PQR}
combined to the density of $H^{\tilde s +2}_D(\Omega )$ in
$H^{s+2}_D(\Omega )$ that $u\in C([0,T];H^s_D(\Omega ))$ for
$v_0\in H^{s+2}_D(\Omega )$.
In particular, $u(T)\in H^s_D(\Omega )$, so that $S$ is an isomorphism from
$H^{s+2}_D(\Omega ) $ onto $H^s_D(\Omega )$. This completes the proof of
Proposition \ref{regularite} and of Theorem \ref{thm3}.
{\hfill$\quad$\qd\\}
\noindent
\subsubsection{Neumann boundary control}
We adopt the following definition.
\begin{defi}
The open set $\Gamma _0\subset\partial \Omega$ is called a {\em Neumann
control domain} if given any $u_0,\ u_1\in L^2(\Omega)$ and
any time $T>0$, one may find a control $h\in L^2(0,T;L^2(\Gamma _0))$
such that the solution $u=u(x,t)$ of
\begin{equation}
\label{B123}
\left\{
\begin{array}{ll}
iu_t+\Delta u = 0\qquad &\text{ in } \Omega\times (0,T)\\
\frac{\partial u}{\partial \nu}=1_{\Gamma _0}h(x,t)\qquad &\text{ on }
\partial\Omega \times (0,T) \\
u(0)=u_0&
\end{array}
\right.
\end{equation}
satisfies $u(T)=u_1$.
\end{defi}
The following result provides Neumann control domains
in {\em any} dimension $n\ge 2$.
\begin{prop}
\label{prop20}
Let $\Omega =(0,\pi )^n$, and
let $\Gamma _0\subset \partial \Omega$ be a side of $\Omega$.
Then $\Gamma _0$ is a Neumann control domain.
\end{prop}
\noindent
{\em Proof.} Assume e.g. that $\Gamma _0= \{ 0 \} \times (0,\pi )^{n-1}$.
By Dolecki-Russell criterion, we only have to check the
following observability inequality
\begin{equation}
||v_0||^2_{L^2(\Omega )} \le C \int_0^T\int_{\Gamma _0}|v(x,t)|^2 d\sigma dt
\end{equation}
where $v_0$ is any function in $L^2(\Omega )$ and $v=v(x,t)$ solves
\begin{equation}
\label{B124}
\left\{
\begin{array}{ll}
iv_t+\Delta v = 0\qquad &\text{ in } \Omega\times (0,T)\\
\frac{\partial v}{\partial \nu}=0\qquad &\text{ on }
\partial\Omega \times (0,T) \\
v(0)=v_0.&
\end{array}
\right.
\end{equation}
Expanding $v_0$ as
$$
v_0(x)=\sum_{k\in \mathbb N ^n} c_k\cos(k_1x_1)\cdots \cos (k_nx_n),
$$
then the corresponding solution $v(x,t)$ reads
$$
v(x,t)=\sum_{k\in \mathbb N ^n} c_ke^{-i|k|^2t}
\cos(k_1x_1)\cdots \cos (k_nx_n).
$$
It follows that
\begin{eqnarray*}
\int_0^T\int_{\Gamma _0}|v(x,t)|^2 d\sigma dt
&= &\int_0^T \int_{(0,\pi )^{n-1}} | \sum_{k\in \mathbb N ^n} c_ke^{-i|k|^2t}
\cos(k_2x_2)\cdots \cos (k_n x_n)|^2 dx_2\cdots dx_n dt \\
&\sim& \sum_{k_2,...,k_n\ge 0}
\int_0^T \left\vert \sum_{k_1\ge 0} c_ke^{-ik_1^2t}\right\vert ^2dt
\sim \sum_{k\in \mathbb N ^n} |c_k|^2 \sim ||v_0||^2_{L^2(\Omega )},
\end{eqnarray*}
where we used the orthogonality of the functions
$\cos(k_2x_2)\cdots \cos(k_n x_n)$
in $L^2(\Gamma _0)$ and Ingham's lemma.{\hfill$\quad$\qd\\}
We now aim to extend Proposition \ref{prop20} to a control result in a space
$H^s(\Omega)$, $s>0$. We define $H^s_N(\Omega )=D(A_N^{\frac{s}{2}})$, where
$A_N$ is the Neumann Laplacian (i.e. $A_N u=u-\Delta u$ with
$D(A_N)=\{ u\in H^2(\Omega),\ \partial u/\partial \nu = 0\text{ on }
\partial \Omega\} \subset L^2(\Omega )$).
A result similar to Theorem \ref{thm3} may be obtained along the same lines.
We limit ourselves to giving a weaker result with a very short proof.
\begin{thm}
\label{thm4}
Let $\Gamma _0$ be a Neumann control domain, $T=2\pi$, $s \in [0,1)$
and $u_0,u_1 \in H^s_N(\Omega )$.
Then there exists a control input
$h\in H^{\frac{s}{2}}(\mathbb T ;L^2(\partial\Omega ))$ such that the solution
$u$ of \eqref{B123} satisfies $u(T)=u_1$.
\end{thm}
\noindent{\em Proof.} Without loss of generality, we may assume that
$u_0=0$.
A direct computation shows that for any (smooth) solution $u$ of
\eqref{B123} emanating from $u_0=0$ and any (smooth) solution
$v$ of \eqref{B124}, it holds
\begin{equation}
\label{identity}
i\int_\Omega u(x,T)\overline{v(x,T)}\, dx
= - \int_0^T\!\!\!\int_{\partial \Omega} 1_{\Gamma_0}h(x,t)
\overline{v} d\sigma dt.
\end{equation}
As usual, for any $h\in L^2(0,T; L^2(\partial \Omega ))$, the solution
$u\in C([0,T];L^2(\Omega ))$ of \eqref{B123} is defined by
\begin{equation}
i(u(t) , v(t) )_{L^2(\Omega )}
= - (h, 1_{\Gamma _0}v)
_{L^2(0,t;L^2(\partial \Omega))}, \qquad\forall t\in [0,T],\ \forall v_0\in
L^2(\Omega)
\end{equation}
where $v(t)$ solves \eqref{B124}.\\
{\sc Claim 1.} If $v_0\in H^{-s}_N(\Omega )$ for some $s\in \mathbb R$, then
$v\in H^{-\frac{s}{2}}(\mathbb T ;L^2(\partial \Omega ))$. \\
Indeed, if we write
$v_0=\sum_{k\in \mathbb N ^n} c_k\cos (k_1x_1)\cdots\cos (k_n x_n)$ and
$$v(x,t)=\sum_{k\in \mathbb N ^n} c_k e^{-i|k|^2t}
\cos (k_1x_1)\cdots\cos (k_n x_n)
$$
then we have that
\begin{equation}
\label{LQR}
||v||^2_{H^{-\frac{s}{2}}(\mathbb T ,L^2(\partial \Omega ))}
\sim \sum_{k}(1+|k|^2)^{-s} |c_k|^2\sim ||v_0||^2_{H^{-s}_N(\Omega )}.
\end{equation}
We may rewrite \eqref{identity} in the form
\begin{equation}
\label{identitybis}
i\langle u(T) , v(T) \rangle _{H^s_N, H^{-s}_N}
= - \langle h, 1_{\Gamma _0} v
\rangle_{H^{\frac{s}{2}}(\mathbb T ;L^2(\partial \Omega)),
H^{-\frac{s}{2}}(\mathbb T ; L^2(\partial \Omega))}.
\end{equation}
Note that $u\in C([0,T]; H^s_N(\Omega ))$ if $0\le s<1$.
It remains to establish the following\\
{\sc Claim 2.} (Observability inequality) The following
estimate holds for the solutions of \eqref{B124}:
\begin{equation}
\label{obs}
||1_{\Gamma _0} v||^2
_{H^{-\frac{s}{2}}(\mathbb T ;L^2(\partial\Omega ))} \ge const||v_0||^2_{H^{-s}_N
(\Omega ) }
\end{equation}
If \eqref{obs} is not true, one can construct a sequence $\{ v_j\}$ such that
\begin{equation}
\label{obs2}
j||1_{\Gamma _0} v_j||^2
_{H^{-\frac{s}{2}}(\mathbb T ;L^2(\partial\Omega ))} <
||v_j(0)||^2_{H^{-s}_N(\Omega ) }=1.
\end{equation}
Let $w_j=(1-\partial _t^2)_p^{-\frac{s}{4}} v_j$, where for
any $\sigma\in \mathbb R$
$$
(1-\partial _t^2)_p^{\sigma} \sum_{l\in \mathbb Z}c_l e^{ilt} =
\sum_{l\in \mathbb Z}(1+|l|^2)^\sigma c_l e^{ilt}.
$$
Then $w_j$ solves
\eqref{B124} with $w_j(0)$ substituted to $v_0$, and from \eqref{obs2} we obtain
\begin{equation}
\label{obs3}
1_{\Gamma _0} w_j \to 0 \ \text{ in } \
L^2(\mathbb T ;L^2(\partial \Omega )).
\end{equation}
As $\Gamma _0$ is a Neumann control domain, we infer that
$w_j(0)\to 0$ in $L^2(\Omega)$, hence
$$
w_j\to 0 \text{ in } L^2(\mathbb T ; L^2 (\partial \Omega )).
$$
This gives
$$
v_j\to 0 \text{ in } H^{-\frac{s}{2}}(\mathbb T ;L^2(\partial \Omega)).
$$
Using \eqref{LQR}, we infer that
$v_j(0) \to 0$ in $H^{-s}_N(\Omega )$, which contradicts
\eqref{obs2}.
This completes the proof of Theorem \ref{thm4}.
{\hfill$\quad$\qd\\}
\section{Nonlinear systems}
\subsection{Internal control}
In this section we consider the following
nonlinear control system
\begin{equation}
\left \{ \begin{array}{l} iu_t + \Delta u + N(u)
=i Gh = ia(x)h(x,t), \ x\in \T ^n, \ t>0,
\\ \\ u(x,0) =\phi (x), \end{array} \right.
\label{4.1}
\end{equation}
where $a\in C^\infty (\mathbb T ^n)$, and the nonlinearity $N(u)$ reads
\begin{equation}
\label{Nu}
N(u)=\lambda u^{\alpha _1} \overline{u}^{\alpha _2},\qquad
\alpha _1 + \alpha _2 =: \alpha + 1\ge 2,
\end{equation}
with $\lambda \in \mathbb R$, and $\alpha , \alpha _1, \alpha _2\in \mathbb N$.
Note that for any $\alpha =2\beta\in 2\mathbb N ^*$,
$|u|^{\alpha} u = u^{\beta +1} \overline{u}^{\beta}$.
We introduce the number
\begin{equation}
\label{salphan}
s_{\alpha ,n}=
\left\{
\begin{array}{ll}
\frac{n}{2}- 1 \quad &\text{if} \ \alpha =1,\\
\frac{n}{2} -\frac{3}{4} -\frac{1}{4(n-1)}\quad &\text{if} \ \alpha =2,\\
\frac{n}{2} -\frac{2}{\alpha} &\text{if} \ \alpha \ge 3.
\end{array}
\right.
\end{equation}
Thus $s_{\alpha,n}=s_c:=\frac{n}{2}-\frac{2}{\alpha}$
(the critical Sobolev exponent obtained by scaling in NLS) for $\alpha \ge 3$,
while $s_{\alpha, n}>s_c$ for $\alpha = 1,2$ (except for $n=\alpha =2$ where
$s_{2,2}=s_c=0$).
By Corollary \ref{cor100} (see below), the system (\ref{4.1}) is locally
well-posed in the space $H^s (\T ^n)$ for $\alpha \ge 1$ and
$s>s_{\alpha , n}$ with
$\phi \in H^s (\T ^n)$ and $h\in L^2_{loc} (\mathbb R , H^s (\T ^n))$.
Our main concern is its exact controllability in the space
$H^s (\T ^n)$.
\begin{thm}
\label{thm10}
For given $n\ge 2$, $\alpha _1, \alpha _2\in \mathbb N$ with
$\alpha _1+\alpha _2 =:\alpha +1\ge 2$, and $a\not\equiv 0$, the system
(\ref{4.1}) is locally exactly controllable in the space $H^s
(\T ^n )$ for any $s> s_{\alpha , n}$. More precisely, for any given
$T>0$, there exists a number $\delta >0$ depending on $\alpha, \ n, \ T $
and $\lambda $ such that if $\phi , \ \psi \in H^s (\T ^n)$
satisfy
\[ \| \phi \|_s\leq \delta ,
\qquad \| \psi \|_s \leq
\delta, \]
then one can choose a control input $h\in L^2 (0,T; H^s
(\T ^n))$ such that the system (\ref{4.1}) admits a solution
$u\in C([0,T]; H^s (\T ^n ))$ satisfying
\[ u(x,0) = \phi (x), \qquad u(x,T) = \psi(x) .\]
\end{thm}
The system (\ref{4.1}) can be rewritten in its equivalent integral form
\begin{equation}
u(t) = W (t) \phi + i \int ^t_0 W (t-\tau )(N(u)(\tau ))
d\tau + \int ^t_0 W(t-\tau )[G h]
(\tau ) d\tau .\label{y-2}
\end{equation}
To prove Theorem \ref{thm10}, a smoothing property is needed for the
operator from $f$ to $u$, where
\[ u(t) = \int ^t_0 W(t-\tau ) f(\tau) d\tau .\]
This needed smoothing property was provided in Bourgain's work
\cite{bourgain-1,bourgain-2} where he dealt with the Cauchy problem
for the periodic Schr\"odinger equation.
For given $s,b\in \mathbb R$, the Bourgain space $X_{s,b}$ is the space of functions
$u:\mathbb T ^n\times \mathbb R\to \mathbb C$ for which the norm
$$||u||_{X_{s,b}}=||W(-t)u(.,t)||_{H^b_t(H^s_x)}$$
is finite. Decomposing $u$ as
$$
u(x,t)=\sum_{k \in \mathbb Z ^n}
\int_\mathbb R {\hat u}(k,\tau )e^{i(k\cdot x + \tau t)} d\tau
$$
we have that
$$
||u||^2_{X_{s,b}}=\sum_{k\in \mathbb Z ^n}\int _\mathbb R
\langle \tau + |k|^2 \rangle ^{2b}
\langle k\rangle ^{2s} |\hat{u}(k , \tau )|^2 d\tau
$$
where $\langle y\rangle :=(1+|y|^2)^{\frac12}$.
For given $T>0$, $X_{s,b}^T$ is the restriction norm space
$$
X_{s,b}^T=\{ u_{|\mathbb T ^n\times (0,T)};\ u\in X_{s,b} \}
$$
with the restriction norm
$$
||u||_{X_{s,b}^T} = \inf \{ ||\tilde u||_{X_{s,b}};
\tilde u \in X_{s,b}, \ \tilde u_{\vert \T ^n\times (0,T)}
=u\}.
$$
Before we proceed to show the exact controllability results, we
present the two following technical lemmas (see e.g. \cite{tao})
which play important roles in the proof of Theorem \ref{thm10}.
\begin{lem}
\label{lem10}
For given $T>0$ and $s,b\in \mathbb R$, there exists a
constant $C>0$ such that
\[ \| W(t) \phi \|_{X_{s,b}^T }\leq C \| \phi \| _s \]
for any $\phi \in H^s(\mathbb T ^n).$
\end{lem}
\begin{lem}
\label{lem11}
For given $T>0$, $b>1/2$, and $s\in \mathbb R$, there exists a
constant $C>0$ such that
\[ \left \| \int ^t_0 W(t-\tau) f(\tau) d \tau \right \|
_{X_{s,b}^T } \leq C \| f\| _{X_{s,b-1}^T} \]
for any $f\in X_{s,b-1}^T$.
\end{lem}
The following multilinear estimate is crucial when applying the
contraction mapping theorem.
\begin{prop}
\label{prop12}
Let $n\geq 2$, $\alpha \in \mathbb N ^*$ and $s > s_{\alpha,n}$.
Then there exist some numbers $b\in (0,\frac{1}{2})$ and $C>0$ such that
\begin{equation}
\label{multilinear}
\| \prod_{i=1}^{\alpha +1} \tilde{u}_i \| _{X_{s,-b}}\leq
C \prod_{i=1}^{\alpha +1} \| u_i\| _{X_{s,b}}\quad
\forall u_1 ,..., u_{\alpha +1} \in X_{s,b},
\end{equation}
where $\tilde{u}_i$ denotes $u_i$ or $\overline{u_i}$.
\end{prop}
\begin{cor}
\label{cor100}
Let $n\ge 2$, $\alpha \in \mathbb N ^*$, and $s>s_{\alpha ,n}$. Pick
$u_0\in H^s(\mathbb T ^n)$ and $h\in X_{s,0}=L^2(\mathbb R; H^s(\mathbb T ^n))$. Then
there exist two numbers $b>\frac{1}{2}$ and
$T=T(||u_0||_{H^s(\mathbb T ^n)},||h||_{X_{s,0}})$ so that the initial-value problem
\eqref{4.1} admits a unique solution $u\in X_{s,b}^T$.
\end{cor}
\begin{rem}
Proposition \ref{prop12}, which is proved in Appendix for the sake
of completeness, is essentially due to Bourgain.
It was proved in
\cite{bourgain-2} when $\alpha =n=2$, and in \cite{bourgain-1}
in Besov-type spaces when $s>s_b$, where
\begin{equation}
\label{sbourgain}
s_b=
\left\{
\begin{array}{ll}
s_c \quad &\text{if} \ n=2,\\
\max (s_c, \frac{3}{4}) \quad &\text{if} \ n=3,\\
\max (s_c,\frac{3n}{n+4}) \quad &\text{if}\ n\ge 4.
\end{array}
\right.
\end{equation}
Notice that $s_b>s_c$ only for $(\alpha ,n )\in \{ (2,3),(2,4),(2,5),(3,4)\}$.
The corresponding values of $s_b,s_c$ and $s_{\alpha, n}$ are reported in
Table \ref{table1}.
\begin{table}[htbp]
\begin{center}
\begin{tabular}{|l|c|c|c|c|}
\hline
&&&&\\
$(\alpha ,n)$ & $(2,3)$ & $(2,4)$ & $(2,5)$ & $(3,4)$ \\
&&&&\\
\hline
&&&&\\
$s_b$ & $\frac{3}{4}$ & $\frac{3}{2}$ & $\frac{5}{3}$ & $\frac{3}{2}$ \\
&&&&\\
\hline
&&&&\\
$s_{\alpha,n}$ & $\frac{5}{8}$& $\frac{7}{6}$& $\frac{27}{16}$ &
$\frac{4}{3}$\\
&&&&\\
\hline
&&&&\\
$s_c$ & $\frac{1}{2}$ & 1 & $\frac{3}{2}$ & $\frac{4}{3}$ \\
&&&&\\
\hline
\end{tabular}
\end{center}
\caption{$s_b$, $s_{\alpha, n}$ and $s_c$ for
$(\alpha, n)\in \{(2,3),(2,4),(2,5),(3,4)\}$}
\label{table1}
\end{table}
On the other hand, $s_b=s_c < s_{\alpha, n}$ for $\alpha =2$ and $n\ge 6$.
Sharp results for the local well-posedness of NLS on $\mathbb T ^n$ are also given
in \cite{KPV96} for $\alpha =n=1$, and
in \cite{grunrock01} for $(\alpha _1,\alpha _2)=(0,2)$ and
$2\le n\le 4$.
\end{rem}
It follows at once from Proposition \ref{prop12} that for any $T>0$,
$s>s_{\alpha, n}$, and some
$b>1/2$, $b'>b-1$ we have
\[
\| N(v) - N(w) \|_{X^T_{s,b'}}
\le C (||v||^\alpha _{X^T_{s,b}} +||w||^\alpha _{X^T_{s,b}})
||v-w||_{X^T_{s,b}}\quad
\forall v,w\in X^T_{s,b}.
\]
We are now in a position to give a proof of Theorem \ref{thm10}.
\medskip
\noindent
{\bf Proof of Theorem \ref{thm10}:} Set
\[ \omega (v, T)= i\int _0^T W (T-\tau ) N(v) (\tau)
d\tau .\]
By Theorem \ref{thm1}, if we choose
\[ h= \Phi (\phi, \psi -\omega (v, T)),\]
then
\begin{eqnarray*}
& & \qquad \quad W (t) \phi + \int ^t_0 W
(t-\tau )\left (iN(v)+ G \Phi (\phi, \psi -\omega (v,T)\right )
(\tau ) d\tau\\ \\
& = &
\left\{
\begin{array}{ll}
\phi (x) \ \hbox{in} \ \mathbb T ^n \ & \hbox{when} \ t= 0; \\ \\
\psi (x) -\omega (v, T)+\omega (v,T)=\psi (x) \ \hbox{in} \ \mathbb T ^n ,
\ & \hbox{when} \ t=T
.\end{array} \right. \end{eqnarray*}
It suggests us to consider the nonlinear map:
\[ \Gamma (v) = W (t) \phi + i \int ^t_0 W
(t-\tau )\left (iN(v)+ G \Phi (\phi, \psi -\omega (v,T)\right )(\tau )
d\tau . \]
The proof would be complete if we can show that this map
$\Gamma $ has a fixed point in the space $X_{s,b}^T$, with
$b\in (\frac{1}{2},1)$.
To this end, note that by using Lemma \ref{lem10}, Lemma \ref{lem11} and
Proposition \ref{prop12}, there
exist a number $b\in (\frac{1}{2},1)$ and some
constants $C_j$, $j=1,2,3$ such that
\[
\| \Gamma (v)\| _{X_{s,b}^T} \leq C_ 1 \left (\| \phi \| _s
+\| \psi \| _s
+\| \omega (v, T)\| _s \right )
+C_2 \| v\|_{X_{s,b}^T}^{\alpha +1}
\]
for any $v\in X_{s,b}^T$ and
\[
\| \Gamma (v_1)-\Gamma (v_2)\| _{X_{s,b}^T}
\leq C_1 \| \omega (v_1, T)-\omega (v_2, T)\| _s
+ C_3 \left ( \|v_1\|
_{X_{s,b}^T}^{\alpha} + \|v_2\| _{X_{s,b}^T}^{\alpha}\right ) \|
v_1-v_2\|_{X_{s,b}^T} \] for any $v_1,v_2 \in X_{s,b}^T$.
Note that there exists a constant $C_4>0$ such that
\begin{eqnarray*}
\|\omega (v,T)\|_s
&\leq &\| \int ^t_0 W(t-\tau )
N(v) (\tau ) d\tau \|_{C([0,T]; H^s (\T ^n ))} \\ \\
& \leq & const \| \int ^t_0 W(t-\tau )N(v)(\tau ) d\tau
\|_{X_{s,b}^T} \\ \\
&\leq & C_4 \|v\|^{\alpha +1} _{X_{s,b}^T} .
\end{eqnarray*}
Similarly
\[ \|\omega (v_1,T)-\omega (v_2,T)\|_s \leq C_5
\left ( \|v_1\| _{X_{s,b}^T}^{\alpha }
+ \|v_2\| _{X_{s,b}^T}^{\alpha}\right ) \|v_1-v_2\|_{X_{s,b}^T} .\]
As a result, by increasing the constants $C_2$ and $C_3$, we obtain
\[ \| \Gamma (v)\| _{X_{s,b}^T}
\leq C_ 1 (\| \phi \|_s + \| \psi \|_s )
+C_2 \| v\|_{X_{s,b}^T}^{\alpha +1} \]
for any $v\in X_{s,b}^T $ and
\[ \| \Gamma (v_1)-\Gamma (v_2)\| _{X_{s.b}^T}
\leq C_3 \left ( \|v_1\| _{X_{s,b}^T}^{\alpha }
+ \|v_2\| _{X_{s,b}^T}^{\alpha} \right )
\| v_1-v_2\|_{X_{s,b}^T} \]
for any $v_1 , v_2 \in X_{s,b}^T$.
Pick $\delta >0$, $\phi , \psi \in H^s (\T ^n )$
with $\|\phi \|_s + \|\psi \|_s \le \delta $, and set
$M=2C_1 \delta$. If $\|v\|_{X_{s,b}^T}\leq M$ and
\[ \mbox{ $\|v_j\|_{X_{s,b}^T} \leq M, \ j=1,2$,}\]
then
\begin{eqnarray*}
\| \Gamma (v)\|_{X_{s,b}^T}
&\le & C_1 \delta + C_2 M^{\alpha + 1}\\
&\le & 2C_1 \delta = M
\end{eqnarray*}
as long as
\[ C_2 M^{\alpha } \le \frac{1}{2}\cdot \]
Choose $\delta >0$ so that $M=2C_1\delta$ fulfills
\[ C_2 M^{\alpha } \le \frac12 \
\mbox{ and }\ C_3M^\alpha \le \frac{1}{4}, \]
and let $B_{M }$ be the ball in the space $X_{s,b}^T$
centered at the origin of radius $M$. For given
$\phi,\psi \in H^s (\T ^n )$ with
$ \| \phi \|_s + \| \psi \|_s \leq \delta$, we have
\[ \| \Gamma (v)\|_{X_{s,b}^T} \leq M \]
for any $v\in B_M$ and
\[
\| \Gamma (v_1)-\Gamma (v_2)\|_{X_{s,b}^T}
\leq \frac12 \| v_1 -v_2\| _{X_{s,b}^T} \]
for any $v_1, v_2 \in B_M$. That is to say, $\Gamma
$ is a contraction in the ball $B_M$. The proof is complete.{\hfill$\quad$\qd\\}
Let us now consider the Schr\"odinger equation posed on a cube
$\Omega =(0,\pi )^n$
\begin{equation}
iu_t + \Delta u + N(u) = ia(x)h(x,t),
\qquad x\in \Omega, \ t\in (0,T) \label{7.4}
\end{equation}
with either the homogeneous Dirichlet boundary conditions
\begin{equation}
u(x,t) =0
\qquad (x,t) \in \partial \Omega \times (0,T) \label{7.5}
\end{equation}
or the homogeneous Neumann boundary conditions
\begin{equation}
\frac{\partial u}{\partial \nu} (x,t) = 0
\qquad (x,t) \in \partial \Omega \times (0,T). \label{7.6}
\end{equation}
The nonlinearity $N(u)$ is still as in \eqref{Nu}.
It is remarkable that internal control results with Dirichlet
(resp. Neumann) homogeneous boundary conditions can be deduced from those
already proved for periodic boundary conditions.
\begin{cor}
\label{dirichlethom}
For given $n\ge 2$, $\alpha _1, \alpha _2\in \mathbb N$ with
$\alpha _1+\alpha _2 =:\alpha +1\ge 2$ and $\alpha$ even,
and $a\not\equiv 0$, the system
(\ref{7.4})-(\ref{7.5}) is locally exactly controllable in the space $H^s_D
(\Omega )$ for any $s> s_{\alpha , n}$. More precisely, for any given
$T>0$, there exists a number $\delta >0$ depending on $\alpha, \ n, \ T $
and $\lambda $ such that if $\phi , \ \psi \in H^s_D (\Omega )$
satisfy
\[ \| \phi \|_{H^s_D(\Omega )}\leq \delta ,
\qquad \| \psi \|_{H^s_D(\Omega )} \leq
\delta, \]
then one can choose a control input
$h\in L^2 (0,T; H^s_D (\Omega ))$ such that the system
(\ref{7.4})-(\ref{7.5}) admits a solution
$u\in C([0,T]; H^s_D (\Omega ))$ satisfying
\[ u(x,0) = \phi (x), \qquad u(x,T) = \psi(x) .\]
\end{cor}
\begin{cor}
\label{neumannhom}
For given $n\ge 2$, $\alpha _1, \alpha _2\in \mathbb N$ with
$\alpha _1+\alpha _2 =:\alpha +1\ge 2$
and $a\not\equiv 0$, the system
(\ref{7.4})-(\ref{7.6}) is locally exactly controllable in the space $H^s_N
(\Omega )$ for any $s> s_{\alpha , n}$. More precisely, for any given
$T>0$, there exists a number $\delta >0$ depending on $\alpha, \ n, \ T $
and $\lambda $ such that if $\phi , \ \psi \in H^s_N (\Omega )$
satisfy
\[ \| \phi \|_{H^s_N(\Omega )}\leq \delta ,
\qquad \| \psi \|_{H^s_N(\Omega )} \leq
\delta, \]
then one can choose a control input
$h\in L^2 (0,T; H^s_N (\Omega ))$ such that the system
(\ref{7.4})-(\ref{7.6}) admits a solution $u\in C([0,T]; H^s_N (\Omega ))$
satisfying
\[ u(x,0) = \phi (x), \qquad u(x,T) = \psi(x) .\]
\end{cor}
We shall say that a function from $(-\pi,\pi)^n$ to $\mathbb C$ is {\em odd}
(resp. {\em even}), if it is odd with respect to each coordinate $x_i$,
$1\le i\le n$.
The proof relies on the basic, but crucial observation that the functions in
$H^s_D(\Omega )$ (resp. $H^s_N(\Omega )$) coincide with the restrictions
to $\Omega$ of the functions in $H^s(\mathbb T ^n)$ which are odd (resp. even).
The issue is therefore reduced to an extension of Theorem \ref{thm10}
in the framework of odd (resp. even) functions in $H^s(\mathbb T ^n)$.
Extending the function $a$ in \eqref{7.4}
to $\mathbb T ^n$ as an even function, we notice that
the control input $h$ in Theorem \ref{thm1} can be chosen odd (resp. even)
if the functions $\phi,\psi$ are odd (resp. even). Indeed, the observability
inequality holds as well in the subspaces
\begin{eqnarray*}
H^s_{odd}(\mathbb T ^n) &=& \{ u\in H^s_p (\mathbb T ^n);\
u(x_1,...,x_{i-1},-x_i,x_{i+1},...,x_n)=-u(x)\quad\forall x\in \mathbb T ^n,
\ \forall i\},\\
H^s_{even}(\mathbb T ^n) &=& \{ u\in H^s_p (\mathbb T ^n);\
u(x_1,...,x_{i-1},-x_i,x_{i+1},...,x_n)=u(x)\quad\forall x\in \mathbb T ^n,
\ \forall i\}
\end{eqnarray*}
of $H^s(\mathbb T ^n)$ for $s\le 0$. On the other hand, since $u$ and $N(u)$ are
simultaneously odd (resp. even), we see that the
contraction mapping theorem can
be applied in a space of odd (resp. even) trajectories to derive
the result in Corollary \ref{dirichlethom}
(resp. \ref{neumannhom}). Full details are provided in
\cite{RZ2007b} for $n=1$.
\subsection{Boundary control}
In this section we consider the Schr\"odinger equation posed on a rectangle
$\Omega =(0,l_1) \times \cdots \times (0,l_n)$
\begin{equation}
iu_t + \Delta u + N(u) = 0,
\qquad x\in \Omega, \ t\in (0,T) \label{4.4}
\end{equation}
with either the Dirichlet boundary conditions
\begin{equation}
u(x,t) =1_{\Gamma _0} h(x,t)
\qquad (x,t) \in \partial \Omega \times (0,T) \label{4.5}
\end{equation}
or the Neumann boundary conditions
\begin{equation}
\frac{\partial u}{\partial \nu} (x,t) = 1_{\Gamma _0} h(x,t)
\qquad (x,t) \in \partial \Omega \times (0,T). \label{4.6}
\end{equation}
When we shall consider a smooth Dirichlet controller $g$, then the boundary condition \eqref{4.5} will be replaced by
\begin{equation}
\label{69bis}
u(x,t) = g(x) h(x,t)
\qquad (x,t) \in \partial \Omega \times (0,T).
\end{equation}
$N(u)$ still stands for the nonlinear term in NLS. We first give a result
(with a small control region) providing precise
informations on the smoothness of the control input
and of the trajectories when $N(u)$ is {\em weakly} nonlinear.
To simplify the exposition, we assume here that
$$\Omega=(0,\pi )^n .$$
We denote by $u=W_D(t)u_0$ the solution of \eqref{B0} for $h=0$.
For given $s,b\in \mathbb R$, $X_{s,b}(\Omega )$ denotes the Bourgain
space of functions
$u:\Omega \times \mathbb R\to \mathbb C$ for which the norm
$$||u||_{X_{s,b}(\Omega ) }= c ||W_D(-t) u(.,t)||_{H^b(\mathbb R ;
H^s_D (\Omega ))}$$
is finite. Decomposing $u$ as
$$
u(x,t)=\sum_{k \in (\mathbb N ^*)^n}
\int_\mathbb R {\hat u}(k,\tau )e^{i\tau t}\sin (k_1x_1)\cdots \sin (k_nx_n) d\tau
$$
we can choose the constant $c$ so that
$$
||u||^2_{X_{s,b}(\Omega )}=\sum_{k\in (\mathbb N ^*) ^n}\int _\mathbb R
\langle \tau + |k|^2 \rangle ^{2b}
\langle k\rangle ^{2s} |\hat{u}(k , \tau )|^2 d\tau <\infty .
$$
The restriction norm space $X_{s,b}^T(\Omega )$ is defined in the usual way
(see above the definition of $X^T_{s,b}$).
For $u\in H^s_D(\Omega )$ given, we denote by $\tilde u$ its odd extension
to $\T ^n =(-\pi , \pi )^n$; i.e., ${\tilde u}_{|(0,\pi )^n}=u$, and
$\tilde u$ is odd with
respect to each coordinate $x_i$. Note that $\tilde u\in H^s(\mathbb T ^n)$
and $||\tilde u||_s\sim ||u||_{H^s_D(\Omega )}$.
Defining $\tilde u(.,t)$ from $u(.,t)$ in a similar way, we observe that
$$
||\tilde u||_{X_{s,b}^T}\sim ||u||_{X^T_{s,b}(\Omega )}.
$$
It is then clear that Lemmas \ref{lem10} and \ref{lem11} hold true with
$W_D(t)$, $H^s_D(\Omega )$ and $X^T_{s,b}(\Omega )$ substituted to
$W(t)$, $H^s(\T ^n )$ and $X^T_{s,b}$, respectively.
We shall assume that the nonlinear term $N(u)$ satisfies the following
multilinear estimate
\begin{equation}
\label{multilinearnu}
||N(u)-N(v)||_{X_{s,b'}(\Omega )} \le
c(u,v)\, ||u-v||_{X_{s,b}(\Omega )}
\end{equation}
where $s\in \mathbb R$, $-1/2 <b'<b\le b'+1$ and $c(u,v)\to 0$ as $u\to 0,\ v\to 0$ in $X_{s,b}(\Omega )$.
Theorem \ref{thm3} can be extended to a semilinear context as follows.
\begin{thm}
\label{thm11}
Let $g$ be a smooth Dirichlet controller,
and let the nonlinearity $N(u)$ satisfy \eqref{Nu} and
\eqref{multilinearnu} with
$s\in [-1,\frac{1}{2})$, $b>0$ and $s+2b<\frac{1}{2}$.
Pick any $T>0$. Then there exists $\delta >0$ such that for any
$u_0,u_T \in H^s_D(\Omega )$ satisfying
$$
||u_0 ||_{H_D^s(\Omega )}\le \delta,\quad
||u_T ||_{H_D^s(\Omega )}\le \delta
$$
one may find a control input
$h\in H^{\frac{s+1}{2}}(\mathbb T ; L^2 (\partial \Omega ))$ and a solution
$u\in C([0,T];H^s_D(\Omega )) \cap X^T_{s,b}$ of \eqref{4.4} and \eqref{69bis} such that $u(0)=u_0$ and $u(T)=u_T$.
\end{thm}
\noindent
{\em Proof. }
For $u_T\in H^{s}_D(\Omega )$, let $h$ be the control
given by HUM which steers \eqref{AA2} from $0$ to $u_T$, namely
$h=\partial v/\partial \nu$ with $v=W_D(t)v_0$ and $v_0=S^{-1}u_T\in H^{s+2}_D(\Omega )$ (cf. Theorem \ref{thm3}). Recall that
$h\in H^{\frac{s+1}{2}}(\mathbb T; L^2(\partial \Omega ))$ by \eqref{h1992}.
We set $u=\Lambda u_T=\Gamma S^{-1}u_T$.
The regularity of $u$ is depicted in the following proposition.
\begin{prop}
\label{prop15} Assume that $-1\le s<1/2$ and $s+2b<1/2$. Then $\Lambda$ maps continuously $H^s_D(\Omega )$ into $C([0,T];H^s_D(\Omega ))\cap
X_{s,b}^T(\Omega )$.
\end{prop}
\noindent
{\em Proof of Proposition \ref{prop15}.} It follows from Proposition \ref{regularite} and Theorem \ref{thm3} that $\Lambda$ maps
continuously $H^s_D(\Omega )$ into $C([0,T];H^s_D(\Omega ))$. Let us turn our attention to the Bourgain space $X_{s,b}^T(\Omega )$. \\
{\em Step 1.} We prove several claims used thereafter.\\
{\sc Claim 3.} For any $\gamma >1/2$, it holds
$$
\sup_{\lambda \in \mathbb R} \sum_{k\in \mathbb Z} \langle \lambda ^2 -k^2\rangle ^{-\gamma}
<\infty.
$$
In what follows, $C$ denotes a constant independent of $\lambda$ and $k$ which
may vary from line to line. Pick $\lambda \in \mathbb R ^+$. For $0\le \lambda \le 1$
$$
\langle \lambda ^2 - k^2 \rangle ^{-\gamma}
\le \langle k^2\rangle ^{-\gamma} + \langle 1-k^2\rangle ^{-\gamma}
$$
and the result is then obvious. For $\lambda > 1$, we have
\begin{eqnarray*}
\sum_{k\in \mathbb Z}\langle \lambda ^2 -k^2 \rangle ^{-\gamma}
&\le & C\left(
\int_0^{\lambda -1} |\lambda ^2 -x^2|^{-\gamma }dx +
\int_{\lambda +1}^\infty |x^2-\lambda ^2|^{-\gamma} dx +1 \right) \\
&=& C \lambda ^{1-2\gamma} \left( \int_0^{1-\lambda ^{-1}} |1-y^2|^{-\gamma}
dy + \int_{1+\lambda ^{-1}}^{+\infty} |y^2-1|^{-\gamma} dy +1\right) \\
&\le& C \lambda ^{1-2\gamma} \left( \int_0^{1-\lambda ^{-1}} |1-y|^{-\gamma}
dy + \int_{1+\lambda ^{-1}}^{2} |y-1|^{-\gamma} dy +1\right) \\
&\le &
\left\{
\begin{array}{ll}
C\lambda ^{1-2\gamma}(\lambda ^{-1+\gamma} +1) &\text{ if } \gamma \ne 1;\\
C\lambda ^{-1}(\ln \lambda + 1) &\text{ if } \gamma =1
\end{array}
\right.
\end{eqnarray*}
and the claim follows. \\
{\sc Claim 4.} If $s\ge -1$, $0<\delta <1$, $s+2\delta <1/2$, and
$k>1+2(s+1)$, then for some constant $C>0$
$$
S(p) := \sum_{q;|q|\ne |p|}
\frac{q_n^{2s+2}}{||q|^2-|p|^2|^{2(1-\delta )}}
\prod_{j=1}^{n-1}\langle p_j-q_j\rangle ^{-k} \le C \langle p\rangle ^{2s+2}.
$$
Write $S(p) = S^1(p) + S^2(p)$, where the sum $S^1(p)$ is restricted to
the $q=(q',q_n)$ with $|q'|\ge |p|$ and $|q|\ne |p|$. Noticing that
$|q|^2-|p|^2 = q_n^2 +|q'|^2 - |p|^2 \ge q_n^2$ for such $q$, we obtain that
$$
S^1(p) \le \sum_{q_n} q_n^{2s+4\delta -2} \sum_{q'}\prod _{j=1}^{n-1}
\langle p_j-q_j \rangle ^{-k} \le C\le C\langle p\rangle ^{2s+2}
$$
To bound $S^2(p)$, we fix any $q'\in (\mathbb N ^*)^{n-1}$ with
$|q'|<|p|$ and set
$$
\lambda = \sqrt{|p|^2-|q'|^2} \ge 1.
$$
We have that
\begin{eqnarray*}
\sum_{q_n;|q_n^2-\lambda ^2|\ge 1}
\frac{q_n^{2s+2}}{|q_n^2-\lambda ^2|^{2(1-\delta )}}
&\le& C\left(
\int_{|x^2-\lambda ^2|\ge 1}
\frac{x^{2s+2}}{|x^2-\lambda ^2|^{2(1-\delta )}} dx + \lambda ^{2s+2}
\right) \\
&\le& C\left( \lambda ^{2s+4\delta -1}
\int_{|y^2-1|\ge \lambda ^{-2}}
\frac{y^{2s+2}}{|y^2-1|^{2(1-\delta)}}dy + \lambda ^{2s+2}
\right) \\
&\le& C(\lambda ^{2s+4\delta -1}\cdot \lambda ^{2-4\delta}\cdot \ln \lambda +
\lambda ^{2s+2})\\
&\le& C \big( p_n^{2s+2} + \sum_{j=1}^{n-1}\langle p_j^2 -q_j^2
\rangle ^{s+1}\big).
\end{eqnarray*}
It follows that
\begin{eqnarray*}
S^2(p) &\le &
C\sum_{q'}(p_n^{2s+2}
+\sum_{j=1}^{n-1}
\langle p_j^2-q_j^2 \rangle ^{s+1}) \prod_{l=1}^{n-1}
\langle p_l-q_l\rangle ^{-k} \\
&\le& C \big( p_n^{2s+2} + \sum_{j=1}^{n-1}\sum_{q_j\ge 1}
\langle p_j^2-q_j^2\rangle ^{s+1} \langle p_j-q_j\rangle ^{-k} \big) \\
&\le& C\big( p_n^{2s+2} +\sum_{j=1}^{n-1}\sum_{q_j\ge 1}
\langle p_j+q_j \rangle ^{s+1} \langle q_j-p_j \rangle ^{-(k-s-1)} \big).
\end{eqnarray*}
To complete the proof of Claim 4, we need the following\\
{\sc Claim 5.} Let $\sigma \ge 0$ and $k>\sigma + 1$. Then there
exists a constant $C>0$ such that
$$
\sum_{m\ge 1} \langle m + n \rangle ^\sigma \langle m - n \rangle ^{-k}
\le C n^\sigma \qquad \forall n\ge 1.
$$
Split the sum into $\Sigma _1 + \Sigma _2$ where
$\Sigma _1 = \sum_{1\le m\le 3n} \langle m+n \rangle ^\sigma
\langle m-n\rangle ^{-k}$. Note that
$$
\Sigma _1 \le \langle 4n \rangle ^\sigma \sum_{l\in \mathbb Z}
\langle l\rangle ^{-k} \le C \langle n\rangle ^\sigma
$$
since $k>1$. On the other hand, noticing that
$m-n>(m+n)/2$ for $m>3n$, we have that
$$
\Sigma_2 \le \sum_{m>3n} \langle 2(m-n)\rangle ^\sigma
\langle m-n\rangle ^{-k}
\le C\sum_{m>3n}\langle m-n\rangle ^{-(k-\sigma)} \le C.
$$
Claim 5 is proved.
Pick $k>1 + 2(s+1)\ge 1$. It follows from Claim 5 that
$$
\sum_{q_j} \langle p_j+q_j\rangle ^{s+1}\langle p_j-q_j\rangle ^{-(k-s-1)}
\le C p_j^{s+1}.
$$
Since $s+1\ge 0$ and $p_j\ge 1$, we conclude that
$$
S(p)\le C (p_n^{2s+2} + \langle p'\rangle ^{s+1} )
\le C \langle p\rangle ^{2s+2}.
$$
This completes the proof of Claim 4. \\
{\em Step 2.} Assume that $s<0$ and $s+2b<1/2$, and pick
any $u_T\in H^s_D(\Omega )$ and any $\eta \in C_0^\infty (\mathbb R )$ with
$\eta (t)=1$ for $0\le t \le T$. Let $v_0=S^{-1}u_T\in H^{s+2}_D(\Omega )$ be decomposed as in \eqref{XY1}. Let us prove that $u=\Lambda u_T\in X_{s,b}^T$.
It is sufficient to prove that
$$
||\eta (t) u||_{X_{s,b}} \le C ||v_0||_{H^{s+2}_D(\Omega )}.
$$
Recall that $u$ is given by \eqref{XY5}-\eqref{XY6}, and that $u(t)$ may be defined
this way for all $t\in \mathbb R$ . Again, we can limit
ourselves to proving that $u_{F_0}\in X_{s,b}^T$, where $u_{F_0}$ is the contribution due to $F_0=\{ x\in \partial \Omega; \ x_n=0 \}$ in $u$.
$u_{F_0}$ is decomposed as
$$
u_{F_0} =\sum_{q\in (\mathbb N ^*)^n} u_q(t) \sin (q_1x_1)\cdots \sin (q_nx_n)
$$
where
$$u_q(t)=-\left( \frac{2}{\pi} \right) ^n
\sum_{p\in (\mathbb N ^*)^n}
v_p\frac{e^{-i|p|^2t} - e^{-i|q|^2t} }{|q|^2-|p|^2}I_{F_0}$$
with the convention \eqref{XY4}.
$\hat .$ denoting time Fourier transform, an application of the elementary
property
$$\widehat{e^{irt}\eta (t)}(\tau )=\hat \eta (\tau -r)$$
yields
$$
\widehat{\eta u_q}(\tau ) = -\left( \frac{2}{\pi}\right) ^n
\left( \sum_{p;|p|\ne |q|}
v_p \frac{\hat\eta (\tau +|p|^2) - \hat\eta (\tau + |q|^2)}{|q|^2-|p|^2} I_{F_0}
+\sum_{p;|p|=|q|}
iv_p \widehat{t\eta (t) }(\tau + |q|^2)I_{F_0} \right) .
$$
For a function $w$ decomposed as
$$
w(x,t)=\sum_{q\in (\mathbb N ^*)^n} w_q(t)\sin (q_1x_1)\cdots \sin (q_nx_n)
$$
we recall that
$$
||w||^2_{X_{s,b}(\Omega )}
=\sum_{q\in (\mathbb N ^*)^n}\int d\tau \langle \tau +|q|^2\rangle ^{2b} \langle q\rangle ^{2s}
| \hat{w}_q (\tau )|^2
$$
Therefore, it is sufficient to check that
$$
I:=\sum_{q\in (\mathbb N ^*)^n}\int d\tau \langle q \rangle ^{2s}
\langle \tau + |q|^2 \rangle ^{2b} | \widehat{\eta u_q}(\tau )|^2
\le c\sum_p \langle p\rangle ^{2s+4} |v_p|^2.
$$
Using \eqref{XYZ}, we may write
$$
I\le c(I_1+I_2+I_3)
$$
where
\begin{eqnarray*}
I_1 &=& \sum_q\int d\tau \langle q\rangle ^{2s} \langle \tau + |q|^2 \rangle ^{2b}
\left( \sum_{p;|p|=|q|} |v_p \, \widehat{t\eta (t)} (\tau + |q|^2)|
p_nq_n \prod_{j=1}^{n-1} \langle p_j-q_j\rangle ^{-k} \right) ^2 \\
I_2 &=& \sum_q\int d\tau \langle q\rangle ^{2s} \langle \tau + |q|^2\rangle ^{2b}
\left( \sum_{p;|p|\ne |q|} \left\vert v_p
\frac{\hat\eta (\tau + |q|^2)}{|q|^2-|p|^2} \right\vert
p_nq_n \prod_{j=1}^{n-1} \langle p_j-q_j\rangle ^{-k} \right) ^2 \\
I_3 &=& \sum_q\int d\tau \langle q\rangle ^{2s} \langle \tau + |q|^2\rangle ^{2b}
\left( \sum_{p;|p|\ne |q|} \left\vert v_p
\frac{\hat\eta (\tau + |p|^2)}{|q|^2-|p|^2} \right\vert
p_nq_n \prod_{j=1}^{n-1} \langle p_j-q_j\rangle ^{-k} \right) ^2
\end{eqnarray*}
We bound separately $I_1$, $I_2$ and $I_3$. \\
1.
\begin{eqnarray*}
I_1 &\le& C (\int d\sigma \langle \sigma \rangle ^{2b}
|\widehat{t\eta (t)} (\sigma)|^2)
\sum _q \langle q\rangle ^{2s} q_n^2
\left( \sum_{p;|p|=|q|}
|v_p| p_n\prod_{j=1}^{n-1}\langle p_j-q_j \rangle ^{-k} \right) ^2 \\
&\le& C\sum _q \langle q\rangle ^{2s} q_n^2
\left( \sum_{p;|p|=|q|} |v_p|^2 p_n^2
\prod_{j=1}^{n-1}\langle p_j-q_j \rangle ^{-k} \right)
\left( \sum_{p;|p|=|q|}\prod _{j=1}^{n-1}
\langle p_j - q_j\rangle ^{-k} \right)
\end{eqnarray*}
where we used successively a change of variables in the integral term,
the fact that $\eta \in {\cal S }(\mathbb R )$ and Cauchy-Schwarz inequality.
From
\begin{equation*}
\sum_{p;|p|=|q|} \prod_{j=1}^{n-1} \langle p_j-q_j\rangle ^{-k}
\le \sum_{p_1,...,p_{n-1}} \left(
\prod_{j=1}^{n-1}\langle p_j-q_j\rangle ^{-k}
\sum_{p_n;|p|=|q|} 1 \right)
\le \prod_{j=1}^{n-1} \sum_{p_j\in \mathbb Z} \langle p_j\rangle ^{-k}
<\infty
\end{equation*}
we deduce that
\begin{eqnarray*}
I_1 &\le& C\sum_p |v_p|^2 |p|^2 \sum_{q;|q|=|p|}
\langle q\rangle ^{2s+2} \prod_{j=1}^{n-1}\langle p_j-q_j\rangle ^{-k} \\
&\le& C\sum_{p} |v_p|^2 |p|^{2s+4}.
\end{eqnarray*}
2.
\begin{eqnarray*}
I_2 &=& C (\int d\sigma \langle \sigma \rangle ^{2b}
|\hat{\eta} (\sigma)|^2)
\sum _q \langle q\rangle ^{2s} q_n^2
\left( \sum_{p;|p|\ne |q|}
\left\vert \frac{v_p}{ |q|^2 - |p|^2 } \right\vert
p_n\prod_{j=1}^{n-1}\langle p_j-q_j \rangle ^{-k} \right) ^2 \\
&\le& c\sum _q \langle q\rangle ^{2s} q_n^2
\left( \sum_{p;|p|\ne |q|} \frac{|v_p|^2 p_n^2}{||q|^2-|p|^2|^{2(1-\delta )}}
\prod_{j=1}^{n-1}\langle p_j-q_j \rangle ^{-k} \right)
\left( \sum_{p;|p|\ne |q|} ||q|^2-|p|^2|^{-2\delta }
\prod _{j=1}^{n-1} \langle p_j - q_j\rangle ^{-k} \right)
\end{eqnarray*}
where we used Cauchy-Schwarz inequality, and $\delta > 1/4$ was chosen
so that $s+2\delta <1/2$. From Claim 3, we obtain that
\begin{equation*}
\sum_{p;|p|\ne |q|} | |q|^2-|p|^2|^{-2\delta} \prod_{j=1}^{n-1}
\langle p_j - q_j\rangle ^{-k}
\le C\sum_{p'}\prod_{j=1}^{n-1} \langle p_j-q_j\rangle ^{-k}
\sum_{p_n;|p|\ne |q|}
\langle |q|^2-|p|^2\rangle ^{-2\delta} <const.
\end{equation*}
Therefore, since $s<0$, we see that
$$
I_2 \le C \sum_q q_n^{2s+2} \sum_{p;|p|\ne |q|}
\frac{|v_p|^2 p_n^2}{||q|^2-|p|^2|^{2(1-\delta )}}
\sum_{j=1}^{n-1} \langle p_j - q_j\rangle ^{-k}
$$
and from Claim 4
$$
I_2 \le C\sum_{p} |v_p|^2 |p|^{2s+4}.
$$
3. From the elementary estimate
$$
\langle \tau + |q|^2 \rangle
\le c \langle \tau + |p|^2 \rangle \langle |q|^2 - |p|^2\rangle
$$
we infer that
\begin{equation}
\label{X10}
I_3\le C\sum_q \int d\tau \langle q\rangle ^{2s} |q_n|^2
\left(
\sum_{p; |p|\ne |q|} |v_p|
\frac{|\hat\eta (\tau + |p|^2)| \langle \tau + |p|^2\rangle ^b}
{||q|^2-|p|^2|^{1-b}} p_n \prod_{j=1}^{n-1} \langle p_j-q_j\rangle ^{-k}
\right) ^2.
\end{equation}
For any fixed $\gamma >1$, we have that for some constant $c>0$
$$
\langle \sigma \rangle ^b |\hat \eta (\sigma )| \le c \langle \sigma \rangle ^{-\gamma}
\qquad \forall \sigma \in \mathbb R .
$$
Expanding the squared term in (\ref{X10}) results in
\begin{eqnarray*}
I_3
&\le& C\sum_q \langle q\rangle ^{2s} |q_n|^2
\sum_{p;|p|\ne |q|}\
\sum_{\tilde p; |\tilde p|\ne |q|}
\frac{|v_p| \, |v_{\tilde p}| p_n\tilde p_n}
{||q|^2 -|p|^2|^{1-b} ||q|^2 -|\tilde p|^2|^{1-b}}\\
&&\quad \times (\prod_{j=1}^{n-1}\langle p_j-q_j\rangle ^{-k}\langle \tilde p_j- q_j\rangle ^{-k})
\int d\tau \langle \tau + |p|^2 \rangle ^{-\gamma }
\langle \tau + |\tilde p|^2 \rangle ^{-\gamma}\\
&\le& C\sum_q \langle q\rangle ^{2s} |q_n|^2
\sum_{p;|p|\ne |q|}\
\sum_{\tilde p; |\tilde p|\ne |q|}
\frac{|v_p| \, |v_{\tilde p}| p_n\tilde p_n}
{||q|^2 -|p|^2|^{1-b} ||q|^2 -|\tilde p|^2|^{1-b}}\\
&&\quad \times (\prod_{j=1}^{n-1}\langle p_j-q_j\rangle ^{-k}\langle \tilde p_j- q_j\rangle ^{-k})
\langle |p|^2 -|\tilde p|^2\rangle ^{-\gamma}
\end{eqnarray*}
where we used the following estimate valid for $\gamma >1$ (see e.g. \cite[Lemma 7.34]{linares-ponce})
$$
\int d\tau \langle \tau + \tau _1 \rangle ^{-\gamma}
\langle \tau + \tau _2 \rangle ^{-\gamma}
\le c \langle \tau _1 -\tau _2\rangle ^{-\gamma}.
$$
Thus
$$
I_3 \le
C \sum_q \langle q\rangle ^{2s} q_n^2
\sum_{p;|p|\ne |q|}
\frac{|v_p|^2p_n^2}{||q|^2 -|p|^2|^{2(1-b)}}(\prod_{j=1}^{n-1} \langle p_j-q_j\rangle ^{-k} )
\sum_{\tilde p;|\tilde p|\ne |q|} \prod_{j=1}^{n-1}
\langle \tilde p_j - q_j \rangle ^{-k} \langle |p|^2 - |\tilde p|^2 \rangle ^{-\gamma}.
$$
Since $\gamma >1/2$, it follows from Claim 3 that
$$
\sum_{\tilde p}\prod_{j=1}^{n-1}
\langle \tilde p_j -q_j \rangle ^{-k} \langle |p|^2 - |\tilde p| \rangle ^{-\gamma}
\le \sum_{{\tilde p}_1,...,{\tilde p}_{n-1}}\prod_{j=1}^{n-1}
\langle \tilde p_j -q_j\rangle ^{-k} \sum_{\tilde p_n}
\langle {\tilde p}_n^2 + |\tilde p'|^2 -|p|^2\rangle ^{-\gamma} <const.
$$
Thus
$$
I_3 \le C \sum_q\langle q\rangle ^{2s} q_n^2
\sum_{p;|p|\ne |q|} \frac{|v_p|^2 p_n^2}{||q|^2-|p|^2|^{2(1-b)}}
\prod_{j=1}^{n-1}\langle p_j-q_j\rangle ^{-k}.
$$
Using Claim 4 and the fact that $s\in [-1,0)$, we have that
$$
I_3 \le C \sum_p |v_p|^2 |p|^2 \sum_{q;|q|\ne |p|}
\frac{q_n ^{2s+2}}{||q|^2-|p|^2|^{2(1-b)}}
\prod_{j=1}^{n-1} \langle p_j-q_j\rangle^{-k}
\le \sum_p |v_p|^2 |p|^{2s+4}.
$$
\noindent{\em Step 3.} Assume that $s+2b<1/2$ with
$s\in [0,1/2)$. Let $u_T$, $v_0$, $u$ and $\eta$ be as in Step 2.
Then
\begin{eqnarray}
||\eta (t) \Gamma v_0||_{X_{s,b}}
&\le& C||\eta D^{s+1}\Gamma v_0||_{X_{-1,b}} \nonumber \\
&\le& C \left( ||\eta (t) \Gamma (D^{s+1}v_0)||_{X_{-1,b}}
+||\eta (t) [\Gamma , D^{s+1}] v_0||_{X_{-1,b}}\right) . \label{AAA1}
\end{eqnarray}
According to Step 2, the first term in the r.h.s. of
\eqref{AAA1} is less than $C||D^{s+1}v_0||_1\le C||v_0||_{s+2}$, for
$-1+2b<1/2$. The contribution due to $F_0=\{ x\in \partial \Omega ; \ x_n=0 \}$
in $||\eta (t) [\Gamma, D^{s+1}]v_0||^2_{-1, b}$ is estimated by
\begin{eqnarray*}
C_{F_0} &\le & \sum_q \int d\tau
\langle q \rangle ^{-2} \langle \tau + |q|^2 \rangle ^{2b}
\left\vert \sum_{p;|p|\ne |q|}
v_p\frac{|p|^{s+1}-|q|^{s+1}}{|q|^2-|p|^2}
(\hat\eta (\tau + |p|^2) -\hat\eta (\tau + |q|^2))I_{F_0} \right\vert ^2 \\
&\le& C(I_2'+I_3')
\end{eqnarray*}
where
\begin{eqnarray*}
I_2' &=& \sum_q\int d\tau \langle q\rangle ^{-2} \langle \tau + |q|^2 \rangle ^{2b}
\left( \sum_{p;|p|\ne |q|}
|v_p\hat\eta (\tau + |q|^2)| \frac{|p|^s + |q|^s}{|p| + |q|} p_nq_n
\prod_{j=1}^{n-1}\langle p_j-q_j\rangle ^{-k} \right) ^2 ,\\
I_3' &=& \sum_q\int d\tau \langle q\rangle ^{-2} \langle \tau + |q|^2 \rangle ^{2b}
\left( \sum_{p;|p|\ne |q|}
|v_p\hat\eta (\tau + |p|^2)| \frac{|p|^s + |q|^s}{|p| + |q|} p_nq_n
\prod_{j=1}^{n-1}\langle p_j-q_j\rangle ^{-k} \right) ^2.
\end{eqnarray*}
We bound separately $I_2'$ and $I_3'$. \\
1. We have that
\begin{eqnarray*}
I_2' &\le &
C \big( \int d\sigma \langle \sigma \rangle ^{2b} |\hat\eta (\sigma )|^2 \big)
\sum_q \langle q\rangle ^{-2} |q_n|^2 \left\vert \sum_{p;|p|\ne |q|}
|v_p| \frac{|p|^s + |q|^s}{|p|+|q|} p_n \prod_{j=1}^{n-1}
\langle p_j-q_j\rangle ^{-k} \right\vert ^2 \\
&\le & C\sum_q \left\vert \sum_{p}
|v_p| \frac{|p|^s + |q|^s}{|p|+|q|} p_n \prod_{j=1}^{n-1}
\langle p_j-q_j\rangle ^{-k} \right\vert ^2 \\
&\le & C\sum_p |p|^{3+2s} |v_p|^2 \\
&\le & C||v_0||^2_{s+\frac{3}{2}}.
\end{eqnarray*}
where we used \eqref{orion1}-\eqref{orion2}. \\
2. Doing computations similar to those performed in Step 2, we obtain that
\begin{eqnarray*}
I_3' &\le &
C\sum_q \langle q\rangle ^{-2} q_n^2 \sum_{p;|p|\ne |q|}
|v_p|^2 p_n^2 \frac{|p|^{2s} + |q|^{2s}}{(|p|+|q|)^2}
\left\vert |q|^2 -|p|^2\right\vert ^{2b} \prod_{j=1}^{n-1}
\langle p_j-q_j\rangle ^{-k} \\
&\le& C\sum_p |v_p|^2 |p|^2 \sum_{q; |q|\ne |p|}(|p|+|q|)^{2s+4b-2}
\prod_{j=1}^{n-1} \langle p_j-q_j\rangle ^{-k} \\
&\le& C||v_0||^2_{1}
\end{eqnarray*}
where we used the fact that $s+2b<1/2$. Since $s+2\ge 1$, we finally have that
$$
C_{F_0}\le C ||v_0||^2_{H^{s+2}_D(\Omega )}.
$$
This completes the proof of Proposition \ref{prop15}.{\hfill$\quad$\qd\\}
We can now complete the proof of Theorem \ref{thm11}. Let
$s,b,u_0$ and $u_T$ be as in the statement of the theorem.
Using Proposition \ref{prop15} and proceeding as in the
proof of Theorem \ref{thm10}, one can show that the map
\begin{equation}
\label{X2}
\Gamma (v) = W_D(t)u_0 + i\int_0^t W_D(t-\tau )N(v)(\tau )\, d\tau
+\Lambda (u_T - W_D(T)u_0 -\omega (v,T))
\end{equation}
has a fixed-point $\Gamma (v)=v$ in some closed ball $B_M\subset X_{s,b}^T(\Omega )$
provided that
$||u_0||_{H^s_D(\Omega )} + ||u_T||_{H^s_D(\Omega )} $ is small enough.
Such a trajectory $v$ fulfills all the requirements of Theorem \ref{thm11}.
In particular, $v\in X_{s,b}^T(\Omega )\cap C([0,T]; H^s_D(\Omega))$.
The smoothness of the last term in \eqref{X2} follows from
Proposition \ref{prop15}. In (\ref{X2}), we used the notation
$$
\omega (v,T) = i\int_0^T W_D(T-\tau )N(v)(\tau )d\tau.
$$
Note that $\int_0^t W_D(t-\tau )N(v)(\tau )\, d\tau\in X_{s,b'+1}^T(\Omega )
\subset C([0,T]; H^s_D(\Omega ))$, by Lemma \ref{lem11},
\eqref{multilinearnu}, and the
fact that $b'> -1/2$. In particular, $\omega (v,T)\in H^s_D(\Omega )$.
The proof of Theorem \ref{thm11} is achieved.
{\hfill$\quad$\qd\\}
\begin{rem}
(a) Using ideas from \cite{bourgain-1}, it is likely that Theorem
\ref{thm11} may be applied when $n\ge 2$, $\Gamma_0$ is a neighborhood of a vertex, and $N(u)=\lambda |u|^{\alpha}u$ with $\alpha >0 $ small enough. \\
(b) The condition $s+2b<1/2$ in Proposition \ref{prop15} is actually sharp.
Indeed, let us take $n=1$ and pick any $p\in \mathbb N^*$ and any $\eta \in {\cal S} (\mathbb R )$
with $|\hat\eta (\tau )|>1$ for $-1\le \tau \le 1$. Set
$v_0(x)=\sin (px)$ for $x\in \Omega =(0,\pi )$. With $\Gamma _0=\{ 0\}$, we have
that $I_{F_0}=pq$ with
$$
\widehat{\eta u_q}(\tau ) =
\left\{
\begin{array}{ll}
-\displaystyle\frac{2i}{\pi}\widehat{t\eta (t)}(\tau + p^2)p^2 \qquad &\text{ if } q=p;\\[3mm]
-\displaystyle\frac{2}{\pi}\displaystyle\frac{\hat\eta (\tau + p^2) -\hat\eta (\tau + q^2)}{q^2-p^2}pq
\qquad &\text{ if } q\ne p.
\end{array}
\right.
$$
Therefore
\begin{eqnarray*}
\frac{\pi ^2}{4} ||\eta u||^2_{X_{s,b}(\Omega )}
&=& \int d\tau \sum_{q;q\ne p}\langle q\rangle ^{2s}
\langle \tau + q^2 \rangle ^{2b}
\left\vert \frac{\hat\eta (\tau + p^2) -\hat\eta (\tau +q^2)}
{q^2-p^2}\right\vert ^2 p^2q^2 \\
&&\qquad + (\int d\tau \langle \tau + p^2\rangle ^{2b} |\widehat{t\eta(t)} (\tau + p^2)|^2
) \langle p\rangle ^{2s} p^{4}\\
&=& \int d\tau \sum_{q;q\ne p}\langle q\rangle ^{2s} \langle \tau + q^2\rangle ^{2b}
\frac{|\hat\eta (\tau + p^2)|^2}{|q^2-p^2|^2} p^2q^2 + J(p)
\end{eqnarray*}
where $|J(p)|\le Cp^{2s+4}\le C||v_0||^2_{s+2}$, according to the
estimations of $I_1$, $I_2$, and the fact that
$$
\int d\tau \langle \tau + q^2 \rangle ^{2b} |\hat\eta (\tau + p^2)
\hat\eta (\tau + q^2)|\ d\tau \le const <\infty.
$$
Since for $q\ne p$
$$
\int d\tau \langle \tau + q^2 \rangle ^{2b} |\hat\eta (\tau + p^2)|^2
\ge \int_{-p^2-1}^{-p^2+1}d\tau \langle \tau + q^2\rangle ^{2b}
\ge C |q^2-p^2|^{2b}
$$
we have that for $s+2b\ge 1/2$,
$$
\int d\tau \sum_{q;q\ne p}\langle q\rangle ^{2s} \langle \tau + q^2\rangle ^{2b}
\frac{|\hat\eta (\tau + p^2)|^2}{|q^2-p^2|^2} p^2q^2
\ge Cp^2\sum_{q;q>p}|q^2-p^2|^{2b-2}\langle q\rangle ^{2s}q^2 =\infty,
$$
therefore $\eta u\not\in X_{s,b}(\Omega )$.
The condition $s+2b<1/2$ seems related to the
fact that any smooth function on $\mathbb T^n$ with nonnull
boundary values belongs to the space
$H^s_D(\Omega )$ for $s<1/2$ only. Better results will
probably require to consider other Bourgain spaces than $X_{s,b}(\Omega )$.
\end{rem}
\begin{cor}
\label{cor11}
Let $n=1$, $\Omega =(0,\pi)$, $\Gamma _0= \{ 0 \}$, and let the nonlinear term $N(u)$ satisfy
$$
|N(u)-N(v)| \le C (|u|^\alpha + |v|^\alpha )|u-v|,\qquad \forall u,v\in \mathbb R.
$$
for some $\alpha \in [0,5/4)$. Let $p=\frac{4}{3}(\alpha +1) <3$.
Then there exists a number $\delta >0$ such that for any $u_0,u_T\in L^2(\Omega )$ satisfying
$$
||u_0||_{L^2(\Omega )}<\delta, \quad ||u_T||_{L^2(\Omega )}<\delta
$$
one may find a function $h\in H^{\frac{1}{2}}(0,T)$ and a solution
$u\in C([0,T];L^2(\Omega ))\cap L^p(0,T;L^p(\Omega ))$
of \eqref{4.4}-\eqref{4.5} such that
$u(0)=u_0$ and $u(T)=u_T$.
\end{cor}
For instance, $N_1(u)=\lambda |u|^\alpha u$ with
$0\le \alpha <5/4$, and $N_2(u)$ of the form \eqref{Nu} with $\alpha =1$ are concerned.\\
{\em Proof.} From the classical Strichartz estimate (see e.g. \cite{tao})
$$
||u||_{L^4(\mathbb R ; L^4(\mathbb T ))} \le C||u||_{X_{0,\frac{3}{8}}}
$$
we obtain at once the following estimates involving the spaces
$X^T_{s,b}(\Omega )$
\begin{eqnarray*}
||u||_{L^4(0,T;L^4(\Omega ))} &\le& C||u||_{X_{0,\frac{3}{8}}^T(\Omega )}\\
||u||_{X^T_{0,-\frac{3}{8}}(\Omega )} &\le&
C||u||_{L^{\frac{4}{3}} (0,T;L^{\frac{4}{3}}(\Omega ))}.
\end{eqnarray*}
Notice that for $v\in L^p(0,T;L^p(\Omega ))$, we have that
$$
\int_0^t W_D (t-\tau )N(v)(\tau ) d\tau
\in X^T_{0,\frac{5}{8}}(\Omega )
\subset C([0,T];L^2 (\Omega )) \cap L^p(0,T;L^p(\Omega )) .
$$
Indeed,
\begin{eqnarray*}
||\int_0^t W_D(t-\tau ) N(v)(\tau ) d\tau ||_{X^T_{0,\frac{5}{8}}
(\Omega )}
&\le& C||N(v)||_{X^T_{0,-\frac{3}{8}}(\Omega )} \\
&\le& C||N(v) ||_{L^\frac{4}{3}(0,T;L^\frac{4}{3}(\Omega ))} \\
&\le& C||v||^{\alpha +1}_{L^p(0,T;L^p(\Omega ))}<\infty \cdot
\end{eqnarray*}
In particular, $\omega (v,T)=i\int_0^T W_D (T-\tau )N(v)(\tau )d\tau
\in L^2(\Omega )$. On the other hand, by Proposition \ref{prop15}, $\Lambda$
maps continuously $L^2(\Omega )$ into $C([0,T];L^2(\Omega ))
\cap X_{0,b}^T (\Omega )$ for any $b<1/4$. Interpolating between
$$
X_{0,\frac{3}{8}} \subset L^4(\mathbb R ;L^4(\mathbb T )) \quad \text{ and } \quad
X_{0,0}=L^2(\mathbb R ; L^2(\mathbb T ))
$$
we obtain that
$$
X_{0,b}\subset L^p(\mathbb R ; L^p(\mathbb T ))\quad \text{ for }
b=\frac{3}{2}(\frac{1}{2}-\frac{1}{p})<\frac{1}{4}\cdot
$$
Therefore
$$
\Lambda (L^2(\Omega )) \subset C([0,T];L^2(\Omega ))
\cap L^p(0,T;L^p(\Omega )).
$$
It follows that the map
\begin{equation*}
\Gamma (v) = W_D(t)u_0 + i\int_0^t W_D(t-\tau )N(v)(\tau )\, d\tau
+\Lambda (u_T - W_D(T)u_0 -\omega (v,T))
\end{equation*}
is well defined from $L^p(0,T;L^p(\Omega ))$ into
$C([0,T];L^2(\Omega ))\cap L^p(0,T;L^p(\Omega ))$. Using the computations above,
one readily sees that $\Gamma$ contracts in some ball
$B_M\subset L^p(0,T;L^p(\Omega ))$, provided that
$||u_0||_{L^2(\Omega )}+||u_T||_{L^2(\Omega )}$ is small enough.{\hfill$\quad$\qd\\}
\begin{cor}
\label{cor12}
Theorem \ref{thm11} may be applied when $n=2$, $\Omega =(0, \pi )^2$, $g$
is a smooth Dirichlet controller,
$N(u)=\overline{u}^2$, $s\in (-\frac{3}{8}, -\frac{1}{3})$,
$b\in (\frac{3}{8}, \frac{1}{2})$ with $s+2b<\frac{1}{2}$, and $b'>-\frac{1}{2}$ is sufficiently close to
$-\frac{1}{2}$.
\end{cor}
Corollary \ref{cor12} is a direct consequence of Theorem \ref{thm11} and of
the following result, whose proof is postponed in Appendix.
\begin{prop}
\label{prop30}
Let $s\in (-\frac{3}{8}, -\frac{1}{3})$ and
$b\in (\frac{3}{8}, \frac{1}{2})$. Then there exists $b'\in (-\frac{1}{2},
-\frac{5}{12})$ and $C>0$ such that
\begin{eqnarray}
||\overline{v}_1\overline{v}_2||_{X_{s,b'}(\mathbb T ^2)}
&\le& C||v_1||_{X_{s,b}(\mathbb T ^2)} ||v_2||_{X_{s,b}(\mathbb T ^2)},
\qquad \forall v_1,v_2\in X_{s,b}(\mathbb T ^2), \label{PP1}\\
||\overline{u}_1\overline{u}_2||_{X_{s,b'}(\Omega )}
&\le& C||u_1||_{X_{s,b}(\Omega )} ||u_2||_{X_{s,b}(\Omega )},
\qquad \forall u_1,u_2\in X_{s,b}(\Omega ). \label{PP2}
\end{eqnarray}
\end{prop}
Notice that if we increase the value of $s$, the state space in which the
controllability result holds has to take
into account the fact that the value (or the normal derivative) of the
function vanishes on $\partial \Omega \setminus \Gamma _0$.
To state a result of this kind,
we limit ourselves to the situation when $\Gamma _0$ is a side, e.g.
$$
\Gamma _0=\{ 0\} \times (0,l_2)\times \cdots \times (0,l_n).
$$
Introduce the domain
$\tilde \Omega =(-1,l_1)\times (0,l_2)\cdots \times (0,l_n)$
and a function $a\in C_0^\infty(\tilde\Omega \setminus \overline{\Omega})$,
and consider the internal control problem
\begin{equation}
iu_t + \Delta u + N(u) = ia(x)h(x,t) ,
\qquad x\in \tilde\Omega, \ t\in (0,T). \label{4.10}
\end{equation}
Taking the restriction to
$\Omega \times (0,T)$ of solutions of \eqref{4.10}, we obtain
as a corollary of Theorem \ref{thm10} that both systems
(\ref{4.4})-(\ref{4.5})
and (\ref{4.4})-(\ref{4.6}) are locally exactly controllable in some
subspace of $H^s (\Omega )$ for any $s> s_{\alpha ,n}$.
\begin{cor} For given $\alpha \geq 1$, $n\geq 2$,
$\lambda \in \mathbb R$, $s> s_{\alpha, n}$
and $T>0$, there exists a constant $\delta >0$ such that for any
$u_0 , \ u_1 \in H^s (\Omega )$ satisfying
\[ \| u_i \|_{H^s (\Omega )} \leq \delta , \ i=0,1 \]
and
\[
u_i=\Delta u_i=\cdots =\Delta ^p u_i=0\quad
x\in \partial\Omega \setminus \Gamma_0,\ p\le \left[\frac{2s-1}{4}\right],
\ i=0,1
\]
\[\text{(resp.}\qquad\qquad
\frac{\partial u_i}{\partial \nu}=
\frac{\partial \Delta u_i}{\partial \nu}=\cdots =
\frac{\partial \Delta ^p u_i}{\partial \nu}=0\quad
x\in \partial \Omega \setminus \Gamma _0, \ p\le \left[\frac{2s-3}{4}
\right] ,\ i=0,1),
\]
then one can choose a control input $h$ such
that system (\ref{4.4})-(\ref{4.5}) (resp. system
(\ref{4.4})-(\ref{4.6})) admits a solution
$u\in C([0,T]; H^s (\Omega ))$ with
\[ u(x,0)= u_0 (x), \qquad u(x,T) =u_1 (x). \]
\end{cor}
\noindent
\begin{rem}
By using the same extension and restriction argument, one can derive
a local controllability result in the space $H^s(\Omega )$ when
$s>s_{\alpha, n}$ and for any given bounded smooth set $\Omega$, provided
that the control is applied on the whole boundary
(i.e. $\Gamma _0=\partial \Omega$). A result of this kind for which
the critical Sobolev exponent $s=s_c=s_{2,2}=0$ is reached, is given in
\cite{RZ2008}.
\end{rem}
\section{Stabilization}
In this section we focus on the internal stabilization of the semilinear
Schr\"odinger equation on the torus $\T ^n$
\begin{equation}
\label{stab}
iu_t + \Delta u + N(u)=-ia^2(x)u,\qquad x\in \T ^n
\end{equation}
where $a$ is any smooth real function with $a\not\equiv 0$.
We have the following local exponential stability result which does not
require the Geometric Control Condition.
\begin{thm}
Let $a\in C^\infty_0(\T ^n)$, $a\not \equiv 0$, and let $s>s_{\alpha ,N}$.
Then there exist some constants $\nu$, $C$ such that every solution $u$ of
\eqref{stab} issued from
the initial state $u_0\in H^s(\T ^n )$ satisfies
\begin{equation}
\label{decay}
||u(t)||_s\le C e^{-\nu t} ||u_0||_s \quad \forall t\ge 0.
\end{equation}
\end{thm}
\noindent
{\em Proof.} We proceed as in \cite{RZ2007b}.
The operator $A_a=i\Delta - a^2$ with domain
${\mathcal D}(A_a)=H^{s+2}(\mathbb T ^n)$ generates a continuous group
$(W_a(t))_{t\in \mathbb R}$ of operators on $H^s(\mathbb T ^n)$. The first
step is to check that the semigroup $(W_a(t))_{t\in \mathbb R ^+}$ is
exponentially stable in $H^s(\mathbb T ^n)$. This is done in the following
\begin{prop}
\label{prop200}
There exist positive constants $C>0$ and $\nu >0$ such that
\begin{equation}
\label{s10}
||W_a(t)u_0||_s\le Ce^{-\nu t} || u_0 ||_s\qquad \forall t\ge 0.
\end{equation}
\end{prop}
\noindent
{\em Proof.} When $s=0$, the exponential
stability of $(W_a(t))_{t\in \mathbb R ^+}$ is a direct consequence of
Theorem \ref{thm1}, according to \cite{liu}. To prove (\ref{s10})
when $s=2$, we pick any $u_0\in H^2(\T ^n )$ and set $v:=u_t$.
Then $v$ solves the system
\begin{equation}
\left\{
\begin{array}{l}
v_t=i\Delta v -a^2(x)v,\qquad x\in \mathbb T ^n,\\
v(x,0)=v_0(x):=i\Delta u_0 (x) -a^2(x)u_0(x).
\end{array}
\right.
\end{equation}
By the property (\ref{s10}) established when $s=0$, we have
$$
||u(t)||_0\le Ce^{-\nu t}||u_0||_0, \qquad
||v(t)||_0\le Ce^{-\nu t}||v_0||_0.
$$
Since $i\Delta u=v+a^2 u$, we conclude that
$$
||u(t)||_2 \le Ce^{-\nu t}||u_0||_2\qquad \forall t\ge 0.
$$
An easy induction yields (\ref{s10}) for any $s\in 2\mathbb N$.
The proposition then follows by a classical interpolation argument. {\hfill$\quad$\qd\\}
Let us now turn our attention to the stability properties of the
nonlinear system
\begin{eqnarray*}
u_t=A_a u +iN(u), \quad
u(.,0)=u_0
\end{eqnarray*}
that we shall write in its integral form
\begin{equation}
\label{integral}
u(t)=W_a(t)u_0 +
i \int_0^t W_a(t-\tau )N(u)(\tau ) d \tau.
\end{equation}
At this point, we need to establish linear estimates
when $W_a$ is substituted to $W$.
\begin{lem}
\label{lem4.1}
Let $T>0$, $s\geq 0$ and $0\leq b\leq 1$ be given. Then there exists a
constant $C>0$ depending only on $T$, $s$ and $b$ such that
\[ \| W_a(t) \phi \|_{X_{s,b}^T} \leq C \| \phi \| _s \]
for any $\phi \in H^s (\mathbb T ^n) $
\end{lem}
\noindent
{\em Proof.} An application of Duhamel formula gives
\begin{equation}
\label{duhamel}
W_a(t)\phi =W(t)\phi - \int_0^t W(t-\tau )(a^2W_a(\tau )\phi)d\tau.
\end{equation}
It follows that
\begin{eqnarray*}
|| W_a(t)\phi ||_{X_{s,b}^T}
&\le& || W(t)\phi ||_{X_{s,b}^T} +
||\int_0^t W(t-\tau )(a^2 W_a(\tau )\phi ) d\tau||_{X_{s,b}^T} \\
&\le& C||\phi||_s + C ||a^2 W_a(t)\phi||_{ X_{s,b-1}^T} \\
&\le& C||\phi||_s + C||W_a(t)\phi||_{ L^2 (0,T;H^s(\mathbb T ^n)) }
\qquad (\hbox{as}\ b-1\le 0)\\
&\le& C||\phi||_s,
\end{eqnarray*}
as desired. {\hfill$\quad$\qd\\}
\begin{lem}
\label{lem4.2}
Let $T>0$, $s\geq 0$, and $b\in (\frac12 , 1)$ be given.
Then there exists a constant $C>0$ depending only on
$T$, $s$ and $b$ such that
\[ \left \| \int ^t_0 W_a(t-\tau) f(\tau) d \tau \right \|_{X_{s,b}^T}
\leq C\| f\| _{X_{s,b-1}^T} \] for any $f\in X_{s,b-1}^T.$
\end{lem}
\noindent
{\em Proof.}
It follows from (\ref{duhamel}) that
$$
\int_0^t W_a(t-\tau )f(\tau )d\tau =
\int_0^t W (t-\tau )f(\tau )d\tau - \int_0^t W(t-\tau ) a^2
\left( \int_0^\tau W_a (\tau -s)f(s)ds\right) d\tau,
$$
hence
\begin{eqnarray*}
||\int_0^t W_a(t-\tau )f(\tau )d\tau ||_{X_{s,b}^T}
&\le& C||f||_{X_{s,b-1}^T}
+ C||a^2 \int_0^t W_a (t-s) f(s)ds||_{X_{s,b-1}^T}\\
&\le& C||f||_{X_{s,b-1}^T}
+ C||\int_0^t W_a (t-s) f(s)ds||_{X_{s,0}^T}\\
&\le&
C||f||_{X_{s,b-1}^T}
+ C T^\alpha ||\int_0^t W_a(t-s)f(s)\, ds ||_{X_{s,b}^T}
\end{eqnarray*}
for some constant $\alpha >0$, by virtue of
Lemmas \ref{lem10} and \cite[Lemma 2.11]{tao}.
The result follows at once if $T$ is small enough,
say $T<T_0$. For $T\ge T_0$, the result
follows from Lemma \ref{lem4.1} and an easy induction.
{\hfill$\quad$\qd\\}
Let us now proceed to the proof of the exponential stability of the
system (\ref{stab}). Pick a number $s\ge 0$. According to
Proposition \ref{prop200}, there exist positive constants
$C,\nu$ such that
$$
||W_a(t) u_0||_s \le Ce^{-\nu t } ||u_0||_s \qquad \forall t\ge 0.
$$
Pick a time $T>0$ such that
$$
Ce^{-\nu T}< \frac{1}{4}
$$
and fix a number $b\in (\frac12,1)$. We seek a solution $u$ of the
integral equation (\ref{integral}) in the form of a fixed point of the map
$$
\Gamma (u)=W_a(t)u_0 + i\int_0^t W_a(t-\tau )N(u)(\tau )d\tau
$$
in some ball $B_M$ of the space $X_{s,b}^T$. This will be done provided that
$||u_0||_s\le \delta$ where $\delta $ is a small number to be determined.
Furthermore, to ensure the exponential stability,
$\delta$ and $M$ will be chosen in such a way that
$||u(T)||_s\le ||u_0||_s/2$. Pick for the moment any $\delta >0$ and $M>0$,
and let $u_0\in H^s(\mathbb T ^n)$ be such that
$||u_0||_s \le \delta $. By computations
similar to those displayed in the proof of Theorem \ref{thm10} with $W_a(t)$
substituted to $W(t)$, we arrive to
$$
||\Gamma (u)||_{X_{s,b}^T} \le c ||u_0||_s + cM^{\alpha +1} \qquad
\forall u\in B_M
$$
and
$$
||\Gamma (u)-\Gamma (v)||_{X_{s,b}^T} \le c M^{\alpha } ||u-v||_{X_{s,b}^T}
\qquad \forall u,v \in B_M$$
for some constant $c>0$ independent of $\delta$, $M$, and $u_0$.
On the other hand, using the estimate of $||\omega (T,u)||_s$ in
the proof of Theorem \ref{thm10}, we obtain
\begin{eqnarray*}
||\Gamma (u)(T)||_s
&\le& ||W_a(T)u_0||_s + ||\int_0^T W_a(T-t) N(u)(t)dt ||_s \\
&\le& \frac{1}{4} ||u_0||_s +cM^{\alpha +1}.
\end{eqnarray*}
Pick $\delta =4cM^{\alpha +1}$ where $M>0$ is chosen so that
$$
(4c^2+c)M^{\alpha +1}\le M \ \hbox{ and } \ c M^\alpha \le \frac{1}{2}.
$$
Then we have
\begin{eqnarray*}
||\Gamma (u)||_{X_{s,b}^T} &\le& M \qquad \forall u\in B_M\\
||\Gamma (u)-\Gamma (v)||_{X_{s,b}^T} &\le& \frac{1}{2}||u-v||_{X_{s,b}^T}
\qquad \forall u,v\in B_M.
\end{eqnarray*}
Thus the map $\Gamma$, which is a contraction in $B_M$, has a
fixed point $u\in B_M$. By construction, $u$ fulfills
$$
||u(T)||_s = ||\Gamma (u)(T) ||_s \le \frac{\delta }{2}.
$$
Assume now that $0<||u_0||_s<\delta$. Changing $\delta$ into
$\delta ':=||u_0||_s$ and $M$ into
$M':=(\delta '/\delta)^{\frac{1}{\alpha +1}}M$, we
obtain that $||u(T)||_s\le ||u_0||_s/2$, and an obvious induction yields
$||u(kT)||_s\le 2^{-k}||u_0||_s$ for any $k\ge 0$. As
$X_{s,b}^T\subset C([0,T];H^s(\mathbb T ^n))$ for $b>1/2$, and
$||u||_{X_{s,b}^T}\le M=(\delta/(4c))^{\frac{1}{\alpha + 1}}$, we
infer by the semigroup property that there exist some constants
$C'>0, \nu'>0$ such that
$$
||u(t)||_s\le C'e^{-\nu 't}||u_0||_s.
$$
The proof is complete. {\hfill$\quad$\qd\\}
\section{Appendix}
\subsection{Proof of Proposition \ref{prop12}.}
We proceed as in \cite[pp. 115-118]{bourgain-2}.
We first introduce some notations.
Let $|x|_\infty :=\sup_{1\le i\le n}|x_i|$
for $x=(x_i)_{1\le i\le n}\in \mathbb R^n$. We introduce a
dyadic partition of $\mathbb R ^n$
$$
\mathbb Z ^n = \cup _{j\ge 0} D_j,
$$
where $D_0 =\{ 0 \}$, and $D_j=\{ k\in \mathbb Z ^n;\ 2^{j-1} \le |k|_\infty <2^j\}$
for $j\ge 1$. For any H\"older exponent $p,q\in [1,+\infty ]$, we write
$L_t^pL_x^q$ for $L^p(\mathbb R _t, L^q (\mathbb T ^n_x))$. The (discrete) cube of center
$x_0\in \mathbb R ^n$ and sidelength $2R>0$ is
$$
Q(x_0,R)=\{ k\in \mathbb Z ^n;\ |k-x_0|_\infty \le R \} .
$$
The Strichartz estimate (\cite{bourgain-1},\cite{grunrock00})
$$
||u||_{L^4_tL^4_x} \le c||u||_{X_{s,b}},
\qquad s>\frac{n}{2} -\frac{n+2}{4}, \ b>\frac{1}{2},
$$
when combined with the standard estimates
\begin{eqnarray*}
|| u ||_{L^\infty _t L^2_x} &\le& c ||u||_{X_{0,b}},\quad b>\frac{1}{2} \\
|| u ||_{L^2_tL^2_x} &=&||u||_{X_{0,0}}
\end{eqnarray*}
and Sobolev embedding theorem, gives by interpolation the following result.
\begin{lem}(\cite[cor. 2.2]{grunrock00})
\label{interpolation}
Let $n\ge 2$.\\
(i) For all $p,q,s$ satisfying
\begin{equation}
\label{S20}
0< \frac{1}{p}\le \frac{1}{4},\ 0< \frac{1}{q} \le \frac{1}{2} -
\frac{1}{p},
\ s>\frac{n}{2} -\frac{2}{p} -\frac{n}{q},
\end{equation}
there exists a number $b\in (0,\frac{1}{2})$ such that for all
$u\in X_{s,b}$, it holds
\begin{equation}
\label{S21}
||u||_{L^p_tL^q_x} \le c ||u||_{X_{s,b}}
\end{equation}
(ii) For all $p,q,s,b$ satisfying
\begin{equation}
\label{S22}
0\le \frac{1}{p}\le \frac{1}{q}\le \frac{1}{2}\le
\frac{1}{p} + \frac{1}{q} \le 1,\
s>(n-2)(\frac{1}{2} -\frac{1}{q}),\
\text{ and } b > 1 - \frac{1}{p}-\frac{1}{q}
\end{equation}
then for all $u\in X_{s,b}$, \eqref{S21} holds.
\end{lem}
Let ${\cal F}_x$ denote the Fourier transform in $x$, and let
$1_{Q}$ denote the characteristic function of the
cube $Q$. The following result,
inspired by an observation made in \cite{bourgain-1}, indicates that
for a function spatially supported in a cube, only the sidelength of
the cube (not its center) comes into play in \eqref{S21}.
\begin{lem} (\cite[Lemma 2.4]{grunrock00})
Assume that for $p,q,s,b$ the estimate \eqref{S21} is valid. Then there exists
a constant $c>0$ such that for any cube
$Q$ of center $x_0\in \mathbb R ^n$ and sidelength $R>0$ it holds
\begin{equation}
||({\cal F}_x^{-1} 1_Q {\cal F}_x) u||_{L^p_t L^q_x}
\le c R^s ||u||_{X_{0,b}}\cdot
\end{equation}
\end{lem}
It follows that if \eqref{S20} (or \eqref{S22}) holds and if $u=u(x,t)$ is
a function decomposed as
$$
u(x,t)=\sum_{|k-x_0|_\infty \le R} \int_{\mathbb R}
{\hat u}(k,\tau ) e^{i(k\cdot x + \tau t)} d\tau
$$
then
\begin{equation}
||u||_{L^p_tL^q_x} \le c R^s ||u||_{X_{0,b}}
=cR^s \left(\sum_{|k-x_0|_\infty \le R}\int_{\mathbb R}
\langle \tau + |k|^2 \rangle ^{2b} |\hat u (k,\tau )|^2 d\tau
\right) ^{\frac{1}{2}}.
\label{S15}
\end{equation}
Let the functions $u_1,...,u_{\alpha + 1}\in X_{s,b}$ be given, where
$s$ and $b$ denote some positive numbers, and let us set
$$
u= \tilde{u}_1 \tilde{u}_2 \cdots \tilde{u}_{\alpha + 1}
$$
where $\tilde{u}_i$ is $u_i$ or $\overline{u_i}$. To estimate
$||u||_{X_{s,-b}}$ we proceed by duality, estimating the integral
$\int_{\mathbb R}\int_{\mathbb T ^n}u\overline{v}dxdt$ for any $v\in X_{-s,b}$ with
$||v||_{X_{-s,b}}\le 1$. By Plancherel theorem
\begin{eqnarray*}
\int_{\mathbb R}\int_{\mathbb T ^n} u\overline{v}\, dxdt
&=& \sum_{k\in \mathbb Z ^n}\int_{\mathbb R} \hat u(k,\tau) \overline{\hat v}(k,\tau )d\tau\\
&=& \sum_{k_1\cdots k_{\alpha +1}}\int_{\tau_1 \cdots \tau _{\alpha +1}}
\langle k\rangle ^s \big(\prod_{i=1}^{\alpha +1} \hat{\tilde u}_i(k_i,\tau_i)\big)
\langle k\rangle ^{-s}\overline{\hat v}(k,\tau)
\end{eqnarray*}
where $k=k_1 + \cdots + k_{\alpha +1}$ and $\tau =\tau _1 + \cdots +
\tau _{\alpha +1}$. Notice that
$\hat{\overline u}(k_i,\tau _i)=\overline{\hat{u}_i (-k_i,-\tau _i)}$.
Writing $k_i\in D_{j_i}$, $j_i\ge 0$, we obtain
$$
\vert \int_{\mathbb R}\int_{\mathbb T ^n} u\overline{v}\, dxdt \vert
\le\sum_{j_1\cdots j_{\alpha +1}} \sum_{k_i\in D_{j_i}}
\int_{\tau _1\cdots \tau_{\alpha +1}}
\langle k\rangle ^s (\prod _{i=1}^{\alpha +1}
\vert\hat{u}_i(k_i,\tau _i)\vert ) \langle k \rangle ^{-s}
|\hat v(k,\tau )|,
$$
where now
$k=\pm k_1 \cdots \pm k_{\alpha +1}$,
$\tau =\pm \tau _1 \cdots \pm \tau_{\alpha +1}$
($+k_i$ if $\tilde {u}_i=u_i$, $-k_i$ if $\tilde{u}_i=
\overline{u_i}$, and the same for $\pm\tau _i$).
We shall
focus on the sum $\Sigma =\sum_{j_1\ge j_2\ge \cdots \ge j_{\alpha +1}}$, the
other contributions leading to similar bounds.
As $|k_i|_\infty \le 2|k_1|_\infty$ for $i\ge 2$,
we have that
$$
\Sigma\le c\sum_{j_1\ge \cdots\ge j_{\alpha +1}}2^{j_1s}
\sum_{k_i\in D_{j_i}}
\int_{\tau _1\cdots \tau_{\alpha +1}}
(\prod _{i=1}^{\alpha +1}
\vert\hat{u}_i(k_i,\tau _i)\vert ) \langle k \rangle ^{-s}
|\hat v(k,\tau )|.
$$
Pick $\gamma \in\mathbb N ^*$ with
$$
\alpha \le 2^{\gamma -2}
$$
and split $\Sigma$ into $\Sigma _1 + \Sigma _2$ where $\Sigma _1$ corresponds
to the $j_1,...,j_{\alpha +1}$ for which
$$
j_1\ge j_2+\gamma +2 \ge j_2\ge j_3 \ge \cdots \ge j_{\alpha +1}.
$$
Consider a ``partition'' of $D_{j_1}$ into a collection of cubes $Q_l$ of
sidelength $2^{j_2}$
$$
D_{j_1}=\cup_{l} Q_l.
$$
Note that each $k\in D_{j_1}$ belongs to at most $2^n$ cubes $Q_l$.
For any $l$, we denote by ${\tilde Q}_l$ the cube of sidelength
$2^{j_2+\gamma}$ with the same center as $Q_l$ if $k=k_1\pm k_2\cdots$, and
with center the opposite of that of $Q_l$ if $k=-k_1\pm k_2\cdots$.
We claim that $k\in {\tilde Q}_l$ when $k_1\in Q_l$ and
$k_i\in D_{j_i}$ for $i\ge 2$. Indeed
\begin{equation}
\label{S30}
|k_2|_\infty +\cdots +|k_{\alpha +1}|_\infty
\le \alpha 2^{j_2} \le 2^{j_2+\gamma -2},
\end{equation}
hence if $Q_l=Q(x_0,2^{j_2-1})$
$$
|\pm x_0-k|_\infty \le |\pm x_0-\pm k_1|_\infty + |k_2|_\infty +\cdots +
|k_{\alpha +1}|_\infty
\le 2^{j_2 -1} + 2^{j_2+\gamma -2} \le 2^{j_2+\gamma -1}.
$$
Notice also that
${\tilde Q}_l\subset D_{j_1-1}\cup D_{j_1} \cup D_{j_1 +1}$ since
the sidelength of $\tilde{Q}_l$ is at most $2^{j_1-2}$ and $Q_l\subset
D_{j_1}$. It follows that
$$
\Sigma _1 \le c \!\!\!\!\!\!\!\!
\sum_{
\begin{array}{c}
\scriptstyle j_1\ge j_2+\gamma + 2\\
\scriptstyle j_2\ge j_3\ge \cdots \ge j_{\alpha +1}
\end{array}}
\!\!\!\!\!\!\!\!
2^{j_1s}
\sum_l
\sum_{k_1\in Q_l}
\!\!\!\!\!\!\!\!
\sum_{
\begin{array}{c}
\scriptstyle k_2\in D_{j_2},\\
\scriptstyle k_{\alpha +1}\in D_{j_{\alpha +1}}
\end{array}}
\int_{\tau _1\cdots \tau _{\alpha +1}}
(\prod_{i=1}^{\alpha +1} | \hat{u}_i(k_i,\tau _i)|)
1_{\tilde{Q}_l}(k) \langle k\rangle ^{-s} |\hat v(k, \tau )|.
$$
Let us introduce the functions
\begin{eqnarray*}
f_l(x,t) &=& \sum_{k\in Q_l}
\int_\mathbb R |\hat{u}_1 (k,\tau )|e^{i(k\cdot x + \tau t)} d\tau \\
g_l(x,t) &=& \sum_{k\in \tilde{Q}_l}
\int_\mathbb R \langle k\rangle ^{-s} |\hat v (k, \tau )| e^{i(k\cdot x + \tau t)} d\tau
\end{eqnarray*}
and
$$
h_i(x,t)=\sum_{k\in D_{j_i}} \int_\mathbb R |\hat{u}_i(k,\tau )|
e^{i(k\cdot x + \tau t)}d\tau \quad \text{ for } i=2,...,\alpha +1.
$$
By Plancherel theorem
$$
\Sigma_1 \le c\!\!\!\!\!\!\!
\sum_{
\begin{array}{c}
\scriptstyle j_1\ge j_2+\gamma +2\\
\scriptstyle j_2\ge j_3\ge \cdots \ge j_{\alpha +1}
\end{array}}\!\!\!\!\!\!\!
2^{j_1s}
\sum_l\int_{\mathbb R} \int_{\mathbb T ^n}
|f_lh_2\cdots h_{\alpha +1}g_l|\, dxdt.
$$
Pick H\"older exponents $p_1,q_1,p_2,q_2\in [1,\infty )$ such that
\begin{eqnarray}
\frac{3}{p_1} + \frac{\alpha -1}{p_2} &=& 1 \label{S50} \\
\frac{3}{q_1} + \frac{\alpha -1}{q_2} &=& 1 \label{S51}
\end{eqnarray}
We have that
$$
\int_{\mathbb R}\int_{\mathbb T ^n}|f_lh_2\cdots h_{\alpha +1}g_l| dxdt
\le
||f_l||_{L_t^{p_1}L_x^{q_1}}
||g_l||_{L_t^{p_1}L_x^{q_1}}
||h_2||_{L_t^{p_1}L_x^{q_1}}
\prod_{i=3}^{\alpha +1}||h_i||_{L_t^{p_2}L_x^{q_2}}.
$$
Assume that for some exponents $s_1,b_1,s_2,b_2$ the following estimates
hold
\begin{eqnarray}
||u||_{L_t^{p_1}L_x^{q_1}} &\le& c ||u||_{X_{s_1,b_1}}, \label{S40} \\
||u||_{L_t^{p_2}L_x^{q_2}} &\le& c ||u||_{X_{s_2,b_2}}. \label{S41}
\end{eqnarray}
Then, by \eqref{S15} and the fact that the sidelength of $Q_l$ (resp.
$\tilde{Q}_l$) is $2^{j_2}$ (resp. $2^{j_2+\gamma}$), we have
\begin{eqnarray}
||f_l||_{L_t^{p_1}L_x^{q_1}} &\le& c2^{j_2s_1}
\big( \sum_{k\in Q_l}\int_\tau \langle \tau + |k|^2\rangle ^{2b_1}
|\hat u_1|^2 \big) ^{\frac{1}{2}} \label{S300} \\
||g_l||_{L_t^{p_1}L_x^{q_1}} &\le& c2^{j_2s_1}
\big( \sum_{k\in \tilde{Q}_l}\int_\tau \langle \tau + |k|^2\rangle ^{2b_1}
\langle k \rangle ^{-2s}
|\hat v|^2 \big) ^{\frac{1}{2}} \label{S301} \\
||h_2||_{L_t^{p_1}L_x^{q_1}} &\le& c2^{j_2s_1}
\big( \sum_{k\in D_{j_2}}\int_\tau \langle \tau + |k|^2\rangle ^{2b_1}
|\hat u_2|^2 \big) ^{\frac{1}{2}} \label{S302}
\end{eqnarray}
and for $i=3,...,\alpha +1$
\begin{eqnarray}
||h_i||_{L_t^{p_2}L_x^{q_2}}
&\le& c2^{j_is_2}
\big( \sum_{k\in D_{j_i}}\int_\tau \langle \tau + |k|^2\rangle ^{2b_2}
|\hat u_i|^2 \big) ^{\frac{1}{2}}\nonumber \\
&\le& c\big( \sum_{k\in D_{j_i}}
\int_\tau \langle \tau + |k|^2 \rangle ^{2b_2} \langle k \rangle ^{2s_2} |\hat u_i |^2
\big) ^{\frac{1}{2}}.
\label{S303}
\end{eqnarray}
Using Cauchy-Schwarz in $\sum_l$, we obtain
\begin{eqnarray*}
\Sigma _1
&\le& c\!\!\!\!\!\!\!\!\!\!
\sum_{
\begin{array}{c}
\scriptstyle j_1\ge j_2+\gamma +2\\
\scriptstyle j_2\ge j_3\ge \cdots \ge j_{\alpha +1}
\end{array}}
\!\!\!\!\!\!\!\!\!\!
2^{j_1s+3j_2s_1}
\big(
\sum_l\sum_{k\in Q_l}
\int_\tau \langle \tau + |k|^2\rangle ^{2b_1}|\hat u_1|^2
\big) ^{\frac{1}{2}}
\big(
\sum_l\sum_{k\in \tilde{Q}_l}
\int_\tau \langle \tau + |k|^2\rangle ^{2b_1}
\langle k\rangle ^{-2s} |\hat v|^2 \big) ^{\frac{1}{2}}\\
&& \qquad
\big(
\sum_{k\in D_{j_2}}\int_\tau \langle \tau + |k|^2\rangle ^{2b_1}|\hat u_2|^2
\big) ^{\frac{1}{2}}
\prod_{i=3}^{\alpha +1}
\big(
\sum_{k\in D_{j_i}}\int_\tau
\langle \tau + |k|^2\rangle ^{2b_2}
\langle k\rangle ^{2s_2} |\hat u_i|^2 \big) ^{\frac{1}{2}}\\
&\le& c\!\!\!\!\!\!\!\!\!\!
\sum_{
\begin{array}{c}
\scriptstyle j_1\ge j_2+\gamma +2\\
\scriptstyle j_2\ge j_3\ge \cdots \ge j_{\alpha +1}
\end{array}}
\!\!\!\!\!\!\!\!\!\!
\big(
\sum_{k\in D_{j_1}}\int_\tau \langle \tau + |k|^2\rangle ^{2b_1}
\langle k\rangle ^{2s}|\hat u_1|^2
\big) ^{\frac{1}{2}}
\big(
\sum_{k\in D_{j_1-1}\cup D_{j_1}\cup D_{j_1+1}}
\int_\tau \langle \tau + |k|^2\rangle ^{2b_1}
\langle k\rangle ^{-2s} |\hat v|^2 \big) ^{\frac{1}{2}}\\
&& \qquad
\big(
\sum_{k\in D_{j_2}}\int_\tau \langle \tau + |k|^2\rangle ^{2b_1}
\langle k\rangle ^{6s_1} |\hat u_2|^2 \big) ^{\frac{1}{2}}
\prod_{i=3}^{\alpha +1}
\big(
\sum_{k\in D_{j_i}}\int_\tau
\langle \tau + |k|^2\rangle ^{2b_2}
\langle k\rangle ^{2s_2} |\hat u_i|^2 \big) ^{\frac{1}{2}}.
\end{eqnarray*}
We used the fact that a point $k\in D_{j_1-1}\cup D_{j_1}\cup D_{j_1+1}$
belongs to (at most) a finite number of cubes ${\tilde Q}_l$, bounded by
$(2^{\gamma +2}+1)^n$.
A sum
$\sum_{j_i\ge 0}\big(
\sum_{k\in D_{j_i}}
\int_\tau \langle \tau + |k|^2 \rangle ^{2b_2} \langle k\rangle ^{2s_2}
|\hat u_i|^2\big) ^{\frac{1}{2}} $
can be estimated by $c||u_i|| _{X_{s_2+\varepsilon , b_2}}$ for any
$\varepsilon >0$ thanks to Cauchy-Schwarz. Summing successively in
$k_{\alpha +1}, ..., k_1$, we arrive at
$$
\Sigma _1
\le c||u_1||_{X_{s,b_1}} ||v||_{X_{-s,b_1}}
||u_2||_{X_{3s_1+\varepsilon, b_1}}
\prod_{i=3}^{\alpha + 1}
||u_i||_{X_{s_2+\varepsilon, b_2}}.
$$
The same bound for $\Sigma _2$ can be obtained by a more simple analysis.
Indeed, as $j_1\le j_2+\gamma +1$ in the sum over $j_1,...,j_{\alpha +1}$,
we obtain
$$
\Sigma _2 \le
c\!\!\!\!\!\!\!\!\!\!
\sum_{
\begin{array}{c}
\scriptstyle j_1\le j_2+\gamma +1 \\
\scriptstyle j_2\ge j_3\ge \cdots \ge j_{\alpha +1}
\end{array}}
\!\!\!\!\!\!\!\!\!\!
2^{j_1s}
\int_{\mathbb R} \int_{\mathbb T ^n} |f h_2\cdots h_{\alpha + 1} g| dxdt,
$$
where
\begin{eqnarray*}
f(x,t) &=& \sum_{k\in D_{j_1}}
\int_\mathbb R |\hat{u}_1 (k,\tau )|e^{i(k\cdot x + \tau t)} d\tau \\
g(x,t) &=& \sum_{|k|\le (2^{\gamma +1} + \alpha) 2^{j_2}}
\int_\mathbb R \langle k\rangle ^{-s} |\hat v (k, \tau )| e^{i(k\cdot x + \tau t)} d\tau
\end{eqnarray*}
and $h_2,...,h_{\alpha +1}$ as above. Since
$2^{j_1s_1}\le c 2^{j_2s_1}$, we still have
\begin{eqnarray*}
||f||_{L_t^{p_1}L_x^{q_1}}
&\le& c2^{j_2s_1} \big( \sum_{k\in D_{j_1}}
\int_\tau \langle \tau + |k|^2\rangle ^{2b_1}
|\hat u_1|^2 \big) ^{\frac{1}{2}} \\
||g||_{L_t^{p_1}L_x^{q_1}}
&\le& c2^{j_2s_1} \big( \sum_{k\in \mathbb Z ^n}\int_\tau \langle \tau + |k|^2\rangle ^{2b_1}
\langle k \rangle ^{-2s}
|\hat v|^2 \big) ^{\frac{1}{2}}
\end{eqnarray*}
Next, $\Sigma _2$ is estimated as $\Sigma _1$ (see above).
At this stage, we have
proved that
\begin{equation}
\label{S500}
\Sigma \le
c||u_1||_{X_{s,b_1}} ||v||_{X_{-s,b_1}}
||u_2||_{X_{3s_1+\varepsilon , b_1}}
\prod_{i=3}^{\alpha +1} ||u_i||_{X_{s_2+\varepsilon}, b_2}
\end{equation}
where $\varepsilon >0$ is arbitrary small, the exponents $s_1,b_1,s_2,b_2$
are taken so that \eqref{S40}-\eqref{S41} are satisfied, with the
H\"older exponents $p_1,q_1,p_2,q_2$ satisfying \eqref{S50}-\eqref{S51}.
The proof will be complete if, in addition, we have
$$
s\ge \sup \{ 3s_1+\varepsilon, s_2+\varepsilon \} ,\quad
b_1<\frac{1}{2}, b_2<\frac{1}{2}.
$$
We distinguish three cases:
(i) $\alpha \ge 3$; (ii) $\alpha =2$; (iii) $\alpha =1$.\\
\underline{(i) $\alpha \ge 3$}\\
We aim to reach any value $s>s_c$. To find the sets of exponents
$(p_1,q_1,s_1,b_1)$, $(p_2,q_2,s_2,b_2)$ satisfying \eqref{S20},
\eqref{S50} and
\eqref{S51}, and leading to the ``smallest'' value of $s$, we are let to
minimize the functional $\sup \{ 3\sigma _1,\sigma _2\}$, where
\begin{eqnarray}
\sigma _1 &=& \frac{n}{2} - (\frac{2}{p_1}+\frac{n}{q_1}) \label{T1}\\
\sigma _2 &=& \frac{n}{2} - (\frac{2}{p_2}+\frac{n}{q_2}) \label{T2}
\end{eqnarray}
under the constraints
\begin{eqnarray}
&&4\le p_1<\infty \label{T3}\\
&&0 < \frac{1}{q_1}\le \frac{1}{2} - \frac{1}{p_1} \label{T30}\\
&&4\le p_2 <\infty \label{T4}\\
&&0 < \frac{1}{q_2}\le \frac{1}{2} - \frac{1}{p_2} \label{T40}\\
&&\frac{3}{p_1} +\frac{\alpha -1}{p_2} = 1 \label{T5}\\
&&\frac{3}{q_1} + \frac{\alpha -1}{q_2} = 1. \label{T6}
\end{eqnarray}
At this point, it is convenient to introduce the numbers $r_1,r_2$ with
\begin{eqnarray}
\frac{1}{r_1} &=& \frac{2}{p_1} + \frac{n}{q_1} \label{T7}\\
\frac{1}{r_2} &=& \frac{2}{p_2} + \frac{n}{q_2}\cdot \label{T8}
\end{eqnarray}
Note that, by \eqref{T5}-\eqref{T6},
\begin{equation}
\frac{3}{r_1} + \frac{\alpha -1}{r_2}=n+2.
\label{T9}
\end{equation}
Therefore,
$3\sigma _1=\frac{n}{2}-2+\frac{\alpha -1}{r_2}$ (resp.
$\sigma _2=\frac{n}{2}-\frac{1}{r_2}$) is a nonincreasing function
(resp. a nondecreasing function) of $r_2$. Thus the least value of
$\sup \{3\sigma_1, \sigma _2\}$
is achieved when $3\sigma _1=\sigma _2$, which yields
\begin{equation}
\label{S1000}
r_2=\frac{\alpha}{2}, \ r_1 = 3 (n+\frac{2}{\alpha})^{-1}, \quad
3\sigma _1=\sigma_2=\frac{n}{2}-\frac{2}{\alpha}\cdot
\end{equation}
It remains to find $p_1,q_1,p_2,q_2$ satisfying \eqref{T3}-\eqref{T8}.
Note first that \eqref{T6} is satisfied whenever \eqref{T5} is, by \eqref{T9}.
Taking $p_1$ as variable, we infer from \eqref{T5}, \eqref{T7} and \eqref{T8}
that
$$
\frac{1}{p_2}=\frac{1}{\alpha -1}(1-\frac{3}{p_1}),\quad
\frac{1}{q_1}=\frac{1}{3}(1+\frac{2}{n\alpha}) -\frac{2}{np_1},\quad
\frac{1}{q_2}=\frac{2}{n(\alpha -1)} (\frac{3}{p_1}-\frac{1}{\alpha}).
$$
The constraints \eqref{T4}, \eqref{T30} and \eqref{T40} are
found to be respectively equivalent to
\begin{equation}
\label{constraints}
p_1\le 3(1-\frac{\alpha -1}{4})^{-1} (\text{for }\
\alpha \le 4), \quad p_1\ge \sup \big\{ 6(n+\frac{2}{\alpha})^{-1},
6(1-\frac{2}{n})(1-\frac{4}{n\alpha})^{-1}\big\}, \quad
p_1 < 3\alpha .
\end{equation}
The value $p_1=6$ fulfills all the requirements in
\eqref{constraints}. Let now
$s>\frac{n}{2}-\frac{2}{\alpha }$ be given. Choose $\varepsilon >0$ such that
$4\varepsilon < s-(\frac{n}{2}-\frac{2}{\alpha})$, and pick
$s_1\in (\sigma _1, \sigma _1 + \varepsilon )$, and
$s_2\in (\sigma _2, \sigma _2 + \varepsilon )$.
Then \eqref{S40} and \eqref{S41} hold for some numbers
$b_1<\frac{1}{2}$, $b_2<\frac{1}{2}$,
according to Lemma \ref{interpolation}. Set finally
$b=\sup \{b_1,b_2\}$. Then we have
$$
\Sigma \le c \big( \prod_{i=1}^{\alpha +1} ||u_i||_{X_{s,b}}\big)
||v||_{X_{-s,b}}
$$
which gives \eqref{multilinear}.\\
\underline{(ii) $\alpha=2$}\\
Observe first that the approach followed in (i) does not work
for $n>2$. Indeed, the constraints \eqref{T3}-\eqref{S1000}
impose $p_1=p_2=q_1=q_2=4$, and the equation $3\sigma _1=\sigma _2$
is then satisfied only for $n=2$.
Assume $n\ge 3$. We now search a couple
$(p_1,q_1)$ satisfying
\begin{equation}
0<\frac{1}{p_1} \le \frac{1}{q_1} \le \frac{1}{2}
\le \frac{1}{p_1}+\frac{1}{q_1} \le 1,\quad
s_1> (n-2)(\frac{1}{2}-\frac{1}{q_1}), \quad
b_1> 1-\frac{1}{p_1}-\frac{1}{q_1},
\label{W1}
\end{equation}
while $(p_2,q_2)$ still satisfies
\begin{equation}
0 < \frac{1}{p_2}\le \frac{1}{4},\quad
0\le \frac{1}{q_2}\le \frac{1}{2}-\frac{1}{p_2},\quad
s_2>\frac{n}{2} -\frac{2}{p_2}-\frac{n}{q_2}\cdot
\label{W2}
\end{equation}
The H\"older exponents $(p_1,q_1)$ and $(p_2,q_2)$ have to satisfy
the relations
\begin{eqnarray}
\frac{3}{p_1}+\frac{1}{p_2} &=& 1, \label{W3}\\
\frac{3}{q_1}+\frac{1}{q_2} &=& 1. \label{W4}
\end{eqnarray}
We still minimize the functional $\sup \{ 3\sigma _1, \sigma _2\}$, where
$$
\sigma _1=(n-2)(\frac{1}{2}-\frac{1}{q_1}), \quad
\sigma _2 = \frac{n}{2} -\frac{2}{p_2}-\frac{n}{q_2}=
\frac{n}{2}-\frac{2}{p_2} -n (1-\frac{3}{q_1})
$$
by solving in $q_1$ the equation
$3\sigma _1=\sigma _2$. Taking $p_2=4$ to
produce the least value of $\sigma _2$, we find as solution
$q_1=3(1+\frac{1}{4n-5})\in (3,4)$, which yields $p_1=4$ and
$q_2=4(n-1)$ by \eqref{W3}-\eqref{W4}, and
$$
3\sigma _1=\sigma _2= \frac{n}{2}-\frac{3}{4}-\frac{1}{4(n-1)}\cdot
$$
The constraints on
$p_1,q_1,p_2,q_2$ in \eqref{W1}-\eqref{W2} are clearly fulfilled,
for $n>2$. Pick now any $s>\frac{n}{2}-\frac{3}{4}-\frac{1}{4(n-1)}$
and $\varepsilon >0$ such that $4\varepsilon < s-(\frac{n}{2}-\frac{3}{4}
-\frac{1}{4(n-1)})$.
We next pick
$s_1\in (\sigma _1, \sigma _1 +\varepsilon )$,
$s_2\in (\sigma _2, \sigma _2 +\varepsilon )$,
$b_1\in (1-\frac{1}{p_1}-\frac{1}{q_1},\frac{1}{2})$, and $b_2<\frac{1}{2}$
so that \eqref{S21} holds.
Then \eqref{multilinear} follows with $b=\sup\{ b_1,b_2\}$. \\
\underline{(iii) $\alpha =1$}\\
In this case, we have with $p_1=q_1=3$
$$
\Sigma \le c
||u_1||_{X_{s,b_1}}
||u_2||_{X_{3s_1+\varepsilon ,b_1}}
||v||_{X_{-s,b_1}}
$$
provided that \eqref{W1} is satisfied, i.e.
$$
s_1>\sigma_1=\frac{n-2}{6},\quad b_1>\frac{1}{3}\cdot
$$
Therefore, if $s>\frac{n}{2}-1$, taking $\varepsilon >0$ such that
$4\varepsilon < s-(\frac{n}{2}-1)$, $s_1\in (\sigma _1, \sigma _1
+\varepsilon)$, and $b=b_1\in (\frac{1}{3},\frac{1}{2})$, we conclude that
$$
\Sigma \le c||u_1||_{X_{s,b}} ||u_2||_{X_{s,b}} ||v||_{X_{-s,b}}
$$
and \eqref{multilinear} follows. {\hfill$\quad$\qd\\}
\subsection{Proof of Proposition \ref{prop30}.}
We begin with the proof of \eqref{PP1} by following closely
\cite{grunrock01}. Note, however, that the main concern here is to have
the condition $s+2b<1/2$ fulfilled.
Let $s,b$ be as in the statement of
Proposition \ref{prop30}, and let $v_1,v_2\in X_{s,b}$ be decomposed as
$$
v_i(x,t)
=\int_\mathbb R \sum_{k\in \mathbb Z ^2} {\cal F} v_i (k, \tau )
e^{i(k\cdot x + \tau t)} d\tau \qquad i=1,2.
$$
(Here, we use the symbol $\cal F$ instead of $\hat \cdot$ to denote Fourier transform in space and time.)
Let
$$
f_i(k, \tau ) = \langle k\rangle ^s \langle \tau - |k|^2 \rangle ^{b}
{\cal F }\, {\overline{ v}_i}(k,\tau ),\quad \ i=1,2.
$$
Then
\begin{equation}
\label{PP3}
||\overline{v}_1\overline{v}_2||_{X_{s,b'}}
= || \langle k\rangle ^s \langle \tau + |k|^2 \rangle ^{b'}
\int_{\tau _1 + \tau _2 = \tau} \sum_{k_1+k_2=k}\
\prod_{i=1}^2 \langle k_i\rangle ^{-s} \langle \tau _i -|k_i|^2\rangle ^{-b} f_i
||_{L^2_{k,\tau}}
\end{equation}
where $\int_{\tau _1 + \tau _2 = \tau} \sum_{k_1+k_2=k}$ stands for
$\int_{\mathbb R} d\tau _1 \sum_{k_1\in \mathbb Z ^2}$ with the relations
$\tau _1 + \tau _2 = \tau$ and $k_1+k_2=k$ satisfied.
Let $A_0$ (resp. $A_i$, $i=1,2$) denote the region where the largest number
among $\langle \tau +|k|^2\rangle$, $\langle\tau _1 -|k_1|^2 \rangle $ and
$\langle \tau _2 -|k_2|^2 \rangle $, is $\langle \tau +|k|^2 \rangle $
(resp. $\langle \tau _i -|k_i|^2\rangle$, $i=1,2$). We infer from the relation
$$
\tau + |k|^2 -\sum_{i=1}^2 (\tau _i - |k_i|^2)
= |k|^2 +\sum_{i=1}^2 |k_i|^2
$$
that
\begin{equation}
\label{PP4}
\langle k\rangle ^2 +\sum_{i=1}^2 \langle k_i \rangle ^2 \le
C\left( \langle \tau + |k|^2 \rangle + \sum_{i=1}^2 \langle \tau _i -|k_i|^2 \rangle
\right)
\end{equation}
Let us begin with the region $A_0$.
\eqref{PP4} gives, with $0<\varepsilon <
\inf \{ \frac{1}{2}(\frac{1}{2}-|s|), 2(b-|s|) \} $
and $-b':=\frac{1}{2}(\frac{1}{2}-s) + \varepsilon <\frac{1}{2}$
$$
\langle k\rangle ^{\frac{1}{2} +s} \prod_{i=1}^2 \langle k_i\rangle ^{-s+\varepsilon}
\le C \langle \tau + |k|^2 \rangle ^{-b'}.
$$
The contribution in \eqref{PP3} due to $A_0$ is therefore bounded by
\begin{eqnarray*}
&&C||\langle k\rangle ^{-\frac{1}{2}}
\int_{\tau _1 + \tau _2=\tau } \sum_{k_1+k_2=k}
\langle k_i\rangle ^{-\varepsilon} \langle \tau _i -|k_i|^2\rangle ^{-b}
|f_i| ||_{L^2_{k,\tau}} \\
&&\qquad =C ||\langle k\rangle ^{-\frac{1}{2}}
\int_{\tau _1 + \tau _2 =\tau}\sum_{k_1+k_2=k}
\langle k_i\rangle ^{s-\varepsilon } |{\cal F}\, \overline{v}_i| ||_{L^2_{k, \tau}} \\
&&\qquad =C||\prod_{i=1}^2 J^{s-\varepsilon}
{\cal F}^{-1} |{\cal F}\, {\overline v_i} |
||_{L^2_tH^{-\frac{1}{2}}_x} \\
&&\qquad \le C||\prod_{i=1}^2 J^{s-\varepsilon}
{\cal F}^{-1} |{\cal F}\, {\overline v_i} |
||_{L^2_tL^q_x}, \qquad q>\frac{4}{3}\\
&&\qquad \le C\prod_{i=1}^2 ||J^{s-\varepsilon}
{\cal F}^{-1} |{\cal F}\, {\overline v_i} |
||_{L^4_tL^{2q}_x}, \qquad q>\frac{4}{3}\\
&&\qquad \le C\prod_{i=1}^2 ||J^{s-\varepsilon}
{\cal F}^{-1} |{\cal F}\, {\overline v_i} |
||_{X^-_{\varepsilon, b}} \\
&&\qquad \le C\prod_{i=1}^2 ||v_i||_{X_{s,b}}
\end{eqnarray*}
where we used the fact that $L^q(\mathbb T ^2 ) \subset H^{-\frac{1}{2}}(\mathbb T ^2)$
for $q>4/3$ (by dualizing the Sobolev embedding $H^\frac{1}{2}(\mathbb T ^2) \subset L^p(\mathbb T ^2)$ for $p<4$), H\"older inequality, and \eqref{S21}-\eqref{S22}.
We also used the notation
$$||u||_{X^-_{s,b}}=(\int_\mathbb R \sum_{k\in \mathbb Z ^2} \langle k\rangle ^{2s}
\langle \tau -|k|^2\rangle ^{2b} |{\cal F}u(k,\tau )|^2 d\tau )^{\frac{1}{2}} =
||\overline{u}||_{X_{s,b}}$$
borrowed from \cite{grunrock00}.
It remains to estimate the contributions in \eqref{PP3} due to the regions
$A_1$ and $A_2$. By symmetry, we can consider only the region $A_1$.
In $A_1$, since $-s+\frac{\varepsilon}{2} <b$, we have that
$$
\langle k_2\rangle ^{-s+\varepsilon} \langle k_1\rangle ^{-s}
\le C \langle \tau _1 - |k_1|^2 \rangle ^{-s+\frac{\varepsilon}{2}}
\le C \langle \tau _1 -|k_1|^2\rangle ^b
$$
and therefore the contribution in \eqref{PP3} is bounded by
$$
||\langle k\rangle ^s \langle \tau + |k|^2 \rangle ^{b'}
\int_{\tau _1 + \tau _2 =\tau }\sum_{k_1+k_2 =k}
|f_1|\langle k_2\rangle ^{-\varepsilon}
\langle \tau _2 -|k_2|^2\rangle ^{-b} |f_2| ||_{L^2_{k,\tau}}
=C||{\cal F}^{-1} |f_1| J^{s-\varepsilon} {\cal F}^{-1}
|{\cal F}\, \overline{v}_2 | ||_{X_{s,b'}}.
$$
By \eqref{S20}-\eqref{S21} with $-s>1/3$ and $-b'$ chosen sufficiently close to
$\frac{1}{2}$, we have that
$$
X_{-s,-b'}\subset L^6(\mathbb R ; L^6(\mathbb T ^2)),\quad \text{ hence }\quad
L^{\frac{6}{5}}(\mathbb R ; L^{\frac{6}{5}}(\mathbb T ^2)) \subset X_{s,b'}.
$$
It follows that
\begin{eqnarray*}
||{\cal F}^{-1} |f_1| J^{s-\varepsilon} {\cal F}^{-1}
|{\cal F}\, \overline{v}_2 | ||_{X_{s,b'}}
&\le&
C||{\cal F}^{-1} |f_1| J^{s-\varepsilon}{\cal F}^{-1}
|{\cal F}\, \overline{v} _2| ||_{L^{\frac{6}{5}}_tL^{\frac{6}{5}}_x}\\
&\le& C||{\cal F}^{-1} |f_1|||_{L^2_tL^2_x}
|| J^{s-\varepsilon}{\cal F}^{-1}
|{\cal F}\, \overline{v}_2| ||_{L^3_t L^3_x}\\
&\le& C||\overline{v}_1||_{X^-_{s,b}}
|| J^{s-\varepsilon }{\cal F}^{-1}|{\cal F}\, \overline{v}_2|
||_{X^-_{\varepsilon ,b}}\\
&\le& C||v_1||_{X_{s,b}}
|| v_2 ||_{X_{s,b}}\\
\end{eqnarray*}
where we used H\"older inequality and \eqref{S21}-\eqref{S22} with $p=q=3$.
This completes the proof of \eqref{PP1}.
To derive \eqref{PP2} from \eqref{PP1}, we consider two functions $u_1,u_2$
in $X_{0,b}(\Omega ) \subset X_{s,b}(\Omega )$, and consider their odd extensions
$v_1,v_2$ to $(-\pi ,\pi)^2$; i.e., $v_i(\epsilon _1 x_1,\epsilon _2 x_2)
=\epsilon_1\epsilon_2 u_i(x_1,x_2)$ for $x=(x_1,x_2)\in \Omega$ and
$\epsilon_i =\pm 1$. Note that $v_1,v_2\in X_{0,b}$ and that
$\overline{u}_1\overline{u}_2=(\overline{v}_1\overline{v}_2)_{\vert _{\Omega }}$.
For any function
$w=\sum_{k\in \mathbb N ^2} \int_\mathbb R {\cal F} w(k, \tau)
e^{i\tau t}\cos (k_1x_1 ) \cos (k_2x_2) d\tau$, we set
$$
||w||^2_{X_{s,b}(\Omega )_N} =
\sum_{k\in \mathbb N ^2} \int_\mathbb R \langle \tau + |k|^2 \rangle ^{2b}
\langle k\rangle ^{2s} |{\cal F} w(k,\tau )|^2 d\tau.
$$
The Bourgain space $X_{s,b}(\Omega )_N$ (with Neumann boundary conditions) is
defined as the space of the $w$'s for which the norm
$||w||_{X_{s,b}(\Omega )_N}$ is finite. Since the function $\overline{v}_1\overline{v}_2$ is even with respect to both $x_1$ and $x_2$, we have that
$$
||\overline{u}_1\overline{u}_2||_{X_{s,b'}(\Omega )_N}
\sim C ||\overline{v}_1\overline{v}_2 ||_{X_{s,b'}}
\le C||v_1||_{X_{s,b}} ||v_2||_{X_{s,b}}
\le C ||u_1||_{X_{s,b}(\Omega )} ||u_2||_{X_{s,b}(\Omega )}\cdot
$$
We claim that $X_{s,b}(\Omega ) =X_{s,b}(\Omega )_N$ for $|s|<1/2$ and
$|b|\le 1$.
Note first that this is true for $|s|<\frac{1}{2}$ and $b=0$, since
$$
X_{s,0}(\Omega )=L^2(\mathbb R ; H^s(\Omega ))=X_{s,0}(\Omega )_N.
$$
The claim is also true for $|s|<1/2$ and $b=1$, since
$$
u\in X_{s,1}(\Omega ) \iff u\in X_{s,0}(\Omega ) \text{ and }
iu_t+\Delta u \in X_{s,0}(\Omega )
$$
and since a similar criterion may be written for $X_{s,1}(\Omega )_N$.
The claim is also true for $|s|<1/2$ and $0\le b\le 1$ by interpolation,
and for $|s|<1/2$ and $|b|\le 1$ by duality. \eqref{PP2} follows for
$u_1,u_2\in X_{0,b}(\Omega )$, and also for
$u_1,u_2\in X_{s,b}(\Omega )$ by density.
This completes the proof of Proposition \ref{prop30}.
\medskip
\addcontentsline{toc}{section}{References}
|
\section{INTRODUCTION}
A detailed description of hadron structure can be accessed through the
generalized parton
distributions (GPD's)
(see, e.g., Refs. \cite{Ji98,diehlpr,pasq}).
We determine hadron vertex functions
by a study of
electromagnetic (em) form factors (ff) in the spacelike (SL) region within
constituent quark models, and then we use the vertex
functions to evaluate the GPD's.
In this contribution we shortly review :
(i) our results for the unpolarized GPD's
of the pion within covariant and light-front (LF) models \cite{FPPS};
and (ii) our preliminary results for the unpolarized longitudinal
and transverse parton momentum distributions (TMD) in the nucleon within
a LF framework \cite{PDFPS}.
In order to study the GPD's in the valence region as well as in the nonvalence (NV) region,
the Fock state decomposition \cite{Bro} of the hadron state
has to be considered : e.g. for the pion
$ ~| \pi \rangle ~~= ~~
|q\bar{q} \rangle ~~ +
~~ |q \bar{q} ~ q \bar{q}\rangle ~ + ~
|q \bar{q} ~g\rangle ~ + ~... ~$.
Isoscalar (IS) and isovector (IV) pion GPD's, $H^{0,1}_{\pi}(x,\xi,t)$, in
the light-cone gauge are
\begin{eqnarray} &&
\hspace{-0.7 cm} H^{0}_{\pi} = \int \frac{dz^-}{4\pi} e^{i x P^+ z^-}
\left . \langle p' | ~\bar \psi_q
(-\frac{z}{2} ) ~ \gamma ^+ \psi_q (\frac{z}{2} ) | p
\rangle \right |_{\tilde z=0}
\hspace{0.6 cm} H^{1}_{\pi} = \hspace{-0.1 cm}\int \frac{dz^-}{4\pi} e^{i x P^+ z^-}
\left . \langle p' | \bar \psi_q
(-\frac{z}{2}) \gamma ^+ \tau_3 \psi_q (\frac{z}{2})
| p \rangle
\right |_{\tilde z=0}
\end{eqnarray}
where $\tilde z \equiv \{z^+ = z^0 + z^3 , {\bf z}_\perp\}$ , and
$\psi_q(z)$ is the quark field isodoublet, while
\begin{figure}[b]
\vspace{-0.7cm}
\hspace{.3cm}
{\includegraphics[width=6.cm]{fig1n.eps}}
\caption{Diagrammatic representation of the pion GPD, with four momentum definitions
(after Ref. \cite{FPPS}).}
\end{figure}
$ ~{{P}}=\frac12(p'+p)$,
$ ~\Delta = p^\prime-p$,
$~x=k^+/{P}^+$ $~\xi=-(p'^+ - p^+)/2 {{P}}^+$,
$~t=\Delta^2~$,
with $k$ is the average
momentum of the active quark, i.e. the one that interacts with the photon (see Fig. 1).
The variable $x$ allows one to single out (i) the valence region (DGLAP \cite{dglap}), for
$~1\ge x\ge |\xi|$ and $-|\xi| \ge x\ge -1$, diagonal in the Fock space;
and (ii) the nonvalence region (ERBL \cite{erbl}), for $|\xi|> x >-|\xi|$,
non diagonal in the Fock space.
Three pion models are used :
\begin{description}
\item
[An analytic covariant pion model with symmetric regulators and a bare photon vertex.]
We use a pion Bethe-Salpeter amplitude (BSA) with a $\gamma^5$ coupling \cite{tob92}
and a constituent quark (CQ) mass $m=220$ MeV.
Two covariant symmetric forms for the momentum dependent part, $\Lambda(k-P,p)$, of the
BSA are considered:
i) a sum form, and ii) a product form,
which depend on a parameter
chosen to fit the pion decay
constant $f_\pi$ in each model.
The $u$-quark GPD in the pion, $~H^u(x,\xi,t) = H^{I=0}_{\pi}$ $+$ $H^{I=1}_{\pi}$~,
is given in {\em{impulse approximation}} by
\begin{eqnarray} &&
\hspace{-.7cm} H^u(x,\xi,t) = -\imath ~N_c~2 ~ \frac {m^2}{f^2_\pi} ~
\int
\frac{d^4k}{2(2\pi)^4} ~ \delta(P^+x-k^+) \; ~ V^+ ~
\Lambda(k-P,p^{\prime})\; \;
\Lambda(k-P,p) \quad ,
\label{jmu}
\end{eqnarray}
where
the $\delta$ function imposes the correct support, $-|\xi| \le x \le 1$,
for the active quark,
$N_c=3~$ is the number of colors,
$S(k)$ the free quark propagator and
$
V^+=Tr\left \{ S\left({k}-{P}\right)
\gamma^5 ~S\left({k}+
{\Delta}/{2}\right)
\gamma^+~S\left({k}-{\Delta}/{2}\right)\gamma^5\right \}
\label{trace}
$.
We adopt a Breit frame with $~~\Delta^+ = -\Delta^- \ge0$. Then in this model
the whole kinematical range $-1 \le \xi \le 1$ can be explored.
\item
[Mandelstam-inspired pion light-front model.]
We extend to GPD's \cite{FPPS} the model \cite{DFPS} for the pion ff,
based on the covariant
Mandelstam formula for the current \cite{mandel}, with
a microscopic Vector Meson (VM) dominance dressing for the photon vertex.
The expression of Eq. (\ref{jmu}) for $ H^u$ holds, but
the bare quark-photon vertex, $\gamma^+$,
is replaced by the VM dominance (VMD) vertex of \cite{DFPS}.
In the $k^-$ integration only the propagators poles are considered, i.e. the BSA
analytic structure is disregarded in i) the pion state and ii)
the photon vertex.
The dynamical inputs are the pion and VM LF wave functions (wf's).
For the tridimensional reduction of
the $n-th$ VM BSA in the valence sector, $0<k^+<P^+_n$, we take the
eigenfunction of a relativistic
mass operator \cite{Fred}, normalized to the probability
of the valence Fock state \cite{DFPS}.
For the pion in the valence sector
the
eigenfunction of the mass operator \cite{Fred} is used, while
for the {\em{NV pion vertex}},
a constant is assumed \cite{Choi}.
All of the parameters of \cite{DFPS} are used, but for
the CQ mass $m= 200$ MeV, instead of $m= 265$ MeV. A parameter, $w=-1$, which
modulates the relative weights of our two instantaneous contributions is used
to fit the pion ff.
For this model we take the Breit frame where $ {\Delta}_\perp = 0$,
and assume $m_\pi = 0$ (see Ref. \cite{DFPS}).
Then $\xi = -1$ and only the NV region
contributes.
\item
[Light-front Hamiltonian dynamics model.]
In the light-front Hamiltonian dynamics (LFHD) model \cite{Chung}
the Drell-Yan $~\Delta^+ = 0~$ reference frame is adopted
and then the variable
$x$ becomes the longitudinal momentum fraction $x_q$, since $\xi=0$ for any $t$.
Within the LFHD
model the pion LF wave function is
\begin{eqnarray} &&
\hspace{-0.8cm}\Psi_\pi (x,{\bf \kappa}_{\perp }; \lambda_q, \lambda_{\bar q})
= \psi_\pi(x,{\bf \kappa}_{\perp }) ~
\sum_{\mu_q,\mu_{\bar q}} \left({\textstyle {1\over 2}}\mu_q{\textstyle {1\over 2}} \mu_{\bar q}|00\right)
{D}^{1/2\,*}_{\mu_q\lambda_q}\left[R({\bf \kappa})\right]
{D}^{1/2\,*}_{\mu_{\bar q}\lambda_{\bar q}}\left[R(-{\bf \kappa})\right] \quad ,
\label{pionwf}
\end{eqnarray}
with $~\lambda_i~$ ($i=q, \bar q$)~ the spin projections,
$~{\bf \kappa}~ \equiv ~\{{\bf \kappa}_\perp, ~ \kappa_z\} $,
$\kappa_z = M_0(x,{\bf \kappa}_\perp)~(x -{1 \over 2})$ and
$M_0(x,{\bf \kappa}_\perp)$ the pion free mass.
The Melosh rotation
${D}^{1/2}_{\lambda\mu}\left[R({\bf \kappa})\right ] =
\bra{\lambda}R({\bf \kappa})\ket{\mu}
$
converts the instant-form spins
into LF spins and ensure the rotational covariance.
For the momentum component of the pion wf we use
a Gaussian form.
The CQ mass $m=250$ MeV and a parameter
in the exponent
are adjusted to fit $f_{\pi}$ and the pion charge radius \cite{FPPS}.
In this model the pion GPD $H^u(x,\xi=0,t) $ in the range $0 \le x \le 1$ is given by
a diagonal contribution with $n=2$ constituents
\begin{eqnarray} &&
\hspace{-0.8cm}H^u
= \sum_{\{\lambda_i\}}
\int
{d{\bf \kappa}_\perp \over 2(2\pi)^3}~
\Psi_\pi^{*}(x,{\bf
\kappa}^\prime_{\perp};\{\lambda_i\})
\Psi_\pi(x,{\bf \kappa}_{\perp};\{\lambda_i\}) ~~.
\label{eq:overlap}
\end{eqnarray}
Initial and final
transverse components of active quark momenta
in the intrinsic frame are related by
${\bf \kappa^\prime}_{\perp } =
{\bf \kappa}_{\perp } + (1-x) ~ { \Delta}_{\perp}$.
Then at large $|t|$, i.e. at large $|\Delta_{\perp}|$, $H^u(x,\xi=0,t) $
is expected to be non vanishing only for $x \sim 1$.
\end{description}
\section{Pion longitudinal and transverse momentum distributions}
\begin{figure}[t]
\vspace{0.5cm}
\includegraphics[width=7.cm]{U_x.eps}
\caption{ Thin dashed line:
covariant model with the sum-form BSA.
Dotted line: covariant model with the product-form BSA.
Thick dashed line: LFHD model
(after Ref. \cite{FPPS}).}
\label{strucx}
\end{figure}
At $~\xi=0~$ one has $~x~= ~ x_q~$ and for $~t=0~$ one gets from $~H^u~$ the longitudinal
momentum distribution
\begin{eqnarray} &&
\hspace{-0.2cm} u(x) = H^u(x,0,0) = 2 ~ H^{I=1}_\pi(x,0,0)
= \int d{\bf k}_\perp~ f_1(x,k_\perp) \quad \quad (x\ge 0)~ ,
\label{f1xk}
~~\end{eqnarray}
where
$f_1(x,{k}_\perp)$ is the TMD.
The $u(x)$ distributions for our models are compared in Fig. 2.
The covariant sum-form model is unable to give a vanishing value
for $u(x)$ at the end points, while
the product-form model describes
both the ff tail and the end-point fall-off of the parton distribution \cite{FPPS}.
Indeed, the product-form model
has a $k_{\perp}^4$ fall-off of the BSA, compatible
with a BSA kernel dominated by the one-gluon-exchange (OGE) and
faster than the ${k}_{\perp}^2$ decay of the sum-form model,
and both at $~-t$ $\rightarrow \infty~$ and at $~x \rightarrow 0~$ or
at $~x \rightarrow 1~$ the high
momentum part of the pion state is probed.
At $~|\xi|=1~$ and at $~\xi=0~$ the
covariant product-form model exhibits the same general behavior of the other models and
for all of the considered models at high $t$
a maximum of GPD's around $x \sim 1$ occurs \cite{FPPS}.
This last fact has a
simple kinematical explanation at $~\xi=0~$.
At $~\xi=-1~$, where only the $q \bar{q}$ pair production contributes,
one has $x = 2k^+/\Delta^+$, while
large
$|t|$ values mean large $\Delta^+ = \Delta_z
\approx 2 k_{zq}$, and $2 k^+ = k^+_q - k^+_{\bar{q}}
\sim 2 E_q = 2 ~ \sqrt{m^2 + {\bf{k}}^2}$, since each quark in the pair is almost on its mass
shell. Then
$ x \sim E_q / k_{zq} ~~ \rightarrow ~~ 1 ~ .$
As noticed in \cite{FPPS}, also at $|\xi|=x$
the maximum of GPD as $-t ~ \rightarrow ~ \infty$ moves from $~x \sim 0.5~$ towards
$~x = 1$. The GPD
at $|\xi|=x$ allows one to
explore the transition from the valence to the NV region and this kinematical
regime should be relevant to study single spin asymmetry \cite{diehlpr}.
\vspace{-0.cm}
\section{Nucleon
parton momentum distributions}
\label{PaceE_sec:10}
We describe the quark-nucleon vertex function through a BSA,
with a Dirac structure suggested
by an effective Lagrangian \cite{de},
and adopt a Breit reference frame where~$~~{\bf q}_{\perp}=0$ and
$q^+=
|q^2|^{1/2}$. Our CQ mass is ~$m=m_u = m_d = 200 ~ {\rm{MeV}}$.
The
current in the SL region
is approximated {\em{microscopically}} by the Mandelstam formula \cite{mandel}
\vspace{0.0cm}
\begin{eqnarray} &&
\hspace{-0.9cm}\langle \sigma',P_N'|j^\mu~|P_N,\sigma\rangle
=3~N_c
\int {d^4k_1 \over (2\pi)^4}\int {d^4k_2 \over (2\pi)^4}
\sum \left \{ \bar \Phi^{\sigma'}_N(k_1,k_2,k'_3,P_N')
S^{-1}(k_1) S^{-1}(k_2)~{\cal I}^\mu(k_3,q)~
\Phi^\sigma_N(k_1,k_2,k_3,P_N)\right \}
\end{eqnarray}
where
$k'_{3} = k_{3} + q$, ${\cal I}^\mu(k_3,q)$
is the quark-photon vertex, and $\sum$ implies a sum over isospin and
spinor indexes.
The
Mandelstam formula is projected out by an analytic integration on $k_1^-$ and $k_2^-$,
taking into account only the poles of the propagators.
Then the current becomes the sum of a purely valence contribution
and a NV, pair-production contribution.
Clearly, after the $k^-$ integrations,
the vertex functions depend only upon the LF three-momenta.
The quark-photon vertex has IS and IV contributions,
$ {\cal I}^\mu=~{\cal I}^\mu_{IS} +\tau_z
{\cal I}^\mu_{IV}
\label{curr}
$,
and each term contains a purely valence contribution
(in the SL region only) and a contribution corresponding to the pair production (Z-diagram).
In turn the Z-diagram contribution can be decomposed in a bare term $+$ a
VMD term, viz
\begin{eqnarray} &&
\hspace{-0.7 cm} {\cal I}^\mu_{i}(k,q) = {\cal N}_{i} ~ \theta(P_N^+-k^+) ~ \theta(k^+) ~
\gamma^\mu + \theta({q}^+ + k^+)
~
\theta(-k^+)~\left \{ Z_B~{\cal N}_{i} ~ \gamma^\mu+
Z^i_{VM}~\Gamma^\mu(k,q,i)\right\}
\end{eqnarray}
with $i = IS, IV$, ${\cal N}_{IS}=1/6$ and ${\cal N}_{IV}=1/2$. The constants $Z_B$
and
$Z^i_{VM}$
are unknown weights to be extracted from
the phenomenological analysis of the experimental data.
According to the label $i$, the VMD term $\Gamma^\mu(k,q,i)$,
which does not involve free parameters,
includes IV or IS mesons. Indeed in \cite{nucleon}
the microscopic model for the VMD, successfully used in
\cite{DFPS} for the pion ff
and based on the mass operator of Ref. \cite{Fred},
was extended to IS mesons.
\begin{figure}[t]
\vspace{-1.5cm}
\hspace{-.0cm}{\includegraphics[width=6.5cm]{f1pnew.eps}}
\hspace{.7cm}
{\includegraphics[width=6.5cm]{f1nnew.eps}}
\vspace{-1.7cm}
\caption{Left panel: transverse-momentum distributions for a $u$ quark in
the proton. $G(k_{\perp}) = (1 ~ + ~ k_{\perp}^2/m_{\rho}^2)^{-5.5}$,
$m_{\rho}$ = 770 MeV and $k_{perp} = k_{\perp} $. Right panel: the same as in
the left panel, but for a $d$ quark inside
the proton (after Ref. \cite{PDFPS}). }
\end{figure}
In the valence vertexes the spectator quarks are on the
$k^-$-shell, and the BSA momentum dependence
is approximated through a nucleon wf PQCD inspired,
which depend on
the free mass of the three-quark system, $M_0(1,2,3)$,
\begin{eqnarray} &&
\hspace{-0.6cm} \Psi_N(\tilde{k}_1,\tilde{k}_2,P_N) = P_N^+ ~
{ {\Lambda}(k_1,k_2,k_3)|_{(k_{1on}^-,k_{2on}^-)} \over [m_N^2 - M^2_0(1,2,3)]^{~}}
= ~ P_N^+ ~ {\cal {N}}~
{~(9~m^2)^{7/2}
\over (\xi_1\xi_2\xi_3)^{p}~ \left[\beta^2 + M^2_0(1,2,3)\right]^{7/2}} \quad ,
\end{eqnarray}
where $\tilde{k}_i \equiv (k_i^+,{\bf k}_{i\perp})$,
$\xi_i = k^+_i/P_N^+$ ~($i=1,2,3$)
and ${\cal N}$ is a normalization constant.
The power $ 7/2 $ and the parameter $ p = 0.13 $ are chosen to have
an asymptotic decrease of the triangle contribution faster than the dipole.
Only the triangle diagram determines
the magnetic moments, which are weakly dependent on $p$. Then
$\beta = 0.645$ can be fixed by the magnetic moments and we obtain $\mu_p = 2.87\pm0.02$ ~
($\mu_p^{exp}$ = 2.793)
and $\mu_n = -1.85\pm0.02$ ~ ($\mu_n^{exp}$ = -1.913).
For the Z-diagram contribution, the NV vertex
is needed.
It can depend on the free mass, $M_0(1,2)$, of the (1,2) quark pair
and on the free mass, $M_0(N,\bar {3})$, of the ( nucleon - quark $\bar {3}$ ) system
entering the NV vertex.
Then in the SL region we approximate the
momentum dependence of the NV vertex
$ {\Lambda}_{NV}^{SL} = {\Lambda}(k_1,k_2,k_3)|_{k^-_{1on},k^{'-}_{3on}}$ by
${\Lambda}_{NV}^{SL}= [g_{12}]^{2}[g_{N\bar {3}}]^{3/2}
\left [{k_{12}^+ / P_N^{\prime +} } \right ]
\left [ P_N^{\prime +} / k_{\overline {3}}^+ \right ]^r
\left [P_N^{+} / k_{\overline {3}}^+ \right ]^{r}
$,
with
$
k_{12}^+ = k_1^+ + k_2^+ $ and $ g_{AB} = (m_A ~ m_B) /\left
[\beta^2+M^2_0(A,B)\right] $.
The power 2 of $[g_{12}]^{2}$ is suggested from counting rules.
The power 3/2 of $[g_{N\bar {3}}]^{3/2}$ and the parameter $r=0.17$ are chosen
to have an asymptotic dipole behavior for the NV contribution,
as suggested by the OGE dominance.
We performed a fit for the ff's of our free parameters,
$Z_B$, $Z^i_{VM}$, $p$, $r$ in the SL region, obtaining
a $\chi ^2$/datum = 1.7.
The Z-diagram turns out to be essential
in our reference frame
with $q^+ > 0$.
In particular, {\em{the possible zero in $G_E^p/G_M^p$ for $q^2<0$ is strongly related
to the pair-production contribution, i.e. to higher Fock state components}}.
The longitudinal distribution $q(x)$ is the limit in
the forward case of the unpolarized
GPD ${H}^q(x,\xi,t)$.
For $P_N' = P_N$, both $q^+$ and $\xi$ are vanishing and $x = k_3^+ / P_N^+ = \xi_3$
is the fraction of the active quark longitudinal momentum.
As a consequence the function ${H}^q(x,\xi,t)$
reduces to the longitudinal parton
distribution function $q(x)$:
\begin{eqnarray}
\hspace{-1cm} {H}^q(x,0,0) = q(x) = \int d{\bf k}_{\perp} ~~ f_1^q(x,k_{\perp}) ~
= ~ \int \frac{dz^-}{4\pi} e^{i x P^+ z^-}
\left . \langle P_N | \bar \psi_q
(-\frac{z}{2}) \gamma ^+ \, \psi_q (\frac{z}{2}) | P_N \rangle \right |_{\tilde{z}=0} ~~ ,
\label{struc1}
\end{eqnarray}
where
an average on the nucleon helicities is understood.
Once all the parameters of the nucleon light-front wf
$\Psi_N(\tilde k_1,\tilde k_2,P_N)$ have been determined, one can easily define
the TMD of the active quark, $f_1^q(x,k_{\perp})$, in terms of the LF wf
and through Eq. (\ref{struc1})
also the longitudinal distribution of the struck quark. From the isospin symmetry
one has
$~u_p(x)=d_n(x)=u(x)$ and
$d_p(x)=u_n(x)= d(x)$.
Our preliminary results for $f_1^{u(d)}(x,k_{\perp})$ in the proton and for $u(x)$
and $d(x)$ are
shown in Fig. 3 and in Fig. 4, respectively.
It can be observed that the decay of our $f_1$ vs $k_{\perp}$
is faster than in diquark models of the nucleon
\cite{Jacob}, while it is slower than in gaussian factorization
models for the TMD \cite{Anselm}.
As far as the longitudinal distributions are concerned,
a reasonable agreement of our $u(x)$ with the CTEQ4 fit to the experimental data
\cite{Lai}
can be seen in Fig. 4.
\section{Conclusions}
\label{PaceE_con}
Microscopical models for pion
and nucleon em form factors
have been investigated
with good results in the SL region.
The Z-diagram (i.e. higher Fock state component) has been shown to be essential,
in reference
frames where $q^+ \ne 0$.
The analysis of the form factors allows us to get hadron vertexes that are used to
evaluate the unpolarized longitudinal and transverse momentum distributions.
The effects of a $k_{\perp}$ fall-off
compatible with the OGE dominance has been explored. Its role
for the description of the ff tail at high $-t$ and of vanishing
longitudinal parton distributions at the end
points has been shown.
The covariant product-form model is able to reproduce the pion GPD's evaluated
by the
Mandelstam-inspired model at $|\xi|=1$ and by the LFHD model at $\xi=0$. Then
one could argue that the product-form
model contains the main
ingredients for the description of the constituents inside the pion and could be applied
to study
experimental data.
Our next step will be the calculation of polarized longitudinal and transverse
momentum distributions.
\vspace{.2cm}
\begin{figure}[t]
\centering
\vspace{.8cm}
\hspace{.0cm}{\includegraphics[width=6.5cm]{PaceE-fig3.eps}}
\hspace{.7cm}{\includegraphics[width=6.5cm]{PaceE-fig4.eps}}
\vspace{.0cm}
\caption{Left panel: longitudinal momentum distribution for a $u$ quark inside
the proton. Dashed lines: our preliminary results; thick solid lines: our results after
evolution to $Q^2$ = 1.6 (GeV/c)$^2$;
thin solid lines: CTEQ4 fit to the experimental data \cite{Lai}. Right panel:
the same as in the left panel, but for a $d$ quark in
the proton (after Ref. \cite{PDFPS}). }
\end{figure}
\bibliographystyle{aipproc}
|
\section{Introduction}
Diffusion in polymer solutions is among the oldest subjects of polymer physics.
\cite{DeGennes79,doi:86} In general, transport by diffusion can be characterized by two
diffusion coefficients: the self-diffusion coefficient $D_s$ and the cooperative diffusion
coefficient $D_c$. $D_s$ describes the motion of one molecule relative to the
surrounding molecules due to thermal motions while $D_c$ describes the motion
of a number of molecules in a density gradient. \cite{adam:77,Cosgrove83,Kanematsu05,Brown91,LeBon99}
The obvious importance of diffusion in polymer physics has led to a rather large number
of studies of $D_c$ by dynamic light scattering (DLS),
\cite{adam:77,Cosgrove83,Pecora85,Brown91,LeBon99,Min03,Kanematsu05} while $D_s$ can be
obtained by pulsed-field gradient nuclear magnetic resonance
\cite{Cosgrove83,LeBon99,Kanematsu05} and label techniques like forced Rayleigh
scattering \cite{Hervet79} or fluorescence correlation spectroscopy (FCS).
\cite{Zettl04,Liu05,Zettl07} However, in many cases $D_s$ and $D_c$ could not be obtained
for the same homopolymer using the same technique.
Such measurements would be very interesting since a central problem in the dynamics
of semidilute entangled polymer solutions is the quantitative understanding of the
interplay of self-diffusion and cooperative diffusion. Very recently it has been
found theoretically that the coupling of self- and cooperative motion due to
topological constraints is also important for rather stiff macromolecules.
\cite{Bier08}
At present, DLS is certainly among the most accurate methods to measure $D_c$ and there is a number of careful studies conducted on polymer solutions. In principle, FCS is the method of choice for studying diffusion of single macromolecules in a matrix of same molecular weight giving $D_s$ or in a solution of polymers of different molecular weight (tracer diffusion \cite{ribe:04,gian:07,Cher09}). In opposite to DLS, FCS requires chains labeled by a stable fluorescing molecule. Moreover, the number of labels per macromolecules should be constant to arrive at results that can be directly compared to theory. Given these problems, the use of FCS for measurements of $D_s$ on synthetic polymers has been scarce so far. \cite{Zettl04,Liu05,Zettl07,Grab08} Moreover, the full potential of this method has not yet fully been exploited yet since FCS should also allow one to obtain $D_c$. \cite{Ricka89, Scalettar89}
Recently, a well-defined polymeric model system has been presented and used for quantitative FCS-measurements in dilute solution: \cite{Zettl04,Zettl07} Nearly monodisperse polystyrene chains have been prepared by anionic polymerization and subsequently labeled by single fluorescent dye. Since the molecular weight of the different samples span a wide range, these polymers provide a nearly ideal model system for exploring the chain dynamics over a wide range of molecular weights and concentrations. Using these labeled chains, we recently presented an in-depth study of the experimental FCS set-up \cite{Zettl04} as well as of the dynamics in dilute solution. \cite{Zettl07}
Here we pursue these studies further by presenting an investigation of polymer diffusion in the semi-concentrated regime by FCS. In order to obtain accurate data of cooperative diffusion, these studies are combine with DLS-measurements on exactly the same molecular weights and concentrations. Thus, $D_c$ and $D_s$ can now be obtained from identical systems and directly be compared. In the course of these studies we found that a second cooperative mode becomes visible in the FCS-experiments if the concentration exceeds a given value. This surprising finding prompted us to conduct a full theoretical analysis of both the FCS- as well of the DLS-data throughout the entire time scale and range of concentrations available by these experiments. In doing so we extend the theoretical modeling beyond the usual scaling laws. The entire study is devoted to a comprehensive understanding of polymer dynamics in solution ranging from the dilute state up to the onset of glassy dynamics.
The paper is organized as follows: After the Section Experimental we first present the FCS-data together with the finding of the new cooperative diffusion. In the subsequent section a quantitative modeling of the data in terms of an analytical theory will be given. In the last section special attention will be paid to possible practical applications of these findings to the spinning of nanofibers. A Conclusion will wrap up the entire discussion.
\section{Experimental Section}
\subsection{Dye Labeled Polystyrene}
All experiments reported here were carried out with linear polystyrenes having a narrow
molecular weight distribution. For details of the synthesis see ref \cite{Zettl04}.
The molecular weight and polydispersity of the polymers are summarized in Table~\ref{PS}.
The solutions for the FCS experiments were prepared in toluene p.\ a.\ grade by blending
a constant concentration of $10^{-8}\;\mathrm{M}$ Rhodamine~B labeled polystyrene with
varying amounts of unlabeled polystyrene from the same synthesis batch. Each labeled
polymer carries only one dye molecule at one of its ends. To verify our results,
additional solutions were prepared with varying labeled polystyrene and a constant amount
of unlabeled polystyrene. We have used preparative gel permeation
chromatography to separate labeled polymer and free dye molecules. \cite{Zettl04,Zettl06}
Therefore, the resulting dye-labled polymer does not contain any measurable amount of
free dye molecules.
\begin{table}[hb]
\begin{center}
\caption{\label{PS} Molecular weight $M_w$, polydispersity index PDI=$M_w/M_n$ and
hydrodynamic radius $R_h$ at infinite dilution of the polystyrenes used in the
present study. The second and the third virial coefficients $A_2$ and $A_3$, respectively,
have been calculated using scaling laws taken from the literature ($A_2$:
ref \cite{Kniewske83} and $A_3$: ref \cite{Min03}). $c^+$ is the concentration at which
the second diffusion time appears in the FCS measurements.}
\vspace*{0.2cm}
\begin{tabular}{cccccc}
\hline\\[-10pt]
$M_w[\frac{kg}{Mol}]$ & PDI & $R_h[nm]$ & $A_2[\frac{cm^3Mol}{g^2}]$ &
$A_3[\frac{cm^6Mol}{g^3}]$ & $c^+[\mathrm{wt\%}]$ \\[-10pt]\\\hline\hline
11.5& 1.03& 1.(4)& 7.4$\cdot 10^{-4}$& 2.1$\cdot 10^{-3}$& -\\
17.3& 1.03& 1.(6)& 6.8$\cdot 10^{-4}$& 2.6$\cdot 10^{-3}$& -\\
67.0& 1.05& 3.(9)& 5.1$\cdot 10^{-4}$& 5.8$\cdot 10^{-3}$& 20\\
264 & 1.02& 7.(3)& 3.8$\cdot 10^{-4}$& 1.3$\cdot 10^{-2}$& 6.5\\
515 & 1.09& 9.(8)& 3.3$\cdot 10^{-4}$& 1.9$\cdot 10^{-2}$& 4.8\\
\hline
\end{tabular}
\end{center}
\end{table}
\subsection{Methods}
For FCS measurements we modified the commercial ConfoCor2 setup (Carl Zeiss, Jena, Germany)
\cite{Rigler01} with a 40$\times$ Plan Neofluar objective (numerical aperture NA=0.9).
The Rhodamine~B labeled PS-chains were excited by a HeNe-Ion laser at 543~nm. The intensity
for all measurements was $4\mu$W in sample space. As second setup we used a MicroTime200
(PicoQuant, Berlin, Germany) \cite{Wahl04} with a 100$\times$ oil immersion objective
(NA=1.45). Here the detection beam path was divided by a 50/50 beam splitter on two
detectors to crosscorrelate the signals. This crosscorrelation is necessary to prevent
distortion of the fluorescence correlation function by detector afterpulsing. \cite{Enderlein05}
For details of the FCS-measurements see refs \cite{Rigler01,Zettl04,Zettl07}.
Cooperative diffusion coefficients
$D_c$ were measured by DLS using an ALV 4000 light scattering goniometer (Peter, Germany).
\subsection{Evaluation of Data}
In FCS \cite{Magde72, Rigler01} a laser beam is focused by an objective with high
numerical aperture (typically $\geq$~0.9) and excites fluorescent molecules entering the
illuminated observation volume. The emitted fluorescent light is collected by the same
optics and separated from scattered light by a dichroic mirror. The emitted light is
detected by an avalanche photo diode. The time dependent intensity fluctuations $\delta
I(\tau)=I(\tau)-\left\langle I(\tau) \right\rangle$ are analyzed by an autocorrelation function,
where $\left\langle\,\,\, \right\rangle$ denotes an ensemble average. This autocorrelation
function can be written as \cite{Ricka89}
\begin{equation} \label{eq1}
G(\tau)=\frac{1}{N}\int d {\bf q}\,\Omega ({\bf q}) C({\bf q},\tau)
\end{equation}
where \mbox{$\Omega ({\bf q})=\pi^{-\frac{3}{2}}w_{x,y}^2w_z
\exp(\!-w_{x,y}^2 (q_x^2\!+\!q_y^2)/4\!-\!w_z^2 q_z^2/4)$}
is a Gaussian filter function characterizing the observation volume in Fourier space
with $\int d {\bf q}\, \Omega ({\bf q})=1$, $N$ is the average number of fluorescently
labeled molecules in the observation volume, and ${\bf q}=(q_x, q_y, q_z)$. Here
$w_{x,y}=296$ nm is the dimension of the observation volume perpendicular to the
optical axis and $w_z=8 w_{x,y}$ is the dimension along the optical axis.
\cite{Zettl04,Zettl07} For an ideal gas consisting of non-interacting molecules the
initial amplitude reduces to the familiar relationship $G(0)=1/N$. \cite{Rigler01}
The time-dependent fluorescence density-density autocorrelation function
$C({\bf q},\tau)$ is expressed in terms of a coupled-mode model \cite{Pusey82,Akcasu:91}
as
\begin{equation} \label{eq2}
C({\bf q},\tau)=\frac{C_c(q,0)e^{-q^2 \phi_c(\tau)/6}+C_s(q,0)e^{-q^2 \phi_s(\tau)/6}}
{C_c(q,0)+C_s(q,0)}
\end{equation}
where $q=|{\bf q}|$. Here the mean square displacements $\phi_c(\tau)$ and $\phi_s(\tau)$
are given by
\begin{eqnarray} \label{eq3}
\phi_c(\tau)&=&6 D_c \tau\,,
\\\phi_s(\tau)&=&6 D_s \tau+ B_s(\tau)\,. \label{eq4}
\end{eqnarray}
The term $B_s(\tau)$ allows one to take into account the contributions from internal
polymer chain motions. \cite{doi:86} If only a few of the molecules are fluorescently
labeled, the self-diffusion coefficient $D_s$ can be measured in the FCS experiment.
\cite{Zettl07} If all of the molecules are fluorescently labeled, the cooperative diffusion
coefficient $D_c$ can be obtained. \cite{Scalettar89} In the case that neither of these
limits applies, both the self mode and the cooperative mode will be present in the spectrum
of the autocorrelation function. The diffusion coefficients can be extracted by fitting
\begin{eqnarray}
\lefteqn{G(\tau)=}\nonumber
\\&&\sum_{i \in \{s,c\}} G_i(0)
\left(1+ \frac{2 \phi_i(\tau)}{3 w_{x,y}^2}\right)^{-1}
\left(1+ \frac{2 \phi_i(\tau)}{3 w_{z}^2}\right)^{-1/2}
\nonumber
\\&& \label{eq5}
\end{eqnarray}
to the experimental data. FCS is not only sensitive to intensity fluctuations due
to the motion of labled molecules but also due to photokinetic processes of the
fluorescent dyes which occur for short times $\tau \lesssim 5\times 10^{-3}$ ms.
This additional relaxation has been taken into account as discussed in refs
\cite{Zettl04,Zettl06,Zettl07}.
DLS allows one to measure the time dependent autocorrelation function of the scattered
electric field which can be expressed in terms of the elements of the fluid
polarizability tensor. \cite{Pecora85} For an incident light wave traveling in the
$x$ direction with a polarization vector in the $z$ direction the intensity of the
scattered electric field can be written as
\begin{eqnarray}
I_{VV}({\bf q},\tau)&\sim&\int d{\bf r}\,d{\bf r}'\,
\left\langle \alpha_{zz}({\bf r}+{\bf r}',\tau) \alpha_{zz}({\bf r}',0)\right\rangle
e^{i{\bf q}\cdot{\bf r}}\,,\nonumber
\\&& \label{eq5a}
\end{eqnarray}
where the absolute value of the scattering vector ${\bf q}$ is given by
$q=|{\bf q}|=(4\pi n/\lambda)\sin(\theta/2)$ in which $n$ is the refractive index
of the medium. $\lambda$ is the incident wavelength and $\theta$ is the scattering
angle. The $zz$ element of the fluid polarizability tensor is denoted as
$\alpha_{zz}({\bf r},\tau)$. The experimentally accessible quantity is the intensity autocorrelation function $g^{(2)}_{VV}({\bf q},\tau)$. For photon counts
obeying Gaussian statistics, the intensity autocorrelation function is related to
the electric field autocorrelation function $g^{(1)}_{VV}({\bf q},\tau)$ according to
\begin{eqnarray} \label{eq5b}
g^{(2)}_{VV}({\bf q},\tau)&=&1+f_{VV} \left(g^{(1)}_{VV}({\bf q},\tau)\right)^2\,,
\end{eqnarray}
where $f_{VV}$ is dependent on the scattering geometry. The electric field correlation
function can be calculated for various systems. For a solution containing purely
diffusing particles the electric field correlation function is given by
$g^{(1)}_{VV}(q,\tau)=\exp(-q^2 D_c \tau)/\sqrt{f_{VV}}$.
\section{Diffusion coefficients measured by FCS}
\begin{figure}[hb]
\begin{center}
\includegraphics[width=7cm]{FCS_fig1.eps}
\caption{Normalized autocorrelation function obtained from FCS for polystyrene of
molecular weight $M_w=67~\mathrm{kg/Mol}$ in toluene for various
polymer concentrations: 0.03~wt\% (-- --), 9.1~wt\% (-- $\cdot$), 20~wt\% (-- $\cdot\cdot$)
and 28~wt\% (---). A second diffusion time appears at 20~wt\% on a shorter timescale compared
to self-diffusion. The thick solid line is the normalized crosscorrelation curve without
detector afterpulsing for the 28~wt\% polymer solution. The dotted vertical line marks the
time scale above which this artefact becomes negligible, i.e., the solid thin and thick lines
coincide for $\tau>0.01$ ms.}
\label{AC}
\end{center}
\end{figure}
Figure \ref{AC} shows normalized autocorrelation functions measured by FCS. The average number of
labeled polymers in the observation volume was kept constant to $N\approx0.8$ whereas the number of unlabeled polymers
increases up to $N_u=3 \times 10^6$ for the \mbox{28 wt~\%} polymer solution. The thin broken
curves are measured at the ConfoCor2 setup and the thick solid curve is measured at the
MicroTime200 setup. The curves obtained at the ConfoCor2 setup have an additional decay on
the time scale less than 10~$\mu s$. This additional decay belongs to detector afterpulsing.
Hence, the evaluation of the correlation curves has been done only for $\tau\ge$ 10~$\mu s$
as indicated by the dotted line in Figure \ref{AC}. For low polymer concentrations we obtained
correlation curves with a single diffusion time. With increasing polymer concentration the
correlation curves shift to higher diffusion times.
As an entirely new finding, Figure \ref{AC} presents a new mode related to a second diffusion time measured with
FCS at higher polymer concentrations. This second diffusion time appears at shorter
time scales than the one related to self-diffusion. The concentration $c^+$ at which the second
diffusion time is detected depends on the molecular weight: The higher the molecular weight,
the lower is $c^+$ (see Table \ref{PS}). In general $c^+$ is about 15$\times$ the overlap
concentration determined in an earlier study. \cite{Zettl07} For the concentration $c^+$
the ratio between these two diffusion times is in the range of 60. From both diffusion times
we calculated the diffusion coefficients from the relations given above.
\begin{figure}[ht]
\begin{center}
\includegraphics[width=7cm]{FCS_fig2.eps}
\caption{Comparison of self-diffusion coefficients ($D_s$, $\bullet$) with cooperative
diffusion coefficients ($D_c$, $\lozenge$) for different molecular weights: $M_w=$ 11 and
17 kg/Mol (top and bottom). Open and solid symbols refer to DLS and FCS measurements,
respectively. The solid lines represent $D_s$ calculated according to
eq \ref{eq6} with $D_c$ as input from DLS measurements. The dashed lines represent $D_c$
calculated vice versa, i.e., with $D_s$ as input from FCS experiments. Insets: Measured
ratio $D_c/D_s$ (symbols) together with the corresponding ratio obtained from
eqs \ref{eq6} and \ref{eq7} within a third order virial approximation (see Table \ref{PS}).}
\label{DSversusCshortMW}
\end{center}
\end{figure}
\begin{figure}[ht]
\begin{center}
\includegraphics[width=7cm]{FCS_fig3.eps}
\caption{Comparison of self-diffusion coefficients ($D_s$, $\bullet$) with
cooperative diffusion coefficients ($D_c$, $\blacklozenge$, $\lozenge$)
for different molecular weights: $M_w=$ 67, 264 and 515 kg/Mol (from top to bottom). Open
and solid symbols refer to DLS and FCS measurements, respectively. The solid lines represent
$D_s$ calculated according to eq \ref{eq6} with $D_c$ as input from DLS measurements.
The dashed lines represent $D_c$ calculated vice versa, i.e., with $D_s$ as input from
FCS experiments. For comparison the dotted lines represents the scaling prediction
$D_s \sim M_w^{-2} c^{-7/4}$ for long polymer chains in the semidilute entangled regime
(see eq \ref{eq11}). Insets: Measured ratio $D_c/D_s$ (symbols) together with the
corresponding ratio obtained from eqs \ref{eq6} and \ref{eq7} within a third order virial
approximation (see Table \ref{PS}).}
\label{DSversusClargeMW}
\end{center}
\end{figure}
In Figures \ref{DSversusCshortMW} and \ref{DSversusClargeMW} all diffusion
coefficients measured with FCS and DLS are compared at identical conditions.
At infinite dilution both diffusion coefficients $D_s$ and $D_c$ have the
same value. In dilute solutions $D_s$ and $D_c$ show a linear dependency on the concentration
as expected according to the Kirkwood-Riseman theory. \cite{Kirkwood48} But $D_s$ decreases
whereas $D_c$ increases with increasing polymer concentration. The decrease of $D_s$ is due
to the friction between the chains and the increase of $D_c$ is due to the increasing osmotic
pressure. \cite{Fujita90,Kim86} At high concentrations $D_c$ exhibits a maximum.
The insets in Figure \ref{DSversusCshortMW} and Figure \ref{DSversusClargeMW} show the ratio ${D_c}/{D_s}$ of measured values. The lines are theoretical values calculated according to \cite{Kanematsu05,LeBon99}
\begin{equation} \label{eq6}
\frac{D_c}{D_s} = \left(1-\bar{v} c\right) \frac{d \Pi}{d c}
\end{equation}
with the partial specific volume of the polymer $\bar{v}$ and the polymer concentration
$c$. The dependence of the osmotic pressure on $c$ can approximated by a virial expansion
\begin{equation} \label{eq7}
\frac{d \Pi}{d c}= 1 + 2A_2 M_w c + 3A_3 M_w c^2 + \ldots\, ,
\end{equation}
where $A_2$ and $A_3$ are the second and third virial coefficients, respectively, and $M_w$
is the molecular weight. For the calculation of $d \Pi\,/dc$ we used the corresponding values
from the literature gathered in Table \ref{PS} and \mbox{$\bar{v}=0.916$ cm$^3$/g}. \cite{Schulz57}
The measured and the calculated ratio are well described as demonstrated by the inset of
Figures \ref{DSversusCshortMW} and \ref{DSversusClargeMW}. The self-diffusion coefficients
$D_s$ can be determined from the cooperative diffusion coefficient $D_c$ obtained by DLS
measurements and vice versa. $D_s$ and $D_c$ can be measured with high accuracy by FCS
and DLS using the same polymers. Their relation is fully understood in terms of eq \ref{eq6}.
For comparison we note that both the molecular dye diffusion coefficient and
the macromolecular tracer diffusion coefficient decrease with increasing
concentration of the matrix polymer. \cite{Cher09}
\begin{figure}[ht]
\begin{center}
\includegraphics[width=7cm]{FCS_fig4.eps}
\caption{Amplitudes $G_s(0)$ ($\bullet$) and $G_c(0)$ ($\square$)
extrapolated from the measured FCS-autocorrelation function $G(\tau)$ as a
function of labeled molecules $N$ for polystyrene with $M_w=67~\mathrm{kg/Mol}$ at 20~wt\%.
For the self-diffusion $G_s(0)\propto 1/N$ (-- --), while $G_c(0)$ exhibits a linear dependence
on $N$ (---) for the cooperative diffusion.}
\label{fig4}
\end{center}
\end{figure}
Figure \ref{fig4} displays the amplitudes $G_i(0)$ (see eq \ref{eq5}) as a function
of $N$ for polystyrene with $M_w=67~\mathrm{kg/Mol}$ at 20~wt\%. The amplitude of the
self-diffusion mode $G_s(0)$ is proportional to $1/N$. In the presence of non-correlated
background signal (scattering, afterpulsing, electronic noise) this is modified to $1/N -
2b/N^2$. \cite{Rigler01} Here $b$ is proportional to the noise
intensity, which is assumed to be significantly smaller than the fluorescence signal.
For the cooperative mode one finds an amplitude scaling of $1 - 2bN$.
For sufficiently small $b$, this will yield a dependence as shown by Figure \ref{fig4} for
the fast correlation component.
The ratio $G_c(0)/G(0)$ is a non-monotonic function of the concentration for a fixed
number of labeled molecules $N$. It increases form 0 to a value below 1 at the concentration
$c^+$. $G_c(0)/G(0)$ slightly decreases upon further increasing the concentration in
the semidilute entangled regime. Finally, it increases upon approaching the glass
transition concentration.
\section{Scaling theory and Langevin equation approach}
In the following section, the findings presented in the previous sections will be compared
to current models of polymer diffusion.
\subsection{Scaling theory and reptation model}
The application of scaling theory and the reptation model to polymer solutions has
been presented in various treatises (see, e.g., refs \cite{DeGennes79,doi:86,lodg:90,mcle:02}).
Hence, we only discuss the equations necessary for this study. Three concentration regimes
can be distinguished: dilute, semidilute unentangled, and semidilute entangled solutions.
Scaling arguments and the reptation model lead to following relations for the self-diffusion
coefficient $D_s$ and the cooperative diffusion coefficient $D_c$:
\begin{eqnarray}
D_s &=& D_c \sim M_w^{-3/5}\, c^{0} \,,\hspace*{0.5cm} c \ll c^*\,,
\label{eq9}
\\D_c &\sim & M_w^0\, c^{3/4} \,,\hspace*{0.5cm} c > c^*\,,
\label{eq10}
\\D_s &\sim & M_w^{-2}\, c^{-7/4} \,,\hspace*{0.5cm} c > c^{**}\,.
\label{eq11}
\end{eqnarray}
Here the overlap concentration $c^*$ is the boundary concentration between the dilute
and semidilute regimes. This concentration depends on molecular weight as
\begin{equation} \label{eq8}
c^* \sim M_w^{1-3\nu}=M_w^{-4/5}\,,
\end{equation}
where the Flory exponent $\nu=3/5$ for a good solvent has been used. The crossover
concentration from the semidilute unentangled to the semidilute entangled regime is
denoted as $c^{**}$.
For very low concentrations in the dilute regime, the self-diffusion coefficient
is indistinguishable from the cooperative diffusion coefficient as is apparent
from Figures \ref{DSversusCshortMW} and \ref{DSversusClargeMW}.
In Figure \ref{fig5} the self-diffusion coefficient is plotted as a function of
the molecular weight $M_w$ for a fixed concentration \mbox{$c=9.1$ wt \%}. The
experimental data (solid squares) follow the scaling laws given by eq \ref{eq9}
(dashed line) and eq \ref{eq11} (solid line) for \mbox{$M_w \le 20$ kg/Mol} and
\mbox{$M_w \ge 264$ kg/Mol}, respectively. Moreover, $D_s$ is rather independent
of concentration for \mbox{$c \lesssim 10$ wt \%} in the case of the low
molecular weight solution (see Figure \ref{DSversusCshortMW} and eq \ref{eq9}).
The concentration dependence of $D_s$ of the higher molecular weight solutions
(\mbox{$M_w \ge 264$ kg/Mol}) is in accord with the scaling prediction for the
reptation model (eq \ref{eq11}) which is represented in Figure \ref{DSversusClargeMW}
by the dotted lines. Hence the FCS measurements verify the basic scaling and reptation
theory for semidilute entangled polymer solutions similar to earlier forced
Rayleigh scattering experiments of polystyrene in benzene. \cite{Hervet79,lege:81}
\begin{figure}[ht]
\begin{center}
\includegraphics[clip=,width=7cm]{FCS_fig5.eps}
\caption{The self-diffusion coefficient $D_s$ ($\bullet$) measured by FCS at the fixed
concentration \mbox{$c=9.1$ wt \%} as a function of the molecular weight $M_w$.
The dashed and solid lines of slope $M_w^{-3/5}$ (see eq \ref{eq9}) and
$M_w^{-2}$ (see eq \ref{eq11}), respectively, represent two asymptotic scaling
regimes.}
\label{fig5}
\end{center}
\end{figure}
In the limit $c \to 0$ the experimental data follow the scaling law given by
eq \ref{eq9} irrespective of the molecular weight, \cite{Zettl07} i.e.,
also the higher molecular weight PS solutions obey the scaling relation
$D_s \sim M_w^{-3/5}\, c^{0}$ . This result is in agreement with earlier
quasi-elastic light scattering experiments for polystyrene in 2-butanone
\cite{king:73} or in benzene. \cite{adam:77}
\subsection{Internal motions of chains}
In order to examine the influence of internal chain motions such as bending
and stretching on the dynamics (see refs \cite{harn:95,harn:98,harn:99a,harn:99b}
and references therein), one may trace out the internal degrees
of freedom of a polymer chain by studying the monomer mean square displacement
$B_s(\tau)$ in eq \ref{eq4} in detail. Various theoretical predictions
on the time dependence of the monomer mean square displacement
of both continuously and single labeled DNA molecules in aqueous solution have been
verified using FCS measurements. \cite{lumm:03,shus:04,gros:06,petr:06,wink:06,shus:08}
In these earlier experimental and theoretical studies the $\Theta$ condition
has been considered. However, for PS in toluene solutions the intramolecular excluded
volume interaction has to be taken into account. In this case scaling arguments
\cite{krem:84,paul:91} lead to the following time dependence of the monomer mean
square displacement:
\begin{eqnarray}
B_s(\tau) = B_s \tau^{1/(1+1/(2\nu))} = B_s \tau^{6/11}\,.
\label{eq12}
\end{eqnarray}
It proves convenient to consider the function \mbox{$1/G(\tau)-1$}, which amplifies
the time dependence of $G(\tau)$ for small times, because $w_z^2=64 w_{x,y}^2$
in eq \ref{eq5}. \cite{wink:06} If the autocorrelation function $G(\tau)$ exhibits a
time dependence according to eqs \ref{eq4}, \ref{eq5}, and \ref{eq12} with $G_c(0)=0$,
a double logarithmic plot will directly yield the exponent $1/(1+1/(2\nu))$ for small
times provided the intramolecular dynamics dominates, i.e., $B_s(\tau) >> 6 D_s \tau$.
Figure \ref{fig6} shows such a representation of the autocorrelation function for
the 515 kg/Mol PS chains in dilute solution. The experimental data (solid squares) follow
the scaling law given by eq \ref{eq12} (dotted line) and the diffusive behavior
(lower dashed line) for short and large times, respectively. Hence for short times
the decay of the autocorrelation function is dominated by intramolecular chain relaxations,
while self-diffusion dominates for large times. Figure \ref{fig6} demonstrates
that the measured autocorrelation function agrees with the calculated
results (solid line) obtained from eqs \ref{eq4}, \ref{eq5}, and \ref{eq12}
with $D_s$ and $B_s$ as input. The mean displacements $\sqrt{\phi_s(\tau)}$ as calculated
from eq \ref{eq5} with $G_s(0)=1$ and $G_c(0)=0$ are given by 131 nm and 598 nm
for $\tau=0.01$ ms and $\tau=1$ ms, respectively.
It is apparent from Figure \ref{fig6} that the contribution of internal chain motions
cannot be observed in the case of the 17 kg/Mol PS chains in dilute solution (solid
triangles) because of the dominating diffusive motion (upper dashed line). The
self-diffusion coefficient $D_s$ increases upon decreasing molecular weight according to
eq \ref{eq9}, while $B_s$ is less dependent on molecular weight. Finally, it is
worthwhile to mention the contribution of internal chain motions to the dynamics decreases
upon increasing the polymer concentration because of the presence of the surrounding
polymer chains. \cite{krem:84,harn:99c}
\begin{figure}[ht!]
\begin{center}
\includegraphics[clip=,width=7cm]{FCS_fig6.eps}
\caption{The autocorrelation function $1/G(\tau)-1$ of a 515 kg/Mol ($\blacksquare$)
and a 17 kg/Mol ($\blacktriangle$) polystyrene solution measured by FCS in the limit
\mbox{$c \to 0$} as a function of the time $\tau$. The dotted and dashed lines of
slope $\tau^{6/11}$ (see eq \ref{eq12}) and $\tau$ (see eq \ref{eq4}), respectively,
represent two asymptotic scaling regimes. The solid line displays the result
for the 515 kg/Mol polystyrene solution as obtained from eq \ref{eq5} with
eqs \ref{eq4} and \ref{eq12} as input. The autocorrelation function of the 17 kg/Mol
polystyrene solution ($\blacktriangle$ and upper dashed line) is shifted up by a factor
of 2.}
\label{fig6}
\end{center}
\end{figure}
\subsection{Cooperative diffusion}
We now turn our attention to the scaling law for the cooperative diffusion coefficient
given by eq \ref{eq10}. Figure \ref{fig7} displays the cooperative diffusion
coefficient $D_c$ of the 515 kg/Mol PS solution (solid squares) together with the
scaling law (dashed line) as a function of the concentration. Several experimental
measurements have yielded the concentration dependence $D_c \sim c^{0.65}$ instead
of the scaling prediction $D_c \sim c^{3/4} = c^{0.75}$.
\cite{adam:77,wilt:84,nemo:84,zhan:99,rauc:03} Various possible explanations for
these deviations from the scaling law have been discussed, \cite{geis:78,brow:90}
such as the countermotion of the solvent induced by the motion of the polymers.
On the basis of our results shown in Figure \ref{fig7} we note that the transition
between the dilute regime with $D_c \sim c^0$ (dotted line and eq \ref{eq9}) and
the semidilute unentangled regime with $D_c \sim c^{3/4}$ (dashed line and
eq \ref{eq10}) is not so abrupt, as has been assumed by scaling theories, but is
a rather smooth crossover that extends over more than one order in magnitude of
concentration.
It has been emphasized that it would be desirable to model
the dynamics both in the dilute regime and the semidilute regimes explicitly within
one theoretical approach. \cite{lodg:90} Successful models should incorporate
the transition region between the dilute regime and the semidilute regimes.
In the next subsection we provide a quantitative basis for such a modelling of
cooperative dynamical properties of polymer chains in good solution.
\begin{figure}[ht!]
\vspace*{0.5cm}
\begin{center}
\includegraphics[clip=,width=7cm]{FCS_fig7.eps}
\caption{The normalized cooperative diffusion coefficient $D_c$ ($\blacksquare$; DLS)
of a 515 kg/Mol polystyrene solution as a function of the concentration $c$.
The dashed and dotted lines of slope $c^{3/4}$ (see eq \ref{eq10}) and
$c^0$ (see eq \ref{eq9}), respectively, represent two asymptotic scaling
regimes. The solid line displays the results as obtained from the Langevin
and generalized Ornstein-Zernike equation according to eqs \ref{eq15} -
\ref{eq21}. The arrow marks the location of the concentration
\mbox{$c^+=0.044$ g/ml} at which the cooperative diffusion mode appears in
the FCS measurements (see Figure \ref{DSversusClargeMW}).}
\label{fig7}
\end{center}
\end{figure}
\subsection{Analytical theory: Langevin and generalized Ornstein-Zernike equation}
We consider a monodisperse polymer solution consisting of $N_{tot}=N+N_u$ polymer
chains and the solvent. Each polymer chain carries $n$ scattering units.
The total dynamic scattering function $S_{tot}(q,\phi,\tau)$ is defined as
\begin{eqnarray} \label{eq13}
S_{tot}(q,\phi,\tau)&=&\frac{1}{N_{tot} \, n^2}\left\langle
\sum\limits_{\alpha, \gamma=1}^{N_{tot}}
\sum\limits_{j,k=1}^{n}
e^{i{\bf q}\cdot \left({\bf r}_{\alpha j}(\tau)-{\bf r}_{\gamma k}(0)\right)}
\right\rangle\,,\nonumber
\\&&
\end{eqnarray}
where $q=|{\bf q}|$ is the magnitude of the scattering vector ${\bf q}$ and
$\langle \,\,\,\, \rangle_\phi$ denotes an ensemble average for a given polymer
volume fraction $\phi$. Here ${\bf r}_{\alpha j}(\tau)$ is the position vector
of the $j$-th scattering unit ($1 \le j \le n$) of the $\alpha$-th
particle ($1 \le \alpha \le N_{tot}$) at time $\tau$. The normalized total
dynamic scattering function is related to the electric field autocorrelation
function measured by DLS according to
$S_{tot}(q,\phi,\tau)/S_{tot}(q,\phi,0)=g^{(1)}_{VV}(q,\tau) \sqrt{f_{VV}}$. (see eq \ref{eq5b}).
The time evolution of the total dynamic scattering function is assumed to be governed
by the Langevin equation \cite{doi:86}
\begin{equation} \label{eq14}
\frac{d}{d\tau} S_{tot}(q,\phi,\tau)=-\Gamma(q,\phi) S_{tot}(q,\phi,\tau)\,.
\end{equation}
The validity of this equation is not obvious since entanglements have
not been taken into account in the derivation of this equation. \cite{doi:86}
However, the short time-scale dynamics can be described by eq \ref{eq14}
since the topological constraints are not so important in the short time-scale
dynamics as is apparent from the fact that the cooperative diffusion
coefficient $D_c$ is considerably larger than the self-diffusion coefficient
$D_s$ in the semidilute entangled regime (see Figure \ref{DSversusClargeMW}).
The decay rate $\Gamma(q,\phi)$ is given by \cite{doi:86}
\begin{eqnarray}
\lefteqn{\Gamma(q,\phi)=}\nonumber
\\&&\frac{k_BT}{4\pi^2 \eta}\int\limits_0^\infty
d q_1\, q_1^2 \frac{S_{tot}(q_1,\phi,0)}{S_{tot}(q,\phi,0)}
\left(\frac{q_1^2+q^2}{2q_1q}
\log\left|\frac{q_1+q}{q_1-q}\right|-1\right)\,,\nonumber
\\&& \label{eq15}
\end{eqnarray}
where the temperature $T$ and the viscosity $\eta$ characterize the solvent.
The volume fraction-dependent cooperative diffusion coefficient $D_c(\phi)$ can
be calculated according to
\begin{equation} \label{eq16}
D_c(\phi)=\lim_{q\to 0}\frac{\Gamma(q,\phi)}{q^2}\,.
\end{equation}
Furthermore, the total static scattering function reads
\begin{equation} \label{eq17}
S_{tot}(q,\phi,0)=1+\phi h(q,\phi)/(V_p P(q,\phi))\,,
\end{equation}
where $V_p$ is the volume of a dissolved polymer chain and $h(q,\phi)$
is a particle-averaged total correlation function. The particle-averaged
intramolecular correlation function
\begin{eqnarray}
P(q,\phi)&=&\frac{1}{N_{tot} \, n^2}\left\langle
\sum\limits_{\alpha=1}^{N_{tot}}
\sum\limits_{j,k=1}^{n}
e^{i{\bf q}\cdot \left({\bf r}_{\alpha j}(0)-{\bf r}_{\alpha k}(0)\right)}
\right\rangle\,,\nonumber
\\&& \label{eq18}
\end{eqnarray}
characterizes the geometric shape of the polymer chains at a given volume
fraction $\phi$. While the particle-averaged intramolecular correlation function
accounts for the interference of radiation scattered from different parts of the same
polymer chain in a scattering experiment, the local order in the fluid is characterized by
$h(q,\phi)$. The particle-averaged total correlation function is related to a
particle-averaged direct correlation function $c(q,\phi)$ by the generalized
Ornstein-Zernike equation of the Polymer Reference Interaction Site Model (PRISM),
which reads (see refs \cite{schw:97,harn:08,yeth:09} and references therein)
\begin{equation} \label{eq19}
h(q,\phi)=P^2(q,\phi)c(q,\phi)/(1-\phi c(q,\phi)P(q,\phi)/V_p)\,.
\end{equation}
This generalized Ornstein-Zernike equation must be supplemented by a closure
relation. If the interaction sites are simply the centers of exclusion spheres,
to account for steric effects, a convenient closure is the Percus-Yevick
approximation. \cite{schw:97} The PRISM integral equation theory has been
successfully applied to various experimental systems such as polymers,
\cite{schw:97,harn:01a} bottle-brush polymers, \cite{boli:07,boli:09}
rigid dendrimers, \cite{rose:06,harn:07} and charged colloids.
\cite{yeth:96,yeth:97,shew:98,harn:00,harn:01,harn:02,webe:07,henz:08}
The overall size of the polymer chains is reduced considerably upon increasing
the volume fraction implying a concentration dependence of the particle-averaged
intramolecular correlation function $P(q,\phi)$. Therefore, we consider
the following particle-averaged intramolecular correlation function \cite{fuch:97}
\begin{equation} \label{eq20}
P(q,\phi)=\left(1+0.549\, q^2 r_g^2(\phi)\right)^{-5/6}
\end{equation}
with the volume fraction dependent radius of gyration
\begin{eqnarray} \label{eq21}
r_g^2(\phi)&=&
\left\{\begin{array}{c@{\quad,\quad}l}
r_g^2(0) & c < c^*
\\r_g^2(0)\left(\frac{\displaystyle c}{\displaystyle c^*}\right)^{-1/8}
&c > c^*
\end{array}\right.\,.
\end{eqnarray}
Here the relation between the volume fraction $\phi$ and the concentration $c$
is given by $\phi= \bar{v} c$, where \mbox{$\bar{v}=0.916$ cm$^3$/g} is the
specific weight of PS. \cite{Schulz57} The scaling law given by eq \ref{eq21}
has been confirmed experimentally for PS in a good solvent using small angle
neutron scattering. \cite{daou:75}
Figure \ref{fig7} demonstrates that the measured cooperative diffusion coefficients
(solid squares) agree with the calculated results (solid line) obtained from
eqs \ref{eq15} - \ref{eq21} both in the dilute and semidilute regimes. In particular,
the features of the broad crossover region between the dilute and the semidilute
regimes are captured correctly by the integral equation theory. In the calculations
the model parameter \mbox{$c^*=0.0032$ g/ml$\,$} \cite{Zettl07} and
\mbox{$r_g(0)=32.8$ nm} for the 515 kg/Mol PS solution has been used. This radius of
gyration is about \mbox{6 \%} larger than corresponding radii of gyration of PS
in various good solvents. \cite{Kniewske83,fett:94,Min03,tera:04} The deviation between
the radius of gyration used in the calculations and the radii of gyration
reported in the literature might be due to the fact that the hydrodynamic interaction
has been taken into account in terms of the Oseen tensor in order to derive
eq \ref{eq15}. Using the Rotne-Prager tensor \cite{rotn:69,harn:96}
as a first correction to the Oseen tensor will improve the results. Moreover, the
size polydispersity $M_w/M_n=1.09$ of the 515 kg/Mol PS solution leads to a diffusion
coefficient which is characteristic for monodisperse polymers of larger radius of
gyration. \cite{harn:99}
Finally, we note that the maximum of $D_c$ in the semidilute entangled regime marks the
onset of glassy dynamics which is discussed in ref \cite{rauc:03}. The friction-controlled
dynamics in this concentration regime is not captured by eqs \ref{eq14} and \ref{eq15}
and will be discussed in subsection \ref{glassydynamic}.
\subsection{Coupling of cooperative fluctuations with single polymer chain motion}
In the following we shall discuss the equation of motion which determines the dynamics
an individual polymer chain. The PS chains are linear chain molecules which are described
by a chain model for macromolecules. \cite{harn:95,harn:96,harn:98} We consider a continuous,
differentiable space curve ${\bf r}(s,\tau)$, where $s \in[-L/2,L/2]$ is a coordinate along the
macromolecule and ${\bf r}(L/2,\tau)$ is the position vector of the labeled end monomer. The
Langevin equation of motion including hydrodynamic interaction is given by \cite{harn:96}
\begin{eqnarray} \label{eq22}
3\pi\eta\frac{\partial}{\partial \tau}{\bf r}(s,\tau)&=&\int\limits^{L/2}_{-L/2}
ds'\,\left(3\pi\eta H(s-s')+\delta(s-s')\right) \nonumber
\\&\times&\!\!\!\left(O(s') {\bf r}(s',\tau) + {\bf f}(s',\tau)\right) + {\bf F}(s,\tau),
\end{eqnarray}
with
\begin{eqnarray} \label{eq23}
O(s)&=&3k_BT p\frac{\partial^2}{\partial s^2} - \frac{3k_BT}{4p}\frac{\partial^4}
{\partial s^4}\,.
\end{eqnarray}
Here $1/(2p)$ is the persistence length, $H(s-s')$ is the hydrodynamic interaction
tensor, and ${\bf f}(s,\tau)$ is the stochastic force. The force ${\bf F}(s,\tau)$
describes the influence of intermolecular forces and is discussed below. The numerical
solution of eq \ref{eq22} allows one to calculate the mean square displacement
(see eq \ref{eq4}) according to
\begin{eqnarray} \label{eq25}
\phi_s(\tau)=\left\langle \left({\bf r}(L/2,\tau)-{\bf r}(L/2,0)\right)^2\right\rangle\,.
\end{eqnarray}
This chain model has been used in the limit \mbox{${\bf F}(s,\tau)=0$} in order to
describe FCS measurements of DNA molecules in dilute solution.
\cite{gros:06,petr:06,wink:06} In particular, the model predicts the observed crossover
from subdiffusive motion \mbox{($B_s(\tau)$ in eq \ref{eq4})} to diffusive motion
\mbox{($6 D_s \tau$ in eq \ref{eq4})} upon increasing the time $\tau$.
Moreover, it has been shown that the chain ends are more mobile than the central
part of the polymer chain for short times. \cite{wink:06} For comparison
we note that the quantity $\phi_s(\tau)$ contributes to the so called incoherent
dynamic structure factor which is accessible by quasielastic neutron
scattering (see ref \cite{harn:97} and references therein).
The key physics determining the dynamics of chain molecules in semidilute entangled solution
arises from the intermolecular interaction which are taken into account in terms of the
force ${\bf F}(s,\tau)$ in eq \ref{eq22}. Various expressions for the force ${\bf F}(s,\tau)$
have been proposed
(see, e.g., refs \cite{schw:89a,seme:91,genz:94,schw:97b,guen:99,seme:98,altu:04,pokr:06,pokr:08}).
These earlier theoretical considerations have demonstrated the coupling of cooperative
fluctuations with single polymer chain motion in the semidilute entangled regime.
This coupling allows one to measure $D_c$ from the dynamics of individual labeled
polymer chains with FCS. Hence, it provides the explanation for the finding of a cooperative
mode in the FCS-experiment. The topological interaction in semidilute entangled polymer
solutions seriously affects dynamical properties since it imposes constraints on the
motion of the polymers. When the motion of a single polymer chain is partly hindered by
the presence of other chains the cooperative diffusion becomes highly correlated and can
be studied using only a small fraction of labeled molecules. Moreover, the number of molecules
statistically involved in the correlated dynamics increases considerably upon approaching
the glass transition concentration.
Figures \ref{fig8} (a) and (b) display the function $1/G(\tau)-1$ for the 17 kg/Mol PS
chains and the 515 kg/Mol PS chains in dilute solution (solid squares, \mbox{$c \to 0$})
and in semidilute solution (solid triangles, \mbox{$c = 13 $ wt \%}). For the
17 kg/Mol PS chains only self-diffusion can be measured using FCS irrespective of
the concentration (see Figure \ref{fig8} (a)) because of insufficient chain overlap.
In the case of the 515 kg/Mol PS chains self-diffusion dominates for large times as is
indicated by the dashed lines in Figure \ref{fig8} (b). The cooperative diffusion observed
in the semidilute entangled solution (solid triangles in Figure \ref{fig8} (b)) dominates
the autocorrelation function on the same time scale as intramolecular chain relaxations
in the case of a dilute solution (solid squares in Figure \ref{fig8} (b)). Hence one may
conclude that upon increasing the polymer concentration the contribution of internal
chain motions to the single chain dynamics decreases while the contribution of the
cooperative motions increases because of the fluctuations of the surrounding polymer
chains. Both types of dynamics are observable on the same time scale but in different
concentration regimes for high molecular weight PS chains. In the case of internal chain
motions the dynamics is driven by fluctuations of the solvent while fluctuations
of the surrounding polymer network induce the cooperative dynamics. The fact that
cooperative diffusion and internal chain motions occur on similar time and length
scales has already been discussed earlier (see ref \cite{jian:96} and references
therein).
\begin{figure}[ht!]
\vspace*{0.5cm}
\begin{center}
\includegraphics[width=7cm]{FCS_fig8.eps}
\caption{The FCS autocorrelation function $1/G(\tau)-1$ of 17 kg/Mol polystyrene
chains (a) and 515 kg/Mol polystyrene chains (b) in dilute solution
($\blacksquare$, \mbox{$c \to 0$}) and in semidilute solution
($\blacktriangle$, \mbox{$c = 13 $ wt \%}). The dashed lines of slope $\tau$
characterize self-diffusion. Intramolecular motions and cooperative
diffusion dominate in dilute and semidilute entangled solution, respectively,
for short times in the case of the high molecular weight polystyrene chains in (b).}
\label{fig8}
\end{center}
\end{figure}
Without entanglements the local concentration fluctuations at low scattering
vectors ${\bf q}$ are suppressed by the osmotic pressure of the solution, and the
total dynamic scattering function $S_{tot}(q,\phi,\tau)$ measured by DLS decays via
cooperative diffusion according to eqs \ref{eq14} - \ref{eq16}. However, in the presence
of entanglements, there is an additional suppression of concentration fluctuations.
Some concentration fluctuations may be frozen in by the entanglements.
\cite{broc:77,doi:92,eina:99} This fraction of light scattering signal may only decay
with the spectrum of relaxation times of the entanglements themselves, leading to a
slow relaxation of the total dynamic scattering function as is shown in Figures \ref{fig9}
(a) and (b) for the 67 kg/Mol and 515 kg/Mol PS chains in semidilute entangled solution
at \mbox{c=13 \% wt} (solid triangles). The corresponding upper solid lines in Figures
\ref{fig9} (a) and (b) have been calculated according to
\begin{eqnarray} \label{eq26}
S_{tot}(q,\phi,\tau)&=&S_{c}(q,\phi)\exp(-q^2 D_c \tau)\nonumber
\\&+&S_{sl}(q,\phi)\exp(-\tau/\tau_{sl})\,,
\end{eqnarray}
where $\tau_{sl}$ is a decay time. For arbitrary values of the magnitude of the
scattering vector $q$ and the volume fraction $\phi$, the shape of the total dynamic
scattering function $S_{tot}(q,\phi,\tau)$ is more complex than the expression
given in eq \ref{eq26}. For large values of $q$ intramolecular motions lead
to a stretched exponential decay of $S_{tot}(q,\phi,\tau)$ for short times
(see e.g., refs \cite{harn:96,harn:99}). Moreover, the contribution
of the slow relaxation to $S_{tot}(q,\phi,\tau)$ is in general given by
a linear combination of exponentially decaying functions, i.e.,
$\sum_i \exp(-\tau/\tau_{i,d})$. \cite{eina:02,take:07}
\begin{figure}[t]
\vspace*{0.5cm}
\begin{center}
\includegraphics[width=7cm]{FCS_fig9.eps}
\caption{The total dynamic scattering function $S_{tot}(q,\phi,\tau)$ of
67 kg/Mol polystyrene chains (a) and 515 kg/Mol polystyrene chains (b) measured by
DLS in semidilute unentangled solution ($\blacksquare$, \mbox{$c = 1 $ wt \%})
and in semidilute entangled solution ($\blacktriangle$, \mbox{$c = 13 $ wt \%}).
The solid lines follow from eq \ref{eq26}. For short times cooperative diffusion
dominates, while the slow relaxation dominates for very large times in
semidilute entangled solution. There is no slow relaxation in semidilute unentangled
solution, i.e., $S_{sl}(q,\phi)=0$ in eq \ref{eq26}. The absolute value of the
scattering vector is given by \mbox{$q=157.6\, \mu$m$^{-1}$}.}
\label{fig9}
\end{center}
\end{figure}
Experiments on PS in various solvents have confirmed that the slow relaxation
can be measured using DLS.
\cite{brow:90a,nico:90a,nico:90b,nico:90c,wang:93,brow:93,sun:94,wang:95,lin:97}
However, the microscopic understanding of the slow relaxation needs to be improved.
\cite{li:08} On the basis of our FCS
and DLS measurements shown in Figures \ref{fig8} and \ref{fig9} we note that
self-diffusion ($D_s$) occurs on an intermediate time scale, i.e.,
\mbox{$1/(q^2D_c)=0.05$ ms}, \mbox{$1/(q^2D_s)=16$ ms}, and \mbox{$\tau_{sl}=1087$ ms}
for $q=157.6 \mu$m$^{-1}$ for the 515 kg/Mol PS chains. For comparison Figures \ref{fig9}
(a) and (b) also display the measured total dynamic scattering function of the
PS chains in semidilute unentangled solution (solid squares). In this case there
is no slow relaxation due to insufficient chain overlap. The corresponding
lower solid lines in Figures \ref{fig9} (a) and (b) have been calculated according
to eq \ref{eq26} with $S_{sl}(q,\phi)=0$.
The direct DLS measurement of the slow relaxation confirms our earlier remark that
cooperative diffusion becomes highly correlated in the transient entanglement network and
can be studied using only a small fraction of labeled polymer chains within FCS.
As is illustrated in Figure \ref{fig12} (b) unlabeled polymer chains (see, e.g., the polymer
chain denoted by the index 1) and labeled polymer chains (see, e.g., the polymer chain
denoted by the index 2) move in a coherent manner due to entanglements into the FCS
observation volume enclosed by the grey ellipsoidal lines. The resulting temporal
fluctuations of fluorescence light emitted by labeled polymer chains can be detected
by FCS in terms of the cooperative diffusion. A spherical volume of mean size equivalent
to the radius of gyration of an individual polymer chain contains about 15 polymer chains
at the concentration $c^+$ at which cooperative diffusion is measured with FCS. Consequently,
neighbouring chains strongly interpenetrate and entangle with each other leading to
highly cooperative motions in this correlated state. Without entanglements
cooperative diffusion cannot be detected if only a small fraction of the polymer chains
are labeled due to insufficient chain overlap. Hence in dilute and semidilute unentangled
solutions the unlabeled polymer chain denoted by 1 in Figure \ref{fig12} (a) moves from
left to right into the FCS observation volume nearly without influencing the remaining
labeled and unlabeled polymer chains.
\begin{figure}[h!]
\vspace*{0.5cm}
\begin{center}
\includegraphics[width=8cm]{FCS_fig12.eps}
\caption{Schematic illustration of the cooperative diffusion process which is related
to the relaxation of the total polymer number density towards the average total number
density. The polymer chain denoted by the index 1 moves in (a) and (b) from left to
right into the FCS observation volume enclosed by the grey ellipsoidal lines. The
polymer chain diffuses into the observation volume nearly without influencing the
locations of the remaining polymer chains in an unentangled solution in (a), while
the motion of the polymer chain leads to coherent movement of the surrounding polymer
chains in semidilute entangled solution in (b). The size of the polymer chains, the
size observation volume, and the number of polymer chains are not drawn to absolute
scale. Only the fact that in (b) the motion of the unlabeled polymer chain denoted by
the index 1 induces a correlated movement of the labeled polymer chain denoted by the
index 2 into the observation volume is relevant. Each labeled polymer chain carries
only one dye molecule at one of its ends which is marked by a black dot. As the
labeled polymer chain denoted by the index 2 diffuses into the observation volume
from left to right in (b), it causes temporal fluctuations of the detected fluorescence
intensity which can be measured by FCS even in the case that the number of labeled
polymer chains is considerably smaller than the number of unlabeled polymer chains.
In addition self-diffusion can be measured using FCS both in (a) and (b) as
discussed in Section IV A. In (b) self-diffusion of polymer chains corresponds
to movements of the polymer chains along their contour through the transient
network.}
\label{fig12}
\end{center}
\end{figure}
\subsection{Onset of glassy dynamics} \label{glassydynamic}
Upon approaching the glass transition concentration \mbox{$c_{gl}\approx 80$ wt \%}
of PS in toluene, \cite{rauc:03,kona:93} the dynamics of the polymer chains slows down
considerably (see ref \cite{pete:09} and references therein). A first signature of
this slowing down is given by the deviations of the measured cooperative diffusion
coefficients $D_c$ from the solid line at high concentrations in fig \ref{fig7}.
The cooperative diffusion coefficient decreases by more than three decades as compared
to its maximum value upon further increasing the concentration (see fig 6 in ref
\cite{rauc:03}). A second signature of the onset of glassy dynamics is given by the
shape of the autocorrelation function $G(\tau)$ measured with FCS. Figure \ref{fig10}
displays measured functions $1/G(\tau)-1$ (solid symbols) for the 515 kg/Mol PS chains
at three concentrations \mbox{$c = 9.1, 13$, and $20 $ wt \%} together with the
autocorrelation function for the highest concentration (solid line) calculated according
to eq \ref{eq5} with eq \ref{eq3} and
\begin{eqnarray} \label{eq27}
\phi_s(\tau)&=&6 D_s \tau+ A_s \tau^\beta\,,\hspace{0.5cm} \beta=0.3\,.
\end{eqnarray}
Subdiffusive motion characterized by the stretching parameter $\beta$ is observed
as an additional mode on an intermediate time scale between the fast cooperative
diffusion ($D_c$) and the slow self-diffusion ($D_s$). The dotted line in fig \ref{fig10}
represents the asymptotic shape of $1/G(\tau)-1$ in the intermediate time regime.
Both the exponent $\beta=0.3$ and the time scale agree with literature values for
PS. \cite{rauc:03,lind:79}
\begin{figure}[h!]
\vspace*{0.5cm}
\begin{center}
\includegraphics[clip,width=7cm]{FCS_fig10.eps}
\caption{The measured FCS autocorrelation function $1/G(\tau)-1$ of a 515 kg/Mol
polystyrene solution at three concentrations: \mbox{$c = 9.1 $ wt \%}, ($\bullet$);
\mbox{$c = 13 $ wt \%}, ($\blacktriangle$); \mbox{$c = 20 $ wt \%}, ($\blacksquare$).
The solid line displays the result for the highest concentration as obtained from
eq \ref{eq5} with eqs \ref{eq4} and \ref{eq27} as input. For short and large times
cooperative diffusion and self-diffusion dominate, respectively. The dotted line
represents the asymptotic shape of the autocorrelation function in the
intermediate time regime.}
\label{fig10}
\end{center}
\end{figure}
\section{An application: Comparison with minimum concentration required to produce nanofibers}
The understanding of dynamical properties of semidilute entangled polymer solutions
is also important for various technological relevant applications. As an example
we discuss the formation of nanofibers from polymer solutions. Polymer nanofibers
are attractive building blocks for functional nanoscale devices. They are promising
candidates for various applications, including filtration, protective clothing,
polymer batteries, sensors, and tissue engineering. \cite{rama:05,stev:05}
Electrospinning is one of the most established fiber fabrication methods and has attracted
much attention due to the ease by which nanofibers can be produced from polymer
solutions. \cite{grei:07} Fibers produced by this approach are at least one or two
orders of magnitude smaller in diameter than those produced by conventional fiber
production methods like melt or solution spinning. In a typical electrospinning
process a jet is ejected from the surface of a charged polymer solution when the
applied electric field strength overcomes the surface tension. The ejected jet travels
rapidly to the collector target located at some distance from the charged polymer
solution under the influence of the electric field and becomes collected in the form
of a solid polymer nanofiber. However, this method requires a dc voltage in the
kV range and high fiber production rates are difficult to achieve because only a
single fiber emerges from the nozzle of the pipet holding the polymer solution.
\cite{grei:07} In order to overcome these deficiencies an efficient procedure
enabling the parallel fabrication of a multitude of polymer fibers with regular
morphology and diameters as small as 25 nm has been reported recently. \cite{weit:08}
It involves the application of drops of a polymer solution onto a standard spin
coater, followed by fast rotation of the chuck, without the need of a mechanical
constriction. The fiber formation relies upon the instability of the spin-coated
liquid film that arises due to a competition of the centrifugal force and the
Laplace force induced by the surface curvature. This Rayleigh-Taylor instability
triggers the formation of thin liquid jets emerging from the outward driven
polymer solution, yielding solid nanofibers after evaporation of the solvent.
\begin{figure}[h!]
\begin{center}
\includegraphics[clip,width=7cm]{FCS_fig11.eps}
\caption{The concentration $c^+$ ($\bullet$) at which the cooperative diffusion
mode appears in the FCS measurements together with the minimum concentration
$c_{fib}$ ($\square$) required to produce nanofibers \cite{weit:08} as a function
of the molecular weight $M_w$. The solid line of slope $M_w^{-4/5}$ represents a
scaling relation valid for polymers in a good solvent.}
\label{fig11}
\end{center}
\end{figure}
The reason why the ejected jets of polymer solution do not further break up into
individual droplets, but rather give rise to continuous, solid nanofibers, is the
related to the dynamic properties of the polymer solutions. In order elucidate this
point in more detail, Figure \ref{fig11} displays the minimum concentration $c_{fib}$
required to produce nanofibers from 200 kg/Mol and 950 kg/Mol poly-(methylmethacrylate)
solution (open squares) \cite{weit:08} together with the concentration $c^+$ at
which the cooperative diffusion mode appears in the FCS measurements of the 67 kg/Mol,
264 kg/Mol, and 515 kg/Mol PS solutions (solid circles). Interestingly, the
concentrations $c_{fib}$ and $c^+$ follow approximately the same scaling relationship
$c_{fib}=c^+ \sim M_w^{-4/5}$ (c.f., eq \ref{eq8}). Hence, the nanofiber
formation requires that the polymer concentration exceeds the concentration $c^+$
where basically all molecules are involved in the correlated cooperative dynamics.
Uniform fibers cannot be obtained for lower concentrations due to insufficient
chain overlap and the dominating self-diffusion which leads to a disentanglement
under the influence of external forces such as the centrifugal force or the
electrostatic force.
\section{Conclusion}
A general analysis of the diffusion in polystyrene solutions obtained by
fluorescence correlation spectroscopy and by dynamic light scattering has been
presented. Two different diffusion coefficients have been obtained with fluorescence
correlation spectroscopy using single-labeled polystyrene in toluene solutions
[Figures \ref{AC} - \ref{fig4}]. The self-diffusion coefficient $D_s$ results from
fluorescence correlation spectroscopy in the limit of small concentrations of
labeled molecules and for arbitrary concentrations of unlabeled molecules. Moreover,
the cooperative diffusion coefficient $D_c$ in the semidilute entangled regime
becomes accessible as well which is ascribed to an {\bf effective} long-range interaction
of the labeled chains in the transient entanglement network. The self-diffusion coefficients
$D_s$ can be determined from the cooperative diffusion coefficient $D_c$ obtained
by dynamic light scattering measurements and vice versa according to eqs \ref{eq6}
and \ref{eq7}.
The measurements verify the basic scaling and reptation theory for semidilute
entangled polymer solutions [Figures \ref{DSversusClargeMW}, \ref{fig5},
\ref{fig6} and eqs \ref{eq9}, \ref{eq11}, \ref{eq12}]. A quantitative basis for
the modelling of the cooperative diffusion coefficient is given by a Langevin and
generalized Ornstein-Zernike equation [eqs \ref{eq13} - \ref{eq21}]. The calculated
cooperative diffusion coefficients agree with the measured results both in the dilute
and semidilute regimes [Figure \ref{fig7}]. In particular the features of the
crossover region between the dilute and the semidilute regimes are captured correctly
by the underlying integral equation theory.
For large times the decay of the fluorescence correlation spectroscopy autocorrelation
function is dominated by self-diffusion, while intramolecular chain relaxations in
dilute solution and cooperative diffusion in semidilute entangled solution dominate for
short times [Figures \ref{fig6} and \ref{fig8}]. An additional slow relaxation in
semidilute entangled solution can be observed by dynamic light scattering [Figure \ref{fig9}].
Moreover, the fluorescence correlation spectroscopy autocorrelation function
exhibits an additional mode on an intermediate time scale upon approaching the glass
transition concentration [Figure \ref{fig10}].
Finally, it has been shown the minimum concentration required to produce solid
nanofibers from a polymer solution follows the same scaling relationship
as the concentration at which the cooperative diffusion mode appears in the fluorescence
correlation spectroscopy measurements [Figure \ref{fig11}]. The nanofiber
formation requires that the polymer concentration exceeds the concentration where
basically all molecules are involved in the correlated cooperative dynamics.
Hence fluorescence correlation spectroscopy is helpful for the understanding of
dynamical properties of semidilute entangled polymer solutions in the case of
technological relevant applications.
\\
We thank A.\ H.\ E.\ M\"uller and A.\ B\"oker for the synthesis of the polymers and the Deutsche Forschungsgemeinschaft, SFB 481 (A11), Bayreuth, for financial support.
|
\section{Introduction} \label{chap:intro} \end{centering}
\setcounter{figure}{0} \setcounter{equation}{0}
We take up the study of a system very familiar from physics (\cite{seitz} for example): a long string of coupled linear oscillators (called `agents' or `cars' from here on) with linear nearest neighbor interactions. In order to model things like traffic we change certain features of the model that is so familiar in the context of statistical physics. The most significant ones are that we do not assume that interaction is symmetric
and that the oscillators are damped. Furthermore we allow the system to have non-zero mean velocity. Finally we assume that the chain is \emph{finite} and we take into account boundary effects from the ends of the chain (no periodic boundary conditions). The class of systems whose study we take up in this paper is illustrated in Figure \ref{fig:chain} and made precise in Equation \ref{example2}.
\begin{figure}[ptbh]
\centering
\includegraphics[width=4.6in]{flocks6figA.eps}
\caption{\emph{ In the upper figure the `symmetric' ($\rho=\frac12$) flock is illustrated.
Each agent is linearly coupled to its nearest neighbor. At $t=0$ the agent labeled 0, the leader, undergoes a forced motion: either an oscillation or a kick in the direction of the arrow above it. It receives no feedback from the flock.
The asymmetric interaction is suggested in the lower figure, where we drew the $k$-agent with its interactions and their weights. The arrows give the direction of the information flow. }}
\label{fig:chain}
\end{figure}
The purpose here is to gain a qualitative understanding of how a perturbation in the movement of one the agents (which we call the \emph{leader}) propagates throughout the flock of agents.
Here is a more precise formulation. Suppose that a flock is moving coherently or ``in formation", ie: each member has the same constant velocity, and inter-agent distances are constant. Now the leader's motion is perturbed. After a while also the agent farthest away from the leader (called the \emph{trailing car}) senses the effect, and its orbit will start to deviate from the in formation orbit. What is the ratio between these two perturbations \emph{as a function of the size $N$ of the flock}, while keeping all other parameters fixed ?
The immediate motivation is to study which kinds of interactions might take place in actual large flocks or might seem desirable to implement in artificial flocks (think of cars on a highway equipped with an automatic pilot). It would seem that in both types of systems the kinds of interaction preferred would be those where the growth rate just mentioned is as low as possible.
Ideally we find the growth rate of perturbations
in many different models, for example where interactions are allowed to include up to 2 or 3 or 10 neighbors on either side. This seems analytically out of reach. Thus the acceleration of the $k$-th car is determined by its observations of the $k-1$-st and the $k+1$-st car. What we chose to do instead is to introduce a \emph{weighted} nearest neighbor interaction, that is: For $\rho\in[0,1]$ we weight the information coming from the $k-1$-st car with $1-\rho$, and the other with $\rho$ (see Equation (\ref{example2})).
This line of thought has many potential applications in technology. Suppose for example one wants to automatically control dense traffic on a single lane road. Each car is equipped with sensors that register the relative velocity and position of its neighbors. This information is then used to regulate the acceleration of the cars.
A typical example is the `canonical traffic problem' proposed in \cite{flocks4} and \cite{flocks5}, where the lead car abruptly accelerates (when, for example, a traffic light turns green), and the other cars try to follow.
Our conclusions for the systems typified by Equation \ref{example2}, can be summarized as follows. In all systems the ratio of the size of perturbation of the trailing car to the perturbation of the leader grows exponentially in the size $N$ of the flock, with one exception: when $\rho=\frac12$. In the last case (called the \emph{symmetric} case because equal attention is paid to the cars on either side) this growth is only \emph{linear} as function of $N$. The study of the symmetric case was reported in \cite{flocks4, flocks5}. Here we continue that program by investigating the asymmetric case.
The curious fact that perturbations in the leader's position or velocity is necessarily unbounded
as the size of the flock grows is of course
of paramount importance in many applications (automated traffic, biological flocks, and so on).
In fact this has been commented upon by several authors already, notably \cite{SPH,BH} in cases
similar to our $\rho=0$ and $\rho=\frac12$ cases. In this note we give precise estimates for
how these perturbations grow as function of the number of agents,
not only in those cases but also for other models (all $\rho\in[0,1]$).
A more distant motivation to study these systems is to attain a general understanding of linear oscillators interacting according to a more general graph (called the ``communication graph", see for instance \cite{flocks2} for definitions). In that context similar questions arise but in a more general context. What we present here is a simple example.
Section \ref{chap:model} defines the model. In Section \ref{chap:stability} the definitions of stability we use are given and briefly discussed. In it we also state the main result of this note. Sections \ref{chap:asymptotic} and \ref{chap:harmonic} prove these results. (Some of the more calculational steps in these proofs are relegated to the Appendix.)
\vskip .2in
\noindent{\bf Notational Conventions:} To avoid confusion, we finally list two important conventions here. The first is that after Section \ref{chap:asymptotic} we assume that both $f$ and $g$ are negative reals to insure asymptotic stability (Theorem \ref{theo:stable}). The second is that in order for certain expressions ($\mu_+(i\omega)$ and $\mu_-(i\omega)$, for $\omega>0$) in Theorem \ref{theo:a_n}) to be continuous functions it is convenient to define the symbol $\sqrt z$ as the root with angle in the interval $[0,\pi)$ (branch cut along the positive real axis).
\vskip .2in
\section*{Acknowledgements:} I am grateful for useful conversations with Folkert Tangerman.
\vskip 1.in
\begin{centering}\section{The Model} \label{chap:model} \end{centering}
\setcounter{figure}{0} \setcounter{equation}{0}
We define the model and propose a notion of stability for flocks. Our strategy is to study qualitative aspects of the solution for fixed $\rho$, $f$, and $g$, and as we let $N$ tend to infinity.
Let $f$ and $g$ be real, and $\rho\in[0,1]$.
The model is given by:
\begin{eqnarray}
\forall i \in \{1,\cdots N-1\} \;:\; \dot x_i &=& u_i \cr
\dot u_i &=&f\left\{(x_{i}-h_{{i}})- (1-\rho)(x_{i-1}-h_{{i-1}})-\rho(x_{i+1}-h_{{i+1}})\right\}
+g\left\{u_i-(1-\rho)u_{i-1}-\rho u_{i+1}\right\}\cr
\dot x_{N} &=& u_{N} \cr \dot u_{N}
&=& f\left\{(x_{N}-h_{{N}})- (x_{N-1}-h_{{N-1}})\right\}+
g\left\{u_{N}-u_{N-1}\right\}\cr
\;\;{\rm and }\;\; \quad \quad x_0 &=& x_0(t) \quad {\rm given}
\label{example2}
\end{eqnarray}
The (constant) parameters $h_i$ determine the desired relative distances between agents $i$ and $i-1$ as $h_i-h_{i-1}$. The feedback parameters $f$ and $g$ are independent of $i$ and time. Note that the total number of agents is in fact also a parameter in this problem. We do not carry this into the notation. The subscript $N$ will always stand for the last agent in a system with agents numbered from 0 to $N$. Similarly we do not carry the dependence on the parameters $f$ and $g$ explicitly into the notation.
One can show that the orbits of this system such that the distances between successive cars are preserved (namely: $-(h_i-h_{i-1})$) form a 2-parameter family, namely $x_k(t)=x_0(0)+v_0(0)t+h_k$ and $\dot x_k(t)= v_0(0)$. We will call these orbits \emph{in formation orbits}.
It is advantageous to write Equation (\ref{example2}) in a more compact form. Introduce the notation
\begin{displaymath}
z\equiv (z_1,\dot z_1, z_2,\dot z_2,\cdots, z_N,\dot z_N) \equiv
(x_1-h_1,u_1,x_2-h_2,u_2, \cdots, x_{N}-h_N, u_{N})^T \quad .
\end{displaymath}
The leading car is not encoded since its orbit is a priori given. The system can now be recast as a first order ODE:
\begin{equation}
\dot z = M z + \Gamma_0(t)\quad .
\label{eq:indepleader}
\end{equation}
The details of this are discussed in \cite{flocks2}, here we just give the relevant definitions.
Let $I$ and $P$ are $N$-dimensional square matrices, where $I$ is the identity and $P$ is given by
\begin{equation}
P= I-Q_\rho \quad \;\;{\rm where }\;\;\quad Q_\rho=\left(\begin{array}{ccccc}
0 & \rho & & & \\
1-\rho & 0 & \rho & & \\
& \ddots & \ddots & \ddots & \\
& & 1-\rho & 0 & \rho \\
& & & 1 & 0
\end{array}\right),
\label{eq:laplacian}
\end{equation}
$P\equiv I-Q_\rho$ is called the reduced graph Laplacian. It describes the flow of information among the agents, with the exception of the leader (hence the word `reduced'). The $2\times2$ matrices $A$ and $K$ are given by:
\begin{equation}
A= \left(\begin{array}{cc}
0 & 1 \\
0 & 0
\end{array}\right) \quad \;\;{\rm and }\;\; \quad
K= \left(\begin{array}{cc}
0 & 0 \\
f & g
\end{array}\right) \quad .
\label{eq:A-and-K}
\end{equation}
The orbit of the leader is assumed to be a priori given and therefore only appears in the forcing term $\Gamma_0(t)$.
We will refer to this agent as an \emph{(independent) leader}.
Analyzing Equation (\ref{example2}) and assuming without loss of generality that $h_0=0$, one gathers that:
\begin{equation}
\Gamma_0(t) = \left(\begin{array}{c}
0 \\
(1-\rho)\left(fz_0(t)+g\dot z_0(t)\right) \\
0\\
\vdots
\end{array}\right) \quad .
\label{eq:Gamma_0}
\end{equation}
To define $M$ of Equation (\ref{eq:indepleader}) in terms of these quantities, we use the Kronecker product ($\otimes$)
\begin{equation}
M\equiv I \otimes A + P \otimes K \quad .
\end{equation}
The advantage of this somewhat roundabout way of defining the matrix $M$ is that in the eigenvalues of the reduced Laplacian $P$ in many cases are known. From that the eigenvalues of $M$ can then be derived.
That is the program followed in the Section \ref{chap:asymptotic}.
\vskip 1.in
\begin{centering}\section{Stability of Flocks} \label{chap:stability} \end{centering}
\setcounter{figure}{0} \setcounter{equation}{0}
\noindent
\begin{defn}
The system given in Equation \ref{eq:indepleader} is called `asymptotically stable' if all eigenvalues of $M$ have negative real part.
\label{defn:asymptotical}
\end{defn}
Suppose for the moment that $\Gamma_0(t)=0$ fot $t>t_0$. Then
the solution of the system tends to 0 exponentially fast (in $t$) if and only if the system is asymptotically stable. This corresponds to the usual notion of asymptotic stability (see for example \cite{arnold}, Section 23). The question when the system is asymptotically stable has a straightforward answer (see Theorem \ref{theo:stable}): it is if and only if $f$ and $g$ in Equation (\ref{example2}) are negative.
We will therefore from now on \emph{assume that $f$ and $g$ are negative}.
One can show (Proposition \ref{prop:a_k}) that if the leader executes an oscillation of the form $e^{i\omega t}$, then $z_k(t)$ tends to $a_k(i\omega)e^{i\omega t}$ as $t$ tends to infinity. The functions $a_k(i\omega)$ are called the frequency response functions.
\begin{defn} Let $A_N\equiv \sup_{\omega\in\mbox{I${\!}$R}}\; |a_N(i\omega)|$. The system is called `harmonically stable' if it is asymptotically stable and if $\limsup_{N\rightarrow \infty}\; \left|A_N\right|^{1/N}\leq 1$. Otherwise the system is called `harmonically unstable'.
\label{defn:harmonic}
\end{defn}
This kind of instability roughly says that certain long-term oscillatory perturbations in the orbit of the leader will have their amplitude magnified by a factor that is exponentially large in $N$.
In Theorem \ref{theo:stable2} we establish that the system is harmonically unstable if $\rho\neq 1/2$; harmonic stability for $\rho=\frac12$ was established in \cite{flocks4}.
In earlier work (\cite{flocks5}) we studied a `fundamental traffic problem' which roughly corresponds to setting the acceleration of the leader equal to the Dirac delta function, $\delta(t)$, and $ z_0(0)=\dot z_0(0)=0$. This gives $z_0(t)=t$ for $t\geq 0$ which can be substituted into Equation (\ref{eq:Gamma_0}).
\begin{defn} Consider Equation (\ref{eq:indepleader}) with forcing determined by $\ddot z_0(t)\equiv \delta(t)$ and subject to the initial conditions $z_k(0)=\dot z_k(0)=0$ and $\dot z_k(0)=0$. Let $Z^{(i)}_N\equiv \sup_{t>0} |\frac{d^i}{dt^i}(z_N(t)-z_0(t))|$. The system is called `impulse stable' if it is asymptotically stable and if for $i$ 0, 1 and 2, we have $\limsup_{N\rightarrow \infty}\; \left|Z^{(i)}_N\right|^{1/N}\leq 1$. (And `impulse unstable' in the other case.)
\label{defn:impulse}
\end{defn}
Loosely interpreted this kind of instability means that if we give the leader a 'unit-kick', then that perturbation travels through the flock and causes $\sup|x_N(t)|$, $\sup |\dot x_N(t)|$, or $\sup |\ddot x_N(t)|$ to grow exponentially in $N$, before eventually dying out (due to asymptotic stability).
Impulse stability is perhaps at first sight more natural or appealing than harmonic stability because it is formulated in the time domain, whereas the latter takes place in the Fourier domain. However mathematically the criterion is much harder to check.
In \cite{flocks5} we proved that the case $\rho=1/2$ is impulse stable. But the case $\rho\neq 1/2$ is much more problematic and will be taken up in a separate work (\cite{flocks7}).
It therefore may be argued that all three kinds of stability are necessary to form large flocks. From the above remarks, it follows that of the systems investigated here only the one with $\rho=\frac12$ satisfies all three.
It is interesting that even in that case \emph{linear} growth of $Z^{(i)}_N$ still takes place (see \cite{flocks4, flocks5}). This seems to be the best case possible. we summarize this discussion with our main result.
\begin{theo} The system given by Equation (\ref{example2}) satisfies all three stability criteria if and only if $f$ and $g$ are negative and $\rho=1/2$.
\label{theo:main}
\end{theo}
\vskip 1.in
\begin{centering}\section{Asymptotic Stability} \label{chap:asymptotic} \end{centering}
\setcounter{figure}{0} \setcounter{equation}{0}
\noindent We derive the criterion for asymptotical stability.
In the statement of the next result we use the following equation, where $\rho \in(0,1)$ and $\phi$ are real variables:
\begin{equation}\label{cot2}
(2\rho - 1)\cot\phi=\cot N\phi\; .
\end{equation}
Recall that the matrix $P$ is defined in Equation (\ref{eq:laplacian}).
\vskip.2in\begin{prop} (\cite{tridiagonal})
For any $\rho\in(0,1)$, the matrix $P$ has $N$ distinct eigenvalues $\{\lambda_\ell\}_{\ell=0}^{N-1}$: \\
\emph{\bf i) If $\rho\in (0,\frac{1}{2}]$:} for $\ell\in\{0,\ldots, N-1\}$, $\lambda_\ell=1-2\sqrt{\rho(1-\rho)}\,\cos \phi_\ell$, where $\phi_\ell \in \left( \frac{\ell\pi}N, \frac{(\ell+1)\pi}N\right)$ solves (\ref{cot2}).\\
\emph{\bf ii) If $\rho\in (\frac12,\frac{N+1}{2N}]$:} Identical to i).\\
\emph{\bf iii) If $\rho\in (\frac{N+1}{2N},1)$:} for $\ell\in\{1,\ldots, N-2\}$, $\lambda_\ell=1-2\sqrt{\rho(1-\rho)}\,\cos \phi_\ell$, where $\phi_\ell \in \left( \frac{\ell\pi}N, \frac{(\ell+1)\pi}N\right)$ solves (\ref{cot2}); $\lambda_0=\frac{(2\rho-1)^2}{2\rho^2}\left(\frac{1-\rho}{\rho} \right)^{N-1} + \mathcal{O}\left(\left(\frac{1-\rho}{\rho} \right)^{2N-2}\right)$ and $\lambda_{N-1}=2-\lambda_0$.
\label{prop:evalsP}
\end{prop}
\noindent One can show (see \cite{flocks2,flocks4, flocks5}) that the eigenvalues of $M$ defined in Equation (\ref{eq:indepleader}) are given by the solutions $\nu_{\ell\pm}$ of
\begin{equation}
\nu^2-\lambda_\ell g \nu-\lambda_\ell f= 0 \quad,
\label{eq:evals2}
\end{equation}
where $\lambda_\ell$ runs through the spectrum of $P$. So:
\begin{theo} The eigenvalues of $M$ are
\begin{displaymath}
\nu_{\ell\pm} = \frac{1}{2}\left(\lambda_\ell g \pm \sqrt{(\lambda_\ell g)^2 + 4 \lambda_\ell f}\right)= \frac{\lambda_\ell g}{2}\left(1\pm \sqrt{1+\frac{4f}{\lambda_\ell g^2}}\right) \quad ,
\end{displaymath}
where $\lambda_\ell$ runs through the spectrum of $P$. Because the $\lambda_\ell$ are contained in the interval $[0,2]$ (see Proposition \ref{prop:evalsP}), the system is stabilized (or asymptotically stable) if and only if both $f$ and $g$ are strictly smaller than zero.
\label{theo:stable}
\end{theo}
\vskip 1.in
\begin{centering}\section{Harmonic Stability} \label{chap:harmonic} \end{centering}\
\setcounter{figure}{0} \setcounter{equation}{0}
In this section we calculate (Proposition \ref{theo:a_n}) and study the properties (Theorem \ref{theo:stable2}) of the frequency response function of the trailing car in detail. The main result of this Section is Theorem \ref{theo:stable2} that says that if $[0,1)\backslash\{\frac12\}$ then the system is harmonically unstable. (If $\rho=1$ the question is moot as the leader's motion goes unperceived by the flock.)
The following constant will frequently simplify formulae:
\begin{displaymath}
\kappa \equiv \frac{1-\rho}{\rho} \quad \;\;{\rm or }\;\; \quad \rho = \frac{1}{1+\kappa} \quad .
\end{displaymath}
\vskip .2in
\begin{prop} If $z_0(t) = e^{\nu t}$ and $\nu\in i\mbox{I${\!}$R}$ and $\nu \not\in \;{\rm spec }\;(M)$, there is a sequence $a=(a_1\cdots, a_N)$ of complex numbers $\{a_k(\nu)\}_{k=1}^N$ so that trajectory of the system is asymptotic to:
\begin{displaymath}
z_k(t) = a_k(\nu)\,e^{\nu t} \quad \;\;{\rm where }\;\; \quad a(\nu)= -(M-\nu I)^{-1}g_0.
\end{displaymath}
\label{prop:a_k}
\end{prop}
\noindent {\bf Proof:} (See also \cite{flocks2}.) Since $z_0(t) = e^{\nu t}$, the non-autonomous term in Equation (\ref{eq:indepleader}) can be written as $\Gamma_0(t)=g_0e^{\nu t}$, where $g_0$ is a constant vector. The general solution of the system is: $e^{Mt}z_0+(M-\nu I)^{-1}\left( e^{Mt}-e^{\nu t}\right)g_0$, and $e^{Mt}$ tends to zero. \hfill \vrule Depth0pt height6pt width6pt \vspace{6mm}
\noindent The complex functions $a_k(\nu)$ (where $\nu \in i\mbox{I${\!}$R}$) are called the \emph{frequency response} (of the $k$-th agent).
\vskip .2in
\begin{prop}
i): For $\rho\in (0,1)\backslash \{\frac{1}{2}\}$ the frequency response function of the $k$-th agent is given by the functions:
\begin{displaymath}
\begin{array}{c}
\displaystyle a_{k}(\nu)=
\kappa^k \frac{\left(\mu_+ -\mu_+^{-1}\right)\mu_+^{N-k}- \left(\mu_- - \mu_-^{-1}\right)\mu_-^{N-k}}
{\left(\mu_+ -\mu_+^{-1}\right)\mu_+^{N}- \left(\mu_- - \mu_-^{-1}\right)\mu_-^{N}} \\
\;\;{\rm where }\;\; \quad \displaystyle \mu_\pm=\mu_\pm(\nu) \equiv \frac{1}{2\rho}\left(\gamma\pm \sqrt{\gamma^2-4\rho(1-\rho)}\right) \quad \;\;{\rm and }\;\; \quad
\gamma = \gamma(\nu)\equiv \displaystyle \frac{f+ g \nu -\nu^2}{f+ g\nu}
\quad .
\end{array}
\end{displaymath}
ii): (\cite{flocks5}) When $\rho=\frac12$ the above expressions simplify to:
\begin{displaymath}
\begin{array}{c}
\displaystyle a_{k}(\nu)=\frac{\mu_+^{N-k}+\mu_-^{N-k}} {\mu_+^{N}+\mu_-^{N}} \\
\;\;{\rm where }\;\; \quad \mu_\pm(\nu) \equiv \gamma\pm \sqrt{\gamma^2-1} \quad \;\;{\rm and }\;\; \quad
\gamma = \gamma(\nu) = \displaystyle \frac{f+ g \nu - \nu^2}{f+ g \nu}
\quad .
\end{array}
\end{displaymath}
iii): (\cite{flocks4}) When $\rho=0$:
\begin{displaymath}
a_k(\nu)= \gamma(\nu)^{-k} \quad .
\end{displaymath}
\label{theo:a_n}
\end{prop}
\noindent {\bf Proof:} The reasoning of i) is identical to that in Lemma 3.2 of \cite{flocks5}. \hfill \vrule Depth0pt height6pt width6pt \vspace{6mm}
\noindent{\bf Remark:} Even though $\mu_\pm$ are not rational functions of $\nu$, one can check that in fact the $a_k(\nu)$ in fact are proper rational.
\vskip .1in
\noindent{\bf Remark:} The coefficients $\mu_\pm$ in the Theorem are the roots of the following quadratic equation:
\begin{displaymath}
\rho \mu^2-\gamma \rho +(1-\rho) = 0 \quad .
\end{displaymath}
\vskip .2in
The most important case of Theorem \ref{theo:a_n} is the frequency response of the trailing car, labeled $N$, when $\rho\in (0,1)\backslash \{\frac12 \}$. The previous result immediately implies:
\begin{cory} For $\rho\in (0,1)\backslash \{\frac{1}{2}\}$ the frequency response function of the last agent is given by
\begin{displaymath}
a_{N}(\nu)= \frac{1+\kappa}{\kappa} \;\kappa^N \; \frac{\mu_+ - \mu_-}
{\left(\mu_+-\mu_+^{-1}\right)\mu_+^{N}- \left(\mu_--\mu_-^{-1}\right)\mu_-^{N}} \quad ,
\end{displaymath}
with $\mu_\pm$ and $\gamma$ as before.
\label{cor:trailing}
\end{cory}
Because the inverse Fourier Transform of $a_N$ equals the real valued function $\ddot z_N(t)$, we are mainly interested in the case where $\nu=i\omega$ and $\omega$ is a positive real number (the \emph{frequency}). Note that then also $a_N(-i\omega)$ must equal the complex conjugate of $a_N(i\omega)$. In fact, using Lemma \ref{lem:taylor} in the previous Proposition yields that $a_N(0)=1$. Thus it is sufficient to study $a_N(i\omega)$ only for $\omega> 0$.
\vskip .2in
\begin{prop}
Suppose $f$, $g$, $\omega>0$ and $\rho\in (0,1/2)\cup(1/2,1)$ are all fixed. Then, for $r\in(0,1)$ as in Lemma \ref{lem:mu+bigger}, as $N$ large tends to infinity:
\begin{displaymath}
a_N(i\omega)= \frac{1+\kappa}{\kappa}\;\mu_-^N\;\frac{\mu_+-\mu_-}{\mu_+-\mu_+^{-1}} \left(1+{\cal O}(r^N)\right) \quad .
\end{displaymath}
\label{prop:rho-small-big}
\end{prop}
\noindent {\bf Proof:}
Use Proposition \ref{theo:a_n} and the fact that $\mu_-\mu_+=\kappa$ to rewrite
\begin{equation}
a_N(\nu)=\frac{1+\kappa}{\kappa}\;\mu_-^N\;\frac{\mu_+-\mu_-}{\mu_+-\mu_+^{-1}} \;\left(1- \frac{\mu_--\mu_-^{-1}}{\mu_+-\mu_+^{-1}} \left(\frac{\mu_-}{\mu_+}\right)^N\right)^{-1} \quad .
\label{eq:aN-to-mu}
\end{equation}
Since
\begin{displaymath}
\frac{\mu_--\mu_-^{-1}}{\mu_+-\mu_+^{-1}}= \frac{-1}{\kappa}\; \frac{\mu_+^2-\kappa^2}{\mu_+^2-1} \quad ,
\end{displaymath}
it suffices to prove that if $\omega>0$, then $\mu_+(\omega)^2\neq 1$.
Now suppose that $\mu_+(\omega)^2= 1$, then the second remark after Proposition \ref{theo:a_n} implies that $\gamma(i\omega)=\pm1$ and so Lemma \ref{lem:gamma} implies that $\omega=0$. \hfill \vrule Depth0pt height6pt width6pt \vspace{6mm}
\noindent{\bf Remark:} It is important to realize that the even for large but \emph{fixed} $N$, the last factor could become large if we let $\mu_+$ approximate 1. The results in the appendix can be used to show that this can happen if and only if $\rho<\frac12$ and $\omega$ very close to zero. We make use of this fact in the proof (for $\rho<\frac12$) of the next and main result of this section.
\begin{theo} For all $\rho \in [0,1)\backslash \{\frac12\}$, $A_N$ grows exponentially in $N$. When $\rho=\frac12$, growth is linear in $N$.
\label{theo:stable2}
\end{theo}
\noindent {\bf Proof:}
The second statement has been proved in \cite{flocks5}. Fix $\rho\in(0,1/2)$ and let $\omega_+$ as in Equation \ref{eq:omega+}. Lemma \ref{lem:omega+} and the remark thereafter now imply that $|\mu_-(i\omega)|>1$ if and only if $\omega\in(0,\omega_+)$.
The result follows directly from the first part of Proposition \ref{prop:rho-small-big}. Finally fix $\rho\in (\frac12,1)$, or $\kappa\in (0,1)$.
First use Lemma \ref{lem:taylor} to see that if $\omega^2=\frac12 |f|(1-\kappa)^2\kappa^{N-1}$, then $\mu_+=1-\frac{(1-\kappa^2)}{2}\,\kappa^{N-1}+{\cal O}(\kappa^{3N/2})$ and $\mu_-=\kappa(1+\frac{(1-\kappa^2)}{2}\,\kappa^{N-1})+{\cal O}(\kappa^{3N/2})$. Substitute this into the denominator of $a_N$ in Corollary \ref{cor:trailing}. The leading order cancels. The next term is of order at least $\kappa^{3N/2}$. Thus $A_N$ is of order at least $\kappa^{-N/2}$. \hfill \vrule Depth0pt height6pt width6pt \vspace{6mm}
\vskip 1.in
\begin{centering}\section{Appendix: Technical Results} \label{chap:tresults} \end{centering}\
\setcounter{figure}{0} \setcounter{equation}{0}
For completeness we collect a number of straightforward results that are necessary for development of the theory, but would clutter the exposition in the main text.
Various relevant quantities are evaluated for $\nu=i\omega$ where $\omega$ is real and non-negative. As observed in the main text, we use $\omega\geq 0$ without loss of generality. The conventions mentioned in the Introduction also hold.
\vskip .2in
\begin{lem} i): For $\rho \in (0,\frac12)$ and $\omega \geq 0$ small:
\begin{displaymath}
\mu_+ = \frac{1-\rho}{\rho}\left(1 + \frac{\omega^2}{(2\rho-1)|f|} - i\; \frac{|g| \;\;\omega^3}{(2\rho-1)f^2} \right) + {\cal O}(\omega^4) \quad \;\;{\rm and }\;\; \quad
\mu_- = 1 - \frac{\omega^2}{(2\rho-1)|f|} + i\; \frac{|g| \;\;\omega^3}{(2\rho-1)f^2} + {\cal O}(\omega^4) \quad .
\end{displaymath}
ii): For $\rho \in (\frac12,1)$ and $\omega \geq 0$ small:
\begin{displaymath}
\mu_+ = 1 - \frac{\omega^2}{(2\rho-1)|f|} + i\; \frac{|g| \;\;\omega^3}{(2\rho-1)f^2} + {\cal O}(\omega^4) \quad \;\;{\rm and }\;\; \quad
\mu_- = \frac{1-\rho}{\rho}\left(1 + \frac{\omega^2}{(2\rho-1)|f|} - i\; \frac{|g| \;\;\omega^3}{(2\rho-1)f^2} \right) + {\cal O}(\omega^4) \quad .
\end{displaymath}
\label{lem:taylor}
\end{lem}
\noindent {\bf Proof:} By sheer calculation. (See \cite{flocks4} for some of the computational details.) \hfill \vrule Depth0pt height6pt width6pt \vspace{6mm}
\noindent {\bf Remark:} This expansion diverges for $\rho=1/2$; in that case we have (\cite{flocks4}):
\begin{displaymath}
\mu_\pm = 1-\frac{\omega^2}{|f|}\pm
\frac{\omega^2|g|}{\sqrt{2}|f|^{3/2}}+{\cal O}(\omega^4) +i\left( \pm
\frac{\sqrt{2}\omega}{|f|^{1/2}} \pm {\cal O}(\omega^3) \right)
\quad .
\end{displaymath}
\begin{lem} $\displaystyle\gamma(i\omega)=1-\frac{\omega^2 |f|}{f^2+\omega^2g^2} + i\,\frac{\omega^3|g|}{f^2+\omega^2g^2}\quad$.
\label{lem:gamma}
\end{lem}
\begin{figure}[ptbh]
\centering
\includegraphics[height=2.3in]{flock6fig2.eps}
\includegraphics[height=2.3in]{flock6fig3.eps}
\includegraphics[height=2.3in]{flock6fig4.eps}
\caption{\emph{ The eigenvalues $\mu_+(i\omega)$ (blue) and $\mu_-(i\omega)$ (red) of $C$ when $f=g=-1$ for $\omega$ positive.
From left to right: $\rho=0.4$, $0.5$, and $0.6$. In addition the circles with radii $\sqrt{\kappa}$ and $\kappa$ are drawn in green and black, resp., where $\kappa\equiv \frac{1-\rho}{\rho}$.}}
\label{fig:rho}
\end{figure}
\begin{lem} For each $\rho \in (0,1)\backslash \{\frac12\}$, the number $r\equiv \sup_\omega \frac{|\mu_-(i\omega)|}{|\mu_+(i\omega)|}$ exists and is contained in $(0,1)$. (See Figure \ref{fig:rho}.)
\label{lem:mu+bigger}
\end{lem}
\noindent {\bf Proof:} From Lemma \ref{lem:gamma}, $\gamma(i\omega)\approx -\frac{i\omega}{g}$ when $\omega$ is large. Substitute this into the expression for $\mu_\pm$ in Theorem \ref{theo:a_n} to see that for large $\omega$, in fact $\frac{|\mu_-(i\omega)|}{|\mu_+(i\omega)|}$ becomes very small.
When $\omega=0$, Lemma \ref{lem:taylor} implies that $\frac{|\mu_-(i\omega)|}{|\mu_+(i\omega)|}= \min\{\kappa,\kappa^{-1}\}$.
It is now sufficient to prove that for $\omega\in\mbox{I${\!}$R}^+$ the absolute values $|\mu_\pm|$ are never equal. So suppose there are $\omega_0$ and $\theta\in\mbox{I${\!}$R}$ so that $\mu_+(i\omega_0)-\mu_-(i\omega_0)e^{i\theta}=0$. The first item in Theorem \ref{theo:a_n} gives:
\begin{displaymath}
\gamma (1-e^{i\theta}) = -\sqrt{\gamma^2-4\rho(1-\rho)}\,(1+e^{i\theta})\quad \quad .
\end{displaymath}
Dividing this by $1+e^{i\theta}$, squaring the equation, and noting that $\frac{(1-e^{i\theta})^2}{(1+e^{i\theta})^2}=-(\tan(\frac\theta2))^2$, we see that
\begin{displaymath}
\gamma^2 \left(1+\left(\tan \frac{\theta}{2}\right)^2\right)=4\rho(1-\rho) \quad .
\end{displaymath}
This implies that $\gamma^2$ is a positive real and therefore $\gamma$ is real for some $\omega\neq 0$, which is impossible by Lemma \ref{lem:gamma}.
\hfill \vrule Depth0pt height6pt width6pt \vspace{6mm}
\begin{lem} For each $\rho \in (0,1/2)$, there is a unique $\omega_+>0$ such that
\begin{displaymath}
\omega\in\left(0,\omega_+\right)\quad \Longrightarrow \quad |\mu_-(i\omega)|>1 \quad \;\;{\rm and }\;\; \quad \omega>\omega_+\quad \Longrightarrow \quad|\mu_-(i\omega)|<1 \quad .
\end{displaymath}
\label{lem:omega+}
\end{lem}
\noindent {\bf Proof:} We know that $\mu_-(0)=1$ and (from the proof of the previous Lemma) for large $\omega$: $|\mu_-(\omega)|$ is small. It is sufficient to prove that $\omega_+$ is the unique solution in $(0,\infty)$ of $|\mu_-(i\omega)|=1$ and that it is simple.
Consider the characteristic equation
$\rho\mu^2-\gamma \mu + (1-\rho)=0$ and suppose that there is a root $\mu=e^{i\theta}$. Then $\gamma=\rho e^{i\theta}+(1-\rho)e^{-i\theta} =\cos(\theta)+i(2\rho-1)\sin(\theta)$. Equate this to the expression given in Lemma \ref{lem:gamma} and use the fact that $\cos^2(\theta)+\sin^2(\theta)=1$ to obtain:
\begin{displaymath}
\left(1-\dfrac{\omega^2 |f|} {f^2+\omega^2g^2}\right)^2+\dfrac{1}{(2\rho-1)^2}\left(\dfrac{\omega^3|g|} {f^2+\omega^2g^2}\right)^2=1
\end{displaymath}
This equation factors as follows:
\begin{displaymath}
\omega^2\left(\dfrac{g^2}{(2\rho-1)^2}\,\omega^4+(f^2-2|f|g^2)\omega^2-2|f|^3 \right)=0
\end{displaymath}
The second factor gives exactly one simple positive root for $\omega^2$, yielding a unique simple positive root $\omega=\omega_+$. \hfill \vrule Depth0pt height6pt width6pt \vspace{6mm}
\noindent {\bf Remark:} In fact,
\begin{equation}
\omega_+^2= (1-2\rho)|f|\left( \left(1-\frac{|f|}{2g^2}\right)(1-2\rho) + \sqrt{\left(1-\frac{|f|}{2g^2}\right)^2(1-2\rho)^2 +\,\frac{2|f|}{g^2}} \right) \quad .
\label{eq:omega+}
\end{equation}
\vskip .35 in
|
\section{Introduction}
\noindent Currently, applying constraint technology to a large, complex problem
requires significant manual tuning by an expert. Such experts are rare. The
central aim of this project is to improve the scalability of constraint
technology, while simultaneously removing its reliance on manual tuning by an
expert. We propose a novel, elegant means to achieve this -- a \emph{constraint
solver synthesiser}, which generates a constraint solver specialised to a given
problem. Constraints research has mostly focused on the incremental improvement
of general-purpose solvers so far. The closest point of comparison is currently
the G12 project~\cite{g12}, which aims to combine existing general constraint
solvers and solvers from related fields into a hybrid. There are previous
efforts at generating specialised constraint solvers in the literature,
e.g.~\cite{minton}; we aim to use state-of-the-art constraint solver technology
employing a broad range of different techniques. Synthesising a constraint
solver has two key benefits. First, it will enable a fine-grained optimisation
not possible for a general solver, allowing the solving of much larger, more
difficult problems. Second, it will open up many new research possibilities.
There are many techniques in the literature that, although effective in a
limited number of cases, are not suitable for general use. Hence, they are
omitted from current general solvers and remain relatively undeveloped. Among
these are for example conflict recording~\cite{nogoods}, backjumping~\cite{cbj},
singleton arc consistency~\cite{sac}, and neighbourhood inverse
consistency~\cite{inversecons}. The synthesiser will select such techniques as
they are appropriate for an input problem. Additionally, it can also vary basic
design decisions, which can have a significant impact on
performance~\cite{survey}.
\smallskip
The system we are proposing in this paper, Dominion, implements a design that is
capable of achieving said goals effectively and efficiently. The design
decisions we have made are based on our experience with Minion~\cite{minion} and
other constraint programming systems.
\smallskip
The remainder of this paper is structured as follows. In the next section, we
describe the design of Dominion and which challenges it addresses in particular.
We then present the current partial implementation of the proposed system and
give experimental results obtained with it. We conclude by proposing directions
for future work.
\section{Design of a synthesiser for specialised constraint solvers}
\noindent The design of Dominion distinguishes two main parts. The
\emph{analyser} analyses the problem model and produces a solver specification
that describes what components the specialised solver needs to have and which
algorithms and data structures to use. The \emph{generator} takes the solver
specification and generates a solver that conforms to it. The flow of
information is illustrated in Figure~\ref{design}.
\begin{figure}[bt]
\begin{center}
\includegraphics{dc-fig1}
\end{center}
\caption{Components and flow of information in Dominion. The part above the
dashed line is the actual Dominion system. The dotted arrow from the problem
model to the specialised solver designates that either the model is encoded
entirely in the solver such that no further information is required to solve the
problem, or the solver requires further input such as problem
parameters.\label{design}}
\end{figure}
Both the analyser and the generator optimise the solver. While the analyser
performs the high-level optimisations that depend on the structure of the
problem model, the generator performs low-level optimisations which depend on
the implementation of the solver. Those two parts are independent and linked by
the solver specification, which is completely agnostic of the format of the
problem model and the implementation of the specialised solver. There can be
different front ends for both the analyser and the generator to handle problems
specified in a variety of formats and specialise solvers in a number of
different ways, e.g.\ based on existing building blocks or synthesised
from scratch.
\subsection{The analyser}
\noindent The analyser operates on the model of a constraint problem class or
instance. It determines the constraints, variables, and associated domains
required to solve the problem and reasons about the algorithms and data
structures the specialised solver should use. It makes high-level design
decisions, such as whether to use trailing or copying for backtracking memory.
It also decides what propagation algorithms to use for specific constraints and
what level of consistency to enforce.
The output of the analyser is a solver specification that describes all the
design decisions made. It does not necessarily fix all design decisions -- it
may use default values -- if the analyser is unable to specialise a particular
part of the solver for a particular problem model.
In general terms, the requirements for the solver specification are that it
\begin{inparaenum}[(a)]
\item describes a solver which is able to find solutions to the analysed problem
model and
\item describes optimisations which will make this solver perform better than a
general solver.
\end{inparaenum}
The notion of better performance includes run time as well as other resources
such as memory. It is furthermore possible to optimise with respect to a
particular resource; for example a solver which uses less memory at the expense
of run time for embedded systems with little memory can be specified.
The solver specification may include a representation of the original problem
model such that a specialised solver which encodes the problem can be produced
-- the generated solver does not require any input when run or only values
for the parameters of a problem class. It may furthermore modify the original
model in a limited way; for example split variables which were defined as one
type into several new types. It does not, however, optimise it like for example
Tailor~\cite{tailor}.
The analyser may read a partial solver specification along with the model of the
problem to be analysed to still allow fine-tuning by human experts while not
requiring it. This also allows for running the analyser incrementally, refining
the solver specification based on analysis and decisions made in earlier steps.
\smallskip
The analyser creates a constraint optimisation model of the problem of
specialising a constraint solver. The decision variables are the design
decisions to be made and the values in their domains are the options which are
available for their implementation. The constraints encode which parts are
required to solve the problem and how they interact. For example, the
constraints could require the presence of an integer variable type and an equals
constraint which is able to handle integer variables. A solution to this
constraint problem is a solver specification that describes a solver which is
able to solve the problem described in the original model. The weight attached
to each solution describes the performance of the specialised solver and could
be based on static measures of performance as well as dynamic ones; e.g.\
predefined numbers describing the performance of a specific algorithm and
experimental results from probing a specific implementation.
This metamodel enables the use of constraint programming techniques for
generating the specialised solver and ensures that a solver specification can be
created efficiently even for large metamodels.
\smallskip
The result of running the analyser phase of the system is a solver specification
which specifies a solver tailored to the analysed problem model.
\subsection{The generator}
\noindent The generator reads the solver specification produced by the analyser
and constructs a specialised constraint solver accordingly. It may modify an
existing solver, or synthesise one from scratch. The generated solver has to
conform to the solver specification, but beyond that, no restrictions are
imposed. In particular, the generator does not guarantee that the generated
specialised solver will have better performance than a general solver, or indeed
be able to solve constraint problems at all -- this is encoded in the solver
specification.
In addition to the high-level design decisions fixed in the solver
specification, the generator can perform low-level optimisations which are
specific to the implementation of the specialised solver. It could for example
decide to represent domains with a data type of smaller range than the default
one to save space.
The scope of the generator is not limited to generating the source code
which implements the specialised solver, but also includes the system to build
it.
The result of running the generator phase of the system is a specialised solver
which conforms to the solver specification.
\section{Preliminary implementation and experimental results}
\noindent We have started implementing the design proposed above in a system
which operates on top of Minion~\cite{minion}. The analyser reads Minion input
files and writes a solver specification which describes the constraints and the
variable types which are required to solve the problem. It does not currently
create a metamodel of the problem. The generator modifies Minion to support only
those constraints and variable types. It furthermore does some additional
low-level optimisations by removing infrastructure code which is not required
for the specialised solver. The current implementation of Dominion sits between
the existing Tailor and Minion projects -- it takes Minion problem files, which
may have been generated by Tailor, as input, and generates a specialised Minion
solver.
The generated solver is specialised for models of problem instances from the
problem class the analysed instance belongs to. The models have to be the same
with respect to the constraints and variable types used.
Experimental results for models from four different problem classes are shown
in Figure~\ref{results}. The graph only compares the CPU time Minion and the
specialised solver took to solve the problem; it does not take into account the
overhead of running Dominion -- analysing the problem model, generating the
solver, and compiling it, which was in the order of a few minutes for all of the
benchmarks.
\begin{figure}[!tb]
\includegraphics{results.pdf}
\vspace*{-2em}
\caption{Preliminary experimental results for models of instances of four
problem classes. The $x$ axis shows the time standard Minion took to solve the
respective instance. The labels of the data points show the parameters of the
problem instance, which are given in parentheses in the legend. The times
were obtained using a development version of Minion which corresponds to release
0.8.1 and Dominion-generated specialised solvers based on the same version of
Minion. Symbols below the solid line designate problem instances where the
Dominion-generated solver was faster than Minion. The points above the line are
not statistically significant; they are random noise. The dashed line designates
the median for all problem instances.\label{results}}
\end{figure}
The problem classes Balanced Incomplete Block Design, Golomb Ruler, $n$-Queens,
and Social Golfers were chosen because they use a range of different
constraints and variable types. Hence the optimisations Dominion can perform are
different for each of these problem classes. This is reflected in the
experimental results by different performance improvements for different
classes.
Figure~\ref{results} illustrates two key points. The first point is that even a
quite basic implementation of Dominion which does only a few optimisations can
yield significant performance improvements over standard Minion. The second
point is that the performance improvement does not only depend on the problem
class, but also on the instance, even if no additional optimisations beyond the
class level were performed. For both the Balanced Incomplete Block Design and
the Social Golfers problem classes the largest instances yield significantly
higher improvements than smaller ones.
At this stage of the implementation, our aim is to show that a specialised
solver can perform better than a general one. We believe that
Figure~\ref{results} conclusively shows that. As the problem models
become larger and take longer to solve, the improvement in terms of absolute run
time difference becomes larger as well. Hence the more or less constant overhead
of running Dominion is amortised for larger and more difficult problem models,
which are our main focus. Generating a specialised solver for problem classes
and instances is always going to entail a certain overhead, making the approach
infeasible for small and quick-to-solve problems.
\section{Conclusion and future work}
\noindent We have described the design of Dominion, a solver generator, and
demonstrated its feasibility by providing a preliminary implementation. We have
furthermore demonstrated the feasibility and effectiveness of the general
approach of generating specialised constraint solvers for problem models by
running experiments with Minion and Dominion-generated solvers and obtaining
results which show significant performance improvements. These results do not
take the overhead of running Dominion into account, but we are confident that
for large problem models there will be an overall performance improvement
despite the overhead.
Based on our experiences with Dominion, we propose that the next step should be
the generation of specialised variable types for the model of a problem
instance. Dominion will extend Minion and create variable types of the sort
``Integer domain ranging from 10 to 22''. This not only allows us to choose
different representations for variables based on the domain, but also to
simplify and speed up services provided by the variable, such as checking the
bounds of the domain or checking whether a particular value is in the domain.
The implementation of specialised variable types requires generating solvers for
models of problem instances because the analysed problem model is essentially
rewritten. The instance the solver was specialised for will be encoded in it and
no further input will be required to solve the problem. We expect this
optimisation to provide an additional improvement in performance which is more
consistent across different problem classes, i.e.\ we expect significant
improvements for all problem models and not just some.
We are also planning on continuing to specify the details of Dominion and
implementing it.
\section{Acknowledgements}
\noindent The authors thank Chris Jefferson for extensive help with the
internals of Minion and the anonymous reviewers for their feedback. Lars
Kotthof\/f is supported by a SICSA studentship.
|
\section{Introduction}
The study of the viscosity of strongly interacting quantum fluids has brought together very different areas of
physics -- black holes and string theory, quark-gluon plasmas, quantum fluids and cold atoms -- which, at first sight,
appear to have little in common~\cite{Son-review,Schafer09}.
This extraordinary development originated with the work of Son, Starinets and coworkers~\cite{Son-review,Policastro01,Kovtun05}
who calculated the shear viscosity in a strongly interacting quantum field theory,
the ${\cal{N}}=4$ supersymmetric Yang-Mills (SYM) theory,
and conjectured a lower bound
\begin{equation}
{\eta}/{s} \ge {\hbar}/{(4\pi k_B)}
\label{bound}
\end{equation}
for the ratio of the shear viscosity $\eta$ to the entropy density $s$ of \emph{any} system.
These results were obtained using the AdS/CFT formalism where certain strongly coupled
field theories can be mapped onto weakly coupled gravity theories.
Although a number of counterexamples have since been proposed~\cite{Cohen07,Brigante08,Kats09,Buchel09}, there are no known experimental violations of the bound given by Eq.~(\ref{bound}). Remarkably, two very different experimental systems come close to saturating the bound:
the quark-gluon plasma at Brookhaven's Relativistic Heavy Ion Collider~\cite{RHIC,Teaney01}, and
ultracold atomic Fermi gases~\cite{Turlapov08,Gelman05}
close to a Feshbach scattering resonance, where the $s$-wave scattering length becomes infinite~\cite{Trentoreview}.
This is the strongly interacting unitary regime that lies at the center of the BCS-BEC crossover. These two systems are amongst the hottest and coldest systems every realized in a laboratory.
\medskip
In this paper we focus on \emph{non-relativistic} quantum fluids, with particular emphasis on \emph{strongly
interacting Fermi gases}. These are systems for which the most controlled experiments should be possible.
A ``perfect fluid'' with the minimum shear viscosity is necessarily in a quantum regime, since the bound involves $\hbar$.
In addition, it must also be in a strongly interacting regime where \emph{well-defined quasiparticle excitations do not exist}.
If the system had sharp quasiparticles, then their mean scattering rate $\tau^{-1}$ would be much less than the average energy
per particle $\epsilon_0$, so that $\hbar/\tau \ll \epsilon_0$. We can then use Boltzmann's kinetic theory approach to obtain
$\eta \sim n \epsilon_0 \tau$, where $n$ is the number density.
Using $s \sim nk_B$, we find a large $\eta/s \sim \epsilon_0 \tau / k_B \gg {\hbar}/{k_B}$.
Thus, in order to find perfect fluids that come close to saturating the lower bound given by Eq.~(\ref{bound}), one must
look at strongly interacting quantum fluids where the quasiparticle approximation fails.
In this paper we use Kubo formulas for the frequency-dependent
spectral functions for shear viscosity $\eta(\omega)$ and
bulk (or second) viscosity $\zeta(\omega)$, and derive several exact, non-perturbative results without making
weak coupling or quasiparticle approximations. Our main results are:
\medskip
$\bullet$ We establish a microscopic connection between the shear viscosity $\eta$ and the normal fluid density $\rho_n$
and show that a non-zero $\rho_n$ is a necessary condition for a non-vanishing $\eta$.
\medskip
$\bullet$ We derive sum rules for $\eta(\omega)$ and $\zeta(\omega)$ of any Bose or Fermi system with
an arbitrary two-body interaction; see Eqs.~(\ref{etasumruleiso}) and (\ref{zetasumruleiso}).
\medskip
$\bullet$ For a dilute two-component Fermi gas, we find the shear viscosity sum rule
\begin{eqnarray}
\frac{1}{\pi}\int^{\infty}_{0}\!\!d\omega \left[\eta(\omega) -\frac{C}{10\pi\sqrt{m\omega}}\right]=
\frac{\varepsilon}{3}-\frac{ C}{10\pi m a},
\label{etasumrule0finite}
\end{eqnarray}
valid for arbitrary temperature and $1/(k_F a)$, where $a$ is the $s$-wave scattering length.
Here, $\varepsilon$ is the energy density and $C$ is the \textit{contact}~\cite{Tan08}. A central quantity in many of our results,
$C = k_F^4 {\cal C}[1/(k_F a), T/\epsilon_F]$ can be defined via the large-$k$ tail of the momentum distribution
$n_{{\bf k}} \simeq C/k^4$ for $k \gg k_F$, and characterizes the short-distance properties of the many-body state.
\medskip
$\bullet$ For the bulk viscosity, we obtain the sum rule
\begin{eqnarray}
\frac{1}{\pi}\int^{\infty}_{0}d\omega\; \zeta(\omega)=\frac{1}{72 \pi m a^2} \left( \frac{\partial C}{\partial a^{-1}} \right)_s,
\label{zetasumrule0}
\end{eqnarray}
where the derivative is at fixed entropy per particle $s \equiv S/N$. [Different from Eq.~(\ref{bound}), in the remainder of this paper we use $s$ to denote this quantity rather than the entropy density.]
Below the superfluid transition, the bulk viscosity that enters Eq.~(\ref{zetasumrule0}) is $\zeta_2$,
associated with the damping of in-phase motions of the superfluid and normal components. The positivity
of $\zeta(\omega)$ implies that $(\partial C / \partial a^{-1})_s \geq 0$.
\medskip
$\bullet$ At unitarity, the bulk viscosity spectral function vanishes at \emph{all} frequencies and \emph{all} temperatures. Quite generally, $\zeta(\omega) \geq 0$,
but the sum rule in Eq.~(\ref{zetasumrule0}) vanishes for $|a|=\infty$ and thus $\zeta(\omega) = 0$ for the unitary Fermi gas.
This generalizes the result~\cite{Son07} that the static bulk viscosity $\zeta(0)$ vanishes at unitarity.
\medskip
$\bullet$ It follows from the previous result that, at unitarity, the shear viscosity spectral function $\eta(\omega)$ can be
related to density-density correlations as
\begin{eqnarray}
\eta(\omega)
= \lim_{q\to 0}\frac{3\omega^3}{4 q^4}\;\mathrm{Im}\chi_{\rho\rho}({\bf q},\omega) \ \ \ \ \ (|a| = \infty).
\end{eqnarray}
Thus, $\eta(\omega)$ for the unitary Fermi gas
can be measured spectroscopically using, for instance, two-photon Bragg spectroscopy.
\medskip
$\bullet$ We show from our sum rules that various dynamic response functions for Fermi gases
have high-frequency tails characterized by odd-integer power laws,
whose magnitudes are controlled by the contact $C$. The tail $C/\sqrt{\omega}$ of $\eta(\omega)$ is evident from
Eq.~(\ref{etasumrule0finite}). Using this we find that the dynamic structure factor has a tail of the form~\cite{Son10}
$S({{\bf q}},\omega) \sim C q^4/\omega^{7/2}$ for
$q \to 0$ and $\omega \to \infty$, which is shown to be a generic feature of short range physics.
\bigskip
In the remainder of this Section we describe how the rest of the paper is organized.
In Section~\ref{Kubosec}, we begin with a careful derivation of Kubo formulas for the spectral functions
$\eta(\omega)$ and $\zeta(\omega)$ in terms of current-current correlation functions,
Eqs.~(\ref{KuboTR}) and (\ref{KuboLR}), and, equivalently, in terms of the stress-stress correlator, Eq.~(\ref{KuboPiFull}).
In Section~\ref{rhon-section} we recall some elementary facts about the shear viscosity of a fluid and why it
is analogous to the resistivity, and not the conductivity, of a metal. We then establish a connection between the viscosity $\eta$ and
the normal fluid density $\rho_n$ using microscopic response functions.
After establishing the positivity of $\eta(\omega)$ and of $\zeta(\omega)$ in Section~\ref{posdefsec},
we derive sum rules for these quantities in Section~\ref{sumrulesec}. The most general sum rules for
the shear and bulk viscosities of any Bose or Fermi system with an arbitrary isotropic interaction potential $V(p)$,
and valid for all temperatures, are given in Eqs.~(\ref{etasumruleiso}) and (\ref{zetasumruleiso}).
In Section~\ref{diluteviscositysec}, we specialize to the dilute Fermi gas, with interparticle spacing
$k_F^{-1}$ and $s$-wave scattering length $a$ both much larger than the characteristic range $r_0$ of the potential.
We obtain the $\zeta$ sum rule in Eq.~(\ref{zetasumrule0}), which is finite in the zero range limit.
The $\eta$ sum rule, however, has an ultraviolet divergence; see Eq.~(\ref{etasumrule2}).
We identify, in Section~\ref{highfrequencytailsec}, the $C/\sqrt{\omega}$ high-frequency tail of the shear viscosity spectral function,
and derive the sum rule given by Eq.~(\ref{etasumrule0finite}), which is manifestly finite for $r_0 \to 0$.
The sum rules given by Eqs.~(\ref{etasumrule0finite}) and (\ref{zetasumrule0}) are valid in both the normal and superfluid phases,
with $\zeta$ replaced by $\zeta_2$ in the latter state.
In Section~\ref{Crossoversec} we show from the $\zeta$ sum rule that, at unitarity, $\zeta(\omega)$ vanishes at all frequencies
and all temperatures.
We also discuss the $1/(k_F a)$-dependence of the $\eta$ and $\zeta$ sum rules across the BCS-BEC crossover, using
available quantum Monte Carlo data for the energy density at $T=0$.
We discuss the connection between viscosity and density-density correlations in Section~\ref{SqomegaSec} and find
two interesting results. First, we show how a density probe such as two-photon Bragg spectroscopy can in principle be used to measure the shear
viscosity spectral function $\eta(\omega)$ at unitarity.
Second, we identify the high-frequency $\omega^{-7/2}$ tail in the dynamic structure factor $S({{\bf q}},\omega)$.
In Section~\ref{comparisonsec}, we briefly compare the sum rules that we have derived for non-relativistic
quantum fluids with those obtained in relativistic quantum field theories. Finally in Section~\ref{conclusionssec}
we conclude with open questions.
There are five Appendices which contain technical details of derivations or review certain results which are used at
various places in the paper. In Appendix~\ref{KuboPi0}, we briefly discuss an alternate stress tensor operator often used to calculate the shear viscosity. Some results from dissipative two-fluid hydrodynamics, which we use in our paper, are reviewed in Appendix~\ref{hydrosec}.
We review in Appendix \ref{ContactAppendix} results related to the contact that are used at several places in the paper,
and also give a detailed derivation of certain equations that involve the contact.
In Appendix~\ref{Pressuresec}, we derive a microscopic expression for the pressure. Finally, in
Appendix \ref{universalthermosec} we give details of the derivation of the $\zeta$ sum rule which make use of the scaling form
of thermodynamic functions across the BCS-BEC crossover.
\section{Kubo formula for viscosity}
\label{Kubosec}
We begin by deriving Kubo formulas for the bulk and shear viscosity. Although the results of this Section
are, for the most part, ``well known'', we could not find a complete derivation at any one place in the literature.
In particular, there are several subtle points not dealt with adequately
elsewhere, not least the definition of the stress tensor operator
$\widehat{\Pi}_{\alpha\beta}$ for non-relativistic systems.
To introduce notation, we start with the Euler equation
\begin{eqnarray}
m\partial_t j_{\alpha}({\bf r},t)= -\partial_{\beta}\Pi_{\alpha\beta}({\bf r},t),
\label{euler}
\end{eqnarray}
where $m$ is the mass of the particles, $j_{\alpha}$ is the (number) current and
$\Pi_{\alpha\beta}$ is the momentum flux density tensor, which we call the \emph{stress tensor}, for short.
Here, $\alpha$ and $\beta$ take on values $x,y,z$ (and there is no difference between upper and lower indices
in our non-relativistic formulation).
In general, the stress tensor is given by~\cite{LLFM}
\begin{eqnarray}
\Pi_{\alpha\beta} = P\delta_{\alpha\beta} + \rho u_{\alpha}u_{\beta} -\sigma'_{\alpha\beta},
\label{stress}
\end{eqnarray}
where $P$ is the pressure, $\rho$ the mass density and ${\bm u}$ the velocity.
The viscous term $\sigma'_{\alpha\beta}$ is given by
\begin{eqnarray}
\sigma'_{\alpha\beta}= \eta\left[\partial_{\beta} u_{\alpha} \!+\! \partial_{\alpha}u_{\beta} \!-\! \frac{2}{3}\delta_{\alpha\beta}({\bm \nabla}\cdot{\bm u})\right] \!+\! \zeta\delta_{\alpha\beta}({\bm \nabla}\cdot{\bm u}),
\label{stresstensor}
\end{eqnarray}
where $\eta$ is the shear viscosity and $\zeta$ the bulk viscosity.
The generalization of Eq.~(\ref{stresstensor}) to the superfluid state is well known~\cite{LLFM} and
involves additional bulk viscosities.
At the end of Section~\ref{currentcorrelatorsec}, we show that the Kubo formula we derive for $\zeta$
describes the bulk viscosity $\zeta_2$ in the superfluid phase.
Our goal is to obtain Kubo formulas for frequency-dependent generalizations of
the long-wavelength viscosities, $\eta$ and $\zeta$, in terms of equilibrium correlation functions
of the many-body system.
The Kubo formulas for viscosities are often written in terms of the stress-stress
correlators; see, e.g., Sec.~90 of Ref.~\cite{LLStatPhysII}.
However, the form of the stress tensor (or momentum flux density) \emph{operator}
$\widehat{\Pi}_{\alpha\beta}$ is not obvious, and many different, complicated
expressions \cite{Forster} which are presumably equivalent can be found in the literature.
Part of the problem is to write down an operator expression
for the pressure $P$ in terms of particle positions and momenta. In high-energy physics, a simple way to calculate the stress-energy tensor $\widehat{\Pi}_{\alpha\beta}$ is to vary the action with respect to the metric in curved space-time. We prefer, however, to describe non-relativistic fluids without going to curved space-time.
To begin with, in~\ref{currentcorrelatorsec}, we adopt an approach that permits us to get around the
complexities of defining the stress operator $\widehat{\Pi}$. We
consider the linear response of a fluid to an externally imposed velocity field
and derive Kubo formulas for the bulk and shear viscosities in terms of \emph{current-current correlators}.
The results of this subsection are the same as those of Kadanoff and Martin~\cite{Kadanoff63}.
In~\ref{Kubocomparisonsec}, we use an operator form of Euler's equation to make the connection between bulk and shear
viscosities and \emph{stress-stress correlators}.
In Appendix~\ref{KuboPi0}, we derive an alternative form of the stress correlator, which works
only for the shear viscosity in the zero-frequency limit, but is often used in calculations.
\subsection{Current correlators}
\label{currentcorrelatorsec}
We calculate within linear response theory~\cite{NozieresPines1,Baymbook} the current flow in a fluid
subjected to an external velocity field ${\bm u}({\bf r},t) = {\bm u}({\bf r}) e^{-i\omega t} e^{0^+t}$ which is
turned on adiabatically.
Our goal is to relate the imaginary part of this response function
to viscosity through the dissipative part of the stress tensor.
The response of a fluid to the ``moving walls'' of its container is a standard concept in the theory of
superfluidity~\cite{Baymbook}. Here, we generalize this analysis to a non-uniform and time-varying
external perturbation ${\bm u}({\bf r},t)$, taking the long wavelength limit at the end.
We write the Hamiltonian of the system $\hat{H}$ plus external perturbation $\hat{H}'$ as~\cite{wallnote}
\begin{eqnarray}
\hat{H}_{\rm total} &=& \frac{1}{2m}\sum_{i=1}^N \int d{\bf r}\;
\left[\hat{{\bf p}}_i - m{\bm u}({\bf r},t)\delta({\bf r}-\hat{{\bf r}}_i)\right]^2 + \hat{V}\nonumber\\
=\hat{H} \! &-& \!\frac{1}{2}\sum_{i=1}^N \int d{\bf r} {\bm u}({\bf r},t)\cdot
\left\{ \hat{{\bf p}}_i,\delta({\bf r}\!-\!\hat{{\bf r}}_i) \right\}
+ {\cal{O}}[{\bm u}^2],\label{Hshift}\end{eqnarray}
where $\hat{{\bf p}}_i$ and $\hat{{\bf r}}_i$ are the momentum and position operators, respectively, for the $i$th particle,
$m$ is the mass, and $\hat{V}$ is the potential energy operator.
The anticommutator $\{ \hat{A},\hat{B} \} = \hat{A}\hat{B} + \hat{B}\hat{A}$ is used to
symmetrize products.
We thus see that to linear order in ${\bm u}$, the external perturbation is
\begin{eqnarray}
\hat{H}'(t) = - m\int d{\bf r} e^{-i\omega t} e^{0^+t} {\bm u}({\bf r})\cdot\hat{\bm{j}}({\bf r},t),
\label{Hpert}
\end{eqnarray}
where
$\hat{\bm{j}} = \sum_{i=1}^N \left\{ \hat{{\bf p}}_i , \delta({\bf r}\!-\!\hat{{\bf r}}_i) \right\}/2m$
is the current density operator.
Linear response theory gives the result~\cite{NozieresPines1,Baymbook}
\begin{eqnarray}
\lefteqn{\langle \hat{j}_{\alpha}({\bf r},t)\rangle = m \int d{\bf r}'\!\int^{\infty}_{-\infty}\!\!\! dt'e^{0^+t'}e^{-i\omega t'}\times}&&\nonumber\\&&\;\;\;\;\;\;\;\;\;\;\;\; \chi^{\alpha\beta}_{J}({\bf r}-{\bf r}',t-t')u_{\beta}({\bf r}').\label{lr1}
\end{eqnarray}
Here and below, we use the standard convention of summing over repeated indices.
The retarded current correlation function
$\chi^{\alpha\beta}_J$ is obtained by using
$\hat{A}=\hat{j}_{\alpha}$ and $\hat{B}=\hat{j}_{\beta}$ in Eq.~(\ref{chiAB}) below.
For later use, we provide a general definition for the \emph{retarded} response function,
or correlator, for operators $\hat{A}$ and $\hat{B}$:
\begin{eqnarray}
\lefteqn{\chi_{A,B}({\bf r}-{\bf r}',t-t') \equiv}&&\nonumber\\&& i\Theta(t-t')\langle[\hat{A}({\bf r},t),\hat{B}^{\dagger}({\bf r}',t')]\rangle.
\label{chiAB} \end{eqnarray}
Here, $\langle \hat{Q}\rangle = {\rm Tr} [\hat{Q} \exp(-\hat{H}/T)]/{\cal{Z}}$ is
the thermal expectation value at temperature $T$ and
${\cal{Z}} = {\rm Tr}[\exp(-\hat{H}/T)]$ is the partition function.
The step-function $\Theta(t-t')$ enforces causality.
We will use the convention of unit volume $\Omega = 1$ and set $\hbar = k_B = 1$,
unless explicitly stated otherwise.
We find the spectral representation for $\chi_{A,B}$ using the exact eigenstates and eigenvalues of the fully interacting
many-body Hamiltonian $\hat{H}|a\rangle = E_a|a\rangle$, and Fourier transform the result to obtain
\begin{eqnarray}
\lefteqn{\chi_{A,B}({\bf q},\omega) = \frac{1}{{\cal{Z}}}\sum_{a,b}e^{-E_a/T}\times}&&\nonumber
\\&&\left[
\frac{\langle a|\hat{B}^{\dagger}_{{\bf q}}|b\rangle\langle b|\hat{A}_{{\bf q}}|a\rangle}{\omega + E_{ba} + i0^+}
- \frac{\langle a|\hat{A}_{{\bf q}}|b\rangle\langle b|\hat{B}^{\dagger}_{{\bf q}}|a\rangle}{\omega - E_{ba} + i0^+}
\right],
\label{FourierchiAB}
\end{eqnarray}
where $E_{ba}\equiv E_b-E_a$.
The quantity of central interest to us in this paper is the imaginary part of $\chi$, given by
\begin{eqnarray}
\lefteqn{\mathrm{Im}\chi_{A,B}({\bf q},\omega) = \pi(1-e^{-\omega/T})}&&\nonumber\\&&
\times\frac{1}{\cal{Z}}\sum_{a,b}e^{-\beta E_a}\langle a|\hat{A}_{{\bf q}}|b\rangle\langle b|\hat{B}^{\dagger}_{{\bf q}}|a\rangle
\delta(\omega -E_{ba}).
\label{ImchiAB}
\end{eqnarray}
Returning to the problem of interest, we find that the induced current,
obtained by Fourier transforming Eq.~(\ref{lr1}), is
\begin{eqnarray}
\langle \hat{j}^{\alpha}({\bf q},\omega)\rangle = m \chi^{\alpha\beta}_{J}({\bf q},\omega)u_{\beta}.
\label{lr}
\end{eqnarray}
$\chi^{\alpha\beta}_J$ is given by Eq.~(\ref{FourierchiAB}) with $\hat{A}_{{\bf q}}=\hat{j}^{\alpha}_{{\bf q}}$ and
$\hat{B}^{\dagger}_{{\bf q}} = \hat{j}^{\beta}_{-{\bf q}}$, where
\begin{eqnarray}
\hat{j}^{\alpha}_{{\bf q}} = \frac{1}{2m}\sum_{{\bf k}\sigma} (2k_{\alpha}+q_{\alpha}) \hat{c}^{\dagger}_{{\bf k}\sigma}\hat{c}_{{\bf k}+{\bf q}\sigma}
\label{jdef}\end{eqnarray}
is the current operator with $\sigma$ denoting the different internal states of interest (e.g., spin).
Next, we need to relate Eq.~(\ref{lr}) to viscosity,
using ``constitutive relations'' between the current and transport coefficients.
For this we use Eqs.~(\ref{stress}) and (\ref{stresstensor}) substituted into Eq.~(\ref{euler}),
where the symbols $j_{\alpha}$ and $\Pi_{\alpha\beta}$,
\emph{without the hats used for operators}, are understood to
denote expectation values. In the long-wavelength limit, the contributions to the stress tensor coming from viscous terms dominate
over contributions from pressure fluctuations, while the convective term $\partial_{\beta}u_{\alpha}u_{\beta}$ is beyond linear order in velocity. We thus get
$m\partial_t j^{\alpha}=\zeta\partial_{\alpha}({\bm \nabla}\cdot{\bm u}) + \eta\left[\nabla^2 u_{\alpha}+ \partial_{\alpha}({\bm \nabla}\cdot{\bm u})/3\right]$.
Fourier transforming and comparing with Eq.~(\ref{lr}), we obtain
\begin{eqnarray}
\zeta q_{\alpha}q_{\beta}u_{\beta} + \eta\left(\!q^2u_{\alpha}\! +\! \frac{1}{3}q_{\alpha}q_{\beta}u_{\beta}\!\right)\! =\!- i\omega m^2 \chi^{\alpha\beta}_{J}({\bf q},\omega)u_{\beta}.\nonumber\\
\label{Kubo1}
\end{eqnarray}
We decompose the current correlation function into its
longitudinal ($\chi_L$) and transverse ($\chi_T$) components:
\begin{eqnarray}
\chi^{\alpha\beta}_{J} = \chi_L\frac{q_{\alpha}q_{\beta}}{q^2} + \chi_T\left(\delta_{\alpha\beta}-\frac{q_{\alpha}q_{\beta}}{q^2}\right)
\label{chiTL}
\end{eqnarray}
By taking appropriate $q \to 0$ limits \cite{limits} of Eq.~(\ref{Kubo1}) we find
\begin{eqnarray}
\eta(\omega) = \lim_{q\to 0}{(-i\omega)}m^2 \chi_T({\bf q},\omega)/{q^2}
\label{KuboT}
\end{eqnarray}
and
\begin{eqnarray}
\zeta(\omega) + {4\eta(\omega)}/{3} = \lim_{q\to 0}{(-i\omega)}m^2 \chi_L({\bf q},\omega)/{q^2}.
\label{KuboL}\end{eqnarray}
These expressions define the \emph{complex} shear and bulk viscosities.
We will be interested in the properties and sum rules of
the \emph{spectral functions}:
\begin{eqnarray}
\mathrm{Re}\; \eta(\omega) = \lim_{q\to 0}{m^2 \omega}\mathrm{Im}\chi_T({\bf q},\omega)/{q^2}
\label{KuboTR}
\end{eqnarray}
and
\begin{eqnarray}
\mathrm{Re}\; \zeta(\omega) + {4\mathrm{Re} \;\eta(\omega)}/{3} = \lim_{q\to 0}
{m^2 \omega}\mathrm{Im}\chi_L({\bf q},\omega)/{q^2}.
\label{KuboLR}
\end{eqnarray}
The static viscosities $\eta$ and $\zeta$ introduced in Eq.~(\ref{stresstensor})
are $\eta \equiv \mathrm{Re}\; \eta(\omega=0)$ and $\zeta \equiv \mathrm{Re}\; \zeta(\omega=0)$.
In closing this subsection, we note that the Kubo formulas for the viscosity derived here and below
are valid in both the normal and superfluid phases, provided we recognize that the bulk viscosity
in the superfluid state refers to $\zeta_2$, which describes damping associated with an in-phase motion of the superfluid and normal fluid components~\cite{LLFM}. To understand this in more detail, we recall Landau's two-fluid hydrodynamics~\cite{LLFM} for the
superfluid state. In this theory, three bulk viscosities, $\zeta_1$, $\zeta_2$, and $\zeta_3$, are required
to describe the dissipation associated with different types of relative motions of the superfluid and normal components.
The \emph{longitudinal} response does not distinguish between the superfluid and normal components \cite{Baymbook} and
thus forces the superfluid and normal fluid velocities to be equal: ${\bf v}_s = {\bf v}_n = {\bm u}$. When both components
flow with the same velocity, the two-fluid hydrodynamic stress tensor [see Eq.~(140.5) in Ref.~\cite{LLFM}] reduces to the expression in
Eq.~(\ref{stresstensor}), with $\zeta$ replaced by $\zeta_2$, the bulk viscosity associated with the damping of the
in-phase motions of the superfluid and normal fluid components.
One can also show by direct application of Eq.~(\ref{KuboLR2}) to the two-fluid
hydrodynamic density response function in Eq.~(\ref{chirhorho}) that the left-hand side of Eq.~(\ref{KuboLR2}) is $\zeta_2+4\eta/3$ in the
low-frequency two-fluid hydrodynamic regime.
\subsection{Stress correlators}
\label{Kubocomparisonsec}
We next derive Kubo formulas equivalent to those derived above but expressed in terms of
the correlators of a suitably defined stress tensor operator $\widehat{\Pi}_{\alpha\beta}$.
These are useful to make connections with the literature~\cite{Kovtun05,Bruun05,Peshier05}.
We will also use these results in connection with the positivity of the bulk viscosity
spectral function and its vanishing for the unitary Fermi gas.
The $\widehat{\Pi}_{\alpha\beta}$ operator must satisfy
\begin{eqnarray}
i m [\hat{j}_{\alpha},\hat{H}] = \partial_{\beta}\widehat{\Pi}_{\alpha\beta},
\label{euler-op}
\end{eqnarray}
which is the operator version of the Euler equation, Eq.~(\ref{euler}), and is simply a statement of momentum conservation.
We go to Fourier space and relate matrix elements of the current operator to those of the stress tensor
by sandwiching Eq.~(\ref{euler-op}) between exact many-body eigenstates.
Using the spectral representation in Eq.~(\ref{FourierchiAB}) we can then relate the current correlator
$\chi^{\alpha\beta}_{J}({\bf q},\omega)$ to the stress correlator
$\chi^{\alpha\beta,\mu\nu}_{\Pi}({\bf q},\omega)$. The latter is defined by choosing
$\hat{A}=\hat{\Pi}^{\alpha\beta}({\bf q})$ and $\hat{B}=\hat{\Pi}^{\mu\nu}(-{\bf q})$ in Eq.~(\ref{FourierchiAB}).
For simplicity we calculate only $\chi^{xx}_{J}$, which will suffice for our purposes.
The final result, after some simple algebra, is
\begin{eqnarray}
m^2\omega^2\chi^{xx}_{J}({\bf q},\omega) &=&q_{\alpha}q_{\beta}\chi^{x\alpha,x\beta}_{\Pi}({\bf q},\omega) -
mq_{\alpha}\langle[\widehat{\Pi}^{x\alpha}_{{\bf q}},\hat{j}^{x}_{-{\bf q}}]\rangle.\nonumber\\ \!\!\!\!\!\!\!
\label{Correlationrelation}
\end{eqnarray}
Note that $\widehat{\Pi}^\prime_{\alpha\beta} = \widehat{\Pi}_{\alpha\beta} + \hat{\Lambda}_{\alpha\beta}$,
with any symmetric tensor $\hat{\Lambda}$ satisfying $\partial_{\beta}\hat{\Lambda}_{\alpha\beta} = 0$, will
also be a solution to the Euler equation, Eq.~(\ref{euler-op}).
This non-uniqueness in the definition of $\widehat{\Pi}$ does not affect our final results
for the viscosity, related to $\chi^{xx}_{J}$, since a symmetric $\hat{\Lambda}$
with $q_{\beta}\hat{\Lambda}_{\alpha\beta} = 0$ makes no contribution to
Eq.~(\ref{Correlationrelation}).
Using the decomposition given by Eq.~(\ref{chiTL}), and taking the appropriate limits, we find
\begin{eqnarray}
m^2\omega^2\lim_{q\to 0}\frac{\chi_T}{q^2} = \lim_{q\to 0}
\left[ \chi^{xy,xy}_{\Pi} - \frac{m}{q} \langle[\widehat{\Pi}^{xy}_{{\bf q}},\hat{j}^{x}_{-{\bf q}}]\rangle \right],
\label{TPi}
\end{eqnarray}
where we have taken $q_x$ and $q_z$ to zero before $q_y$, and
\begin{eqnarray} m^2\omega^2\lim_{q\to 0}\frac{\chi_L}{q^2} = \lim_{q\to 0} \left[ \chi^{xx,xx}_{\Pi}- \frac{m}{q} \langle[\widehat{\Pi}^{xx}_{{\bf q}},\hat{j}^{x}_{-{\bf q}}]\rangle \right],
\label{LPi}
\end{eqnarray}
where we have taken $q_y$ and $q_z$ to zero before $q_x$. We note that the commutators on the right hand sides
of Eqs.~(\ref{TPi}) and (\ref{LPi}) only affect the real parts of $\chi_T$ and $\chi_L$ and not the
spectral functions of interest, shown in the next two equations.
Using the Kubo formulas given by Eqs.~(\ref{KuboTR}) and (\ref{KuboLR})
that were derived above, we find
\begin{eqnarray}
\mathrm{Re}\; \eta(\omega) = \lim_{q\to 0}{\mathrm{Im}\chi^{xy,xy}_{\Pi}({\bf q},\omega)}/{\omega}
\label{KuboPiT}
\end{eqnarray}
and
\begin{eqnarray}
\mathrm{Re}\; \zeta(\omega) + {4\mathrm{Re} \;\eta(\omega)}/{3}
= \lim_{q\to 0}{\mathrm{Im}\chi^{xx,xx}_{\Pi}({\bf q},\omega)}/{\omega}.
\label{KuboPiL}
\end{eqnarray}
In an isotropic system, in the $q \to 0$ limit, the only fourth rank tensor allowed by
symmetry is of the form $A \delta_{\alpha\beta} \delta_{\mu\nu}
+ B\left( \delta_{\alpha\mu} \delta_{\beta\nu} + \delta_{\alpha\nu} \delta_{\beta\mu} \right)$.
We can thus combine Eqs.~(\ref{KuboPiT}) and (\ref{KuboPiL}) to write
\begin{eqnarray}
\left[\mathrm{Re}\; \zeta - \frac{2}{3} \mathrm{Re}\; \eta \right] \delta_{\alpha\beta} \delta_{\mu\nu}
&+& \mathrm{Re}\; \eta \left( \delta_{\alpha\mu} \delta_{\beta\nu} + \delta_{\alpha\nu} \delta_{\beta\mu} \right)
\nonumber
\\
&=& \lim_{q\to 0}\frac{\mathrm{Im}\chi^{\alpha\beta,\mu\nu}_{\Pi}({\bf q},\omega)}{\omega}.
\label{KuboPiFull}
\end{eqnarray}
A very useful formula for the bulk viscosity follows from Eq.~(\ref{KuboPiFull})
by looking at its $(xx,yy)$ component and combining it with the $(xx,xx)$ component in Eq.~(\ref{KuboPiL}).
Using the summation convention, we thus obtain
\begin{eqnarray}
\mathrm{Re}\; \zeta(\omega) = \lim_{q\to 0}\frac{\mathrm{Im}\chi^{\alpha\alpha,\beta\beta}_{\Pi}({\bf q},\omega)}{9\;\omega}.
\label{KuboPiZeta}
\end{eqnarray}
We emphasize again that the Kubo formulas for the bulk and shear viscosities expressed in terms
of the stress-stress correlation function are equivalent to
those expressed in terms of current-current correlations, Eqs.~(\ref{KuboTR}) and (\ref{KuboLR}).
The two sets of equations are simply related by the exact conservation law, Eq.~(\ref{euler-op}).
Above, we focused on the \emph{dissipative} parts of the response, i.e., the \emph{real} parts of the viscosities.
Comparing Eqs.~(\ref{KuboT}) and (\ref{KuboL}) with Eqs.~(\ref{TPi}) and (\ref{LPi}), we
see that the imaginary part of $\eta$ and the imaginary part of $\left(4\eta/3 + \zeta\right)$ are \textit{not}
given by $\lim_{\omega\to 0}\lim_{q\to 0}\mathrm{Re}\chi^{xy,xy}_{\Pi}/\omega$ and $\lim_{\omega\to 0}\lim_{q\to 0}\mathrm{Re}\chi^{xx,xx}_{\Pi}/\omega$, respectively. $\mathrm{Im}\; \eta$ and $\mathrm{Im}\; \zeta$,
when written in terms of stress correlators, also involve the frequency-independent, equal-time commutator terms in Eqs.~(\ref{TPi}) and (\ref{LPi}).
This point seems to be missed in treatments that start out with the stress correlator formalism.
The imaginary parts of the transport coefficients
are most simply expressed in terms of the current correlation functions, Eqs.~(\ref{KuboT}) and (\ref{KuboL}).
In the $\omega \to 0$ limit, the validity of this assertion can be seen quite independently from
hydrodynamics (see Appendix~\ref{hydrosec}). Allowing $\eta$ to be complex in the hydrodynamic expression for the transverse current correlation function in Eq.~(\ref{chiT}), for instance, one can readily confirm that the imaginary part of the shear viscosity is indeed given by Eq.~(\ref{KuboT}).
\section{Shear viscosity and normal fluid density}
\label{rhon-section}
In this Section we discuss the relation between the static shear viscosity $\mathrm{Re}\; \eta(\omega=0)$
and the normal fluid density $\rho_n$, both of which can be written in terms of the \emph{transverse} current-current
correlation function. This allows us to prove that a non-zero
normal fluid density $\rho_n$ is a necessary condition for a non-vanishing shear viscosity $\eta$.
This is, perhaps, not entirely unexpected on physical grounds, but we are unaware of
a microscopic proof, valid for all Galilean invariant Bose or Fermi quantum fluids,
that does not rely on a quasiparticle approximation.
Before turning to the calculation, it may be useful to review some elementary facts
about the shear viscosity $\eta$. Given that there is a Kubo formula for $\eta(\omega)$
in terms of the current-current correlation function, Eq.~(\ref{KuboTR}), and that in kinetic theory
$\eta$ is proportional to the mean free path, it may seem natural to assume
that the shear viscosity of a fluid is the analog of metallic conductivity.
This, however, is completely misleading. The shear viscosity is, in fact, the analog of the \emph{resistivity}.
This is clear, e.g., from the classical formula of Poiseuille
for the flow rate $Q = {\pi R^4 \Delta P}/(8 \eta L)$, with a pressure difference $\Delta P$ across
a cylindrical pipe of radius $R$ and length $L$.
We will see below that zero viscosity in a superfluid is the analog of zero resistance in
a superconductor.
We begin by rewriting the Kubo formula for the shear viscosity, given by Eq.~(\ref{KuboTR}), using the spectral representation in Eq.~(\ref{ImchiAB}):
\begin{eqnarray}
\mathrm{Re}\eta(\omega) = \lim_{(T)} \frac{\pi m^2}{\cal{Z}}
\sum_{a,b}\left[e^{-\beta E_a}\;-\;e^{-\beta E_b}\right] E_{ba}
\nonumber\\
\times\frac{\vert \langle b|\hat{j}^x_{{\bf q}}|a\rangle \vert^2}{q^2}
\delta(\omega -E_{ba}),
\label{etaomega}
\end{eqnarray}
Here and below, the ``transverse limit'', denoted by $ \lim_{(T)}$, means that for $\chi_J^{xx}$
we first set $q_x = 0$ and then take the limit $q_y \to 0$.
The normal fluid density $\rho_n$ characterizes the response of a fluid to
moving walls and determines the moment of inertia of a cylinder containing the fluid; see, e.g.,
the detailed discussion in Refs.~\cite{Baymbook,NozieresPines2}. It is defined in terms of the
real part of the static transverse current correlator:
\begin{eqnarray}
\rho_n = \lim_{q \to 0} m^2 \mathrm{Re}\chi_T({\bf q},\omega=0).
\label{rhon1}
\end{eqnarray}
Using the spectral representation in Eq.~(\ref{FourierchiAB}) for $\chi_J^{xx}$, we
can rewrite this result as
\begin{eqnarray}
\rho_n = \lim_{(T)} \frac{m^2}{\cal{Z}}
\sum_{a,b}\frac{\left[e^{-\beta E_a}\;-\;e^{-\beta E_b}\right]}{E_{ba}}
{\vert \langle b|\hat{j}^x_{{\bf q}}|a\rangle \vert^2}.
\label{rhon2}
\end{eqnarray}
Our goal now is to understand the connection between the shear viscosity
$\eta$, which is obtained by taking the $\lim_{\omega \to 0}\lim_{q \to 0}$ of $\mathrm{Im}\chi_T$
in Eq.~(\ref{etaomega}), and the normal fluid density $\rho_n$, which is
the $\lim_{q \to 0}\lim_{\omega \to 0}$ of $\mathrm{Re}\chi_T$ in Eq.~(\ref{etaomega}).
In lattice models of superconductors, it has been suggested \cite{Scalapino} that the order of the
$q$ and $\omega$ limits can be safely interchanged for the \emph{transverse} current correlator,
because all ``transverse'' excitations are gapped (unlike longitudinal excitations such as
phonons in charge-neutral systems). However, this argument is \emph{not} valid for the systems
of interest to us. This can be seen, e.g., from the hydrodynamic form of $\chi_T$ in Eq.~(\ref{chiT})
which has a ``diffusion pole'' that makes the order of limits quite different.
To prove the result stated at the beginning of this Section, we will show that
$\rho_n = 0$ implies $\eta = 0$. The starting condition $\rho_n = 0$ makes sense
only at $T=0$, since at any finite temperature there will necessarily be some thermal excitations.
Furthermore, the vanishing of the normal fluid density
\begin{eqnarray}
\rho_n = \lim_{(T)} 2 m^2
\sum_{b}\frac{\vert \langle b|\hat{j}^x_{{\bf q}}|0\rangle \vert^2}{E_{b0}}
\label{rhon3}
\end{eqnarray}
at $T=0$ implies that each term in the sum $\sum_b$ over states vanishes.
This means that, for each state $|b\rangle $, if the excitation energy varies as $\lim_{(T)}E_{b0} \sim q^{\alpha_b}$, with
$\alpha_b \geq 0$, then the matrix element of the current operator vanishes even faster:
$\lim_{(T)}\vert \langle b|\hat{j}^x_{{\bf q}}|0\rangle \vert \sim q^{\alpha_b + \beta_b}$
with $\beta_b > 0$. Note that we are not making any assumptions about the nature of the
spectrum since both gapless ($\alpha_b > 0$) and gapped ($\alpha_b = 0$) excitations are permitted.
In either case, the matrix element of $\hat{j}^x_{{\bf q}}$ vanishes, since the $q \to 0$ limit
of $\hat{j}^x_{{\bf q}}$ is the total momentum, which commutes with the Hamiltonian in a Galilean invariant system.
It is only in such a system that $\rho_n$ vanishes at $T=0$~\cite{NozieresPines2,Paramekanti}.
Now that we have constrained the matrix elements for any form of the excitation spectrum given $\rho_n = 0$,
we now ask how these constraints impact the shear viscosity. We look separately at the contribution from
gapless and gapped states to Eq.~(\ref{etaomega}), which at $T=0$ can be written as
\begin{eqnarray}
\eta(\omega) = \lim_{(T)} {\pi m^2}
\sum_{b} E_{b0}\frac{\vert \langle b|\hat{j}^x_{{\bf q}}|0\rangle \vert^2}{q^2}
\delta(\omega -E_{b0}).
\label{etaomega2}
\end{eqnarray}
Each gapless state $b$, with $\alpha_b > 0$, will contribute a term
$\lim_{(T)} q^{2\alpha_b + \beta_b - 2}\delta(\omega - A_b q^{\alpha_b})$,
which gives a vanishing contribution \cite{eta-zero} in the limit $q \to 0$ for all $\omega > 0$.
Finally taking the $\omega \to 0$ limit, we find that the contribution of the gapless states
to $\eta$ vanishes.
Next, consider the gapped states with $\alpha_b = 0$, so that
$\lim_{(T)}E_{b0} \equiv \Delta_b > 0$. Their contribution to Eq.~(\ref{etaomega2})
yields an expression of the form
$\eta(\omega) = \lim_{(T)}
\sum_{b}^\prime C_b q^{\beta_b - 2} \delta(\omega -\Delta_{b})$,
where the prime indicates a sum over all gapped states. This result contributes
to both the $\eta$ sum rule and the high-frequency tail that we will
derive later in the paper. The important point here is that
for $0< \omega < \min_b^\prime\left\{\Delta_b\right\}$, i.e., below the minimum gap
of all excitations, $\eta(\omega) = 0$.
Thus, we conclude that
the vanishing of the normal fluid density implies that the static limit of the shear viscosity
vanishes as well: $\eta = 0$. This means that the Galilean invariant ground state of a superfluid
has zero shear viscosity~\cite{zeroviscosity}. This is similar to the zero d.c. resistivity of a charged superconductor,
as already mentioned at beginning of this Section.
There is, however, an important difference in that the vanishing resistivity persists all the way
up to the transition temperature $T_c$. Even though there are normal fluid excitations in a
superconductor, the infinite conductivity of the condensate ``shorts out'' the normal fluid in
a superconductor. In marked contrast, in a neutral superfluid,
$\eta$ vanishes only at $T=0$. For $0 < T < T_c$, even though a condensate exists,
the normal fluid excitations give rise to a non-zero shear viscosity.
\section{Positivity of spectral functions}
\label{posdefsec}
We simplify notation and write from now on
\begin{eqnarray}
\eta(\omega) \equiv \mathrm{Re}\; \eta(\omega)
\; \; \mathrm{and} \; \;
\zeta(\omega) \equiv \mathrm{Re}\; \zeta(\omega),
\label{simplifyReal}
\end{eqnarray}
unless explicitly stated otherwise. This should cause no confusion since we will not be dealing
with the corresponding imaginary parts.
Before deriving sum rules for
$\eta(\omega)$ and $\zeta(\omega)$ in Section~\ref{sumrulesec},
it is important to discuss here their positivity properties.
Every time we say `positive' we actually mean
`non-negative', a term we find awkward for repeated use.
We will show that
\begin{eqnarray}
\eta(\omega)\ge 0\;\; \mathrm{and}\;\; \zeta(\omega) \ge 0
\;\;\;\; \forall \omega.
\label{Work}
\end{eqnarray}
The simplest approach is to make explicit use of the spectral representation. We will see that this
is sufficient to prove the positivity of $\eta(\omega)$, but \emph{not} that of $\zeta(\omega)$.
To prove the latter, we will calculate the power absorbed by the
fluid from an external velocity perturbation with $\nabla \cdot {\bm u} \neq 0$.
Let us begin with Eqs.~(\ref{KuboTR}) and (\ref{KuboLR}) and
use the spectral representation given by Eq.~(\ref{ImchiAB}) with $\hat{A}_{{\bf q}}=\hat{j}^x_{{\bf q}}$
and $\hat{B}^{\dagger}_{{\bf q}} = \hat{j}^x_{-{\bf q}}$. The transverse and longitudinal components
are obtained, as usual, by taking suitable $q \to 0$ limits~\cite{limits}.
Using $|\langle n|\hat{j}^x_{-{\bf q}}|m\rangle|^2 \geq 0$ and $\omega[1-\exp(-\beta\omega)]\geq 0$ for all $\omega$,
we see that both
$\omega\mathrm{Im}\chi_T({\bf q},\omega)$ and $\omega\mathrm{Im}\chi_L({\bf q},\omega)$ are positive.
Thus we obtain
\begin{eqnarray}
\eta(\omega)\ge 0\;\; \mathrm{and}\;\; \zeta(\omega) +4\eta(\omega)/3\ge 0
\;\;\;\; \forall \omega.
\label{Work2}
\end{eqnarray}
The inequality for $\zeta(\omega)$ is much weaker than what we wish to prove.
One reason to expect that a stronger result should exist for $\zeta(\omega)$ is that
it is known from hydrodynamics (see Sec.~49 of Landau and Lifshitz~\cite{LLFM}) that
the \emph{static} bulk viscosity $\zeta(0)$ must be positive.
To generalize this to all frequencies, we exploit the idea that
the time-averaged power absorbed by the system from an external perturbation
is necessarily positive.
The rate at which the external velocity perturbation given by Eq.~(\ref{Hpert}) does work on the fluid is given
by
\begin{eqnarray}
\frac{d W}{dt} = i\omega m\int d{\bf r} e^{-i\omega t} e^{0^+t}{\bm u}({\bf r})\cdot\langle \hat{\bm{j}}({\bf r},t)\rangle. \label{Power}
\end{eqnarray}
Following Ref.~\cite{ChaikinLubensky}, one finds that the time average of the
power absorbed by the fluid is
\begin{eqnarray} \overline{\frac{dW}{dt}} = \frac{m^2}{2} \sum_{\bf q}
u_{\alpha}(-{\bf q})\left[\omega\mathrm{Im}\chi^{\alpha\beta}_J({\bf q},\omega)\right]u_{\beta}({\bf q})>0.
\label{Power2}\end{eqnarray}
$\overline{dW}/dt > 0$ follows from the fact that energy can only be dissipated
for \emph{any} choice of the external velocity field.
This implies that the real, symmetric matrix $\omega\mathrm{Im}\chi^{\alpha\beta}({\bf q},\omega)$ must be positive definite,
which is equivalent to the positivity of its eigenvalues.
Using Eq.~(\ref{chiTL}), we see that these eigenvalues are precisely
$\omega \mathrm{Im}\chi_L({\bf q},\omega)$ and $\omega \mathrm{Im}\chi_T({\bf q},\omega)$,
so that we simply rederive Eq.~(\ref{Work2}), and do not obtain $\zeta(\omega) \ge 0$.
To constrain $\zeta(\omega)$, without any $\eta(\omega)$ contribution, we must
look at an external velocity field ${\bm u}({\bf r},t) = {\bm u}({\bf r})e^{-i\omega t}$ with
${\bm u}({\bf r}) = a{\bf r}$, where $a = \left(\nabla \cdot {\bm u}\right)/3$ is
spatially uniform.
To analyze the effect of such a perturbation, we first need to rewrite Eq.~(\ref{Power2})
in terms of the stress correlator so that $\partial_\alpha u_\beta$ is directly involved. Second,
${\bm u}({\bf r}) = a{\bf r}$ is not Fourier transformable, so we must work in ${\bf r}$-space, rather than
${\bf q}$-space used elsewhere in the paper.
We use the same derivation that led from the operator Euler equation given by Eq.~(\ref{euler-op}) to Eq.~(\ref{Correlationrelation}),
to get
\begin{eqnarray}
m^2\omega^2\mathrm{Im}\chi^{\alpha\beta}_{J}({\bf q},\omega) = q_{\mu}q_{\nu}\mathrm{Im}\chi^{\alpha\mu,\beta\nu}_{\Pi}({\bf q},\omega).
\label{ImCorrelationrelation}
\end{eqnarray}
Using this in Eq.~(\ref{Power2}) and rewriting the resulting expression in real space, we get
\begin{eqnarray}
\lefteqn{\overline{\frac{dW}{dt}} = \frac{1}{2}\int \!d{\bf r} \!\int \!d{\bf r}' \times}&&\nonumber\\&&\partial_{\alpha}u_{\mu}({\bf r})\left[\frac{\mathrm{Im}\chi^{\alpha\mu,\beta\nu}_{\Pi}({\bf r}-{\bf r}',\omega)}{\omega}\right]\!\partial_{\beta}u_{\nu}({\bf r}'),
\label{Power3}
\end{eqnarray}
which must hold for arbitrary velocity fields ${\bm u}({\bf r})$.
To isolate the contribution of the bulk viscosity, we choose the velocity field ${\bm u} = a{\bf r}$,
for which the shear term (in square brackets)
in the viscous stress tensor, Eq.~(\ref{stresstensor}), vanishes.
Using $\partial_\alpha u_\beta = a \delta_{\alpha\beta}$ in Eq.~(\ref{Power3}) we get
$\mathrm{Im}\chi^{\alpha\alpha,\beta\beta}_{\Pi}({\bf q} \to 0,\omega)/\omega \ge 0$. From the result
given by Eq.~(\ref{KuboPiZeta}) for the bulk viscosity, it immediately follows that $\zeta(\omega)\ge 0$ for all $\omega$.
\section{Sum rules}
\label{sumrulesec}
We now derive sum rules for the shear and bulk viscosities,
$\int^{\infty}_0 d\omega \eta(\omega)$ and $\int^{\infty}_0 d\omega \zeta(\omega)$.
We will first show that
\begin{eqnarray}
\lefteqn{\frac{1}{\pi}\int^{\infty}_0 \!\!d\omega
\lim_{q\to 0}\!\frac{\omega}{q^2}\mathrm{Im}\chi^{xx}_{J}({\bf q},\omega)=}&&\nonumber
\\&&\!\!\!\!\!\!\!\!\lim_{q\to 0}
\frac{\langle[\hat{j}^{x}_{-{\bf q}},[\hat{H},\hat{j}^{x}_{{\bf q}}]]\rangle}{2 q^2}
+\lim_{\omega\to 0}\lim_{q\to 0}\frac{\omega^2}{2q^2}\mathrm{Re}\chi^{xx}_{J}({\bf q},\omega).
\label{noncommutator}
\end{eqnarray}
Then we will simplify the two terms on the right hand side of Eq.~(\ref{noncommutator}):
the first term by explicit evaluation of the commutators, and the second by appealing to
hydrodynamics.
To see what is involved in deriving Eq.~(\ref{noncommutator}), let us first be na\"ive and
ignore the $q \to 0$ limit.
Evaluating the integral on the left hand side by using the
spectral representation in Eq.~(\ref{ImchiAB}) for $\mathrm{Im}\chi^{xx}_{J}$,
we only obtain the first commutator term on the right.
But taking the $q \to 0$ limit after doing the $\omega$-integration
is \emph{not} the same as interchanging the order of these operations!
In order to do it correctly ($q \to 0$ limit before the $\omega$-integration),
we exploit the Kramers-Kronig (K-K) relations to evaluate the
integral in Eq.~(\ref{noncommutator}).
The only subtle point in this approach is that we need to ensure that the
analytic functions which we K-K transform
decay sufficiently rapidly for $\omega \to \infty$.
Using the expression in Eq.~(\ref{FourierchiAB}), it is straightforward to expand
the current correlator in powers of $\omega^{-1}$ for large frequencies.
One finds~\cite{Pitaevskiibook},
\begin{eqnarray}
\lim_{\omega\to \infty}\chi^{xx}_J({\bf q},\omega) &=& \frac{\langle[\hat{j}^{x}_{-{\bf q}},\hat{j}^{x}_{{\bf q}}]\rangle}{\omega} - \frac{\langle[\hat{j}^{x}_{-{\bf q}},[\hat{H},\hat{j}^{x}_{{\bf q}}]]\rangle}{\omega^2} + \ldots.
\nonumber\\
\label{chiJasymptote}
\end{eqnarray}
The $\omega^{-1}$ term vanishes since $\langle[\hat{j}^{x}_{-{\bf q}},\hat{j}^{x}_{{\bf q}}]\rangle = -(2q_{x}/m^2)\sum_{{\bf k}\sigma}n_{{\bf k}\sigma} k_{x} = 0$ in a uniform system.
We further note that this expansion is strictly valid only for a smooth potential~\cite{divergence}, a point
which we will elaborate on in later Sections.
Let us define a function $F(\omega)$ as
\begin{eqnarray} F(\omega) \equiv \lim_{q\to 0}\frac{\omega^2}{q^2}\left[\chi^{xx}_J({\bf q},\omega) + \frac{\langle[\hat{j}^{x}_{-{\bf q}},[\hat{H},\hat{j}^{x}_{{\bf q}}]]\rangle}{\omega^2}\right],
\label{F} \end{eqnarray}
where the $q \to 0$ limit is defined appropriately~\cite{limits} for the longitudinal and
transverse cases.
From Eq.~(\ref{chiJasymptote}), we see that $\lim_{\omega\to\infty}F(\omega)$ vanishes
at least as fast as $\omega^{-1}$ and we can K-K transform it. We thus obtain
\begin{eqnarray}
\lim_{\omega\to 0}\mathrm{Re} F(\omega) &=& \frac{{\cal{P}}}{\pi}\int^{\infty}_{-\infty}d\omega'\frac{\mathrm{Im}F(\omega')}{\omega'}\\
&=&\frac{2}{\pi}\int^{\infty}_{0}d\omega'
\lim_{q\to 0}\frac{\omega'}{q^2}\mathrm{Im}\chi^{xx}_J({\bf q},\omega').\nonumber
\label{noncommutator2}
\end{eqnarray}
where we have used the fact that $\omega \mathrm{Im}\chi^{xx}_J({\bf q},\omega)$ is an even function
of $\omega$.
Using Eq.~(\ref{F}) on the left-hand side of this expression immediately leads to
the result, Eq.~(\ref{noncommutator}), quoted above.
As mentioned earlier, $\lim_{\omega\to 0}\lim_{q\to 0}(\omega^2/2q^2)\mathrm{Re}\chi^{xx}_J$
in Eq.~(\ref{noncommutator}) arises from the noncommutativity of the $\omega\to 0$ and $q\to 0$ limits.
Since this term involves the zero-frequency, long-wavelength limit where hydrodynamics is applicable, we can use hydrodynamic expressions for the current correlation function to evaluate it.
In Appendix~\ref{hydrosec}, we review such expressions and show that for any simple hydrodynamic liquid, one has
\begin{eqnarray}
\lim_{\omega\to 0}\lim_{q\to 0}\frac{m^2\omega^2}{2q^2}\mathrm{Re}\chi_T({\bf q},\omega) &=& 0,\nonumber\\
\label{difference2}\\
\lim_{\omega\to 0}\lim_{q\to 0}\frac{m^2\omega^2}{2q^2}\mathrm{Re}\chi_L({\bf q},\omega) &=& - \frac{\rho c_s^2}{2},\nonumber
\end{eqnarray}
where the adiabatic sound speed is $c_s \equiv (\partial P/\partial\rho)^{1/2}$ at fixed $s=S/N$.
Equation~(\ref{difference2}) is valid for both normal fluids and superfluids
(within two-fluid hydrodynamics).
Combining Eqs.~(\ref{KuboTR}), (\ref{KuboLR}), (\ref{noncommutator}), and (\ref{difference2}), we find the following sum rules:
\begin{eqnarray} \frac{1}{\pi}\int^{\infty}_0 \!\!d\omega
\eta(\omega)=\lim_{q\to 0}\frac{m^2\langle[\hat{j}^{x}_{-{\bf q}},[\hat{H},\hat{j}^{x}_{{\bf q}}]]\rangle_T}{2 q^2},\label{etasumrule}\end{eqnarray}
\begin{eqnarray} \frac{1}{\pi}\!\int^{\infty}_0 \!\!\!d\omega\!\!
\left[\!\zeta(\omega)\!+\!\frac{4\eta(\omega)}{3}\right] \!=\! \lim_{q\to 0}\frac{m^2\langle[\hat{j}^{x}_{-{\bf q}},[\hat{H},\hat{j}^{x}_{{\bf q}}]]\rangle_L}{2 q^2}\! -\! \frac{\rho c_s^2}{2}. \nonumber\\ \label{zetasumrule}\end{eqnarray}
Here, $\langle\cdots\rangle_{T(L)}$ denotes the $q \to 0$ limit appropriate to the transverse (longitudinal) case~\cite{limits}.
The last remaining step in our derivation is
to evaluate the commutators in Eqs.~(\ref{etasumrule}) and (\ref{zetasumrule}).
We consider a system of fermions or bosons described by the Hamiltonian
\begin{eqnarray}
\hat{H}\! &=& \hat{K} + \hat{V}
\\
&=& \! \sum_{{\bf k}\sigma}\varepsilon_{{\bf k}}\hat{c}^{\dagger}_{{\bf k}\sigma}\hat{c}_{{\bf k}\sigma}\! +\!\frac{1}{2}\!\sum_{\substack{{\bf k}\bk'{\bf p}\\ \sigma\sigma'}}\!V(p)\hat{c}^{\dagger}_{{\bf k}+{\bf p}\sigma}\hat{c}^{\dagger}_{{\bf k}'\!-{\bf p}\sigma'}\hat{c}_{{\bf k}'\!\sigma'}\hat{c}_{{\bf k}\sigma}.\nonumber
\label{H}\end{eqnarray}
For a single-component Bose system, $\sigma=\sigma'$ assumes one value;
for fermions, $\sigma = \uparrow,\downarrow$ can take one of two ``spin'' values.
It is straightforward, but tedious, to evaluate the commutator in Eq.~(\ref{noncommutator}) for this Hamiltonian.
One finds, for both fermions and bosons,
\begin{eqnarray}
\lefteqn{\frac{m^2}{2}\langle[\hat{j}^{x}_{-{\bf q}},[\hat{H},\hat{j}^{x}_{{\bf q}}]]\rangle=}&&
\label{Jsumrule}
\\
&&\frac{\langle \hat{K} \rangle}{3}\left(2q^2_{x} + q^2\right) + n\frac{ q^2q^2_{x}}{8m}-
\frac{1}{2} \langle\!\langle 2V(p)p^2_{x}
\nonumber
\\
&-&V\!(|{\bf p}-{\bf q}|)(p_{x}\!-\!q_{x})^2
- V\!(|{\bf p}+{\bf q}|)(p_{x}\!+\!q_{x})^2 \rangle\!\rangle.
\nonumber
\end{eqnarray}
Here, $\langle \hat{K} \rangle = \sum_{{\bf k}\sigma}\varepsilon_{{\bf k}} n_{{\bf k}\sigma}$ is the kinetic energy density,
and we have introduced the shorthand notation
\begin{eqnarray}
\left\langle\!\left\langle {\cal Q} \right\rangle\!\right\rangle \equiv \frac{1}{2}\sum_{\substack{{\bf k}\bk'{\bf p}\\ \sigma\sigma'}}
{\cal Q} \langle \hat{c}^{\dagger}_{{\bf k}+{\bf p}\sigma}\hat{c}^{\dagger}_{{\bf k}'-{\bf p}\sigma'}
\hat{c}_{{\bf k}'\sigma'}\hat{c}_{{\bf k}\sigma}\rangle.
\label{intaverage}
\end{eqnarray}
Related expressions specific to Bose liquids are given in Ref.~\cite{Dalfovo92}. We also note in passing that the longitudinal component of Eq.~(\ref{Jsumrule}) is related by Eq.~(\ref{Nconserv}) to the so-called ``$\langle \omega^3\rangle $ sum rule" discussed for electronic systems~\cite{Puff65}.
The right-hand side of Eq.~(\ref{Jsumrule}) varies as $q^2$ as $q\to 0$,
which cancels the $1/q^2$ in Eqs.~(\ref{etasumrule}) and (\ref{zetasumrule}).
Evaluating the transverse and longitudinal limits of Eq.~(\ref{Jsumrule}), one finds the following viscosity sum rules:
\begin{eqnarray}
\frac{1}{\pi}\int^{\infty}_{0}\!\!d\omega \eta(\omega)= \frac{\varepsilon}{3}-\frac{\langle \hat{V} \rangle}{3} + \frac{2\overline{V}'}{15}
+ \frac{\overline{V}''}{30}
\label{etasumruleiso}
\end{eqnarray}
and
\begin{eqnarray}
\frac{1}{\pi}\int^{\infty}_{0}\!\!d\omega \zeta(\omega) =
\frac{5\varepsilon}{9}+\frac{4\langle \hat{V} \rangle}{9} +
\frac{5\overline{V}'}{9}+
\frac{\overline{V}''}{18}-\frac{\rho c_s^2}{2}.
\label{zetasumruleiso}
\end{eqnarray}
Here,
$\varepsilon=\langle \hat{K} \rangle+\langle \hat{V} \rangle$ is the total energy density,
$\langle \hat{V} \rangle$ the potential energy density, and the terms
$\overline{V}'$ and $\overline{V}''$ are defined using Eq.~(\ref{intaverage}) as
\begin{eqnarray}
\overline{V}' \equiv \left\langle\!\left\langle p\left({\partial V}/{\partial p}\right) \right\rangle\!\right\rangle \ \
{\rm and} \ \
\overline{V}'' \equiv \left\langle\!\left\langle p^2\left({\partial^2V}/{\partial p^2}\right) \right\rangle\!\right\rangle.
\label{Vpp}
\end{eqnarray}
These sum rules are valid at all temperatures (i.e., in the superfluid as well as normal phase) for any Bose or Fermi system with an arbitrary,
spin-independent, isotropic interaction potential $V(p)$. We emphasize that these are exact results obtained without making
any quasiparticle approximations.
In the next Section (Sec.~\ref{diluteviscositysec}), we simplify these sum rules
for the case of a two-component Fermi gas with short range interactions,
which is of relevance to experiments on ultracold atomic Fermi gases with Feshbach scattering resonances.
Before closing this Section, let us briefly discuss viscosity sum rules using the stress correlator representation.
For the shear viscosity spectral function, the sum rule
\begin{eqnarray}
\int^{\infty}_0 d\omega \eta(\omega) &=& \frac{1}{\pi}\int^{\infty}_0 d\omega \lim_{q\to 0}
\frac{\mathrm{Im}\chi^{xy,xy}_{\Pi}({\bf q},\omega)}{\omega} \nonumber\\ &=&
\lim_{\omega\to 0}\lim_{q\to 0}\frac{1}{2}\mathrm{Re}\chi^{xy,xy}_{\Pi}({\bf q},\omega)
\label{etasumrulePi}\end{eqnarray}
follows trivially from the Kramers-Kronig relation.
To show that this is the same as Eq.~(\ref{etasumrule}), we
use Eq.~(\ref{TPi}) in the second line of Eq.~(\ref{etasumrulePi}).
One can rewrite the commutator in Eq.~(\ref{TPi}) using the Fourier transform of
Eq.~(\ref{euler-op}), and set $\omega^2 \mathrm{Re}\chi_T/q^2$ to zero
using the hydrodynamic result, Eq.~(\ref{difference2}), to obtain Eq.~(\ref{etasumrule}).
\section{Dilute two-component Fermi gas}
\label{diluteviscositysec}
We now specialize to the case of a two-component Fermi gas in the dilute limit, where the effective range $r_0$ of the
potential (van der Waals at ``long" distances, with $r_0\sim 100 a_0$) is much smaller than the $s$-wave scattering length $a$ and the mean interparticle spacing $k^{-1}_F$. (In typical experiments, $k^{-1}_F\sim 1\mu m$ and $500 a_0\lesssim |a| \lesssim \infty$.)
In the zero range limit $r_0\to 0$, all physical observables are
universal ($r_0$-independent) functions of the
energy scale $\epsilon_F$ (or length scale $k^{-1}_F$) and the dimensionless parameters
$T/\epsilon_F$ (temperature) and $1/(k_F a)$ (interaction).
We will show that for Fermi gases, the results given by
Eqs.~(\ref{etasumruleiso}) and (\ref{zetasumruleiso}) of the previous Section,
reduce to the simple expressions given by Eqs.~(\ref{etasumrule0finite}) and (\ref{zetasumrule0})
in the Introduction.
Our main task is to calculate the terms $\overline{V}'$ and $\overline{V}''$, involving \textit{gradients}
of the interaction potential, defined in Eq.~(\ref{Vpp}).
We use the real-space approach developed by Zhang and Leggett~\cite{Zhang08}, which is a simple
way to derive results first obtained by Tan~\cite{Tan08,Braaten08}.
Using the two-body density matrix
\begin{eqnarray}
\lefteqn{{\cal F}({\bf r}) = }&&\nonumber\\&&\!\!\!\!\!\int\!\! d^3{\bf R} \Big\langle \hat{\psi}^{\dagger}_{\uparrow}({\bf R}\!+\!\frac{{\bf r}}{2})
\hat{\psi}^{\dagger}_{\downarrow}({\bf R}\!-\!\frac{{\bf r}}{2})
\hat{\psi}_{\downarrow}({\bf R}\!-\!\frac{{\bf r}}{2})\hat{\psi}_{\uparrow}({\bf R}\!+\!\frac{{\bf r}}{2})\Big\rangle
\nonumber
\\
\label{rho2def}
\end{eqnarray}
we rewrite $\overline{V}'$ and $\overline{V}''$ in real space as
\begin{eqnarray}
\overline{V}' = \int d^3{\bf r} \; r \frac{\partial V(r)}{\partial r} {\cal F}({\bf r})
\label{moment1}
\end{eqnarray}
and
\begin{eqnarray}
\overline{V}''
= \int d^3{\bf r}\; r^2 \frac{\partial^2 V(r)}{\partial r^2} {\cal F}({\bf r}).
\label{moment2}
\end{eqnarray}
Since $V(r)$ is short-ranged, these expressions are only sensitive to the short-distance ($r_0\lesssim r \ll k^{-1}_F$) structure of the two-body density matrix. (The non-universal contribution from distances smaller than $r_0$ is assumed to be small.) For a two-component dilute Fermi gas, at these short distances, the two-body density matrix is~\cite{Zhang08}
\begin{eqnarray}
{\cal F}({\bf r}) = \frac{C}{16\pi^2}\left(\frac{1}{r}-\frac{1}{a}\right)^2.
\label{rho2}
\end{eqnarray}
Here, $C$ is the contact~\cite{Tan08,Zhang08,Braaten08} mentioned in the Introduction.
In Appendix \ref{ContactAppendix} we remind the reader how the contact $C$ governs both the short-distance
behavior of the two-body density matrix in Eq.~(\ref{rho2}), and the
large-$k$ tail of the momentum distribution function
$\lim_{k\to \infty} n_{{\bf k}\sigma} = {C}/{k^4}$.
Using integration by parts, we transform gradients of the potential $V(r)$
in Eqs.~(\ref{moment1}) and (\ref{moment2}) into gradients of the
two-body density matrix, Eq.~(\ref{rho2}). We thus find
\begin{eqnarray}
\overline{V}' = \frac{C}{4\pi}\int dr V(r)\left(-1 + {4r}/{a}\right)\label{moment1b}
\end{eqnarray}
and
\begin{eqnarray} \overline{V}'' =\frac{C}{2\pi}\int dr V(r)\left(1 - {6r}/{a}\right).\label{moment2b}
\end{eqnarray}
All that remains is to evaluate the two integrals
$X_n = C\int dr V(r) (r/a)^n/4\pi$ with $n=0,1$ in the limit
where the range of the potential $r_0 \to 0$. The Tan relations
are precisely what we need to evaluate such (possibly divergent) integrals.
The details of this analysis are described in Appendix \ref{ContactAppendix}.
We use the potential energy density~\cite{Tan08,Braaten08}
\begin{eqnarray}
\langle \hat{V} \rangle = - \frac{C \Lambda}{2\pi^2 m} + \frac{ C}{4\pi m a},
\label{eint}
\end{eqnarray}
where $\Lambda \equiv 1/r_0$ is the ultraviolet cutoff ,
and the pressure
\begin{eqnarray}
P = 2\varepsilon/3 + {C}/(12\pi m a)
\label{tanP}
\end{eqnarray}
to determine $X_0$ and $X_1$.
In deriving these results, we also use an expression for the pressure
$P$ in terms of $\varepsilon$, $\langle \hat{V} \rangle$ and $\overline{V}'$
which is derived in Appendix~\ref{Pressuresec} using the Feynman-Hellmann theorem.
Our final results for $\overline{V}'$ and $\overline{V}''$,
derived in Appendix \ref{ContactAppendix}, are
\begin{eqnarray}
\overline{V}' =
-\langle \hat{V} \rangle -2\varepsilon + 3P = C\Lambda/{2\pi^2 m}
\label{identity1}
\end{eqnarray}
and
\begin{eqnarray} \overline{V}''
= 2\langle \hat{V}\rangle + 8\varepsilon - 12P = - \frac{C \Lambda}{\pi^2 m} - \frac{ C}{2\pi ma}
\label{identity2}
\end{eqnarray}
Using these results in the general sum rules given by Eqs.~(\ref{etasumrule}) and (\ref{zetasumrule}), we
obtain the $\eta$ and $\zeta$ sum rules for the two-component dilute
Fermi gas which are valid for \emph{all} values of $1/(k_F a)$ throughout the BCS-BEC crossover,
so long as $a,k^{-1}_F\gg r_0$, and at all temperatures, both in the superfluid and normal phases, so long as $T\ll 1/mr^2_0$.
For the shear viscosity, we find
\begin{eqnarray}
\int^{\Lambda^2/m}_{0}d\omega\; \eta(\omega)/{\pi} &=& {\varepsilon}/{3} - {2\langle \hat{V} \rangle}/{5}
\nonumber
\\
&=& \frac{\varepsilon}{3} - \frac{C}{10 \pi m a} + \frac{C \Lambda}{5\pi^2 m},
\label{etasumrule2}
\end{eqnarray}
where we have imposed the energy cutoff $\Lambda^2/m = 1/mr_0^2$~\cite{uv-cutoff}.
In the zero-range limit as $\Lambda = 1/r_0 \to \infty$, the right hand side diverges.
(Strictly speaking, every physical potential has a small non-zero effective range $r_0$,
which leads to a well-defined, finite results, but one that is ``non-universal'' in that it depends
on short distance physics.) We will see in the following Section, Sec.~\ref{highfrequencytailsec},
how to make sense of this divergence and find a modified sum rule that remains finite as $r_0\to 0$.
For the bulk viscosity we find
\begin{eqnarray}
\int^{\infty}_{0}d\omega\; \zeta(\omega)/{\pi} &=& P - \varepsilon/9 - {\rho c_s^2}/{2}
\\
&=& \frac{5\varepsilon}{9} + \frac{C}{12 \pi m a} - \frac{\rho c_s^2}{2}.
\label{zetasumrule2}
\end{eqnarray}
Below the superfluid transition, the bulk viscosity $\zeta$ that enters Eq.~(\ref{zetasumrule0}) is the bulk viscosity $\zeta_2$,
as explained earlier.
We can rewrite the right hand side of this sum rule in a useful way using
simple facts about the scaling form of thermodynamic functions across the entire BCS-BEC crossover,
as described in detail in Appendix~\ref{universalthermosec}. The final result is
\begin{eqnarray}
\int^{\infty}_{0}d\omega\; \zeta(\omega)/{\pi} =
\frac{1}{72 \pi m a^2} \left( \frac{\partial C}{\partial a^{-1}} \right)_s,
\label{zetasumrule2A}
\end{eqnarray}
where the derivative is taken at constant entropy per particle $s = S/N$.
The positivity of the sum rule, given that of its integrand, implies that the contact is a
monotonically increasing function of $1/(k_Fa)$ through the BCS-BEC crossover.
We will discuss this further in Section~\ref{Crossoversec}.
\section{High-frequency tails}
\label{highfrequencytailsec}
In this Section we derive a modified shear viscosity sum rule that is manifestly finite
in the $\Lambda = 1/r_0 \to \infty$ limit. This is obtained by relating the
linear (in $\Lambda$) divergence in the sum rule, Eq.~(\ref{etasumrule2}), to a high-frequency tail
in $\eta(\omega)\sim 1/\sqrt{\omega}$, and then ``subtracting out'' the contribution of this tail.
We use ``high frequency'' or $\omega \to \infty$, to mean
$\epsilon_F \ll \omega \lesssim 1/mr_0^2$.
We also argue that a high-frequency tail of the form
$\omega^{-n/2}$, with odd integer $n$, in a variety of spectral functions
is a generic feature of short-range physics. As discussed below, it shows up in many contexts,
even outside dilute quantum gases.
We can rewrite the $\eta$ sum rule in Eq.~(\ref{etasumrule2}) as
\begin{eqnarray}
\frac{1}{\pi}\int^{\infty}_{0}\!\!d\omega \left[\eta(\omega) -\frac{C\Theta(\omega - \Omega_0)}{10\pi\sqrt{m\omega}}\right]
\nonumber \\
= \frac{\varepsilon}{3} - \frac{ C}{10\pi m a} + \frac{C}{5\pi^2}\sqrt{\frac{\Omega_0}{m}},
\label{etasumrule3}
\end{eqnarray}
where $\Omega_0$ is an arbitrary energy scale. If we choose $\Omega_0$ to be $\Lambda^2/m$
we recover Eq.~(\ref{etasumrule2}). But for any finite $\Omega_0$,
subtracting out the $\omega^{-1/2}$ tail makes the integral ultraviolet
convergent and we can take the cutoff $\Lambda$ to infinity.
If we choose $\Omega_0 = 0$, we obtain the
finite sum rule
\begin{eqnarray}
\frac{1}{\pi}\int^{\infty}_{0}\!\!d\omega \left[\eta(\omega) -\frac{C}{10\pi\sqrt{m\omega}}\right]
= \frac{\varepsilon}{3} - \frac{ C}{10\pi m a}.
\label{etasumrule3A}
\end{eqnarray}
The price we pay for using this finite, $r_0$-independent result
(in the $r_0 \to 0$ limit) is that we sacrifice the positivity
of the integrand. At sufficiently small $\omega$, we must necessarily have
$\eta(\omega) < C/(10\pi\sqrt{m\omega})$ since $\eta(0)$ is finite.
One can, in principle, exploit the freedom in Eq.~(\ref{etasumrule3}) and
choose $\Omega_0$ to be large enough so
that the integrand is always positive, however.
The finiteness of the right hand side of Eq.~(\ref{etasumrule3A})
implies that the integrand on the left must vanish at least as fast as $\omega^{-3/2}$ for the integral
to converge at large $\omega$.
Thus the asymptotic behavior of the spectral function
$\eta(\omega)$ is of the form
\begin{eqnarray}
\eta(\omega \to \infty) \simeq \frac{C}{10\pi\sqrt{m\omega}}.
\label{etahighomega}
\end{eqnarray}
We note that a high-frequency tail in the imaginary part of a
retarded correlation function which goes like
$\omega^{-n/2}$, with positive integer $n$, is a general feature of short-range two-body physics.
Suppose that for some operator $\hat{A}$, the corresponding $n$-th moment sum rule has the form
\begin{eqnarray}
\frac{1}{\pi}\int^{\infty}_0 d\omega \omega^n \mathrm{Im}\chi_{A,A}(\omega) = \alpha \langle V\rangle + \cdots,\label{sumrulediv}
\end{eqnarray}
where we only show the divergent term explicitly;
the ellipses denote regular terms. $\alpha$ is some combination of parameters and is not, in general, dimensionless.
In addition to the current correlation function ($n=1$), diverging sum rules of the form given by Eq.~(\ref{sumrulediv}) arise for the radio frequency (RF) spectral function ($n=1$)~\cite{RFsumrule}, and, as we show below, the density response function ($n=3$). Using the same reasoning as above, a divergence of the form given by Eq.~(\ref{sumrulediv})
implies a high-frequency tail.
For a dilute two-component Fermi gas with $a\gg r_0$,
the high-frequency tail is given by
\begin{eqnarray}
\mathrm{Im}\chi_{A,A}(\omega \to \infty) \simeq \frac{\alpha C}{4\pi m^{1/2}}\frac{1}{\omega^{n+1/2}}.
\end{eqnarray}
As seen from the above arguments, an $\omega^{-3/2}$ tail arises in the
radio-frequency spectroscopy response function $I(\omega)$~\cite{Schneider09,Strinati09} for Fermi gases.
Another interesting example is the $\omega^{-7/2}$ tail in the density response of a dilute Fermi gas which we derive
in Section~\ref{SqomegaSec}.
There, we also point out that an identical asymptotic behavior is found for the dense Bose liquid $^4$He, which further
emphasizes the generality of the short-distance physics in all quantum fluids.
\section{Sum rules through the BCS-BEC crossover}
\label{Crossoversec}
In this Section we consider the bulk and shear viscosity sum rules through the BCS-BEC crossover, going from the weakly attractive BCS limit
($a$ small and negative) with large Cooper pairs to the BEC limit ($a$ small and positive) with weakly interacting, tightly bound molecules.
The crossover can be traversed by changing $x = 1/(k_F a)$ from $x = -\infty$ (BCS limit) to $x = +\infty$ (BEC limit).
In experiments, the scattering length $a$ is varied by tuning a magnetic field about a Feshbach resonance. Precisely at resonance, $x=0$, the
scattering length diverges and the Fermi gas is in a very strongly interacting ``unitary regime" where the pair size is of the order of the interparticle
spacing.
To actually compute the viscosity sum rules given by Eqs.~(\ref{zetasumrule2A}) and (\ref{etasumrule3A})
for arbitrary coupling $x = 1/(k_F a)$ and temperature $T$,
we need to know the energy density $\varepsilon = n \epsilon_F {\cal{E}}(x,T/\epsilon_F)$, from which we can determine the contact $C$ as described below.
In general, we will need to use quantum Monte Carlo (QMC) data for the energy density to evaluate the sum rules.
However, as shown below, we are able to analytically constrain the bulk viscosity spectral function at unitarity.
\begin{figure}
\begin{center}
\epsfig{file=bulksumrule.eps, angle=0,width=0.47\textwidth}
\caption{(Color
online) The value of the bulk viscosity sum rule given by the right-hand side of Eq.~(\ref{zetasumrule2A}) at $T=0$ in units of $n\epsilon_F$ through the BCS-BEC crossover. }
\label{bulksumrulefig}
\end{center}
\end{figure}
We see from Eq.~(\ref{zetasumrule2A}) that the bulk viscosity sum rule vanishes at unitarity:
\begin{eqnarray}
\frac{1}{\pi}\int^{\infty}_{0}d\omega\; \zeta(\omega)= 0 \ \ \ \ (|a| = \infty).
\label{ZetaSumUnitarity}
\end{eqnarray}
We are using here the fact that $(\partial C / \partial a^{-1})_s$ is finite (i.e., non-infinite) at
$x = 1/(k_F a) = 0$ at all temperatures. One can also see this using
elementary arguments that do not involve the contact. From ``universal thermodynamics"~\cite{Ho04}, the only energy scales at unitarity are $\epsilon_F$ and the temperature, and
we can directly show that $P - \varepsilon/9- \rho c_s^2/2=0$ (see Appendix~\ref{universalthermosec}).
The vanishing sum rule, Eq.~(\ref{ZetaSumUnitarity}), together with the positivity condition
$\zeta(\omega) \geq 0$ derived in Section~\ref{posdefsec}, implies
\begin{eqnarray}
\zeta(\omega) = 0 \ \ \ \forall \omega \ \ \ (|a| = \infty).
\label{ZetaUnitarity}
\end{eqnarray}
That the \emph{static} bulk viscosity $\zeta(0)$ vanishes is a well-known
consequence~\cite{Son07} of scale or conformal invariance at unitarity~\cite{Castin04}. Our result generalizes this to arbitrary frequencies.
As discussed below in Section~\ref{SqomegaSec}, our result actually has important implications for measuring the
frequency dependent shear viscosity of a unitary Fermi gas using a density probe such as two-photon Bragg scattering.
Another general consequence of $\zeta(\omega) \geq 0$ is that its sum rule must be positive for
all $x = 1/(k_F a)$ and $T$. Equation~(\ref{zetasumrule2A}) then implies that
\begin{eqnarray}
\left(\frac{\partial C}{\partial a^{-1}}\right)_{{s}}\geq 0\;\;\;\;\forall a,
\end{eqnarray}
so that the contact must be a monotonically increasing function of $1/(k_Fa)$ through the BCS-BEC crossover
at fixed entropy per particle. We can understand this inequality intuitively as follows: the contact $C$, which
is related to the probability of finding two particles of opposite spin close to each other, can only increase with
increasing attraction $a^{-1}$.
In Fig.~\ref{bulksumrulefig} we show the bulk viscosity sum rule in Eq.~(\ref{zetasumrule2A}) at $T=0$
calculated using QMC data~\cite{Astrakharchik04} for the energy density $\varepsilon$.
The contact $C$ is obtained from $\varepsilon$ using Tan's ``adiabatic relation"~\cite{Tan08}
\begin{eqnarray}
\left(\partial \varepsilon / \partial a^{-1} \right)_{s} = - C/(4\pi m),
\label{tanA}
\end{eqnarray}
where the derivative is taken at fixed entropy per particle $s\equiv S/N$.
We fitted the QMC data and took numerical derivatives with respect to $a^{-1}$.
Since the $\zeta$ sum rule involves the second derivative of QMC data
for the energy density, the results may not be very accurate far from unitarity in either direction.
\begin{figure}
\begin{center}
\epsfig{file=shearsumrule.eps, angle=0,width=0.47\textwidth}
\caption{(Color online) The value of the finite shear viscosity sum rule, with the contribution from the high-frequency tail in Eq.~(\ref{etahighomega}) subtracted out, given by the right-hand side of Eq.~(\ref{etasumrule3A}) at $T=0$ in units of $n\epsilon_F$ through the BCS-BEC crossover.}
\label{shearsumrulefig}
\end{center}
\end{figure}
Both the vanishing of the $\zeta$ sum rule at $x= 1/(k_F a) =0$ and its positivity away from
unitarity are apparent in Fig.~\ref{bulksumrulefig}. This is due to the
$1/a^2$ dependence of the sum rule in the vicinity of unitarity.
We emphasize the nontriviality of the result given by Eq.~(\ref{zetasumrule2A}) in the unitarity region.
In the form first derived in Eq.~(\ref{zetasumrule2}), the right hand side is
$(P - \varepsilon/9- \rho c_s^2/2)$. Each term in this expression has both constant and order $x$
contributions, which must all cancel to give a final result which goes like $x^2$ at small $x$.
In the BCS limit, the $\zeta$ sum rule vanishes as $2n\epsilon_F/(27\pi|x|$) since $C\to 4\pi^2n^2a^2_s$~\cite{Tan08}. In the BEC limit, the energy density is dominated by the negative molecular binding energy, $\varepsilon\approx nE_b/2$, with $E_b = -1/(ma^2)$. Thus, $C\to 4\pi n/a$ and the sum rule grows as $n|E_b|/18$.
Next, in Fig.~\ref{shearsumrulefig}, we plot the shear viscosity sum rule given by Eq.~(\ref{etasumrule3A}) at $T = 0$
again using the QMC data of Ref.~\cite{Astrakharchik04}.
Because of the $1/\sqrt{\omega}$ subtraction extending all the way down to $\omega = 0$, the $\eta$ sum rule
in Eq.~(\ref{etasumrule3A}) is \emph{not} constrained to be positive. Using the above analytic result for the contact in the BCS limit, one finds that the $\eta$ sum rule asymptotes to $0.2 n\epsilon_F$
in the BCS limit.
At unitarity, $|a| = \infty$ and the $\eta$ sum rule is $\varepsilon/3 \simeq 0.4 \times (3 n\epsilon_F/5) \times (1/3) = 0.08 n\epsilon_F$.
On the BEC side of the resonance the sum rule changes sign, tending to $(17/30)nE_b$ in the BEC limit.
\section{Dynamic Structure Factor}
\label{SqomegaSec}
We now discuss the connection between viscosity and the density-density correlator or dynamic structure factor.
This analysis leads to two interesting results for the two-component Fermi gas. First, we predict that
a density probe such as two-photon Bragg spectroscopy~\cite{Iacuppo} can in principle be used to measure the
frequency dependent $\eta(\omega)$ at unitarity:
\begin{eqnarray}
\eta(\omega)
= \lim_{q\to 0}\frac{3\omega^3}{4 q^4}\;\mathrm{Im}\chi_{\rho\rho}({\bf q},\omega) \ \ \ \ \ (|a| = \infty).
\label{bragg}
\end{eqnarray}
Second, we derive the high-frequency tail~\cite{Son10}
\begin{eqnarray}
\lim_{\omega\to\infty}\lim_{q\to 0}S({\bf q},\omega) = \frac{2q^4C}{15 \pi^2 m^{1/2}}\frac{1}{\omega^{7/2}},
\label{sqomegatail}
\end{eqnarray}
a result that is valid for all $1/(k_F a)$ and all temperatures. As discussed below, such non-analytic
tails are also known in other strongly interacting quantum fluids like $^4$He.
We start with the operator form of the continuity equation
\begin{eqnarray}
i [\hat{\rho},\hat{H}] = m \partial_{\alpha}\hat{j}_{\alpha},
\label{continuity-op}
\end{eqnarray}
where $\hat{\rho} = m\hat{n}$ is the mass density operator, and take its matrix elements
between exact many-body eigenstates. Using the spectral representation, Eq.~(\ref{FourierchiAB}),
we relate the density correlator $\chi_{\rho\rho}$ to the the \emph{longitudinal} current correlator [see Eq.~(\ref{Nconserv})].
The latter is related to the viscosity as shown in Eq.~(\ref{KuboLR}),
namely $\zeta(\omega) + 4 \eta(\omega)/3 = \lim_{q\to 0} {m^2 \omega}\mathrm{Im}\chi_L({\bf q},\omega)/{q^2}$.
We thus obtain
\begin{eqnarray}
\zeta(\omega) + {4\eta(\omega)}/{3} = \lim_{q\to 0}{\omega^3}\mathrm{Im}\chi_{\rho\rho}({\bf q},\omega)/{q^4}.
\label{KuboLR2}
\end{eqnarray}
We discuss two situations where the contribution of $\zeta(\omega)$ vanishes and we can obtain
interesting results connecting $\eta(\omega)$ and density correlations.
First, we focus on the unitary Fermi gas where $\zeta(\omega)$ vanishes at all $\omega$ (as shown in
Section \ref{Crossoversec}) and Eq.~(\ref{KuboLR2}) simplifies to Eq.~(\ref{bragg}).
Thus, the frequency-dependent shear viscosity $\eta(\omega)$ in a unitary Fermi gas can in principle be measured
using an experiment like Bragg scattering, which directly probes $\mathrm{Im}\chi_{\rho\rho}$.
Second, let us look at the high-frequency regime $\epsilon_F \ll \omega \lesssim 1/(mr_0^2)$.
The $\zeta$ sum rule in Eq.~(\ref{zetasumrule2}) is convergent in the $r_0\to 0$ limit,
and thus $\zeta(\omega)$ must decay faster than $1/\omega$, while $\eta(\omega) \sim 1/\sqrt{\omega}$
(see Section \ref{highfrequencytailsec}). Thus, as $\omega \to \infty$, the bulk viscosity $\zeta(\omega)$
is much smaller than the shear viscosity $\eta(\omega)$ for all $1/(k_F a)$ and all $T/\epsilon_F$.
Using Eq.~(\ref{KuboLR2}) we thus find
\begin{eqnarray}
\eta(\omega \to \infty) &\simeq& \lim_{\omega\to\infty}\lim_{q\to 0}\frac{3\omega^3}{4q^4}\mathrm{Im}\chi_{\rho\rho}({\bf q},\omega)
\nonumber\\ &=&
\lim_{\omega\to\infty}\lim_{q\to 0}\frac{3\pi\omega^3}{4q^4}S({\bf q},\omega),
\label{etahighomegaSqw}
\end{eqnarray}
The dynamic structure factor $S({\bf q},\omega)$ is related to $\mathrm{Im}\chi_{\rho\rho}$ via the
fluctuation-dissipation theorem
\begin{eqnarray}
S({\bf q},\omega) = \frac{\mathrm{Im}\chi_{\rho\rho}({\bf q},\omega)}{\pi[1-\exp(-\beta\omega)]}.
\end{eqnarray}
Our final result, given by Eq.~(\ref{sqomegatail}), for the high-frequency tail of $S({\bf q},\omega)$ is obtained
by using the high-frequency tail, Eq.~(\ref{etahighomega}), of $\eta(\omega)$ in Eq.~(\ref{etahighomegaSqw}).
The high-frequency $\omega^{-7/2}$ tail of the dynamic structure factor result is a universal feature of short-range two-body interactions.
Remarkably, such a tail was first noticed in deep inelastic neutron scattering studies of superfluid $^4$He~\cite{Wong77}
and was subsequently understood in terms of hard-sphere gases~\cite{Kirkpatrick84}.
The high-frequency neutron scattering experiments probe the
short distance properties of the two-body pair distribution function.
[In dilute Fermi gases, this is directly related to the contact $C$; see Eq.~(\ref{rho2})].
It may seem surprising that such anomalous high-frequency tails arise even in dense systems like $^4$He.
Recall that this behavior should be visible in a frequency range $n^{2/3}/m < \omega < 1/mr^2_0$
in which the interaction ``looks" short-range. Even in $^4$He, where $nr^3_0\lesssim 1$, such a frequency range can be found using deep inelastic neutron scattering, although the range is obviously much smaller than in dilute gases with $nr^3_0\ll 1$.
\section{Comparison with sum rules for Relativistic Field Theories}
\label{comparisonsec}
There has been a considerable effort in the high-energy literature to understand the
properties of viscosity spectral functions and their sum rules; see, e.g., Refs.~\cite{Teaney06,Romatschke09,Moore08}).
In addition to understanding the transport coefficients within the AdS/CFT framework,
this work seems to be motivated in part by an interest in reliably extracting transport coefficients
of the quark-gluon plasma from lattice QCD calculations of Euclidean correlation functions.
We briefly discuss here some similarities and differences between the results for
relativistic quantum field theories and those derived in this paper for
non-relativistic Fermi gases: Eqs.~(\ref{zetasumrule2A}) and (\ref{etasumrule3A}).
There exist a number of Boltzmann calculations of the viscosity spectral functions in
weak coupling QCD~\cite{Teaney06,Schafer09}.
For the shear viscosity, the authors of Ref.~\cite{Schafer09} find the shear viscosity sum rule
\begin{eqnarray}
\frac{1}{\pi}\int^{\omega_c}_0d\omega \eta(\omega) = \frac{\varepsilon+P}{5},
\label{QCDshearsumrule}
\end{eqnarray}
where $g^4T\ll \omega_c \ll g^2T$ is a cutoff that removes a diverging contribution from a high-frequency tail.
For the ${\cal{N}}=4$ supersymmetric Yang-Mills theory (SYM), Romatschke and Son~\cite{Romatschke09}
derived the following shear viscosity sum rule:
\begin{eqnarray}
\frac{1}{\pi}\int^{\infty}_0 d\omega [\eta(\omega)-\eta_{T=0}(\omega)] = \frac{\varepsilon}{5}.
\label{SUSY}
\end{eqnarray}
Here, a diverging vacuum contribution from a $T$-independent high-frequency tail has been subtracted out.
We note that our $\eta$ sum rule in Eq.~(\ref{etasumrule3A}), though similar in structure, has one key difference.
The high-frequency tail for the Fermi gas is in general $T$-\emph{dependent}, because its coefficient is
set by the contact $C = k_F^4 {\cal C}[1/(k_F a), T/\epsilon_F]$.
A non-perturbative calculation of the bulk viscosity sum rule in ${\cal{N}}=4$ supersymmteric Yang-Mills theory and pure Yang-Mills theory (QCD with no quarks) has been given recently by Romatschke and Son~\cite{Romatschke09}:
\begin{eqnarray}
\lefteqn{\frac{1}{\pi}\int^{\infty}_0 d\omega [\zeta(\omega)-\zeta_{T=0}(\omega)] =}&&\nonumber\\&&(3\varepsilon+P)(1-3c^2)- 4(\varepsilon-3P),
\label{QCDbulksumrule}
\end{eqnarray}
where $c\equiv \sqrt{\partial P/\partial \varepsilon}$ is the sound speed in relativistic hydrodynamics (with the speed of light equal to unity)~\cite{LLFM}. There are some differences and one very
interesting similarity with our $\zeta$ sum rule in Eq.~(\ref{zetasumrule2A}). In contrast to the Fermi gas spectral function $\zeta(\omega)$, there is a need
to subtract out a divergent tail in Eq.~\ref{QCDbulksumrule} and this tail appears to be $T$-independent.
The interesting similarity is that in the ``conformal limit" $P = \varepsilon/3$, the right hand
side of Eq.~(\ref{QCDbulksumrule}) vanishes, analogous to the unitary Fermi gas.
\section{Conclusions}
\label{conclusionssec}
In this paper we have derived various exact, non-perturbative results for the shear and bulk viscosities
of non-relativistic quantum fluids, focusing on the strongly interacting Fermi gas. Our main results were already
summarized in the Introduction. To conclude, we discuss some open questions and how our results relate to them.
Most calculations~\cite{Bruun05,Schafer07} of the viscosity in strongly interacting Fermi gases
have so far been restricted to solving Boltzmann equations or using diagrammatic perturbation theory, in essence making a
quasiparticle approximation. Such results are valid in the high and low temperature regimes, but
not in the most interesting regime near and above $T_c$ where a quasiparticle approximation is questionable
and the shear viscosity is known to be the smallest. It was recognized some time back~\cite{Randeria-reviews}
that there is a breakdown of Fermi liquid theory in the normal (i.e., non-superfluid) state of the
strongly interacting regime of the BCS-BEC crossover. It was shown that precursor pairing correlations
lead to a pseudogap \cite{Randeria-reviews}, which is a strong suppression of low-energy spectral weight
in various response functions. It is likely that no sharp quasiparticle excitations exist in this regime
near unitarity and just above $T_c$, but controlled calculations of dynamic quantities are very difficult.
Quantum Monte Carlo methods have played an important role in determining the equilibrium thermodynamic properties
of the unitary Fermi gas. However, results for transport coefficients are much less common, since they require
analytic continuation of imaginary time (Euclidean) data to the real axis~\cite{Aarts02}.
The sum rules we derive could serve as useful constraints on similar calculations for
strongly interacting Fermi gases.
From an experimental point of view, the (static) shear viscosity for strongly interacting Fermi gases
has been estimated from studies of the damping of collective oscillations~\cite{Turlapov08}. We have shown
above that, at unitarity, the full frequency dependence of the shear viscosity spectral function
$\eta(\omega)$ can be obtained from two-photon Bragg spectroscopy. While it would be a challenging experiment (the density response being very small for small-$q$), this would give extremely important insights
into the strongly interacting Fermi gas, analogous to optical conductivity measurements of solids.
Finally, we return to the conjectured bound~\cite{Son-review} on the shear viscosity, Eq.~(\ref{bound}).
Proving or disproving the existence of a bound~\cite{bound} for non-relativistic quantum fluids like the strongly interacting Fermi gas
remains a challenging open problem. We hope that the spectral functions and sum rules derived here
constitute a step in this direction, just as they have for other well known inequalities
in quantum many-body physics.
\begin{acknowledgments}
ET would like to thank Shizhong Zhang, Georg Bruun, Joaqu\'in Drut, Vijay Shenoy, and Jason Ho for stimulating discussions.
MR would like to thank the participants at the International Conference on Recent Progress in Many Body Theories
(RPMBT15) last summer for spurring his interest in this problem. We thank Eric Braaten, Dam Son, and Sandip Trivedi for comments on the manuscript and Stefano Giorgini for sharing with us the Monte-Carlo data of Ref.~\cite{Astrakharchik04}.
We gratefully acknowledge support from NSF-DMR 0706203, ARO W911NF-08-1-0338, and NSF-DMR 0907366.
\end{acknowledgments}
|
\section{Introduction}
For many years, the dominant method for discovering massive clusters
of galaxies at $z>1$, has been via X-ray emission from the hot gas in
their dark matter potential wells. Over the years, the \emph{ROSAT}
and \emph{XMM-Newton} Telescopes, in particular, have yielded a
handful of well-studied examples e.g., RDCS J0910+5422 \citep{shr02},
RDCS J1252.9-2927 \citep{rte04}, RX J0848+4452 (Lynx E;
\citealt{rse99}), XMMU J2235-25 \citep{mrl05}, and XMMXCS J2215.9-1738
\citep{srs06}. In recent years, it has been realised that, by
incorporating deep infrared (IR) observations, existing optical
imaging techniques can be adapted to successfully detect clusters at
redshifts competitive to the existing X-ray surveys. However, because
IR observations covering only modest areas (120 arcmin$^2$ - 7.25
deg$^2$) have been available, the (cluster or candidate) systems
detected to date have been less massive than those discovered from the
X-ray surveys (\citealt{seb05, bba06, vB07, mcC07, zat07, and09,
ebg08, goto08, kri08, kurk08, mwl08}; although see \citealt{see97}).
Massive clusters of galaxies are rare, and one requires as widefield a
survey as possible to detect them. The largest area \emph{Spitzer}
Space Telescope Survey is the 50 square degree seven passband (3.6,
4.5, 5.8, 8.0, 24, 70, 160 $\micron$) Spitzer Wide-Area Infrared
Extragalactic Survey (SWIRE) Legacy Survey \citep{lon03, sur05,
shu08}. We have obtained deep $\ensuremath{z^{\prime}}-$band imaging \citep{mwy09,
wmy09}, and combined this with the pre-existing InfraRed Array
Camera (IRAC; \citealt{faz04}) observations from the SWIRE survey,
aiming to select clusters at $z > 1$ using a two filter
$\ensuremath{z^{\prime}}-[3.6]$ infrared adaptation of the well-proven optical cluster
red sequence (RS) method \citep{gy00,gy05,gil07}.
The SpARCS\footnote{http://www.faculty.ucr.edu/$\sim$gillianw/SpARCS}
survey \citep{wmy09,mwy09}, is now complete and has an effective area
(defined as the usable overlap with SWIRE after excluding chip gaps,
regions near bright stars, etc), of 41.9 deg$^{2}$. The SpARCS catalog
contains several hundred cluster candidates at $z > 1$, by far the
largest, homogeneously selected sample of its kind. In
\citeauthor{mwy09} and \citeauthor{wmy09} we presented an overview of
the SpARCS survey, and reported spectroscopic confirmation of three
clusters at $z=1.18, 1.20$ and $1.34$. In this paper, we report
spectroscopic confirmation of three additional clusters at $z=0.87,
1.16$ and $1.21$.
The paper is organized as follows: in \S\ref{observations}, we
describe the imaging and spectroscopic observations, and the data
reduction; in \S\ref{analysis}, we describe the spectroscopic catalog;
in \S\ref{results}, we present our results for the three individual
clusters, and estimate the cluster velocity dispersions and dynamical
masses; and in \S\ref{discussion}, we present a discussion and the
main conclusions based on our results. We assume a $\Lambda$-CDM
cosmology with $\Omega_M=0.3$ and $\Omega_{\Lambda}=0.7$, and H$_0=70$
km s$^{-1}$ Mpc$^{-1}$.
\section{Observations}\label{observations}
\subsection{Photometry}\label{photometry}
The SpARCS/SWIRE survey is comprised of six fields \citep*[see table 1
of][]{wmy09}. The three clusters described in this paper were
identified as high significance cluster candidates in the 7.9 deg$^2$
ELAIS-N1 field. A full description of the SpARCS data reduction,
cluster candidate detection algorithm and catalogs will appear in
Muzzin et al.\ 2010, in preparation. We provide only a brief overview
of the main details here. SpARCS $z^{\prime}$ observations of the
Northern fields were observed by CFHT/MegaCam for a total integration
time of 6000s per pointing. Photometry was performed on both the
$\ensuremath{z^{\prime}}$ and IRAC mosaics using the Sextractor photometry package
\citep{ba96}. The total 3.6 $\micron$ magnitude of all sources in the
field was computed using a large aperture, equal to the geometric mean
radius of the Sextractor isophotal aperture. Further details of the
photometric pipeline may be found in \citet{lwm05} and
\citet{mwl08}. The depth of the $z^{\prime}$ data varies from pointing
to pointing depending on the seeing and the sky background, however,
the mean 5 $\sigma$ depth for extended sources in the ELAIS-N1 field
was $\sim23.7$ Vega (24.2 AB).
The $\ensuremath{z^{\prime}} - 3.6$ color of all sources was computed using an
aperture of diameter three IRAC pixels ($3.66 \arcsec$). No aperture
corrections were applied to the $\ensuremath{z^{\prime}}$ photometry because the $3.66
\arcsec$ color aperture was much larger than the median seeing.
However, the IRAC point spread function (PSF) has broad wings compared
to a typical ground-based seeing profile, and it was necessary to
apply aperture corrections to the measured 3.6 $\micron$ magnitude
before computing the color of each galaxy.
\subsection{Cluster selection}\label{selection}
Galaxy clusters were identified in the SpARCS survey, using an
algorithm very similar to that described in detail in \citet{mwl08},
as applied to the Spitzer First Look Survey (FLS;
\citealt{lwm05}). For cluster detection in both the FLS and SpARCS
surveys, we used an infrared adaptation of the cluster red-sequence
(CRS) technique \citep{gy00,gy05}. This algorithm maps the density of
galaxies in a survey within narrow color slices, giving greater weight
to brighter galaxies, and flagging the highest overdensities as
candidate clusters. The one important difference here, compared to
\citet{mwl08}, is that the latter paper used an $R-[3.6]$ color to
detect clusters at $0 < z < 1.3$ in the FLS. The deeper IRAC data in
SWIRE ($4\times30$s frames), compared to the FLS ($5\times12$s
frames), combined with the $\ensuremath{z^{\prime}}-[3.6]$ color choice, allow SpARCS
to detect clusters to higher redshift than was possible with the FLS
dataset.
From analysis of the richnesses and colors, \CLc\ and \CLb\ were
identified as rich cluster candidates at $z>1$, and \CLa\ as an
unusually rich cluster candidate at $z\sim0.9$
(Table~\ref{tab_clusters}).
\subsection{Spectroscopy}
Spectroscopic observations of \CLa, \CLc\ and \CLb\ were obtained with
the Low Resolution Imaging Spectrograph (LRIS; \citealt{occ95}) on the
Keck I telescope on the nights of April 3th and April 4th, 2008. The
seeing on both nights was about $1$\arcsec. LRIS has beam-splitters
that separate the light between two sides: red \citep{occ95} and blue
\citep{mcb98}. On the red side, we used the 400 line mm$^{-1}$
grating which gives a coverage of $\Delta \lambda = 3810$
\AA\ centered at 8500 \AA, with a dispersion of 1.86 \AA/pixels. All
the masks were designed with slits of 1\farcs1 width, which gave a
resolution of $\sim$7-9 \AA\ (FWHM) over the wavelength interval
$\sim$6600-9600 \AA. For the redshift range of these particular
clusters, all of the prominent spectral features e.g., the 4000
\AA-rest-frame break or [O$\mathrm{II}$] emission line, fall in the
LRIS red side wavelength window.
We observed one mask each in the case of clusters \CLa\ and \CLc, and
two in the case of \CLb. The total exposure time was 6000s ($5\times
1200$) for \CLa, 8400s ($7\times 1200$) for \CLc, and 8400s ($7\times
1200$) and 7000s ($6\times 1200$) for mask $a$ and mask $b$ of
\CLb\ (Table~1). The masks contained about 30 slits each (five of
these were alignment stars and the remaining 25 were galaxies,
selected according to a sliding scale of priority).
To reduce the number of slits placed on obvious foreground galaxies,
we prioritized slits on galaxies with colors near each cluster's
red-sequence. To avoid significant selection bias, we used a very
broad cut around the red-sequence, intended to include both star
forming and non-star forming systems. Slits were placed on galaxies
with priorities from 1 to 4. Priority 1 was galaxies with colors
within 0.2 magnitude of the RS, and with $[3.6] < 17.0$ (Vega).
Priority 2 was galaxies with colors between 0.2 and 0.6 magnitudes
bluer than the RS, and with $17.0 < [3.6] < 18.7$. Priority 3 was
galaxies with same colors as priority 1, but with $17.0 < [3.6] <
18.7$, and priority 4 was galaxies with same colors as priority 2, but
with $17.0 < [3.6] < 18.7$. Priorities 1 to 4 roughly correspond to
bright red-sequence, blue cloud, faint red-sequence, and faint blue
cloud galaxies respectively. Approximately ten priority 1 galaxies
could be accomodated per mask for \CLa, and five per mask for
\CLc\ and \CLb.
To reduce the LRIS data, we adapted custom software, based on that
developed by \citet{drl05,drl07} to reduce VLT/FORS2 data. Overscan
and bias corrections were applied to both the calibration (bias, flats
and lamp arcs) and to the science (clusters and standard stars) frames
using the IRAF {\it lrisbias} and {\it lccdproc} tasks developed by
G. D. Wirth and C. Fassnacht\footnote{The LRIS software for IRAF is
available at:
http://www2.keck.hawaii.edu/inst/lris/kecklris.html}. After these
corrections had been applied, the data were next corrected for
geometric distortions across the dispersion axis, and then separated
into individual slitlets and reduced using standard long-slit
techniques. Each individual slitlet was processed using an algorithm
similar to that implemented in {\it bogus} (developed by D. Stern,
A. J. Bunker and S. A. Stanford)\footnote{This software can be
obtained from:
https://zwolfkinder.jpl.nasa.gov/$\sim$stern/homepage/bogus.html}.
The flatfield slitlets were normalized and applied to the
corresponding science slitlets. The individual 2D spectra were then
background-subtracted and fringe-corrected before being stacked to
produce a final co-added 2D spectrum. Slits longer than 90 pixels
($\geq19$\arcsec) were further corrected for ``long-slit" distortions.
The 1D spectrum of the source (or sources) in each slitlet was then
extracted using standard IRAF tasks.
A set of NeAr lamp exposures were used to wavelength-calibrate the
observations. LRIS spectra can suffer significant flexure which may
introduce wavelength offsets as large as $\sim16$ \AA\ in the
wavelength solution, in the most extreme cases. The offsets in the
wavelength calibration due to flexure were corrected by comparing to
the wavelengths of known skylines. We estimate the residual
uncertainties in wavelength calibration to be at most 1 \AA, which
correspond to a redshift uncertainty of $\delta z \sim 1 \times
10^{-4}$ at $\lambda=8500$ \AA. Finally, a relative flux calibration
was performed, using a sensitivity function spanning the range 5600 -
9400 \AA\, which was obtained from long-slit observations of the
standard stars Feige 34 and Feige 67 \citep{o90}.
\section{Analysis}\label{analysis}
\subsection{Redshift catalogs}\label{z_cat}
A redshift solution was sought for each object by cross-correlating
\citep{td79} each of the LRIS spectra with observed galaxy templates
\citep{kcb96,ssp03} using the task {\it XCSAO} in the IRAF {\it RVSAO}
package \citep{kmw92}. If a redshift could be obtained for a spectrum,
it was assigned a quality flag, ``Q'', with one of two possible
values: 0 or 1. High-confidence redshifts, unambiguously determined by
two or more clear features in the continuum or in absorption, or by
obvious emission lines such as [O$\mathrm{II}$] ($\lambda3727$) were
assigned $Q=0$. Lower-confidence redshifts based on ambiguous
identification of a few spectral features, often in low S/N spectra,
were assigned $Q=1$.
In addition to a quality flag, an emission line flag, ``E'', was also
assigned to each galaxy, based on the presence or absence of emission
line features. Spectra displaying one or more emission feature such
as H$\beta$, [O$\mathrm{III}$]($\lambda$4959,$\lambda$5007) or
H$\alpha$ in the case of low-redshift ($z<0.8$) galaxies, or
[O$\mathrm{II}$]($\lambda$3727) in the case of higher redshift
galaxies, were assigned $E=1$. Spectra showing no sign of excess
emission above the stellar continuum, were assigned $E=0$. The total
number of galaxies (cluster members and foreground or background
galaxies, excluding stars) extracted from the fields of \CLa,
\CLb\ and \CLc\ were 22, 19 and 34 (see Tables \ref{tab_EN1_240},
\ref{tab_EN1_349} and \ref{tab_EN1_381}). Occasionally, one or more
serendipitous sources fell within a slit. The breakdown by quality
flag was 20, 13 and 14 $Q=0$ galaxies, and 2, 6 and 20 $Q=1$ galaxies,
respectively.
Example spectra for a subsample of cluster members (see
Section~\ref{cl240} for the definition of a ``cluster member'') are
shown in Figure~\ref{cl_specs}. The left column shows examples of
\CLa\ members, the center column shows examples of \CLc\ members, and
the right column shows exampels of \CLb\ members. The ID of each
object is indicated (Tables \ref{tab_EN1_240}, \ref{tab_EN1_349} and
\ref{tab_EN1_381}). Fluxes are in relative units, and smoothed by 7
pixels (1 pix $\sim 2$ \AA). Prominent spectral features are indicated
by vertical lines.
\section{Results}\label{results}
\label{results}
\subsection{\it \CLa}\label{cl240}
In determining cluster membership, and calculating a velocity
dispersion and mass, only high confidence, $Q=0$, galaxies were
considered. Definitive cluster membership was determined using the
code of \citet{b06}, which is based on the shifting-gap technique of
\citet{fgg96}. This procedure uses both galaxy angular position and
radial velocity information to exclude near-field interlopers.
The squares in the left panel of Figure~\ref{z_histo_240} show the 14
$Q=0$ galaxies identified as members of cluster \CLa\ (see
Table~\ref{tab_EN1_240}) by the shifting-gap technique. For \CLa,
these $Q=0$ cluster members fall in the range $0.84 < z < 0.90$,
indicated by the vertical dashed lines shown in the right panel of
Figure~\ref{z_histo_240}, which shows the redshift histogram for the
\CLa\ field.
The properties of the cluster members and non-members are summarized
in Table~\ref{tab_EN1_240}. The total 3.6 $\micron$ magnitude (column
4 in Table~\ref{tab_EN1_240}), and $\ensuremath{z^{\prime}} - 3.6$ color (column 5 in
Table~\ref{tab_EN1_240}) were calculated as described in
Section~\ref{photometry}. In determining membership, we did not
require that any $Q=1$ galaxies satisfy the shifting-gap criteria; if
their redshifts fell in the range $0.84 < z < 0.90$ we included them
as ``cluster members" in Table~\ref{tab_EN1_240}, but re-emphasize
that we do not utilize them in estimating the cluster mean redshift or
velocity dispersion. The total number of cluster members is 16 (14
with $Q=0$ and 2 with $Q=1$), and six foreground/background galaxies.
All confirmed members of cluster \CLa\ were passive ($E=0$) galaxies.
In some cases, a redshift was obtained which did not correspond to any
galaxy in our photometric catalog. In the case of faint galaxies, this
was because a spectroscopic redshift was obtained for a strong
emission line galaxy whose continuum fell below the detection
threshold of the $\ensuremath{z^{\prime}}$ catalog. In the case of bright galaxies,
this was because of blending issues with the IRAC $1.8\arcsec$ $3.6
\micron$ PSF. These galaxies were assigned both a $[3.6]$-band
magnitude and a $z^{\prime}-[3.6]$ color of 99 in
Table~\ref{tab_EN1_240}. Figure~\ref{proj_dist_240} shows
$r^{\prime}z^{\prime}[3.6]$ color composites of cluster \CLa. The
$r^{\prime}$ data were obtained from WFC on the Isaac Newton
Telescope, and are available with the SWIRE public data release
\citep{sur05}. The white squares (green circles) overlaid on the right
panel show the 16 cluster members (and foreground/background galaxies
in the $7 \arcmin$ FOV) with spectroscopically-confirmed redshifts
from Keck/LRIS (see Table~\ref{tab_EN1_240}).
Once cluster membership was established, both the redshift and
rest-frame velocity dispersion, $\sigma_z$, of SpARCS J003550-431224
were then calculated iteratively using the ``robust estimator'',
$\sigma_{rob}$ \citep{bfg90}. The robust estimator has been shown to
be less sensitive than the standard deviation to outliers which may
persist even after rejecting interlopers using the shifting-gap
technique. The actual estimator used depends on the number of cluster
members and is either the biweight estimator for datasets with at
least 15 members, or, as here in the case of \CLa\ with 14 $Q=0$
members, the gapper estimator. The gapper estimator is discussed more
fully in \citet{bfg90}, \citet{gbg93} and \citet{b06}.
The line-of-sight rest-frame velocity dispersion, $\sigma_v$, was
calculated directly from the vector of spectroscopic redshift
measurements, \overrightarrow{z}, as
\begin{equation}
\sigma_v=\frac{\sigma_z(\overrightarrow{z})\times c}{1+z_{cl}} \ ,
\label{veldisp}
\end{equation}
\noindent
where $c$ is the speed of light, and $\sigma_z(\overrightarrow{z})$ is
the estimated dispersion of the measured redshifts with respect to the
center of the distribution, $z_{cl}$.
A mean redshift of $z_{cl}=0.871\pm0.002$ and a velocity dispersion of
$\sigma_v = 1230\pm320$ km s$^{-1}$ were calculated for
\CLa\ (Table~\ref{tab_clusters}). The uncertainty on the latter was
determined using Jackknife resampling of the data. For comparison, a
Gaussian with an rms of 1230 km s$^{-1}$ has been overlaid on the
redshift histogram in the right panel of Figure~\ref{z_histo_240}.
The line-of-sight rest-frame velocity dispersion can be used to
calculate a dynamical estimate of $R_{200}(z)$, the radius at which
the mean interior density is 200 times the critical density,
$\rho_{crit}$, and $M_{200}(z)$, the mass contained within
$R_{200}(z)$. In the spherical collapse model, at redshift $z$,
$R_{200}$ can be calculated from
\begin{equation}
R_{200} = \frac{\sqrt{3}\sigma}{10H(z)} \,
\end{equation}
\noindent
where
$H(z)=H_0\sqrt{(\Omega_M(1+z)^3+\Omega_k(1+z)^2+\Omega_{\Lambda})}$,
is the Hubble parameter at redshift $z$, and
\begin{equation}
M_{200}(z)=3 \frac{\sigma^2_v R_{200}(z)}{G} \ .
\end{equation}
\noindent
Based on its velocity dispersion of $\sigma_v = 1230\pm320$ km
s$^{-1}$, we estimate an $R_{200}= 1.9\pm0.5$ Mpc and a dynamical mass
of $M_{200}=(2.0^{+2.0}_{-1.2})\times10^{15}$ M$_{\odot}$ for
\CLa\ (Table~\ref{tab_clusters}). Although this mass is preliminary
(see section~\ref{richness}), it seems likely that \CLa\ is an
unusually massive cluster, perhaps the most massive cluster in the
entire SpARCS survey, and comparable in mass to that of cluster MS
1054-03 at $z=0.83$ \citep{hoe00, gio04, jee05, tran07}.
The $z^{\prime}-[3.6]$ vs. $z^{\prime}$ color-magnitude diagram for
all galaxies (gray circles) within a radius of $R_{200}$ ($=1.9$ Mpc)
of the center of \CLa\ is shown in Figure~\ref{colmag_240}. This
radius is approximately equal to the virial radius.
Spectroscopically-confirmed galaxies are shown by colored symbols. In
this figure (and in Figures~\ref{colmag_349} and \ref{colmag_381}),
cluster members ($Q=0$ or $Q=1$) are shown by red circles and
foreground/background galaxies by blue squares. Note that there are
several cluster members or foreground/background galaxies shown in
Table~\ref{tab_EN1_240}, which fall within a projected radius of
$R_{200}$ of the cluster center, but for which a color could not be
determined. These galaxies do not appear in Figure~\ref{colmag_240}
(or in Figures~\ref{colmag_349} or \ref{colmag_381}, in the cases of
Tables~\ref{tab_EN1_349} and Tables~\ref{tab_EN1_381}).
\subsection{\it \CLc}\label{cl349}
Seven $Q=0$ galaxies were determined to be cluster members of \CLc\ by
the shifting-gap technique. These galaxies are shown by squares in the
left panel of Figure~\ref{z_histo_349}. These galaxies lie in the
redshift range $1.14 < z < 1.19$, indicated by the vertical dashed
lines in the right panel of
Figure~\ref{z_histo_349}. Table~\ref{tab_EN1_349} summarizes the
``cluster members", the ten $Q=0$ and $Q=1$ galaxies with redshifts in
this range, and the nine foreground/background galaxies. The right
panel of Figure~\ref{z_histo_349} shows the redshift histogram for the
field of \CLc. A Gaussian with an rms of 950 km s$^{-1}$ has been
overlaid.
Figure~\ref{proj_dist_349} shows $r^{\prime}z^{\prime}[3.6]$ color
composites of cluster \CLc. The white squares (green circles)
overlaid on the right panel show the cluster members
(foreground/background galaxies) with spectroscopically-confirmed
redshifts (see Table~\ref{tab_EN1_349}) which fall within the
$5\times5\arcmin$ FOV of the image.
Of the seven $Q=0$ cluster members, five were classified as passive
($E=0$), one was classified as emission line ($E=1$), and one was
classifed as an AGN ($E=2$). The upper left white square in
Figure~\ref{proj_dist_349} corresponds to the AGN. There were also
three passive $Q=1$ cluster members (Table~\ref{tab_EN1_349}). The
spectrum of the confirmed AGN (object ID $\#$ 854892 in
Table~\ref{tab_EN1_349}) is shown in the lowest panel of the center
column in Figure~\ref{cl_specs}. This source shows
NeV($\lambda$3346,$\lambda$3426) and
NeIII($\lambda$3869,$\lambda$3968) in emission, which are common in
AGN, as well as prominent [OII]($\lambda$3727) emission and high-order
balmer lines (H$\delta$, H$\epsilon$ and H6) also in emission.
The cluster mean redshift and velocity dispersion, were estimated
iteratively from the seven $Q=0$ members, using the gapper estimator.
The mean redshift of \CLc\ was calculated to be
$z_{cl}=1.161\pm0.003$. The velocity dispersion was calculated to be
$\sigma_v = 950\pm330$ km s$^{-1}$, which corresponds to $R_{200}=
1.2\pm0.4$ Mpc, and a dynamical mass of $M_{200}=(7.7^{+11}_{-5.5})
\times10^{14}$ M$_{\odot}$(Table~\ref{tab_clusters}).
The $z^{\prime}-[3.6]$ vs. $z^{\prime}$ color-magnitude diagram for
all galaxies (gray circles) within a radius of $R_{200}$ ($=1.2$ Mpc)
of the center of \CLc\ is shown in Figure~\ref{colmag_349}. The red
circles and blue squares indicate those cluster members and
foreground/background galaxies which lie within a projected distance
of $R_{200}$ of the cluster center, and for which a color could be
determined.
\subsection{\it \CLb}\label{cl381}
Seven $Q=0$ galaxies were also determined to be cluster members of
\CLb\ by the shifting-gap technique. These galaxies are indicated by
squares in the left panel of Figure~\ref{z_histo_381}. Two galaxies,
indicated by crosses in Figure~\ref{z_histo_381} (ID $\#$'s 727869 and
734082 in Table~\ref{tab_EN1_381}), were identified as near-field
interlopers.
The seven $Q=0$ members lie in the redshift range $1.19 < z < 1.22$,
shown by the vertical dashed lines in the right panel of
Figure~\ref{z_histo_381}. Table \ref{tab_EN1_381} summarizes the
``cluster members", the ten $Q=0$ or $Q=1$ galaxies with redshifts in
this range, and the 24 foreground/background galaxies. The right
panel of Figure~\ref{z_histo_381} shows the redshift histogram for the
field of \CLb. A Gaussian with an rms of 410 km s$^{-1}$ has been
overlaid.
Figure~\ref{proj_dist_381} shows $r^{\prime}z^{\prime}[3.6]$ color
composites of cluster \CLb. The white squares (green circles)
overlaid on the right panel show the cluster members
(foreground/background galaxies) with spectroscopically-confirmed
redshifts (see Table~\ref{tab_EN1_381}) which fall within the
$6\times6 \arcmin$ FOV of the image. Of the seven $Q=0$ cluster
members, all were classified as emission line ($E=1$). Of the three
$Q=1$ cluster members, one was classified as passive ($E=0$), and two
as emission line ($E=1$). The cluster mean redshift and velocity
dispersion, were calculated iteratively from the seven $Q=0$ members,
using the gapper estimator. The mean redshift of \CLb\ was estimated
to be $z_{cl}=1.210\pm0.002$ and the velocity dispersion, $\sigma_v =
410\pm300$ km s$^{-1}$, which corresponds to $R_{200}= 0.51\pm0.38$
Mpc and a dynamical mass of $M_{200}=(0.60^{+2.5}_{-0.59})
\times10^{14}$ M$_{\odot}$ (Table~\ref{tab_clusters}).
\CLb\ is significantly less massive than \CLa\ or \CLc. The much
smaller $R_{200}$ estimated for \CLb, combined with the geometric
limitations of LRIS with respect to the redshift number density yield
measurable from a single mask, resulted in a yield of only four
spectroscopically confirmed members within $R_{200}$ from two masks.
Moreover, because of blending issues with IRAC'S $1.8\arcsec$ $3.6
\micron$ PSF, a reliable $z^{\prime}-[3.6]$ color could not be
determined for one of these galaxies (ID $\#$ 4000013 in
Table~\ref{tab_EN1_381}). Figure~\ref{colmag_381} shows the
$z^{\prime}-[3.6]$ vs. $z^{\prime}$ color-magnitude diagram for all
galaxies (gray circles) within a radius of $R < R_{200}$ ($=510$ kpc)
of the center of cluster \CLb. The three red circles denote the
cluster members with ID $\#$'s 723814, 722784, and 722712 in
Table~\ref{tab_EN1_381}.
\section{Discussion and Conclusions}\label{discussion}
\subsection{Red-Sequence Photometric Redshifts}\label{RS}
Color-magnitude diagrams for \CLa, \CLb\ and \CLb\ were presented in
Figures~\ref{colmag_240}, \ref{colmag_349}, and
\ref{colmag_381}. Column 2 of Table~\ref{tab_properties} shows the
$\ensuremath{z^{\prime}}-[3.6]$ color of the red sequence for \CLa, \CLb\ and
\CLb. The color is calculated from the mean color of the
spectroscopically confirmed red-sequence cluster members. Also shown
in column 2 of Table~\ref{tab_properties} is the color of the RS for
three additional clusters, previously reported in \citet{mwy09} and
\citet{wmy09}. These clusters are \CLd\ at $z=1.180$, \CLe\ at
$z=1.196$, and \CLf\ at $z=1.335$ (column 6). For all six clusters a
systematic uncertainty in the RS color of 0.15 magnitude has been
assumed (This uncertainty reflects the fact that, at present, we are
using the zeropoints provided by
ELIXIR\footnote{http://www.cfht.hawaii.edu/Science/CFHTLS-DATA/elixirhistory.html}
for the \ensuremath{z^{\prime}}\ observations. We expect, in the future, to be able
to reduce these photometric uncertainties, using our own internal
calibration).
Columns 3, 4 and 5 of Table~\ref{tab_properties} show the redshift
that would be estimated for each cluster based on the measured RS
color (column 2), assuming a solar metallicity single burst BC03 model
and a formation redshift of either $z_{f}=3$, 4 or 10. As can be seen
from Table~\ref{tab_properties} (and the left panel of
Figure~\ref{zsandmasses}), the photometric redshift inferred from the
measured $z^{\prime}-[3.6]$ color has a slight dependence on one's
choice of formation redshift, although differences in color between
the models at $z\sim1$ are fairly small ($\Delta m =0.1$ between
$z_f={4}$ and $z_f={10}$). Utilizing the colors from the $z_{f}=4$
model, the three new clusters presented here were assigned
preliminarily redshift estimates of $z_{\rm phot}=0.84$, $z_{\rm
phot}=1.09$, and $z_{\rm phot}=1.20$ (Table~\ref{tab_properties}).
These photometrically estimated redshifts are very similar to, albeit
slightly lower than, the spectroscopically determined values of
$z_{spec}=0.871$, $z_{spec}=1.161$, and $z_{spec}=1.210$ (column 6 in
Table~\ref{tab_properties}).
The $\ensuremath{z^{\prime}}- [3.6]$ color vs. spectroscopic redshift for all six
SpARCS clusters in Table~\ref{tab_properties} is plotted in the left
panel of Figure~\ref{zsandmasses}. The solid, dotted and and dashed
lines show the BC03 model colors as a function of redshift for
formation redshifts of $z_{f}=3$, 4 and 10. It is clear from
Figure~\ref{zsandmasses} that the agreement between the model colors
and the observations is very good.
With a larger sample of clusters, there may turn out to be small but
real discrepancies between the models and the measured RS colors. In
the left panel of Figure~\ref{zsandmasses}, the $z_{f}=4$ model can be
seen to be slightly redder than the observations in the case of five
clusters (or equivalantly, the inferred photometric redshift can be
seen to be slightly lower than the spectroscopic redshift), but
slightly bluer in the case of one cluster (\CLf). These small offsets
in color between the models and the observations, can also be seen
directly from Figures~\ref{colmag_240}, \ref{colmag_349}, and
\ref{colmag_381}. The solid lines in these three figures show the RS
color predicted by the BC03 $z_{f}=4$ model \emph{at the spectroscopic
redshift of the cluster}, and can be seen to be slightly redder than
the observed color. A more detailed comparison between the model
predictions and the observations will be made in a future paper
employing a larger sample of SpARCS clusters.
Despite the aforementioned caveats, and the issue of degeneracies
between the photometric redshift of the clusters and the formation
redshifts of their galaxies, our overall conclusion is that the
inferred one-color photometric redshifts and the spectroscopic
redshifts are in excellent ($\Delta z \lesssim 0.1$) general
agreement.
\subsection{Cluster Masses estimated from the Richness Parameter, $B_{gc,R}$}\label{richness}
In addition to estimating the masses for \CLa, \CLc, and \CLb\ from
the galaxy line-of-sight velocity dispersion, we also estimated the
masses from the richness of the clusters, using the $B_{gc,R}$
richness parameter. \citet{gy05} introduced $B_{gc,R}$, an adaptation
of the $B_{gc}$ richness parameter, intended to utilize two-band
photometry to increase the contrast of the cluster with the
background, and therefore provide a measurement of the richness that
is less sensitive to foreground/background large scale structures.
$B_{gc}$ is the amplitude of the three-dimensional, cluster
center-galaxy spatial correlation function, $\xi(r) \sim B_{gc}
r^{-1.8}$ \citep{yl99}.
Instead of counting galaxies in a single passband, $B_{gc,R}$ is
obtained by counting galaxies in a color slice centered on the
location of each cluster's red-sequence in the $z^{\prime}-[3.6]$
color-magnitude diagram. In computing $B_{gc,R}$, we used a slice
bounded in color by $z^{\prime}-[3.6] = \pm0.3$ of the best-fit RS
color returned by the cluster finding algorithm (\citealt{mwl08}), and
bounded in magnitude by $(\ensuremath{M_{\star}}+1)$, where $\ensuremath{M_{\star}}$ is the BC03
$z_{f}=4$ model prediction of the characteristic magnitude of a galaxy
\emph{at the photometric redshift} corresponding to that RS color
(Table~\ref{tab_properties}). The background galaxy counts were
determined from the color distribution in the entire 7.9 deg$^2$
ELAIS-N1 field, minus the regions known to contain galaxy clusters.
The $B_{gc,R}$ richnesses of the three clusters were computed to be
$2452\pm422$ Mpc$^{1.8}$ (\CLa), $1762\pm358$ Mpc$^{1.8}$ (\CLc), and
$819\pm246$ Mpc$^{1.8}$ (\CLb). Based on the empirical calibration of
$B_{\rm gc}$ vs. $M_{200}$ determined by \citet{myh07} in the K-band
for 15 CNOC1 clusters at $z \sim 0.3$, these richnesses correspond to
$M_{200} = (22.4\pm4.2) \times 10^{14} \ensuremath{M_\odot}$ for \CLa, $M_{200} =
(13.1\pm2.8) \times 10^{14} \ensuremath{M_\odot}$ for \CLc, and $M_{200} = (3.8\pm1.2)
\times 10^{14} \ensuremath{M_\odot}$ for \CLb.
For comparison, columns 7 and 8 of Table~\ref{tab_properties} show the
dynamical mass, $M_{200}^{\sigma}$, and richness mass,
$M_{200}^{B_{gc}}$, estimates for all six SpARCS clusters. Although
the uncertainties associated with both of the mass estimators are
large, they are consistent with each other at the 1$-\sigma$ level for
five out of the six clusters and at the 2$-\sigma$ level for the
sixth. The agreement between the two mass estimators can be seen in
the right panel of Figure~\ref{zsandmasses}. Based on all six SpARCS
clusters spectroscopically confirmed to date, our conclusion is that,
in addition to there being excellent agreement between the photometric
and spectroscopic redshifts, there is also reasonable agreement
between the cluster dynamical and richness mass estimates. The
dynamical masses estimates should be considered preliminary at this
stage. Uncertainties will reduce as more data becomes available from
a large spectroscopic follow-up program of SpARCS clusters currently
being carried out at the Gemini telescopes.
\subsection{$z>1$ Cluster Surveys}\label{surveys}
Collectively, the six SpARCS clusters confirmed to date (the three
clusters presented in \citealt{mwy09} and \citealt{wmy09}, plus the
three clusters presented here), demonstrate that, given the
availability of infrared observations, the RS technique is an
efficient and effective method of detecting \emph{bona fide} massive
galaxy clusters at $z \gtrsim 1$. Moreover, our studies of these six
clusters are showing that it is possible to infer fundamental
parameters such as cluster redshift and mass \emph{from the survey
data itself} (see also \citealt{ebg08}).
At $z<1$, both the optical Red-sequence Cluster Surveys, RCS-1
\citep{gy00,gy05} and RCS-2 \citep{y07}, and The Sloan Digital Sky
Survey (SDSS; \citealt{york00, ko07b}) have shown that it is feasible
to measure cosmological parameters from the evolution of the cluster
mass function \citep{gl07, roz09b}. In order to do this efficiently,
the survey data themselves are used to detect clusters, and also to
estimate the redshift and the mass of those clusters. The redshift is
estimated from the red sequence color \citep{gil07, gl07, ko07b}, and
the mass is estimated from the optical richness \citep{ye03,
gil07,bec07, roz09a}. The fact that SpARCS is also now
demonstrating the practicality of estimating redshifts and masses at
$z \gtrsim 1$ from the survey data alone is heartening for the current
generation of surveys aiming to utilize optical-infrared high redshift
cluster observations to constrain cosmological parameters e.g.,
SpARCS, The UKIRT Infrared Deep-Sky Survey Deep Extragalactic Survey
(UKIDSS DXS; \citealt{law07}),and the IRAC Shallow Cluster Survey
(ISCS; \citealt{ebg08}). These optical-IR surveys will provide
complementary samples to those selected using the Sunyaev-Zel'dovich
effect, e.g., The South Pole Telescope Survey (SPT;
\citealt{ruhl04,caa09}), The Atacama Cosmology Telescope (ACT;
\citealt{kos03}), and The Atacama Pathfinder EXperiment (APEX;
\citealt{dob06}).
The complete SpARCS catalog contains several hundred cluster
candidates at $z > 1$. With new large, homogeneous, reliable $z>1$
catalogs becoming available from SpARCS and other surveys in the very
near future, the prospects look bright for high redshift cluster and
cluster galaxy evolution studies in the coming years.\\
\acknowledgments
We thank the referee for useful comments. The authors wish to
recognize and acknowledge the very significant cultural role and
reverence that the summit of Mauna Kea has always had within the
indigenous Hawaiian community. We are most fortunate to have the
opportunity to conduct observations from this mountain. This work is
based in part on archival data obtained with the Spitzer Space
Telescope, which is operated by the Jet Propulsion Laboratory,
California Institute of Technology under a contract with NASA. Support
for this work was provided by an award issued by JPL/Caltech. GW also
gratefully acknowledges support from NSF grant AST-0909198, and from
the College of Natural and Agricultural Sciences at UCR.
{\it Facilities:} \facility{Keck (LRIS)}, \facility{Spitzer (IRAC)}, \facility{ CTIO (MOSAIC)}, \facility{CFHT (MegaCam)}.
|
\section{I. Introduction}
One-dimensional nanostructures have attracted much attention since they provide potential applications for nanoelectronics and nanophotonics \cite{1Dnano,Yang01,Nobis,Vugt,zhchen}.
For example, zinc oxide nanorods with hexagonal cross-section can be applied as whispering gallery resonators, in which the coupling between the resonant modes and free excitons depends sensitively on the cross-sectional radius \cite{Nobis,Vugt,zhchen}. Growth of nanorods have been extensively reported thus far, yet there are few studies concentrating on the underlying atomistic mechanisms of growth, especially on the understanding and controlling the cross-sectional radius of nanorods from atomic point of view \cite{wang04,rzf02,Kong,variT,variP,WangRC,Umar06,PLD}.
It is known that surface kinetics plays an important role in determining the morphology and size of nanostructures \cite{Zhang97,Tersoff94,Krug,Villain00}. The deposited adatoms can either diffuse within the topmost layer and aggregate to form a new layer nucleus, or hop downward across the step edges and contribute to the lateral growth of the topmost layer.
Under a certain deposition condition, the kinetics-controlled competition between the growth in the normal direction to the substrate and the lateral growth is expected to determine the growth modes and morphologies \cite{Tersoff94,Krug}.
The surface kinetics can be described by intralayer and interlayer hopping rates of adatoms, $\nu =\nu _{0}\exp (-E_{d}/kT)$ and $\nu ^{\prime }=\nu'_{0}\exp (-E_{s}/kT)$, respectively, where the prefactors $\nu_0$ and $\nu'_0$ are the attempt rates which are approximately of the same value; $k$ is the Boltzmann's constant and $T$ the temperature. The interlayer diffusion barrier ($E_{s}$) is normally larger than the intralayer one ($E_{d}$). The difference of these two values is denoted as $E_{es}$, which is known as Ehrlich-Schwoebel barrier (ESB) \cite{ES1,ES2}. The ratio of $\nu/\nu'$ increases with ESB as $\exp(E_{es}/kT)$. The ESB is reported to increase with the step height and saturate in several atomic layers, with a value usually referred as three-dimensional (3D) ESB. The conventional ESB is applicable for a monolayer step and is hereafter termed as two-dimensional (2D) ESB \cite{LiuH,Zhang02}.
When ESB is small enough to allow sufficient inter-layer diffusion, layer-by-layer growth occurs, whereas larger ESB leads to multilayer growth \cite{Tersoff94}. In the latter case, islands initiate from each nucleus, which approach nanorods if the cross-sections remain the same along the longitudinal direction. The key point to select and control the nanorod growth is to understand how the radius of the cross-section varies when atomic layers add up under specific growth conditions. Since there is a large difference between the 2D- and 3D- ESBs, it is also crucial to identify the specific roles of 2D- and 3D-ESBs in the nanorod growth.
In this paper, we study the influence of deposition rate and ESB on the growth process of nanorods. A characteristic radius has been identified, which increases proportional to one fifth power of the ratio of the 2D-ESB limited hopping rate of adatoms to the deposition rate. Both the growth modes and the nanorod diameter can be selected by tuning this characteristic radius.
We demonstrate that when the radius of the initial island is larger than the characteristic radius, the growth morphology evolves from a taper-like structure to a nanorod with radius equal to the characteristic radius. However, if the characteristic radius becomes larger than the radius of the initial island, by increasing the 2D-ESB limited hopping rate or decreasing the deposition rate, nanorod morphology can be maintained during the growth, with a stable radius being limited by both the radius of the initial island and 3D-ESB. The theoretical predictions of the characteristic radius are demonstrated with experimental observations of ZnO growth, and good consistency has been found.
\section{II. Nucleation on top of an island}
Let us consider a nanostructure with thickness of $n$ atomic layers. For simplicity, the cross section is taken as circular. The radius of the $i$-th layer is denoted as $R_i$, measured in terms of the surface cell parameter $a_0$. The dimensionless area and perimeter are $A_i=\pi R_i^2$ and $L_i=2\pi R_i$, respectively.
Assuming the growth units are deposited in the normal direction of the surface, at rate $F$ per surface cell of area $a_0^2$. The number of adatom $\eta$ per surface cell is determined by the diffusion euqation, $d\eta /dt = \nu \nabla^2 \eta +F $. By solving the diffusion equation, the distribution of the number density of adatom can be obtained,
\begin{equation}
\eta = \frac{F}{4\nu}(R_n^2-r^2)+\eta_e, \label{eq:eta}
\end{equation}
where $\eta_e$ is the dimensionless number density of adatom at the edge of $A_n$. Before nucleation occurs on top of $A_n$, a deposited adatom on $A_{n}$ has no other choice but to hop across the step edges after a survival time of $\tau$. At steady state, the number of atoms leaving the surface per unit time, $L_n\eta_e\nu'$, is balanced by that of the atoms deposited on the surface per unit time, $FA_{n}$. It gives the number density of adatom on the boundary,
\begin{equation}
\eta_e= (FR_n)/(2\nu'). \label{eq:eta-e}
\end{equation}
The total number of adatom on $A_n$ can be obtained by integrating,
\begin{equation}
N= \int_0^{R_n} \eta 2\pi r dr=\frac{F\pi R_n^3}{2\nu'}(\frac{R_n}{4\nu/\nu'}+1).
\end{equation}
The average survival time of an adatom is thus $\tau = N\Delta t$, where $\Delta t=1/(FA_{n})$ is the time interval between subsequent deposition events on $A_{n}$. In addition to $\Delta t$ and $\tau$, another concerned time scale is the traversal time for an atom to visit all the sites of $A_{n}$, $\tau _{tr}=A_n/\nu $.
As $A_n$ grows, the probability of nucleation on $A_n$ increases. Once a new nucleus forms on $A_n$, the number of atomic layers $n$ increases by one, which leads to the growth in the normal direction to the substrate. For the simplest case that a dimer is the smallest stable island, the nucleation rate on $A_n$ can be given by $\Omega =p_{1}p_{2}/\Delta t$, where $p_1=1-\exp (-\tau /\Delta t)$ is the probability that an atom is deposited during the presence of another atom on the surface, and $p_2=1-\exp (-\tau /\tau _{tr})$ is the encounter probability \cite{Rottler}.
It has been reported that with slow deposition the total number of adatoms is usually much less than unity, $i.e.$ $N\ll 1$, which means that $\tau \ll \Delta t$ \cite{Krug,Rottler}. Furthermore, in typical island growth with large ESB, $\nu/\nu'$ is much larger than dimensionless $R_n$, so the first term in Eq.~(\ref{eq:eta}) is much smaller than the second term, and the number density of adatom on $A_n$ is approximately uniform, $\eta \simeq \eta_e$. The total number of adatom becomes $N= \frac{F\pi R_n^3}{2\nu'}$, which gives
\begin{equation}
\tau=N\Delta t = \frac{R_n}{2\nu'}.
\end{equation}
It indicates that $\tau \gg \tau_{tr}$ when $\nu/\nu' \gg R_n$, and the encounter probability of two adatoms simultaneously present on the island $p_2$ is approximately unity. The nucleation rate therefore can be approximated as
\begin{equation}
\Omega = \frac{F^{2}A_{n}^{3}}{L_n\nu ^{\prime }} = \frac{F^2 \pi ^2 R_n^5 }{2\nu ^{\prime }}. \label{eq:omega}
\end{equation}
Depending on the height of the steps across which the adatoms hop to the lower layer, 2D-ESB or 3D-ESB plays the role to influence the interlayer atomic diffusion, respectively. Correspondingly subscripts $2D$ and $3D$ will be added to $\nu'$ or $\Omega$ in the context to show such a difference.
\section{III. Characteristic Radius}
For a buried layer, $i.e.$ the atomic layer above which a second layer has formed, the condition $\tau \gg \tau_{tr}$ means that an adatom can always be trapped by the ascending steps before getting chance to hop to the lower layer. The adatoms on $A_n$ before second-layer nucleation, however, has no other choice but to hop to $A_{n-1}$. The lateral growth of the topmost layer $A_n$ is thus contributed by the atoms deposited on area $A_{n-1}$,
\begin{equation}
dA_{n}/dt=FA_{n-1}= F\pi R_{n-1}^2. \label{eq:dA}
\end{equation}
We assume for the moment that $A_{n-1}$ is large enough so that $A_n$ is always smaller than $A_{n-1}$.
The probability $f$ that a second layer has nucleated on $A_n$, according to $df/dt=\Omega(1-f)$, increases with $A_n$ as the following, \begin{equation}
f=1-\exp (-I), \label{eq:f}
\end{equation}
where \begin{equation}
I=\int_{0}^{A_{n}}\Omega_{2D} \frac{dt}{dA}dA \label{eq:I0}
\end{equation}
can be regarded as the average number of nucleus on $A_{n}$. Substituting Eqs.\ (\ref{eq:omega}) and (\ref{eq:dA}) into Eq.(\ref{eq:I0}),
\begin{equation}
I=\frac{F}{\nu_{2D}^{\prime }}\frac{\pi ^{2}R_{n}^{7}}{7R_{n-1}^{2}}=%
\frac{R_{n}^{7}}{R_{c}^{5}R_{n-1}^{2}}, \label{eq:I}
\end{equation}
where $R_c$ is defined as
\begin{equation}
R_{c}=\left( \frac{7\nu_{2D}^{\prime }}{F\pi ^{2}}\right) ^{1/5}. \label{eq:L2d}
\end{equation}
$\nu_{2D}^{\prime }$ in Eq.\ (\ref{eq:L2d}) denotes the hopping rate of adatom from $A_n$ to $A_{n-1}$, with subscript ${2D}$ emphasizing that $R_{c}$ is determined by the 2D-ESB across the monolayer step edges. Since the probability $f$ goes rapidly from nearly zero to nearly unity when $I=1$,
the condition that $I=1$ can be used as a criterion for the formation of the $(n+1)$-th layer \cite{Tersoff94}.
According to Eq.~(\ref{eq:I}), the second layer nucleus has formed before $R_{n}$ reaches $R_{n-1}$ if $R_{n-1}>R_{c}$. If $R_{n-1}<R_{c}$, however, the probability that second-nucleus forms on top of $A_n$ is still nearly zero even though $R_n$ reaches $R_{n-1}$.
As $R_{c}$ is determined just by the 2D-ESB and the growth conditions, such as the deposition rate and the growth temperature, it can be regarded as a characteristic radius of the growth system.
In heterogeneous growth, it is known that the effect of the foreign substrate on surface kinetics properties depends strongly on the thickness of the grown layers. The interlayer hopping rate $\nu_{2D}'$ is therefore variable, especially in the first two layers. Accordingly, we denote hereafter the characteristic radius of the first layer and the other layers as $R_{c0}$ and $R_c$, respectively, in order to distinguish their difference.
It is known that $R_{c0}$ is critical in determination of the growth mode, such as layer-by-layer growth or island growth in the beginning of heterogenous growth. Here we propose that, once the island growth sets in, it is the characteristic radius $R_{c}$ that plays a key role during the development of a separate island in selecting growth mechanisms and the lateral size of the island.
\section{IV. Two Growth Scenarios}
In heterogenous growth, the average radius of the foreign substrate occupied by each nucleus is denoted as $R_{0}$. If $R_0$ is larger than $R_{c0}$, a second-layer nucleus forms when the radius of the first layer approaches $R_{1}=R_{c0}\left( {R_{0}}/{R_{c0}}\right) ^{2/7}<R_0$, which means that the second layer nucleus forms before the first layers coalescence and thus island growth sets in. We define $R_1$ as the radius of the initial island from which the island growth starts.
\begin{figure}[t]
\includegraphics[width=8.5cm]{Fig1}
\caption{Schematic of the two different growth modes from the initial island (grey colors) with radius of $R_1$: (a) $R_1>R_c$; (b) $R_1<R_{c}$. $R_0$ denotes the average radius of the substrate occupied by per island, and $R_n$ is the radius of the topmost layer in the nanorod with $n$ grown atomic layers.}
\label{fig1}
\end{figure}
As discussed in previous section, in heterogenous growth the characteristic radius for the second layer changes from $R_{c0}$ to $R_{c}$. In homogeneous growth, or in late stage of the heterogenous growth where substrate effect is negligible, the characteristic radius can vary by changing the deposition rate $F$ or the temperature $T$, according to Eq.\ (\ref{eq:L2d}). In these cases, $R_{c0}$ and $R_{c}$ are defined as the characteristic radius before and after changing the growth conditions respectively, $R_0$ and $R_1$ correspond to the radius of the topmost two layers at the moment that a nucleus forms on $R_1$ after changing the growth conditions.
Two growth scenarios can be identified according to the way that the characteristic radius changes. The first scenario occurs when $R_c$ decreases from $R_{c0}$, $i.e.$, $R_{c}< R_{c0}$. Since $R_{1}>R_{c0}$, it is still larger than $R_{c}$. Thus the third layer forms atop when ${R_{2}}$ approaches $R_{c}\left( {R_{1}}/{R_{c}}\right) ^{2/7}$. If we assume all the buried layers cease growing, the radius of each layer is fixed at the moment when it is buried by a new layer. Therefore the radius of the $i$-th layer in a nanostructure with $n$ grown atomic layers, $R_i$, is determined by setting $I(R_{i})=1$,
\begin{equation}
\frac{R_{i}}{R_{c}}= \left( \frac{R_{i-1}}{R_{c}}\right) ^{\frac{2}{7}}
=\left( \frac{R_{1}}{R_{c}}\right) ^{(\frac{2}{7})^{i-1}}, (i=2,3,...,n). \label{eq:Rn2d}
\end{equation}
Eq.~(\ref{eq:Rn2d}) indicates that $R_i$ decreases rapidly with increasing $i$, until it approaches $R_{c}$. Correspondingly, the growth morphology changes from a tapered one to a nanorod in several transient layers, as schematically shown in Fig.~1(a).
Obviously this is merely a limiting case in which some strong screening effects exist so that the topmost layer dominates the deeper layers in capturing the deposit atoms \cite{steer}. The opposite limiting case is that in which the growth units are equally deposited on the exposed area, thus the buried layers can also grow. In the latter case island growth leads to the well-known wedding-cake morphology \cite{Krug}.
The realistic situation is that between these two cases, in which only some finite topmost layers are involved in capturing the deposit atoms as a result of a mediate screening effect. To illustrate the screening effect on the island growth, we consider that only the topmost finite $N_g$ layers keep growing laterally. The quantity $1/N_g$, which is in the range of $0$ to $1$, can be taken as a measure of the {\it screening strength}. We have carried out numerical calculations of the rate equations with different values of $N_g$, the details of which will be reported elsewhere. We find that when the number of the atomic layers of the island $n$ is smaller than $N_g$, the island grows with the well-known wedding-cake shape. When $n$ increases larger than $N_g$, the radii $R_i$ for $i<N_g$ grow gradually to $R_0$, while for $N_g<i<n-N_g$ $R_i$ approach their stable values after sufficient growth,
\begin{equation}
\frac{R_{i}}{R_{c}}=X^{\left(\frac{2}{7}\right)^{(i/N_g^2)}}, (N_g<i \leq n-N_g). \label{eq:Ng}
\end{equation}
$X$ is a constant determined by $(R_1/R_c)$ and $N_g$. For $N_g=1$, $X=(R_1/R_c)^{3.5}$, consistent with Eq.~(\ref{eq:Rn2d}).
The radius $R_i$ decreases with $i$, until it approaches $R_c$ after some transient layers and then remains this value. The number of the transient layers is proportional to $N_g^2$, with coefficient $a$ of the order of magnitude of $10$. We therefore show that, under a certain screening strength, wedding-cake morphologies ($n<N_g$), tapered morphologies ($N_g<n<aN_g^2$) and nanorods ($n>aN_g^2 $) can be observed successively during the island growth.
It is thus clear that if the radius of the initial island $R_1>R_c$, the structure finally approaches a nanorod with a tapered base beneath. The {\it screening strength} only influences the number of the transient layers of convergence, $i.e.$, the atomic layers of the tapered base. Since the nucleation takes place on the topmost monolayer, we refer the nanorod converged from $R_{1}> R_{c}$ as the 2D-ESB-limited nanorod, with radius $R_{2D}$ equals to the characteristic radius of the system, $R_{2D}=R_{c}$. We term this growth mode as the 2D-ESB-limited one.
The second scenario occurs when $R_c$ increases from $R_{c0}$ to a value much larger than $R_{1}$. In this case the average number of nucleus on top of $A_2$ when $A_2$ covers $A_1$ is $(R_{1}/R_{c})^{5}\simeq 0$. Therefore the topmost two layers can bunch to a bilayer, which then grows laterally from $R_{1}$ to $R_{2}$ till a new nucleus finally forms atop. The process repeats and the growth morphology remains rod-like as shown in Fig.~1(b). The whole nanorod grows laterally as the number of the atomic layers $n$ increases, fed by the deposited adatom on the top of the nanorod. Therefore,
\begin{equation}
\frac{dA_n}{dt}=\frac{1}{n}FA_n=\frac{1}{n} F\pi R_n^2.
\end{equation}
The average number of nucleus can be obtained by integrating Eq.~(\ref{eq:I0}),
\begin{eqnarray}
I&=& \frac{R^5_{n-1}}{R^5_c} +\int_{A_{n-1}}^{A_{n}} \Omega_{3D} \frac{dt}{dA} dA \nonumber \\
&=& \frac{R^5_{n-1}}{R^5_c}+ \frac{7n}{5} \frac{\nu'_{2D}}{\nu'_{3D}}\frac{R_n^5-R^5_{n-1}}{R_c^5}, \label{eq:T3d}
\end{eqnarray}
where $R_c$ is the same characteristic radius defined in Eq.~(\ref{eq:L2d}). Note $\Omega_{3D}$ denotes the nucleation rate on $A_n$ after $A_n$ approaches $A_{n-1}$, and $\nu _{3D}^{\prime }$ represents the interlayer hopping rate across the multilayer step edges. As discussed above, when $R_{c}$ is much larger than $R_{n-1}$, the first term in Eq.\ (\ref{eq:T3d}) is nearly zero. The radius of nanorod with $n$ atomic layers $R_n$ can be therefore determined by setting $I=1$ in Eq.\ (\ref{eq:T3d}),
\begin{eqnarray}
R_n^5
= R_{n-1}^5 + \frac{5\alpha }{7n} R_c^5, \label{eq:rn5}
\end{eqnarray}
where $\alpha =\nu _{3D}^{\prime }/\nu _{2D}^{\prime }=\exp[-(E_{es}^{3D}-E_{es}^{2D})/kT]$ and $0<\alpha <1$. The radii of the $i-th$ layer $R_i$ can be obtained by iterating Eq.\ (\ref{eq:rn5}),
\begin{eqnarray}
R_i = R_n= \left[ R_{1}^5 + \frac{5}{7}\alpha R_{c}^5 \sum_{h=2}^{n} \frac{1}{h} \right]^{1/5},(i=2,3,...n) \label{eq:Rn3d}
\end{eqnarray}
The nanorod radius increases until the difference of $R_{n}$ and $R_{n-1}$ is smaller than one lattice parameter for sufficient large $n$, when it approaches the stable radius of the 3D-ESB-limited nanorod $R_{3D}$. According to Eq. (\ref{eq:rn5}), this happens when
\begin{equation}
n=n_s=\frac{\alpha R_c^5}{7R_{3D}^4}. \label{eq:nst}
\end{equation}
Substituting Eq.~(\ref{eq:nst}) into Eq.\ (\ref{eq:Rn3d}) and replacing the summation with logarithm for large enough $n$, the stable radius of the 3D-ESB-limited nanorod can be obtained as,
\begin{equation}
R_{3D}=\left[ R_{1}^{5}+\frac{5}{7}\alpha R_{c}^{5}(\gamma-1+\ln \frac{\alpha R_{c}^{5}%
}{7R_{3D}^{4}})\right] ^{1/5}, \label{eq:R3d}
\end{equation}%
wheren $\gamma$ is the Euler's constant.
Since $\alpha R_c^5$ is proportional to $\nu'_{3D}/F$, it is evident that the nanorod growth in this case is determined by the radius of the initial island $R_1$, the deposition rate and the 3D-ESB.
Larger 3D-ESB ( smaller $\alpha $) facilitates the convergence ($i.e.$ smaller $n_s$) and leads to smaller $R_{3D}$, which is consistent with a recent Monte-Carlo simulation on copper nanorod growth \cite{Huang08}.
\begin{figure}[t]
\includegraphics[width=8.5cm]{Fig2}
\caption{The radius of the $n$-th layer $R_n$ in an island with $n$ grown atomic layers, for the cases of $R_1>R_c$ with $N_g=1$ (above the dotted line), and $R_1<R_c$ (below the dotted line), as a function of $n$. The values are calculated according to Eqs.\ (\ref{eq:Rn2d}) and (\ref{eq:Rn3d}). The dotted line guides the characteristic radius $R_{c}$. $\alpha=\nu'_{3D}/\nu'_{2D}$ is the ratio of the hopping rate across the multilayer step to the one across the monolayer step. }
\label{fig2}
\end{figure}
It is worthy to emphasize that Eq.~(\ref{eq:rn5}) is valid only when the first term in Eq.\ (\ref{eq:T3d}) is negligible, and it is physically invalid to extrapolate from Eq.\ (\ref{eq:R3d}) to a $R_{3D}$ larger than $R_c$. Actually, the 2D-ESB-limited growth sets in once $R_{n}$ is increased to $R_{c}$. Therefore the real radius $R_{3D}$ may never exceed $R_{c}$.
For comparison, we show in Fig.~2 the radius $R_{n}$ in a nanostructure with $n$ grown atomic layers for the two scenarios, according to Eqs.\ (\ref{eq:Rn2d}) and (\ref{eq:Rn3d}). It is clear that when the radius of the initial island $R_{1}$ is larger than the characteristic radius $R_{c}$, $R_{n}$ decreases rapidly until it approaches $R_{c}$. Consequently, the growth morphology evolves from a taper-like structure to a nanorod with uniform radius of $R_{c}$. In this scenario, the converged radius is limited by 2D-ESB. When $R_{1}<R_{c}$, $R_n$ corresponds to the nanorod radius. It increases relatively slowly, with a stable radius smaller than $R_c$, determined by the 3D-ESB and the radius of the initial island $R_1$.
\section{V. Experiment Verifications}
\begin{figure}[t]
\includegraphics[width=8.5cm]{Fig3}
\caption{ (a) Schematic of the temperature change in experiments, where the three dashed lines correspond to the moments the growth is cut off by large flux of nitrogen. The ZnO products obtained at the three cut off moments are shown in (b), (c) and (d), respectively. }
\label{fig3}
\end{figure}
Experimentally we take zinc oxide (ZnO) vapor growth as an example to verify the selection of the nanorod radius via varying $R_c$ by changing the growth temperature. For this purpose, unlike conventional ZnO nanorod growth systems where catalysts are usually introduced, here we establish a physical growth system without using additive chemicals.
The nanorods of ZnO were synthesized catalyst-free in a horizontal tube furnace with programmable temperature control. The pure zinc powder (99.9\% Alfa Aesar) and polished Si(100) substrate were arranged in the same quartz boat and 1.0 cm apart. The growth was carried out with flux of nitrogen controlled as 300 standard cubic centimeter per minute (sccm) and oxygen gas flux as 5 sccm. The temperature in the central section of the furnace was homogeneous, where the quartz boat was placed. In each run of the experiment, the temperature was changed as shown in Fig.~3(a) to control the deposition rate. The growth of nanorods was terminated at different time by sudden increasing the nitrogen flux and cutting off the oxygen supply, as illustrated by the three dashed lines in Fig.~3(a). In this way, we expect to preserve the growth morphology at the moment that the growth has been terminated.
In our experiments, the temperature is varied while the flux
of N$_2$ and O$_2$ are kept as constants for ZnO nanorod growth. The evaporated zinc atoms react with oxygen molecules, and the partial pressure of the production ZnO is proportional to that of zinc, $ P_\mathrm {ZnO} \propto P_\mathrm {Zn}/K_p $, where $K_P$ is the reaction constant. Both $P_\mathrm{Zn}$ and $K_p$ exponentially depend on temperature,
\begin{eqnarray}
&& P_\mathrm{Zn}\propto \exp(-\frac{B_\mathrm{Zn}}{kT}),~~
K_P \propto \exp(-\frac{B_K}{kT}),
\end{eqnarray}
where $B_\mathrm{Zn}= 0.58$ $\mathrm{eV}$ $(6776 \mbox{K})$ \cite{crc} and $B_K=0.21$ $\mathrm{eV}$ $(2474 \mbox{ K})$ \cite{wzl}.
The partial pressure of ZnO can be therefore written as,
\begin{equation}
P_\mathrm{ZnO}= P_0 \exp(-B/kT)
\end{equation}
where $B=B_\mathrm{Zn}-B_K=0.37 \mathrm{eV}$, and $P_0$ is a constant determined by the other growth conditions except of temperature. The temperature-dependent deposition rate per lattice site can be written as $F=a_{0}^{2}P_{\mathrm{ZnO}}/\sqrt{2\pi mkT}$.
According to Eq.\ (\ref{eq:L2d}), the characteristic radius $R_c$ is
\begin{equation}
R_{c}= sb(kT)^{1/10}\exp (-\Delta E/kT), \label{eq:zno}
\end{equation}%
where $b=\left[ 7\nu _{0}\sqrt{2\pi m}/(P_{0}a_{0}^{2})\right] ^{1/5}$, $\Delta E=(E_{s}-B)/5$, and $s$ is the geometrical factor associated with the different cross-sectional shapes of nanorod. Equation (\ref{eq:zno}) shows obviously that $R_c$ can be tuned by changing temperature. If $\Delta E$ is positive, $R_c$ decreases with decreasing temperature, then the first (2D-ESB limited) growth scenario is realized, and the radius of nanorod corresponds to the characteristic radius $R_c$. Otherwise if $\Delta E$ is negative, $R_c$ increases with decreasing temperature, and so does the radius of nanorods.
\begin{figure}[t]
\includegraphics[width=8cm]{Fig4}
\caption{
The diameter of the second segment nanorod as a function of the corresponding temperature. The dashed curve gives the theoretical fitting results according to Eq.\ (\ref{eq:zno}).}
\label{fig4}
\end{figure}
Now let us check the variation of morphologies of ZnO nanorods with step-decreased temperature as shown in Fig.~3(a).
The morphologies of the nanorods were characterized by a field emission scanning electron microscopy (LEO 1530VP). Figure 3(b) illustrates ZnO structures grown at constant temperature of $600^{\circ }$C. There is no evident variation of the cross-sectional diameter along the longitudinal direction of the nanorods. When the temperature was decreased from 600 to 550 $^{\circ }$C, a second segment of nanorods appear, the cross-section area of which shrink to smaller values, as shown in Fig.\ 3(c). The morphologies obtained after two temperature drops are shown in Fig.~3(d), where two evident changes of cross-sectional diameter can be identified on the nanorods as guided by the dashed line. The cross-section of the ZnO nanorods are of the typical hexagonal one in all cases.
The experimental observation that the radii of nanorods decrease from one segments to the successive ones with decreasing temperature suggests that the growth mode is the 2D-ESB-limited one. Therefore the nanorod radius under a certain temperature is expected to be equal to the corresponding $R_c$. In order to eliminate the influence of substrates, the temperature dependence of the radii of the nanorods in the second sections are explored, while keeping all the other growth conditions as the same.
In Fig.~4, we plot the circum diameters of the cross-section of the nanorods in the second segments of the structures as shown in Fig.~3(c), as a function of the corresponding temperature. The dashed curve gives the theoretical fitting results according to Eq.\ (\ref{eq:zno}). It shows that the theoretical model is in good consistency with the experimental data. The fit value of $\Delta E$ is $0.59$ eV, which leads to $E_{s}$ of about $3.3$ eV. We should point out that this value is only a rough estimate of the effective barrier against the adatom diffusing across a monolayer step edge in zinc oxide system.
\section{VI. Summary}
We have demonstrated that two nanorod growth modes can be realized, depending on a characteristic radius which is determined by the 2D-ESB and the deposition rate. When the radius of the initial island is larger than this characteristic radius, the nanorod radius is 2D-ESB-limited and approaches the characteristic value. Otherwise the nanorod is 3D-ESB limited with a stable radius smaller than the characteristic radius. We suggest that our results is helpful to select the desired growth modes and control the diameter of nanorods and nanowires.
Although experimental studies on growth of ZnO nanorod with hexagonal cross-section has been reported before, to the best of our knowledge, a quantitative study considering the kinetics in the interfacial growth remains rare. Moreover, the theoretical model proposed here in fact is a generic one that is not limited to ZnO nanorod growth only. Experimentally, if one can precisely tune the characteristic radius $R_{c}$ for a specific growth system, or choose a desired initial radius by using suitable seeds, either kinds of growth modes can be selected in order to obtain different morphologies and nanorod size. By this means, microscopic informations of 2D-ESB and 3D-ESB can also be inferred from the experiments.
\section{Acknowledgement}
This work was supported by NSF of China (10974079 and 10874068) and Jiangsu Province (BK2008012), MOST of China (2004CB619005 and 2006CB921804). Z. Zhang acknowledges partial support by USDOE (grant No. DE-FG02-05ER46209, the Division of Materials Sciences and Engineering, Office of Basic Energy Sciences), and USNSF grant No. DMR-0906025.
|
\section{Introduction}
The principal curvatures of a surface or lamination smoothly embedded in a hyperbolic 3-manifold are related to the topology of the surface and the 3-manifold. For example in \cite{Breslin} we show that incompressible surfaces and strongly irreducible Heegaard surfaces embedded in hyperbolic 3-manifolds can always be isotoped to a surface with principal curvatures bounded in absolute value by a fixed constant that does not depend on the surface or the 3-manifold. In \cite{Breslin3} we show that laminations in hyperbolic 3-manifolds with principal curvatures everywhere close to zero have boundary leaves with non-cyclic fundamental group and that laminations in hyperbolic 3-manifolds with principal curvatures everywhere in the interval $(-1,1)$ have boundary leaves with non-trivial fundamental group.
This note was motivated by a question about surfaces with principal curvatures near the interval $(-1,1)$. It is well known that a closed orientable surface smoothly embedded in a finite-volume complete hyperbolic 3-manifold with principal curvatures everywhere in the interval $(-1,1)$ is incompressible and lifts to a quasi-plane in $\mathbb{H}^3$ (see Thurston's notes \cite{tnotes} or Leininger \cite{Leininger} for a proof). Thus Heegaard surfaces and fibers in hyperbolic 3-manifolds cannot have principal curvatures everywhere in the interval $(-1,1)$. We are interested in finding obstructions to isotoping Heegaard surfaces and fibers in hyperbolic 3-manifolds to have principal curvatures close to the interval $(-1,1)$.
See Rubinstein \cite{Rubminimal} or Krasnov-Schlenker \cite{Krasnov} for more on surfaces in hyperbolic 3-manifolds with principal curvatures in the interval $(-1,1)$.
It follows from Freedman-Hass-Scott \cite{FHS} that an incompressible surface in a closed Riemannian 3-manifold can be isotoped to a minimal surface. It follows from work of Pitts-Rubinstein that a strongly irreducible Heegaard surface in a closed Riemannian 3-manifold can be be isotoped to either a minimal surface or the boundary of a regular neighborhood of a minimal surface (see \cite{Rubminimal} for a sketch of the proof). We show that given an upper bound on the genus of a minimally embedded fiber or Heegaard surface and a lower bound on the injectivity radius of the hyperbolic 3-manifold, there exists a $\delta > 0$ such that the fiber or Heegaard surface must contain a point at which one of the principal curvatures is greater than $1 + \delta$ in absolute value.
\begin{thm}\label{fiber}
For each $g \ge 2$, $\epsilon > 0$, there exists $\delta := \delta(g,\epsilon)$ such that if $S$ is a genus $g$ minimally embedded fiber in a closed hyperbolic mapping torus $M$ with $\operatorname{inj}(M) > \epsilon$, then $S$ contains a point at which one of the principal curvatures is at least $1 + \delta$ in absolute value.
\end{thm}
\begin{thm}\label{heegaard}
For each $g \ge 2$, $\epsilon > 0$, there exists $\delta := \delta(g,\epsilon)$ such that if $S$ is a genus $g$ minimally embedded Heegaard surface in a closed hyperbolic 3-manifold $M$ with $\operatorname{inj}(M) > \epsilon$, then $S$ contains a point at which one of the principal curvatures is at least $1 + \delta$ in absolute value.
\end{thm}
The proofs of Theorem \ref{fiber} and Theorem \ref{heegaard} both use geometric limit arguments. Assuming that no such $\delta > 0$ exists, we consider a sequence of hyperbolic 3-manifolds as in the statment with minimally embedded fibers or Heegaard surfaces whose principal curvatures are closer and closer to the interval $[-1,1]$. After possibly passing to a subsequence, the sequence of manifolds converges geometrically to a hyperbolic 3-manifold $M$ and the surfaces converge to an incompressible surface $S$ in $M$ with principal curvatures everywhere in the interval $[-1,1]$. This implies that the limit set of a lift of $S$ to $\mathbb{H}^3$ is a proper subset of $\partial\mathbb{H}^3$.
In either case, we show that the cover of $M$ corresponding to the image of $\pi_1 (S)$ in $\pi_1 (M)$ has a doubly degenerate hyperbolic structure contradicting that the limit set of a lift of $S$ to $\mathbb{H}^3$ is a proper subset of $\partial\mathbb{H}^3$.
\section{Preliminaries}
Let $M$ be a hyperbolic 3-manifold with no cusps and finitely generated fundamental group. By a result of Scott, $M$ has a \textit{compact core} which is a compact submanifold $C$ of $M$ whose inclusion into $M$ is a homotopy equivalence. The connected components of $M \setminus C$ are called the \textit{ends} of $M$. It follows from the positive solution of the Tameness Conjecture by Agol \cite{agol} and Calegari-Gabai \cite{Cal-Gabai} that an end of $M$ is homeomorphic to $\Sigma \times [0,\infty )$ where $\Sigma$ is a closed orientable surface.
The convex core, $CC(M)$, of $M$ is the smallest convex submanifold of $M$ whose inclusion is a homotopy equivalence. An end $E$ of $M$ is \textit{convex-cocompact} if $E \cap CC(M)$ is compact and $E$ is \textit{degenerate} otherwise. Given a closed orientable surface $\Sigma$ of genus greater than one, a hyperbolic structure on $\Sigma \times \mathbb{R}$ such that both ends are degenerate is called \textit{doubly degenerate}.
A sequence of pointed hyperbolic $n$-manifolds $(M_i ,p_i )$ \textit{converges geometrically} to the pointed hyperbolic $n$-manifold $(M,p)$ if for every sufficiently large $R$ and each $\epsilon > 0$, there exists $i_0$ such that for every $i \ge i_0$, there is a $(1 + \epsilon )$-bilipschitz pointed diffeomorphism $\kappa_i : (B(p,R) , p) \rightarrow M_i$, where $B(p,R) \subset M$ is the ball of radius $R$ centered at $p$ and $B(p_i ,R) \subset M_i$ is the ball of radius $R$ centered at $p_i$. We call the maps $\kappa_i$ \textit{almost isometries}.\\
We will use the fact that minimal surfaces have bounded diameter in the presence of a lower bound on injectivity radius. See Rubinstein \cite{Rubminimal} or Souto \cite{souto} for more on minimal surfaces in hyperbolic 3-manifolds.
\begin{lemma}\label{mindiam}
Let $S$ be a connected minimal surface in a complete hyperbolic 3-manifold $M$ with $\operatorname{inj}(M) \ge \epsilon$. Then the diameter of $S$ is at most $4|\chi(F)|/\epsilon + 2\epsilon$.
\end{lemma}
We will also use the following Lemma in the proofs of Theorem \ref{fiber} and Theorem \ref{heegaard}.
\begin{lemma}\label{limitset}
If $S$ is a closed orientable surface smoothly immersed with principal curvatures everywhere in the interval $[-1,1]$ in a complete hyperbolic 3-manifold $M$ with no cusps, then the limit set of a lift of $S$ to $\mathbb{H}^3$ is a proper subset of $\partial\mathbb{H}^3$.
\end{lemma}
\begin{proof}
Let $\tilde{S}$ be a lift of $S$ to $\mathbb{H}^3$. Assume that $\tilde{S}$ is not a horosphere, as otherwise we are done. Thus the principal curvatures of $S$ cannot be everywhere equal to $1$ or everywhere equal to $-1$.
If the principal curvatures at every point of $S$ are $-1$ and $1$, then there is a pair of line fields defined on the entire surface, implying that $S$ is a torus. Since closed surfaces in $M$ with all principal curvatures in $[-1,1]$ are incompressible and $M$ has no cusps, $S$ cannot be a torus.
Thus there is a point $p$ in $\tilde{S}$ at which one of the principal curvatures is in $(-1,1)$. Assume that the other principal curvature at $p$ is in $[-1,1)$. Let $H$ be a horosphere tangent to $\tilde{S}$ at $p$. Use an upper half space model of $\mathbb{H}^3$ in which $H$ is a horizontal plane and $\tilde{S}$ is below $H$. Let $l$ be a simple loop in $\tilde{S}$ which contains $p$ such that the principal curvatures at each point on $l$ are in $[-1,1)$ with at least principal curvature in $(-1,1)$. At each point $x$ in $l$, let $H_x$ be the horosphere above $\tilde{S}$ tangent to $\tilde{S}$ at $x$. For each $x$ in $l$, let $c_x \in \partial\mathbb{H}^3$ be the center of the horosphere $H_x$. The set of points $C = \{ c_x | x \in l \}$ forms a closed curve in $\partial\mathbb{H}^3$. Since the principal curvatures of $\tilde{S}$ are everywhere in the interval $[-1,1]$, $\tilde{S}$ cannot transversely intersect any of the horospheres $H_x$. Thus, the limit set of $\tilde{S}$ cannot cross the closed curve $C$, so that the limit set of $\tilde{S}$ is a proper subset of $\partial\mathbb{H}^3$.
\end{proof}
It is well-known that the limit set of a lift to $\mathbb{H}^3$ of a fiber $\Sigma$ in a doubly degenerate hyperbolic $\Sigma \times \mathbb{R}$ is the entire boundary $\partial\mathbb{H}^3$. By Lemma 2, such a fiber $\Sigma$ cannot be smoothly embedded with principal curvatures everywhere in the interval $[-1,1]$.
\section{Principal curvatures of fibers}
In the proof of Theorem \ref{fiber}, we will use the following well-known fact about geometric limits of hyperbolic mapping tori.
\begin{thma}\label{thma}
Let $(M_i,p_i)$ be a sequence of pairwise distinct pointed hyperbolic mapping tori with genus $g$ fibers and $\operatorname{inj}(M_i) > \epsilon$ for all $i$. Then a subsequence of $(M_i,p_i)$ converges geometrically to a pointed hyperbolic 3-manifold $(M,p)$ homeomorphic to $\Sigma \times \mathbb{R}$ where $\Sigma$ is a closed genus $g$ surface and $M$ has a doubly degenerate hyperbolic structure.
\end{thma}
\noindent\textit{Proof of Theorem \ref{fiber}.} Suppose, for contradiction, that Theorem \ref{fiber} does not hold. Then there exists a sequence of hyperbolic mapping tori $(M_i)$ with $\operatorname{inj}(M_i) > \epsilon$ such that $M_i$ has a genus $g$ minimal surface fiber with principal curvatures less than $1 + 1/i$ in absolute value. For each $i$, let $p_i$ be a point in $S_i$. By Theorem A the sequence $(M_i,p_i)$ has a subsequence, say the entire sequence, which converges to a doubly degenerate pointed hyperbolic 3-manifold $(M,p)$ homeomorphic to $\Sigma \times \mathbb{R}$ where $\Sigma$ is a genus $g$ closed surface. By Lemma \ref{mindiam}, the diameters of the surfaces $S_i$ are uniformly bounded. Thus we can find a compact subset $K$ of $M$ homeomorphic to $\Sigma \times [-1,1]$ such that for $i$ large enough, say for all $i$, $S_i$ is contained in $\kappa_i (K)$.
The surface $S := \Sigma\times\{0\}$ in $M$ is isotopic to $\kappa_i^{-1} (S_i)$ for each $i$.
Since the surfaces $\kappa_i^{-1} (S_i)$ have bounded area and curvature, a subsequence converges to a smoothly immersed surface with principal curvatures in $[-1,1]$ which is homotopic to $S$. Lemma \ref{limitset} implies that the limit set of a lift of $S$ to $\mathbb{H}^3$ is a proper subset of $\partial\mathbb{H}^3$, contradicting the fact that $M$ is doubly degenerate. \hfill$\Box$
\section{Principal curvatures of Heegaard surfaces}
In the proof of Theorem \ref{heegaard}, we will use the following well-known fact about geometric limits.
\begin{thmb}\label{thmb}
Every sequence $(M_i,p_i)$ of pointed hyperbolic 3-manifolds with $\operatorname{inj}(M_i,p_i)$ bounded away from 0 has a geometrically convergent subsequence.
\end{thmb}
We also need a Lemma from Souto (Lemma 2.1 from \cite{souto2}).
\begin{lemma}\label{sequence}
Let $(M_i)$ be a sequence of hyperbolic 3-manifolds converging to a hyperbolic manifold $M$. Assume that there is a compact subset $K\subset M$ such that for all sufficiently large $i$ the homomorphism $\pi_1(K) \rightarrow \pi_1(M_i)$ provided by geometric convergence is surjective. Then, if the cover of $M$ corresponding to the image of $\pi_1(K)$ into $\pi_1(M)$ has a convex-cocompact end, so does $M_i$ for all but finitely many $i$.
\end{lemma}
\noindent\textit{Proof of Theorem \ref{heegaard}.} Suppose for contradiction that Theorem \ref{heegaard} does not hold. Then there exists a sequence $(M_i)$ of closed hyperbolic 3-manifolds with $\operatorname{inj}(M_i) > \epsilon$ such that $M_i$ has a genus $g$ minimal Heegaard surface $S_i$ with principal curvatures less than $1 + 1/i$ in absolute value. For each $i$ let $p_i$ be a point in $S_i$. By Theorem B the sequence $(M_i,p_i)$ has a convergent subsequence, say the entire sequence, which converges geometrically to a pointed hyperbolic 3-manifold $(M,p)$. By Lemma \ref{mindiam}, the diameters of the surfaces $S_i$ are uniformly bounded. Thus each $M_i$ contains a compact subset $K_i$ homeomorphic to $S_i \times [-1,1]$ with uniformly bounded diameter. For $i$ large enough the pull-back $\kappa_i^{-1}(K_i)$ of $K_i$ through the almost isometries provided by geometric convergence are embedded compact subsets homeomorphic to $\Sigma \times [-1,1]$ where $\Sigma$ is a closed surface of genus $g$. For $i$ large enough the surfaces $\kappa_i^{-1}(S_i)$ are all isotopic to a fixed embedded genus $g$ surface $S$ in $M$.
Since the surfaces $\kappa_i^{-1} (S_i)$ have bounded area and curvature, a subsequence converges to a smoothly immersed surface with principal curvatures in $[-1,1]$ which is homotopic to $S$.
Thus the surface $S$ is incompressible in $M$ and by Lemma \ref{limitset} the limit set of a lift of $S$ to $\mathbb{H}^3$ is a proper subset of $\partial\mathbb{H}^3$.
To arrive at a contradiction we will show that the cover of $M$ corresponding to the image of $\pi_1(S)$ into $\pi_1(M)$ is doubly degenerate, implying that the limit set of a lift of $S$ to $\mathbb{H}^3$ is all of $\partial\mathbb{H}^3$. For $i$ large enough $\kappa_i(S)$ is isotopic to the Heegaard surface $S_i$ in $M_i$, so that the homomorphism $(\kappa_i)_* : \pi_1(S) \rightarrow \pi_1(M_i)$ provided by geometric convergence is surjective. By Lemma \ref{sequence}, if the cover of $M$ corresponding to the image of $\pi_1(S)$ into $\pi_1(M)$ has a convex-cocompact end, so does $M_i$ for all but finitely many $i$. Since each $M_i$ is closed we have that the cover of $M$ corresponding to the image of $\pi_1(S)$ into $\pi_1(M)$ cannot have a convex-cocompact end. Thus the cover of $M$ corresponding to the image of $\pi_1(S)$ into $\pi_1(M)$ is doubly degenerate contradicting the fact that $S$ is isotopic to a surface with principal curvatures everywhere in $[-1,1]$. \hfill$\Box$\\
\textbf{Acknowledgement.} This work was partially supported by the NSF RTG grant 0602191.
\bibliographystyle{amsalpha}
|
\section{Introduction}
\label{sec:introduction}
While string theory has had remarkable successes over the last several
years, accelerated by the revolutions in understanding its
non--perturbative properties, it is still very much the case that we
do not yet know what the theory is. We cannot state unambiguously what
the basic degrees of freedom are (it is highly context dependent in a
way that depends upon the dynamics themselves), and even the
backgrounds in which the theory propagates are themselves open to
interpretation. For example, in some descriptions and situations, the
theory contains gravity, and in others, it does not. From some
perspectives there are open strings present, and from others, only
closed. Ironically, several of these frustrating (from the point of
view of finding simple definitions) features are also among the
theory's most powerful positive traits, allow an ever--widening range
of applications of the theory to diverse problems, often of a strongly
coupled nature.
While applications continue, it is still important to try to get to
grips with what the theory is. At the very least, this is important
from a pragmatic standpoint, since perhaps a useful definition or
characterization of string theory might be put to use as, for example,
a diagnostic device in identifying when a physical problem may have
some aspect of it that is amenable to solution by string theory
methods. More generally, if string theory ultimately plays some
fundamental role in the understanding of physics beyond the standard
model, and/or in cosmology and other origins questions about the
universe at large, a more profound understanding of the nature, power,
and scope of the theory would seem to be highly desirable.
At best, to date, as a result of various dualities, we know that it is
probably part of some larger physical framework which itself is only
string theoretic in various corners of its parameter space. This
physical setting, called M--theory, remains profoundly mysterious well
over a decade after the first clear glimpses of
it\cite{Hull:1994ys,Townsend:1995kk,Witten:1995ex}.
Historically, problems pertaining to such essential matters of
understanding in physics are greatly illuminated by having a rich set
of examples that are simple, but yet complex enough to contain all the
important phenomena in question. For the problems outlined above, it
would be rather excellent to have the simplest possible string
theories that still contain some of the marvellous non-perturbative
physics we know and love, and be able to follow them as they
connect to each other in ways that are entirely invisible in
perturbation theory. Further icing on the cake would be to have the
physics all captured in terms of relatively familiar structures
for which there is an existing technology for its study.
This is the subject of this paper (and a follow--up to appear
later\cite{companion}), at least in part. The simplest known strings
with tractable non--perturbative physics that contain a rich set of
phenomena (such as holography and open--closed dualities) are the
minimal strings\cite{Gross:1989vs,Brezin:1990rb,Douglas:1989ve}, and
in particular (where non--perturbative physics is concerned) the
type~0 strings (formulated in
refs.\cite{Morris:1990bw,Dalley:1991qg,Dalley:1991vr,Johnson:1992pu,Dalley:1992br}
and refs.\cite{Crnkovic:1990ms,Hollowood:1991xq}, and recognized as
type~0 strings in ref.\cite{Klebanov:2003wg}). The non--perturbative
formulation of the type 0A and type 0B strings can be done rather
beautifully in terms of certain integrable systems, as we will review
later: Type 0A has the Korteweg--de Vries system while type 0B has
Zakharov--Shabat. The non--perturbative physics of each is formulated
in terms of associated non--linear ordinary differential equations
often called ``string equations''. While much of the language of the
two is similar, these are very different systems, and except for
various accidental (from the perspective of those separate
formulations) perturbative coincidences (and a non--trivial
non--perturbative equivalence for one model --- see later), the
physics of each are quite separate indeed.
This paper builds on all of these results, taking them much
further. We have found that there is a larger framework into which the
type 0A and type 0B string theories can be naturally embedded and
within which they are connected as parts of a larger theory. We found
further that the two string theories are merely special points in a
much larger tapestry of possibilities. When perturbation theory is
examined, other special points suggest themselves, and they turn out
to be just as ``stringy'' as the original type~0 theories, deserving
to be thought of as string theories as well. We begin the program of
trying to identify some of these theories, with some success. We also
find that the larger framework provides natural definitions of regimes
of the type~0 string theories that are hard to define using
perturbation theory, and we will report more fully on
non--perturbative aspects in a follow--up paper\cite{companion}.
In this sense, we have a precise analogue of M--theory. We have a
larger physical framework that is not itself a string theory, but that
can be readily specialized to yield string theories as special
limits. We can move between different theories in a quite natural way,
which is nonetheless outside the framework of any of its daughter
string theories. We find this encouraging and exciting.
At the base of our infinite family of string equations, organizing
much of this remarkable structure, is a non--linear differential
equation known as Painlev\'e~IV. This well--known equation from the
classical mathematics literature\footnote{See for example the lovely
monograph of ref.\cite{Noumi:2004} and references therein.}, part of
a celebrated family of six equations, has two arbitrary constants,
usually denoted $\alpha$ and $\beta$. (Actually, two copies of
Painlev\'e IV turn up in our story, intertwined in an interesting
way.) It turns out that the type 0A and 0B points in the tapestry of
theories occur at the vanishing of one or other of these constants for
one of the copies of Painlev\'e~IV. The vanishing of the constants of
the second Painlev\'e IV hint at interesting new special points.
After reviewing crucial aspects of the type~0A and~0B string theories
in section~\ref{sec:type0}, we unpack the dispersive water wave
hierarchy and present the infinite family of equations we propose as
the string equations in
section~\ref{sec:dww}. Section~\ref{sec:painleve} highlights the role
of Painlev\'e~IV. In section~\ref{sec:connectAB} we show how the
structures of section~\ref{sec:type0} arise as special points in this
larger framework, while section~\ref{sec:beyond} is a detailed study
of the rich properties of the string equations and the types of
solutions available. We organize and classify a great deal of the
physics that appears, and notice in section~\ref{sec:square} that much
of the physics can be organized in terms of a square. The square is
reminiscent of the main square organizing the moduli space of ${\hat
c}=1$ (two--dimensional) string theories, discovered in
ref.\cite{Seiberg:2005bx}, and we contemplate a possible connection,
perhaps induced by Renormalization Group
flow\cite{Gross:1990ub,Hsu:1992cm} to the ${\hat c}<1$ context of the
work in question. The possible relation between the squares helps us
make a conjecture about the nature of two new special points we find:
They might be type~IIA and~IIB minimal string theories. (Note that
these are type~II theories in the sense of the structure of the GSO
projection used to formulate them. There is no spacetime
supersymmetry\cite{Seiberg:2005bx}.) In section~\ref{sec:new} we carry
out a comparison of the new structures we found to some continuum
computations for one--loop partition functions. From this we
strengthen aspects of our type~II suggestion. We conclude in
section~\ref{sec:conclusion} with a brief summary and discussion.
\section{The $(A,A)$ Type~0 Theories: Review}
\label{sec:type0}
We will start with a brief review of the type~0 string theories
coupled to the $(2,4k)$ superconformal minimal models and show how
these theories can be elegantly described within the framework of an
integrable hierarchy of partial differential equations (PDEs)
accompanied by an hierarchy of ordinary differential equations
(ODEs). These models exhibit novel and interesting physics, all of
which and more will be seen to be embedded in the DWW system that we
will describe in the next section. This review will help establish our
notation and the framework upon which we can readily build the more
general structure.
\subsection{Type 0A Strings}
\label{sec:0A}
We begin with the following ordinary differential equation (known in
the old days as a ``string equation'')
\begin{equation}\label{streqn0A}
w\mathcal{R}^2 - \frac{1}{2}\mathcal{R}\mathcal{R}'' + \frac{1}{4} \mathcal{R}'^{2} = \nu^2 \Gamma^2 \quad .
\end{equation}
This equation (or, really, family of equations) and its properties
have been studied in several papers. It was first derived and studied
as a fully non--perturbative definition of a string theory in
refs.\cite{Dalley:1991qg,Dalley:1991vr,Johnson:1992pu,Dalley:1992br},
and evidence that it defines a type~0A string theory was presented
first in ref\cite{Klebanov:2003wg}. Further properties of the
equation, in particular concerning how branes and fluxes are encoded
by it and the underlying integrable (KdV) system, were presented in
refs.\cite{Carlisle:2005mk,Carlisle:2005wa}. Here $w(z)$ is a real
function of the real variable $z$, a prime denotes $\nu
{\partial}/{\partial z}$, and $\Gamma$ and $\nu$ are real constants.
The quantity $\mathcal R$ is defined by
\begin{equation}\label{GDpoly}
\mathcal{R} = \sum_{k = 0}^{\infty} \Big ( k + \frac{1}{2}\Big ) t_k P_k \quad ,
\end{equation}
where the $P_k[w]$ are polynomials in $w(z)$ and its $z$--derivatives,
called the {Gel'fand--Dikii} polynomials\cite{Gelfand:1975rn}. They are
related by a recursion relation (defining a recursion operator $R_2$)
\begin{equation}\label{KdVrec}
P'_{k+1} = \frac{1}{4}P'''_{k} - wP'_{k} - \frac{1}{2}w'P_{k}\equiv R_2 P'_k\ ,
\end{equation}
and fixed by the value of the constant $P_{0}$ and the requirement
that the rest vanish for vanishing $w$. Some of them are:
\begin{eqnarray}\label{0APolys}
P_{0} = \frac{1}{2}; \quad P_{1} = -\frac{1}{4} w; \quad P_{2} = \frac{1}{16}(3 w^2 - w''); \nonumber\\
P_{3} = -\frac{1}{64}(10w^3 - 10ww'' - 5(w')^2 + w'''); \cdots
\end{eqnarray}
The $k$th model is chosen by setting all the $t_j$ to zero except
$t_{0} \equiv z$ and
\begin{equation}
t_{k}=\frac{(-4)^{k+1}(k!)^2}{(2k+1)!}\ .
\end{equation}
This number is chosen so that the coefficient\footnote{This gives $w=z^{1/k} +\ldots$ as $z\rightarrow+\infty$. If we had instead chosen $t_0=-z$,
we would have chosen the coefficient of $w^k$ to be unity.} of $w^k$ in $\mathcal{R}$ is set
to $-1$.
The function $w(z)$ defines the partition function $Z = \exp (-F)$ of
the string theory $via$
\begin{equation}\label{0AFreeEnergy}
w(z) = 2 \nu^2 \frac{\partial^2 F}{\partial \mu^2}\Big{|}_{\mu = z} \quad ,
\end{equation}
where $\mu$ is the coefficient of the lowest dimension operator in the
world--sheet theory. So $w(z)$ is a two--point function of the theory.
From the point of view of the $k$th theory, all the other $t_{j}$
represent couplings of closed string operators $\mathcal{O}_j$. It is
well known\cite{Douglas, Banks} that the insertion of each operator
is captured in terms of the integrable KdV hierarchy of flows describing how $w(z,t_j)$ evolves in $t_j$:
\begin{equation}\label{KdVflow}
\frac{\partial w}{\partial t_{j}} = P'_{j+1} =R_2 P'_j\quad .
\end{equation}
For the $k$th model, equation (\ref{streqn0A}), which has remarkable
properties\cite{Carlisle:2005mk, Carlisle:2005wa}, is known to furnish
a complete non--perturbative definition of a family of spacetime
bosonic string theories \cite{Dalley:1992br}. The models are actually
type 0A strings \cite{Klebanov:2003wg}, based upon the $(2,4k)$
superconformal minimal models coupled to super--Liouville theory. As
superconformal theories, they have central charge
\begin{equation}
\label{eq:central}
{\hat c}=1-\frac{(2k-1)^2}{k}\ .
\end{equation}
The asymptotic expansions for the first two $k$ are:
\medskip
\noindent
{$k = 1$}
\begin{eqnarray}\label{0Aexpnsk=1}
w(z) &=& z + \frac{\nu \Gamma}{z^{1/2}} - \frac{\nu^2 \Gamma^2}{2 z^2} + \frac{5}{32} \frac{\nu^3}{z^{7/2}}\Gamma\left(4\Gamma^2 + 1\right) +\cdots \quad (z \rightarrow \infty) \\
w(z) &=& 0 + \frac{\nu^2 (4 \Gamma^2 - 1)}{4 z^2} + \frac{\nu^4}{8} \frac{(4\Gamma^2 - 1)(4 \Gamma^2 - 9)}{z^5}+ \cdots \quad (z \rightarrow -\infty) \nonumber
\end{eqnarray}
\noindent
{$k = 2$}
\begin{eqnarray}\label{0Aexpnsk=2}
w(z) &=& z^{1/2} + \frac{\nu \Gamma}{2 z^{3/4}} - \frac{1}{24}\frac{\nu^2}{z^2}\left(6 \Gamma^2 + 1\right) + \cdots \quad (z \rightarrow \infty) \\
w(z) &=& (4 \Gamma^2 - 1)\left(\frac{\nu^2}{4 z^2} + \frac{1}{32}\frac{\nu^6}{z^7}(4 \Gamma^2 - 9)(4 \Gamma^2 - 25) + \cdots\right) \quad (z \rightarrow -\infty) \nonumber
\end{eqnarray}
It should be noted that the solution for $z > 0$ can be numerically
and analytically shown to match onto the solution for $z<0$, providing
a unique\cite{Dalley:1991vr,Johnson:1992pu,Dalley:1992br}
non--perturbative completion of the theory. (See figure~\ref{fig:plot}
for an example of a solution found using numerical methods.)
\begin{figure}[ht]
\begin{center}
\includegraphics[width=120mm]{type0Aplot}\\
\caption{\small A plot of the $k=2$ type~0A solution showing how the
perturbative regimes at large $|z|$ are smoothly connected.
Section~\ref{sec:dww} discusses a function $v(x)$ ($x\propto z$),
which has a number of different classes of behaviour distinguished
by choice of boundary condition. The type~0A theory has class
$v_1(z)$ in the $+z$ perturbative regime and class $v_2(z)$ in the
$-z$ perturbative regime. Here we have set $\nu=1$ and
$\Gamma=0$.}\label{fig:plot}
\end{center}
\end{figure}
As instructed in equation~\reef{0AFreeEnergy}, integrating twice the
asymptotic expansions (such as those in equations~\reef{0Aexpnsk=1}
and~\reef{0Aexpnsk=2}) furnishes the free energy $F(\mu)$, and it can
be seen to define a perturbative expansion in the dimensionless string
coupling
\begin{equation}\label{0Astrcoupling}
g_{s} = \frac{\nu}{\mu^{1 + \frac{1}{2k}}} \quad .
\end{equation}
For all models, in the $\mu \rightarrow +\infty$ regime, $\Gamma$
represents\cite{Dalley:1992br,Klebanov:2003wg} the number of
background ZZ D--branes \cite{Zamolodchikov:2001ah} in the model, with
a factor of $\Gamma$ for each boundary in the worldsheet
expansion. These are point--like branes localized at infinity in the
Liouville direction $\phi$, deep in the strong coupling region. In the
$\mu \rightarrow -\infty$ regime, $\Gamma$ represents the number of
units of RR--flux in the background, with $g_s^2 \Gamma^2$ appearing
when there is an insertion of pure RR--flux
\cite{Klebanov:2003wg}. Since there is a unique non--perturbative
solution connecting the two regimes, the string equation
(\ref{streqn0A}) supplies a non-perturbative completion of the theory
that is a very clear example of a geometric transition between these
two distinct (D-branes $vs$ RR--fluxes) spacetime descriptions of the
physics.
The function $w(z)$ is the potential in the Hamiltonian ${\cal H}\equiv
-\nu^2\partial^2_z + w(z)$ of the well--known (in the inverse
scattering literature) associated Sturm--Liouville problem connected
to the integrable KdV hierarchy. The wavefunctions of that problem
define the partition functions of
FZZT\cite{Fateev:2000ik,Teschner:2000md} D--branes stretched along the
Liouville direction $\phi$, ending at a finite $\phi_c$ set by the
eigenvalue. The zero--energy problem is
interesting\cite{Carlisle:2005mk,Carlisle:2005wa}, since there the
FZZT D--branes stretch to infinity, and the Hamiltonian's
factorization, ${\cal H}=-(\nu\partial_z\pm g(z))(\nu\partial_z\mp
g(z))$ where $w(z)=g(z)^2\pm g(z)^\prime$, is highly convenient. The
function $g(z)$ (its definition in the equation before is termed a
Miura map in the integrable literature) satisfies\cite{Dalley:1992br}
an infinite hierarchy of equations sometimes called the Painlev\'e~II
hierarchy since the equation at $k=1$ is the Painlev\'e~II
equation\footnote{Those equations were derived in a string theory
context by studying unitary matrix
models\cite{Periwal:1990gf,Periwal:1990qb}. Painlev\'e~II
hierarchies have a mathematical life independent of this physical
context, however. See {\it e.g.,} refs\cite{airault,clarkson}.}. The
asymptotic expansion of $y(z)$ generated by these equations is in
terms of worldsheets involving ZZ D--branes (or fluxes) and FZZT
D--branes. The entire problem defines a toy supersymmetric quantum
mechanics problem within which the celebrated B\"acklund
transformations of the KdV system can be made manifest. In this
language the ZZ D--branes are identified with the number of threshold
bound states (formally, zero--velocity solitons) of the system, and
the B\"acklund transformations change their number by an
integer. These more recently established
features\cite{Carlisle:2005mk,Carlisle:2005wa}, together with the
earlier identification\cite{Douglas, Banks} of the role of the KdV
flows in organizing the close string operators, show how the
integrable model and inverse--scattering technology of the
mathematical physics literature comes to life in organizing the open
and closed string content of minimal string theory.
\subsection{Type 0B Strings}
\label{sec:0B}
Type 0B string theory coupled to the $(2, 4k)$ superconformal minimal
models \cite{Klebanov:2003wg} is described succinctly by the following
{string} equations\cite{Crnkovic:1990ms,Hollowood:1991xq}:
\begin{eqnarray}\label{streqn0B}
\sum_{l = 0}^{\infty} t_{l}(l + 1)R_{l} = 0 \ ,\qquad
\sum_{l = 0}^{\infty}t_{l}(l + 1)H_{l} + \nu q = 0 \ ,
\end{eqnarray}
where the $R_l$ and $H_l$ are polynomials of functions $r(x)$ and
$\omega (x)$ (and their derivatives), and~$\nu$ and $q$ are real
constants.
The differential polynomials satisfy the following recursion relations
\begin{eqnarray}\label{ZSrec}
R_{l + 1} = \omega R_{l} -\left(\frac{H_l^\prime}{r}\right)^\prime+ rH_{l}\ ,\qquad
H_{l + 1}' = \omega H_{l}' - r R_{l}' \ ,
\end{eqnarray}
where a prime denotes $\nu {\partial}/{\partial x}$. Some of them are:
\begin{eqnarray}\label{ZSpoly}
H_{-1} &=& 1, \quad R_{-1} = 0;\nonumber\\
H_{0} &=& 0, \quad R_{0} = r;\nonumber\\
H_{1} &=& -\frac{r^2}{2}, \quad R_{1} = \omega r;\nonumber\\
H_{2} &=& -r^2\omega, \quad R_{2} = -\frac{r^3}{2} + r\omega^2 + r'' ; \\
H_{3} &=& \frac{3}{8}r^4 -\frac{3}{2}r^2\omega^2 + \frac{1}{2}r'^2 - rr'' \quad ,\nonumber\\
R_{3} &=&-\frac{3}{2}r^3\omega + r\omega^3 + 3r'\omega' + 3\omega r'' + r\omega'' \quad .\nonumber
\end{eqnarray}
The function ${\widetilde w}(x) = {r^2}/{4}$ defines the partition function of the
theory $via$
\begin{equation}\label{0BFreeEnergy}
{\widetilde w}(x) = \frac{r^2}{4} = \nu^2 \frac{d^2F}{dx^2} \quad .
\end{equation}
The $n$th model is chosen by setting all $t_l$ to zero except $t_0
\sim x$ and $t_n$, analogous to what was done in the previous section
concerning the 0A case. Note that these models have an interpretation
as type~0B strings coupled to the $(2,2n)$ superconformal minimal
models only for even\footnote{For odd $n$, lack of modular invariance
of the partition function rules out the interpretation as type~0B
strings coupled to superconformal matter \cite{Klebanov:2003wg}.}
$n$. Writing $n = 2k$, we again have a set of models connected to
the $(2,4k)$ superconformal minimal models, this time type~0B.
As in the 0A case, from the point of view of the $k$th theory, all the
other $t_j$ represent coupling to closed string operators
$\mathcal{O}_j$. Again the insertion of each operator can be expressed
in terms of the Zakharov--Shabat\cite{Zakharov:1979zz} hierarchy of flows, the underlying integrable system in this case:
\begin{eqnarray}\label{0Bflow}
\frac{\partial \beta}{\partial t_k} = R_{k+1}\ ,\qquad
\frac{\partial r}{\partial t_k} = -\frac{H^{'}_{k+1}}{r} \ ,
\end{eqnarray}
where $\beta^{'} \equiv \omega$.
The asymptotic expansions of the string equations (\ref{streqn0B}) for
the first even $n=2k$ are:
\noindent
$n = 2 \,\,\, (k=1)$
\begin{eqnarray}\label{0Bexpnsm=2}
{\widetilde w}(x) &=& \frac{x}{4} + \left(q^2 - \frac{1}{4}\right)\left[\frac{\nu^2}{2x^2} + \left(q^2 - \frac{9}{4}\right)\left(\frac{-2\nu^4}{x^5} + \cdots\right)\right]\ , \quad (x \rightarrow \infty) \nonumber\\
{\widetilde w}(x) &=& \frac{\nu q\sqrt{2}}{4|x|^{1/2}} - \frac{\nu^2 q^2}{4 |x|^2} + \frac{\nu^3}{|x|^{7/2}}\frac{5\sqrt{2}}{64} q\left(1 + 4q^2\right)+\cdots \quad (x \rightarrow -\infty)
\end{eqnarray}
\noindent
$n = 4 \,\,\, (k=2)$
\begin{eqnarray}\label{0Bexpnsm=4}
{\widetilde w}(x) &=& \frac{\sqrt{x}}{4} + \frac{\nu^2}{144 x^2} \left(64q^2 - 15\right) + \cdots ; \quad (x \rightarrow \infty) \\
{\widetilde w}(x) &=& \frac{\sqrt{|x|}}{2\sqrt{14}} + \frac{\nu}{2 |x|^{3/4}}\frac{q}{\sqrt{3}\cdot7^{1/4}} + \cdots \quad (x \rightarrow -\infty) \nonumber
\end{eqnarray}
Upon integrating twice, the asymptotic expansions in equations
(\ref{0Bexpnsm=2}) and (\ref{0Bexpnsm=4}) furnish the free energy
perturbatively as an expansion in the dimensionless string coupling,
given by the same expression as before in
equation~\reef{0Astrcoupling}.
For these models, in the $\mu \rightarrow -\infty$ regime, $q$
represents the number of background ZZ D--branes in the model, with a
factor of $q$ for each boundary in the world sheet expansion, while in
the $\mu \rightarrow \infty$ regime it counts the number of units of
RR-flux in the background\cite{Klebanov:2003wg}. The asymptotic
expansions in the two directions can be argued in
ref.\cite{Klebanov:2003wg} to match onto each other analytically in a
particular ('t Hooft) limit. For the case $k=1$, the full
non--perturbative solution is known since it can be mapped directly to
the solution known for the $k=1$ type~0A case, as will be discussed below
in section~\ref{sec:connectingAB}.
For later reference, we briefly discuss the structure of these
solutions with increasing~$n$. As argued in ref.\cite{Klebanov:2003wg}
the $n=2$ expansions are deformations of the solutions of the
equation with $q=0$. However things get interesting for $n=4$. As before,
the $x>0$ solution is a deformation of the solution with
$q=0$. For $x<0$, $q=0$ allows for the trivial solution $r(x)
= 0$, but trying to deform this for $q \neq 0$ leads only to a
complex solution. Additionally, $q=0$ allows for two nontrivial solutions
with $r(x) \neq 0$ and $\omega(x) \neq 0$. These two are related by
$\omega \rightarrow -\omega$ and are interpreted as $\IZ_2$ symmetry
breaking solutions. In the interpretation of
ref.\cite{Klebanov:2003wg}, this is due to the presence of R--R
fields. Both these solutions have a real extension to the case with $q
\neq 0$. The requirement of matching a $x<0$ solution to the $x>0$
solution picks out one of these for $q > 0$ and the other for $q <
0$. (The $x < 0$ solution listed above is for $q > 0$.) For higher $n$, more
such symmetry--breaking solutions arise. We will see how this is
organized explicitly in section~\ref{sec:beyond}.
\subsection{A Non--perturbative Connection Between 0A and 0B}
\label{sec:connectingAB}
It turns out that the simplest case of $n = 2$ ($k=1$), the 0A and 0B
theories are non--perturbatively related in a very special way. In
this case the conformal model is trivial ({\it i.e.} ${\hat c}=0$)
and we simply have the pure world--sheet supergravity sector. The
strings are unencumbered by a spacetime embedding (not
counting the ubiquitous Liouville direction,~$\phi$).
The string equation for the $k=1$ 0A theory,
equation~(\ref{streqn0A}) with $\mathcal{R} = w(z) - z$, is
\begin{equation}\label{streqn0Ak=1}
w\left(w-z \right)^2 - \frac{1}{2} \nu^2 \frac{\partial^2 w}{\partial z^2} \left(w-z\right) + \frac{1}{4} \nu^2 \left(\frac{\partial w}{\partial z} - 1\right)^2 = \nu^2 \Gamma^2 \quad .
\end{equation}
On the other hand, the string equation (\ref{streqn0B}) for the $k=1$
0B theory can be written succinctly as
\begin{equation}\label{streqn0Bm=2}
\nu^{2} \frac{\partial^2 r}{\partial x^2} -\frac{1}{2} r^3 + \frac{1}{2} x r + \nu^{2} \frac{q^2}{r^3} = 0 \quad .
\end{equation}
Notice that the perturbative expansion for the $k=1$ 0A theory as $z
\rightarrow \infty$ (equation~\reef{0Aexpnsk=1}) looks similar to the
perturbative expansion for the $n=2$ 0B theory as $x \rightarrow
-\infty$ (equation~\reef{0Bexpnsm=2}) up to a (non--universal) sphere
term, once one identifies $\Gamma$ with $q$. The two expansions are
just offset by various powers of $2$. In fact there exists a
non--perturbative map between the two
equations~\cite{Morris:1990bw,Morris:1992zr} that can be seen as
follows. First define a function $f(z)$ {\it via}
\begin{equation}\label{Morrismap}
w(z) = f(z)^{2} + z \quad ,
\end{equation}
for which the string equation for the 0A theory (\ref{streqn0Ak=1})
becomes
\begin{equation}\label{0Ak1to0Bm2}
\nu^2 \partial^{2}_{z} f - f^3 - z f + \nu^2 \frac{\Gamma^2}{f^3} = 0 \quad .
\end{equation}
After rescaling using
\begin{eqnarray*}
f = 2^{-1/6} r ; \qquad z = 2^{-1/3} x \quad .
\end{eqnarray*}
equation (\ref{0Ak1to0Bm2}) becomes the string equation for the 0B
theory (\ref{streqn0Bm=2}), but with the sign of $x$ reversed. The
physics of pure 0A supergravity and pure 0B supergravity ({\it i.e.}
${\hat c}=0$) are non--perturbatively related, with the brane and flux
perturbative regions exchanged.
This non--perturbative connection between the 0A and 0B theories (for
$k=1$) follows, mathematically, from the fact that the same basic
string equation appears at the base of the two separate (KdV {\it vs}
ZS) hierarchies of equations. This connection is partly in the spirit
of a much larger set of connections that we are reporting on in this
paper. We find that the KdV and ZS structures are embedded in a much
larger structure, the dispersive water wave hierarchy of
equations, and find a class of connections (of a different sort)
between 0A and 0B for all~$k$ and see that they define two special
corners of a larger tapestry of physical theories.
\section{The Dispersive Water Wave Hierarchy}
\label{sec:dww}
The standard dispersive water wave (DWW) hierarchy\cite{Kuper}, which
will play a central role in the new physics we uncover, is a
two--component system\footnote{See refs.\cite{Broer,Kaup1,Kaup2} for
earlier studies of the properties of the dispersive water wave
equations, and we will use the notation of
refs.\cite{Kuper,Gordoa:2001}.}. It is described by:
\begin{equation}\label{DWWPDE}
\mathbf u_{t_{n}}= R^{n}\mathbf u_{x} \equiv \nu\partial_x \mathbf{L}_{n+1}[\mathbf{u}] \quad ,
\end{equation}
where $\mathbf{u}_{t_n}\equiv \partial_{t_n}\mathbf{u}$, $\mathbf{u}_x \equiv \nu\partial_x\mathbf{u}$
(note that here and in the rest of the paper, for any function $G(x)$, $G_{x}$ will denote $\nu\,
\partial G/\partial x$), and we adopt a matrix notation:
\begin{eqnarray*}
\mathbf u=\left(\begin{array}{c}
u\\
v\end{array}\right) ,
\quad \quad
\mathbf{L}_n[\mathbf{u}]=\left(\begin{array}{c}
L_n[u,v]\\
K_n[u,v]\end{array}\right) \ .
\end{eqnarray*}
Here,
\begin{equation}\label{DWWRecOp}
R\equiv \frac{1}{2}\left(\begin{array}{cc}
\partial_{x}u\partial_{x}^{-1}-\partial_{x}&2\\
2v+v_{x}\partial_{x}^{-1}&u+\partial_{x}\end{array}\right) \quad ,
\end{equation}
is the recursion operator for the DWW hierarchy. The operator $R$ can
be written as the quotient of two Hamiltonian operators $B_{1}$ and
$B_{2}$,
\begin{eqnarray*}
R = B_{2}\circ B_{1}^{-1} \quad ,
\end{eqnarray*}
where $B_{1}$ and $B_{2}$ are given by
\begin{eqnarray} \label{formofBs}
B_{2}=\frac{1}{2}\left(\begin{array}{cc}
2\partial_{x}&\partial_{x}u-\partial_{x}^{2}\\
u\partial_{x}+\partial_{x}^{2}&v\partial_{x}+\partial_{x}v\end{array}\right)\ , \qquad
B_{1}=\left(\begin{array}{cc}
0&\partial_{x}\\
\partial_{x}&0\end{array}\right) \ .
\end{eqnarray}
The $\mathbf{L}_n$ obey the recursion relation
\begin{equation}\label{Lrec}
\mathbf{L}_{n+1,x}=R\,\mathbf{L}_{n,x}\, ,
\end{equation}
which follows immediately from (\ref{DWWPDE}). The first few $L_{n}$ and $K_{n}$ are as follows:
\begin{eqnarray}\label{DWWPolys}
L_0 &=& 2; \quad K_0 = 0; \nonumber\\
L_1 &=& u; \quad K_1 = v; \nonumber\\
L_2 &=& \frac{1}{2}u^2 + v - \frac{1}{2}u_x; \quad K_2 = uv + \frac{1}{2}v_x;\nonumber\\
L_3 &=& \frac{1}{4} u^3 + \frac{3}{2} uv - \frac{3}{4} u u_x + \frac{1}{4} u_{xx};\\
K_3 &=& \frac{3}{4}u^2 v + \frac{3}{4}v^2 + \frac{3}{4} uv_x + \frac{1}{4}v_{xx} \quad .\nonumber
\end{eqnarray}
The normalization of $\mathbf{L}_0$ is chosen so as to reproduce (\ref{DWWPDE}) for $n=0$.\\
\indent The DWW hierarchy can be reduced to a one--component system by
demanding that one of the two independent functions vanish. If we set
$u(x)=0$, we actually reduce to the KdV hierarchy: two operations of
the (reduced) DWW recursion operator give
\begin{equation}\label{KdVRecOpfrmDWWRecOp}
R\circ R\equiv R^{2}=\left(\begin{array}{cc}
R_{2}&0\\
\frac{1}{4}(2v_{x}+v_{xx}\partial_{x}^{-1})&R_{2}\end{array}\right) \quad ,
\end{equation}
where $R_{2}=\frac{1}{4}(\partial_{x}^{2}+4v+2v_{x}\partial_{x}^{-1})$
is the recursion operator of the KdV hierarchy, shown in
equation~\reef{KdVrec}. Thus we obtain a reduction of the even flows
of the original hierarchy (\ref{DWWPDE}) to
\begin{equation}\label{KdVPDE}
v_{t_{2n}}= R_{2}^{n}v_{x} \ ,
\end{equation}
which is the KdV system in $-v(x)$ with independent variable $x$, and
the even times of DWW map to the times of the KdV $t_{2n}\to
t_n$. (Compare with equation~\reef{KdVflow}). \\
\indent In addition, the recursion relation~\reef{Lrec} reduces to
\begin{equation}\label{KdVrec2}
L_{2n+2,x}=R_2\circ L_{2n,x},
\end{equation}
which is exactly the KdV recursion relation~\reef{KdVrec}. The relative normalizations
are $L_0=2$ and $P_0=\frac{1}{2}$ so we conclude that $L_{2n}=4P_k$. Moreover, with $u=0$, it
is immediate that $L_{2n+1}=0$.
The other obvious reduction, $v(x)=0$, reduces the system to
the Burgers hierarchy. We do not explore if there are any string
theory consequences of that in this paper, since $v(x)$ is used to
define the partition function of our theories in all our examples.
\subsection{Scaling Reductions and New String Equations}
Integrable hierarchies of partial differential equations (PDEs) can be reduced to ordinary differential equations
(ODEs) through an additional condition on the variables. In our
context, these ODEs are sometimes to be thought of as defining string
theories; they are the ``string equations'' of a family of theories,
forming an hierarchy themselves. As outlined around
equation~\reef{KdVflow}, string equations can be thought of as
supplying the initial conditions for the partition function, and then
the PDEs of flows describe how the partition function evolves as a
function of the operators that couple to the $t_k$\cite{Douglas}. The
original example of all
this\cite{Gross:1989vs,Brezin:1990rb,Douglas:1989ve} was a hierarchy
of equations that have the Painlev\'e~I equation as the non--trivial
equation at their base, indexed by an integer $k$. They defined the
bosonic $c<1$ string theories coupled to the $(2,2k-1)$ conformal
minimal models.
It was later realized\cite{Dalley:1991vr} that another rich family of
string equations (those in equations~\reef{streqn0A}) can be obtained
by imposing certain scaling relations on the variables of the KdV
system (note however that the equations were originally
derived\cite{Morris:1990bw,Dalley:1991qg} directly from matrix model
constructions analogous to the original route). With this in mind, we
explore a similarity reduction of the DWW hierarchy, expecting to
obtain new string equations at the end of the day.
We follow the approach originally used to derive the string equations
of type~0A~\reef{streqn0A} for the KdV hierarchy
\cite{Dalley:1991vr,Dalley:1992br}. To that end, assign $v$ mass
dimension 1. The dimensions of the other terms uniquely follow from
\reef{DWWPDE} and are $[u]=\frac{1}{2}$, $[x]=-\frac{1}{2}$ and
$[t_n]=-\frac{1}{2}(n+1)$ Thus we can write down two Callan--Symanzik
equations expressing the scaling symmetry,
\begin{equation}
\begin{split}
\frac{1}{2}u+\frac{1}{2}x u_x+\sum_{n=0}^{\infty}\frac{1}{2}(n+1)t_n u_{t_n} & = 0 \\
v+\frac{1}{2}x v_x+\sum_{n=0}^{\infty}\frac{1}{2}(n+1)t_n v_{t_n} & =0 \quad .
\end{split}
\end{equation}
Using~\reef{DWWPDE} and~\reef{Lrec} we can rewrite these equations,
\begin{equation}\label{CS1}
\begin{split}
\frac{1}{2}u+\frac{1}{2}x u_x+\sum_{n=0}^{\infty}\frac{1}{2}(n+1)t_n \left(\frac{1}{2}uL_{n,x}+K_{n,x}-\frac{1}{2}L_{n,xx}+\frac{1}{2}u_x L_n \right) & = 0 \\
v+\frac{1}{2}x v_x+\sum_{n=0}^{\infty}\frac{1}{2}(n+1)t_n\left(\frac{1}{2}v_x L_n+\frac{1}{2}K_{n,xx}+\frac{1}{2}u K_{n,x}+ v L_{n,x} \right) & =0\quad.
\end{split}
\end{equation}
Defining,
\begin{equation}\label{LK}
\left(\begin{array}{c}
\mathcal{L}\\
\mathcal{K}\end{array}\right) = \sum_{n=0}^{\infty}\frac{1}{2}(n+1)t_n \mathbf{L}_n,
\end{equation}
we can rewrite~\reef{CS1},
\begin{eqnarray}
\label{CS21}
\frac{1}{2}u\mathcal{L}_x+\frac{1}{2}u_x \mathcal{L}+\mathcal{K}_x-\frac{1}{2}\mathcal{L}_{xx}&=& 0 \\
\label{CS22}
v\mathcal{L}_x+\frac{1}{2}v_x \mathcal{L}+\frac{1}{2}\mathcal{K}_{xx}+\frac{1}{2}u\mathcal{K}_x&=&0\quad,
\end{eqnarray}
where we have used the fact that we will take $t_0=x$ and the other
$t_n$ to be independent of $x$. Equation~\reef{CS21} can readily be
integrated. Moreover, solving~\reef{CS21} for $\mathcal{K}_x$ and
substituting the result into~\reef{CS22} yields an expression, which,
after multiplying by $\mathcal{L}$, can also be integrated. The
results are our new coupled string equations,
\begin{eqnarray}\label{DWWstring1}
-\frac{1}{2}\mathcal{L}_x+\frac{1}{2}u\mathcal{L}+\mathcal{K} &=& \nu c\quad \quad \quad \quad \quad \\
\label{DWWstring2}
\left(-v+\frac{1}{4}u^2+\frac{1}{2}u_x\right)\mathcal{L}^2-\frac{1}{2}\mathcal{L}\mathcal{L}_{xx}+\frac{1}{4}\mathcal{L}_x^2&=&\nu^2 \Gamma^2\quad ,\quad \quad \quad \quad \quad
\end{eqnarray}
where we have introduced two integration constants, $c$ and
$\Gamma$. We stress that the simplest possibility is for $c$ and
$\Gamma$ to be independent of $x$ \emph{and} $t_i$, though only
independence of $x$ is strictly necessary.
\\
\indent The $n$th model is chosen by setting all $t_i$ equal to zero
except for $t_0=x$ and $t_n$ which is chosen to be a numerical factor
to fix the normalization. We choose to parameterize $t_n$ as
\begin{equation}\label{tntogn}
g_n \equiv \frac{1}{\frac{1}{2}(n+1)t_n}
\end{equation}
in order to make direct contact with some recent literature which
discusses this system in a much different (mathematical)
context\cite{Gordoa:2001}.
\section{The Organizing Role of Painlev\'e IV}
\label{sec:painleve}
Let us focus on the case $n=1$, which forms the bottom of the
hierarchy of string equations from which all others follow using the
recursion relations. The string equations in this
case reduce to the Painlev\'e IV equation, an important equation from
the mathematical literature. Its appearance at the bottom of the
ladder of string equations we're presenting here is significant. Note
that the entire family of string equations can be generated from this
$n=1$ case by use of the recursion operator, and so structures at this
level will be reflected at higher $n$, even while the complexity of
the equations increases. Also notable is that this is the first time
that a role for this equation has been uncovered in this context of
non--perturbative string theory, and it takes its place alongside the
Painlev\'e~I and~II equations whose roles (mentioned earlier) have
been established in this context already. In fact, part of the
motivation that led to the discoveries upon which we report here was
the question as to the further role of the Painlev\'e equations in
such systems. Painlev\'e~IV emerged naturally as a candidate equation
to play a role and this led to our studying of the DWW system that we
found connected to Painlev\'e~IV in the literature\cite{Gordoa:2001}.
Let us see more explicitly how the equation emerges. Remarkably, it
will naturally appear in two different ways\cite{Gordoa:2001,Gordoa:2005}.
\subsection{Painlev\'e~IV: First Movement}
The string equations for $n=1$ are:
\begin{eqnarray}\label{DWWn=1}
2v - u_x + u^2 + g_1 x u = 2 \nu g_1 (c + \frac{1}{2}) \quad , \hspace{30mm} \\
\left(-v + \frac{1}{4}u^2 + \frac{1}{2}u_x\right)(u + g_1 x)^2 -\frac{1}{2}u_{xx}(u + g_1 x) + \frac{1}{4}\left(u_x + \nu g_1\right)^2 = \nu^2 g_1^2 \Gamma^2 \quad . \nonumber
\end{eqnarray}
where we have used the relation $g_1 = \frac{1}{t_1}$.
Solving the first of these for $v$ gives
\begin{equation}\label{DWWn=1v}
v = \frac{1}{2}\left(u_x - u^2 - g_1 x u + 2 \nu g_1 (c + \frac{1}{2})\right) \quad ,
\end{equation}
and substituting this into the second yields a second order ODE in $u$
which, under the change of variables
\begin{equation}
u(x) = y(x) - g_1 x \quad
\end{equation}
becomes
\begin{equation}
y_{xx} = \frac{1}{2}\frac{y_{x}^2}{y} + \frac{3}{2}y^3 -2 g_1 x y^2 + 2 \left[\left(\frac{g_1^2x^2}{4}\right) - \nu \alpha_1\right] y - \nu^2 \frac{1}{2} \frac{\beta^2_1}{y} \quad .
\end{equation}
$\alpha_1$ and $\beta_1$ are constants related to $c$ and $\Gamma$ in
the DWW string equations through
\begin{eqnarray}\label{alphabetaDefn}
\alpha_{1} = g_{1}(c + \frac{1}{2})\ , \qquad
\beta_{1} = \pm 2 g_{1} \Gamma \ ,
\end{eqnarray}
Setting $g_1 = -2$, and dropping the subscripts on the constants, gives
\begin{eqnarray}
\label{eq:painleve}
y_{xx} = \frac{1}{2}\frac{y_{x}^2}{y} + \frac{3}{2}y^3 +4 x y^2 + 2 \left(x^2 - \nu \alpha\right) y - \nu^2 \frac{1}{2} \frac{\beta^2}{y} \ ,
\end{eqnarray}
which is the fourth Painlev\'e equation $P_{IV}$ in standard form, and
\begin{equation}
\alpha=-(2c+1)\ ,\quad \beta=-8\Gamma^2\ . \label{eq:painleveconstants1}
\end{equation}
We will see in the next section, specific constraints yielding the 0A and 0B
theories that require $c = -\frac{1}{2} $ and $\Gamma = 0$ respectively. Notably,
these are precisely the values for which the constants $\alpha$ and
$\beta$ in the standard form of Painlev\'e~IV vanish.
\subsection{Painlev\'e~IV: Second Movement}
In fact, there is another natural appearance of the Painlev\'e~IV
equation in this system\cite{Gordoa:2001,Gordoa:2005}, at $n=1$. There is a
natural generalization\cite{Kuper} of the Miura map (that we saw for
KdV in section~\ref{sec:type0}) to the DWW system, defining
new variables $U$ and $V$:
\begin{equation}
u=U\ , \qquad v=UV-V^2+V'\ .
\end{equation}
Now, as we saw above, $y(x)= u(x)-2x = U(x)-2x$ satisfies
Painlev\'e~IV with constants $\alpha$ and $\beta$ given in
equation~\reef{eq:painleve}. Well, additionally, $-V(x)$ satisfies a
copy of Painlev\'e~IV (equation~\reef{eq:painleve}) also, but with
constants related to our physical parameters by
\begin{equation}
\label{eq:painleveconstants2}
\alpha=\mp 3\Gamma+c+1\ ,\quad \beta=(c\pm\Gamma)^2\ .
\end{equation}
We take this seriously, not the least because the variables described
by the Miura map in the case of KdV (type~0A) were seen to be
physically very natural, pertaining as they do to the FZZT and ZZ
D--branes. (See the end of section~\ref{sec:0A} for a brief review.)
We expect therefore (but this needs more exploration) that this DWW
Miura map also leads to rich physics. The cases $c=\pm\Gamma$ and
$c=-1\pm3\Gamma$ imply vanishing of $\alpha$ and $\beta$ and may well
have some special significance in this context. (We will, for example,
find special solutions for all $n$ corresponding to $c=\pm\Gamma$
points. It is also interesting to note that the equations together
point to the values $c=\pm\frac12$, $\Gamma=\pm\frac12$, values which
do feature prominently in what is to follow.)
\section{Connecting the Type~0 String Theories}
\label{sec:connectAB}
Having introduced the DWW hierarchy, we now show how {both} the
type~0A and type~0B string theories coupled to the $(2,4k)$
superconformal minimal models can be found embedded in this system. We
show that by placing appropriate constraints on the full system of
string equations, one can recover the respective string equations for
both of the type~0 theories. Quite beautifully, these constraints require one of the
two integration constants $(c,\Gamma)$ to freeze to particular
values. The remaining unfixed constant then acts as the parameter that
counts the number of ZZ--branes or units of R--R flux in each theory,
depending on which asymptotic region (positive or negative large $x$)
is under consideration.
\subsection{Reduction to 0A}
It was seen in section~\reef{sec:dww} that setting $u$=$0$ reduces the DWW
hierarchy to the KdV hierarchy. We therefore expect that this constraint also
reduces our new string equations to the~0A string equations. That this indeed
occurs can be seen as follows. Equation~\reef{LK} gives,
\begin{equation}
\begin{split}
\mathcal{L}[u\mathrm{=}0,v\mathrm{=}-w]&=\sum_{n=0}^{\infty}\frac{1}{2}(n+1)t_n^{\mathrm{DWW}}L_n \\
&=\sum_{n=0}^{\infty}\frac{1}{2}(2n+1)t_{2n}^{\mathrm{DWW}}L_{2n} \\
&=4\sum_{j=0}^{\infty}(j+\frac{1}{2})t_{2j}^{\mathrm{DWW}}P_{j} \\
&=\mathcal{R}[w]
\end{split}
\end{equation}
where we have used that $L_{2n+1}[u$=$0,v]=0$ and $L_{2n}[u$=$0,v$=$-w]=4P_n[w]$ (see below~\reef{KdVrec2}). The last equality holds provided that we make the identification,
\begin{equation}\label{tn}
t_{2n}^{DWW}= \frac{1}{4}t_n^{\mathrm{KdV}}=\frac{(-1)^{n+1}4^n(n!)^2}{(2n+1)!}
\; \Rightarrow \;g_{2n} = 2\frac{(-1)^{n+1}(2n)!}{4^n(n!)^2}\quad .
\end{equation}
Finally, we see that when $u=0$ and $v=-w$, equation~\reef{DWWstring2}
exactly reduces to equation~\reef{streqn0A}, \emph{i.e.} our new
string equations encode 0A string theory coupled to the $(2,4k)$
superconformal minimal models. For even flows it is easy to show that
$u=0$ is only consistent with the other string
equation~\reef{DWWstring1} if $c$ is frozen:
\begin{equation}
c = -\frac{1}{2} \ .
\end{equation}
So one of the parameters in the original DWW equations becomes fixed
when recovering type~0A coupled to the $(2,4k)$ superconformal minimal
models, while the other parameter~$\Gamma$ counts the number of branes
or units of RR--flux in the type 0A theory. We will see this behaviour
again in the case that we recover the type~0B theory.
\subsection{Reduction to 0B}\label{sec:0Breduc}
Consider the following redefinition of the DWW variables $\{u(x), v(x), x\}$ to the ZS
variables $\{r(y), \omega(y), y\}$:
\begin{eqnarray}\label{DWWto0B}
y=2x,\quad
u(y)=2\left(\omega(y) - \frac{r_{y}}{r(y)}\right), \quad
v(y)= -r^2(y) \ .
\end{eqnarray}
The recursion relation~\reef{Lrec} becomes,
\begin{equation}
\left(\begin{array}{c}
L_{n+1}\\
K_{n+1,y}\end{array}\right) =
\left(\begin{array}{c}
(\omega-\frac{r_y}{r})L_n-L_{n,y}+K_{n,y}\\
-r(rL_n-\frac{K_{n,y}}{r})_y+\omega K_{n,y}\end{array}\right),
\end{equation}
or,
\begin{equation}
\left(\begin{array}{c}
R_{n+1}\\
H_{n+1,y}\end{array}\right) =
\left(\begin{array}{c}
\omega R_n-\left(\frac{H_{n,y}}{r}\right)_y+rH_{n,y}\\
-r R_{n,y}+\omega H_{n,y}\end{array}\right),
\end{equation}
where we have defined,
\begin{equation}
R_n =\frac{1}{2}\left( r L_n-\frac{K_{n,y}}{r}\right),\quad\quad H_n =\frac{1}{2}K_n.
\end{equation}
These are precisely the recursion relations of the ZS
hierarchy~\reef{ZSrec}. Moreover, the $H_n$ and $R_n$ just defined
actually agree with those presented in~\reef{ZSpoly}. It suffices to
check $n=0$: from~\reef{DWWPolys} we have $L_0=2$ and $K_0=0$ which
imply $H_0=0$ and $R_0=r$, as expected.
\\
\indent Finally, we may ask how we can produce the~0B string
equations~\reef{streqn0B} from our new string
equations~\reef{DWWstring1} and~\reef{DWWstring2}. The answer turns
out to be simple and elegant: all we must do is set
\begin{equation}
\mathcal{L}=0\quad .
\end{equation}
Equation~\reef{DWWstring1} then requires,
\begin{equation}
\sum_{n=0}^{\infty}t_n(n+1)H_n-\nu c=0\quad,
\end{equation}
which further implies,
\begin{equation}
\sum_{n=0}^{\infty}t_n(n+1)R_n = 0\quad .
\end{equation}
The $t_n$ required here to consistently produce the equations of \cite{Klebanov:2003wg} are identical to the values we determined earlier~\reef{tn}. So, upon identifying $c=-q$, we have exactly produced the~0B string equations. \\
\indent Again notice how the consistency of our constraint
$\mathcal{L}=0$ with~\reef{DWWstring2} forces one of our parameters to
vanish
\begin{equation}
\Gamma = 0 \quad ,
\end{equation}
leaving the parameter $c = -q$ to count the number of ZZ branes or
RR--fluxes in the type 0B theory. Finally, we remark that the partition function
of the~0B theory is determined \emph{via}
\begin{equation}
\begin{split}
F &= \frac{1}{\nu^2}\int d^2y\; \frac{r(y)^2}{4}\\
&=-\frac{1}{\nu^2}\int d^2x\; v(x)
\end{split}
\end{equation}
so that $-v(x)$ encodes the partition function for both~0A and~0B.
\section{DWW Unconstrained --- Beyond the Familiar}
\label{sec:beyond}
We have seen that constraining the DWW string
equations~\reef{DWWstring1} appropriately leads to the 0A and 0B
theories (coupled to the superconformal $(2,4k)$ series),
respectively. The constraints take the system with two free parameters
$(c,\Gamma)$ and define two special points: 0A with $(c=-1/2, \Gamma$ free)
and 0B with $(c$ free, $\Gamma{=}0)$. We can also consider the fully
unconstrained system with both parameters $(c,\Gamma)$ unfixed, and
general $\{v(x), u(x)\}$. Interestingly, we get {\it multiple}
asymptotic expansions for the variable $v(x)$. The structure of the
equations and the corresponding asymptotic expansions gets richer as
$n$ increases. Since in both cases, asymptotic expansions of $v(x)$
gave us, upon integrating twice, an expansion of a partition function
for a string theory, we look again for it to define an interesting
partition function in the new cases we will encounter. While this is
an assumption, we shall see it bear fruit presently.
In this section we will first list the asymptotic expansions for the
first few $n$ obtained from the corresponding string equations. We
will explain the organizational rules we use to group these
expansions into various classes. A careful analysis of the patterns
we uncover in what follows allows us to extrapolate to higher $n$ and
predict the structure of the expansions for any $n$. We will see that a
subset of these, when appropriately combined, reproduce the type~0
expansions that we have already encountered. In addition, we obtain
{\it completely new} expansions which have not been presented in the
literature before. Our key observation here is that these also
resemble perturbative sectors of string theories (either with branes
or fluxes present). We take these seriously as new string theories and
our task after this section will be to identify what string theories
they might be.
The number of expansions grows large as $n$ increases (we will see
later that the number of expansions is $(n+1)^2$). To deal with this
proliferation of expansions, we classify them into classes whose
members are related to one another by simple symmetries. The classes
themselves are distinguished by a number of salient features, many of
which we explore in what follows. We choose to define the classes
based on their behavior at order $\nu^0$ (this is equivalent to the
leading behavior in $g_s^{-2}$, the sphere level of closed string
perturbation theory, as we will see later). Since DWW is a two
component system, we must consider the leading behavior of both
functions, $u$ and $v$. We adopt the following
classification scheme:\\
\begin{equation}\label{classes}
\begin{split}
\textrm{Class 1:}\quad&u_1 \sim 0,\quad\quad\; v_1\sim x^{2/n}\\
\textrm{Class 2:}\quad&u_2 \sim 0,\quad\quad\; v_2\sim 0\\
\textrm{Class 3:}\quad&u_3 \sim x^{1/n},\quad v_3\sim 0\\
\textrm{Class 4:}\quad&u_4 \sim x^{1/n},\quad v_4\sim x^{2/n},\quad u_4^2/v_4 \sim 1/4\\
\textrm{Class 5:}\quad&u_5 \sim x^{1/n},\quad v_5\sim x^{2/n},\quad u_5^2/v_5 \sim a\neq 1/4\\
\end{split}
\end{equation}
\\We postpone the detailed study of $u$ to subsequent work. Here we
mention its leading behavior only to complete the classification; in
what follows, we focus exclusively on $v$ and its asymptotic
expansions.
The details of sections~\ref{sec:expandone}--\ref{sec:expandfour},
being a list of examples that we found instructive, might be a little
dry on first reading and so the reader is encouraged to skip to
section~\ref{sec:general} for the general case.
\subsection{$n = 1$}
\label{sec:expandone}
The string equations for this case were already written in
equations~\reef{DWWn=1}. Solving the first of these for $v$ gave
equation~\reef{DWWn=1v}, and substituting into the second yields a
scalar second order ODE in $u(x)$ (equivalent to Painlev\'e~IV), which
can be used to produce the expansions. Asymptotic expansions for $u(x)$
can then be used to yield asymptotic expansions for $v(x)$ using
equation~\reef{DWWn=1v}.
We obtain three classes of expansions for $v(x)$,
\begin{eqnarray}\label{DWWn=1expn}
v_2&=&\frac{\nu^2}{x^2}(c^2-\Gamma^2)\left(1-\frac{\nu}{ g_1 x^2}6c+\frac{\nu^2}{g_1^2 x^4}(45c^2-5\Gamma^2+5)-\cdots\right) \quad ,\nonumber\\
v_3&=&\nu(c-\Gamma)\left(1-\frac{\nu}{g_1 x^2}2\Gamma-\frac{\nu^2}{g_1^2 x^4}6\Gamma(c-3\Gamma)-\cdots\right) \quad ,\\
v_4&=&\frac{1}{9}g_1^2 x^2+\nu\frac{2 g_1 c}{3}-\frac{\nu^2}{x^2}\frac{1}{3}(3c^2+9\Gamma^2-1)+\frac{\nu^3}{g_1
x^4}6c(c^2-9\Gamma^2) - \cdots \quad .\nonumber
\end{eqnarray}
Upon integrating twice (following what we learned from the type~0
theories in sections~\ref{sec:0A} and~\ref{sec:0B}), one can obtain the
free energy for a genus expansion of a
string theory which allows us to identify the string coupling to be
$g_{s} = {\nu}/{x^2}$.
\subsubsection{Symmetries for $n=1$}
The other expansions within each class can be obtained by the following
symmetry operation:
\begin{eqnarray*}
f_1: \Gamma \to -\Gamma \quad .
\end{eqnarray*}
Since $v_2$ and $v_4$ contain only even powers of $\Gamma$, the are
invariant under this map; however, $f_1\circ~v_3~\neq~v_3$. Thus there are
two expansions in the $v_3$ class, and, together with $v_2$ and $v_4$,
these comprise {four} total $n=1$ expansions.
\subsection{$n = 2$}
The string equations are:
\begin{eqnarray}\label{DWWn=2}
u_{xx} &=& 3u u_x - u^3 -6u v -2g_2 x u + 4 \nu g_2\left(c + \frac{1}{2}\right) \quad ,\nonumber\\
v_{xx} &=& 2\left( \frac {(u v + \frac{1}{2}v_x - \nu g_2 c)^2 - \nu^2 {g_2}^2 {\Gamma}^2}{v+ \frac{1}{2}u^2 -\frac{1}{2}u_x + g_2 x}\right) \\
&&- 2v \left(v+\frac{1}{2}u^2-\frac{1}{2}u_x + g_2 x\right)- 2(u v)_x \quad . \nonumber
\end{eqnarray}
where again $g_2 = \frac{1}{t_2}$. Solving the first of these for $v$ gives,
\begin{equation}\label{DWWn=2v}
v = \frac{1}{6 u}\left(u_{xx} -3u u_x + u^3 + 2g_2 x u -4 \nu g_2 \left(c + \frac{1}{2}\right)\right) \quad .
\end{equation}
Substituting this into the second yields a scalar fourth order ODE in
$u$, which can then be used to produce the expansions for $v$. In this
case, there are four relevant classes of expansions
\begin{eqnarray}\label{DWWn=2expn}
v_1&=&-g_2x-\frac{\nu g_2^{1/2}}{x^{1/2}}\Gamma+\frac{\nu^2}{x^2}\frac{1}{8}\left(-4c^2+4\Gamma^2+1\right) + \cdots \quad ,\nonumber\\
v_2&=&\frac{\nu^2}{x^2}\left(c^2-\Gamma^2\right)\left(1-\frac{2\nu^2}{g_2 x^3}(5c^2-\Gamma^2+1)+ \cdots \right) \quad ,\\
v_3&=&\frac{g_2^{1/2}\nu}{x^{1/2}}(c-\Gamma)\left(\frac{i}{\sqrt{2}}+\frac{\nu}{g_2^{1/2}x^{3/2}}\frac{1}{4}(c-5\Gamma)- \cdots\right) \quad ,\nonumber\\
v_4&=&-\frac{g_2x}{5}+\frac{\nu g_2^{1/2}}{x^{1/2}}\frac{ic}{\sqrt{5}}-\frac{\nu^2}{ x^2}\frac{1}{4}(2c^2+10\Gamma^2-1) -\cdots \quad .\nonumber
\end{eqnarray}
Here, we see that the string coupling is $g_{s} = {\nu}/{x^{\frac{3}{2}}}$.
\subsubsection{Symmetries for $n=2$} The other expansions within each class can
be obtained by the following operations,
\begin{eqnarray*}
f_1:\Gamma\to-\Gamma \ ,\qquad
f_2:c\to-c \ ,
\end{eqnarray*}
and compositions thereof. Some of the $v_i$ are invariant under one
or both of these maps.
Altogether, there are {nine} distinct expansions. Here are the
four classes of expansions together with the number of distinct
expansions within each class and the maps that lead to them:
\begin{eqnarray*}
v_1(2) : \{1, f_1\}; \quad v_2(1): \{1\}; \quad v_3(4):\{1, f_1, f_2, f_1\circ f_2\}; \quad v_4(2):\{1, f_2\} \quad .
\end{eqnarray*}
\subsection{$n=3$}
The string equations are too long to be written down explicitly here
and so we omit them. Four classes of expansions are produced in this
case, and one sees expansions in Class 5 appearing for the first time here.
\begin{eqnarray}
v_2 &=&\frac{\nu^2}{x^2}(c^2-\Gamma^2)\left(1-\frac{\nu^3}{g_3 x^4}\frac{5}{2}c(7c^2-3\Gamma^2+5) + \cdots \right) \quad , \nonumber\\
v_3&=&\frac{\nu}{x^{2/3}}(c-\Gamma)\left(\frac{(-2)^{2/3} g_3^{1/3}}{3}+\frac{\nu}{x^{4/3}}\frac{1}{3}(c-3\Gamma)- \cdots\right) \quad , \\
v_4&=&\frac{2\cdot 2^{1/3}}{35^{2/3}}g_3^{2/3}x^{2/3}+\frac{\nu g_3^{1/3}} {x^{2/3}}\frac{2\cdot2^{2/3} c}{3\cdot35^{1/3}} - \frac{\nu^2}{ x^2}\frac{1}{9}(3c^2+21\Gamma^2-2) + \cdots \quad , \nonumber\\
v_5&=&-\frac{2\cdot2^{1/3}}{5^{2/3}}g_3^{2/3}x^{2/3}-\frac{\nu g_3^{1/3}}{x^{2/3}}\frac{2^{2/3}}{3\cdot5^{1/3}}(c+\sqrt{5}\Gamma)-\frac{\nu^2}{x^2}\frac{1}{9}(3c^2-3\Gamma^2-1) + \cdots \quad .\nonumber
\end{eqnarray}
Here the string coupling turns out to be $g_{s} = {\nu}/{x^{\frac{4}{3}}}.$
\subsubsection{Symmetries for $n=3$}
The other expansions within each class can
be obtained by the following operations,
\begin{equation}
f_1:\Gamma\to-\Gamma \ , \quad f_3:x\to-x \ , \quad f_4:g_3\to-g_3\ ,
\end{equation}
and any compositions of these maps.
A quick calculation shows that there 16 different expansions
\begin{eqnarray*}
v_2(1)&:&\{1\} ; \nonumber\\
v_3(6)&:& \{1, f_1, f_3=f_4, f_1\circ f_3 = f_1\circ f_4, f_1\circ f_3 \circ f_4, f_3 \circ f_4\}; \nonumber\\
v_4(3)&:& \{1, f_3=f_4, f_3\circ f_4\}; \nonumber\\
v_5(6)&:& \{1, f_1, f_3=f_4, f_1\circ f_3 = f_1\circ f_4, f_1\circ f_3 \circ f_4, f_3 \circ f_4\} \quad .\nonumber
\end{eqnarray*}
\subsection{$n = 4$}
\label{sec:expandfour}
Our explicit string equations are rather complicated and so we will
not list them here. The following five classes of expansions are
obtained:
\begin{eqnarray}\label{DWWn=4expn}
v_1(x) &=& -\frac{2 i }{\sqrt{3}}\sqrt{g_4}\sqrt{x} - \frac{\nu g_4^{1/4}}{x^{3/4}} \frac{(1+i)\Gamma }{2 \cdot 3^{1/4}} -\frac{\nu^2}{x^2}\frac{1}{48}\left(12 c^2-12\Gamma^2-5\right) + \cdots \quad ,\nonumber \\
v_2(x) &=&-\frac{\nu^2}{x^2}(\Gamma^2-c^2)\left(1-\frac{3\nu^4}{2 g_4 x^5}(21c^4-14 c^2 \Gamma^2 + \Gamma^4 + 35 c^2 - 5\Gamma^2 + 4)\right) + \cdots \quad ,\nonumber\\
v_3(x) &=& -\frac{g_4^{1/4} \nu}{x^{3/4}}(c-\Gamma)\left(\frac{1}{2^{3/4}(1+i)}+\frac{\nu}{g_4^{1/4} x^{5/4}}\frac{7\Gamma-3c}{8} + \cdots \right) \quad ,\\
v_4(x)&=&-\frac{2i}{3\sqrt{7}}\sqrt{g_4}\sqrt{x}-\frac{g_4^{1/4}\nu}{x^{3/4}}\frac{c}{\sqrt{3}\cdot7^{1/4}(1+i)}-\frac{\nu^2}{ x^2}\frac{1}{24}\left(6c^2+54\Gamma^2-5\right) + \cdots \quad ,\nonumber\\
v_5(x)&=&-2\sqrt{\frac{2}{21}}\sqrt{g_4}\sqrt{x}-\frac{g_4^{1/4}\nu}{x^{3/4}}\frac{21^{1/4}}{2^{5/4}}\left(-\frac{c}{\sqrt{7}}+\frac{\Gamma}{\sqrt{3}}\right)-\frac{\nu^2}{ x^2}\frac{1}{48}\left(12c^2-12\Gamma^2-5\right) + \cdots\nonumber \quad .
\end{eqnarray}
Here, the string coupling is $g_{s} = {\nu}/{x^{\frac{5}{4}}}$.
\subsubsection{Symmetries for $n=4$}
The other expansions within each class can
be obtained by the following operations,
\begin{equation}
f_1: \Gamma \to -\Gamma \ , \quad f_2:c \to -c \ , \quad f_{3,4}: (x, g_4) \to (-x, -g_4) \ ,
\end{equation}
and any arbitrary composition of those maps.
A quick calculation shows that there are 25 distinct
expansions. Here are the five classes of expansions together with the
number of distinct expansions within each class and the maps that lead
to them:
\begin{eqnarray*}
v_1(4) &:& \{1, f_1, f_{3,4}, f_1 \circ f_{3,4}\}; \nonumber\\
\quad v_2(1) &:& \{1\}; \nonumber\\
\quad v_3(8) &:& \{1, f_1, f_2, f_1\circ f_2, f_{3,4}, f_1\circ f_{3,4}, f_2 \circ f_{3,4}, f_1\circ f_2 \circ f_{3,4}\}; \nonumber\\
v_4(4) &:& \{1, f_2, f_{3,4}, f_2 \circ f_{3.4}\}; \nonumber\\
\quad v_5(8) &:& \{1, f_1, f_2, f_1\circ f_2, f_{3,4}, f_1\circ f_{3,4}, f_2 \circ f_{3,4}, f_1\circ f_2 \circ f_{3,4}\} \quad . \nonumber
\end{eqnarray*}
\subsection{Patterns and Asymptotia}
\label{sec:general}
The expansions displayed above for $n=1$ to $n=4$ exhibit a rich
structure which we explore shortly. First we briefly review the
interpretation given to the parameter $\Gamma$ of the~0A theory (and
also to $q$ of the~0B theory). In the $\mu$ (or $x$) $\rightarrow
+\infty$ regime, $\Gamma$ represents the number of background ZZ
D--branes in the model, with a factor of $\Gamma$ for each boundary in
the worldsheet expansion. Since an orientable surface with odd (even)
Euler characteristic must contain an odd (even) number of boundaries,
$\Gamma$ must be raised to an odd (even) power if $g_s$ is. In
addition, the power of $\Gamma$ must be less than or equal to the
power of $g_s$. On the other hand, in the $\mu$ (or $x$)~$\rightarrow
-\infty$ regime, $\Gamma$ represents the number of units of RR--flux
in the background, with $g_s^2 \Gamma^2$ appearing when there is an
insertion of pure RR--flux. So in this case both $\Gamma$ and~$g_s$
should appear with even powers.
\\
\indent In applying these observations to our DWW expansions, we
immediately notice the remarkable fact that the various expansions
have powers of the parameters which somehow allow for interpretations
as counting branes or fluxes. This is by no means guaranteed, and
indeed its occurrence was one of our main motivations for in--depth
study of the system. The presence of two parameters, however, leads to
a few subtleties. For example, in some expansions an interpretation in
terms of branes is only possible if one of the two parameters is set
to zero. With keep such observations in mind as we begin the study of
the various expansions.
Finally we note that the asymptotic direction ({\it i.e.,} positive or
negative $x$) of each expansion can be fixed by requiring that once we
fix the value of $g_n$, the expansion must be real (which is an
important constraint since $v$ encodes the free energy). The value of
$g_n$, in turn, can be fixed using the values listed in
equation~(\ref{tn}) since we must reproduce the 0A theory. With all of
these observations we are ready to begin analyzing our expansions.
\subsubsection{Class 1}
\begin{itemize}
\item $v_1$ contains powers of $\Gamma$ consistent with those of a parameter
counting branes. This remains true for any value of $c$.
\item $v_1$ contains powers of $c$ consistent with those of a parameter
counting fluxes. However, for arbitrary $\Gamma$, $g_s$ appears with
odd powers, inconsistent with our requirements for a description of
fluxes, as mentioned above. This problem is avoided if we set
$\Gamma=0$ since this forces the odd powers of $g_s$ to vanish.
\item In particular, setting $\Gamma = 0$ with $g_n$ given by
equation~\reef{tn} reduces this expansion to the $x>0$
flux-expansion in $c$ (or $q$) for the type~0B theory seen in
equations~(\ref{0Bexpnsm=2},\ref{0Bexpnsm=4}) for $n = 2,4$
respectively.
\item Alternatively, setting $c = -\frac{1}{2}$ with the same value of
$g_n$ reduces these expansions to those of the type~0A for
$x>0$. (We listed them in equations~(\ref{0Aexpnsk=1} and
\ref{0Aexpnsk=2}) for $k=1,2$ respectively\footnote{Recall that the
DWW hierarchy index $n$ is related to KdV hierarchy index $k$ by
$n=2k$.}.)
\item With the values of $g_n$ need to reduce to~0A and~0B, one
obtains real expansions in this class {\it only} if $x > 0$. Hence
we fix this class of expansions to be $x \rightarrow +\infty$
asymptotic expansions.
\item This class of expansions only exists for even $n$.
\end{itemize}
\subsubsection{Class 2}
\begin{itemize}
\item For even $n$, $v_2$ contains powers of $\Gamma$ and $g_s$
consistent with those of a parameter counting fluxes. This is true
for all values of $c$. For odd $n$, the powers of $\Gamma$ are still
consistent with the flux interpretation, but there are odd powers of
$g_s$ which are inconsistent with fluxes. These odd powers can be
removed by setting $c=0$.
\item For even $n$, $v_2$ contains powers of $c$ and $g_s$ consistent
with a parameter counting fluxes. This is true for all values of
$\Gamma$. For odd $n$, the powers of $c$ are consistent with those
of a parameter counting branes. In this interpretation, there are no
contributions from surfaces with only one boundary.
\item Setting $c =-\frac{1}{2}$ with $g_n$ as chosen in
equation~\reef{tn}, reduces the $v_2$ expansions to the type~0A
expansions for $x<0$. (We listed them in equations~\reef{0Aexpnsk=1}
and \reef{0Aexpnsk=2} for $k=1,2$ respectively.)
\item Only even powers of $g_n$ and $x$ appear, so the
requirement of reality does not fix the direction of these
expansions.
\item Consistency with the 0A expansions forces us to consider
the expansions in this class as $x \rightarrow -\infty$
expansions. We note the possibility
that these might appear as $x \rightarrow +\infty$ expansions
outside of the simple type~0A context we've seen so far\footnote{In
fact, we can already think of an example. There are rational
solutions of the type~0A string equations that were considered in
a string theory context in ref.\cite{Johnson:2006ux}. The rational
solutions have $v_2$ type expansions (for $c=-1/2$) in both
asymptotic directions for $x$. Clearly there are analogous
rational solutions for the full DWW equations that have $v_2$
asymptotic expansions that generalize the known cases. We have
constructed large families of them, and leave their study for a
later publication.}.
\item These expansions vanish when $c^2
= \Gamma^2$. (This is a likely special point(s) in parameter space. We got a
first hint of this point in section~\ref{sec:painleve} where the
second copy of Painlev\'e~IV has $\beta=0$.)
\item This class of expansions appears for all $n$.
\end{itemize}
\subsubsection{Class 3}
\begin{itemize}
\item $v_3$ contains powers of $\Gamma$ consistent with those of a
parameter counting branes. This is true only for $c=0$.
\item $v_3$ contains powers of $c$ consistent with those of a
parameter counting branes. This is true only for $\Gamma=0$.
\item We notice that associating one boundary to each factor of $c$
\emph{and} $\Gamma$ also produces a consistent worldsheet
expansion. We might speculate about whether, in general, these
expansions might capture $c$ and $\Gamma$ simultaneously counting
branes.
\item Setting $\Gamma = 0$ with $g_2 = 1$ reduces these expansions to
the $x<0$ brane-expansions for the type~0B theory for $n=2$ as seen
in equation~(\ref{0Bexpnsm=2}). Hence we fix the expansions in this
class to be $x \rightarrow -\infty$ expansions.
\item At $n=4$, the value $g_4 = -\frac{3}{4}$ with $\Gamma = 0$
renders this expansion complex for $x < 0$. This fits in nicely with
the structure of expansions observed in the 0B case,
reviewed\footnote{Recall that the trivial solution with $r(x) = 0$
in that case did not have a real deformation for $q \neq 0$. The
$v_3$ class is exactly the analogue of this trivial solution.} in
section~\ref{sec:0B}.
\item $c = \Gamma$ causes these expansions to vanish (we got a first hint of this
point in section~\ref{sec:painleve} where the second copy of
Painlev\'e~IV has $\beta=0$.)
\item This class of expansions exists for all $n$, but is real as an
$x<0$ expansion only for $n=2$~mod~$4$.
\end{itemize}
\subsubsection{Class 4}
\begin{itemize}
\item The $v_4$ class of expansions has not, to our knowledge, made a
previous appearance in the literature, as it does not appear until
encountering the DWW system.
\item $v_4$ contains powers of $c$ consistent with those of a
parameter counting branes. This remains true for any value of
$\Gamma$.
\item $v_4$ contains powers of $\Gamma$ consistent with those of a
parameter counting fluxes. However, for arbitrary $c$, $g_s$ appears
with odd powers, inconsistent with our requirements for a
description of fluxes, as mentioned above. This problem is avoided
if we set $c=0$ since this forces the odd powers of $g_s$ to vanish.
\item The direction of $v_4$ is not immediately determined by the
consistency conditions we have used so far. Compatibility with the
type~0 theories at $n=2$ requires $g_2=1$ which renders $v_4$ real
for $x \rightarrow -\infty$. On the other hand, compatibility with
the type~0 theories at $n=4$ requires $g_4=-\frac{3}{4}$ which
renders $v_4$ real for $x \rightarrow +\infty$.
\item We will later provide evidence in favor of $v_4$ existing for $x
> 0$.
\item In general, for the type~0 choices~\reef{tn} for $g_n$, $v_4$
remains real for $x>0$ when $n = 0$ mod $4$ and becomes complex when
$n = 2$ mod $4$.
\item In the special case $n=2$, $v_4$ can be made real by setting $c=0$.
\end{itemize}
\subsubsection{Class 5}
\begin{itemize}
\item $v_5$ contains powers of $\Gamma$ consistent with those of a parameter
counting branes. This is true only for $c=0$.
\item $v_5$ contains powers of $c$ consistent with those of a parameter
counting branes. This is true only for $\Gamma=0$.
\item As for $v_3$, we notice that associating one boundary to each
factor of $c$ \emph{and} $\Gamma$ also produces a consistent
worldsheet expansion. We might speculate that, in general, these
expansions might capture $c$ and $\Gamma$ simultaneously counting
branes.
\item These expansions do not exist for $n<3$.
\item With $\Gamma = 0$ and $g_4 =
-\frac{3}{4}$ this class reduces to the $x<0$ brane--expansion in
$c$ seen for the 0B theory at $n=4$ in equation
(\ref{0Bexpnsm=4}). (Recall that there $q=-c$.)
\item These are the non--trivial broken--symmetry solutions obtained in
the 0B theory as one increases $n$. We reviewed this at the end of
section~\ref{sec:0B}.
\item As $n$ increases, further expansions in this class arise for
every odd $n$, which we generically label $v_{i \geq 5}$. These are
distinguished by the different values of $a$ in~\reef{classes}, but
since their behavior is identical for our purposes we often group
them together.
\item For odd $n$, reality imposes no restrictions on the direction of
$v_{i \geq 5}$.
\item For $n=4$, reality requires that we fix $v_5$ to be an $x
\rightarrow -\infty$ expansion. For the subsequent even $n$, some of
the $v_{i \geq 5}$ are real for $+x$, while the remaining are real
for~$-x$.
\end{itemize}
\subsection{The Structure at Higher $n$}
We can extrapolate the pattern observed for the first few $n$ and make
predictions for the structures that should appear at higher $n$. The
first observation is that there are $(n+1)^2$ expansions in all
(taking into account the various expansions related by symmetries in
each class) at each $n$. The counting can be broken down as follows.
\begin{table}[!h!!]
\begin{center}
\begin{tabular}[t]{|c|c|c|c|c|c|c|c|r|}
\hline
&$v_2$&$v_4$&$v_1$&$v_3$&$v_5$&$v_6$&$v_7$&${\rm total}$\\
\hline
$n=1$&1&1&&2&&&&4\\
\hline
$n=2$&1&2&2&4&&&&9\\
\hline
$n=3$&1&3&&6&6&&&16\\
\hline
$n=4$&1&4&4&8&8&&&25\\
\hline
$n=5$&1&5&&10&10&10&&36\\
\hline
$n=6$&1&6&6&12&12&12&&49\\
\hline
$n=7$&1&7&&14&14&14&14&64\\
\hline
$n=8$&1&8&8&16&16&16&16&81\\
\hline
\end{tabular}
\caption{\small The number and types of expansion classes for $v(x)$,
as a solution to the string equations~\reef{DWWstring1}
and~\reef{DWWstring2}, with increasing $n$ from 1 to 8. See text
for further discussion.}
\label{table:expansions}
\end{center}
\end{table}
Class 2 has exactly one member for each $n$, while Class 1 and Class 4
each have $n$ members. (Recall that Class 1 only exists for even $n$.)
As previously mentioned, for every odd $n$, new expansions in Class 5
(the $v_{i\geq 5}$) appear. These reduce to the $x<0$ broken symmetry
expansions of the 0B theory (for $\Gamma=0$ and $g_n$ in
equation~\reef{tn}) when $n$ is even. These expansion classes each
contain $2n$ members. The appearance of these new expansions is
consistent with the counting provided in ref.\cite{Klebanov:2003wg}
for the 0B expansions, as reviewed at the end of section~\ref{sec:0B}.
The counting is tabulated in Table~\ref{table:expansions}. All
together, we see that for odd $n$, adding across the rows gives a
total of $1 + n + \frac{n+1}{2} \cdot 2n = (n+1)^2$ expansions, while
for even $n$ we get $1 + n + n + \frac{n}{2} \cdot 2n = (n + 1)^2$
expansions.
\section{An Organizing Square}
\label{sec:square}
To construct a full solution for $v(x)$, we need to specify its
behaviour in the two asymptotic directions, positive and negative
$x$. Consider the example of type~0A discussed in
section~\ref{sec:0A}. The string equation at a given $k$ was shown to
have a solution connecting these two perturbative regimes with a full
non--perturbative completion, plotted for $k=2$ in
figure~\ref{fig:plot}. (Recall that $z\propto x$ and $w=-v$.) This
solution is in fact made of two expansions, $v_1(x)$ for the positive
$x$ regime and $v_2(x)$ for negative $x$. The parameter $c$ is frozen
to $-\frac12$ in this case, leaving the parameter $\Gamma$ to count
D--branes at $+x$ and fluxes at $-x$, as described
in~\ref{sec:general}.
This is the organizing scheme we follow in order to construct more
theories, with type~0B being another working example, this time with
$\Gamma=0$ and using $v_1$ for positive $x$ and
$v_3,v_5$, or the higher $v_i$ (not $v_4$) for negative $x$, as
already discussed.
In constructing new theories, matching perturbative expansions
does not guarantee that a full non--perturbative solution exists with
the desired properties. Further work is needed, using both analytic
and numerical techniques, in order to demonstrate the non--perturbative
existence of the proposed theories. This is the subject of our
companion paper, where we find several non--perturbative solutions
numerically, and present analytical arguments in favour of several new
non--perturbatively complete theories. For the rest of this paper, our
analysis will be concerned with the various perturbative regimes that
appear from our DWW string equations.
Much of this structure can be organized neatly into the shape of a
square, with the string theory special points we know so far at two of
the corners. The $v_1$ and $v_2$ pair form two edges with the type~0A
(with $c=-\frac12$) theory where they join. Then $v_3$ (at $n = 2$)
(or $v_5$ at $n = 4$, and so on) make another edge, with type~0B
($\Gamma=0$) at the corner where that edge meets the $v_1$ edge. See
figure~\ref{square1}.
Using our observations from section~\reef{sec:general}, we conjecture
that $v_4$ and $v_2$ form a physical pair when $\Gamma$ is fixed, with
$c$ counting either branes of fluxes in the perturbative
regimes. Similarly, $v_4$ with $v_3$ (or $v_{i \geq 5}$) may form
physical pair for fixed $c$, with $\Gamma$ counting fluxes or
branes. Since $v_2, v_3$, and $v_{i \geq 5}$ appeared as $x<0$
expansions, it is natural to fix the direction of $v_4$ to be $+x$. It
fits elegantly at the bottom of the square, at least when $n = 0 \mod
4$, ({\it i.e.,} when $v_4$ is real for positive $x$).
\begin{figure}[ht]
\begin{center}
\includegraphics[width=100mm]{SquareNo1}\\
\caption{\small DWW Expansions forming a square. See text for
explanation.}\label{square1}
\end{center}
\end{figure}
We summarize all of this in figure~\ref{square1}. The special points in
parameter space with $c^2 = \Gamma^2$ or $c=\Gamma$, where $v_2$ and $v_3$ vanish,
are represented by the dark squares on the vertical edges.
This way of organizing things immediately suggests that there are two
new special points, corresponding to the lower two corners of the
square, and we've called them Theory~A and Theory~B. We will need to
determine what the special values of $c$ and $\Gamma$ might be for
these corners, and the nature of the new theories. (The special values
$c=0$ and $\Gamma^2 = \frac{1}{4}$, complementary to the known values
for the type~0 theories, are suggestive, but so far this is a
guess. We will find several pieces of evidence to support this
suggestion in later sections.)
The lines connecting the special points are not (at this stage) to be
taken too literally, since we do not have a clear statement of the
nature of the theory (stringy or not) away from the special
points. However, the structure is highly suggestive, and reminiscent
of the square discovered in ref.\cite{Seiberg:2005bx} organizing the
moduli space of ${\hat c}=1$ strings. We reproduce it here in
figure~\ref{fig3}. ${\hat c}=1$ strings are fully two dimensional,
having in addition to the Liouville direction $\phi$ an extra
direction $X$. This direction can be compactified on a circle, and in
the square the lines represent values of radii varying between 0 and
$\infty$. The relations between theories then arise as a result of
T--dualities (horizontally) possibly combined with discrete twists by
discrete fermionic symmetry operations (vertically).
\begin{figure}[ht
\begin{center}
\includegraphics[width=105mm]{SquareNo3}\\
\caption{\small The moduli space of two--dimensional string
theories\cite{Seiberg:2005bx}. The four corners of the square
represent the four string theories 0B, 0A, IIB and IIA. The lines
labelled~1--8 represent different compactifications. The points on
each line represent compactifications with different radii
$R$. Lines~1, 4, 7 and 8 interpolate between different
non--compact theories as $R$ varies between $0$ and~$\infty$. The
points marked with black squares on lines 7 and 8 represent the
non--critical superstrings of ref.\cite{Kutasov:1990ua}. See
ref.\cite{Seiberg:2005bx} for discussion of other features of the
diagram.}\label{fig3}
\end{center}
\end{figure}
In the case under study here, there is generically no compact circle,
and so the analogy is limited, but it is possible that it is not
entirely coincidental that a square emerges. Two dimensional string
theories can descend to ${\hat c}<1$
theories by Renormalization Group flow\cite{Gross:1990ub,Hsu:1992cm},
and so an organizing square at ${\hat c}=1$ may well leave an imprint
at ${\hat c}<1$ that still is an organizing square. The fact that
there are two special points on the vertical lines on the ${\hat c}=1$
square that match our $c^2=\Gamma^2$ and $c=\Gamma$ points is suggestive.
Inspired by the similarity between our square and that of
ref.\cite{Seiberg:2005bx}, we explored whether the two unknown
theories at the bottom of our square could actually be type~IIA and
type~IIB string theories, coupled to superconformal minimal models.
We present the details of our explorations in section~\ref{sec:new}.
\section{A Search for New Theories}
\label{sec:new}
So far, we have demonstrated that the DWW hierarchy has an extremely
rich structure of asymptotic expansions that naturally contain both
the 0A and 0B string theories coupled to the $(A,A)$ $(2,4k)$
superconformal minimal models. We've also pointed out that there are
new expansions that seem to have perfectly stringy interpretations, in
terms of backgrounds containing D--branes or fluxes once either $c$ or
$\Gamma$ has been fixed. It is natural to wonder whether a sensible
interpretation as string theories coupled to some matter minimal
(super--minimal) models can be given to these new expansions. We will
find some success with this for some corners of parameter space.
The square in figure~\ref{fig3} motivates the conjecture that the new
string theories are type~II string theories coupled to some
superconformal minimal models. This has some physical motivation
since type II string theories can be obtained as twisted orbifolds of
type 0 theories. Also, the $(A,D)$ series of superconformal minimal
models can be obtained as orbifolds of the $(A,A)$ series $via$ a
twist in the matter sector. In descending from ${\hat c}=1$ to ${\hat
c}<1$, the remnants of the twisted T--dualities connecting the
type~0 and type~II sectors could well be a combination of these
orbifold actions. In what follows, we argue that type~II string
theories coupled to $(A,D)$ $(4,4k-2)$ superconformal minimal models
are natural candidates for the physics encoded by the new special
points of our string equations.
One method of partially checking which theories are being captured by
our asymptotic expansions is to compare the (putative) torus
contributions (terms at order~$g_s^0$ in the free energy) with a
continuum calculation ({\it i.e.,} results of a traditional
world--sheet string one--loop computation) for these models. Such a
comparison will enable us to specialize to various points in parameter
space and provide further consistency checks to determine the exact
underlying models.
\subsection{The $g_s^0$ terms}
We begin by listing the terms that appear at order $g_s^0$ in
the expansion for the free energy for each class of the expansions
studied in section~\ref{sec:beyond}, for all $n$:
\begin{eqnarray}\label{DWWtorusterms}
v_1(x)&:& \frac{n+1}{12n}-\frac{c^2}{n}+\frac{\Gamma^2}{n} \quad ,\nonumber\\
v_2(x)&:& c^2-\Gamma^2 \quad ,\nonumber\\
v_3(x)&:& \frac{n-1}{2n}c^2-\frac{n+1}{n}c\Gamma+\frac{n+3}{2n}\Gamma^2 \quad ,\\
v_4(x)&:& \frac{n+1}{6n}-\frac{c^2}{n}-\frac{2n+1}{n}\Gamma^2 \quad ,\nonumber\\
v_i(x)&:& \frac{n+1}{12n}-\frac{c^2}{n}+\frac{\Gamma^2}{n} \quad , \nonumber \quad (i \geq 5)\ .
\end{eqnarray}
Notice that the torus term in the expansion classes labeled $v_i$ is
identical to that in $v_1$. This is a generalization of the curious
observation that was made in ref.\cite{Klebanov:2003wg}, that the
symmetry breaking solutions (for $x<0$) of the 0B theory have the same
torus terms (for $x>0$) as those of the 0B theory.
In each case, these terms are at order $x^{-2}$ in the expansion, and
multiplied by $\nu^2$. Following {\it e.g.,}
equations~\reef{0AFreeEnergy} and~\reef{0BFreeEnergy}, the free energy
is obtained by integrating twice and dividing by $\nu^2$, yielding the
same terms above multiplied by $\ln(|x|)$, which is part of a standard
Liouville theory volume factor that is common to everything we will do
at this order at perturbation theory\footnote{The Liouville direction
$\phi$ is effectively a box of volume $V_{\rm
L}=-\ln(|\mu|/\Lambda)/\alpha_{\rm min}$ where $\alpha_{\rm min}$
is the Liouville dressing of the lowest dimension operator in the
theory (see {\it e.g.,}
refs.\cite{Ginsparg:1993is,DiFrancesco:1993nw} for a review). In a
unitary theory, $\mu$ is the cosmological constant, the coefficient
of the puncture operator, which measures worldsheet area. Recall
that the dilaton and hence the local string coupling increases with
$\phi$, and so there is a natural cutoff at the point where
perturbation theory begins to break down, denoted
$\Lambda$. Standard conventions are to choose a scale such that
$\Lambda$ is unity and we will write $\mu=x$ in much of what
follows, differing slightly from our notation in {\it e.g.,}
equation~\reef{0AFreeEnergy}.}.
\subsection{The Continuum Partition Functions}\label{continuum}
We now present several continuum partition functions in the even spin
structures sector for both type~0 and type~II theories coupled to
$(A,A)$ and $(A,D)$ modular invariants. Ref.~\cite{Saleur}, presents
the modular invariant partition functions in the even spin structures
$(-,-)$, $(-,+)$ and $(+,-)$ for all the ${\cal{N}} = 1$
superconformal minimal models, as classified in
ref.~\cite{Cappelli}. Ref.~\cite{Bershadsky:1991zs} combines these
results with Liouville theory to compute some of the string theory
partition functions and we follow their methods to present the type~II
expressions that we suggest at the end of this section.
\subsubsection{The Type 0 Theories}
{$\bullet$} The $(A_{p-1}, A_{q-1})$ modular invariants. \\
The contribution of the {\it even} spin structures to the genus one
path integral for the $(A_{p-1}, A_{q-1})$ superconformal minimal
models coupled to supergravity has been calculated in
ref.\cite{Bershadsky:1991zs}:
\begin{eqnarray}\label{PFnAAmins1}
Z^{(A,A)}_{\rm even} &=& - \frac{1}{16} \frac{(p - 1)(q - 1)}{(p + q - 1)} \ln |x| \quad , \quad (p,q \quad {\rm odd})\\
\label{PFnAAmins2}
Z^{(A,A)}_{\rm even} &=& - \frac{1}{16} \frac{(p - 1)(q - 1) + 1}{(p + q - 2)} \ln |x| \quad . \quad (p,q \quad {\rm even})
\end{eqnarray}
\\
{$\bullet$} The $(A_{p-1}, D_{q/2 + 1})$ modular invariants.\\
The superconformal minimal model partition functions may be written in terms of
the partition functions for fields on a circle at special radii, as
shown in ref.\cite{Saleur}. These can be combined with partition
functions for affinized compact circle theories to yield the desired
one--loop string theory expressions\cite{Bershadsky:1991zs}. Using
this technique, it is easy to show that the partition functions in the
even spin structures for the $(A_{p-1}, D_{q/2 + 1})$ modular
invariants are:
\begin{eqnarray}\label{PFnsADmins1}
Z^{(A,D)}_{\rm even} &=& - \frac{1}{64} \frac{(3p - 4)(q + 2)}{(p + q - 2)} \ln |x|\ , \quad (q = 2 \mod 4)\ ,\\
\label{PFnsADmins2}
Z^{(A,D)}_{\rm even} &=& - \frac{1}{32} \frac{(p - 2)(q + 3) + 2}{(p + q - 2)} \ln |x|\ , \quad (q = 0 \mod 4)\ .
\end{eqnarray}
\subsubsection{The Type II Theories}\label{sec:type-ii-z-even}
By analogy with the previous section, similar procedures can be used
to propose partition functions in the even spin structures for
superconformal minimal models coupled to the type II string theories, using as
starting point the partition functions for the corresponding circle
theories given in ref.\cite{Seiberg:2005bx}. Our results are:
\\
\newline
{$\bullet$} The $(A_{p-1}, A_{q-1})$ modular invariants. \\
\begin{eqnarray}\label{PFnAAminstypeII1}
\tilde{Z}^{(A,A)}_{\rm even} &=& \frac{1}{32} \frac{(p - 1)(q - 1)}{(p + q - 1)} \ln |x| \quad , \quad (p,q \quad {\rm odd})\\
\label{PFnAAminstypeII2}
\tilde{Z}^{(A,A)}_{\rm even} &=& \frac{1}{32} \frac{(p - 1)(q - 1) + 1}{(p + q - 2)} \ln |x| \quad . \quad (p,q \quad {\rm even})
\end{eqnarray}
{$\bullet$} The $(A_{p-1}, D_{q/2 + 1})$ modular invariants.\\
\begin{eqnarray}\label{PFnsADminstypeII1}
\tilde{Z}^{(A,D)}_{\rm even} &=& \frac{1}{64} \frac{p(q + 2)}{(p + q - 2)} \ln |x|\ , \quad (q = 2 \mod 4)\ ,\\
\label{PFnsADminstypeII2}
\tilde{Z}^{(A,D)}_{\rm even} &=& \frac{1}{64} \frac{p(q + 2) - 4}{(p + q - 2)} \ln |x|\ , \quad (q = 0 \mod 4)\ .
\end{eqnarray}
While this is a natural extension of the definitions for bosonic
strings and type~0 strings {\it via} combinations circle partition
functions, as we have already stated, an independent direct definition
of the type~II strings coupled to minimal models (explicitly coupling
to super--Liouville and defining the appropriate GSO projection) would
be desirable. This method only produces $Z_{\rm even}$, whereas a
direct definition would give explicit expressions for type~IIA and
type~IIB.
\subsection{Searching for Special Values of Parameters}\label{secSpecial}
\subsubsection{Comparison of Torus Terms --- The Known}
Before proceeding to the general story, we briefly review the
comparison\cite{Klebanov:2003wg} between the torus terms supplied by
the asymptotic expansions and the continuum calculations for the
type~0 theories coupled to the $(2,4k)$ (A,A) series of superconformal
minimal models.
For the $(2,4k)$ series, equation~\reef{PFnAAmins2} becomes
\begin{equation}\label{PFn2,4kmins}
Z_{\rm even}^{(2,4k)} = \frac{1}{2}\left(Z_{\rm 0A}(x) + Z_{\rm 0B}(x)\right) = -\frac{1}{16} \ln |x| \quad .
\end{equation}
Let us compare this with the results we have from our string equations
presented in section~\ref{sec:type0}. The torus terms in each
direction are as listed below:
\begin{eqnarray}\label{PFns0A0B}
Z_{\rm 0A} &=& -\frac{k-1}{24k} \ln |x|\ , \qquad Z_{\rm 0B} = -\frac{2k+1}{24k} \ln |x|\ , \qquad (x > 0) \ ,\nonumber\\
Z_{\rm 0A} &=& -\frac{1}{8} \ln |x|\ , \hskip0.75cm \qquad Z_{\rm 0B} = 0 \ , \qquad \hskip2.35cm (x < 0) \ .
\end{eqnarray}
These can be read off from the expansions given in
sections~\ref{sec:0A} and~\ref{sec:0B}, or alternatively by starting
with our DWW string equations and expansions given in
section~\ref{sec:beyond}, and specializing to either $c=-\frac12$
(type 0A) or $\Gamma=0$ (type 0B). It follows from the torus
terms~\reef{PFns0A0B} that
\begin{equation}\label{PFneven2,4k}
Z_{\rm even}^{(2,4k)} = \frac{1}{2}\left(Z_{\rm 0A}(x) + Z_{\rm 0B}(x)\right) = -\frac{1}{16} \ln |x| \quad ,
\end{equation}
for either sign of $x$, in agreement with the worldsheet computation
above. As argued in ref.\cite{Klebanov:2003wg}, this is strong
evidence that indeed these asymptotic expansions represent the
$(2,4k)$ super--minimal models coupled to supergravity.
Let us try to systematize the above procedure in the context of the
DWW string equations. Recall that specializing to the~0A~(0B) theory
requires $c = -\frac{1}{2}$ ($\Gamma = 0$). We seek to discover which
properties of our one--loop partition functions are general conditions
that produce these values.
To this end, turn again to the square of figure~\ref{square1}. It
suggests three possible theories: Theory $B$, which has $\Gamma$ fixed
and $v_2$ governing the $x\rightarrow -\infty$ asymptotics and $v_4$
governing the $x\rightarrow+\infty$ asymptotics (for which we
introduce the notation $(v_2|v_4)$); $\widetilde{A}$, which has $c$
fixed and $(v_3|v_4)$ asymptotics; and $\widehat{A}$, which has $c$
fixed and $(v_{i\ge 5}|v_4)$ asymptotics. We summarize these
possibilities and the resulting torus terms in table~\ref{table:Torus
terms}. These terms are obtained from~\reef{DWWtorusterms} by
eliminating $c$ or $\Gamma$ whenever its appearance represents an
insertion of a worldsheet boundary or a flux vertex operator, which
would change the topology. So, {\it e.g.}, $\Gamma$ cannot appear in
the type 0A torus terms while~$c$ cannot appear for type 0B.
\begin{table}[!h!!]
\begin{center}
\begin{tabular}{|c|c|c|}
\hline
$$ & $x>0$ & $x<0$ \\
\hline
$Z_{\rm 0A}\phantom{\bigg(}$ &$a\left(\frac{n+1}{12n}-\frac{c^2}{n}\right)\ \quad\quad \left(v_1\right)$ & $a{c^2}\ \;\quad\quad\quad\quad \left(v_2\right)$ \\
\hline
$Z_{\rm 0B}\phantom{\bigg(} $& $b\left(\frac{n+1}{12n} + \frac{\Gamma^{2}}{n}\right)\ \quad\quad \left(v_1\right)$ & $b\left(\frac{n+3}{2n}\right)\Gamma^{2}\ \quad \left(v_3\right)$ \\
\hline
$Z_{\rm \widetilde{A}}\phantom{\bigg(}$ & $\tilde{a}\left(\frac{n+1}{6n} - \frac{c^{2}}{n}\right)\ \quad\quad \left(v_4\right) $ & $\tilde{a}\left(\frac{n-1}{2n}\right)c^2\ \quad \left(v_3\right)$ \\
\hline
$Z_{\rm \widetilde{B}}\phantom{\bigg(}$& $\tilde{b}\left(\frac{n+1}{6n}-\frac{2n+1}{n}\Gamma^2\right)\ \quad \left(v_4\right)$ & $-\tilde{b}\Gamma^2\ \;\quad\quad \quad \left(v_2\right)$ \\
\hline
$Z_{\rm \widehat{A}}\phantom{\bigg(}$ & $\hat{a}\left(\frac{n+1}{6n} - \frac{c^{2}}{n}\right)\ \quad \quad \left(v_4\right) $ &$\hat{a}\left(\frac{n+1}{12n}-\frac{c^2}{n}\right)\ \quad \left(v_{i\ge 5}\right)$ \\
\hline
$Z_{\rm \widehat{B}}\phantom{\bigg(}$& $\hat{b}\left(\frac{n+1}{6n}-\frac{2n+1}{n}\Gamma^2\right)\ \quad \left(v_4\right)$ & $-\hat{b}\Gamma^2\ \;\quad\quad \quad \left(v_2\right)$ \\
\hline
\end{tabular}
\caption{\small Unnormalized torus terms for string theories, before
fixing parameters $c$ and $\Gamma$. We have also indicated from
which expansion each term arises.}\label{table:Torus terms}
\end{center}
\end{table}
As table~1 indicates, we have multiplied the torus terms by
unknown normalizations
$\{a,b\}$, $\{\tilde{a},\tilde{b}\}$ and $\{\hat{a},\hat{b}\}$, which we will later attempt to fix. Notice that we have distinguished two possibilities for $Z_B$: $Z_{\widetilde{B}}$ and $Z_{\widehat{B}}$, to match the ${\widetilde A}, {\widehat A}$ choices. We now construct the three possible $Z_{\mathrm{even}} = \frac{1}{2}\left(Z_A+Z_B\right)$ from these torus terms.\\
\newline 1. The known theories: 0A and 0B
\begin{eqnarray}\label{Zevenknown}
Z^{(\rm 0A, 0B)}_{\rm even} &=& \frac{1}{2}\left[\left(\frac{a+b}{12}\right)\left(\frac{n+1}{n}\right) - \frac{\left(ac^2 - b\Gamma^2\right)}{n}\right] \quad \ \quad (x > 0) \nonumber\\
Z^{(\rm 0A, 0B)}_{\rm even} &=& \frac{1}{2}\left[a c^2 + \left(\frac{n+3}{2n}\right)b\Gamma^{2}\right]\;\;\quad\quad\quad\quad\quad \quad\quad\quad (x < 0)
\end{eqnarray}
2. The unknown theories: $\widetilde{\rm A}$ and $\widetilde{\rm B}$
\begin{eqnarray}\label{Zevenunknown2}
Z^{(\rm \widetilde{A}, \widetilde{B})}_{\rm even} &=& \frac{1}{2}\left[\left(\frac{\tilde{a} + \tilde{b}}{6}\right)\left(\frac{n+1}{n}\right) - \frac{1}{n}\tilde{a}c^2 - \left(\frac{2n+1}{n}\right)\tilde{b}\Gamma^2\right] \quad \quad (x > 0) \nonumber\\
Z^{(\rm \widetilde{A}, \widetilde{B})}_{\rm even} &=& \frac{1}{2}\left[\left(\frac{n-1}{2n}\right)\tilde{a}c^2 - \tilde{b}\Gamma^2\right] \quad\qquad\qquad\qquad\quad\qquad \qquad\quad (x < 0)
\end{eqnarray}
3. The unknown theories: $\widehat{\rm A}$ and $\widehat{\rm B}$
\begin{eqnarray}\label{Zevenunknown3}
Z^{(\rm \widehat{A}, \widehat{B})}_{\rm even} &=& \frac{1}{2}\left[\left(\frac{\hat{a} + \hat{b}}{6}\right)\left(\frac{n+1}{n}\right) - \frac{1}{n}\hat{a}c^2 - \left(\frac{2n+1}{n}\right)\hat{b}\Gamma^2\right] \quad , \quad (x > 0) \nonumber\\
Z^{(\rm \widehat{A}, \widehat{B})}_{\rm even} &=& \frac{1}{2}\left[\hat{a}\left(\frac{n+1}{12n}\right) - \frac{1}{n}\hat{a}c^2 - \hat{b}\Gamma^2\right] \qquad \qquad\qquad \qquad\quad \quad\quad (x < 0)
\end{eqnarray}
\subsubsection{The known theories: 0A and 0B}
We will now see what conditions are required to deduce the special
parameter values, $c = -\frac{1}{2}$ and $\Gamma = 0$, for type~0A
and~0B respectively. We begin by observing that the continuum
partition functions listed in section~\ref{continuum} are independent
of the sign of $x$. Therefore, we will impose this as a constraint on
every $Z_{\rm even}$ that we construct,
\begin{equation}\label{Prop1}
\textrm{Condition 1:}\quad\quad
Z_{\rm even}(x > 0) = Z_{\rm even}(x < 0)\quad.
\qquad\qquad\quad
\end{equation}
We impose this condition by equating the two expressions in
equation~\reef{Zevenknown}. Interestingly, the $n$--dependence
factorizes completely leading to
\begin{equation}\label{RelncGammaAA}
ac^{2} + \frac{b}{2}\Gamma^{2} = \frac{a + b}{12} \quad .
\end{equation}
\indent While it is indeed possible for $c$ and $\Gamma$ to inherit
$n$ dependence from dependence on $t_n$, the simplest possibility is
that these parameters are independent of $n$. That the $n$ dependence
factors out of equation~\reef{RelncGammaAA} allows this simplicity to
be realized. This suggests that we also impose
\begin{equation}
\textrm{Condition 2:}\quad\quad
c \textrm{ and } \Gamma \textrm{ are independent of } n . \\
\qquad\qquad\quad
\end{equation}
Substituting~\reef{RelncGammaAA} back into~\reef{Zevenknown} we
obtain,
\begin{equation}\label{Z0A0B}
Z^{(\rm 0A,0B)}_{\rm even} = \frac{a+b}{24}+\frac{3b}{4n}\Gamma^2
\end{equation}
Finally, we impose one more condition,
\begin{equation}
\textrm{Condition 3:}\quad\quad
Z_{\mathrm{even}} \textrm{ is independent of } n,
\qquad\qquad\quad
\end{equation}
which, together with Condition 2, forces us to conclude
\begin{equation}\label{cGammaAAvalues}
c^{2} = \frac{a+b}{12a}\ , \quad \Gamma = 0 \quad \Rightarrow \quad Z^{(\rm 0A,0B)}_{\rm even} = \frac{a+b}{24}\; .
\end{equation}
\indent Condition 3 is motivated by the remarkable fact that
$Z_{\rm{even}}$ \emph{is} independent of $n$, which is nontrivial
given the form of equation~\reef{PFnAAmins2}, and also by the
observation that this constraint correctly produces the required
parameter values in this case. This is as far as we can go without
some extra information. Fortunately, for the type~0 theories, we
actually know $a= -\frac{1}{2}$ and $b = -1$: the factor of half is
because of the doubling of the free energy the 0B theory relative to
the 0A theory and the negative sign is because it is $-v$ which
defines the two--point function for these theories. (Recall the
observation at the end of section~\reef{sec:0Breduc}). Thus we see
that we obtain the correct parameter values $c=-\frac{1}{2}$ and
$\Gamma=0$ for the theories under consideration. Moreover, we
correctly obtain the known value\footnote{A subtle but important point
should be mentioned here. We have used the torus term from $v_3$ to
be the $x < 0$ contribution to the 0B theory leading us to $\Gamma =
0$. This means that the $x < 0$ contribution to the 0B theory is
actually zero. While $v_3$ is real as a $x < 0$ expansion for $n = 2
\mod 4$, it is complex for $n = 0 \mod 4$. The resolution is that
the zero contribution is to be understood as coming from the {\it
trivial} $v = 0$ solution of the 0B theory obtained by setting $c
= 0$ in the 0B string equations. It is interesting that our
procedure using $v_3$ should give parameter values that are
consistent with the {\it trivial} solution.} of
$Z_{\rm{even}}=-\frac{1}{16}$.
Next, we turn to exploring where our new conditions take us in
investigating the unknown corners.
\subsubsection{The unknown corners: $\widetilde{\rm A}$ and $\widetilde{\rm
B}$}\label{sec:unknown corners}
Using equation~\reef{Zevenunknown2}, we see that Condition~1 gives,
\begin{equation}\label{RelncGammaAD}
\frac{\tilde{a}}{2}c^{2} + \tilde{b}\Gamma^{2} = \frac{\tilde{a} + \tilde{b}}{6} \quad \Rightarrow \quad Z^{(\rm \widetilde{A}, \widetilde{B})}_{\rm even}= -\frac{\tilde{a}+\tilde{b}}{12} +\tilde{a}c^2\left(\frac{1}{2}-\frac{1}{4n}\right)\, .
\end{equation}
Conditions~2 and~3 give,
\begin{equation}\label{GammaAD}
\Gamma^{2} = \frac{\tilde{a} + \tilde{b}}{6\tilde{b}} \quad \mathrm{and} \quad
c = 0 \ ,
\end{equation} and so\footnote{This $c = 0$ result means that the
contribution to the torus terms coming from $v_3$ is zero. In a
sense, this result is the analog of what
was seen for the type~0B theories. Our argument again is that the torus term
for $x<0$ at the $(v_4|v_3)$ corner comes from the trivial $v=0$
solution, analogous to the 0B case.} we have
\begin{equation}\label{ABtilde}
Z^{(\rm \widetilde{A}, \widetilde{B})}_{\rm even} = -\frac{\tilde{a} + \tilde{b}}{12} \quad .
\end{equation}
Remarkably enough, everything seems to work as before, except now we
obtain new values for our parameters and a new expression for $Z_{\rm
even}$. We have yet to make an numerical prediction for $Z_{\rm
even}$ as we still must fix the normalizations. We will return to
this point in next section~\ref{sec:new-theories}.
\subsubsection{The unknown corners: $\widehat{\rm A}$ and $\widehat{\rm
B}$}\label{sec:unknown-tilde}
Condition~1 gives:
\begin{equation}
\Gamma^2=\frac{\hat{a}+2\hat{b}}{12\hat{b}}\quad \Rightarrow \quad Z^{(\rm \widehat{A}, \widehat{B})}_{\rm even} = -\frac{\hat{b}}{12}+\frac{\hat{a}(1-12c^2)}{24n}
\end{equation}
Conditions~2 and~3 imply,
\begin{equation}
c^2=\frac{1}{12}, \quad\Gamma^2=\frac{\hat{a}+2\hat{b}}{12\hat{b}}\quad\Rightarrow\quad Z^{(\rm \widehat{A}, \widehat{B})}_{\rm even} = -\frac{\hat{b}}{12}
\end{equation}
This result is not as encouraging. $Z_{\rm even}$ does not depend on
$\hat{a}$ which implies that it does not depend on theory $\hat{A}$ at
all. We find this unacceptable and therefore conclude that
Conditions~2 and~3 are incompatible constraints in this case. We
nevertheless take seriously the possibility that theories like this
may exist, but with $n$--dependent parameters and an $n$--dependent
$Z_{\rm even}$. We will briefly consider this possibility in
section~\reef{sec:ndep}.
\subsection{The New Theories}\label{sec:new-theories}
In the above we have explored the crucial and quite constraining
assumption that $Z_{\rm even}$ is independent of the index $n =
2k$. Theories $\tilde{\rm A}$ and $\tilde{\rm B}$ seemed particularly
suited to produce such a $Z_{\rm even}$. In an effort to better
understand what these theories may be, we next ask which continuum
theories are capable of producing an $n$--independent $Z_{\rm
even}$. Looking at the various partition functions listed in
section~\ref{continuum}, it is easy to see that only the following
choices give
rise to an $n$--independent partition function:\\
\newline 1. $Z^{(A,A)}_{\rm even}$ in equation~\reef{PFnAAmins2} with
$p=2$ and $q=4k$.\footnote{The two positive integers $p$ and $q$
labeling the super--minimal models must obey: $q > p$; $q - p = 0
\mod 2$; if both are odd, they are coprime and if both are even,
then $p/2$ and $q/2$ must be coprime. There is also a standard
restriction that if $p$ and $q$ are even, then $(q-p)/2$ must be
odd. It follows that if $p=2$ then $q=4k$.} These theories are
type~0 string theories coupled to the $(2,4k)$ $(A,A)$ models and are
already described
at the known corners.\\
\newline 2. $Z^{(A,D)}_{\rm even}$ in equation~\reef{PFnsADmins1} with
$p=4$ and $q = 4k - 2$.\footnote{ Strictly speaking, $q = 4k \pm
2$. The choice $q = 4k + 2$ would suggest that the $(4,6)$ $(A,D)$
model exists at at $k=1$; however, all combinations of real
expansions at $k=1$ have been exhausted in describing the type~0
$(A,A)$ $(2,4)$ model. We therefore expect
the $(4,6)$ model to appear at $k=2$ where more combinations of expansions exist since $v_4$ is real. Thus we choose $q=4k-2$ with $k\ge 2$.} These theories are the type~0 strings theories coupled to the $(4, 4k-2)$ $(A,D)$ models with $Z^{(A,D)}_{\rm even} = - \frac{1}{8} \ln |x|$.\\
\newline 3. $\tilde{Z}^{(A,A)}_{\rm even}$ in
equation~\reef{PFnAAminstypeII2} with $p=2$ and $q=4k$. These are the
type II strings coupled to the $(2,4k)$ $(A,A)$
models with $\tilde{Z}^{(2,4k)}_{\rm even} = \frac{1}{32} \ln |x|$.\\
\newline 4. $\tilde{Z}^{(A,D)}_{\rm even}$ in
equation~\reef{PFnsADminstypeII2} with $p=2$ and $q=4k$. These are the
type II strings coupled to the $(2,4k)$ $(A,D)$ models with
$\tilde{Z}^{(2,4k)}_{\rm even} = \frac{1}{32} \ln |x|$. This is the
same partition function as case~$(3)$ above.
\\
\newline 5. $\tilde{Z}^{(A,D)}_{\rm even}$ in
equation~\reef{PFnsADminstypeII1} with $p=4$ and $q = 4k-2$. These
theories are the type~II string theories coupled to the $(4,4k-2)$
$(A,D)$ models with
$\tilde{Z}^{(4,4k-2)}_{\rm even} = \frac{1}{16} \ln |x|$.\\
\newline We seek to identify $Z^{(\rm \widetilde{A},
\widetilde{B})}_{\rm even} = -\frac{\tilde{a} + \tilde{b}}{12}$ with
one of the above partition functions, but to do so we will need some
way of determining $\tilde{a}$ and $\tilde{b}$. We consider the
possibility that the theories $\rm \widetilde{A}$ and $\rm
\widetilde{B}$ are type~IIA and type~IIB string theories,
respectively. As we will see, this will uniquely fix the
normalizations and allow us to identify the resulting $Z^{(\rm
\widetilde{A}, \widetilde{B})}_{\rm even}$ as that of case~(5)
above, {\it i.e.} type~II string theories coupled to the $(4,4k-2)$
$(A,D)$ models.
\subsubsection{Type~II Theories Coupled to the Superconformal Minimal Models}
Pursuing the apparent similarity with the moduli space of ${\hat c}=1$
theories\cite{Seiberg:2005bx} a little further, (recall our discussion
in section~\ref{sec:square}, and figure~\ref{fig3}) we can study the
partition functions of those theories (compactified on a circle) to
get clues as to the possible relative normalizations. We list them
here\cite{Seiberg:2005bx}:
\begin{eqnarray}
{\rm 0A:} \qquad && \frac{Z}{V_L}=\frac{1}{12\sqrt{2}}\left(\frac{\sqrt{\alpha^\prime}}{R}+2\frac{R}{\sqrt{\alpha^\prime}}\right) \ , \nonumber \\
{\rm 0B:} \qquad && \frac{Z}{V_L}=\frac{1}{12\sqrt{2}}\left(2\frac{\sqrt{\alpha^\prime}}{R}+\frac{R}{\sqrt{\alpha^\prime}}\right)\ , \nonumber\\
{\rm IIA:} \qquad && \frac{Z}{V_L}=-\frac{1}{24\sqrt{2}}\left(2\frac{\sqrt{\alpha^\prime}}{R}+\frac{R}{\sqrt{\alpha^\prime}}\right)\ , \nonumber \\
{\rm IIB:} \qquad && \frac{Z}{V_L}=-\frac{1}{24\sqrt{2}}\left(\frac{\sqrt{\alpha^\prime}}{R}+2\frac{R}{\sqrt{\alpha^\prime}}\right)\ .
\end{eqnarray}
An important clue is to be found in the $1/R$ behaviour of each
theory. This is the physics of the field theory sector (Kaluza--Klein
states) that propagate on the circle, and the relative normalizations
are of interest to us. Type~0B has twice as much energy in this sector
as type~0A, and type~IIA has double that of type~IIB. Now, as
expected, the type~0 theories exchange under the T--duality operation
$R\to\alpha^\prime/R$, as do the type~II theories. Now T--duality
vertically should take the type~II theories to the type~0 theories,
but this needs to be done on a circle twisted by $(-1)^{f_l}$. The
partition functions for IIA/B on such a circle is:
\begin{eqnarray}
{\rm IIA:} \qquad && \frac{Z}{V_L}=-\frac{1}{24\sqrt{2}}\left(\frac{\sqrt{\alpha^\prime}}{R}-\frac{R}{\sqrt{\alpha^\prime}}\right) \ , \nonumber \\
{\rm IIB:} \qquad && \frac{Z}{V_L}=-\frac{1}{12\sqrt{2}}\left(\frac{\sqrt{\alpha^\prime}}{R}-\frac{R}{\sqrt{\alpha^\prime}}\right)\ .
\end{eqnarray}
Notice that the field theory term of type~IIA is now actually half
that of type~IIB. Now we can compare the overall type~II normalization
to the type~0 by doing the vertical T--duality. Under
$R\to\alpha^\prime/R$ we see that type~IIB field theory term matches
that of the type~0A theory, while for type~IIA, the field theory term
is 1/4 that of the type~0B value.
In summary, we see that comparing squares would suggest that $Z_{\rm
A}=Z_{\rm IIA}$ and $Z_{\rm B}=Z_{\rm IIB}$, and the chain of
relationships above gives the relative normalization of 0B and 0A as 2
to 1, while that of (twisted) IIB and IIA is 2 to 1, so ${\tilde
b}=2{\tilde a}$. Finally, T--duality (with a twist) between 0A
and IIB gives them the same relative normalization, implying,
\begin{equation}
\tilde{b}=-\frac{1}{2}, \quad \tilde{a}=-\frac{1}{4}\; .
\end{equation}
In this case we get therefore that for our putative type~II theories,
\begin{equation}
Z^{(\rm \tilde{A}, \tilde{B})}_{\rm even}=\frac{1}{16}\ln |x| \; ,
\end{equation}
which is equal to that of the type~0 case, but with an opposite
sign. This is the same as the partition
function~\reef{PFnsADminstypeII1} for the type~II theories coupled to
the $(4,4k-2)$ $(A,D)$ super--minimal models. Interestingly,
equation~\reef{GammaAD} implies that $\Gamma^2=1/4$ for IIB and $c=0$
for IIA, which nicely mirrors the values $c^2=1/4$ for 0A and
$\Gamma=0$ for 0B.
Further work is required to strengthen our conjecture that we have
here the type~IIA and~IIB theories coupled to the superconformal
minimal models. As stated, an explicit definition of ${\hat c}<1$
type~II strings (separately for types~A and~B, and not just our
suggestion for $Z_{\rm even}$ given in
section~\ref{sec:type-ii-z-even}) does not yet seem to exist in the
literature, and our attempts to directly define them so far are
incomplete. Note that we have assumed that the relative normalizations
of the partition functions that follow from ${\hat c}=1$ really
descend to the ${\hat c}<1$ case, and while reasonable, this needs to
be proven. The explicit $k$ dependence of the individual partition
functions that we get by taking the DWW one--loop expressions
seriously have a (so far) unilluminating form at positive $x$ that we
have not been able to check against a continuum computation. On the
other hand the negative $x$ result neatly mirrors the type~0 case. We
list the results here, and leave this
line of investigation for the future\footnote{We also considered the
possibility that our $\tilde{A}$ and $\tilde{B}$ theories are again
type~0A and type~0B string theories, respectively, perhaps coupled
to an $(A,D)$ modular invariant. This would again fix the relative
normalization of the partition functions such that $a=-\frac{1}{2}$
and $b=-1$, yielding $Z^{\rm (\widetilde{A},\widetilde{B})}_{\rm
even} = \frac{1}{8} \ln |x| $, which unfortunately differs from
the continuum partition function for the $(4, 4k-2)$ $(A,D)$
super--minimal models by a sign. To match, we would have to assume
that $+v$ encodes the partition function instead of $-v$, which
contradicts our earlier establishment of the $(A,A)$ type~0
points.}:
\begin{eqnarray}\label{PFnsIIAIIB}
Z_{\rm IIA}&=&-\frac{2k+1}{48 k}\ln|x|\ , \qquad Z_{\rm IIB}=\frac{8k+1}{48k}\ln|x|\ , \qquad (x>0)\ ,\nonumber \\
Z_{\rm IIA}&=&0\ , \qquad \hskip2.3cm Z_{\rm IIB}=\frac18\ln|x|\ , \hskip0.9cm\qquad (x<0)\ .
\end{eqnarray}
This should be compared with the type~0 cases displayed in
equation~\reef{PFns0A0B}.
\subsubsection{$n$--dependence}\label{sec:ndep}
Our main successes so far have involved theories in which $Z_{\rm
even}$ and the values for $\Gamma$ and $c$ have been independent of
$n$. While these are the simplest possibilities, there is little
reason to suspect that these theories are the only types amenable to
description by DWW. We observed in section~\ref{sec:unknown-tilde}
that there might be new theories, {\it e.g.,} the ones we called
$\widehat{A}$ and~$\widehat{B}$, whose description will most likely
require such $n$--dependence for $(\Gamma,c)$. Unfortunately, the
availability of continuum calculations to compare to is limited, so
obtaining convincing evidence for any particular proposal is
difficult. Nevertheless, there is much more evidence to accumulate
from the differential equations themselves. We hope to report upon
this issue in later work.
We briefly mention one way that $n$--dependence might emerge
naturally. The $Z_{\rm even}$ that we have constructed from the torus
terms of our perturbation expansions depend on three variables: $n$,
$c$, and $\Gamma$. We must however impose sign independence which
reduces the total number of independent variables describing $Z_{\rm
even}$ to two. In special cases the algebra conspires and we find
that $Z_{\rm even}$ actually depends on no variables at all. This
behavior is to be compared to $Z_{\rm even}$ as computed from the
continuum partition functions. For generic $p$ and $q$, $Z_{\rm even}$
is a function of two variables, but in special cases this dependence
completely vanishes. This motivates us to attempt to express for $n$,
$c$, and $\Gamma$ as functions of $p$ and $q$.
Equation~\reef{Z0A0B} indicates that $Z_{\rm even}^{\rm 0A, 0B}$ is
inversely proportional to $n$. The form of equation~\reef{PFnAAmins2} then
suggests that we take $n\sim p+q-2$. Since $q=4k=2n$ in this case, we
can predict
\begin{equation}
n=\frac{1}{2}(p+q-2)\ .
\end{equation}
Now equating $Z_{\rm even}^{\rm 0A, 0B}$ of equation~\reef{Z0A0B} with
the quantity in equation~\reef{PFnAAmins2} and using the relation
$c^2+\Gamma^2=\frac{1}{4}$, which follows from sign independence,
gives
\begin{equation}
\Gamma^2 =\frac{1}{12}(p-2)(q-2)\ .
\end{equation}
Note that the models already studied had $p=2$ which gave us
$\Gamma=0$. We can perform the same exercise for $Z^{(\rm \tilde{A},
\tilde{B})}_{\rm even}$ which we have argued describes type II
strings coupled to $(A,D)$ $(4,4k-2)$ models. Again we have
$n=\frac{1}{2}(p+q-2)$, which holds for $p=4$ and
$q=4k-2=2n-2$. Equating $Z^{(\rm \tilde{A}, \tilde{B})}_{\rm even}$ of
equation~\reef{RelncGammaAD} with the quantity in
equation~\reef{PFnsADminstypeII1} and using $\frac{1}{4}c^2+\Gamma^2 =
\frac{1}{4}$ gives,
\begin{equation}
c^2=-\frac{(p-4)(q-2)}{8(p+q-3)} \ .
\end{equation}
For the models we've previously
considered, $p=4$, so we correctly reduce to the condition $c=0$. Note
here that we would run into problems if we considered $p>4$ because
then $c$ would become imaginary.
\section{Discussion}
\label{sec:conclusion}
As discussed at length in the introduction, we find very significant
this rich framework into which we can embed so many string theories and
discover how they intertwine with each other. We suspect that there
may be many more theories to be found in this framework, and that we
have only just begun to learn how to extract and identify the various
theories using the limited comparisons we can do to existing continuum
computations.
We have been able to gather evidence for a square of connected
theories, firmly establishing the top corners as type~0A and~0B
theories coupled to the $(A,A)$ modular invariant $(2,4k)$ minimal
models, and finding several strong pieces of evidence that the special
points at the bottom are string theories, the
(non--supersymmetric\cite{Seiberg:2005bx}) type~IIA and~IIB theories
coupled to the $(A,D)$ modular invariant $(4,4k-2)$ minimal
models. See figure~\ref{square2}.
Notice that the $v_3$ side of the square is generically understood as
defining the theory when the number background branes, {\it i.e.,}
$c$ or $\Gamma$, is set to zero. This is the case for type~0B (already established
in ref.~\cite{Klebanov:2003wg}) and is inherited by the type~IIA
theory as well. Away from $c=0$ $v_3$ is complex for $k$ even, and one
of the symmetry breaking $v_i (i\geq5)$ may supply the $x<0$ physics
instead. (For $k$ odd, $v_3$ is real for $c\neq0$ and so may well
furnish the $x<0$ physics in those cases.)
It is also important to note that for $k$ odd, $v_4$ is no longer real
(with the exception of $c=0$ at $k=1$). This suggests that this regime
of perturbation theory at $x>0$ (containing fluxes for type~IIA and
branes for type~IIB) is ill--defined at those values of $k$, even
while the opposite regime at $x<0$ exists. This may suggest
non--perturbative instabilities for those values of $k$. An
instructive prototype of this possibility is the original bosonic
family of string theories defined non--perturbatively in terms of a
Painlev\'e~I hierarchy of string
equations\cite{Gross:1989vs,Brezin:1990rb,Douglas:1989ve}. At large
$x$, the leading behaviour for the two point function was $w(x)=x^k$
for the $k$th model, and for positive $x$ the physics was the
$(2,2k-1)$ conformal minimal models coupled to Liouville theory. For
even $k$, the the $x<0$ regime gives complex values, signalling the
non--perturbative problems. Something analogous may be going on here
for odd $k$ (except $k=1$, $c=0$) for the type~II $(A,D)$ $(4,4k-2)$
models. We will report further non--perturbative aspects in a
follow--up paper where we discuss numerical and analytic studies of
solutions that connect different perturbative regimes\cite{companion}.
\begin{figure}[ht]
\begin{center}
\includegraphics[width=100mm]{SquareNo2v3}\\
\caption{A family of string theories, forming a square. See text for
details.}\label{square2}
\end{center}
\end{figure}
There are many more interesting issues to understand. While we
compared to our proposal for a continuum definition of the even sector
partition function of the type~II theories, a direct continuum
definition of the type~IIA and~IIB string theories would be valuable
to have, in order to provide another check on our conjecture for the new
points. Whether or not they are type~II string theories, it is clear
that these new theories are of interest, and may (as already stated)
be only the first in a very large family of new theories that our
string equations define. They are all nicely interconnected with the
known type~0 theories, and have a rich non--perturbative sector. We
will explore much of this non--perturbative physics in a follow--up
paper\cite{companion}.
We have seen several signs that the organizing square is inherited
from the organizing square of theories seen at ${\hat c}=1$ in
ref.\cite{Seiberg:2005bx}. We suspect this inheritance arises
physically by RG flow, and it would be of value to explore this
further (there are bosonic investigations in the
literature\cite{Gross:1990ub,Hsu:1992cm}). In particular, the special
points with $\Gamma{=}\pm c$, where our partition functions vanish
(also suggested by the underlying Painlev\'e~IV structures we saw in
section~\ref{sec:painleve}, which are also worth understanding
better), may well be related to the supersymmetric points identified
in ref.\cite{Seiberg:2005bx}.
In summary, we have found a rich laboratory of solvable string theory
models with several non--perturbative connections between them by
realizing them as special points of a larger physical system, the
theory of dispersive water waves. In some respects it is an analogue
of what we would like to see in studies of M--theory. It will be
interesting to learn whether the larger framework of dispersive water
waves can yield any new insights about the non--perturbative nature of
string theories and related theories.
\section*{Acknowledgements}
This work was supported by the US Department of Energy. RI thanks
Tameem Albash, Nikolay Bobev, Arnab Kundu, Hubert Saleur, and Nicholas
Warner for useful conversations. CVJ thanks the Aspen Center for
Physics for hospitality while some of this work was carried out.
\providecommand{\href}[2]{#2}\begingroup\raggedright |
\section{Introduction}
\qquad In \cite{DK1} the concept of $harmonic-$ geometric polynomials and $%
harmonic-$ exponential polynomials are introduced and $hyperharmonic$
generalizations of these polynomials and numbers are obtained. Furthermore
it is shown that these polynomials are quite useful to obtain closed forms
of some series related to harmonic numbers. In this paper, we extend this
analysis to $r-versions$ of these polynomials and numbers.
Boyadzhiev\ $\cite{B}$ has presented\ and discussed the following
transformation formula:%
\begin{equation}
\sum_{n=0}^{\infty }\frac{g^{\left( n\right) }\left( 0\right) }{n!}f\left(
n\right) x^{n}=\sum_{n=0}^{\infty }\frac{f^{\left( n\right) }\left( 0\right)
}{n!}\sum_{k=0}^{n}\QATOPD\{ \} {n}{k}x^{k}g^{\left( k\right) }\left(
x\right) \label{1}
\end{equation}%
where $f$, $g$ are appropriate functions and $\QATOPD\{ \} {n}{k}$ are
Stirling numbers of the second kind.
One of the principal objectives of the present paper is to give closed forms
of some series related to harmonic numbers as well. To this end, we give a
useful generalization of $\left( \ref{1}\right) $ which contains $r-$%
Stirling numbers of the second kind instead of Stirling numbers of the
second kind as:%
\begin{equation}
\sum_{n=r}^{\infty }\frac{g^{\left( n\right) }\left( 0\right) }{n!}\binom{n}{%
r}\frac{r!}{n^{r}}f_{r}\left( n\right) x^{n}=\sum_{n=r}^{\infty }\frac{%
f^{\left( n\right) }\left( 0\right) }{n!}\sum_{k=0}^{n}\QATOPD\{ \}
{n}{k}_{r}x^{k}g^{\left( k\right) }\left( x\right) \text{,} \label{mf}
\end{equation}%
where $f_{r}\left( x\right) $ denotes the Maclaurin series of $f\left(
x\right) $ exclude the first $r-1$ terms.
Thanks to formula $\left( \ref{mf}\right) $ we introduce the concept of $r-$
geometric and $r-$ exponential polynomials and numbers. We obtain explicit
relations between the $r-$versions and the classical versions of these
polynomials and numbers. Besides, we present harmonic (and hyperharmonic)
versions of $r-$ geometric and $r-$ exponential polynomials and numbers as
well. The short lists of all these polynomials and numbers are given.
On the other hand formula $\left( \ref{mf}\right) $ and harmonic $r-$
geometric polynomials enables us to obtain closed forms of the following
series%
\begin{equation}
\sum_{n=r}^{\infty }\binom{n}{r}r!n^{m-r}H_{n}x^{n}\text{,} \label{mfr}
\end{equation}%
where $m$ and $r$ are integers such that $m\geq r$ and $H_{n}$ is the $n$-th
partial sum of the harmonic series
In the rest of this section we introduce some important notions.
\textbf{Stirling numbers of the first and second kind}
Stirling numbers of the first kind $\QATOPD[ ] {n}{k}$ and Stirling numbers
of the second kind $\QATOPD\{ \} {n}{k}$ are quite important in
combinatorics $\left( \cite{BG, BQ, C, Ri}\right) $. Briefly for the
integers $n\geq k\geq 0;$ $\QATOPD[ ] {n}{k}$ counts the number of
permutations of $n$ elements with exactly $k$ cycles and $\QATOPD\{ \} {n}{k}
$ counts the number of ways to partition a set with $n$ elements into $k$
disjoint, nonempty subsets $\left( \cite{C}\right) $.
We note that for $n\geq k\geq 1$, the following identity holds for Stirling
numbers of the second kind%
\begin{equation}
\QATOPD\{ \} {n}{k}=\QATOPD\{ \} {n-1}{k-1}+k\QATOPD\{ \} {n-1}{k}.
\label{4}
\end{equation}
There is a certain generalization of these numbers namely $r$-Stirling
numbers $\left( \cite{Br}\right) $ which are similar to the weighted
Stirling numbers $\left( \cite{CA1, CA2}\right) $. Represantation and
combinatorial meanings of these numbers are as follows $\left( \cite{Br}%
\right) $:
$r-$Stirling numbers of the first kind;%
\begin{eqnarray*}
\QATOPD[ ] {n}{k}_{r} &=&\text{The number of permutations of the set }%
\left\{ 1,2,\ldots ,n\right\} \text{ having} \\
&&k\text{ cycles, such that the numbers }1,2,...,r\text{ are in distinct
cylecs,}
\end{eqnarray*}
$r-$Stirling numbers of the second kind;%
\begin{eqnarray*}
\QATOPD\{ \} {n}{k}_{r} &=&\text{ The number of partitions of the set }%
\left\{ 1,2,\ldots ,n\right\} \text{ into } \\
&&k\text{ non-empty disjoint subsets, such that the numbers } \\
&&1,2,...,r\text{ are in distinct subsets.}
\end{eqnarray*}
Specializing $r=0$ gives the classical Stirling numbers.
The $r-$Stirling numbers of the second kind satisfy the same recurrence
relation as $\left( \ref{4}\right) $ except for the initial conditions, i.e $%
\left( \cite{Br}\right) $%
\begin{eqnarray}
\QATOPD\{ \} {n}{k}_{r} &=&0\text{, \ \ }n<r, \notag \\
\QATOPD\{ \} {n}{k}_{r} &=&\delta _{k,r}\text{, \ \ }n=r, \label{5} \\
\QATOPD\{ \} {n}{k}_{r} &=&\QATOPD\{ \} {n-1}{k-1}_{r}+k\QATOPD\{ \}
{n-1}{k}_{r}\text{, \ \ }n>r. \notag
\end{eqnarray}%
\textbf{Exponential polynomials and numbers}
Exponential polynomials (or single variable Bell polynomials) $\phi
_{n}\left( x\right) $ are defined by$\left( \cite{BL1, B2, G, Ri}\right) $%
\begin{equation}
\phi _{n}\left( x\right) :=\sum_{k=0}^{n}\QATOPD\{ \} {n}{k}x^{k}. \label{6}
\end{equation}
The first few exponential polynomials are:%
\begin{equation}
\begin{tabular}{|l|}
\hline
$\phi _{0}\left( x\right) =1\text{,}$ \\ \hline
$\phi _{1}\left( x\right) =x\text{,}$ \\ \hline
$\phi _{2}\left( x\right) =x+x^{2}\text{,}$ \\ \hline
$\phi _{3}\left( x\right) =x+3x^{2}+x^{3}\text{,}$ \\ \hline
$\phi _{4}\left( x\right) =x+7x^{2}+6x^{3}+x^{4}\text{.}$ \\ \hline
\end{tabular}
\label{7}
\end{equation}
The well known exponential numbers (or Bell numbers)\ are obtained by
setting $x=1$ in $\phi _{n}\left( x\right) $, i.e $\left( \cite{BL2, C, CG}%
\right) $%
\begin{equation}
\phi _{n}:=\phi _{n}\left( 1\right) =\sum_{k=0}^{n}\QATOPD\{ \} {n}{k}.
\label{8}
\end{equation}%
Hence the first few exponential numbers are:%
\begin{equation}
\phi _{0}=1\text{, }\phi _{1}=1\text{, }\phi _{2}=2\text{, }\phi _{3}=5\text{%
, }\phi _{4}=15\text{.} \label{9}
\end{equation}
In \cite{DK} the authors obtained some fundemental properties of the
exponential polynomials and numbers using Euler-Seidel matrices method as:%
\begin{equation}
\phi _{n+1}\left( x\right) =x\sum_{k=0}^{n}\binom{n}{k}\phi _{k}\left(
x\right) \label{10}
\end{equation}%
and%
\begin{equation}
\phi _{n+1}=\sum_{k=0}^{n}\binom{n}{k}\phi _{k}. \label{11}
\end{equation}
Recently, Mezo $\left( \cite{M}\right) $ has defined the "$r-$Bell
polynomials and numbers" as:%
\begin{equation}
B_{n,r}\left( x\right) =\sum_{k=0}^{n}\QATOPD\{ \} {n+r}{k+r}_{r}x^{k}
\label{12}
\end{equation}%
and%
\begin{equation}
\text{ }B_{n,r}=\sum_{k=0}^{n}\QATOPD\{ \} {n+r}{k+r}_{r} \label{12+}
\end{equation}%
respectively. $r-$Exponential polynomials and numbers which we discuss in
the present paper are slightly different than the $r-$Bell polynomials and
numbers in \cite{M}.
\textbf{Geometric polynomials and numbers}
Geometric polynomials are defined in \cite{B, S, ST} as follows:%
\begin{equation}
F_{n}\left( x\right) :=\sum_{k=0}^{n}\QATOPD\{ \} {n}{k}k!x^{k}. \label{13}
\end{equation}
The first few geometric polynomials are:%
\begin{equation}
\begin{tabular}{|l|}
\hline
$F_{0}\left( x\right) =1\text{,}$ \\ \hline
$F_{1}\left( x\right) =x\text{,}$ \\ \hline
$F_{2}\left( x\right) =x+2x^{2}\text{,}$ \\ \hline
$F_{3}\left( x\right) =x+6x^{2}+6x^{3}\text{,}$ \\ \hline
$F_{4}\left( x\right) =x+14x^{2}+36x^{3}+24x^{4}\text{.}$ \\ \hline
\end{tabular}
\label{14}
\end{equation}
Specializing $x=1$ in $\left( \ref{13}\right) $ we get geometric numbers (or
ordered Bell numbers) $F_{n}$ as $\left( \cite{B, ST, W}\right) $:%
\begin{equation}
F_{n}:=F_{n}\left( 1\right) =\sum_{k=0}^{n}\QATOPD\{ \} {n}{k}k!.
\label{i14}
\end{equation}
The first few geometric numbers are:%
\begin{equation}
F_{0}=1\text{, }F_{1}=1\text{, }F_{2}=3\text{, }F_{3}=13\text{, }F_{4}=75%
\text{.} \label{i15}
\end{equation}
Boyadzhiev $\left( \cite{B}\right) $\ introduced the "general geometric
polynomials" as%
\begin{equation}
F_{n,r}\left( x\right) =\frac{1}{\Gamma \left( r\right) }\sum_{k=0}^{n}%
\QATOPD\{ \} {n}{k}\Gamma \left( k+r\right) x^{k}\text{,} \label{16+}
\end{equation}%
where Re$\left( r\right) >0$. In the fifth section we will deal with the
general geometric polynomials.
Exponential and geometric polynomials are connected by the following
integral relation $\left( \cite{B}\right) $%
\begin{equation}
F_{n}\left( z\right) =\int_{0}^{\infty }\phi _{n}\left( z\lambda \right)
e^{-\lambda }d\lambda . \label{16}
\end{equation}
In \cite{DK} the authors also obtained some fundemental properties of the
geometric polynomials and numbers using Euler- Seidel matrices method as:%
\begin{equation}
F_{n+1}\left( x\right) =x\sum_{k=0}^{n}\binom{n+1}{k}F_{k}\left( x\right)
\label{15}
\end{equation}%
and%
\begin{equation}
F_{n}=\sum_{k=0}^{n-1}\binom{n}{k}F_{k}. \label{15+}
\end{equation}
By means of $r-$Stirling numbers Nyul $\left( \cite{GN}\right) $ introduced "%
$r-$geometric polynomials and numbers (or $r-$Fubini or ordered $r-$Bell
polynomials and numbers)" are respectively as follows:%
\begin{equation}
F_{n,r}\left( x\right) =\sum_{k=0}^{n}\left( k+r\right) !\QATOPD\{ \}
{n+r}{k+r}_{r}x^{k} \label{17}
\end{equation}%
and%
\begin{equation*}
F_{n,r}=\sum_{k=0}^{n}\left( k+r\right) !\QATOPD\{ \} {n+r}{k+r}_{r}.
\end{equation*}
In this work $r-$geometric polynomials come up naturally in an application
of the generalized transformation formula as well.
Notice that, our concept of $r-$geometric polynomials is slightly different
than in \cite{GN}.
\textbf{Harmonic and Hyperharmonic numbers}
The $n$-th harmonic number is the $n$-th partial sum of the harmonic series:
\begin{equation}
H_{n}:=\sum_{k=1}^{n}\frac{1}{k}\text{,} \label{18}
\end{equation}
where $H_{0}=0.$
For an integer $\alpha >1$, let
\begin{equation}
H_{n}^{(\alpha )}:=\sum_{k=1}^{n}H_{k}^{(\alpha -1)}\text{,} \label{i18}
\end{equation}%
with $H_{n}^{(1)}:=H_{n}$, be the $n$-th hyperharmonic number of order $%
\alpha $ $\left( \cite{BG, CG}\right) $.
These numbers can be expressed in terms of binomial coefficients and
ordinary harmonic numbers as $\left( \cite{CG, MD}\right) $:%
\begin{equation}
H_{n}^{(\alpha )}=\binom{n+\alpha -1}{\alpha -1}(H_{n+\alpha -1}-H_{\alpha
-1}). \label{20}
\end{equation}
Well-known generating functions of the harmonic and hyperharmonic numbers
are given by%
\begin{equation}
\sum_{n=1}^{\infty }H_{n}x^{n}=-\frac{\ln \left( 1-x\right) }{1-x}
\label{21}
\end{equation}%
and%
\begin{equation}
\sum_{n=1}^{\infty }H_{n}^{\left( \alpha \right) }x^{n}=-\frac{\ln \left(
1-x\right) }{\left( 1-x\right) ^{\alpha }}. \label{22}
\end{equation}%
respectively $\left( \cite{DM}\right) $.
The following relations connect harmonic and hyperharmonic numbers with the
Stirling and $r-$Stirling numbers of the first kind $\left( \cite{BG}\right)
$:%
\begin{equation}
\QATOPD[ ] {k+1}{2}=k!H_{k}\text{,} \label{23}
\end{equation}%
and%
\begin{equation}
k!H_{k}^{\left( r\right) }=\QATOPD[ ] {n+r}{r+1}_{r}\text{.} \label{24}
\end{equation}
\section{Generalization of the transformation formula}
\qquad In this section firstly we mention Boyadzhiev's Theorem $4.1$ in \cite%
{B} and give a useful generalization of it. As a result of this
generalization we introduce $r-$geometric polynomials and numbers.
Suppose we are given an entire function $f$\ and a function $g$, analytic in
a region containing the annulus $K=\{z:r<|z|<R\}$ where $0<r<R$. Hence these
functions have following series expansions,%
\begin{equation*}
f\left( x\right) =\sum_{n=0}^{\infty }p_{n}x^{n}\text{ and }g\left( x\right)
=\sum_{n=-\infty }^{\infty }q_{n}x^{n}.
\end{equation*}%
Now we are ready to state Boyadzhiev's theorem.
\begin{theorem}
\label{Main Theorem}$\left( \cite{B}\right) $ Let the functions $f$ and $g$
be described as above. If the series%
\begin{equation*}
\sum_{n=-\infty }^{\infty }q_{n}f\left( n\right) x^{n}
\end{equation*}%
converges absolutely on $K$, then%
\begin{equation}
\sum_{n=-\infty }^{\infty }q_{n}f\left( n\right) x^{n}=\sum_{m=0}^{\infty
}p_{m}\sum_{k=0}^{m}\QATOPD\{ \} {m}{k}x^{k}g^{\left( k\right) }\left(
x\right) \label{0}
\end{equation}%
holds for all $x\in K$.
\end{theorem}
Stirling numbers of the second kind appear in the formula $\left( \ref{0}%
\right) $ due to $\left( xD\right) $ operator. Our aim is to get a more
general formula than $\left( \ref{0}\right) $\ which contains $r-$Stirling
numbers of the second kind instead of Stirling numbers of the second kind.
Accordingly, first we generalize the operator $\left( xD\right) $.
\subsection{Generalization of the\textbf{\ operator} $\left( xD\right) $}
The operator $\left( xD\right) $ operates a function $f\left( x\right) $ as;%
\begin{equation}
\left( xD\right) f\left( x\right) :=xf^{\prime }\left( x\right) \text{,}
\label{o1}
\end{equation}%
where $f^{\prime }$ is the first derivative of the function $f.$
For any $m$-times differentiable function $f$ we have $\left( \cite{B}%
\right) $,%
\begin{equation}
\left( xD\right) ^{m}f\left( x\right) =\sum_{k=0}^{m}\QATOPD\{ \}
{m}{k}x^{k}f^{\left( k\right) }\left( x\right) . \label{o2}
\end{equation}%
This fact can be easily proven with induction on $m$ by the help of $\left( %
\ref{4}\right) $.
Our first aim is to generalize the operator $\left( xD\right) $.\ Later we
use this generalization to obtain $r-$ geometric and $r-$exponential
polynomials and numbers. In the light of this motivation and after a plenty
of observations we arrive the following definition.
\begin{definition}
\label{GO}Let $f$ be a function which is at least $m-$times differentiable
and $r$ be a nonnegative integer. Then the action $\left( xD_{r}\right) $ is%
\begin{equation}
\left( xD_{r}\right) ^{m}f\left( x\right) :=\left\{
\begin{array}{cc}
0 & ,m<r \\
\left( xD\right) ^{m-r}x^{r}f^{\left( r\right) }\left( x\right) & ,m\geq r%
\end{array}%
\right. . \label{g1}
\end{equation}
\end{definition}
An equivalent statement of this definition is given by the following
proposition.
\begin{proposition}
By applying $\left( xD_{r}\right) $ $m-$times to a function $f$ which is at
least $m-$times differentiable, then the following \ equation holds%
\begin{equation}
\left( xD_{r}\right) ^{m}f\left( x\right) =\sum_{k=0}^{m}\QATOPD\{ \}
{m}{k}_{r}x^{k}f^{\left( k\right) }\left( x\right) \text{,} \label{g2}
\end{equation}%
where $m\geq r$.
\end{proposition}
\begin{proof}
It follows from induction on $m$, in the light of Definition $\ref{GO}$\ and
recurrence relation $\left( \ref{5}\right) $.
\end{proof}
The equation $\left( \ref{g2}\right) $\ is a generalization of the equation $%
\left( \ref{o2}\right) $\ since setting $r=0$ in $\left( \ref{g2}\right) $\
gives the equation $\left( \ref{o2}\right) $.
\begin{corollary}
Let $n$ be an integer, then%
\begin{equation}
\left( xD_{r}\right) ^{m}x^{n}=n^{m-r}\binom{n}{r}r!x^{n}. \label{g3}
\end{equation}
\end{corollary}
\subsection{Generalization of the transformation formula}
\qquad Now we give our main theorem that is a generalization of Theorem 1.
\begin{theorem}
\label{MT}Let $f\left( x\right) $ be an entire function and $g\left(
x\right) $ be an analytic function on the annulus $K=\{z,s<|z|<S\}$, where $%
0\leq s<S$. Suppose that their power series be given as%
\begin{equation*}
f\left( x\right) =\sum_{m=0}^{\infty }p_{m}x^{m}\text{ and }g\left( x\right)
=\sum_{n=-\infty }^{\infty }q_{n}x^{n}.
\end{equation*}%
If the series
\begin{equation}
\sum_{n=-\infty }^{\infty }q_{n}\binom{n}{r}\frac{r!}{n^{r}}f_{r}\left(
n\right) x^{n} \label{ac}
\end{equation}
where $r$ is a nonnegative integer and $f_{r}\left( x\right) $ denotes the
power series $\sum_{m=r}^{\infty }p_{m}x^{m}$, converges absolutely on $K$,
then%
\begin{equation}
\sum_{n=-\infty }^{\infty }q_{n}\binom{n}{r}\frac{r!}{n^{r}}f_{r}\left(
n\right) x^{n}=\sum_{m=r}^{\infty }p_{m}\sum_{k=0}^{m}\QATOPD\{ \}
{m}{k}_{r}x^{k}g^{\left( k\right) }\left( x\right) \label{g4}
\end{equation}%
holds for all $x\in K$.
\end{theorem}
\begin{proof}
By considering the power series expansion of $g\left( x\right) $ with $%
\left( \ref{g2}\right) $ and $\left( \ref{g3}\right) $ we have%
\begin{equation}
\sum_{n=-\infty }^{\infty }q_{n}\binom{n}{r}n^{m-r}r!x^{n}=\sum_{k=0}^{m}%
\QATOPD\{ \} {m}{k}_{r}x^{k}g^{\left( k\right) }\left( x\right) \label{g5}
\end{equation}%
where $m$ and $r$\ are integer such that $m\geq r\geq 0$. If we multiply
both sides of the equation $\left( \ref{g5}\right) $ by $p_{m}$ and sum on $m
$ from $r$ to infinity we get%
\begin{equation*}
\sum_{n=-\infty }^{\infty }q_{n}\binom{n}{r}\frac{r!}{n^{r}}%
\sum_{m=r}^{\infty }p_{m}n^{m}x^{n}=\sum_{m=r}^{\infty
}p_{m}\sum_{k=0}^{m}\QATOPD\{ \} {m}{k}_{r}x^{k}g^{\left( k\right) }\left(
x\right) \text{,}
\end{equation*}%
since $\left( \ref{ac}\right) $\ is converges absolutely on $K$. This
completes proof.
\end{proof}
\begin{corollary}
\label{gsv}Let $g$ be an analytic function on the disk $D=\{z,0\leq |z|<S\}$
then%
\begin{equation}
\sum_{n=r}^{\infty }\frac{g^{\left( n\right) }\left( 0\right) }{n!}\binom{n}{%
r}\frac{r!}{n^{r}}f_{r}\left( n\right) x^{n}=\sum_{n=r}^{\infty }\frac{%
f^{\left( n\right) }\left( 0\right) }{n!}\sum_{k=0}^{n}\QATOPD\{ \}
{n}{k}_{r}x^{k}g^{\left( k\right) }\left( x\right) . \label{g6}
\end{equation}
\end{corollary}
Most of the results in the subsequent sections depend on the Corollary $\ref%
{gsv}$.
\begin{remark}
Specializing $r=0$ in the Theorem $\ref{MT}$\ we turn back to the Theorem $%
4.1$ of Boyadzhiev $\left( \cite{B}\right) $. Therefore from now on we are
interested in the case $r\geq 1$.
\end{remark}
\section{$r-$exponential and $r-$geometric polynomials and numbers}
\qquad Stirling numbers of the first and second kind are notable in many
branches of mathematics, especially in combinatorics, computational
mathematics and computer sciences $\left( \cite{AS, BQ, C, CG, GKP}\right) $%
. Importance of the exponential polynomials and numbers are substantially
because of their direct connection with Striling numbers. $r-$Stirling
numbers $\left( \cite{Br}\right) $ are one of the reputable generalizations
of Stirling numbers. Therefore introduction of the concepts of the $r-$
exponential and $r-$ geometric polynomials and numbers are good motivation
for us.
\subsection{$r-$exponential polynomials and numbers}
Firstly we consider $g\left( x\right) =e^{x}$ in the equation $\left( \ref%
{g6}\right) $. Hence we get%
\begin{equation}
\sum_{n=r}^{\infty }\binom{n}{r}\frac{r!}{n^{r}}f_{r}\left( n\right) \frac{%
x^{n}}{n!}=e^{x}\sum_{n=r}^{\infty }\frac{f^{\left( n\right) }\left(
0\right) }{n!}\sum_{k=0}^{n}\QATOPD\{ \} {n}{k}_{r}x^{k}. \label{g7}
\end{equation}%
The finite sum on the RHS is a generalization of exponential polynomials. We
call these polynomials as "$r-$exponential polynomials" and indicate them
with $_{r}\phi _{n}\left( x\right) $. Hence%
\begin{equation}
_{r}\phi _{n}\left( x\right) :=\sum_{k=0}^{n}\QATOPD\{ \} {n}{k}_{r}x^{k}.
\label{g8}
\end{equation}
The first few $r-$exponential polynomials are:
\begin{equation}
\begin{tabular}{|l|l|l|l|}
\hline
$_{r}\phi _{n}\left( x\right) $ & $r=1$ & $r=2$ & $r=3$ \\ \hline
$n=0$ & $0$ & $0$ & $0$ \\ \hline
$n=1$ & $x$ & $0$ & $0$ \\ \hline
$n=2$ & $x+x^{2}$ & $x^{2}$ & $0$ \\ \hline
$n=3$ & $x+3x^{2}+x^{3}$ & $2x^{2}+x^{3}$ & $x^{3}$ \\ \hline
$n=4$ & $x+7x^{2}+6x^{3}+x^{4}$ & $4x^{2}+5x^{3}+x^{4}$ & $3x^{3}+x^{4}$ \\
\hline
\end{tabular}
\label{Lrep}
\end{equation}
Similar to the classical case, "$r-$exponential numbers" can be defined by
setting $x=1$ in $\left( \ref{g8}\right) $\ i.e,%
\begin{equation}
_{r}\phi _{n}:=\sum_{k=0}^{n}\QATOPD\{ \} {n}{k}_{r}. \label{g8+}
\end{equation}
Hence the first few $r-$exponential numbers are:
\begin{equation}
\begin{tabular}{|l|l|l|l|}
\hline
$_{r}\phi _{n}$ & $r=1$ & $r=2$ & $r=3$ \\ \hline
$n=0$ & $0$ & $0$ & $0$ \\ \hline
$n=1$ & $1$ & $0$ & $0$ \\ \hline
$n=2$ & $2$ & $1$ & $0$ \\ \hline
$n=3$ & $5$ & $3$ & $1$ \\ \hline
$n=4$ & $15$ & $10$ & $4$ \\ \hline
\end{tabular}
\label{Lren}
\end{equation}
Now we give an explicit formula which connects $r-$exponential polynomials
with the classical exponential polynomials. Also this formula allows us to
calculate $_{r}\phi _{n}\left( x\right) $ easily.
\begin{proposition}
We have%
\begin{equation}
_{r}\phi _{n+r}\left( x\right) =x^{r}\sum_{k=0}^{n}\binom{n}{k}r^{n-k}\phi
_{k}\left( x\right) \text{,} \label{g9}
\end{equation}%
where $n$ and $r$ are nonnegative integers.
\end{proposition}
\begin{proof}
Let $m$ be an integer such that $m\geq r\geq 0$ and we specialize $f\left(
x\right) =x^{m}$ in $\left( \ref{g7}\right) $.Then we get%
\begin{equation*}
_{r}\phi _{m}\left( x\right) e^{x}=\sum_{n=r}^{\infty }\binom{n}{r}\frac{r!}{%
n!}n^{m-r}x^{n}.
\end{equation*}%
RHS of this equation can be written as%
\begin{equation*}
x^{r}\sum_{n=0}^{\infty }\left( n+r\right) ^{m-r}\frac{x^{n}}{n!}%
=x^{r}\sum_{k=0}^{m-r}\binom{m-r}{k}r^{m-r-k}\sum_{n=0}^{\infty }n^{k}\frac{%
x^{n}}{n!}.
\end{equation*}%
Considering the definition of the operator $\left( xD\right) $ this becomes%
\begin{equation*}
x^{r}\sum_{k=0}^{m-r}\binom{m-r}{k}r^{m-r-k}\left( xD\right) ^{k}e^{x}.
\end{equation*}%
The equation $\left( \ref{o2}\right) $ enables us to write%
\begin{equation*}
x^{r}e^{x}\sum_{k=0}^{m-r}\binom{m-r}{k}r^{m-r-k}\phi _{k}\left( x\right) .
\end{equation*}%
Comparision of the LHS and the RHS completes the proof.
\end{proof}
Similar relation can be given between classical exponential numbers and $r-$%
exponential numbers as a corollary.
\begin{corollary}
\begin{equation}
_{r}\phi _{n+r}=\sum_{k=0}^{n}\binom{n}{k}r^{n-k}\phi _{k}. \label{g10}
\end{equation}
\end{corollary}
The following corollary shows that the equation $\left( \ref{g9}\right) $ is
a generalization of the equation $\left( \ref{10}\right) $.
\begin{corollary}
\begin{equation*}
\phi _{n+1}\left( x\right) =x\sum_{k=0}^{n}\binom{n}{k}\phi _{k}\left(
x\right) \text{.}
\end{equation*}
\end{corollary}
\subsection{$r-$geometric polynomials and numbers}
By considering $g\left( x\right) =\frac{1}{1-x}$ in the equation $\left( \ref%
{g6}\right) $ we get%
\begin{equation}
\sum_{n=r}^{\infty }\binom{n}{r}\frac{r!}{n^{r}}f_{r}\left( n\right) x^{n}=%
\frac{1}{1-x}\sum_{n=r}^{\infty }\frac{f^{\left( n\right) }\left( 0\right) }{%
n!}\sum_{k=0}^{n}\QATOPD\{ \} {n}{k}_{r}k!\left( \frac{x}{1-x}\right) ^{k}.
\label{rFg}
\end{equation}%
We call the finite sum of the RHS as "$r-$geometric polynomials" and
indicate them with $_{r}F_{n}\left( x\right) $. Hence%
\begin{equation}
_{r}F_{n}\left( x\right) :=\sum_{k=0}^{n}\QATOPD\{ \} {n}{k}_{r}k!x^{k}.
\label{g11}
\end{equation}
The first few $r-$geometric polynomials are:
\begin{equation}
\begin{tabular}{|l|l|l|l|}
\hline
$_{r}F_{n}\left( x\right) $ & $r=1$ & $r=2$ & $r=3$ \\ \hline
$n=0$ & $0$ & $0$ & $0$ \\ \hline
$n=1$ & $x$ & $0$ & $0$ \\ \hline
$n=2$ & $x+2x^{2}$ & $2x^{2}$ & $0$ \\ \hline
$n=3$ & $x+6x^{2}+6x^{3}$ & $4x^{2}+6x^{3}$ & $6x^{3}$ \\ \hline
$n=4$ & $x+14x^{2}+36x^{3}+24x^{4}$ & $8x^{2}+30x^{3}+24x^{4}$ & $%
18x^{3}+24x^{4}$ \\ \hline
\end{tabular}
\label{Lrgp}
\end{equation}
We define "$r-$geometric numbers" by specializing $x=1$ in $\left( \ref{g11}%
\right) $\ as%
\begin{equation}
_{r}F_{n}:=\sum_{k=0}^{n}\QATOPD\{ \} {n}{k}_{r}k!. \label{g12}
\end{equation}
The first few $r-$geometric numbers are:
\begin{equation}
\begin{tabular}{|l|l|l|l|}
\hline
$_{r}F_{n}$ & $r=1$ & $r=2$ & $r=3$ \\ \hline
$n=0$ & $0$ & $0$ & $0$ \\ \hline
$n=1$ & $1$ & $0$ & $0$ \\ \hline
$n=2$ & $3$ & $2$ & $0$ \\ \hline
$n=3$ & $13$ & $10$ & $6$ \\ \hline
$n=4$ & $75$ & $62$ & $42$ \\ \hline
\end{tabular}
\label{Lrgn}
\end{equation}
The following proposition gives an explicit formula between $r-$geometric
polynomials and generalized geometric polynomials which have given by the
equation $\left( \ref{16+}\right) $.
\begin{proposition}
For any nonnegative integers $n$ and $r$ we have%
\begin{equation}
_{r}F_{n+r}\left( x\right) =x^{r}r!\sum_{k=0}^{n}\binom{n}{k}%
r^{n-k}F_{k,r+1}\left( x\right) \label{g13}
\end{equation}
\end{proposition}
\begin{proof}
Let $m$ be a nonnegative integer such that $m\geq r$. By setting $f\left(
x\right) =x^{m}$ in $\left( \ref{rFg}\right) $ we get%
\begin{equation*}
\frac{1}{1-x}\text{ }_{r}F_{m}\left( \frac{x}{1-x}\right)
=\sum_{n=r}^{\infty }\binom{n}{r}r!n^{m-r}x^{n}\text{.}
\end{equation*}%
Rearranging RHS gives%
\begin{equation*}
x^{r}r!\sum_{k=0}^{m-r}\binom{m-r}{k}r^{m-r-k}\sum_{n=0}^{\infty }\binom{n+r%
}{r}n^{k}x^{n}\text{.}
\end{equation*}%
We can write this by means of $\left( xD\right) $\ operator as%
\begin{equation*}
x^{r}r!\sum_{k=0}^{m-r}\binom{m-r}{k}r^{m-r-k}\left( xD\right) ^{k}\frac{1}{%
\left( 1-x\right) ^{r+1}}.
\end{equation*}%
Considering the fact that (equation $\left( 3.26\right) $ in \cite{B})%
\begin{equation*}
\left( xD\right) ^{k}\frac{1}{\left( 1-x\right) ^{r+1}}=\frac{1}{\left(
1-x\right) ^{r+1}}F_{k,r+1}\left( \frac{x}{1-x}\right)
\end{equation*}%
completes the proof.
\end{proof}
A similar result between numbers is as follows.
\begin{corollary}
\begin{equation}
_{r}F_{n+r}=r!\sum_{k=0}^{n}\binom{n}{k}r^{n-k}F_{k,r+1} \label{g14}
\end{equation}
\end{corollary}
Owing to $\left( \ref{g14}\right) $,\ we give the following relations for
classical geometric polynomials and numbers as a corollary.
\begin{corollary}
\begin{equation}
F_{n+1}\left( x\right) =x\sum_{k=0}^{n}\binom{n}{k}F_{k,2}\left( x\right)
\text{,} \label{g15}
\end{equation}%
\begin{equation}
F_{n+1}=\sum_{k=0}^{n}\binom{n}{k}F_{k,2}. \label{g16}
\end{equation}
\end{corollary}
\section{Harmonic $r-$geometric and harmonic $r-$exponential polynomials and
numbers}
\qquad We introduce the concepts of $harmonic-$geometric and $harmonic-$%
exponential polynomials and numbers in $\cite{DK1}$. Along this section we
follow similar approach in $\left( \cite{DK1}\right) $\ to investigate
harmonic $r-$geometric and harmonic $r-$exponential polynomials and numbers.
\subsection{Harmonic $r-$geometric polynomials and numbers}
\qquad We consider the generating function of harmonic numbers\ as the
function $g$\ in the transformation formula $\left( \ref{g6}\right) .$ From
\cite{DK1} we have
\begin{equation}
g^{\left( k\right) }\left( z\right) =\frac{k!\left( H_{k}-\ln \left(
1-z\right) \right) }{\left( 1-z\right) ^{k+1}} \label{r1}
\end{equation}%
and
\begin{equation}
g^{\left( k\right) }\left( 0\right) =k!H_{k}. \label{r2}
\end{equation}
With the help of Theorem $\ref{MT}$ we state the following transformation
formula for harmonic numbers.
\begin{proposition}
Let $r$ be a nonnegative integer and $f$ be an entire function. Then we have%
\begin{eqnarray}
&&\sum_{n=r}^{\infty }\binom{n}{r}H_{n}\frac{r!}{n^{r}}f_{r}\left( n\right)
x^{n} \notag \\
&=&\frac{1}{1-x}\sum_{n=r}^{\infty }\frac{f^{\left( n\right) }\left(
0\right) }{n!}\sum_{k=0}^{n}\QATOPD\{ \} {n}{k}_{r}k!H_{k}\left( \frac{x}{1-x%
}\right) ^{k} \label{r3} \\
&&-\frac{\ln \left( 1-x\right) }{1-x}\sum_{n=r}^{\infty }\frac{f^{\left(
n\right) }\left( 0\right) }{n!}\sum_{k=0}^{n}\QATOPD\{ \} {n}{k}_{r}k!\left(
\frac{x}{1-x}\right) ^{k}. \notag
\end{eqnarray}
\end{proposition}
\begin{proof}
Employing $\left( \ref{r1}\right) $ and $\left( \ref{r2}\right) $ in $\left( %
\ref{g6}\right) $ gives the statement.
\end{proof}
Second part of the RHS of the equation $\left( \ref{r3}\right) $ contains $r-
$geometric polynomials which are familiar to us from the previous section.
But the first part contains a new family of polynomials which is a
generalization of $harmonic-$geometric polynomials $\left( \cite{DK1}\right)
$. We call them as "harmonic $r-$geometric polynomials and indicate them
with $_{r}F_{n}^{h}\left( x\right) $. Thus
\begin{equation}
_{r}F_{n}^{h}\left( x\right) :=\sum_{k=0}^{n}\QATOPD\{ \}
{n}{k}_{r}k!H_{k}x^{k}. \label{r4}
\end{equation}
The first few harmonic $r-$geometric polynomials are:
\begin{equation}
\begin{tabular}{|l|l|l|l|}
\hline
$_{r}F_{n}^{h}\left( x\right) $ & $r=1$ & $r=2$ & $r=3$ \\ \hline
$n=0$ & $0$ & $0$ & $0$ \\ \hline
$n=1$ & $x$ & $0$ & $0$ \\ \hline
$n=2$ & $x+3x^{2}$ & $3x^{2}$ & $0$ \\ \hline
$n=3$ & $x+9x^{2}+11x^{3}$ & $6x^{2}+11x^{3}$ & $11x^{3}$ \\ \hline
$n=4$ & $x+21x^{2}+66x^{3}+50x^{4}$ & $12x^{2}+55x^{3}+50x^{4}$ & $%
33x^{3}+50x^{4}$ \\ \hline
\end{tabular}
\label{Lhrgp}
\end{equation}
"Harmonic $r-$geometric numbers" can be defined by setting $x=1$ in $\left( %
\ref{r4}\right) $, i.e%
\begin{equation}
_{r}F_{n}^{h}:=\sum_{k=0}^{n}\QATOPD\{ \} {n}{k}_{r}k!H_{k}. \label{r4+}
\end{equation}
The first few harmonic $r-$geometric numbers are:
\begin{equation}
\begin{tabular}{|l|l|l|l|}
\hline
$_{r}F_{n}^{h}$ & $r=1$ & $r=2$ & $r=3$ \\ \hline
$n=0$ & $0$ & $0$ & $0$ \\ \hline
$n=1$ & $1$ & $0$ & $0$ \\ \hline
$n=2$ & $4$ & $3$ & $0$ \\ \hline
$n=3$ & $21$ & $17$ & $11$ \\ \hline
$n=4$ & $138$ & $117$ & $83$ \\ \hline
\end{tabular}
\label{Lhrgn}
\end{equation}
Hence with this notation we state the formula $\left( \ref{r3}\right) $
simply as%
\begin{eqnarray}
&&\sum_{n=r}^{\infty }\binom{n}{r}H_{n}\frac{r!}{n^{r}}f_{r}\left( n\right)
x^{n} \notag \\
&=&\frac{1}{1-x}\sum_{n=r}^{\infty }\frac{f^{\left( n\right) }\left(
0\right) }{n!}\left\{ _{r}F_{n}^{h}\left( \frac{x}{1-x}\right)
-_{r}F_{n}\left( \frac{x}{1-x}\right) \ln \left( 1-x\right) \right\} .
\label{r5}
\end{eqnarray}
Due to the following corollary we obtain closed forms of some series related
to harmonic numbers and binomial coefficients.
\begin{corollary}
\begin{equation}
\sum_{n=r}^{\infty }\binom{n}{r}r!n^{m-r}H_{n}x^{n}=\frac{1}{1-x}\left\{
_{r}F_{m}^{h}\left( \frac{x}{1-x}\right) -_{r}F_{m}\left( \frac{x}{1-x}%
\right) \ln \left( 1-x\right) \right\} \text{,} \label{r6}
\end{equation}%
where $m$ and $r$ are integers such that $m\geq r$.
\end{corollary}
\begin{proof}
It follows by setting $f\left( x\right) =x^{m}$ in the equation $\left( \ref%
{r5}\right) $.
\end{proof}
\begin{remark}
Formula $\left( \ref{r6}\right) $ allow us to calculate closed forms of
several harmonic number series. The case $r=1$ in $\left( \ref{r6}\right) $\
has been analyzed in \cite{DK1} already.
\end{remark}
The case $r=2$ gives%
\begin{equation}
\sum_{n=2}^{\infty }n^{m-1}\left( n-1\right) H_{n}x^{n}=\frac{1}{1-x}\left\{
_{2}F_{m}^{h}\left( \frac{x}{1-x}\right) -_{2}F_{m}\left( \frac{x}{1-x}%
\right) \ln \left( 1-x\right) \right\} . \label{gs1}
\end{equation}%
Hence some series and their closed forms that we get from $\left( \ref{gs1}%
\right) $\ are as follows:
For $m=2$ we have%
\begin{equation}
\sum_{n=2}^{\infty }n\left( n-1\right) H_{n}x^{n}=\frac{x^{2}\left\{ 3-2\ln
\left( 1-x\right) \right\} }{\left( 1-x\right) ^{3}}\text{.} \label{gs1a}
\end{equation}
For $m=3$ we have%
\begin{equation}
\sum_{n=2}^{\infty }n^{2}\left( n-1\right) H_{n}x^{n}=\frac{x^{2}\left\{
6+5x-\left( 4+2x\right) \ln \left( 1-x\right) \right\} }{\left( 1-x\right)
^{4}}\text{,} \label{gs1b}
\end{equation}%
and so on.
The case $r=3$ gives%
\begin{eqnarray}
&&\sum_{n=3}^{\infty }n^{m-2}\left( n-1\right) \left( n-2\right) H_{n}x^{n}
\notag \\
&=&\frac{1}{1-x}\left\{ _{3}F_{m}^{h}\left( \frac{x}{1-x}\right)
-_{3}F_{m}\left( \frac{x}{1-x}\right) \ln \left( 1-x\right) \right\} \text{.}
\label{gs2}
\end{eqnarray}%
Hence some series and their closed forms that we get from $\left( \ref{gs2}%
\right) $\ are as follows:
For $m=3$ we have%
\begin{equation}
\sum_{n=3}^{\infty }n\left( n-1\right) \left( n-2\right) H_{n}x^{n}=\frac{%
x^{3}\left\{ 11-6\ln \left( 1-x\right) \right\} }{\left( 1-x\right) ^{4}}%
\text{.} \label{gs3}
\end{equation}
For $m=4$ we have%
\begin{equation}
\sum_{n=3}^{\infty }n^{2}\left( n-1\right) \left( n-2\right) H_{n}x^{n}=%
\frac{x^{3}\left\{ 33+17x-\left( 18+6x\right) \ln \left( 1-x\right) \right\}
}{\left( 1-x\right) ^{5}}\text{,} \label{gs4}
\end{equation}%
and so on.
Now we give a summation formula for the multiple series.
\begin{proposition}
\begin{eqnarray}
&&\sum_{n=r}^{\infty }\left( \sum_{k=0}^{n-r}\binom{k}{r}\binom{n+s-k}{s}%
r!k^{m-r}H_{k}\right) x^{n} \notag \\
&=&\sum_{n=r}^{\infty }\left( \sum_{0\leq k_{1}\leq k_{2}\leq \cdots \leq
k_{s+1}\leq n}\binom{k_{1}}{r}r!k_{1}^{m-r}H_{k_{1}}\right) x^{n}
\label{gs5} \\
&=&\frac{1}{\left( 1-x\right) ^{s+2}}\left\{ _{r}F_{m}^{h}\left( \frac{x}{1-x%
}\right) -_{r}F_{m}\left( \frac{x}{1-x}\right) \ln \left( 1-x\right) \right\}
\notag
\end{eqnarray}
\end{proposition}
\begin{proof}
Multiplying both sides of the equation $\left( \ref{r6}\right) $ with the
Newton binomial series and considering that%
\begin{equation*}
\sum_{k=0}^{n-r}\binom{k}{r}\binom{n+s-k}{s}r!k^{m-r}H_{k}=\sum_{0\leq
k_{1}\leq k_{2}\leq \cdots \leq k_{s+1}\leq n}\binom{k_{1}}{r}%
r!k_{1}^{m-r}H_{k_{1}}
\end{equation*}%
we get the statement.
\end{proof}
By setting $r=2$ and $s=0$ in the formula $\left( \ref{gs5}\right) $ we can
give the following applications:
For $m=2$ we have%
\begin{equation}
\sum_{n=2}^{\infty }\left( \sum_{k=2}^{n}k\left( k-1\right) H_{k}\right)
x^{n}=\frac{x^{2}\left\{ 3-2\ln \left( 1-x\right) \right\} }{\left(
1-x\right) ^{4}}\text{.} \label{gs5a}
\end{equation}
For $m=3$ we have%
\begin{equation}
\sum_{n=2}^{\infty }\left( \sum_{k=2}^{n}k^{2}\left( k-1\right) H_{k}\right)
x^{n}=\frac{x^{2}\left\{ 6+5x-\left( 4+2x\right) \ln \left( 1-x\right)
\right\} }{\left( 1-x\right) ^{5}}\text{,} \label{gs5b}
\end{equation}%
and so on.
\begin{remark}
By the help of $\left( \ref{23}\right) $ we can state $_{r}F_{n}^{h}\left(
x\right) $ and $_{r}F_{n}^{h}$ in terms of $r-$Stirling numbers of the
second kind and Stirling numbers of the first kind as%
\begin{equation}
_{r}F_{n}^{h}\left( x\right) =\sum_{k=0}^{n}\QATOPD\{ \} {n}{k}_{r}\QATOPD[ ]
{k+1}{2}x^{k} \label{r7}
\end{equation}%
and%
\begin{equation}
_{r}F_{n}^{h}=\sum_{k=0}^{n}\QATOPD\{ \} {n}{k}_{r}\QATOPD[ ] {k+1}{2}.
\label{r8}
\end{equation}
\end{remark}
\subsection{Harmonic $r-$exponential polynomials and numbers}
\qquad Bearing in mind the similarity of exponential and geometric
polynomials and being inspried by the definition of harmonic exponential
polynomials and numbers we arrive the following definition.
\begin{definition}
\label{hrb}For the nonnegative integers $n$ and $r,$ "harmonic $r-$%
exponential polynomials" and "harmonic $r-$exponential numbers" are defined
respectively as%
\begin{equation}
_{r}\phi _{n}^{h}\left( x\right) :=\sum_{k=0}^{n}\QATOPD\{ \}
{n}{k}_{r}H_{k}x^{k} \label{r9}
\end{equation}%
and%
\begin{equation}
_{r}\phi _{n}^{h}:=\sum_{k=0}^{n}\QATOPD\{ \} {n}{k}_{r}H_{k}. \label{r10}
\end{equation}
\end{definition}
The first few harmonic $r-$exponential polynomials are:
\begin{equation}
\begin{tabular}{|l|l|l|l|}
\hline
$_{r}\phi _{n}^{h}\left( x\right) $ & $r=1$ & $r=2$ & $r=3$ \\ \hline
$n=0$ & $0$ & $0$ & $0$ \\ \hline
$n=1$ & $x$ & $0$ & $0$ \\ \hline
$n=2$ & $x+\frac{3}{2}x^{2}$ & $\frac{3}{2}x^{2}$ & $0$ \\ \hline
$n=3$ & $x+\frac{9}{2}x^{2}+\frac{11}{6}x^{3}$ & $3x^{2}+\frac{11}{6}x^{3}$
& $\frac{11}{6}x^{3}$ \\ \hline
$n=4$ & $x+\frac{21}{2}x^{2}+11x^{3}+\frac{25}{12}x^{4}$ & $6x^{2}+\frac{55}{%
6}x^{3}+\frac{25}{12}x^{4}$ & $\frac{11}{2}x^{3}+\frac{25}{12}x^{4}$ \\
\hline
\end{tabular}
\label{Lhrep}
\end{equation}
The first few harmonic $r-$exponential numbers are:
\begin{equation}
\begin{tabular}{|l|l|l|l|}
\hline
$_{r}\phi _{n}^{h}$ & $r=1$ & $r=2$ & $r=3$ \\ \hline
$n=0$ & $0$ & $0$ & $0$ \\ \hline
$n=1$ & $1$ & $0$ & $0$ \\ \hline
$n=2$ & $\frac{5}{2}$ & $\frac{3}{2}$ & $0$ \\ \hline
$n=3$ & $\frac{22}{3}$ & $\frac{29}{6}$ & $\frac{11}{6}$ \\ \hline
$n=4$ & $\frac{295}{12}$ & $\frac{69}{4}$ & $\frac{91}{12}$ \\ \hline
\end{tabular}
\label{Lhren}
\end{equation}
\begin{remark}
Definition $\ref{hrb}$ enables us to extend the relation $\left( \ref{16}%
\right) \ $as%
\begin{equation}
_{r}F_{n}^{h}\left( z\right) =\int_{0}^{\infty }\text{ }_{r}\phi
_{n}^{h}\left( z\lambda \right) e^{-\lambda }d\lambda . \label{r11}
\end{equation}
\end{remark}
\section{Hyperharmonic $r-$geometric and hyperharmonic $r-$exponential
polynomials and numbers}
\qquad For the completeness of this work, now we consider hyperharmonic
numbers and their transformations. In this way we could generalize almost
all results of \cite{DK1} and in previous sections of the present paper.
\subsection{Hyperharmonic $r-$geometric polynomials and numbers}
\qquad Similar to the previous section, let us consider the function $g$ in
the transformation formula $\left( \ref{g6}\right) $\ as the generating
function of the hyperharmonic numbers. From \cite{DK1} we have
\begin{equation}
g^{\left( k\right) }\left( x\right) =\frac{\Gamma \left( k+\alpha \right) }{%
\Gamma \left( \alpha \right) }\frac{1}{\left( 1-z\right) ^{\alpha +k}}\left(
H_{k+\alpha -1}-H_{\alpha -1}-\ln \left( 1-x\right) \right) \label{r12}
\end{equation}%
and
\begin{equation}
g^{\left( k\right) }\left( 0\right) =k!H_{k}^{\left( \alpha \right) }.
\label{r13}
\end{equation}%
Now we give a transformation formula for hyperharmonic numbers.
\begin{proposition}
\label{phht}For integers $r\geq 0$ and $\alpha \geq 1$ we have%
\begin{eqnarray}
&&\sum_{n=r}^{\infty }\binom{n}{r}H_{n}^{\left( \alpha \right) }\frac{r!}{%
n^{r}}f_{r}\left( n\right) x^{n} \notag \\
&=&\frac{1}{\left( 1-z\right) ^{\alpha }}\sum_{n=r}^{\infty }\frac{f^{\left(
n\right) }\left( 0\right) }{n!}\sum_{k=0}^{n}\QATOPD\{ \}
{n}{k}_{r}k!H_{k}^{\left( \alpha \right) }\left( \frac{x}{1-x}\right) ^{k}
\label{r14} \\
&&-\frac{\ln \left( 1-x\right) }{\left( 1-z\right) ^{\alpha }}%
\sum_{n=r}^{\infty }\frac{f^{\left( n\right) }\left( 0\right) }{n!}\frac{1}{%
\Gamma \left( \alpha \right) }\sum_{k=0}^{n}\QATOPD\{ \} {n}{k}_{r}\Gamma
\left( k+\alpha \right) \left( \frac{x}{1-x}\right) ^{k}. \notag
\end{eqnarray}
\end{proposition}
\begin{proof}
Consideration $\left( \ref{r12}\right) $ and $\left( \ref{r13}\right) $ in $%
\left( \ref{g6}\right) $ give the statement.
\end{proof}
The first part of the RHS is a generalization of harmonic $r-$geometric
polynomials which contains hyperharmonic numbers instead of harmonic
numbers. We call these polynomials as "hyperharmonic $r-$geometric
polynomials" and indicate them with $_{r}F_{n,\alpha }^{h}\left( x\right) $.
Thus%
\begin{equation}
_{r}F_{n,\alpha }^{h}\left( x\right) =\sum_{k=0}^{n}\QATOPD\{ \}
{n}{k}_{r}k!H_{k}^{\left( \alpha \right) }x^{k} \label{r15}
\end{equation}
The first few hyperharmonic $r-$geometric polynomials are:
Case $\alpha =2$
\begin{equation}
\begin{tabular}{|l|l|l|}
\hline
$_{r}F_{n,2}^{h}\left( x\right) $ & $r=1$ & $r=2$ \\ \hline
$n=0$ & $0$ & $0$ \\ \hline
$n=1$ & $x$ & $0$ \\ \hline
$n=2$ & $x+5x^{2}$ & $5x^{2}$ \\ \hline
$n=3$ & $x+15x^{2}+26x^{3}$ & $10x^{2}+26x^{3}$ \\ \hline
$n=4$ & $x+35x^{2}+156x^{3}+154x^{4}$ & $20x^{2}+130x^{3}+154x^{4}$ \\ \hline
\end{tabular}
\label{hhrgp2}
\end{equation}
Case $\alpha =3$
\begin{equation}
\begin{tabular}{|l|l|l|}
\hline
$_{r}F_{n,3}^{h}\left( x\right) $ & $r=1$ & $r=2$ \\ \hline
$n=0$ & $0$ & $0$ \\ \hline
$n=1$ & $x$ & $0$ \\ \hline
$n=2$ & $x+7x^{2}$ & $7x^{2}$ \\ \hline
$n=3$ & $x+21x^{2}+47x^{3}$ & $14x^{2}+47x^{3}$ \\ \hline
$n=4$ & $x+49x^{2}+282x^{3}+342x^{4}$ & $28x^{2}+235x^{3}+342x^{4}$ \\ \hline
\end{tabular}
\label{hhrgp3}
\end{equation}
The second part of the RHS of $\left( \ref{r14}\right) $\ contains also a
generalization of the polynomials, "general geometric polynomials",\ which
we mention with the equation $\left( \ref{16+}\right) $. We call these
polynomials as "general $r-$geometric polynomials" and indicate them with $%
_{r}F_{n,\alpha }\left( x\right) $. Hence%
\begin{equation}
_{r}F_{n,\alpha }\left( x\right) =\frac{1}{\Gamma \left( \alpha \right) }%
\sum_{k=0}^{n}\QATOPD\{ \} {n}{k}_{r}\Gamma \left( k+\alpha \right) x^{k}.
\label{r16}
\end{equation}
The first few general $r-$geometric polynomials are:
Case $\alpha =2$
\begin{equation}
\begin{tabular}{|l|l|l|}
\hline
$_{r}F_{n,2}\left( x\right) $ & $r=1$ & $r=2$ \\ \hline
$n=0$ & $0$ & $0$ \\ \hline
$n=1$ & $2x$ & $0$ \\ \hline
$n=2$ & $2x+6x^{2}$ & $6x^{2}$ \\ \hline
$n=3$ & $2x+18x^{2}+24x^{3}$ & $12x^{2}+24x^{3}$ \\ \hline
$n=4$ & $2x+42x^{2}+144x^{3}+120x^{4}$ & $24x^{2}+120x^{3}+120x^{4}$ \\
\hline
\end{tabular}
\label{grgp2}
\end{equation}%
and
Case $\alpha =3$
\begin{equation}
\begin{tabular}{|l|l|l|}
\hline
$_{r}F_{n,3}\left( x\right) $ & $r=1$ & $r=2$ \\ \hline
$n=0$ & $0$ & $0$ \\ \hline
$n=1$ & $3x$ & $0$ \\ \hline
$n=2$ & $3x+12x^{2}$ & $12x^{2}$ \\ \hline
$n=3$ & $3x+36x^{2}+60x^{3}$ & $24x^{2}+60x^{3}$ \\ \hline
$n=4$ & $3x+84x^{2}+360x^{3}+360x^{4}$ & $48x^{2}+300x^{3}+360x^{4}$ \\
\hline
\end{tabular}
\label{grgp3}
\end{equation}
With the help of these notations we can state $\left( \ref{r14}\right) $
simply as%
\begin{eqnarray}
&&\sum_{n=r}^{\infty }\binom{n}{r}H_{n}^{\left( \alpha \right) }\frac{r!}{%
n^{r}}f\left( n\right) x^{n} \label{r14+} \\
&=&\frac{1}{\left( 1-z\right) ^{\alpha }}\sum_{n=r}^{\infty }\frac{f^{\left(
n\right) }\left( 0\right) }{n!}\left[ _{r}F_{n,\alpha }^{h}\left( \frac{x}{%
1-x}\right) -_{r}F_{n,\alpha }\left( \frac{x}{1-x}\right) \ln \left(
1-x\right) \right] . \notag
\end{eqnarray}
\begin{remark}
Specializing $x=1$ in $\left( \ref{r15}\right) $\ we get "hyperharmonic $r-$
geometric numbers" as%
\begin{equation}
_{r}F_{n,\alpha }^{h}=\sum_{k=0}^{n}\QATOPD\{ \} {n}{k}_{r}k!H_{k}^{\left(
\alpha \right) }. \label{r17}
\end{equation}
\end{remark}
The first few hyperharmonic $r-$geometric numbers are:
Case $\alpha =2$
\begin{equation}
\begin{tabular}{|l|l|l|}
\hline
$_{r}F_{n,2}^{h}$ & $r=1$ & $r=2$ \\ \hline
$n=0$ & $0$ & $0$ \\ \hline
$n=1$ & $1$ & $0$ \\ \hline
$n=2$ & $6$ & $5$ \\ \hline
$n=3$ & $42$ & $36$ \\ \hline
$n=4$ & $346$ & $304$ \\ \hline
\end{tabular}
\label{hhrfn2}
\end{equation}
Case $\alpha =3$
\begin{equation}
\begin{tabular}{|l|l|l|}
\hline
$_{r}F_{n,3}^{h}$ & $r=1$ & $r=2$ \\ \hline
$n=0$ & $0$ & $0$ \\ \hline
$n=1$ & $1$ & $0$ \\ \hline
$n=2$ & $8$ & $7$ \\ \hline
$n=3$ & $69$ & $61$ \\ \hline
$n=4$ & $674$ & $605$ \\ \hline
\end{tabular}
\label{hhrfn3}
\end{equation}
and specializing $x=1$ in $\left( \ref{r16}\right) $\ gives "general $r-$
geometric numbers" as%
\begin{equation}
_{r}F_{n,\alpha }=\frac{1}{\Gamma \left( \alpha \right) }\sum_{k=0}^{n}%
\QATOPD\{ \} {n}{k}_{r}\Gamma \left( k+\alpha \right) . \label{r18}
\end{equation}
The first few general $r-$geometric numbers are:
Case $\alpha =2$
\begin{equation}
\begin{tabular}{|l|l|l|}
\hline
$_{r}F_{n,2}$ & $r=1$ & $r=2$ \\ \hline
$n=0$ & $0$ & $0$ \\ \hline
$n=1$ & $2$ & $0$ \\ \hline
$n=2$ & $8$ & $6$ \\ \hline
$n=3$ & $44$ & $36$ \\ \hline
$n=4$ & $308$ & $264$ \\ \hline
\end{tabular}
\label{grgn2}
\end{equation}%
and
Case $\alpha =3$
\begin{equation}
\begin{tabular}{|l|l|l|}
\hline
$_{r}F_{n,3}$ & $r=1$ & $r=2$ \\ \hline
$n=0$ & $0$ & $0$ \\ \hline
$n=1$ & $3$ & $0$ \\ \hline
$n=2$ & $15$ & $12$ \\ \hline
$n=3$ & $99$ & $84$ \\ \hline
$n=4$ & $807$ & $708$ \\ \hline
\end{tabular}
\label{grgn3}
\end{equation}
Thanks to the following corollary of Proposition $\left( \ref{phht}\right) $
we have closed forms of some series related to hyperharmonic numbers and
binomial coefficients.
\begin{corollary}
\begin{equation}
\sum_{n=r}^{\infty }\binom{n}{r}r!n^{m-r}H_{n}^{\left( \alpha \right) }x^{n}=%
\frac{1}{\left( 1-z\right) ^{\alpha }}\left[ _{r}F_{m,\alpha }^{h}\left(
\frac{x}{1-x}\right) -_{r}F_{m,\alpha }\left( \frac{x}{1-x}\right) \ln
\left( 1-x\right) \right] . \label{r19}
\end{equation}
\end{corollary}
\begin{proof}
For a positive integers $m\geq r,$ setting $f\left( x\right) =x^{m}$ in $%
\left( \ref{r14+}\right) $ gives $\left( \ref{r19}\right) $.
\end{proof}
\begin{remark}
Specializing the values of $r$, $m$ and $\alpha $ in $\left( \ref{r19}%
\right) $\ one can get closed forms of several hyperharmonic numbers series.
\end{remark}
Now we extend the formula $\left( \ref{gs5}\right) $\ to hyperharmonic
number series.
\begin{proposition}
\begin{eqnarray}
&&\sum_{n=r}^{\infty }\left( \sum_{k=0}^{n-r}\binom{k}{r}\binom{n+s-k}{s}%
r!k^{m-r}H_{k}^{\left( \alpha \right) }\right) x^{n} \notag \\
&=&\sum_{n=r}^{\infty }\left( \sum_{0\leq k_{1}\leq k_{2}\leq \cdots \leq
k_{s+1}\leq n}\binom{k_{1}}{r}r!k_{1}^{m-r}H_{k_{1}}^{\left( \alpha \right)
}\right) x^{n} \label{mhhrs} \\
&=&\frac{1}{\left( 1-x\right) ^{\alpha +s+1}}\left\{ _{r}F_{m}^{h}\left(
\frac{x}{1-x}\right) -_{r}F_{m}\left( \frac{x}{1-x}\right) \ln \left(
1-x\right) \right\} \notag
\end{eqnarray}
\end{proposition}
\begin{proof}
Multiplying both sides of the equation $\left( \ref{r19}\right) $ with the
Newton binomial series gives the statement.
\end{proof}
\begin{remark}
Also special values of $r$, $m$, $s$ and $\alpha $ in $\left( \ref{mhhrs}%
\right) $\ gives closed forms of several mutiplicative hyperharmonic numbers
series.
\end{remark}
\begin{remark}
Using $\left( \ref{24}\right) $ we get an alternative expression of
hyperharmonic $r-$ geometric polynomials and numbers as%
\begin{equation}
_{r}F_{n,\alpha }^{h}\left( x\right) =\sum_{k=0}^{n}\QATOPD\{ \} {n}{k}_{r}%
\QATOPD[ ] {k+\alpha }{\alpha +1}_{\alpha }x^{k}, \label{r20}
\end{equation}%
\begin{equation}
_{r}F_{n,\alpha }^{h}=\sum_{k=0}^{n}\QATOPD\{ \} {n}{k}_{r}\QATOPD[ ] {%
k+\alpha }{\alpha +1}_{\alpha }. \label{r21}
\end{equation}
\end{remark}
\subsection{Hyperharmonic $r-$exponential polynomials and numbers}
\begin{definition}
For positive integers $m$ and $r$ "hyperharmonic $r-$exponential
polynomials" are defined as%
\begin{equation}
_{r}\phi _{n,\alpha }^{h}\left( x\right) =\sum_{k=0}^{n}\QATOPD\{ \}
{n}{k}_{r}H_{k}^{\left( \alpha \right) }x^{k}\text{.} \label{r22}
\end{equation}
\end{definition}
The first few hyperharmonic $r-$exponential polynomials are:
Case $\alpha =2$
\begin{equation}
\begin{tabular}{|l|l|l|}
\hline
$_{r}\phi _{n,2}^{h}\left( x\right) $ & $r=1$ & $r=2$ \\ \hline
$n=0$ & $0$ & $0$ \\ \hline
$n=1$ & $x$ & $0$ \\ \hline
$n=2$ & $x+\frac{5}{2}x^{2}$ & $\frac{5}{2}x^{2}$ \\ \hline
$n=3$ & $x+\frac{15}{2}x^{2}+\frac{13}{3}x^{3}$ & $5x^{2}+\frac{13}{3}x^{3}$
\\ \hline
$n=4$ & $x+\frac{35}{2}x^{2}+26x^{3}+\frac{77}{12}x^{4}$ & $10x^{2}+\frac{65%
}{3}x^{3}+\frac{77}{12}x^{4}$ \\ \hline
\end{tabular}
\label{hhrep2}
\end{equation}
Case $\alpha =3$
\begin{equation}
\begin{tabular}{|l|l|l|}
\hline
$_{r}\phi _{n,3}^{h}\left( x\right) $ & $r=1$ & $r=2$ \\ \hline
$n=0$ & $0$ & $0$ \\ \hline
$n=1$ & $x$ & $0$ \\ \hline
$n=2$ & $x+\frac{7}{2}x^{2}$ & $\frac{7}{2}x^{2}$ \\ \hline
$n=3$ & $x+\frac{21}{2}x^{2}+\frac{47}{6}x^{3}$ & $7x^{2}+\frac{47}{6}x^{3}$
\\ \hline
$n=4$ & $x+\frac{49}{2}x^{2}+47x^{3}+\frac{171}{12}x^{4}$ & $14x^{2}+\frac{%
235}{6}x^{3}+\frac{171}{12}x^{4}$ \\ \hline
\end{tabular}
\label{hhrep3}
\end{equation}
Hence "hyperharmonic $r-$exponential numbers" are defined as%
\begin{equation}
_{r}\phi _{n,\alpha }^{h}=\sum_{k=0}^{n}\QATOPD\{ \} {n}{k}_{r}H_{k}^{\left(
\alpha \right) }. \label{r23}
\end{equation}
The first few hyperharmonic $r-$exponential numbers are:
Case $\alpha =2$
\begin{equation}
\begin{tabular}{|l|l|l|}
\hline
$_{r}\phi _{n,2}^{h}$ & $r=1$ & $r=2$ \\ \hline
$n=0$ & $0$ & $0$ \\ \hline
$n=1$ & $1$ & $0$ \\ \hline
$n=2$ & $\frac{7}{2}$ & $\frac{5}{2}$ \\ \hline
$n=3$ & $\frac{77}{6}$ & $\frac{28}{3}$ \\ \hline
$n=4$ & $\frac{611}{12}$ & $\frac{457}{12}$ \\ \hline
\end{tabular}
\label{hhren2}
\end{equation}
Case $\alpha =3$
\begin{equation}
\begin{tabular}{|l|l|l|}
\hline
$_{r}\phi _{n,3}^{h}$ & $r=1$ & $r=2$ \\ \hline
$n=0$ & $0$ & $0$ \\ \hline
$n=1$ & $1$ & $0$ \\ \hline
$n=2$ & $\frac{9}{2}$ & $\frac{7}{2}$ \\ \hline
$n=3$ & $\frac{58}{3}$ & $\frac{89}{6}$ \\ \hline
$n=4$ & $\frac{347}{4}$ & $\frac{809}{12}$ \\ \hline
\end{tabular}
\label{hhren3}
\end{equation}
\begin{remark}
We extend the relation $\left( \ref{r11}\right) $ as
\end{remark}
\begin{equation*}
_{r}F_{n,\alpha }^{h}\left( z\right) =\int_{0r}^{\infty }\phi _{n,\alpha
}^{h}\left( z\lambda \right) e^{-\lambda }d\lambda .
\end{equation*}
|
\section{Introduction}
\label{kievsky_intro}
Realistic nucleon-nucleon (NN) potentials reproduce the
experimental NN scattering data up to energies
of $350$ MeV with a $\chi^2$ per datum close to 1.
However, the use of these potentials in the description
of the three- and four-nucleon bound and scattering states gives a
$\chi^2$ per datum much larger than 1 (see for example
Refs.\cite{walter,kiev01}).
In order to improve that situation, different three-nucleon
force (TNF) models have been introduced so far. Widely
used in the literature are the Tucson-Melbourne (TM) and the
Urbana IX (URIX) models \cite{tm,urbana}. These models are based
on the exchange mechanism of two pions between three nucleons with the
intermediate excitation of a $\Delta$ resonance.
The TM model has been revisited
within a chiral symmetry approach~\cite{friar:a},
and it has been demonstrated that the contact term present in it
should be dropped. This new TM potential, known as TM$'$, has been
subsequently readjusted~\cite{tmp}. The final operator structure
coincides with that one
given in the TNF of Brazil already derived many years
ago~\cite{brazil}.
Recently, TNFs have been derived based on chiral effective field theory at
next-to-next-to-leading order~\cite{epelbaum02}.
A local version of these interactions (hereafter referred as N2LOL)
can be found in Ref.~\cite{N2LO}.
All these models contain a certain number of parameters that fix the
strength of the interaction. It is a common
practice to determine these parameters from
the three- and four-nucleon binding energies. A particular TNF is in
general associated to a specific NN potential and the sum of the two interactions
forms the nuclear potential energy. The two- and three-nucleon interactions
derived using chiral effective field theory are consistently constructed.
However a particular TNF can be used associated with different NN interactions.
As a consequence, the parametrization of a particular TNF could change
since different NN potentials predict different $A=3,4$ binding energies.
More recently, a new class of two-nucleon interactions has been
obtained ($V_{low-k}$ potentials). With the purpose of
eliminating the high-momentum part of the interaction,
the Hilbert space has been separated
into low and high momentum regions and the renormalization group
method has been used to integrate out the high momentum components
above a cutoff $\Lambda$~\cite{Bog07}.
The value for $\Lambda$ is typically chosen to reproduce
the triton binding energy.
All these potential models can be used to study bound and
scattering states in the $A=3,4$ systems in order to extract
information about their capability to describe the nuclear
dynamics. Besides the bound state energies,
in the $A=3$ system,
the $n-d$ doublet scattering length $^2a_{nd}$ can give valuable information.
In principle this quantity is correlated, to some extent, to the $A=3$ binding
energy through the so-called Phillips line~\cite{phillips,bedaque}.
However the presence of TNFs of the type studied here
breaks this correlation. Therefore $^2a_{nd}$ emerges as an independent
observable. Due to the lack of excited states in the
$A=3$ system, the zero energy state is the first one above the ground state.
In the case of $n-d$ scattering at zero energy, the $J={\frac{1}{2}}^+$ state
is orthogonal to the
triton ground state and, for this reason, it presents a node in the relative
distance between the incident nucleon and the deuteron. The position of the
node is related to the scattering length and it is also
sensitive to the relation between the overall attraction and repulsion of the
interaction. Several of the realistic NN potentials underestimate
the triton binding energy. Therefore by adding a TNF,
with the strength fixed for example to reproduce the triton binding energy,
the balance between the total attraction and repulsion in the potential
changes. This leads to a modification in $^2a_{nd}$ and
this modification depends on the parameters in the TNF.
The determination
of the TNF parametrization able to describe the triton binding energy
$B$($^3$H), the $\alpha$-particle binding energy $B$($^4$He) and $^2a_{nd}$ has
been analyzed in Ref.~\cite{epelbaum02} for a TNF derived from chiral effective
field theory. A similar analysis has not been done for the local TNF models
URIX, TM' and N2LOL.
In Refs.~\cite{report,marcucci09} results for $B$($^3$H), $B$($^4$He), $^2a_{nd}$
are given using different combinations of NN interactions
(see Table~\ref{tb:table1}). Those results indicate
that the models are not able to describe simultaneously the $A=3,4$
binding energies and $^2a_{nd}$.
In order to analyze further the mentioned discrepancies, here we study
potential models constructed summing to the AV18 NN potential~\cite{av18}
the three-nucleon interactions of TM', URIX and N2LOL.
Parametrizations of the URIX and TM'
models already exist in conjunction with the AV18 potential. Conversely the
N2LOL force has been constructed using the N3LO-Idaho potential
from Entem et al.~\cite{entem}.
So, here we adapt its parametrization to reproduce,
when summed to the AV18 interaction, the triton binding energy.
Different parametrizations of the three TNF models are analyzed studying
the description of $B$($^3$H), $B$($^4$He) and $^2a_{nd}$ and
some polarization observables in $p-d$ scattering.
The calculations have been done using the hyperspherical
harmonic (HH) method as given in Refs.~\cite{phh,kiev97,hh4b,hh4s}
to describe bound and scattering states in $A=3,4$ systems using local
potentials. The extension
to treat nonlocal potentials was given in Refs.~\cite{marcucci09,viv06}.
In a different application devoted to study scattering states
in few-nucleon systems, a discussion of the use of the integral
relations derived in Ref.~\cite{intrel} from the Kohn Variational
principle (KVP) is given.
It has been shown that starting from the KVP, the tangent of the phase-shift
can be put in a form of a quotient where both, the numerator and the denominator,
are given in the form of an integral relation. This is similar to what was proposed
in Ref.~\cite{harris}, however its strict relation with the KVP has not been
recognized. To be noticed that a
general formulation of the scattering theory using surface-integrals is given
in Ref.~\cite{kadyrov}.
Here we would like to discuss some specific examples of the integral relations
derived from the KVP. Starting the analysis in the simplest case,
the $A=2$ system, we show that they can be used to compute phase-shifts from
bound state like functions. A second application of the integral
relations regards the possibility of determining
phase-shifts from a calculation in which the Coulomb
potential has been screened. All these examples serve
to demonstrate the general validity of the KVP formulated in terms
of integral relations. Due to their short-range nature, they are determined
by the wave function in the interaction region and not from its
explicit asymptotic behaviour. This means that each wave function
solving $(H-E)\Psi=0$ in the interaction region can be used to determine
the corresponding scattering amplitude even if its asymptotic behaviour is not
the physical one.
\section{The HH expansion for $A=3,4$ systems}
\label{sec:form}
In this section we briefly review the HH method for bound and
scattering states.
\subsection{The HH Method for Bound States}
\label{subsec:bs}
The nuclear wave function for the three-body system can be written as
\begin{equation}
|\Psi\rangle=\sum_\mu c_\mu |\Psi_\mu\rangle \ ,
\label{eq:psi}
\end{equation}
where $|\Psi_\mu\rangle$ is a suitable complete set of
states, and $\mu$ is an index denoting the set of quantum numbers
necessary to completely specify the basis elements.
The coefficients of the expansion can be calculated using the
Rayleigh-Ritz variational principle, which states that
\begin{equation}
\langle\delta_c \Psi\,|\,H-E\,|\Psi\rangle
=0 \ ,
\label{eq:rrvar}
\end{equation}
where $\delta_c \Psi$ indicates the variation of
$\Psi$ for arbitrary infinitesimal
changes of the linear coefficients $c_\mu$. Where
the Hamiltonian of the system consists in the kinetic part plus
two- and three-nucleon interaction terms
\begin{equation}
H=-\frac{\hbar^2}{2m}\sum_i\nabla^2_i +\sum_{i<j}V(i,j)+\sum_{i<j<k}W(i,j,k)
\end{equation}
The problem of determining $c_\mu$ and the energy $E$
is reduced to a generalized eigenvalue problem,
\begin{equation}
\sum_{ \mu'}\,\langle\Psi_\mu\,|\,H-E\,|\, \Psi_{\mu'}\,\rangle \,c_{\mu'}=0
\ .
\label{eq:gepb}
\end{equation}
The main difficulty of the method is to compute the
matrix elements of the Hamiltonian $H$ with respect to the basis states
$|\Psi_\mu\rangle$. Usually $H$ is given as a sum of terms (kinetic energy,
two-body potential, etc.). The calculation of the matrix elements of
some parts of $H$ can be more conveniently performed in coordinate
space, while for other parts it could be easier to work in momentum
space. Therefore, it is important that the basis states
$|\Psi_\mu\rangle$ have simple expressions in both spaces. The
HH functions indeed have such a property.
In the case of three nucleons of mass
$m$ the Jacobi vectors ${\bm x}_{1p},{\bm x}_{2p}$
correspond to a given particle permutation denoted with $p$, which
specifies the particle order
$i,j,k$,
\begin{eqnarray}
{\bm x}_{2p}&=&\frac{1}{\sqrt{2}}({\bm r}_j-{\bm r}_i) \ , \nonumber \\
{\bm x}_{1p}&=&\sqrt{\frac{2}{3}}({\bm r}_k-\frac{1}{2}({\bm r}_i+{\bm r}_j)) \ .
\label{eq:jacc3}
\end{eqnarray}
Here $p=1$ corresponds to the order 1,2,3.
It is convenient to replace the modulii of ${\bm x}_{2p}$ and
${\bm x}_{1p}$ with the so-called hyperradius and
hyperangle, defined as
\begin{eqnarray}
\rho&=&\sqrt{{\bm x}_{1p}^2+{\bm x}_{2p}^2} \ , \label{eq:rho} \\
\tan{\phi_{p}}&=&\frac{x_{1p}}{x_{2p}} \ . \label{eq:hypera}
\end{eqnarray}
Note that $\rho$ does not depend on
the particle permutation $p$.
The complete set of hyperspherical coordinates is then
given by $\{\rho,\Omega^{(\rho)}_p\}$, with
\begin{equation}
\Omega^{(\rho)}_p=[{\hat{{\bm x}}}_{1p},{\hat{{\bm x}}}_{2p};\phi_{p}] \ ,
\label{eq:omegar}
\end{equation}
and the suffix $(\rho)$ recalls the use of the coordinate space.
The expansion states $|\Psi_\mu\rangle$ of
Eq.~(\ref{eq:psi}) are then given by
\begin{equation}
|\,\Psi_\mu^{(\rho)}\,\rangle
= f_l(\rho) {\cal Y}_{ \{G\} }(\Omega^{(\rho)})\ ,
\label{eq:rexp}
\end{equation}
where $f_l(\rho)$ for $l=1,\ldots\,M$ is a complete set of hyperradial
functions, chosen of the form
\begin{equation}
f_l(\rho)=\gamma^{3} \sqrt{\frac{l!}{(l+5)!}}\,\,\,
L^{(5)}_l(\gamma\rho)\,\,{\rm e}^{-\frac{\gamma}{2}\rho} \ .
\label{eq:fllag}
\end{equation}
Here $L^{(5)}_l(\gamma\rho)$ are Laguerre polynomials,
and the non-linear parameter $\gamma$ is
variationally optimized. As an example, for the N3LO-Idaho potential,
it can be chosen in the interval 6--8 fm$^{-1}$.
The functions ${\cal Y}_{ \{G\} }(\Omega^{(\rho)})$ are written
as
\begin{equation}
{\cal Y}_{ \{G\} }(\Omega^{(\rho)})= \sum_{p=1}^{3}\bigg[
Y^{LL_z}_{ [G] }(\Omega^{(\rho)}_p)
\otimes [S_2\otimes \frac{1}{2}]_{S S_z} \bigg]_{J J_z}\,
[T_2\otimes\frac{1}{2}]_{T T_z} \ , \label{eq:hha3}
\end{equation}
where the sum is performed over the three even permutations.
The spin (isospin) of particles $i$ and $j$ are coupled to
$S_2$ ($T_2$), which is itself coupled to the spin (isospin)
of the third particle to give
the state with total spin $S$ (isospin $T,T_z$).
The total orbital angular momentum $L$ and the total
spin $S$ are coupled to the total angular momentum
$J,J_z$.
The functions $Y^{LL_z}_{[G]}(\Omega^{(\rho)}_p)$,
having a definite value of
$L,L_z$, are the HH functions:
\begin{equation}
Y^{LL_z}_{ [G] }(\Omega^{(\rho)}_p) =
\biggl[Y_{\ell_2}({\hat{{\bm x}}}_{2p}) \otimes
Y_{\ell_1}({\hat{{\bm x}}}_{1p}) \biggr]_{LL_z}
N_{[G] }\, ^{(2)}P_{n}^{\ell_1,\ell_2}(\phi_p) \ .
\label{eq:hh3}
\end{equation}
Here $Y_{\ell_1}({\hat{{\bm x}}}_{1p})$ and $Y_{\ell_2}({\hat{{\bm x}}}_{2p})$ are
spherical harmonics, $N_{[G]}$ is a normalization factor and
$^{(2)}P_{n}^{\ell_1,\ell_2}(\phi_p)$ is an hyperspherical polynomial.
The grand angular quantum number $G$ is
defined as $G=2n+\ell_1+\ell_2$.
The notations $[G]$ and $\{G\}$ of Eqs.~(\ref{eq:hh3})
and~(\ref{eq:hha3}) stand for $[\ell_1,\ell_2;n]$ and
$\{\ell_1,\ell_2,L,S_2,T_2$, $S,T;n\}$, respectively,
and $\mu$ of Eq.~(\ref{eq:rexp}) is $\mu=\{G\},l$. Note that
each set of quantum numbers $\{\ell_1,\ell_2,L,S_2,T_2,S,T\}$
is called ``channel'', and the antisymmetrization of
${\cal Y}_{ \{G\} }(\Omega^{(\rho)})$ requires $\ell_2+S_2+T_2$
to be odd. In addition, $\ell_1+\ell_2$ must be even (odd)
for positive (negative) parity.
The HH functions having grand angular quantum number $G$ constructed in
terms of a given set of Jacobi vectors
${\bm x}_{1p},{\bm x}_{2p}$, defined starting from the particle
order $i,j,k$, can always be expressed in terms of
the HH functions constructed, for instance, in terms of
${\bm x}_{1 (p=1)},{\bm x}_{2 (p=1)}$ with the same value of $G$.
In fact, the following relation holds
\begin{equation}
Y^{LL_z}_{[\ell_1,\ell_2;n]}(\Omega^{(\rho)}_p) =
\sum_{\ell_1',\ell_2',n'}
a^{(p),L}_{\ell_1,\ell_2,n;\,\ell_1',\ell_2',n'}
Y^{LL_z}_{[\ell_1',\ell_2';n']}(\Omega^{(\rho)}_{(p=1)})\ ,
\label{eq:rr3}
\end{equation}
where the sum is
restricted to the values $\ell_1'$, $\ell_2'$, and $n'$
such that $\ell_1'+\ell_2'+2n'=G$.
The coefficients $ a^{(p),L}_{\ell_1,\ell_2,n;\,\ell_1',\ell_2',n'}$
relating the two sets of HH functions are known as the Raynal-Revai
coefficients~\cite{RR70}.
Also the spin-isospin states can be recoupled to obtain states where the
spin and isospin quantum numbers are coupled in a given order of the particles.
The result is that the antisymmetric functions
${\cal Y}_{ \{G\} }$ can be expressed as a
superposition of functions constructed in terms of a given order of particles
$i,j,k$, each one having the pair $i$,$j$ in a definite spin and
angular momentum state. When the two-body potential acts on the pair of
particles $i$,$j$, the effect of the projection is easily taken into
account.
The expansion states of Eq.~(\ref{eq:psi})
in momentum space can be obtained as follows.
Let $\hbar{\bm k}_{1p},\hbar{\bm k}_{2p}$ be the conjugate Jacobi momenta
of the Jacobi vectors, given by
\begin{eqnarray}
\hbar{\bm k}_{2p}&=&\frac{1}{\sqrt{2}}({\bm p}_j-{\bm p}_i) \ , \nonumber \\
\hbar{\bm k}_{1p}&=&\sqrt{\frac{2}{3}}({\bm p}_k-\frac{1}{2}({\bm p}_i+{\bm p}_j)) \ ,
\label{eq:jacm3}
\end{eqnarray}
${\bm p}_i$ being the momentum of the $i$-th particle.
We then define a hypermomentum $Q$ and a set of
angular-hyperangular variables as
\begin{eqnarray}
Q&=&\sqrt{{\bm k}_{1p}^2+{\bm k}_{2p}^2} \ , \nonumber \\
\Omega^{(Q)}_p&=&[{\hat{{\bm k}}}_{2p},{\hat{{\bm k}}}_{1p};\varphi_{p}] \ ,
\label{eq:hyperq}
\end{eqnarray}
where
\begin{equation}
\tan{\varphi_{p}}=\frac{k_{1p}}{k_{2p}} \ .
\label{eq:hyperaq}
\end{equation}
Then, the momentum-space version of the wave function
given in Eq.~(\ref{eq:rexp}) is
\begin{equation}
|\,\Psi_\mu^{(Q)}\,\rangle=
g_{ G,l }(Q) {\cal Y}_{ \{G\} }(\Omega^{(Q)}) \ ,
\label{eq:qexp}
\end{equation}
where ${\cal Y}_{ \{G\} }(\Omega^{(Q)})$ is the same as
${\cal Y}_{ \{G\} }(\Omega^{(\rho)})$ of Eq.~(\ref{eq:hha3})
with ${\bm x}_{ip}\rightarrow{\bm k}_{ip}$,
and
\begin{equation}
g_{G,l}(Q)=(-i)^G\,\int_0^\infty d\rho\,
\frac{\rho^{3}}{Q^{2}}\,
J_{G+2}(Q\rho)\, f_{l}(\rho) \ .
\label{eq:vg}
\end{equation}
With the adopted form of $f_l(\rho)$ given in Eq.~(\ref{eq:fllag}),
the corresponding functions $g_{G,l}(Q)$ can be easily calculated,
and they are explicitly given in Ref.~\cite{viv06}.
\subsection{The HH Method for Scattering States Below Deuteron
Breakup Threshold}
\label{subsec:ss}
We consider here the extension of the
HH technique to describe $N-d$ scattering states
below deuteron breakup threshold, when both local
and non-local interaction models are considered.
The wave function $\Psi_{N-d}^{L S J J_z}$ describing the $N-d$
scattering state with incoming orbital angular momentum $L$ and channel spin
$S$, parity $\pi=(-)^L$,
and total angular momentum $J, J_z$,
can be written as
\begin{equation}
\Psi_{N-d}^{LSJJ_z}=\Psi_C^{LSJJ_z}+\Psi_A^{LSJJ_z} \ ,
\label{eq:psica}
\end{equation}
where $\Psi_C^{LSJJ_z}$ describes the system in the region where the particles
are close to each other and their mutual interactions are strong,
while $\Psi_A^{LSJJ_z}$ describes the relative motion between the nucleon $N$
and the deuteron in the asymptotic region, where the $N-d$ nuclear
interaction is negligible. The function $\Psi_C^{LSJJ_z}$, which has to
vanish in the limit of large intercluster
separations, can be expanded on the HH basis as it has been done
in the case of bound states. Therefore,
applying Eq.~(\ref{eq:psi}),
the function $\Psi_C^{LSJJ_z}$ can be casted in the form
\begin{equation}
|\Psi^{LSJJ_z}_C\rangle=\sum_{\mu}\, c_\mu\,
|\Psi_\mu \rangle \ ,
\label{eq:psis}
\end{equation}
where $|\Psi_\mu\rangle$ is defined in Eqs.~(\ref{eq:rexp})
and~(\ref{eq:qexp}) in coordinate- and momentum-space, respectively.
The function $\Psi_A^{LSJJ_z}$ is the appropriate
asymptotic solution of the relative $N-d$ Schr\"odinger equation.
It is written as a linear combination of the following functions,
\begin{equation}
\Omega_{LSJJ_z}^{\lambda}=\sum_{p=1}^3\Omega_{LSJJ_z}^{\lambda}(p) \ ,
\label{eq:psiomp}
\end{equation}
where the sum over $p$ has to be done over the three even permutations
and
\begin{eqnarray}
\Omega_{LSJJ_z}^{\lambda}(p)&=& \sum_{l=0,2}w_l(x_{2p})\,R^\lambda_L(y_p)
\Bigl\{\Bigl[ [Y_l(\hat{{\bm x}}_{2p})\otimes S_2]_1\otimes \frac{1}{2}\Bigr]_S
\nonumber \\ & & \otimes Y_{L}(\hat{{\bm y}}_p) \Bigr \}_{JJ_z}
[ T_2\otimes \frac{1}{2} ]_{TT_z} \ .
\label{eq:psiom}
\end{eqnarray}
Here the spin and isospin quantum numbers of particles $i$ and $j$
have been coupled to $S_2$ and $T_2$, with $S_2=1$, $T_2=0$ for the
deuteron,
$w_l(x_{2p})$ is the deuteron wave function component in the waves $l=0,2$,
${{\bm y}}_p$ is the distance
between $N$ and the center of mass of the deuteron, i.e.
${\bm y}_p=\sqrt{\frac{3}{2}}{\bm x}_{1p}$,
$Y_l(\hat{{\bm x}}_{2p})$ and $Y_{L}(\hat{{\bm y}}_p)$ are the standard spherical
harmonic functions,
and the functions $R^\lambda_L(y_p)$ are the regular ($\lambda\equiv R$)
and irregular ($\lambda\equiv I$) radial solutions of the relative two-body
$N-d$ Schr\"odinger equation without the nuclear interaction.
These regular and irregular functions, denoted as
${\cal F}_L(y_p)$ and ${\cal G}_L(y_p)$ respectively, have the form
\begin{eqnarray}
{\cal F}_L(y_p)&=&\frac{1}{(2L+1)!!q^L C_L(\eta)}\,{F_L(\eta,\xi_p)\over \xi_p} \ ,
\nonumber \\
{\cal G}_L(y_p)&=&(2L+1)!! q^{L+1}C_L(\eta)f_R(y_p){G_L(\eta,\xi_p)\over \xi_p}
\ ,
\label{eq:risol}
\end{eqnarray}
where $q$ is the
modulus of the $N-d$ relative momentum
(related to the total kinetic energy in the center of mass system by
$T_{cm}={q^2\over 2\mu}$, $\mu$ being the $N-d$ reduced mass),
$\eta=2\mu e^2/q$ and $\xi_p=qy_p$ are the usual Coulomb parameters,
and the regular (irregular) Coulomb function $F_L(\eta,\xi_p)$
($G_L(\eta,\xi_p)$) and the
factor $C_L(\eta)$ are defined in the standard
way~\cite{chen:b}. The factor $(2L+1)!! q^L C_L(\eta)$
has been introduced so that ${\cal F}$ and ${\cal G}$
have a finite limit for $q\rightarrow 0$.
The function $f_R(y_p)=[1-\exp(-b y_p)]^{2L+1}$
has been introduced to regularize $G_L$ at small values of $y_p$.
The trial parameter $b$ is
determined by requiring that $f_R(y_p)\rightarrow 1$ outside
the range of the nuclear interaction,
thus not modifying the asymptotic behaviour of the
scattering wave function. A value of $b=0.25$ fm$^{-1}$
has been found appropriate.
The non-Coulomb case of Eq.~(\ref{eq:risol}) is
obtained in the limit $e^2\rightarrow 0$. In this case, $F_L(\eta,\xi_p)/\xi_p$ and
$G_L(\eta,\xi_p)/\xi_p$ reduce to the regular and irregular Riccati-Bessel
functions and
the factor $(2L+1)!!C_L(\eta)\rightarrow 1$ for $\eta\rightarrow 0$.
With the above definitions, $\Psi_A^{LSJJ_z}$ can be written in the form
\begin{equation}
\Psi_A^{LSJJ_z}= \sum_{L^\prime S^\prime}
\bigg[\delta_{L L^\prime} \delta_{S S^\prime}
\Omega_{L^\prime S^\prime JJ_z}^R
+ {\cal R}^J_{LS,L^\prime S^\prime}(q)
\Omega_{L^\prime S^\prime JJ_z}^I \bigg] \ ,
\label{eq:psia}
\end{equation}
where the parameters ${\cal R}^J_{LS,L^\prime S^\prime}(q)$ give the
relative weight between the regular and irregular components
of the wave function. They
are closely related to the reactance matrix (${\cal K}$-matrix)
elements, which can be written as
\begin{eqnarray}
& {\cal K}^J_{LS,L^\prime S^\prime}(q)= &\cr
& (2L+1)!!(2L'+1)!!&q^{L+L'+1}C_L(\eta)C_{L^\prime}(\eta)
{\cal R}^J_{LS,L^\prime S^\prime}(q) \;\;\ .
\end{eqnarray}
By definition of the ${\cal K}$-matrix, its eigenvalues are
$\tan\delta_{LSJ}$, $\delta_{LSJ}$ being the phase shifts.
The sum over $L^\prime$ and $S^\prime$ in Eq.~(\ref{eq:psia}) is over all
values compatible with a given $J$ and parity $\pi$. In particular, the sum
over $L^\prime$ is limited to include either even or odd values since
$(-1)^{L^\prime}=\pi$.
The matrix elements ${\cal R}^J_{LS,L^\prime S^\prime}(q)$ and
the linear coefficients $c_\mu$ occurring in the expansion of $\Psi^{LSJJ_z}_C$
of Eq.~(\ref{eq:psis})
are determined applying the Kohn variational principle,
which states that the functional
\begin{eqnarray}
[{\cal R}^J_{LS,L^\prime S^\prime}(q)]&=&
{\cal R}^J_{LS,L^\prime S^\prime}(q)
- \left \langle \Psi^{L^\prime S^\prime JJ_z }_{N-d} \left |
{\cal L} \right |
\Psi^{LSJJ_z}_{N-d}\right \rangle \ , \nonumber \\
{\cal L}&=&\frac{m}{2\sqrt{3}\hbar^2}(H-E) \ , \label{eq:kohn}
\end{eqnarray}
has to be stationary with respect to variations of the trial parameters
in $\Psi^{LSJJ_z}_{N-d}$.
Here $E$ is the total energy of the system, $m$ is the nucleon mass,
and ${\cal L}$ is chosen so that
\begin{equation}
\langle \Omega^R_{LSJJ_z}| {\cal L} | \Omega^I_{LSJJ_z} \rangle
-\langle \Omega^I_{LSJJ_z}| {\cal L} | \Omega^R_{LSJJ_z} \rangle =1 \ .
\end{equation}
As described in Ref.~\cite{kiev97},
using Eqs.~(\ref{eq:psis}) and~(\ref{eq:psia}),
the variation of the diagonal functionals of Eq.~(\ref{eq:kohn}) with
respect to the linear parameters $c_\mu$ leads to the following
system of linear inhomogeneous equations:
\begin{equation}
\sum_{\mu'} \langle \Psi_\mu| {\cal L} |\Psi_{\mu'}\rangle c_{\mu'} =
-D^\lambda_{LSJJ_z}(\mu) \ .
\label{eq:set1}
\end{equation}
Two different terms $D^\lambda$ corresponding to
$\lambda\equiv R,I$ are introduced and are defined as
\begin{equation}
D^\lambda_{LSJJ_z}(\mu)= \langle \Psi_\mu| {\cal L} |
\Omega^\lambda_{LSJJ_z}\rangle \ .
\label{eq:dlm}
\end{equation}
The matrix elements ${\cal R}^J_{LS,L'S'}(q)$ are obtained
varying the diagonal functionals of Eq.~(\ref{eq:kohn}) with respect to them.
This leads to the following set of algebraic equations
\begin{equation}
\sum_{L'' S''} {\cal R}^J_{LS,L''S''}(q) X_{L'S',L''S''}= Y_{LS,L'S'} \ ,
\label{eq:set2}
\end{equation}
with the coefficients $X$ and $Y$ defined as
\begin{eqnarray}
X_{LS,L'S'}&= \langle
\Omega^I_{LSJJ_z}+\Psi^{LSJJ_z,I}_C| {\cal L} |\Omega^I_{L'S'JJ_z}\rangle \ ,
\nonumber \\
Y_{LS,L'S'}&=-\langle
\Omega^R_{LSJJ_z}+\Psi^{LSJJ_z,R}_C| {\cal L} |\Omega^I_{L'S'JJ_z}\rangle \ .
\label{eq:xy}
\end{eqnarray}
Here $\Psi^{LSJJ_z,\lambda}_C$ is the solution of the set of
Eq.~(\ref{eq:set1}) with the corresponding $D^\lambda$ term. A
second order estimate of ${\cal R}^J_{LS,L'S'}(q)$ is
given by the quantities $[{\cal R}^J_{LS,L'S'}(q)]$, obtained by
substituting in Eq.~(\ref{eq:kohn}) the
first order results. Such second-order calculation provides a symmetric
reactance matrix. This condition is not {\it a priori} imposed,
and therefore it is a useful test of the numerical accuracy.
In the particular case of $q=0$ (zero-energy scattering),
the scattering can occur only in the channel $L=0$ and the observables
of interest are the scattering lengths. Within the
present approach, they can be easily obtained from the relation
\begin{equation}
^{(2J+1)}a_{Nd}=-\lim_{q\rightarrow 0}{\cal R}^J_{0J,0J}(q)\ .
\label{eq:scleng}
\end{equation}
An alternative way to solve the scattering problem, used when
$q\neq 0$, is to apply the complex Kohn variational principle
to the ${\cal S}$-matrix, as in Ref.~\cite{kiev97}.
The approach presented so far for bound and scattering states
does not have too many differences compared to the method
presented for instance in Ref.~\cite{phh}, and known as
pair-correlated hyperspherical harmonics (PHH) method. In fact,
in the PHH method a correlation factor is included in the HH
expansion of Eq.~(\ref{eq:psis}) to take into account the strong
short-range correlations induced by the realistic two-body potentials,
like the AV18. The presence of correlation
functions makes the convergence of the expansion much faster than in the
uncorrelated case. However, the PHH method cannot be simply
implemented when non-local two-body interactions are considered,
unless the Fourier transform of the potential is
performed. The calculation involving $\Psi_C^{LSJJ_z}$
can be performed with the HH or PHH expansions
in coordinate- or in momentum-space, depending
on what is more convenient.
\section{Three Nucleon Force Models}
\label{kievsky_sec:1}
In Ref.~\cite{report} the description of bound states and zero-energy
states for $A=3,4$ has been reviewed in the context of the HH method.
In Table~\ref{tb:table1} we report results for
the triton and $^4$He binding energies as well as for the doublet
$n-d$ scattering length $^2a_{nd}$ using the
AV18 and the N3LO-Idaho NN potentials and using the following
combinations of two- and three-nucleon interactions:
AV18+URIX, AV18+TM', N3LO-Idaho+N2LOL and N3LO-Idaho+URIXp. In this last
model the parameter in front of the spin-isospin independent part of the
URIX potential has been rescaled by a factor of 0.384 to fit the triton binding
energy~\cite{marcucci09} (we call this model URIXp). We have considered also
the $V_{low k}$ model, obtained from the AV18 interaction with a cutoff
parameter $\Lambda=2.2$ fm$^{-1}$.
The results are compared to the experimental values reported
in the table. Worthy of notice is
the recent very accurate datum for $^2a_{nd}$~\cite{doublet}.
\begin{table}[h]
\caption{The triton and $^4$He binding energies $B$ (in MeV),
and doublet scattering length $^2a_{nd}$ (in fm)
calculated using the indicated two- and three-nucleon interactions.
The experimental results are also reported.}
\label{tb:table1}
\begin{tabular}{@{}llll}
\hline
Potential & $B$($^3$H) & $B$($^4$He) & $^2a_{nd}$ \cr
\hline
AV18 & 7.624 & 24.22 & 1.258 \cr
N3LO-Idaho & 7.854 & 25.38 & 1.100 \cr
AV18+TM' & 8.440 & 28.31 & 0.623 \cr
AV18+URIX & 8.479 & 28.48 & 0.578 \cr
N3LO-Idaho+N2LOL & 8.474 & 28.37 & 0.675 \cr
N3LO-Idaho+URIXp & 8.481 & 28.53 & 0.623 \cr
$V_{low-k}$ & 8.477 & 29.15 & 0.572 \cr
\hline
Exp. & 8.482 & 28.30 & 0.645$\pm$0.003$\pm$0.007 \cr
\hline
\end{tabular}
\end{table}
From the table we may observe that only the results obtained
using an interaction model that includes a TNF are close to the
corresponding experimental values. In the case of the AV18+TM', the
strength of the TM' potential has been fixed to reproduce the
$^4$He binding energy and, as can be seen from the table, the
triton binding energy is underpredicted. Conversely,
the strength of the URIX potential
has been fixed to reproduce the triton binding energy giving too much
binding for $^4$He. The strength of the N2LOL potential has been
fixed to reproduce simultaneously the triton and the $^4$He binding
energies whereas the N3LO-Idaho+URIXp model overbinds $^4$He. These two
models give a better description of $^2a_{nd}$. The $V_{low-k}$ interaction
reproduces the triton binding energy but overbinds $^4$He appreciably
and $^2a_{nd}$ is not well described. In conclusion a simultaneous
correct description of the three quantities is not achieved by any of
the models considered.
To analyze further this fact,
we give a brief description of the TM' (or Brazil),
URIX and N2LOL models. They can be put in the following way:
\begin{eqnarray}
W(1,2,3) & = & aW_a(1,2,3)+bW_b(1,2,3)+dW_d(1,2,3) \nonumber \\
W(1,2,3) & = & aW_a(1,2,3)+bW_b(1,2,3)+dW_d(1,2,3) \nonumber \\
& & +c_DW_D(1,2,3)+c_EW_E(1,2,3) \; .
\label{eq:w123}
\end{eqnarray}
Each term corresponds to a different source and has a different operator
structure.
The first three terms arise from the exchange of two pions between three nucleons.
The $a$-term is coming from $\pi N$ $S$-wave scattering
whereas the $b$-term and $d$-term, which are the most important,
come from $\pi N$ $P$-wave scattering. The specific form of these three terms
in configuration space is the following:
\begin{eqnarray}
W_a(1,2,3) & = & \frac{W_0}{c^2\hbar^2}
(\bm\tau_1\cdot\bm\tau_2)(\bm\sigma_1\cdot \bm r_{31})
(\bm\sigma_2\cdot \bm r_{23}) y(r_{31})y(r_{23}) \nonumber \\
W_b(1,2,3) & = & W_0 (\bm\tau_1\cdot\bm\tau_2) [(\bm\sigma_1\cdot\bm\sigma_2)
y(r_{31})y(r_{23}) \nonumber \\
&+ & (\bm\sigma_1\cdot \bm r_{31})
(\bm\sigma_2\cdot \bm r_{23})(\bm r_{31}\cdot \bm r_{23})
t(r_{31})t(r_{23}) \nonumber \\
&+ & (\bm\sigma_1\cdot \bm r_{31})(\bm\sigma_2\cdot \bm r_{31})
t(r_{31})y(r_{23}) \nonumber \\
&+ & (\bm\sigma_1\cdot \bm r_{32})(\bm\sigma_2\cdot \bm r_{32})
y(r_{31})t(r_{23})] \\
W_d(1,2,3) & = & W_0(\bm\tau_3\cdot\bm\tau_1\times\bm\tau_2)
[(\bm\sigma_3\cdot \bm\sigma_2\times\bm\sigma_1)y(r_{31})y(r_{23}) \nonumber \\
&+ & (\bm\sigma_1\cdot \bm r_{31})
(\bm\sigma_2\cdot \bm r_{23})(\bm\sigma_3\cdot\bm r_{31}\times \bm r_{23})
t(r_{31})t(r_{23}) \nonumber \\
&+ & (\bm\sigma_1\cdot \bm r_{31})(\bm\sigma_2\cdot \bm r_{31}\times\bm\sigma_3)
t(r_{31})y(r_{23}) \nonumber \\
&+ & (\bm\sigma_2\cdot \bm r_{32})(\bm\sigma_3\cdot \bm r_{32}\times
\bm\sigma_1) y(r_{31})t(r_{23})] \nonumber \;\; ,
\end{eqnarray}
with $W_0$ an overall strength.
The $b$- and $d$-terms are present in the three models whereas the $a$-term
is present in the TM' and N2LOL and not in URIX. In the first two models,
the radial functions $y(r)$ and $t(r)$ are obtained from the following function
\begin{equation}
f_0(r)=\frac{12\pi}{m_\pi^3}\frac{1}{2\pi^2}
\int_0^\infty dq q^2 \frac{j_0(qr)}{q^2+m_\pi^2}F_\Lambda(q)
\label{eq:f0r}
\end{equation}
where $m_\pi$ is the pion mass and
\begin{equation}
\left\{ \begin{array}{lll}
y(r) & = & \frac{1}{r} f^\prime_0(r) \\
&\mbox{}& \\
t(r) & = & \frac{1}{r} y^\prime(r) \,\,\ .
\end {array}
\right.
\label{eq:y0r}
\end{equation}
The cutoff function $F_\Lambda$
in the TM' or Brazil models is taken as
$[(\Lambda^2-m_\pi^2)/(\Lambda^2+q^2)]^2$. In the N2LOL model it is taken as
$\exp(-q^4/\Lambda^4)$. The momentum cutoff $\Lambda$ is a parameter of the model
fixing the scale of the problem in momentum space.
In the N2LOL, it has been taken $\Lambda=500$ MeV, whereas in the TM' model the quantity
$\Lambda/m_\pi$ has been varied to describe the triton or $^4$He binding energy
at fixed values of the constants $a$,$b$ and $d$. In the literature several cases
have been explored with typical
values around $\Lambda= 5 m_\pi$.
In the URIX model the radial dependence of the $b$- and $d$-terms is given in terms
of the functions
\begin{equation}
\left\{ \begin{array}{lll}
Y(r)& = &{\rm e}^{-x}/x\,\xi_Y \\
&\mbox{}& \\
T(r)& = &(1+3/x+3/x^2)Y(r)\,\xi_T
\end {array}
\right.
\label{eq:Y0r}
\end{equation}
with $x=m_\pi r$ and the cutoff functions are defined as
$\xi_Y=\xi_T=(1-{\rm e}^{-cr^2})$, with $c=2.1$ fm$^{-2}$.
This regularization has been used in the AV18 potential
as well. Since the parameters in the URIX model has been determined in
conjunction with the AV18 potential, the use of the same
regularization was a choice of consistency.
The relation between the functions $Y(r),T(r)$
and those of the previous models is
\begin{equation}
\left\{ \begin{array}{lll}
Y(r) & =& y(r)+T(r) \\
&\mbox{}& \\
T(r) & =& \frac{r^2}{3}t(r)\,\,\, .
\end {array}
\right.
\label{eq:T0r}
\end{equation}
With the definition given in Eq.(\ref{eq:f0r}), the asymptotic behaviour of
the functions $f_0(r)$, $y(r)$ and $t(r)$ is:
\begin{eqnarray}
&f_0(r\rightarrow\infty)&\rightarrow \frac{3}{m_\pi^2}\frac{{\rm e}^{-x}}{x}
\nonumber \\
&y(r\rightarrow\infty)&\rightarrow -\frac{3{\rm e}^{-x}}{x^2}
\left(1+\frac{1}{x}\right) \\
&t(r\rightarrow\infty)&\rightarrow \frac{3}{r^2}\frac{{\rm e}^{-x}}{x}
\left(1+\frac{3}{x}+\frac{3}{x^2}\right) \;\; . \nonumber
\label{eq:f0rasymp}
\end{eqnarray}
In fact, with the normalization chosen for $f_0$, the functions $Y$ and
$T$ defined from $y$ and $t$ in Eq.~(\ref{eq:T0r}) and those ones defined
in the URIX model in Eq.~(\ref{eq:Y0r})
coincide at large separation distances. Conversely, they have a
different short range behavior.
The last two terms in Eq.~(\ref{eq:w123}) correspond to a 2N contact term
with a pion emitted or absorbed ($D$-term) and to a 3N contact interaction
($E$-term). Their local form, in configuration space,
derived from Ref.~\cite{N2LO}, are
\begin{eqnarray}
W_D(1,2,3) & = & W_0^D (\bm\tau_1\cdot\bm\tau_2) \times \nonumber \\
& \{ & \!\! (\bm\sigma_1\cdot\bm\sigma_2)
[y(r_{31})Z_0(r_{23})+Z_0(r_{31})y(r_{23})] \nonumber \\
& + & (\bm\sigma_1\cdot \bm r_{31})(\bm\sigma_2\cdot \bm r_{31})
t(r_{31})Z_0(r_{23}) \nonumber \\
& + &(\bm\sigma_1\cdot \bm r_{32})(\bm\sigma_2\cdot \bm r_{32})
Z_0(r_{31})t(r_{23})\} \\
W_E(1,2,3) & = & W_0^E(\bm\tau_1\cdot\bm\tau_2) Z_0(r_{31})Z_0(r_{23}) \,\, .
\nonumber
\end{eqnarray}
The constant $W_0^D,W_0^E$ fix the strength of these terms.
In the case of the URIX model the $E$-term is present without the isospin operator
structure and it has been included as purely phenomenological, without
justifying its form from a particular exchange mechanism. Its radial dependence
has been taken as $Z_0(r)=T^2(r)$.
In the N2LOL model, the function $Z_0(r)$ is defined as
\begin{equation}
Z_0(r)=\frac{12\pi}{m_\pi^3}\frac{1}{2\pi^2}
\int_0^\infty dq q^2 j_0(qr) F_\Lambda(q)
\label{eq:z0r}
\end{equation}
with the same cutoff function used in the definition of $f_0$ in Eq.(~\ref{eq:f0r}),
$F_\Lambda(q)=\exp(-q^4/\Lambda^4)$.
In the TM' model the $D$- and $E$-terms are absent.
Each model is now identified from the values assigned to the different
constants $a,b,d,c_D,c_E$.
Following Refs.~\cite{tmp,nogga02}, in the case of the TM' model,
the values of the constants have been chosen as $a=-0.87\; m^{-1}_\pi$,
$b=-2.58\; m^{-3}_\pi$, and $d=-0.753\; m^{-3}_\pi$; the strength
$W_0=(gm_\pi/8\pi m_N)^2\;m_\pi^4$ and
the cutoff has been fixed to $\Lambda=4.756\;m_\pi$ in order to describe
correctly $B$($^4$He).
In Table~\ref{tb:table1} the calculations have been
done using these values with $g^2=197.7$, $m_\pi=139.6$ MeV,
$m_N/m_\pi=6.726$ ($m_N$ is the nucleon mass)
as given in the original derivation of the TM potential.
As mentioned before, this model does not include the $D$- and $E$-terms.
In the URIX model the $b$- and $d$-terms are present, however with
a fix relative value. The strength of these terms is:
$bW_0=4\;A^{PW}_{2\pi}$ and $d=b/4$, with $A^{PW}_{2\pi}=-0.0293$ MeV.
The model includes a purely central repulsive term introduced to
compensate the attraction of the previous term, which
by itself would produce a large overbinding in infinite nuclear matter.
It is defined as
\begin{equation}
W_E^{URIX}(1,2,3)=A_R T^2(r_{31})T^2(r_{23})
\end{equation}
with $A_R=0.0048$ MeV.
In the N2LOL potential the constants of the
$a$-, $b$-, $d$-, $D$- and $E$-terms are defined in the following way:
\begin{eqnarray}
& W_0=\frac{1}{12\pi^2}\left(\frac{m_\pi}{F_\pi}\right)^4g^2_A m_\pi^2 \nonumber \\
& W_D^0=\frac{1}{12\pi^2}\left(\frac{m_\pi}{F_\pi}\right)^4
\left(\frac{m_\pi}{\Lambda_x}\right) \frac{g_A m_\pi}{8} \\
& W_E^0=\frac{1}{12\pi^2}\left(\frac{m_\pi}{F_\pi}\right)^4
\left(\frac{m_\pi}{\Lambda_x}\right) m_\pi \nonumber
\end{eqnarray}
with $a= c_1 m^2_\pi$, $b= c_3/2$, $d= c_4/4$, and
$c_1=-0.00081$ MeV$^{-1}$, $c_3=-0.0032$ MeV$^{-1}$, $c_4=-0.0054$ MeV$^{-1}$
taken from Ref.~\cite{entem}. The other two constants, $c_D=1.0$ and $c_E=-0.029$,
have been determined in Ref.~\cite{N2LO} from a fit to $B$($^3$H)
and $B$($^4$He) using the N3LO-Idaho+N2LOL potential model.
The numerical values of the constant entering in
$W_0$, $W^0_D$ and $W^0_E$ are taken as $m_\pi=138$ MeV, $F_\pi=92.4$ MeV,
$g_A=1.29$, and the chiral symmetry breaking scale $\Lambda_x=700$ MeV.
In order to analyze the different short range structure of the TNF models,
in Fig.~\ref{fig:functions} we compare the non-dimensional functions
$Z_0(r)$, $y(r)$ and $T(r)$ for the three models under consideration.
In the TM' model using the definition
of Eq.(\ref{eq:z0r}) and using the corresponding cutoff function we can define:
\begin{eqnarray}
Z^{TM}_0(r) & = & \frac{12\pi}{m_\pi^3}\frac{1}{2\pi^2}
\int_0^\infty dq q^2 j_0(qr) \left(\frac{\Lambda^2-m_\pi^2}{\Lambda^2+q^2}
\right)^2 \nonumber \\
& = & \frac{3}{2}\left(\frac{m_\pi}{\Lambda}\right)
\left(\frac{\Lambda^2}{m_\pi^2}-1\right)^2 {\rm e}^{-\Lambda r} \;\; .
\label{eq:z0rtm}
\end{eqnarray}
This function is showed in the first panel of Fig.~\ref{fig:functions}
as a dashed line.
From the figure we can see that, in the case of the URIX model, the functions
$Z_0(r)$ and $y(r)$ go to zero as $r\rightarrow 0$. This is not the case for the
other two models and is a consequence of the regularization choice of the $Y$
and $T$ functions adopted in the URIX.
\begin{figure*}[!hbt]
\begin{center}
\includegraphics[scale=0.6,angle=0]{functions.eps}
\caption{ The $Z_0(r)$, $y(r)$ and $T(r)$ functions as functions of the
interparticle distance $r$ for the URIX (solid line), TM' (dashed line) and
N2LOL (dotted line) models.}
\label{fig:functions}
\end{center}
\end{figure*}
\section{Parametrization Study of the Three Nucleon Forces}
\label{kievsky_sec:2}
In this section we study possible variations to the parametrization of
the TNF models in order to describe the $A=3,4$ binding energies and
$^2a_{nd}$.
\subsection{Tucson-Melbourne Force}
\label{kievsky_subsec2:1}
We first study the TM' potential and
we would like to see if, using the AV18+TM' interaction, it is possible
to reproduce simultaneously the triton binding energy and the
doublet $n-d$ scattering length for some values of the parameters.
The $a$-term gives a very small contribution to these quantities,
therefore, in the following analysis we maintain it fixed at the value
$a=-0.87\; m^{-1}_\pi$.
In Fig.~\ref{fig:tucson2}, left panel, the doublet $n-d$ scattering length
is given as a function of the parameter $b$ (in units of
its original value $b=-2.58\; m^{-3}_\pi$) for
different values of the cutoff $\Lambda$ (in units of $m_\pi$). The box
in the figure includes values compatible with the experimental results.
The value of the constant $d$ has been fixed to reproduce the triton
binding energy. The corresponding values of
the parameter $d$ (in units of its original value $d=-0.753\; m^{-3}_\pi$)
are given in the right panel as a function of $b$.
Each point of the curves in both panels
corresponds to a set of parameters that, in connection with the AV18 potential,
reproduces the triton binding energy.
The variations of the parameters given in Fig.~\ref{fig:tucson2} do not
exhaust all the possibilities. However we can observe that,
with the AV18+TM' potential, there is a very small region in the parameter's
phase space available for a simultaneous description of
the triton binding energy and the doublet scattering length.
This small region corresponds to a big value of $b$ and $d$ results to be
almost zero. Moreover, the value of the cutoff $\Lambda$ around $3.8m_\pi$ is smaller
than the values usually used with the TM' potential ($\Lambda\approx5m_\pi$).
To be noticed that, for negative values of
the parameters $a$, $b$ and $d$, the TM' potential is attractive. It
does not include explicitly a repulsive term. Added to a specific NN
potential that underpredicts the three-nucleon binding energy,
it supplies the extra binding by fixing appropriately its strength.
However, as mentioned in the Introduction, the scattering length is sensitive to
the balance between the attractive part and the repulsive part of the
complete interaction. Therefore, it seems that supplying only an attraction,
fixed to reproduce the triton binding energy, in the case of the TM'
interaction it is difficult to reproduce correctly this balance.
As discussed before, the TM' potential is a modification of the original
TM potential compatible with chiral symmetry. At the same order
(next-to-next-to-leading order) in the chiral effective field theory
the $D$- and $E$-terms appear (see Ref.~\cite{epelbaum02} and references therein)
as given in Eq.(\ref{eq:w123}).
Here we introduce the following additional term to the TM' potential
based on a contact term of three nucleons
\begin{equation}
W^{TM}_E(1,2,3)=W^0_E\sum_{cyc}Z^{TM}_0(r_{31})Z^{TM}_0(r_{23}) \,\, .
\end{equation}
This term is similar to the repulsive term of the URIX model and,
for the sake of simplicity,
we do not include the $({\bm \tau}_1\cdot{\bm \tau}_2)$ operator.
The function $Z_0^{TM}$ is a positive function, therefore, for positive values of
$c_E$, the new term is repulsive. We include it in the following
analysis of the TM' potential.
The analysis of the new term is given in Fig.~\ref{fig:tucson3}. In the left panel
the doublet $n-d$ scattering length
is given as a function of the parameter $b$ (in units of
its original value $b=-2.58\; m^{-3}_\pi$) for
different values of the strength of the $W_E^{TM}$-term. The value of the
cutoff $\Lambda$ has been fixed to $4.8\;m_\pi$. The box
in the figure includes values compatible with the experimental results.
Moreover, the value of the constant $d$ has been fixed to reproduce the triton
binding energy. The corresponding values of the $^4$He binding energy,
$B(^4{\rm He})$, is given in the right panel.
Comparing the left panels in Figs.~\ref{fig:tucson2} and~\ref{fig:tucson3}, the
effect of the new term is clear. In Fig.~\ref{fig:tucson2} we see that using
$\Lambda=4.8\;m_\pi$, $^2a_{nd}$ is not well reproduced. Conversely,
in Fig.~\ref{fig:tucson3}, the inclusion of the new term moves this curve in
the correct direction and with values of its strength around $c_E=1.6$ it is
possible to reproduce the experimental value of $^2a_{nd}$. There is also an
improvement in the description of $B(^4{\rm He})$. In fact, the AV18+TM'
model with $\Lambda=4.8\;m_\pi$ reproduces the triton binding energy
as can be seen from Fig.~\ref{fig:tucson2}. However it predicts
$B(^4{\rm He})=28.55$ MeV, which is slightly too high. With the $W_E^{TM}$-term,
at $c_E=1.6$, the description of $B(^4{\rm He})$ improves. For example with
$b=-3.87\; m^{-3}_\pi$, $d=-3.375\; m^{-3}_\pi$ and $\Lambda=4.8\;m_\pi$,
we obtain $B(^4{\rm He})=28.36$ MeV, very close to the experimental value.
\begin{figure*}[!htb]
\begin{center}
\vspace{1.5cm}
\includegraphics[scale=0.6,angle=0]{tucson2.eps}
\caption{ The doublet scattering length $a_{n-d}$ as a function of
the parameter $b$ of the TM' potential (right panel) for different values
of the cutoff. The corresponding values of the parameter $d$ used to
reproduce the triton binding energy (left panel).}
\label{fig:tucson2}
\end{center}
\end{figure*}
\begin{figure*}[!htb]
\begin{center}
\vspace{1cm}
\includegraphics[scale=0.6,angle=0]{tucsonL4_8.eps}
\caption{ The doublet scattering length $a_{n-d}$ as a function of
the parameter $b$ of the TM' potential including the $W_E^{TM}$-term,
for different values
of the strength $c_E$ (right panel). The corresponding values of $B(^4{\rm He})$
(left panel).}
\label{fig:tucson3}
\end{center}
\end{figure*}
\subsection{UrbanaIX Force}
\label{kievsky_subsec2:2}
In the following we analyze the URIX potential which has two parameters,
$A^{PW}_{2\pi}$ and $A_R$.
In this model the strength of the $d$-term was related to the strength of
the $b$-term as $b=4d$. The original values of the parameters were fixed in
Ref.~\cite{urbana} in conjunction with the AV18 NN potential and,
from Table~\ref{tb:table1}, we observe that the model correctly describes
the triton binding energy. However, it overestimates $B$($^4$He) and underestimates
$^2a_{nd}$. In order to improve the description of these quantities,
we have varied the constants $A^{PW}_{2\pi}$, $A_R$ and
the relative strength $D^{PW}_{2\pi}=d/b$ between the $b$- and $d$-terms.
For a given value of $A^{PW}_{2\pi}$, the values of
$A_R$ and $D^{PW}_{2\pi}$ has been chosen to reproduce $B(^3{\rm H})$ and
$^2a_{nd}$. The results are given in Fig.~\ref{fig:urbana1}. In panel (a),
$A^{PW}_{2\pi}$ is given as a function of $D^{PW}_{2\pi}$ with $A_R$
varying from $0.0176$ MeV at $A^{PW}_{2\pi}=-0.02$ to
$0.0210$ MeV at $A^{PW}_{2\pi}=-0.050$ MeV. These values of $A_R$ are
more than three times greater than the original value. In panel
(b) and (c) the results for $^2a_{nd}$ and $B(^4{\rm He})$ are given
respectively. The latter has not been included in the determination
of the parameters, however we observe a rather good description in particular
for values of $D^{PW}_{2\pi}>0.7$.
With a modification of the parameters in the URIX force, we were able to describe
$B$($^3$H), $^2a_{nd}$ and $B$($^4$He). This has been achieved with a
substantial increase of the repulsive term. Also $D^{PW}_{2\pi}$ is
quite far from its original value. For example,
at the original value of $A^{PW}_{2\pi}=-0.0293$ MeV,
the relative strength is $D^{PW}_{2\pi}=1$ and $A_R=0.0181$ MeV. This is
four times and more than three of the original values, respectively. As
$D^{PW}_{2\pi}$ diminishes, $A_R$ tends to increase further with the consequence
that the mean value of the repulsive part of $W$ results to be more than three times
the original AV18+URIX value. This is compensated by a lower mean value of the kinetic
energy.
A further analysis of the effects of the new parametrizations
is done in the next section studying selected $p-d$ polarization
observables.
\begin{figure*}[!htb]
\begin{center}
\vspace{1.5cm}
\includegraphics[scale=0.6,angle=0]{urb1.eps}
\caption{ (a) The relative strength $D^{PW}_{2\pi}$ as a function of $A^{PW}_{2\pi}$.
In each point of the curve the triton binding energy and $^2a_{nd}$ are
well described. (b) Values of $^2a_{nd}$ for the seven combinations of the parameters
indicated as solid points in panel (a). (c) The corresponding predictions for
$B$($^4$He). The crosses indicate the results using the parameters defined in the URIX
model}
\label{fig:urbana1}
\end{center}
\end{figure*}
\subsection{N2LOL Force}
The parameters $c_1$, $c_3$ and $c_4$ of the N2LOL have been taken from
the chiral N3LO NN force of Ref.~\cite{entem}, whereas the $c_D$ and $c_E$
parameters have been determined in Ref.~\cite{N2LO}, in conjunction with
that NN force, by fitting $B$($^3$H) and
$B$($^4$He). Here we are going to use the N2LOL force in conjunction with
the AV18 NN interaction, so we have to modify its parametrization since
the amount of attraction to be gained is now different
(see Table~\ref{tb:table1}). Moreover, the modification has to be done in such
a way that $B$($^3$H) and $^2a_{nd}$ are well reproduced. As an example,
in Fig.~\ref{fig:n2lo1}, $^2a_{nd}$ is shown as a function of the parameter
$c_3$ (in units of its original value $c_3=0.0032$ MeV$^{-1}$) fixing
$c_D=0.4,c_E=0.1$ and varying $c_4$ in order to reproduce $B$($^3$H).
With the values $c_3=-0.0048$ MeV$^{-1}$, $c_4=0.0043$ MeV$^{-1}$,
$^2a_{nd}$ fall inside the box and matches the experimental value.
In this case, the$^4$H binding energy results $B(^4{\rm H})=28.36$ MeV.
\section{Polarization observables with the new parametrizations}
In the previous section we have analyzed different parametrizations of the
TM', URIX and N2LOL TNFs determined in conjunction with the AV18 NN potential.
With the new parametrizations the three quantities under observation,
$B$($^3$H), $^2a_{nd}$ and $B(^4{\rm He})$, are well reproduced. However,
some substantial modifications to the first two models were necessary.
In the case of the TM' interaction, we found necessary to include a repulsive term.
In the analysis of the URIX interaction, the strength of the repulsive
term resulted to be more than three times larger. In the case of the
N2LOL interaction, a minor adjustment of the parameters was necessary. Now we would
like to analyze the effects of the new parametrizations in observables that
are not correlated to the binding energies or to $^2a_{nd}$. Some polarization
observables in $p-d$ scattering have this characteristic, in particular the
vector and tensor analyzing powers.
In Fig.~\ref{fig:all1}, the differential cross section $d\sigma/d\Omega$,
the vector polarization observables $A_y$ and $iT_{11}$ and the tensor polarization
observables $T_{20}$, $T_{21}$ and $T_{22}$ are shown at the laboratory energy
$E_{lab}=3$ MeV, for the different potential
models. As a reference we use the AV18+URIX interaction given in the figure
as a blue line.
In the figure, the other three curves corresponds to particular parametrizations
of the models that reproduce $^2a_{nd}$ and $B$($^3$H) and approximate, as much
as possible, $B(^4{\rm He})$. The parametrizations of the models selected for the figure
are the following: the AV18+URIX$^*$ model is defined with
$A_{2\pi}^{PW}=-0.0293$ MeV, $D_{2\pi}^{PW}=1$ and $A_R=0.018$ MeV.
In the AV18+TM$^*$ model we have used
$a=-0.87\; m^{-1}_\pi$, $b=-9.804\; m^{-3}_\pi$, $d=-3.1657\; m^{-3}_\pi$,
$c_E=1$, and $\Lambda=4 m_\pi$.
In the AV18+N2LO$^*$ model the parametrization corresponds to
$c_1=-0.00081$ MeV$^{-1}$ (its original value),
$c_3=-0.0048$ MeV$^{-1}$, $c_4=-0.0043$ MeV$^{-1}$, $c_D=0.4$ and $c_E=0.1$.
From the figure we can observe that the models
describe equally well the differential cross section and the tensor analyzing powers
$T_{20},T_{22}$. Differences are observed in the vector analyzing powers $A_y$ and
$iT_{11}$. Taking as a reference the results of the AV18+URIX model,
in both cases the AV18+URIX$^*$ model produces a noticeable worse description
whereas the AV18+N2LOL$^*$ slightly improves the description. The new
parametrizations of the TNF models overpredict $T_{21}$ in all cases,
in particular the AV18+TM$^*$ model.
\begin{figure}[htb]
\begin{center}
\vspace{2cm}
\includegraphics[scale=0.6,angle=0]{n2lo_CE-0.1.eps}
\caption{ $^2a_{nd}$ as a function of the $c_3$ parameter in the N2LOL model.}
\end{center}
\label{fig:n2lo1}
\end{figure}
\begin{figure*}[htb]
\begin{center}
\vspace{2cm}
\includegraphics[scale=0.7,angle=0]{all1.eps}
\caption{ Cross section, vector and tensor analyzing powers for $p-d$
scattering at $E_{lab}=3$ MeV. Experimental points are for Ref.~\protect\cite{shimizu}}
\end{center}
\label{fig:all1}
\end{figure*}
\section{The Kohn Variational Principle in terms of Integral Relations}
Recently two integral relations have been derived from the KVP~\cite{intrel}.
It has been shown that starting from the KVP, the tangent of the phase-shift
can be expressed in a form of a quotient where both, the numerator and the
denominator, are given as two integral relations.
Let us first consider a two-body system interacting
through a short-range potential $V(r)$ at the center of mass energy $E$
in a relative angular momentum state $l=0$.
The solution of the Schr\"odinger equation in configuration space
($m$ is twice the reduced mass),
\begin{equation}
(-\frac{\hbar^2}{m}\nabla^2+V-E)\Psi(\bm r)=0 \;\; ,
\end{equation}
can be obtained after specifying the corresponding boundary conditions.
For $E>0$, with $k^2=E/(\hbar^2/m)$ and assuming a short-range potential $V$,
$\Psi(\bm r)=\phi(r)/\sqrt{4\pi}$ and
\begin{equation}
\phi(r\rightarrow\infty)\longrightarrow
\sqrt{k} \left[A\frac{\sin(kr)}{kr}+B\frac{\cos(kr)}{kr}\right ] \;\; .
\end{equation}
With the above normalization, the solution $\Psi$ verifies the following
integral relations:
\begin{eqnarray}
-&\frac{m}{\hbar^2} <\Psi|H-E|F>=B
&{\rm with} \hspace{0.5cm} F=\sqrt{\frac{k}{4\pi}} \frac{\sin(kr)}{kr} \cr
&\frac{m}{\hbar^2} <\Psi|H-E|G>=A
&{\rm with} \hspace{0.5cm} G=\sqrt{\frac{k}{4\pi}} \frac{\cos(kr)}{kr} \cr
& \tan\delta = \frac{B}{A} \;\; .
\label{rel1}
\end{eqnarray}
Explicitly they are
\begin{eqnarray}
&-& \frac{m}{\hbar^2\sqrt{k}}\int_0^\infty dr
\sin(kr)V(r)[r\phi(r)]=B \cr
& & \cr
& & \frac{m}{\hbar^2\sqrt{k}}\int_0^\infty dr
\cos(kr)V(r)[r\phi(r)]+\frac{\phi(0)}{\sqrt{k}} =A,
\label{origin}
\end{eqnarray}
where in the last integral we have used the property
$\nabla^2(1/r)=-4\pi\delta({\bm r})$.
In practical cases the solution of the Schr\"odinger equation is obtained
numerically. Then, $\tan\delta$ is extracted from $\phi(r)$ analyzing its
behavior
outside the range of the potential. The equivalence between the extracted
value and that one obtained from the integral relations defines the accuracy
of the numerical computation. A relative difference of the order of
$10^{-7}$ of the two values is usually achieved using standard numerical
techniques to solve the differential equation and to compute the two
one-dimensional integrals. To be noticed the short range character of the
integral relations. This means that the phase-shift is determined by
the internal structure of the wave function.
The last relation in Eq.~\refeq{origin} shows a dependence on the value of the
wave function at the origin. It could be convenient to eliminate this explicit
dependence since the numerical determination of $\phi(0)$ might be problematic,
as we will show. To this end we introduce a regularized function
$\tilde G=f_{reg}G$ with the property $|\tilde G(r=0)|<\infty$ and
$\tilde G=G$ outside the interaction region. A possible choice is
\begin{equation}
\tilde G=\sqrt{\frac{k}{4\pi}}\frac{\cos(kr)}{kr}(1-{\rm e}^{-\gamma r})\;\; ,
\end{equation}
where the regularization function $f_{reg}=(1-{\rm e}^{-\gamma r})$ has been
introduced
with $\gamma$ being a non linear parameter which will be discussed below.
Values verifying $\gamma>1/r_0$,
with $r_0$ the range of the potential, could be appropriate.
The regularized function $\tilde G$ (as well as the irregular function $G$),
verifies the normalization condition
\begin{equation}
\frac{m}{\hbar^2}\left[<F|H-E|\tilde G>-<\tilde G|H-E|F>\right] =1 \;\; .
\label{norm}
\end{equation}
Therefore the second integral relation in Eq.~\refeq{rel1} remains valid using
$\tilde G$ in place of $G$,
\begin{equation}
\frac{m}{\hbar^2} <\Psi|H-E|\tilde G>=A \;\; ,
\label{reg1}
\end{equation}
with the explicit form:
\begin{equation}
\frac{m}{\hbar^2\sqrt{k}}\int_0^\infty d{r}
\cos(kr)V(r)[r\phi(r)]+ I_\gamma =A
\label{reg2}
\end{equation}
where in $I_\gamma$ all terms depending on $\gamma$, introduced by
$f_{reg}$, are included. Comparing Eq.~\refeq{reg2} to Eq.~\refeq{origin}
we identify $I_\gamma=\phi(0)/\sqrt{k}$.
In the following we demonstrate that the relation $\tan\delta=B/A$, which is
an exact relation when the exact wave function $\Psi$ is used in Eq.~\refeq{rel1},
can be considered accurate up to second order when a trial wave function is
used, as it has a strict connection with the Kohn variational principle.
The connection of the integral relations with the KVP is straightforward.
Defining a trial wave function $\Psi_t$ as
\begin{equation}
\Psi_t=\Psi_c +AF+B\;\tilde G \;\; ,
\label{psic}
\end{equation}
with $\Psi_c\rightarrow 0$ as
$r\rightarrow\infty$, the condition
$\Psi_t\rightarrow A F+B \; G$ as $r\rightarrow\infty$ is fulfilled. The KVP
states that the second order estimate for $\tan\delta$ is
\begin{equation}
[\tan\delta]^{2^{nd}}=\tan\delta - \frac{m}{\hbar^2}<(1/A)\Psi_t|H-E|(1/A)\Psi_t> \,\ .
\label{kohnn}
\end{equation}
The above functional is stationary with respect to variations
on $\Psi_c$ and $\tan\delta$. Without loosing generality $\Psi_c$ can be
expanded in a (square integrable) complete basis
\begin{equation}
\Psi_c=\sum_n a_n \phi_n(r) \;\; .
\end{equation}
The variation of the functional with respect to the linear
parameters $a_n$
and $\tan\delta$ leads to the following equations
\begin{eqnarray}
&<\phi_n|H-E|\Psi_t>=0 \cr
& \cr
&<\tilde G|H-E|\Psi_t>=0 \;\;\; .
\label{first}
\end{eqnarray}
To obtain the last equation, the normalization relation of Eq.~\refeq{norm}
has been used.
From these two equations, $\Psi_c$ and the first order
estimate of the phase shift $(\tan\delta)^{1^{st}}$ can be determined.
To be noticed that the first equation implies
$<\Psi_c|H-E|\Psi_t>=0$. Furthermore, from the general relation
$(m/{\hbar^2})\left[<\Psi_t|H-E|\tilde G>-
<\tilde G|H-E|\Psi_t>\right]=A$, and using
the second equation in Eq.~\refeq{first}, the following integral relation results
\begin{equation}
\frac{m}{\hbar^2}<\Psi_t|H-E|\tilde G>=A \;\; .
\end{equation}
Replacing the two relations of Eq.\refeq{first} into the functional of
Eq.\refeq{kohnn}, a second order estimate of the phase shift is obtained
\begin{equation}
[\tan\delta]^{2^{nd}}=(\tan\delta)^{1^{st}}
- \frac{m}{\hbar^2}<F|H-E|(1/A)\Psi_t> \,\ .
\label{second1}
\end{equation}
Multiplying Eq.~\refeq{second1} by $A$ one gets
\begin{equation}
B^{2^{nd}}=B^{1^{st}}
- \frac{m}{\hbar^2}<F|H-E|\Psi_t> \,\ .
\label{second2}
\end{equation}
On the other hand, a first order estimate for the coefficient $B$ can be obtained
from the general relation
\begin{equation}
\frac{m}{\hbar^2}\left[<F|H-E|\Psi_t>-
<\Psi_t|H-E|F>\right]=B^{1^{st}} \,\,\ .
\label{firstb}
\end{equation}
Therefore,
replacing Eq.\refeq{firstb} in Eq.\refeq{second2}, a second order
integral relation for $B$ is obtained. The above results can be
summarized as follow
\begin{eqnarray}
B^{2^{nd}}& = & -\frac{m}{\hbar^2}<\Psi_t|H-E|F> \cr
&& \cr
A & = & \frac{m}{\hbar^2}<\Psi_t|H-E|\tilde G> \cr
&& \cr
[\tan\delta]^{2^{nd}} & = & B^{2^{nd}}/A \,\, .
\label{relint}
\end{eqnarray}
These equations extend the validity of the integral relations,
given in Eq.\refeq{rel1} for the exact wave functions, to trial
wave functions. To be noticed that $F,\tilde G$ are solutions
of the Schr\"odinger equation in the asymptotic region, therefore
$(H-E)F\rightarrow 0$ and $(H-E)\tilde G\rightarrow 0$
as the distance between the particles increases.
As a consequence the decomposition of $\Psi_t$ in
the three terms of Eq.~\refeq{psic} can be considered formal since,
due to the short-range character
of the relation integrals, it is sufficient that the trial wave function
be a solution of $(H-E)\Psi_t=0$ in the interaction region, without
an explicit indication of its asymptotic behavior. This fact,
together with the variational character
of the relations allows for a number of applications to be discussed
in the next sections.
\section{Integral Relations for $A=2,3$ systems}
Applications of the integral relations to systems with $A=2,3$
are given. We first consider the following central,
$s$-wave gaussian potential
\begin{equation}
V(r)=-V_0\exp{(-r^2/r_0^2)} \;\; ,
\end{equation}
with $V_0=-51.5$ MeV, $r_0=1.6$ fm and $\hbar^2/m=41.4696$ MeV fm$^2$.
This potential has a shallow $L=0$ bound state with energy
$E_{2B}=-0.397743$ MeV.
In the $A=2$ system, the orthogonal basis
\begin{equation}
\phi_m={\cal L}_m^{(2)}(z)\exp{-(z/2)} \;\; ,
\end{equation}
with ${\cal L}_m$ a (normalized)
Laguerre polynomial and $z=\beta r$, being $\beta$
a nonlinear parameter, is used to expand the wave function of the system
\begin{equation}
\Psi_0=\sum_{m=0}^{M-1} a^0_m \phi_m \,\, .
\end{equation}
The Schr\"odinger equation is transformed to an eigenvalue problem that can
be solved for different values of the dimension $M$ of the
basis. The variational principle states that
\begin{equation}
E_0=\bra \Psi_0|H|\Psi_0 \ket \ge E_{2B} \;\; ,
\end{equation}
with the equality obtained for $M\rightarrow\infty$.
The nonlinear parameter $\beta$ can be fixed to make improve the
convergence properties of the basis. In fact,
for each value of $M$ there is a value of $\beta$ that minimizes the energy.
Increasing $M$, the minimum of the energy becomes less dependent
on $\beta$ resulting in a plateau.
Increasing further the dimension of the basis, the extension
of the plateau increases as well, without any appreciable improvement in
the eigenvalue, indicating that the convergence has been reached up to
certain accuracy. At each step $\Psi_0$
represents a first order estimate of the bound state exact wave function.
In the proposed example the system has only one bound state. So,
with proper values of $M$ and $\beta$,
the diagonalization of $H$ results in one negative eigenvalue
$E_0$ and $M-1$ positive eigenvalues $E_j$ ($j=1,....,M-1$). The
corresponding wave functions
\begin{equation}
\Psi_j=\sum_{m=0}^{M-1} a^j_m \phi_m \hspace{0.5cm} j=1,....,M-1 \;\; ,
\end{equation}
are approximate solutions of $(H-E_j)\Psi_j=0$ in the interaction region.
As $r\rightarrow\infty$
they go to zero exponentially and therefore they do not represent a physical
scattering state.
The negative energy $E_0$ and the first three positive energy eigenvalues
($E_j$, $j=1,3$)
are shown in Fig.~\ref{fig:itfig1} as a function of $\beta$ in the case of $M=40$.
We observe the plateau already reached by $E_0$ for the values
of $\beta$ showed in the figure. We observe also the monotonic
behavior of the positive eigenvalues toward zero as $\beta$ decreases.
The corresponding eigenvectors
can be used to compute the integral relations of Eq.~\refeq{relint} and to
calculate the second order estimate of the phase-shifts
$\delta_j$ at the specific energies $E_j$.
This analysis is shown in Table~\ref{tab:ittab1} in which the
non linear parameter $\beta$ of the Laguerre basis has been fixed to $1.2$ fm$^{-1}$.
In the first row of the
table the ground state energy is given for different values of the
number $M$ of Laguerre polynomials. The stability of $E_0$ at the
level of $1$ keV is achieved already with $M=20$.
For a given value of $M$, $E_j$, with $j=1,2,3$,
are the first three positive eigenvalues.
The eigenvectors corresponding to positive energies approximate
the scattering states at the specific energies. Since the lowest
scattering state appears at zero energy, none of the positive eigenvalues
can reach this value for any finite values of $M$.
Defining
$ k^2_j=\frac{m}{\hbar^2}E_j$, the second order estimate for the phase
shift at each energy and at each value of $M$ is obtained as
\begin{eqnarray}
-&\frac{m}{\hbar^2} <\Psi_j|H-E|F_j>=B_j
&{\rm with} \hspace{0.2cm} F_j=\sqrt{\frac{k_j}{4\pi}} \frac{\sin(k_jr)}{k_jr} \cr
\cr
&\frac{m}{\hbar^2} <\Psi_j|H-E|\tilde G_j>=A_j
&{\rm with} \hspace{0.2cm} \tilde G_j=f_{reg}\sqrt{\frac{k_j}{4\pi}}
\frac{\cos(k_jr)}{k_jr} \cr
\cr
& [\tan\delta_j]^{2^{nd}} = B_j/A_j.
\label{rel2}
\end{eqnarray}
\begin{table}[h]
\caption{The two-nucleon bound state $E_0$ and the first three positive eigenvalues
$E_j$ $(j=1,3)$, as a function of the number of Laguerre polynomials $M$.
The second order estimates, $[\tan\delta_j]^{2^{nd}}$, obtained applying the integral
relations are given in each case and compared to exact results, $\tan\delta_j$.}
\begin{tabular}{lcccc}
\hline
M & 10 & 20 & 30 & 40 \cr
\hline
$E_0$ &-0.395079 &-0.397740 &-0.397743 &-0.397743 \cr
\hline
$E_1$ & 0.536349 & 0.116356 & 0.048091 & 0.026008 \cr
$[\tan\delta_1]^{2^{nd}}$
&-1.507280 &-0.622242 &-0.392005 &-0.286479 \cr
$\tan\delta_1$
&-1.522377 &-0.621938 &-0.392021 &-0.286480 \cr
\hline
$E_2$ & 1.984580 & 0.449655 & 0.190019 & 0.103503 \cr
$[\tan\delta_2]^{2^{nd}}$
&-5.919685 &-1.353736 &-0.812313 &-0.584389 \cr
$\tan\delta_2$
&-5.703495 &-1.354691 &-0.812270 &-0.584388 \cr
\hline
$E_3$ & 4.512635 & 0.994433 & 0.423117 & 0.231645 \cr
$[\tan\delta_3]^{2^{nd}}$
&13.998124 &-2.451174 &-1.302799 &-0.908128 \cr
$\tan\delta_3$
&12.684474 &-2.448343 &-1.302887 &-0.908131 \cr
\hline
\end{tabular}
\label{tab:ittab1}
\end{table}
\begin{figure}
\begin{center}
\vspace{1.0cm}
\includegraphics[scale=0.35,angle=0]{result.eps}
\caption{The two-nucleon bound state energy $E_0$ and the first three positive
eigenvalues $E_j$ as a function of $\beta$ in the case of $M=40$}
\label{fig:itfig1}
\end{center}
\end{figure}
On the other hand, as we are considering the $A=2$ system,
at each specified energy $E_j$ the phase shift $\tan\delta_j$
can be obtained
by solving the Schr\"odinger equation numerically. The two values,
$[\tan\delta_j]^{2^{nd}}$ and $\tan\delta_j$, are given in the Table~\ref{tab:ittab1}
at the corresponding energies as a function of $M$. We observe that, as
$M$ increases, the relative difference between the variational estimate
and the exact value reduces, for example at $M=40$ is around $10^{-6}$. In fact,
as $M$ increases, each eigenvector gives a better representation of
the exact wave function in the internal region and the second order
estimates, $[\tan\delta_j]^{2^{nd}}$ approach the exact result.
In a different application, the integral relations can be used to calculate
the phase-shift of a process in which the two particles
interact through a short range potential plus the Coulomb potential,
imposing free asymptotic conditions to the wave function.
As an example we use the same two body potential used in the previous analysis
and add the Coulomb potential:
\begin{equation}
V(r)=-V_0\exp{-(r/r_0)^2}+ \frac{e^2}{r} \,\,\, .
\label{potc}
\end{equation}
For positive energies and $l=0$, the wave function behaves asymptotically as
\begin{equation}
\Psi^{(c)}(r\rightarrow\infty)= AF_c(r)+BG_c(r)\;\; ,
\label{asympc}
\end{equation}
with $F_c(r),G_c(r)$ the regular and irregular Coulomb functions, respectively.
The phase-shift is $\tan\delta_c=B/A$. The KVP remains valid when the long range
Coulomb potential is considered and its form in terms of the integral
relations results:
\begin{eqnarray}
-&\frac{m}{\hbar^2} <\Psi^{(c)}_t|H-E|F_c>=B \cr
\cr
&\frac{m}{\hbar^2} <\Psi^{(c)}_t|H-E|\tilde G_c>=A \cr
\cr
& [\tan\delta_c]^{2^{nd}} = \frac{B}{A}\;\; .
\label{rel3}
\end{eqnarray}
with $\tilde G_c=f_{reg}G_c$ and $\Psi^{(c)}_t$ a trial wave function
behaving asymptotically as $\Psi^{(c)}$. Since
$(H-E)|F_c>$ and $(H-E)|\tilde G_c>$ go to
zero outside the range of the short range potential, the integrals in
Eq.~\refeq{rel3} are negligible outside that region. Therefore, for
the computation of the phase-shift it is enough to require that
$\Psi^{(c)}_t$ verifies
$(H-E)\Psi^{(c)}_t=0$, inside that region. To exploit this fact, we introduce the
following screened potential:
\begin{equation}
V_{sc}(r)
=-V_0\exp{[-(r/r_0)^2]}+ \left[{\rm e}^{-(r/r_{sc})^n}\right]\frac{e^2}{r}\;\; .
\end{equation}
For specific values of $n$ and $r_{sc}$ it has the property of being
extremely close to the potential $V(r)$ of Eq.~\refeq{potc}
for $r<r_0$, with $r_0$ the range of the short
range potential. The screening factor ${\rm e}^{-(r/r_{sc})^n}$ cuts the
Coulomb potential for $r>r_{sc}$.
Using the potential $V_{sc}$ to describe a scattering process,
the wave function behaves asymptotically as
\begin{equation}
\Psi_{n,r_{sc}}(r\rightarrow\infty)= AF(r)+BG(r)
\end{equation}
with $F,G$ from Eq.~\refeq{rel2},
since $V_{sc}$ is a short range potential. Solving the Schr\"odinger
equation for this potential,
it is possible to obtain the wave function $\Psi_{n,r_{sc}}$
for different values of $n$ and $r_{sc}$.
This wave function can be considered as a trial wave function for the problem
in which the Coulomb potential is unscreened. Accordingly it can be used as input
in Eq.~\refeq{rel3} to obtain a second order estimate
of the Coulomb phase-shift,
\begin{eqnarray}
-&\frac{m}{\hbar^2} <\Psi_{n,r_{sc}}|H-E|F_c>=B \cr
\cr
&\frac{m}{\hbar^2} <\Psi_{n,r_{sc}}|H-E|\tilde G_c>=A \cr
\cr
& [\tan\delta_c]^{2^{nd}} = \frac{B}{A}
\label{rel4}
\end{eqnarray}
where in $H$ the unscreened Coulomb potential is considered.
This estimate depends
on $n$ and $r_{sc}$ as the wave function does. In Fig.~\ref{fig:itfig3} the
second order estimate $[\tan\delta_c]^{2^{nd}}$ is shown as a function of
$r_{sc}$ for different values of $n$. The straight line is the exact value
of $\tan\delta_c$ obtained solving the Schr\"odinger equation.
We can observe that for $n\ge4$ and $r_{sc}>30$ fm
the second order estimate coincides with the exact results. In this example
the integral relations derived from the Kohn Variational Principle have
been used to extract a phase-shift in presence of the Coulomb potential
using wave functions with free asymptotic conditions.
\begin{figure}
\begin{center}
\vspace{1.0cm}
\includegraphics[scale=0.35,angle=0]{example2.eps}
\caption{The two-nucleon second order estimate $[\tan\delta_c]^{2^{nd}}$
as a function of $r_{sc}$ for different values of $n$. As a reference
the exact value for $\tan\delta_c$ is given as a straight line. }
\label{fig:itfig3}
\end{center}
\end{figure}
Finally an application of the integral relations to the $A=3$ system is
discussed. To this end we give the generalization of the integral
relations to the case in which more than one channel is open.
The coefficients $A$ and $B$ of Eq.~\refeq{rel2}
correspond to matrices
\begin{eqnarray}
B_{ij}&=&-\frac{m}{\hbar^2}<\Psi_i|H-E|F_j> \cr
\cr
A_{ij}&=&\frac{m}{\hbar^2} <\Psi_i|H-E|{\widetilde G}_j> \cr
\cr
R^{2^{nd}}&=&A^{-1}B. \; \;\;\
\label{secondij}
\end{eqnarray}
with $R^{2^{nd}}$ the second order estimate of the scattering matrix
whose eigenvalues are the phase shifts and the indices $(i,j)$ indicate
the different asymptotic configurations accessible at the specific energy
under consideration. We consider $p-d$ scattering at $E_{lab}=3$ MeV
using the AV18 potential in the $J=1/2^+$ state. The corresponding
scattering matrix is a $2\times 2$ matrix. The corresponding phase-shift
and mixing parameters have been calculated using the PHH expansion
and are given in Table~\ref{tab:irtab2}. From the
previous discussion we have shown that it is possible to solve an
equivalent problem with a screened Coulomb potential, so with free
asymptotic conditions, and then use the integral relations to extract
the scattering matrix corresponding to the unscreened problem. This has
been done using Eq.~\refeq{secondij} and the results are given in
Table~\ref{tab:irtab2} using $r_{sc}=50$ fm and $n_{sc}=5$.
We observe a complete agreement between the two procedures.
\begin{table}[h]
\caption{Phase-shift and mixing parameters for $p-d$ scattering
at $E_{lab}=3$ MeV using the AV18 potential. Results using the
PHH expansion (second column) and using the integral relations
(last column)}
\begin{tabular}{lcc}
\hline
& $p-d$ & Int.Rel. \\
\hline
$ ^4D_{1/2}$&$-3.563^\circ$ & $-3.562^\circ $ \\
$ ^2S_{1/2}$&$-32.12^\circ$ & $-32.12^\circ $ \\
$\eta_{1/2+}$&$1.100^\circ $ & $ 1.101^\circ $ \\
\hline
\end{tabular}
\label{tab:irtab2}
\end{table}
\section{Conclusions}
Stimulated by the fact that the commonly used TNF models
do not reproduce simultaneously
the triton and $^4$He binding energy and the $n-d$ doublet scattering length, we
have analyzed possible modifications of some of the TNF models usually used
in the description of light nuclei: the TM' and the URIX models. We have also
considered the recent N2LOL model. In each of these models we have
varied the original parameters so as to improve the
description of the mentioned quantities. Furthermore we have studied the
description of some $p-d$ polarization observables at $E_{lab}=3$ MeV. We have
observed that the modification of the URIX produces a worse description of the
vector polarization observables due to the artificial increase of the strength
of the repulsive term. The analysis of the TM' model has put in evidence the
necessity of including a repulsive term. In the case of the N2LOL model a
fine tuning of the parameters was possible in order to have an acceptable
description of the triton and $^4$He binding energies and the $n-d$ doublet scattering
length. Moreover, in the polarization observables we observe an improvement
in the vector analyzing powers and a slightly worse description of $T_{21}$.
From this analysis we have established a connection between the short-range structure of
the TNF and the polarization observables at low energies.
In a different application, we have discussed the use of the integral
relations derived from the KVP in the description of scattering states.
Firstly we have shown the use of bound state like wave functions to compute
the scattering matrix and, in the case of charged particles, the possibility
of computing
phase-shifts using scattering wave functions with free asymptotic conditions,
obtained after screening the Coulomb interaction. Both problems are of
interest in the study of light nuclei.
\section{Acknowledgments}
This work has been done in collaboration with my colleagues in Pisa
M. Viviani, L. Girlanda and L.E. Marcucci, with C. Romero-Redondo
and E. Garrido (CSIC) and P. Barletta (UCL).
\section{Bibliography}
\label{biblio}
|
\section{Introduction}
\label{sec:intro}
The analysis of the early exercise boundary and the optimal stopping time for American put options on assets paying zero dividends has attracted a lot of attention from both theoretical as well as practical point of view. An American put option is a financial contract between the writer and the holder of the option. It gives the holder the right, but not the obligation, to sell the underlying asset at the prescribed strike price any time before expiration. Under the standard assumptions made on the underlying stock process and completeness of the financial market (c.f. \cite{H,Kw}) the American put option can be priced using the Black-Scholes equation (c.f. \cite{BS}) on a time dependent domain of the underlying asset price. More precisely, the early exercise boundary problem for the American put option can be formulated as follows: find a solution $V=V(S,t)$ and the early exercise boundary position $S_f=S_f(t)$ satisfying
\begin{eqnarray}
&&\frac{\partial V}{\partial t} + r S\frac{\partial V}{\partial S} + {\sigma^2\over 2} S^2 \frac{\partial^2 V}{\partial S^2} - r V =0\,,
\qquad 0<t<T,\ S_f(t) < S <\infty \,,
\nonumber
\\
&&V(+\infty ,t)=0,\ V(S_f(t), t)= E - S_f(t)\,, \ \frac{\partial V}{\partial S}(S_f(t),t)=-1\,,
\\
&&V(S,T)=(E-S)^+\,.
\nonumber
\label{amer-put}
\end{eqnarray}
The solution $V(S,t)$ is defined on a time-dependent domain $S\in(S_f(t), \infty )$, where $t\in(0,T)$ (cf. Kwok \cite{Kw}).
Here $S>0$ stands for the underlying stock price, $E>0$ is the exercise (strike) price, $r>0$ is the risk-free rate, $\sigma>0$ is the volatility of the underlying stock process and $T$ denotes the time of maturity. In what follows, we denote by $\tau= T-t$ the time to maturity. The function $[0,T]\ni t \mapsto S_f(t)\in \R$ represents the early exercise boundary position. The above mathematical formulation of the problem of pricing the American put option by means of a solution to the free boundary problem is a basis for development of various integral equations for describing the early exercise boundary position $S_f(t)$. The analytical approximation formulae are often based on approximation of a solution to such an integral equation. Notice that there are also other numerical methods for approaching the free boundary problem (\ref{amer-put}) like e.g. front-fixing and transformation methods. We refer the reader to papers by Kwok and Wu \cite{KW}, \v{S}ev\v{c}ovi\v{c} \cite{Se1,Se2}, Ankudinova and Ehrhardt \cite{AE1} and references therein.
Following Kwok \cite{Kw}, a solution $V=V(S,t)$ to the problem of pricing the American put option fulfills the following variational inequality:
\begin{eqnarray}
\label{var-amer-put}
&&\frac{\partial V}{\partial t} + r S\frac{\partial V}{\partial S} + {\sigma^2\over 2} S^2 \frac{\partial^2 V}{\partial S^2} - r V \le 0\,,\quad V(S,t) \ge V(S,T)\,,
\nonumber
\\
&& \left(\frac{\partial V}{\partial t} + r S\frac{\partial V}{\partial S} + {\sigma^2\over 2} S^2 \frac{\partial^2 V}{\partial S^2} - r V \right) \Big(V(S,t)-V(S,T)\Big) = 0,
\nonumber
\\
&& \hbox{for all}\ \ 0<t<T,\ 0< S< \infty,
\\
&& V(0, t)= E, \quad V(+\infty,t)=0,\quad \hbox{for} \ \ 0<t<T\,,
\nonumber
\\
&&V(S,T)=(E-S)^+\,,\quad \hbox{for} \ \ 0< S< \infty.
\nonumber
\end{eqnarray}
The formulation of the problem of pricing American put option as a variational inequality is often used when we need to compute not only the free boundary position $S_f(t)$ but also the entire solution $V(S,t)$. The above variational inequality can be effectively solved by means of the so-called projected successive over relaxation (PSOR) method by Elliot and Ockendon \cite{EO}.
In the last decades, many different, but equivalent, integral equations for pricing the American put option have been derived by Barone-Adesi and Whaley \cite{BW}, Bunch and Johnson \cite{BJ}, Carr, Jarrow and Mynemi \cite{CJM}, MacMillan \cite{Mac} and others. The asymptotic analysis often leads to an approximate expression of the free boundary close to expiry. Since the closed form analytical formula for the early exercise boundary position is not known, many authors (see e.g. Geske, Johnson and Roll \cite{GJ,GR}, Johnson \cite{J}, Karatzas \cite{K1}, Evans, Kuske and Keller \cite{KK,EKK}, Mynemi \cite{M} and recent papers by Alobaidi \emph{et al.} \cite{A,MA}, Stamicar \emph{et al.} \cite{SSC}, the survey paper by Chadam \cite{Ch} and other references therein) investigated various approximation models and derived different approximate expressions for valuing American call and put options. We also refer to the books by Kwok \cite{Kw} and Wilmott \emph{et al.} \cite{WDH} for a survey of classical theoretical and computational results in the field of pricing the American put option.
In this paper, we focus on comparison of the valuation formulae due to Evans, Kuske and Keller \cite{KK,EKK}, Stamicar, \v{S}ev\v{c}ovi\v{c} and Chadam \cite{SSC} and the recent analytic approximation formula by Zhu \cite{Zhu2006} (see also \cite{Zhu2007,Zhu2008}). Our main goal is to present qualitative and quantitative comparison of the above mentioned analytical and numerical approximation methods for calculating the early exercise boundary position. In the first part of the paper, we analyze and compare asymptotic behavior of the early exercise boundary close to expiry for analytical approximations developed by Stamicar, \v{S}ev\v{c}ovi\v{c} and Chadam \cite{SSC}, Evans, Kuske, Keller \cite{KK,EKK} and Zhu \cite{Zhu2006,Zhu2007,Zhu2008}.
We show that the approximation formulae due to Evans, Kuske and Keller \cite{KK,EKK} and Stamicar, \v{S}ev\v{c}ovi\v{c} and Chadam \cite{SSC} have the same asymptotic behavior of $S_f(t)$ as $t\to T$. We also show that the analytic approximation formula due to Zhu has an asymptotic behavior differing from the previous ones by a logarithmic factor.
In the second part we propose a new numerical scheme for computation of the entire function $S_f(T), t\in[0,T]$, based on a solution to the nonlinear integral equation from \cite{SSC}. We compare numerical results obtained by the new numerical method to those of the projected successive over relaxation method by Elliot and Ockendon \cite{EO} for solving the variational inequality (\ref{var-amer-put}) and the analytical approximation formula recently developed by Zhu \emph{et al.} in \cite{Zhu2006,Zhu2007,Zhu2008}.
\section{Analytical approximate valuation formulae}
In this section we present a survey of analytical, implicit integral and numerical approximation schemes for computing the early exercise boundary for the American put option. First we focus on the recent result due to Zhu who in \cite{Zhu2006} derived a closed analytic approximation formula for the early exercise boundary position $S_f(t)=\varrho(T-t)$. We also derive the asymptotic behavior of $S_f(t)$ for $t\to T$.
Next we concentrate on implicit representation formulae for $\varrho(\tau)$ expressed in the form of a single nonlinear integral equation for the function $\varrho$. We recall implicit integral equation derived by Stamicar, \v{S}ev\v{c}ovi\v{c} and Chadam in \cite{SSC}. We again derive the asymptotic behavior of the early exercise boundary position as $t\to T$. In the last subsection we present another approximations derived by Evans, Kuske and Keller \cite{KK,EKK}.
\subsection{Analytical approximation valuation formula by Zhu}
In this section we recall a recent interesting result due to Zhu.
In \cite{Zhu2006} Zhu derived a new analytical approximation formula of the early exercise boundary by application of the Laplace and inverse Laplace integral transforms to a dimensionless form of the governing parabolic PDE and successfully obtained a closed analytic approximation formula for the early exercise boundary position as a sum of a perpetual option and integral that valuates early exercise boundary position. The resulting formula for the early exercise boundary $S_f(t) = \varrho(T-t)$ reads as follows:
\begin{equation}
\varrho^{Zhu}(\tau) = \frac{\gamma E}{1+\gamma} + \frac{2 E}{\pi} \int_{0}^{\infty} \frac{\zeta e^{-\tau\frac{\sigma^2}{2}(a^2+\zeta^2)}}{a^2+\zeta^2} e^{-f_1^*(\zeta)}\sin(f_2^*(\zeta)) d\zeta ,
\label{eq:zhu}
\end{equation}
where $ \gamma = \frac{2r}{\sigma^2}, \ \ a = \frac{1+\gamma}{2}, \ \ b = \frac{1-\gamma}{2}$, and
\begin{eqnarray}
\label{eq:f12}
f_1^*(\zeta) &=& \frac{1}{b^2+\zeta^2} \left[ b \ln \left(\frac{1}{\gamma}\sqrt{a^2+\zeta^2}\right) + \zeta \arctan(\zeta/a) \right],
\nonumber \\
f_2^*(\zeta) &=& \frac{1}{b^2+\zeta^2} \left[ \zeta \ln \left( \frac{1}{\gamma}\sqrt{a^2+\zeta^2}\right)
- b \arctan(\zeta/a) \right].
\\
\nonumber
\end{eqnarray}
Notice that the first summand in (\ref{eq:zhu}) represents the constant value of a perpetual put option i.e. the limit $\lim_{\tau\to\infty} \varrho(\tau)=\gamma E/(1+\gamma)$.
\subsubsection*{Early exercise boundary asymptotic close to expiry}
Next we examine the asymptotic behavior of the function $\varrho^{Zhu}(\tau)$ for $\tau\to 0$.
Notice that we have $\varrho(0)=S_f(T)=E$ (c.f. Kwok \cite{Kw}). We shall prove that
\[
\lim_{\tau\to 0^+}
\frac{ E -\varrho^{Zhu}(\tau)}{\sqrt{\tau}(-\ln\tau)} = \frac{1}{\sqrt{2\pi}} E \sigma.
\]
Indeed, if we introduce the change of variables: $s= \tau\frac{\sigma^2}{2}(a^2+\zeta^2)$ we obtain
\[
\frac{E -\varrho^{Zhu}(\tau)}{\sqrt{\tau}(-\ln\tau)} =
\frac{2 E}{\pi} \int_{ \tau\frac{\sigma^2}{2}a^2}^{\infty} \frac{1-e^{-s}}{2s} e^{-f_1^*} \frac{\sin(f_2^*)}{\sqrt{\tau}(-\ln\tau)}ds, \quad\hbox{for any}\ \tau\in (0,T],
\]
where $f_i^*= f_i^*((\frac{2s}{\tau\sigma^2} -a^2)^{\frac{1}{2}}), i=1,2$. It is easy to verify that
\begin{eqnarray*}
&&\lim_{\tau\to0^+} f_1^* = 0,\ \ \lim_{\tau\to0} f_2^* = 0,
\\
&&\lim_{\tau\to0^+} \frac{\sin(f_2^*)}{\sqrt{\tau}(-\ln\tau)} =\lim_{\tau\to0} \frac{f_2^*}{\sqrt{\tau}(-\ln\tau)} = \frac{\sigma}{2\sqrt{2s}}\,,
\end{eqnarray*}
for any $s>0$. Using the Lebesgue dominated convergence theorem we finally obtain
\[
\lim_{\tau\to 0^+}
\frac{E -\varrho^{Zhu}(\tau)}{\sqrt{\tau}(-\ln\tau)} =
\frac{E\sigma }{\pi} \int_{0}^{\infty} \frac{1-e^{-s}}{(2s)^{\frac{3}{2}}} ds
=\frac{1}{\sqrt{2\pi}} E \sigma\,,
\]
as claimed. As a consequence of the previous result we can conclude the following asymptotic approximation of the formula by Zhu:
\begin{equation}
\varrho^{Zhu}(\tau) \approx E \left(1 - \frac{\sigma}{\sqrt{2\pi}} \sqrt{\tau}(-\ln\tau) \right) \ \hbox{for} \ 0<\tau\ll 1,
\label{eq:asymptotic-Zhu}
\end{equation}
i.e. $\varrho^{Zhu}(\tau) = E \left(1 - \frac{\sigma}{\sqrt{2\pi}} \sqrt{\tau}(-\ln\tau) \right) + o(\sqrt{\tau}(-\ln\tau))$ as $\tau\to0^+$.
In Fig.~\ref{fig:asymptotic-Zhu} we present a comparison of the analytic solution $\varrho^{Zhu}(\tau)$ and its asymptotic approximation (\ref{eq:asymptotic-Zhu}) for $\tau\in[0, T]$ and $E=100,\sigma=0.3, r=0.1, T=10^{-4}$.
\begin{figure}
\begin{center}
\includegraphics[width=6.5cm]{figures/approximationZhu.ps}
\end{center}
\caption{Comparison of the analytic solution $\varrho^{Zhu}$ (solid curve) and its asymptotic approximation (\ref{eq:asymptotic-Zhu}) (dashed curve).}
\label{fig:asymptotic-Zhu}
\end{figure}
\subsubsection*{Convexity of the early exercise boundary obtained from Zhu's formula}
One of the important features of the early exercise boundary for the American put option is the convexity of the function $\varrho(\tau) = S_f(T-\tau)$ for $\tau\in (0,T]$. The analytic proof of the convexity of $\varrho$ has been presented just recently by Chadam \emph{et al.} in \cite{CCJZ}. We also recall that the early exercise boundary is log-concave as a function of log of the underlying asset price (cf. Ekstr\"om and Tysk \cite{E,ET}).
A relatively simple proof of the convexity of $\varrho=\varrho^{Zhu}$ follows directly from the analytic valuation formula (\ref{eq:zhu}). Indeed, for any $0<\tau\le T$, we have the following expression for the second derivative of the function $\varrho^{Zhu}(\tau)$:
\[
\frac{d^2}{d\tau^2}\varrho^{Zhu}(\tau) = \frac{2 E \sigma^4}{4 \pi} \int_{0}^{\infty}
(a^2+\zeta^2)\zeta e^{-\tau\frac{\sigma^2}{2}(a^2+\zeta^2)} e^{-f_1^*(\zeta)}\sin(f_2^*(\zeta)) d\zeta.
\]
In what follows, we shall prove $f_2^*(\zeta)\equiv f_2^*(\zeta; \gamma)\in[0,\pi]$ provided that $\gamma\ge\gamma_0$ where $\gamma_0>0$ is a constant.
\begin{equation}
\gamma_0= \min(\gamma>0\ | \ \max_{\zeta>0}f_2^*(\zeta,\gamma) \le \pi )
\end{equation}
\begin{figure}
\begin{center}
\includegraphics[width=5.5cm]{figures/functionf2.ps}
\includegraphics[width=5.5cm]{figures/functionG.ps}
\end{center}
\caption{A graph of the function $f_2^*=f_2^*(\zeta; \gamma)$ for various values of the parameter $\gamma$ (left). A graph of the function $G(\gamma)=\max_{\zeta>0}f_2^*(\zeta; \gamma)$ (right).}
\label{fig:functionf2}
\end{figure}
The numerical value of $\gamma_0$ can be estimated as $\gamma_0\approx 0.0167821$.
\begin{corollary}
If $\frac{2r}{\sigma^2}=\gamma \ge \gamma_0$ where $\gamma_0\approx 0.0167821$ then $f_2^*=f_s^*(\zeta,\gamma)\in[0,\pi]$ for any $\zeta>0$. As a consequence, we have $\frac{d^2}{d\tau^2}\varrho^{Zhu}(\tau)>0$, i.e. the function $\varrho^{Zhu}(\tau)$ as well the early exercise boundary $S_f(t)$ for the American put option are convex functions.
\end{corollary}
\begin{remark}
Notice that the condition $\frac{2r}{\sigma^2}=\gamma \ge \gamma_0$ is fulfilled for typical market-based choices of the model parameters $r$ and $\sigma$. For example, if $r=0.01$ (i.e. $r=1\%$ p.a.) then $\frac{2r}{\sigma^2}\ge \gamma_0$ provided that the condition $\sigma^2 <1.19$ (i.e. $\sigma^2 \le 119\%$ p.a.) is satisfied.
In Fig.\ref{fig:functionf2} we present graphs of the function $\zeta\mapsto f_2^*(\zeta; \gamma)$ for various values of the parameter $\gamma$, including the critical value $\gamma=\gamma_0\approx 0.0167821$ for which the function $G(\gamma)=\max_{\zeta>0}f_2^*(\zeta; \gamma)$ attains the critical value $G(\gamma_0)=\pi$.
\end{remark}
\subsection{Approximation formula due to Stamicar, \v{S}ev\v{c}ovi\v{c} and Chadam}
In \cite{SSC} Stamicar, \v{S}ev\v{c}ovi\v{c} and Chadam derived a single nonlinear integral equation for the early exercise boundary position. Based on this integral equation the authors derived improved analytical approximation of free boundary near the expiry. Asymptotic behavior and justification of the early exercise behavior close to expiry have been recently analyzed by Chadam \emph{et al.} in \cite{CCJZ,CC}. They proved that the right asymptotic expansion can be obtained from the nonlinear integral equation developed by Stamicar, \v{S}ev\v{c}ovi\v{c} and Chadam in \cite{SSC}. We briefly recall key steps of derivation of the nonlinear integral equation for the early exercise boundary position $\varrho(\tau) = S_f(T-\tau)$ for the free boundary problem (\ref{amer-put}). Let us introduce the following change of variables
$x=\ln\left(S/\ro(\tau)\right)$ where $\tau=T-t, \ro(\tau)=S_f(T-\tau)$.
Similarly as in the case of a call option (see \cite{Se1}) we define a synthetised portfolio $\Pi$ for the put option $\Pi(x,\tau) = V(S,t) - S \frac{\partial V}{\partial S}(S,t)$. Then it is easy to verify that $\Pi$ is a solution to the following parabolic equation:
\begin{eqnarray}
&&\frac{\partial \Pi}{\partial \tau}
- a(\tau )\frac{\partial \Pi}{\partial x}
- \frac{\sigma^2}{2}\frac{\partial^2 \Pi}{\partial x^2}
+ r \Pi = 0,\quad x>0,\tau\in(0,T), \nonumber \\
&&\Pi (0,\tau ) = E, \quad \Pi (\infty ,\tau ) = 0, \quad
\Pi (x,0) = 0, \quad x>0, \tau\in(0,T),
\label{bc1} \\
&&\frac{\sigma^2}{2} \frac{\partial \Pi}{\partial x}(0,\tau ) = - r E,\quad \hbox{for}\ \tau\in(0,T),\nonumber
\end{eqnarray}
where $a(\tau ) =\frac{\dot{\varrho}(\tau )}{\varrho (\tau )}+r - \frac{\sigma^2}{2}$ (see Stamicar \emph{et al.} \cite{SSC}, or \v{S}ev\v{c}ovi\v{c} \cite{Se1,Se2}).
Applying the Fourier transform one can find the Fourier image of the function $\Pi$ in terms of the free boundary position $\varrho$. The resulting equation for the free boundary position reads as
$\frac{\sigma^2}{2} \frac{\partial \Pi}{\partial x}(0,\tau ) = - r E$, from which the weakly singular integral equation for the function $\varrho$ can be found by using the inverse Fourier transform (see \cite{SSC} for details). More precisely, the function $\varrho(\tau)$ fulfills the equation:
\begin{equation}
\varrho (\tau )=Ee^{-(r-\frac{\sigma^2}{2})\tau + \sigma \sqrt{2 \tau }\eta (\tau )},
\label{eq:SSC}
\end{equation}
where the auxiliary function $\eta (\tau )$ is a solution to the following nonlinear integral equation
\begin{equation}
\eta (\tau ) =-\sqrt{-\ln \left[ \frac{r\sqrt{2 \pi \tau}}{\sigma} e^{ r\tau }\left(
1-\frac{F_\eta(\tau )}{\sqrt{\pi }}\right) \right] },\qquad \hbox{for}\ \tau\in[0,T].
\label{eq:SSCintegralequation}
\end{equation}
Here the function $F_\eta$ depends on $\eta$ via the expression
\begin{eqnarray}
F_\eta(\tau ) &=&2\int_0^{\pi /2}e^{-r\tau \cos ^2\theta
- g^2_\eta(\tau ,\theta)}
\left(
\frac{\sigma\sqrt{\tau }}{\sqrt{2}}\sin \theta +g_\eta(\tau ,\theta )\tan \theta
\right)
\,d\theta , \label{F}
\\
g_\eta(\tau ,\theta ) &=&\frac 1{\cos \theta }\left[ \eta (\tau )- \eta (\tau \sin ^2\theta )
\sin\theta \right], \label{g}
\end{eqnarray}
for $\tau\in[0,T], \theta\in [0,\frac{\pi}{2}]$. According to \cite{SSC}, the asymptotic analysis of the above integral equation for the unknown function $\eta(\tau)$ enables us to conclude the asymptotic approximation formula for $\eta(\tau)$ as $\tau\to 0$.
The early exercise behavior of $\varrho(\tau)$ for $\tau\to 0$ can be then deduced from the second order iteration to the system (\ref{eq:SSCintegralequation}) and (\ref{F}) when starting from the initial guess $\eta_0(\tau) = (r-\frac{\sigma^2}{2}) \frac{\tau^{\frac{1}{2}}}{\sigma\sqrt{2}}$ corresponding to the constant early exercise boundary $S_{f0}(t) \equiv E$. One can iteratively compute $F_{\eta_0}$, $\eta_1$ and $F_{\eta_1}$, $\eta_2$. It turned out from calculation performed in \cite{SSC} that the second consecutive iterate $\eta_2$ is the lowest order (in $\tau$) approximation of $\eta$. Namely,
\begin{equation}
\eta (\tau )\sim -\sqrt{-\ln \left[ \frac{2 r}{\sigma} \sqrt{2\pi \tau } e^{r \tau
}\right] } \quad \hbox{as}\ \tau\to 0^+.
\label{canad1-eta2}
\end{equation}
Interestingly enough, it has been shown just recently by Chen \emph{et al.} \cite{CCJZ} that the early exercise boundary function $\varrho$ is convex (see also \cite{Ch,CC}). Moreover, the approximation formula (\ref{canad1-eta2}) derived by Stamicar, \v{S}ev\v{c}ovi\v{c} and Chadam \cite{SSC} provides the right asymptotic behavior for $\tau\to 0^+$. Furthermore, Chen and Chadam \cite{CC} derived sixth-th order expansion of the function
\begin{equation}
\alpha(\tau) = - \xi - \frac{1}{2\xi} + \frac{1}{8\xi^2} + \frac{17}{24\xi^3}
-\frac{51}{64\xi^4}-\frac{287}{120\xi^5}+\frac{199}{32\xi^6}+O(\xi^{-7}),
\label{eq:asymptotics}
\end{equation}
for $\xi = \ln {\sqrt{\frac{8\pi r^2 \tau}{\sigma^2}}}\to -\infty$ as $\tau\to 0^+$ where
\begin{equation}
\varrho(\tau) = E e^{ - \sigma \sqrt{2\tau \alpha(\tau)}}.
\label{eq:ro_ssh}
\end{equation}
\subsubsection*{Early exercise boundary asymptotic close to expiry}
Similarly as in the case of the analytic approximation formula by Zhu, we examine the asymptotic behavior of the function $\varrho(\tau)$ for $\tau\to 0$ where $\varrho(\tau)\equiv\varrho^{SSC}(\tau)$ is given by the equation:
\begin{equation}
\varrho (\tau )=Ee^{-(r-\frac{\sigma^2}{2})\tau + \sigma \sqrt{2 \tau }\tilde\eta (\tau )},\quad
\hbox{where}\ \ \tilde\eta (\tau ) = -\sqrt{-\ln \left[ \frac{2 r}{\sigma} \sqrt{2\pi \tau } e^{r \tau
}\right] }.
\label{eq:SSC-new}
\end{equation}
Employing expression (\ref{eq:SSC-new}) it is straightforward to verify that
\[
\lim_{\tau\to 0^+}
\frac{E-\varrho^{SSC}(\tau)}{\sqrt{\tau}\sqrt{-\ln\tau}} = E \sigma.
\]
Again, as a consequence of the above limit we conclude the following asymptotic approximation of the analytic valuation formula due to Stamicar, \v{S}ev{c}ovi\v{c} and Chadam:
\begin{equation}
\varrho^{SSC}(\tau) \approx E \left(1 - \sigma\sqrt{\tau}\sqrt{-\ln\tau} \right) \quad \hbox{for} \ 0<\tau\ll 1,
\label{eq:asymptotic-SSC}
\end{equation}
i.e. $\varrho^{SSC}(\tau) = E \left(1 - \sigma\sqrt{\tau}\sqrt{-\ln\tau} \right) + o(\sqrt{\tau}\sqrt{-\ln\tau})$ as $\tau\to0^+$.
Notice that the asymptotic formula (\ref{eq:asymptotic-SSC}) differs from the one obtained from Zhu's formula (\ref{eq:asymptotic-Zhu}) by a logarithmic factor $\sqrt{-\ln\tau}$.
\subsection{Approximation formulae by Evans, Kuske and Keller}
In \cite{KK} Kuske and Keller proposed another analytic approximation of the early exercise boundary for times close to expiration. Then, in the paper with Evans \cite{EKK}, they improved and extended the formula for the case of dividend-paying asset.
We begin with the approximation formula by Kuske and Keller \cite{KK}. Their approximation formula for the position of the early exercise boundary close to expiry $t\to T$ reads as follows:
\begin{equation}
\varrho^{KK}(\tau) \approx E\left(
1 -\sigma\sqrt{2\tau}\sqrt{-\ln{ \left[ \frac{2r}{\sigma} \sqrt{\frac{9 \pi \tau}{2} } \right] }}
\right),\qquad \hbox{as}\ \ \tau\to0^+.
\label{eq:KK}
\end{equation}
In \cite{EKK} Evans, Kuske and Keller derived an improved asymptotic formula:
\begin{equation}
\varrho^{EKK}(\tau) \approx E\left(
1 -\sigma\sqrt{2\tau}\sqrt{-\ln{\left[ \frac{2r}{\sigma} \sqrt{2\pi\tau} \right] }}
\right),\qquad \hbox{as}\ \ \tau\to0^+.
\label{eq:EKK}
\end{equation}
Although, asymptotic formulae (\ref{eq:KK}), (\ref{eq:EKK}) by Evans, Kuske and Keller and Stamicar, \v{S}ev\v{c}ovi\v{c} and Chadam (\ref{eq:SSC-new}) differ in higher order terms of $\tau$, it holds
\begin{equation}
\lim_{\tau\to 0^+}
\frac{E-\varrho^{SSC}(\tau)}{\sqrt{\tau}\sqrt{-\ln\tau}} = \lim_{\tau\to 0^+}
\frac{E-\varrho^{KK}(\tau)}{\sqrt{\tau}\sqrt{-\ln\tau}} =\lim_{\tau\to 0^+}
\frac{E-\varrho^{EKK}(\tau)}{\sqrt{\tau}\sqrt{-\ln\tau}} =
E \sigma.
\label{eq:alllim}
\end{equation}
It means that approximation formulae due to Evans, Kuske and Keller \cite{KK,EKK} and Stamicar, \v{S}ev\v{c}ovi\v{c} and Chadam \cite{SSC} have the same asymptotic behavior close to expiry $t\approx T$, i.e. $0<\tau\ll 1$.
\section{Numerical methods for calculation of the early exercise boundary}
\label{sec:numerical}
The early exercise boundary function $\varrho(\tau)$ for the entire time interval $\tau\in[0,T]$, can be approximated by using numerical methods as well. In this section we present two approaches: 1) a new local iterative algorithm based on the integral equation due to Stamicar, \v{S}ev\v{c}ovi\v{c} and Chadam \cite{SSC}; 2) the well-known PSOR method (c.f. Kwok \cite{Kw}).
\subsection{A new numerical algorithm based on a solution to the integral equation}
\label{sec:ssch-method}
The aim of this section is to introduce a new numerical algorithm for computation of the early exercise boundary of the American put option. It is based on a solution to the system of implicit equations (\ref{eq:SSCintegralequation}), (\ref{F}), (\ref{g}) derived by Stamicar \emph{et al.} in \cite{SSC}. The idea of the proposed algorithm is to sequentially compute values of the auxiliary function $\eta=\eta(\tau)$ in nodal points $\tau_i\in [0,T]$. In contrast to global iterative algorithms which iteratively compute the entire solution $\varrho(\tau), \tau\in[0,T],$ (see e.g. \v{S}ev\v{c}ovi\v{c} \cite{Se1}) we only need to find a root of a real valued function at each nodal point $\tau_i$. This is due to the form of functions $F_\eta, g_\eta$ (see (\ref{F}) and (\ref{g})) whose values at $\tau\in (0,T]$ depend only on the value $\eta(\tau)$ and the history path $\{\eta(\xi), 0\le \xi<\tau\}$.
Our new algorithm for computation of the approximation of the early exercise boundary $\varrho(\tau)=S_f(T-\tau)$ reads as follows:
\begin{enumerate}
\item{} Construct a division $0=\tau_0 < \tau_1 < ... < \tau_m=T$ of the interval $[0, T]$.
To this end we can employ either equidistant partition $\tau_i = (i/m) T$, or we can use $\tau_i = (i/m)^2 T$ in order to adjust the discretization mesh to desired behavior (\ref{eq:asymptotic-SSC}) of $\varrho(\tau)$ close to expiry $\tau\approx 0$. We take $m\gg 1$ sufficiently large such that $\frac{2 r}{\sigma} \sqrt{2\pi \tau_1 } e^{r \tau_1 }<1$.
\item{} Compute the value of $\eta_1\approx \eta(\tau_1)$ from the analytic approximation formula (\ref{canad1-eta2}), i.e.
\[
\eta_1 = -\sqrt{-\ln \left[ \frac{2 r}{\sigma} \sqrt{2\pi \tau_1 } e^{r \tau_1
}\right] }.
\]
\item{} for $i=2, ..., m$, compute the value $\eta_i\approx \eta(\tau_i)$ as follows:
\begin{itemize}
\item[3-1] Construct the mapping ${\mathcal G}_{\eta_i}(\tau_i, \theta) = \frac 1{\cos \theta }\left[ \eta_i -\tilde\eta (\tau_i \sin ^2\theta ) \sin \theta\right],$
where $\tilde\eta(\tau_i \sin ^2\theta)$ is a linear interpolation function between the points
$(\tau_j,\eta_j)$ and $(\tau_{j+1},\eta_{j+1})$ if $\tau_j \leq \tau_i \sin^2 \theta < \tau_{j+1}$ for some $1\le j<i$.
If $0<\tau_i \sin^2 \theta< \tau_1$ then $\tilde\eta(\tau_i \sin ^2\theta)$ is given by the analytic approximation formula (\ref{canad1-eta2}).
\item[3-2] Construct the mapping ${\mathcal F}_{\eta_i}(\tau_i)$:
\[
{\mathcal F}_{\eta_i}(\tau_i)= 2\int_0^{\pi /2}e^{-r\tau_i \cos ^2\theta
- {\mathcal G}^2_{\eta_i}(\tau_i ,\theta)}
\left(
\frac{\sigma\sqrt{\tau_i }}{\sqrt{2}}\sin \theta +{\mathcal G}_{\eta_i}(\tau_i ,\theta )\tan \theta
\right)
\,d\theta .
\]
As for the numerical quadrature of the above integral we can employ the composed Newton-Cotes method of the fourth order with, at least, 1000 subintervals.
\item[3-3] Find the root $\eta_i$ of the equation:
\[
\eta_i =-\sqrt{-\ln \left[ \frac{r\sqrt{2 \pi \tau_i}}{\sigma} e^{ r\tau_i }\left(
1-\frac{{\mathcal F}_{\eta_i}(\tau_i )}{\sqrt{\pi }}\right) \right] }.
\]
The above equation can be solved using either bisection method, Newton's method or any other numerical iterative method for finding roots of real valued functions. In order to speed-up convergence we can use already constructed value $\eta_{i-1}$ as a starting point for iterations at the time level $\tau_i$.
\end{itemize}
\item{} Go to step 3 and repeat the calculation of $\eta_i$ for the next value of $i$ until $i\le m$.
\item{}
From discrete values $\eta_i, i=1,2, ..., m,$ we compute the approximation $\varrho_i$ of the early exercise boundary position $\varrho(\tau_i)$ as follows:
\[
\varrho_i = E e^{-(r-\frac{\sigma^2}{2})\tau_i + \sigma \sqrt{2 \tau_i }\eta_i},
\]
We set $\varrho_0= E$.
The entire profile $\varrho(\tau)=S_f(T-\tau), \tau\in[0,T]$, is then computed as a linear interpolation function between discrete values $\{ (\tau_i,\varrho_i), i=0, ..., m\}$.
\end{enumerate}
\subsection{Approximate solution using the PSOR method}
\label{sec:psor}
In this section, we present a brief overview how the early exercise boundary can be found using a finite difference numerical approximation method applied to the variational inequality (\ref{var-amer-put}). The method consits in computation the option price $V(S,t)$ using the so-called projected successive over relaxation (PSOR) method introduced by Ockendon and Elliot in \cite{EO}.
Having computed a solution $V(S,t)$ to the variational inequality (\ref{var-amer-put}) we can calculate the early exercise boundary position. Indeed, given a time $t$, the critical stock price $S_f(t)$ is equal to the maximal stock price $S=S_f(t)$ for which the option price is equal to the payoff, i.e.
\[
S_f(t) = \max \{ S>0 \ | \ V(S,t) = (E-S)^+ \}.
\]
Following Kwok \cite{Kw}, the idea of the PSOR method is to transform (\ref{var-amer-put}) by introducing new variables $x=\ln(S/E), \quad \tau=T-t, \quad u(x,\tau) = e^{\alpha x + \beta \tau } V(E e^x, T-\tau)$
where constants $\alpha,\beta$ are defined by $\alpha = \frac{r}{\sigma^2} -\frac{1}{2}, \beta = \frac{r}{2}+ \frac{\sigma^2}{8} + \frac{r^2}{2\sigma^2}$.
We denote $u_i^j \approx u(i h, j k)$ the finite difference approximation of a solution to the transformed variational inequality for $i=-n, ..., -1,0,1, .., n$, $j=1, ..., m$. The spatial and time discretization
steps $h, k > 0$ are chosen such that $h=L/n$, $k=T/m$, respectively. Here $T$ represents expiration time and $L$ is a sufficiently large bound for the interval $x\in (-L,L)$. For practical purposes we can take $L\approx 1$ .
In each time step $j=1,2, ..., m$, a linear complementarity problem for the finite difference approximation vector $u^j\in \R^{2n+1}$ is solved by using the iterative successive over relaxation (SOR) method where iterates are projected to the transformed pay-off diagram. This is done by taking the maximum of the transformed pay-off and computed iteration of a solution obtained by the SOR successive iteration. For details we refer the reader to \cite[pp. 212--224]{Kw}.
\section{Numerical comparison of the early exercise boundary approximations}
\label{sec:comparison}
\subsection{Comparison of approximations close to expiry}
\begin{figure}
\begin{center}
\subfigure[$T = 5\times10^{-5}$ (1 min)]{
\includegraphics[height=4.5cm]{figures/analyticke_1_vysoke.ps}
}
\subfigure[$T = 4\times 10^{-3}$ (1 day)]{
\includegraphics[height=4.5cm]{figures/analyticke_2_vysoke.ps}
}
\subfigure[$T = 0.08 $ (1 month)]{
\includegraphics[height=4.5cm]{figures/analyticke_3_vysoke.ps}
}
\subfigure[$T = 0.25 $ (3 months)]{
\includegraphics[height=4.5cm]{figures/analyticke_4_vysoke.ps}
}
\end{center}
\caption{Comparison of analytic approximation formulae for various maturities $T$ on a yearly basis.}
\label{fig:analyticke}
\end{figure}
This section focuses on numerical comparison of analytic approximations due to Stamicar, \v{S}ev\v{c}ovi\v{c} and Chadam (\ref{eq:SSC}), Kuske and Keller (\ref{eq:KK}), Evans, Kuske and Keller (\ref{eq:EKK}), Zhu (\ref{eq:zhu})
and our new local iterative algorithm from section 3.1 for the early exercise boundary for times $0<\tau=T-t\ll 1$ close to expiry.
For computational purposes we chose the volatility $\sigma=30\%$, risk-free interest rate $r=10\%$ p.a., and the strike price $E=100\$$.
\begin{table}
\caption{\small
Comparison of the early exercise boundary obtained by analytic approximation formulae and the iterative algorithm to the benchmark PSOR method.}
\begin{center}
\scriptsize
\begin{tabular}{l|ccccc|cccc} \hline
\multirow{2}{*}& \multicolumn{5}{|c|}{\bf Early exercise boundary} & \multicolumn{4}{|c}{\bf Relative error} \\
\multirow{2}{*} {$\tau$}& \multicolumn{5}{|c|}{$\varrho(\tau)=S_f(T-\tau)$ } & \multicolumn{4}{|c}{\bf w.r. to the PSOR method} \\
& {\bf EKK } & {\bf Zhu } & {\bf SSCh-A } & {\bf SSCh } & {\bf PSOR } & {\bf EKK } & {\bf Zhu } & {\bf SSCh-A } & {\bf SSCh } \\ \hline\hline
0.000 01 & 99.69 & 99.51 & 99.69 & 99.690 & 99.7 & 0.01\% & 0.19\% & 0.01\% & 0.01\% \\
0.000 05 & 99.37 & 99.03 & 99.37 & 99.358 & 99.4 & 0.03\% & 0.37\% & 0.03\% & 0.04\% \\
0.000 1 & 99.14 & 98.72 & 99.15 & 99.111 & 99.2 & 0.06\% & 0.49\% & 0.06\% & 0.09\% \\
0.000 5 & 98.28 & 97.57 & 98.29 & 98.270 & 98.31 & 0.03\% & 0.76\% & 0.02\% & 0.04\% \\
0.001 & 97.7 & 96.83 & 97.72 & 97.660 & 97.73 & 0.03\% & 0.92\% & 0.01\% & 0.07\% \\
0.01 & 95.62 & 94.27 & 95.69 & 95.502 & 95.6 & 0.02\% & 1.39\% & 0.09\% & 0.10\% \\
0.01 & 94.33 & 92.73 & 94.43 & 94.070 & 94.18 & 0.16\% & 1.54\% & 0.27\% & 0.11\% \\
0.04 & 91.12 & 88.66 & 91.31 & 90.205 & 90.3 & 0.90\% & 1.82\% & 1.12\% & 0.11\% \\
0.1 & 89.29 & 85.25 & 89.42 & 86.762 & 86.94 & 2.70\% & 1.93\% & 2.86\% & 0.20\% \\
\hline
\end{tabular}
\end{center}
{\scriptsize
Legend: EKK - Evans, Kuske and Keller \cite{EKK},
SSCh-A - Stamicar, \v{S}ev\v{c}ovi\v{c} and Chadam \cite{SSC},
SSCh - our new local iterative algorithm from section 3.1,
ZHU - Zhu \cite{Zhu2006}, PSOR - Projected SOR method \cite{Kw}.
}
\label{tab:analyticke}
\end{table}
In Fig.~\ref{fig:analyticke} we present quantitative comparison of analytic approximation formulae by Kuske and Keller (KK), Evans, Kuske and Keller (EKK), Stamicar, \v{S}ev\v{c}ovi\v{c} and Chadam (SSCh-A), Zhu's formula (ZHU). As a numerical benchmark solution we chose the PSOR method with $n=1000$ spatial grid points and $m=1000$ time steps (see section \ref{sec:psor}). It should be obvious that the approximation formulae KK, EKK and SSCh-A exhibit similar behavior w.r. to PSOR for the time close to expiry (see Fig.~\ref{fig:analyticke} a), b)). On the other hand, on a larger time horizon KK, EKK as well as SSCh-A become nondecreasing and Zhu's formula (ZHU) better approximates the PSOR solution (see Fig.~\ref{fig:analyticke} c), d)). It is also worthwile to note that Zhu's formula undershoots the early exercise boundary for small values of $\tau$ when compared to KK, EKK, SSCh-A and PSOR. This phenomenon can be easily justified by calculating the limit
\begin{equation}
\lim_{\tau\to 0^+}
\frac{E-\varrho^{SSC}(\tau)}{E-\varrho^{Zhu}(\tau)} \sqrt{-\ln\tau}
= 1.
\label{eq:undershoot}
\end{equation}
In Table~\ref{tab:analyticke} we calculated the early exercise boundary position for EKK, ZHU, SSCh-A, PSOR methods and our new local iterative algorithm described in section 3.1 which is labeled as SSCh.
We also calculated the relative error $\Delta^{method}(\tau)$ defined as
\[
\Delta^{method}(\tau) = \frac{\left|S_f^{method}(T-\tau)-S_f^{PSOR}(T-\tau)\right|}{S_f^{PSOR}(T-\tau)}, \quad \hbox{for} \ \tau\in [0,T],
\]
where $S_f^{PSOR}$ is the early exercise boundary computed by the PSOR method. For $\tau \approx 1$ minute EKK, SSCh-A and PSOR methods have almost identical values (close to $99.4\$$) whereas Zhu's boundary position has been calculated as $99.03\$$. On the other hand, other approximations (EKK and SSCh-A) differs significantly from early exercise boundary obtained by the PSOR method as we enlarge time to expiration $\tau>0.02$. The relative error in the early exercise boundary position calculated by Zhu's formula w.r. the PSOR method is less than $2\%$. The best approximation of the early exercise boundary has been achieved by our local iterative algorithm SSCh.
In summary, SSCh-A, KK and EKK analytic approximation formulae are suitable for approximation of the early exercise boundary close to expiration whereas, for a longer time horizon, it is recommended to use the analytical approximation formula by Zhu. The new local iterative approximation derived in section 3.1 can be used for both small as well as large time horizon.
\begin{table}
\caption{\small Comparison of the early exercise boundary on a long time horizon.}
\begin{center}
\scriptsize
\begin{tabular}{l | ccc | cc} \hline
\multirow{2}{*}{ $\tau$ }
&
\multicolumn{3}{|c|}{ \bf Early exercise boundary } & \multicolumn{2}{|c}{ \bf Rel. error w.r. to} \\
&
\multicolumn{3}{|c|}{ $S_f(T-\tau)$} & \multicolumn{2}{|c}{ \bf PSOR method} \\
& {\bf PSOR} & {\bf Zhu }& {\bf SSCh} & { \bf Zhu} & {\bf SSCh} \\
\hline \hline
0 & 100. & 100. & 100. & 0\% & 0\% \\
0.02 & 92.8672 & 90.8575 & 92.3461 & 2.16\% & 0.56\% \\
0.04 & 90.7707 & 88.6563 & 90.2088 & 2.33\% & 0.62\% \\
0.06 & 89.3300 & 87.2160 & 88.7771 & 2.37\% & 0.62\% \\
0.08 & 88.2350 & 86.1300 & 87.6695 & 2.39\% & 0.64\% \\
0.1 & 87.3279 & 85.2538 & 86.7636 & 2.38\% & 0.65\% \\
0.2 & 84.2962 & 82.3766 & 83.7476 & 2.28\% & 0.65\% \\
0.4 & 81.0179 & 79.3593 & 80.4793 & 2.05\% & 0.66\% \\
0.6 & 79.0571 & 77.5961 & 78.5391 & 1.85\% & 0.66\% \\
0.8 & 77.6986 & 76.3752 & 77.1895 & 1.7\% & 0.66\% \\
1 & 76.6695 & 75.4580 & 76.1632 & 1.58\% & 0.66\% \\
1.5 & 74.9137 & 73.8879 & 74.4094 & 1.37\% & 0.67\% \\
2 & 73.8107 & 72.8731 & 73.2722 & 1.27\% & 0.73\% \\
3 & 72.5786 & 71.6205 & 71.8735 & 1.32\% & 0.97\% \\
4 & 72.0121 & 70.8778 & 71.0464 & 1.58\% & 1.34\% \\
5 & 71.7966 & 70.3925 & 70.5100 & 1.96\% & 1.79\% \\
\hline
\end{tabular}
\end{center}
{\scriptsize
Legend:
SSCh - our new local iterative algorithm from section 3.1, ZHU - Zhu \cite{Zhu2006}, PSOR - \cite{Kw}.
}
\label{tab:comparison}
\end{table}
\subsection{The long term horizon}
In the long term horizon, i.e. $\tau=T-t \approx 1 $ year or even more, we can no longer use the analytical approximations SSCh-A, KK, EKK designed for $0<\tau\ll 1$ any more. These solutions loose monotonicity for $\tau \approx 0.1$ and they become even undefined for large values of $\tau$ because of the sign change in the logarithm.
This is why only Zhu's analytical approximation formula for the early exercise boundary (\ref{eq:zhu}) can be used in the long term horizon. We compared Zhu's approximation with two numerical methods described in section \ref{sec:numerical}. The first method is our new numerical method (labeled by SSCh) based on the integral equation (\ref{eq:SSCintegralequation}) which was described in section~\ref{sec:ssch-method}. The second method is the classical PSOR method described in section~\ref{sec:psor}.
As a time horizon, we chose the large expiration time $T=5$ years. Other model parameters are the same as in the previous section, i.e. $E=100\$ $, $\sigma=30\%$, $r=10\%$ p.a.
The computational results are shown in Fig.~\ref{fig:years5} (left) and Table~\ref{tab:comparison}. We can observe that the shape of all solutions is very similar.
In Fig.~\ref{fig:years5} (right) we plotted the relative error with respect to the PSOR method, which was used as a benchmark. Zhu's analytic approximation formula has the relative error between $1$ and $2.5\%$ and it attained local minimum around $\tau \approx 2$ years. This is due to the fact that Zhu's method is slightly undershooting the solution close to expiry, i.e. for $\tau\approx 0$.
The solution computed by our new SSCh scheme shows nearly constant error term until $\tau \approx 2.5$ years, then the relative error starts to grow up. For $\tau\approx 5$ years, the numerical solution SSCh is approaching Zhu's approximation. This is due to loose of precision of the PSOR method itself when the exact early exercise boundary could be closer to SSCh and Zhu's approximation than to the PSOR solution.
\setcounter{subfigure}{0}
\begin{figure}
\begin{center}
\includegraphics[height=4.5cm]{figures/longterm_zhu_ssch_psor1.ps}
\hskip 5truemm
\includegraphics[height=4.5cm]{figures/longterm_zhu_ssch_error1.ps}
\end{center}
\caption{\small Comparison of the early exercise boundary position in the long time horizon.
The early exercise boundary position (left). The relative error with respect to the PSOR method (right).
}
\label{fig:years5}
\end{figure}
\section{Comparison of options prices}
In this section we address the question concerning the difference between the American put option price and the approximative option price computed with an approximation of the early exercise boundary. More precisely, let $V^{am}(S,t)$ be the solution to the free boundary problem (\ref{amer-put}) with the early exercise boundary profile $S_f$. Let us consider a given function $S^{app}_f$ representing an approximation of the early exercise boundary profile $S_f$. We denote by $V^{app}$ the unique solution to the parabolic equation:
\begin{figure}
\begin{center}
\includegraphics[width=7.5cm]{figures/comparison.ps}
\end{center}
\caption{A profile $S\mapsto V^{am}(S,t)$ of the American option price and its comparison to the option price $V^{app}$ computed with respect to the approximative early exercise boundary $S^{app}_f(t)=\varrho^{app}(T-t)$. The corresponding European style of an option is labeled by $V^{eu}$. }
\label{fig:comparison}
\end{figure}
\begin{eqnarray}
&&\frac{\partial V^{app}}{\partial t} + r S\frac{\partial V^{app}}{\partial S} + {\sigma^2\over 2} S^2 \frac{\partial^2 V^{app}}{\partial S^2} - r V^{app} =0\,,
\quad t\in(0,T),\ S^{app}_f(t) < S \,,
\nonumber
\\
&&V^{app}(+\infty ,t)=0,\ V^{app}(S^{app}_f(t), t)= E - S^{app}_f(t)\,,
\\
&&V^{app}(S,T)=(E-S)^+\,.
\nonumber
\label{amer-put-appr}
\end{eqnarray}
Notice that we do not require the solution $V^{app}$ to satisfy the $C^1$ smooth pasting contact condition $ \frac{\partial V^{app}}{\partial S}(S^{app}_f(t),t)=-1$. In fact, $V^{app}$ is a solution to the put barrier option (cf. Kwok \cite{Kw}) with a given down-and-out barrier $t\mapsto S^{app}(t)$. For asset prices $0<S<S^{app}_f(t)$ we set
\[
V^{app}(S,t) = E- S.
\]
A comparison of the profile $S\mapsto V^{am}(S,t)$ of the American option price and the approximative option price $V^{app}$ is shown in Fig.~\ref{fig:comparison}. We also plot the common lower bound for both put option prices represented by the plain vanilla European put option labeled by $V^{eu}$.
Knowing the functions $t\mapsto S_f(t)$ and $t\mapsto S^{app}_f(t)$ it is not difficult to calculate the difference $V^{am}(S,t) - V^{app}(S,t)$ between option prices. Indeed, using the standard transformation (see e.g. Kwok \cite{Kw})
\[
V^{am}(S,t) = E e^{\alpha x+\beta \tau} u^{am}(x,\tau), \quad
V^{app}(S,t) = E e^{\alpha x+\beta \tau} u^{app}(x,\tau),
\]
where
\[
x=\ln(S/E), \ \tau=T-t, \qquad \alpha=\frac{1}{2} -\frac{r}{\sigma^2}, \quad \beta = -\frac{r}{2} - \frac{r^2}{2\sigma^2} - \frac{\sigma^2}{8},
\]
taking into account the fact that $V^{am}(S,t) = E-S$ for $0<S<S_f(t)$ and $V^{app}(S,t) = E-S$ for $0<S<S^{app}_f(t)$ we can conclude that $u^{am}, u^{app}$ are solution to the following Cauchy problems:
\[
\frac{\partial u^{am}}{\partial \tau} - \frac{\sigma^2}{2} \frac{\partial^2 u^{am}}{\partial x^2} =
\left\{
\begin{array}{cc}
0& \hbox{for}\ x> \ln(\varrho(\tau)/E), \\
r e^{-\alpha x-\beta \tau} & \hbox{for}\ x\le \ln(\varrho(\tau)/E),
\end{array}
\right.
\]
\[
\frac{\partial u^{app}}{\partial \tau} - \frac{\sigma^2}{2} \frac{\partial^2 u^{app}}{\partial x^2} =
\left\{
\begin{array}{cc}
0& \hbox{for}\ x> \ln(\varrho^{app}(\tau)/E), \\
r e^{-\alpha x-\beta \tau} & \hbox{for}\ x\le \ln(\varrho^{app}(\tau)/E),
\end{array}
\right.
\]
defined for $-\infty<x<\infty, 0<\tau<T,$ where $\varrho(\tau)=S_f(T-\tau), \varrho^{app}(\tau)=S^{app}_f(T-\tau)$. Notice that the difference $v(x,\tau)=u^{am}(x,\tau)-u^{app}(x,\tau)$ satisfies $v(x,0)=0$ for each $x\in\R$. Using Green's representation formula for a solution to a linear parabolic equation we obtain, after some calculations, the explicit expression for the difference of option prices
\begin{eqnarray}
&&V^{am}(S,t) - V^{app}(S,t)
\\
&&
\quad
=
r E \int_0^{\tau} \left| \int_{\ln(\varrho^{app}(\xi)/E)}^{\ln(\varrho(\xi)/E)}
G(x - s, \tau -\xi )e^{\alpha (x-s)+\beta (\tau-\xi)} ds \right| d\xi,
\nonumber
\label{difference-prices}
\end{eqnarray}
where $G(x,\tau) = e^{-x^2/(2\sigma^2\tau)}/\sqrt{2\pi\sigma^2\tau}$ is the Green function. The above difference in option prices is always nonnegative because the American option price is greater or equal to the price of a down-and-out barrier option with the prescribed barrier $S^{app}_f(t)=\varrho^{app}(T-t)$ (see Kwok \cite{Kw}).
If we evaluate this difference at the American option early exercise boundary position $S_f(t)$ then we obtain a slightly simplified expression:
\begin{eqnarray}
&&V^{am}(S_f(t),t) - V^{app}(S_f(t),t)
\\
&& \quad =
r E \int_0^{\tau} e^{-r (\tau-\xi)}
\left|
N(\tilde\gamma(\tau,\xi)) - N(\gamma(\tau,\xi))
\right|\, d\xi \nonumber
\label{difference}
\end{eqnarray}
where $\tau=T-t$ and
\[
\tilde \gamma(\tau,\xi) = \frac{\ln\frac{\varrho(\tau)}{\varrho^{app}(\xi)} + (r-\sigma^2/2) (\tau-\xi)}{\sigma\sqrt{\tau-\xi}},
\quad
\gamma(\tau,\xi) = \frac{\ln\frac{\varrho(\tau)}{\varrho(\xi)} + (r-\sigma^2/2) (\tau-\xi) }{\sigma\sqrt{\tau-\xi}}.
\]
Notice that the difference $V^{am}(S,t) - V^{eu}(S,t)$ of the American and European style of put options is rather small. Similarly, the difference $V^{am}(S,t) - V^{app}(S,t)$ is small. Therefore it is reasonable to calculate the mispricing error $V^{am}(S,t) - V^{app}(S,t)$ with respect to the benchmark mispricing difference $V^{am}(S,t) - V^{eu}(S,t)$ evaluated at $S=S_f(t)$. To this end, let us introduce the following relative mispricing error function:
\begin{equation}
err(T-t) = \frac{V^{am}(S_f(t),t) - V^{app}(S_f(t),t)}{V^{am}(S_f(t),t) - V^{eu}(S_f(t),t)}.
\label{error}
\end{equation}
The denominator of (\ref{error}) can be easily calculated by recalling that
\[
V^{am}(S_f(t),t)= E - S_f(t)
\ \ \hbox{and}\ \
V^{eu}(S,t)= E e^{-r(T-t)} N(-d_2) - S N(-d_1),
\]
where
\[
d_1 = \frac{\ln\frac{S}{E} + (r+\sigma^2/2)(T-t) }{\sigma\sqrt{T-t}}, \quad
d_2 = \frac{\ln\frac{S}{E} + (r-\sigma^2/2)(T-t) }{\sigma\sqrt{T-t}}
\]
(see Kwok \cite{Kw}).
In our practical experiment, we evaluated the relative misspricing error function $err(\tau)$ for the approximation of the early exercise boundary obtained by Zhu, i.e. we set $\varrho^{app}\equiv\varrho^{Zhu}$. In Fig.~\ref{fig:comparison2} (left) we plotted the relative error $\epsilon(\tau)$ in the early exercise boundary position
\[
\epsilon(T-t) = \frac{S_f(t)-S^{Zhu}_f(t)}{S_f(t)}
\]
between the true early exercise position $S_f(t)=\varrho(T-t)$ and Zhu's approximation $S^{Zhu}_f(t) = \varrho^{Zhu}(T-t)$. We can see that the maximal relative error in the early exercise boundary position is only $0.32\%$ and it is attained six hours prior expiration.
The relative error function $err(\tau)$ for times $\tau=T-t$ close to expiry (less than two days) is depicted in Fig.~\ref{fig:comparison2} (right). We can see that the error rapidly increases when the time $t$ approaches expiration $T$. For one day to expiration ($\tau=4\times 10^{-3}$) the error is $15\%$. It increases beyond $70\%$ as $t\to T$. This is due to the fact that Zhu's approximation underestimates the free boundary potion as $\tau=T-t\to 0^+$ (see (\ref{eq:undershoot}) ).
\begin{figure}
\begin{center}
\includegraphics[width=6.5cm]{figures/ZhuSSCh-error.ps}
\includegraphics[width=6.5cm]{figures/ZhuSSChRelPricingError.ps}
\end{center}
\caption{Comparison of the early exercise boundary $\varrho(\tau)$ and the approximative early exercise boundary $\varrho^{app}\equiv\varrho^{Zhu}$ obtained from Zhu's formula. The model parameters were chosen as: $E = 1, r = 0.1, \sigma = 0.3$ for the time $\tau=T-t\in(0, 0.006)$ close to expiration.}
\label{fig:comparison2}
\end{figure}
\section{Conclusions}
We presented qualitative and quantitative comparison of analytical approximations and numerical methods for computation the early exercise boundary position of the American put option paying zero dividends. We also proposed a new local iterative numerical scheme for construction of the entire early exercise boundary which is based on a solution to a nonlinear integral equation. We derived asymptotic behavior of approximation formulae for the time close to expiry. We proved that the asymptotic formulae by Evans, Kuske and Keller \cite{KK,EKK}, Stamicar, \v{S}ev\v{c}ovi\v{c} and Chadam \cite{SSC} have the same asymptotic behavior close to expiry. We also showed that the analytic approximation formula by Zhu \cite{Zhu2006} has a different asymptotic behavior. On the other hand, for a long time horizon, Zhu's formula yields quantitatively the same results as those of our new local iterative numerical scheme and the numerical benchmark PSOR method.
\paragraph{Acknowledgments:}
We thank the anonymous referees for their valuable comments and suggestions. This research was supported by the bilateral Slovak--Bulgarian project APVV SK-BG-0034-08.
|
\section{\label{sec:level1}Introduction}
The recent discovery of superconductivity in layered 1111 phase
quaternary compounds $R$$T$$Pn$O ($R$ = lanthinides, $T$ = Fe and
Ni, $Pn$ = P and As) has attracted tremendous attention to this
class of materials
\cite{Hosono-LaP,Hosono-LaF,ChenXH-SmOF,WangNL-CeOF,ZhaoZX-LnOD,WenHH-LaSr,WangC-GdTh,LuoJL-LaNi}.
Besides this 1111 phase oxy-pnictides, superconductivity was
subsequently discovered in other iron(nickel)-based layered
compounds with similar Fe(Ni)As layers
\cite{Johrendt-BaK,LiFeAs,WuMK-FeSe,Ogino42622}. In all the
FeAs-based parent compounds, there is a structural phase
transition in the temperature range 100-200 K, and a spin-density
wave (SDW) type antiferromagnetic (AFM) ordering associated with
Fe ions accompanies the structural transition
\cite{WangNL-SDW,DaiPC-LaNeutron,BaoW-BaNeutron}. Various chemical
doping approaches or application of high pressure can suppress the
structural transition and AFM order, and high-$T_c$
superconductivity consequently appears. For example, in the
FeAs-based La-1111 system, superconductivity has been achieved by
the chemical doping at four different crystallographic sites
\cite{Hosono-LaF,WenHH-LaSr,WangC-GdTh,Co1,Co2,LaP}. Meanwhile,
low-$T_c$ superconductivity has been observed in FeP-based
\cite{Hosono-LaP} and NiAs(P)-based \cite{LuoJL-LaNi,Hosono-NiAs}
compounds with similar layered structure, but there is neither
structural transition nor AFM ordering associated with Fe(Ni) ions
in these compounds. Furthermore, many reports have found that both
normal state properties and superconductivity of the NiAs-based
systems are likely of conventional type
\cite{LuoJL-LaNi,Hosono-NiAs}. This result suggests that there
could be a close relationship between structural transition/AFM
order and high-$T_c$ superconductivity. Moreover, the origin of
the AFM order in the parent FeAs-based pnictides is still an open
issue theoretically. Regardless of the origin of the AFM ordering,
several theories have suggested that the superconductivity is tied
to the magnetism in the FeAs-based
materials\cite{Mazin,Yildirim,XiangT,Seo2009}. All these results
imply that the mechanism of superconductivity in FeAs-based
systems could be fundamentally different to NiAs-based systems.
Systematic investigation on the physical properties of these
parent pnictides can shed light on the mechanisms of
superconductivity.
Thermoelectric effects are very sensitive to subtle changes in
electronic structure, and can provide information on the ground
state and low energy excitations. Especially, the transverse
magneto-thermoelectric effect, i.e., Nernst effect, which is
defined as the appearance of a transverse electric field $E_{y}$
in response to a temperature gradient $\nabla T || x$ in the
presence of a perpendicular magnetic field $H||z$ and under open
circuit conditions, has becomes a powerful probe in studying
exotic superconductors like high-$T_c$ cuprates \cite{zhan
nernst}, charge-density wave (CDW) superconductor NbSe$_2$
\cite{NbSe2}, heavy fermion superconductors \cite{heavy fermion},
and $p$-wave superconductor Sr$_2$RuO$_4$ \cite{SrRuO} etc. The
first report on the Nernst effect in F-doped LaFeAsO
superconductor has discovered that the vortex liquid regime below
$T_c$ is quite large \cite{ZhuZW}, consistent with the simulation
result\cite{JPLv}, and there is an enhanced anomalous Nernst
signal just above $T_c$.
Here we report the systematic investigation on the thermopower and
Nernst effect of FeAs-based and NiAs-based parent oxypnictides.
Both thermopower and Nernst coefficient of the undoped LaFeAsO
system are significantly large compared to usual metals and the
strcutural/AFM transition causes anomalous changes in
thermoelectric properties. However, the thermopower and Nernst
effect of the NiAs-based LaNiAsO system are likely of conventional
type, implying usual Fermi liquid behavior. The fundamental
difference in the thermoelectric properties of the two systems
suggests that there is a close relationship between high-$T_c$
superconductivity and anomalous thermoelectric properties in
FeAs-based systems.
\section{\label{sec:level2}Experimental}
The polycrystalline samples of LaFeAsO and LaNiAsO were prepared
by the solid state reaction using LaAs, Fe$_{2}$O$_{3}$/NiO, Fe/Ni
and LaF$_{3}$ as starting materials. The sample preparation
details can be found in the previous report \cite{LuoJL-LaNi}. The
powder X-ray diffraction patterns indicate that the resultant is
single phase and all the diffraction peaks can be well indexed
based on the tetragonal ZrCuSiAs-type structure with the space
group P4/nmm.
The resistivity was measured by usual four-probe method. The Hall
effect was measured by scanning magnetic field at fixed
temperatures. The thermoelectric properties were measured by a
steady-state technique. The temperature gradient used for the
thermoelectric measurements, measured by a pair of differential
Type E thermocouples, was around $0.5$ K/mm. All the measurements
were performed in a Quantum Design PPMS-$9$ system. The Nernst
signal $e_{y}$ is defined as $e_{y}\equiv\frac{E_{y}}{|\nabla T|}$
. The Nernst signal was measured at positive and negative field
polarities, and the difference of the two polarities was taken to
remove any thermopower contributions. The Nernst coefficient
$\nu_N$ is equal to $e_{y}$/$B$. At very low temperatures, $e_{y}$
is not strictly linear with magnetic field $H$, the Nernst
coefficient is then taken as the initial slope of the $e_{y}$
versus $\mu_0H$ curves.
\section{\label{sec:level3}Results and discussion}
\subsection{\label{sec:level1} Thermoelectric effects of LaFeAsO}
Traces of Nernst signal as a function of magnetic field up to
$\mu_0$$H$ of 8 T for the parent oxypnictide LaFeAsO is displayed
in Figure 1 at various temperatures. At high temperatures (see the
lower panel of Fig.1) , the Nernst signal ($e_{y}$) is positive
and changes linearly with magnetic field. The Nernst signal
reaches a maximum around 100 K. Below 100 K, the Nernst signal
decreases, and becomes a little nonlinear with magnetic field as
$T <$ 80 K. Between 40 to 60 K, the Nernst signal changes from
positive to negative. Such a sign change may not suggest the
change in the charge carrier type. In multi-band systems, the
contributions of different bands to the Nernst signal could have
different signs.
The temperature dependence of Nernst coefficient, $\nu_N(T)$,
together with its thermopower, $S(T)$, is shown in Figure 2. The
temperature dependence of thermopower is consistent with previous
reports \cite{Mandrus,LiLJ,LiyYK}, which exhibits a pronounced
hump as the system undergoes the structural phase transition/AFM
ordering at $T^*$ of about 160 K, where $T^*$ is the structural
phase transition temperature and it can be determined by the
resistivity measurements. The Nernst coefficient is defined as the
initial slope of $e_{y}$-$H$ curves for $T<80 K$. The Nernst
coefficient is positive and shows weak temperature dependence as
$T > T^*$ of about 160 K. However $\nu_N$ starts to increase below
$T^*$ of 160 K, reaches a maximum around 100 K, and then decreases
with decreasing temperatures. With further cooling, it changes
sign around 50 K, and then reaches a minimum (valley) around 20 K
before it finally goes to zero. Compared to the usual metals and
the normal state of high-$T_c$ cuprates, $\nu_N(T)$ is much
larger.
In order to get more insight on the thermoelectric effect, the
ratio of $\nu_N$/$S$ is calculated and plotted in the upper panel
of Fig.3. There is a very clear sharp drop in $\nu_N$/$S$ as the
system becomes AFM ordered below $T^*$. But the ratio increases
monotonously with further cooling. By measuring the Hall effect
and thermopower simultaneously, the off-diagonal thermoelectric
(Peltier conductivity) term can be separated from the Nernst
signal and then both the Hall angle tan$\theta$ (defined as
$\rho_{xy}$/$\rho$) and "thermal" Hall angle tan$\theta_{th}$
(defined as $\alpha_{xy}$/$\alpha$) can be obtained. The
temperature dependence of Hall coefficient ($R_{H}$) measured
under magnetic field ($\mu_0H$) of 8 T is also shown in the upper
panel of Fig. 3. $R_H$ is negative, suggesting that the dominant
charge carrier is electron-like, consistent with previous reports
\cite{4NLWang,Wen,23Oak}. The $R_H$ is about -4.8$\times 10^{-9}$
m$^3$/C at $T$ = 300 K, corresponding to a Hall number ($n_H$ =
1/($eR_H$)) of about 0.18 electrons per unit cell, which could be
an upper limit for the electron concentration $n_e$ because
LaFeAsO is a nearly compensated metal according to the band
calculations \cite{BandCal}. For comparison, the more metallic
BaFe$_2$As$_2$ has a Hall number of about 0.56 electrons per unit
cell at room temperature according to the Hall effect
measurements, but band calculations predict a Hall number of 0.15
electrons per unit cell \cite{Wen122}.
The normal state Nernst signal is comprised of two terms, viz.
\begin{equation}
e_{y}=\rho\alpha_{xy}-S\tan\theta=S(\tan\theta_{th}-\tan\theta),
\end{equation}
where $\alpha_{xy}$ is the off-diagonal Peltier coefficient, $S =
\frac{\alpha}{\sigma}$ the thermopower ( $\sigma$ the diagonal
conductivity and $\alpha$ the diagonal Peltier coefficient),
$\rho$ the resisitivity, tan$\theta$ the Hall angle, and
tan$\theta_{th}$ the "thermal" Hall angle. For usual simple
metals, the two terms related to the Hall angle and "thermal" Hall
angle in Eq. (1) cancel each other, which is so-called "Sondheimer
cancellation" \cite{Cancellation,Cancellation1}. It is known that
Hall angle, rather than Hall coefficient, is directly related to
the scattering rate of the quasiparticle scattering. The Hall
angle can be calculated from the Hall coefficient, i.e.,
tan$\theta$ = $\rho_{xy}$/$\rho$, and then the "thermal" Hall
angle can be obtained from the Nernst signal by using Eq.(1). The
obtained Hall angle and "thermal" Hall angle are plotted in the
lower panel of Fig.3 as a function of temperature. The two angles
exhibit very different temperature dependence, especially the
decrease in the "thermal" Hall angle is much more significant
below $T^*$. It becomes clear that the Sondheimer cancellation of
the Nernst signal is no longer held for LaFeAsO because of the
different temperature dependence of the two angles below $T^*$.
Even above $T^*$, the Nernst coefficient is also larger than that
of usual metals. The enhanced Nernst signal above $T^*$ could
result from the multi-band effect. The presence of two types of
charge carriers in multi-band systems such as
NbSe$_2$\cite{NbSe2}, could invalidate the Sondheimer
cancellation, resulting in a large Nernst signal. However it is
hard to understand the anomalous temperature dependence of $\nu_N$
below $T^*$ even in the frame of multi-band effect. There might be
significant changes in the scattering mechanism below $T^*$, and
these changes seem to have more pronounced influence on the
thermal channel. We propose that the SDW order or SDW fluctuations
could affect the spin-dependent scattering process and thus cause
significant changes in the scattering rates which could be
band-dependent. Subtle changes in the scattering mechanism could
causes anomalous Nernst effect, as observed in the p-wave
superconductor Sr$_2$RuO$_4$\cite{SrRuO}.
The first-principles band calculations have proposed that the
nesting between the electron-type Fermi surface (FS) and hole-type
FS could account for the AFM transition in the parent compounds
such as LaFeAsO, and that the superconductive pairing might be
mediated by spin fluctuations \cite{WangNL-SDW,Mazin,BandCal}.
Moreover, the studies of angle-resolved photoemission spectroscopy
(ARPES) discovered that the coexistence of hole and electron
pockets connected via the AFM wave vector is essential to
high-$T_c$ superconductivity \cite{ARPES1,ARPES2}. The AFM order
and structural phase transition in the parent compounds causes
significant changes in the electronic structure. Based on our
measurements of resistivity and Hall effect, the charge carrier
concentration in LaFeAsO estimated by the Hall number decreases by
a factor of about 100 and meanwhile the scattering rate (in
inverse proportion to the mobility) decreases by a factor of about
160 as temperature decreases from 300 K to 10 K. For the parent
compound BaFe$_2$As$_2$, similar results were reported
\cite{Wen122,Hall}. Such changes in the charge carrier
concentration and scattering rates in the parent compound
BaFe$_2$As$_2$, SrFe$_2$As$_2$, and EuFe$_2$As$_2$ were also
reported by the optical spectroscopy measurements \cite{Wang,
Eu122}. All the results suggest that the charge carrier transport
are dominated by the AFM fluctuations in the parent compounds.
When the system undergoes the AFM order below about 160 K, the
charge carrier concentration deceases sharply due to the SDW
gapping on FSs, but the scattering rates decreases even more
drastically due to the AFM order which suppresses the spin
fluctuations. It can be seen from Fig.3, the AFM order may have
more pronounced influence on the thermal channel. Thus our results
imply that the anomalous changes in the thermopower and Nernst
coefficient in LaFeAsO should have close relation with the
inter-band scattering between electron-type and hole-type bands.
\subsection{\label{sec:level2} Thermoelectric effects of LaNiAsO}
The thermopower and Hall coefficient (measured under $\mu_0H$ of 5
T) of LaNiAsO are plotted as a function of temperature in Fig.4.
The negative $R_H$ implies that the charge carriers are dominantly
electron type, same as in LaFeAsO. The absolute value of $R_H$ for
LaNiAsO is more than 1 order of magnitude smaller than that of
LaFeAsO, possibly indicating that the LaNiAsO system has a
relatively higher carrier density. The Hall number $n_H$ at $T$ =
300 K is about 8.3 electrons per unit cell, about 50 times larger
compared to LaFeAsO. Since Ni$^{2+}$3$d^8$ contributes two more
electrons than Fe ion does, the Fermi energy shifts up in LaNiAsO,
and the hole bands tend to be fully filled. As a result predicted
by band calculations\cite{BandCal}, the electron bands dominate
the conductivity in LaNiAsO \cite{LuoJL-LaNi}. The much smaller
thermopower also suggests that LaNiAsO is a good metal with high
density of charge carriers compared to LaFeAsO. The small hollow
in $S$ at low temperature could be caused by phonon drag effect.
It is interesting that $R_H$ also exhibits a sharp decrease below
50 K. Such a change in $R_H$ at low temperature has not been well
understood presently.
Traces of Nernst signal as a function of magnetic field for
LaNiAsO is displayed in Fig.5 at several selected temperatures .
The Nernst signal ($e_{y}$) is very small, comparable to the noise
level (the noise voltage is about 10 nV in our measurement
system). This Nernst signal is as small as usual metals. $e_{y}$
at $\mu_0H$ of 6 T is only in the range of $\pm$30 nV/K, almost
two order of magnitude smaller than that of LaFeAsO. The Nernst
coefficient is obtained by fitting the $e_y$ vs. $H$ curves with a
linear function, and it is plotted in Fig.6. The Hall angle
tan$\theta$ and the term $S$tan$\theta$ are also plotted in Fig.6
for comparison. It can seen that the Nernst coefficient is
comparable to the term $S$tan$\theta$, implying that the
Sondheimer cancellation is partially held and the system could be
dominated by only one type charge carrier (electron). Please note
that the temperature dependence of tan$\theta$ is very weak and
its magnitude is much smaller that that of LaFeAsO, especially at
low temperatures.
The result of Nernst effect suggests that LaNiAsO is like usual
metal and only one type (electron-type) charge carrier dominates.
Xu et al. \cite {BandCal} compared the band structures of LaFeAsO
and LaNiAsO by first-princeples calculations. It was found that
the electron FS cylinders around M point becomes larger for
LaNiAsO, and the hole-type FS cylinders around $\Gamma$ point
disappear. Thus the nesting between hole-type FSs and
electron-type FSs proposed in LaFeAsO is no longer held in
LaNiAsO. Recall that the multi-band effect could account for the
enhanced Nernst signals in both parent and F-doped LaFeAsO which
has relatively high superconducitng transition temperatures, the
inter-band scattering should indeed play an important role in the
occurrence of high $T_c$. On the other hand, our result suggests
that the LaNiAsO system lacks of such an inter-band scattering
mechanism, and thus it has a low $T_c$. The relationship between
high-$T_c$ superconductivity and the inter-band scattering is
worthy of further experimental investigations.
\section{\label{sec:level4}Conclusions}
In summary, we report the thermeropower and Nernst effect of
FeAs-based and NiAs-based parent oxypnictides. For the LaFeAsO
system, it is found that both thermopower and Nernst coefficient
are significantly large compared to usual metals and the
structural/AFM transition associated with Fe ions causes
significant enhancements in the thermoelectric coefficients. We
propose that the unique strong inter-band scattering as well as
electron correlation in FeAs-based systems may account for these
anomalies in thermoelectric properties. Meanwhile, the thermopower
and Nernst effect of the NiAs-based LaNiAsO system are of
conventional type, implying usual Fermi liquid behavior. The
fundamental difference in the thermoelectric properties of the two
systems suggests that there is a close relationship between
high-$Tc$ superconductivity and anomalous thermoelectric
properties in FeAs-based pnictides.
\section*{Acknowledgments}
This project is supported by the National Science Foundation of
China (Grant. No.10628408 and 10931160425) and the National Basic
Research Program of China (Grant No. 2006CB601003 and
2007CB925001).
\section*{References}
|
\section{Introduction}
Testing whether a signal lies within a subspace is a problem arising in a wide range of applications including medical~\cite{medimag} and hyperspectral~\cite{hyper} imaging, communications~\cite{multiaccess}, radar~\cite{radar}, and anomaly detection~\cite{hyper2}. The classical formulation of this problem is a binary hypothesis test of the following form. Let $v \in \mathbb{R}^n$ denote a signal and let $x = v + w$, where $w$ is a noise of known distribution. We are given a subspace $S \subset \mathbb{R}^n$ and we wish to decide if $v \in S$ or not, based on $x$. Tests are usually based on some measure of the energy of $x$ in the subspace $S$, and these `matched subspace detectors' enjoy optimal properties \cite{scharf, scharfpap}.
This paper considers a variation on this classical problem, motivated by high-dimensional applications where it is prohibitive or impossible to measure $v$ completely. We assume that only a small subset $\Omega \subset \{1,\dots,n\}$ of the elements of $v$ are observed (with or without noise), and based on these observations we want to test whether $v \in S$. For example, consider monitoring a large networked system such as a portion
of the Internet. Measurement nodes in the network may have software that
collects measurements such as upload and download rate, number of packets,
or type of traffic given by the packet headers. In order to monitor the
network these measurements will be collected in a central place for compilation,
modeling and analysis. The effective dimension of the state of such systems
is often much lower than the extrinsic dimension of the network itself.
Subspace detection, therefore, can be a useful tool for detecting changes or
anomalies. The challenge is that it may be impossible to obtain every
measurement from every point in the network due to resource constraints,
node outages, etc.
The main result of this paper answers the following question. Given a subspace $S$ of dimension $r\ll n$, how many elements of $v$ must be observed so that we can reliably decide if it belongs to $S$? The answer is that, under some mild incoherence conditions,
the number is $O(r \log r)$. This means that reliable matched subspace detectors can be constructed from very few measurements, making them scalable and applicable to large-scale testing problems.
The main focus of this paper is an estimator of the energy of $v$ in $S$ based on only observing the elements $\{v_i\}_{i \in \Omega}$. Section~\ref{sec:estimator} proposes the estimator. Section~\ref{sec:main} presents a theorem giving quantitative bounds on the estimator's performance and the proof using three lemmas that are proved in the Appendix. Section~\ref{sec:exp} presents numerical experiments. Section~\ref{sec:detect} applies the
main result to the subspace detection problem, both with and without noise.
\section{Energy Estimation from Incomplete Data}
\label{sec:estimator}
Let $v_\Omega$ be the vector of dimension $|\Omega|\times 1$ comprised of the elements $v_{i}$, $i\in \Omega$, ordered lexigraphically; here $|\Omega|$ denotes the cardinality of $\Omega$. The energy of $v$ in the subspace $S$ is $\|P_S v||_2^2$, where $P_S$ denotes the projection operator onto $S$. There are two natural estimators of $\|P_S v||_2^2$ based on $v_\Omega$. The first is simply to form the $n\times 1$ vector $\widetilde v$ with elements $v_i$ if $i \in \Omega$ and zero if $i \not \in \Omega$, for $i=1,\dots,n$.
This `zero-filled' vector yields the simple estimator $\|P_S \widetilde v\|_2^2$. Filling missing elements with zero is a fairly common, albeit na\"{i}ve, approach to dealing with missing data. Unfortunately, the estimator $\|P_s \widetilde v\|_2^2$ is fundamentally flawed. Even if $v \in S$, the zero-filled vector $\widetilde v$ does not necessarily lie in $S$.
A better estimator can be constructed as follows. Let $U$ be an $n \times r$ matrix whose columns span the $r$-dimensional subspace $S$. Note that for any such $U$, $P_S = U (U^T U)^{-1} U^T$. With this representation in mind, let $U_\Omega$ denote the $|\Omega| \times r$ matrix, whose rows are the $|\Omega|$ rows of $U$ indexed by the set $\Omega$, arranged in lexigraphic order. Since we only observe $v$ on the set $\Omega$, another approach to estimating its energy in $S$ is to assess how well $v_\Omega$ can be represented in terms of the rows of $U_\Omega$. Define the projection operator $P_{S_\Omega} := U_\Omega (U^T_\Omega U_\Omega)^{\dagger}U_\Omega^T$, where $^\dagger$ denotes the pseudoinverse. It follows immediately that if $v \in S$, then $\|v - P_S v||_2^2 = 0$ and $\|v_\Omega - P_{S_\Omega} v_\Omega\|_2^2 = 0$, whereas $\|\widetilde v - P_S \widetilde v\|_2^2$ can be significantly greater than zero. This property makes $\| P_{S_\Omega} v_\Omega\|_2^2$ a much better candidate estimator than $\|P_S \widetilde v\|_2^2$. However,
if $|\Omega| \leq r$, then it
it is possible that
$\|v_\Omega - P_{S_\Omega} v_\Omega\|_2^2 = 0$,
even if $\|v - P_S v||_2^2 >0$.
Our main result shows that if $|\Omega|$ is just slightly greater than
$r$, then with high probability $\|v_\Omega - P_{S_\Omega} v_\Omega\|_2^2$
is very close to $\frac{|\Omega|}{n} \|v - P_S v||_2^2$.
\section{Main Theorem}
\label{sec:main}
Let us now focus on our main goal of detecting from a very small number of samples whether there is energy in a vector $v$ outside the $r$-dimensional subspace $S$. In order to do so, we must first quantify how much information we can expect each sample to provide. The authors in~\cite{candesrecht} defined the \emph{coherence} of a subspace $S$ to be the quantity
$$\mu(S) := \frac{n}{r} \max_j \| P_{S} e_j \|_2^2\,.$$ That is, $\mu(S)$ measures the maximum magnitude attainable by projecting a standard basis element onto $S$. Note that $1 \leq \mu(S) \leq \tfrac{n}{r}$. The minimum $\mu(S)=1$ can be attained by looking at the span of any $r$ columns of the discrete Fourier transform. Any subspace that contains a standard basis element will maximize $\mu(S)$. For a vector $z$, we let $\mu(z)$ denote the coherence of the subspace spanned by $z$. By plugging in the definition, we have $$\mu(z) = \frac{n \|z\|_\infty^2}{\|z\|_2^2}\,.$$
To state our main theorem, write $v=x+y$ where $x \in S$ and $y \in S^{\perp}$. Let the entries of $v$ be sampled uniformly with replacement. Again let $\Omega$ refer to the set of indices for observations of entries in $v$, and denote $|\Omega| = m$. Given these conventions, we have the following.
\begin{theorem}
Let $\delta > 0$ and $m \geq \frac{8}{3} r\mu(S) \log\left(\frac{2r}{\delta}\right)$. Then with probability at least $1-4\delta$,
\begin{equation*}
\frac{m(1-\alpha) - r\mu(S)\frac{(1+\beta)^2}{(1-\gamma)}}{n} \|v - P_S v\|_2^2 \leq \|v_\Omega - P_{S_\Omega} v_\Omega \|_2^2
\end{equation*} and
\begin{equation*}
\|v_\Omega - P_{S_\Omega} v_\Omega \|_2^2 \leq (1+\alpha)\frac{m}{n}\|v - P_S v\|_2^2
\end{equation*} where $\alpha = \sqrt{\frac{2\mu(y)^2}{m}\log\left(\frac{1}{\delta}\right)}$, $\beta = \sqrt{2\mu(y)\log\left(\frac{1}{\delta}\right)}$, and $ \gamma = \sqrt{\frac{8r\mu(S)}{3m} \log\left(\frac{2r}{\delta}\right)}$.
\label{mainthm}
\end{theorem}
\begin{proof}
In order to prove the theorem, we split the quantity of interest into three terms and bound each with high probability. Consider $\|v_\Omega - P_{S_\Omega}v_\Omega\|_2^2 = \|y_\Omega - P_{S_\Omega} y_\Omega \|_2^2$. Let the $r$ columns of $U$ be an orthonormal basis for the subspace $S$. We want to show that
\begin{equation}
\|y_\Omega - P_{S_\Omega} y_\Omega \|_2^2 = \|y_\Omega\|_2^2 - y_\Omega^T U_\Omega\left(U_\Omega^TU_\Omega\right)^{-1}U_\Omega^Ty_\Omega
\label{term}
\end{equation}
is near $\frac{m}{n} \|y\|_2^2$ with high probability. To proceed, we need the following three Lemmas whose proofs can be found in the Appendix.
\begin{lemma}
With the same notations as Theorem~\ref{mainthm},
\begin{equation*}
(1-\alpha)\frac{m}{n}\|y\|_2^2 \leq \|y_\Omega\|_2^2 \leq (1+\alpha)\frac{m}{n}\|y\|_2^2
\label{normy}
\end{equation*}
with probability at least $1-2\delta$. \label{normybound}
\end{lemma}
\begin{lemma}
With the same notations as Theorem~\ref{mainthm},
$$ \|U_\Omega^Ty_\Omega\|_2^2 \leq (\beta+1)^2 \frac{m}{n} \frac{r\mu(S)}{n} \|y\|_2^2$$ with probability at least $1-\delta$.
\label{normutybound}
\end{lemma}
\begin{lemma}
With the same notations as Theorem~\ref{mainthm},
$$\|\left(U_\Omega^TU_\Omega\right)^{-1}\|_2 \leq \frac{n}{(1-\gamma)m}$$
with probability at least $1-\delta$, provided that $\gamma<1$.
\label{uuinvbound}
\end{lemma}
To apply these three Lemmas, write the second term of Equation (\ref{term}) as
$$ y_\Omega^TU_\Omega\left(U_\Omega^TU_\Omega\right)^{-1}U_\Omega^Ty_\Omega = \|W_\Omega U_\Omega^T y_\Omega\|_2^2$$ where $W_\Omega^TW_\Omega = \left(U_\Omega^TU_\Omega\right)^{-1}$. By Lemma~\ref{uuinvbound}, $U_\Omega^T U_\Omega$ is invertible under the assumptions of our theorem, and hence $W_\Omega$ is well-defined and has spectral norm bounded by the square root of the inverse of the smallest eigenvalue of $U_\Omega^T U_\Omega$. That is, we have
\begin{eqnarray*}
\|W_\Omega U_\Omega^T y_\Omega\|_2^2 &\leq& \|W_\Omega\|_2^2 \|U_\Omega^T y_\Omega\|_2^2\\
&=& \|W_\Omega^TW_\Omega\|_2 \|U_\Omega^T y_\Omega\|_2^2 \\
&=& \|\left(U_\Omega^TU_\Omega\right)^{-1}\|_2 \|U_\Omega^Ty_\Omega\|_2^2\,.
\end{eqnarray*}
$\|\left(U_\Omega^TU_\Omega\right)^{-1}\|_2$ is bounded by Lemma~\ref{uuinvbound} and $ \|U_\Omega^Ty_\Omega\|_2$ is bounded by Lemma~\ref{normutybound}. Putting these two bounds together with the bounds in Lemma~\ref{normybound} and using the union bound, we have that with probability at least $1-4\delta$
\begin{eqnarray*}
(1+\alpha)^2\frac{m}{n}\|y\|_2^2 \geq \|y_\Omega\|_2^2 - \|\left(U_\Omega^TU_\Omega\right)^{-1}\|_2 \|U_\Omega^Ty_\Omega\|_2^2 \\
\geq (1-\alpha)^2 \frac{m}{n}\|y\|_2^2 - \frac{(\beta+1)^2r\mu(S)}{(1-\gamma)n} \|y\|_2^2
\end{eqnarray*}
giving us our bound.
\end{proof}
\section{Discussion and Numerical Experiments}
\label{sec:exp}
In this section we wish to give some intuition for the lower bound in Theorem~\ref{mainthm} and show simulations of the estimate $\|v_\Omega - P_{S_\Omega} v_\Omega\|_2$. If the parameters $\alpha, \beta, \gamma$ are very near $0$, our lower bound is approximately equal to $$\frac{m-r\mu(S)}{n} \|v - P_S v\|_2$$ For an incoherent subspace, the parameter $\mu(S) = 1$. In this case, for $m\leq r$ the bound is $\leq 0$, which is consistent with the fact that $\|v_\Omega - P_{S_\Omega} v_\Omega\|_2 = 0$ always for $m \leq r$. Once $m \geq r+1$, linear algebraic reasoning tells us that $\|v_\Omega - P_{S_\Omega} v_\Omega\|_2$ will be strictly positive with positive probability; Theorem~\ref{mainthm} goes further to say the norm is strictly positive with high probability once $m \sim O(r log r)$.
The parameters $\alpha, \beta, \gamma$ all depend on $\sqrt{\log\left(\frac{1}{\delta}\right)}$; these parameters grow as $\delta$ gets very small. Increasing the number of observations $m$ will counteract this behavior for $\alpha$ and $\gamma$, but this does not hold for $\beta$. In fact, even if the vector $y$ is incoherent and $\mu(y)=1$, its minimum value, then $\beta = 2$ for $\delta \approx .135$. To get $\beta$ very near zero, $\delta$ must be \emph{very} near one, but this is not a useful regime.
We can see, however, that in simulations these large constants are somewhat irrelevant; The large deviations analysis needed for the proof is overly conservative in most cases.
This plays out in the simulations shown in Figure~\ref{fig_sim}, where we see that for very incoherent subspaces, $ \|v_\Omega - P_{S_\Omega} v_\Omega\|_2 $ is always positive for $m > r \mu(S) \log r$. The plots show the minimum, maximum and mean value of $ \|v_\Omega - P_{S_\Omega} v_\Omega\|_2 $ over 100 simulations, for fixed $S$ and fixed $v$ such that $\|v\|_2^2 = 1$ and $v \in S^\perp$. For each value of the sample size $m$, we sampled 100 different instances of $\Omega$\emph{without} replacement, giving us a realistic idea of how much energy of $v$ is captured by $m$ samples. Our simulations for the Fourier basis and a basis made of orthogonalized Gaussian random vectors always showed the estimate to be positive for $m > r \mu(S) \log r$, even for the worst-case simulation run. For more coherent subspaces, we often (but not always) see that the norm is positive as long as $m > r\mu(S) \log r$.
\begin{figure}[!t]
\centerline{\subfloat[Incoherent subspace (random Gaussian basis). $\mu(S)\approx 1.5$, $\mu(y) \approx 13.6$.]{\includegraphics[width=1.5in]{fig_simprojerr_gauss.pdf}\label{fig_first_case}}
\hfil
\subfloat[Coherent subspace. $\mu(S)\approx4.1$, $\mu(y)\approx47.0$.]{\includegraphics[width=1.5in]{fig_simprojerr_fives.pdf}\label{fig_second_case}}}
\caption{These plots show the projection residual $\|v_\Omega - P_{S_\Omega}v_\Omega\|_2^2$ over 100 simulations. Each of the simulations has a fixed subspace, vector $v \in S^\perp$ and sample size $m$, but different sample set $\Omega$ drawn \emph{without} replacement. The problem size is $n=10000$, $r=50$. }
\label{fig_sim}
\end{figure}
\begin{figure}
\centering
\includegraphics[width=1.5in]{fig_zerofilled}
\caption{Simulation results for the zero-filling approach, $v \in S$, $\|v\|_2^2 = 1$. The basis used is a random Gaussian basis, $r=50$, $n=10000$, $\mu(S)\approx1.5$, $\mu(y)\approx17.9$.
Note that the zero-filled residuals can be made arbitrarily large by increasing $\|v\|_2^2$.}
\label{fig_zerofill}
\end{figure}
\section{Matched Subspace Detection}
\label{sec:detect}
We have the following detection set up. Our hypotheses are $\mathcal{H}_0 : v \in S$ and
$\mathcal{H}_1 : v \notin S$
and the test statistic we will use is $$t(v_\Omega) = \|v_\Omega - P_{S_\Omega} v_\Omega \|_2^2 \overset{{{\cal H}_1}}{\underset{{{\cal H}_0}}{\gtrless}} \eta$$
In the noiseless case, we can let $\eta = 0$; our result in Theorem~\ref{mainthm} shows for $\delta>0$, the probability of detection is $\mathit{P}_D = \mathbb{P}\left[t(v_\Omega) > 0 | \mathcal{H}_1 \right] \geq 1-4\delta$ as long as $m$ is large enough, and we also have that the probability of false alarm is zero, $\mathit{P}_{FA} = \mathbb{P} \left[ t(v_\Omega) > 0 | \mathcal{H}_0 \right] = 0$ since the projection error will be zero when $v \in S$.
When we introduce noise we have the same hypotheses, but we compute the statistic on $\widetilde{v}_\Omega = v_\Omega + w$ where $w \sim \mathcal{N}(0,1)$ is Gaussian white noise: $$t(\widetilde{v}_\Omega) = \|\widetilde{v}_\Omega - P_{S_\Omega} \widetilde{v}_\Omega \|_2^2 \overset{{{\cal H}_1}}{\underset{{{\cal H}_0}}{\gtrless}} \eta_\lambda$$ We choose $\eta_\lambda$ to fix the probability of false alarm: $$\mathbb{P} \left[ t(\widetilde{v}_\Omega) > \eta_\lambda | \mathcal{H}_0 \right] \leq \lambda = \mathit{P}_{FA}$$ Then we have from \cite{scharf} that $t(\tilde{v}_\Omega)$ is distributed as a non-central $\chi^2$ with $r$ degrees of freedom and non-centrality parameter $\|v_\Omega - P_{S_\Omega} v_\Omega \|_2^2$, and that $\mathit{P}_D$ is monotonically increasing with the non-centrality parameter. Putting this together with Theorem~\ref{mainthm} we see that as $m$ grows, $\|v_\Omega - P_{S_\Omega} v_\Omega \|_2^2$ grows and thus the probability of detection grows.
We now show why the heuristic approach of zero-filling the incomplete vector $v_\Omega$ does not work.
As we described in Section~\ref{sec:estimator}, the zero-filling approach is to fill the vector $v$ with zeros and then project onto the full subspace $S$. We denote the zero-filled vector as $v_0$ and then calculate the projection energy only on the observed entries: $$t_0(v_\Omega) = \|v_\Omega - \left( P_S v_0 \right)_\Omega \|_2^2 \overset{{{\cal H}_1}}{\underset{{{\cal H}_0}}{\gtrless}} \eta$$
Simple algebraic consideration reveals that $t_0(v_\Omega)|\mathcal{H}_0$ is positive. In fact, even in the absence of noise, the probability of false alarm can be arbitrarily large as $\|v\|_2^2$ increases. The value of $t_0(v_\Omega)|\mathcal{H}_0$, based on noiseless observations, is plotted as a function of the number of measurements in Figure~\ref{fig_zerofill}.
We note that for unknown noise power or structured interference, these results can be extended using the GLRT \cite{scharfpap}.
\section{Conclusion}
We have shown that it is possible to detect whether a highly incomplete vector has energy outside a subspace. This is a fundamental result to add to a burgeoning collection of results for incomplete data analysis given a low-rank assumption. Missing data are the norm and not the exception in any massive data collection system, so this result has implications on many other areas of study.
One of our reviewers shared an insight that the process by which we observe some components and observe erasures in other components can be expressed as a projection operator. It may be possible to extend the results of Theorem~\ref{mainthm} to a wide class of models of random projection operators beyond the class of deletion operators studied here.
\section*{Acknowledgments}
The authors would like to thank the reviewers for their thoughtful comments. This work was supported in part by AFOSR grant FA9550-09-1-0140.
|
\section{Introduction}
The nuclear symmetry energy is the energy needed per nucleon to
convert all protons in a symmetric nuclear matter to neutrons.
Knowledge on the density dependence of nuclear symmetry energy is
important for understanding the dynamics of heavy ion collisions
induced by radioactive beams, the structure of exotic nuclei with
large neutron or proton excess, and many important issues in nuclear
astrophysics~\cite{LCK08,ireview98,ibook,baran05}. At normal nuclear
matter density, the nuclear symmetry energy has long been known to
have a value of about $30$~MeV from fitting the binding energies of
atomic nuclei with the liquid-drop mass formula. Somewhat stringent
constraints on the nuclear symmetry energy below the normal nuclear
density have also been obtained during past few years from studies
of the isospin diffusion~\cite{Tsa04,Liu07,Che05a,LiBA05c} and
isoscaling~\cite{She07} in heavy-ion reactions, the size of neutron
skins in heavy nuclei~\cite{Ste05b}, and the isotope dependence of
giant monopole resonances in even-A Sn isotopes~\cite{Gar07}. For
nuclear symmetry energy at high densities, transport model studies
have shown that the ratio of negatively to positively charged pions
produced in heavy ion collisions with neutron-rich nuclei is
sensitive to its stiffness~\cite{bali02,Ferini:2006je}. Comparison
of this ratio from an isospin-dependent Boltzmann-Uehling-Ulenbeck
(IBUU) transport model based on the non-relativistic
momentum-dependent (MDI) nuclear effective
interactions~\cite{xiao09} with measured data from heavy ion
collisions by the FOPI Collaboration~\cite{FOPI} at GSI seems to
indicate that the nuclear symmetry energy at high density might be
very soft. Although this study does not include the relativistic
effects, which may affect the charged pion ratio as shown in
Ref~\cite{Ferrini:2005jw}, it provides an important step in the
determination of the nuclear symmetry energy at high densities.
The transport model used in Ref.~\cite{xiao09} neglects, however,
medium effects on pions, although it includes those on nucleons and
produced $\Delta$ resonances through their isospin-dependent
mean-field potentials and scattering cross sections. It is
well-known that pions interact strongly in nuclear medium as a
result of their $p$-wave couplings to the
nucleon-particle--nucleon-hole and delta-particle--nucleon-hole
($\Delta$-hole) excitations, leading to the softening of their
dispersion relations or an increased strength of their spectral
functions at low
energies~\cite{weise75,friedmann81,oset82,xia94,hees05,korpa08}.
Including pion medium effects in the transport model has previously
been shown to enhance the production of low energy pions in high
energy heavy ion collisions, although it does not affect the total
pion yield~\cite{xiong93}. Since pions of different charges are
modified differently in asymmetric nuclear matter that
has unequal proton and neutron fractions~\cite{korpa99},
including such isospin-dependent medium effects is expected to
affect the ratio of negatively to positively charged pions produced
in heavy ion collisions.
\section{Pion $p$-wave interactions in nuclear medium}
Considering only the dominant $\Delta$-hole excitations as in
Ref.~\cite{ko89}, as the contribution from the nucleon particle-hole
excitations is known to be small, the self-energy of a pion of
isospin state $m_t$, energy $\omega$, and momentum $k$ in a hot
nuclear medium due to its $p$-wave interaction is given by
\begin{eqnarray}\label{pi}
\Pi_0^{m_t} &\approx& \frac{4}{3} \left(
\frac{f_\Delta^{}}{m_\pi^{}} \right)^2 k^2 F_\pi^2(k)
\sum_{m_\tau,m_T^{}} \left|\left\langle {\textstyle\frac{3}{2}} \,
m_T^{} | 1\, m_t\, {\textstyle\frac{1}{2}} \, m_\tau \right\rangle\right|^2\notag\\
&\times& \int \frac{d^3p}{(2\pi)^3}\frac{1}{e^{(m_N^{} + p^2/2m_N^{}
+U_N^{m_\tau}-\mu_B^{}-2m_\tau\mu_Q^{})/T}+1}
\left(\frac{1}{\omega-\omega_{m_T^{}}^{+}}+\frac{1}{-\omega-\omega_{m_T^{}}^{-}}\right),\notag\\
\end{eqnarray}
with $\omega_{m_T^{}}^{\pm} \approx m_\Delta^{} +U_\Delta^{m_T^{}} +
(\vec{k} \pm \vec{p})^2/2m_\Delta^{}-i\Gamma_\Delta^{m_T^{}}/2
-m_N^{}-U_N^{m_\tau} - p^2/2 m_N^{}$. In the above, $m_\pi \simeq
138$~MeV, $m_N^{} \simeq 939$~MeV, and $m_\Delta^{} \simeq 1232$~MeV
are the masses of pion, nucleon, and $\Delta$ resonance,
respectively; $f_\Delta^{} \simeq 3.5$ is the $\pi N\Delta$ coupling
constant and $F_\pi(k) = [1+0.6(k^2/m^2_\pi)]^{-1/2}$~\cite{art} is
the $\pi N\Delta$ form factor determined by fitting the decay width
$\Gamma_\Delta\simeq 118$~MeV of $\Delta$ resonance in free space.
The summation in Eq.~(\ref{pi}) is over the nucleon isospin state
$m_\tau$, and the $\Delta$ resonance isospin state $m_T^{}$; and the
factor $\langle {\textstyle\frac{3}{2}} \, m_T^{} | 1\,
m_t\,{\textstyle\frac{1}{2}} \, m_\tau \rangle$ is the
Clebsch-Gordan coefficient from the isospin coupling of pion with
nucleon and $\Delta$ resonance. The momentum integration is over
that of nucleons in the nuclear matter given by a Fermi-Dirac
distribution with $\mu_B$ and $\mu_Q$ being, respectively, the
baryon and charge chemical potentials determined by charge and
baryon number conservations; $\rho_N^{m_\tau}$ and $U_N^{m_\tau}$
are, respectively, the density and mean-field potential of nucleons
of isospin state $m_\tau$ in asymmetric nuclear matter; and
$\Gamma_\Delta^{m_T^{}}$ and $U_\Delta^{m_T^{}}$ are, respectively,
the width and mean-field potential of $\Delta$ resonance of isospin
state $m_T^{}$.
For the nucleon mean-field potential $U_N^{m_\tau}$, we have used
the one obtained from the momentum-independent (MID)
interaction~\cite{LCK08}, i.e., $U_N^{m_\tau}(\rho_B^{},\delta_{\rm like}) =
\alpha(\rho_B^{}/\rho_0^{}) + \beta(\rho_B^{}/\rho_0^{})^\gamma +
U_{\text{asy}}^{m_\tau}(\rho_B^{} ,\delta_{\rm like})$, with
$U_{\text{asy}}^{m_\tau}(\rho_B^{} ,\delta_{\rm like})= -4\{ F(x)(\rho_B^{}/\rho_0^{}) + [18.6 -
F(x)] (\rho_B^{}/\rho_0^{})^{G(x)}\} m_\tau \delta_{\rm like} + [18.6 -
F(x)][G(x) - 1] (\rho_B^{}/\rho_0^{})^{G(x)} {\delta_{\rm
like}}^{2}$ being the nucleon symmetry potential. The parameters
$\alpha=-293.4$~MeV, $\beta=240.1$~MeV, and $\gamma=1.216$ are
chosen to give a compressibility of $212$~MeV and a binding energy
per nucleon of $-16$~MeV for symmetric nuclear matter at the
saturation or normal nuclear density $\rho_0^{} = 0.16~{\rm
fm}^{-3}$. The nucleon symmetry potential
$U_{\rm asy}^{m_\tau}(\rho_B^{} ,\delta_{\rm like})$
depends on the baryon density $\rho_B^{} = \rho_n^{} + \rho_p^{} +
\rho_{\Delta^-}^{} + \rho_{\Delta^0}^{} + \rho_{\Delta^+}^{} +
\rho_{\Delta^{++}}^{}$ and the isospin asymmetry $\delta_{\rm
like}=(\rho_n^{} - \rho_p^{} + \rho_{\Delta^-}^{} -
\rho_{\Delta^{++}}^{}+ \rho_{\Delta^0}^{}/3 -
\rho_{\Delta^+}^{}/3)/\rho_B^{}$ of the asymmetric hadronic matter,
which is a generalization of the isospin asymmetry $\delta =
(\rho_n^{} - \rho_p^{})/(\rho_n^{} + \rho_p^{})$ usually defined for
asymmetric nuclear matter without $\Delta$ resonances~\cite{bali02}.
The nucleon mean-field potential also depends on the stiffness of
nuclear symmetry energy through the parameter $x$ via the functions
$F(x)$ and $G(x)$. We consider the three cases of $x=0$, $x=0.5$,
and $x=1$ with corresponding values $F(x=0) = 129.98$ and $G(x=0) =
1.059$, $F(x=0.5) = 85.54$ and $G(x=0.5) = 1.212$, and $F(x=1) =
107.23$ and $G(x=1) = 1.246$. The resulting nuclear symmetry energy
becomes increasingly softer as the value of $x$ increases, with
$x=1$ giving a nuclear symmetry energy that becomes negative at
about 3 times the normal nuclear matter density. These symmetry
energies reflect the uncertainties in the theoretical predictions on
the stiffness of nuclear symmetry energy at high densities. For the
mean-field potentials of $\Delta$ resonances, their isoscalar
potentials are assumed to be the same as those of nucleons, and
their symmetry potentials are taken to be the average of those for
neutrons and protons with weighting factors depending on the charge
state of $\Delta$ resonance~\cite{art}, i.e., $U_{\rm
asy}^{\Delta^{++}} = U_{\rm asy}^p$, $U_{\rm asy}^{\Delta^+} =
{\textstyle\frac{2}{3}} U_{\rm asy}^p + {\textstyle\frac{1}{3}}
U_{\rm asy}^n$, $U_{\rm asy}^{\Delta^0} = {\textstyle\frac{1}{3}}
U_{\rm asy}^p + {\textstyle\frac{2}{3}} U_{\rm asy}^n$, and $U_{\rm
asy}^{\Delta^-} = U_{\rm asy}^n$.
Including the short-range $\Delta$-hole repulsive interaction via
the Migdal parameter $g^\prime$, which has values $1/3 \le g^\prime
\le 0.6$~\cite{weise75,friedmann81,oset82,xia94,hees05,korpa08},
modifies the pion self-energy to $\Pi^{m_t} =
\Pi_0^{m_t}/(1-g^\prime\Pi_0^{m_t}/k^2)$. The pion spectral function
$S_\pi^{m_t}(\omega,k)$ is then related to the imaginary part of its
in-medium propagator $D^{m_t}(\omega,k) =
1/[\omega^2-k^2-m_\pi^2-\Pi^{m_t}(\omega,k)]$ via
$S_\pi^{m_t}(\omega,k) = -(1/\pi)\,\mbox{Im}\,D^{m_t}(\omega,k)$.
The modification of the pion properties in nuclear medium affects
the decay width and mass distribution of $\Delta$ resonance. For a
$\Delta$ resonance of isospin state $m_T^{}$ and mass $M$ and at
rest in nuclear matter, its decay width is then given by~\cite{ko89}
\begin{eqnarray}\label{gamma}
&&\Gamma_\Delta^{m_T^{}}(M)\approx -2 \sum_{m_\tau,m_t} |\langle {\textstyle\frac{3}{2}} \,
m_T^{} | 1\, m_t\, {\textstyle\frac{1}{2}} \, m_\tau \rangle|^2 \int \frac{d^3{\bf k}}{(2\pi)^3} \left(
\frac{f_\Delta^{}}{m_\pi^{}} \right)^2 F_\pi^2(k)\nonumber\\
&&\times\left[\frac{1}{z_\pi^{-1}e^{(\omega-m_t\mu_Q^{})/T}-1}+1\right]\left[1-\frac{1}{e^{(m_N^{} + k^2/2m_N^{} +U_N^{m_\tau}-\mu_B^{}-2m_\tau\mu_Q^{})/T}+1}\right]\\
&&\times\mbox{Im}\, \left[\frac{k^2}{3}\frac{D^{m_t}(\omega,k)}
{(1-g^\prime\Pi_0^{m_t}(\omega,k)/k^2)^2} +
{g^\prime}^2\frac{\Pi^{m_t}(\omega,k)}{k^2} \right].\nonumber
\end{eqnarray}
In the above, the first term in the last line is due to the decay of
the $\Delta$ resonance to pion but corrected by the contact
interaction at the $\pi N\Delta$ vertex, while the second term
contains the contribution from its decay to the $\Delta$-hole state
without coupling to pion. The first two factors in the momentum
integral take into account, respectively, the Bose enhancement for
the pion and the Pauli blocking of the nucleon. To
include possible chemical non-equilibrium effect, a fugacity
parameter $z_\pi^{}$ is introduced for pions.
The pion energy $\omega$ is determined from energy conservation,
i.e., $M + U_\Delta^{m_T^{}} = \omega + m_N^{} + k^2/2m_N^{} +
U_N^{m_\tau}$. The resulting mass distribution of $\Delta$
resonances is then given by $P_\Delta(M) =
A[\Gamma_\Delta^{m_T^{}}(M)/2]/[(M-m_\Delta^{})^2
+{\Gamma_\Delta^{m_T^{}}}^2(M)/4]$, where $A$ is a normalization
constant to ensure the integration of $P_\Delta(M)$ over $M$ is one.
\begin{figure}[h]
\centerline{\includegraphics[width=4.5in,height=4.5in,angle=0]{fig1.EPS}}
\caption{(Color online) Spectral functions of pions in asymmetric
nuclear matter of density $2\rho_0^{}$ and isospin asymmetry
$\delta_{\rm like}=0.133$ as functions of pion energy for different
pion momenta of (a) $m_\pi$, (b) $2 m_\pi$, (c) $3 m_\pi$, and (d)
$4 m_\pi$. All are calculated with the Migdal parameter
$g^\prime=1/3$.}\label{pion}
\end{figure}
\begin{figure}[h]
\centerline{\includegraphics[width=3.5in,height=3.5in,angle=0]{fig2.EPS}}
\caption{(Color online) Mass distributions of $\Delta$ resonances at
rest in asymmetric nuclear matter of density $2\rho_0^{}$ and
isospin asymmetry $\delta_{\rm like}=0.133$. The solid line
corresponds to that in free space. The distributions near the
threshold and at the peak are enlarged in the insets.} \label{delta}
\end{figure}
We have solved Eqs.~(\ref{pi}) and (\ref{gamma}) self-consistently
to obtain the pion spectral functions and the mass distributions of
$\Delta$ resonances in asymmetric nuclear matter. The results
obtained with the Migdal parameter $g^\prime = 1/3$ are illustrated
in Fig.~\ref{pion} and Fig.~\ref{delta} for an asymmetric nuclear
matter of isospin asymmetry $\delta_{\rm like}\simeq 0.133$, twice
the normal nuclear matter density $\rho_B^{} = 2\rho_0^{}$,
temperature $T\simeq 43.6~{\rm MeV}$, and chemical potentials
$\mu_B^{} \simeq 941.89~{\rm MeV}$ and $\mu_Q^{} \simeq -18.26~{\rm
MeV}$, corresponding to those to be used in our thermal model and
also similar to those reached in the transport model with the
nuclear symmetry energy $x=1$ for central Au+Au collisions at the
beam energy of $0.4~{\rm AGeV}$~\cite{xiao09}. Shown in
Fig.~\ref{pion} are the pion spectral functions as functions of pion
energy for different values of pion momentum. It is seen that for
low pion momenta the spectral function at low energies has a larger
strength for $\pi^-$ (dotted line) than for $\pi^0$ (solid line),
which has a strength larger than that for $\pi^+$ (dashed line).
This behavior is reversed for high pion energies. Fig.~\ref{delta}
shows the mass distributions of $\Delta$ resonances at rest in
asymmetric nuclear matter as functions of mass. One sees that they
are similar to that in free space (solid line) as a result of the
cancelation between the pion in-medium effects, which
enhance the strength at low masses, and the Pauli-blocking
of the nucleon from delta decay, which reduces the strength at low
masses. This is consistent with the observed similar energy
dependence of the photo-proton and photo-nucleus absorption cross
sections around the $\Delta$ resonance mass~\cite{vanPee:2007tw}.
Furthermore, the strength around the peak and near the threshold of
the $\Delta$ resonance mass distribution slightly decreases with
increasing charge of the $\Delta$ resonance due to nonzero isospin
asymmetry of the nuclear medium.
\section{Charged pion ratio in hot dense asymmetric nuclear matter}
To see the above isospin-dependent pion in-medium effects on the
$\pi^-/\pi^+$ ratio in heavy ion collisions, we have used a thermal
model which assumes that pions are in thermal equilibrium with
nucleons and $\Delta$ resonances~\cite{bertsch}. In terms of the
spectral function $S_i(\omega,k)$, the density of a particle species
$i$ is then given by
\begin{eqnarray}\label{density}
\rho_i^{} \approx g_i^{} \int \frac{d^3{\bf k}}{(2\pi)^3}
d\omega^{n_i^{}} S_i(\omega,k)\frac{1}{z_i^{-1}
e^{(\omega - B_i\mu_B^{}-Q_i\mu_Q^{} )/T} \pm 1}.
\end{eqnarray}
In the above, $g_i^{}$, $B_i$, and $Q_i$ are the degeneracy, baryon
number, and charge of the particle. The fugacity parameter
$z_i^{}$ is introduced to take into account possible chemical non-equilibrium effect.
The exponent $n_i^{}$ is $2$ for pions and $1$ for nucleons
and $\Delta$ resonances. For the spectral functions of $\Delta$
resonances, we neglect their momentum dependence and replace the
integration over the energy $\omega$ by that over mass. The $\omega$
in the Fermi-Dirac distribution for $\Delta$ resonances is then
simply $\omega=M+k^2/2M+U_\Delta^{m_T}$. For nucleons, their
spectral functions are taken to be delta functions if we neglect the
imaginary part of their self-energies, i.e., $S_N^{m_\tau}(\omega,k)
= \delta ( \omega^{} - m_N^{} - k^2/2m_N^{} - U_N^{m_\tau} ).$
According to studies based on the transport
model~\cite{bali02,xiao09,xiong93}, the total number of pions and
$\Delta$ resonances in heavy ion collisions reaches a maximum value
when the colliding matter achieves the maximum density, and remains
essentially constant during the expansion of the matter. For Au+Au
collisions at the beam energy of $0.4~{\rm AGeV}$, for which the
$\pi^-/\pi^+$ ratio has been measured by the FOPI Collaboration at
GSI~\cite{FOPI}, the IBUU transport model gives a maximum density
that is about twice the normal nuclear matter density and is
insensitive to the stiffness of the nuclear symmetry energy, as it
is mainly determined by the isoscalar part of the nuclear equation
of state~\cite{xiao09}. This density is thus used in the thermal
model. The temperature in the thermal model is determined by fitting
the measured pion to nucleon ratio, which is about $0.014$ including
pions and nucleons from the decay of $\Delta$
resonances~\cite{FOPI}, without medium effects and with unity
fugacity parameters for all particles, and the value is $T \simeq
43.6$~MeV. The assumption that pions and $\Delta$ resonances are in
chemical equilibrium is consistent with the short chemical
equilibration times estimated from the pion and $\Delta$ resonance
production rates. The isospin asymmetry of the hadronic matter is
then taken to be $\delta_{\rm like} \simeq 0.080$, $0.106$, and
$0.143$, corresponding to net charge densities of $0.920\rho_0^{}$,
$0.894\rho_0^{}$ and $0.857\rho_0^{}$, for the three symmetry
energies given by $x=0$, $0.5$, and $1$, respectively, in order to
reproduce the $\pi^-/\pi^+$ ratios of $2.20$, $2.40$, and $2.60$
predicted by the IBUU transport model of Ref.~\cite{xiao09} using
corresponding symmetry energy parameters without pion in-medium
effects. Since the medium effects enhance the pion and $\Delta$
resonance densities, to maintain the same pion to nucleon ratio as
the measured one requires the fugacity parameters for pions and
$\Delta$ resonances to be less than one. Also, the pion in-medium
effects have been shown to affect only slightly the pion and the
$\Delta$ resonance abundance~\cite{xiong93}, indicating that both
pions and $\Delta$ resonances are out of chemical equilibrium with
nucleons when medium effects are included, as expected from the
estimated increasing pion and $\Delta$ resonance chemical
equilibration times as a result of the medium effects. Because of
the small number of pions (about 0.3\%) and $\Delta$ resonances
(about 1.1\%) in the matter, the density, temperature, and net
charge density of the hadronic matter are expected to remain
unchanged when the pion in-medium effects are introduced. They lead
to, however, a slight reduction of the isospin asymmetry to
$\delta_{\rm like} \simeq 0.073$, $0.097$, and $0.133$ for the three
symmetry energies, respectively. We note that with the fugacity of nucleons
kept at $z_N=1$, the fugacity parameters of about $z_\pi^{} = 0.061$ and
$z_\Delta^{}= 0.373$ are needed to maintain same total number of pions and $\Delta$ resonances
as in the case without pion in-medium effects, and that the required values
for the fugacity parameters increase only slightly for the other two
symmetry energies considered here.
\begin{figure}[h]
\centerline{\includegraphics[width=3.5in,height=3.5in,angle=0]{fig3.EPS}}
\caption{(Color online) The $\pi^-/\pi^+$ ratio in Au+Au collisions
at the beam energy of $0.4~{\rm AGeV}$ for different values of
nuclear symmetry energy ($x=0$, $0.5$, and $1$) and the Migdal
parameter $g^\prime=1/3$, $0.4$, $0.5$, and $0.6$ in the
$\Delta$-hole model for the pion $p$-wave interaction. Results for
$g^\prime=\infty$ correspond to the case without the pion in-medium
effects.} \label{ratio}
\end{figure}
Results on the $\pi^-/\pi^+$ ratio in Au+Au collisions at the beam
energy of $0.4~{\rm AGeV}$ are shown in Fig.~\ref{ratio}. With the
value $g^\prime = 1/3$ for the Migdal parameter, values for the
$\pi^-/\pi^+$ ratio are $2.32$, $2.60$, and $2.94$ for the
three symmetry energy parameters $x=0$, $0.5$, and $1$,
respectively, which are larger than corresponding values for the
case without including the pion in-medium effects as shown by those
for $g^\prime = \infty$ in Fig.~\ref{ratio}. These results indicate
that the isospin-dependent pion in-medium effects on the charged
pion ratio are comparable to those due to the uncertainties in the
theoretically predicted stiffness of the nuclear symmetry energy.
The measured $\pi^-/\pi^+$ ratio of about $3$ by the FOPI
Collaboration, shown in Fig.~\ref{ratio} by the dash-dotted line
together with the error bar, which without the pion in-medium
effects favors a nuclear symmetry energy softer than the one given
by $x=1$, is now best described by a less softer one.
Fig.~\ref{ratio} further shows the results obtained with larger
values of $g^\prime=0.4$, $0.5$ and $0.6$ for the Migdal parameter.
It is seen that the isospin-dependent pion in-medium effects are
reduced in these cases compared to the case of $g^\prime = 1/3$ as
the repulsive interaction between $\Delta$-hole states becomes
stronger, thus reducing the pion in-medium effects. With these
larger values of $g^\prime$, symmetry energies softer than that
given by $x=1$ are then needed to describe the measured
$\pi^-/\pi^+$ ratio.
\section{Pion $s$-wave interactions in nuclear medium}
The above study does not include the $s$-wave interactions of
pions with nucleons. Calculations based on the chiral perturbation
theory have shown that the pion $s$-wave interaction modifies the
mass of a pion in nuclear medium, and for asymmetric nuclear matter
this effect depends on the charge of the pion~\cite{Kaiser:2001bx}.
Up to the two-loop approximation in chiral perturbation
theory~\cite{Kaiser:2001bx}, the self energies of $\pi^-$, $\pi^+$,
and $\pi^0$ in asymmetric nuclear matter of proton density $\rho_p$
and neutron density $\rho_n$ are given, respectively, by
\begin{eqnarray}
\Pi^-(\rho_p,\rho_n)&=&\rho_n[T^-_{\pi N}-T^+_{\pi N}]-\rho_p[T^-_{\pi N}+T^+_{\pi N}]+\Pi^-_{\rm rel}(\rho_p,\rho_n)+\Pi^-_{\rm cor}(\rho_p,\rho_n)\notag\\
\Pi^+(\rho_p,\rho_n)&=&\Pi^-(\rho_n,\rho_p)\notag\\
\Pi^0(\rho_p,\rho_n)&=&-(\rho_p+\rho_n)T^+_{\pi N}+\Pi^0_{\rm cor}(\rho_p,\rho_n).
\end{eqnarray}
In the above, $T^\pm$ are the isospin-even and isospin-odd $\pi
N$-scattering $T$-matrices which have the empirical values $T^-_{\rm
\pi N}\approx 1.847~{\rm fm}$ and $T^+_{\rm \pi N}\approx
-0.045~{\rm fm}$ extracted from the energy shift and width of the 1s
level in pionic hydrogen atom. The term $\Pi^-_{\rm rel}$ is due to
the relativistic correction, whereas the terms $\Pi^-_{\rm cor}$ and
$\Pi^0_{\rm cor}$ are the contributions from the two-loop order in
chiral perturbation theory. Numerically, it was found in
Ref.~\cite{Kaiser:2001bx} that changes of pion masses in asymmetric
nuclear matter of density $\rho=0.165~{\rm fm}^{-3}$ and isospin
asymmetry $\delta=0.2$ are $\Delta m_{\pi^-}=13.8~{\rm MeV}$,
$\Delta m_{\pi^+}=-1.2~{\rm MeV}$, and $\Delta
m_{\pi^0}=6.1~{\rm MeV}$.
\begin{figure}[h]
\centerline{\includegraphics[width=3.5in,height=3.5in,angle=0]{fig4.EPS}}
\caption{(Color online) Similar to Fig.~\ref{ratio} with both pion
$s$-wave and $p$-wave interactions included.}\label{ratiosp}
\end{figure}
Taking into account the isospin-dependent pion self energies due to
pion $s$-wave interactions in asymmetric nuclear matter changes the
results shown in Figs.~\ref{pion} and \ref{delta}. For the pion
spectral function, the one for $\pi^+$ now has a larger strength at
low energies than that for $\pi^-$. Similarly, the strength near the
threshold of the $\Delta$ resonance mass distribution now increases
with increasing charge of the $\Delta$ resonance, although that around
the peak still decreases with increasing $\Delta$ resonance charge. As a result, the
$\pi^-/\pi^+$ ratio in Au+Au collisions at the beam energy of
$0.4~{\rm AGeV}$ is slightly reduced after the inclusion of both
pion $s$-wave and $p$-wave interactions in asymmetric nuclear matter
as shown in Fig.~\ref{ratiosp}.
\section{Summary}
The pion spectral function in asymmetric nuclear matter becomes
dependent on the charge of a pion. For the $p$-wave interaction of
the pion, modeled by its couplings to the $\Delta$-hole excitations
in nuclear medium, it leads to an increased strength of the $\pi^-$
spectral function at low energies relative to that of the $\pi^+$
spectral function in dense asymmetric nuclear matter. In a thermal
model, this isospin-dependent effect increases the $\pi^-/\pi^+$
ratio from heavy ion collisions, and the effect is comparable to
that due to the uncertainties in the theoretically predicted
stiffness of the nuclear symmetry energy at high densities. However,
including also the pion $s$-wave interaction based on results from
the chiral perturbation theory reverses the isospin-dependent pion
in-medium effects, leading instead to a slightly reduced
$\pi^-/\pi^+$ ratio in neutron-rich nuclear matter. Taking
into consideration of the isospin-dependent pion in-medium effects
in the transport model thus would have some influence on the extraction of the
nuclear symmetry energy from measured $\pi^-/\pi^+$ ratio.
\section*{Acknowledgments}
This talk was based on work supported in part by the US National Science
Foundation under Grant No. PHY-0758115 and the Welch Foundation
under Grant No. A-1358.
|
\section{$(n,m)$-pure exact sequences}
\label{S:pure}
By using a standard technique, (see for instance \cite[Chapter I, Section 8]{FuSa01}), we can prove the following theorem, and similar results hold if we replace $n$ or $m$ with $\aleph_0$.
\begin{theorem}
\label{T:pure} Assume that $R$ is an algebra over a commutative ring $S$ and $E$ is an injective $S$-cogenerator. Then, for each exact sequence $(\Sigma)$ of left $R$-modules $0\rightarrow A\rightarrow B\rightarrow C\rightarrow 0$, the following conditions are equivalent:
\begin{enumerate}
\item $(\Sigma)$ is $(n,m)$-pure;
\item for each $(m,n)$-presented left module $G$ the sequence $\mathrm{Hom}_R(G,(\Sigma))$ is exact;
\item every system of $n$ equations over $A$
\[\sum_{j=1}^{m}r_{i,j}x_j=a_i\in A\qquad (i=1,\dots,n)\]
with coefficients $r_{i,j}\in R$ and unknowns $x_1,\dots,x_m$ has a solution in $A$ whenever it is solvable in $B$;
\item the exact sequence of right $R$-modules $\mathrm{Hom}_S((\Sigma),E)$ is $(m,n)$-pure.
\end{enumerate}
\end{theorem}
Propositions~\ref{P:PureProj} and \ref{P:PureInj} can be deduced from \cite[Theorem 1]{Warf69}.
A left $R$-module $G$ is called $(n,m)$-{\it pure-projective} if for each $(n,m)$-pure exact sequence $0\rightarrow A\rightarrow B\rightarrow C\rightarrow 0$ the sequence
\[0\rightarrow \mathrm{Hom}_R(G,A)\rightarrow \mathrm{Hom}_R(G,B)\rightarrow \mathrm{Hom}_R(G,C)\rightarrow 0\]
is exact. Similar definitions can be given by replacing $n$ or $m$ by $\aleph_0$. From Theorem~\ref{T:pure} and by using standard technique (see for instance \cite[Chapter VI, Section 12]{FuSa01}) we get the following proposition in which $n$ or $m$ can be replaced by $\aleph_0$:
\begin{proposition}
\label{P:PureProj} Let $G$ be a left $R$-module. Then the following assertions hold:
\begin{enumerate}
\item there exists a $(n,m)$-pure exact sequence of left modules \[0\rightarrow K\rightarrow F\rightarrow G\rightarrow 0\] where $F$ is a direct sum of $(m,n)$-presented left modules;
\item $G$ is $(n,m)$-pure projective if and only if it is a summand of a direct sum of $(m,n)$-presented left modules.
\end{enumerate}
\end{proposition}
A left $R$-module $G$ is called $(n,m)$-{\it pure-injective} if for each $(n,m)$-pure exact sequence $0\rightarrow A\rightarrow B\rightarrow C\rightarrow 0$ the sequence
\[0\rightarrow \mathrm{Hom}_R(C,G)\rightarrow \mathrm{Hom}_R(B,G)\rightarrow \mathrm{Hom}_R(A,G)\rightarrow 0\]
is exact.
If $M$ is a left module we put $M^{\sharp}=\mathrm{Hom}_{\mathbb{Z}}(M,\mathbb{Q}/\mathbb{Z})$. Thus $M^{\sharp}$ is a right module. It is the {\it character module} of $M$.
If $A$ is a submodule of a left $R$-module $B$, we say that $B$ is a $(n,m)$-{\it pure essential extension} of $A$ if $A$ is a $(n,m)$-pure submodule of $B$ and for each nonzero submodule $K$ of $B$ such that $A\cap K=0$, $(A+K)/K$ is not a $(n,m)$-pure submodule of $B/K$. If, in addition, $B$ is $(n,m)$-pure injective, we say that $B$ is a $(n,m)$-{\it pure injective hull} of $A$. In these above definitions and in the following proposition $n$ or $m$ can be replaced by $\aleph_0$.
\begin{proposition}
\label{P:PureInj} The following assertions hold:
\begin{enumerate}
\item each left $R$-module is a $(n,m)$-pure submodule of a $(n,m)$-pure injective left module;
\item each left $R$-module has a $(n,m)$-pure injective hull which is unique up to an isomorphism.
\end{enumerate}
\end{proposition}
\begin{proof}
(1). Let $M$ be a left $R$-module. By Proposition~\ref{P:PureProj} there exists a $(m,n)$-pure exact sequence of right $R$-modules $0\rightarrow K\rightarrow F\rightarrow M^{\sharp}\rightarrow 0$ where $F$ is a direct sum of $(n,m)$-presented right modules. From Theorem~\ref{T:pure} it follows that $(M^{\sharp})^{\sharp}$ is a $(n,m)$-pure submodule of $F^{\sharp}$. By \cite[Corollary 1.30]{Fac98} $M$ is isomorphic to a pure submodule of $(M^{\sharp})^{\sharp}$. So, $M$ is isomorphic to a $(n,m)$-pure submodule of $F^{\sharp}$. By using the canonical isomorphism $(F\otimes_R-)^{\sharp}\cong\mathrm{Hom}_R(-,F^{\sharp})$ we get that $F^{\sharp}$ is $(n,m)$-pure injective since $F$ is a direct sum of $(n,m)$-presented modules.
(2). Since (1) holds and every direct limit of $(n,m)$-pure exact sequences is $(n,m)$-pure exact too, we can adapt the method of Warfield's proof of existence of pure-injective hull to show (2)(see \cite[Proposition 6]{War69}). We can also use \cite[Proposition 4.5]{Ste67}.
\end{proof}
\begin{proposition}
\label{P:PureLocal} Let $R$ be a commutative ring and let $(\Sigma)$ be a short exact sequence of $R$-modules. Then $(\Sigma)$ is $(n,m)$-pure if and only if, for each maximal ideal $P$ $(\Sigma)_P$ is $(n,m)$-pure.
\end{proposition}
\begin{proof}
Assume that $(\Sigma)$ is $(n,m)$-pure and let $M$ be a $(n,m)$-presented $R_P$-module where $P$ is a maximal ideal. There exists a $(n,m)$-presented $R$-module $M'$ such that $M\cong M'_P$ and $M\otimes_{R_P}(\Sigma)_P\cong (M'\otimes_R(\Sigma))_P$. We deduce that $(\Sigma)_P$ is $(n,m)$-pure.
Conversely, suppose that $(\Sigma)$ is the sequence $0\rightarrow A\rightarrow B\rightarrow C\rightarrow 0.$ Let $M$ be a $(n,m)$-presented $R$-module. Then, for each maximal ideal $P$, $(\Sigma)_P$ is $(n,m)$-pure over $R$ since $M\otimes_R(\Sigma)_P\cong M_P\otimes_{R_P}(\Sigma)_P$. On the other hand, since $M\otimes_R(\prod_{P\in\mathrm{Max}\ R}(\Sigma)_P)\cong (\prod_{P\in\mathrm{Max}\ R} M\otimes_R(\Sigma)_P)$, $\prod_{P\in\mathrm{Max}\ R}A_P$ is a $(n,m)$-pure submodule of $\prod_{P\in\mathrm{Max}\ R}B_P$. By \cite[Lemme 1.3]{Cou82} $A$ is isomorphic to a pure submodule of $\prod_{P\in\mathrm{Max}\ R}A_P$. We successively deduce that $A$ is a $(n,m)$-pure submodule of $\prod_{P\in\mathrm{Max}\ R}B_P$ and $B$.
\end{proof}
\section{Comparison of purities over a semiperfect ring}
\label{S:semiperfect}
In this section we shall compare $(n,m)$-purities for different pairs of integers $(n,m)$. In \cite{PPR99} some various purities are also compared. In particular some necessary and sufficient conditions on a ring $R$ are given for the $(1,1)$-purity to be equivalent to the $(\aleph_0,\aleph_0)$-purity.
The following lemma is due to Lawrence Levy, see \cite[Lemma 1.3]{WiWi75}. If $M$ be a finitely generated left (or right) $R$-module, we denote by $\mathrm{gen}\ M$ its minimal number of generators.
\begin{lemma}
\label{L:Levy} Let $R$ be a ring. Assume there exists a positive integer $p$ such that $\mathrm{gen}\ A\leq p$ for each finitely generated left ideal $A$ of $R$. Then $\mathrm{gen}\ N\leq p\times\mathrm{gen}\ M$, if $N$ is a finitely generated submodule of a finitely generated left $R$-module $M$.
\end{lemma}
From this lemma and Theorem~\ref{T:pure} we deduce the following:
\begin{proposition}
\label{P:genIdeal} Let $R$ be a ring. Assume there exists a positive integer $p$ such that $\mathrm{gen}\ A\leq p$ for each finitely generated left ideal $A$ of $R$. Then, for each positive integer $n$:
\begin{enumerate}
\item each $(n,np)$-pure exact sequence of right modules is $(n,\aleph_0)$-pure exact;
\item each $(np,n)$-pure exact sequence of left modules is $(\aleph_0,n)$-pure exact.
\end{enumerate}
\end{proposition}
\begin{corollary}
\label{C:Artinian} Let $R$ be a left Artinian ring. Then there exists a positive integer $p$ such that, for each positive integer $n$:
\begin{enumerate}
\item each $(n,np)$-pure exact sequence of right modules is $(n,\aleph_0)$-pure exact;
\item each $(np,n)$-pure exact sequence of left modules is $(\aleph_0,n)$-pure exact.
\end{enumerate}
\end{corollary}
\begin{proof}
Each finitely generated left $R$-module $M$ has a finite length denoted by $\mathrm{length}\ M$, and $\mathrm{gen}\ M\leq\mathrm{length}\ M$. So, for each left ideal $A$ we have $\mathrm{gen}\ A\leq\mathrm{length}\ R$.
We choose $p=\sup\{\mathrm{gen}\ A\mid A\ \mathrm{left\ ideal\ of}\ R\}$ and we apply the previous proposition.
\end{proof}
Let $R$ be a ring and $J$ its Jacobson radical. Recall that $R$ is {\it semiperfect} if $R/J$ is semisimple and idempotents lift modulo $J$.
\begin{theorem}
\label{T:semiperfect} Let $R$ be semiperfect ring. Assume that each indecomposable finitely presented cyclic left $R$-module has a local endomorphism ring. The following conditions are equivalent:
\begin{enumerate}
\item there exists an integer $p>0$ such that, for each integer $n>0$, each $(n,np)$-pure exact sequence of right modules is $(n,\aleph_0)$-pure exact;
\item there exists an integer $p>0$ such that, for each integer $n>0$, each $(np,n)$-pure exact sequence of left modules is $(\aleph_0,n)$-pure exact;
\item there exists an integer $q>0$ such that each $(1,q)$-pure exact sequence of right modules is $(1,\aleph_0)$-pure exact;
\item there exists an integer $q>0$ such that each $(q,1)$-pure exact sequence of left modules is $(\aleph_0,1)$-pure exact;
\item there exists an integer $q>0$ such that each indecomposable finitely presented cyclic left module is $q$-related;
\item there exists an integer $p>0$ such that $\mathrm{gen}\ A\leq p$ for each finitely generated left ideal $A$ of $R$.
\end{enumerate}
Moreover, if each indecomposable finitely presented left $R$-module has a local endomorphism ring, these conditions are equivalent to the following:
\begin{itemize}
\item[(7)] there exist two positive integers $n,\ m$ such that each $(n,m)$-pure exact sequence of right modules is $(n,\aleph_0)$-pure exact;
\item [(8)] there exist two positive integers $n,\ m$ such that each $(m,n)$-pure exact sequence of left modules is $(\aleph_0,n)$-pure exact;
\end{itemize}
\end{theorem}
\begin{proof}
By Proposition~\ref{P:genIdeal} $(6)\Rightarrow (1)$. By Theorem~\ref{T:pure} $(1)\Leftrightarrow (2)$, $(3)\Leftrightarrow (4)$ and $(7)\Leftrightarrow (8)$. It is obvious that $(2)\Rightarrow (4)$ and $(2)\Rightarrow (7)$.
$(4)\Rightarrow (5)$. Let $C$ be an indecomposable finitely presented cyclic left module. Then $C$ is $(q,1)$-pure-projective. So, $C$ is a direct summand of a finite direct sum of $(1,q)$-presented left modules. Since $R$ is semiperfect, we may assume that these $(1,q)$-presented left modules are indecomposable. So, by Krull-Schmidt theorem $C$ is $(1,q)$-presented.
$(5)\Rightarrow (6)$. Let $A$ be a finitely generated left ideal. Then $R/A=\oplus_{i=1}^tR/A_i$ where, for each $i=1,\dots,t$, $A_i$ is a left ideal and $R/A_i$ is indecomposable. We have the following commutative diagram with exact horizontal sequences:
\[\begin{matrix}
0 & \rightarrow & \oplus_{i=1}^tA_i & \rightarrow & R^t & \rightarrow & \oplus_{i=1}^tR/A_i & \rightarrow & 0 \\
{} & {} & \downarrow & {} & \downarrow & {} & \downarrow & {} & {} \\
0 & \rightarrow & A & \rightarrow & R & \rightarrow & R/A & \rightarrow & 0 \\
\end{matrix}\]
Since the right vertical map is an isomorphism, we deduce from the snake lemma that the other two vertical homomorphisms have isomorphic cokernels. It follows that $\mathrm{gen}\ A\leq tq+1$ because $\mathrm{gen}\ A_i\leq q$ by $(5)$. On the other hand, let $P$ be a projective cover of $R/A$. Then $P$ is isomorphic to a direct summand of $R$. We know that the left module $R$ is a finite direct sum of indecomposable projective modules. Let $s$ the number of these indecomposable summands. It is easy to show that $t\leq s$. So, if $p=sq+1$, then $\mathrm{gen}\ A\leq p$.
$(8)\Rightarrow (5)$. Let $C$ be an indecomposable finitely presented cyclic left module. Then $C$ is $(m,n)$-pure-projective. So, $C$ is a direct summand of a finite direct sum of $(n,m)$-presented left modules. Since $R$ is semiperfect, we may assume that these $(n,m)$-presented left modules are indecomposable. So, by the Krull-Schmidt theorem $C$ is $(1,m)$-presented.
\end{proof}
A ring $R$ is said to be \textit{strongly $\pi$-regular} if, for each $r\in R$, there exist $s\in R$ and an integer $q\geq 1$ such that $r^q=r^{q+1}s$. By \cite[Theorem 3.16]{Fac98} each strongly $\pi$-regular $R$ satisfies the following condition: for each $r\in R$, there exist $s\in R$ and an integer $q\geq 1$ such that $r^q=sr^{q+1}$. Recall that a left $R$-module $M$ is said to be {\it Fitting} if for each endomorphism $f$ of $M$ there exists a positive integer $t$ such that $M=\ker\ f^t\oplus f^t(M)$.
\begin{lemma}
\label{L:Fitting} Let $R$ be a strongly $\pi$-regular semiperfect ring. Then:
\begin{enumerate}
\item each finitely presented cyclic left (or right) $R$-module is Fitting;
\item each indecomposable finitely presented cyclic left (or right) $R$-module has a local endomorphism ring.
\end{enumerate}
\end{lemma}
\begin{proof} In \cite[Lemma 3.21]{Fac98} it is proven that every finitely presented $R$-module is a Fitting module if $R$ is a semiperfect ring with $\mathrm{M}_n(R)$ strongly $\pi$-regular for all $n$. We do a similar proof to show $(1)$.
$(2)$. By \cite[Lemma 2.21]{Fac98} each indecomposable Fitting module has a local endomorphism ring.
\end{proof}
\begin{corollary}
\label{C:semiperfect} Let $R$ be a strongly $\pi$-regular semiperfect ring. The following conditions are equivalent:
\begin{enumerate}
\item there exists an integer $p>0$ such that, for each integer $n>0$, each $(n,np)$-pure exact sequence of right modules is $(n,\aleph_0)$-pure exact;
\item there exists an integer $p>0$ such that, for each integer $n>0$, each $(np,n)$-pure exact sequence of left modules is $(\aleph_0,n)$-pure exact;
\item there exists an integer $q>0$ such that each $(1,q)$-pure exact sequence of right modules is $(1,\aleph_0)$-pure exact;
\item there exists an integer $q>0$ such that each $(q,1)$-pure exact sequence of left modules is $(\aleph_0,1)$-pure exact;
\item there exists an integer $q>0$ such that each indecomposable finitely presented cyclic left module is $q$-related;
\item there exists an integer $p>0$ such that $\mathrm{gen}\ A\leq p$ for each finitely generated left ideal $A$ of $R$.
\end{enumerate}
Moreover, if $\mathrm{M}_n(R)$ is strongly $\pi$-regular for all $n>0$, these conditions are equivalent to the following:
\begin{itemize}
\item[(7)] there exist two positive integers $n,\ m$ such that each $(n,m)$-pure exact sequence of right modules is $(n,\aleph_0)$-pure exact;
\item [(8)] there exist two positive integers $n,\ m$ such that each $(m,n)$-pure exact sequence of left modules is $(\aleph_0,n)$-pure exact;
\end{itemize}
\end{corollary}
\begin{proof}
By Lemma~\ref{L:Fitting} each indecomposable finitely presented cyclic left $R$-module has a local endomorphism ring.
If $\mathrm{M}_n(R)$ is strongly $\pi$-regular for all $n$, then by \cite[Lemmas 3.21 and 2.21]{Fac98} each indecomposable finitely presented left $R$-module has a local endomorphism ring. So, we apply Theorem~\ref{T:semiperfect}.
\end{proof}
Recall that a ring $R$ is {\it right perfect} if each flat right $R$- module is projective.
\begin{corollary}
\label{C:perfect} Let $R$ be a right perfect ring. The following conditions are equivalent:
\begin{enumerate}
\item there exists an integer $p>0$ such that, for each integer $n>0$, each $(n,np)$-pure exact sequence of right modules is $(n,\aleph_0)$-pure exact;
\item there exists an integer $p>0$ such that, for each integer $n>0$, each $(np,n)$-pure exact sequence of left modules is $(\aleph_0,n)$-pure exact;
\item there exists an integer $q>0$ such that each $(1,q)$-pure exact sequence of right modules is $(1,\aleph_0)$-pure exact;
\item there exist two positive integers $n,\ m$ such that each $(n,m)$-pure exact sequence of right modules is $(n,\aleph_0)$-pure exact;
\item there exist two positive integers $n,\ m$ such that each $(m,n)$-pure exact sequence of left modules is $(\aleph_0,n)$-pure exact;
\item there exists an integer $p>0$ such that $\mathrm{gen}\ A\leq p$ for each finitely generated left ideal $A$ of $R$.
\end{enumerate}
\end{corollary}
\begin{proof}
For all $n>0$, $\mathrm{M}_n(R)$ is right perfect. Since each right perfect ring satisfies the descending chain condition on finitely generated left ideals, then $\mathrm{M}_n(R)$ is strongly $\pi$-regular for all $n>0$. We apply Corollary~\ref{C:semiperfect}.
\end{proof}
\section{Comparison of purities over a commutative ring}
\label{S:commutative}
In the sequel of this section $R$ is a commutative local ring, except in Theorem~\ref{T:ComPur}. We denote respectively by $P$ and $k$ its maximal ideal and its residue field
Let $M$ be a finitely presented $R$-module. Recall that $\mathrm{gen}\ M=\dim_k\ M/PM$. Let $F_0$ be a free $R$-module whose rank is $\mathrm{gen}\ M$ and let $\phi:F_0\rightarrow M$ be an epimorphism. Then $\ker\ \phi\subseteq PF_0$. We put $\mathrm{rel}\ M=\mathrm{gen}\ \ker\ \phi$. Let $F_1$ be a free $R$-module whose rank is $\mathrm{rel}\ M$ and let $f:F_1\rightarrow F_0$ be a homomorphism such that $\mathrm{im}\ f=\ker\ \phi$. Then $\ker\ f\subseteq PF_1$. For any $R$-module $N$, we put $N^*=\mathrm{Hom}_R(N,R)$. Let $f^*:F_0^*\rightarrow F_1^*$ be the homomorphism deduced from $f$. We set $\mathrm{D}(M)=\mathrm{coker}\ f^*$ the Auslander and Bridger's dual of $M$. The following proposition is the version in commutative case of \cite[Theorem 2.4]{War75}:
\begin{proposition}
\label{P:AusBrid} Assume that $M$ has no projective summand. Then:
\begin{enumerate}
\item $\ker\ f^*\subseteq PF_0^*$ and $\mathrm{im}\ f^*\subseteq PF_1^*$;
\item $M\cong \mathrm{D}(\mathrm{D}(M))$ and $\mathrm{D}(M)$ has no projective summand;
\item $\mathrm{gen}\ \mathrm{D}(M)=\mathrm{rel}\ M$ and $\mathrm{rel}\ \mathrm{D}(M)=\mathrm{gen}\ M$;
\item if $M=M_1\oplus M_2$ then\\ $\mathrm{gen}\ M=\mathrm{gen}\ M_1 + \mathrm{gen}\ M_2$ and $\mathrm{rel}\ M=\mathrm{rel}\ M_1 + \mathrm{rel}\ M_2$.
\item $\mathrm{End}_R(\mathrm{D}(M))$ is local if and only if so is $\mathrm{End}_R(M)$.
\end{enumerate}
\end{proposition}
\begin{lemma}
\label{L:iso} Let $M$ be a finitely generated $R$-module, $s$ an endomorphism of $M$ and $\bar{s}$ the endomorphism of $M/PM$ induced by $s$. Then $s$ is an isomorphism if and only if so is $\bar{s}$.
\end{lemma}
\begin{proof}
If $s$ is an isomorphism it is obvious that so is $\bar{s}$. Conversely, $\mathrm{coker}\ s=0$ by Nakayama lemma. So, $s$ is surjective. By using a Vasconcelos's result (see \cite[Theorem V.2.3]{FuSa01}) $s$ is bijective.
\end{proof}
\begin{proposition}
\label{P:locEnd} Assume that there exists an ideal $A$ with $\mathrm{gen}\ A=p+1$ where $p$ is a positive integer. Then, for each positive integers $n$ and $m$ with $(n-1)p+1\leq m\leq np+1$, there exists a finitely presented $R$-module $W_{p,n,m}$ whose endomorphism ring is local and such that $\mathrm{gen}\ W_{p,n,m}=n$ and $\mathrm{rel}\ W_{p,n,m}=m$.
\end{proposition}
\begin{proof}
Suppose that $A$ is generated by $a_1,\dots,a_p,a_{p+1}$. Let $F$ be a free module of rank $n$ with basis $e_1,\dots,e_n$ and let $K$ be the submodule of $F$ generated by $x_1,\dots,x_m$ where these elements are defined in the following way: if $j= pq+r$ where $1\leq r\leq p$, $x_j=a_re_{q+1}$ if $r\ne 1$ or $q=0$, and $x_j=a_{p+1}e_q+a_1e_{q+1}$ else; when $m=pn+1$, $x_m=a_{p+1}e_n$. We put $W_{p,n,m}=F/K$.
We can say that $W_{p,n,m}$ is named by the following $n\times m$ matrix, where $r=m-p(n-1)$:
\bigskip
\(\left( \begin{matrix}
a_1 & .. & a_p & a_{p+1} & 0 & \hdotsfor[0.7]{4}\\
0 & .. & 0 & a_1 & a_2 & \dots & a_{p+1} & 0 & .. \\
\vdots & & & & & & & & \ddots\\
0 & \hdotsfor[0.7]{7} & \\
0 & \hdotsfor[0.7]{8}
\end{matrix}
\begin{matrix}
\hdotsfor[0.7]{7} & 0 \\
\hdotsfor[0.7]{7}& 0 \\
& & & & & & & \vdots \\
0 & a_1 & .. & a_p & a_{p+1} & 0 & .. & 0 \\
\hdotsfor[0.7]{3} & 0 & a_1 & a_2 & .. & a_r
\end{matrix}\right)
\)
\bigskip
Since $K\subseteq PF$, $\mathrm{gen}\ W_{p,n,m}=n$. Now we consider the following relation: $\sum_{j=0}^{m}c_jx_j=0$. From the definition of the $x_j$ we get the following equality:
\[\sum_{q=0}^{n-2}\left( \sum_{i=1}^{p+1}c_{pq+i}a_i\right) e_{q+1}+\left( \sum_{i=1}^{r}c_{p(n-1)+i}a_i\right) e_n=0.\]
Since $\{e_1,\dots,e_n\}$ is a basis and $\mathrm{gen}\ A=p+1$ we deduce that $c_j\in P,\ \forall j,\ 1\leq j\leq m$. So, $\mathrm{rel}\ W_{p,n,m}=m$.
Let $s\in \mathrm{End}_R(W_{p,n,m})$. Then $s$ is
induced by an endomorphism $\tilde{s}$ of $F$ which satisfies
$\tilde{s}(K)\subseteq K$. For each $j,\ 1\leq j\leq n,$ there exists a family $(\alpha_{i,j})$ of elements of $R$ such that:
\begin{equation} \label{eq:s1}
\tilde{s}(e_j)=\sum_{i=1}^{n}\alpha_{i,j}e_i
\end{equation}
Since $\tilde{s}(K)\subseteq K,\ \forall j, 1\leq j\leq m,\ \exists$ a
family $(\beta_{i,j})$ of elements of $R$ such that:
\begin{equation} \label{eq:s2}
\tilde{s}(x_j)=\sum_{i=1}^{m}\beta_{i,j}x_i
\end{equation}
From (\ref{eq:s1}), (\ref{eq:s2}) and the equality $x_1=a_1e_1$
if follows that:
\[\sum_{q=1}^{n}\alpha_{q,1}a_1e_q=\sum_{q=0}^{n-2}\left( \sum_{i=1}^{p+1}\beta_{pq+i,1}a_i\right) e_{q+1}+\left( \sum_{i=1}^{r}\beta_{p(n-1)+i,1}a_{i}\right) e_n.\]
Then, we get:
\[\forall q,\ 1\leq q\leq n-1,\qquad \alpha_{q,1}a_1=\sum_{i=1}^{p+1}\beta_{p(q-1)+i,1}a_i\]
\[\mathrm{and}\qquad\alpha_{n,1}a_1=\sum_{i=1}^{r}\beta_{p(n-1)+i,1}a_{i}.\]
We deduce that: $\forall q,\ 2\leq q\leq n,\ \beta_{p(q-2)+p+1,1}\in P$ and $\beta_{p(q-1)+1,1}\equiv \alpha_{q,1}\ [P]$. So,
\begin{equation}\label{eq:s3}
\forall q,\ 2\leq q\leq n,\ \alpha_{q,1}\in P.
\end{equation}
Now, let $j=p\ell+1$ where $1\leq \ell\leq (n-1)$. In this case, $x_j=a_{p+1}e_{\ell}+a_1e_{\ell+1}$. From (\ref{eq:s1}) and (\ref{eq:s2}) it follows that:
\[\sum_{q=1}^{n}(\alpha_{q,\ell}a_{p+1}+\alpha_{q,\ell+1}a_1)e_q=\sum_{q=0}^{n-2}\left( \sum_{i=1}^{p+1}\beta_{pq+i,j}a_i\right) e_{q+1}+\left( \sum_{i=1}^{r}\beta_{p(n-1)+i,j}a_{i}\right) e_n.\]
Then, we get:
\[\forall q,\ 1\leq q\leq n-1,\qquad \alpha_{q,\ell}a_{p+1}+\alpha_{q,\ell+1}a_1=\sum_{i=1}^{p+1}\beta_{p(q-1)+i,j}a_i\]
\[\mathrm{and}\qquad\alpha_{n,\ell}a_{p+1}+\alpha_{n,\ell+1}a_1=\sum_{i=1}^{r}\beta_{p(n-1)+i,j}a_{i}.\]
We deduce that \[\forall q,\ell,\ 1\leq q,\ell\leq (n-1),\ \alpha_{q,\ell}\equiv \beta_{p(q-1)+p+1,j}\ [P]\ \mathrm{and}\ \alpha_{q+1,\ell+1}\equiv \beta_{pq+1,j}\ [P],\] whence $\alpha_{q,\ell}\equiv\alpha_{q+1,\ell+1}\ [P]$. Consequently, $\forall q,\ 1\leq q\leq n,\ \alpha_{q,q}\equiv \alpha_{1,1}\ [P]$ and $\forall t,\ 1\leq t\leq (n-1),\ \forall q,\ 1\leq q\leq (n-t),\ \alpha_{q+t,q}\equiv \alpha_{1+t,1}\equiv 0\ [P]$ by (\ref{eq:s3}). Let $\bar{s}$ be the endomorphism of $W_{p,n,m}/PW_{p,n,m}$ induced by $s$. If $\alpha_{1,1}$ is a unit then $\bar{s}$ is an isomorphism, else $\overline{\mathbf{1}_{W_{p,n,m}}-s}$ is an isomorphism. By Lemma~\ref{L:iso} we conclude that either $s$ or $(\mathbf{1}_{W_{p,n,m}}-s)$ is an isomorphism. Hence, $\mathrm{End}_R(W_{p,n,m})$ is local. \end{proof}
\begin{remark}
Observe that $\mathrm{D}(W_{1,n-1,n})$ is isomorphic to the indecomposable module built in the proof of \cite[Theorem 2]{War70}.
\end{remark}
\begin{theorem}
\label{T:ComPur} Let $R$ be a commutative ring. The following assertions hold:
\begin{enumerate}
\item Assume that, for any integer $p>0$, there exists a maximal ideal $P$ and a finitely generated ideal $A$ of $R_P$ such that $\mathrm{gen}_{R_P}\ A\geq p+1$. Then, if $(n,m)$ and $(r,s)$ are two different pairs of integers, the $(n,m)$-purity and the $(r,s)$-purity are not equivalent.
\item Assume that, there exists an integer $p>0$ such that, for each maximal ideal $P$, for any finitely generated ideal $A$ of $R_P$, $\mathrm{gen}_{R_P}\ A\leq p$. Then:
\begin{enumerate}
\item for each integer $n>0$ the $(\aleph_0,n)$-purity (respectively $(n,\aleph_0)$-purity) is equivalent to the $(np,n)$-purity (respectively $(n,np)$-purity);
\item if $p>1$, then, for each integer $n>0$, for each integer $m,\ 1\leq m\leq n(p-1)$, the $(n,m)$-purity (respectively $(m,n)$-purity) is not equivalent to the $(n,m+1)$-purity (respectively $(m+1,n)$-purity).
\end{enumerate}
\end{enumerate}
\end{theorem}
\begin{proof} By Proposition~\ref{P:PureLocal} we may assume that $R$ is local with maximal $P$. By Theorem~\ref{T:pure} the $(n,m)$-purity and the $(r,s)$-purity are equivalent if and only if so are the $(m,n)$-purity and the $(s,r)$-purity.
(1). Suppose that $r>n$ and let $t=\min(m,s)$. Let $q$ be the greatest divisor of $(r-1)$ which is $\leq t$ and $p=(r-1)/q$. Let $A$ be a finitely generated ideal such that $\mathrm{gen}\ A>p$. By way of contradiction, suppose that $W_{p,q,r}$ is $(n,m)$-pure-projective. By Proposition~\ref{P:PureProj} $W_{p,q,r}$ is a summand of $\oplus_{i\in I}F_i$ where $I$ is a finite set and $\forall i\in I,\ F_i$ is a $(m,n)$-presented $R$-module. Since its endomorphism ring is local, $W_{p,q,r}$ is an exchange module (see \cite[Theorem 2.8]{Fac98}). So, we have $W_{p,q,r}\oplus(\oplus_{i\in I}G_i)\cong(\oplus_{i\in I}H_i)\oplus(\oplus_{i\in I}G_i)$ where $\forall i\in I$, $G_i$ and $H_i$ are submodules of $F_i$ and $F_i=G_i\oplus H_i$. Let $G=\oplus_{i\in I}G_i$. Then $G$ is finitely generated. By \cite[Proposition V.7.1]{FuSa01} $\mathrm{End}_R(G)$ is semilocal. By using Evans's theorem (\cite[Corollary 4.6]{Fac98}) we deduce that $W_{p,q,r}\cong(\oplus_{i\in I}H_i)$. Since $W_{p,q,r}$ is indecomposable, we get that it is $(m,n)$-presented. This contradicts that $\mathrm{rel}\ W_{p,q,r}=r>n$.
(2)(a) is an immediate consequence of Proposition~\ref{P:genIdeal}.
(2)(b). There exist two integers $q,t$ such that $m+1=(q-1)(p-1)+t$ with $n\geq q\geq 1$ and $1\leq t\leq p$. As in (1) we prove that $W_{p-1,q,m+1}$ is not $(m,n)$-pure-projective.
\end{proof}
\begin{remark}
In the previous theorem, when there exists an integer $p>1$ such that, for any finitely generated ideal $A$ $\mathrm{gen}\ A\leq p$, we don't know if the $(n,m)$-purity and the $(n,m+1)$-purity are equivalent when $n(p-1)+1\leq m\leq np-1$. If $R$ is a local ring with maximal $P$ with residue field $k$ such that $P^2=0$ and $\dim_kP=p$ it is easy to show that each finitely presented $R$-module $F$ with $\mathrm{gen}\ F=n$ and $\mathrm{rel}\ F=np$ is semisimple. So, the $(np,n)$-purity is equivalent to the $(np-1,n)$-purity.
\end{remark}
\section{$(n,m)$-flat modules and $(n,m)$-injective modules}
\label{S:flat}
Let $M$ be a right $R$-module. We say that $M$ is $(n,m)${\it -flat} if for any $m$-generated submodule $K$ of a $n$-generated free left $R$-module $F$, the natural map: $M\otimes_RK\rightarrow M\otimes_RF$ is a monomorphism. We say that $M$ is $(\aleph_0,m)${\it -flat} (respectively $(n,\aleph_0)${\it -flat}) if $M$ is $(n,m)$-flat for each integer $n>0$ (respectively $m>0$). We say that $M$ is $(n,m)${\it -injective} if for any $m$-generated submodule $K$ of a $n$-generated free right $R$-module $F$, the natural map: $\mathrm{Hom}_R(F,M)\rightarrow \mathrm{Hom}_R(K,M)$ is an epimorphism. We say that $M$ is $(\aleph_0,m)${\it -injective} (respectively $(n,\aleph_0)${\it -injective}) if $M$ is $(n,m)$-injective for each integer $n>0$ (respectively $m>0$). A ring $R$ is called {\it left self $(n,m)$-injective} if $R$ is $(n,m)$-injective as left $R$-module.
If $R$ is a commutative domain, then an $R$-module is $(1,1)$-flat (respectively $(1,1)$-injective) if and only if it is torsion-free (respectively divisible).
The following propositions can be proved with standard technique: see \cite[Theorem 4.3 and Proposition 2.3]{ZhChZh05}. In these propositions the integers $n$ or $m$ can be replaced with $\aleph_0$.
\begin{proposition}
\label{P:flat} Assume that $R$ is an algebra over a commutative ring $S$ and let $E$ be an injective $S$-cogenerator. Let $M$ be a right $R$-module. The following conditions are equivalent:
\begin{enumerate}
\item $M$ is $(n,m)$-flat;
\item each exact sequence $0\rightarrow L\rightarrow N \rightarrow M\rightarrow 0$ is $(n,m)$-pure, where $L$ and $N$ are right $R$-modules;
\item for each $(m,n)$-presented right module $F$, every homomorphism $f:F\rightarrow M$ factors through a free right $R$-module;
\item $\mathrm{Hom}_S(M,E)$ is a $(n,m)$-injective left $R$-module.
\end{enumerate}
\end{proposition}
\begin{proposition}
\label{P:inj} Let $M$ be a right module. The following conditions are equivalent:
\begin{enumerate}
\item $M$ is $(n,m)$-injective;
\item each exact sequence $0\rightarrow M\rightarrow L\rightarrow N\rightarrow 0$ is $(m,n)$-pure, where $L$ and $N$ are right $R$-modules;
\item $M$ is a $(m,n)$-pure submodule of its injective hull.
\end{enumerate}
\end{proposition}
\begin{proposition}
\label{P:localflat} Let $R$ be a commutative ring. Then an $R$-module $M$ is $(n,m)$-flat if and only if, for each maximal ideal $P$, $M_P$ is $(n,m)$-flat over $R_P$.
\end{proposition}
\begin{lemma}
\label{L:p-gen} Let $M$ be $p$-generated right $R$-module where $p$ is a positive integer. Then $M$ is flat if and only if it is $(1,p)$-flat.
\end{lemma}
\begin{proof}
Only ``if'' requires a proof. Let $A$ be a left ideal. Assume that $M$ is generated by $x_1,\dots,x_p$. So, if $z\in M\otimes_RA$, $z=\sum_{i=1}^px_i\otimes a_i$ where $a_1,\dots,a_p\in A$. Suppose that the image of $z$ in $M\otimes_RR$ is $0$. If $A'$ is the left ideal generated by $a_1,\dots,a_p$, if $z'$ is the element of $M\otimes_RA'$ defined by $z'=\sum_{i=1}^px_i\otimes a_i$, then $z$ (respectively $0$) is the image of $z'$ in $M\otimes_RA$ (respectively $M\otimes_RR$). Since $M$ is $(1,p)$-flat we successively deduce that $z'=0$ and $z=0$.
\end{proof}
\bigskip
It is well known that each $(1,\aleph_0)$-flat right module is $(\aleph_0,\aleph_0)$-flat. {\bf For each positive integer $p$, is each $(1,p)$-flat right module $(\aleph_0,p)$-flat?}
The following theorem and Theorem~\ref{T:parfait} give a partial answer to this question.
\begin{theorem}
\label{T:p-flatideal} Let $p$ be a positive integer and let $R$ be a ring. For each positive integer $n$, assume that, for each $p$-generated submodule $G$ of the left $R$-module $R^n\oplus R$, $(G\cap R^n)$ is the direct limit of its $p$-generated submodules.Then a right $R$-module $M$ is $(1,p)$-flat if and only if it is $(\aleph_0,p)$-flat.
\end{theorem}
\begin{proof}
We shall prove that $M$ is $(n,p)$-flat by induction on $n$. Let $G$ be a $p$-generated submodule of the left $R$-module $R^{n+1}=R^n\oplus R$. Let $\pi$ be the projection of $R^{n+1}$ onto $R$ and $G'=\pi(G)$. Then $G'$ is a $p$-generated left module. We put $H=G\cap R^n$.
We have the following commutative diagram with exact horizontal sequences:
\[\begin{matrix}
{} & {} & M\otimes_RH &\rightarrow & M\otimes_RG &\xrightarrow{1_M\otimes\pi} & M\otimes_RG'& \rightarrow & 0\\
{} & {} & \downarrow & {} & \downarrow & {} & \downarrow & {} & {} \\
0 & \rightarrow & M\otimes_R R^n &\rightarrow & M\otimes_RR^{n+1} &\xrightarrow{1_M\otimes\pi} & M\otimes_RR& \rightarrow & 0
\end{matrix}\]
Let $u:G\rightarrow R^{n+1},\ u':G'\rightarrow R,\ w:R^n\rightarrow R^{n+1}$ be the inclusion maps and let $v=u\vert_{H}$. Then $(1_M\otimes u')$ is injective. Let $H'$ be a $p$-generated submodule of $H$. By the induction hypothesis $M$ is $(n,p)$-flat. So, $(1_M\otimes(v\vert_{H'}))$ is injective. It follows that $(1_M\otimes v)$ is injective too. We conclude that $(1_M\otimes u)$ is injective and $M$ is $(\aleph_0,p)$-flat.
\end{proof}
\begin{corollary}
\label{C:p-flatideal} Let $p$ be a positive integer and let $R$ be a ring such that each left ideal is $(1,p)$-flat. Then, for each positive integer $q\leq p$, a right $R$-module $M$ is $(1,q)$-flat if and only if it is $(\aleph_0,q)$-flat.
\end{corollary}
\begin{proof}
Let the notations be as in the previous theorem. Since $G'$ is a flat left $R$-module by Lemma~\ref{L:p-gen}, $H$ is a pure submodule of $G$. Let $\{g_1,\dots,g_q\}$ be a spanning set of $G$ and let $h_1,\dots,h_t\in H$. For each $k$, $1\leq k\leq t$ there exist $a_{k,1},\dots,a_{k,q}\in R$ such that $h_k=\sum_{i=1}^qa_{k,i}g_i$. It follows that there exist $g'_1,\dots,g'_q\in H$ such that $\forall k,\ 1\leq k\leq t,\ h_k=\sum_{i=1}^qa_{k,i}g'_i$. So, each finitely generated submodule of $H$ is contained in a $q$-generated submodule. We conclude by applying Theorem~\ref{T:p-flatideal}.
\end{proof}
\begin{corollary}
\label{C:OneAlepFlat} Let $R$ be a commutative local ring with maximal $P$. Assume that $P^2=0$. Let $q$ a positive integer. Then:
\begin{enumerate}
\item each $(1,q)$-flat module is $(\aleph_0,q)$-flat;
\item each $(1,q)$-injective module is $(\aleph_0,q)$-injective.
\end{enumerate}
\end{corollary}
\begin{proof} Let the notations be as in the previous theorem. We may assume that $G\subseteq PR^{n+1}$. Then $G$ is a semisimple module and $H$ is a direct summand of $G$. So, $(1)$ is a consequence of Theorem~\ref{T:p-flatideal}.
$(2)$. Let $M$ be a $(1,q)$-injective module. We shall prove by induction on $n$ that $M$ is $(n,q)$-injective. We have the following commutative diagram:
\[\begin{matrix}
0 & \rightarrow & \mathrm{Hom}_R(R,M)& \rightarrow & \mathrm{Hom}_R(R^{n+1},M)& \rightarrow & \mathrm{Hom}_R(R^n,M)& \rightarrow & 0\\
{} & {} & \downarrow & {} & \downarrow & {} & \downarrow & {} & {} \\
0 & \rightarrow & \mathrm{Hom}_R(G',M)& \rightarrow & \mathrm{Hom}_R(G,M)& \rightarrow & \mathrm{Hom}_R(H,M)& \rightarrow & 0
\end{matrix}\]
where the horizontal sequences are exact. By the induction hypothesis the left and the right vertical maps are surjective. It follows that the middle vertical map is surjective too.
\end{proof}
\bigskip
By \cite[Example 5.2]{Sha01} or \cite[Theorem 2.3]{Jon71}, for each integer $n>0$, there exists a ring $R$ for which each finitely generated left ideal is $(1,n)$-flat (hence $(\aleph_0,n)$-flat by Corollary~\ref{C:p-flatideal}) but there is a finitely generated left ideal which is not $(1,n+1)$-flat. The following proposition gives other examples in the commutative case.
\begin{proposition}
\label{P:ExFlat} Let $R$ be a commutative local ring with maximal ideal $P$ and residue field $k$. Assume that $P^2=0$ and $\dim_k\ P>1$. Then, for each positive integer $p<\dim_k\ P$, there exists:
\begin{enumerate}
\item a $(p+1,1)$-presented $R$-module which is $(\aleph_0,p)$-flat but not $(1,p+1)$-flat;
\item a $(\aleph_0,p)$-injective $R$-module which is not $(1,p+1)$-injective.
\end{enumerate}
\end{proposition}
\begin{proof}
(1). Let $F$ be a free $R$-module of rank $(p+1)$ with basis $\{e_1,\dots,e_p,e_{p+1}\}$, let $(a_1,\dots,a_p,a_{p+1})$ be a family of linearly independent elements of $P$, let $K$ be the submodule of $F$ generated by $\sum_{i=1}^{p+1}a_ie_i$ and let $M=F/K$. Then $M\cong D(W_{p,1,p+1})$ (see the proof of Proposition~\ref{P:locEnd}). First, we show that $K$ is a $(1,p)$-pure submodule of $F$. We consider the following equation:
\begin{equation}\label{eq:s4}
\sum_{j=1}^pr_jx_j=s(\sum_{i=1}^{p+1}a_ie_i)
\end{equation}
where $r_1,\dots,r_p,s\in R$ and with unknowns $x_1,\dots,x_p$. Assume that this equation has a solution in $F$. Suppose there exists $\ell,\ 1\leq\ell\leq p$, such that $r_{\ell}$ is a unit. For each $j,\ 1\leq j\leq p$, we put $x'_j=\delta_{j,\ell}r_{\ell}^{-1}s(\sum_{i=1}^{p+1}a_ie_i)$. It is easy to check that $(x'_1,\dots,x'_p)$ is a solution of (\ref{eq:s4}) in $K$. Now we assume that $\ r_j\in P,\ \forall j,\ 1\leq j\leq p$. Suppose that $(x_1,\dots,x_p)$ is a solution of (\ref{eq:s4}) in $F$. For each $j,\ 1\leq j\leq p$, $x_j=\sum_{i=1}^{p+1}c_{j,i}e_i$, where $c_{j,i}\in R$. We get the following equality:
\begin{equation}
\sum_{i=1}^{p+1}\left( \sum_{j=1}^pr_jc_{j,i}\right) e_i=\sum_{i=1}^{p+1}sa_ie_i
\end{equation}
We deduce that:
\begin{equation}
\forall i,\quad 1\leq i\leq p+1,\quad\sum_{j=1}^pr_jc_{j,i}=sa_i
\end{equation}
So, if $s$ is a unit, $\forall i,\ 1\leq i\leq p+1,\ a_i\in\sum_{j=1}^pRr_j$. It follows that \[\dim_k\left( \sum_{i=1}^{p+1}Ra_i\right) \leq p\] that is false. So, $s\in P$. In this case (\ref{eq:s4}) has the nil solution. Hence $M$ is $(\aleph_0,p)$-flat by Proposition~\ref{P:flat}(2) and Corollary~\ref{C:OneAlepFlat}.
By way of contradiction suppose that $M$ is $(1,p+1)$-flat. It follows that $K$ is a $(1,p+1)$-pure submodule of $F$ by Proposition~\ref{P:flat}. Since $M$ is $(1,p+1)$-pure-projective we deduce that $M$ is free. This is false.
(2). Let $E$ be an injective $R$-cogenerator. Then $\mathrm{Hom}_R(M,E)$ is $(\aleph_0,p)$-injective but not $(1,p+1)$-injective by Proposition~\ref{P:flat}(4).
\end{proof}
\medskip
In a similar way we show the following proposition.
\begin{proposition}
Let $R$ be a commutative local ring with maximal ideal $P$. Assume that $P^2=0$. Let $M$ be a $(m,1)$-presented $R$-module with $m>1$, let $\{x_1,\dots,x_m\}$ be a spanning set of $M$ and let $\sum_{j=1}^ma_jx_j=0$ be the relation of $M$, where $a_1,\dots,a_m\in P$. If $p=\mathrm{gen}\ (\sum_{j=1}^mRa_j)-1>0$, then:
\begin{enumerate}
\item $M$ is $(\aleph_0,p)$-flat but not $(1,m)$-flat;
\item $\mathrm{Hom}_R(M,E)$ is $(\aleph_0,p)$-injective but not $(1,m)$-injective, where $E$ is an injective $R$-cogenerator.
\end{enumerate}
\end{proposition}
When $R$ is an arithmetical commutative ring, i.e. its lattice of ideals is distributive, each $(1,1)$-flat module is flat and by \cite[Theorem VI.9.10]{FuSa01} the converse holds if $R$ is a commutative domain (it is also true if each principal ideal is flat). However we shall see that there exist non-arithmetical commutative rings for which each $(1,1)$-flat module is flat. Recall that a left (or right) $R$-module $M$ is {\it torsionless} if the natural map $M\rightarrow (M^*)^*$ is injective.
\begin{proposition}
\label{P:AlephInj} For each ring $R$ the following conditions are equivalent:
\begin{enumerate}
\item $R$ is right self $(\aleph_0,1)$-injective;
\item each finitely presented cyclic left $R$-module is torsionless;
\item each finitely generated left ideal $A$ satisfies $A=\mathrm{l-ann}(\mathrm{r-ann}(A)).$
\end{enumerate}
\end{proposition}
\begin{proof}
We prove $(1)\Leftrightarrow (2)$ as \cite[Theorem 2.3]{Ja73} and $(2)\Leftrightarrow (3)$ is easy.
\end{proof}
\begin{theorem}
\label{T:parfait} Let $R$ be a right perfect ring which is right self $(\aleph_0,1)$-injective. Then each $(1,1)$-flat right module is projective.
\end{theorem}
\begin{proof}
Let $M$ be a $(1,1)$-flat right $R$-module. It is enough to show that $M$ is flat. Let $A$ be a finitely generated left ideal of $R$. Assume that $\{a_1,\dots,a_n\}$ is a minimal system of generators of $A$ with $n>1$. Let $z\in M\otimes_RA$ such that its image in $M$ is $0$. We have $z=\sum_{i=1}^ny_i\otimes a_i$, where $y_1,\dots,y_n\in M$, and $\sum_{i=1}^ny_ia_i=0$. For each $i,\ 1\leq i\leq n$, we set $A_i=\sum_{\binom{j=1}{j\ne i}}^nRa_j$. Then, $\forall i,\ 1\leq i\leq n$, $A_i\subset A$. For each finitely generated left ideal $B$ we have $B=\mathrm{l-ann}(\mathrm{r-ann}(B))$. It follows that, $\forall i,\ 1\leq i\leq n$, $\mathrm{r-ann}(A)\subset \mathrm{r-ann}(A_i)$. Let $b_i\in \mathrm{r-ann}(A_i))\setminus \mathrm{r-ann}(A)$. Then $y_ia_ib_i=0$. From the $(1,1)$-flatness of $M$ we deduce that $y_i=\sum_{k=1}^{m_i}y'_{i,k}c_{i,k}$, where $y'_{i,1},\dots,y'_{i,m_i}\in M$ and $c_{i,1},\dots,c_{i,m_i}\in R$ with $c_{i,k}a_ib_i=0,\ \forall k,\ 1\leq k\leq m_i$. It follows that $z=\sum_{i=1}^n(\sum_{k=1}^{m_i}y'_{i,k}\otimes c_{i,k}a_i)$. Let $A^{(1]}$ be the left ideal generated by $\{c_{i,k}a_i\mid 1\leq i\leq n,\ 1\leq k\leq m_i\}$. Then $A^{(1)}\subset A$; else, $\forall i,\ 1\leq i\leq n$, $a_i=\sum_{j=1}^{n}(\sum_{k=1}^{m_j}d_{i,j,k}c_{j,k}a_j)$ with $d_{i,j,k}\in R$; we get that $a_ib_i=\sum_{j=1}^{n}(\sum_{k=1}^{m_j}d_{i,j,k}c_{j,k}a_jb_i)$; but $a_jb_i=0$ if $j\ne i$ and $c_{i,k}a_ib_i=0$; so, there is a contradiction because the second member of the previous equality is $0$ while $a_ib_i\ne 0$ . Let $\{a^{(1)}_1,\dots,a^{(1)}_{n_1}\}$ be a minimal system of generators of $A^{(1)}$. So, $z=\sum_{i=1}^{n_1}y^{(1)}_i\otimes a^{(1)}_i$ where $y^{(1)}_1,\dots,y^{(1)}_{n_1}\in M$, and $z$ is the image of $z^{(1)}\in M\otimes_RA^{(1)}$ defined by $z^{(1)}=\sum_{i=1}^{n_1}y^{(1)}_i\otimes a^{(1)}_i$. If $n_1\leq 1$ we conclude that $z^{(1)}=0$ since $M$ is $(1,1)$-flat, and $z=0$. If $n_1>1$, in the same way we get that $z^{(1)}$ is the image of an element $z^{(2)}\in M\otimes_RA^{(2)}$ where $A^{(2)}$ is a left ideal such that $A^{(2)}\subset A^{(1)}$. If $\mathrm{gen}\ A^{(2)}>1$ we repeat this process, possibly several times, until we get a left ideal $A^{(l)}$ with $\mathrm{gen}\ A^{(l)}\leq 1$; this is possible because $R$ satisfies the descending chain condition on finitely generated left ideals since it is right perfect (see \cite[Th\'eor\`eme 5 p.130]{Ren75}). The $(1,1)$-flatness of $M$ implies that $z^{(l)}=0$ and $z=0$. So, $M$ is projective.
\end{proof}
Let $\mathcal{P}$ be a ring property. We say that a commutative ring $R$ is {\it locally $\mathcal{P}$} if $R_P$ satisfies $\mathcal{P}$ for each maximal ideal $P$.
The following corollary is a consequence of Theorem~\ref{T:parfait} and Proposition~\ref{P:localflat}.
\begin{corollary}
\label{C:LocPer} Let $R$ be a commutative ring which is locally perfect and locally self $(\aleph_0,1)$-injective. Then each $(1,1)$-flat $R$-module is flat.
\end{corollary}
\section{$(n,m)$-coherent rings}
\label{S:coherent}
We say that a ring $R$ is left $(n,m)${\it -coherent} if each $m$-generated submodule of a $n$-generated free left $R$-module is finitely presented. We say that $R$ is left $(\aleph_0,m)${\it -coherent} (respectively $(n,\aleph_0)${\it -coherent}) if for each integer $n>0$ (respectively $m>0$) $R$ is left $(n,m)$-coherent. The following theorem can be proven with standard technique: see \cite[Theorems 5.1 and 5.7]{ZhChZh05}. In this theorem the integers $n$ or $m$ can be replaced with $\aleph_0$.
\begin{theorem}\label{T:coh}
Let $R$ be a ring and $n, m$ two fixed positive integers. Assume that $R$ is an algebra over a commutative ring $S$. Let $E$ be an injective $S$-cogenerator. Then the following conditions are equivalent:
\begin{enumerate}
\item $R$ is left $(n,m)$-coherent;
\item any direct product of right $(n,m)$-flat $R$-modules is $(n,m)$-flat;
\item for any set $\Lambda$, $R^{\Lambda}$ is a $(n,m)$-flat right $R$-module;
\item any direct limit of a direct system of $(n,m)$-injective left $R$-modules is $(n,m)$-injective;
\item for any exact sequence of left modules $0\rightarrow A\rightarrow B\rightarrow C\rightarrow 0$, $C$ is $(n,m)$-injective if so is $B$ and if $A$ is a $(\aleph_0,m)$-pure submodule of $B$;
\item for each $(n,m)$-injective left $R$-module $M$, $\mathrm{Hom}_S(M,E)$ is $(n,m)$-flat.
\end{enumerate}
\end{theorem}
\bigskip
It is well known that each left $(1,\aleph_0)$-coherent ring is left $(\aleph_0,\aleph_0)$-coherent. {\bf For each positive integer $p$, is each left $(1,p)$-coherent ring left $(\aleph_0,p)$-coherent?}
Propositions~\ref{P:PF=PP} and \ref{P:CommParf} and Theorem~\ref{T:parfait2} give a partial answer to this question.
\begin{proposition}
\label{P:PF=PP} Let $p$ be a positive integer and let $R$ be a ring. For each positive integer $n$, assume that, for each $p$-generated submodule $G$ of the left $R$-module $R^n\oplus R$, $(G\cap R^n)$ is the direct limit of its $p$-generated submodules. Then the following conditions are equivalent:
\begin{enumerate}
\item $R$ is left $(1,p)$-coherent;
\item $R$ is left $(\aleph_0,p)$-coherent.
\end{enumerate}
Moreover, when these conditions hold each $(1,p)$-injective left module is $(\aleph_0,p)$-injective.
\end{proposition}
\begin{proof}
It is obvious that $(2)\Rightarrow (1)$.
$(1)\Rightarrow (2)$. Let $\Lambda$ be a set. By Theorem~\ref{T:coh} $R^{\Lambda}$ is a $(1,p)$-flat right module. From Theorem~\ref{T:p-flatideal} we deduce that $R^{\Lambda}$ is a $(\aleph_0,p)$-flat right module. By using again Theorem~\ref{T:coh} we get $(2)$.
Let $M$ be a $(1,p)$-injective left module. By Theorem~\ref{T:coh} $M^{\sharp}$ is a $(1,p)$-flat right $R$-module. Then it is also $(\aleph_0,p)$-flat. We deduce that $(M^{\sharp})^{\sharp}$ is a $(\aleph_0,p)$-injective left module. Since $M$ is a pure submodule of $(M^{\sharp})^{\sharp}$, it follows that $M$ is $(\aleph_0,p)$-injective too.
\end{proof}
\begin{proposition}
\label{P:CommParf} Let $R$ be a commutative perfect ring. Then $R$ is Artinian if and only if it is $(1,1)$-coherent.
\end{proposition}
\begin{proof}
Suppose that $R$ is $(1,1)$-coherent. Since $R$ is perfect, $R$ is a finite product of local rings. So, we may assume that $R$ is local with maximal $P$. Let $S$ be a minimal non-zero ideal of $R$ generated by $s$. Then $P$ is the annihilator of $s$. So, $P$ is finitely generated and it is the sole prime ideal of $R$. Since all prime ideals of $R$ are finitely generated, $R$ is Noetherian. On the other hand $R$ satisfies the descending chain condition on finitely generated ideals. We conclude that $R$ is Artinian.
\end{proof}
\bigskip
Except in some particular cases, we don't know if each $(1,p)$-injective module is $(\aleph_0,p)$-injective, even if we replace $p$ by $\aleph_0$.
\begin{theorem}
\label{T:parfait2} Let $R$ be a ring which is right perfect, left $(1,1)$-coherent and right self $(\aleph_0,1)$-injective. Then each $(1,1)$-injective left module is $(\aleph_0,\aleph_0)$-injective and $R$ is left coherent.
\end{theorem}
\begin{proof}
Let $M$ be a left $(1,1)$-injective module. By Theorem~\ref{T:coh} $M^{\sharp}$ is $(1,1)$-flat. Whence $M^{\sharp}$ is projective by Theorem~\ref{T:parfait}. We do as in the proof of Proposition~\ref{P:PF=PP} to conclude that $M$ is $(\aleph_0,\aleph_0)$-injective.
For each set $\Lambda$, $R^{\Lambda}$ is a $(1,1)$-flat right module by Theorem~\ref{T:coh}. It follows that $R^{\Lambda}$ is a projective right module by Theorem~\ref{T:parfait}.
\end{proof}
Recall that a ring is {\it quasi-Frobenius} if it is Artinian and self-injective.
\begin{corollary}
\label{C:Frob} Let $R$ be a quasi-Frobenius ring. Then, for each right (or left) $R$-module $M$, the following conditions are equivalent:
\begin{enumerate}
\item $M$ is $(1,1)$-flat;
\item $M$ is projective;
\item $M$ is injective;
\item $M$ is $(1,1)$-injective.
\end{enumerate}
\end{corollary}
\begin{proof}
It is well known that $(2)\Leftrightarrow (3)$. By Theorem~\ref{T:parfait} $(1)\Leftrightarrow (2)$ because $R$ satisfies the conditions of this theorem, and it is obvious that $(3)\Rightarrow (4)$ and the converse holds by Theorem~\ref{T:parfait2}.
\end{proof}
We prove the following theorem as \cite[Th\'eor\`eme 1.4]{Cou82}.
\begin{theorem}\label{T:locCoh}
\label{T:pure-projective} Let $R$ be a commutative ring and $n, m$ two fixed positive integers. The following conditions are equivalent:
\begin{enumerate}
\item $R$ is $(n,m)$-coherent;
\item for each multiplicative subset $S$ of $R$, $S^{-1}R$ is $(n,m)$-coherent, and for each $(n,m)$-injective $R$-module $M$, $S^{-1}M$ is $(n,m)$-injective over $S^{-1}R$;
\item For each maximal ideal $P$, $R_P$ is $(n,m)$-coherent and for each $(n,m)$-injective $R$-module $M$, $M_P$ is $(n,m)$-injective over $R_P$.
\end{enumerate}
\end{theorem}
Recall that a ring $R$ is a {\it right IF-ring} if each right injective $R$-module is flat.
\begin{theorem}
\label{T:LocPerIF} Let $R$ be a commutative ring which is locally perfect, $(1,1)$-coherent and self $(1,1)$-injective. Then:
\begin{enumerate}
\item $R$ is coherent, self $(\aleph_0,\aleph_0)$-injective and locally quasi-Frobenius;
\item each $(1,1)$-flat module is flat;
\item each $(1,1)$-injective module is $(\aleph_0,\aleph_0)$-injective.
\end{enumerate}
\end{theorem}
\begin{proof}
By Theorem~\ref{T:locCoh} $R_P$ is $(1,1)$-coherent and $(1,1)$-injective for each maximal ideal $P$. Let $a$ be a generator of a minimal non-zero ideal of $R_P$. Then $PR_P$ is the annihilator of $a$ and consequently $PR_P$ is finitely generated over $R_P$. Since all prime ideals of $R_P$ are finitely generated, we deduce that $R_P$ is Artinian for each maximal ideal $P$. Moreover, the $(1,1)$-injectivity of $R_P$ implies that the socle of $R_P$ (the sum of all minimal non-zero ideals) is simple. It follows that $R_P$ is quasi-Frobenius for each maximal ideal $P$.
Let $M$ be a $(\aleph_0,\aleph_0)$-injective $R$-module. By Theorem~\ref{T:locCoh} $M_P$ is $(1,1)$-injective for each maximal ideal $P$. By Corollary~\ref{C:Frob} $M_P$ is injective for each maximal ideal $P$. We conclude that $R$ is self $(\aleph_0,\aleph_0)$-injective and it is coherent by Theorem~\ref{T:locCoh}.
If $M$ is $(1,1)$-injective, we prove as above that $M_P$ is injective for each maximal ideal $P$. It follows that $M$ is $(\aleph_0,\aleph_0)$-injective.
The second assertion is an immediate consequence of Corollary~\ref{C:LocPer}.
\end{proof}
The following proposition is easy to prove:
\begin{proposition}
\label{P:1coh} A ring $R$ is left $(\aleph_0,1)$-coherent if and only if each finitely generated right ideal has a finitely generated left annihilator.
\end{proposition}
\begin{example}
Let $V$ be a non-Noetherian (commutative) valuation domain whose order group is not the additive group of real numbers and let $R=V[[X]]$ be the power series ring in one indeterminate over $V$. Since $R$ is a domain, $R$ is $(\aleph_0,1)$-coherent. But, in \cite{AnWa87} it is proven that there exist two elements $f$ and $g$ of $R$ such that $Rf\cap Rg$ is not finitely generated. By using the exact sequence $0\rightarrow Rf\cap Rg\rightarrow Rf\oplus Rg\rightarrow Rf+Rg\rightarrow 0$ we get that $Rf+Rg$ is not finitely presented. So, $R$ is not $(1,2)$-coherent.
\end{example}
|
\section{Introduction}
Though the CKM paradigm~\cite{NC63, KM73} of CP violation in the Standard Model (SM) has been extremely successful in
describing a multitude of experimental data, in the past few years some indications of
deviations have surfaced, specifically in the flavor sector~\cite{LS07,LS08,LS09,uli_lenz,bona}. An intriguing aspect
of these deviations is that so far they have more prominently,
though not exclusively,
occurred in CP violating observables
only. While many beyond the standard model (BSM) scenarios
can account for such effects~\cite{APS1, APS2, AJB081, AJB082, MN08,lang,paridi}, a very simple extension of the SM
that can cause these anomalies is the addition of an
extra family as we emphasized in a recent study~\cite{SAGMN08,as_moriond09}. In this paper,
we will extend our previous work and study the implications of the standard model
with four generations (SM4) in rare B and K decays.
Although our initial motivation for studying SM4 was triggered
by the deviations in the CP violating observables in B, $B_s$ decays,
we want to stress that actually SM4 is, in fact, a very simple and interesting
extension of the three generation SM (SM3). The fact that the heavier
quarks and leptons in this family can play a crucial role in dynamical
electroweak-symmetry breaking (DEWSB) as an economical way to address the hierarchy
puzzle renders this extension of SM3 especially interesting. In addition,
whereas, as is widely recognized SM3 does not have enough CP
to facilitate baryogenesis, that difficulty is readily and significantly ameliorated in SM4~\cite{Hou08,CJ88, GK08}. Besides, given that three families exist, it is
clearly important to search for the fourth.
That rare B-decays are particularly sensitive to the fourth generation was in fact emphasized
long ago~\cite{AS_olds1,AS_olds2,AS_olds3,AS_olds4,ge_olds}. The potential role of heavy quarks in DEWSB was also another reason
for the earlier interest~\cite{norton,symp_SM4_8789,Holdom:1986rn,Hung:2009ia,Hashimoto:2009ty}. LEP/SLC discovery that a fourth family
(essentially) massless neutrino does not exist was one reason that caused
some pause in the interest
on SM4. A decade later discovery of neutrino oscillations and of neutrino mass
managed to off-set to some degree this concern about the 4th family's
necessarily involving massive neutrino.
Electroweak precision tests provide a very important constraint on the mass difference of the 4th family isodoublet. In this context the PDG reviews
for a number of years may have been declaring a ``prematured death" of the fourth
family~\cite{erler}; careful studies show in fact that while mass difference between the isodoublet quarks is constrained to be less than $\approx 75$ GeV, an extra generation of quarks is not excluded by the current data. In fact,
it is also claimed that for
certain values of particle masses the quality of the fit with four generations
is comparable to that of the SM3~\cite{novikov1,novikov2,Kribs_EWPT,chanowitz}.
The addition of fourth generation to the SM means that the quark
mixing matrix will now become a $4 \times 4$ matrix ($V_{CKM4}$) and the parametrization of
this unitary matrix requires six real parameters and three phases. The two extra phases imply the possibility
of extra sources of CP violation \cite{AS_olds3}.
In \cite{SAGMN08}, it was shown that a fourth family of quarks with $m_{t'}$ in the range of
(400 - 600)
GeV provides a simple explanation for the several indications of new
physics that have been observed involving CP asymmetries in the B, $B_s$
decays~\cite{LS07,LS08,LS09,uli_lenz,bona}. The
built-in hierarchy of $V_{CKM4}$ is such that the $t'$
readily provides a needed perturbation ($\approx 15\%$) to $\sin 2 \beta$ as measured in
$B \to \psi K_s$ and simultaneously is the
dominant source of CP asymmetry in $B_s \to \psi \phi$.
While most of the B, $B_s$ CP-anomalies are easily accommodated and explained
by SM4, we note that, in contrast, EW precision tests constrain the
mass-splitting between $t'$ and $b'$ to be small,
around $70$ GeV \cite{NEED, novikov1,novikov2, Kribs_EWPT}; so for $m_{t'}$ of O(500 GeV) their masses have to be degenerate
to O(15\%). As far
as the lepton sector is concerned, it is clear that the 4th family lepton has
to be quite different from the previous three families in that the neutral
lepton has to be rather massive, with mass $> m_Z/2$. This may also be a
clue that the underlying nature of the 4th family may be quite different
from the previous three families \cite{DM4}.
In this paper we extend our previous work~\cite{SAGMN08} on the implications of SM4, to
study the direct CP asymmetry in $B\to X_s \gamma$, $B\to X_s\, l^+ \, l^-$ and in
$B_s\to X_s\ell\nu$, forward-backward (FB) asymmetry in $B\to X_s\ (K^*) l^+ \ l^-$,
decay rates of $B\to X_s \nu \bar \nu $, $B_s\to \mu^+\mu^-, \tau^+\tau^-$ and
$K_L \to \pi^0\nu \bar \nu$ and
CP violation in $B \to \pi K$ and $B^0 \to \pi^0 \pi^0$ modes. We show that
SM4 can ameliorate the difficulty in understanding the large
difference, O(15\%), between the direct CP asymmetries in neutral B decays
to $K^+ \pi^-$ versus that of the charged B-decays to $K^+ \pi^0$ partly due to the
enhanced isospin violation that SM4 causes in flavor-changing penguin
transitions due to the heavy $m_{t'}$~\cite{AS_olds1} originating from the
evasion of the decoupling theorem and partly if the corresponding strong phase(s) are large
in SM4. The enhanced electroweak penguin amplitude
provides a color-allowed ($Z\to \pi^0$) contribution which is not present for
$\pi^{\pm}$ case. However, we want to emphasize that the prediction obtained
using the QCD factorization approach~\cite{QCDF1,QCDF,LS07} depends on many
input parameters therefore it has large theoretical uncertainties. Apart
from the SM parameters such as CKM matrix, quark masses, the strong
coupling constant and hadronic parameters there are large
theoretical uncertainties related to the modeling of power corrections
corresponding to weak annihilation effects and the chirally-enhanced power
corrections to hard spectator scattering. Therefore the numerical results for
the direct CP asymmetries are not reliable.
Several of these observables like FB asymmetry in $B\to K^* l^+ \ l^-$
\cite{Hovhannisyan:2007pb}, CP asymmetry in $B_s\to \psi\phi$ \cite{hou03}
and the decay rate of $K_L \to \pi^0\nu \bar \nu$
\cite{Hou:2005yb} have also been studied before, as well as
many other interesting aspects of SM4 by Hou and collaborators \cite{Hou:2005hd,
Arhrib:2006pm, Hou:2006zza, Hou:2006jy}, see also \cite{lenz4}.
However, their analysis was generally restricted to $m_{t'}$ of $\sim\,$ 300 {\it GeV}.
On the other hand, our analysis seems to favor $m_{t'}$ in the range of
(400 - 600) GeV to explain the observed CP asymmetries in the B, $B_s$ decays.
We note also that recent analysis by Chanowitz seems to disfavor most of the
parameter space they have used \cite{chanowitz} whereas our parameter space is
largely unaffected \cite{NTU}.
We identify several processes wherein SM4 causes large deviations from the expectations of SM3;
for example, $B \to X_s \nu \bar \nu$, $B_s\to \mu^+\mu^-$, $A_{SL}(B_s\to X_s\ell\nu )$,
$a_{CP} (B \to \pi K)$, $a_{CP} (B \to \pi^0 \pi^0)$,
$K_L \to \pi^0 \nu \bar \nu$ and of course mixing-induced CP in
$B_s \to \psi \phi$ etc.
These observables will be measured with higher statistics at the upcoming
high intensity K, B, $B_s$ experiments at CERN, FERMILAB, JPARC facilities etc
and in particular at the LHCb experiment and possibly also at the Super-B factories and hence may provide further indirect evidence for an additional family of quarks.
The paper is arranged as follows. After the introduction, we provide constraints on the 4$\times$4 CKM matrix by incorporating oblique corrections along
with experimental data from important observables
involving Z, B and K decays as well as $B_d$ and $B_s$ mixings etc.
In Sec. \ref{results}, we present the estimates of many useful observables in the SM4. Finally in Sec. \ref{concl}, we present our summary.
\section{Constraints on the CKM4 matrix elements }
\label{constraint_ckm}
In our previous article \cite{SAGMN08}, to find the limits on $V_{CKM4}$
elements, we concentrated mainly on the constraints that will come from
vertex correction to $Z\to b\bar{b}$, $Br(B\to X_s \gamma)$,
$Br(B\to X_s \, l^+ \, l^-)$, $B_d - \bar{B_d}$ and
$B_s - \bar{B_s}$ mixing, $Br(K^+\to \pi^+\nu\nu)$ and
indirect {\it CP} violation in $K_L \to \pi\pi$ described
by $|\epsilon_k|$. We did not consider $\epsilon'/\epsilon$ as a constraint
because of its large hadronic uncertainties. Chanowitz \cite{chanowitz}
has shown that as $m_{t'}$ becomes very large more important constraint is from non decoupling oblique
corrections rather than the vertex correction to $Z\to b\bar{b}$. In this
article we have extended our analysis by including the constraint form
non decoupling oblique corrections as well; we note that for $m_{t'}\, \:\raisebox{-0.5ex}{$\stackrel{\textstyle<}{\sim}$}\: 500$ {\it GeV} our
previous constraints are largely unaffected
but for $m_{t'}\approx 600$ {\it GeV} the oblique corrections start to have effect.
With the inputs given in Table. (\ref{tab3})
we have made the scan over the entire parameter space by a flat
random number generator and obtained the constraints on various parameters of
the 4$\times$4 mixing matrix. In the following subsections we briefly
discuss the various input parameters used in our analysis.
\subsection{Oblique correction}
The $Z$ pole, $W$ mass, and low-energy data can be used to search for and set limits on
deviations from the SM. Most of the effects on precision measurements can be described
by the three gauge self-energy parameters $S$, $T$ and $U$. We assume these parameters to
be arising from new physics only i.e
they are equal to zero exactly in SM, and do not include any contributions
from $m_t$ and $M_H$.
The effects of non-degenerate multiplets of chiral fermions can be
described by just three parameters, $S$, $T$ and $U$ at the one-loop level
\cite{peskin,novikov1,sher,erler,novikov3}. $T$ is proportional
to the difference between the $W$ and $Z$ self-energies at $Q^2=0$, while $S$
is associated with the difference between the $Z$ self-energy at
$Q^2=M^2_{Z}$ and $Q^2=0$ and $(S+U)$ is associated with the difference
between $W$ self-energy at $Q^2=M^2_{W}$ and $Q^2=0$.
A non-degenerate $SU(2)$ doublet $\binom{f1}{f2}$ with masses $m_1$ and $m_2$
respectively yields the contributions \cite{peskin}
\begin{eqnarray}
S &=&\frac{1}{6\pi}\Big[1- Y \ln({m^2_1/ m^2_2})\Big],\\ \nonumber
T &=& \frac{1}{16 \pi s_W^2 c_W^2 M^2_Z}\Big[m^2_1+m^2_2-\frac{2 m^2_1 m^2_2}
{m^2_1-m^2_2}\ln({m^2_1/ m^2_2})\Big],\\ \nonumber
U &=&\frac{1}{6\pi}\Big[- \frac{5 m^4_1-22 m^2_1 m^2_2 + 5 m^4_2}{3(m^2_1-m^2_2)^2}+ \frac{m^6_1-3 m^4_1 m^2_2 - 3 m^2_1 m^4_2+ m^6_2}{(m^2_1-m^2_2)^3}
\ln({m^2_1/ m^2_2})\Big],
\end{eqnarray}
where $Y$ is the hypercharge of the doublet.
A heavy non-degenerate doublet of fermions contributes positively
to $T$ as
\begin{equation}
\rho^{\ast}_0-1= \frac{1}{1-\alpha T}-1 \approx \alpha T,
\end{equation}
where $\rho^{\ast}_0$ denotes the low-energy ratio of neutral to charged current
couplings in neutrino interactions.
The parameter $U$ plays a fairly unimportant role, all the neutral current and
low energy observables depend only on $S$ and $T$ \cite{peskin}.
In addition $U$ is often predicted to be very small. In most of the models $U$
should differ from zero by only a percent of $T$.
In the case of an extra family with the doublet $\binom{t'}{b'}$, the
contribution to $T$ and $S$ parameters are given by \cite{chanowitz}
\begin{eqnarray}
T_4&=&\frac{1}{8 \pi x_W (1- x_W)}\Big[3 \Big(|V_{t'b'}|^2 \delta{m}_{t'b'}+
|V_{t'b}|^2 \delta{m}_{t'b} + |V_{tb'}|^2 \delta{m}_{tb'} - |V_{t'b}|^2
\delta{m}_{tb} \\ \nonumber
&{}& + |V_{t's}|^2 \delta{m}_{t's}\Big) + \delta{m}_{l_4\nu_4}\Big],\\
S_4&=&\frac{3}{6\pi}\Big(1- \frac{1}{3} \ln{\frac{m^2_{t'}}{m^2_{b'}}}\Big),
\end{eqnarray}
with
\begin{equation}
\delta{m}_{12}= \frac{1}{2 M^2_Z}\Big({m^2_1+m^2_2}-\frac{2 m^2_1 m^2_2}
{m^2_1-m^2_2} \ln({m^2_1/ m^2_2})\Big).
\end{equation}
\subsection{Vertex corrections to $Z\to b\bar{b}$}
Including QCD and QED corrections, the $Z\to b\bar{b}$ decay width can be
written as \cite{zbb1}
\begin{eqnarray}
\Gamma(Z\to q\bar{q}) &=& {\frac{N_c}{48}} {\frac{\alpha} {s^2_W c^2_W}} m_Z\left(|a_q|^2 + |v_q|^2\right)
\nonumber \\
&{}& \times\big(1 + \delta^{(0)}_b \big) \big(1 + \delta^q_{QED}\big)\big(1 + \delta^q_{QCD}\big)\big(1 + \delta^q_{\mu}\big)\big(1 + \delta^q_{tQCD}\big)\big(1 + \delta_q\big),
\end{eqnarray}
where
\begin{equation}
v_q = \Big(2 I^q_3 - 4 |Q_q| s^2_W\Big), \hskip 20pt a_q = 2 I^q_3,
\end{equation}
and $\delta$'s are various corrections which are discussed below.
In the decay of the $Z\to b\bar{b}$, the top quark mass
enters in the loop correction to the vertex mediated by the W gauge boson. Due
to spontaneous symmetry breaking effects the top mass can not be neglected in
the calculation. In fact there is a top mass dependence that grows like
$\frac{m^2_t}{m^2_Z}$ as in many other one-loop weak processes such as $K-\bar{K}$,\, $B-\bar{B}$\, ($\Delta F = 2$ mixings),\, $b\to s\ell^+\ell^-$\, etc.
The additional contribution to the $Zb\bar{b}$ vertex,
due to nonzero value of the top quark mass can be written as:
\begin{equation}
\delta_b \approx 10^{-2}\left(\Big(-\frac{m^2_t}{2 m^2_Z} + 0.5\Big)|V_{tb}|^2
+ \Big(- \frac{m^2_{t'}}{2 m^2_Z} + 0.5\Big)|V_{t'b}|^2\right).
\end{equation}
$\delta^{q}_{QED}$ gives small final-state QED corrections
that depend on the charge of final fermion,
\begin{equation}
\delta^q_{QED} = \frac{3\alpha}{4\pi} Q^2_q.
\end{equation}
It is very small (0.2\% for charged leptons, 0.8\% for u-type quarks and 0.02\%
for d-type quarks).
$\delta_{QCD}$ gives the QCD corrections common to all quarks and it is given
by
\begin{equation}
\delta_{QCD} = \frac{\alpha_s}{\pi} + 1.41 \Big(\frac{\alpha_s}{\pi} \Big)^2.
\end{equation}
$\alpha_s$ is the QCD coupling constant taken at the $m_Z$ scale, i.e. $\alpha_s= \alpha_s(m^2_Z) = 0.12$.
$\delta^q_{\mu}$ contains the kinematical effects of the external fermion masses, including some mass-dependent QCD radiative corrections. It is only important
for the b-quark (0.5\%) and to a lesser extent for the $\tau$-lepton (0.2\%)
and the c-quark (0.05\%). It is given by
\begin{equation}
\delta^q_{\mu} = \frac{3 \mu^2_q} {v^2_q + a^2_q}\left(- \frac{1}{2} a^2_q\left(1 + \frac{8 \alpha_s} {3\pi}\right) + v^2_q \frac{\alpha_s}{\pi}\right),
\end{equation}
where $\mu^2_q \equiv 4\bar{m}^2_q(m^2_Z)/ m^2_Z$.
By taking appropriate branching ratios it is possible to isolate the large top
mass dependent $Zb\bar{b}$ vertex $\delta_b$ \cite{zbb1},
\begin{equation}
R_h \equiv \frac{\Gamma(Z\to b\bar{b})} {\Gamma(Z\to hadrons)} = \big( 1 + 2/R_s
+ 1/R_c + 1/R_u \big)^{-1},
\end{equation}
where $R_q \equiv \frac{\Gamma(Z\to b\bar{b})} {\Gamma(Z\to q\bar{q})}$.
All other corrections cancel exactly in this branching ratio except the
correction to the $Zb\bar{b}$ vertex which only depends on the top quark mass.
\subsection {$B \to X_s \gamma$ decay}
Radiative B decays have been a topic of great theoretical and experimental interest for long.
Although the inclusive radiative decay $B \to X_s \gamma$ is
loop suppressed within the SM, it has relatively large
branching ratio making it statistically favorable from the experimental
point of view and hence it serves as an important
probe to test SM and its possible extensions. The present world
average of $Br(B\to X_s \gamma)$ is $(3.55 \pm 0.25)\times 10^{-4}$ \cite{Barberio:2008fa}
which is in good agreement with its SM prediction \cite{Misiak:2006ab,Misiak:2006zs}.
Apart from the branching ratio of $B \to X_s \gamma$,
direct $CP$ violation in $B \to X_s \gamma$, $A_{CP}^{B \to X_s \gamma}$ ,
can serve as an important observable to search physics beyond SM;
therefore we will also study this direct CP asymmetry in this paper
(see Section \ref{dacpbsg}).
The quark level transition $b \to s \gamma$ induces the
inclusive $B \to X_s \gamma$ decay. The effective
Hamiltonian for $b \to s \gamma$ can be written in the following form
\begin{equation}
{\cal H}_{eff} = \frac{4 G_F}{\sqrt{2}} V_{ts}^{*} V_{tb}
\sum_{i=1}^{8} C_i(\mu) \, Q_i(\mu)\;,
\end{equation}
where the form of operators $O_i(\mu)$ and the expressions for calculating the Wilson
coefficients $C_i(\mu)$ are given in \cite{Buras:1994dj}.
The introduction of fourth generation changes the values of
Wilson coefficients $C_7$ and $C_8$ via the virtual exchange
of the $t'$-quark and can be written as
\begin{equation}
C_{7,8}^{\rm tot}(\mu) = C_{7,8}(\mu) + \frac{ V_{t' s}^{*} V_{t' b}}
{ V_{ts}^{*} V_{tb}} C_{7,8}^{t'} (\mu)\;.
\label{wtot_78}
\end{equation}
The values of $C_{7,8}^{t'}$ can be calculated from the expression
of $C_{7,8}$ by replacing the mass of $t$-quark by $m_{t'}$.
In order to reduce the uncertainties arising from $b$--quark
mass, we consider the following ratio
\begin{equation}
R = \frac{Br(B \to X_s \gamma)}{Br(B \to X_c e \bar \nu_e)}\;. \nonumber
\end{equation}
In leading logarithmic approximation this ratio can be written as \cite{Buras:1997fb}
\begin{equation}
R = \frac{\left| V_{ts}^{*} V_{tb} \right|^2}{\left| V_{cb} \right|^2} \,\,
\frac{6 \alpha \left| C_7^{\rm tot}(m_b) \right|^2}{\pi f(\hat m_c) \kappa(\hat m_c)}\;.
\label{R}
\end{equation}
Here the Wilson coefficient $C_7$ is evaluated at the scale $\mu=m_b$.
The phase space factor $f(\hat{m_c})$ in $Br(B \to X_c e {\bar \nu})$
is given by \cite{Nir:1989rm}
\begin{equation}
f(\hat{m}_c) = 1 - 8\hat{m}^2_c + 8\hat{m}_c^6 - \hat{m}_c^8 -
24\hat{m}_c^4 \ln \hat{m}_c \;.
\end{equation}
$\kappa(\hat{m_c})$ is the $1$-loop QCD correction factor
\cite{Nir:1989rm}
\begin{equation}
\kappa(\hat{m_c})=1-\frac{2\alpha_s(m_b)}{3\pi}\left[\left(\pi^2-\frac{31}{4}\right)(1-\hat{m_c})^2+\frac{3}{2}\right]\;.
\end{equation}
Here $\hat{m_c}=m_c/m_b$.
\subsection {$B \to X_s\, l^+ \,l^-$ decay}
The quark level transition $b \to s\, l^+ \, l^-$ is responsible for the inclusive decay $B \to X_s \, l^+ \, l^-$.
We apply the same approach introduced for $b \to s \gamma$.
The effective Hamiltonian for the decay $b \to s\, l^+\, l^-$ is given by
\begin{equation}
{\cal H}_{eff} = \frac{4 G_F}{\sqrt{2}} V_{ts}^{*} V_{tb}
\sum_{i=1}^{10} C_i(\mu) \, Q_i(\mu)\;.
\end{equation}
In addition to the operators relevant for $b \to s \gamma$, there are two new
operators:
\begin{equation}
Q_{9} = (\bar{s}b)_{V-A}(\bar{l}l)_V, \hskip 20pt Q_{10} = (\bar{s}b)_{V-A}(\bar{l}l)_V\;.
\end{equation}
The amplitude for the decay $B \to X_s \, l^+ \, l^-$
in SM4 is given by
\begin{eqnarray}
M &~=~& \frac{G_F \alpha}{\sqrt{2} \pi}
\, V_{ts}^*V_{tb} \, \Bigl[ C_9^{\rm tot} \, \bar{s}
\gamma_\mu P_L b \, \bar{l} \gamma_\mu l+ C^{\rm tot}_{10}
\, \bar{s} \gamma_\mu P_L b \, \bar{l} \gamma_\mu \gamma_5
l \nonumber \\
&& \hskip2.5truecm
-~2m_b \, \frac{C^{\rm tot}_7}{q^2} \, \bar{s} i
\sigma_{\mu\nu} q^\nu P_R b \,
\bar{l} \gamma_\mu l \Bigr] ~,
\label{matrix}
\end{eqnarray}
where $P_{L,R} = (1 \mp \gamma_5)/2$ and $q$ is the sum of
$l^+$ and $l^-$ momenta. Here the Wilson coefficients are evaluated at $\mu$=$m_b$.
The differential branching ratio is given by
\begin{equation}
\frac{{\rm d}Br(\bs)}{{\rm d}z}
= \frac{\alpha^2 B(B\rightarrow X_c e {\bar \nu})}
{4 \pi^2 f(\hat{m_c})\kappa(\hat{m}_c)} (1-z)^2\left(1-\frac{4t^2}{z}\right)^{1/2}
\frac{|V_{tb}^{*}V_{ts}|^2}{|V_{cb}|^2} D(z)\,,
\label{eq:dbr_bsll}
\end{equation}
where
\begin{eqnarray}
D(z) &=& |C_9^{\rm tot}|^2\left(1+\frac{2t^2}{z}\right)(1+2z)
+ 4|C_7^{\rm tot}|^2\left(1+ \frac{2t^2}{z}\right)\left(1+\frac{2}{z}\right) \nonumber \\
& & + |C_{10}^{\rm tot}|^2 \left[ ( 1 + 2z) + \frac{2t^2}{z}(1-4z)\right]
+12 {\rm Re}(C_7^{\rm tot} C_{9}^{\rm tot*})\left(1+\frac{2t^2}{z}\right)\;.
\label{bsll_Dz}
\end{eqnarray}
Here $z \equiv q^2/m_b^2$, $t \equiv m_{l}/m_{b}$ and
$\hat{m}_q=m_q/m_b$ for all quarks $q$.
In the framework of SM4, the Wilson coefficients $C_{7}^{\rm tot }$,
$C_{9}^{\rm tot}$ and $C_{10}^{\rm tot}$ are given by
\begin{eqnarray}
C_{7,10}^{\rm tot }& = & C_{7,10}(m_b)\,+\, \frac{V_{t's}^*V_{t'b}}{V_{ts}^*V_{tb}}\,C_{7,10}^{t' }(m_b)\;, \\
C_{9}^{\rm tot }& = & C_9(m_b) \, +\, Y(z)\,+\,
\frac{V_{t's}^*V_{t'b}}{V_{ts}^*V_{tb}}\,C_{9}^{t' }(m_b)\;,
\label{wcsm4}
\end{eqnarray}
where the function $Y(z)$ is given in \cite{Buras:1994dj}.
The measurements of the $B\to X_s\ell^+ \ell^-$ in the two regions, so called
low $q^2$ $(q^2\protect\:\raisebox{-0.5ex}{$\stackrel{\textstyle<}{\sim}$}\:\, 6 GeV^2)$ and high $q^2$
$(q^2\protect\:\raisebox{-0.5ex}{$\stackrel{\textstyle>}{\sim}$}\: \,14 {GeV}^2)$, are complementary as they have different
sensitivities to the short distance physics. Compared to small $q^2$, the rate
in the large $q^2$ region has a smaller renormalization scale dependence and
$m_c$ dependence. Although the rate is smaller at large $q^2$, the experimental
efficiency is better. Large $q^2$ constrains the $X_s$ to have small invariant
mass, $m_{X_s}$, which suppresses the background from $B\to X_c \ell^-\bar{\nu}
\to X_s \ell^+\ell^- \nu \bar{\nu}$. To suppress this background at small $q^2$
region an upper cut on $m_{X_s}$ is required, complicating the theoretical
description due to the dependence of the measured rate on the shape function,
which is absent at large $q^2$. In the low $q^2$ region the
dominant contribution to $B_s\to X_s\ell^+\ell^-$ comes from virtual photon and
much less from $Z$. It is the $Z$ that is very sensitive to $m_{t'}$ as
that amplitude grow with $m^2_{t'}$. The photonic contribution cares only
about the electric charge, modulo logarithmic QCD corrections. For these
reasons we will be using the branching ratio only in the high $q^2$ region to constrain SM4.
The theoretical calculations shown above for the branching ratio of
$B \to X_s\, l^+ \,l^-$ are rather uncertain in the intermediate $q^2$ region
($7$~GeV$^2 < q^2 < 12$~GeV$^2$) owing to the vicinity of
charmed resonances. The predictions are relatively more robust
in the low-$q^2$ ($1 \,{\rm GeV^2} < q^2 < 6\, {\rm GeV^2}$)
and the high-$q^2$ ($14.4\, {\rm GeV^2} < q^2 < m_b^2$) regions.
For $m_{t'}> 300\, {\rm GeV}$, $Br(B\to X_s \,l^+ \, l^-)$ is completely dominated by the
Wilson coefficient $C_{10}^{\rm tot}$. Hence in our numerical analysis, we neglect the small $z$-dependence in
$C_{9}^{\rm tot}$.
\subsection {$B_{q}-\bar B_{q}$ mixing}
Within SM, $B_{q}-\bar B_{q}$ mixing ($q=d,s$) proceeds to an excellent approximation only
through the box diagrams with internal top quark exchanges. In case of
four generations there is an additional contribution to $B_{q}-\bar B_{q}$ mixing
coming from the virtual exchange of the fourth generation up quark $t'$.
The mass difference $\Delta M_q$ in SM4 is given by
\begin{equation}
\Delta M_q = 2|M_{12}|\;,
\end{equation}
where
\begin{eqnarray}
M_{12} &=& \frac{G_F^2 m_W^2}{12\pi^2} m_{B_q} B_{bq} f_{B_q}^2 \Big\{
\eta_t \left( V_{tq} V_{tb}^{*} \right)^2 S_0(x_t) +
\eta_{t'} \left( V_{t' q} V_{t' b}^{*}\right)^2
S_0(x_{t'})
\nonumber \\&&
+ 2\eta_{tt'} \left(V_{tq} V_{tb}^{*} \right) \left( V_{t' q} V_{t' b}^{*}\right) S_0(x_t,x_{t'}) \Big\}~,
\end{eqnarray}
where $x_t=m_t^2/m_W^2$, $x_{t'}=m_{t'}^2/M_W^2$ and
\begin{eqnarray}
S_0(x_t) &=& \frac{4 x_t - 11 x_t^2 + x_t^3}{4 (1-x_t)^2}
-\frac{3}{2} \frac{x_t^3 \mbox{\rm ln} x_t}{(1-x_t)^3}~,\\
S_0(x_{t'}) &=& S_0(x_t \to x_{t'})~, \\
S_0(x_t,x_{t'}) &=& x_t x_{t'} \Bigg\{
\frac{\mbox{\rm ln} x_{t'}}{x_{t'}-x_t} \Bigg[
\frac{1}{4} + \frac{3}{2} \frac{1}{1-x_{t'}} -
\frac{3}{4} \frac{1}{(1-x_{t'})^2} \Bigg]
\nonumber \\
&& - \frac{\mbox{\rm ln} x_t}{x_{t'}-x_t} \Bigg[
\frac{1}{4} + \frac{3}{2} \frac{1}{1-x_t} -
\frac{3}{4} \frac{1}{(1-x_t)^2} \Bigg]
\nonumber \\
&& -\frac{3}{4} \frac{1}{(1-x_t) (1-x_{t^\prime})} \Bigg\}\;.
\end{eqnarray}
Here $\eta_t$ is the QCD correction factor and its value is $0.5765\pm0.0065$ \cite{buras1}.
The QCD correction factor $\eta_{t'}$ is given by \cite{Hattori:1999ap}
\begin{eqnarray}
\eta_{t'} = \Big(\alpha_s(m_t)\Big)^{6/23}
\left( \frac{\alpha_s(m_{b^\prime})}{\alpha_s(m_t)} \right)^{6/21}
\left( \frac{\alpha_s(m_{t^\prime})}{\alpha_s(m_{b^\prime})} \right)^{6/19}\;.
\end{eqnarray}
$\alpha_s(\mu)$ is the running coupling constant at the scale $\mu$ at NLO \cite{Buchalla:1995vs}.
Here we assume $\eta_{t'}=\eta_{tt'}$ for simplicity. The numerical values of
the structure functions $S_0(x_{t'})$, $S_0(x_t,x_{t'})$
and the QCD correction factor $\eta_{t'}$ are given in Table~\ref{tab1} and Table~\ref{tab2} respectively for various $t'$ mass.
\begin{table}[t]
\begin{center}
\begin{tabular}{|c||c|c|}
\hline
$m_{t'}$(GeV) &400 & 600 \\
\hline
$S_0 (x_{t'})$ &9.225 &17.970 \\
\hline
$S_0 (x_t, x_{t'})$ &4.302 &5.225 \\
\hline
\end{tabular}
\caption{The structure functions $S_0(x_{t'})$ and $S_0(x_t,x_{t'})$.}
\label{tab1}
\end{center}
\end{table}
\begin{table}[t]
\begin{center}
\begin{tabular}{|c||c|c|}
\hline
$m_{t'}$(GeV) &400 & 600 \\
\hline
$\eta_{t'}$ &0.522 &0.514 \\
\hline
\end{tabular}
\caption{The QCD correction factor $\eta_{t'}$.}
\label{tab2}
\end{center}
\end{table}
\subsection{Indirect {\it CP} violation in $K_L \to \pi\pi$}
Indirect CP violation in $K_L\to \pi \pi$ is described by the parameter
$\epsilon_K$, the working formula for it is given by \cite{buras2}
\begin{equation}
\epsilon_K = {\rm exp}( i \phi_{\epsilon})\sin\phi_{\epsilon} \Big( {{\rm Im}
{M^{k}_{12}}/\Delta M_k} + \zeta \Big),
\label{epsilon}
\end{equation}
where $\zeta = {{\rm Im}{A_0}\over {\rm Re}{A_0}}$ with $A_0\equiv A\big(K\to (\pi\pi)_{I=0}\big)$
and $\Delta M_K$ denoting the $K_L-K_S$ mass difference. The off-diagonal
element $M_{12}$ in the neutral $K$-meson mass matrix represents $K^0-\bar{K^0}$ mixing and
is given by
\begin{equation}
M^{*}_{12} = {{\langle\bar{K^0}|{\cal{H}}_{eff}(\Delta S = 2)|K^0\rangle} \over 2 m_K}
\end{equation}
The phase $\phi_{\epsilon}$ is given by
\begin{equation}
\phi_{\epsilon}=(43.51 \pm 0.05)^{\circ}
\end{equation}
The second term in eq. \ref{epsilon} constitutes a ${\cal{O}}(5)$\% correction
to $\epsilon_K$. In most of the phenomenological analysis
$\phi_{\epsilon}$ is taken as $\pi/4$ and $\zeta$ is taken as zero. However
$\zeta\not= 0 $ and $\phi_{\epsilon} < \pi/4$ results in a suppression effect in
$\epsilon_k$ relative to the approximate formula with $\zeta = 0 $ and
$\phi_{\epsilon} = \pi/4$. In order to include these corrections we have used
the parametrization
\begin{equation}
\kappa_{\epsilon}=\sqrt{2}\sin{\phi_{\epsilon}}\bar{\kappa_{\epsilon}},
\label{kappa}
\end{equation}
where $\bar{\kappa_{\epsilon}}=0.94 \pm 0.02$ and consequently
${\kappa_{\epsilon}}= 0.92\pm0.02$, $\bar{\kappa_{\epsilon}}$ parameterizing
the effect of $\zeta\ne 0$ \cite{buras2}.
After some calculations it can be shown that \cite{Buras:1997fb}
\begin{eqnarray}
M_{12} &=& {G^2_F\over 12\pi^2} f^2_K B_K m_K M^2_W \Big[{\lambda^{*}_c}^2\eta_c S_0(x_c)+ {\lambda^{*}_t}^2\eta_t S_0(x_t)
+ 2 {\lambda^{*}_c}{\lambda^{*}_t}\eta_{ct} S_0(x_c, x_t)
\nonumber \\ &&
+ {\lambda^{*}_{t'}}^2\eta_{t'} S_0(x_{t'})
+ 2 {\lambda^{*}_c}{\lambda^{*}_{t'}}\eta_{c{t'}} S_0(x_c, x_{t'})
+ 2 {\lambda^{*}_t}{\lambda^{*}_{t'}}\eta_{t{t'}} S_0(x_t, x_{t'}) \Big]\,,
\label{m12}
\end{eqnarray}
where $\lambda_i = \lambda^{*}_{is}\lambda_{id}$ and $x_q=(m^2_q/M^2_W)$
for all quarks $q$.
Inserting (\ref{m12}) and (\ref{kappa}) in (\ref{epsilon}) one finds
\begin{eqnarray}
\epsilon_K &=& {{G^2_F\over 12\pi^2 \sqrt{2} {\Delta M_K}}} {\kappa_{\epsilon}}
f^2_K B_K m_K M^2_W\, {\rm Im}\Big[{\lambda^{\ast}_c}^2\eta_c S_0(x_c)
+ {\lambda^{\ast}_t}^2\eta_t S_0(x_t)
\nonumber \\ && +
2 {\lambda^{\ast}_c}{\lambda^{\ast}_t}\eta_{ct} S_0(x_c, x_t)
+ {\lambda^{\ast}_{t'}}^2\eta_{t'} S_0(x_{t'}) + 2 {\lambda^{\ast}_c}{\lambda^{\ast}_{t'}}\eta_{c{t'}} S_0(x_c, x_{t'})
\nonumber \\&&
+ 2 {\lambda^{\ast}_t}{\lambda^{\ast}_{t'}}\eta_{t{t'}} S_0(x_t, x_{t'}) \Big]\;,
\label{epsilon1}
\end{eqnarray}
where $f_K= 160\, \rm MeV$. The value for $B_K$ has
been taken from Ref. \cite{latticeold}, in a recent analysis
\cite{alv,rbc2} the error has been reduced to $\:\raisebox{-0.5ex}{$\stackrel{\textstyle<}{\sim}$}\: \,4\%$, however, in our analysis
we use the more conservative value mentioned in Table. \ref{tab3} from \cite{latticeold}.
\begin{table}[t]
\begin{center}
\begin{tabular}{|c|c|}
\hline
$B_K = 0.72 \pm 0.05$ \cite{latticeold} & $f_{bs}\sqrt{B_{bs}} =
0.281 \pm 0.021$ GeV \cite{lattice3} \\
$\Delta{M_s} = (17.77 \pm 0.12) ps^{-1}$ \cite{cdf}& $\Delta{M_d} = (0.507 \pm 0.005) ps^{-1}$ \\
$\xi_s = 1.2 \pm 0.06$ \cite{lattice3} & $\gamma = (75.0 \pm 22.0)^{\circ} $ \\
$|\epsilon_k|\times 10^{3} = 2.32 \pm 0.007$ & $\sin 2\beta_{\psi K_s} = 0.672 \pm 0.024$\\
$Br(K^+\to \pi^+\nu\nu) = (0.147^{+0.130}_{-0.089})\times 10^{-9}$ & $Br(B\to X_c \ell \nu) = (10.61 \pm 0.17)\times 10^{-2}$\\
$Br(B\to X_s \gamma) = (3.55 \pm 0.25)\times 10^{-4}$ & $Br(B\to X_s \ell^+ \ell^-) = (0.44 \pm 0.12)\times 10^{-6}$ \\
$R_{bb} = 0.216 \pm 0.001$ & ( High $q^2$ region )\\
$|V_{ub}| = (37.2 \pm 2.7)\times 10^{-4}$ & $|V_{cb}| = (40.8 \pm 0.6)\times 10^{-3} $\\
$\eta_c = 1.51\pm 0.24$ \cite{uli1} & $\eta_t = 0.5765\pm 0.0065$
\cite{buras1}\\
$\eta_{ct} = 0.47 \pm 0.04$ \cite{uli2} & $m_t = 172.5$ GeV\\
$T_4 = 0.11 \pm 0.14$ & \\
\hline
\end{tabular}
\caption{Inputs that we use in order to constrain the SM4 parameter space,
we have considered the 2$\sigma$ range for $V_{ub}$.}
\label{tab3}
\end{center}
\end{table}
\subsection{$K^+\to \pi^+\nu \bar{\nu}$ decay}
The effective Hamiltonian for $K^+\to \pi^+\nu \bar{\nu}$ can be written as
\begin{eqnarray}
{\cal{H}}_{eff} &=& {G_F \over \sqrt{2}}{\alpha\over {2\pi\sin^2{\Theta_w}}}
\sum_{l=e, \mu, \tau}\Big[ V_{cs}^{\ast}V_{cd} X^l_{NL} + V_{ts}^{\ast}V_{td} X(x_t)
\nonumber \\ &&
+ V_{t's}^{\ast}V_{t'd} X(x_{t'}) \Big](\bar{s}d)_{V-A}(\bar{\nu}_l\nu_l)_{V-A}\;.
\end{eqnarray}
First term is the contribution from the charm sector.
The function $X(x)$ is relevant for the top part,
\begin{equation}
X(x) = X_0(x) + {\alpha_s\over 4\pi} X_1(x)\;,
\end{equation}
where $x_q=(m^2_q/M^2_W)$ for all quarks $q$.
Here $X_0(x)$ is the leading contribution given by
\begin{equation}
X_0(x) = {x\over 8}\left[- {2 + x \over 1-x} + {3x-6 \over (1-x)^2} \ln{x}\right]\;,
\end{equation}
and $X_1(x)$ is the QCD correction. The expression for $X_1(x)$ is given in \cite{Buras:1997fb}.
The function $X$ can also be written as
\begin{equation}
X(x_{t/t'}) = \eta_X . X_0(x_{t/t'}), \hskip 20pt \eta_X = 0.994\;.
\end{equation}
Here $\eta_X$ represents the NLO corrections.
The function $X^l_{NL}$ is the function corresponding to $X(x_t)$ in the charm sector. It results
from the NLO calculations and its explicit form is given in \cite{Buchalla:1995vs,Buchalla:1993wq}.
The branching fraction of $K^+\to \pi^+\nu\bar{\nu}$ can be written as follows
\begin{eqnarray}
Br(K^+\to \pi^+\nu\bar{\nu}) &=& \kappa_+ \Big[\left({{\rm Im}\lambda_t \over \lambda^5}X(x_t)
+ {{\rm Im}\lambda_{t'} \over \lambda^5} X(x_{t'}) \right)^2
\nonumber \\ &&
+ \Big({{\rm Re}{\lambda_c}\over \lambda}P_0(X) + {{\rm Re}{\lambda_t} \over \lambda^5}X(x_t)
+ {{\rm Re}{\lambda_{t'}} \over \lambda^5} X(x_{t'})\Big)^2 \Big],
\label{brkpip}
\end{eqnarray}
where
\begin{equation}
\kappa^+ = r_{K+} {{3\alpha^2 \,Br(K^+\to \pi^0 e^+ \nu )}\over {2\pi^2 \sin^4{\Theta_W}}}\lambda^8\;,
\end{equation}
\begin{equation}
P_0(X) = {1\over\lambda^4}\left[{2\over 3} X^e_{NL} + {1\over 3} X^{\tau}_{NL}\right]\;,
\end{equation}
and $r_{K+} = 0.901$ summarizes the isospin breaking corrections in relating
the $K^+\to \pi^+\nu\bar{\nu}$ to the well measured leading decay $K^+\to \pi^0 e^+ \nu $.
\begin{table}[t]
\begin{center}
\begin{tabular}{|c||c|c|c|c|}
\hline
$m_{t'}$ ({\it GeV}) & 300 & 400 & 500 & 600 \\
\hline
$\lambda^s_{t'}$& (0.09 - 2.5) & (0.08 - 1.4)& (0.06 - 0.9) &(0.05 - 0.6) \\
\hline
$\phi_s'$ & 0 $\to$ 80 & 0 $\to$ 80 & 0 $\to$ 80 & 0 $\to$ 80 \\
\hline
\end{tabular}
\caption{Allowed ranges for the parameters, $\lambda^s_{t'}$ ($\times 10^{-2}$)
and phase $\phi_s'$ (in degree) for different masses $m_{t'}$ ( GeV), that has
been obtained from the fitting with the inputs in Table \ref{tab3} and allowed by the
present experimental bound for {\it CP} asymmetry in $B_s\to J/\psi\phi$ \cite{SAGMN08}.}
\label{tab4}
\end{center}
\end{table}
\section{Predictions in the SM4}
\label{results}
\begin{figure}[htbp]
\includegraphics[width=8cm,height=6cm,clip]{scp.eps}%
\hspace{0.2cm}%
\includegraphics[width=8cm,height=6cm,clip]{s_bs.eps}
\caption{(a) Correlation between $S_{\phi K_s}$ and $S_{\psi\phi}$
(left panel) \, and \, (b) Variation of $S_{\psi\phi}$ with the phase $\phi_{s}'$ of
$\lambda^s_{t'}$ (right panel), for $m_{t'}= 300$ (magenta), $400$ (red), $500$ (green) and
$600$ (blue) GeV respectively. The horizontal lines (left panel) represent the experimental
$1\sigma$ range for $S_{\phi K_s}$ whereas the vertical lines (black
1-$\sigma$ and red 2-$\sigma$ ) represent that for $S_{\psi\phi}$; in the right panel
the horizontal lines are for $S_{\psi\phi}$}.
\label{figpap1}
\end{figure}
Fig. \ref{figpap1} (left panel) shows the correlations between the {\it CP}
asymmetries in $B_d\to \phi K_s$ and $B_s\to \psi \phi$ whereas right panel shows the
variation $S_{\psi\phi}$ with the new phase $\phi_{s}'$ \footnote{Soon after we posted
version 1 of our paper , \cite{buras4th} appeared which also discusses about the
phenomenology of SM4. To facilitate direct comparision with that work we are
adding few extra figures in this revised version.}; which has already
been shown in our previous article \cite{SAGMN08} for $m_{t'} = 400, \, 500\,$
and $600\, {\it GeV}$; here, we have also included in the plot
$m_{t'} = 300\, {\it GeV}$. This is to clarify the fact
that the present data on {\it CP} asymmetries tends to favor a fourth family of
quarks with $m_{t'}$ in the range $(400\, - 600\,) {\it GeV}$.
In this article therefore, we will focus mostly on $m_{t'}\approx 400 - 600$
{\it GeV} when we provide numerical results for SM4 for some
interesting observables related to $B$ and $K$ system which could be tested
experimentally.
\subsection{Direct CP asymmetry in $B \to X_s \gamma$}
\label{dacpbsg}
$A_{CP}$ in $B \to X_s \gamma$ is defined as
\begin{equation}
A_{CP}^{B \to X_s \gamma} = \frac{\Gamma(\bar{B}\to X_s \gamma)-\Gamma(B \to X_{\bar{s}} \gamma)}
{\Gamma(\bar{B}\to X_s \gamma)+\Gamma(B \to X_{\bar{s}} \gamma)}
\end{equation}
Within the SM, $A_{CP}^{B \to X_s \gamma}$ is predicted to be less than $1\%$ \cite{Kagan:1998bh,Kiers:2000xy,Soares:1991te}.
The most recent SM prediction is \cite{Hurth:2003dk} (Here we have calculated the errors by adding all
errors given in the mentioned reference in quadrature )
\begin{equation}
A_{CP}^{B \to X_s \gamma}|_{E_{\gamma}>1.6\,{\rm GeV}} = \big(0.44^{+0.24}_{
-0.13} \big)\%\;.
\end{equation}
The current world average of $A_{CP}^{B \to X_s \gamma}$ is $(-1.2\pm 2.8)\%$ \cite{Barberio:2008fa},
which is consistent with zero or a very small direct CP asymmetry as we have in the SM.
The present experimental uncertainty is still an order of magnitude
greater than the theoretical error. However a dramatic improvement in the experimental
sensitivity is possible at the upcoming Super-B factories and sensitivity of
about
$0.4\%-0.5\%$ can be achieved \cite{Browder:2008em}.
As the CP asymmetry within the SM is less than $1\%$,
observation of a sizable CP asymmetry would be a clean signal of new physics.
It is expected that the new physics models with non-standard CP-odd phases
can enhance $A_{CP}^{B \to X_s \gamma}$ and hence we study $A_{CP}^{B \to X_s \gamma}$
within the framework of SM4.
\begin{figure}[t]
\includegraphics[width= 0.60 \linewidth]{fbsg.eps}
\caption{Correlation between CP asymmetry in $B\to X_s \gamma$ and
$S_{\psi\phi}$, the CP asymmetry in $B_s\to J/\psi \phi$; where the red
and blue regions correspond to $m_{t'}$ = 400 and 600 GeV whereas horizontal
lines represent the SM limit for CP asymmetry and the vertical lines
represent the $2\sigma$ limit for CP asymmetry in $B_s\to J/\psi \phi$.}
\label{fig:acp_bsg}
\end{figure}
The general expression for the CP asymmetry in $B \to X_s \gamma$ is \cite{Kiers:2000xy}
\begin{eqnarray}
A_{CP}^{B \to X_s \gamma} & \simeq & \frac{10^{-2}}{|C^{\rm tot}_7(m_b)|^2}
\Big\{ -1.82\;{\rm Im}\left[C_7^{\rm new}\right]
+\,1.72\;{\rm Im}\left[C_8^{\rm new}\right]
-\,4.46\;{\rm Im}\left[C_8^{\rm new}
C_7^{{\rm new} *}\right]
\nonumber \\ &&
+\,3.21\;{\rm Im}\left[\epsilon_{s}\left(
1-2.18 \; C_7^{{\rm new} *}
-0.26\; C_8^{{\rm new} *}\right)\right]\Big\} \;,
\label{acpgen4}
\end{eqnarray}
where
\begin{equation}
\epsilon_s = \frac{V_{us}^* V_{ub}}{V_{ts}^* V_{tb}} \;,
\end{equation}
Here the new physics Wilson coefficients $C_{7,8}^{\rm new}$ are at scale $M_W$.
In SM4,
\begin{equation}
C_{7,8}^{\rm new}=\frac{V_{t's}^*V_{t'b}}{V_{ts}^*V_{tb}}C_{7,8}^{t'}(M_W)\;.
\end{equation}
In the Fig. \ref{fig:acp_bsg} we have shown the correlation between CP
asymmetries in $(B\to X_s \gamma)$ and $B_s\to J/\psi\phi$
($S_{\psi\phi}$). The current $2\sigma$ experimental range for $S_{\psi\phi}$
is given by $[-0.90,-0.17]$ \cite{cdfd0}. The SM value for $A_{CP}(B\to X_s \gamma)$
corresponds to $S_{\psi\phi}\approx 0$ or in other words
$\phi^{t'}_s\approx 0$. It is easy to understand the nature of the plot i.e
decrease of $A_{CP}(B\to X_s \gamma)$
with increase of $S_{\psi\phi}$. From the expression for
$A_{CP}(B\to X_s \gamma)$ (eq.~(\ref{acpgen4})), it is clear that in SM the
only contribution to $A_{CP}$ will come from the first part of the fourth term.
In the presence of new phase and new coupling, the first two terms
and the fourth term will contribute to $A_{CP}$.
Contribution from the first two term is always negative and increases
(mod value) with the new physics coupling ( within the NP region
we are interested) whereas the fourth term is always positive and it has
very small increase with the new physics coupling or phase.
\subsection{{\it CP} asymmetry in $B_s\to X_s\ell\nu$ }
In this section we shall concentrate on semileptonic {\it CP} asymmetry
($A_{SL}$) in $B_s$ system \footnote{We were about to post a short
paper reporting our study of $A_{SL}$ in SM4 when the paper \cite{buras4th} appeared wherein
this topic is also discussed-consequently we are making a very breif addition of this in version
\,2 of our paper. Our results agree with Buras {\rm et. al} \cite{buras4th}.}. In general the
{\it CP} asymmetry in semileptonic $B_s$ decays defined as,
\begin{align}
A_{SL} &= \frac{\Gamma[\bar{B}^{phys}_{s}(t)\to \ell^+ X]-\Gamma[B^{phys}_{s}(t)\to \ell^- X]}{\Gamma[\bar{B}^{phys}_{s}(t)\to \ell^+ X]+\Gamma[B^{phys}_{s}(t)\to \ell^- X]},
\end{align}
depends on the relative phase between the absorptive and dispersive parts of
$B_s-\bar{B_s}$ mixing amplitude \cite{hw},
\begin{align}
A_{SL} &={\it Im}\left(\frac{\Gamma_{12}}{M_{12}}\right) =
\frac{|\Gamma^{s}_{12}|}{|M^{SM}_{12}|} \frac{\sin\phi_s}{|\Delta_s|},
\label{asld1}
\end{align}
with $\phi_s = arg\left(-\frac{M^s_{12}}{\Gamma^s_{12}}\right)$, the relative phase
between $B_s-\bar{B_s}$ mixing and the corresponding $b\to c\bar{c} s$ decays and $|\Delta_s|$
parametrises the NP effect in $M^s_{12}$ \cite{uli_lenz}.
$|\Gamma_{12}/M_{12}|=O(m_b^2/M_W^2)$ suppresses $A_{SL}$ to the percent level,
apart from this there is a GIM suppression factor $m_c^2/m_b^2$ reducing
$A_{SL}$ by another order of magnitude. Because of these suppression factors
it is very small in SM, for $B_s$ system it is ${\cal{O}}(10^{-5})$. The GIM
suppression is lifted if new physics contributes to $\arg (M_{12})$. Therefore
$A_{SL}$ is very sensitive to new {\it CP} phases \cite{run2,llnp}.
The situation where new physics could enhance $A_{SL}$ by a factor
$\cal{O}$(10-100) makes this asymmetry a sensitive probe of new physics.
Recently the search for {\it CP} violation in semileptonic $B_s$ decays
achieved a much more improved sensitivity \cite{d07,cdf7}:
\begin{align}
A_{SL} &=(2.45 \pm 1.96)\times 10^{-2} \hskip 50pt {\rm D0}
\nonumber \\
&=(2.00\pm 2.79)\times 10^{-2} \hskip 50pt {\rm CDF} .
\end{align}
Present world average is given by \cite{hfag},
\begin{align}
A_{SL} &=(-0.37 \pm 0.94)\times 10^{-2} \hskip 50pt {\rm HFAG} .
\end{align}
In near future more precise measurements can exclude SM prediction if it is much
enhanced then the SM prediction.
It is important to note that the scenarios like SM4 can significantly
affect $M_{12}^s$, but not $\Gamma_{12}^s$, which is dominated by the
CKM-favoured $b\to c\ov{c}s$ tree-level decays. The leading contribution to $\Gamma_{12}^s$ was
obtained in \cite{hw,LO}. At present $\Gamma_{12}^s$ is known to
next-to-leading-order (NLO) in both $\ov{\Lambda}/m_b$ \cite{bbd1} and
$\alpha_s(m_b)$ \cite{bbgln1,rome03,bbln}, later in 2006 Nierste and Lenz
\cite{uli_lenz} have improved the NLO calculation
for $\ensuremath{\Delta \Gamma}_s$ and updated the value for $\ensuremath{\Delta \Gamma}_s$.
\begin{figure}[htbp]
\includegraphics[width=8cm,height=6cm,clip]{asl_coup.eps}%
\hspace{0.2cm}%
\includegraphics[width=8cm,height=6cm,clip]{asl_psi.eps}
\caption{Left panel shows the semileptonic {\it CP} asymmetry $A_{SL}$ as a function of
$|\lambda^s_{t'}|$ whereas in the right panel correlation between
$A_{SL}$ and $S_{\psi\phi}$ is shown; red and blue region corresponds to $m_{t'}$ = 400
and 600 $\rm GeV$ respectively, the SM value of $A_{SL}$ (of order $10^{-5}$) is too close to
zero to be visible in the plot whereas the SM value for $S_{\psi\phi}$ is $-0.04$.}
\label{fig_asl}
\end{figure}
In Fig. \ref{fig_asl} the sensitivity of semileptonic {\it CP} asymmetry
to SM4 is shown and we note an enhancement by a factor of 100 from its SM predicion
of order $10^{-5}$. It could have a value $-0.4\%$ and $-0.3\%$ corresponding to maximum
values of $S_{\psi\phi}$ for $m_{t'} = $ 400 and 600 \,{\it GeV} respectively.
\subsection{CP asymmetry in $\bs$}
\label{acpbq}
It is very useful to consider new physics effects in the
observables which are either zero or highly suppressed in the SM as they constitute null test of the SM \cite{soni_gers} .
The reason is that any finite or large measurement of
such an observable may signal the existence of new physics.
The CP asymmetry in $B \to X_s \, l^+ \, l^-$ is one such
observable.
In the SM, the CP asymmetry in
$B \to X_s\, l^+ \,l^-$ is $\sim 10^{-3}$ \cite{Du:1995ez,Ali:1998sf}.
In the SM,
the only source of CP violation
is the unique phase in the CKM
quark mixing matrix.
However in many possible extensions of the SM,
there can be extra phases contributing to the
CP asymmetry. Hence the CP asymmetry in $B \to X_s\, l^+ \,l^-$ is sensitive to SM4.
The CP asymmetry in $\bs$ is defined as
\begin{equation}
A_{\rm CP}(z)=\frac{(dBr/dz)-(d\overline{Br}/dz)}{(dBr/dz)+(d\overline{Br}/dz)}=
\frac{D(z)-\overline{D(z)}}{D(z)+\overline{D(z)}}\;,
\end{equation}
where $Br$ and $\overline{Br}$ represent the branching ratio of $\bar{B} \to X_{s}l^+ l^-$
and its complex conjugate $B \to \bar{X_{s}} l^+ l^- $ respectively. $dBr/dz$ is given in eq.~(\ref{eq:dbr_bsll}).
The Wilson coefficients $C_{7}^{\rm tot }$, $C_{9}^{\rm tot}$, and $C_{10}^{\rm tot}$ can be written as
\begin{eqnarray}
C_{7}^{\rm tot }& = & C_{7}(m_b)\,+\, \lambda_{tt'}^s\,C_{7}^{t' }(m_b)\;, \\
C_{9}^{\rm tot }& = & \xi_1\,+\, \lambda_{tu}^s\xi_2\,+\,
\lambda_{tt'}^s\,C_{9}^{t' }(m_b)\;, \\
C_{10}^{\rm tot } & = & C_{10}(m_b)+ \lambda_{tt'}^s\,C_{10}^{t' }(m_b)\;,
\label{c10sm4}
\end{eqnarray}
where
\begin{equation}
\lambda_{tu}^s=\frac{\lambda_u^s}{\lambda_t^s}=\frac{V_{ub}^*V_{us}}{V_{tb}^*V_{ts}}\;,
\end{equation}
\begin{equation}
\lambda_{tt'}^s=\frac{\lambda_{t'}^s}{\lambda_t^s}=\frac{V_{t'b}^*V_{t's}}{V_{tb}^*V_{ts}}\;,
\end{equation}
so that all three relevant Wilson coefficients are complex in general.
The parameters $\xi_i$ are given by \cite{Buras:1994dj}
\begin{eqnarray}
\xi_1 & = & C_9(m_b) \, +\, 0.138 \,\omega(z)\,+\,g(\hat{m}_{c},z) (3 C_1 + C_2 + 3 C_3 +
C_4 + 3 C_5 + C_6)\nonumber\\&&- \frac{1}{2}g(\hat{m}_{d},z)
(C_3 + 3C_4) - \frac{1}{2}
g(\hat{m}_{b},z)(4 C_3 + 4 C_4 + 3C_5 + C_6) \nonumber\\
& & +\frac{2}{9} (3 C_3 + C_4 + 3C_5 + C_6)\;, \\
\xi_2 & = & [ g(\hat{m}_{c},z)- g(\hat{m}_{u},z)](3C_1 + C_2)\; .
\end{eqnarray}
Here
\begin{eqnarray}
\omega(z) & = & - \frac{2}{9} \pi^2 - \frac{4}{3}\mbox{Li}_2(z) - \frac{2}{3}
\ln z \ln(1-z) - \frac{5+4z}{3(1+2z)}\ln(1-z) \nonumber \\
& & - \frac{2 z (1+z) (1-2z)}{3(1-z)^2 (1+2z)} \ln z + \frac{5+9z-6z^2}{6
(1-z) (1+2z)} \; ,
\end{eqnarray}
with
\begin{equation}
\mbox{Li}_2(z)\,=\,-\int_0^z dt\, \frac{{\rm ln}(1-t)}{t}\;.
\end{equation}
The function $g(\hat m,z)$ represents the one loop
corrections to the four-quark operators $O_1-O_6$ and is given by \cite{Buras:1994dj}
\begin{eqnarray}
g(\hat m, z) & = & -\frac{8}{9}\ln\frac{m_b}{\mu_b} - \frac{8}{9}\ln \hat m +
\frac{8}{27} + \frac{4}{9} x \\
& & - \frac{2}{9} (2+x) |1-x|^{1/2}
\left\{\begin{array}{ll}
\left( \ln\left| \frac{\sqrt{1-x} + 1}{\sqrt{1-x} - 1}\right| - i\pi \right), &
\mbox{for } x \equiv \frac{4\hat m^2}{z} < 1 \nonumber \\
2 \arctan \frac{1}{\sqrt{x-1}}, & \mbox{for } x \equiv \frac
{4\hat m^2}{z} > 1,
\end{array}
\right.
\end{eqnarray}
For light quarks, we have $\hat{m}_{u}\simeq \hat{m}_{d}\simeq0$.
In this limit,
\begin{equation}
g(0, z) = \frac{8}{27} -\frac{8}{9} \ln\frac{m_b}{\mu_b} - \frac{4}{9} \ln
z + \frac{4}{9} i\pi\;.
\end{equation}
We compute $g(\hat{m},z)$ at $\mu_b = m_b$.
$d\overline{Br}/dz$ can be obtained from $dBr/dz$
by making the following replacements:
\begin{eqnarray}
C_{7}^{\rm tot } =C_{7}(m_b)\,+\, \lambda_{tt'}^s\,C_{7}^{t' }(m_b)
& \to & \overline {C_{7}^{\rm tot }}=C_{7}(m_b)\,+\,
\lambda_{tt'}^{s*}\,C_{7}^{t' }(m_b)\;, \\
C_{9}^{\rm tot }=\xi_1\,+\, \lambda_{tu}^s\xi_2\,+\, \lambda_{tt'}^s\,C_{9}^{t' }(m_b)
& \to &
\overline {C_{9}^{\rm tot }}=\xi_1\,+\, \lambda_{tu}^{s*}\xi_2\,+\, \lambda_{tt'}^{s*}\,C_{9}^{t' }(m_b)\;, \\
C_{10}^{\rm tot }=C_{10}(m_b)+ \lambda_{tt'}^s\,C_{10}^{t' }(m_b)
& \to &
\overline {C_{10}^{\rm tot }}=C_{10}(m_b)+ \lambda_{tt'}^{s*}\,C_{10}^{t' }(m_b)\;.
\end{eqnarray}
Then we get \cite{alok}
\begin{eqnarray}
D(z) - \overline{D(z)}
&=& 2 \left( 1 + \frac{2 t^2}{z} \right)
\bigg[ {\rm Im}(\lambda_{tu}^s) \left\{ 2(1+2z) {\rm Im}(\xi_1 \xi_2^\ast) - 12 C_7 {\rm Im}(\xi_2) \right\} \nonumber \\
&& \phantom{2 \left( 1 + \frac{2 t^2}{z} \right) }
+ X_{im} \left\{ (1+2 z) C_9^{t'} + 6 C_{7}^{t'} \right\} \bigg] \; ,
\label{acp_num} \\
D(z) + \overline{D(z)}
&=& \left( 1 + \frac{2 t^2}{z} \right)
\Bigl[ (1 + 2z) \left\{ B_1 + 2 C_9^{t'} \left( | \lambda_{t t'}^s |^2 C_9^{t'} + X_{re} \right) \right\} \Bigr. \nonumber \\
&& \Bigl. + 12 \left\{ B_2 + 2 C_7 C_9^{t'} {\rm Re}(\lambda_{t t'}^s)
+ C_{7}^{t'} \left( 2 | \lambda_{t t'}^s |^2 C_9^{t'} + X_{re} \right) \right\}\Bigr] \nonumber \\
&&+ 8 \left( 1 + \frac{2 t^2}{z} \right) \left( 1 + \frac{2}{z} \right) |C_7^{\rm tot}|^2 \nonumber \\
&&+2 \left[ \left( 1 + 2 z \right) + \frac{2 t^2}{z} \left( 1 - 4 z \right) \right] |C_{10}^{\rm tot}|^2 \; ,
\label{acp_den}
\end{eqnarray}
where
\begin{eqnarray}
X_{re} &=& 2 \left\{ {\rm Re}\left( \lambda_{t t'}^s \right) {\rm Re}\left( \xi_1 \right)
+ {\rm Re}\left( \lambda_{t t'}^s {\lambda_{tu}^s}^\ast \right) {\rm Re}\left( \xi_2 \right)\right\} \; , \\
X_{im} &=& 2 \left\{ {\rm Im}\left( \lambda_{t t'}^s \right) {\rm Im}\left( \xi_1 \right)
+ {\rm Im}\left( \lambda_{t t'}^s {\lambda_{tu}^s}^\ast \right) {\rm Im}\left( \xi_2 \right)\right\} \; ,
\label{xim}\\
B_1 &=& 2 \left\{ |\xi_1|^2 + {|\lambda_{tu}^s \xi_2|}^2 + 2 {\rm Re}\left( \lambda_{tu}^s \right) {\rm Re}\left( \xi_1 \xi_2^\ast \right) \right\} \; , \\
B_2 &=& 2 C_{7} \left\{ {\rm Re}(\xi_1) + {\rm Re}(\lambda_{tu}^s) {\rm Re}(\xi_2) \right\} \; , \\
|C_{10}^{\rm tot}|^2
&=& {\left( C_{10} \right)}^2
+ |\lambda_{t t'}^s |^2 {\left( C_{10}^{t'} \right)}^2 + 2 C_{10} C_{10}^{t'} {\rm Re}\left( \lambda_{t t'}^s \right) \; , \\
|C_{7}^{\rm tot}|^2
&=&{\left( C_{7} \right)}^2
+ |\lambda_{t t'}^s |^2 { \left( C_{7}^{t'} \right)}^2
+ 2 C_{7} C_{7}^{t'} {\rm Re}\left( \lambda_{t t'}^s \right) \; .
\end{eqnarray}
\begin{figure}
\includegraphics[width= 0.60 \linewidth]{fbslacp.eps}
\caption{Correlation between CP asymmetry in $B\to X_s \ell^+\ell^-$
(high-$q^2$ region) and $S_{\psi\phi}$. In the SM both the values are very
small and in the plot they correspond to the point $[-0.04,0.0]$ .
The red and blue regions correspond to $m_{t'}$ = 400 and 600 GeV whereas the
vertical lines represent $2\sigma$ experimental range for $S_{\psi\phi}$. }
\label{fig:acp_bsll}
\end{figure}
From the expression for $g(\hat{m}, z)$ it is clear that the strong phase in
$g(\hat{m}_{u/d}, z)$ and $g(\hat{m}_{c}, z)$ is responsible for CP
asymmetry in $B\to X_s \ell^+\ell^-$ within the SM. $g(\hat{m}_{u/d}, z)$ is
complex in both high and low-$q^2$ region whereas $g(\hat{m}_{c}, z)$ is
complex only in the high-$q^2$ region. On the other hand $g(\hat{m}_{b}, z)$
is always real. The SM CP asymmetry in high-$q^2$ region is almost zero
since Im($\xi_2$) is very small, almost one order in magnitude relative to
its value in low-$q^2$ region, due to the relative cancellations of strong
phases in $\xi_2$. In the presence of new physics $\xi_2$ is unaffected but
$\xi_1$ increases with the new physics coupling . On the other hand we have
contribution from the second term of eq.~(\ref{acp_den}) as a whole the CP
asymmetry will increase with $S_{\psi\phi}$, as shown in the figure \ref{fig:acp_bsll}.
\subsection{FB asymmetry in $B \to X_s\, l^+ \, l^-$}
The quark level transition $b \to s\, l^+\, l^-$ is forbidden at the
tree level within the SM and can occur only via one or more loops. Hence it has the potential
to test higher order corrections to the SM and also to constrain many of its
possible extensions. It gives rise to the inclusive decay $B \to X_s\, l^+ \, l^-$
which has been experimentally observed \cite{Aubert:2004it,Iwasaki:2005sy}
with a branching ratio close to its
SM predictions, $Br(B\to X_s \ell^+ \ell^-)(1 < q^2 < 6$ ${\rm GeV}^2)$=
$(1.63 \pm 0.20)\times 10^{-6}$ and $Br(B\to X_s \ell^+ \ell^-)(q^2 > 14.4$
${\rm GeV}^2)$=
$(3.84 \pm 0.75) \times 10^{-7}$ \cite{Ghinculov:2003bx,Ali:2002jg,Ali:2002ik}.
Apart from the branching ratio of semi-leptonic decay, there are other
observables
which are sensitive to new physics contribution to $b \to s$ transition.
One such observable is forward-backward (FB) asymmetry of leptons in
$B \to X_s\, l^+ \, l^-$. The FB asymmetry of leptons in
$B(p_b) \to X_s(p_s)\, l^+(p_{l^+})\, l^-(p_{l^-})$ is obtained by
integrating the double differential branching ratio
($d^2 Br/dz dcos\theta$)
with respect to the angular variable $cos\theta$ \cite{Ali:1991is}
\begin{eqnarray}
A_{FB}(z)= \frac{\int_0^{1}dcos\theta \frac{d^2Br}{dz \
dcos\theta}-\int_{-1}^{0}dcos\theta \frac{d^2Br}{dz\ dcos\theta}}
{\int_0^{1}dcos\theta \frac{d^2Br}{dz\ dcos\theta}+\int_{-1}^{0}dcos\theta \frac{d^2Br}{dz
\ d\cos\theta}}\;,
\end{eqnarray}
where $z \equiv q^2/m_{b}^{2}\equiv (p_{l^+}+p_{l^-})^2/m_{b}^{2} $ and
$\theta$
is the angle between the momentum of the $B$-meson (or the outgoing $s$-quark)
and that of $l^+$ in the center of mass frame of the dileptons $l^+ l^-$.
FB asymmetry
measures the difference in the right-chiral and left-chiral couplings of the
leptonic current. FB asymmetry is driven by the top quark \cite{Ali:1991is}
and hence it is sensitive to the fourth generation up type quark $t'$.
\begin{figure}[t]
\includegraphics[width= 0.60 \linewidth]{fbasy.eps}
\caption{Forward backward (FB) asymmetry in $B\to X_s \ell^+\ell^-$ has been
plotted with $z={q^2 \over m^2_b}$ , the red and blue regions correspond to
$m_{t'}$ = 400 and 600 GeV respectively and the black thick line represents
that for SM and the green line represents the zero of $A_{FB}$.}
\label{fig:fb}
\end{figure}
Within the framework of SM4, the FB asymmetry in $B \to X_s\, l^+ \, l^-$ is given by
\begin{equation}
A_{FB}(z) =- 3 \left(1-\frac{4t^2}{z}\right)^{1/2}\,\frac {E(z)}{D(z)}\;,
\end{equation}
where
\begin{equation}
E(z)= {\rm Re}(C_9^{\rm tot}C_{10}^{\rm tot*})z
+ 2{\rm Re}(C_7^{\rm tot}C_{10}^{\rm tot*})\;,
\end{equation}
and $D(z)$ is given in eq.~(\ref{bsll_Dz}).
The FB asymmetry in $B \to X_s\, l^+ \, l^-$ becomes zero for a particular
value of the dilepton invariant mass.
Within SM, the zero of $A_{FB}(q^2)$ appears in the low $q^2$ region, sufficiently away
from the charm resonance region to allow the precise prediction of
its position in perturbation theory. The value of the zero of the FB asymmetry
is one of the most precisely calculated observables in flavor physics
with a theoretical error of order $5\%$. The NNLO prediction for the
zero of FB asymmetry is with $m_b=4.8$ $\rm GeV$\cite{Huber:2007vv}
\begin{equation}
(q^2)_0= (3.5 \pm 0.12)\, {\rm GeV}^2 \,.
\end{equation}
This zero varies from model to model.
Thus it can serve as an important probe
to test SM4 experimentally.
As far as experiments are concerned, this quantity has not been
measured as yet. But estimates show that a precision of about $5\%$ could
be obtained at Super-B factories \cite{Browder:2008em}.
From Fig. \ref{fig:fb} one can see that the value of $z = \frac{q^2}{m^2_b}$,
for which $A_{FB}(z)$-asymmetry is zero, could be shifted to a lower value
than its SM value (although it is consistent with the SM within the
uncertainty). For $m_{t'}$ = 400 and 600 GeV, one could have the value for
$(q^2)_0$ ranging between $(3.09 \to 3.57)$ ${\rm GeV}^2$ for
$m_b = 4.8$ $\rm GeV$.
\subsection{FB asymmetry in $B\to K^* \ell^+\ell^-$}
The quark level transition $b \to s \ell^+\ell^-$ is responsible for the exclusive decay $B\to K^* \ell^+\ell^-$.
The exclusive decay $B\to K^* \ell^+\ell^-$ has relatively large theoretical errors as compared to the inclusive decay $b \to s \ell^+\ell^-$
due to the uncertainty in the determination of the hadronic form factors appearing in the transition amplitude $B\to K^*$. However the exclusive decays are more readily accessible in the experiments. Therefore despite the large theoretical errors, the precise measurement of the exclusive decays could provide hints for possible deviations from the SM.
The decay $B\to K^* \ell^+\ell^-$ has been observed at the Babar and Belle experiments \cite{babar-03,babar-06,belle-03}. Within the present experimental
and theoretical precisions, the measured branching ratio is in agreement with the SM prediction \cite{Ali:2002jg,lunghi}.
However the measurements of the invariant dilepton mass is sparse. It is expected that the precise measurements
of the Dalitz distributions in $B\to K^* \ell^+\ell^-$
is possible at the LHCb and at the Super B factories.
In particular, the measurement of FB asymmetry in $B\to K^* \ell^+\ell^-$ is of great importance. This is because the uncertainty due to the form factors is minimal \cite{ali-00}.
\begin{figure}[t]
\includegraphics[width=8cm,height=6cm,clip]{fbxl.eps}%
\hspace{0.2cm}%
\includegraphics[width=8cm,height=6cm,clip]{fbxh.eps}
\caption{FB asymmetry in $B\to K^* \ell^+\ell^-$
in the low-$q^2$ (left panel) and the high-$q^2$ region (right panel). The
red and the blue regions correspond to
$m_{t'}$ = 400 and 600 GeV respectively and the grey region represents the SM
prediction.}
\label{fb_excld}
\end{figure}
Within the SM4, the normalized FB-asymmetry in $B\to K^* \ell^+\ell^-$ is
given by \cite{ali-00}
\begin{eqnarray}
A_{FB}(z) &=&- \frac{G^2_F\alpha^2m^4_B}{2^8 \pi^5 (d\Gamma/dz)}|V^*_{ts}V_{tb}|^2 z \lambda \left(1-\frac{4\hat{m}^2_l}{z}\right)
\times \Bigg[{\rm Re}(C^{\rm tot}_9 C^{\rm tot *}_{10})VA_1
\nonumber\\ &&
+ \frac{\hat{m}_b}{z} {\rm Re}(C^{\rm tot}_7 C^{\rm tot *}_{10})\Big\{VT_2(1-\hat{m}_{K^*})+A_1T_1(1+\hat{m}_{K^*})\Big\} \Bigg]\;,
\end{eqnarray}
where
\begin{eqnarray}
\lambda &=&1+\hat{m}^4_{K^*} + z^2-2 z - 2\hat{m}^2_{K^*} (1+z) \; ,\\
z&=&\frac{q^2}{m^2_B}\;, \\
\hat{m}_{K^*} &=& \frac{m_{K^*}}{m_B}\;.
\end{eqnarray}
Here $(d\Gamma/dz)$ is the $B\to K^* \ell^+\ell^-$ differential decay
distributions and its detailed expression can be seen from Ref. \cite{ali-00}.
The form factors $A_i, \, V,\, T_i$ are calculated in the light cone QCD approach and their values are given in \cite{ali-00}.
The zero of FB-asymmetry is determined by the equation,
\begin{equation}
Re\Big(C^{eff}_9({z}_0)\Big)= - 2 \frac{\hat{m}_b}{{z}_0} C^{eff}_7
\frac{1 - {z}_0}{1 + m^2_{K^*}-{z}_0},
\end{equation}
where ${z}_0$ corresponds to the value of ${z}$ for which FB-asymmetry is
zero, within SM the value of $(q^2)_0$ for $m_b= 4.8$ $\rm GeV$ is given by
\cite{ali-00}
\begin{equation}
(q^2)_0 = z_0 M^2_B = 2.88^{+0.44}_{-0.28} \hskip 5pt {\rm GeV}^2.
\end{equation}
\begin{table}[t]
\begin{center}
\begin{tabular}{|c|c|c|c|c|}
\hline
$q^2 ({\rm GeV}^2/c^2)$ & \multicolumn{1}{c}{} & \multicolumn{1}{c}{} {$A_{FB}$}&
\multicolumn{1}{c}{} & \\
\cline{2-5}
& exp & SM & $m_t'=400\, {\rm GeV}$ & $m_t'=600\, {\rm GeV}$ \\
\hline
$0.6 - 1.0$ & $0.47^{+0.26}_{-0.33}$ & $(-0.18 \to -0.19)$ &
$(-0.13 \to -0.19)$ & $(-0.08 \to -0.19 )$ \\
\hline
$1.0 - 6.0$ & $0.26^{+0.28}_{-0.31}$ & $(-0.2 \to 0.2)$ &
$(-0.2 \to 0.2)$ & $(-0.2 \to 0.2 )$ \\
\hline
$6.0 - 8.0$ & $0.45^{+0.21}_{-0.26}$ & $(0.19 \to 0.30)$ &
$(0.17 \to 0.28)$ & $(0.11 \to 0.30 )$ \\
\hline
$16.5 - 18.0$ & $0.66^{+0.12}_{-0.16}$ & $(0.28 \to 0.49)$
& $(0.25\to 0.45)$ & $0.15 \to 0.47$ \\
\cline{3-5}
$18.0 - 19.5$ & For $ (q^2 > 16)$ & $(0.003 \to 0.30)$ &
$(0.003\to 0.27)$ & $0.003 \to 0.28$ \\
\hline
\end{tabular}
\caption{Values of FB-asymmetry in different $q^2$ region.}
\label{tabfb}
\end{center}
\end{table}
From the left panel of Fig. \ref{fb_excld}, it is clear that within the
uncertainty, the zero of the FB asymmetry in the SM4
is consistent with the SM prediction.
In Table \ref{tabfb} we have made a comparative study between SM, SM4 and
experimental ranges for $A_{FB}(q^2)$ in different $q^2$ region and one could
see that the SM and SM4 predictions are within the present experimental bound.
One interesting feature of data is that for low $q^2$ (first two bins), the
central value (with appreciable errors) of $A_{FB}$ is positive whereas SM
predicts negative $A_{FB}$ for these bins. Note also that there are deviations between
SM and SM4 predicted FB-asymmetries in some regions of $q^2$, for example
$q^2$ (${ \rm GeV}^2$) with values in between $(0.6\to 1.0)$, $(6.0 \to 8.0)$
and $(16.5 \to 18.0)$ the lower limit of SM4 predicted values are lower in
magnitude than that for SM predictions; these differences are more prominent for
$m_{t'} = 600$ GeV (see Table. \ref{tabfb}).
\subsection {$B_s \to l^+ l^-$ decay}
The purely leptonic decays $B_s \to l^+ l^-$, where $l=e,\,\mu,\,\tau$, are chirally suppressed within the SM and hence
have appreciably smaller branching ratios as compared to that of the semi-
leptonic decays. The helicity suppression is more dominant in the case of
$B_s \to e^+ e^-$ and $B_s \to \mu^+ \mu^-$ which have branching ratio of
$\sim\,(7.7\pm 0.74)\times 10^{-14}$ and $\sim\,(3.35 \pm 0.32)\times 10^{-9}$ respectively \cite{buras03}, within the SM.
However the suppression is evaded to some extent in the case of
$B_s \to \tau^+ \tau^-$ due to the large $m_{\tau}$, which has a branching ratio of $\sim 10^{-7}$. These decays are yet to be observed experimentally. The present upper bound on $B_s \to e^+ e^-$ and $B_s \to \mu^+ \mu^-$ are \cite{Barberio:2008fa}
\begin{eqnarray}
Br(B_s \to e^+ e^-) &<& 0.28 \times 10^{-6}\;,
\nonumber \\
Br(B_s \to \mu^+ \mu^-) &<& 3.60 \times 10^{-8}\;.
\end{eqnarray}
As far as the $\tau$ channel is concerned, the current experimental information is rather poor. Using the LEP data on $B \to \tau \nu$ decays, the indirect bound on $Br(B_s \to \tau^+ \tau^-)$ is obtained to be \cite{Grossman:1996qj}
\begin{equation}
Br(B_s \to \tau^+ \tau^-) < 5\% \;.
\end{equation}
Though the decay $B_s \to \tau^+ \tau^-$ has relatively larger branching ratio
compared to $B_s \to e^+ e^-$ and $B_s \to \mu^+ \mu^-$, its
observation will also be extremely difficult as the reconstruction of $\tau$
is a very challenging task. However, the upcoming experiments at the LHC can
reach the SM sensitivity of $B_s \to \mu^+ \mu^-$ and hence it can serve as an
important probe to test the SM and constrain many new physics models. The LHCb
will be able to probe the SM predictions for $B_s \to \mu^+ \mu^-$ at
$3 \sigma$ with $2\,fb^{-1}$ of data \cite{Lenzi:2007nq} whereas the ATLAS and
CMS will be able to reconstruct the $B_s \to \mu^+ \mu^-$ signal at
$3 \sigma$ with $30\,fb^{-1}$ of data collection \cite{Smizanska:2008qm}.
Here we study the decay $B_s \to \mu^+ \mu^-$ and $B_s \to \tau^+ \tau^-$ in
the context of SM4. Within the SM4, the branching ratio of $B_s \to l^+ l^-$
is given by
\begin{equation}
Br(B_s \to l^+ l^-) = \frac{G^2_F \alpha^2 m_{B_s }m_l^2 f_{B_s}^2 \tau_{B_s}}{16 \pi^3}
|V_{tb}^{}V_{ts}^{\ast}|^2 \sqrt{1 - \frac{4 m_l^2}{m_{B_s}^2}}\, \Big| C_{10}^{\rm tot}\Big|^2\;.
\end{equation}
The branching ratio of $B_s \to l^+ l^-$ can be predicted with higher accuracy
by correlating it with the $B_{s}-\bar B_{s}$ mixing and then considerable
uncertainty due to mixing angle and $f_{B_s}$ gets removed. We have
\begin{equation}
Br(B_s \to l^+ l^-) = \frac{3\alpha^2\tau_{B_s}m_l^2}{8\pi B_{bs}m_W^2}\, \sqrt{1 - \frac{4 m_l^2}{m_{B_s}^2}}\,
\frac{\Big| C_{10}^{\rm tot}\Big|^2}{|\Delta'|}\Delta M_s\;,
\label{bsmu1}
\end{equation}
where $B_{bs}$ is the ``Bag-parameter" for $B_s$ mesons for which lattice
result is given by \cite{latticebag},
\begin{equation}
B_{bs} = 1.33 \pm 0.06,
\end{equation}
however, in order to be conservative we use the value $1.33 \pm 0.15$ .
In eq. \ref{bsmu1} the parameter $\Delta'$ is defined as,
\begin{equation}
\Delta' = \Big[\eta_t S_0(x_t) + \eta_{t'} \frac{\left( V_{t' s} V_{t' b}^{*}\right)^2}{\left( V_{ts} V_{tb}^{*} \right)^2}S_0(x_{t'}) + 2\eta_{tt'} \frac{\left( V_{t' s} V_{t' b}^{*}\right)}{\left( V_{ts} V_{tb}^{*} \right)}S_0(x_t,x_{t'})\Big]\;.
\label{del2}
\end{equation}
\begin{figure}[t]
\includegraphics[width=8cm,height=6cm,clip]{brbsmu.eps}%
\hspace{0.2cm}%
\includegraphics[width=8cm,height=6cm,clip]{brbstau.eps}
\caption{Correlation between branching fraction in $B_s\to \mu^+\mu^- $
(left panel) and $B_s\to \tau^+\tau^- $ (right panel) with $S_{\psi\phi}$,
where the red and blue regions correspond to $m_{t'}$ = 400 and 600 GeV
respectively, the horizontal lines represent the SM limit for
$Br(B_s \to \ell^+\ell^-)$ whereas the vertical lines represent the
$2\sigma$ experimental range for $S_{\psi\phi}$.}
\label{fig_brbsmu}
\end{figure}
In fig. \ref{fig_brbsmu} we have shown the correlation between the branching
fraction $Br(B_s\to \ell^+\ell^-)$ and {\it CP} asymmetry in $B_s\to \psi\phi$,
it is clear that there are possibilities for appreciably different predictions
in SM4 compared to SM, enhanced or diminished by a factor of ${\cal{O}}(3)$.
Note also that enhanced branching fractions correspond to a large {\it CP}
asymmetry in $B_s\to \psi\phi$ and smaller branching fractions correspond to
smaller asymmetry. The corresponding upper limit on the branching fractions
are given by,
\begin{align}
Br(B_s\to \mu^+\mu^-) & <\, 8.0\times 10^{-9} \hskip 20pt m_{t'} = 400\,
{\it GeV}, \nonumber \\
& < 1.2\, \times 10^{-8}, \hskip 20pt m_{t'} = 600\,
{\it GeV},\nonumber \\
Br(B_s\to \tau^+\tau^-) & <\, 1.8\times 10^{-6} \hskip 20pt m_{t'} = 400\,
{\it GeV}, \nonumber \\
& < 2.4\, \times 10^{-6}, \hskip 20pt m_{t'} = 600\,
{\it GeV}.
\end{align}
However, when $S_{\psi\phi}$ is close to its SM value i.e when the {\it CP}
violating phase, $\phi^s_{t'}$, of $V_{t's}$ is close to zero, the branching
fractions reduce from their SM value since $|C_{10}^{\rm tot}|$ and $\delta'$ in
eq. \ref{del2} are reduced from its SM value due to destructive interference with
SM4 counterpart.
\subsection{Branching fraction $B\to X_s \nu \bar{\nu}$}
The decays $B\to X_s \nu \bar{\nu}$ are the theoretically cleanest decays in
the field of rare $B$-decays. They are dominated by the same $Z^0$-penguin and
box diagrams involving top quark exchanges which we encounter in the
case of $K_L\to \pi^0 \nu \bar{\nu}$ ,
since the change of the external quark flavors has no impact on the
$m_{t/t'}$ dependence, the later is fully described by the function
$X(x_{t/t'})$ which includes the NLO corrections. The charm contribution is
negligible here. The effective Hamiltonian for the decay
$B\to X_s\nu\bar{\nu}$ is given by
\begin{equation}
{\cal{H}}_{eff} = {G_F \over \sqrt{2}}{\alpha\over {2\pi\sin^2{\Theta_w}}}\left( V^{\ast}_{tb}V_{ts} X(x_t) + V^{\ast}_{t's}V_{t'd} X(x_{t'}) \right)(\bar{b}s)_{V-A}(\bar{\nu}\nu)_{V-A} + h.c.
\end{equation}
with
\begin{eqnarray}
X(x)={x\over8}\Big[{{2+x}\over{x-1}}+{{3 x-6}\over(x-1)^2}\ln x\Big]
\end{eqnarray}
\begin{figure}
\includegraphics[width= 0.60 \linewidth]{brbsnu.eps}
\caption{Correlation between branching fraction in $B\to X_s \nu \bar{\nu}$
and $S_{\psi\phi}$, where the red and blue regions correspond to $m_{t'}$ = 400
and 600 GeV respectively, the horizontal lines represent the SM limit
for $Br(B\to X_s \nu \bar{\nu})$ whereas the vertical lines represent the
$2\sigma$ experimental range for $S_{\psi\phi}$.}
\label{fig_brbsnu}
\end{figure}
The calculation of the branching fractions for $B\to X_s \nu \bar{\nu}$ can be
done in the spectator model corrected for short distance QCD effects.
Normalizing it to $Br\big(B\to X_c \nu \bar{\nu}\big)$ and summing over
three neutrino flavors one finds \cite{Buras:1997fb,grossman}
\begin{eqnarray}
Br\big(B\to X_s \nu \bar{\nu}\big)\over {Br\big(B\to X_c e
\bar{\nu}\big)} &=& {3 \alpha^2 \over 4\pi^2 \sin^4\Theta_W} {\bar{\eta}\over
f(z)\kappa(z)}{1\over |V_{cb}|^2} \Big|\lambda_t X(x_t) +
\lambda_{t'} X(x_{t'})\Big|^2 \nonumber \\
{} &=& {{\tilde{C}^2 \bar{\eta}}\over{|V_{cb}|^2 f(z) \kappa(z)}},
\label{brbsnu1}
\end{eqnarray}
where
\begin{equation}
{\tilde{C}}^2=({\tilde{C}}^{SM})^2\Big|1+ {V^{\ast}_{t'b}V_{t's}\over
V^{\ast}_{tb}V_{ts}}{X_0(x_{t'})\over X_0(x_{t})}\Big|^2,
\label{brbsnu2}
\end{equation}
with
\begin{equation}
({\tilde{C}}^{SM})^2={{\alpha}^2\over{2 \pi^2\sin^4\Theta_W}}
\Big|V^{\ast}_{tb}V_{ts}X_0(x_{t})\Big|^2.
\label{brbsnu3}
\end{equation}
The factor $\bar\eta$ represents the QCD correction to the matrix element of
the $b\to s\nu{\bar{\nu}}$ transition due to virtual and bremsstrahlung
contributions and is given by the well known expression
\begin{equation}
{\bar\eta} = \kappa(0)= 1 + {2\alpha_s(m_b) \over 3\pi}\Big({25\over 4} - \pi^2\Big) \approx 0.83.
\end{equation}
The SM4 predicted branching fraction $Br(B\to X_s \nu \bar{\nu})$ could be
sufficiently larger than its SM limit, $(3.66 \to 4.01)\times 10^{-5}$
\cite{Buras:1997fb} within the uncertainties, for values of $S_{\psi\phi}$
sufficiently away from its SM predictions. We are constraining
$\lambda^{s}_t = V_{tb} V^{\ast}_{ts}$ using CKM4 unitarity with $\lambda^{s}_{t'} =
V_{t'b} V^{\ast}_{t's}$ as free parameter, with the change
of phase and amplitude of $\lambda^{s}_{t'}$, $|\lambda^s_t|$
increases from its SM value resulting an overall enhancement of $Br(B\to X_s \nu \bar{\nu})$ from
its SM prediction.
For values of $\phi^s_{t'}$ close to $80^{\circ}$, the terms within modulus in
eq. \ref{brbsnu2} and eq. \ref{brbsnu3} have their maximum values and so the
branching fraction is sufficiently larger than its SM prediction and
reach its maximum value $4.8\times 10^{-5}$. In passing, we note incidently that the upper
limit that we have obtained for SM4 is consistent with that obtained in Ref. \cite{burasewerth},
in models with minimal flavor violation (MFV), and with the present
experimental bound $6.4\times 10^{-4}$ \cite{expbsnu}.
\subsection{Branching fraction $K^+ \to \pi^+ \nu \bar{\nu}$}
Although we have taken branching fraction for $K^+ \to \pi^+ \nu \bar{\nu}$ as
a constrain to fit $V_{CKM4}$, in Fig. (\ref{fig_ktopip}) we show the effect of SM4;
note that in the left panel only the $1\sigma$ range
for the branching fraction using the constraints given in the Table. \ref{tab3}
(except $Br(K^+ \to \pi^+ \nu \bar{\nu})$) is shown\footnote{Right panel is added in our version
2 to facilitate direct comparision with \cite{buras4th}.}.
\begin{figure}
\includegraphics[width=8cm,height=6cm,clip]{brkpnu.eps}
\hspace{0.2cm}%
\includegraphics[width=8cm,height=6cm,clip]{brknufull.eps}
\caption{Plot between the branching fraction of $K^+ \to \pi^+ \nu \bar{\nu}$
with $\phi^{ds}_{t'} = \phi^{d}_{t'}-\phi^{s}_{t'}$ bounded by the present experimental limit,
red and blue region corresponds to $m_{t'}$ = 400 and 600 $\rm GeV$
respectively, the green and black horizontal lines represent $1\sigma$ limit
for SM and experimental value respectively. Left panel shows only 1\,$\sigma$ range expected
in SM4; full range is shown in the right panel.}
\label{fig_ktopip}
\end{figure}
From Fig. \ref{fig_ktopip} one could see that the
$Br(K^+ \to \pi^+ \nu \bar{\nu})$ could be enhanced to its present experimental upper limit.
In order to understand the nature of the plot one needs to concentrate on eq. (\ref{brkpip}),
and it is important to note that
$Br(K^+ \to \pi^+ \nu \bar{\nu})$ is dominated by the second term of
the expression i.e the term proportional to $Re(\lambda_q)$ it should also be
noted that the SM and SM4 part for each term has a relative sign difference.
When $\phi^{ds}_{t'}$ is negative (i.e when $\phi^{d}_{t'}$ has values in
between $(0-80)^{\circ}$) and $\phi^{ds}_{t'} > 270^{\circ}$ the
branching fraction will decrease because of the destructive interference
between SM and SM4 part in the second term of eq. (\ref{brkpip}). For
$\phi^{ds}_{t'}$ in between $(90 - 180)^{\circ}$ the branching fraction have
values above the SM value it is due to constructive interference between SM
and SM4 in the second term of eq. (\ref{brkpip}).
Present NNLO predictions for branching fraction for
$K^+ \to \pi^+ \nu \bar{\nu}$ within SM is given by \cite{uliburas}
\begin{equation}
Br(K^+ \to \pi^+ \nu \bar{\nu}) = (8.5 \pm 0.7)\times 10^{-11},
\end{equation}
and the SM4 $1\sigma$ limit on $Br(K^+\to \pi^+ \nu \bar{\nu})$ is given by
\begin{eqnarray}
Br(K^+\to \pi^+ \nu \bar{\nu}) = (4.0 \to 12.0)\times 10^{-11}; \hskip 20pt
m_{t'}= 400 \hskip 5pt {\rm GeV}, \nonumber \\
Br(K^+\to \pi^+ \nu \bar{\nu}) = (4.0 \to 13.0)\times 10^{-11}; \hskip 20pt
m_{t'}= 600 \hskip 5pt {\rm GeV}.
\end{eqnarray}
Again these upper limits are consistent with the 95\% confidence level limit obtained
in Ref. \cite{burasewerth} calculated in MFV model.
\subsection{Branching fraction $K_L\to \pi^0\nu \bar{\nu}$}
\begin{figure}
\includegraphics[width=8cm,height=6cm,clip]{brk0nu.eps}%
\hspace{0.2cm}%
\includegraphics[width=8cm,height=6cm,clip]{brk0nufull.eps}
\caption{The branching fraction of $K_L \to \pi^0 \nu \bar{\nu}$ versus
$\phi^{ds}_{t'} = \phi^{d}_{t'}-\phi^{s}_{t'}$ in SM4,
red and blue region corresponds to $m_{t'}$ = 400 and 600 $\rm GeV$
respectively, the black horizontal lines represent $1\sigma$ SM limit; left panel
shows only $1\sigma$ range expected in SM4, full range for SM4 is shown in the right panel.}
\label{fig_kltopi}
\end{figure}
The effective Hamiltonian for $K_L\to \pi^0 \nu \bar{\nu}$ can be written as
\begin{equation}
{\cal{H}}_{eff} = {G_F \over \sqrt{2}}{\alpha\over {2\pi\sin^2{\Theta_w}}}\left( V^{\ast}_{ts}V_{td} X(x_t) + V^{\ast}_{t's}V_{t'd} X(x_{t'}) \right)(\bar{s}d)_{V-A}(\bar{\nu}\nu)_{V-A} + h.c.
\end{equation}
Within SM $K_L\to \pi^0\nu \bar{\nu}$ decay, proceeds almost entirely through
{\it CP} violation, is completely dominated by short-distance loop diagrams with top quark exchanges, here the charm contribution can be fully neglected.
The branching fraction of $K_L\to \pi^0\nu\bar{\nu}$ can be written as follows
\begin{equation}
Br(K_L\to \pi^0\nu\bar{\nu}) = \kappa_L .\left[\left({Im\lambda_t \over \lambda^5} X(x_t) + {Im\lambda_{t'} \over \lambda^5} X(x_{t'}) \right)^2 \right]
,
\label{brkpi0}
\end{equation}
with
\begin{equation}
\kappa_L = {r_{K_L}\over r_{K^+}}{\tau(K_L)\over\tau(K^+)}\kappa_+ = 1.80 \times 10^{-10},
\end{equation}
$\kappa_+$ and $r_{K_L}=0.944$ summarizing isospin breaking corrections in
relating $K_L\to \pi^0\nu\bar{\nu}$ to $K^+\to \pi^0 e^+ \nu$. The current
value of branching fraction for $K_L\to \pi^0\nu\bar{\nu}$ with SM is given
by \cite{uliburas}
\begin{equation}
Br(K_L \to \pi^{0} \nu \bar{\nu}) = (2.76 \pm 0.40)\times 10^{-11}.
\end{equation}
In Fig. (\ref{fig_kltopi}) the variation of branching fraction
$Br(K_L\to \pi^0\nu\bar{\nu})$ with the phase $\phi^{ds}_{t'}$ is shown\footnote{Right panel is
added in our revised version to facilitate direct comparision with \cite{buras4th}.}.
We note that with the constraint on $Br(K^+ \to \pi^+ \nu \bar{\nu})$ (Table. \ref{tab3}),
while, in principle $Br(K_L\to \pi^0\nu\bar{\nu})$ could be enhanced as much as
$1.2\times 10^{-9}$ (right panel Fig. \ref{fig_kltopi}), the expected $1\, \sigma$ range in SM4
(left panel Fig. \ref{fig_kltopi}) is only to $7\times 10^{-11}$, however, at 95\% CL the value
could be enhanced to $8\times 10^{-10}$. The branching fraction has its
maximum value when the phase $\phi^{ds}_{t'}$ has the value $\pm 90^{\circ}$ and
$270^{\circ}$ since SM4 contribution picks up its maximum value at those points
(eq. \ref{brkpi0}).
The SM4 $1\sigma$ limit on $Br(K_L\to \pi^0\nu\bar{\nu})$ is given by
\begin{eqnarray}
Br(K_L\to \pi^0\nu\bar{\nu}) = (1.0 \to 5.2)\times 10^{-11}; \hskip 20pt
m_{t'}= 400 \hskip 5pt {\rm GeV}, \nonumber \\
Br(K_L\to \pi^0\nu\bar{\nu}) = (1.0 \to 6.2)\times 10^{-11}; \hskip 20pt
m_{t'}= 600 \hskip 5pt {\rm GeV},
\end{eqnarray}
the upper limits are consistent with the limit calculated in
Ref. \cite{burasewerth}.
\subsection{CP violation in $B \to \pi K$ modes}
The observed data from the currently running two asymmetric $B$
factories are almost consistent with the SM
predictions and till now there is no compelling evidence for new physics.
However there are some interesting deviations from the SM
associated with the $b \to s$ transitions, which
provide us with possible indication of new physics.
For example the mixing induced CP asymmetries in many
$b \to s \bar q q$ penguin dominated modes do not seem to agree with
the SM expectations. The measured values in such modes
follow the trend $S_{s
\bar q q} < \sin 2 \beta $ \cite{Barberio:2008fa,LS09}, whereas in the SM they are
expected to be similar \cite{Grossman:1996ke,London:1997zk}. In this context $B \to \pi K$
decay modes, which receive dominant contributions from $b \to s$
mediated QCD penguins in the SM, provide another testing
ground to look for new physics.
The first one is the difference in direct CP asymmetries in $B^- \to
\pi^0 K^-$ and $\bar B^0 \to \pi^+ K^-$ modes. These two modes
receive similar dominating contributions from tree and penguin
diagrams and hence one would naively expect that these two channels
will have the same direct CP asymmetries i.e., ${\cal A}_{\pi^0
K^-}= {\cal A}_{\pi^+ K^-}$. In the QCD factorization approach, the
difference between these asymmetries is found to be \cite{LS07}
\begin{equation} \Delta A_{CP} ={\cal A}_{ K^- \pi^0} -{\cal A}_
{ K^- \pi^+} =(2.5 \pm
1.5)\%
\end{equation}
whereas the corresponding experimental value \cite{Barberio:2008fa} is
\begin{equation} \Delta A_{CP} =(14.8 \pm
2.8)\%\;,
\end{equation}
which yields nearly $4 \sigma$ deviation.
The second anomaly is associated with the mixing induced
CP asymmetry in $B^0 \to \pi^0 K^0$ mode.
The time dependent CP asymmetry in this mode is defined
as \begin{eqnarray} \frac{\Gamma(\bar B^0(t) \to \pi^0 K_s)-\Gamma(B^0(t) \to
\pi^0 K_s)}{\Gamma(\bar B^0(t) \to \pi^0 K_s)+\Gamma(B^0(t) \to
\pi^0K_s)}= A_{\pi^0 K_s} \cos(\Delta M_d t)+S_{\pi^0 K_s}
\sin(\Delta M_d t)\;, \end{eqnarray} and in the pure QCD penguin limit one
expects $A_{\pi^0 K_s} \approx 0$ and $S_{\pi^0 K_s} \approx \sin(2
\beta)$. Small non-penguin contributions do provide some corrections
to these asymmetry parameters and it has been shown in Ref.
\cite{Beneke:2005pu,Cheng:2006dk,Buchalla:2005us} that these corrections
generally tend to increase $S_{K
\pi^0}$ from its pure penguin limit of ($\sin 2 \beta$) by a
modest amount i.e., $ S_{\pi^0 K_s} \approx 0.8$.
Recently, using isospin symmetry it has been shown in \cite{rf, mg1,beak}
that the standard model favors a large $S_{\pi^0 K_s} \approx 0.99$.
However, the recent results from Belle
\cite{belle} and Babar \cite{babar} are \begin{eqnarray} A_{\pi^0 K_s}&=&
0.14 \pm 0.13 \pm
0.06,~~~~~S_{\pi^0 K_s}=0.67 \pm 0.31 \pm 0.08~~~~({\rm
Belle})\nonumber\\
A_{\pi^0 K_s}&=& -0.13 \pm 0.13 \pm 0.03,~~~~~S_{\pi^0 K_s}=0.55 \pm
0.20 \pm 0.03~~~~({\rm Babar}) \end{eqnarray} with average \begin{equation} A_{\pi^0 K_s}=
-0.01 \pm 0.10,~~~~~S_{\pi^0 K_s}=0.57 \pm 0.17\;.\label{av} \end{equation}
As seen from (\ref{av}), the observed value of
$S_{\pi^0 K_s}$ is found to be
smaller than the present world average value of $\sin 2 \beta=0.672
\pm 0.024 $ measured in $b \to c \bar c s$ transitions
\cite{Barberio:2008fa} by nearly $1\sigma$ and the deviation from the
SM expectation given above is possibly even larger.
This deviation which is opposite to the SM expectation, implies the
possible presence of new physics in the $ B^0 \to K^0 \pi^0 $ decay
amplitude. In the SM, this decay mode receives contributions from
QCD penguin ($P$), electroweak penguin ($P_{EW})$ and color
suppressed tree ($C$) diagrams, which follow the hierarchical
pattern $P:P_{EW}:C= 1:\lambda : \lambda^2$, where $\lambda \approx
0.2257$ is the Wolfenstein expansion parameter. Thus, accepting the
above discrepancy seriously one can see that the electroweak penguin
sector is the best place to search for new physics.
To account for these discrepancies here we consider the effect of
sequential fourth generation quarks \cite{Hou:2006zza,Hou:2005hd,Arhrib:2006pm,Hou:2006jy,AS_olds1}.
In the SM, the relevant
effective Hamiltonian describing the decay modes $B \to \pi K$
is given by
\begin{equation}
{\cal H}_{eff}^{SM} = \frac{G_F}{\sqrt{2}}\left[ V_{ub}
V_{us}^*(C_1O_1+C_2 O_2)- V_{tb}V_{ts}^*\sum_{i=3}^{10} C_i O_i
\right].
\end{equation}
With a sequential fourth generation, the Wilson coefficients $C_i$'s
will be modified due to the new contributions from $t^\prime$ quark
in the loop. Furthermore, due to the presence of the $t'$ quark the
unitarity condition becomes $\lambda_u+\lambda_c+\lambda_t+
\lambda_{t'}=0$, where $\lambda_q= V_{qb}V_{qs}^*$.
Thus, including the fourth generation and replacing $\lambda_t
=-(\lambda_u+\lambda_c +\lambda_{t'})$, the modified Hamiltonian
becomes \begin{eqnarray} {\cal H}_{eff}& = &\frac{G_F}{\sqrt{2}}\left[
\lambda_u(C_1O_1+C_2 O_2)- \lambda_t \sum_{i=3}^{10} C_i O_i
-\lambda_{t'}\sum_{i=3}^{10} C_i^{t'} O_i\right]\nonumber\\
&=&\frac{G_F}{\sqrt{2}}\left[ \lambda_u(C_1O_1+C_2 O_2
+\sum_{i=3}^{10} C_i O_i)+
\lambda_c\sum_{i=3}^{10} C_i O_i -\lambda_{t'}\sum_{i=3}^{10} \Delta
C_i O_i\right]\;, \end{eqnarray} where $\Delta C_i$'s are the effective (t
subtracted) $t'$ contributions.
Thus, one can obtain the transition amplitudes in the QCD
factorization approach as \cite{QCDF1,QCDF}
\begin{eqnarray} \sqrt2 A(B^- \to
\pi^0 K^-)&=&\lambda_u \Big(A_{\pi \bar K}(\alpha_1+\beta_2)+A_{\bar
K \pi}\alpha_2 \Big)\nonumber\\
&+&\sum_{p=u,c}\lambda_p\Big(A_{\pi \bar
K}(\alpha_4^p+\alpha_{4,EW}^p+\beta_3^p+\beta_{3,EW}^p)+\frac{3}{2}A_{\bar
K \pi}\alpha_{3,EW}^p \Big)\nonumber\\
&-& \lambda_{t'}\Big(A_{\pi \bar K}(\Delta \alpha_4+\Delta
\alpha_{4,EW}+\Delta \beta_3+\Delta \beta_{3,EW})+\frac{3}{2}A_{\bar
K \pi}\Delta\alpha_{3,EW} \Big),\nonumber\\ A(\bar B^0 \to \pi^+
K^-)&=&\lambda_u \Big(A_{\pi \bar K}~\alpha_1 \Big) +
\sum_{p=u,c}\lambda_p A_{\pi
\bar
K}\Big(\alpha_4^p+\alpha_{4,EW}^p+\beta_3^p-\frac{1}{2}
\beta_{3,EW}^p \Big)\nonumber\\
&-& \lambda_{t'}A_{\pi \bar K}\Big(\Delta \alpha_4+\Delta
\alpha_{4,EW}+\Delta \beta_3-\frac{1}{2}\Delta \beta_{3,EW} \Big),\nonumber\\
\sqrt 2 A(\bar B^0 \to \pi^0
\bar K^0)&=&\lambda_u A_{\bar
K \pi}\alpha_2 +\sum_{p=u,c}\lambda_p\Big[A_{\pi \bar
K}\Big(-\alpha_4^p+\frac{1}{2}\alpha_{4,EW}^p-
\beta_3^p+\frac{1}{2}\beta_{3,EW}^p \Big)\nonumber\\
&+&\frac{3}{2}A_{\bar K \pi}\alpha_{3,EW}^p \Big]
-\lambda_{t'}\Big[A_{\pi \bar
K}\Big(-\Delta \alpha_4+\frac{1}{2}\Delta \alpha_{4,EW}-
\Delta \beta_3+\frac{1}{2}\Delta \beta_{3,EW} \Big)\nonumber\\
&+&\frac{3}{2}A_{\bar K \pi}\Delta \alpha_{3,EW} \Big],
\end{eqnarray} where \begin{equation} A_{\pi \bar K}= i\frac{G_F}{\sqrt 2}M_B^2 F_0^{B\to
\pi}f_K~~~~~{\rm and}~~~~A_{ \bar K \pi}= i\frac{G_F}{\sqrt 2}M_B^2
F_0^{B\to K}f_{\pi}\;. \end{equation}
\begin{table}[t]
\begin{tabular}{|ccccccc|}
\hline $m_{t'}$ (in GeV) &&& 400 &&& 600 \\
\hline
$\Delta C_3(m_b) $ &&& 0.628 &&&1.471 \\
$\Delta C_4(m_b) $ &&& $-0.274$ &&& $-0.578$ \\
$\Delta C_5(m_b) $ &&& 0.042 &&& 0.086 \\
$\Delta C_6(m_b) $ &&& $-0.206$ &&& $-0.362$ \\
$\Delta C_7(m_b) $ &&& 0.443 &&& 1.072 \\
$\Delta C_8(m_b) $ &&& $0.168$ &&& 0.407 \\
$\Delta C_9(m_b) $ &&& $-1.926$ &&& $-4.465$ \\
$\Delta C_{10}(m_b) $ &&& 0.433 &&& 1.005 \\
$\Delta C_{7 \gamma}^{eff}(m_b) $ &&& $-5.667$ &&& $-7.239$ \\
$\Delta C_{8g}^{eff}(m_b) $ &&& $-1.452$ &&& $-1.728$ \\
\hline
\end{tabular}
\caption{Values of the Wilson coefficients $\Delta C_i$'s
at different $b$-mass scale.}
\label{tab5}
\end{table}
These amplitudes can be symbolically represented as \begin{eqnarray} Amp=
\lambda_u A_u +\lambda_c A_c -\lambda_{t'}A_{t'} .\end{eqnarray} $\lambda$'s
contain the weak phase information and $A_i$'s are associated with
the strong phases. Thus one can explicitly separate the strong and
weak phases and write the amplitudes as \begin{eqnarray} Amp=\lambda_c
A_c\Big[1+ r a e^{i(\delta_1-\gamma)}-r' b e^{i(\delta_2+\phi_s)}],
\end{eqnarray} where $a=|\lambda_u/\lambda_c|$, $b=|\lambda_{t'}/\lambda_c|$,
$-\gamma$ is the weak phase of $V_{ub}$ and $\phi_s$ is the weak
phase of $\lambda_{t'}$. $r=|A_u/A_c|$, $r'=|A_{t'}/A_c|$, and
$\delta_1$ ($\delta_2$) is the relative strong phases between $A_u$
and $A_c$ ($A_{t'}$ and $A_c$). From these amplitudes one
can obtain the direct and mixing induced CP asymmetry parameters as
\begin{eqnarray}
A_{\pi K}&=&\frac{2\Big[ra \sin
\delta_1 \sin \gamma +r' b \sin \delta_2 \sin \phi_s + r r' a b
\sin(\delta_2-\delta_1) \sin
(\gamma+\phi_s)\Big]}{\Big[{\cal{R}}+2ra \cos \delta_1 \cos
\gamma-2r' b \cos \delta_2\cos \phi_s - 2 r r' a b
\cos(\delta_2-\delta_1)\cos(\gamma+\phi_s)\Big]}\;,\nonumber\\
S_{\pi K}&= &\frac{X}{{\cal R}+2ra \cos \delta_1 \cos \gamma-2r'
b \cos \delta_2 \cos \phi_s- 2 r r' a b\cos(\delta_2-\delta_1)
\cos(\gamma+\phi_s)
}\;,\end{eqnarray} where
${\cal{R}}=1+(ra)^2+(r'b)^2$ and
\begin{eqnarray} X &=& \sin 2 \beta+ 2 r a \cos
\delta_1 \sin(2 \beta+\gamma)- 2 r' b \cos \delta_2
\sin(2 \beta- \phi_s) +(r a)^2\sin(2 \beta +2 \gamma) \nonumber\\
&+&(r' b)^2\sin(2 \beta -2 \phi_s)-2 rr' a b \cos(\delta_2-
\delta_1) \sin(2 \beta+\gamma -\phi_s). \end{eqnarray}
To find out the new contributions due to the fourth generation
effect, first we have to evaluate the new Wilson coefficients
$C_i^{t'}$. The values of these coefficients at the $M_W$ scale can
be obtained from the corresponding contributions from the $t$ quark
by replacing the mass of $t$ quark in the Inami-Lim functions
\cite{lim} by $t'$ mass. These values can then be evolved to the
$m_b$ scale using the renormalization group equation \cite{Buchalla:1995vs}
\begin{equation}
{\vec{C}} (m_b) = U_{5} (m_{b},M_{W}, \alpha) {\vec C} (M_{W})
\end{equation}
where $C$ is the $10 \times 1$ column vector of the Wilson
coefficients and $U_{5}$ is the five flavor $10 \times 10$ evolution
matrix. The explicit forms of ${\vec C}(M_W)$ and $U_{5} (m_b, M_W,
\alpha)$ are given in \cite{Buchalla:1995vs}. The values of $\Delta
C_{i=1-10} (m_b)$ in the NLO approximation and the coefficients of
the dipole operators $C_{7 \gamma}^{eff}$ and $C_{8g}^{eff}$ in the
LO for different $m_{t'}$ values are presented in Table \ref{tab5}.
\begin{figure}[t]
\centerline{\epsfysize 2.5 truein \epsfbox{kpi.eps}} \caption{The
allowed range of the CP asymmetry difference ($\Delta A_{CP}$) in
the ($\Delta A_{CP}-\lambda_t'$) plane, where the red and blue
regions correspond to $m_{t'}$ = 400 and 600 GeV; grey shaded regions correspond
to the uncertainties due to hadronic parameters.} \label{kpi2}
\end{figure}
\begin{figure}[t]
\includegraphics[width=8cm,height=6cm,clip]{fig1-sm.eps}%
\hspace{0.2cm}%
\includegraphics[width=8cm,height=6cm,clip]{fig1.eps}
\caption{Correlation plots between the mixing induced CP asymmetry
$S_{\pi^0 K_s}$ and the direct CP asymmetry $A_{\pi^0 K_s}$ in the
SM (left panel) and in the fourth generation model (right panel)
where the red and blue regions correspond to $m_{t'}=$400 and 600
GeV . The horizontal and vertical lines represent $1\sigma$ experimental
allowed ranges.}
\label{correlation2}
\end{figure}
For numerical evaluation, we use input parameters as follows. For
the form factors and decay constants we use $F_0^{B \to K}(0)=0.34
\pm 0.05$, $F_0^{B \to \pi}(0)=0.28 \pm 0.05$, $f_{\pi}=0.131$ GeV,
$f_K=0.16$ GeV and for Gegenbauer moments we use $\lambda_B = 350 \pm 150$\, MeV
\cite{QCDF}. We varied the hard spectator and
annihilation phases $\phi_{A,H}$ in the entire range i.e., between
$[-\pi, \pi]$, imposing the constraint that the corresponding
branching ratios should be within the three sigma experimental
range. Also we have included $20\%$ uncertainty in $\Lambda_{QCD}$ i.e we varied
$\Lambda_{QCD}= 225$\, MeV from its nominal value in SM3 \cite{QCDF} by $\pm 45$\, MeV,
which enters in the hard spectator contribution \footnote{The corresponding
choices in the scenario S4 of \cite{QCDF} are given by $F_0^{B \to K}(0)=0.31$,
$F_0^{B \to \pi}(0)=0.25$, $f_{\pi}=0.131$ GeV, $f_K=0.16$ GeV, $\lambda_B = 200$\, MeV,
$\phi_{A,H} = - 55^{\circ}$ and $\Lambda_{QCD} = 225$ MeV}. Since $\lambda_B$ and
$\Lambda_{QCD}$ were previously fixed to 200\, MeV and 225\, MeV respectively to
fit the data interpreted in SM3, it may not be unreasonable to assume small
changes for SM4. For the CKM matrix elements we use values as given in the
Table \ref{tab1}. We have also used the range of $\lambda_{t'}$ and $\phi_s$ as obtained
from the fit for different $m_{t'}$.
Using these values we show the allowed regions in the
$\Delta A_{CP}-\lambda_{t'}$ plane for different values of $m_{t'}$ in
figure \ref{kpi2} and we note that an enhancement in $\Delta A_{CP}$ upto
the current $1\,\sigma$ experimental upper bound ($\approx 17.6 \%$) is possible
for largish strong phases, $\phi_{A,H} \sim (-45\, \to \,- 90)^{\circ}$.
The correlation plots between mixing induced and
direct CP asymmetry parameters in $B^0 \to \pi^0 K^0$ modes are
shown in figure \ref{correlation2}.
\subsection{CP violation in $B^0 \to \pi^0 \pi^0$ modes}
As discussed earlier there exists several hints for the
possible existence of new physics in the
$b \to s$ sector. So the next obvious question is:
Do the $b \to d$ penguin amplitudes also have significant new physics
contribution? The present data does not provide any conclusive answer
to it. The obvious example is the $B \to \pi \pi $ processes, which receive
dominant contribution from $b \to u$ tree and from $b \to d$ penguin diagrams.
The present data \cite{Barberio:2008fa} are presented in Table \ref{tab6}.
\begin{table}[htbp]
\begin{center}
\begin{tabular}{|c|c|c|}
\hline
Decay mode & HFAG Average \\
\hline
$10^6 \times {\rm Br}(B^0 \to \pi^+ \pi^-)$ & $5.16 \pm 0.22 $\\
$10^6 \times {\rm Br}(B^- \to \pi^- \pi^0)$ & $5.59 \pm 0.41 $\\
$10^6 \times {\rm Br}(B^0 \to \pi^0 \pi^0)$ & $1.55 \pm 0.19
$ \\
$S_{\pi^+ \pi^-}$& $-0.65 \pm 0.07 $\\
{ $A_{\pi^+ \pi^-}$ } & { $ 0.38 \pm 0.06 $ }\\
$A_{\pi^- \pi^0}$ & $0.06 \pm 0.05 $ \\
{ $A_{\pi^0 \pi^0}$} & { $0.43_{-0.24}^{+0.25} $ }\\
\hline
\end{tabular}
\end{center}
\caption{Experimental results for $B \to \pi \pi$ processes}
\label{tab6}
\end{table}
Thus, it can be seen that the measured value of Br$(B^0
\to \pi^0 \pi^0)$ is nearly two times larger than the corresponding
theoretical predictions \cite{QCDF,lisanda}.
Also the measured values of direct CP asymmetry parameters
$A_{\pi^+ \pi^-}$ and $A_{\pi^0 \pi^0}$ are higher than the corresponding
SM predictions \cite{QCDF}.
Thus, the discrepancy between the theoretical and the measured quantities
imply that there may also be some new physics effect in the $b \to
d$ penguins as speculated in $b \to s$ penguins.
Let us first write down the most general topological amplitudes
for $B \to \pi \pi$ modes as
\begin{eqnarray}
\sqrt 2 A(B^+ \to \pi^+ \pi^0)& =& -(T +C +P_{ew}),\nonumber\\
A(B^0 \to \pi^+ \pi^-)& =&-(T+P),\nonumber\\
\sqrt 2 (B^0\to \pi^0 \pi^0) &=& -(C-(P-P_{ew})).
\end{eqnarray}
From the above relations it can be seen that if there will be additional
new contribution to the penguin sector
with other amplitudes as expected in SM4 then that may explain $B \to \pi \pi$
observations.
As discussed earlier, due to the presence of the additional
generation of quarks the unitarity condition becomes
$\lambda_u + \lambda_c
+\lambda_t +\lambda_{t'}=0$.
Thus, including the new contributions one can symbolically represent
these amplitudes as \begin{eqnarray} Amp = \lambda_u^d A_u^d +\lambda_c^d A_c^d
-\lambda_{t'}^d~A_{new} =\lambda_u^d A_u^d\Big[1- r_1~ a_1~
e^{i(\delta_1^d+\gamma)}-r'_1~ b_1~ e^{i(\delta_2^d+\phi_d)}\Big],
\end{eqnarray} where
$b_1=|\lambda_{t'}^d/\lambda_u^d|$, $\phi_d$ is the weak phase of
$\lambda_t'^d$.
$r'_1=|A_{new}/A_u^d|$, and $\delta_2^d$ is
the relative strong phases between $A_{new}$ and
$A_u^d$.
Thus from the
above amplitude one can obtain the CP averaged branching ratio,
direct and mixing induced CP asymmetry parameters as
\begin{eqnarray}
{\rm Br}&=& \frac{|p_{c.m}| \tau_B}{8 \pi M_B^2}
\Big[{{\cal{R}}_1}-2r_1a_1 \cos \delta_1^d \cos
\gamma-2r'_1 b_1 \cos \delta_2^d \cos (\phi_d+\gamma)
\nonumber\\&&
+2 r_1 r'_1 a_1 b_1
\cos(\delta_2^d-\delta_1^d)\cos\phi_d\Big]\;,\nonumber\\
A_{\pi \pi}&=&\frac{2\Big[r_1a_1 \sin
\delta_1^d \sin \gamma +r'_1 b_1 \sin \delta_2^d \sin (\phi_d+\gamma)
+ r_1 r'_1 a_1 b_1
\sin(\delta_1^d-\delta_2^d) \sin
\phi_d \Big]}{\Big[{\cal{R}}_1-2r_1a_1 \cos \delta_1^d \cos
\gamma-2r'_1 b_1 \cos \delta_2^d \cos (\phi_d+\gamma) +
2 r_1 r'_1 a_1 b_1
\cos(\delta_2^d-\delta_1^d)\cos\phi_d\Big]}\;,\nonumber\\
S_{\pi \pi}&=&\frac{X_1}{\Big[{\cal R}_1
-2r_1a_1 \cos \delta_1^d \cos
\gamma-2r'_1 b_1 \cos \delta_2^d \cos (\phi_d+\gamma) +
2 r_1 r'_1 a_1 b_1
\cos(\delta_2^d-\delta_1^d)\cos\phi_d\Big]
}\;,
\end{eqnarray}
where
\begin{eqnarray} X_1 &=&-\Big[ \sin (2 \beta+2\gamma)- 2 r_1 a_1 \cos
\delta_1^d \sin(2 \beta+\gamma)+ 2 r'_1 b_1 \cos \delta_2^d
\sin(\phi_d-(2 \beta+ \gamma))\nonumber\\ &+&(r_1 a_1)^2\sin(2 \beta )
+(r'_1 b_1)^2\sin(2\phi_d-2 \beta)-2 r_1r'_1 a_1 b_1 \cos(\delta_1^d-
\delta_2^d) \sin(\phi_d-2 \beta)\Big].
\end{eqnarray}
and ${\cal{R}}_1=1+(r_1a_1)^2+(r'_1b_1)^2$.
\begin{figure}[htb]
\centerline{\epsfysize 2.5 truein \epsfbox{pi0-600.eps}} \caption{The
correlation plot between the direct CP asymmetry and the CP-averaged
branching ratio for the $B^0 \to \pi^0 \pi^0 $ process where the
grey region corresponds to the SM result and
the red and blue regions correspond to
$m_{t'}=$400 and 600 GeV respectively. The horizontal and vertical lines represent the
1-$\sigma$ experimental range of the corresponding observables.}
\label{pi02}
\end{figure}
Now varying $\lambda_{t'}^d$ between 0 and $1.5\times 10^{-4}$ and $\phi_d$
between $(0-360)^\circ$ we present the correlation plot between the
direct CP asymmetry parameter and branching ratio in Fig. \ref{pi02}.
From the figure one can see that the observed data could be
accommodated in the SM with four generations.
\section{Summary and Outlook}
\label{concl}
Standard Model with four generations should be considered seriously.
We do not have a good understanding of fermion generations. We have
already seen three; why not the fourth? Electroweak precision tests
do not rule out the existence of a fourth family, though they do require
that the mass difference between the $t^\prime$ and the $b^\prime$ be less than about 75 GeV.
This degeneracy amounting to O(10\%) for $\approx 500$ GeV masses does not
seem so serious. Of course, the electroweak precision tests suggest then
a possible heavy Higgs particle but this actually may be hinting at a very
interesting resolution to the hierarchy puzzle. This is because
heavier quarks of the 4th generation can play a significant role in dynamical
electroweak-symmetry breaking, i.e. a composite Higgs particle .
Another extremely interesting implication of a 4th family is the gigantic improvement over the three generation case in the context of baryogenesis,
as in particular emphasized by Hou~\cite{Hou08}.
These two implications of a 4th family are in themselves so interesting, if not
profound, that even though at this time the repercussions for dark matter and/or unification are not quite clear, the idea should be given a serious consideration.
Although one of us (A.S.) had gotten already interested and involved in the physics of the 4th generation over twenty years ago, our recent interest
was instigated by the fact that this obvious extension of the Standard Model
offers a simple solution to many of the anomalies that have been seen in B,
$B_s$ decays. For one thing the predicted value of $\sin 2 \beta$ in the SM
is coming out to be too high from the one directly measured via the gold-plated
$\psi K_s$ mode. Besides, the value of $\sin 2 \beta$ measured via many of the penguin-dominated modes is systematically coming out to be smaller than the
predicted value. Then there is the very large difference in the direct CP
asymmetry between $K^+ \pi^-$ and $K^+ \pi^0$ decays of the $B^0$ and $B^+$.
Finally, there is the fact that both CDF and D0 find that $B_s \to \psi \phi$
decays are exhibiting O(2$\sigma$) non-vanishing CP asymmetries whereas
SM predicts vanishing small asymmetry.
The effect seen in $B_s \to \psi \phi$ at Fermilab is doubly significant.
First of all two of the anomalies discussed above that were seen at B-factories
taken seriously suggest a non-standard CP-odd phase in $b \to s$ transitions. That then makes it extremely difficult, if not impossible, for new physics
not to show up as well in $B_s$ mixing; thus the B-factory anomalies basically
imply non-standard CP effects in mixing induced CP-asymmetry in $B_s \to \psi \phi$. The second crucial aspect of the CP asymmetry in $B_s \to \psi \phi$
is that it is a gold-plated effect; that is the fact that in the SM CP asymmetry in that mode
should be vanishingly small is a very clean prediction with no serious hadronic uncertainty.
Therefore it is extremely important that Fermilab gives very high
priority to confirming or refuting this effect. In fact very soon the LHCb
experiment at CERN should also be able to study this mode and clarify this issue.
In an earlier paper we had focused on studying the CP anomalies seen in B, $B_s$
decays in SM4 mentioned above; we found that the SM4 offers a simple explanation for most of the
anomalies with the heavy quarks of mass around 400 - 600 GeV.
This paper is a follow-up wherein we further explore the implications
of SM4 for K and B, $B_s$ decays. By using a host of measurements in K, B, $B_s$
decays such as indirect CP violation parameter $\epsilon_K$, $K^+ \to \pi^+ \nu \bar \nu$, mixing induced CP asymmetry in $B \to \psi K_s$, Br ($B \to X_s \gamma$), semi-leptonic decays of B etc along with oblique parameters and Br( $Z \to b \bar b$), we first constrained the enlarged 4$\times$4 CKM-matrix.
We then explored the implications of the SM4 for a variety of processes such
as $a_{CP}(B \to X_s \gamma)$, $Br(B_s\to \mu^+\mu^-)$, $a_{CP} (B \to X_s l^+ l^-)$, $A_{SL}(B_s\to X_s \ell\nu)$, $A_{FB}(B \to X_s l^+ l^-)$, $A_{FB}(B \to K^* l^+ l^-)$,
$Br(B \to X_s \nu \bar \nu)$, CP asymmetries in $B\to \pi^0 K_s$ and in $B\to \pi^0 \pi^0$ etc.
We identified many processes wherein SM4 predicts significant differences from
SM3, {\rm e.g} $S(B_s\to \psi\phi)$, $a_{CP}(B \to X_s \gamma)$, $a_{CP} (B \to X_s l^+ l^-)$,
$A_{SL}(B_s\to X_s \ell\nu)$, $Br(B \to X_s \nu \bar \nu)$, $Br(B_s\to \mu^+\mu^-)$,
$Br(K_L\to \pi^0\nu\bar{\nu})$ {\rm etc}; thus studies therein should especially
provide further understanding of the parameter space of SM4.
One of the most interesting aspect of the 4th generation hypothesis is that
it is testable relatively easily in the LHC experiments where in fact it has
distinctive
signatures \cite{Hou08}. In the coming few years not only we should be able to learn about the existence or lack thereof of quarks and leptons of the 4th family, the heavier Higgs that is also favored
in SM4 scenario should be easier to search for in the LHC experiments via
the gold-plated mode: $H \to Z Z$. Also the heavy Higgs has interesting implications for
flavour-diagonal and flavour-changing final states involving $t'$ and/or $b'$ \cite{shaouly}.
Therefore, LHC should shed significant light on the question of SM4 in the next few years.
\begin{acknowledgments}
We want to thank Andrzej Buras, Martin Beneke, Thorsten Feldmann, Tillmann Heidsieck,
Alexander Lenz and Giovanni Punzi for many discussions.
SN would also like to thank Carlo Giunti for discussion regarding numerical
analysis and the theory division of Saha Institute of Nuclear Physics (SINP)
, in particular to Gautam Bhattacharyya, for hospitality. The work of AKA is
financially supported by
NSERC of Canada. The work of AS is suppported in part by the US DOE grant
\# DE-AC02-98CH10886(BNL). The work of AG is supported in part by CSIR and DST, Govt. of India and the work of RM is supported in part by DST, Govt. of India.
SN's work is supported in part by MIUR under contract 2008H8F9RA$\_$002 and by the European
Community’s Marie-Curie Research Training Network under contract MRTN-CT-2006-035505 ‘Tools and Precision Calculations for Physics Discoveries at Colliders’.
\end{acknowledgments}
|
\section{Introduction}
Recently, there are remarkable developments in experiments for the nucleon
electromagnetic form factors:\\
1) For the space-like momentum, the ratio of the
electric and magnetic form factors of proton, $G_E^p$ and $G_M^p$ respectively,
was shown to be a decreasing function of the
squared momentum transfer $Q^2$ and the experimental results imply that the
proton electric form factor vanishes for $Q^2\approx7 ({\rm GeV/c})^2$
\cite{jonesGE/GM}-\cite{ronGEGM}. \\
2) For the neutron magnetic
form factor, $G_M^n$, very accurate experimental data
were obtained and it approximately satisfies $G_M^n(Q^2)/\mu_n \approx
G_D(Q^2)=(1+Q^2/0.71)^{-2}$, with $Q$ being represented in terms of
GeV/c, for fairly wide range of squared momentum transfer
$Q^2=1.4-4.8$ (GeV/c)$^2$ \cite{andersionGMn},
\cite{lachnietGMn} (CLAS collaboration).\\
3) For the time-like momentum the ratio $|G_E^p/G_M^p|$ was obtained
\cite{bardinGptm}, \cite{aubert|GEGM|} (BABAR collaboration), while
previously the data of form factors had been analyzed under the assumption
$G_E^p=0$ or $G_E^p=G_M^p$ .\\
Asymptotically, the experimental data of nucleon magnetic factors decrease more
rapidly than the dipole formula for large $Q^2$ and the decrease has been
understood as
a realization of perturbative QCD \cite{brodsky}, the behavior of which can be
formulated in terms of the dispersion theory with appropriate conditions on the
absorptive parts;
we assumed unsbtracted dispersion relations for the charge
and magnetic moment form factors. To realize the asymptotic form of QCD we
imposed the superconvergence conditions.
As the data for the time-like momentum have become accurate, it is necessary
to investigate the form factors for the space-like and time-like momentums
systematically. For this purpose the dispersion theory is effective.
The dispersion theoretical calculations performed so far, the value of $G_M^n$
turned out to be larger than the above mentioned new experimental data for
$Q^2=1.4-4.8$ (GeV/c)$^2$ (see Ref. \cite{lachnietGMn}).
It is of vital importance to
investigate if it is
possible to realize the experimental data simply by the
adjustment of parameters or by the refinement of absorptive parts in the
dispersion relation.
It is the purpose of this paper to analyze experimental data of nucleon form
factors by
the dispersion theory, with the QCD constraints imposed, taking account of the above
mentioned new experimental results.
Organization of the paper is given as follows: In Sec.\ 2 we explain the
superconvergent dispersion relation and give conditions
which are used in this paper. We summarize the absorptive
parts, which are broken up into three parts: Low, intermediate and asymptotic
momentum regions. For each momentum region the imaginary parts are given.
The asymptotic part is expressed as an expansion in terms of the analytically
regularized running
coupling constant in the renormalization group for QCD. In Sec.\ 3 we remark
on
the numerical analysis.
In Sec.\ 4 numerical results are summarized. The final section is
devoted to general discussions.
\section{Dispersion Relation for the Electromagnetic Form Factors}
We assume the unsubtracted dispersion relations for the charge and magnetic
moment form factors, $F_1^{I}$ and $F_2^I$, respectively, with $I$
denoting the isospin state $I=0,1$. That is,
\begin{equation}
F_i^{I}=\frac{1}{\pi}\int_{t_0}^{\infty}dt'
\frac{{\rm{Im}}\, F_i^{I}(t')}{t'-t},\quad(i=1,\,\,2) \label{unsubtracted}
\end{equation}
where the threshold is $t_0=4\mu^2$. Here $\mu$ is the pion mass being
taken as the average of the neutral and charged pion masses. We impose
conditions on ${\rm{Im}}\, F_i^{I}$ to realize the QCD conditions.
\subsection{Superconvergence Condition and QCD}
Experimental data imply that the magnetic form factors of nucleon decrease
more rapidly than the dipole formula for large squared momentum transfer.
The decrease agrees with the prediction of perturbative QCD, where magnetic
form factors of nucleon
decrease for $Q^2\to \infty$ as
\begin{equation}
G_M(q^2)\to {\rm const}\frac{\alpha_S(Q^2)^2}{Q^4}
\left(\ln \frac{Q^2}{\Lambda^2}\right)^{4/3\beta_0},
\end{equation}
where $\alpha_S$ is the running coupling constant of QCD and
$\beta_0=11-2n_f/3$ with $n_f$ being the number of flavor. $\Lambda $ is the
QCD scale parameter having the dimension of momentum.
To realize the QCD predictions we impose the following conditions on
the charge and magnetic moment form factors:
\begin{eqnarray}
F_1(Q^2)&\to& {\rm const}/[Q^2 (\ln Q^2/{\Lambda}^2)^{\gamma}], \nonumber \\
F_2(Q^2)&\to& {\rm const}/[Q^4 (\ln Q^2/{\Lambda}^2)^{\gamma}], \label{QCD12}
\end{eqnarray}
for $Q^2 \to \infty$ with $\gamma\ge2$.
We briefly summarize the asymptotic
theorems which are used to incorporate the
constraints of QCD \cite{brodsky}, where the proof is given in
Ref.\ \cite{nw1}.
Let $F(t)$ satisfy the dispersion relation (\ref{unsubtracted}),
and ${\rm Im}F$ is given as
\begin{equation}
{\rm Im}F(t') = \frac{c}{[\ln (t'/\Lambda^2)]^{\gamma+1}}
+O\left(\frac{1}{[\ln(t'/\Lambda^2)]^{\gamma+2}}\right)
\end{equation}
for $t\to \infty$ with $\gamma>1$. Then $F(t)$ becomes
\begin{eqnarray}
F(t)&=&\frac{1}{\pi}\int_{t_0}^{\infty}dt'
\frac{c}{(t'-t)[\ln (t'/\Lambda^2)]^{\gamma+1}}\\ \nonumber
&\to & \frac{c}{\pi\gamma\ln(|t|/\Lambda^2)]^{\gamma}} \label{disp1}
\end{eqnarray}
for $t\to \pm \infty$. Generally, when $F(t')$ satisfies
\begin{equation}
t^{\prime\,n+1}{\rm Im}F(t')\,\,\to \frac{c}{[\ln(t'/\Lambda^2)]^{\gamma+1}}
+O\left(\frac{1}{[\ln(t'/\Lambda^2)]^{\gamma+2}}\right) \label{asymp2}
\end{equation}
for $t' \to \infty$ and the superconvergence conditions
\begin{equation}
\int_{t_0}^{\infty}dt't^{\prime k}{\rm Im}F(t')=0,\quad k=0,1,\cdots,n,
\label{asym4}
\end{equation}
$F(t)$ given by (\ref{unsubtracted}) approaches for
$t \to \pm\infty$ to the following formula:
\begin{equation}
F(t)= \frac{1}{\pi}\int_{t_0}^{\infty}dt'\frac{{\rm Im}F(t')}{t'-t}
\to \frac{1}{t^{n+1}}\frac{c}{\pi\gamma[\ln(|t|/\Lambda^2)^{\gamma}},
\label{asym5}
\end{equation}
which can be proved by using (\ref{disp1}) and (\ref{asym4})
together with the identity
$$
\frac{1}{t'-t}=-\frac{1}{t}\Big\{1+\frac{t'}{t}+\cdots
+\Big(\frac{t'}{t}\Big)^n\Big\}+\frac{1}{t^{n+1}}
\frac{t^{\prime n+1}}{t'-t}.
$$
Indeed, by using (\ref{asym4}) we have
\begin{equation}
\int_{t_0}^{\infty}dt'\frac{{\rm Im}F(t')}{t'-t}
=\frac{1}{t^{n+1}}\int_{t_0}^{\infty}dt'
\frac{t^{\prime n+1}{\rm Im}F(t')}{t'-t},
\end{equation}
which leads to (\ref{asym5}) as $t^{\prime\,n+1}{\rm Im}F(t')$ satisfies
(\ref{asymp2}).
To obtain the asymptotic formulas (\ref{QCD12}), therefore, we impose the
superconvergence conditions on the imaginary part of form factors,
${\rm Im}F_i^I(t)$ ($i=1,\,\,2$; $I$ denotes isospin) in the unsubtracted
dispersion relation (\ref{unsubtracted}):\\
\begin{eqnarray
\frac{1}{\pi}\int_{t_0}^{\infty}dt'\,{\rm Im}F_1^I(t')
&=&\frac{1}{\pi}\int_{t_0}^{\infty}dt't'\,{\rm Im}F_1^I(t')=0,
\label{sup1}\\
\frac{1}{\pi}\int_{t_0}^{\infty}dt'\,{\rm Im}F_2^I(t') \nonumber
&=& \frac{1}{\pi}\int_{t_0}^{\infty}dt't'\,{\rm Im}F_2^I(t')
= \frac{1}{\pi}\int_{t_0}^{\infty}dt't^{\prime\,2}\,{\rm Im}F_2^I(t')
=0, \label{sup2}
\end{eqnarray
where ${\rm Im}F_i^{I}(t')$ satisfies the asymptotic conditions for $t' \to
\infty$
\begin{equation}
t^{\prime\, i}{\rm Im}F_i^{I}(t') \to {\rm const}/[\ln (t'/\Lambda^2)]^{\gamma+1}
\quad (i=1,2). \label{asymptotic}
\end{equation}
In addition to the conditions (\ref{sup1}) and (\ref{sup2}) we impose the
normalization conditions at $t=0$:
\begin{eqnarray
\frac{1}{2} &=& \frac{1}{\pi}\int_{t_0}^{\infty}dt'\,{\rm Im}F_1^{I}(t')/t',
\label{norm1} \\
g^I &=& \frac{1}{\pi}\int_{t_0}^{\infty}dt'\,{\rm Im}F_2^{I}(t')/t',
\label{norm2}
\end{eqnarray
where $g^I$ is the anomalous magnetic moments of nucleons with the
isospin $I$.
\subsection{Imaginary part of the form factors}
Let us discuss the imaginary parts of nucleon form factors, which are broken
up into three parts:
The low momentum, the intermediate, and the asymptotic regions.
\subsubsection{Low momentum region}
The imaginary parts of the charge and magnetic moment form factors,
${\rm Im}F_i^{V}$, are given in terms of two pion contribution as follows:
\begin{eqnarray*
{\rm Im}[F_1^{V}(t)/e] &=& \frac{m}{2}\frac{(t-4\mu^2)}{4m^2-t} \left(\frac{t-4\mu^2}{t}\right)^{1/2}\\% \nonumber \\
&\times& {\rm Re}\Big[M^{*}(t) \Big\{f_{+}^{(-)1}(t) -\frac{t}{4m^2}\frac{m}{\sqrt 2}f_{-}^{(-)1}(t)\Big\}\Big]
\end{eqnarray*
\begin{eqnarray
{\rm Im}[2mF_2^{V}(t)/e] &=& \frac{m}{2}\frac{(t-4\mu^2)}{(4m^2-t)}
\left(\frac{t-4\mu^2}{t}\right)^{1/2} \label{ImH}\\%
\nonumber\\
&\times& {\rm Re}\Big[M^{*}\Big\{\frac{m}{\sqrt 2}f_{-}^{(-)1}(t)
-f_{+}^{(-)1}(t)\Big\}\Big], \nonumber
\end{eqnarray
where $f_{\pm}^{(-)1}(t)$ are helicty
amplitudes for $ \pi\pi\leftrightarrow N\bar{N}$, $M(t)$ is the pion
form factor and $\mu$ is the pion mass. The superscript $V$ denotes
the iso-vector part.
For the helicity amplitudes we use the numerical values given by H\"ohler and
Schopper \cite{hoe} and parameterize $M(t)$ according to them.
\begin{equation}
M(t) =
t_{\rho}\{1+(\Gamma_{\rho}/m_{\rho}d)\}
[t_{\rho}-t-im_{\rho}^2\Gamma_{\rho}(q_t/q_{\rho})^3\sqrt{t}]^{-1},
\end{equation}
where $m_{\rho}$ and $\Gamma_{\rho}$ are the $\rho$ meson mass and width
respectively and
\begin{eqnarray}
t_{\rho}&=&m_{\rho}^2,\quad q_{\rho}=\sqrt{t_{\rho}-\mu^2},\quad \\
d&=& \frac{3\mu^2}{\pi t_{\rho}}\ln\frac{m_{\rho}+2q_{\rho}}{2\mu}
+\frac{m_{\rho}}{2\pi q_{\rho}}\Big(1-\frac{2\mu^2}{t_{\rho}}\Big).
\end{eqnarray}
The imaginary parts thus obtained are denoted as ${\rm Im}F_i^{H}\,\,(i=1,2)$
hereafter. It must be remarked that the $\rho$ meson contribution is included in the helicity amplitudes of Ref.\ \cite{hoe}. The uncorrelated kaon pair is
neglected here
as the effect was estimated to be small \cite{fw2}.
\subsubsection{Intermediate region}
The intermediate states $4\mu^2\le t \le \Lambda_1^2$ are approximated
by the addition of the Breit-Wigner terms, with the imaginary part
parameterized as follow:
\begin{equation}
{\rm Im}f_R^{BW}(t) = \frac{g}{(t-M_R^2)^2+g^2},
\end{equation}
where
\begin{equation}
g=\frac{\Gamma M_R^2(M_R^2+t_{res})^3}{t_{res}^2(M_R^2-t_0)^{3/2}}
\sqrt{\frac{(t-t_0)^3}{t}}\frac{t^2}{(t+t_{res})^3}.
\end{equation}
Here $M_R$ and $\Gamma$ are the mass and width of resonance, respectively, the threshold $t_0$ is
$t_0=4\mu^2$ and $t_{res}$ is treated as an
adjustable parameter. g is introduced to cut-off the Breit-Wigner
formula.
We write the intermediate part as the summation of resonances
\begin{equation}
{\rm Im} F_i^{BW,I}=\sum_n a_n^{I,i}f_{nR}^{I}, \label{ImBW}
\end{equation}
where $I$ is the isospin and $n$ is the labeling of resonances (see Table I).
Here the suffix $i$ denotes $i=1,\,\,2$, corresponding to the charge and
magnetic moment form factors $F_1^N$ and $F_2^N$ ($N$ = n or p). The same formulas for
$f_{nR}^{I}$ are used for $i=1$ and $i=2$.
\subsubsection{Asymptotic region}
We express the form factors as power series in the running coupling constant
of QCD, $\alpha_S$. To calculate the absorptive part, it is necessary to
perform analytic continuation to the time-like momentum. Here we give
only the necessary procedure for the analytic continuation of
the running coupling constant to the time-like momentum by using the
analytic regularization \cite{dok1} \cite{dok2}, as the
formulation is given in Ref.\ \cite{nw1}.
Let $\alpha_S(Q^2)$ be the running coupling constant in the renormalization
group calculated by the
perturbative QCD as the function of the squared momentum $Q^2$ for the
space-like momentum. We use the three loop approximation for
$\alpha_S(Q^2)$, which is expressed in the Pad\'e form.
\begin{equation}
\alpha_S(Q^2) = \frac{4\pi}{\beta_0}\Big[\ln(Q^2/\Lambda^2)
+a_1\ln\{\ln(Q^2/\Lambda^2)\}
+a_2\frac{\ln\{\ln(Q^2/\Lambda^2)\}}{\ln(Q^2/\Lambda^2)}
+\frac{a_3}{\ln(Q^2/\Lambda^2)}+\cdots\Big]^{-1}. \label{PQCD}
\end{equation}
$\Lambda$ is the QCD scale parameter, and $a_i$ are expressed in terms of
the $\beta$ function of QCD,
\begin{equation}
a_1=2\beta_1/\beta_0^2, \quad a_2=4\frac{\beta_1^2}{\beta_0^4}, \quad
a_3=
\frac{4\beta_1^2}{\beta_0^4}\left(1-\frac{\beta_0\beta_2}{8\beta_1^2}\right),
\end{equation}
where
\begin{equation}
\beta_0 = 11-\frac{2n_f}{3},\quad \beta_1=51-\frac{19n_f}{3},
\beta_2 = 2357-\frac{5033}{9}n_f+\frac{325}{27}n_f^2
\end{equation}
with $n_f$ being the number of flavor.
We perform the analytic continuation of the squared momentum to the time-like
region, $s$, by the replacement in (\ref{PQCD})
\begin{equation}
Q^2\to e^{-i\pi}s.
\end{equation}
Then $\alpha_S(e^{-i\pi}s)$ becomes complex and is expressed as follows:
\begin{eqnarray
\alpha_S(e^{-i\pi}s) &=& \,1/(u-iv)=\frac{u+iv}{D}, \\
D &=& \,u^2+v^2,
\end{eqnarray
where $u$ and $v$ are given as
\begin{eqnarray
u &=& \ln(s/\Lambda^2)+\frac{a_1}{2}\ln\{\ln^2(s/\Lambda^2)+\pi^2\}
\nonumber \\
&&+\frac{a_2}{\ln^2(s/\Lambda^2)+\pi^2}
\Big[\frac{1}{2}\ln(s/\Lambda^2)\ln\{\ln^2(s/\Lambda^2)
+\pi\theta\}\Big] \nonumber \\
&&+\frac{a_3\ln(s/\Lambda^2)}{\ln^2(s/\Lambda^2)+\pi^2},
\end{eqnarray
\begin{eqnarray
v &=& \pi+a_1\theta \nonumber \\
&&-\frac{a_2}{\ln^2(s/\Lambda^2)+\pi^2}
\Big[\frac{\pi}{2}\ln\{\ln^2(s/\Lambda^2)+\pi^2\}
-\theta\ln(s/\Lambda^2)\Big] \nonumber \\
&&-\frac{\pi a_3}{\ln\{\ln^2(s/\Lambda^2)+\pi^2\}},
\end{eqnarray
with
\begin{equation}
\theta=\tan^{-1}\{\pi/\ln(s/\Lambda^2)\}.
\end{equation}
The running coupling constant is given by the dispersion integral
both for the space-like and the time-like momentum
\begin{equation}
\alpha_R(t)=\int_0^{\infty}dt'\frac{\sigma(t')}{t'-t} \label{regular}
\end{equation}
with
\begin{equation}
\sigma(t')={\rm Im}\alpha_S(e^{-i\pi}s)=4\pi v/\beta_0D.
\end{equation}
$\alpha_R(t)$ represented by (\ref{regular}) is called analytically
regularized running coupling constant as it has no singular point for
$t=-Q^2<0$.
The regularization eliminates the ghost
pole of $\alpha_S(Q^2)$, given by (\ref{PQCD}), appearing at
\begin{equation}
Q^2=Q^{*2}=\Lambda^2 e^{u^{*}},
\end{equation}
where $u^{*}=0.7659596\cdots$ for the number of flavor $n_f=3$.
Calculating (\ref{regular}), we find that $\alpha_R(t)$ is approximately given by the simple formula with the ghost pole subtracted
\begin{equation}
\alpha_R(Q^2)\approx \alpha_S(Q^2)-A^{*}/(Q^2-Q^{*2}) \label{regular-1},
\end{equation}
where the residue $A^{*}$ is
\begin{equation}
A^{*}=4\pi\Lambda^2e^{u^{*}}/\Big\{\beta_0
\Big(1+\frac{a_1}{u^{*}}
-a_2\frac{\ln u^{*}}{u^{*2}}+\frac{a_2-a_1}{u^{*2}}\Big)\Big\}.
\end{equation}
We use (\ref{regular-1}) as the regularized coupling constant; for the
time-like momentum we replace $Q^2 \to e^{-i\pi}s$ in (\ref{regular-1}) as
was mentioned before.
The QCD parts, $F_i^{QCD,\,I}$ ($i=1,2$; $I$ = 0,1) for the squared time-like
momentum, are written as follows:
\begin{equation}
F_i^{QCD,\,I}(s)= \hat{F}_i^{QCD,\,I}(s)h_i(s), \label{qcd1}
\end{equation}
where $\hat{F}_i^{QCD,\,I}$'s are given as expansion in terms of the
running coupling constant
\begin{equation}
\hat{F}_i^{QCD,\,I}(s)=\sum_{j\ge 2}c_j^{QCD,\,I}
\{\alpha_R(s)\}^j \label{qcd2}
\end{equation}
for the time-like squared momentum $s$. We multiply by the function $h(s)$ in
(\ref{qcd1}) to assure the convergence of the superconvergence
conditions (\ref{sup1}) and (\ref{sup2}).
The following formula is assumed for $h_i(s)$:
\begin{equation}
h_i(s)=\left(\frac{s-t_{Q}}{s+t_1}\right)^{3/2}
\left(\frac{t_2}{{s+t_2}}\right)^{i+1},
\end{equation}
which may be interpreted as the form factor for $\gamma \to q\bar{q}$ with
$t_Q$ being the threshold of the quark antiquark pair.
The parameters $t_Q$, $t_1$ and $t_2$ are
taken as adjustable parameters and will be determined by the analysis of
experimental data.
For the time-like momentum, we perform the analytic
continuation of the regularized effective coupling constant $\alpha_R(Q^2)$
to $\alpha_R(s)$ through the equation
\begin{equation}
\alpha_R(s)=\alpha_R(Q^2e^{-i\pi})={\rm Re}[\alpha_R(s)]
+i\,{\rm Im}[\alpha_R(s)].
\end{equation}
We express the QCD part as the power series expansion in $\alpha_R(s)$
\begin{equation}
\hat{F}_i^{QCD,I}(s)=\sum_{2\le j}c_{i,j}^{QCD,\,{\rm I}}
\{\alpha_R(s)\}^j. \label{cQCD}
\end{equation}
The summation in (\ref{cQCD}) begins in the second order in the effective
coupling constant so as to realize the logarithmic decrease of the nucleon
form factors.
Imaginary part of (\ref{cQCD}) is obtained to be
\begin{eqnarray
&&{{\rm Im}\hat{F}}_i^{QCD,I}
=\,2c_{i,2}^{QCD,\,{\rm I}}{\rm Re}\,\alpha_R
{\rm Im}\,\alpha_R \nonumber \\
&&\quad+ c_{i,3}^{QCD,\,{\rm I}}[3({\rm Re}\,\alpha_R)^2{\rm Im}\,\alpha_R
-({\rm Im}\,\alpha_R)^3] \nonumber \\
&&\quad+c_{i,4}^{QCD,\,I}[4({\rm Re}\,\alpha_R)^3{\rm Im}\,\alpha_R
-4{\rm Re}\,\alpha_R({\rm Im}\,\alpha_R)^3] \nonumber \\
&&\quad+\cdots, \label{ImQCD}
\end{eqnarray
and
\begin{equation}
{\rm Im}F_i^{QCD,\,I}(s)={\rm Im}\hat{F}_i^{QCD\,I}(s)h_i(s).\label{ImQCD1}
\end{equation}
We write the low energy part, intermediate resonance part and asymptotic
QCD parts of form factors as $F_i^{\rm{ H}}$, $F_i^{BW, I}$ and $F_i^{QCD,\,I}$,
respectively, which are
given by the dispersion integral with the imaginary parts (\ref{ImH}),
(\ref{ImBW}) and (\ref{ImQCD1}). The form factors $F_i^{I}$
are defined by adding them up. We impose the conditions (\ref{sup1}) and
(\ref{sup2}) on ${\rm Im}F_i^{I}$ so that the QCD conditions are satisfied.
\section{Numerical Analysis}
We analyzed the experimental data of nucleon electromagnetic form factors
$G_M^p/\mu_pG_D$, $G_E^p/G_D$, $G_M^n/\mu_nG_D$
$G_E^n$ and the ratio $\mu_p G_E^p/G_M^p$ for the space-like momentum transfer, and
$|G^p|$ and $|G^n|$ in Refs.\ \cite{bostedGMp}-
\cite{madeyGEn/GMn} for the time-like momentum transfer and the above mentioned
recent experimental data
$G_M^n$ for the space-ike and
$|G_E^p/\mu_p G_M^p|$ for the time-like momentum transfer.
The parameters appearing in the
formulas are determined so as to minimize $\chi^2$.
As was mentioned in the introduction we analyze by taking account of
the recent experimental data: (a) $G_M^n$ for $Q^2 = 1-4.8$ (GeV/c)$^2$
(CLAS collaboration) and (b) $|\mu_pG_E^p|/|G_M^p|$ (BABAR collaboration).
In order to see how the situation changes by taking account of these new
experiments in addition to the other data,
we perform analysis for the following two cases in the $\chi^2$ analysis:\\
Case I: Both of the experimental data, (a) $|\mu_pG_E^p/G_M^p|$ for the
time-like
momentum and
(b) new data for $G_M^n$ for the space-like momentum, are added.\\
Case II: Only the data (a) $|\mu_pG_E^p/G_M^p|$ for the time-like
momentum are added.\\
Let us remark on the experiments for the time-like momentum
\cite{bardinGptm}, \cite{ablikimGptm}, \cite{antonelliGntm}, where
the form factors
$|G^p|$ and $|G^n|$ are determined by using the
formula for the cross section $\sigma_0$ for the processes
$e+\bar{e} \to N+\bar{N}$ or $N+\bar{N} \to e+\bar{e}$, which is given as
\begin{equation}
\sigma_0=\frac{4\pi\alpha^2\nu}{3s}\left(1+\frac{2m_p^2}{s}\right)
|G(s)|^2.
\label{exp}
\end{equation}
Here $\alpha$ is the fine structure constant and $\nu$ is the nucleon
velocity.
$|G_M^N|$ are estimated from $|G|$ under the assumption $G_M=G_E$ or $G_E=0$.
$\sigma_0$ is expressed in terms of $G_M^N$ and $G_E^N$ as follows:
\begin{equation}
\sigma_0=\frac{4\pi\alpha^2\nu}{3s}
\left(|G_M^N|^2+\frac{2m^2}{s}|G_E^N|^2\right). \label{theory}
\end{equation}
Equating (\ref{exp}) and (\ref{theory}), we have
\begin{equation}
|G|^2=\frac{|G_M^N|^2+2m^2|G_E^N|^2/s}{1+2m^2/s}. \label{timedata}
\end{equation}
Substituting our calculated result of form factors to the right hand side
of (\ref{timedata}), we obtain
the theoretical value for $|G|$, which is compared with the experimental
data for the magnetic form factor obtained under the assumption $G_M=G_E$.
The parameters appearing in our analysis are the following:
Residues at resonances, coefficients appearing in the expansion by the
QCD effective coupling constants, cut-offs for the intermediate region
$\Lambda_1$.
In addition to them we have parameters in the Breit-Wigner formula and
the convergence factor $h$ of QCD contribution, $t_Q$, $t_{res}$, $t_1,\,\,
t_2,\,\, t_3$.
We have taken the masses and the widths of resonances as adjustable parameters.
As the superconvergence constraints impose very stringent conditions on the
form factors, it was necessary to take the masses and widths as parameters. \\
\section{Numerical Results}
We give in Tables 1, 2 and 3 the results for the parameters
for the cases I and II obtained by the $\chi^2$ analysis; in Table 1 the masses
and widths of
resonances and in Table 2 residues at
resonance poles and in Table 3 the coefficients
$c_{i,j}^{QCD,\,I}$ $(i=1,2; \,\,j=2,3,4;\,\,I=0,1)$
in the expansion in terms of the
effective coupling constant $\alpha_R$ of QCD defined by (\ref{cQCD}).
The number of flavor is taken as $n_f = 3$. ${\rm Im} F_i^H$ is cut-off at
$\Lambda_0^2=0.779$
GeV$^2$ and the Breit-Wigner formulas at $\Lambda_1$ = 26.0 GeV.
The QCD parameter is fixed at $\Lambda$ = 0.216 GeV. The other parameters
are determined as follows:\\
Case I: $t_0$ = 4$\mu^2$, $t_1$ = 0.243$\times 10^3$ GeV$^2$,
$t_2$ = 0.237$\times 10^3$ GeV$^2$, $t_{res}$ = 0.2260$\times10^3$ GeV$^2$,
$t_Q$ = 0.202$\times 10^2$ GeV$^2$.\\
Case II: The same as in the case I except for $t_{res}$ = 0.2253$\times10^3$
GeV$^2$.
The value of
$\chi^2$ is obtained to be $\chi^2_{tot}=393.4$ for the case I and
$\chi^2_{tot}= 308.7$ for the case II, which includes both
the data of space-like and time-like regions.
The total number of data is 245 for the Case I and 236 for the
Case II. Number of parameters is 36 so that DOF/$\chi_{min}$ = 1.88 for the
Case I and 1.54 for the Case II.
\begin{table}
\caption{Masses and widths determined by the $\chi^2$ analysis for
the cases I and II.\hspace{64pt}$\quad$}
\begin{tabular}{cc|cl|ccll}\hline
{}& {}& $\hspace{20pt}$ case I &{} & $\hspace{20pt}$ case II&{}\\ \hline
{}&{}& mass & width & mass & width \\
isospin & $n$ &(GeV/c$^2$)&(GeV)&(GeV/c$^2$)&(GeV) \\\hline
{} & 1 & 1.341 & 0.3221 & 1.352 & 0.325 \\
{} & 2 & 1.379 & 0.2204 & 1.370 & 0.220 \\
$I=1$ & 3 & 1.599 & 0.2636 & 1.587 & 0.264 \\
{} & 4 & 1.824 & 0.3679 & 1.826 & 0.368 \\
{} & 5 & 2.048 & 0.3848 & 2.100 & 0.398 \\ \hline
{} & 1 & 0.78256 & 0.844$\times 10^{-2}$ & 0.78256
& 0.844$\times 10^{-2}$ \\
{} & 2 & 1.01945 & 0.426$\times 10^{-2}$ & 1.01945 & 0.426$\times 10^{-2}$ \\
$I=0$ & 3 & 1.212 & 0.1582 & 1.206 & 0.1584 \\
{} & 4 & 1.437 & 0.2102 & 1.440 & 0.2104 \\
{} & 5 & 1.505 & 0.1281 & 1.510 & 0.1285 \\ \hline
\end{tabular}
\end{table}
\begin{table}
\caption{The coefficients $a_i^{I,n}$, residues at the resonance poles,
determined by the $\chi^2$ analysis for the cases I and II.}
\begin{tabular}{cc|cc|cc}\hline
{} & {} & $\hspace{20pt}$ case I & {} & $\hspace{20pt}$ case II {} \\ \hline
isospin & $n$ & $a_1^{I,n}$(GeV$^2$)&$a_2^{I,n}$(GeV$^2$) &
$a_1^{I,n}$(GeV$^2$)&$a_2^{I,n}$(GeV$^2$) \\ \hline
{} & 1 & $-$4.66 & 8.45 & $-$4.47 & 8.37 \\
{} & 2 & $\,\,\,\,$8.489277 &$\,-$17.50252 &$\quad\,$7.4739945 &$-$15.8900\\
$I=1$ & 3 & $\,-$9.623356 &$\,\,\,\,$ 13.46278 & $-$8.050218 &
$\quad\,$10.93951 \\
{} & 4 & $\,\,\,\,$ 7.065036 &$\,\,-$7.310033 &$\quad$6.091668 &
$\quad\,-$6.071918 \\
{} & 5 & $-$0.140 & 1.36 & $-$0.118 & 1.11 \\ \hline
{} & 1 &$\quad$0.899887 & 0.02568286 & 0.8762127 & 0.09468219 \\
{} & 2 & $-$3.625433 & 0.5913331 & $-$3.514823 &$\,\,\,\,$0.3151743 \\
$I=0$ & 3 &$\,\,\,\,$7.385961 &$-$2.033127 &$\,\,\,\,\,$6.954618 &
$-$1.526721
\\
{} & 4 & $-$3.934473 & $-$1.019970 & $-$3.579443 & $-$2.125879 \\
{} & 5 & $-$1.028184 & 2.582630 & $-$1.038278 &$\,\,\,\,\,$3.394527
\\ \hline
\end{tabular}
\end{table}
\begin{table}
\caption{The coefficients $c_{i,j}^{QCD,I}$ of the QCD terms for the cases
I and II determined by the $\chi^2$ analysis.}
\begin{tabular}{cc|cll}\hline
{} & {} & {} &case I & {} \\ \hline
isospin & $i$ & $c_{i,2}^{QCD,I}$ & $c_{i,3}^{QCD,I}$ & $c_{i,4}^{QCD,I}$ \\
\hline
$I=1$ & 1 &$\,\,$ 0.5505731 & $-$4.12 & $-$6.50 \\
{} & 2 & 3.758361 & $-$0.4002$\times 10^2$ & 0.6224$\times 10^2$ \\ \hline
$I=0$ & 1 & $\,\,$1.108707 & $-$2.76 & $-$0.7045 $\times 10^2$ \\
{} & 2 & $-$5.215940 & 0.6706$\times 10^2$ & $-$0.19908$ \times 10^3$ \\
\hline
\hline
{} & {} & {} & case II & {} \\ \hline
isospin & $i$ & $c_{i,2}^{QCD,I}$ & $c_{i,3}^{QCD,I}$ & $c_{i,4}^{QCD,I}$ \\
\hline
$I=1$ & 1 & $-$0.7186148$\times 10^{-1}$ & $\quad$1.48 & $-$6.99 \\
{} & 2 & 4.252983 & $-$0.4375 $\times 10^2$ & 0.5543 $\times 10^2$ \\
\hline
$I=0$ & 1 &$\quad$ 0.8918455 & $-$1.10 & $-$0.6787000 $\times 10^2$ \\
{} & 2 & $-$5.617625 & 0.7029$\times 10^2$ & $-$0.19551$ \times 10^3$ \\
\hline
\end{tabular}
\end{table}
We illustrate in Figs.\ 1 - 9 the calculated results for the form factors.
The results
for the Case I is given by the solid curve and Case II by the dashed one.
Figs.\ 1 - 4 the results for the space-like momentum are illustrated: Fig.\ 1
the proton magnetic form factors $G_M^p/\mu_pG_D$, Fig.\ 2 proton electric
form factor $G_E^p/G_D$, Fig.\ 3 the neutron magnetic form factor
$G_M^n/\mu_n$ and Fig.\ 4 the neutron electric form factor. In Fig.\ 5 we
illustrate the ratio of proton electric and proton magnetic form factors
$\mu_pG_E^p/G_M^p$. We find that $G_E^p=0$ at $Q^2=6.57$ (GeV/c)$^2$ for the
case I and $Q^2=6.79$ (GeV/c)$^2$ for the case II. The form factor for the
time-like momentum $|G|$ is
given in Fig.\ 6 for the proton and in Fig.\ 7 for the neutron. The result for
the proton form factor agrees with the experimental data, but for the neutron
the calculated one becomes larger than the experiments for large $Q^2$.
In Fig.\ 8 we compare the calculated result for the neutron magnetic form
factor $G_M^n/\mu_nG_D$ with the recent experiments. The solid curve agrees
with the experimental data very well. The dashed one becomes a little
larger than the result obtained by the CLAS collaboration. However, the
deviation is not very large. In Fig.\ 9 we illustrate the result for
$|G_E^p/\mu_pG_M^p|$ for the time-like momentum. There seems to be some
discrepancy between the experimental data: The ratio obtained by
Bardin el al. \cite{bardinGptm} is smaller than that of Aubert et al.
\cite{aubert|GEGM|}. Our result coincides with the result of Bardin et al.
for small $Q^2$ and that of Aubert el al. for large $Q^2$.
\begin{figure}
\begin{center}
\includegraphics[width=.5\linewidth]{Fig1.eps}
\end{center}
\caption{Proton magnetic form factor for the space-like momentum. The solid
curve is the result for case I and the dashed one for the case II.}
\end{figure}%
\begin{figure}
\begin{center}
\includegraphics[width=.5\linewidth]{Fig2.eps}
\end{center}
\caption{Proton electric form factor for the space-like momentum. The solid
curve is the result for case I and the dashed one for the case II.}
\end{figure}%
\begin{figure}
\begin{center}
\includegraphics[width=.5\linewidth]{Fig3.eps}
\end{center}
\caption{Neutron magnetic form factors for the space-like momentum. The solid
curve is the result for case I and the dashed one for the case II.}
\end{figure}%
\begin{figure}
\begin{center}
\includegraphics[width=.5\linewidth]{Fig4.eps}
\end{center}
\caption{Neutron electric form factor for the space-like momentum. The solid
curve is the result for case I and the dashed one for the case II.}
\end{figure}%
\begin{figure}
\begin{center}
\includegraphics[width=.5\linewidth]{Fig5.eps}
\end{center}
\caption{ Ratio of the electric and magnetic form factors of proton for the
space-like momentum. The solid
curve is the result for case I and the dashed one for the case II.}%
\end{figure}
\begin{figure}
\begin{center}
\includegraphics[width=.5\linewidth]{Fig6.eps}
\end{center}
\caption{Proton form factors for the time-like momentum. The solid
curve is the result for case I and the dashed one for the case II.}
\end{figure}%
\begin{figure}
\begin{center}
\includegraphics[width=.5\linewidth]{Fig7.eps}
\end{center}
\caption{The neutron form factor for the time-like momentum. The solid
curve is the result for case I and the dashed one for the case II.}
\end{figure}%
\begin{figure}
\begin{center}
\includegraphics[width=.5\linewidth]{Fig8.eps}
\end{center}
\caption{Neutron magnetic form factor for the space-like momentum
in the few (GeV/c)$^2$ region. The solid
curve is the result for case I and the dashed one for the case II.}
\end{figure}%
\begin{figure}
\begin{center}
\includegraphics[width=.5\linewidth]{Fig9.eps}
\end{center}
\caption{$|G_E^p/\mu_pG_M^p|$ for the time-like momentum. The solid
curve is the result for case I and the dashed one for the case II.}
\end{figure}%
\section{Concluding Remarks}
The experimental data for the neutron magnetic form factor for the
space-like momentum with $Q^2=1.4 - 4.8$ (GeV/c)$^2$ \cite{lachnietGMn},
mentioned in Sec.\ 1, are reproduced very well by our calculation.
The absorptive parts of the form factors for the asymptotic region
are approximated by the
power series in the effective coupling constant of QCD, which begins
$O(\alpha_R^2)$ as is given in (\ref{qcd2}). We have taken three terms in
the expansions; the terms of order up to $O(\alpha_R^4)$ are
necessary to reproduce the experiments as in the case of deep inelastic
electron scattering processes.
It is remarked here that the electromagnetic form factor of bosons, both
for the space-like and time-like momentums, can be explained with recourse
to the superconvergent dispersion relation with the QCD constraints \cite{nw1}.
For the electric form factor of proton there are deviation of the dispersion
theoretical calculation from the
experimental data for large $Q^2$, where the data were obtained
by using the Rosenbluth formula. The discrepancy may imply the necessity of
correction
of two photon processes to the experimental data
\cite{borisyuk} \cite{belushkin}.
We used the experimental data for the helicity amplitudes obtained by
H\"oher and Schopper in which the contribution from the $\rho$ meson
is included. As their data are limited to low $t \,\,(\le 0.779$ (GeV/c)$^2$),
we do not have sufficient
data for the region $s \le 4m_N^2$. We
supplemented the unphysical region for $I=1$ state by introducing vector bosons
with the mass, $m_V \stackrel{\large <}{_{\sim}}1.4$ GeV/c$^2$. For the isoscalar state
we also introduced a vector boson with the mass about $1.2$ GeV/c$^2$.
In our calculation we treated all of the vector boson masses and widths as
parameters. If they are kept
at experimental values, we get poor results. The superconvergence
conditions are so strong that the value of $\chi^2$ is very sensitive
to the mass and width. The masses are obtained to be smaller
than the experimental value and the existence of vector bosons with the masses
around 1.2 $\sim$ 1.4 GeV/c$^2$ are necessary both for the $I = 1$ and
$I = 0$ states.
To conclude the paper we remark on the mass around 1.2 GeV/c$^2$. We have
introduced the vector boson to supplement the lack of information on the
the small $Q^2$. However, both
for $I=0$ and $I=1$ states there are indications of resonances observed
by the processes $e^{+}e^{-}\to \eta \pi^{+}\pi^{-}$,
$\gamma p \to \omega \pi^{0} p$ and $B \to D^{*}\omega\pi^{-}$ \cite{komada}.
Incorporation of further resonances may improve results for
the time-like momentum. \\
The authors wish to express gratitude to Professor M. Ishida for the valuable
discussions and comments. We also would like to thank Dr. T. Komada for the
information on the vector bosons with the mass around 1.2 GeV/c$^2$.
|
\section{Introduction}
We study a general class of $K$-dimensional coding methods for data drawn
from a distribution $\mu $ on the unit ball of a Hilbert space $H$. These
methods encode a data point $x\sim \mu $ as a vector $\hat{y}\in \mathbb{R}^{K}$, according to the formula
\begin{equation*}
\hat{y}=\arg \min_{y\in Y}\left\Vert x-Ty\right\Vert ^{2},
\end{equation*}
where $Y\subseteq \mathbb{R}^{K}$ is a prescribed set of \textit{codes}
(called the \emph{codebook}), which we can always assume to span $\mathbb{R}^{K}$, and $T:\mathbb{R}^{K}\rightarrow H$ is a linear map, which defines a
particular \textit{implementation} of the codebook. It embeds the codebook $Y
$ in $H$ and yields the set $T\left( Y\right) $ of exactly codable patterns.
If $\hat{y}$ is the code found for $x$ then $\hat{x}=T\hat{y}$ is the
reconstructed data point. The quantity
\begin{equation*}
f_{T}\left( x\right) =\min_{y\in Y}\left\Vert x-Ty\right\Vert ^{2}
\end{equation*}%
is called the reconstruction error.
Given a codebook $Y$ and a finite number of independent observations $%
x_{1},\dots ,x_{m}\sim \mu $, a common sense approach searches for an
implementation ${\hat T}$ which is optimal on average over the
observed points, that is
\begin{equation}
{\hat T}=\arg \min_{T\in \mathcal{T}}\frac{1}{m}\sum_{i=1}^{m}f_{T}%
\left( x_{i}\right) , \label{main algorithm}
\end{equation}%
where $\mathcal{T}$ denotes some class of linear maps $T:\mathbb{R}%
^{K}\rightarrow H$. As we shall see, this framework is general enough to
include principal component analysis, $K$-means clustering, non-negative
matrix factorization \cite{Lee 1999} and the sparse coding method as
proposed in \cite{Olshausen 1996}.
Whenever the codebook $Y$ is compact and $\mathcal{T}$ is bounded in the
operator norm this approach is justified by the following high-probability,
uniform bound on the expected reconstruction error.
\begin{theorem}
\label{Theorem general} Suppose that $Y$ is a closed subset of the unit ball
of $\mathbb{R}^{K}$, that there is $c\geq 1$ such that $\left\Vert
T\right\Vert _{\infty }\leq c$ for all $T\in \mathcal{T}$ and that $\delta
\in \left( 0,1\right) $. Then with probability at least $1-\delta $ in the
observed data $x_1,\dots,x_m \sim \mu$ we have for every $T\in \mathcal{T}$
that
\begin{equation*}
\mathbb{E}_{x\sim \mu }f_{T}\left( x\right) -\frac{1}{m}\sum_{i=1}^{m}f_{T}%
\left( x_{i}\right) \leq 6c^{2}K^{2}\sqrt{\frac{\pi }{m}}+c^2\sqrt{\frac{%
8\ln 1/\delta }{m}}.
\end{equation*}%
The bound is two-sided in the sense that also with probability at least $%
1-\delta $ we have for every $T\in \mathcal{T}$ that%
\begin{equation*}
\frac{1}{m}\sum_{i=1}^{m}f_{T}\left( x_{i}\right) -\mathbb{E}_{x\sim \mu
}f_{T}\left( x\right) \leq 6c^{2}K^{2}\sqrt{\frac{\pi }{m}}+c^{2}\sqrt{\frac{%
8\ln 1/\delta }{m}}.
\end{equation*}
\end{theorem}
Any compact subset of $\mathbb{R}^{K}$ can of course be down-scaled to be
contained in the unit ball, and the scaling factor can be absorbed in $c$,
so that the above result is applicable to any compact codebook.
The theorem implies a bound on the excess risk: let $T_{0}\in
\mathcal{T} $ be a minimizer of the expected reconstruction error
within the set $\mathcal{T}$. It
follows from the definition of ${\hat T}$ and the above result that
the expected reconstruction error of ${\hat T}$ is with high
probability not more than $O\left( 1/\sqrt{m}\right) $ worse than
that of $T_{0}$.
This order in $m$ is optimal, as we know from existing lower bounds for $K$%
-means clustering \cite{Bartlett Lugosi 1998}. The above dependence on $K$
is, however, generally not optimal, and can be considerably improved with a
more careful analysis, if we are prepared to accept the slightly inferior
rate of $\sqrt{\ln m/m}$ in the sample size. To state this improvement define%
\begin{equation*}
\left\Vert \mathcal{T}\right\Vert _{Y}=\sup_{T\in \mathcal{T}}\left\Vert
T\right\Vert _{Y}=\sup_{T\in \mathcal{T}}\sup_{y\in Y}\left\Vert
Ty\right\Vert .
\end{equation*}%
We then have the following result.
\begin{theorem}
\label{Theorem main} Assume that $\left\Vert \mathcal{T}\right\Vert _{Y}\geq
1$ and that the functions $f_{T}$ for $T\in \mathcal{T}$, when restricted to
the unit ball of $H$, have range contained in $\left[ 0,b\right] $. Fix $%
\delta >0$.
Then with probability at least $1-\delta $ in the observed data $x_1,\dots,x_m \sim \mu$ we have for every $T\in \mathcal{T}$ that
\begin{equation*}
\mathbb{E}_{x\sim \mu }f_{T}\left( x\right) -\frac{1}{m}\sum_{i=1}^{m}f_{T}%
\left( x_{i}\right) \leq \frac{K}{\sqrt{m}}\left( 14\left\Vert \mathcal{T}%
\right\Vert _{Y}+\frac{b}{2}\sqrt{\ln \left( 16m\left\Vert \mathcal{T}%
\right\Vert _{Y}^{2}\right) }\right) +b\sqrt{\frac{\ln 1/\delta }{2m}}.
\end{equation*}%
The bound is two sided in the same sense as the previous result.
\end{theorem}
Both results immediately imply uniform convergence in probability. We are
not aware of other results for nonnegative matrix factorization \cite{Lee
1999} or the sparse coding techniques as in \cite{Olshausen 1996}.
Before proving our results, we will illustrate their implications in some cases
of interest. It turns out that the dependence on $K$ in Theorem \ref{Theorem
main} adapts to the specific situation under consideration.
A preliminary version of this paper appeared in the proceedings of the
2008 Algorithmic Learning Theory Conference \cite{MauPon}. The new version contains Theorem
\ref{Theorem general} and a simplified proof of
Theorem \ref{Theorem main} with improved constants.
\section{Examples of coding schemes\label{section examples}}
Several coding schemes can be expressed in our framework. We describe some of these methods and how our result applies.
\subsection{Principal component analysis\label{subsection pca}}
Principal component analysis (PCA) seeks a $K$-dimensional orthogonal
projection which maximizes the projected variance and then uses this
projection to encode future data. A projection $P$ can be expressed as $%
TT^{\ast }$ where $T$ is an isometry which maps $\mathbb{R}^{K}$ to the
range of $P$. Since
\begin{equation*}
\left\Vert Px\right\Vert ^{2}=\left\Vert x\right\Vert ^{2}-\Vert x-Px\Vert
^{2}=\Vert x\Vert ^{2}-\min_{y\in \mathbb{R}^{K}}\left\Vert x-Ty\right\Vert
^{2}
\end{equation*}%
finding $P$ to maximize the true or empirical expectation of $\Vert
Px\Vert ^{2}$ is equivalent to finding $T$ to minimize the
corresponding expectation of $\min_{y\in \mathbb{R}^{K}}\left\Vert
x-Ty\right\Vert ^{2}$. We see that PCA is described by our framework
upon the identifications $Y=
\mathbb{R}^{K}$ and $\mathcal{T}$ is restricted to the class of isometries $%
T:\mathbb{R}^{K}\rightarrow H$. Given $T\in \mathcal{T}$ and $x\in H$ the
reconstruction error is
\begin{equation*}
f_{T}\left( x\right) =\min_{y\in \mathbb{R}^{K}}\left\Vert x-Ty\right\Vert
^{2}.
\end{equation*}%
If the data are constrained to be in the unit ball of $H$, as we generally
assume, then it is easily seen that we can take $Y$ to be the unit ball of $%
\mathbb{R}^{K}$ without changing any of the encodings. We can therefore
apply Theorem \ref{Theorem main} with $\left\Vert \mathcal{T}\right\Vert _{Y}=1$ and $b=1$.
This is besides the point however, because in the simple case of PCA much
better bounds are available (see \cite{ShaweTaylor 2005}, \cite{Zwald} and
Lemma \ref{Lemma PCA} below). In \cite{Zwald} local Rademacher averages are
used to give faster rates under certain circumstances.
An objection to PCA is, that generic codes have $K$ nonzero components,
while for practical and theoretical reasons sparse codes with much less than
$K$ nonzero components may be preferable \cite{Olshausen 1996}.
\subsection{$K$-means clustering or vector quantization\label{subsection
kmeans clustering}}
Here $Y=\left\{ e_{1},\dots ,e_{K}\right\} $, where the vectors $e_{k}$ form an
orthonormal basis of $\mathbb{R}^{K}$. An implementation $T$ now defines a
set of centers $\left\{ Te_{1},\dots ,Te_{K}\right\} $, the reconstruction
error is $\min_{k=1}^{K}\left\Vert x-Te_{k}\right\Vert ^{2}$ and a data
point $x$ is coded by the $e_{k}$ such that $Te_{k}$ is nearest to $x$. The
algorithm (\ref{main algorithm}) becomes
\begin{equation*}
{\hat T}=\arg \min_{T\in \mathcal{T}}\frac{1}{m}\sum_{i=1}^{m}%
\min_{k=1}^{K}\left\Vert x_{i}-Te_{k}\right\Vert ^{2}.
\end{equation*}%
It is clear that every center $Te_{k}$ has at most unit norm, so that $%
\left\Vert \mathcal{T}\right\Vert _{Y}=1$. Since all data points are in the
unit ball we have $\left\Vert x-Te_{k}\right\Vert ^{2}\leq 4$ so we can set $%
b=4$ and the bound in Theorem \ref{Theorem main} becomes
\begin{equation*}
\left( 14+2\sqrt{\ln \left( 16m\right) }\right) \frac{K}{\sqrt{m}}+\sqrt{%
\frac{8\ln \left( 1/\delta \right) }{m}}.
\end{equation*}
The order of this bound matches up to $\sqrt{\ln m}$ the order given in \cite%
{Biau Devroye Logosi 2006} or \cite{ShaweTaylor 2007}. To illustrate our
method we will also prove the bound
\begin{equation*}
\sqrt{18\pi }\frac{K}{\sqrt{m}}+\sqrt{\frac{8\ln \left( 1/\delta \right) }{m}%
}
\end{equation*}%
(Theorem \ref{Theorem Kmeans clustering}), which is essentially the same as
those in \cite{Biau Devroye Logosi 2006} or \cite{ShaweTaylor 2007}. There
is a lower bound of order $\sqrt{K/m}$ in \cite{Bartlett Lugosi 1998}, and
it is unknown which of the two bounds (upper or lower) is tight.
In $K$-means clustering every code has only one nonzero component, so that
sparsity is enforced in a maximal way. On the other hand this results in a
weaker approximation capability of the coding scheme.
\subsection{Nonnegative matrix factorization\label{subsection nonnegative
matrix factorization}}
Here $Y$ is the positive orthant in $\mathbb{R}^K$, that is the cone
\begin{equation*}
Y=\left\{y: y=(y_1,\dots,y_K),~y_k \geq 0, 1 \leq k \leq K \right\}.
\end{equation*}
A chosen map $T$ generates a cone $T\left( Y\right) \subset H$ onto which
incoming data is projected. In the original formulation by Lee and Seung
\cite{Lee 1999} it is postulated that both the data and the vectors $Te_{k}$
be contained in the positive orthant of some finite dimensional space, but
we can drop most of these restrictions, keeping only the requirement that $%
\left\langle Te_{k},Te_{l}\right\rangle \geq 0$ for $1\leq k,l\leq K$.
No coding will change if we require that $\left\Vert Te_{k}\right\Vert =1$
for all $1\leq k\leq K$ by a suitable normalization. The set $\mathcal{T}$
is then given by
\begin{equation*}
\mathcal{T}=\{T:T\in \mathcal{L}(\mathbb{R}^{K},H),~\left\Vert
Te_{k}\right\Vert =1,~\left\langle Te_{k},Te_{l}\right\rangle \geq 0,~1\leq
k,l\leq K\}.
\end{equation*}%
We can restrict $Y$ to its intersection with the unit ball in $\mathbb{R}^{K}
$ (see Lemma \ref{Little Lemma} below). We obtain that $\left\Vert \mathcal{T
}\right\Vert _{Y}=\sqrt{K}$. Hence, Theorem \ref{Theorem main} yields the
bound
\begin{equation*}
\frac{K}{\sqrt{m}}\left( 14\sqrt{K}+\frac{1}{2}\sqrt{\ln \left( 16mK\right) }%
\right) +\sqrt{\frac{\ln \left( 1/\delta \right) }{2m}}
\end{equation*}%
on the estimation error. We do not know of any other generalization bounds
for this coding scheme.
Nonnegative matrix factorization appears to encourage sparsity, but cases
have been reported where sparsity was not observed \cite{Li 2001}. In fact
this undesirable behavior should be generic for exactly codable data.
Various authors have therefore proposed additional constraints (\cite{Li
2001}, \cite{Hoyer 2004}). It is clear that additional constraints on $%
\mathcal{T}$ can only improve estimation and that the passage from $Y$ to a
subset can only improve our bounds, because the quantity $\|\mathcal{T}\|_Y$
would decrease.
\subsection{Sparse coding\label{subsection sparse coding}}
Another method arises by choosing the $\ell _{p}$-unit ball as a codebook.
Let $Y=\{y:y\in \mathbb{R}^{K},~\Vert y\Vert _{p}\leq 1\}$ and $\mathcal{T}
=\{T:\mathbb{R}^{K}\rightarrow H:\Vert Te_{k}\Vert \leq 1,1\leq k\leq K\}$.
We have
\begin{equation*}
\Vert Ty\Vert =\Vert \sum_{k=1}^{K}y_{k}Te_{k}\Vert \leq
\sum_{k=1}^{K}|y_{k}|\Vert Te_{k}\Vert \leq \left( \sum_{k=1}^{K}\Vert
Te_{k}\Vert ^{q}\right) ^{1/q}\leq K^{1/q}=K^{1-1/p}
\end{equation*}%
implying that $\Vert \mathcal{T}\Vert _{Y}\leq K^{1-1/p}$.
By the same argument as above all the $f_{T}$ have range contained in $\left[
0,1\right] $, so Theorem \ref{Theorem main} can be applied with $b=1$ to
yield the bound
\begin{equation*}
\frac{K}{\sqrt{m}}\left( 14K^{1-1/p}+\frac{1}{2}\sqrt{\ln \left(
16mK^{2-2/p}\right) }\right) +\sqrt{\frac{\ln \left( 1/\delta \right) }{2m}}
\end{equation*}%
on the estimation error. The best bound is obtained when $p=1$, and the
order in $K$ matches that of the bound for $K$-means clustering described
earlier.
The method for $p=1$ is similar to the sparse-coding method proposed by
Olshausen and Field \cite{Olshausen 1996}, with the difference that the term
$\Vert y\Vert _{1}$ is used as a penalty term instead of the hard constraint
$\Vert y\Vert _{1}\leq 1$. The method of Olshausen and Field \cite{Olshausen
1996} approximates with a compromise of geometric proximity and sparsity and
our result asserts that the observed value of this compromise generalizes to
unseen data if enough data have been observed.
\section{Proofs\label{section proofs}}
We first introduce some notation, conventions and auxiliary results. Then we
set about to prove Theorems \ref{Theorem general} and \ref{Theorem main}.
\subsection{Notation, definitions and auxiliary results\label{subsection
notation}}
Throughout $H$ denotes a Hilbert space. The term \textit{norm} and the
notation $\left\Vert \cdot \right\Vert $ and $\left\langle \cdot,\cdot
\right\rangle$ always refer to the Euclidean norm and inner product on $%
\mathbb{R}^{K}$ or on $H$. Other norms are characterized by subscripts. If $%
H_{1}$ and $H_{2}$ are any Hilbert spaces $\mathcal{L}\left(
H_{1},H_{2}\right)$ denotes the vector space of bounded linear
transformations from $H_{1}$ to $H_{2}$. If $H_{1}=H_{2}$ we just write $%
\mathcal{L}\left( H_{1}\right) = \mathcal{L}\left( H_{1},H_{1}\right)$. With
$\mathcal{U}\left(H_{1},H_{2}\right)$ we denote the set of isometries in $%
\mathcal{L}\left(H_{1},H_{2}\right)$, that is maps $U$ satisfying $%
\left\Vert Ux\right\Vert_{H_2}=\left\Vert x\right\Vert_{H_1}$ for all $x\in
H_{1}$.
We use $\mathcal{L}_{2}\left( H\right) $ for the set of Hilbert-Schmidt
operators on $H$, which becomes itself a Hilbert space with the inner
product $\left\langle T,S\right\rangle _{2}=$tr$\left( T^{\ast }S\right) $
and the corresponding (Frobenius) norm $\left\Vert \cdot \right\Vert _{2}$.
For $x\in H$ the rank-one operator $Q_{x}$ is defined by $%
Q_{x}z=\left\langle z,x\right\rangle x$. For any $T\in \mathcal{L}_{2}\left(
H\right) $ the identity
\begin{equation*}
\left\langle T^{\ast }T,Q_{x}\right\rangle _{2}=\left\Vert Tx\right\Vert
^{2}
\end{equation*}
is easily verified.
Suppose that $Y\subseteq \mathbb{R}^{K}$ spans $\mathbb{R}^{K}$.
It is easily verified that the quantity
\begin{equation*}
\left\Vert T\right\Vert _{Y}=\sup_{y\in Y}\left\Vert Ty\right\Vert
\end{equation*}
defines a norm on $\mathcal{L}\left(\mathbb{R}^{K},H\right)$.
We use the following well known result on covering numbers (see, for
example, Proposition 5 in \cite{Cucker Smale 2001}).
\begin{proposition}
\label{Proposition covering}Let $B$ be a ball of radius $r$ in an $N$%
-dimensional Banach space and $\epsilon >0$. There exists a subset $%
B_{\epsilon }\subset B$ such that $\left\vert B_{\epsilon }\right\vert \leq
\left( 4r/\epsilon \right) ^{N}$ and $\forall z\in B,\exists z^{\prime }\in
B_{\epsilon }$ with $d(z,z^{\prime }) \leq \epsilon $, where $d$ is the
metric of the Banach space.
\end{proposition}
The following concentration inequality, known as the bounded difference
inequality \cite{McDiarmid 1998}, goes back to the work of Hoeffding \cite%
{Hoeffding 1963}.
\begin{theorem}
\label{Theorem Bded Difference}Let $\mu _{i}$ be a probability measure on a
space $\mathcal{X}_{i}$, for $i=1,\dots,m$. Let $\mathcal{X} =\prod_{i=1}^{m}%
\mathcal{X}_{i} $ and $\mu =\otimes _{i=1}^{m}\mu_{i}$ be the product space
and product measure respectively. Suppose the function $\Psi :\mathcal{X}
\rightarrow \mathbb{R}$ satisfies
\begin{equation*}
\left\vert \Psi \left( \mathbf{x}\right) -\Psi \left( \mathbf{x}^{\prime
}\right) \right\vert \leq c_{i}
\end{equation*}
whenever $\mathbf{x}$ and $\mathbf{x}^{\prime } \in \mathcal{X}$ differ only in the $i$-th
coordinate, where $c_1,\dots,c_m$ are some positive parameters. Then
\begin{equation*}
\Pr_{\mathbf{x}\sim \mu }\left\{ \Psi \left( \mathbf{x}\right) -\mathbb{E}_{
\mathbf{x}^{\prime }\sim \mu }\Psi \left( \mathbf{x}^{\prime }\right) \geq
t\right\} \leq \exp \left( \frac{-2t^{2}}{\sum_{i=1}^{m}c_{i}^{2}}\right) .
\end{equation*}
\end{theorem}
Throughout $\sigma _{i}$ will denote a sequence of mutually independent
random variables, uniformly distributed on $\left\{-1,1\right\} $ and $%
\gamma_{i}$, $\gamma _{ij}$ will be (multiple indexed) sequences of mutually
independent Gaussian random variables, with zero mean and unit standard
deviation.
If $\mathcal{F}$ is a class of real-valued functions on a space $\mathcal{X}$
and $\mu $ a probability measure on $\mathcal{X}$ then for $m\in \mathbb{N} $
the Rademacher and Gaussian complexities of $\mathcal{F}$ w.r.t. $\mu $ are
defined (\cite{Ledoux 1991},\cite{Bartlett 2002}) as
\begin{eqnarray*}
\mathcal{R}_{m}\left( \mathcal{F},\mu \right) &=&\frac{2}{m}\mathbb{E}_{
\mathbf{x}\sim \mu ^{m}}\mathbb{E}_{\sigma }\sup_{f\in \mathcal{F}
}\sum_{i=1}^{m}\sigma _{i}f\left( x_{i}\right) \text{, } \\
\Gamma _{m}\left( \mathcal{F},\mu \right) &=&\frac{2}{m}\mathbb{E}_{\mathbf{%
\ x}\sim \mu ^{m}}\mathbb{E}_{\gamma }\sup_{f\in \mathcal{F}
}\sum_{i=1}^{m}\gamma _{i}f\left( x_{i}\right)
\end{eqnarray*}
respectively.
Appropriately scaled Gaussian complexities can be substituted for Rademacher
complexities, by virtue of the next Lemma. For a proof see, for example,
\cite[p. 97]{Ledoux 1991}.
\begin{lemma}
\label{Lemma Gauss dominates Rademacher}For $Y\subseteq \mathbb{R}^{k}$ we
have $\mathcal{R}\left( Y\right) \leq \sqrt{\pi /2}~\Gamma \left( Y\right) $.
\end{lemma}
The next result is known as Slepian's lemma (\cite{Slepian}, \cite{Ledoux
1991}).
\begin{theorem}
\label{Slepian Lemma}Let $\Omega $ and $\Xi $ be mean zero, separable
Gaussian processes indexed by a common set $\mathcal{S}$, such that%
\begin{equation*}
\mathbb{E}\left( \Omega _{s_{1}}-\Omega _{s_{2}}\right) ^{2}\leq \mathbb{E}%
\left( \Xi _{s_{1}}-\Xi _{s_{2}}\right) ^{2}\text{ for all }s_{1},s_{2}\in
\mathcal{S}\text{.}
\end{equation*}%
Then%
\begin{equation*}
\mathbb{E}\sup_{s\in \mathcal{S}}\Omega _{s}\leq \mathbb{E}\sup_{s\in
\mathcal{S}}\Xi _{s}.
\end{equation*}
\end{theorem}
The following result, which generalizes Theorem 8 in \cite{Bartlett 2002},
plays a central role in our proof.
\begin{theorem}
\label{Corollary Bound Finite Max Expectation} Let $\left\{ \mathcal{F}%
_{n}:1\leq n\leq N\right\} $ be a finite collection of $\left[ 0,b\right] $%
-valued function classes on a space $\mathcal{X}$, and $\mu $ a probability
measure on $\mathcal{X}$. Then $\forall \delta \in \left( 0,1\right) $ we
have with probability at least $1-\delta $ that%
\begin{equation*}
\max_{n\leq N}\sup_{f\in \mathcal{F}_{n}}\left[ \mathbb{E}_{x\sim \mu
}f\left( x\right) -\frac{1}{m}\sum_{i=1}^{m}f\left( x_{i}\right) \right]
\leq \max_{n\leq N}\mathcal{R}_{m}\left( \mathcal{F}_{n},\mu \right) +b\sqrt{%
\frac{\ln N+\ln \left( 1/\delta \right) }{2m}}.
\end{equation*}
\end{theorem}
\begin{proof}
Denote with $\Psi _{n}$ the function on $\mathcal{X}^{m}$ defined by%
\begin{equation*}
\Psi _{n}\left( \mathbf{x}\right) =\sup_{f\in \mathcal{F}_{n}}\left[ \mathbb{%
E}_{x\sim \mu }f\left( x\right) -\frac{1}{m}\sum_{i=1}^{m}f\left(
x_{i}\right) \right] ,\text{ }\mathbf{x}\in \mathcal{X}^{m}.
\end{equation*}%
By standard symmetrization (see, for example, \cite{van der Vaart 1996}) we
have $\mathbb{E}_{\mathbf{x}\sim \mu ^{m}}\Psi _{n}\left( \mathbf{x}\right)
\leq \mathcal{R}_{m}\left( \mathcal{F}_{n},\mu \right) \leq \max_{n\leq N}%
\mathcal{R}_{m}\left( \mathcal{F}_{n},\mu \right) $. Modifying one of the $%
x_{i}$ can change the value of any $\Psi _{n}\left( \mathbf{x}\right) $ by
at most $b/m$, so that by a union bound and the bounded difference
inequality (Theorem \ref{Theorem Bded Difference})
\begin{equation*}
\Pr \left\{ \max_{n\leq N}\Psi _{n}>\max_{n\leq N}\mathcal{R}_{m}\left(
\mathcal{F}_{n},\mu \right) +t\right\} \leq \sum_{n}\Pr \left\{ \Psi _{n}>%
\mathbb{E}\Psi _{n}+t\right\} \leq Ne^{-2m\left( t/b\right) ^{2}}.
\end{equation*}%
Solving $\delta =Ne^{-2m\left( t/b\right) ^{2}}$ for $t$ gives the result.%
\qed
\end{proof}
Notice that replacing the functions $f\in \mathcal{F}_{n}$ by $b-f$ does not
affect the Rademacher complexities, so the above result can be used in a
two-sided way.
The following lemma was used in Section \ref{subsection nonnegative matrix
factorization}.
\begin{lemma}
\label{Little Lemma}Suppose $\left\Vert x\right\Vert \leq 1$, $\left\Vert
c_{k}\right\Vert =1,$ $\left\langle c_{k},c_{l}\right\rangle \geq 0$, $y\in
\mathbb{R} ^{K}$, $y_{i}\geq 0$. If $y$ minimizes%
\begin{equation*}
h\left( y\right) =\left\Vert x-\sum_{k=1}^{K}y_{k}c_{k}\right\Vert ^{2},
\end{equation*}
then $\left\Vert y\right\Vert \leq 1$.
\end{lemma}
\begin{proof}
Assume that $y$ is a minimizer of $h$ and $\left\Vert y\right\Vert >1$.Then
\begin{equation*}
\left\Vert \sum_{k=1}^{K}y_{k}c_{k}\right\Vert ^{2}=\left\Vert y\right\Vert
^{2}+\sum_{k\neq l}y_{k}y_{l}\left\langle c_{k},c_{l}\right\rangle >1.
\end{equation*}
Let the real-valued function $f$ be defined by $f\left( t\right) =h\left(
ty\right) $. Then
\begin{eqnarray*}
f^{\prime }\left( 1\right) &=&2\left( \left\Vert
\sum_{k=1}^{K}y_{k}c_{k}\right\Vert ^{2}-\left\langle
x,\sum_{k=1}^{K}y_{k}c_{k}\right\rangle \right) \\
&\geq &2\left( \left\Vert \sum_{k=1}^{K}y_{k}c_{k}\right\Vert
^{2}-\left\Vert \sum_{k=1}^{K}y_{k}c_{k}\right\Vert \right) \\
&=&2\left( \left\Vert \sum_{k=1}^{K}y_{k}c_{k}\right\Vert -1\right)
\left\Vert \sum_{k=1}^{K}y_{k}c_{k}\right\Vert \\
&>&0\text{.}
\end{eqnarray*}
So $f$ cannot have a minimum at $1$, whence $y$ cannot be a minimizer of $h$%
. \qed
\end{proof}
\subsection{Proof of the main results}
We now fix a spanning codebook $Y\subseteq \mathbb{R}^{K}$ and recall that,
for $T\in \mathcal{L}\left( \mathbb{R}^{K},H\right) $, we had introduced the
notation
\begin{equation*}
f_{T}\left( x\right) =\inf_{y\in Y}\left\Vert x-Ty\right\Vert ^{2},x\in H%
\text{.}
\end{equation*}%
Our principal object of study is the function class
\begin{equation*}
\mathcal{F}=\left\{ f_{T}:T\in \mathcal{T}\right\} \text{,}
\end{equation*}%
where $\mathcal{T}\subset \mathcal{L}\left( \mathbb{R}^{K},H\right) $ is
some fixed set of candidate implementations of our coding scheme. We first
address the rather general Theorem \ref{Theorem general} which can be
treated in parallel to the case of $K$-means clustering. We begin with a
technical lemma.
\begin{lemma}
\label{Lemma Gaussian Complexity bounds}Suppose that
\begin{enumerate}
\item $\left( e_{k}: 1\leq k\leq K \right) $ is an orthonormal basis of $%
\mathbb{R}^{K}$;
\item $\mathcal{T}$ is the class of linear operators $T:\mathbb{R}%
^{K}\rightarrow H$ with $\left\Vert Te_{k}\right\Vert \leq c$;
\item $\left( x_{i}: 1 \leq i \leq m \right) $ is a sequence $x_{i}\in H$, $%
\left\Vert x_{i}\right\Vert \leq 1$;
\item $\left( \gamma _{ik}: 1 \leq i \leq m,~1\leq k\leq K \right) $ and $%
\left( \gamma _{ikl}:1\leq i\leq m,~1\leq k,l\leq K \right)$ are
orthogaussian sequences.
\end{enumerate}
Then the following three inequalities hold
\begin{eqnarray*}
\mathbb{E}_{\gamma }\sup_{T\in \mathcal{T}}\sum_{i=1}^{m}\sum_{k=1}^{K}%
\gamma _{ik}\left\langle x_{i},Te_{k}\right\rangle &\leq &cK\sqrt{m} \\
\mathbb{E}_{\gamma }\sup_{T\in \mathcal{T}}\sum_{i=1}^{m}\sum_{k=1}^{K}%
\gamma _{ik}\left\Vert Te_{k}\right\Vert ^{2} &\leq &c^{2}K\sqrt{m} \\
\mathbb{E}_{\gamma }\sup_{T\in \mathcal{T}}\sum_{i=1}^{m}\sum_{k,l=1}^{K}%
\gamma _{ikl}\left\langle Te_{k},Te_{l}\right\rangle &\leq &c^{2}K^{2}\sqrt{m%
}.
\end{eqnarray*}
\end{lemma}
\begin{proof}
Using Cauchy-Schwarz' and Jensen's inequalities and the orthogaussian
properties of the $\gamma _{ik}$, we get%
\begin{equation*}
\mathbb{E}_{\gamma }\sup_{T\in \mathcal{T}}\sum_{k=1}^{K}\sum_{i=1}^{m}%
\gamma _{ik}\left\langle x_{i},Te_{k}\right\rangle \leq c\mathbb{E}_{\gamma
}\sum_{k=1}^K\left\Vert \sum_{i=1}^m\gamma _{ik}x_{i}\right\Vert \leq cK%
\sqrt{m}
\end{equation*}%
which is the first inequality. Similarly we obtain%
\begin{eqnarray*}
\mathbb{E}_{\gamma }\sup_{T\in \mathcal{T}}\sum_{k=1}^{K}\sum_{i=1}^{m}%
\gamma _{ik}\left\Vert Te_{k}\right\Vert ^{2} &\leq &c^{2}\mathbb{E}_{\gamma
}\sum_{k=1}^{K}\left\vert \sum_{i=1}^{m}\gamma _{ik}\right\vert \leq c^{2}K%
\sqrt{m} \\
\mathbb{E}_{\gamma }\sup_{T\in \mathcal{T}}\sum_{k,l=1}^{K}\sum_{i=1}^{m}%
\gamma _{ikl}\left\langle Te_{k},Te_{l}\right\rangle &\leq &c^{2}\mathbb{E}%
_{\gamma }\sum_{k,l=1}^{K}\left\vert \sum_{i=1}^{m}\gamma _{ikl}\right\vert
\leq c^{2}K^{2}\sqrt{m}.
\end{eqnarray*}
\qed
\end{proof}
\begin{proposition}
\label{Proposition general complexities}Suppose that the probability measure
$\mu $ is supported on the unit ball of $H$, that $\left\{ e_{k}: 1 \leq k
\leq K\right\} $ is an orthonormal basis of $\mathbb{R}^{K}$ and that $%
\mathcal{T}$ is a class of linear operators $T:\mathbb{R}^{K}\rightarrow H$
with $\left\Vert Te_{k}\right\Vert \leq c$ for $1 \leq k \leq K$, with $%
c\geq 1$. Let $Y$ be a nonempty closed subset of the unit ball in $\mathbb{R}%
^{K}$ and
\begin{equation*}
\mathcal{F}_{Y}=\left\{ x\in H\mapsto \min_{y\in Y}\left\Vert
x-Ty\right\Vert ^{2}:T\in \mathcal{T}\right\} .
\end{equation*}%
Then
\begin{equation*}
\mathcal{R}\left( \mathcal{F}_{Y},\mu \right) \leq 6c^{2}K^{2}\sqrt{\frac{%
\pi }{m}}.
\end{equation*}%
and if $Y=\left\{ e_{k}: 1 \leq k \leq K\right\} $ then the bound improves to%
\begin{equation*}
\mathcal{R}\left( \mathcal{F}_{Y},\mu \right) \leq c^{2}K\sqrt{\frac{18\pi }{%
m}}.
\end{equation*}
\end{proposition}
\begin{proof}
By Lemma \ref{Lemma Gauss dominates Rademacher} it suffices to bound the
corresponding Gaussian averages, which we shall do using Slepian's Lemma
(Theorem \ref{Slepian Lemma}). First fix a sample $\mathbf{x}$ and define
Gaussian processes $\Omega $ and $\Xi $ indexed by $\mathcal{T}$
\begin{eqnarray*}
\Omega _{T} &=&\sum_{i}\gamma _{i}\min_{y}\left\Vert x_{i}-Ty\right\Vert ^{2}%
\text{ and} \\
\Xi _{T} &=&\sqrt{8}\sum_{ik}\gamma _{ik}\left\langle
x_{i},Te_{k}\right\rangle +\sqrt{2}\sum_{ilk}\gamma _{ilk}\left\langle
Te_{l},Te_{k}\right\rangle .
\end{eqnarray*}%
Suppose $T_{1},T_{2}\in \mathcal{T}$. For any $x\in H$ we have, using $%
\left( a+b\right) ^{2}\leq 2a^{2}+2b^{2}$ and Cauchy-Schwarz
\begin{eqnarray*}
&&\left( \min_{y\in Y}\left\Vert x-T_{1}y\right\Vert ^{2}-\min_{y}\left\Vert
x-T_{2}y\right\Vert ^{2}\right) ^{2} \\
&\leq &\left( \max_{y\in Y}\left\Vert x-T_{1}y\right\Vert ^{2}-\left\Vert
x-T_{2}y\right\Vert ^{2}\right) ^{2} \\
&\leq &8\max_{y\in Y}\left( \sum_{k}y_{k}\left\langle x,\left(
T_{1}-T_{2}\right) e_{k}\right\rangle \right) ^{2}+2\max_{y\in Y}\left(
\sum_{kl}y_{k}y_{l}\left\langle e_{k},\left( T_{1}^{\ast }T_{1}-T_{2}^{\ast
}T_{2}\right) e_{l}\right\rangle \right) ^{2} \\
&\leq &8\sum_{k}\left( \left\langle x,T_{1}e_{k}\right\rangle -\left\langle
x,T_{2}e_{k}\right\rangle \right) ^{2}+2\sum_{kl}\left( \left\langle
T_{1}e_{k},T_{1}e_{l}\right\rangle -\left\langle
T_{2}e_{k},T_{2}e_{l}\right\rangle \right) ^{2}.
\end{eqnarray*}%
We therefore have%
\begin{eqnarray*}
\mathbb{E}\left( \Omega _{T_{1}}-\Omega _{T_{2}}\right) ^{2}
&=&\sum_{i}\left( \min_{y}\left\Vert x_{i}-T_{1}y\right\Vert
^{2}-\min_{y}\left\Vert x_{i}-T_{2}y\right\Vert ^{2}\right) ^{2} \\
&\leq &8\sum_{ik}\left( \left\langle x_{i},T_{1}e_{k}\right\rangle
-\left\langle x_{i},T_{2}e_{k}\right\rangle \right) ^{2}+2\sum_{ikl}\left(
\left\langle T_{1}e_{k},T_{1}e_{l}\right\rangle -\left\langle
T_{2}e_{k},T_{2}e_{l}\right\rangle \right) ^{2} \\
&=&\mathbb{E}\left( \Xi _{T_{1}}-\Xi _{T_{2}}\right) ^{2}.
\end{eqnarray*}%
So, by Slepian's Lemma and the first and last inequalities in Lemma \ref%
{Lemma Gaussian Complexity bounds}%
\begin{eqnarray*}
\mathbb{E}\sup_{T\in \mathcal{T}}\Omega _{T} &\leq &\mathbb{E}\sup_{T\in
\mathcal{T}}\Xi _{T} \\
&\leq &\sqrt{8}\mathbb{E}\sup_{T\in \mathcal{T}}\sum_{ik}\gamma
_{ik}\left\langle x_{i},Te_{k}\right\rangle +\sqrt{2}\mathbb{E}\sup_{T\in
\mathcal{T}}\sum_{ilk}\gamma _{ilk}\left\langle Te_{l},Te_{k}\right\rangle \\
&\leq &cK\sqrt{8m}+c^{2}K^{2}\sqrt{2m}.
\end{eqnarray*}%
Multiply by $\sqrt{2\pi }/m$ to get a bound on the Rademacher complexity of
\begin{equation*}
\mathcal{R}\left( \mathcal{F}_{Y},\mu \right) \leq 4cK\sqrt{\frac{\pi }{m}}%
+2c^{2}K^{2}\sqrt{\frac{\pi }{m}}\leq 6c^{2}K^{2}\sqrt{\frac{\pi }{m}}.
\end{equation*}%
To obtain the second conclusion we improve the bound on the Gaussian
average. With $\Omega _{T}$ as above we set
\begin{equation*}
\Xi _{T}=\sum_{i=1}^{m}\sum_{k=1}^{K}\gamma _{ik}\left\Vert
x_{i}-Te_{k}\right\Vert ^{2}\text{.}
\end{equation*}
Now we have for $T_{1},T_{2}\in \mathcal{T}$ that
\begin{eqnarray*}
\mathbb{E}\left(\Omega_{T_{1}}-\Omega _{T_{2}}\right)^{2} &=&
\sum_{i=1}^{m}\left( \min_{k=1}^K \Vert x_{i}-T_{1}e_{k}\Vert
^{2}-\min_{k=1}^K\Vert x_{i}-T_{2}e_{k}\Vert ^{2}\right) ^{2} \\
&\leq &\sum_{i=1}^{m}\max_{k=1}^{K}\left( \left\Vert
x_{i}-T_{1}e_{k}\right\Vert ^{2}-\left\Vert x_{i}-T_{2}e_{k}\right\Vert
^{2}\right) ^{2} \\
&\leq &\sum_{i=1}^{m}\sum_{k=1}^{K}\left( \left\Vert
x_{i}-T_{1}e_{k}\right\Vert ^{2}-\left\Vert x_{i}-T_{2}e_{k}\right\Vert
^{2}\right) ^{2} \\
&=&\mathbb{E}\left( \Xi _{T_{1}}-\Xi _{T_{2}}\right) ^{2}\text{.}
\end{eqnarray*}%
Again with Slepian's Lemma and the triangle inequality
\begin{align*}
\mathbb{E}_{\gamma }\sup_{T\in \mathcal{T}}\Omega _{T}& \leq \mathbb{E}%
_{\gamma }\sup_{T\in \mathcal{T}}\Xi _{T}=\mathbb{E}_{\gamma }\sup_{T\in
\mathcal{T}}\sum_{i=1}^{m}\sum_{k=1}^{K}\gamma _{ik}\left\Vert
x_{i}-Te_{k}\right\Vert ^{2} \\
& \leq 2\mathbb{E}_{\gamma }\sup_{T\in \mathcal{T}}\sum_{i=1}^{m}%
\sum_{k=1}^{K}\gamma _{ik}\left\langle x_{i},Te_{k}\right\rangle +\mathbb{E}%
_{\gamma }\sup_{T\in \mathcal{T}}\sum_{i=1}^{m}\sum_{k=1}^{K}\gamma
_{ik}\left\Vert Te_{k}\right\Vert ^{2} \\
& \leq 3c^{2}K\sqrt{m},
\end{align*}%
where the last inequality follows from the first two inequalities in Lemma %
\ref{Lemma Gaussian Complexity bounds}. Multiply by $\sqrt{2\pi }/m$ as above%
\qed
\end{proof}
Theorem \ref{Theorem general} follows from observing that the functions in $%
\mathcal{F}$ map to $\left[ 0,4c^{2}\right] $ and combining the above bound
on the Rademacher complexity with Theorem \ref{Corollary Bound Finite Max
Expectation} with $N=1$ and $b=4$.
The second conclusion of the proposition yields a bound for $K$-means
clustering, corresponding to the choices $Y=\left\{ e_{1},\dots
,e_{K}\right\} $ and $\mathcal{T}=\left\{ T:\left\Vert Te_{k}\right\Vert
\leq 1,~1\leq k\leq K\right\} $. As already noted in Section \ref{subsection
kmeans clustering} the vectors $Te_{k}$ define the cluster centers. With
Theorem \ref{Corollary Bound Finite Max Expectation} we obtain
\begin{theorem}
\label{Theorem Kmeans clustering}For every $\delta >0$ with probability
greater $1-\delta $ in the sample $\mathbf{x}\sim \mu ^{m}$ we have for all $%
T\in \mathcal{T}$
\begin{equation*}
\mathbb{E}_{x\sim \mu }\min_{k=1}^{K}\left\Vert x-Te_{k}\right\Vert ^{2}\leq
\frac{1}{m}\sum_{i=1}^{m}\min_{k=1}^{K}\left\Vert x_{i}-Te_{k}\right\Vert
^{2}+K\sqrt{\frac{18\pi }{m}}+\sqrt{\frac{8\ln \left( 1/\delta \right) }{m}}.
\end{equation*}%
\bigskip
\end{theorem}
To prove Theorem \ref{Theorem main} a more subtle approach is necessary. The
idea is the following: every implementing map $T\in \mathcal{T}$ can be
factored as $T=US$, where $S$ is a $K\times K$ matrix, $S\in \mathcal{L}%
\left( \mathbb{R}^{K}\right) $, and $U$ is an isometry, $U\in {\mathcal{U}}(%
\mathbb{R}^{K},H)$. Suitably bounded $K\times K$ matrices form a compact,
finite dimensional set, the complexity of which can be controlled using
covering numbers, while the complexity arising from the set of isometries
can be controlled with Rademacher and Gaussian averages. Theorem \ref%
{Corollary Bound Finite Max Expectation} then combines these complexity
estimates.
For fixed $S\in \mathcal{L}\left(\mathbb{R}^{K}\right) $ we denote
\begin{equation*}
\mathcal{G}_{S}=\left\{ f_{US}:U\in \mathcal{U}\left(\mathbb{R}^{K},H\right)
\right\} .
\end{equation*}
Recall the notation $\left\Vert \mathcal{T}\right\Vert _{Y}=\sup_{T\in
\mathcal{T}}\left\Vert T\right\Vert _{Y}=\sup_{T\in \mathcal{T}}\sup_{y\in
Y}\left\Vert Ty\right\Vert $. With $\mathcal{S}$ we denote the set of $%
K\times K$ matrices
\begin{equation*}
\mathcal{S}=\left\{ S\in \mathcal{L}\left( \mathbb{R}^{K}\right) :\left\Vert
S\right\Vert _{Y}\leq \left\Vert \mathcal{T} \right\Vert _{Y}\right\} \text{.%
}
\end{equation*}
\begin{lemma}
\label{Lemma Key} Assume $\left\Vert \mathcal{T}\right\Vert _{Y}\geq 1$,
that the functions in $\mathcal{F}$, when restricted to the unit ball of $H$%
, have range contained in $\left[ 0,b\right] $, and that the measure $\mu $
is supported on the unit ball of $H$. Then with probability at least $%
1-\delta $ we have for all $T\in \mathcal{T}$ that
\begin{multline*}
\mathbb{E}_{x\sim \mu }f_{T}\left( x\right) -\frac{1}{m}\sum_{i=1}^{m}f_{T}
\left( x_{i}\right) \\
\leq \sup_{S\in \mathcal{S}}\mathcal{R}_{m}\left( \mathcal{G}_{S},\mu
\right) +\frac{bK}{2}\sqrt{\frac{\ln \left( 16m\left\Vert \mathcal{T}
\right\Vert _{Y}^{2}\right) }{m}}+\frac{8\left\Vert \mathcal{T}\right\Vert
_{Y}}{\sqrt{m}}+b\sqrt{\frac{\ln \left( 1/\delta \right) }{2m}}.
\end{multline*}
\end{lemma}
\begin{proof}
Fix $\epsilon >0$. The set $\mathcal{S}$ is the ball of radius $\left\Vert
\mathcal{T}\right\Vert _{Y}$ in the $K^{2}$-dimensional Banach space $\left(
\mathcal{L}\left(\mathbb{R}^{K}\right) ,\left\Vert .\right\Vert _{Y}\right) $
so by Proposition \ref{Proposition covering} we can find a subset $\mathcal{S%
}_{\epsilon }\subset \mathcal{S}$, of cardinality $\left\vert \mathcal{S}%
_{\epsilon }\right\vert \leq \left( 4\left\Vert \mathcal{T}\right\Vert
_{Y}/\epsilon \right) ^{K^{2}} $ such that every member of $\mathcal{S}$ can
be approximated by a member of $\mathcal{S}_{\epsilon }$ up to distance $%
\epsilon $ in the norm $\left\Vert .\right\Vert _{Y}$.
We claim that for all $T\in \mathcal{T}$ there exist $U\in {\mathcal{U}}(
\mathbb{R}^{K},H)$ and $S_{\epsilon }\in \mathcal{S}_{\epsilon }$ such that
\begin{equation*}
\left\vert f_{T}\left( x\right) -f_{US_{\epsilon }}\left( x\right)
\right\vert <4\left\Vert \mathcal{T}\right\Vert _{Y}\epsilon,
\end{equation*}
for all $x$ in the unit ball of $H$. To see this write $T=US$ with $U\in {\
\mathcal{U}} (\mathbb{R}^{K},H)$ and $S\in \mathcal{L}(\mathbb{R}^{K})$.
Then, since $U$ is an isometry, we have
\begin{equation*}
\left\Vert S\right\Vert _{Y}=\sup_{y\in Y}\left\Vert Sy\right\Vert
=\sup_{y\in Y}\left\Vert Ty\right\Vert =\left\Vert T\right\Vert _{Y}\leq
\left\Vert \mathcal{T}\right\Vert _{Y}
\end{equation*}
so that $S\in \mathcal{S}$. We can therefore choose $S_{\epsilon }\in
\mathcal{S}_{\epsilon }$ such that $\left\Vert S_{\epsilon }-S\right\Vert
_{Y}<\epsilon $. Then for $x\in H$, with $\left\Vert x\right\Vert \leq 1$,
we have
\begin{eqnarray*}
\left| f_{T}\left( x\right) -f_{US_{\epsilon }}\left( x\right) \right|
&=&\left| \inf_{y\in Y}\left( \left\Vert x-USy\right\Vert^{2} \right)
-\inf_{y\in Y}\left( \left\Vert x-US_{\epsilon }y\right\Vert^{2} \right)
\right| \\
&\leq &\sup_{y\in Y}\left|\left( \left\Vert x-USy\right\Vert ^{2}-\left\Vert
x-US_{\epsilon }y\right\Vert ^{2}\right)\right| \\
&=&\sup_{y\in Y}\left|\left\langle US_{\epsilon }y-USy,2x-\left( USy+US_{\epsilon
}y\right) \right\rangle \right|\\
&\leq &\left( 2+2\left\Vert \mathcal{T}\right\Vert _{Y}\right) \sup_{y\in
Y}\left\Vert \left( S_{\epsilon }-S\right) y\right\Vert \leq 4\left\Vert
\mathcal{T}\right\Vert _{Y}\epsilon .
\end{eqnarray*}
Apply Theorem \ref{Corollary Bound Finite Max Expectation} to the finite
collection of function classes $\left\{ \mathcal{G}_{S}:S\in \mathcal{S}
_{\epsilon }\right\} $ to see that with probability at least $1-\delta $
\begin{eqnarray*}
&&\sup_{T\in \mathcal{T}}\mathbb{E}_{x\sim \mu }f_{T}\left( x\right) -\frac{%
1 }{m}\sum_{i=1}^{m}f_{T}\left( x_{i}\right) \\
&\leq &\max_{S\in \mathcal{S}_{\epsilon }}\sup_{U\in \mathcal{U}\left(
\mathbb{R} ^{K},H\right) }\mathbb{E}_{x\sim \mu }f_{US}\left( x\right) -%
\frac{1}{m} \sum_{i=1}^{m}f_{US}\left( x_{i}\right) +8\left\Vert \mathcal{T}%
\right\Vert _{Y}\epsilon \\
&\leq &\max_{S\in \mathcal{S}_{\epsilon }}\mathcal{R}_{m}\left( \mathcal{G}
_{S},\mu \right) +b\sqrt{\frac{\ln \left\vert \mathcal{S}_{\epsilon
}\right\vert +\ln \left( 1/\delta \right) }{2m}}+8\left\Vert \mathcal{T}
\right\Vert _{Y}\epsilon \\
&\leq &\sup_{S\in \mathcal{S}}\mathcal{R}_{m}\left( \mathcal{G}_{S},\mu
\right) +\frac{bK}{2}\sqrt{\frac{\ln \left( 16m\left\Vert \mathcal{T}
\right\Vert _{Y}^{2}\right) }{m}}+\frac{8\left\Vert \mathcal{T}\right\Vert
_{Y}}{\sqrt{m}}+b\sqrt{\frac{\ln \left( 1/\delta \right) }{2m}},
\end{eqnarray*}
where the last line follows from the known bound on $\left\vert \mathcal{S}
_{\epsilon }\right\vert $, subadditivity of the square root and the choice $%
\epsilon =1/\sqrt{m}$.\qed\bigskip
\end{proof}
\begin{remark}
If $H$ is finite dimensional the above result may be improved to
\begin{equation}
\mathbb{E} f_{T} - {\hat{\mathbb{E}}} f_{T} \leq \frac{b}{2}\sqrt{\frac{dK
\ln \left( 16m\left\Vert \mathcal{T} \right\Vert _{Y}^{2}\right) }{m}}+\frac{%
8\left\Vert \mathcal{T}\right\Vert _{Y}}{\sqrt{m}}+b\sqrt{\frac{\ln \left(
1/\delta \right) }{2m}}. \label{eq:FD}
\end{equation}
To see this, follow the same lines as in Lemma \ref{Lemma Key} to note that
\begin{equation*}
\sup_{T\in \mathcal{T}} \mathbb{E} f_T - {\hat{\mathbb{E}}}f_T \leq \max_{T
\in \mathcal{T}_\epsilon} \mathbb{E} f_T - {\hat{\mathbb{E}}}f_T + 8 \|%
\mathcal{T}\|_Y \epsilon,
\end{equation*}
where $\mathcal{T}_{\epsilon }$ is a subset of $\mathcal{T}$ such that every
member of $\mathcal{T}$ can be approximated by a member of $\mathcal{T}%
_{\epsilon }$ up to distance $\epsilon $ in the norm $\left\Vert\cdot\right%
\Vert_{Y}$.
By Proposition \ref{Proposition covering}, $\left\vert \mathcal{T}_{\epsilon
}\right\vert \leq \left( 4\left\Vert \mathcal{T}\right\Vert _{Y}/\epsilon
\right)^{dK}$. Inequality \eqref{eq:FD} now follows from Theorem \ref%
{Corollary Bound Finite Max Expectation} with $N=|\mathcal{T}_\epsilon|$ and
$\epsilon =1/\sqrt{m}$.
\end{remark}
To complete the proof of Theorem \ref{Theorem main} we now fix some $S\in
\mathcal{S}$ and focus on the corresponding function class $\mathcal{G}_{S}$.
\begin{lemma}
For any $S\in \mathcal{L}\left(\mathbb{R}^{K}\right) $ we have
\begin{equation*}
\mathcal{R}\left( \mathcal{G}_{S},\mu \right) \leq 2 \sqrt{2\pi} \left\Vert
S\right\Vert _{Y} \frac{K}{\sqrt{m}}.
\end{equation*}
\end{lemma}
\begin{proof}
Let $\left\Vert x_{i}\right\Vert \leq 1$ and define Gaussian processes $%
\Omega _{U}$ and $\Xi _{U}$ indexed by ${\mathcal{U}}(\mathbb{R}^{K},H)$
\begin{eqnarray*}
\Omega _{U} &=&\sum_{i=1}^{m}\gamma _{i}\inf_{y\in Y}\left\Vert
x_{i}-USy\right\Vert ^{2} \\
\Xi _{U} &=&2\left\Vert S\right\Vert _{Y}\sum_{k=1}^{K}\sum_{i=1}^{m}\gamma
_{ik}\left\langle x_{i},Ue_{k}\right\rangle \text{,}
\end{eqnarray*}%
where the $e_{k}$ are the canonical basis of $\mathbb{R}^{K}$. For $%
U_{1},U_{2}\in {\mathcal{U}}({\mathbb{R}}^{K},H)$ we have
\begin{eqnarray*}
\mathbb{E}\left( \Omega _{U_{1}}-\Omega _{U_{2}}\right) ^{2} &\leq
&\sum_{i=1}^{m}\left( \sup_{y\in Y}\Vert x_{i}-U_{1}Sy\Vert ^{2}-\Vert
x_{i}-U_{2}S\Vert ^{2}\right) ^{2} \\
&\leq &\sum_{i=1}^{m}\sup_{y\in Y}4\langle x_{i},(U_{2}-U_{1})Sy\rangle ^{2}
\\
&\leq &4\sum_{i=1}^{m}\sup_{y\in Y}\Vert U_{2}^{\ast }x_{i}-U_{1}^{\ast
}x_{i}\Vert ^{2}\Vert Sy\Vert ^{2} \\
&=&4\left\Vert S\right\Vert _{Y}^{2}\sum_{i=1}^{m}\sum_{k=1}^{K}\left(
\left\langle x_{i},U_{1}e_{k}\right\rangle -\left\langle
x_{i},U_{2}e_{k}\right\rangle \right) ^{2} \\
&=&\mathbb{E}\left( \Xi _{U_{1}}-\Xi _{U_{2}}\right) ^{2}.
\end{eqnarray*}%
It follows from Lemma \ref{Lemma Gauss dominates Rademacher} and Slepians
lemma (Theorem \ref{Slepian Lemma}) that
\begin{equation*}
\mathcal{R}_{m}\left( \mathcal{G}_{S},\mu \right) \leq \mathbb{E}_{\mathbf{x}%
\sim \mu ^{m}}\frac{2}{m}\sqrt{\frac{\pi }{2}}\mathbb{E}_{\mathbf{\gamma }%
}\sup_{U}\Xi _{U},
\end{equation*}%
so the result follows from the following inequalities, using Cauchy-Schwarz'
and Jensen's inequality, the orthonormality of the $\gamma _{ik}$ and the
fact that $\left\Vert x_{i}\right\Vert \leq 1$ on the support of $\mu $.
\begin{eqnarray*}
\mathbb{E}_{\mathbf{\gamma }}\sup_{U}\Xi _{U} &=&2\left\Vert S\right\Vert
_{Y}\mathbb{E}\sup_{U}\sum_{k=1}^{K}\left\langle \sum_{i=1}^{m}\gamma
_{ik}x_{i},Ue_{k}\right\rangle \\
&\leq &2\left\Vert S\right\Vert _{Y}\sum_{k=1}^{K}\mathbb{E}\left\Vert
\sum_{i=1}^{m}\gamma _{ik}x_{i}\right\Vert \\
&\leq &2\left\Vert S\right\Vert _{Y}K\sqrt{m}.
\end{eqnarray*}%
\qed
\end{proof}
Substitution of the last result in Lemma \ref{Lemma Key} and noting that,
for $K\geq 1$, $2\sqrt{2\pi }K+8\leq 14K$, gives Theorem \ref{Theorem main}.
Observe that when the set $\mathcal{S}$ contains only the identity matrix,
the function class $\mathcal{G}_S$ is the class of reconstruction errors of
PCA. In this case, the result can be improved as shown by the next lemma.
\begin{lemma}
\label{Lemma PCA}$\mathcal{R}\left( \mathcal{D},\mu \right) \leq 2\sqrt{K/m}$%
.
\end{lemma}
\begin{proof}
Recall, for every $z\in H$, that the outer product operator $Q_{z}$ is
defined by $Q_{z}x=\left\langle x,z\right\rangle z$. With $\left\langle
\cdot,\cdot\right\rangle _{2}$ and $\left\Vert \cdot \right\Vert _{2}$
denoting the Hilbert-Schmidt inner product and norm respectively we have for
$\left\Vert x_{i}\right\Vert \leq1$
\begin{eqnarray*}
\mathbb{E}_{\sigma }\sup_{f\in \mathcal{D}}\sum_{i=1}^{m}\sigma _{i}f\left(
x_{i}\right) &=&\mathbb{E}_{\sigma }\sup_{U\in \mathcal{U}
}\sum_{i=1}^{m}\sigma _{i}\left( \left\Vert x_{i}\right\Vert ^{2}-\left\Vert
UU^{\ast }x_{i}\right\Vert ^{2}\right) \\
&=&\mathbb{E}_{\sigma }\sup_{U\in \mathcal{U}}\left\langle
\sum_{i=1}^{m}\sigma _{i}Q_{x_{i}},UU^{\ast }\right\rangle _{2} \\
&\leq &\mathbb{E}_{\sigma }\left\Vert \sum_{i=1}^{m}\sigma
_{i}Q_{x_{i}}\right\Vert _{2}\sup_{U\in \mathcal{U}}\left\Vert UU^{\ast
}\right\Vert _{2} \\
&\leq &\sqrt{mK},
\end{eqnarray*}
since the Hilbert-Schmidt norm of a $K$-dimensional projection is $\sqrt{K}$%
. The result follows upon multiplication with $2/m$ and taking the
expectation in $\mu ^{m}$.\qed
\end{proof}
An application of Theorem \ref{Corollary Bound Finite Max Expectation} with $%
N=1$ and $b=1$ also give a generalization bound for PCA of order $\sqrt{K/m}$%
.
\section{Concluding remarks}
We have analyzed a general method to encode random vectors in a Hilbert
space $H$. The method searches for an operator $T:\mathbb{R}^{K}\rightarrow H
$ which minimizes, within some prescribed class $\mathcal{T}$, the empirical
average of the reconstruction error, which is defined as the minimum
distance between a given point in $H$ and an image of the operator $T$
acting on a prescribed codebook $Y$.
We have presented two approaches to upper bound the estimation error of the
method in terms of the parameter $K$, the sample size $m$ and the properties
of the sets $\mathcal{T}$ and $Y$. The first approach is based on a direct
bound for the Rademacher average of the loss class induced by the
reconstruction error. The bound matches the best known bound for $K$-means
clustering in a Hilbert space \cite{Biau Devroye Logosi 2006} but also
applies to other interesting coding techniques such as sparse coding and
non-negative matrix factorization. The second approach uses a decomposition
of the function class as a union of function classes parameterized by
$K$-dimensional isometries. The main idea is to approximate the union with a
finite union via covering numbers and then bound the complexity of each
class under the union with Rademacher averages. This second result is more
complicated than the first one, however it provides in certain cases a better
dependency of the bound on the parameter $K$ at the expense of an additional logarithmic
factor in $m$.
We conclude with some open problems and possible extensions which are
suggested by this study. Firstly, it would be valuable to investigate the
possibility of removing the logarithmic term in $m$ in the bound of Theorem
\ref{Theorem main}. Secondly, it would be important to elucidate whether the
dependency in $K$ in the same bound is optimal. The latter problem is
also mentioned in \cite{Biau Devroye Logosi 2006} in the case of
$K$-means clustering. Finally, in would be interesting to study
possible improvements of our results in the case that additional
assumptions on the probability measure $\mu$ are introduced. For
example, in the case of $K$-means clustering in a finite dimensional
Hilbert space \cite{antos2005} shows that for certain classes of
probability measures the rate of convergence can be improved to
$O(\log(m)/m)$ and it may be possible to obtain similar improvements
in our general framework.
\subsection*{Acknowledgments}
This work was supported by EPSRC Grants GR/T18707/01 and EP/D071542/1.
|
\section{Introduction}
\label{sec:intro}
Low dimensional organic molecular compounds based on the BEDT-TTF, TTF and TCNQ molecules are narrow band systems,
where the Coulomb repulsion is large compared to the bandwidth. This leads to many-body effects which are manifest,
e.g., in the spin susceptibility \cite{Torrance} or angular-resolved photoemission experiments
\cite{Jerome,Ito05,Sing}. While theoretical works suggest the importance of Coulomb parameters in these materials
\cite{Torrance,Mazumdar}, only rough theoretical estimates of these parameters do exist \cite{Hubbard}. The use of
modern techniques, such as density-functional theory (DFT), gives accurate results of the parameters for individual
molecules \cite{TTF-TCNQ,Scriven}, improving previous quantum chemistry calculations \cite{Castet,Mori}. However, there is a lack of accurate theoretical estimates of the Coulomb repulsion screened inside the crystal. These are needed for a more realistic description of organic molecular compounds.
In the present paper we introduce a systematic method for calculating these Coulomb parameters in low-dimensional organic molecular crystals. In Sec.~\ref{sec:model} we construct the minimal model and give details of the general formalism. This is applied to TTF-TCNQ salts in Sec.~\ref{sec:submolecular}, where we discuss the distributed-dipole approach. In Sec.~\ref{sec:screening} we give our results for Coulomb parameters.
\section{Model and Formalism}
\label{sec:model}
The minimal model generally used to describe the low-energy electronic properties of low-dimensional organic systems
is the Hubbard model \cite{Anderson}, where transfer integrals between neighboring sites and on-site Coulomb
interactions are taken into account. Such a simple Hubbard model is, however, usually not sufficient to describe
strongly correlated organics, as long-range Coulomb repulsion energies can rarely be neglected \cite{Hubbard,Horsch}. In the generalized model
\begin{eqnarray}
\label{eq:HubbardModel}
H=-t\!\!\sum_{ \langle ij\rangle,\sigma}
(c^{\dagger}_{i,\sigma} c_{j,\sigma} + h.c.)
+ U\!\sum_i n_{i\uparrow} n_{i\downarrow}
+ \sum_{i,j} V_{ij} n_i n_j,
\end{eqnarray}
we include hopping integrals between nearest-neighbors, $t$, on-site Coulomb energies, $U$, and longer-range Coulomb interactions, $V_{ij}$. While transfer integrals can be obtained from {\em ab initio} DFT calculations \cite{TTF-TCNQ,Valenti,Nakamura}, Coulomb parameters are more difficult to obtain, as they include the screening correction due to different high-energy processes inside the crystal, the main mechanism being the polarization of the surrounding, usually highly polarizable, molecules. By realizing that organic molecules preserve their identity inside the crystal, we may estimate the screening energy, by describing the molecular interactions through classical electrostatics.
In BEDT-TTF and TTF-TCNQ families of organic conductors \cite{Ishiguro}, flat molecules crystallize forming low-dimensional anisotropic structures. Considering such a lattice of organic molecules, we ask what happens when a charge is added to one molecule. The electric field of that charge will induce a dipole moment in the surrounding highly polarizable molecules. Solving this electrostatic problem assuming fixed polarizabilities, we obtain the screening of the charged molecule in linear response. This gives the screening contribution of the lattice that is needed for renormalizing the bare Coulomb integral.
Modelling each molecule as a polarizable point with charge $q_i$, dipole moment ${\bf p}_i$ and polarizability $\alpha_i$ the energy of the lattice in the above situation is given by
\begin{eqnarray}
\label{eq:Wq}
W&=&\sum_{i} \left[ \frac{|{\bf {p}}_{i}|^2}{2{\alpha_i}}-{\bf {p}}_{i}\cdot{\bf {E}}_{i}^{ext}+
\Phi_{i}q_{i}+T(q_{i}) \right]
\nonumber \\
&& + \sum_{i,j>i} \left[ U_{1}({\bf p}_{i},{\bf p}_{j})+
U_{2}({\bf p}_{i},q_j)+U_{3}(q_{i},q_{j}) \right],
\end{eqnarray}
where ${\bf E}_{i}^{ext}$ is the electric field and $\Phi({\bf r})$ the potential associated with the external
charge, while the first and fourth terms correspond to the resistance to polarization of each lattice point and to
the internal energy of the molecules as a function of charge state, respectively. The terms $U_{1}$, $U_{2}$ and
$U_{3}$ are the dipole-dipole, dipole-monopole and monopole-monopole interactions \cite{Jackson}. For molecules
without net charge (the electrons involved in charge transfer are explicitly included in the Hubbard model), we can
simplify Eq.~(\ref{eq:Wq}). Expressing it in the notation of Allen \cite{Allen}, defining the dipolar field by $|{\bf E}^{dip}\rangle=\Gamma |{\bf p}\rangle$, where $\Gamma$ is the dipole-dipole interaction matrix:
\begin{equation}
\label{eq:Gammamatrix}
\Gamma_{i\mu,j\nu} = \frac {3{r}_{ij\mu}{r}_{ij\nu}-\delta_{\mu \nu}{r}_{ij}^2}{r_{ij}^5}
\quad,\quad i\neq j, \quad \mu,\nu \equiv x,y,z,
\end{equation}
the energy of the system is then given by
\begin{equation}
\label{eq:WDirac}
W=\frac {1}{2} \langle{{\bf p}| {\alpha}^{-1} - \Gamma |{\bf p}}\rangle-
\langle{{\bf p}|{\bf E}^{ext}}\rangle.
\end{equation}
By applying the variational principle, $\delta W=0$, we obtain the expression for the set of dipoles which minimize the energy of the whole lattice,
\begin{equation}
\label{eq:bracketp}
|{\bf p}\rangle= \left( {\alpha}^{-1} - \Gamma \right)^{-1} |{\bf E}^{ext}\rangle.
\end{equation}
Inserting Eq.~(\ref{eq:bracketp}) into Eq.~(\ref{eq:WDirac}) yields
\begin{equation}
\label{eq:deltaW}
\delta W = -\frac {1}{2} \langle{{\bf E}^{ext}| \left( {\alpha}^{-1} - \Gamma \right)^{-1} |{\bf
E}^{ext}}\rangle.
\end{equation}
Therefore, placing two charges at a lattice point or two different lattice points, we are able to calculate the correction to the total energy due to the polarization of the rest of the molecules in the crystal. Defining the external field as the composition of two fields, corresponding to charges placed at points $n$ and $m$ in the lattice, ${\bf E}^{ext}={\bf E}_n+{\bf E}_m$, the energy of the system is given by
\begin{equation}
\delta W=-\langle{{\bf E}_m| \left( {\alpha}^{-1} - \Gamma
\right)^{-1}|{\bf E}_n}\rangle- \langle{{\bf E}_n| \left( {\alpha}^{-1} -
\Gamma \right)^{-1} |{\bf E}_m}\rangle.
\end{equation}
The equation above gives the on-site Coulomb screening, $\delta U$, when $n=m$, and the inter-site, $\delta V$, $\delta V'$, ..., for $n \neq m$. Thus, defining $|n-m|=l$, the general equation for the screening Coulomb parameters reads
\begin{equation}
\label{eq:deltaVl}
\delta V^l=-\langle{{\bf E}_0| \left( {\alpha}^{-1} - \Gamma \right)^{-1}|{\bf E}_l}\rangle\quad
\Longrightarrow\quad V^l=V_{0}^l+\delta V^l,
\end{equation}
where $\delta U$ corresponds to $l=0$, $\delta V$ to $l=1$, ... and $V_{0}^l$ are the Coulomb parameters for a single molecule/dimer, calculated with DFT \cite{TTF-TCNQ,Scriven}.
To illustrate the method we apply it to a cubic lattice of polarizable point dipoles, in
which an antiferroelectric instability occurs above a critical polarizability $\alpha_c$
\cite{Allen,Luttinger}, due to the anisotropy in the dipole interactions.
Such instability influences the screening of a charge in the lattice.
For small polarizabilities $\alpha$ the induced dipoles simply arrange along the field-lines of
the central point charge. Increasing $\alpha$ we, however, observe quite unconventional dipole arrangements as shown
in Fig.~\ref{dipoleSC}. This reflects the fact that for this lattice the matrix
$(\alpha^{-1}-\Gamma)$ becomes singular at a critical polarizability $\alpha_c$, with the eigenvector of the
vanishing eigenvalue having non-zero momentum (in this case ($\pi,\pi,0$)). Hence the ferroelectric instability happens before the critical value $\alpha_{CM}$ obtained from the Clausius-Mossotti relation \cite{Allen}.
\begin{figure*}
\center
\vspace{3ex}
\includegraphics[scale=1.0]{dipoleSCalpha05.eps}
\includegraphics[scale=1.0]{dipoleSCalpha075.eps}
\includegraphics[scale=1.0]{dipoleSCalpha08.eps}
\caption{Dipole arrangement in the $x$-$y$-plane of a simple-cubic lattice of point dipoles to a charge added at the center. As the polarizability increases from $\alpha=0.5 \alpha_{CM}$ (left), $0.75 \alpha_{CM}$ (middle) to $\alpha=0.8 \alpha_{CM}$ (right), the dipoles start to deviate strongly from pointing along the field-lines of the external field.}
\label{dipoleSC}
\end{figure*}
\section{Submolecular approach}
\label{sec:submolecular}
In a real molecular crystal the molecules typically form stacks, i.e., they are quite closely packed. Consequently,
approximating the polarizability of a molecule by just one point-polarizability is not a good approximation. It only
works outside a sphere containing the charge density of the molecule, which for typical $\pi$-bonded molecules will
intersect the neighboring molecules, but it becomes accurate at large distances. To obtain a better description of
the molecular response, we model it as point-polarizabilities distributed over the non-hydrogen atoms of the
molecule \cite{Mazur}. The accuracy of this submolecular approach is tested for TTF-TCNQ salts.
We consider a small system of only three or five molecules in a TTF chain. For such small systems we can perform constrained density-functional calculations \cite{TTF-TCNQ,Behler}, constraining the added charge to the central molecule. This gives us the screening due to the molecules in this small assembly. We compare to the results of the screening calculated using the distributed point-polarizability approach. Since this approach works least well in the near-field, this is a critical test of the method. Nevertheless, both approaches agree within $5\%$. If we included more distant molecules, for which the computational cost of constrained density-functional calculations would be prohibitive, the agreement would improve further, as the multipole approximation becomes exact in the far-field. Thus, using polarizability tensors obtained with DFT \cite{TTF-TCNQ} in our submolecular approach essentially reproduces the results of full quantum mechanical treatment of the screening.
\section{Screening parameters in TTF-TCNQ}
\label{sec:screening}
To obtain the screened parameters for a TTF-TCNQ crystal, we perform calculations for crystal fragments (clusters)
of increasing sizes to extrapolate to the infinite-size limit. To obtain the parameters for the extended Hubbard
model (\ref{eq:HubbardModel}) we consider the Coulomb screening between two charges at the same or at neighbor
molecules placed in the center of the cluster. For simplicity, the clusters of $N$ molecules are constructed as
spheres of radius $R$, where $R$ is the distance between the doped molecule (or center of the pair of molecules) and the farthest
neighbor in the cluster, hence $\Omega=\frac {4\pi}{3}R^3$ being the volume of the sphere \cite{Pederson}. For large
enough $R$ we can linearly extrapolate the screening energy $\delta U$ or $\delta V$, respectively, versus $1/R$ to
the thermodynamic limit ($R \to \infty$). This is shown in Fig.~\ref{UTTF}, where clusters of up to 400 molecules
are studied. Calculations for TTF-TCNQ salts up to the third nearest-neighbors are performed and results are given
in Table \ref{screening}. The ratio of the radius of the sphere, $R$, to the distance between
third nearest-neighbors, $d$, where we still can make the extrapolation to the thermodynamic limit is
$d/R = 0.43$. However for longer range interactions, the cluster size should be increased
towards this convergence ratio, reaching a linear dependence with the inverse of the volume, which
allows the extrapolation to the bulk value.
\begin{figure}
\center
\includegraphics[scale=0.35,clip]{deltaUVTTF.eps}
\caption{Screening of on-site Coulomb interaction, $\delta U$, in a TTF molecule (left panel), and screening of
inter-site Coulomb interaction, $\delta V$, for nearest-neighbor TTF molecules (right panel), versus
$\Omega^{-1/3}$, for clusters of increasing radius. Extrapolation to the thermodynamic limit is shown by the dotted
line.}
\label{UTTF}
\end{figure}
\begin{table}
\center
\begin{tabular}{l@{\hspace{3ex}}r@{\hspace{3ex}}r@{\hspace{3ex}}r@{\hspace{3ex}}r}
& $\delta U$ & $\delta V$ & $\delta V'$ & $\delta V''$\\ \hline
TTF & 2.7 & 1.9 & 1.25 &0.8\\
TCNQ & 2.6 & 1.9 & 1.3 &0.9\\
\end{tabular}
\caption{\label{screening}
Screening Coulomb energies for TTF-TCNQ.
$\delta U$ is the on-site screening and $\delta V$, $\delta V'$ and $\delta V''$
correspond to first, second and third nearest-neighbor screening energies between molecules
in the same TTF or TCNQ stack. All energies are in eV. For comparison, the
band-width is about 0.7~eV.}
\end{table}
\section{Conclusions}
\label{sec:conclusions}
We have presented a general method for obtaining accurate values of Coulomb parameters in organic molecular
compounds, where molecules interact weakly, and shown how it works for TTF-TCNQ.
We are currently extending this method to other crystals like
quasi-bidimensional salts based on the BEDT-TTF molecule.
\section*{Acknowledgements}
JM and LC acknowledge financial support from MICINN under contract
CTQ2008-06720-C02-02.
|
\section{Introduction}
The two fundamental asymptotically flat, Schwarzschild (S) and Kerr (K) [1] solutions, of General Relativity, were derived almost half a century apart, due to the latter's complexity--it is still daunting when
first encountered. Given K's physical importance, our aim is to provide a transparent, physically instructive, derivation. We will use the labor-saving Weyl method that obviates first wading through the full array of Einstein's equations, then inserting the desired special features of the candidate solution. Instead, we first specify the metric as extensively as possible, using physical and symmetry arguments, then get just the two relevant field equations from the correspondingly reduced Einstein action. This procedure is perfectly legitimate [2], despite appearances. Of course, the equations must still be solved; fortunately they are quite easy. While the metric ansatz, and its plausibility, of course stand on the shoulders of K, the process provides useful insight into its physics.
We will begin by reviewing the oblate spheroidal (OS) coordinates first introduced in this context by [3], then narrow to our candidate metric in this frame. The mechanics of obtaining, and solving, the reduced field equations follows. Separately, we will exhibit the S metric in OS, and that of K in ordinary S, frames. The latter in particular allows one to get a different perspective on K and its limit to S than in OS.
\section{Derivation}
In OS coordinates, the optimal framework for axial symmetry, K has the form [3]
\begin{eqnarray}\label{KerrBL}
ds^2 &= & - dt^2 + \Sigma \, dr^2/\Delta + \Sigma\, d\theta^2 + \of{r^2 + a^2} \, \sin^2\theta \, d\phi^2 + 2 \, m \, r/\Sigma \, \of{dt - a \, \sin^2\theta \, d\phi}^2 ; \\
\Delta &\equiv& r^2 - 2 \, m \, r + a^2, \hskip 1 cm \Sigma \equiv r^2 + a^2 \, \cos^2\theta, \nonumber
\end{eqnarray}
where $(a, m)$ are (the only) constant parameters. [These coordinates are not to be confused with their Cartesian namesakes; as always, they are defined by the interval's form and by their ranges, though the latter are the usual ones.] Deriving this metric from a simple ansatz will be our end-product.
First, consider some limits for orientation. Clearly $m=0=a$ represents flat space in spherical coordinates. The parameter $m$ is aptly named, being the usual ADM mass of the solution (1), that is the system's total energy as measured by an observer at spatial infinity. This is obvious since there (1) coincides with the familiar asymptotic form of S,
\begin{equation}
ds^2 \sim -\of{1 - 2 m/r} \, dt^2 + (1+2 m/r) \, dr^2 + r^2 \, \of{d\theta^2 + \sin^2\theta \, d\phi^2}.
\end{equation}
We know, by the positive energy theorems of GR, that the vanishing of this single parameter is sufficient
to imply flat spacetime. Indeed, at $m=0$, (1) is the direct product of time with the textbook metric for Cartesian three-space expressed in OS coordinates. We must next face K's off-diagonal components, with their linear rather than quadratic dependence on $a$, and attendant loss of reflection symmetry. Fortunately, (sub-)intervals of this type are familiar already in flat space, describing rotating systems with angular velocity $a$, and total ADM angular momentum $J=\pm a\, m$. Note finally that the remaining, angular elements $(d \theta^2, d \phi^2)$ retain their flat space (OS) forms, precisely as they do in the S solution in S coordinates: these frames are in fact defined as keeping ``flat" surface area. The differences between K and flat space, then, are entirely contained in the coefficients of $dr^2$ and $(dt - a\, \sin^2 \theta d\phi)^2$, just as they are (without the extra rotation term) for S. [It is, however, surprising that both coefficients in~\refeq{KerrBL} depend only on $r$ and not on $\theta$, as would be expected {\it a priori}.]
The above de-construction of the K metric provides our basis for its re-construction, starting with a metric ansatz with maximal physical and gauge information and minimal number of unknown functions. We propose
\begin{eqnarray}
ds^2 &= & - dt^2 + \Sigma \, dr^2/D + \Sigma\, d\theta^2 + \of{r^2 + a^2} \, \sin^2\theta \, d\phi^2 + Z/\Sigma \, \of{dt - a \, \sin^2\theta \, d\phi}^2 \label{Kansatz} \\
&=&-dt^2 + \Sigma \, dr^2/D + \Sigma \, \of{4 \, q\, (1 - q)}^{-1} \, dq^2 + (r^2 + a^2) \, (1-q) \, d\phi^2 +Z/\Sigma \, \of{dt - a \, (1 - q) \, d\phi}^2, \nonumber
\end{eqnarray}
where we have replaced the angle $\theta$ by $q= \cos^2\theta$ for notational convenience. The kinematical OS factor $\Sigma$ is defined in~\refeq{KerrBL}, and $(Z, D)$ are the two unknowns, to be determined by the field equations. We follow the (simpler) construction of S, in the spirit of [4], keeping $\Sigma$ in $g_{rr}$ (and its reciprocal in $dt^2$), but modify it by the unknown function $D$. Still following the example of S, the coefficient of the whole rotation term, including $dt^2$, involves a different unknown, $Z$. [Time-independence, the hallmark of K, is also assumed,although it might be derivable by the methods of [4].] Pursuing the analogy still further, we take both unknowns to depend only on $r$; this assumption greatly simplifies the derivation by turning the field equations into ordinary differential ones\footnote{Allowing $(Z,D)$ to depend on $q$ greatly complicates the field equations, contrary to our pedagogical motivation; the more defensible option, letting just $Z$ be general is not too complicated. Note that (as always, [5]) our variables' $q$-independence is only to be invoked in the field equations, after varying the action; however, it is safe to assume time-independence {\it ab initio}.}. While one may argue that introduction of spin should not affect the spherical symmetry of the $dt^2$ coefficient, this is not as defensible here as for S. Rather, its virtue is pragmatic: it is too simple a guess not to be tried first, even if we didn't know it would work.
Evaluating, and varying, the Einstein action with respect to $(Z,D)$ is best done using an algebraic program, unlike for S, whose explicit action is simple; some irreducible, if purely mechanical, complexity remains. The resulting two equations are
\begin{eqnarray}
\frac{\delta S}{\delta D} &=& r^2 \, \sqof{D \, \of{Z + r \, (r - Z')} - r^2 \, \of{a^2 + r^2 - Z}} \label{dsdD} \\
&-& a^2 \, \sqof{ r \, D \, (Z'-r) + (r^2 - Z) \, (a^2 + r^2 - Z)} \, q = 0 \nonumber \\
\frac{\delta S}{\delta Z} &=& r^2 \, \sqof{D \, (Z - r^2) - r \, \of{a^2 + r^2 - Z} \, \of{r -D'}} \label{dsdZ} \\
&+& a^2 \, \sqof{D \, \of{2 \, a^2 + r^2 - 2 \, Z} - \of{a^2 + r^2 - Z} \, \of{2 \, a^2 + 3 \, r^2 - 3 \, Z - r \, D'}} \, q = 0 \nonumber
\end{eqnarray}
It is easy to solve these two sets of equations (note that each requires separate vanishing of the $q^0$ and $q^1$ coefficients). Either set is just an ordinary first order equation for $D'$ or $Z'$ , an artifact of the OS system similar to that in S frame, and specifies the correct answer. For example, from~\refeq{dsdD} we learn
\begin{eqnarray}
D &=& \frac{r^2 \, \of{a^2 + r^2 - Z}}{r^2 + Z - r\, Z'} \\
0 &=& \of{a^2 + r^2 - Z}^2 \, Z \, \of{Z - r \, Z'}
\end{eqnarray}
Of the $3$ possible $Z$-roots, only $Z= 2\, m\, r$ (seting the integration constant to agree with that of~\refeq{KerrBL}) is acceptable, as either $Z=0$ or $Z=r^2+a^2=0=D$ lead to singular metrics; this in turn means $D = r^2+a^2- 2\, m\, r$, which is just (1). Likewise, the two equations in~\refeq{dsdZ} are equivalent to the algebraic $D =r^2+a^2- Z$,
plus an easily solved differential one for $D$ reproducing the result of~\refeq{dsdD}, again modulo
a singular choice, $D=0$.
This completes the derivation of K from our ansatz~\refeq{Kansatz}. We emphasize that the theorems of [2] guarantee the validity of the Weyl procedure, neither missing any correct solutions nor introducing any spurious ones.
\section{Schwarzschild in Oblate Spheroidal; Kerr in Spherical}
In this section, we record the forms (which do not seem available elsewhere) of the S and K solutions in their ``reciprocal" frames: S in OS and K in S. This will shed light on the effects of coordinate choices and more important, permit a better understanding of their relations, especially of S as the non-rotating limit of K. Inserting the spherical(barred)-OS(unbarred) coordinate relations
\begin{equation}\label{barstoun}
\bar r^2 = r^2 + a^2 \, \sin^2\theta\hskip 1 cm
\bar \theta = \hbox{arctan}\of{\frac{\sqrt{r^2 + a^2} \, \tan\theta}{r}} \hskip 1 cm \bar \phi = \phi.
\end{equation}
into the usual S-interval in S-frame, and using our $q=\cos^2 \theta$ instead of $\theta$, we find for S in OS the form
\begin{eqnarray}\label{SinOS}
ds^2&=& -\of{1 - \frac{2 \, m}{\bar r}} \, dt^2 + \sqof{\frac{r^2/\bar r^2}{1 - \frac{2 \, m}{\bar r}} + \frac{a^4 \,(1-q) \, q}{\of{a^2 + r^2} \, \bar r^2} } \, dr^2 + 2 \, \frac{a^2 \, m \, r}{\bar r^2 \, \of{2 \, m - \bar r}} \, dr \, dq \\
&+& \sqof{ \frac{r^2 \, (a^2 + r^2) \, (2 \, m -\bar r) - a^4 \, (1 - q) \, q \, \bar r}{4 \, (1 - q) \, q \, \bar r^2 \, (2 \, m - \bar r)} } \, dq^2 + \of{a^2 + r^2} \, (1-q) \, d\phi^2 \nonumber
\end{eqnarray}
here $\bar r$ stands for its value in~\refeq{barstoun}. As a check on the correctness of this transformation, we have verified that $R_{\mu\nu}=0$ and that the Riemann invariant matches its usual S value:
\begin{equation}
R_{\alpha\beta\mu\nu} \, R^{\alpha\beta\mu\nu} = \frac{48 \, M^2}{\bar r^6}.
\end{equation}
Note that, surprisingly, the S metric has now acquired an-off diagonal, $dr\, dq$, component. Indeed, were we to ansatz a diagonal metric in OS coordinates for S, the resulting field equations would force us back to S-frame\footnote{While the $2$-dimensional $(r, q)$ subspace can of course always be diagonalized, that would takes us outside the OS frame. }, by stating that the only such solutions require $a=0$. It is clear that the $a=0$ limit of~\refeq{SinOS} is the initial S, while $m=0$ is indeed flat space: $g_{rq}=0$ and $(g_{rr} , g_{qq})$ limit to their flat OS forms.
Next, we transform K to S-frame, which is better suited to understanding the S-limit of K, as we now see. The required coordinate relation, inverse to~\refeq{barstoun}, is
\begin{equation}\label{rqrbqb}
2\, \left\{ r^2, a^2 \, q \right\} = \sqrt{a^4 - 2 \, a^2 \, (1 - 2 \, \bar q) \, \bar r^2 + \bar r^4} + (\bar r^2 - a^2) \, \left\{ 1 , -1\right\}
\end{equation}
(where $\bar q = \cos^2\bar\theta$ for the spherical $\bar\theta$) with $(t, \phi)$ unchanged. We then obtain K as
\begin{eqnarray}\label{KDSS}
ds^2 &=& -\of{1 - \frac{2 \, m \, r}{\rho^2}} \, dt^2 - \frac{4 \, a \, m \, r (1-q)}{\rho^2} \, dt \, d\phi + \frac{r^2 \, (r^2 + a^2)^2 + a^4 \, (1 - q) \, q \, \Delta}{\Delta\, \bar r^2 \, \rho^2} \, dr^2 \\
&+& \frac{2 \, a^2 \, m \, r \, \bar r}{\Delta \, \rho^2} \, dr \, dq + \frac{\of{ a^4 \, (1 - q) \, q + r^2 \, \Delta}\, \bar r^4}{4 \, (1 - q) \, q \, r^2 \, \Delta} \, dq^2 + \frac{(1 - q) \, \of{2 \, a^2 \, m \, r \, (1-q) + (r^2 + a^2) \, \rho^2}}{\rho^2} \, d\phi^2 \nonumber \\
\rho^2 &\equiv& r^2 + a^2 \, q \nonumber
\end{eqnarray}
keeping the unbarred coordinates for notational simplicity; they merely mean their values in terms of the barred ones given in~\refeq{rqrbqb}. Note the appearance of the off-diagonal $dr\, dq$ in addition to the spinning part. The reason it is more illuminating to take the K$\longrightarrow$ S limit here than in OS is that one wants to shut off the angular momentum $\sim a\, m$, while keeping $m$ finite. But this forces setting $a=0$, which is overkill in OS, involuntarily also taking us from OS to just S. Instead, since the K of~\refeq{KDSS} is already in S-frame, the parameter $a$ no longer refers to the OS kinematics, but more properly just to the rotation itself. It is clear that the barred and unbarred coordinates coincide in this limit, so that~\refeq{KDSS} reduces to the elementary S. Expansion of K$=$S $+0(a) +0(a^2)$ can then be used to understand the cumulative effects of rotation on the ``base" S, beyond the simple linearized inclusion of $J=a\, m$, an exercise we omit here.
\section{Summary}
We have derived the Kerr [1] solution using physical--axial symmetry and constant rotation--information,
expressed in the OS [3] gauge. The resulting metric ansatz was further reduced by hints from similar derivation of S, and demanding maximal simplicity. Thanks to the power of the Weyl method [2], this led
directly to just the two field equations for the unknown metric components, which were then easily solved. Separately, we shed additional light on the K$\longrightarrow$ S limit by first transforming K to S-frame.
We thank Matt Visser for useful correspondence. SD acknowledges support from grants NSF PHY 07-57190 and DOE DE-FG02-164 92ER40701.
|
\section{Introduction}
Due to their important potential as environmentally benign alternatives to conventional toxic organic solvents,
room-temperature ionic liquids (RTILs) have attracted considerable attention recently.~\cite{seddon:rtil:CPP,keim:rtil,review2:rtil}
According to theoretical~\cite{shim:letter,*shim2,shim:rtil:rot,shim:rtil:review,margulis:rtil:dyn,bhargava:rtil,jeong:pol,jeong:1/f,kobrak:rtil:dyn1,*kobrak:rtil:dyn2,*kobrak:rtil:dyn3} and experimental~\cite{samanta:rtil1,*samanta:rtil2,*samanta:rtil3,*samanta:rtil4,maroncelli:rtil:solrot1,*maroncelli:rtil:complete1, *maroncelli:rtil:complete2,*maroncelli:rtil:solrot2, castner:il1,*castner:il2,vauthey:rtil:dyn,fayer:rtil:oke} studies on solvation and rotational dynamics of RTILs, their long-time behaviors are characterized by nonexponential decay.
This implies that RTILs are dynamically inhomogeneous and their local relaxation is widely distributed in time and space.~\cite{shim:rtil:rot,jeong:1/f} The clustered mobile and immobile ions observed in recent molecular dynamics (MD) simulation studies are ascribed to inhomogeneous dynamics in RTILs.~\cite{margulis:rtil:ree,Ngai:hetero}
Dynamic heterogeneity often invoked to explain many peculiar properties of supercooled liquids~\cite{Richert:review:hetero,Ediger:review:hetero} refers to the enhanced temporal correlation of their local dynamic states with a decrease in temperature. From the viewpoint of facilitated motions, dynamics of supercooled liquids are dominated by fluctuations.~\cite{Chandler:prl,jung:exc} According to several studies based on lattice models, called kinetically-constrained models (KCMs),~\cite{FA,East} non-trivial structures in the space-time trajectory arising from dynamic constraints in the KCM description accurately reproduces many of the dynamical properties of supercooled liquids.~\cite{Chandler:prl,Chandler:spacetime} At the molecular level, it is found that trajectories of individual particles in atomistic models of supercooled liquids are in general governed by dynamic fluctuations and thus cannot be predicted from static properties, such as structures.~\cite{Berthier:predict} Meanwhile, recent studies have attempted to correlate length-scale dependent heterogeneous dynamics with liquid structures on the basis of the dynamic propensity calculated from the isoconfigurational ensemble.~\cite{Harrowell:propensity1,*Harrowell:propensity2,*Harrowell:propensity3,*Daekeon:rtil:propen} Despite these efforts, the origin of persisting dynamic correlations and the potential link of the dynamic correlations to structures still remain open questions in understanding dynamics of glassy liquids.
Glassy dynamics of supercooled liquids are characterized by many unique features such as the stretched exponential decay of time correlation functions, subdiffusive behavior in the intermediate time scale, decoupling of exchange and persistence times,~\cite{jung:exc} and breakdown of the Stokes-Einstein (SE) relation.~\cite{Kivelson:SE} A variety of models of supercooled liquids, e.g., binary Lennard-Jones (LJ) mixture,~\cite{Chaudhuri:glass} supercooled water~\cite{Mazza:SE}, Weeks-Chandler-Anderson (WCA) mixture,~\cite{Hedges:decoupl} and KCMs,~\cite{jung:SE,jung:exc,Berthier:lifetime} reproduce these interesting features, which are generally believed to be intimately related to the dynamic heterogeneity. This seemingly universal nature of dynamic heterogeneity and related phenomena is a main motivation to study glassy dynamics of ionic liquids, the molecular details and interaction potentials of which are quite different from those of supercooled liquids. Specifically, strong electrostatic interactions in RTILs lead to an alternating structure of cations and anions, which is believed to generate a significant memory effect on their dynamics.~\cite{jeong:1/f} Since the cage structure would exert a substantial influence on correlations between local excitation events, it would be both interesting and important to see if dynamic behaviors of RTILs bear any resemblance to those of supercooled liquids.
One of the main challenges in studying dynamics of RTILs (and more generally glassy liquid systems) is the time scale. As is well known, many glassy systems exhibit extremely slow dynamics, in particular, at low temperatures. Therefore it is very difficult, if not impossible, to probe their long-time behaviors directly via atomistic MD simulations. However, it is precisely these dynamics at long times and their variations with temperatures that are of especial interest. We thus employ a coarse-grained model to make simulations efficient and analyze long-time dynamics. We note that this approach is not new. Previous efforts\cite{jeong:thesis,voth:cg,*voth:am-cg,Chiappe:cg} based on similar descriptions have provided useful insights into RTIL properties, both structure and dynamics. In the present paper, we focus on the characterization of glassy dynamics and dynamic heterogeneity of RTILs. As a prototypical RTIL, we study 1-ethyl-3-methylimidazolium hexafluorophosphate (${\rm EMI}^+{\rm PF_6}^-$).
To understand translational dynamics of EMI$^+$PF$_6^-$, we analyze the self-intermediate scattering function and mean square ion displacements of our model RTIL system. We also examine the temperature dependence of its structural relaxation time and related fragility.\cite{Angell:fragility} Our results suggest that EMI$^+$PF$_6^-$ is a fragile glass former and it violates the SE relation at low temperatures. We regard correlations between local dynamic states as an essential aspect of the aforementioned dynamic heterogeneity of glassy liquids. Accordingly, we investigate distributions of the persistence and exchange times, i.e., waiting times for the first and subsequent excitations, by monitoring the motions of individual ions. To understand the influence of Coulombic interactions on glassy dynamics, we also make contact with non-ionic model systems of supercooled liquids.
The outline of this paper is as follows: In Sec.~\ref{sec:model}, we introduce the coarse-grained model of an ionic liquid and briefly describe the simulation methods. In Sec.~\ref{sec:results}, we present the MD results on glassy dynamics, including fragility, nonexponential relaxation, subdiffusion, and violation of the SE relation. The correlation of local events together with the decoupling of persistence and exchange times is demonstrated in Sec.~\ref{sec:anal}, while the roles of the Coulombic interactions are discussed in Sec.~\ref{sec:coul}. Concluding remarks are offered in Sec.~\ref{sec:conclusion}.
\section{RTIL models}
\label{sec:model}
\begin{figure}
\centering
\resizebox{\textwidth}{!}{
\includegraphics{./EPS/fig1}
}
\caption{Coarse-graining scheme. ${\rm EMI}^+$ is reduced to a 4-atom cation, where 5 atoms of the imidazolium ring of ${\rm EMI}^+$ and 3 hydrogen atoms directly attached to it are collapsed to a single united atom (T1) in the coarse-grained model. The methyl and ethyl groups of ${\rm EMI}^+$ are represented, respectively, as one-atom (M1) and two-atom (M3 and E1) moieties and ${\rm PF}_6^-$ is simplified as a united atom as in Ref.~\citenum{shim1}.}
\label{fig:cg}
\end{figure}
In this section, we give a brief explanation of our coarse-grained model\cite{jeong:thesis} for ${\rm EMI}^+{\rm PF_6}^-$
and compare its results for structure and translational dynamics with those of a more atomistic description.
\subsection{Coarse-grained model}
\label{sec:model:cgm}
Our coarse-grained model is based on the RTIL description used by Kim and coworkers\cite{shim1} to study various processes in RTILs.\cite{shim:letter,*shim2,shim:rtil:rot,jeong:1/f,shim:rtil:et,*shim:rtil:sn1,*shim:rtil:diex} Specifically, they employed the united atom representation for the methyl group (M1) as well as the CH$_2$ (E1) and CH$_3$ (M3) moieties of the ethyl group of cations. They used the AMBER force field\cite{amber} for the Lennard-Jones (LJ) interactions and partial charge assignments\cite{lynden-bell:parameter} proposed by Lynden-Bell and coworkers for Coulombic interactions. ${\rm PF_6}^-$ was also described as a united atom. Hereafter, this model will be referred to as the AM description. In our coarse-grained model (CGM), we further simplify the cation description by representing the imidazole ring and H atoms directly attached to it as a single atomic site T1 (Fig.~\ref{fig:cg}). Thus each EMI$^+$ ion consists of 4 united-atom sites, M1, M3, E1 and T1 in CGM. The LJ parameters of T1 were adjusted, so that CGM reproduces the liquid structure of the atomistic AM description reasonably well (see below). For M1, M3, E1 and PF$_6^-$, we used the parameters of AM without any further adjustment.
\begin{table}[!t]
\centering
\caption{The LJ parameters, partial charges, and masses of coarse-grained atoms}
\vspace*{15pt}
\begin{tabular}{c|cccc}
\hline
atom & $\sigma_{ii}$ (\AA) & $\epsilon_{ii}$ (kJ/mol) & $q_i$ $(e)$ & mass (amu) \\
\hline
M1 & 3.905 & 0.7330 & 0.316 & 15.04092 \\
T1 & 4.800 & 1.5000 & 0.368 & 67.08860 \\
E1 & 3.800 & 0.4943 & 0.240 & 14.03298 \\
M3 & 3.800 & 0.7540 & 0.076 & 15.04092 \\
PF6 & 5.600 & 1.6680 & -1.000 & 144.97440 \\
\hline
\end{tabular}
\label{table:param}
\end{table}
We performed MD simulations of EMI$^+$PF$_6^-$ in both the AM and CGM representations using the DL\_POLY program.\cite{dlpoly}. Atoms $i$ and $j$ at positions ${\bf r}_i$ and ${\bf r}_j$ interact with each other through the LJ and Coulomb potentials: \begin{equation}
U_{ij}=4\epsilon_{ij}\left[ \left(\frac{\sigma_{ij}}{r_{ij}}\right)^{12}-
\left(\frac{\sigma_{ij}}{r_{ij}}\right)^{6} \right]+\frac{q_iq_j}{r_{ij}},
\end{equation}
where $r_{ij}\equiv |{\bf r}_i -{\bf r}_j|$ is the distance between the two atoms. The parameters of CGM employed in the present study are compiled in Table~\ref{table:param}. For the AM parameters, the reader is referred to Ref.~\citenum{shim1}. For LJ interactions between unlike atoms, the Lorentz-Berthelot combining rules were used.~\cite{Allen}
The simulation cell of the CGM ionic liquid comprises 512 pairs of rigid cations and anions. We performed simulations in the canonical ensemble at six different temperatures, $T=300, 350, 400, 475, 600$ and $800\,{\rm K}$, using the Nos{\'e}-Hoover thermostat and at density $\rho = 1.31\,{\rm g/cm^{3}}$. Periodic and cubic boundary conditions were employed and long-range electrostatic interactions were computed via the Ewald method. Starting from a crystal configuration obtained by alternating the cations and anions, we equilibrated the system for $2\,{\rm ns}$ prior to production runs at $800\,{\rm K}$. At lower temperatures, we used as an initial configuration one of the equilibrated configurations at higher $T$ that is closest to the temperature of the system under consideration and equilibrated the system for $2\,{\rm ns}$ at $600\,{\rm K}$ and $475\,{\rm K}$, $5\,{\rm ns}$ at $400\,{\rm K}$, and $10\,{\rm ns}$ at $350\,{\rm K}$ and $300\,{\rm K}$. Production runs following equilibration were $5\,{\rm ns}$ in length for $800\,{\rm K}$ and $600\,{\rm K}$, $10\,{\rm ns}$ for $475\,{\rm K}$ and $400\,{\rm K}$, $50\,{\rm ns}$ for $350\,{\rm K}$, and $60\,{\rm ns}$ for $300\,{\rm K}$. At each thermodynamic condition, we carried out six independent production runs, from which the averages were computed. Thus we used, for example, six $60\,{\rm ns}$ trajectories to analyze various dynamic quantities at $300\,{\rm K}$ in CGM.
In the case of AM, we considered 112 pairs of EMI$^+$ and PF$_6^-$ ions with $\rho = 1.31\,{\rm g/cm^{3}}$ at $T=350$ and $400\,{\rm K}$. At either temperature, we performed three independent simulations, each of which was carried out with $10\,{\rm ns}$ equilibration, followed by a $40\,{\rm ns}$ trajectory. Ensemble averages were calculated using three trajectories thus obtained.
\begin{figure}
\centering
\resizebox{\textwidth}{!}{
\includegraphics{./EPS/fig2a}
\includegraphics{./EPS/fig2b}
}
\caption{Radial distributions of ions in EMI$^+$PF$_6^-$ at (a) $350\,{\rm K}$ and (b) $400\,{\rm K}$. The results of CGM and AM are plotted in solid and dashed lines, respectively. The center of mass position is used to represent cation locations in CGM and AM.}
\label{fig:str}
\end{figure}
\subsection{Structure}
\label{sec:model:structure}
Here we consider MD results of the CGM and AM for structure to gain insight into how realistic the former description is. In Fig.~\ref{fig:str}, their results for radial distributions of ions at $T = 350\,{\rm K}$ and $400\,{\rm K}$ are compared. There we employed the center of mass to represent the cation positions for CGM and AM. We notice that CGM captures the RTIL structure of AM very well. The two models yield a excellent agreement in both the peak positions and heights, including minor secondary peaks, e.g., structure near $7\,{\textrm \AA}$ in the cation-anion distribution. Even in the case of the main peak of the cation-anion distribution which shows the largest deviation between the two, the discrepancy in peak location is only $\sim\!0.02\,{\textrm \AA}$. Considering the drastic approximation of the planar imidazole ring as a spherical atom, we think that overall the coarse-grained model fares very well in reproducing the RTIL structure of AM.
\subsection{Translational dynamics}\label{sec:model:translation}
To understand the effect of our coarse-graining (Fig.~\ref{fig:cg}) on system dynamics, we examine ion translational motions by considering their mean square displacement,
$\Delta (t)=\langle N^{-1}\sum_{i=1}^{N}|{\bf r}_i(t)-{\bf r}_i(0)|^2 \rangle$,
and self-intermediate scattering function,
$F_s(q_0,t)\equiv \langle N^{-1} \sum_{i=1}^{N} e^{i {\bf q}_0\cdot[{\bf r}_i(t)-{\bf r}_i(0)]}\rangle$.
Here $\langle \cdots \rangle$ denotes the equilibrium ensemble average, $N$ the number of ions, and ${\bf q}_0$ the wave vector corresponding to the position of the first peak in the static structure factor.
\begin{figure}
\centering
\resizebox{\textwidth}{!}{
\includegraphics{./EPS/fig3}
}
\caption{Mean square displacement $\Delta(t)$ of (a) cations and (b) anions in CGM and AM at $T=350\,{\rm K}$ and $400\,{\rm K}$.}
\label{fig:msd_c}
\end{figure}
\begin{figure}
\centering
\resizebox{\textwidth}{!}{
\includegraphics{./EPS/fig4}
}
\caption{Self-intermediate scattering function $F_s(q_0,t)$ for (a) cations and (b) anions in CGM and AM at $T=350\,{\rm K}$ and $400\,{\rm K}$. The wave vector employed in (a) and (b) are, respectively, $q_0=0.858\,\textrm{\AA}^{-1}$ and $0.878 \,\textrm{\AA}^{-1}$, corresponding to the positions of the first peak in their static structure factor.
}
\label{fig:isf_c}
\end{figure}
\begin{table}[!t]
\centering
\caption{MD results for the diffusion constant $D$ and the structural relaxation time $\tau_{\alpha}$ for CGM}
\vspace*{15pt}
\begin{tabular}{c cc cc}
\hline
$T$ (K) & \multicolumn{2}{c}{$D$ ($\textrm{\AA}^2/{\rm ps}$) } &
\multicolumn{2}{c}{$\tau_{\alpha}$ (ps)} \\
& cation & anion & cation & anion \\
\hline
\noalign{\vspace*{5pt}}
300 & $5.4\times 10^{-5}$& $9.6\times 10^{-6}$& 59600 & 244000 \\
350 & $4.3\times 10^{-4}$ & $9.7\times 10^{-5}$& 5030 & 12700 \\
400 & $2.2\times 10^{-3}$ & $7.2\times 10^{-4}$& 779 & 1610 \\
475 & $8.8\times 10^{-3}$ & $4.0\times 10^{-3}$& 154 & 261 \\
600 & $3.3\times 10^{-2}$ & $2.0\times 10^{-2}$ & 36.1 & 50.2 \\
800 & $9.9\times 10^{-2}$ & $6.8\times 10^{-2}$ & 11.4 & 13.5 \\
\noalign{\vspace*{5pt}}
\hline
\end{tabular}
\label{table:MDresults}
\end{table}
MD results in Figs. \ref{fig:msd_c} and~\ref{fig:isf_c} show that the coarse-grained model exhibits subdiffusive behavior and nonexponential relaxation, consistent with the atomistic model. We, however, notice that dynamics of the former proceed faster than the latter. To quantify this, we calculated the diffusion constant
$D=\lim_{t\rightarrow \infty} [6(t-t_0)]^{-1}{\langle N^{-1}\sum_{i=1}^{N}[{\bf r}_i(t)-{\bf r}_i(t_0)]^2\rangle}$,
where $t_0$ denotes the time when the Fickian behaviors appear in $\Delta (t)$,
and determined the structural relaxation time $\tau_{\alpha}$ via
\begin{equation}
\label{eq:taualpha}
F_s(q_0,\tau_{\alpha})={\rm e}^{-1}\ .
\end{equation}
The MD results for $D$ and $\tau_{\alpha}$ are compiled in Table~\ref{table:MDresults}. At $400\,{\rm K}$, we obtained $D=2.2 \times 10^{-3}\, {\textrm \AA}^2$/ps and 7.2 $\times 10^{-4}\, {\textrm \AA}^2$/ps for cations and anions, respectively, for CGM, while the corresponding values for AM were 6.5 $\times 10^{-4}\, {\textrm \AA}^2$/ps and 1.4 $\times 10^{-4}\, {\textrm \AA}^2$/ps. As for $\tau_{\alpha}$, the coarse-grained model yields 779\,ps and 1610\,ps for cations and anions, respectively, whereas the AM description results in significantly longer relaxation times, 3680\,ps and 9010\,ps. Because of this discrepancy, i.e., CGM is faster than AM in dynamics, $\Delta(t)$ and $F_s(q,t)$ of AM at $400\,{\rm K}$ are in better accord with those of CGM at $350\,{\rm K}$ than at $400\,{\rm K}$ (Figs. \ref{fig:msd_c} and~\ref{fig:isf_c}). (For CGM at $350\,{\rm K}$, simulations yield $D=4.3\times10^{-4}{\textrm \AA}^2$/ps and $9.7\times10^{-5} {\textrm \AA}^2$/ps, and $\tau_{\alpha}=5030$\,ps and $12700 \,{\rm ps}$ for cations and anions, respectively.) This is not surprising in that structure and dynamics of liquid systems, including ionic liquids, depend on molecular shape (i.e., LJ interactions) and charge distributions (viz., electrostatic interaction) of constituent particles.~\cite{Chiappe:cg,voth:cg} In addition to the difference in cation geometry, negative partial charges of nitrogen atoms of EMI$^+$ present in the AM description are completely absent in CGM because of the united atom representation of the imidazole ring. This simplification reduces charge anisotropy of cations and thus frustration in the structure and dynamics of CGM. We therefore expect that both rotational and translational dynamics would be accelerated in CGM, compared to AM. Here we consider only the translational dynamics of cations and anions and postpone the rotational dynamics for a future study.
\section{Glassy dynamics}
\label{sec:glassy}
In this section, we analyze the glassy behavior of the coarse-grained RTIL with the aid of MD simulation results. We examine structural relaxation and ion diffusion at various temperatures and demonstrate that our model belongs to the class of fragile glass formers and violates the Stokes-Einstein relation.
\label{sec:results}
\subsection{Structural relaxation and fragility}
\label{sec:fragility}
The structural relaxation through ion translational dynamics is usually described by the self-intermediate scattering function $F_s(q_0,t)$. In Fig.~\ref{fig:ISF}, the CGM predictions for $F_s(q_0,t)$ at six different temperatures are displayed. At high $T$, the ionic liquid behaves almost like a normal liquid; the thermal fluctuations dominate over the constraints of caged structures of ions, mainly formed by counterions. As the temperature decreases, characteristics of glassy dynamics, such as subdiffusivity and nonexponential relaxation, become more apparent.
\begin{figure}
\centering
\resizebox{\textwidth}{!}{
\includegraphics{./EPS/fig5}
}
\caption{Self-intermediate scattering function $F_s(q_0,t)$ for (a) cations and (b) anions in a coarse-grained ionic liquid at various temperatures. The wave vector $q_0$ is the same as in Fig.~\ref{fig:isf_c}.
In the $\alpha$-relaxation regime, $F_s(q_0,t)$ exhibits nonexponential decay, which is well described by $c\exp[-(t/\tau_0)^\beta]$ for all temperatures. }
\label{fig:ISF}
\end{figure}
At low $T$, the presence of the plateau regime ($\beta$ relaxation) and slow $\alpha$ relaxation, which are hallmarks of supercooled liquids, is rather prominent. The contribution of inertial dynamics in the first $0.1\,{\rm ps}$ or so accounts for less than 10\% of the entire relaxation of $F_s(q_0,t)$, and structural correlation persists for several decades in time, thereby indicating the highly viscous RTIL environment. The slow nonexponential relaxation subsequent to the plateau regime is well described by a stretched exponential function, $c \exp[-(t/\tau_0)^{\beta}]$. At $300\,{\rm K}$, the exponent $\beta$ is found to be 0.64 and 0.59 for cations and anions, respectively. A substantial deviation of these $\beta$ values from unity is another good indicator of the glassy dynamics in the RTIL system. As $T$ increases, so does $\beta$. At $800\,{\rm K}$, the highest temperature we studied, the $\beta$ values are 0.89 and 0.92. Even though greatly enhanced thermal fluctuations at this temperature accelerate structural relaxation immensely by more than four orders of magnitude compared to that of $300\,{\rm K}$, $F_s(q_0,t)$ still maintains a nonexponential character, presumably due to its high pressure condition. We thus expect that if we lower the pressure by reducing its density, the structural relaxation would become a single-exponential decay.
\begin{figure}
\centering
\resizebox{0.95\textwidth}{!}{
\includegraphics{./EPS/fig6a}
\includegraphics{./EPS/fig6b}
}
\caption{Temperature dependence of structural relaxation time $\tau_\alpha$, Eq.~\ref{eq:taualpha}. Dashed and dotted lines are the fits to MD results using (a) $\tau_\alpha=c_1 \exp(d_1/T^2)$ and (b) $\tau_\alpha=c_2 \exp(d_2/T)$. Error bars represent the maximum and the minimum values of $\tau_\alpha$ among six independent trajectories.}
\label{fig:SA}
\end{figure}
We turn to the temperature dependence of $\tau_\alpha$ in the CGM description presented in Figure~\ref{fig:SA} and Table~\ref{table:MDresults}. Since slow structural relaxation at $300\,{\rm K}$ does not allow the determination of its $\tau_\alpha$ directly from the simulation results, we estimated it by employing a stretched exponential fit to $F_s(q_0,t)$. At all other temperatures, $\tau_\alpha$ was determined using the MD results for $F_s(q_0,t)$ in Eq.~\ref{eq:taualpha}. The most salient aspect of our results in Figure~\ref{fig:SA} is that $\tau_{\alpha}$ does not follow the Arrhenius law $\tau_{\alpha}\propto\exp(d_2/T)$. Rather the structural relaxation time shows a super-Arrhenius behavior; specifically, it varies with the temperature as $\tau_{\alpha}\propto\exp(d_1/T^2)$. This means that the CGM RTIL studied here resembles a fragile glass former in the temperature dependence.
For clarity, we make a couple of remarks here. First, while the RTIL density is assumed to be fixed in the present study, it tends to decrease with increasing temperature for real ionic liquids. This density variation, if incorporated, would lead to acceleration of structural relaxation in CGM at high $T$, compared to the results in Fig.~\ref{fig:SA}. This would in turn strengthen the super-Arrhenius character of $\tau_\alpha$ and thus enhance the fragile behavior of CGM. Second, as pointed out in Sec.~\ref{sec:model:translation} above, temperatures of CGM and AM do not coincide. In other words, the temperature of CGM does not correspond to the actual temperature of the atomistic system. As a consequence, our finding that EMI$^+$PF$_6^-$ is a fragile glass former based on the CGM description might not be directly applicable to the real ionic liquid. We however note that a super-Arrhenius temperature dependence was observed in recent measurements in a similar RTIL.\cite{richert:fragile} We thus believe that our result on the fragility of EMI$^+$PF$_6^-$ is robust.
\subsection{Breakdown of the Stokes-Einstein Relation}
\label{sec:SE}
In a normal liquid, the diffusion constant $D$ is usually related to the viscosity $\eta$ via the Stokes-Einstein relation
\begin{equation}
D \propto \frac{T}{\eta}\ .
\label{eq:SE}
\end{equation}
For convenience, we consider another relation
\begin{equation}
D \propto \frac{1}{\tau_\alpha}\ ,
\label{eq:SE2}
\end{equation}
which is equivalent to Eq.~\ref{eq:SE} if the structural relaxation time is proportional to $\eta/T$. Eq.~\ref{eq:SE2} results when translational motions of constituent particles are described by a normal diffusion equation, the Gaussian solution of which is $F_s(q_0,t)=\exp (-q_0^2 Dt) \equiv \exp (-t/\tau_\alpha)$.
\begin{figure}
\centering
\resizebox{\textwidth}{!}{
\includegraphics{./EPS/fig7}
}
\caption{Mean square displacement of (a) cations and (b) anions in the CGM description. }
\label{fig:MSD}
\end{figure}
Simulation results for the mean square displacement of cations and anions at different temperatures are displayed in Fig.~\ref{fig:MSD}. Subdiffusion in the intermediate time scale, another common feature of supercooled liquids, is quite pronounced, especially at low $T$. At long times, ion translations tend to become normal diffusion. We notice that
the time scale associated with the transition from non-Fickian to Fickian dynamics generally coincides with the structural relaxation time.~\cite{szamel:Fick_time} Another noteworthy feature is that
for both $\Delta(t)$ and $F_s(q_0,t)$, motions of PF$_6^-$ tend to be slower than those of EMI$^+$. This is attributed to the fact that anions are heavier than cations.
\begin{figure}
\centering
\resizebox{\textwidth}{!}{
\includegraphics{./EPS/fig8a}
\includegraphics{./EPS/fig8b}
}
\caption{Breakdown of the SE relation in the coarse-grained ionic liquid. (a) Dashed and dotted lines are the fits with the scaling relation $D\sim\tau_\alpha^{-\xi}$, where the exponent $\xi$ is given by 0.87 and 0.92 for cations and anions, respectively. (b) $D\tau_{\alpha}$ deviates from
the constant value as the temperature decreases. Lines are drawn merely as a guide for the eyes.}
\label{fig:SE}
\end{figure}
For additional insight, we have analyzed the relation between $D$ and $\tau_\alpha$ via
\begin{equation}
\label{eq:Dxi}
D\sim\tau_\alpha ^{-\xi},
\end{equation}
where $\xi=1$ corresponds to the SE relation.
We found that $\xi=0.87$ and $0.92$ for cations and anions, respectively (Fig.~\ref{fig:SE}(a)). Thus the product $D\tau_\alpha$ develops a positive deviation from a constant value as $T$ decreases (Fig.~\ref{fig:SE}(b)).
This reveals a weak violation of the SE relation for our model RTIL system
The breakdown of the SE relation indicates that translational dynamics of the ionic liquid can not be described by the conventional diffusion equation, which is a continuity equation combined with a constitutive relation given by Fick's law. Specifically, it is assumed that the diffusion current is proportional to the spatial gradient of the particle concentration and obtains stationarity instantaneously in response to the external perturbation. Accordingly, the diffusion equation is valid only in the limit where the time is sufficiently coarse-grained to ensure instantaneous establishment of stationarity. If there is a significant delay in time before the system reaches stationarity, a crossover from the non-Fickian regime, characterized by anomalous diffusion and $\beta$ relaxation, to the Fickian regime happens. We attribute the main cause of the delay to cage dynamics, i.e., immobile cages that last for a long time (see below).
In this sense, the violation of the SE relation is a manifestation of dynamic heterogeneity,~\cite{jung:SE} which we turn to next.
\section{Correlated local excitations}
\label{sec:anal}
\subsection{Decoupling of exchange and persistence times}
\label{sec:decoupl}
\begin{figure}
\centering
\resizebox{\textwidth}{!}{
\includegraphics{./EPS/fig9}
}
\caption{Displacement of a cation from simulation trajectories at (a) $300\,{\rm K}$ and (b) $600\,{\rm K}$.
Occurrences of excitations at which the cation moves beyond $d$(=3.0 \AA) are marked with triangles. For a given initial position, the waiting time until the occurrence of the first excitation defines the persistence time $\tau_{\mathrm p}$, while the time
interval between subsequent excitations yields the exchange times $\tau_{\mathrm x}$.
}
\label{fig:traj}
\end{figure}
In ionic liquids, due to electrostatic interactions, a central ion is surrounded mainly by counterions in its immediate neighborhood, termed first solvation shell. Thus dynamics of the central ion will be influenced by those of the first solvation shell (cage) and vice versa. (In normal electrolytes, for example, this counter-ion cage, referred to as an ion atmosphere, tends to reduce the diffusion constant of the central ion at low concentration.) Suppose that the central ion undergoes thermal motions in the cage. Once in a while a large fluctuation enables it to escape the cage and subsequently the cage reorganizes. As the temperature is lowered, thermal fluctuations become suppressed and as a result, the frequency of the ion escape from the cage diminishes. While this picture may need quantitative elaboration, it is nonetheless useful to obtain a qualitative understanding of ``excitations'' introduced below and their properties in glassy environments.
We first introduce an excitation as a local event that an ion moves over a distance exceeding a threshold value $d$. The persistence and exchange times associated with excitations are then defined as follows:~\cite{jung:exc,Hedges:decoupl} the persistence time $\tau_{\rm p}$ is the waiting time $t_1$ for an ion $i$ to undergo its first excitation such that $|{\bf r}_i(t_1)-{\bf r}_i(0)|\geq d$, and the exchange time $\tau_{\rm x}$ includes a set of time intervals ${t_2, t_3,\cdots}$ between subsequent excitation events, i.e., $|{\bf r}_i(t_1+t_2)-{\bf r}_i(t_1)|\geq d$, $|{\bf r}_i(t_1+t_2+t_3)-{\bf r}_i(t_1+t_2)|\geq d$, etc. Accordingly, the frequency of excitations gauges the mobility of ions. In the present study, we employ 3.0~\AA\ for $d$. We will return to discuss other possibilities for $d$ later on.
In a glassy environment where cage reorganization is slow, the likelihood of an ion undergoing a second excitation after its first one is higher than that of the initial excitation because the liquid structure disturbed by the first excitation generally provides a favorable environment for another excitation.
In other words, the excitations are not governed by a Poisson process.
In Fig.~\ref{fig:traj}, typical trajectories of an ion at $300\,{\rm K}$ and $600\,{\rm K}$ are displayed. On each trajectory, the events of local excitations are marked with triangular symbols. As expected, excitations at $300\,{\rm K}$ are rare events; their occurrences are irregular and intermittent. This, for instance, leads to a non-exponential tail on the long-time end of the distribution of exchange times, i.e., time intervals between two consecutive excitations (see below). By contrast, the trajectory at $600\,{\rm K}$ is characterized by more regular and frequent excitations than that at 300\,K.
\begin{figure}
\resizebox{\textwidth}{!}{
\centering
\includegraphics{./EPS/fig10}
}
\caption{Decoupling of persistence and exchange times for (a) cations and (b) anions in a coarse-grained ionic liquid.
The persistence time $\tau_{\mathrm p}$ and the exchange time $\tau_{\mathrm x}$ are defined as the waiting times for an ion to produce a displacement exceeding $d$, taken as $3.0\,{\textrm \AA}$, for the first time and thereafter, respectively.
The solid and dashed lines correspond to the probability distributions of $\log \tau_{\mathrm p}$ and $\log \tau_{\mathrm x}$, respectively.}
\label{fig:decoupl}
\end{figure}
Figure~\ref{fig:decoupl} exhibits the probability distributions of the logarithms of persistence and exchange times for cations and anions. At $800\,{\rm K}$, the persistence and exchange times show nearly identical distributions. This clearly indicates that the excitation events follow the Poisson process, viz., they are not correlated. As $T$ decreases, two distributions become distinct and their difference increases. The center of the distribution for $\tau_{\rm p}$ shifts to longer time much more rapidly than the corresponding distribution for $\tau_{\rm x}$. As a result, the deviation between the average exchange time $\langle\tau_{\rm x}\rangle$ and average persistence time $\langle\tau_{\rm p}\rangle$ rises markedly with lowering $T$ (see Fig.~\ref{fig:txtp}). This decoupling of $\tau_{\rm x}$ and $\tau_{\rm p}$ observed here shows that the excitations become increasingly more correlated as $T$ decreases, exposing the dynamically heterogeneous nature of our RTIL system at low $T$. We also notice dramatic enhancement of $\langle\tau_{\rm p}\rangle$ at low $T$, compared to $\langle\tau_{\rm x}\rangle$. For instance, $\langle\tau_{\rm p}\rangle$ is longer than $\langle\tau_{\rm x}\rangle$ by one order of magnitude at 300\,K. This suggests the development of persisting immobile regions. According to a recent MD study in a similar RTIL,\cite{Zhao:ionpair} the life time of stable contact ion pairs that seldom move a large distance as a pair can exceed a few nanoseconds. Such long-lived pairs could potentially be a candidate for immobile regions associated with long $\tau_{\rm p}$. Finally, as mentioned above, we notice that the distributions of both $\tau_{\rm x}$ and $\tau_{\rm p}$ develop a long non-exponential tail on the long-time end as $T$ diminishes (note that the logarithmic time scale is employed in Fig.~\ref{fig:decoupl}).
This is another indicator that the excitations at low $T$ do not obey Poisson statistics and thus are correlated.
\begin{figure}
\centering
\resizebox{0.5\textwidth}{!}{
\includegraphics{./EPS/fig11}
}
\caption{Average persistence and exchange times versus the inverse temperature for $d= 3.0\,\textrm{\AA}$. Lines are drawn as a guide for the eyes.}
\label{fig:txtp}
\end{figure}
\subsection{Analysis of threshold distance dependence}
In the previous subsection, we chose $d=3.0\,{\textrm \AA}$ as the threshold distance in the definition of excitations.
The decoupling of the exchange and persistence times is present in a glassy environment with a physically meaningful value of $d$ chosen, because it is attributed to strong correlations of local events.
Here we examine how a different choice for the $d$ value would influence our analysis.
For example, if we choose a $d$ value considerably less than the distance between neighboring ions, which is approximately $5\,{\textrm \AA}$ for our system (Fig.~\ref{fig:str}), excitations will correspond to small fluctuations of a central ion inside its counterion cage. We can easily imagine that such excitations seldom induce a considerable structural change in the local environment.
Excitations exceeding $\sim5.0\,{\textrm \AA}$ on the other hand describe delocalized hopping or gradual drift, which is usually accompanied by irreversible structural changes. Thus, too large or too small a value for $d$ in the working definition of excitations would not be able to capture, e.g., the decoupling of persistence and the exchange times even though the decoupling occurs irrespective of our choice of $d$.
\begin{figure}
\centering
\resizebox{\textwidth}{!}{
\includegraphics{./EPS/fig12a}
\includegraphics{./EPS/fig12b}
}
\caption{Log-log plot of (a) the average exchange time $\langle \tau_{\mathrm x} \rangle$ and (b) the average persistence time $\langle \tau_{\mathrm p} \rangle$ for cations versus the threshold distance $d$. Lines are drawn as a guide for the eyes in (a) and (b).}
\label{fig:d}
\end{figure}
In Fig.~\ref{fig:d}, we display $\langle\tau_{\rm x}\rangle$ and
$\langle\tau_{\rm p}\rangle$ of the cation versus $d$ at various temperatures.
We notice that $\langle\tau_{\rm x}\rangle$ and $\langle\tau_{\rm p}\rangle$ increase with $d$ according to the power law, $\langle\tau_{\rm x,p}\rangle\sim d^{\delta}$, where $\delta$ ranges approximately from 2 at $800\,{\rm K}$ to 4 at $300\,{\rm K}$.
In the scaling relation, the value $\delta\simeq2$ at $800\,{\rm K}$ corresponds to the diffusive regime, whereas larger values at lower temperatures indicate anomalous diffusion. Also the $d$-dependence of $\langle\tau_{\rm x}\rangle$ and $\langle\tau_{\rm p}\rangle$ becomes stronger as $T$ decreases.
\begin{figure}
\centering
\resizebox{\textwidth}{!}{
\includegraphics{./EPS/fig13a}
\hspace*{-20pt}
\includegraphics{./EPS/fig13b}
\hspace*{-20pt}
\includegraphics{./EPS/fig13c}
}
\caption{(a) Average exchange time $\langle\tau_{\mathrm x}\rangle$ vs. average persistence time $\langle\tau_{\mathrm p}\rangle$; (b) $\langle\tau_{\mathrm x}\rangle$ vs. inverse diffusion constant $D^{-1}$; (c) $\langle\tau_{\mathrm p}\rangle$ vs. structural relaxation time $\tau_{\alpha}$ for $d=5.0\,{\textrm \AA}$, $3.0\,{\textrm \AA}$, and $\sqrt{5}\,{\textrm \AA}$.
Lines represent results fitted to (a) Eq.~\ref{eq:tauxtaup}; (b) and (c) the scaling behaviors $\sim x^{\gamma}$, where $x$ corresponds to $D^{-1}$ and $\tau_{\alpha}$, respectively. In all cases, the value of the scaling exponents $\nu$ and $\gamma$, given for each $d$, is shown to increase with $d$.
}
\label{fig:ddep}
\end{figure}
In Sec.~\ref{sec:SE}, the violation of the SE relation has been demonstrated via Eq.~\ref{eq:Dxi}
with $\xi=0.87$ for cations.
Likewise, the scaling relation
\begin{equation}
\label{eq:tauxtaup}
\langle\tau_{\rm x}\rangle\sim\langle\tau_{\rm p}\rangle^\nu,
\end{equation}
is characterized by the exponent $\nu$, which is also smaller than unity.
As shown in Fig.~\ref{fig:ddep}(a), with $d=5.0\,\textrm{\AA}$, $\nu$ is found to be $0.80$ for cations. With $d=3.0$\,\AA, the corresponding $\nu$ value is
0.67.
We point out that the decoupling between $\langle\tau_{\rm x}\rangle$ and $\langle\tau_{\rm p}\rangle$ appears greater than that of the SE violation, that is, $\nu$ is smaller than $\xi$ in Eq.~\ref{eq:Dxi}. It is known that $\tau_{\alpha}$ determined from $F_s(q_0, t)$ is identified with the average persistence time, while the exchange processes contribute to the diffusion if $d$ is comparable to the size of the ions ($d \gtrsim 5.0\,{\textrm \AA}$).~\cite{Chandler:prl,jung:exc,berthier:Fick_length}
In this perspective, we analyzed the scaling behaviors of $\langle \tau_{\rm x}\rangle$ with $D^{-1}$ and $\langle \tau_{\rm p}\rangle$ with $\tau_{\alpha}$.
The results in Fig.~\ref{fig:ddep}(b) and (c) show that the former is characterized by weaker variations than the latter.
We also notice that the coupling behavior becomes stronger with $d$, i.e., the exponent in the scaling relation increases, for both cases.
But regardless of $d$, $\langle\tau_{\rm x}\rangle$ shows a significant sublinear behavior in $D^{-1}$.
This is due to the fragile characteristic of our model system, where the exchange events are correlated.~\cite{jung:exc}
Therefore we expect that its scaling exponent would not reach unity even if a significantly larger value for $d$ is employed in the definition of $\langle\tau_{\rm x}\rangle$.
By contrast, proportionality between $\langle\tau_{\rm p}\rangle$ and $\tau_{\alpha}$ is expected in the limit $d \rightarrow 2\pi/q_0$.
Note that the largest value of $d$ employed in Fig.~\ref{fig:ddep}(c) is $5.0\,{\textrm \AA}$, which is still less than $2\pi/q_0 (=7.32$\,\AA).
\subsection{Comparison between excitation and brachiation}
\label{sec:brach}
\begin{figure}
\centering
\resizebox{\textwidth}{!}{
\includegraphics{./EPS/fig14.eps}
}
\caption{Displacements of a cation and associated numbers of local excitation and brachiation events, $m(t)$ and $b(t)$. $m(t)$ and $b(t)$ were determined by counting the excitations and brachiation events for the central cation occurring during a time window of $500\,{\rm ps}$ along its MD trajectory. If neighboring anions revert back to original configurations within $200\,{\rm ps}$ of their initial changes, the corresponding events were not considered as brachiation according to Ref.~\citenum{Wu:brach}.
}
\label{fig:brach}
\end{figure}
In an effort to understand ion displacements and related diffusion in RTILs, a mechanism based on brachiation processes, viz., a central ions moves by forming and breaking links to neighboring counterions through Coulomb interactions, has been proposed recently.~\cite{Wu:brach} While this has some similarity to local excitations and cage dynamics, it is essentially a structure-based description that lacks dynamic correlation effects. To see this, we have performed a comparative analysis of brachiation and local excitations. In view of the brachiation length scale,\cite{Wu:brach} we reduced the cutoff distance $d$ slightly to $\sqrt{5}\,{\textrm \AA}$. In Fig.~\ref{fig:brach}, the MD results for a typical cation at $350\,{\rm K}$ together with the numbers of local excitations and brachiation events, $m(t)$ and $b(t)$, are displayed. We notice that while $m(t)$ and $b(t)$ show a significant overlap in certain part of the trajectory, there is in general no strong correlation between the two. For instance, periods like $11 \,{\rm ns} < t< 13$\,ns and
$30\, {\rm ns} < t< 36$\,ns are characterized by large $b(t)$ but vanishing $m(t)$. This means that neighboring ions reorganize in the presence of an immobile central ion. The opposite situation where $m(t)$ is larger than $b(t)$, e.g., $23 \,{\rm ns} < t< 25$\,ns and
$27 \, {\rm ns} < t< 29$\,ns, also occurs frequently, indicating ion translations without any considerable change of its neighbors. The weak correlation between local excitations and brachiation seems to suggest that the Coulomb interaction with neighboring ions does not play a major role in translational dynamics of individual ions in RTILs (see below).
\section{Role of Coulomb interaction}
\label{sec:coul}
Finally, we consider roles played by Coulomb interactions in the glassy behaviors of RTILs. \cite{shim:rtil:ver}
To this end, we compare with a model supercooled liquid that has dynamic characteristics similar to our EMI$^+$PF$_6^-$ system but does not have long-range Coulombic interactions.
We employ a model liquid system with the WCA potentials of Ref.~\citenum{Hedges:decoupl} as a reference to quantify the influence of Coulomb interactions. For easy comparison, we follow Ref.~\citenum{Hedges:decoupl} to measure time and temperature in units of ${(m_{i}\sigma_{ii}^2/\epsilon_{ii})}^{1/2}$ and $\epsilon_{ii}/k_{\rm B}$, respectively. We use the anion parameters, i.e., $i=$ PF$_6^-$, so that ${(m_{i}\sigma_{ii}^2/\epsilon_{ii})}^{1/2}=27.3$ ps and $\epsilon_{ii}/k_{\rm B}=180\,{\rm K}$. Thus $\tau_\alpha=5030\,{\rm ps}$ at $T=350\,{\rm K}$ corresponds to $\tilde\tau_{\alpha}=184$ at $\tilde T=1.94$ in the new scaled units. The reader is reminded that the RTIL system at this temperature is characterized by strong nonexponential relaxation (Fig.~\ref{fig:ISF}). By contrast, the WCA mixture does not exhibit
glassy dynamics at all at this temperature. In fact, dynamics comparable to our RTIL at $\tilde T=1.94$ occur at a much lower temperature $\tilde T=0.4$ for the WCA system.\cite{Hedges:decoupl} This reveals that at a given (scaled) temperature, structural relaxation dynamics in EMI$^+$PF$_6^-$ in the CGM description are considerably slower than those in supercooled liquids characterized by short-range interactions only. This indicates that the Coulomb interactions in effect suppress structural fluctuations and enhance trapping in cage structures.
To gain insight at the microscopic level, we briefly analyze instantaneous power inputs from the RTIL environment to an ion $i$
via the Coulomb and LJ forces, i.e., ${\bf F}^{\rm Coul}_i\cdot {\bf v}_i$ and ${\bf F}^{\rm LJ}_i\cdot {\bf v}_i$ (${\bf v}_i=$ ion velocity).
For simplicity, we consider only the united atom T1 for cations because it is the heaviest atom, located close to the cation center of mass. We examined the static correlation between the two instantaneous powers using
the Pearson correlation coefficient $\rho_{X,Y}$:~\cite{Kanji}
\begin{equation}
\rho_{X,Y}=\frac{\langle (X-\langle X\rangle)(Y-\langle Y \rangle)\rangle}
{\sigma_X \sigma_Y},
\end{equation}
where $X$ and $Y$ denote ${\bf F}^{\rm Coul}_i\cdot {\bf v}_i$ and ${\bf F}^{\rm LJ}_i\cdot {\bf v}_i$, respectively, and $\sigma_X$ and $\sigma_Y$ are their standard deviations.
\begin{figure}
\centering
\resizebox{0.5\textwidth}{!}{
\includegraphics{./EPS/fig15.eps}
}
\caption{Correlation coefficient $\rho_{X,Y}$ versus the inverse temperature, where $X$ and $Y$ denotes
the instantaneous powers by the Coulomb force and the Lennard-Jones force.
As the temperature is lowered, the negativity of linear correlations between $X$ and $Y$ becomes apparent.
}
\label{fig:cc}
\end{figure}
The results in Fig.~\ref{fig:cc} disclose that power inputs arising from the LJ and Coulomb forces are anti-correlated.
Therefore, LJ and Coulomb interactions play antagonistic roles in energy relaxation of individual ions. If we calculate the time integration of ${\bf F}^{\rm Coul}_i\cdot {\bf v}_i$ and ${\bf F}^{\rm LJ}_i\cdot {\bf v}_i$, however, the contribution from the former nearly vanishes and work to individual ions is delivered primarily by the LJ force(data not shown). This result together with our comparative analysis above paints the picture that while the liquid structure of RTILs and thus the energy scale, i.e., $\tilde T$, relevant for glassy dynamics are mainly determined by Coulomb interactions, their relaxation is generally governed by the LJ interactions. We believe this explains why dynamic aspects of glassy RTILs are very similar to those of non-ionic supercooled liquids despite their major difference in long-range interactions. We note that the anti-correlation of power inputs from the LJ and Coulomb forces and the dominance of the former in relaxation dynamics of the ionic liquid observed here also apply to vibrational energy relaxation in RTILs.\cite{shim:rtil:ver}
\section{Conclusions}
\label{sec:conclusion}
We have introduced the coarse-grained model of ionic liquids and probed its dynamical behaviors.
Coarse-graining has simplified the geometry of the system and made dynamics accelerated, compared with the atomistic model. Nevertheless, the overall liquid structure and glassy dynamic properties such as nonexponential structural relaxation and subdiffusive behavior are preserved.
Owing to the reduced number of atoms by coarse-graining, we have been able to perform extensive MD simulations at long times over a wide range of temperature, and investigated the temperature dependence of structural relaxation and diffusion.
We found that our model for ionic liquids belongs to fragile glass formers, where $\tau_{\alpha}$ exhibits strong non-Arrhenius dependence on the temperature.
In addition, the SE relation is violated, which implies that diffusion is enhanced when compared with structural relaxation.
We pay attention to the apparent universality of the abovementioned dynamic features observed in a variety of glassy liquid systems, regardless of the nature of their molecular interactions.
In previous studies of supercooled liquids, kinetic constraints have been successfully employed to explain the peculiar dynamic properties of supercooled liquids in view of facilitated dynamics. Dynamic facilitation emphasizes the dominant role of dynamic correlations in glassy dynamics rather than static properties such as the structure factor and potential energy landscape.~\cite{Stillinger:landscape}
In a similar fashion, we have defined a local excitation in the dynamics of our model and observe dynamic correlations between them. Decoupling of persistence and exchange times has been shown to be highly correlated with local excitations.
Such decoupling behavior exists regardless of the threshold distance $d$, which defines local excitations. However, $d$ determines the physical meaning of local excitation events, especially at low temperatures.
To understand the influence of the Coulomb interactions, we compared structural relaxation dynamics of the RTIL with those of non-ionic models of supercooled liquids. We found that glassy dynamics occur at a considerably higher temperature (in scaled dimensionless units) for the former than for the latter.
We have also investigated instantaneous powers arising from the Coulomb and LJ forces. It was found that they are anti-correlated and their time integration is dominated by the latter. These results seem to indicate that relaxation dynamics of RTILs are dominated by the LJ interactions, while the Coulomb interactions exert a strong influence on the liquid structure and thus set the temperature scale for glassy dynamics.
At low temperatures, immobile regions persist for a long time due to sparse excitations. Once a local excitation occurs, subsequent displacements of ions are more probable, which tends to introduce mobile regions. Thus, decoupling of persistence and exchange times is a plausible explanation for dynamic heterogeneity of glassy liquids.
In the future, we plan to investigate spatiotemporal correlations of local excitation events in order to characterize in more detail dynamic heterogeneity in the RTIL.
\section*{Acknowledgments}
This work was supported by the National Research Foundation of Korea (Grant Nos. R11-2007-012-03003-0, 2009-0070517, and 2009-0080791), the KISTI Supercomputing Center (KSC-2007-500-3008 and KSC-2009-502-0003), and the BK21 Program.
\bibliographystyle{rsc}
|
\section{Introduction}
According to the holographic principle \cite{tHooft},
only a finite amount of information
is allowed to be stored in a region with given
bounding area,
scaling the limit as the area itself.
Its string-theoretical realization was considered
in \cite{Susskind}.
The original motivation has been very general,
since the principle was introduced through combination
of gravitational collapse and
the basic tenets of quantum mechanics.
Somehow, the gravitational context is found to highlight
a fundamental redundancy,
not visible otherwise,
in the quantum-mechanical degrees of freedom
used to describe the systems.
Through a semiclassical discussion,
the aim of this note is
to take the reverse path
and to spot
consequences, if any, on basic physics,
once the holographic principle
is assumed as primeval starting point.
The most general formulation of the holographic principle
at semiclassical level --i.e. with matter degrees of freedom living
in a continuous background spacetime--
is perhaps the generalized covariant entropy bound \cite{FMW},
which states that the matter entropy $S$ on a terminated lightsheet $L$
is bounded by (in Planck units, the units we will use
throughout this note)
\begin{eqnarray}\label{gceb}
S \leq \frac{\Delta A}{4},
\end{eqnarray}
where ${\Delta A}$ is the area difference
between the start- and end-surfaces of $L$,
and $S$ is calculated through the entropy current $s^a$
(assumed to exist),
$S = \int_{L} s^a \epsilon_{abcd}$,
being $\epsilon_{abcd}$
the spacetime (which we take 4-dim)
4-volume form.
In this work, inequality (\ref{gceb}) is assumed to be the precise
mathematical formulation of the holographic principle.
We recall a condition for (\ref{gceb})
--or a consequence of it when (\ref{gceb}) is assumed as starting point--
we shall use throughout the paper,
which shows up
in circumstances in which the effects of spacetime curvature
are extremely tiny.
The condition,
derived and discussed in \cite{Pesci2, Pesci3},
is as follows :
In a spacetime with Einstein's equation,
inequality (\ref{gceb}) is
universally true
if and only if
a local (i.e. depending on the point)
lower-limiting spatial
scale $l^*$, unrelated to gravity,
is assumed to exist in the description
of statistical systems,
with
\begin{eqnarray}\label{lstar}
l^* \equiv \frac{1}{\pi} \frac{s}{\rho + p} =
\frac{1}{\pi T} \ \left( 1 - \frac{\mu n}{\rho + p} \right)
\end{eqnarray}
(where, for the last expression, use of Gibbs-Duhem relation is made).
Here
$s$, $\rho$, $n$, $p$, $T$, $\mu$ are respectively
local entropy, mass-energy and number densities,
local pressure, temperature and chemical potential
(having, this latter, any rest energy included).
For thin plane layers of thickness $l$
--actually, that geometric configuration which to the utmost
challenges the bound--
this becomes
\begin{eqnarray}\label{condition1}
l^* \leq
l,
\end{eqnarray}
meaning that below $l^*$ the notions themselves of energy, entropy and
pressure in the layer become somehow undefined, or, equivalently,
that layers thinner than $l^*$ cannot be cutted physically if the assigned
values of the thermodynamic parameters have to remain unchanged after
cutting.
The reason why this limiting scale $l^*$ should exist
can be recognized through consideration
of the trivial lightsheets associated with thin plane layers;
for them,
$\Delta A$ in (\ref{gceb})
is, from Raychaudhuri's equation \cite{Wald_book1},
quadratic in $l$, whereas $S$ is (obviously) linear,
so that, if a lower limit would not be envisaged for $l$,
when $l \rightarrow 0$ inequality (\ref{gceb}) would be definitely violated.
The bound (\ref{gceb})
is attained
iff i) we consider plane layers
and ii) their thickness just attains the bound (\ref{condition1}).
$l^*$, considered as a time, instead of space, lower-limiting scale
(i.e. a lower-limiting time scale in the evolution
of statistical systems given by
the time it takes light to travel a distance $l^*$),
leads also to foresee \cite{Pesci4}
the universal bound
to the relaxation times \cite{Hod}
of perturbed thermodynamic systems.
As discussed in \cite{Pesci5},
from a quantum-mechanical standpoint
relation (\ref{condition1}) can be re-expressed as
\begin{eqnarray}\label{condition2}
l^* \leq
\lambda,
\end{eqnarray}
where $\lambda$ is the `typical'
quantum wavelength of constituent particles;
a notion which will be sharpened later.
What is meant here is that,
assuming that quantum mechanics allows
for unaltered thermodynamic potentials
in physically cutted slices as thin as the $\lambda$ itself
of the constituent particles,
i) $\lambda$ is the minimum $l$ quantum-mechanically allowed
and ii) (\ref{condition1}) leads to (\ref{condition2}).
For a given system,
$\lambda$ in (\ref{condition2})
provides the tightest bound coming from holography
to that combination
of thermodynamic potentials which we denote with $l^*$.
Only material media with $l^* = \lambda$
can attain the bound (\ref{gceb}), which is indeed attained
when (lightsheets trivially constructed on)
plane layers are considered with thickness $l = \lambda$.
\section{A holographic law in basic thermodynamics}
In condition (\ref{condition2}) any connection with gravity, or
curvature, has disappeared.
We are left with
a flat-spacetime condition,
which compares
a combination of the thermodynamic potentials of a system
with the quantum size of constituent particles;
still, this condition is
an implication of or a pre-requisite for holography.
The aim of this Section
is just to emphasize that, due to this,
(\ref{condition2})
can be read
as a sort of
basic law of thermodynamics of holographic origin,
and discuss its nature.
Let us look, first,
at that
a well-known basic thermodynamic bound
connecting energy $E$, entropy $S$ and size
--the circumscribing radius $R$ actually--
of a system
has already been long since proposed.
The Bekenstein universal bound to specific entropy \cite{BekSE}
\begin{eqnarray}\label{ueb}
\frac{S}{E} \leq 2 \pi R
\end{eqnarray}
indeed,
though
originally found through an argument involving black hole physics,
was since the beginning recognized as
a fundamental thermodynamic bound having nothing to do
with gravity.
In \cite{Pesci5} the relation of bound (\ref{ueb})
with holography has been discussed
(see also \cite{Boussobek}).
At first sight the bounds (\ref{ueb}) and
the holographic relation (\ref{condition1})
seem quite different.
One difference is that
in (\ref{condition1})
$p$ appears also (in $l^*$).
As this regard
we point out that,
considering the conditions of the original argument
bringing to (\ref{ueb}) through the use of the
generalized second law \cite{Bekgen1,Bekgen2},
in circumstances
more general than those originally considered
a contribution from
the work done by pressure should also
be present.
The basic fact used in the derivation of (\ref{ueb})
is, indeed, that if a body with energy $E$
and circumscribing radius $R$ (and negligible self-gravity)
is swallowed by a black hole,
a lower limit definitely exists to the increase
of surface area of the hole, given by $8 \pi E R$
\cite{Bekgen1,Bekgen2}.
Then, from this, and imagining that given a body
a process can always be found
for which this limit is attained,
through the use of the generalized second law for such a process
(\ref{ueb}) is obtained.
Now,
if we consider for simplicity a static black hole
and, instead of assuming that a whole body is swallowed,
we dump in,
just when it
is at its first contact with the horizon,
a small element
of proper thickness $l$ and
cross-sectional area $A$
of an indefinitely extended fluid
(i.e. if, contrary to \cite{Bekgen1,Bekgen2},
we no longer require the stress tensor $T_{ab}$ to be vanishing
outside the body), for a perfect fluid
assumed momentarily at rest in the local static frame
of the metric
to first order in $T_{ab}$
(cf. \cite{Wald_book2})
we get
\begin{eqnarray}\label{deltam}
\Delta M =
\int_0^l dl^\prime \int d^2 S \ T_{ab} \xi^a k^b =
\kappa \int_0^l dl^\prime A l^\prime (\rho + p) =
\kappa A (\rho + p) \frac{l^2}{2} =
\kappa \frac{l}{2} (E + pV),
\end{eqnarray}
where $V$ is the proper volume of the fluid element,
$\kappa$ is the surface gravity of the hole,
$l^\prime$, the proper length in the fluid local frame,
is the chosen affine parameter for the null geodesics
(with tangent $k^a$) on the horizon,
$d^2 S$ is the cross-section element of the horizon at $l^\prime$
and $\xi^a$ is the Killing field, orthogonal to the horizon,
which is timelike at infinity.
In (\ref{deltam}),
use of the relation $\kappa l^\prime k^a = \xi^a$
with $\kappa = $ const
is made, which is appropriate in so far as
the change in the black hole geometry in the process
can be neglected,
which always is the case
provided we choose $M$ large enough.
Here we are also assuming $V$ small enough
to allow for thermodynamic potentials approximately
constant in it.
From (\ref{deltam}) we get
the minimum horizon area increase,
which is
$\frac{8\pi}{\kappa} \Delta M = 4\pi l (E + p V)$,
from which,
through use of the generalized second law,
we obtain
\begin{eqnarray}\label{mod_ueb}
\frac{S}{E + p V} \leq \pi l.
\end{eqnarray}
So,
if the element we drop in is, say,
a gas contained in a box,
and if the thermodynamic system we are considering
consists of both the gas {\it and} the constraining walls,
we are led to $S/E \leq \pi l$
(as in the derivation in \cite{Bekgen2}),\footnote{We are grateful
to R. Bousso for correspondence on this point.}
if instead our system
consists of the gas alone,
our argument says that
the fundamental bound
should be given
by expression (\ref{mod_ueb}),
i.e. with the term $pV$.
Inequality (\ref{mod_ueb}) manifestly coincides
with condition (\ref{condition1}).
In \cite{Pesci5},
it has been shown that,
even if we start from (\ref{mod_ueb}),
for macroscopic bodies we are lead anyway to (\ref{ueb})
(for whichever strength, indeed, of the gravitational effects).
Still,
the bound expressed by condition (\ref{condition2})
strongly differs from the Bekenstein bound,
being the former actually enormously tighter than the latter.
For
a spherical homogeneous system with radius $R$, for example,
the Bekenstein bound says that the ratio $S/E$ is bounded by something
orders of $R$, while according to (\ref{condition2}) this same ratio
is bounded by orders of $\lambda$, the `typical' wavelength of
constituent particles, with
always $\lambda \leq R$,
and
in general $\lambda \ll R$.
Thus,
holography, which, too, has among its consequences the Bekenstein bound,
can be seen to imply for basic thermodynamics
a bound, condition (\ref{condition2}),
in general extremely stronger than Bekenstein's one.
Bound (\ref{condition2})
seems thus could be considered as the fundamental requirement
in basic thermodynamics
of the whole gravity-thermodynamics connection.
It is the basic-thermodynamic imprint
of holography.
Let us give a closer look to bound (\ref{condition2}).
In the transition
from (\ref{condition1})
to (\ref{condition2}),
the particle `typical' wavelength $\lambda$
is, we said, a sort of minimum thickness
below which, in view of the uncertainty relations,
the value of thermodynamic potentials in a physically-cutted slice
are found different than before cutting.
It can be defined more precisely
considering that a
limiting thickness $l_{min}$ should exist
for which the quantum spread in momentum $\Delta p_x^{ind}$, induced
by constraining the particles
in the given thickness,
becomes equal to
the intrinsic momentum spread $\Delta p_x^{int}$
(dictated by the assigned thermodynamic conditions)
\begin{eqnarray}\label{spread}
\Delta p_x^{ind} =
\Delta p_x^{int}
\end{eqnarray}
(here $x$ labels the direction orthogonal to the slice).
For constituent particles
all with a same
quantum spatial uncertainty
which still we denote $\lambda$,
condition (\ref{spread})
will be reached
by definition
just when $l_{min} = \lambda$.
In the general case we define $\lambda$
as that thickness which gives (\ref{spread}).
For a Boltzmann gas the value of a so-defined $\lambda$
is close to thermal de Broglie wavelength.
When checked on actual systems,
condition (\ref{condition2}) is found
in general satisfied by far \cite{Pesci3}.
For the most entropic systems
it appears, instead,
practically attained.
For black body radiation
--and in general for ultrarelativistic constituent particles--,
for example,
an argument described in \cite{Pesci3}
suggests
\begin{eqnarray}\label{blackbody}
\lambda = 1/\pi T,
\end{eqnarray}
and thus
$l^* =
\frac{1}{\pi T} \ \left( 1 - \frac{\mu n}{\rho + p} \right) =
\lambda$,
being $\mu = 0$.
This prompts to consider
the uncertainty relations
as the mechanism which leads
to (\ref{condition2}), i.e.
to (\ref{condition1}).
From this perspective,
condition (\ref{condition2}) (and, thus, (\ref{condition1}))
appears to be a basic thermodynamic relation
arising from quantum mechanics alone.
This relation,
when combined with Einstein's lensing,
leads inexorably to the generalized
covariant bound to entropy,
that is to holography,
and vice versa is
the unavoidable consequence
of the latter in basic thermodynamics.
As known,
the generalized second law
follows from the generalized entropy bound,
assuming the validity of the ordinary second law \cite{FMW}.
From the above this means that,
besides ordinary second law,
just another ingredient from basic
thermodynamics is
needed, and is enough,
to give in a gravitational context
the generalized second law: relation (\ref{condition2}).
We see that this relation turns out to be
that component of basic thermodynamics
responsible for the `generalized part'
of the generalized second law,
i.e. that part which deals with processes
involving horizons.
If we imagine to start from the holographic principle
without any notion of quantum mechanics,
bound (\ref{gceb}) (which, consistently, should thus be understood as
the log
of the number of allowed different microscopic configurations
(instead of the log of the number of allowed orthogonal quantum states))
requires, we have seen,
the existence of a lower-limiting scale $l^*$.
The existence of this spatial limit
would suggest that a `size' should be assigned
to the elementary constituents of matter,
while the value of the spatial limit (expression (\ref{lstar}))
would imply that,
at least for ultrarelativistic systems,
this `size' of the constituents should be related to their momentum
by the uncertainty relations.
That is,
holography demands for
a microscopic description of matter,
discrete in itself anyhow
if finite values are to be assigned to the entropy
of generic systems,
which too is driven
by what we know as
the uncertainty relations,
i.e. it somehow demands for quantum mechanics.\footnote{Cf. \cite{Boussoqm};
see also \cite{Pesci4, Pesci5}.}
\section{A statistical origin for gravity}
At the end, what we have seen so far is that,
if the bundles of light rays actually shrink gravitationally,
the lower-limiting length $l^*$,
which manifests quantum mechanics,
is the effect of bound (\ref{gceb}).
But, what about the source of this shrinking ?
Looking at (\ref{gceb}),
we see that holography
demands that,
even in absence of interactions of any kind,
some mechanism must be at work
which shrinks the bundles of light rays
when going through matter.
The mere existence of some entropy
in a region requires, there, a focusing.
Bound (\ref{gceb}) moreover knows nothing else than
entropy and focusing, so that it is quite natural,
as far as we take (\ref{gceb}) as our primeval starting point,
to somehow suspect entropy as the source of focusing.
Massive bodies can be considered
for which the entropy, when temperature is near absolute zero,
can be negligibly small,
but the focusing sizeable.
So, if entropy is to be responsible of focusing, which entropy
should we take ?
It is clear that if,
for assigned $\rho$, $p$ and $\lambda$,
the rate of shrinking could be determined by a value of $s$ as high
as the limit $\frac{s}{\rho + p} = \pi \lambda$ in (\ref{condition2}),
any matter would comply with the bound (\ref{gceb}).
Thus,
the rate of shrinking could be set by the request that
bound (\ref{gceb}) be always satisfied,
and exactly attained by the most entropic systems.
The perspective we advocate here is that
in giving some piece of matter what we are really doing
is to allocate some maximum amount of information or entropy,
let us call it `intrinsic' information/entropy,
associated with it.
The bound then says that this `intrinsic'
information/entropy must focus light rays
at a precise rate.
We can derive the value of this rate
through consideration of material systems
which do attain the limit in (\ref{condition2}).
For the terminated lightsheet
of a plane layer of a photon gas
with thickness $l$ as small as the limit
$l = l_{min} = \lambda = 1/\pi T$,
entropy just attains the limit in (\ref{gceb}),
as well as in (\ref{condition1}) and (\ref{condition2}).
This means that at these conditions the shrinking
of the null congruence traversing the layer
is given by
\begin{eqnarray}\label{shrinking}
-\Delta A =
4 S =
4 s A \lambda =
4 \pi (\rho + p) A \lambda^2 =
4 \pi A \lambda^2 T_{ab} k^a k^b,
\end{eqnarray}
where $T_{ab}$ is the stress energy of the photon gas
and $k^a$ are the tangents to the null congruence
with respect to the parametrization given by $l$,
and use of (\ref{condition2})
has been made.
On the other hand, from geometry
the shrinking is connected,
through Raychaudhuri's equation,
to the Ricci tensor $R_{ab}$.
The Raychaudhuri equation, in our circumstances of
vanishing shear
and initially vanishing expansion $\theta = \frac{1}{A} \frac{dA}{dl}$,
for very small $l$ reads
\begin{eqnarray}\label{ray}
\frac{d\theta}{dl} =
-R_{ab} k^a k^b =
\text {const},
\end{eqnarray}
from which we get
$\theta = -l R_{ab} k^a k^b$
and thus the mentioned quadratic dependence
of $\Delta A$ on $l$ has the form
\begin{eqnarray}\label{geometry}
-\Delta A =
-\int_0^l \theta A dl^\prime =
A \frac{l^2}{2} R_{ab} k^a k^b.
\end{eqnarray}
From equation (\ref{shrinking}) this gives
\begin{eqnarray}\label{lensing}
R_{ab} k^a k^b = 8 \pi T_{ab} k^a k^b.
\end{eqnarray}
This is not a surprise.
The lensing turns out to be
just that given by Einstein's equation. As it must be,
since bound (\ref{gceb}) has
Einstein's lensing built-in.
The real point here is the perspective: the focusing
is determined by (`intrinsic') information/entropy
through holography (bound (\ref{gceb})).
That is to say,
starting from holography without any notion of gravity
and knowing only of information/entropy,
we end up with what we call gravity,
and this points
to a direction akin to
\cite{Padma1, Padma2}
(last Section of both) and
\cite{Verlinde},
to some extent.\footnote{We notice
that in \cite{LeeKimLee} another derivation of gravity is given,
independent, like the present attempt,
of \cite{Padma1, Padma2, Verlinde}.
In it, gravity is derived
from the Laundauer principle
in quantum information theory as applied to horizons.
In \cite{CaravelliModesto}, moreover,
the Einstein-Hilbert action,
or more general actions, are derived
starting from black hole entropy
as calculated within loop quantum gravity.}
In \cite{Jacobson},
a thermodynamic interpretation of Einstein's equation
(in which this reveals itself as an equation of state)
has already been given from an assumed
proportionality of horizon entropy and area.
The present attempt is supposed to provide a step forward
in that,
through consideration of information as fundamental,
the existence and strength of what we call gravity
is reduced to a principle of
maximum allowed amount of information
inside any closed surface.
That is, what really matters here
is not just the thermodynamic nature of Einstein's equation
(a point, this, of paramount importance indeed,
as for its implications on the opportunity of any attempt to quantize
this equation as well as for its accounting of the occurrence
of thermodynamic laws for classical black holes \cite{Jacobson}),
but that the occurrence itself of gravity
is understood as
what must happen in order that
a certain primeval property of entropy be preserved.
Looking at (\ref{shrinking}),
we see that
the role played by $\pi(\rho + p)$
in determining the shrinking
of the congruence by the photon gas,
can be viewed as played by $s/\lambda$.
This suggests that, given some matter,
it is the `intrinsic' entropy density
over wavelength,
namely the maximum entropy density over wavelength
at the given energy density+pressure,
what should be considered
as the proper source of focusing,
and what determines
the gravitational acceleration.
Considering light rays traversing orthogonally
a thin plane layer of matter,
using the focusing equation \cite{MTW},
which, since the shear is vanishing, reads
\begin{eqnarray}\label{focusing}
\frac{d^2 A^{1/2}}{dl^2} =
- \frac{1}{2} R_{ab} k^a k^b A^{1/2},
\end{eqnarray}
we have that, assuming rotational symmetry
around the propagation axis (as it is the case if local matter
is assumed to be the only source of the field),
the local-frame acceleration $a_t$ felt by the photons of the
congruence
while going through matter
being a distance $d$ apart can be expressed as
\begin{eqnarray}\label{tidal}
a_t =
- \frac{1}{2} R_{ab} k^a k^b d =
- 4 \frac{s}{\lambda} d,
\end{eqnarray}
denoting with $s/\lambda$ the `intrinsic' entropy density
per unit wavelength.
This same expression can be used to determine also
the local acceleration with respect to the origin,
taken halfway between the photons, if $d$ changes its meaning
becoming the
distance from the origin.
Expression (\ref{tidal}) fixes the gravitational
acceleration felt by photons in the local frame,
due to the presence
of local matter,
as determined by its `intrinsic' entropy.
The operational meaning is that
given a material medium
with some local values $\rho_m$ and $p_m$ of energy density
and pressure,
the local matter affects through holography the motion of photons
in the way expressed by (\ref{tidal})
(and this effect is what we call gravity),
where $s/\lambda$ is the entropy density
per unit wavelength
of a photon gas having energy density
$\rho = \frac{3}{4} (\rho_m + p_m)$
and $\lambda = 1/\pi T$, where $T$
is its black body temperature.
The acceleration in (\ref{tidal}) is thus expressed
in terms of entropy density per unit wavelength of
that black body radiation which gives
the `intrinsic' informational content of the local matter
we are considering.
\section{Concluding remarks}
In conclusion, what we have tried to show in the paper is that
the holographic principle,
assumed as primeval starting point,
implies
both a basic relation in flat-spacetime thermodynamics,
relation (\ref{condition2})
(argued to be more fundamental than the Bekenstein bound),
and the curvature effects we call gravity,
with a new entropy
--different from actual thermodynamic entropy--,
the `intrinsic' entropy (per $\lambda$) of a body, playing the role
of source of the curvature.
An expression of the gravitational acceleration in terms
of it has also been given (relation (\ref{tidal})).
We can summarize what we have seen as follows.
The Einstein's focusing we have obtained,
or the explicit expression
for the gravitational acceleration in equation (\ref{tidal}),
permits to view
what we call `gravitational effects'
as actually holography at work,
and the `gravitational' acceleration as a `holographic' acceleration.
Gravity is merely all what is needed
for the `holographic' property of entropy to be preserved.
In particular in equation (\ref{tidal})
we can read directly the strength
of a gravitational acceleration
per given amount of `intrinsic' information associated with matter.
Relation (\ref{condition2})
(with expression (\ref{lstar})),
in the form of uncertainty-like relations,
establishes the rule
for finding the `intrinsic' information allocated with the assigned matter
(so that the meaning of the uncertainty relations would be
in their being what provides the informational content of matter).
This information (per $\lambda$) is defined as
the maximum entropy per $\lambda$ we can associate to that matter,
i.e. the value which just attains (\ref{condition2});
and it turns out to be the entropy per $\lambda$ of
`equivalent' (in a definite sense) black body radiation.
This choice is dictated by the request
that
the strength of the holographic focusing
be just that needed for the entropy in a (terminated) lightsheet
to be universally bounded
(i.e. for every lightsheet geometry)
by the number $\Delta A/4$,
and just attained for the most challenging geometric choices.
Thus, holography,
by saying that the number of allowed
degrees of freedom
inside a given closed surface is bounded
(by a value proportional to the area of the boundary),
induces
curvature effects determined by the `intrinsic' degrees of freedom
carried by matter
(effects with a strength depending on the value of the bound),
and this constitutes
what we call gravity.
To speak of entropy per wavelength of black body radiation
means to speak of entropy per 1-bit-of-information thickness,
since the single bits of information are carried
by the single constituent photons
and we have $\simeq 1$ photon every $\lambda^3$ volume.
This suggests a deeper description of what
we have discussed.
Indeed, holography can be stated to imply that
the allowed number of elementary bits of information
in a layer of 1 bit thickness
at given sum of energy and pressure energy in the layer
is bounded.
To the extent that concepts like bit of information
and energy and pressure energy of a bit
can be regarded as primeval and, as such,
meaningful even in absence of space,
holography is pre-existing to space
(cf. \cite{Padma1, Padma2, Verlinde},
and \cite{MarSmolin}).
In this perspective,
when space is introduced as the information on
`where' information is,
the energy in the bit should spread
to keep unchanged
the elementary amount of information for the bit,
and this would be quantum mechanics.
When expressed in terms of this notion of
space, holography would then become the metric theory
which describes gravity.
I am grateful to Alessio Orlandi for fruitful discussions
on some of the arguments considered in the note.
|
\section{Introduction}
A minimal model for strongly correlated materials including the most complex systems such
as the high temperature cuprates is the Hubbard or t-J model. However, even within the simplified Hubbard model
away from the half-filling, there has been no consensus about the ground states, and various theoretical proposals
have been made for the phase diagram of the high temperature cuprates.
One of such states, the staggered flux phase (d-density wave), was discussed in
exitonic condensation\cite{halperinSSP68}, t-J model, and Hubbard model\cite{kotliarPRB88,marstonPRB89,
hsuPRB91,lee-review07},
and further suggested as a pseudogap phase of the cuprates\cite{chakravartyPRB02}.
The spin-triplet version of the staggered flux phase (spin-triplet d-density wave) was also discussed in the context of high temperature cuprates.\cite{doraPRB07,virosztekIJMP02,keeEPL09}.
Another proposed particle-hole condensate state of the angular momentum $l$=2 channel
with broken rotational symmetry is the electronic nematic state.\cite{kivelsonNATURE98,yamaseJPSJ00,metznerPRL00,oganesyanPRB01,keePRB03,khavkinePRB04,kivelsonRMP03,fradkin09104166,vojtaAP09}.
The spontaneous formation of nematicity has been invoked to explain the anisotropic transport
observed in a two-dimensional electron gas in a high
magnetic field\cite{lillyPRL99,duSSC99} and in Ru-based oxides\cite{borziSCIENCE07}.
Its relevance to high temperature cuprates was evidenced by a neutron scattering measurement
where anisotropic scattering patterns have been observed in YBa$_2$Cu$_3$O$_{6.5}$.\cite{hinkovSCIENCE08}
From a weak-coupling point of view, it is sometimes also called Pomeranchuk instability\cite{pomeranchuk}.
Given various proposed order parameters, the interplay between s- and d-wave order parameters has been of intensive theoretical study.
Examples for s- and d-wave order parameters are s- and d-wave superconductors, charge density wave,
spin density wave, d-density wave, spin-triplet d-density wave, and nematic states.
Among them, it was reported that nematicity plays an important role in the interplay between
s- and d-wave superconductors\cite{keeJPCM08}, and the spin density wave and spin-triplet d-density wave states
\cite{keeEPL09}. However, a full set of the relations between them is still lacking.
In this paper, we offer a complete theory on how
s- and d-wave orders transform via the nematic order, and relations between them.
We found that the order parameters listed above can be organized in two independent six-dimensional vectors.
One vector is composed
of d-wave superconducting, d-density wave, and spin density wave order parameters, while the other vector
contains s-wave superconducting, charge density wave, and spin-triplet d-density wave order parameters.
Each vector transforms under the action of SO(6).
Charge, spin, $\eta$-pairing
\cite{yangPRL89}, $\pi$-pairing\cite{zhangSCIENCE97}, and spin-triplet nematic\cite{wuPRB07} operators
together satisfy the SO(6) Lie algebra.
The nematic order parameter commutes with all generators, and hence is not a part of the SO(6) group.
However, it transforms the two independent vectors connecting s- and d-wave order parameters.
Our findings imply that nematic order does not interfere the competition between the order parameters within
the same vector, but strongly affects the interplay between two order parameters belonging to different vectors. For example, it allows a linear coupling between s- and d-wave order parameters
in the Ginzburg-Landau (GL) free energy, which modifies the physical properties of both phases.
We found that the conditions for non-zero linear coupling differ for particle-particle and particle-hole condensate
states.
Below we will review an SO(6) group theory and
present the relations between the nematic order, the generators, and superspins.
We will show the role of nematic order for particle-particle condensate states,
and elaborate a similar process for particle-hole cases.
We will also discuss the implications of such relations
in the context of GL free energy, and superconducting transition temperature
related to high temperature cuprates.
\section{ nematic order parameter and SO(6) group}
It was first pointed out by Yang\cite{yangPRL89} that the $\eta$-pairing state is an eigenstate of the Hubbard model.
The $\eta$-pairing operator is defined as
$\hat{\eta}^+ =-i\sum_{{\bf k}\sigma\sigma'} \cy{{\bf k}}{\sigma}\sigma^y_{\sigma\sigma'}\dy{-{\bf k}}{\sigma'}$ and $\hat{\eta}^-=(\hat{\eta}^+)^\dagger$, where ${\bf Q} =(\pi,\pi)$ and $\sigma^y$ is a Pauli matrix.
The $\eta$ operator carries charge $2e$ and spin 0,
and commutes with the Hubbard Hamiltonian
at half filling $\mu = U/2$, where $\mu$ is the chemical potential and $U$ is the on-site Hubbard interaction.
It is also an eigenstate of the momentum operator with the eigenvalue ${\bf Q}$.
Later it was found that the $\eta$-pairing operators combined with the charge operator satisfy an SU(2) algebra (named pseudospin)\cite{zhangPRL90},
and further recognized that the Hubbard model has two sets of commuting SU(2) symmetries. One set is characterized by
the pseudospin of $\eta$-pairing and charge operators, and the other is conventional spin operator.\cite{yang-zhang}
The three-dimensional vector transforming under the action of pseudospin SU(2) forms a superspin,
where its three components are
s-wave superconductor, and charge density wave with the ordering wave vector ${\bf Q}$.
It was also reported that the pseudospin SU(2) rotates another superspin
composed of d-wave superconducting and d-density wave order parameters.\cite{nayakPRB00}
On the other hand,
the vector transforming under the spin SU(2) is the spin density wave with the ordering
wave vector ${\bf Q}$. Under a particle-hole transformation for one spin species, $c^\dagger_{i \downarrow} \rightarrow (-1)^i c_{i \downarrow}$, the role of the
two sets of SU(2) generators is interchanged. The same particle-hole transformation maps the positive Hubbard model
to the negative Hubbard model, and it also maps the $\eta$-pairing to the Nagaoka ferromagnetic state.\cite{singhPRL91}
A further generalization of the concept of the exact SO(4) symmetry of the Hubbard model
to a unified theory of antiferromagnetism and d-wave superconductivity based on
SO(5) symmetry was later proposed to understand the physics of the high temperature cuprates.\cite{zhangSCIENCE97,demlerRMP04}
Here we present a full list of relations between s- and d-wave order parameters including those mentioned above.
The ground state order parameters transform as a six-dimensional vector under the action of SO(6).
They can be organized into a vector ${\hat n}_a$ ($a = 1...6$) which should satisfy
\begin{equation}
[{\hat L}_{ab}, {\hat n}_c] = -i \left( \delta_{bc} {\hat n}_a -\delta_{ac} {\hat n}_b \right).
\end{equation}
where ${\hat L}_{ab}$ are generators of SO(6) based on the following
operators:
\begin{eqnarray}
\label{eq:generators}
\hat{Q} &=&-\frac{1}{2}\sum_{{\bf k}\sigma}\left( \cy{{\bf k}}{\sigma}\cn{{\bf k}}{\sigma} +\dy{{\bf k}}{\sigma}\dn{{\bf k}}{\sigma}-1\right), \cr
\hat{S}_\alpha &=&\frac{1}{2}\sum_{{\bf k}\sigma\sigma'}\left( \cy{{\bf k}}{\sigma}\sigma^\alpha_{\sigma\sigma'}\cn{{\bf k}}{\sigma'} +\dy{{\bf k}}{\sigma}\sigma^\alpha_{\sigma\sigma'}\dn{{\bf k}}{\sigma'}\right), \cr
\hat{R}_\alpha &=&\frac{1}{2}\sum_{{\bf k}\sigma\sigma'} d({\bf k}) \left( \cy{{\bf k}}{\sigma}\sigma^\alpha_{\sigma\sigma'}\cn{{\bf k}}{\sigma'} -\dy{{\bf k}}{\sigma}\sigma^\alpha_{\sigma\sigma'}\dn{{\bf k}}{\sigma'}\right), \cr
\hat{\Pi}^+{_\alpha} &=&\sum_{{\bf k}\sigma\sigma'} d({\bf k}) \cy{{\bf k}}{\sigma}\left(\sigma^\alpha \sigma^y\right)_{\sigma\sigma'}\dy{-{\bf k}}{\sigma'},\ \ \ \hat{\Pi}^-=(\hat{\Pi}^+)^\dagger, \cr
\hat{\eta}^+ &=&-i\sum_{{\bf k}\sigma\sigma'} \cy{{\bf k}}{\sigma}\sigma^y_{\sigma\sigma'}\dy{-{\bf k}}{\sigma'},\ \ \ \ \ \ \hat{\eta}^-=(\hat{\eta}^+)^\dagger,
\label{generators}
\end{eqnarray}
where ${\bf k}$ runs over the reduced Brillouin zone, $d({\bf k}) = \cos{k_x} -\cos{k_y}$,
$\alpha$ takes the values $x$, $y$, $z$, and
${\sigma}^{\alpha}$ are the Pauli matrices.
The generators of SO(6) can be represented by an antisymmetric 6x6 matrix,
$\hat{L}_{ab}=-\hat{L}_{ba}$.
\begin{eqnarray}
\hat{L}_{ab}=\left(\begin{array}{c c c c c c}
0 & \hat{Q} & \Re\,\hat{\Pi}_x & \Re\,\hat{\Pi}_y & \Re\,\hat{\Pi}_z & \Re\,\hat{\eta} \\
\, & 0 & \Im\,\hat{\Pi}_x & \Im\,\hat{\Pi}_y & \Im\,\hat{\Pi}_z & \Im\,\hat{\eta} \\
\, & \, & 0 & \hat{S}_z & -\hat{S}_y & \hat{R}_x \\
\, & \, & \, & 0 & \hat{S}_x & \hat{R}_y \\
\, & \, & \, & \, & 0 & \hat{R}_z \\
\, & \, & \, & \, & \, & 0
\end{array}
\right),
\label{eq:genMatrix}
\end{eqnarray}
where $\Re\,\hat{\cal O}\equiv\frac{1}{2}(\hat{\cal O}^-+\hat{\cal O}^+)$ and $\Im\,\hat{\cal O}\equiv\frac{1}{2i}(\hat{\cal O}^--\hat{\cal O}^+)$.
It satisfies the correct SO(6) Lie algebra,\cite{footnote}
\begin{eqnarray}
\label{eq:LLComm}
\comm{\hat{L}_{ab}}{\hat{L}_{cd}}=-i\left(\delta_{ad}\hat{L}_{bc}+\delta_{bc}\hat{L}_{ad}-\delta_{bd}\hat{L}_{ac}-\delta_{ac}\hat{L}_{bd}\right).\nonumber
\label{so-6-group}
\end{eqnarray}
Here $L_{12}$ is the charge operator, and $L_{34}$,$L_{35}$, and $L_{45}$ are the three components of the spin operator.
$L_{16}$ and $L_{26}$ represent real and imaginary part of the $\eta$-pairing.\cite{yangPRL89}
$L_{13}$, $L_{14}$, and $L_{15}$ ($L_{23}$, $L_{24}$ and $L_{25}$) denote $x$, $y$ and $z$ component of the real (imaginary) part of
the $\pi$-pairing which carries charge $2e$ and spin $1$ and represents a broken translational symmetry.\cite{zhangSCIENCE97}
$L_{36}$, $L_{46}$, and $L_{56}$ correspond to the spin-triplet nematic order parameter carrying spin 1
and representing a broken x-y symmetry on a square lattice.\cite{wuPRB07}
There exist two independent vectors. Each vector acts as a superspin,
and transforms under the SO(6). This observation was first reported in Ref. \cite{markiewicz97},
and a similar SO(6) symmetry was found in Fe-pnictide superconductors\cite{podolskyEPL09}.
One superspin (superspin-1) consists of spin-density wave ($\Delta_{sdw}$), d-density wave ($\Delta_{ddw}$), and d-wave superconducting ($\Delta_{dsc}$)
order parameters:
\begin{eqnarray}
\hat{\Delta}_{sdw}^\alpha&=&\frac{1}{2}\sum_{{\bf k}\sigma\sigma'}\left(\cy{{\bf k}}{\sigma}\sigma^\alpha_{\sigma\sigma'}\dn{{\bf k}}{\sigma'}+\dy{{\bf k}}{\sigma}\sigma^\alpha_{\sigma\sigma'}\cn{{\bf k}}{\sigma}\right), \nonumber\\
\hat{\Delta}_{dsc}^+ &= &\sum_{\bf k} d({\bf k}) \left(\cy{{\bf k}}{\uparrow}\cy{-{\bf k}}{\downarrow}-\dy{{\bf k}}{\uparrow}\dy{-{\bf k}}{\downarrow}\right),\nonumber\\\hat{\Delta}_{dsc}^-&=& (\hat{\Delta}_{dsc}^+)^\dagger,\nonumber \\
\hat{\Delta}_{ddw}&=&-\frac{i}{2}\sum_{{\bf k}\sigma} d({\bf k}) \left(\cy{{\bf k}}{\sigma}\dn{{\bf k}}{\sigma}-\dy{{\bf k}}{\sigma}\cn{{\bf k}}{\sigma}\right),
\end{eqnarray}
where ${\hat n}_1 = \Re\, \hat{\Delta}_{dsc}$, ${\hat n}_2 = \Im\, \hat{\Delta}_{dsc}$, $ {\hat n}_3 = \hat{\Delta}_{sdw}^x$, $ {\hat n}_4
=\hat{\Delta}_{sdw}^y$, ${\hat n}_5 = \hat{\Delta}_{sdw}^z$, and ${\hat n}_6 = \hat{\Delta}_{ddw}$.
The other superspin (superspin-2) rotated by the same 15 generators ${\hat L}_{ab}$
is composed of spin-triplet d-density wave ($\Delta_{tsf}$), charge density wave ($\Delta_{cdw}$), and s-wave superconducting ($\Delta_{ssc}$) order
parameters:
\begin{eqnarray}
\hat{\Delta}_{tsf}^\alpha&=& \frac{i}{2} \sum_{{\bf k}\sigma\sigma'} d({\bf k}) \left(\cy{{\bf k}}{\sigma}\sigma^\alpha_{\sigma\sigma'}\dn{{\bf k}}{\sigma'}-\dy{{\bf k}}{\sigma}\sigma^\alpha_{\sigma\sigma'}\cn{{\bf k}}{\sigma}\right), \nonumber\\
\hat{\Delta}_{ssc}^+ &= &\sum_{\bf k} \left(\cy{{\bf k}}{\uparrow}\cy{-{\bf k}}{\downarrow}+\dy{{\bf k}}{\uparrow}\dy{-{\bf k}}{\downarrow}\right), \nonumber\\
\hat{\Delta}_{ssc}^-&=&(\hat{\Delta}_{ssc}^+)^\dagger,\nonumber \\
\hat{\Delta}_{cdw}&=&-\frac{1}{2}\sum_{{\bf k}\sigma} \left(\cy{{\bf k}}{\sigma}\dn{{\bf k}}{\sigma}+\dy{{\bf k}}{\sigma}\cn{{\bf k}}{\sigma}\right),
\end{eqnarray}
where the superspin is arranged as
${\hat n}_1 = \Re\, \hat{\Delta}_{ssc}$, ${\hat n}_2 = \Im\, \hat{\Delta}_{ssc}$, $ {\hat n}_3 = \hat{\Delta}_{tsf}^x$, $ {\hat n}_4
=\hat{\Delta}_{tsf}^y$, ${\hat n}_5 = \hat{\Delta}_{tsf}^z$, and ${\hat n}_6 = \hat{\Delta}_{cdw}$.
What is the role of the nematic order parameter within the SO(6) group?
The nematic order parameter is given by
\begin{equation}
{\hat N} = \sum_{{\bf k}\sigma} d({\bf k}) \left( \cy{{\bf k}}{\sigma} \cn{{\bf k}}{\sigma} - \dy{{\bf k}}{\sigma} \dn{{\bf k}}{\sigma} \right).
\end{equation}
When $\langle {\hat N} \rangle \equiv {\cal N}_0 \neq 0$, the phase is characterized by a broken x-y symmetry of the square lattice, and it is trivial to generalize to a broken point group symmetry in other lattices.
Note that the nematic order parameter commutes with all 15 generators:
\begin{equation}
[\hat{N}, {\hat L}_{ab}]=0.
\label{nematic-generator}
\end{equation}
This means that nematicity is not an SO(6) symmetry breaking field, and does not
interfere with the competition between the order parameters within the superspin.
For example, the phase diagram between antiferromagnetism and d-wave superconductivity
(both belong to superspin-1)
studied in the t-J model based on SO(5) symmetry\cite{zhangSCIENCE97} (a subset of the SO(6) in this study)
is not modified by the presence of nematicity.
However, the nematic operator does not commute with the following conventional quantum rotor model.
\begin{equation}
H_{QR} = \frac{1}{2\chi} \sum_{i, a< b} {\hat L}_{i,ab}^2 + \sum_{<i j>,a} r_a {\hat n}_i^a {\hat n}_j^a,
\end{equation}
where the first term is the kinetic term and $\chi$ is the moment of intertia, and the second term
is the potential term. The Hamiltonian has SO(6) symmetry when $r_a$ is idential to all $a$.
Note that the nematic operator commutes with the first term, but not the second term.
On the other hand, the competition between the order parameters in the same superspin ${\hat n}^a$
is determined by difference in $r_a$.
What are relations between nematic order and other order parameters?
The nematic operator transforms the components of the two independent superspins as follows:
\begin{eqnarray}
\left[ \hat{\Delta}_{dsc}^+ , \hat{\Delta}_{ssc}^- \right] &=& 2 \hat{N},
\;\;\;\;\; \left[ \hat{\Delta}_{dsc}^-, \hat{\Delta}_{ssc}^+ \right] = 2 \hat{N},\nonumber
\\
\left[ \hat{\Delta}_{ddw},\hat{\Delta}_{cdw} \right] &=& i \hat{N},
\;\;\;\;\;
\left[ \hat{\Delta}_{tsf}^z,\hat{\Delta}_{sdw}^z \right] = i \hat{N}, \nonumber
\\
\left[ \hat{\Delta}_{tsf}^+,\hat{\Delta}_{sdw}^- \right] &=& \frac{i}{2} \hat{N},
\;\;\;\;\;\;
\left[ \hat{\Delta}_{tsf}^-,\hat{\Delta}^+_{sdw} \right] = \frac{i}{2} \hat{N}.
\label{nematic-orders}
\end{eqnarray}
The nematic operator transforms s- to d-wave order parameters which belong to
different superspins.
The above results are summarized in
the table below.
\begin{table}[htbp]
\centering
\begin{tabular}{@{} cccc @{}}
\hline
SO(6) generators & \;\; ${\hat Q}$, ${\hat S}$, ${\hat \eta}$-, ${\hat \pi}$-pairing, spin nematic operators \\
\hline
nematic operator & commutes with generators \&\\
& transforms superspin-1 and -2 \\
\hline
superspin-1 & dSC, dDW, SDW \\
superspin-2 & sSC, CDW, spin-triplet dDW\\
\hline
\end{tabular}
\caption{ A summary of the SO(6) group and the relations to nematic order.}
\label{tab:label}
\end{table}
In the following section, we discuss the physical implications of the commutation relations using a GL free energy theory
assuming that nematic order is present.\cite{note}
\section{Ginzburg Landau theory}
The commutation relations in Eq. \ref{nematic-orders},
$[A,B]= {\hat N}$, indicate that if $\langle {\hat N} \rangle \equiv {\cal N}_0$
is finite, a linear coupling between A and B phases, such as $\gamma \; \Phi \; \Psi$ with $\Phi =\langle A \rangle$ and
$\Psi = \langle B \rangle $, may be present in the GL free energy.
The GL free energy is then given by
\begin{equation}
{\cal F} = \frac{a}{2} \Psi^2 + \frac{b}{2} \Phi^2 + \gamma \; \Psi \; \Phi + u \Psi^4 + v \Phi^4 + ....
\label{gl-free-energy}
\end{equation}
Assuming that $a > 0$, $b > 0$, and $a b > \gamma^2$ (none of the phases represented by $\Phi$ and $\Psi$
is ordered), the solutions of the two coupled equations for $\Phi$ and $\Psi$ leads to
the following dispersion of modes:\cite{footnote2}
\begin{equation}
\chi \omega^2 ({\bf k})= \frac{\epsilon_1({\bf k})+\epsilon_2({\bf k})}{2} \pm \frac{1}{2} \sqrt{
(\epsilon_1({\bf k})-\epsilon_2({\bf k}))^2 + 4 \gamma^2},
\end{equation}
where
\begin{eqnarray}
\epsilon_1({\bf k}) &= & a + \rho \{(1+ {\cal N}_0) k_x^2+(1-{\cal N}_0) k_y^2\},
\nonumber\\
\epsilon_2({\bf k}) &= & b + \rho \{(1+{\cal N}_0) k_x^2+(1-{\cal N}_0) k_y^2\}.
\end{eqnarray}
Here we have used the effective Lagrangian
$L_{eff} = \frac{\chi}{2} (\partial_t \Phi)^2
-\frac{\rho}{2} \{(1+{\cal N}_0) (\partial_a \Phi)^2 + (1-{\cal N}_0) (\partial_y \Phi)^2 \} - \frac{a}{2} \Phi^2
+\frac{\chi}{2} (\partial_t \Psi)^2
-\frac{\rho}{2} \{(1+{\cal N}_0) (\partial_a \Psi)^2 + (1-{\cal N}_0) (\partial_y \Psi)^2 \} - \frac{b}{2} \Psi^2
-\gamma \Psi \Phi$, where $\rho$ is the stiffness.
Note that the excitations are anisotropic due to nematicity\cite{kaoPRB05}, and $k_x$ and $k_y$ are deviations from
an ordering wave-vector which is either 0 or ${\bf Q}$ depending on the nature of $\Psi$ (or $\Phi$).
One of the solutions becomes 0 when $\gamma= \sqrt{a b}$, leading to an ordered state.
The condensed state is a linear combination of $\Psi$ and $\Phi$, and the dominant
contribution depends on $a$ and $b$. Also, if one of them, say $\Psi$, is finite (when $a < 0$), the other, $\Phi$, is always induced as long as $\gamma$ is finite.
Is $\gamma$ always finite if nematic order exists?
For example, consider a system in the nematic state with SO(6) symmetry at high temperatures.
At low energy, the system spontaneously breaks the SO(6) symmetry, and one of the phases represented by
$\Psi$ is stabilized. If $\Psi$ represents the d-wave superconducting state, $\Phi$ is the s-wave component.
Similarly if $\Psi$ is the spin density wave, $\Phi$ should be the spin triplet d-density wave.\cite{keeEPL09}
Does nematicity always lead to an induced order parameter of $\Phi$
without any extra condition?
Below we show that it requires another condition for a non-zero linear coupling (in addition to the nematic order), and
that the condition for a finite $\gamma$ differs for particle-particle and
particle-hole condensates.
\section{ Difference between particle-particle and particle-hole pairs}
Let us compute $\gamma$ for particle-particle condensate states.
To check the condition for a non-zero linear coupling coefficient $\gamma$ between d-wave and s-wave superconducting cases
($\Psi =Re\langle \Delta_{dsc} \rangle $ and $\Phi =Re \langle \Delta_{ssc} \rangle$),
we introduce $ \psi^\dagger_{{\bf k}} =
(c^{\dagger}_{{\bf k} \sigma}, c_{-{\bf k} -\sigma})$. Then the order parameter is written as
$\Delta_{ssc} = \sum_{\bf k} \psi^{\dagger}_{\bf k} \tau_1 \psi_{\bf k}$.
Inside the nematic state, the quasiparticle Green's function is written as
\begin{equation}
G^{-1}({\bf k}, i\omega_n)= -i \omega_n + \epsilon_{\bf k} -\mu,
\end{equation}
where
\[
\epsilon_{\bf k}= -2 t (\cos{k_x}+\cos{k_y}) + 2t d({\bf k}) {\cal N}_0 -4 t^\prime \cos{k_x}\cos{k_y},
\]
and $\mu$ is the chemical potential.
$t$ and $t^{\prime}$ represent the nearest neighbor and second nearest neighbor hoppings, respectively.
Assuming that d- and s-wave superconducting fluctuations couple to fermions
with interactions of $g_1$ and $g_2$,
the $\gamma$ coefficient becomes
\begin{eqnarray}
\gamma_{dsc-ssc} & = & g_1 g_2 T \sum_{{\bf k}} d({\bf k})\sum_{i\omega_n} Tr \left( G({\bf k},i\omega_n) \tau_1 G({\bf k}, i\omega_n) \tau_1 \right)
\nonumber\\
& = & g_1 g_2 \sum_{{\bf k}} d({\bf k}) \frac{ n_F(\xi_k) -n_F(-\xi_k)}{2 \xi_k},
\end{eqnarray}
where $\xi_{\bf k} = \epsilon_{\bf k} -\mu$.
$\gamma$ is always finite as long as $\mu$ and/or $t^\prime$ is finite.
In other words, when the particle-hole symmetry is broken and nematic order is present,
the linear coupling term induces d- or s-wave superconducting order as we discussed
in Eq. \ref{gl-free-energy}.
However, the above result is not true for particle-hole pairs. $\gamma$ then is zero
independent of particle-hole symmetry.
To examine the condition for particle-hole cases, let us introduce $ \psi^\dagger_{{\bf k} \sigma} =
(c^{\dagger}_{{\bf k} \sigma}, c^{\dagger}_{{\bf k}+{\bf Q} \sigma})$.
In this basis, the Green's function becomes
\begin{equation}
G^{-1}({\bf k}, i\omega_n)= -i \omega_n I + \tilde{\epsilon_{\bf k}} \tau_3 -\mu_{\bf k} I ,
\end{equation}
where $\tilde{\epsilon}_{\bf k} = -2 t (\cos{k_x}+\cos{k_y}) + 2t d({\bf k}) {\cal N}_0
= -\tilde{\epsilon}_{{\bf k}+{\bf Q}}$
and $\mu_{\bf k} = 4 t^\prime \cos{k_x}\cos{k_y} + \mu = \mu_{{\bf k}+{\bf Q}}$.
The $\gamma$ coefficient for example between the charge density wave and the d-density wave order is then
obtained as
\begin{eqnarray}
\gamma_{cdw-ddw} &\propto & T \sum_{{\bf k}} d({\bf k}) \sum_{i\omega_n} Tr \left( G({\bf k},i\omega_n) \tau_1 G({\bf k}, i\omega_n) \tau_2 \right)
\nonumber\\
& = &0.
\end{eqnarray}
This is similar to the coupling between $Re \Delta_{dsc}$ and $Im \Delta_{ssc}$. This linear coupling
is not allowed in the free energy due to symmetry.
Let us consider the coupling between different directions of spin density wave and spin-triplet d-density wave.
For example, the coupling between antiferromagnetic fluctuations along the $x$-direction
and spin-triplet d-density wave fluctuations along the $y$-direction are given by
$\delta \Delta_{sdw}^{x} \propto \psi_{\bf k}^\dagger \tau_1 \psi_{\bf k}$ and
$\delta \Delta_{tsf}^{y} \propto \psi_{\bf k}^\dagger \tau_1 \psi_{\bf k}$
where $\psi^{\dagger}_{\bf k} = \left( c^{\dagger}_{\bf k, \uparrow}, c^{\dagger}_{{\bf k}+{\bf Q}, \downarrow}
\right)$.
Then the coefficient $\gamma$ is obtained as
\begin{eqnarray}
\gamma_{tsf-sdw} & \propto & \sum_{{\bf k} i\omega_n} d({\bf k}) {\rm Tr}
\left( \tau_1 G({\bf k} i\omega_n) \tau_1 G({\bf k} i\omega_n) \right)
\nonumber\\
&=& \sum_{\bf k} \frac{d({\bf k})}{2 \tilde{\epsilon}_{\bf k}} \left(
n_F(\tilde{\epsilon}_{\bf k} -\mu_{\bf k})- n_F(-\tilde{\epsilon}_{\bf k} -\mu_{\bf k}) \right)
\nonumber\\
&=& 0.
\end{eqnarray}
Note that the coupling is also 0, because both $d({\bf k})$ and $\tilde{\epsilon}_{\bf k}$
change sign under ${\bf k} \rightarrow {\bf k}+{\bf Q}$, while $\mu_{\bf k}$ does not.
The physical reason is that the spin density wave state breaks time reversal symmetry, while
the triplet staggered flux does not. This fact is reflected in the commutation relations,
where Eq. \ref{nematic-orders} has the imaginary factor $i$.
Therefore, a linear coupling is not allowed between s- and d-wave particle-hole condensate states.
However, in the presence of a magnetic field $h$, the result alters.
Note that the particle-particle and particle-hole order parameters are related
by the particle-hole transformation, which also maps the chemical potential to the magnetic field
to be discussed in detail below.
In the presence of an external magnetic field, $\gamma$ changes to
\begin{eqnarray}
& & \gamma_{tsf-sdw} (h \neq 0) \\
& \propto & \sum_{\bf k} \frac{d({\bf k})}{2 (\tilde{\epsilon}_{\bf k} +h) } \left(
n_F(\tilde{\epsilon}_{\bf k} +h -\mu)- n_F(-\tilde{\epsilon}_{\bf k}-h -\mu) \right).
\nonumber
\end{eqnarray}
$\gamma$ between the spin-triplet d-density wave and spin density wave is finite when ${\cal N}_0$ and $h$ are finite.
The leading contribution of $h$ and ${\cal N}_0$ to $\gamma({\cal N}_0,h)$ can be written as
$\gamma({\cal N}_0,h) = \gamma_0 {\cal N}_0 h$, where $\gamma_0$ depends on the interactions
between fermions and the fluctuations of the order parameters. \cite{keeEPL09}
To understand the difference between particle-particle and particle-hole condensates, let us consider the particle-hole
transformation. The particle-hole transformation
mapping the positive to the negative Hubbard model discussed above
maps each component of the superspins as follows:
\begin{eqnarray}
\hat{\Delta}_{sdw}^{\alpha} &\rightarrow& \left(\hat{\Delta}^{\pm}_{ssc}, \hat{\Delta}_{cdw}\right),
\nonumber\\
\hat{\Delta}_{tsf}^{\alpha} &\rightarrow& \left( \hat{\Delta}^{\pm}_{dsc}, \hat{\Delta}_{ddw} \right),
\end{eqnarray}
where $\alpha= x,y,z$.
In addition to the known result that the antiferromagnetic order
transforms to the s-wave superconducting and charge density wave orders,
we found that the spin-triplet d-density wave phase transforms to the
d-wave superconducting and d-density wave orders,
while nematic order is invariant.
Since the chemical potential maps to the magnetic field under the particle-hole transformation,
the conditions for a finite $\gamma$ between particle-particle and particle-hole condensates
are also related by the particle-hole transformation -- $\gamma_0 {\cal N} \mu
\langle \Re \Delta_{dsc} \rangle \langle \Re \Delta_{ssc} \rangle$ maps to
$\gamma_0 {\cal N} h \langle \Delta_{tsf}^y \rangle \langle \Delta_{sdw}^x \rangle$
under the particle-hole transformation.
Therefore, the linear coupling between
d- and s-wave superconducting order parameters requires a finite chemical potential, while
the coupling between spin-triplet d-density wave and spin density wave requires a magnetic
field. Note that both breaks SO(6) symmetry, as the chemical potential and magnetic field appear
as $\mu L_{12}$ and $h L_{34}$ in Hamiltonian, respectively.
\section{effect on superconducting transition temperature}
Let us reexamine the GL free energy, Eq. \ref{gl-free-energy}, to see if the superconducting transition temperature
is modified by the coupling between the d- and s-wave superconducting order parameters.
We consider $\Psi = \langle \Re \Delta_{dsc} \rangle $ and $\Phi = \langle \Re \Delta_{ssc}\rangle$, and $\gamma$ is finite and proportional to the nematic strength ${\cal N}_0$, and particle-hole symmetry is assumed to be broken.
Since the chemical potential couples to the charge operator $L_{12}$, it favors the d-wave superconducting
state over the antiferromangetic and d-density wave states.
The competition between antiferromagnetism, d-wave superconductor,
and d-density wave is determined by SO(6) symmetry breaking terms, where
nematicity does not affect the interplay between them.
Assuming that the superconducting state is stabilized in a finite window of phase space,
and the transition temperature is set by $T_c^0$ ( $a < 0$ below $T_c^0$ and assume $b > 0$),
we are interested in the effect of nematicity on the superconducting transition temperature.
It is straightforward to check that the effective mass term
$a_{eff}$ is modified by $a - \frac{\gamma^2}{4b}$ after integrating out the $\Phi$ field.
Note that $a \propto (T-T_{c}^0)$ and $a_{eff} \propto (T-T_c)$ where $T_c$ is the transition
temperature modified by the coupling $\gamma$.
Since the effective mass gets smaller due to the coupling to the s-wave component, the transition
temperature $T_c$ is higher than $T_c^0$.
However, one should note that the current description is based on a classical theory, and quantum fluctuations
beyond the present study should be taken into account to see if the result may qualitatively change.
\section{Discussion and Summary}
We have studied the role of the nematic order parameter in the interplay between s- and d-wave particle-particle
or particle-hole condensate states. These condensate states include d- and s-wave superconductors,
d-density wave, spin-triplet d-density wave, spin density wave, and charge density wave phases.
We found that the nematic operator transforms d- to s-wave superconductors,
spin-triplet d-density wave to (s-wave) spin-density wave, and d-density wave to (s-wave) charge-density wave
operators. This can be summarized as a transformation between two different six-dimensional vectors.
One vector is composed of d-wave superconductor, d-density wave, and spin-density wave
order parameters, while the other vector consists of s-wave superconductor, charge-density wave, and spin-triplet
d-density wave order parameters.
Each vector acts as a superspin and transforms under the action of SO(6). There exist 15 generators,
which correspond to charge, spin, spin-triplet nematic, $\eta$- and $\pi$-pairing operators, which form the SO(6)
group.
The transformation between the two superspins via nematicity implies that a linear coupling between two order parameters that belong to two different vectors
can be present in the GL free energy. Such a linear coupling allows induced ordering when one of them is condensed.
However, we found that there is an additional condition for a non-zero linear coupling, which
differs for particle-particle and particle-hole condensates. For example,
when d-wave superconductor (particle-particle condensate) and nematic order coexist,
s-wave superconducting order is induced, only when the particle-hole symmetry is broken. On the other hand, when spin-density wave
(particle-hole condensate)
and nematic order coexist, a similar transformation allows an induced spin-triplet d-density wave,
only when time-reversal symmetry is broken. These results are consistent with symmetry considerations. Since
the spin-triplet d-density wave does not break time reversal symmetry, while spin-density
wave does, a linear coupling between the two order parameters is allowed
when time reversal symmetry is broken by an external magnetic field.
It is also interesting to notice that the nematic operator commutes with the generators.
When the Hamiltonian contains a term $- g \sum_{ij} {\hat N}_i {\hat N}_j$ which favors
nematic ordering, it does not act as an SO(6) symmetry breaking field.
It means that the nematic order can exist without interfering the competition among the six different order parameters
within a superspin. It is merely a spectator.
However, it affects the interplay between order parameters which belong to two different superspins.
Nematicity allows a linear coupling between the two order parameters, and
affects the physical properties
of both phases. As an example, we showed that the d-wave superconducting transition temperature is modified by
the coupling
to the s-wave superconducting order parameters which happens when nematicity is present and particle-hole
symmetry is broken.
The nematic order parameter has been widely discussed in the context of strongly correlated materials.
In particular, the phase diagram of the high temperature cuprates is complex
and its complete understanding requires further experimental and theoretical investigation. Our results indicate
that the proposed nematic phase affects phenomena in the superconducting phase such as an
anisotropy in the spin susceptibility and an increase in superconducting transition temperature. It also affects
antiferromagnetism via the coupling to the spin-triplet d-density wave when a magnetic field is applied.
We do not attempt to find a microscopic Hamiltonain with SO(6) symmetry which is beyond the scope of the current study.
However, we emphasize that
Eq. \ref{nematic-generator} and \ref{nematic-orders} are exact independent of symmetry of Hamiltonian,
and SO(6) symmetry is useful to identify the compact relations between the nematic
and other order parameters suggested in the context of high temperature cuprates. The GL free energy analysis
hints the importance of nematicity for the phase diagram of antiferromagnetism and d-wave superconducting phase.
{\it Acknowledgement} : I thank D. Podolsky, M. Norman, S. Kivelson, and especially E. Fradkin
for fruitful discussions. This work has been supported by
NSERC of Canada, Canadian Institute for Advanced Research, and Canada Research Chair.
|
\section{Introduction}
The analysis of the dependence of the angular size of some sources
with redshift was for many decades one of the most important
geometric tests of cosmological models. Different cosmologies
predict different dependences for a given linear size and this
can be compared with the data from observations. The test, first
conceived by Hoyle\cite{Hoy59},
is simple in principle but its application is not so simple
because of the difficulty in finding a standard rod, a type of object
with no evolution in linear size over the lifetime of the Universe.
It is well known that the application of the
angular size ($\theta $) vs. redshift ($z$) test
gives a rough dependence of $\theta \propto z^{-1}$ for QSOs and
radio galaxies at radio wavelengths\cite{Mil71,Kel72,War74,Kap77,Kap87,Uba93},
for first ranked cluster galaxies in the
optical\cite{San72,Djo81,Pas96,Sch97}, and for the
separation of brightest galaxies in clusters \cite{LaV86}
or in QSO-galaxy pairs of the same redshift\cite{Sap99}.
The deficit of large objects at high redshifts with respect to
the predictions of an expanding Universe is believed to
be an evolutionary effect by which galaxies were smaller in the
past (e.g., \cite{Mil71}), or a selection effect (e.g., \cite{Jac73}). Thus,
the $\theta \propto z^{-1}$ relationship, as predicted
by a static Euclidean Universe, would be just
a fortuitous coincidence of the superposition of the $\theta (z)$
dependence in the expanding Universe and evolutionary/selection effects.
Some other studies have tried to find better standard rods.
Ultra-compact radio sources\cite{Kel93,Jac97,Gur99,Jac04,Jac06}
were used to carry out an angular size test: a dependence
different from $\theta \propto z^{-1}$ and
closer to the predictions of expanding Universe models was found. The test
was even used to ascertain not only whether or not the Universe is expanding
but also to constrain the different cosmological parameters. However,
these applications are not free from selection effects\cite{Jac04} and,
as will be discussed in \S \ref{.ultracom}, interpretation of
the results of these tests is not so straightforward.
Another proposal\cite{Mar08a} used the
rotation speed of high redshift galaxies as a standard size indicator
since there is a correlation between size and rotational velocity of galactic
disks. This method was indeed applied\cite{Mar08b} over
a sample of emission-line galaxies with $0.2<z<1$, with the result that
Einstein--de Sitter cosmology is excluded to within 2-$\sigma $, and that
small galaxies (with fixed rotation velocity lower than 100 km/s)
should show a strong evolution in size (at $z=1$ a factor two smaller than
at $z=0$) and no evolution in luminosity within the concordance cosmology,
while large galaxies do not evolve significantly, either in size or in
luminosity. Unfortunately, this method requires spectroscopy, so the
sample has an upper limit of $z\sim 1$ even with very large telescopes,
and the uncertainties are so large that they cannot be interpreted directly
without certain assumptions concerning evolution models. In principle, looking
at figures 4 and 5 of Marinoni {\it et al.}\cite{Mar08b}, one sees no reason to
exclude a $\theta \propto z^{-1}$ dependence.
The aim of this paper is to repeat the angular size test
for galaxies within a wide range of redshifts ($z=0.2-3.2$)
in optical--near infrared surveys (equivalent to the optical at rest).
Data from high spatial resolution surveys available nowadays, such as those
carried out with Hubble Telescope or FIRES, provide useful input for
this old test of the angular size with new analyses and interpretations.
Recent analysis of these data\cite{McI05,Bar05,Tru06}
has shown that the linear size of the galaxies
with the same luminosity should be much lower than locally.
However, this is true only if we considered the standard cosmological model
as correct. In the present paper, I will do a reanalysis of these
data and consider different cosmological scenarios,
which will shed further light on the degeneracy between
expansion + evolution and non-expansion.
\section{Data}
Angular effective radii, defined as circularized
S\'ersic half-light radii within which 50\% of the light is present, are
taken from \cite{McI05,Bar05} for galaxies with $0.2<z<1$ and
from Trujillo {\it et al.}\cite{Tru06} for galaxies with $z>1$.
Both samples separate approximately
early-type and late-type galaxies by means of the exponent ($n_S$)
of their S\'ersic profile: $n_S>2.5$ for early-type galaxies and $n_S<2.5$ for
late type galaxies.
\cite{McI05,Bar05} provide data and angular size
measurement of the GEMS survey with two Hubble Space Telescope colors (F606W and
F850LP). In total, they have 929 galaxies.
The data processing and photometry are discussed by Rix {\it et al.}\cite{Rix04}.
Trujillo {\it et al.}\cite{Tru06} use near-infrared
FIRES data of $z>1$ galaxies in the HDF-S and
MS 1054-03 fields to derive the angular size in the rest-frame V filter.
In total, there are 248 galaxies with $1<z<3.2$. There are 14 more galaxies
with $z>3.2$, but they are very few, very luminous, and their
photometric redshift determination was not very accurate; indeed,
the available on-line data through the FIRES Website,
http://www.strw.leidenuniv.nl/~fires, gives new recalculated
redshifts and some of them are very different.
The data processing and photometry are discussed
in detail by Labb\'e {\it et al.}\cite{Lab03} for HDF-S and F\"orster-Schreiber
{\it et al.}\cite{For06} for the MS 1054-03 field.
From these galaxies, I take only those with $3.4\times
10^{10}<L_{V,rest}(L_{\odot ,V})<2.5\times 10^{11}$,
a total of 393 galaxies (271 galaxies from GEMS with $z<1$, and
122 galaxies from HDF-S/MS 1054-03 fields with $z>1$).
The lower limit is the same as that adopted by
Trujillo {\it et al.}\cite{Tru06}, avoiding the
faintest ones in order to have a more homogeneous sample.
The maximum limit is to avoid the galaxies away from the range
of local galaxies for which the relationship between radius and
luminosity was explored (explained in \S \ref{.astest}).
This is an almost
complete sample for redshifts up to $\approx 2.5$ \cite{Tru06},
and is incomplete for $2.5<z<3.2$. As we will see
later, it is unimportant whether the sample is complete or
not since the test is independent of the luminosity of the
galaxies, but within this restriction I will concentrate on
the analysis of the brightest galaxies.
A higher limit at the lowest luminosity (the sample would be complete up to
$z=3.2$ for $L_{V,rest}>6.7\times 10^{10}$ L$_\odot $)
would reduce the number of galaxies too severely and the statistics
would be poorer. In any case, I will also comment on the results
for these higher luminosity lower limits (see \S \ref{.results}).
\section{Angular size test}
\label{.astest}
I am going to analyze the variation of the angular size,
$\theta _{V,rest}$, of the galaxy rest-frame V with the redshift.
The angular size in the $z<1$ sample was measured at $\lambda _0=9450$
\AA \ (filter F850LP), $\theta_{\lambda _0}$, which,
due to the color gradients, is slightly different from the
angular size at V-rest. This difference is small\cite{McI05,Bar05}
but I apply the following correction to it:
\begin{equation}
\theta_{V,rest}=\left \{ \begin{array}{ll}
\theta_{\lambda _0}
\left[1-0.11\left(\frac{\lambda _V(1+z)}{\lambda _0}-1\right)\right],
& \mbox{ $n_S>2.5$} \\
\theta_{\lambda _0}
\left[1-0.08\left(\frac{\lambda _V(1+z)}{\lambda _0}-1\right)\right],
& \mbox{ $n_S<2.5$}
\end{array}
\right \}
,\end{equation}
with $\lambda_0=9450$ \AA , and $\lambda _V=5500$ \AA .
In any case, this correction is very small (less than 4\%)
and the results would not change significantly if it were not applied.
In HDF-S/MS 1054-03 data, the size was already measured
in the filter which gives V at rest\cite{Tru06}, so no correction is necessary.
Since we have a wide range of luminosities and types of galaxies,
there is a huge dispersion of sizes for a given redshift, with
Malmquist bias, but this dispersion and bias can
be reduced by defining $\theta _*$ as the equivalent angular size if
the galaxy were early-type with a given
V-rest luminosity (I take $10^{10}$ L$_{\odot ,V}$; however,
the variation of this number does not affect any result, just
the calibration in size):
\begin{equation}
\theta _*\equiv
\theta \frac{R(10^{10}\ {\rm L_{\odot ,V}},n_S>2.5)}
{R(L_{V,rest},n_S)}
\label{thetaequiv}
,\end{equation}
where $R(L_{V,rest},n_S)$ is the average radius of
a galaxy for a given luminosity and exponent ($n_S$)
of the S\'ersic profile. The number $n_S$ is affected by an important
uncertainty for galaxies with very small angular size ($\theta <0.125"$)
\cite{Tru06}, which produces some extra dispersion.
In any case, the dispersion of $R$ values is moderate\cite{She03},
and, given that $\theta _*(z)$ does not contain the dispersion of
luminosities, it will present a much lower dispersion than $\theta (z)$.
Certainly, $\theta _*(z)$ contains the spread of $\theta $ and $R$ so the
dispersion due to random errors in these quantities is larger for $\theta _*$ than
for $\theta $; but the dispersion due to the spread of luminosities dominates,
so, as said, $\theta _*$ will present a dispersion much lower than $\theta (z)$.
This definition also avoids selection effects due to Malmquist bias,
since $\theta _*$ for a given redshift should be nearly
independent of the luminosity of the galaxy, at least on average.
Shen {\it et al.}\cite{She03}(Eqs. 14-15) give the median
radius of a galaxy of given $r'_{SDSS}$-luminosity (K-correction applied)
and $n_S$ for local SDSS galaxies.
In total, the radius (measured for $z\approx 0.1$ in the r-band,
which is more or less equivalent to V-band at rest) as a function of the
absolute magnitude in $r'_{SDSS}$-rest $M_{r'-SDSS}$ is:
\begin{equation}
R(M_{r'-SDSS},n_S)
\label{shen}
\end{equation}\[
=\left \{ \begin{array}{ll}
10^{-0.260M_{r'-SDSS}-5.06},& \mbox{ $n_S>2.5$} \\
10^{-0.104M_{r'-SDSS}-1.71}[1+10^{-0.4(M_{r'-SDSS}+20.91)}]^{0.25},
& \mbox{ $n_S<2.5$}
\end{array}
\right \} \ {\rm kpc}
\]
We are not considering the evolution in this Eq. (\ref{shen});
all discussion of the effects of evolution will be considered in
\S \ref{.evol}.
To translate this relationship into a
V-rest luminosity, we must make a color correction, as
in McIntosh {\it et al.}\cite{McI05} [however Trujillo {\it et al.}\cite{Tru06}
interpolate between the rest-frame g-band and
r-band]. Taking into account that
$\langle (V-r'_{SDSS,AB})\rangle$=0.33 for early-type galaxies
and $=0.30$ for late-type galaxies\cite{Fuk95}
(this already includes the transformation of Vega to AB system;
no evolution is considered here),
and that $M_{V,\odot }=+4.79$ (Vega system) we will
have ($L_V$ in units of $10^{10}$ L$_{\odot, V}$)
\begin{equation}
R(L_V,n_S)=\left \{ \begin{array}{ll}
A_0L_V^{0.65},& \mbox{ $n_S>2.5$} \\
B_0L_V^{0.26}(1+B_1L_V)^{0.25},& \mbox{ $n_S<2.5$}
\end{array}
\right \} \ {\rm kpc}
\label{shen2}
,\end{equation}
with $A_0=1.91$, $B_0=2.63$, $B_1=0.692$.
These parameters are valid assuming the concordance cosmological model
($H_0=70$ km/s/Mpc, $\Omega _m=0.3$, $\Omega _\Lambda =0.7$).
For other cosmologies, we have to calibrate the relationship
between luminosities and radii with the corresponding luminosity
and angular distances for $z_{SDSS}\approx 0.1$. See Table \ref{Tab:shenpar}
for the values of $A_0$, $B_0$ and $B_1$ with other cosmologies.
This relationship of Shen {\it et al.}\cite{She03} is fitted with galaxies of
$-19>M_r>-24$ for early types and $-16>M_r>-24$ for late types,
which is equivalent to
$2.5\times 10^9<L_V<2.5\times 10^{11}$ L$_{\odot,V}$ for early types
and $1.6\times 10^8<L_V<2.5\times 10^{11}$ L$_{\odot,V}$ for late types.
Our galaxies are within these limits.
In order to derive the luminosity, $L_{V,rest}$, for a given redshift ($z$) and the
rest flux for the filter V, $F_{V,rest}$,
we need the distance luminosity $d_L(z)$ from different
cosmologies, such that (without considering neither evolution nor extinction)
\begin{equation}
d_L\equiv \sqrt{\frac{L_{V,rest}}{4\pi F_{V,rest}}}
\label{lrest}
.\end{equation}
And the average equivalent angular size evolution with $z$ should be
compared with the prediction of the same cosmology, which, in principle,
without taking into account the evolution, should be given by:
\begin{equation}
d_A(z)\equiv \frac{R_*}{\theta _{*,pred}(z)}
\label{da}
,\end{equation}
where $d_A(z)$ is the angular distance, and $R_*$ is the equivalent
physical radius of the galaxy associated with the equivalent
angular size $\theta _*$; that is, if the galaxy were early-type
of V-rest luminosity $10^{10}$ L$_{\odot, V}$.
\subsection{Different cosmological scenarios}
\label{.cosmomodels}
\begin{enumerate}
\item Concordance model with Hubble constant $H_0=70$ km/s/Mpc,
$\Omega _m=0.3$, $\Omega _\Lambda =0.7$:
\begin{equation}
d_A(z)=\frac{c}{H_0(1+z)}
\int _0^z\frac{dx}{\sqrt{\Omega _m(1+x)^3
+\Omega _\Lambda }}
,\label{concordance}\end{equation}
\begin{equation}
d_L(z)=(1+z)^2d_A(z)
\label{concordance2}
.\end{equation}
\item Einstein--de Sitter model [Eq. (\ref{concordance}) with $\Omega
_\Lambda =0$, $\Omega _m=1$]:
\begin{equation}
d_A(z)=\frac{2c}{H_0(1+z)}\left[1-\frac{1}{\sqrt{1+z}}\right]
,\end{equation}
\begin{equation}
d_L(z)=(1+z)^2d_A(z)
.\end{equation}
Although this is not the standard model nowadays, there are
some researchers who still consider it more appropriate than the concordance
model (e.g., \cite{Vau03,Bla06,And06}).
About the compatibility with type Ia supernovae data, see discussion in \S \ref{.hubble}.
Since the distances for objects at $z=0.1$ are 0.943 times the distances
of the concordance model, the corrected relationship between
luminosity and radius is Eq. (\ref{shen2}) with $A_0=1.90$,
$B_0=2.61$, $B_1=0.699$.
\item Friedmann model of negative curvature with $\Omega =0.3$,
$\Omega _\Lambda=0$, which implies a term of curvature $\Omega _K=0.7$
[\cite{Nab08}, Eq. (26)].
\begin{equation}
d_A(z)=\frac{c}{H_0(1+z)\sqrt{\Omega _K}}
\sinh \left(\sqrt{\Omega _K}\int _0^z\frac{dx}{\sqrt{\Omega _m(1+x)^3
+\Omega _K(1+x)^2 }}\right)
,\end{equation}
\begin{equation}
d_L(z)=(1+z)^2d_A(z)
.\end{equation}
I will use this model of an open universe to check that a model different
to a flat universe does not change the results significantly.
Since the distances for objects at $z=0.1$ are 0.970 times the distances
of the concordance model, the relationship between
luminosity and radius with corrected calibration is
Eq. (\ref{shen2}) with $A_0=1.93$,
$B_0=2.59$, $B_1=0.736$.
\item Quasi-Steady State Cosmology (QSSC),
$\Omega _m=1.27$, $\Omega _\Lambda =-0.09$, $\Omega _c=-0.18$
(C-field density) \cite{Ban99}
\begin{equation}
d_A(z)=\frac{c}{H_0(1+z)}
\int _0^z\frac{dx}{\sqrt{\Omega _c(1+x)^4+\Omega _m(1+x)^3
+\Omega _\Lambda }}
,\end{equation}
\begin{equation}
d_L(z)=(1+z)^2d_A(z)
.\end{equation}
This cosmology is not the standard model, but it can also fit
many data on angular size tests\cite{Ban99}
or Hubble diagrams for SNe Ia\cite{Ban00,Nar02,Vis02}.
The expansion with an oscillatory term gives a dependence
of the luminosity and angular distance similar to the standard model,
adding the effect of matter creation (C-field) with
slight changes depending on the parameters.
The parameters of this cosmology are not as well constrained as
those in the standard model. Here, I use the
best fit for a flat ($K=0$) cosmology given by Banerjee {\it et al.}\cite{Ban99}:
$\Omega _m=1.27$, $\Omega _\Lambda =-0.09$, $\Omega _c=-0.18$, which
corresponds to $\eta =0.887$ (amplitude of the oscillation relative to 1),
$x_0=0.797$ (ratio between actual size of the Universe and the average
size in the present oscillation) and maximum allowed redshift of a galaxy
$z_{max}=6.05$ (note however that the maximum observed redshift has
risen above 8 nowadays according to some authors, \cite{Tan09}). Other
preferred sets of parameters give results that are close.
The values most used are $K=0$, $\Omega _\Lambda =-0.36$, $\eta =0.811$,
$z_{max}=5$ \cite{Ban00,Nar02,Vis02,Nar07},
which imply $\Omega _m=1.63$, $\Omega _c=-0.27$,
but I avoid them because they do not allow galaxies to be fitted with $z>5$.
Parameters with a curvature different from zero ($K\ne 0$) also give results
that are very close in the angular size test\cite{Ban99}.
Since the distances for objects at $z=0.1$ are 0.943 times the distances
of the concordance model, the relationship between
luminosity and radius with corrected calibration is
Eq. (\ref{shen2}) with $A_0=1.94$,
$B_0=2.56$, $B_1=0.778$.
\item Static euclidean model with linear Hubble law for all redshifts:
\begin{equation}
d_A(z)=\frac{c}{H_0}z
\label{angdistst}
,\end{equation}
\begin{equation}
d_L(z)=\sqrt{1+z}d_A(z)
.\end{equation}
These simple relations
indicate that redshift is always proportional
to angular distance, a Hubble law. We assume in this scenario that the
Universe is static; the factor $\sqrt{1+z}$ in the
luminosity distance stems from the loss of energy due to redshift
without expansion. There is no time dilation, and precisely because
of that the factor is not $(1+z)$.
The caveat is to explain
the mechanism different from the expansion/Doppler effect, which gives rise
to the redshift. This cosmological model is not a solution which
has been explored theoretically/mathematically.
However, from a phenomenological point of view, we can
consider this relationship between distance and redshift
as an ad hoc extrapolation from the observed dependence on the low redshift
Universe. Our goal here is to see how
well it fits the data, and forget for the moment the theoretical
derivation of this law.
Since the luminosity distance for objects at $z=0.1$ is 0.966 times
the luminosity distance of the concordance model
and the angular distance is 1.115 times the angular distance of
the concordance model, we must adopt approximately Eq. (\ref{shen2})
with $A_0=2.23$, $B_0=2.98$, $B_1=0.741$.
\item Tired-light/simple static euclidean model :
\begin{equation}
d_A(z)=\frac{c}{H_0}{\rm ln} (1+z)
\label{angdisttl}
,\end{equation}
\begin{equation}
d_L(z)=\sqrt{1+z}d_A(z)
.\end{equation}
This is again a possible ad hoc phenomenological representation
which stems from considering that the photons
lose energy along their paths due to some interaction, and the
relative loss of energy is proportional to the length of that path
(e.g. \cite{LaV86}), i.e.
\begin{equation}
\frac{dE}{dr}=-\frac{H_0}{c}E
.\end{equation}
Of course, as in the previous case, this ansatz is very far
from being considered as the correct one by most cosmologists,
but it is interesting to analyze its compatibility with the angular
size test too.
For the calibration of Eq. (\ref{shen2}), I use the fact that
the luminosity distance for objects at $z=0.1$ is 0.921 times
the luminosity distance of the concordance model,
and the angular distance is 1.063 times the angular distance of
the concordance model, so $A_0=2.26$, $B_0=2.92$, $B_1=0.816$.
\item Tired-light/``Plasma redshift'' static euclidean model:
\begin{equation}
d_A(z)=\frac{c}{H_0}{\rm ln} (1+z)
,\end{equation}
\begin{equation}
d_L(z)=(1+z)^{3/2}d_A(z)
.\end{equation}
The plasma redshift application\cite{Bry04a}(\S 5.8)
used $d_L(z)=(1+z)^{3/2}d_A(z)$ instead of
$d_L(z)=(1+z)^{1/2}d_A(z)$ to take into
account an extra Compton scattering which is double that of the plasma redshift
absorption. For the calibration of Eq. (\ref{shen2}):
$A_0=2.00$, $B_0=2.78$, $B_1=0.674$.
\end{enumerate}
The volume element in a static uiverse is different from
the volume element in the standard concordance model.
Particularly for the standard concordance model the comoving volume element in a
solid angle $d\omega $ and redshift interval $dz$ is (for null curvature, which is
the case of the concordance model)
\begin{equation}
dV_{\rm concordance}=\left(\frac{c}{H_0}\right)^3
\frac{\left[\int _0^z\frac{dx}{\sqrt{\Omega _m(1+x)^3
+\Omega _\Lambda }}\right]^2}{\sqrt{\Omega _m(1+z)^3
+\Omega _\Lambda }}d\omega dz
\label{vol1}
\end{equation}
while for the two first static models it is
\begin{equation}
dV_{\rm static-lin.Hub.}=\left(\frac{c}{H_0}\right)^3z^2d\omega dz
,\end{equation}
\begin{equation}
dV_{\rm static-tir.l.}=\left(\frac{c}{H_0}\right)^3\frac{[{\rm ln}(1+z)]^2}
{(1+z)}d\omega dz
\label{vol2}
.\end{equation}
Hence, if we wanted to evaluate the evolution of some quantity
per unit comoving volume for the static universes, we must multiply
the result in the concordance model by the factor $\frac{dV_{\rm concordance}}
{dV_{\rm static}}$.
\subsection{Results}
\label{.results}
The results of the test are plotted in Fig. \ref{Fig:sizes0} for the
concordance model, with the equivalent angular size of each individual
galaxy, and Fig. \ref{Fig:angsize} for the different models,
with the representation of the average value of
$\log _{10}\theta _*$ in bins of $\Delta \log _{10}(z)=0.10$.
I do a weighted linear fit in the log--log plot.
Since I am calculating an average of the logarithm for the angular
sizes, the possible error in individual galaxies should not significantly affect
the value of the average.
The error bars in each bin of the plot are the statistical errors.
Note that there are values of $\theta _*$ lower than 0.03$''$, but they do indeed
correspond to measured values of $\theta \ge 0.03''$;
$\theta _*$ is lower than $\theta $ at the highest redshifts because
the luminosity in V-filter of those galaxies is higher than $10^{10}$ L$_{\odot ,V}$.
As can be observed, the average
equivalent angular size gives a good fit to a $\theta _*(z)=Kz^{-\alpha }$ law,
with values of $K$, $\alpha $ given in Table \ref{Tab:fit}.
No fit is totally in agreement with the cosmological
prediction without evolution and extinction.
The static model with a linear Hubble
law is not very far from being compatible
with the data: I get $\theta _*(z)=0.136z^{-0.97}$,
near the expected dependence ($\theta _*(z)=0.109z^{-1}$)
although with a slightly larger size.
{\it {\Large \bf NOTE}:
the normalization of the angular size in any model
prediction (solid line in Fig. \ref{Fig:angsize})
stems from the Shen {\it et al.}\cite{She03} calibration with
the corresponding parameters $A_0$, $B_0$ and $B_1$ in each
cosmology. It does not stem from a fit.}
If we separate elliptical ($n_S>2.5$) and disk galaxies
($n_S<2.5$) for the concordance model,
the plots in Fig. \ref{Fig:angsizetype} are obtained, with
a higher slope for elliptical galaxies ($\alpha =1.18\pm 0.04$) than for
disk galaxies ($\alpha =0.80\pm 0.07$). This might be due to a different
evolution of disk and elliptical galaxies, but they
are likely to be due to different systematic errors for elliptical galaxies and
disk galaxies due to systematic errors in the value of
$n_S$. The galaxies with $2.0<n_S<2.5$
are suspected of being strongly contaminated by ellipticals since
the obtained $n_S$ tends to be lower than the real one
(\cite{Tru06}, Fig. 1). If we take disk galaxies only
with $n_S<2.0$, then $\alpha =0.92\pm 0.06$, closer to unity.
On the other hand, elliptical galaxies
with $\theta <0.125"$ are also strongly contaminated by
disk galaxies (\cite{Tru06}, Fig. 2) which are compacted by a wrong measure
of $n_S$ and consequently give a smaller radius than the real one.
If we take elliptical ($n_S>2.5$) galaxies only with $\theta >0.125"$:
$\alpha =1.03\pm 0.05$. Therefore, from the present
analysis and within the systematic errors, we cannot be
sure that the angular size test gives different results
for elliptical and disk galaxies. However, when
all the elliptical and disk galaxies are put together, the excesses
and deficits more or less compensate; there are approximately
as many elliptical galaxies misclassified as disk galaxies as
disk galaxies misclassified as elliptical galaxies.
See further discussion on the systematic errors in \S \ref{.select}.
In Fig. \ref{Fig:angsizelum}, I analyze the dependence of
$\alpha $ on the luminosity of the galaxies, and we see that
there is no significant dependence. $\alpha =1$ is
more or less compatible with all the subsamples of different
luminosity.
\begin{figure}
\vspace{1cm}
{\par\centering \resizebox*{8cm}{8cm}{\includegraphics{fig1.eps}}}
\caption{Log--log plot of the equivalent angular size ($\theta _*$)
vs. redshift ($z$) in the concordance cosmology. Left: ($z<1$)
GEMS data. Right: ($z>1$) HDF-S and MS 1054-03 data.
The average of $\log _{10} \theta _*$ in bins of $\Delta \log _{10}(z)=0.10$
is represented with asterisks and squares with statistical error bars.}
\label{Fig:sizes0}
\end{figure}
\begin{table}
\caption{}
Values of the parameters $A_0$, $B_0$ and $B_1$ in the
relationship between radius and luminosity of a galaxy, Eq.
(\ref{shen2}), calibrated
with different cosmological models.
\begin{center}
\begin{tabular}{cccc}
\label{Tab:shenpar}
Cosmology & $A_0$ & $B_0$ & $B_1$ \\ \hline
Concord. $\Omega _m=0.3$, $\Omega _\Lambda =0.7$ & 1.91 & 2.63 & 0.692 \\
Einstein-de Sitter & 1.94 & 2.56 & 0.778 \\
Friedmann $\Omega _m=0.3$ & 1.93 & 2.59 & 0.736 \\
QSSC $\Omega _m=1.27$, $\Omega _\Lambda =-0.09$, $\Omega _c =0.18$ & 1.94 & 2.56& 0.778 \\
Static linear Hubble law & 2.23 & 2.98 & 0.741 \\
Static, tired light/simple & 2.26 & 2.92 & 0.816 \\
Static, tired light/plasma & 2.00 & 2.78 & 0.674 \\ \hline
St. lin. Hub. law, ext. $a_V=1.6\times 10^{-4}$ Mpc$^{-1}$ & 2.21 & 3.03 & 0.696 \\
St. tired light, ext. $a_V=3.4\times 10^{-4}$ Mpc$^{-1}$ & 2.22 & 3.01 & 0.719 \\
\end{tabular}
\end{center}
\end{table}
\begin{table*}
\caption{}
Weighted fit of the average
equivalent angular size for the data in Fig. \ref{Fig:angsize}
to a law $\theta _*(z)=Kz^{-\alpha }$. Error bars give only statistical
errors and do not include systematic errors as explained in \S
\protect{\ref{.select}}.
\begin{center}
\begin{tabular}{ccc}
\label{Tab:fit}
Cosmology & K & $\alpha $ \\ \hline
Concord. $\Omega _m=0.3$, $\Omega _\Lambda =0.7$ & $0.1251\pm 0.0041$ &
$0.957\pm 0.045$ \\
Einstein-de Sitter & $0.1721\pm 0.0054$ & $0.834\pm 0.044$ \\
Friedmann $\Omega _m=0.3$ & $0.1412\pm 0.0046$ & $0.956\pm 0.045$ \\
QSSC $\Omega _m=1.27$, $\Omega _\Lambda =-0.09$, $\Omega _c =0.18$ & $0.1810\pm 0.0057$ & $0.817\pm 0.044$ \\
Static linear Hubble law & $0.1363\pm 0.0044$ & $0.969\pm 0.045$ \\
Static, tired light/simple & $0.2132 \pm 0.0066$ & $0.717\pm 0.043$ \\
Static, tired light/plasma & $0.0915 \pm 0.0031$ & $1.177\pm 0.047$ \\ \hline
Concord., early type ($n_S>2.5$) & $0.1059\pm 0.0044$ & $1.181\pm 0.043$ \\
Concord., late type ($n_S\le 2.5$) & $0.1441\pm 0.0070$ & $0.805\pm 0.067$ \\
Concord., early type ($n_S>2.5$), $\theta >0.125"$ & $0.1226\pm 0.0060$ & $1.031\pm 0.051$ \\
Concord., late type ($n_S\le 2.0$) & $0.1600\pm 0.0073$ & $0.915\pm 0.057$ \\
Concord., $L_V>6.8\times 10^{10}$ L$_{\odot ,V}$ & $0.1315\pm 0.0083$ & $0.995\pm 0.098$ \\
Concord., $L_V>1.02\times 10^{11}$ L$_{\odot ,V}$ & $0.1106\pm 0.0117$ & $1.313\pm 0.186$ \\ \hline
St. lin. Hub. law, ext. $a_V=1.6\times 10^{-4}$ Mpc$^{-1}$ & $0.1055\pm 0.0035$ & $1.113\pm 0.046$ \\
St. tired light, ext. $a_V=3.4\times 10^{-4}$ Mpc$^{-1}$ & $0.1565\pm 0.0050$ &$0.801\pm 0.044$ \\
\end{tabular}
\end{center}
\end{table*}
\begin{figure*}[htb]
\vspace{1cm}
{\par\centering \resizebox*{4cm}{4cm}{\includegraphics{fig2a.eps}}
\hspace{1cm}\resizebox*{4cm}{4cm}{\includegraphics{fig2b.eps}}\par}
\vspace{1cm}
{\par\centering \resizebox*{4cm}{4cm}{\includegraphics{fig2c.eps}}
\hspace{1cm}\resizebox*{4cm}{4cm}{\includegraphics{fig2d.eps}}\par}
\vspace{1cm}
{\par\centering \resizebox*{4cm}{4cm}{\includegraphics{fig2e.eps}}
\hspace{.2cm}\resizebox*{4cm}{4cm}{\includegraphics{fig2f.eps}}
\hspace{.2cm}\resizebox*{4cm}{4cm}{\includegraphics{fig2g.eps}}\par}
\caption{}
Log--log plot of the average of $\log _{10} \theta _*$, where
$\theta _*$ is the equivalent angular size, vs. redshift ($z$).
Bins of $\Delta \log _{10}(z)=0.10$. Error bars only represent
statistical errors; for the systematic errors, see text in \S
\ref{.select}.
The seven plots are for the seven different cosmologies
described in \S \protect{\ref{.cosmomodels}}. Solid lines are
the model predictions (the normalization stems from the
Shen {\it et al.}\cite{She03} calibration with
the corresponding parameters $A_0$, $B_0$ and $B_1$ in each
cosmology). Dashed lines are the best weighted linear fits.
\label{Fig:angsize}
\end{figure*}
\begin{figure*}
\vspace{1cm}
{\par\centering \resizebox*{5cm}{5cm}{\includegraphics{fig3a.eps}}
\hspace{1cm}\resizebox*{5cm}{5cm}{\includegraphics{fig3b.eps}}\par}
\caption{Log--log plot of the average of $\log _{10} \theta _*$, where
$\theta _*$ is the equivalent angular size, vs. redshift ($z$).
Bins of $\Delta \log _{10}(z)=0.10$.
The two plots are for the concordance cosmology separating
elliptical galaxies ($n_S>2.5$) from disk galaxies
($n_S\le 2.5$).}
\label{Fig:angsizetype}
\end{figure*}
\begin{figure*}
\vspace{1cm}
{\par\centering \resizebox*{5cm}{5cm}{\includegraphics{fig4a.eps}}
\hspace{1cm}\resizebox*{5cm}{5cm}{\includegraphics{fig4b.eps}}\par}
\caption{Log--log plot of the average of $\log _{10} \theta _*$, where
$\theta _*$ is the equivalent angular size, vs. redshift ($z$).
Bins of $\Delta \log _{10}(z)=0.20$ and $\Delta \log _{10}(z)=0.30$
respectively.
The two plots are for the concordance cosmology only for
galaxies with $L_V>6.8\times 10^{10}$ L$_{\odot ,V}$ (122 galaxies)
and $L_V>1.02\times 10^{11}$ L$_{\odot ,V}$ (48 galaxies)
respectively.}
\label{Fig:angsizelum}
\end{figure*}
\subsection{Selection effects, and errors in the angular size
measurement and luminosity-radius relationship}
\label{.select}
As said previously, the definition of $\theta _*$ avoids
selection effects due to Malmquist bias,
since $\theta _*$ for a given redshift should be nearly
independent on how luminous the galaxy is, at least on average.
Nonetheless, although $R_*=\theta _*d_A$ is
independent of the luminosity and type of the galaxy
at $z=0$, its possible evolution might depend on both parameters and this
would give different average $\theta _*$ when the
selection of galaxies is different.
In any case, whatever the sample of galaxies is, if we
are going to interpret the factor
between the cosmological prediction and the observed average $\theta _*$
as a product of size evolution, this factor would
reflect the average evolution of $R_*$ in that sample.
In our case, we are observing the factors for the average population
with $3.4\times 10^{10}<L_{V,rest}(L_\odot )<2.5\times 10^{11}$.
For the concordance model the value of $R_*$ at $z=3.2$ is on
average $\approx 6$ lower than $R_*$ at $z=0$.
At low redshift we know that the
relationship of Eq. (\ref{shen2}) is correct (hence, the average size $R_*$
should not depend on the luminosity of these galaxies). Therefore,
the ratio of sizes between high/low redshift objects
will depend only on the variation of Eq. (\ref{shen2}) in
galaxies selected at high redshift.
We can say that the average size of the selected galaxies at
$z=3.2$ is $\approx 6$ times lower than the galaxies at
low redshift with the equivalent luminosity and S\'ersic profile.
In other words, if there is a (very strong) evolution in
the size of the galaxies for a fixed luminosity (or a variation in
luminosity for a fixed radius),
the factor by which the galaxies are smaller
will depend on which galaxies are selected, but in any case this
factor will represent an average galaxy shrinking factor.
An average factor of 6 in size reduction will mean that there
are galaxies with size reduction by a factor larger than 6, and other
galaxies with size reduction by a factor smaller than 6.
Trujillo {\it et al.}\cite{Tru06}(\S 4.3) discussed the robustness of the
``average'' angular size measurement, whether it is affected
by some other biases or selection effects. Their tests showed that the
results, as presented here, are more or less robust and
the difference in the results if we apply some minor corrections due to
uncertainties or incompleteness in the radius distribution
of the sources are $\Delta (\langle\log _{10}R\rangle)\sim 0.1$,
thus $\Delta \theta _*/\theta _*\sim 20$\% .
There is a minimum angular size that Hubble telescope or FIRES can observe
below which the uncertainties of $\theta $ and $n_S$
are very high, with important systematic deviations.
Normally, the angular sizes are overestimated for these
low angular sizes, so the observed average $\theta _*$
increases with $z$ as more and more of the
smaller angular size galaxies are included in the average.
Moreover, the very compact galaxies are not included in the
sample because they might be classified as stars.
This affects mainly the $z>1$ data, where the limit for
low systematic errors is around $\theta =0.125"$ \cite{Tru06}.
Suppression of these galaxies
would change the average of $\theta _*$ at $z>1$ by 10--30\%.
Systematic errors for $\theta <0.125''$ are up to 50\% or lower (\cite{Tru06},
Fig. 2), which produces a systematic error
on the average $\theta _*$ lower than $\approx 15$\%.
Another effect to investigate is the error in the relationship of Eq.
(\ref{shen2}). This represents the average radius for a given luminosity
and there is some dispersion of values with respect to it (given
by Shen {\it et al.}\cite{She03}). I am not interested here in these statistical errors
of $R$ because this would just affect a dispersion of
values of $\theta _*$ without changing its average value. Our concern
is about the systematic errors in Eq. (\ref{shen2}). I will
analyze five sources of systematic errors:
\begin{enumerate}
\item The color correction of $\langle (V-r'_{SDSS,AB})\rangle$=0.33
for early-type galaxies and $=0.30$ for late-type galaxies \cite{Fuk95}
might change if there is a preferential galactic
type within the group of early- or late-type galaxies is given for some redshift.
Indeed, Fukugita {\it et al.}\cite{Fuk95} give a value of
$\langle (V-r'_{SDSS,AB})\rangle$ of
0.36 for E type, 0.31 for S0, 0.32 for S$_{ab}$, 0.29 for S$_{bc}$,
0.28 for S$_{cd}$. That is, there are variations up to $\approx 0.03$
with respect to the average in early types, and 0.02 for late types.
In the worst cases, these systematic errors in colors would produce
a systematic error of $\approx 2$\% in $\theta _*$, which is negligible.
\item A misclassification of an elliptical galaxy as disk galaxy or vice versa
produces an error up to a factor 2 in the value of $R$ derived from
Eq. (\ref{shen2}), which leads to an error of factor $<2$ in $\theta _*$.
The effect of the uncertainty in $n_S$ for small angular sizes
has already been checked by Trujillo {\it et al.}\cite{Tru06}(\S 4.3) and the error of the
average of $\theta _*$ is lower than $\sim 20$\% for the sample with
all of the galaxies.
However, when I divide the galaxies into elliptical and disk types
(see \ref{Fig:angsizetype}) the effect might be larger, as
was noted previously concerning Fig. \ref{Fig:angsizetype}.
\item Systematic errors in the luminosity at V-rest due to
errors in the photometric redshift amount less than 4\% \cite{Rud06},
which implies errors in $\theta _*$ less
than 2.5\%. However, the systematic error of the luminosity
measurement in faint elliptical galaxies amounts around 15\% \cite{Tru06},
which translates into 6\% of error in $\theta _*$.
\item The calibration of eq. (\ref{shen2}) (calibration of
the parameters $A_0$, $B_0$ and $B_1$) is done
for cosmologies different to concordance model
assuming that the SDSS galaxies have redshift $z_{SDSS}=0.1$. There is indeed
a dispersion of redshift values in SDSS galaxies around this average.
If we set $z_{SDSS}=0.15$ we would obtain, for instance for the
static linear Hubble law ansatz, $A_0=2.32$, $B_0=3.15$, $B_1=0.721$;
the variations in the average measured $\theta _*$ are negligible
($<<1$\%), but the predictions of the cosmological model for $\theta _*$
(proportional to $A_0$) increase by $\approx 6$\%. That is, the ratio
between measured and predicted $\theta _*$ decreases by 6\%.
An error of 0.05 in the average redshift is among the worst cases,
so we can say that the systematic error in the calibration should
be less than $\approx 5$\%.
\item Another systematic error comes from the $V_{max}$
correction of the selection effects in the SDSS data made by
Shen {\it et al.}\cite{She03}(\S 3.1). Shen {\it et al.} \cite{She03}
make a correction of selection
effects by assigning a weight to each galaxy inversely proportional
to the maximum comoving volume within which galaxies identical to one under
consideration can be observed\cite{Qin97}.
This correction mainly affects
faint elliptical galaxies, as can be
observed in fig. 2 of Shen {\it et al.}\cite{She03} and amounts
to $\frac{\Delta R}{R}<\sim 0.7\times exp(-0.7L_V)$
in the average radius of the galaxies
for elliptical galaxies of luminosity $L_V$ (in units of
$10^{10}$ L$_{\odot, V}$).
The weights depend on the cosmological
model, since the comoving volumes at a given redshift change in the
different cosmological models. Moreover, there is an intrinsic
systematic error in using the standard model:
Shen {\it et al.}\cite{She03} used the integration of physical volumes instead
of comoving volumes [see Eq. (8) of Shen{\it et al.}\cite{She03}], which
leads to systematic errors of $V_{max}^*$ up to $\sim 50$\%
[there is a factor $(1+z)^3$ between comoving and physical volumes].
In order to calculate the effect of a change of cosmology, apart
from the recalibration of $A_0$, $B_0$ and $B_1$, I would need to have
all their SDSS data at hand and repeat the full analysis of their
fit with the new conditional maximum volume $V_{max}^*$ values for
each cosmology [see Eq. (9) of Shen {\it et al.}\cite{She03}] using the correct
comoving volume $V_{max}$ instead of the integration of physical
volumes, which is not possible for us. Assuming a
total systematic error in the volume of $\Delta V_{max}^*\sim 0.5V_{max}^*$,
\begin{equation}
\Delta (\langle\log _{10}R\rangle)<\sim 0.15\times exp(-0.7 \langle L_V\rangle)
.\end{equation}
Since $L_V>3.4$ in our selected sample,
this may justify systematic errors in $R(L)$ up to 3\% .
\end{enumerate}
Summing up, apart from the statistical errors plotted in
Fig. \ref{Fig:angsize}, there are systematic errors in the average
$\theta _*$ that may amount up to 30--40\% for the general
case with all the galaxies
\subsection{Relationship with the surface brightness test}
The average surface brightness of a galaxy with total flux
$F_{\lambda ,rest}$ and half-light circularized angular size $\theta $
is:
\begin{equation}
SB=\frac{F_{\lambda ,rest}/2}{\pi \theta ^2}
.\end{equation}
Using eqs. (\ref{lrest}), (\ref{da}); and $d_L=(1+z)^{i/2}d_A$,
with $i=1$ if static, and $i=4$ if expanding, it is
\begin{equation}
SB=\frac{L_{\lambda ,rest}}
{8\pi ^2R^2(1+z)^i}
\label{sb}
.\end{equation}
The intrinsic surface brightness [without the $(1+z)^i$ dimming factor]
is
\begin{equation}
SB_0=\frac{L_{\lambda ,rest}}{8\pi ^2R^2}
\label{sb0}
.\end{equation}
Given the relationship between radius and luminosities of eq.
(\ref{shen2}), we find that the average intrinsic surface
brightness in V-rest (assuming no size evolution) should follow:
\begin{equation}
SB_{0,V-rest}(L_V,n_S)=\left \{ \begin{array}{ll}
\frac{L_V^{-0.30}}{8\pi ^2A_0^2},& \mbox{ $n_S>2.5$} \\
\frac{L_V^{0.48}}{8\pi ^2B_0^2(1+B_1L_V)^{0.50}},& \mbox{ $n_S<2.5$}
\end{array}
\right \}
.\end{equation}
The surface brightness for a given luminosity should
be independent of the redshift if there is no evolution in size.
I can also define an equivalent surface brightness to avoid the
dependence on the luminosity:
\begin{equation}
SB_*\equiv \frac{F_{V,rest}}{2\pi \theta ^2}
\frac{SB_0(10^{10}\ {\rm L_{\odot ,V}},>2.5)}
{SB_0(L_{V,rest},n_S)}=\frac{10^{10}\ {\rm L_{\odot ,V}}}
{8\pi ^2R_*^2(1+z)^i}
.\end{equation}
This is indeed the surface brightness test, or Tolman test\cite{Hub35}.
In Fig. \ref{Fig:angsize} for the static models,
it is observed that the data of $\theta _*(z)$ without size
evolution are fitted more or less (within
the statistical+systematic error). Thus,
$R_*$ is nearly constant in a static model for all redshifts and we obtain
$\langle SB_*\rangle \propto (1+z)^{-1}$, while for the expanding
model we would need a strong evolution in $R_*(z)$ and
the average surface brightness would be
$\langle SB_*\rangle \propto R(z)^{-2}(1+z)^{-4}$. In order to give
the same $\langle SB_*\rangle (z)$ as in the static Universe,
assuming that $d_L(z)$ is approximately the same (which is
nearly true for the concordance and the linear Hubble law luminosity
distances; see \S \ref{.hubble}), $R(z)\approx R(z=0)(1+z)^{-3/2}$.
Lubin \& Sandage\cite{Lub01} obtained
$\langle SB\rangle \propto (1+z)^i$ with $i=1.6-3.2$ for a
sample with $z<0.9$ depending on the filter
(R or I), and the initial hypothesis (expanding or static).
Lerner\cite{Ler06} obtained $i=1.03\pm 0.15$ for $z<\sim 5$,
with data in wavelengths
from ultraviolet to visible from Hubble Space Telescope
(see also Lerner\cite{Ler09}),
while Nabokov \& Baryshev\cite{Nab08} obtained $i=3-4$ with the
same type of data but without including K-corrections.
Andrews\cite{And06} obtained $i=0.99\pm 0.38$, $i=1.15\pm 0.34$,
with two different samples of cD galaxies.
Lubin \& Sandage argue that their data are compatible with
an expanding Universe if we take evolution into account.
Lerner criticizes Lubin \& Sandage
for not using the same range of wavelengths at rest for the
high and low redshift galaxies but instead using K-corrections with
many free parameters, which is less direct and
more susceptible to errors. Moreover, Lerner argue against the
evolution that the intrinsic ultraviolet surface brightness
of high redshift galaxies would be extremely large, with impossible
values (see \S \ref{.uv}).
\section{Evolution of galaxies in expansion models}
\label{.evol}
From the plot in Fig. \ref{Fig:angsize}(concord. model), we see that, in order
to make the concordance model compatible with the data on angular size,
we must assume an evolution such that the galaxies at high redshift are
much smaller than at low redshift ($z<\sim 0.2$). For instance, at redshift
$z=2.5$, the galaxies should be on average $\approx 4.6$ times smaller.
Trujillo {\it et al.}\cite{Tru06} obtained a lower average factor ($\sim 3$)
because they used only galaxies with the constraint
$\theta >0.125''$; however in the error bars of Trujillo et al.
result, this bias effect of avoiding $\theta \le
0.125''$ galaxies is included, and within this error bar
it is compatible with our result.
The most massive galaxies are thought to be contracted by a factor
of $\approx 4$ up to $z=1.5$ \cite{Tru07} and
of $5.5$ up to $z=2.3$\cite{van08}, but Mancini {\it et al.}\cite{Man09} think that
the extra compactness of these galaxies can be understood in terms of fluctuations
due to noise preventing the recovery of the extended low surface brightness halos
in the light profile, so those factors might be not real.
For galaxies with redshift $z=3.2$ in our Fig.
\ref{Fig:angsize}(concord. model),
the average ratio of sizes would increase to a factor 6.1
(comparing the linear fit with the theoretical prediction).
Separating by types, the ratio would be 4.5--6 for disk galaxies
and 6--10 for elliptical galaxies, taking into account the systematic
uncertainties commented in \S \ref{.results}.
Other authors\cite{Fer04,Bou04,Tru04,Wik08,Cim08,vand08,Nab08}
find similar results at high redshift.
If we have galaxies that are on average 6 times smaller than the
equivalent galaxies (of same type and luminosity) at low redshift,
this means that the V-rest luminosity density (inversely proportional to the
cube of the size) of these galaxies is $\approx 200$ times higher
than at low redshift. The surface brightness in V-rest is increased by a
factor $\sim 40$. Is this situation possible?
\subsection{Effect of the expansion of the Universe}
\label{.expansion}
In the standard picture of the expanding models, the expansion does not
affect the galaxies because there is no local effect on particle dynamics
from the global expansion of the universe: the tendency to separate
is a kinematic initial condition, and once this is removed, all memory
of the expansion is lost\cite{Pea08}.
Note, however, that there are other views on the topic of whether there is
expansion of galaxies due to the expansion of the Universe
and the nature of the expansion itself \cite{Fra07,Bar08}.
Dark energy or cosmological vacuum might have some influence
on the size of the galaxies \cite{Now01,Ser07}, but this effect would be small.
Lee\cite{Lee08} suggested instead that the
size of the galaxies increases as the scale factor of the universe
assuming dark matter models based on a Bose--Einstein condensate or
scalar field of ultra-light scalar particle, which is another
heterodox idea to explore. Nonetheless, apart from this
kind of proposals, within standard scenarios, the galaxies do not expand
with the Universe.
An indirect way in which the expansion has an effect is in the formation of galaxies.
In the theoretical $\Lambda $CDM hierarchical scenarios, galaxies formed
at higher redshift should be denser\cite{Mo98} since, to decouple
from expansion, structures must have a given density ratio with the
surrounding density, which is larger at higher redshift. Some authors
(e.g., \cite{Tru06,Fer04}) have used this argument to explain the apparent
size evolution.
However, the observed redshift in galaxies is not its formation redshift,
so the application of this idea is not straightforward.
As a matter of fact, the stellar populations
of most local massive elliptical galaxies are very old\cite{Jim07},
and formed before the age corresponding to
$z\sim 3$, so we cannot say that galaxies now are larger
because of this effect since most of them were formed $>10$ Gyr ago.
The difference of formation epoch of the galaxies
observed now and at $z\sim 3$ is not large.
Moreover, the hierarchical scenario in which massive galaxies form first do
not represent the observed Universe appropriately, as I argue in \S \ref{.mergers}.
Furthermore, the theoretical claim by Mo {\it et al.}\cite{Mo98} derived from the models
that galaxies which formed earlier are denser is not in general observed.
If it were true, we should observe that at low redshift the youngest galaxies
(formed later) should be much larger for a given mass
than the oldest galaxies of age 12--13 Gyr. There
is already evidence that this is not the case: the densest galaxies are young
instead of old\cite{Tru09}. And we can check with our own sample within $0.2<z<3.2$ that
the color of elliptical galaxies is not correlated with size: Fig. \ref{Fig:bmv}.
Redder elliptical galaxies are older and, for a given redshift, indicate earlier
formation, which should be equivalent to smaller size, at least statistically.
This correlation is not observed at all: linear fits in the four redshift ranges
of Fig. \ref{Fig:bmv} all give slopes compatible with zero within $1\sigma$ except
for the range $-0.4<\log_{10}(z)<-0.1$, which gives
$\frac{d(log_{10}(R_*/A_0))}{d(B-V)_{\rm rest}}=+1.0\pm 0.5$,
a $2\sigma $ correlation but in a direction opposite to prediction that galaxies
are larger when redder=older (formed earlier).
Therefore, it is not a question of comparing
the formation of galaxies at different redshifts but the evolution of galaxies
already formed, either isolated or in interaction/merging with other systems.
\begin{figure}
\vspace{1cm}
{\par\centering \resizebox*{8cm}{8cm}{\includegraphics{bmv.eps}}}
\caption{}
Equivalent linear size (normalized with Shen {\it et al.}'s\cite{She03} calibration)
vs. $(B-V)$ color at rest for elliptical galaxies
in the concordance cosmology.
\label{Fig:bmv}
\end{figure}
\subsection{Younger population of high redshift galaxies}
\label{.brighter}
The main argument in favour of the evolution in size for a fixed
luminosity is that the younger a galaxy is the brighter it is, and
we expect to see younger galaxies at high redshift. Therefore,
galaxies with radius smaller than $R_*$ in the past will
produce the same luminosity as galaxies with that radius at
present. How much brighter?
Using Vazdekis {\it et al.}'s\cite{Vaz96} synthesis model,
with revised Kroupa IMF, we can derive the mass--luminosity ratio in a passively
evolving elliptical galaxy as a function of its intrinsic (B-V) color and its
luminosity. The metallicity degeneracy is
broken with an iterative method which uses the relationship between
stellar mass and metallicity\cite{Lop09}. With this method,
the average mass--luminosity ratio of
elliptical galaxies (not affected by extinction) at the last bin of
Fig. \ref{Fig:angsize} ($z=2.5-3.2$) is $\langle M_*/L_V\rangle =0.5$
[$N=6$, $\langle L_V\rangle =10^{11}$ L$_\odot $, $\langle (B-V)\rangle =0$]. Rudnick
{\it et al.}'s\cite{Rud06}(\S 4.4) analysis of the same galaxies
also obtain a quite similar mass--luminosity ratio.
For the elliptical galaxies with $\langle L_V\rangle >5\times 10^{10}$ L$_\odot $
of the two first bins ($z=0.20-0.32$):
$\langle M_*/L_V\rangle =2.2$ [$N=7$, $\langle L_V\rangle =7\times 10^{10}$
L$_\odot $, $\langle (B-V)\rangle =0.94$], that is,
a mass--luminosity ratio 4.4 times larger. Assuming a variation of the luminosity
linearly dependent on the time, the variation of this ratio would be by a
factor 5.8 between z=0.1 (the average redshift of SDSS galaxies, which
are the reference of the size calibration in Shen {\it et al.}\cite{She03}) and $z=3.2$.
This is an acceptable estimation for elliptical galaxies.
Kauffmann {\it et al.}\cite{Kau03}(Fig. 13) showed that SDSS late-type galaxies have
a mass--luminosity ratio around 2 times lower than early-type ones,
and Rudnick {\it et al.}\cite{Rud06}(Fig. 9) showed that blue galaxies at $z\approx 3$
also have mass--luminosity ratios around 0.2-0.3, so
we also keep this number of $\approx 6$ for disk galaxies.
This factor of 6 might be affected by important errors (see \S \ref{.statothers}),
and could be much lower, though not much larger.
Given that we are comparing galaxies with the same luminosity,
the galaxies at $z=3.2$ would have 6 times lower stellar mass than at $z=0.1$.
Hence, from Eqs. (17)-(18) of Shen {\it et al.}\cite{She03}, we find that galaxies at
$z=3.2$ should be 2.7 times smaller than at $z=0.1$ if they are elliptical,
or 2.0 if they are disk galaxies.
These factors are much lower than the measured values of 6--10 and 4.5--6
respectively. A factor in size 2--4 for elliptical or 2--3 for disk galaxies,
including this luminosity evolution correction
remains (in rough agreement with the results by Trujillo {\it et al.}\cite{Tru06}).
Therefore, the argument of ``younger population in higher redshift''
does not serve to justify the present data. In order to explain
the observed size increase in terms only of luminosity evolution of
the stellar population, we would need to set $M_*/L_V$
with respect to the SDSS galaxies a factor 25--60 for elliptical
galaxies, or 50--100 for disk galaxies, too much!
\subsection{Mergers}
\label{.mergers}
If the luminosity density is not due
to an increase of the luminosity of each star, it must be due to
a redistribution of the mass density to make it more compact.
Why? Explanations in terms of mergers proliferate in
the literature (see discussion by Refs. \cite{Tru06,Tru07} and references
therein). Refs. \cite{Tru06,Tru07}
support the merger scenario calculating the size vs. mass ratio and
observing how it also decreases with redshift\footnote{In
Refs. \cite{Tru06,Rud06}, the masses were estimated
from the colors of the galaxies assuming solar metallicity
for the determination of the mass--light ratio and a
Salpeter IMF model. Trujillo {\it et al.}\cite{Tru07}
multiply the masses by a factor 0.5 to correct for
the difference of Kroupa and Salpeter IMF, which is only
a very rough approximation. The masses of Trujillo{\it et al.}\cite{Tru07} are more
accurately calculated, with an uncertainty of a factor two.
All these assumptions introduce significant errors into the calculation
of the mass--light ratio. Therefore, we must take these results with care.},
but this calculation also depends on the cosmological model used
and can give different results for different cosmologies.
Furthermore, Khochfar {\it et al.}\cite{Kho06} point out that the presence
of higher amounts of cold gas at high redshift mergers of ellipticals
also produces size evolution.
Trujillo {\it et al.}\cite{Tru07} cites the paper of Boylan-Kolchin {\it et
al.}\cite{Boy06} as a possible powerful mechanism to increase the radius
in massive elliptical galaxies.
Each merger would follow a law $R\propto M_*^{-\alpha }$ \cite{Boy06}
with $\alpha =1.0$ for a pericenter
of 15 kpc and lower for higher pericenter distances or higher
for lower pericenter distance. So when
two galaxies of the same luminosity approach each other reaching
a minimum distance of 15 kpc and merge,
the total radius will be $\approx 2.00$ times the radius of each individual
galaxy, while Shen {\it et al.}\cite{She03} for local galaxies
give $\alpha =0.56$: a factor $\approx 1.47$ in radius on average
when we double the mass.
This means that each major merger (fusion of galaxies of the same luminosity)
gives an extra factor of 1.36 for ellipticals
in angular size with respect to eq. (\ref{shen2}) relationship.
We would need an average of 2.3--3.6
($=\frac{ln(2-3)}{ln(1.36)}$, i.e. $1.36^{2.3-3.6}=2-3$)
major mergers along the life of each elliptical galaxy to justify
a factor 2--3 in radius.
The observed rate of mergers is indeed much lower than the necessary rate
to justify these numbers. Lin {\it et al.}\cite{Lin04}
with statistics of close galaxy pairs ($r<20 h^{-1}$kpc)
up to z=1.2 show that only $\sim 9$\% of the luminous galaxies
($-21<M_B<-19$) would have
a major merger during their lives since $z=1.2$.
De Propis {\it et al.}\cite{DeP07} get similar merger ratios for local galaxies
($z<0.12$) than Lin {\it et al.}\cite{Lin04} for $0.5<z<1$.
The number of cumulative mergers increases up to 22\%
\cite{Lin08} if we allow larger pair separation ($r<30 h^{-1}$kpc)
and up to 54\% \cite{Lin08} if we allow
a broader definition of major merger including pairs of galaxies
with mass ratio up to four.
Ryan {\it et al.}\cite{Rya08} calculated a number of 42\% galaxies
undergoing some merger up to $z=1$ for
more luminous galaxies ($M_B>-20.5$) [and 62\% for all $z$],
also for mass ratios up to four, and $r<20 h^{-1}$ kpc.
In our case, comparison with major mergers of equal mass
galaxies and with average pericenter around 15 kpc should be made,
so a number of 10--20\% (up to $z=3.2$ is $\sim 50$\% larger than up to
$z=1$) of mergers would be the amount
to compare with the 2--4 mergers (200--400\%) we need for the size evolution.
Or we may account for mergers with mass ratios up to 4 and pericenter
radii larger than 15 kpc on average, although these
would not produce an increase of an extra factor of 1.36 for ellipticals
in angular size with respect to the Eq. (\ref{shen2}) relationship but
lower. Therefore, we would have to compare the number of $\sim 50$\%
with the large number of mergers of this kind to produce the
observed evolution in radius ($\sim 10$ mergers per galaxy;
probability of merging $\sim 1000$\%).
Moreover, I suspect that the number of mergers is
overestimated. First because not all galaxy pairs
become mergers. Second, because of the contamination of
interlopers. There is an uncertainty of radial distance due
to proper motion of the galaxies apart from the Hubble flow,
and many of the identified pairs of galaxies in the projected
sky are not real pairs in 3D space. There may be a chance superposition
in the line of sight of galaxies. The situation is worst
for Ryan {\it et al.}\cite{Rya08}, who use
spectrophotometric redshifts; and interlopers do not introduce
a statistical error, as they say, but a systematic one.
However, the order of magnitude of the pairs with or
without interlopers should be more or less the same,
the interlopers being at $z<3$ lower than $\sim 30$\% \cite{Ber06}.
Also, De Ravel {\it et al.}\cite{DeR08}, for instance, using
spectroscopically confirmed pairs, get similar merger ratios.
Third, because timescales of mergers increase slightly with redshift
and are longer than assumed in most observational studies\cite{Kit08}.
About the CAS method of estimation of merger ratios based
on the identification of major mergers with highly asymmetric
galaxies\cite{Con00,Con08} two things may be commented:
1) The authors consider that all asymmetries in principle associated
with starbursts are due to major mergers, but the mechanisms
which trigger important amounts of star formation might be
different from major (ratio 1:1) mergers; in particular,
minor mergers may also trigger asymmetric star formation.
Strongly disturbed systems, indicative of recent strong
interactions and mergers, account for only a small fraction
of the total star formation rate density\cite{Jog08}.
2) Interlopers, galaxies and stars
with very different redshifts projected as background
or foreground objects in the line of sight
produce an important amount of apparent distortion in the galaxies,
especially at high redshift, where Conselice {\it et al.}\cite{Con08} claim
that a high fraction of galaxies are major mergers. We must bear in mind
that Hubble images may detect very faint galaxies and
there are more than 5 million galaxies per square
degree with $m_z<30$ (\cite{Ell07}; extrapolated from
the counts up to magnitude 28),
one galaxy in each square of $1.6''\times 1.6''$ on average,
which, mixed with the main galaxy ($m_z<27$ in Conselice {\it et al.}\cite{Con08}),
produce apparently distorted galaxies.
Also, De Propis {\it et al.}\cite{DeP07} checked that the contamination is
happening in low redshift galaxies with foreground stars.
Moreover, we see the galaxies once they are
``presumably'' merged but we do not see enough galaxy pairs at $1<z<2$
when both galaxies of the merger are separated (e.g., according to
the already overestimated ratio by Ryan {\it et al.}\cite{Rya08}). This
indicates at least that the asymmetries are not produced by major
mergers but by minor mergers (ratio of masses larger than 4) or other
effects.
Stellar population analyses do not even agree with these merger rates.
Mergers or captures of smaller galaxies can occasionally occur,
but hierarchical-scheme
subunits fusing together and made of gas and stars is not the dominant one by which
massive elliptical galaxies are made, at least for $z<2$
\cite{Chi02}. Most massive elliptical galaxies
have a passive evolution since their creation according to stellar population
analyses\cite{Chi02}. Mergers are beautiful, spectacular
events, but not the dominant mechanism by which elliptical galaxies
are assembled.
Most early-type galaxies with a velocity dispersion exceeding 200 km/s
formed more than 90\% of their current stellar mass at redshift $z>2.5$
\cite{Jim07}.
Elliptical galaxies formed in a process similar to monolithic
collapse, even though their structural and dynamical properties are
compatible with a small number of dry mergers\cite{Cio07},
far from the number of mergers necessary in our case.
Dry mergers do not decrease the galaxy stellar-mass surface
density enough to explain the observed size evolution\cite{Nip09}, and the high density
of the high-z elliptical galaxies does not allow them to evolve into present-day
elliptical galaxies\cite{Nip09}.
Mergers are searched for in the Local Group galaxies too.
In the case of the Milky Way, some authors try
to find evidence of major fusion events of big galaxies,
but up to now we do not see evidence in favour but against such scenarios\cite{Ham07}.
There are minor mergers, of course, and
absorption of small clouds of the intergalactic medium, but
the presence of intermediate mass galaxies
at short distances from the center, at present or
in the past, have yet to be identified. There are certain attempts
to find something, for instance the recent discovery of a galaxy
with a relatively large diameter at only 13 kpc from the
Galactic center called Canis Major, but
that discovery resulted in a fiasco\cite{Lop07}.
In any case, two to four major mergers on average per galaxy is too much.
Another argument against the merger scenario of the hierarchical
CDM cosmology is that galaxy formation is controlled
by a single parameter\cite{Dis08}.
One would expect in the merger scenario that the properties of individual
galaxies be determined by a number of factors related to the star
formation history, merger history (masses, spins and gas content of
the individual merging galaxies), etc., but that is not so; all the
different parameters of the galaxies are correlated\cite{Dis08} and
there is only one single independent parameter based on their mass.
On the other hand, major mergers of disk galaxies of comparable
mass should give place to elliptical galaxies, so it is not easy to understand
in this scenario how the radius of disk galaxies grows.
Conselice {\it et al.}\cite{Con05} show in fact that there is little to no evolution
for disk galaxies at $z<1.2$, for the K-band, in the stellar-mass
Tully--Fisher relation, and in the ratio of stellar/total mass.
Ferguson {\it et al.}\cite{Fer01} also
find that accretion flows play only a minor role in determining the evolution
of the disk scalelength. In models in which the main infall phase precedes
the onset of star formation and viscous evolution, they find the exponential
scalelength to be rather invariant with time. On the other hand,
models in which star formation/viscous evolution and infall occur
concurrently result in a smoothly increasing scalelength with time,
reflecting the mean angular momentum of material which has fallen in at
any given epoch.
Furthermore, selection effects go apparently in the opposite
direction of observing pre-merger galaxies at high redshift.
At very high redshift, we are observing only galaxies with
stellar masses over $10^{11}$ M$_\odot $ and some of them
over $10^{12}$ M$_\odot $ \cite{Tru06}.
And at low redshifts there are galaxies of all masses but the average
stellar mass is much lower than that.
Thinking that very massive galaxies
at high $z$ are building blocks of even more massive low $z$ galaxies
is counterintuitive. After 2--4 mergers of equal mass
galaxies in average, the galaxies should be
4--16 times more massive than the original
building blocks at high redshift. We should be observing some galaxies at
low redshift with stellar masses of $\sim 10^{13}$ M$_\odot $.
With a mass-luminosity ratio of $M_*/L_V=6$
(for a very old population \cite{Vaz96};
if it were younger than 12 Gyr it would be lower so the
luminosity would be even higher),
this would mean galaxies with $L_V=2\times 10^{12}$ L$_{\odot ,V}$,
or absolute magnitude $M_V=-26$. Even cD galaxies in the centers of
the clusters are not as bright as that. Where are these galaxies, then?
Thinking that low- to intermediate-mass galaxies are
the final stage of major merger processes is a reasonable possibility,
since we cannot see their building blocks at high redshift
(they are very faint). However, for only high luminous
galaxies I get more or less the same shrinking factors at high-z
with respect to low-z (see Fig. \ref{Fig:angsizelum}).
It is not a question of some merging which affects low
luminosity galaxies more. I could even concentrate our analysis on galaxies with
$L_V>1.02\times 10^{11}$ L$_{\odot ,V}$ and, in spite of the poorer statistic,
a very strong size evolution between galaxies
at $z=0.5-1$ and $z>2$ can be appreciated. Refs. \cite{Tru07,Bui08}
even get higher evolution for higher stellar masses, but
this has been criticized\cite{Man09}.
Another element that is not consistent with these hierarchical merging scenarios
is that superdense massive galaxies should be common in the early
universe ($z>1.5$), and a non-negligible fraction (1-10\%) of them
should have survived since that epoch without any merging process
retaining their compactness and presenting old stellar populations in
the present universe. However, Trujillo {\it et al.}\cite{Tru09} find only a tiny
fraction of galaxies ($\sim 0.03$\%) of these superdense massive galaxies
in the local Universe ($z<0.2$) and they are relatively young ($\sim 2$ Gyr)
and metal-rich ($[Z/H]\sim 0.2$). Clearly a case of how some authors
(Trujillo {\it et al.}\cite{Tru09})
try to find proofs in favour of the hierarchical merging
scenario, and when they find that the observations point out exactly
the opposite thing of what is expected, instead of claiming that the
hierarchical merging scenario is wrong, they try to deviate attention by
giving less importance to the observed facts for the validation of
the standard theory.
\subsection{Quasar feedback}
Fan {\it et al.}\cite{Fan08} realized that the evolution+merger luminosity
solution is not enough to explain the strong size evolution and they
claimed, ``no convincing mechanism able to account for such
size evolution has been proposed so far''. Nonetheless,
they have proposed a new mechanism to explain this extraordinary
size evolution. Fan {\it et al.}\cite{Fan08} propose that in elliptical galaxies it
is directly related to a quasar feedback:
part of the energy released by the QSO would be spent to produce
outflows of huge amounts of cold gas expelled from the central regions,
a rapid (few tens of Myr) mass loss which induces an expansion of
the stellar distribution.
Although this mechanism might explain
some small part of the expansion of elliptical galaxies,
there are some aspects in this hypothesis which do not fit
with the observed facts well, at least while we do not clarify some points.
First, we do not see such supermassive outflows which are necessary to
maintain the Fan {\it et al.}\cite{Fan08} idea, although since
their life is very short only a few high redshift
QSOs would show it. Second, elliptical galaxies expend their gas to
produce stellar formation which gives rise to their stellar mass.
If a QSO swept the gas away, no young stellar populations
would be observed, but there is now compelling evidence for a
significant post-starburst population in many luminous AGN\cite{Ho05}.
There is also detection of large amounts of warm, extended
molecular gas indicating that QSOs have vigorous star formation\cite{Wal07},
and that the gas is not being expelled.
The blue color of host galaxies, $(B-V)_{rest}\sim 0$ \cite{Sch08},
indicates a young population too.
Third, if QSOs produced such massive outflows ejecting
the gas of the galaxies, this would also apply to disk galaxies.
Around 40\% of host galaxies in QSOs are disk galaxies\cite{Guy06},
and it is clear that disk galaxies still have gas
and active star formation in their disks.
Fourth, a continuous increase in the average size of ellipticals,
as shown in Fig. \ref{Fig:angsizetype}/left, would require the continuous
expulsion of gas, but most massive elliptical galaxies
have had a passive evolution since their creation at $z>2$
according to stellar population analyses\cite{Chi02}, which
indicate that the gas was already drained in them at $z>2$.
Fifth, for an average increase of a factor 3 in size in the elliptical
galaxies, we would need to assume that most elliptical galaxies have
hosted very luminous QSOs during their lives. With a minimum QSO lifetime of
around 40 Myr, as required for the massive outflow mechanism of
Fan {\it et al.}\cite{Fan08}, the number of very luminous QSOs should be
around 1/300 of the number of elliptical galaxies.
This number is too large, given that the density of QSOs
with $L_{bol}>\sim 10^{48}$ erg/s (the necessary luminosity,
5\% of which is spent to produce outflows larger than 1000 M$_\odot /yr$, as posited by
Fan {\it et al.}\cite{Fan08}; $L\sim \frac{M\dot{M}}{0.05R}$ with
$M>2\times 10^{10}$ M$_\odot $ and $R=\frac{1}{3}R_{SDSS}$)
is $\sim 10^{-8}$ Mpc$^{-3}$ \cite{Hop07}.
Nonetheless, I would not dare to say that Fan {\it et al.}'s\cite{Fan08} hypothesis
is incorrect. I think it is an elegant and interesting solution to the problem,
and it must be considered as a serious possibility, provided it is
able to solve their caveats, and give some support to the scenario
with some observations directly interpretable as massive outflows.
\subsection{Rotation or dispersion velocity analysis}
If this hypothesis of lower radius at high redshift
for a given mass were true, due either to mergers or quasar feedback,
we would expect a significant increase in the
rotation speed or dispersion velocity
of those galaxies at high redshift with respect to the
local ones. For a constant mass within a radius $R$, one would roughly
expect rotation speed in a galaxy
$v_{rot}\propto R^{-1/2}$, and something similar for the
dispersion velocities in ellipticals.
Let us analyze whether this change of velocity is taking place.
An analysis of Marinoni {\it et al.}'s \cite{Mar08b}
data with galaxies with $z<1.2$ does not show (see Fig. \ref{Fig:marinoni})
a significant change in the rotation velocity--size relationship or
the rotation velocity--absolute
magnitude relationship. According to the Saintonge {\it et al.}\cite{Sai08} analysis for
low redshift galaxies, the rotation velocity is proportional
to $R^{0.98\pm 0.01}$, and is also related to the absolute magnitude in the
I-band by their eq. (3), so we could derive an average expected
velocity from the luminosity of the galaxy, $v(M_i)$. In Fig.
\ref{Fig:marinoni}, I see that these relationships remain more or less
constant: the variations in $v_{rot}/R^{0.98}$ and $v_{rot}/v(M_i)$
are compatible within 1-$\sigma $ to be null. Particularly,
the best linear fits give:
\begin{equation}
\frac{v_{rot}}{R^{0.98}}=(26.4\pm 5.9)+(8.9\pm 8.8)z
\label{vmar1}
,\end{equation}
\begin{equation}
\frac{v_{rot}}{v(M_i)}=(1.04\pm 0.20)-(0.09\pm 0.16)z
\label{vmar2}
.\end{equation}
Within the error bars, I cannot exclude an increase
in these ratios with redshift compatible with the hypothesis
of radius decrease. An interesting test would be to measure the
rotational velocity in some of the very compact galaxies with
redshift 3. Although getting a spectrum of these faint galaxies
is technically difficult, this would give a proof of whether either they
are really so compact or the cosmological parameters are wrong.
\begin{figure*}
\vspace{1cm}
{\par\centering \resizebox*{5cm}{5cm}{\includegraphics{fig5a.eps}}
\hspace{1cm}\resizebox*{5cm}{5cm}{\includegraphics{fig5b.eps}}\par}
\caption{}
Plot of $v_{rot}/R^{0.98}$ and $v_{rot}/v(M_i)$ for the 39
galaxies of the sample from Marinoni {\it et al.}\cite{Mar08b}. Squares with error
bars are the average of the data (asterisks) with steps of $\delta z=0.2$.
The solid line is the best linear fit, given by eqs. (\ref{vmar1})
and (\protect{\ref{vmar2}}).
\label{Fig:marinoni}
\end{figure*}
The critical assumption of a variable effective
radius is also counter-argued by the proofs in favour of a constant
radius\cite{van98} showing that high redshift first-rank
elliptical galaxies, with similar absolute magnitudes,
have the same velocity dispersions as low redshift
first-rank elliptical galaxies.
Van Dokkum {\it et al.} \cite{van98} point out, however, that
these galaxies are predicted by galaxy formation models to be those
whose formation finished at very high redshift and so it would not be
surprising that these galaxies had the same radius as at low
redshift, specifically because they are in clusters where mergers
may not be likely.
A surprising new result also points in this direction\cite{Cen09}:
the velocity dispersion of giant elliptical galaxies
with average redshift $z\approx 1.7$ from Cimatti {\it et al.}'s \cite{Cim08} sample is
similar or very slightly larger than the dispersion for the same kind
of galaxies with the same stellar mass in the local Universe at
240 km/s, while at $z=0$ it is around 180 km/s. Since
Cimatti {\it et al.}\cite{Cim08} galaxies at $z\approx 1.7$
are more compact than the average size with that luminosity at that redshift,
it is normal to have a slightly higher velocity dispersion.
It should be much higher (over 400 Km/s) at high redshift since a
much lower radius is attributed to them \cite{Cim08}, but it is not.
This result points directly to the conclusion that the galaxies have
not strongly changed their radii. Other alternative ad hoc ideas in terms of
a conspiracy of effects in which the dark matter ratio has increased
the amount necessary to compensate for the radius increase\cite{Cen09}
sound like a queer coincidence,
and have no clear basis in terms of galaxy formation scenarios.
\subsection{Discussion on expansion+evolution models}
All these considerations may make us think that the concordance model
cannot explain the present data. The other expanding models present
similar problems: a factor in average size evolution up to $z=3.2$ equal
to 5.8 for Einstein--de Sitter, 5.6 for Friedmann
with $\Omega _m=0.3$, and 5.9 for the QSSC model.
Phenomenologically, it is possible
to fit the data to any of the expanding models with appropriate evolution of
galaxies. In practice, this evolution for a constant
luminosity galaxy should be so strong (up to
a factor 200 in average in the luminosity density up to $z\approx 3.2$;
systematic errors may change up to a factor 2 this number, but
this does not change the situation)
that the explanations for it seem unrealistic.
The situation for the expanding models becomes even more dramatic
if we go to higher redshift. At redshift 6, the linear
size of the galaxies assuming a concordance model is even lower,
approximately a factor two lower than at $z=3.2$ (\cite{Bou04} Fig. 4).
If we were going to consider a reduction in size by
a factor 12, all the arguments given in this section would become
even stronger.
\section{Analysis of the static universe cases}
In the first two static Universe cases,
there is an excess of size ($\sim 20-30$\%)
for most redshifts.
A possible interpretation of this discrepancy is that there
may be a systematic error in the calculation of the ratio between measured
and predicted average $\theta _*$. In \S \ref{.select}, I have
discussed the possible systematic errors and I concluded that
they should be lower than $\sim 30-40$\%.
These systematic errors might be enough to justify the departures
of the data with respect to the prediction in Fig. \ref{Fig:angsize}
for these static models.
The third case of plasma redshift is much more discrepant
and can only fit the data with a significant evolution of galaxies, although
less strong than in the expanding models: a factor 3 instead of a factor 6 in the
concordance model for a given luminosity from $z=0$ to $z=3.2$.
\subsection{Including extinction}
\label{.extinc}
Apart from the systematic error considerations,
another solution to make the two first static cosmological models
compatible with our data would be related to extinction
rather than the evolution.
Extinction would make the galaxies look fainter,
which means that, through Eqs. (\ref{thetaequiv}) and
(\ref{shen2}), when the corrections
of extinctions are made, the luminosity is larger and
$\theta _*$ is smaller than their values without
corrections. Let us check this hypothesis with a rough calculation.
Instead of Eq. (\ref{lrest}), the inclusion of the
IGM extinction with absorption
coefficient $\kappa $ (area per unit mass) will give the following
relationship
\begin{equation}
L_{V,rest}=4\pi F_{V,rest}d_L^2e^{\rho _{dust}\int _0^{d_A}dr\ \kappa
[\lambda _V\frac{1+z(d_A)}{1+z(r)}]}
.\end{equation}
I have assumed a constant dust density $\rho _{dust}$ along the line
of sight, which is an appropriate approximation for a homogeneous
Universe without expansion and moderate amounts of dust ejection by the galaxy. If
we considered an expanding Universe with a strong dust emission rate
by the galaxies, we should include $\rho _{dust}(r)$ within the integral,
but it is not the case here.
The absorption coefficient can approximately be described with a
wavelength dependence:
\begin{equation}
\kappa (\lambda )=\kappa (\lambda _V)\left(\frac{\lambda }{\lambda _V}
\right)^{-\alpha }
.\end{equation}
Hence, and using Eqs. (\ref{angdistst}) and (\ref{angdisttl}),
\begin{equation}
L_{V,rest}=4\pi F_{V,rest}d_L^2
e^{\frac{c\ a_V}
{H_0(\alpha +m)}[(1+z)^m-(1+z)^{-\alpha}]}
,\end{equation}
with $m=1$ for the cosmology with linear Hubble law, and
$m=0$ for the tired light case. $a_V\equiv \kappa (\lambda _V)\rho _{dust}$
is the absorption in V per unit length, which means
there are $1.086a_V$ magnitudes in V of extinction per unit length.
The value of the exponent
$\alpha $ is not totally independent of $\lambda $ but I take it
approximately as constant. I adopt $\alpha =2$, as observed
in near-infrared bands in our Galaxy\cite{Nis06}.
For lower wavelengths (optical, ultraviolet) $\alpha $ would be lower.
Since most of the sources have the
wavelength equivalent to V-rest in the near-infrared, and
the extinction curve of dust in the intervening QSO absorbers
resembles the SMC extinction curve\cite{Kha05},
this approximation is reasonable.
The absolute value of $a_V$ is not well known and neither do we
know whether it is significant or null. There are only
some constraints for the maximum value (e.g., \cite{Ino04},
although based on standard cosmology).
The values of $a_V$ which give the best fit
to our data are: $a_V=1.6\times 10^{-4}$ Mpc$^{-1}$ for the linear
Hubble law case, and $a_V=3.4\times 10^{-4}$ Mpc$^{-1}$ for the
simple tired light case. Assuming
$\kappa (\lambda _V)\sim 10^5$ cm$^2$/gr \cite{Wic96},
the value for the dust density necessary to produce such an extinction
would be $\rho _{dust}\sim 6\times 10^{-34}$ g/cm$^3$,
and $\rho _{dust}\sim 1.2\times 10^{-33}$ g/cm$^3$ respectively,
which is within the range of possible values. Vishwakarma\cite{Vis02} gives values of
$\rho _{dust}=3-5\times 10^{-34}$ g/cm$^3$, but for the QSSC model.
Inoue \& Kamaya\cite{Ino04} allow values as high as $\rho _{dust}\sim 10^{-33}$
g/cm$^3$ for the high z IGM within the standard concordance cosmology.
For comparison, the average baryonic density of
the Universe is (taking $\Omega _b=0.042$; \cite{Spe07})
$\rho _b=3.9\times 10^{-31}$ g/cm$^3$, so this would mean that
IGM dust constitutes 0.15 or 0.30\% of the total baryonic matter, reasonable
amounts.
Whether the extinction used in the models with extinction would be grey or would
introduce some small reddening is not totally clear, but this is not
a question for the present analysis. I just note that reddening in the optical would
depend on the value of $\alpha $ in the visible at intermediate to high redshift
and the variability of $\kappa (\lambda )$ with respect to $\lambda $ in the UV.
The features of dust extinction in the UV are not easy to model in an IGM with
unknown composition.
With this simple correction for extinction,
the results are significantly improved,
as shown in Fig. \ref{Fig:angsizeext}. The tired light case
gives a better fit.
\begin{figure*}
\vspace{1cm}
{\par\centering \resizebox*{5cm}{5cm}{\includegraphics{fig7a.eps}}
\hspace{1cm}\resizebox*{5cm}{5cm}{\includegraphics{fig7b.eps}}\par}
\caption{Log--log plot of the average of $\log _{10} \theta _*$, where
$\theta _*$ is the equivalent angular size, vs. redshift ($z$).
Bins of $\Delta \log _{10}(z)=0.10$.
The two plots are for the first two static cosmologies with the inclusion
of a constant IGM extinction which gives the best fit.}
\label{Fig:angsizeext}
\end{figure*}
\subsection{Comparison with angular size test for ultra-compact radio
sources}
\label{.ultracom}
Compact radio sources have been used by several authors to carry out
the angular size test because these sources were thought to be
free of evolutionary effects. However, the different results obtained
with these sources has raised the suspicion that they may not be
such good standard rods. Apparently, these rods are somewhat flexible.
For example, Kellermann\cite{Kel93} claimed that the angular size test
for these sources fitted Einstein--de Sitter expectations very well,
when Einstein--de Sitter was the fashionable model.
Jackson \& Dodgson\cite{Jac97} claimed the opposite: that it was not compatible with
Einstein--de Sitter, and that, given that $\Omega _m=0.2$,
the best fit for the cosmological constant term was
$\Omega _\Lambda =-3.0$; flat cosmological models were excluded
with $>70$\% C.L. Jackson\cite{Jac04},
in the era of the concordance model as the fashionable cosmology,
again carried out the analysis of the same data used
by Jackson \& Dodgson\cite{Jac97},
doing some new corrections due to selection effects and bias,
and they get the best fit for $\Omega _m=0.29$, $\Omega _\Lambda =0.37$,
compatible within 1-$\sigma $ with the concordance model.
With further data, Jackson \& Jannetta\cite{Jac06} get the
best fit for $\Omega _m = 0.25^{+0.04}_{-0.03}$,
$\Omega _\Lambda = 0.97^{+0.09}_{-0.13}$ (68\% C.L.).
It seems that the general trend is to obtain the result
expected from fashionable cosmologies on the date in which the test is
carried out, and when incompatibilities appear, some selections
effects, biases, small evolution effects are sought to try to make the
results compatible. In my opinion, this is not a very
objective way to do science, but let us leave the discussion of the
methodology of cosmology aside.
One important selection effect is derived from the fact that
linear sizes depend on radio luminosities. Jackson\cite{Jac04} tries to
take this effect into account and suggests a method of correcting
it: binning the data into groups of 42 points in the redshift distribution
and taking as representative of each bin the mean of points between
11 to 17 within each one, counting from the smallest objects.
This is supposed to be done to compensate for the dependence
$R_{rad}\propto L_{rad}^{-1/3}$ for a given redshift
and for the fact that the lowest luminosity points
cannot be observed at high redshift. In my opinion, this is
not the right way to correct the selection effect. Jackson\cite{Jac04}
is just doing a median which gives more weight to smaller objects,
but this median is shifted at high redshift by the lack of low
luminosity objects.
Nevertheless, my concern is not about Jackson's
method of correcting for the Malmquist bias, but about the
relationship between radius and luminosity. The relationship in Fig. 1
of Jackson\cite{Jac04} is applicable only to the concordance model
and it will be far different for static Universes. Particularly
in Figs. 1 and 2 of Jackson\cite{Jac04} we see that the linear size is
almost independent of redshift between $z=0.5$ and $z=4$: around
10 pc, with some scattering due in part to the range of luminosities for
each redshift. However, if I use eq. (\ref{angdistst}) of the static
Universe with linear Hubble law instead
of eq. (\ref{concordance}) of the concordance model
to calculate the angular size distances, I see that the
linear sizes at $z\approx 3.5$ should be a factor of 10 larger
($\sim 100$ pc instead of $\sim 10$ pc). Therefore, the linear
sizes would not be independent of redshift but highly dependent
on it. In such a case, Figs. 1 and 2 of Jackson\cite{Jac04} would be
transformed into a plot with a continuous increase in linear size with
radio luminosity for all ranges of luminosity for all redshifts:
roughly $R_{rad}\propto L_{rad}^{1/2}$.
Since at high redshift we cannot observe
objects with low $L_{rad}$, this means that we would be losing
objects of low $R_{rad}$ at high $z$, so
the median of $\theta $ would be highly overestimated with respect
to low to intermediate redshift objects.
This would explain why the
angular size test for ultra-compact radio sources does not
give a $z^{-1}$ dependence. Therefore, a static Universe is not
excluded by this test unless we demonstrate with data at low redshift
that the radii of the ultra-compact sources does not depend on their
luminosities.
Ultra-compact objects could be
used to carry out a right angular size test, but we cannot
directly compare objects of low luminosity at low redshift
with objects of high luminosity at high redshift
because we have no guarantee that the linear size is the
same in both cases (and it is not valid to assume a cosmological
model a priori to prove that it is good a standard rod,
because the method should be independent of any cosmological
assumption if we want to derive from it which is the best cosmological
model). We should either i) compare objects of the same luminosity
(different for each cosmology), or ii) define a $\theta _*$ as in
the present paper in which we need to calibrate the size--luminosity
relationship in the low-z Universe. This second option has the
caveat that we do not have very high luminosity compact radio sources
at low redshifts so we need to extrapolate the local radius--luminosity
relationship for high luminosities.
\subsection{Hubble diagram for the different cosmologies}
\label{.hubble}
In Fig. \ref{Fig:hubble}, I show the different distance moduli
for the different cosmologies without extinction, together with some real
data of SNe Ia compiled by Kowalski {\it et al.}\cite{Kow08}:
\begin{equation}
m_{V,rest}-M_{V,rest}=5\log _{10}[d_L(z)({\rm Mpc})]+25
.\end{equation}
In the first two static models, if we wanted to include the extinction, we should sum
$A_{V,rest}(z)=\frac{1.086c\ a_V}
{H_0(\alpha +m)}[(1+z)^m-(1+z)^{-\alpha}]$
to the distance modulus.
One aspect is remarkable: the value of the distance modulus for
the concordance model is very similar to its value for the static model
with a linear Hubble law, and it can be seen in Fig. \ref{Fig:hubble} how
the data of SNe Ia are approximately compatible with this scenario.
The fit for the concordance model over the data gives a reduced $\chi ^2$:
$\chi _r^2=3.34$;
while for the static model with linear Hubble law $\chi _r^2=4.20$, slightly
worst but not by much: the concordance model reduces the $\chi ^2$ by only 20\%.
Lerner\cite{Ler09} also show this by comparing the residuals of the Hubble diagram in the
concordance model and in the static universe ansatz and realizing that they are
both similar.
Is it a coincidence\footnote{The degree of coincidence depends on
the maximum redshift, $z_{max}$, we use. For instance,
the value of $\Omega _\Lambda$ in a flat Universe which
gives a best minimum square fit to
the Hubble diagram for a static model with linear Hubble law and
without extinction is $\Omega _\Lambda \approx
0.39+0.108z_{max}-0.0085z_{max}^2$.
For $z_{max}=6$, as in our plot of Fig. \ref{Fig:hubble}, it gives
$\Omega _\Lambda=0.74$, but it falls to 0.55 if we only consider
it up to redshift 1.7, as usually for SNe Ia data.
If we set $\Omega _m=0.3$ and we searched for the best value
of $\Omega _\Lambda $ with any value of $\Omega _{total}$, we would
get the best fit for $\Omega _\Lambda \approx
0.76+0.171z_{max}-0.053z_{max}^2+0.0038z_{max}^3$. For $z_{max}=6$,
it gives the best fit for $\Omega _\Lambda=0.70$,
and it increases to 0.91 if we only consider
it up to redshift 1.7.}? With a slight extinction of
$a_V\sim 1-2\times 10^{-4}$ Mpc$^{-3}$ (the range of the best fit
obtained in \S \ref{.extinc}), the agreement is also quite conspicuous
at least for $z<1$. For the
simple tired light model, only with extinction of the order
$a_V\sim 5\times 10^{-4}$ Mpc$^{-3}$ (not very far from the best fit
obtained in \S \ref{.extinc}) would get the coincidence for $z<1$.
For the plasma redshift tired light model without extinction,
the agreement with SNe Ia for $z<1$ is also acceptable, as noted by
Brynjolfsson\cite{Bry04a}.
This means that, with the static models,
we can fit nearly the same Hubble diagrams as
the concordance model with its cosmological constant, particularly
for supernovae fits. It is not the
place here to extend the discussion on
the analysis of the compatibility of the static model with a linear Hubble law and
the supernovae Ia data; this would require a discussion of the systematic
errors, the selection effects, etc. At present,
I just want to emphasize that there are no major problems to make compatible
the static model of linear Hubble law with SNe Ia data.
\begin{figure*}
\vspace{1cm}
{\par\centering \resizebox*{5cm}{5cm}{\includegraphics{fig9a.eps}}
\hspace{1cm}\resizebox*{5cm}{5cm}{\includegraphics{fig9b.eps}}\par}
\vspace{1cm}
{\par\centering \resizebox*{5cm}{5cm}{\includegraphics{hubble.eps}}
\hspace{1cm}\resizebox*{5cm}{5cm}{\includegraphics{hubble_zoom.eps}}\par}
\vspace{1cm}
\caption{}
Distance modulus as a function of redshift for different
cosmological models. Data of supernovae Ia compiled by Kowalski {\it et al.}\cite{Kow08}
were added. Right plots are zooms of the left plots.
\label{Fig:hubble}
\end{figure*}
For other cosmologies, and without extinction,
the difference from the concordance model is larger,
especially for the highest redshifts. This does not mean that they are discarded,
because the objects used as standard candles (particularly supernovae)
might have an absolute magnitude which is not constant with redshift
or some extinction along their lines of sight.
Several authors\cite{Dom00,Agu00,Goo02,Row02,Sha02,Pod06,Bal06,And06,Sch07} also
think that SNe Ia data used to derive $\Omega _\Lambda =0.7$
are affected by several systematic uncertainties that make the
$\Lambda $-CDM cosmology uncertain.
\subsection{Evolution, and other tests}
\label{.statothers}
Another topic to discuss is the apparent evolution of galaxies
at different redshift. I have not included
any evolution correction for the static models in the plots of
Fig. \ref{Fig:angsizeext} although
some slight evolution might be compatible with them,
in the sense that brighter
populations are present at higher redshifts.
Some evolution might be present because,
as discussed in \S \ref{.results}, elliptical galaxies
get lower angular sizes at high redshifts
than disk galaxies, possibly due to the different mean ages
of their populations or merger rate; although some of these differences
might be due to systematic errors too, as said in \S \ref{.select}.
Also, the most massive galaxies present a higher ratio of
angular size with respect to local galaxies\cite{Tru07,van08,Bui08},
which may be interpreted here as
a real higher compactness with respect to the least massive galaxies.
Note, however, that, as said above, this extra-compactness can be understood in terms
of fluctuations due to noise preventing the recovery of the extended low surface
brightness halos in the light profile\cite{Man09}.
There is also a slight color at rest evolution\cite{Lop09} and a mass--luminosity ratio
evolution, as said in \S \ref{.brighter}; but these mass calculations are subject to important
errors depending on the synthesis model, IMF assumption, etc.; hence, there is a wide
range of possible values of mass--luminosity ratio evolution. The bluer (at rest) color of
high redshift galaxies might be due to bias, because bluer galaxies with
younger populations are brighter. We must also bear in mind that photometric errors are
larger at higher redshift (because of the fainter fluxes), the
photometric redshifts have higher uncertainties and consequently
might affect the determination of the luminosities at rest, etc.
At present, with the data used from Refs. \cite{Tru06,Rud06} at $z\ge 1$,
the correlation between $z$ and $(B-V)_{rest}$ is:
$\langle z(B-V)_{rest}\rangle - \langle z\rangle \langle (B-V)_{rest}\rangle
=-0.048\pm 0.021$ for elliptical galaxies and $-0.033\pm 0.015$
for disk galaxies. The variation in color with the variation in redshift is
much lower than the dispersion of colors and possible systematic effects.
Therefore, we cannot exclude the possibility that the luminosity evolution is small enough to be within the
error bars of our data with the static models.
Note however that the evolution of some quantities per unit comoving volume
[for instance, the star formation ratio, expressed in mass per unit
time per unit comoving volume, is claimed to be
significantly higher in the past\cite{Hop04}] must be corrected with
respect to the concordance cosmology with Eqs. (\ref{vol1}), (\ref{vol2}),
which reduces by a factor of 10 at $z=2$ the ratio in the static model with a linear
Hubble law with respect to the concordance one.
In any case, there may be a real evolution which should be explained, either
in expanding or static models. Note that ``static'' does not mean ``no evolution'';
``static'' means ``no expansion''; there is not necessarily a contradiction in observing
evolution in a static Universe, provided that the creation of galaxies is not
a continuous process.
Explaining the cosmic microwave background radiation (CMBR) and its anisotropies
is not the purpose of the present paper. Note, however, that there exist alternative
scenarios to explain them apart from the model of standard hot Big Bang
(e.g., \cite{Nar03,Nar07}; see the discussion in another paper of mine\cite{Lop08}, \S 1).
Concerning other arguments/tests in favor of the expansion (e.g.,
\cite{Lub01,Gol01,Mol02}), we must bear in mind that they
are usually a matter of discussion. Tests such as the time dilation
in SNe Ia, which were claimed to be a definitive proof of the expansion
of the Universe, find counterarguments and criticisms from opponents\cite{Bry04b,Lea06,Cra09},
who claim that a static Universe is compatible with the
data. The same thing happens with any other test, including
the present one of angular size. Apart from the present angular size test,
there are other tests that also present results in
favor of a static Universe and against an expanding Universe
(e.g., \cite{LaV86,Mol91,Jaa93,Tro93,Tro96,And01,And06,Ler06,Ler09}).
Perhaps the most immediate problem with the static Universe
is understanding the cause of the redshift of galaxies, but
there are several proposals for alternative mechanism to produce
redshifts without expansion or Doppler effect,
so the hypothesis of a static Universe
is not an impossible one. Other facts, such as the formation of
the large scale structure, the creation of the light elements,
etc., also provide alternative explanations different from the standard model.
Further discussion of all the questions raised in this paragraph are given in my
review\cite{Lop03}.
\subsubsection{UV surface brightness test}
\label{.uv}
The main discrepancy in the different tests such as angular size, surface brightness,
Hubble diagram, etc., is the evolution of galaxy luminosities,
which is very large for the defenders of the expansion and not so
large for the defenders of the static
models. Lerner\cite{Ler06} proposes a test of the evolution hypothesis
that is also useful in the present case. There is a limit on the UV surface
brightness of a galaxy, because when the surface density of hot bright stars and
thus supernovae increases large amounts of dust are produced to absorb
all the UV except that from a thin layer.
Further increase in surface density of hot bright stars beyond a given
point just produces more dust, and a thinner surface layer, not an
increase in UV surface brightness. Based on this principle, there
should be a maximum surface brightness in UV-rest wavelengths independent
of redshift. Scarpa {\it et al.}\cite{Sca07} measured in low redshift galaxies
a maximum FUV (1550 \AA \ at rest) emission of 18.5 mag$_{AB}$/arcsec$^2$
(the average is $24-25$ mag$_{AB}$/arcsec$^2$); no galaxy should
be brighter per unit angular area than that.
Using eqs. (\ref{sb}), (\ref{sb0}) with
the flux at wavelength 1550 \AA \ at rest\footnote{An
interpolation with the publicly available fluxes in filters U, B, V,
I$_{814}$ and J was used to get it. These fluxes are given
in flux per unit frequency so the dimming factor must be multiplied
by a factor $(1+z)$ with respect to the flux in the whole filter.
That is, $SB_0=SB(1+z)^3$ for the expanding case, and
$SB_0=SB$ for the static case.} from the subsample
MS 1054-03 in Trujillo {\it et al.}\cite{Tru06} galaxies, and
the angular sizes $\theta _{1550 \AA }=1.14\theta _V$ \cite{McI05}
for $n_S>2.5$ and $\theta _{1550 \AA }=1.10\theta _V$ \cite{Bar05}
for $n_S<2.5$, I get
the values of $SB_0$ for all the galaxies in this subsample
for the expanding or static cases (Fig. \ref{Fig:surfaceb}).
For the expanding universe,
many galaxies have average intrinsic surface brightness ($SB_0$)
lower(brighter) than 18.5 mag$_{AB}$/arcsec$^2$,
the galaxy MS 1054-03/1356 being the brightest one per unit angular
area: 14.8 mag$_{AB}$/arcsec$^2$ (30 times brighter than the limit).
The angular size of this galaxy MS 1054-03/1356 (0.027$''$ circularized in V;
0.031$''$ in FUV) might be affected by some error since it is below 0.125$''$ in V,
but even an error of 100\% in angular size would produce an error of 1.5
magnitudes in surface brightness, not 3.7 magnitudes as we observe
here. The dispersion (r.m.s.) of
$\frac{\theta _{1550 \AA }}{\theta _V}$
might be around 20\% (\cite{Bar05}, Fig. 2), which means an
uncertainty of around 0.4 mag/arcsec$^2$ (1-$\sigma $) in $SB_0$,
much lower than the differences between $SB_0$ and its limit of 18.5.
Moreover, even avoiding galaxies with angular size less than 0.125$''$,
there are some galaxies with surface brightness over the limit,
up to 6 times brighter than the limit. Too high intrinsic FUV surface
brightness. Lerner\cite{Ler06} also argues why other
alternative explanations (lower production of dust at high redshift, winds
or others) are not consistent.
However, for the static models, all galaxies have average intrinsic
surface brightness ($SB_0$) within $\theta $
higher(fainter) than the limit 18.5 mag$_{AB}$/arcsec$^2$, a result which
may be interpreted again in favor of the static scenario.
\begin{figure*}
\vspace{1cm}
{\par\centering \resizebox*{5cm}{5cm}{\includegraphics{fig6.eps}}
\hspace{1cm}\resizebox*{5cm}{5cm}{\includegraphics{fig8.eps}}\par}
\caption{Intrinsic FUV-rest surface brightness
of the galaxies in the subsample MS 1054-03 in an expanding (left) or static
(right) universe.
The dashed line stands for the minimum value
of 18.5 mag/arcsec$^2$ over which all galaxies should be located.
Points with circles stand for data with $\theta <0.125"$.}
\label{Fig:surfaceb}
\end{figure*}
\subsection{Is a static model theoretically impossible?}
Apart from the discussion on the observations, which are inconclusive,
a static model is usually rejected by most cosmologists on the grounds
of a belief/prejudice that a static model is impossible.
However, from a purely theoretical point of view, without
taking into account the astronomical observations, the representation
of the Cosmos as Euclidean and static is not excluded. Both expanding
and static space are possible for the description of the Universe, even with
evolution.
Before Einstein and the rise of Riemannian and other non-Euclidean
geometries to the stage of physics, attempts to describe the
known Universe with a Euclidean Universe were given, but with
the problem of justifying a stable equilibrium. Within a relativistic context,
Einstein\cite{Ein17} proposed a static model including a cosmological
constant, his biggest blunder according to himself, to avoid a collapse.
This model still has problems to guarantee the stability, but it
might be solved somehow. Narlikar \& Arp\cite{Nar93} solve it
within some variation of the Hoyle--Narlikar conformal theory
of gravity, in which small perturbations of the flat Minkowski
spacetime would lead to small oscillations about the line element
rather than to a collapse.
Boehmer {\it et al.}\cite{Boe07} analyze the stability of the Einstein static
universe by considering homogeneous scalar perturbations in the context
of $f(R)$ modified theories of gravity and it is found
that a stable Einstein cosmos with a positive cosmological constant
is possible.
Other authors solve it with the
variation of fundamental constants \cite{Van84,Tro87}.
Another idea by Van Flandern\cite{Van93} is that hypothetical gravitons responsible
for the gravitational interaction
have a finite cross-sectional area, so that they can only travel a finite
distance, however great, before colliding with another graviton.
So the range of the force of gravity would necessarily be limited in this way
and collapse is avoided.
As said in \S \ref{.expansion}, the very concept of space expansion has its own
problems\cite{Fra07,Bar08}. The curved geometry (general relativity and its
modifications) has no conservation of the energy-momentum of the gravity field
(the well-known problem of the pseudo-tensor character of the energy-momentum of
the gravity field in general relativity). However, Minkowski space follows the
conservation of energy-momentum of the gravitational field. One approach with a
material tensor field in Minkowski space is given in Feynman's gravitation\cite{Fey95}, where
the space is static but matter and fields can be expanding
in a static space. It is also worth mentioning a model related to modern
relativistic and quantum field theories of basic fundamental interactions
(strong, weak, electro-magnetic): the relativistic field gravity theory and fractal matter
distribution in static Minkowski space\cite{Bar08b}.
Olber's paradox for an infinite Universe also needs subtle
solutions, but extinction, absorption and reemission of light, fractal
distribution of density and
the mechanism which itself produces the redshift of the galaxies might
have something to do with its solution.
These are old questions discussed in many classical books on
cosmology (e.g., \cite{Bon61}, ch. 3) and do not warrant further
discussion here.
\section{Conclusions}
Summing up, the main conclusions of this paper are the following:
\begin{itemize}
\item The average angular size of the galaxies for a given luminosity
with redshifts between $z=0.2$ and 3.2 is approximately
proportional to $z^{-\alpha }$, with
$\alpha $ between 0.7 and 1.2, depending on the assumed cosmology.
\item Any model of an expanding Universe without evolution is totally
unable to fit the angular size vs. $z$ dependence.
The hypothesis that galaxies which formed earlier have much
higher densities does not work because it is not observed here that the smaller
galaxies are precisely those which formed earlier; in any case, the galaxies
observed today were formed mostly at redshifts not very different from the
galaxies observed at higher redshifts.
A very strong evolution in size would be able to get an agreement with the data but
there appear caveats to justify it in terms of age variation of the
population and/or mergers and/or ejection of massive outflows in the
quasar feedback. An average of the necessary
two to four major mergers per galaxy during
its lifetime is excessive, and neither is it understood how
massive elliptical galaxies may present passive evolution in this scenario
or how spiral galaxies can become larger during their lifetimes.
The depletion of gas in ellipticals by a QSO feedback mechanism
does not appear to be in agreement with the observed star formation and other
facts. Moreover, no evolution is observed in the rotation/dispersion
velocities and, the FUV surface brightness turns out to be prohibitively high in some
galaxies at high redshift.
\item Static Euclidean models with a linear Hubble law or simple tired light
fit the shape of the
angular size vs. $z$ dependence very well: there is a difference
in amplitude of 20--30\%, which is within the possible systematic errors.
An extra small intergalactic extinction may also explain this difference
of 20--30\% . Some weak evolution of very high redshift sources is allowed,
although non-evolution is a possible solution too. For the plasma redshift
tired light static model, a strong (albeit weaker than in expanding models)
evolution in galaxy size is necessary to fit the data.
The SNe Ia Hubble diagram can also be explained in terms of
these models.
\item It is also remarkable that the explanation of test
results with an expanding Universe requires four coincidences:
\begin{enumerate}
\item The combination of expansion and (very strong) size evolution gives
nearly the same result as a static Euclidean universe
with a linear Hubble law alone: $\theta \propto z^{-1}$.
\item This hypothetical evolution in size for galaxies is the same
in normal galaxies as in QSOs, as in radio galaxies, as in
first ranked cluster galaxies, as the separation among bright
galaxies in clusters. Everything evolves in the same way to
produce approximately a dependence $\theta \propto z^{-1}$.
\item The concordance model gives approximately the same
(differences of less than 0.2 mag within $z<4.5$) distance modulus
in a Hubble diagram as the static Euclidean universe with a linear Hubble law.
\item The combination of expansion, (very strong) size evolution,
and dark matter ratio variation gives the same result for the velocity
dispersion in elliptical galaxies (the result is that it is nearly constant
with $z$) as for a simple static model with no evolution in size and no
dark matter ratio variation.
\end{enumerate}
These four coincidences might make us think that possibly we
should apply Occam's razor {\it ``Entia non sunt multiplicanda
praeter necessitatem''}, we should use the simplest models
that can reproduce the same things as a complex model with
many more free parameters does.
\end{itemize}
It would be an irony of
fate that, after all the complex solutions pursued by cosmologists
for the last century, we had to come back to simple scenarios such as a
Euclidean static Universe without expansion. None the less, we cannot
at present defend any of these simple models apart from the standard one
because this would require other analyses. The conclusion of this paper
is just that the data on angular size vs. redshift
present some conflict with the standard model, and that they are
in accordance with a very simple phenomenological extrapolation of the Hubble
relation that might ultimately be linked to a static model of the universe.
\section*{Acknowledgments}
Thanks are given to: the anonymous referee for very helpful suggestion to
improve this paper;
Ignacio Trujillo and Eric J. Lerner
for useful discussions on many topics related to this paper
and comments on a draft of it; Riccardo Scarpa,
Carlos M. Guti\'errez, Juan Betancort-Rijo also
for their comments on a draft of this paper; Daniel
MacIntosh for providing me the GEMS data from the
papers \cite{McI05,Bar05}; Terry J. Mahoney for proof-reading this paper.
MLC was supported by the {\it Ram\'on y Cajal} Programme
of the Spanish Ministry of Science.
|
\section{Introduction}
\label{intro}
In the standard model (SM) of particle physics, \ensuremath{C\!P}\xspace\ violation occurs
due to an irreducible phase appearing in the quark-flavor mixing matrix,
called the Cabibbo-Kobayashi-Maskawa (CKM) matrix, which relates the weak
interaction eigenstates to that of mass. The study of $B$ meson decays
allows us to carry out a multitude of measurements involving
the angles and sides of the so-called unitarity triangle (UT), a graphical
sketch of the unitarity of the CKM matrix in the complex plane. The
{\em raison d'\^{e}tre} of the two $B$-factory experiments -- Belle at
KEK, Japan and BaBar at SLAC, USA -- was to precisely measure various UT
parameters. By doing so, they were designed to verify the \ensuremath{C\!P}\xspace\ violation
mechanism within the SM, as suggested by Kobayashi and Maskawa~\cite{km},
and to set constraints on potential new physics contributions in the flavor
sector.
In these proceedings, we summarize recent results on \ensuremath{C\!P}\xspace\ violation,
involving three angles of the unitarity triangle, and describe a number
of hints for new physics observed with the $B$-factory experiments.
\section{Angles of the unitarity triangle}
\label{ang}
The UT angles are determined through the measurement of the time dependent
\ensuremath{C\!P}\xspace\ asymmetry, $A_{\ensuremath{C\!P}\xspace}(t)$, defined as
\begin{equation}
A_{\ensuremath{C\!P}\xspace}(t) = \frac{N[\kern 0.18em\overline{\kern -0.18em B}{}\xspace^0(t)\to f_{\ensuremath{C\!P}\xspace}]-N[B^0(t)\to f_{\ensuremath{C\!P}\xspace}]}
{N[\kern 0.18em\overline{\kern -0.18em B}{}\xspace^0(t)\to f_{\ensuremath{C\!P}\xspace}]+N[B^0(t)\to f_{\ensuremath{C\!P}\xspace}]},
\label{eq1}
\end{equation}
where $N[\kern 0.18em\overline{\kern -0.18em B}{}\xspace^0/B^0(t)\to f_{\ensuremath{C\!P}\xspace}]$ is the number of $\kern 0.18em\overline{\kern -0.18em B}{}\xspace^0/B^0$s that
decay into a \ensuremath{C\!P}\xspace\ eigenstate $f_{\ensuremath{C\!P}\xspace}$ after time $t$. The asymmetry, in
general, can be expressed in terms of two components:
\begin{equation}
A_{\ensuremath{C\!P}\xspace}(t)=S_f\sin(\Delta mt)+A_f\cos(\Delta mt),
\label{eq2}
\end{equation}
where $\Delta m$ is the difference in mass of $B^0$ mass eigenstates.
The sine coefficient $S_f$ is related to the UT angles, while the cosine
coefficient $A_f$ is a measure of direct \ensuremath{C\!P}\xspace\ violation. For the latter
to have a nonzero value, we need at least two competing amplitudes with
different weak and strong phase to contribute to the decay final state.
As an example, for the decay $B^0\to\ensuremath{{J\mskip -3mu/\mskip -2mu\psi\mskip 2mu}}\xspace\KS$, where mostly one diagram
contributes, the cosine term is expected to vanish and the sine term is
proportional to the UT angle $\phi_1$\footnote{An alternative notation
of $\beta$, $\alpha$ and $\gamma$, that correspond to $\phi_1$, $\phi_2$
and $\phi_3$, respectively, is equally abundant in the literature.}. The
time-dependent \ensuremath{C\!P}\xspace\ asymmetry is, therefore, given as
\begin{equation}
A_{\ensuremath{C\!P}\xspace}(t) = -\xi_f\sin(2\phi_1)\sin(\Delta mt),
\label{eq3}
\end{equation}
where $\xi_f$ is the \ensuremath{C\!P}\xspace\ eigenvalue of the final state. In the case of
$B$ factories, the measurement of $A_{\ensuremath{C\!P}\xspace}(t)$ utilizes decays of the
\Y4S\ into two neutral $B$ mesons, of which one can be completely
reconstructed into a \ensuremath{C\!P}\xspace\ eigenstate, while the decay products of the
other (called the tag $B$) identify its flavor at decay time. The time
difference $t$ between the two $B$ decays is determined by reconstructing
their decay vertices. Finally the \ensuremath{C\!P}\xspace\ asymmetry amplitudes, proportional
to the UT angles, are obtained from an unbinned maximum likelihood fit to
the proper time distributions separately for events tagged as $\kern 0.18em\overline{\kern -0.18em B}{}\xspace^0$
and $B^0$.
\subsection{The angle {\boldmath $\phi_1$}}
The most precise measurement of the angle $\phi_1$ is obtained from a
study of the decays $B^0\to$ charmonium $+K^{(*)0}$. These decays, known
as ``golden modes'', mainly proceed via the CKM-favored tree diagram
$b\to c\bar{c}s$ with an internal $W$ boson emission. The subleading
penguin (loop) contribution to the final state, that has a different weak
phase compared to the tree diagram, is suppressed by almost two orders of
magnitude. This makes $A_f=0$ in Eq.~\ref{eq2} to a very good approximation.
Besides the theoretical simplicity, these channels also offer experimental
advantages because of the relatively large branching fractions ($\sim 10^{-3}$)
and the presence of narrow resonances in the final state, which provides
a powerful rejection against combinatorial background. The \ensuremath{C\!P}\xspace\ eigenstates
considered for this analysis include $\ensuremath{{J\mskip -3mu/\mskip -2mu\psi\mskip 2mu}}\xspace\KS$, $\ensuremath{\psi{(2S)}}\xspace\KS$, $\chi_{c0}\KS$,
$\eta_c\KS$ and $\ensuremath{{J\mskip -3mu/\mskip -2mu\psi\mskip 2mu}}\xspace\KL$. The measured world-average value of $\sin(2\phi_1)$
is $0.67\pm 0.02$. Figure~\ref{sin2phi1} shows the impact of this measurement
by Belle and BaBar, that eventually led to half of the 2008 physics Nobel
prize~\cite{nobel} being awarded to Kobayashi and Maskawa, when compared to
other experiments.
\begin{figure}[!htb]
\center
\includegraphics[width=.7\columnwidth]{phi1}
\caption{Average of $\sin(2\phi_1)$ from all experiments, as compiled by the
HFAG~\cite{hfag}.}
\label{sin2phi1}
\end{figure}
\subsection{The angle {\boldmath $\phi_2$}}
Decays of $B$ mesons to the final states $hh$ ($h=\rho$ or $\pi$),
dominated by the CKM-suppressed $b\to u$ transition, are sensitive
to the UT angle $\phi_2$. The presence of $b\to d$ penguin diagrams,
however, complicates the situation by introducing additional phases
such that the measured parameter is no more $\phi_2$ alone, rather
an effective value $\phi^{\rm eff}_2 =\phi_2+\delta\phi_2$. (Note that
the same prescription {\it vis-a-vis} penguin pollution also applies
to other UT angles, wherever appropriate.) At present, the most
precise measurement of this angle is obtained in the analysis of
the decays $B\to\rho\rho$. Combining with additional constraints
coming from $B\to\rho\pi$ and $B\to\pi\pi$, we measure $\phi_2=
\left(89.0^{+4.4}_{-4.2}\right)^\circ$~\cite{ckmfitter}.
\subsection{The angle {\boldmath $\phi_3$}}
The angle $\phi_3$ is measured by exploiting the interference between the
decays $B^-\to D^{(*)0}K^{(*)-}$ and $B^-\to\ensuremath{\Dbar}\xspace^{(*)0}K^{(*)-}$, where
both $D^0$ and $\kern 0.2em\overline{\kern -0.2em D}{}\xspace^0$ decay to a common final state. This measurement
can be performed in three different ways: utilizing decays of $D$ mesons
to \ensuremath{C\!P}\xspace\ eigenstates~\cite{glw}, making use of doubly Cabibbo-suppressed
decays of the $D$ meson~\cite{ads}, and exploiting the interference
pattern in the Dalitz plot of $D\to\KS\pi^+\pi^-$ decays~\cite{ggsz}.
Currently, the last method provides the strongest constraint on $\phi_3$.
Combining all related measurements from Belle and BaBar, the world-average
value is found to be $\phi_3=\left(73^{+19}_{-24}\right)^\circ$~\cite{ckmfitter}.
\begin{figure}[!htb]
\center
\includegraphics[width=.48\columnwidth]{rhoeta_small_angles}
\includegraphics[width=.48\columnwidth]{rhoeta_small_global}
\caption{Constraints on the UT coming from the measurements of angles
only (left) and using all relevant experimental inputs (right).}
\label{ckmfitresult}
\end{figure}
In Fig.~\ref{ckmfitresult} we summarize the constraints on the UT coming
from the measurements of angles only, as well as after including other
experimental inputs. To a very good approximation, the Kobayashi-Maskawa
formalism is found to be the right description of \ensuremath{C\!P}\xspace\ violation in the SM.
Needless to say that we still need to improve the precision on the third
angle $\phi_3$ -- one has just made a head-start! Similarly, we expect
the errors on other two angles to shrink further, {\it e.g.}, once Belle
analyzes its full \Y4S\ dataset.
\section{Search for physics beyond the SM}
In this section we attempt to enumerate various hints for, or constraints
on, potential new physics contributions, as observed with the $B$ factories.
\subsection{Measured {\boldmath $\sin(2\phi_1)$} with the penguins}
As $\sin(2\phi_1)$ is the most precisely measured observable concerning
\ensuremath{C\!P}\xspace\ violation in $B$ decays, we can use it as a ``Standard Candle'' to
set constraints on new physics by looking for possible deviations from
this value in a number of ways. One such is the comparison of the values
of $\sin(2\phi^{\rm eff}_1)$ measured in penguin dominated decays with
the world-average value of $\sin(2\phi_1)$, coming from decays involving
charmonium final states. The caveat to making such a comparison is that
the penguin modes may have additional topologies that could lead to a
difference between $\sin(2\phi_1)$ and $\sin(2\phi^{\rm eff}_1)$. If these
SM corrections, $\Delta_{\rm SM}$, are well known then any residual difference
$\Delta S=\sin(2\phi^{\rm eff}_1)-\sin(2\phi_1)-\Delta_{\rm SM}$ would be
from new physics. It has been recently pointed out~\cite{lunghisoni} that
by comparing the penguin to tree channels one remains insensitive to possible
new physics contribution common to both. Therefore, it is important to compare
the directly measured values of $\sin(2\phi^{\rm eff}_1)$ with the predictions
of SM-based constraints for the same observable. Figure~\ref{adrian}
summarizes the different constraints on $\sin(2\phi^{\rm eff}_1)$, where
the maximum difference between the measured and indirect values has a
significance above $2$ standard deviations.
\begin{figure}[!htb]
\center
\includegraphics[width=.7\columnwidth]{sin2phi1-peng}
\caption{Measured values of $\sin(2\phi_1)$ in (yellow/light-shaded)
charmonium decays, (blue/dark-shaded) penguin decays, and
(green/medium-shaded) inferred from indirect measurements~\cite{lunghisoni}.}
\label{adrian}
\end{figure}
\subsection{Direct {\boldmath $\ensuremath{C\!P}\xspace$} violation in {\boldmath $B$} decays}
Both Belle and BaBar have carried out a number of sensitive \ensuremath{C\!P}\xspace\ violation
measurements in various $B$ decays. Most notable of them is the decay
$B^0\to K^+\pi^-$, where direct \ensuremath{C\!P}\xspace\ violation has been established beyond
any doubt -- the measured \ensuremath{C\!P}\xspace\ asymmetry is $(-9.8^{+1.2}_{-1.1})\%$.
There are a number of interesting evidences at the level of 3 standard
deviations in the decays $B^0\to\eta K^{*0}$, $B^-\to\eta K^-$,
$B^-\to\rho^0 K^-$, $B^0\to\rho^+\pi^-$ and $B^-\to\ensuremath{\Dbar}\xspace^{(*)0}K^-$. Another
important result has come out from $B^-\to K^-\pi^0$, with the \ensuremath{C\!P}\xspace\ asymmetry
$(+5.0\pm 2.5)\%$. This in contrast to the result of $B^0\to K^+\pi^-$~\cite{nature},
where similar Feynman diagrams contribute at the tree level, tells us that
it could be either due to a large contribution from the color-suppressed
tree diagram, or from possible new physics contribution in the electroweak
penguin, or from both. Before firmly concluding anything, it is
suggested~\cite{sumrule} to check the \ensuremath{C\!P}\xspace\ violation result from the decay
$B^0\to K^0\pi^0$, with a larger dataset.
\subsection{Polarization puzzle in {\boldmath $B\to VV$}}
For a $B$ meson decaying to two vector particles, $B\to VV$, theoretical
models based on QCD factorization~\cite{vvqcdf} or perturbative
QCD~\cite{vvpqcd} predict the fraction of longitudinal fraction $f_L$
to be approximately $1-(m^2_V/m^2_B)$, where $m_{V(B)}$ is the mass of
the vector ($B$) meson, for tree-dominated decays. As an example, in
the case of $B\to\rho\rho$ the prediction for $f_L$ is close to $0.9$,
which matches well with the measurement~\cite{hfag}. For
decays dominated by the penguin transition, however, there is a large
discrepancy between predictions ($\sim 0.75$) and observations, that
tend to cluster around $0.5$. This unexpected result on polarization,
mostly driven by the measurement of $B\to\phi K^*$, has motivated
several further studies.
\subsection{Constraints on the charged Higgs}
\begin{figure}[!htb]
\center
\includegraphics[width=.5\columnwidth]{taunu}
\caption{Purely leptonic $B$ decays proceed via the annihilation of
quark-antiquark into a $W$ boson (or, potentially into a charged
Higgs boson).}
\label{taunu}
\end{figure}
The purely leptonic decay $B^-\to\tau^-\ensuremath{\nub_\tau}\xspace$ provides an excellent
probe for the charged Higgs that could potentially appear in the
annihilation of $b$ and $\bar{u}$ quarks similar to the SM diagram,
where a $W^-$ boson is created in the annihilation process (see
Fig.~\ref{taunu}). For instance, if we take the prediction of the
two-Higgs doublet model~\cite{2hdm}, the observed branching fraction could
be enhanced or suppressed by a factor of $(1-m^2_B\tan^2\beta/m^2_H)^2$,
where $m_H$ is the mass of the charged Higgs and $\tan\beta$ is the
ratio of the two Higgs vacuum expectation values. On the experimental
side, identifying the decay $B^-\to\tau^-\nu_\tau$, which involves at
least two neutrinos in the final state, is a real challenge. Both
Belle and BaBar have made the best use of their detector hermiticity
and particle identification capability, and in doing so they
obtain~\cite{taunuresult} a branching fraction world-average of
$(1.73\pm 0.35)\times 10^{-4}$ for the decay. The SM prediction is
$(1.20\pm 0.25)\times 10^{-4}$, where the dominant uncertainties
come from the error in the CKM matrix element $V_{ub}$ and the
$B$-meson decay constant. Comparing the SM expectation with the
measurement, we derive a constraint on $m_H$ as a function of
$\tan\beta$. This constraint is well complimented by the measurement
of $B\to D^{(*)}\tau\nu_\tau$~\cite{dtaunuresult} and the inclusive
$b\to s\gamma$ measurement~\cite{bsgamma}. It is worth noting
that the combined result~\cite{iijima}, which excludes a charged
Higgs up to a mass of $600\gevcc$ for $\tan\beta>60$ and $300\gevcc$
for $\tan\beta>30$, is already comparable to what is expected for a
direct search~\cite{atlas} using a $30\ensuremath{\mbox{\,fb}^{-1}}\xspace$ data sample at the LHC.
\subsection{{\boldmath $B\to K^{(*)}\ell^+\ell^-$}: Any smoking gun?}
The decay channel $b\to s\ell^+\ell^-$ is an experimenters delight, since
it offers many interesting observables that can be measured in the decays
of $B$ mesons to both inclusive and exclusive $s\ell^+\ell^-$ final
states, where $s$ denotes a strangeness-one meson. In particular, for the
exclusive mode $K^{(*)}\ell^+\ell^-$ the observables include $f_L$,
the forward-backward asymmetry $A_{FB}$, the isospin asymmetry $A_I$, and
the ratio of rates to $e^+e^-$ and $\mu^+\mu^-$ final states (lepton flavor
ratio). Recent measurements at the $B$ factories~\cite{sllbelle,sllbabar}
show that the branching fraction and the lepton flavor ratio agree with
SM expectations. However, a deviation from the SM is indicated in $A_{FB}$
(Fig.~\ref{afb-belle}), albeit with large statistical uncertainty. We
need more statistics than currently available, which would be possible
with the future experiments~\cite{superb}, to either confirm or refute
this tantalizing hint. If it is finally turned out to be real, it would
be a clean signature of new physics~\cite{wilson,amoletc}.
\begin{figure}[!htb]
\center
\includegraphics[width=.7\columnwidth]{afb-belle}
\caption{Results for (top) $f_L$ and (middle) $A_{FB}$ in $K^*\ell^+\ell^-$
as a function of $q^2$, together with the solid (dotted) curve representing
the SM ($C^{\rm NP}_7=-C^{\rm SM}_7$) prediction. (Bottom) The plot of
$A_I$ {\it vs.} $q^2$ for the $K^*\ell^+\ell^-$ (filled circles) and
$K\ell^+\ell^-$ (open circles) modes. The two shaded regions are veto
windows to reject events containing a $\ensuremath{{J\mskip -3mu/\mskip -2mu\psi\mskip 2mu}}\xspace$ or a $\ensuremath{\psi{(2S)}}\xspace$.}
\label{afb-belle}
\end{figure}
\section{Conclusions}
The two $B$-factory experiments have performed exceptionally well, each
producing an average over 400 high-quality journal publications within
only ten years of their inception. What we present here, is a small sampling
of their recent highlighted results. It is fair to say that the SM
continues to hold its ground in the flavor sector, though there are some
hints of new physics available, which should be investigated with more data.
\section{Acknowledgements}
I thank the organizers for their kind invitation to this interesting
conference. Thanks are also due to my Belle colleagues for their
valuable help. This work is supported in part by the Department of
Atomic Energy and the Department of Science and Technology of India.
|
\section{Introduction} \label{section:intro}
Davis and Januskiewicz~\cite{Da-Ja-1991}
introduced the notion of what is now
called a \emph{quasitoric manifold} which is a real $2d$-dimensional closed
smooth manifold $M$ with a locally standard smooth action of $T:=(S^1)^d$ whose
orbit space can be identified with a simple polytope. We note that there is a
natural bijection between the set of simple polytopes and the set of simplicial
polytopes via the dual operation. A quasitoric manifold $M$ is said to be
\emph{over} a simplicial polytope $P$ if the orbit space of $M$ can be
identified with the dual of $P$.\footnote{Many toric topologists prefer to use
the terminology `simple' instead of `simplicial'. However, to simplify the
arguments, we define all notions in terms of simplicial polytopes throughout
this paper.} By Davis and Januskiewicz \cite{Da-Ja-1991}, the equivariant
cohomology ring $H^\ast_T(M) = H^\ast(ET \times_T M)$ with $\mathbb{Z}$-coefficient is
isomorphic to the face ring $\mathbb{Z}[P]$ of $P$ as a graded ring, where $ET$ is a
contractible space which admits a free $T$-action. We note that the natural
projection $p:ET \times_T M \to BT$ induces a $H^\ast(BT)$-module structure of
$H^\ast_T(M)$, where $BT:=ET/T$. They also showed that it is a free-module,
i.e., $H^\ast_T(M) \cong H^\ast(M) \otimes H^\ast(BT)$. Hence, we deduce that
$H^\ast(M) \cong H^\ast_T(M)/p^\ast(H^\ast(BT)) \cong \mathbb{Z}[P]/p^\ast(H^\ast(BT))$,
where $p^\ast : H^\ast(BT) \to H^\ast_T(M)$ is the induced map of $p$. See
\cite[Section 5]{Bu-Pa-2002} for more details. Thus $H^\ast(M)$ contains some
information of the orbit space $P$. With this viewpoint, Choi et
al. \cite{Ch-Pa-Su-2010} defined the cohomological rigidity of $P$ as
follows.\footnote{The original definition of cohomological rigidity was firstly
introduced by Masuda and Suh in \cite{Ma-Su-2008} in terms of toric manifolds
and fans.} A simplicial polytope $P$ is \emph{cohomologically rigid} if there
exists a quasitoric manifold $M$ over $P$, and whenever there exists a
quasitoric manifold $N$ over another polytope $Q$ with a graded ring isomorphism
$H^*(M)\cong H^*(N)$, then $P=Q$ up to isomorphism. Choi et al.~
\cite{Ch-Pa-Su-2010} also showed that $H^*(M)\cong H^*(N)$ implies
$\beta_{i,j}(P)=\beta_{i,j}(Q)$ for all $i,j$, where $\beta_{i,j}(P)$ is the
\emph{$(i,j)$th graded Betti number} of $P$. One can define the graded Betti numbers
$\beta_{i,j}(P)$ using a finite free resolution of the face ring of $P$, for
example see \cite{Ch-Ki-2010}. Instead of doing this, we will simply take the
following Hochster's formula as the definition of $\beta_{i,j}(P)$:
\begin{equation}
\label{eq:1}
\beta_{i,j}(P)=\sum_{\substack{W\subset V(P)\\ |W|=j}}
\dim_{\mathbf{k}}\widetilde H_{j-i-1}(P|_W;\mathbf{k}),
\end{equation}
where $V(P)$ is the set of vertices of $P$, $\mathbf{k}$ is an arbitrary field and
$P|_W$ is the realization of the simplicial complex $\{F \cap W : F \in
\Delta(P)\}$, where $\Delta(P)$ is the boundary complex of $P$.
With this motivation, we define the
following.
\begin{defn}
A simplicial polytope $P$ is \emph{combinatorially rigid} (or simply
\emph{rigid}) if we have $P=P'$ for any simplicial polytope $P'$ satisfying
$\beta_{i,j}(P)=\beta_{i,j}(P')$ for all $i, j\geq0$.
\end{defn}
Hence, if $P$ supports a quasitoric manifold and $P$ is combinatorially rigid,
then $P$ is cohomologically rigid.
In the present paper, we investigate the combinatorial rigidity of
$3$-dimensional simplicial polytopes. Remark that since all $3$-dimensional
simplicial polytopes support quasitoric manifolds, $3$-dimensional
combinatorially rigid polytopes are cohomologically rigid.
Let $P$ be a $3$-dimensional simplicial polytope with $n$ vertices. Then
$\dim_{\mathbf{k}}\widetilde H_{j-i-1}(P|_W;\mathbf{k}) = 0$ if $j-i-1\geq2$ for $W \subsetneq
V(P)$ and $\dim_{\mathbf{k}}\widetilde H_{j-i-1}(P|_W;\mathbf{k})=\delta_{j-i-1,2}$ for
$W=V(P)$, where $\delta_{x,y}=1$ if $x=y$ and $0$ otherwise. Thus it is enough
to consider $\beta_{i-1,i}(P)$ and $\beta_{i-2,i}(P)$. By the Poincar\'{e}
duality $\beta_{i,j}(P) = \beta_{n-3-i,n-j}(P)$, we have
$\beta_{i-2,i}(P)=\beta_{n-i-1,n-i}(P)$. Thus we only need to consider
$\beta_{i-1,i}(P)$, which we will call \emph{the $i$th special graded Betti
number} and we denote $b_i(P):=\beta_{i-1,i}(P)$. By \eqref{eq:1}, we can
interpret $b_i(P)$ in a purely combinatorial way as follows:
\begin{equation}\label{eq:bk}
b_i(P)=\sum_{\substack{W\subset V(P)\\|W|=i}}
\left(\cc(P|_W)-1\right),
\end{equation}
where $\cc(P|_W)$ denotes the number of connected components of $P|_W$.
From the above observation, we get the following proposition.
\begin{prop}
Let $P$ be a $3$-dimensional simplicial polytope. Then $P$ is combinatorially
rigid if and only if $P$ is determined by $b_i(P)$'s for $i\geq 0$.
\end{prop}
So far, several polytopes are proved to be cohomologically rigid. In
$3$-dimensional case, Choi et al.~\cite{Ch-Pa-Su-2010} classified all
cohomologically rigid polytopes with at most $9$ vertices using computer and
proved that the icosahedron is cohomologically rigid. Since they only used the
graded Betti numbers, what they found are combinatorially rigid as well.
To state our main results we need to define connected sum which is a simple
operation to get a $d$-dimensional simplicial polytope from two $d$-dimensional
simplicial polytopes.
Let $P_1$ and $P_2$ be simplicial polytopes. A \emph{connected sum} of $P_1$ and
$P_2$ is a polytope obtained by attaching a facet of $P_1$ and a facet of
$P_2$. It depends on the way of choosing the two facets and identifying their
vertices. Let $\mathcal{C}(P_1\# P_2)$ denote the set of connected sums of $P_1$ and
$P_2$. If there is only one connected sum of $P_1$ and $P_2$ up to isomorphism,
then we will write the unique polytope as $P_1\# P_2$.
If a simplicial polytope $P$ can be expressed as a connected sum of two
polytopes, then $P$ is called \emph{reducible}. Otherwise, $P$ is called
\emph{irreducible}.
Let $T_4$, $C_8$, $O_6$, $D_{20}$ and $I_{12}$ be the five Platonic solids: the
tetrahedron, the cube, the octahedron, the dodecahedron and the icosahedron
respectively.
In this paper, we prove the following necessary condition to be combinatorially
rigid for $3$-dimensional reducible simplicial polytopes. See
Section~\ref{sec:necessary} or Figures~\ref{fig:cube}, \ref{fig:dodeca} and
\ref{fig:prism} for the definition of $\xi_1(C_8)$, $\xi_2(C_8) $,
$\xi_1(D_{20})$, $\xi_2(D_{20})$ and $B_n$, the bipyramid with $n$ vertices.
\begin{thm}\label{thm:necessary}
Let $P$ be a $3$-dimensional simplicial polytope. If $P$ is reducible and
combinatorially rigid, then $P$ is either $T_4 \# T_4 \# T_4$ or $P_1\#P_2$,
where
\begin{align*}
P_1 & \in \{T_4,O_6,I_{12}\},\\
P_2 &\in \{T_4,O_6,I_{12},\xi_1(C_8),\xi_2(C_8)
,\xi_1(D_{20}),\xi_2(D_{20})\} \cup \{B_n:n\geq7\}.
\end{align*}
\end{thm}
Note that $B_n$ is defined for $n\geq5$ and we have $B_5=T_4\#T_4$ and $B_6=O_6$.
In fact, $T_4 \# T_4 \# T_4$ is known to be rigid, see \cite{Ch-Pa-Su-2010}. We
also prove that $P_1\#P_2$ is rigid for some $P_1$ and $P_2$ in
Theorem~\ref{thm:necessary}.
\begin{thm}\label{thm:rigid}
The following polytopes are combinatorially rigid:
$$T_4\# T_4, T_4\# O_6, T_4\# I_{12}, T_4\# B_n, O_6\# O_6, O_6\# B_n,$$
where $n\geq 7$. See Table~\ref{tab:rigid}.
\end{thm}
\begin{table}
\begin{tabular}{c|c|c|c|c|c|c|c|c}
$\#$ & $T_4$ & $O_6$ & $I_{12}$ & $B_n$, $n\geq7$ & $\xi_1(C_8)$& $\xi_2(C_8)$&
$\xi_1(D_{20})$ & $\xi_2(D_{20})$\\ \hline
$T_4$ & rigid & rigid & rigid & rigid & ?& ?& ?& ?\\ \hline
$O_6$ & - & rigid & ? & rigid & ?& ?& ?& ?\\ \hline
$I_{12}$ & - & - & ? & ? & ?& ?& ?& ?\\
\end{tabular}
\caption{Combinatorial rigidity of a connected sum of
two irreducible polytopes.}
\label{tab:rigid}
\end{table}
The rest of this paper is organized as follows. In Section~\ref{sec:necessary}
we prove Theorem~\ref{thm:necessary}. In Section~\ref{sec:max} we find the
maximum of $b_{n-4}(P)$ for a simplicial polytope $P$ with $n$ vertices. In
Section~\ref{sec:rigid} we prove Theorem~\ref{thm:rigid}.
\section{A necessary condition for rigid reducible polytopes
}\label{sec:necessary}
The authors \cite{Ch-Ki-2010} found the following formula for the special graded
Betti numbers of $P\in\mathcal{C}(P_1\# P_2)$ for $d$-dimensional simplicial polytopes
$P_1$ and $P_2$ with $n_1$ and $n_2$ vertices respectively:
\begin{equation}\label{thm:connected_sum}
b_k (P) = \sum_{i=0}^k \left( b_i (P_1) \binom{n_2 -
d}{k-i} + b_i (P_2) \binom{n_1 - d}{k-i}\right) + \binom{n_1 +
n_2 -2d}{k}.
\end{equation}
The above formula says that the special graded Betti numbers of a connected sum
of two polytopes do not depend on the ways of choosing the two facets and
identifying them. Thus we get the following.
\begin{prop}\label{thm:unique}
Let $P\in\mathcal{C}(P_1\# P_2)$ for $d$-dimensional simplicial polytopes $P_1$ and
$P_2$. If $P$ is rigid, then $P$ is the only element in $\mathcal{C}(P_1\# P_2)$.
\end{prop}
\emph{From now on, all polytopes that we consider are $3$-dimensional and
simplicial unless otherwise stated.} As usual for $3$-dimensional polytopes,
we will call $0$, $1$, $2$-dimensional face, respectively, \emph{vertex},
\emph{edge} and \emph{face}. We will sometimes identify a polytope $P$ with its
graph which is also called the $1$-skeleton of $P$. For a set $B$ of vertices,
$P|_B$ is the subgraph of $P$ induced by $B$.
Let $P$ be a polytope with vertex set $V$. A \emph{$k$-belt} of $P$ is a set
$B=\{v_1,v_2,\ldots,v_k\}$ of $k$ vertices such that $P|_B$ is a $k$-gon and
$P|_{V\setminus B}$ is disconnected. Let $|V|=n$. It is easy to see that if
$b_{n-k}(P)\ne 0$ for $k>0$, then $P$ has a $t$-belt for some $t\leq k$.
Note that $P$ has a $3$-belt if and only if $P$ is reducible. If
$P\in\mathcal{C}(P_1\#P_2)$, then the vertices of the attached face of $P_1$ (or
equivalently $P_2$) form a $3$-belt. Using this observation we can prove the
following.
\begin{prop}
If $P$ is a connected sum of at least $3$ irreducible polytopes and $P\ne
T_4\# T_4\# T_4$, then $P$ is not rigid.
\end{prop}
\begin{proof}
Let $P=P_1 \# \cdots \# P_\ell$ for irreducible polytopes
$P_1,\ldots,P_\ell$. By Proposition~\ref{thm:unique}, it is enough to show
that there are two different polytopes in $\mathcal{C}(P_1 \# \cdots \#
P_\ell)$.
Let $Q\in\mathcal{C}(P_1 \# \cdots \# P_\ell)$ be a polytope satisfying the following
condition $(*)$: there is an edge contained in all $3$-belts. We can construct
such $Q$ as follows. Let us fix an edge $\{a,b\}$ of $P_1$. Let $Q_1=P_1$ and
for $2\leq i\leq \ell$, let $Q_i\in\mathcal{C}(Q_{i-1}\# P_i)$ be a polytope obtained
by attaching a face of $Q_{i-1}$ containing $\{a,b\}$ and a face of
$P_i$. Then $Q=Q_\ell$ satisfies the condition $(*)$.
It is sufficient to construct $Q'\in\mathcal{C}(P_1 \# \cdots \# P_\ell)$ which
does not satisfy the condition $(*)$.
Assume $\ell=3$. Since $P\ne T_4\# T_4\# T_4$, we can assume that $P_1\ne
T_4$. Let $R\in\mathcal{C}(P_1\#P_2)$ and $\{a,b,c\}$ be the unique $3$-belt of
$R$. Note that $R$ has at least $7$ faces. Since there are only $6$ faces in
$R$ containing an edge of the $3$-belt $\{a,b,c\}$, we can find a face $F$ of
$R$ which does not contain any such edge. Let $Q'$ be a polytope in
$\mathcal{C}(R\#P_3)$ obtained by attaching $F$ and a face of $P_3$. Then the vertices
of $F$ form a $3$-belt of $Q'$. Clearly $Q'$ does not satisfy the condition
$(*)$.
Assume $\ell\geq4$. Let $R\in\mathcal{C}(P_1 \# \cdots \# P_{\ell-1})$ be a
polytope satisfying the condition $(*)$. Since $R$ has $\ell-2$ $3$-belts,
there is a unique edge $\{a,b\}$ contained in all $3$-belts of $R$. Let
$F$ be a face of $R$ which does not contain $\{a,b\}$. Let
$Q'\in\mathcal{C}(R\#P_\ell)$ be a polytope obtained by attaching $F$ and a face of
$P_\ell$. Then $Q'$ does not satisfy the condition $(*)$.
\end{proof}
Let $P$ be a polytope which is not necessarily simplicial. The
\emph{first-subdivision}, denoted $\xi_1(P)$, of $P$ is the simplicial polytope
obtained from $P$ by adding one vertex at the center of each face and connecting
it to all vertices of the face. The \emph{second-subdivision} (or
\emph{barycentric subdivision}) denoted $\xi_2(P)$, of $P$ is the simplicial
polytope obtained from $P$ by adding one vertex at the center of each face and
connecting it to all vertices of the face and all mid-points of the edges of the
face. See Figures~\ref{fig:cube} and \ref{fig:dodeca}.
\begin{figure}
\centering
\begin{pspicture}(0,0)(3,1.5) \vput(0,1)0 \vput(0,0)0 \vput(1,0)0
\vput(1.7,0.5)0 \vput(1.7,1.5)0 \vput(0.7,1.5)0
\pspolygon(0,1)(0,0)(0,0)(1,0)(1,0)(1.7,0.5)(1.7,0.5)(1.7,1.5)(1.7,1.5)(0.7,1.5)
\psline(1,1)(0,1) \psline(1,1)(1.7,1.5) \psline(1,1)(1,0)
\end{pspicture}
\begin{pspicture}(0,0)(3,1.5) \vput(0,1)0 \vput(0,0)0 \vput(1,0)0
\vput(1.7,0.5)0 \vput(1.7,1.5)0 \vput(0.7,1.5)0
\pspolygon(0,1)(0,0)(0,0)(1,0)(1,0)(1.7,0.5)(1.7,0.5)(1.7,1.5)(1.7,1.5)(0.7,1.5)
\psline(1,1)(0,1) \psline(1,1)(1.7,1.5) \psline(1,1)(1,0)
\vput(0.5,0.5)0 \psline(0.5,0.5)(1,1) \psline(0.5,0.5)(0,1)
\psline(0.5,0.5)(0,0) \psline(0.5,0.5)(1,0) \vput(0.85,1.25)0
\psline(0.85,1.25)(1,1) \psline(0.85,1.25)(0,1)
\psline(0.85,1.25)(0.7,1.5) \psline(0.85,1.25)(1.7,1.5)
\vput(1.35,0.75)0 \psline(1.35,0.75)(1,1) \psline(1.35,0.75)(1,0)
\psline(1.35,0.75)(1.7,0.5) \psline(1.35,0.75)(1.7,1.5)
\end{pspicture}
\begin{pspicture}(0,0)(2,1.5) \vput(0,1)0 \vput(0,0)0 \vput(1,0)0
\vput(1.7,0.5)0 \vput(1.7,1.5)0 \vput(0.7,1.5)0
\pspolygon(0,1)(0,0)(0,0)(1,0)(1,0)(1.7,0.5)(1.7,0.5)(1.7,1.5)(1.7,1.5)(0.7,1.5)
\psline(1,1)(0,1) \psline(1,1)(1.7,1.5) \psline(1,1)(1,0)
\vput(0.5,0.5)0 \psline(0.5,0.5)(1,1) \psline(0.5,0.5)(0,1)
\psline(0.5,0.5)(0,0) \psline(0.5,0.5)(1,0) \vput(0.5,1)0
\psline(0.5,0.5)(0.5,1) \vput(0,0.5)0 \psline(0.5,0.5)(0,0.5)
\vput(0.5,0)0 \psline(0.5,0.5)(0.5,0) \vput(1,0.5)0
\psline(0.5,0.5)(1,0.5) \vput(0.85,1.25)0 \psline(0.85,1.25)(1,1)
\psline(0.85,1.25)(0,1) \psline(0.85,1.25)(0.7,1.5)
\psline(0.85,1.25)(1.7,1.5) \vput(0.5,1)0 \psline(0.85,1.25)(0.5,1)
\vput(0.35,1.25)0 \psline(0.85,1.25)(0.35,1.25) \vput(1.2,1.5)0
\psline(0.85,1.25)(1.2,1.5) \vput(1.35,1.25)0
\psline(0.85,1.25)(1.35,1.25) \vput(1.35,0.75)0
\psline(1.35,0.75)(1,1) \psline(1.35,0.75)(1,0)
\psline(1.35,0.75)(1.7,0.5) \psline(1.35,0.75)(1.7,1.5)
\vput(1,0.5)0 \psline(1.35,0.75)(1,0.5) \vput(1.35,0.25)0
\psline(1.35,0.75)(1.35,0.25) \vput(1.7,1)0
\psline(1.35,0.75)(1.7,1) \vput(1.35,1.25)0
\psline(1.35,0.75)(1.35,1.25)
\end{pspicture}
\caption{$C_8$, $\xi_1(C_8)$ and $\xi_2(C_8)$.}
\label{fig:cube}
\end{figure}
\begin{figure}
\centering
\psset{unit=.8cm}
\begin{pspicture}(-2,-2)(2,2) \vput(1.14127,0.37082)0 \vput(0,1.2)0
\vput(-1.14127,0.37082)0 \vput(-0.705342,-0.97082)0
\vput(0.705342,-0.97082)0
\pspolygon(1.14127,0.37082)(0,1.2)(-1.14127,0.37082)(-0.705342,-0.97082)(0.705342,-0.97082)
\vput(1.80701,0.587132)0 \vput(1.11679,1.53713)0 \vput(0,1.9)0
\vput(-1.11679,1.53713)0 \vput(-1.80701,0.587132)0
\vput(-1.80701,-0.587132)0 \vput(-1.11679,-1.53713)0 \vput(0,-1.9)0
\vput(1.11679,-1.53713)0 \vput(1.80701,-0.587132)0
\pspolygon(1.80701,0.587132)(1.11679,1.53713)(0,1.9)(-1.11679,1.53713)(-1.80701,0.587132)(-1.80701,-0.587132)(-1.11679,-1.53713)(0,-1.9)(1.11679,-1.53713)(1.80701,-0.587132)
\psline(1.14127,0.37082)(1.80701,0.587132) \psline(0,1.2)(0,1.9)
\psline(-1.14127,0.37082)(-1.80701,0.587132)
\psline(-0.705342,-0.97082)(-1.11679,-1.53713)
\psline(0.705342,-0.97082)(1.11679,-1.53713)
\end{pspicture}
\begin{pspicture}(-2,-2)(2,2) \vput(1.14127,0.37082)0 \vput(0,1.2)0
\vput(-1.14127,0.37082)0 \vput(-0.705342,-0.97082)0
\vput(0.705342,-0.97082)0
\pspolygon(1.14127,0.37082)(0,1.2)(-1.14127,0.37082)(-0.705342,-0.97082)(0.705342,-0.97082)
\vput(1.80701,0.587132)0 \vput(1.11679,1.53713)0 \vput(0,1.9)0
\vput(-1.11679,1.53713)0 \vput(-1.80701,0.587132)0
\vput(-1.80701,-0.587132)0 \vput(-1.11679,-1.53713)0 \vput(0,-1.9)0
\vput(1.11679,-1.53713)0 \vput(1.80701,-0.587132)0
\pspolygon(1.80701,0.587132)(1.11679,1.53713)(0,1.9)(-1.11679,1.53713)(-1.80701,0.587132)(-1.80701,-0.587132)(-1.11679,-1.53713)(0,-1.9)(1.11679,-1.53713)(1.80701,-0.587132)
\psline(1.14127,0.37082)(1.80701,0.587132) \psline(0,1.2)(0,1.9)
\psline(-1.14127,0.37082)(-1.80701,0.587132)
\psline(-0.705342,-0.97082)(-1.11679,-1.53713)
\psline(0.705342,-0.97082)(1.11679,-1.53713) \vput(0,0)0
\psline(0,0)(1.14127,0.37082) \psline(0,0)(0,1.2)
\psline(0,0)(-1.14127,0.37082) \psline(0,0)(-0.705342,-0.97082)
\psline(0,0)(0.705342,-0.97082) \vput(0.822899,1.13262)0
\psline(0.822899,1.13262)(0,1.2)
\psline(0.822899,1.13262)(1.14127,0.37082)
\psline(0.822899,1.13262)(1.80701,0.587132)
\psline(0.822899,1.13262)(1.11679,1.53713)
\psline(0.822899,1.13262)(0,1.9) \vput(-0.822899,1.13262)0
\psline(-0.822899,1.13262)(-1.14127,0.37082)
\psline(-0.822899,1.13262)(0,1.2) \psline(-0.822899,1.13262)(0,1.9)
\psline(-0.822899,1.13262)(-1.11679,1.53713)
\psline(-0.822899,1.13262)(-1.80701,0.587132)
\vput(-1.33148,-0.432624)0
\psline(-1.33148,-0.432624)(-0.705342,-0.97082)
\psline(-1.33148,-0.432624)(-1.14127,0.37082)
\psline(-1.33148,-0.432624)(-1.80701,0.587132)
\psline(-1.33148,-0.432624)(-1.80701,-0.587132)
\psline(-1.33148,-0.432624)(-1.11679,-1.53713) \vput(0,-1.4)0
\psline(0,-1.4)(0.705342,-0.97082)
\psline(0,-1.4)(-0.705342,-0.97082)
\psline(0,-1.4)(-1.11679,-1.53713) \psline(0,-1.4)(0,-1.9)
\psline(0,-1.4)(1.11679,-1.53713) \vput(1.33148,-0.432624)0
\psline(1.33148,-0.432624)(1.14127,0.37082)
\psline(1.33148,-0.432624)(0.705342,-0.97082)
\psline(1.33148,-0.432624)(1.11679,-1.53713)
\psline(1.33148,-0.432624)(1.80701,-0.587132)
\psline(1.33148,-0.432624)(1.80701,0.587132)
\end{pspicture}
\begin{pspicture}(-2,-2)(2,2) \vput(1.14127,0.37082)0 \vput(0,1.2)0
\vput(-1.14127,0.37082)0 \vput(-0.705342,-0.97082)0
\vput(0.705342,-0.97082)0
\pspolygon(1.14127,0.37082)(0,1.2)(-1.14127,0.37082)(-0.705342,-0.97082)(0.705342,-0.97082)
\vput(1.80701,0.587132)0 \vput(1.11679,1.53713)0 \vput(0,1.9)0
\vput(-1.11679,1.53713)0 \vput(-1.80701,0.587132)0
\vput(-1.80701,-0.587132)0 \vput(-1.11679,-1.53713)0 \vput(0,-1.9)0
\vput(1.11679,-1.53713)0 \vput(1.80701,-0.587132)0
\pspolygon(1.80701,0.587132)(1.11679,1.53713)(0,1.9)(-1.11679,1.53713)(-1.80701,0.587132)(-1.80701,-0.587132)(-1.11679,-1.53713)(0,-1.9)(1.11679,-1.53713)(1.80701,-0.587132)
\psline(1.14127,0.37082)(1.80701,0.587132) \psline(0,1.2)(0,1.9)
\psline(-1.14127,0.37082)(-1.80701,0.587132)
\psline(-0.705342,-0.97082)(-1.11679,-1.53713)
\psline(0.705342,-0.97082)(1.11679,-1.53713) \vput(0,0)0
\psline(0,0)(1.14127,0.37082) \psline(0,0)(0,1.2)
\psline(0,0)(-1.14127,0.37082) \psline(0,0)(-0.705342,-0.97082)
\psline(0,0)(0.705342,-0.97082) \vput(0.570634,0.78541)0
\psline(0,0)(0.570634,0.78541) \vput(-0.570634,0.78541)0
\psline(0,0)(-0.570634,0.78541) \vput(-0.923305,-0.3)0
\psline(0,0)(-0.923305,-0.3) \vput(0,-0.97082)0
\psline(0,0)(0,-0.97082) \vput(0.923305,-0.3)0
\psline(0,0)(0.923305,-0.3) \vput(0.822899,1.13262)0
\psline(0.822899,1.13262)(0,1.2)
\psline(0.822899,1.13262)(1.14127,0.37082)
\psline(0.822899,1.13262)(1.80701,0.587132)
\psline(0.822899,1.13262)(1.11679,1.53713)
\psline(0.822899,1.13262)(0,1.9) \vput(0.570634,0.78541)0
\psline(0.822899,1.13262)(0.570634,0.78541) \vput(1.47414,0.478976)0
\psline(0.822899,1.13262)(1.47414,0.478976) \vput(1.4619,1.06213)0
\psline(0.822899,1.13262)(1.4619,1.06213) \vput(0.558396,1.71857)0
\psline(0.822899,1.13262)(0.558396,1.71857) \vput(0,1.55)0
\psline(0.822899,1.13262)(0,1.55) \vput(-0.822899,1.13262)0
\psline(-0.822899,1.13262)(-1.14127,0.37082)
\psline(-0.822899,1.13262)(0,1.2) \psline(-0.822899,1.13262)(0,1.9)
\psline(-0.822899,1.13262)(-1.11679,1.53713)
\psline(-0.822899,1.13262)(-1.80701,0.587132)
\vput(-0.570634,0.78541)0
\psline(-0.822899,1.13262)(-0.570634,0.78541) \vput(0,1.55)0
\psline(-0.822899,1.13262)(0,1.55) \vput(-0.558396,1.71857)0
\psline(-0.822899,1.13262)(-0.558396,1.71857)
\vput(-1.4619,1.06213)0 \psline(-0.822899,1.13262)(-1.4619,1.06213)
\vput(-1.47414,0.478976)0
\psline(-0.822899,1.13262)(-1.47414,0.478976)
\vput(-1.33148,-0.432624)0
\psline(-1.33148,-0.432624)(-0.705342,-0.97082)
\psline(-1.33148,-0.432624)(-1.14127,0.37082)
\psline(-1.33148,-0.432624)(-1.80701,0.587132)
\psline(-1.33148,-0.432624)(-1.80701,-0.587132)
\psline(-1.33148,-0.432624)(-1.11679,-1.53713)
\vput(-0.923305,-0.3)0 \psline(-1.33148,-0.432624)(-0.923305,-0.3)
\vput(-1.47414,0.478976)0
\psline(-1.33148,-0.432624)(-1.47414,0.478976) \vput(-1.80701,0)0
\psline(-1.33148,-0.432624)(-1.80701,0) \vput(-1.4619,-1.06213)0
\psline(-1.33148,-0.432624)(-1.4619,-1.06213)
\vput(-0.911067,-1.25398)0
\psline(-1.33148,-0.432624)(-0.911067,-1.25398) \vput(0,-1.4)0
\psline(0,-1.4)(0.705342,-0.97082)
\psline(0,-1.4)(-0.705342,-0.97082)
\psline(0,-1.4)(-1.11679,-1.53713) \psline(0,-1.4)(0,-1.9)
\psline(0,-1.4)(1.11679,-1.53713) \vput(0,-0.97082)0
\psline(0,-1.4)(0,-0.97082) \vput(-0.911067,-1.25398)0
\psline(0,-1.4)(-0.911067,-1.25398) \vput(-0.558396,-1.71857)0
\psline(0,-1.4)(-0.558396,-1.71857) \vput(0.558396,-1.71857)0
\psline(0,-1.4)(0.558396,-1.71857) \vput(0.911067,-1.25398)0
\psline(0,-1.4)(0.911067,-1.25398) \vput(1.33148,-0.432624)0
\psline(1.33148,-0.432624)(1.14127,0.37082)
\psline(1.33148,-0.432624)(0.705342,-0.97082)
\psline(1.33148,-0.432624)(1.11679,-1.53713)
\psline(1.33148,-0.432624)(1.80701,-0.587132)
\psline(1.33148,-0.432624)(1.80701,0.587132) \vput(0.923305,-0.3)0
\psline(1.33148,-0.432624)(0.923305,-0.3) \vput(0.911067,-1.25398)0
\psline(1.33148,-0.432624)(0.911067,-1.25398)
\vput(1.4619,-1.06213)0 \psline(1.33148,-0.432624)(1.4619,-1.06213)
\vput(1.80701,0)0 \psline(1.33148,-0.432624)(1.80701,0)
\vput(1.47414,0.478976)0
\psline(1.33148,-0.432624)(1.47414,0.478976)
\end{pspicture}
\caption{$D_{20}$, $\xi_1(D_{20})$ and $\xi_2(D_{20})$.}
\label{fig:dodeca}
\end{figure}
Let $P$ be a simplicial polytope. The \emph{type} of a face $F$ of $P$ is
defined to be $\type(F)=(x,y,z)$, where $x,y,z$ are the degrees of the three
vertices of $F$ with $x\geq y\geq z$. If all faces of $P$ have the same type,
then $P$ is called \emph{face-transitive}\footnote{The usual definition is that
$P$ is face-transitive if for any two faces $F_1$ and $F_2$ of $P$ there is an
automorphism on $P$ sending $F_1$ to $F_2$. It is not difficult to see that
our definition is equivalent to this.} In this case, we define $\type(P)$ to
be the type of a face of $P$. If $P$ is face-transitive and $\type(P)=(x,x,x)$
for an integer $x$, then $P$ is called \emph{regular}. Note that $T_4$, $O_6$
and $I_{12}$ are the only regular simplicial polytopes.
\begin{lem}\label{lem:face-transitive}
If $|\mathcal{C}(P\#Q)|=1$ for irreducible polytopes $P$ and $Q$, then one of $P$, $Q$
is regular and the other is face-transitive.
\end{lem}
\begin{proof}
If $P$ is not face-transitive, then $P$ has two faces $F_1$, $F_2$ with
different types. Let $F$ be any face of $Q$. Let $P_1$ (resp.~$P_2$) be a
polytope in $\mathcal{C}(P\#Q)$ obtained by identifying $F_1$ (resp.~$F_2$) with
$F$. Then $P_1$ and $P_2$ can not be the same, which is a contradiction to
$|\mathcal{C}(P\#Q)|=1$. Thus $P$ is face-transitive and so is $Q$ by the same
argument.
Let $\type(P)=(x,y,z)$ and $\type(Q)=(a,b,c)$. We can identify a face $F'$ of
$P$ with a face $F''$ of $Q$ in the following two ways: identify the vertices
of degree $x,y,z$ in $F'$ with (1) the vertices of degree $a,b,c$ in $F''$ and
(2) the vertices of degree $c,b,a$ in $F''$ respectively. Then the resulting
polytope has a unique $3$-belt with vertices of degree $x+a-2, y+b-2, z+c-2$
in the first case and $x+c-2,y+b-2,z+a-2$ in the second case. Since two
polytopes are the same, we have $x+a-2=x+c-2$ or $x+a-2=z+a-2$. Thus $a=c$ or
$x=z$. Since $a\geq b\geq c$ and $x\geq y\geq z$, we have $a=b=c$ or $x=y=z$,
which implies that either $P$ or $Q$ is regular.
\end{proof}
\begin{figure}
\centering
\psset{unit=20pt}
\begin{pspicture}(-2,-2)(5,2)
\rput(3.5,0){$\stackrel{dual}{\longleftrightarrow}$}
\vput(1.9563,0.792088){x0} \vput(1.9563,-1.20791){y0} \vput(1.33826,0.256855){x1} \vput(1.33826,-1.74314){y1} \vput(0.209057,0.0054781){x2} \vput(0.209057,-1.99452){y2} \vput(-1,0.133975){x3} \vput(-1,-1.86603){y3} \vput(-1.82709,0.593263){x4} \vput(-1.82709,-1.40674){y4} \vput(-1.9563,1.20791){x5} \vput(-1.9563,-0.792088){y5} \vput(-1.33826,1.74314){x6} \vput(-1.33826,-0.256855){y6} \vput(-0.209057,1.99452){x7} \vput(-0.209057,-0.0054781){y7} \vput(1,1.86603){x8} \vput(1,-0.133975){y8} \vput(1.82709,1.40674){x9} \vput(1.82709,-0.593263){y9} \edge{x0}{x1} \edge{x1}{x2} \edge{x2}{x3} \edge{x3}{x4} \edge{x4}{x5} \edge{x5}{x6} \edge{x6}{x7} \edge{x7}{x8} \edge{x8}{x9} \edge{x9}{x0} \edge{y0}{y1} \edge{x0}{y0} \edge{y1}{y2} \edge{x1}{y1} \edge{y2}{y3} \edge{x2}{y2} \edge{y3}{y4} \edge{x3}{y3} \edge{y4}{y5} \edge{x4}{y4} \edge{x5}{y5} \edge{x0}{y0} \psset{linestyle=dotted}\edge{y5}{y6} \edge{x5}{y5} \edge{y6}{y7} \edge{x6}{y6} \edge{y7}{y8} \edge{x7}{y7} \edge{y8}{y9} \edge{x8}{y8} \edge{y9}{y0} \edge{x9}{y9}
\end{pspicture}
\begin{pspicture}(-2,-2)(2,2)
\vput(0,2){x} \vput(0,-2){y} \vput(2,0){0} \vput(1.61803,-0.587785){1} \vput(0.618034,-0.951057){2} \vput(-0.618034,-0.951057){3} \vput(-1.61803,-0.587785){4} \vput(-2,0){5} \vput(-1.61803,0.587785){6} \vput(-0.618034,0.951057){7} \vput(0.618034,0.951057){8} \vput(1.61803,0.587785){9} \edge{8}{-1} \edge{9}{0} \edge{0}{1} \edge{1}{2} \edge{2}{3} \edge{3}{4} \edge{4}{5} \edge{5}{6} \edge{x}{1} \edge{y}{1} \edge{x}{2} \edge{y}{2} \edge{x}{3} \edge{y}{3} \edge{x}{4} \edge{y}{4} \edge{9}{x} \edge{5}{x} \edge{0}{x} \edge{6}{x} \psset{linestyle=dotted} \edge{4}{y} \edge{5}{y} \edge{6}{y} \edge{7}{y} \edge{8}{y} \edge{9}{y} \edge{0}{y} \edge{1}{y} \edge{7}{x} \edge{6}{7} \edge{8}{x} \edge{7}{8} \edge{9}{x} \edge{8}{9}
\end{pspicture}
\caption{A prism and a bipyramid. They are dual to each other.}
\label{fig:prism}
\end{figure}
A \emph{prism} is the product of an $n$-gon and an interval. A \emph{bipyramid}
is the dual of a prism. Let $B_n$ denote the bipyramid with $n$ vertices. See
Figure~\ref{fig:prism}.
Fleischner and Imrich \cite[Theorem 3]{Fleischner1979} classified all
face-transitive $3$-dimensional polytopes. The following is a consequence of
their result.
\begin{prop}
Let $P$ be a face-transitive simplicial polytope. Then $P$ is either
a bipyramid, a Platonic solid, the first-subdivision of a Platonic
solid or the second-subdivision of a Platonic solid.
\end{prop}
One can check that $\xi_2(T_4) = \xi_1(C_8)$, $\xi_2(C_8) = \xi_2(O_6)$ and
$\xi_2(D_{20})=\xi_2(I_{12})$. Note that $\xi_1(T_4)$, $\xi_1(O_6)$ and
$\xi_1(I_{12})$ are reducible. Thus we get Theorem~\ref{thm:necessary}.
\section{The maximum of $b_{n-4}(P)$ for irreducible polytopes} \label{sec:max}
Let $P$ be a polytope with $n$ vertices. Since $b_{n-3}(P)$ is the number of
$3$-belts, we get the following.
\begin{prop}
Let $P_1,\ldots,P_\ell$ be irreducible polytopes and let
$P\in\mathcal{C}(P_1\#\cdots\#P_\ell)$. If $P$ has $n$ vertices, then
$b_{n-3}(P)=\ell-1$.
\end{prop}
We will first find an upper bound of $b_{n-4}(P)$ for an irreducible polytope
$P$ with $n$ vertices, which is established when $P=B_n$, the bipyramid with
$n$ vertices. Our first step is to find $b_k(B_n)$.
\begin{prop}
For $k\leq n-3$, we have
$$b_k(B_n) =\frac{(n-2)(k-1)}{n-2-k}\binom{n-4}{k}+\delta_{k,2}.$$
\end{prop}
\begin{proof}
Observe that $B_n$ is the graph obtained from an $(n-2)$-gon by adding two
vertices $v$, $u$ connected to all vertices of the $(n-2)$-gon. Let $W$ be a
set of $k$ vertices. If $v\in W$ or $u\in W$, then $B_n|_W$ is connected
unless $k=2$ and $W=\{u,v\}$. Thus it is sufficient to consider the vertices
in the $(n-2)$-gon. Then it follows from the result in
\cite[Example~2.1.(c)]{Br-Hi-1998}; see also \cite[Corollary~3.7]{Ch-Ki-2010}.
\end{proof}
\begin{table}
\psset{unit=4pt}
\centering
\begin{tabular}{|c|c|c|c|c|c|c|c|c|c|c|} \hline
$n$ & $4$ & $6$ & $7$ &
\multicolumn{2}{c|}{$8$} &
\multicolumn{5}{c|}{$9$} \\
\hline
$P^*$
& \raisebox{-9pt}{\begin{pspicture}(0,-1)(6,6.5)
\rput{180}(3,2.4){\rput(-3,-3){\pspolygon(3,0)(0.4,4.5)(5.6,4.5) \psline(0.4,4.5)(3,3)
\psline(3,0)(3,3) \psline(5.6,4.5)(3,3)}}
\end{pspicture}}
& \raisebox{-9pt}{\begin{pspicture}(0,-1)(6,6.5)
\pspolygon(5.12,5.12)(0.88,5.12)(0.88,0.88)(5.12,0.88)
\pspolygon(4.06,4.06)(1.94,4.06)(1.94,1.94)(4.06,1.94)
\psline(5.12,5.12)(4.06,4.06) \psline(0.88,5.12)(1.94,4.06)
\psline(0.88,0.88)(1.94,1.94) \psline(5.12,0.88)(4.06,1.94)
\end{pspicture}}
& \raisebox{-9pt}{\begin{pspicture}(0,-1)(6,6.5)
\pspolygon(3,6)(0.15,3.93)(1.24,0.57)(4.76,0.57)(5.85,3.93)
\pspolygon(3,4.5)(1.57,3.46)(2.12,1.79)(3.88,1.79)(4.43,3.46)
\psline(3,6)(3,4.5) \psline(0.15,3.93)(1.57,3.46)
\psline(1.24,0.57)(2.12,1.79) \psline(4.76,0.57)(3.88,1.79)
\psline(5.85,3.93)(4.43,3.46)
\end{pspicture}}
& \raisebox{-9pt}{\begin{pspicture}(0,-1)(6,6.5)
\pspolygon(0,3)(1.5,0.4)(4.5,0.4)(6,3)(4.5,5.6)(1.5,5.6)
\pspolygon(1.5,3)(2.25,1.7)(3.75,1.7)(4.5,3)(3.75,4.3)(2.25,4.3)
\psline(0,3)(1.5,3) \psline(1.5,0.4)(2.25,1.7)
\psline(4.5,0.4)(3.75,1.7) \psline(6,3)(4.5,3)
\psline(4.5,5.6)(3.75,4.3) \psline(1.5,5.6)(2.25,4.3)
\end{pspicture}}
& \raisebox{-9pt}{\begin{pspicture}(0,-1)(6,6.5)
\pspolygon(3,6)(0.15,3.93)(1.24,0.57)(4.76,0.57)(5.85,3.93)
\pspolygon(3,4.5)(1.57,3.46)(2.12,2.09)(3.88,2.09)(4.43,3.46)
\pspolygon(2.12,2.09)(3.88,2.09)(3.88,1.49)(2.12,1.49)
\psline(3,6)(3,4.5) \psline(0.15,3.93)(1.57,3.46)
\psline(1.24,0.57)(2.12,1.49) \psline(4.76,0.57)(3.88,1.49)
\psline(5.85,3.93)(4.43,3.46)
\end{pspicture}}
& \raisebox{-9pt}{\begin{pspicture}(0,-1)(6,6.5)
\pspolygon(3,6)(0.65,4.87)(0.08,2.33)(1.69,0.30)(4.31,0.30)(5.92,2.33)(5.35,4.87)
\pspolygon(3,4.5)(1.83,3.93)(1.54,2.67)(2.35,1.65)(3.65,1.65)(4.46,2.67)(4.17,3.93)
\psline(3,6)(3,4.5) \psline(0.65,4.87)(1.83,3.93)
\psline(0.08,2.33)(1.54,2.67) \psline(1.69,0.30)(2.35,1.65)
\psline(4.31,0.30)(3.65,1.65) \psline(5.92,2.33)(4.46,2.67)
\psline(5.35,4.87)(4.17,3.93)
\end{pspicture}}
& \raisebox{-9pt}{\begin{pspicture}(0,-1)(6,6.5)
\pspolygon(0,3)(1.5,0.4)(4.5,0.4)(6,3)(4.5,5.6)(1.5,5.6)
\pspolygon(1,3)(1.5,3.87)(2,3)(1.5,2.13)
\pspolygon(4,3)(4.5,3.87)(5,3)(4.5,2.13) \psline(2,3)(4,3)
\psline(0,3)(1,3) \psline(5,3)(6,3) \psline(1.5,0.4)(1.5,2.13)
\psline(1.5,5.6)(1.5,3.87) \psline(4.5,0.4)(4.5,2.13)
\psline(4.5,5.6)(4.5,3.87)
\end{pspicture}}
& \raisebox{-9pt}{\begin{pspicture}(0,-1)(6,6.5)
\pspolygon(0,3)(1.5,0.4)(4.5,0.4)(6,3)(4.5,5.6)(1.5,5.6)
\pspolygon(1.5,3)(2.25,2)(3.75,2)(4.5,3)(3.75,4.3)(2.25,4.3)
\pspolygon(2.25,1.4)(3.75,1.4)(3.75,2)(2.25,2) \psline(0,3)(1.5,3)
\psline(1.5,0.4)(2.25,1.4) \psline(4.5,0.4)(3.75,1.4)
\psline(6,3)(4.5,3) \psline(4.5,5.6)(3.75,4.3)
\psline(1.5,5.6)(2.25,4.3)
\end{pspicture}}
& \raisebox{-9pt}{\begin{pspicture}(0,-1)(6,6.5)
\pspolygon(0,3)(1.5,0.4)(4.5,0.4)(6,3)(4.5,5.6)(1.5,5.6)
\pspolygon(1.5,3)(2.25,1.7)(3,2)(3.75,1.7)(4.5,3)(3.75,4.3)(3,4)(2.25,4.3)
\psline(3,2)(3,4) \psline(0,3)(1.5,3) \psline(1.5,0.4)(2.25,1.7)
\psline(4.5,0.4)(3.75,1.7) \psline(6,3)(4.5,3)
\psline(4.5,5.6)(3.75,4.3) \psline(1.5,5.6)(2.25,4.3)
\end{pspicture}}
& \raisebox{-9pt}{\begin{pspicture}(0,-1)(6,6.5)
\pspolygon(3,6)(0.15,3.93)(1.24,0.57)(4.76,0.57)(5.85,3.93)
\pspolygon(3,4.5)(2.49,3.70)(1.57,3.46)(2.12,1.79)(3,2.24)(3.88,1.79)(4.43,3.46)(3.52,3.66)
\psline(3,6)(3,4.5) \psline(0.15,3.93)(1.57,3.46)
\psline(1.24,0.57)(2.12,1.79) \psline(4.76,0.57)(3.88,1.79)
\psline(5.85,3.93)(4.43,3.46) \psline(3,3)(3,2.24)
\psline(3,3)(2.49,3.70) \psline(3,3)(3.52,3.66)
\end{pspicture}}
\\ \hline
$b_{n-4}(P)$ & $-1$ & $3$ & $5$ & $9$ & $5$ & $14$ & $12$ & $8$ & $6$ & $3$
\\ \hline
\end{tabular}
\caption{The complete list of simple polytopes $P^*$ with $n$
faces
for $n\leq 9$ such that $P$ is an irreducible simplicial polytope
with $n$ vertices,
and the numbers $b_{n-4}(P)$.} \label{table:list}
\end{table}
Choi et al. \cite{Ch-Pa-Su-2010} computed the graded Betti numbers of all
simplicial polytopes with at most $9$ vertices. We need some of their result as
shown in Table~\ref{table:list}.
\begin{thm}\label{thm:irr n-4}
Let $n\geq 4$ be a fixed integer. Let $P$ be an irreducible polytope with $n$
vertices. Then
$$b_{n-4}(P)\leq \binom{n-3}{2}-1 + \delta_{n,6}.$$ The equality
holds if and only if $P=B_n$ or $P=T_4$.
\end{thm}
\begin{proof}
Induction on $n$. By Table~\ref{table:list}, it is true for $n\leq 7$. Assume
that $n\geq 8$ and it is true for all integers less than $n$. Since $P$ is
irreducible, $b_{n-4}(P)$ is the number of $4$-belts of $P$. If there is no
4-belt in $P$, the theorem is true since $b_{n-4}(P)=0$ and $P$ is not a
bipyramid. Otherwise, take a 4-belt $B=\{v_1,v_2,v_3,v_4\}$ such that $v_i$ is
connected to $v_{i+1}$ for $i=1,2,3,4$, where $v_5=v_1$.
Now assume that the graph $P$ is embedded in a plane. Let $V$ be the vertex set
of $P$. There are two connected components in $P|_{V\setminus B}$. Let $X_1$
(resp.~$X_2$) be the set of vertices in the connected component in
$P|_{V\setminus B}$ which is outside (resp.~inside) of the $4$-gon consisting of
the vertices in $B$.
We can assume that $|X_1|>1$ and $|X_2|>1$, because otherwise we can assign the
unique vertex of $X_1$ or $X_2$ to $B$, which implies that $b_{n-4}$ is less
than the number of vertices, and hence, $b_{n-4}(P)\leq n < \binom{n-3}2-1$.
For $i=1,2$, let $P_i$ be the polytope obtained from $P$ by contracting all
vertices in $X_i$ to a single vertex $x_i$. Note that $x_i$ is connected to all
vertices of $B$ in $P_i$.
Let $n_1$ and $n_2$ be the number of vertices of $P_1$ and $P_2$
respectively. Then $n_1+n_2=n+6$ and $n_1,n_2\geq 7$.
For $i=1,2$, let $A_i$ (resp.~$B_i$) be the set of vertices $u$ of $P_i$ such
that $\{x_i,v_1,v_3,u\}$ (resp.~$\{x_i,v_2,v_4,u\}$) is a $4$-belt. We claim
that $A_1 =\emptyset$ or $B_1 =\emptyset$. Assume that both $A_1$ and $B_1$ are
nonempty. Note that $x_1$ is the only vertex in $P_1$ which lies outside of the
$4$-gon $\{v_1,v_2,v_3,v_4\}$. If $u\in A_1$ and $u'\in B_1$, then $\{u,v_1\}$,
$\{u,v_3\}$, $\{u',v_2\}$ and $\{u',v_4\}$ are edges. Since $P_1$ is a planar
graph, we must have $u=u'$. Then the edges $\{u,v_i\}$ for $i=1,2,3,4$ divide
the $4$-gon $\{v_1,v_2,v_3,v_4\}$ into four triangular regions. If we have a vertex inside of the
triangle $\{u,v_i,v_{i+1}\}$, then this forms a $3$-belt of $P_1$, and thus of
$P$. Thus we do not have any vertex except $u$ inside of the $4$-gon
$\{v_1,v_2,v_3,v_4\}$. Then we get $n_1=6$ which is a contradiction. Thus we
have $A_1 =\emptyset$ or $B_1 =\emptyset$. For the same reason, we also have
$A_2 =\emptyset$ or $B_2 =\emptyset$.
Let $a_i=|A_i|$ and $b_i=|B_i|$ for $i=1,2$. Then we have
\begin{align*}
b_{n-4}(P)&=b_{n_1-4}(P_1)+b_{n_1-4}(P_1)-1-a_1-a_2-b_1-b_2+a_1a_2+b_1b_2\\
&=b_{n_1-4}(P_1)+b_{n_1-4}(P_1)+(a_1-1)(a_2-1)+(b_1-1)(b_2-1)-3.
\end{align*}
Since $a_1+b_1\leq n_1-5$ and $a_2+b_2\leq n_2-5$, we have
$(a_1-1)(a_2-1)+(b_1-1)(b_2-1)\leq (n_1-6)(n_2-6)+1$, where the equality holds
if and only if $(a_1,b_1,a_2,b_2)$ is equal to $(n_1-5,0,n_2-5,0)$ or
$(0,n_1-5,0,n_2-5)$. Note that $a_i=n_i-5$ or $b_i=n_i-5$ if and only if $P_i$
is a bipyramid. Moreover, $(a_1,b_1,a_2,b_2)$ is equal to $(n_1-5,0,n_2-5,0)$ or
$(0,n_1-5,0,n_2-5)$ if and only if $P=B_n$.
Since $n_1,n_2<n$, by the
induction hypothesis, we get
\begin{align*}
b_{n-4}(P)&\leq \binom{n_1-3}2-1+\binom{n_2-3}2-1 +(n_1-6)(n_2-6)+1-3\\
&=\binom{n-3}2-1,
\end{align*}
where the equality holds if and only if $P=B_n$.
\end{proof}
Using a similar argument, we can find the second largest value of
$b_{n-4}(P)$ for an irreducible polytope $P$ with $n$ vertices.
\begin{figure}
\centering
\psset{unit=20pt}
\begin{pspicture}(-2,-2)(5,2)
\rput(3.5,0){$\stackrel{dual}{\longleftrightarrow}$}
\vput(1.9563,0.792088){x0} \vput(1.9563,-1.20791){y0}
\vput(1.33826,0.256855){x1} \vput(1.33826,-1.74314){y1}
\vput(0.434898,0.0557535){x2} \vput(-1.16542,0.225833){x3}
\vput(0.209057,-.3445219){xx2} \vput(-1,-.216025){xx3} \edge{xx2}{x2}
\edge{xx3}{x3} \edge{xx3}{xx2} \vput(0.209057,-1.99452){y2}
\vput(-1,-1.86603){y3} \vput(-1.82709,0.593263){x4}
\vput(-1.82709,-1.40674){y4} \vput(-1.9563,1.20791){x5}
\vput(-1.9563,-0.792088){y5} \vput(-1.33826,1.74314){x6}
\vput(-1.33826,-0.256855){y6} \vput(-0.209057,1.99452){x7}
\vput(-0.209057,-0.0054781){y7} \vput(1,1.86603){x8}
\vput(1,-0.133975){y8} \vput(1.82709,1.40674){x9}
\vput(1.82709,-0.593263){y9} \edge{x0}{x1} \edge{x1}{x2} \edge{x2}{x3}
\edge{x3}{x4} \edge{x4}{x5} \edge{x5}{x6} \edge{x6}{x7} \edge{x7}{x8}
\edge{x8}{x9} \edge{x9}{x0} \edge{y0}{y1} \edge{x0}{y0} \edge{y1}{y2}
\edge{x1}{y1} \edge{y2}{y3} \edge{xx2}{y2} \edge{y3}{y4}
\edge{xx3}{y3} \edge{y4}{y5} \edge{x4}{y4} \edge{x5}{y5} \edge{x0}{y0}
\psset{linestyle=dotted}\edge{y5}{y6} \edge{x5}{y5} \edge{y6}{y7}
\edge{x6}{y6} \edge{y7}{y8} \edge{x7}{y7} \edge{y8}{y9} \edge{x8}{y8}
\edge{y9}{y0} \edge{x9}{y9}
\end{pspicture}
\begin{pspicture}(-2,-2)(2,2)
\vput(0,2){x} \vput(0,-2){y} \vput(2,0){0} \vput(1.61803,-0.587785){1}
\vput(0.618034,-0.951057){2} \vput(-0.618034,-0.951057){3}
\vput(-1.61803,-0.587785){4} \vput(-2,0){5}
\vput(-1.61803,0.587785){6} \vput(-0.618034,0.951057){7}
\vput(0.618034,0.951057){8} \vput(1.61803,0.587785){9}
\vput(-.3,.5){z}
\edge zx \edge z4 \edge z3 \edge z2
\edge{8}{-1} \edge{9}{0} \edge{0}{1} \edge{1}{2} \edge{2}{3}
\edge{3}{4} \edge{4}{5} \edge{5}{6} \edge{x}{1} \edge{y}{1}
\edge{x}{2} \edge{y}{2} \edge{y}{3} \edge{x}{4} \edge{y}{4}
\edge{9}{x} \edge{5}{x} \edge{0}{x} \edge{6}{x}
\psset{linestyle=dotted}
\edge{4}{y} \edge{5}{y} \edge{6}{y} \edge{7}{y} \edge{8}{y}
\edge{9}{y} \edge{0}{y} \edge{1}{y} \edge{7}{x} \edge{6}{7}
\edge{8}{x} \edge{7}{8} \edge{9}{x} \edge{8}{9}
\end{pspicture}
\caption{An edge-cut-prism and a semi-bipyramid. They are dual to each other.}
\label{fig:edgecut}
\end{figure}
Let $P$ be a prism which is a product of a $k$-gon and an interval. Let $e$ be
an edge of one of the two $k$-gons of $P$. Then we can obtain another simple
polytope from $P$ by `cutting' the edge $e$. We will call such a polytope an
\emph{edge-cut-prism}. A \emph{semi-bipyramid} is the dual of an edge-cut-prism.
See Figure~\ref{fig:edgecut}.
\begin{thm}\label{thm:second max}
Let $P$ be an irreducible polytope with $n$ vertices. If $P\ne B_n$, then
$$b_{n-4}(P)\leq \binom{n-5}{2}+2.$$ The equality holds if and only if $P$
is a semi-bipyramid.
\end{thm}
\begin{proof}
Induction on $n$. If $n\leq 9$, then by Table~\ref{table:list}, it holds.
Assume that $n\geq 10$ and it holds for all integers less than $n$. We will
use the same notations in the proof of Theorem~\ref{thm:irr n-4}.
\begin{description}
\item[Case 1] $a_i= n_i-5$ or $b_i= n_i-5$ for $i=1,2$. Then $P_1$ and $P_2$
are bipyramids. Recall that if $(a_1,b_1,a_2,b_2)$ is equal to
$(n_1-5,0,n_2-5,0)$ or $(0,n_1-5,0,n_2-5)$, then $P=B_n$. Thus
$(a_1,b_1,a_2,b_2)$ must be equal to $(n_1-5,0,0,n_2-5)$ or
$(0,n_1-5,n_2-5,0)$. Hence,
\begin{align*}
b_{n-4}(P) &= b_{n_1-4}(P_1)+b_{n_1-4}(P_1)-1-a_1-a_2-b_1-b_2+a_1a_2+b_1b_2\\
&= \binom{n_1-3}2 -1+\binom{n_2-3}2 -1 -1 -(n_1-5)-(n_2-5)\\
&=\binom{n_1-3}2 +\binom{n_2-3}2 -n+1.
\end{align*}
Recall that $n\geq 10$, $n_1,n_2\geq7$ and $n_1+n_2=n+6$. It is easy to check
that if $n_1=7$ or $n_2=7$, then $P$ is a semi-bipyramid and
$b_{n-4}(P)=\binom{n-5}{2}+2$. Otherwise $\binom{n_1-3}2 +\binom{n_2-3}2
-n+1\leq \binom{n-5}{2}+\binom{5}{2} -n+1< \binom{n-5}{2}+2$ because $n\geq10$.
\item[Case 2] Otherwise. We can assume that $1\leq a_1< n_1-5$. We claim that
$a_1\ne n_1-6$. For contradiction, suppose $a_1 = n_1-6$. Let
$A_1=\{u_1,u_2,\ldots,u_{n_1-6}\}$ and assume that in the embedding of $P_1$,
the $4$-gon $\{v_1,v_2,v_3,v_4\}$ is divided into the $4$-gons
$\{v_1,u_i,v_3,u_{i+1}\}$ for $i=0,1,\ldots,n_1-6$, where $u_0=v_2$ and
$u_{n_1-5}=v_4$. Let $w$ be the unique vertex inside of the $4$-gon
$\{v_1,v_2,v_3,v_4\}$ which is not contained in $A_1$. Then $w$ is contained
in the $4$-gon $\{v_1,u_i,v_3,u_{i+1}\}$ for some $i$. If $\{u_i,u_{i+1}\}$ is
an edge, then $\{v_1,u_i,u_{i+1}\}$ or $\{v_3,u_i,u_{i+1}\}$ is a
$3$-belt. Thus $\{u_i,w\}$ and $\{u_{i+1},w\}$ are edges. Since $P_1$ is
simplicial, $\{v_1,w\}$ and $\{v_3,w\}$ are also edges. Then
$\{v_1,w,v_3,x_1\}$ is a $4$-belt of $P_1$, which is a contradiction to
$w\not\in A_1$. Thus we must have $1\leq a_1\leq n_1-7$. Using a similar
argument, we can check that $a_1=n_1-7$ if and only if $P_1$ is a
semi-bipyramid.
Since $P_1$ is not a bipyramid and $n_1<n$, by the induction
hypothesis, we get
\begin{align*}
b_{n-4}(P) &=b_{n_1-4}(P_1)+b_{n_1-4}(P_1)-1-a_1-a_2-b_1-b_2+a_1a_2+b_1b_2\\
&\leq b_{n_1-4}(P_1)+b_{n_2-4}(P_2)-1- a_1-a_2+a_1a_2\\
&= b_{n_1-4}(P_1)+b_{n_2-4}(P_2)+(a_1-1)(a_2-1)-2\\
&\leq \binom{n_1-5}{2}+2+\binom{n_2-3}{2}-1+(n_1-8)(n_2-6)-2\\
&=\binom{n-5}{2}+2,
\end{align*}
\end{description}
where the equality holds if and only if $a_1=n_1-7$ and $a_2=n_1-5$,
equivalently, $P$ is a semi-bipyramid.
\end{proof}
Now we will find the maximum of $b_{n-4}(P)$ when $P$ is a connected sum of
$\ell$ irreducible polytopes.
Let $P\in\mathcal{C}(P_1\# P_2)$ be a polytope with $n$ vertices, where $P_1$ and $P_2$
are polytopes with $n_1$ and $n_2$ vertices respectively. Additionally, we
assume that $P_2$ is irreducible. Note that $n=n_1+n_2-3$. By
\eqref{thm:connected_sum}, we have
$$b_{n-4} (P) = \sum_{i=0}^{n-4} \left( b_i (P_1) \binom{n_2 -
3}{n-4-i} + b_i (P_2) \binom{n_1 - 3}{n-4-i}\right) + \binom{n-3}{n-4}.$$
Since $\binom{n_2 - 3}{n-4-i}=0$ unless $i\geq n_1-4$ and $\binom{n_1 -
3}{n-4-i}=0$ unless $i\geq n_2-4$, we get
\begin{equation}\label{eq:n-4}
b_{n-4}(P)=b_{n_1-4}(P_1)+b_{n_2-4}(P_2)+(n_2-3)b_{n_1-3}(P_1) +(n-3).
\end{equation}
\begin{lem}\label{thm:mainlemma}
Let $P\in\mathcal{C}(P_1\#\cdots \# P_\ell)$, where $P_i$ is an irreducible polytope
with $n_i$ vertices. Let $n$ be the number of vertices of $P$. Then
$$b_{n-4}(P) = \sum_{i=1}^{\ell} b_{n_i -4}(P_i) + (n-3)(\ell-1).$$
\end{lem}
\begin{proof}
Induction on $\ell$. If $\ell=1$ then it is clear. Let $\ell\geq2$ and assume
that it holds for all integers less than $\ell$.
Let $P\in\mathcal{C}(P_1\#\cdots\#P_\ell)$. Then $P\in\mathcal{C}(Q\# P_\ell)$ for
some $Q\in\mathcal{C}(P_1\#\cdots\#P_{\ell-1})$. Let $n'$ be the number of
vertices of $Q$. Then by \eqref{eq:n-4},
$$b_{n-4}(P)=b_{n'-4}(Q)+b_{n_\ell-4}(P_\ell)+(n_\ell-3)b_{n'-3}(Q)+(n-3).$$
By the induction hypothesis,
$$b_{n'-4}(Q)=\sum_{i=1}^{\ell-1}b_{n_i-4}(P_i)+(n'-3)(\ell-2).$$ Since
$b_{n'-3}(Q)=\ell-2$ and $(n'-3)+(n_\ell-3)=n-3$, we get
$$b_{n-4}(P)=\sum_{i=1}^{\ell}b_{n_i-4}(P_i)+(n-3)(\ell-1).$$
\end{proof}
For an integer $n$, let $f(n)=\binom{n-3}{2}-1+\delta_{n,6}$. Thus $f(n)$ is the
maximum of $b_{n-4}(P)$ as shown in Theorem~\ref{thm:irr n-4}.
\begin{lem}\label{thm:flemma}
Let $m\geq n>4$. Then we have $f(m)+f(n)<f(m+2)+f(n-2)$, if $(m,n)=(6,6)$,
and $f(m)+f(n)<f(m+1)+f(n-1)$, otherwise.
\end{lem}
\begin{proof}
If $(m,n)=(6,6)$, then it is clear. Otherwise,
\begin{align*}
f(m+1)+f(n-1)-f(m)-f(n) &= (m-3)-(n-4)+\delta_{m+1,6}+\delta_{n-1,6}
-\delta_{m,6}-\delta_{n,6}\\
&\geq (m-n)+1 -\delta_{m,6}-\delta_{n,6}\geq1.
\end{align*}
\end{proof}
\begin{lem}\label{thm:fflemma}
Let $n,\ell$ be fixed integers. Let $n_1,\ldots,n_\ell$ be integers satisfying
$n_i\geq 4$ for all $i$ and $\sum_{i=1}^\ell n_i -3(\ell-1)=n$. Then
$$f(n_1)+\cdots+f(n_\ell)\leq \binom{n-2}{2}
-n\ell + \frac{\ell(\ell+3)}{2} +\delta_{\ell,n-5},$$ where the equality holds
if and only if for some $j$, $n_j=n-\ell+1$ and $n_i=4$ for $i\ne j$.
\end{lem}
\begin{proof}
The equality condition is straightforward to check.
Since the number of sequences $n_1,\ldots,n_\ell$ satisfying the conditions are
finite, there exists a sequence such that $f(n_1)+\cdots+f(n_\ell)$ is maximal.
Thus it is sufficient to show that if there are two integers $i,j$ with
$n_i\geq n_j>4$, then there is a sequence $n_1',n_2',\ldots,n_\ell'$
satisfying the conditions and $f(n_1)+\cdots+f(n_\ell)<
f(n_1')+\cdots+f(n_\ell')$. By Lemma~\ref{thm:flemma}, we can obtain such a
sequence by replacing $n_i$ and $n_j$ with $n_i+1$ and $n_j-1$ (or $n_i+2$ and
$n_j-2$ if $n_i=n_j=6$).
\end{proof}
The following is the main theorem of this section.
\begin{thm}\label{thm:n-4 ell}
Let $P$ be a connected sum of $\ell$ irreducible polytopes. Let $n$ be the
number of vertices of $P$. If $\ell \leq n-4$, then
$$b_{n-4}(P) \leq \binom{n-3}{2} + \frac{\ell(\ell-3)}2
+\delta_{\ell,n-5},$$ where the equality holds if and only if $P$ is a
connected sum of a bipyramid and $\ell-1$ tetrahedrons.
\end{thm}
\begin{proof}
Let $P=P_1 \# \cdots \# P_\ell$. Let $P_i$ have $n_i$ vertices. We can assume
$n_1\geq\cdots\geq n_\ell\geq 4$. By Theorem~\ref{thm:irr n-4},
Lemma~\ref{thm:mainlemma} and Lemma~\ref{thm:fflemma}, we have
\begin{align*}
b_{n-4}(P) &= \sum_{i=1}^{\ell} b_{n_i -4}(P_i) + (n-3)(\ell-1)\\ &\leq
\sum_{i=1}^{\ell} f(n_i)+(n-3)(\ell-1)\\
&\leq \binom{n-2}{2} -n\ell
+ \frac{\ell(\ell+3)}{2} +\delta_{\ell,n-5} + (n-3)(\ell-1)\\
&=\binom{n-3}{2} + \frac{\ell(\ell-3)}2 +\delta_{\ell,n-5}.
\end{align*}
The equality holds if and only if $P_1$ is a bipyramid and $P_i$ is the
tetrahedron for $i>1$.
\end{proof}
\section{Combinatorial rigidity of simplicial polytopes}
\label{sec:rigid}
Recall that two polytopes $P$ and $Q$ are combinatorially rigid if
$b_k(P)=b_k(Q)$ for all $k$ implies $P=Q$.
If $b_k(P)=b_k(Q)$ for all $k$, then the numbers of vertices of $P$ and $Q$ are the
same. In fact, the number of vertices of $P$ is determined by $b_2(P)$ as follows.
\begin{prop}\label{thm:b2}
Let $P$ be a polytope with $n$ vertices. Then
$$n=\frac{7+\sqrt{8 \cdot b_2(P)+1}}2.$$
\end{prop}
\begin{proof}
Let $e$ (resp.~$f$) be the number of edges (resp.~faces) of $P$. Then by the
Euler's theorem, we have $n-e+f=2$. Since $P$ is simplicial, we have
$2e=3f$. Thus $e=3n-6$. Observe that $b_2(P)$ is the number of ways of
choosing two vertices of $P$ which are not connected by an edge. Thus
$b_2(P)=\binom{n}{2}-e=\binom{n}{2}-3n+6$. Solving this equation, we get the
theorem.
\end{proof}
\begin{prop}\label{thm:rigid prism}
A bipyramid and a semi-bipyramid are rigid.
\end{prop}
\begin{proof}
Let $P$ be a polytope with $b_k(P)=b_k(B_n)$ for all $n$. By
Proposition~\ref{thm:b2}, $P$ has $n$ vertices. Since
$b_{n-3}(P)=b_{n-3}(B_n)=0$, $P$ is irreducible. By Theorem~\ref{thm:irr n-4}
with $b_{n-4}(P)=b_{n-4}(B_n)$, we get $P=B_n$. Thus $B_n$ is rigid. We can
prove the rigidity of a semi-bipyramid in the same way using
Theorem~\ref{thm:second max}.
\end{proof}
Theorem~\ref{thm:rigid} follows from the following propositions. Note that
$T_4\#T_4$ is rigid because it is the only polytope with $5$ vertices. Note also
that $O_6=B_6$.
\begin{prop}\label{thm:tet}
For $n\geq6$, the connected sum $T_4\# B_n$ is rigid.
\end{prop}
\begin{proof}
This is an immediate consequence of Theorem~\ref{thm:n-4 ell}.
\end{proof}
\begin{prop}\label{thm:cube}
For $n\geq6$, the connected sum $O_6\# B_n$ is rigid.
\end{prop}
\begin{proof}
Let $n'$ be the number of vertices of $O_6\# B_n$, i.e., $n'=n+3$. Let $P$
be a polytope with $b_k(P)=b_k(O_6\# B_n)$ for all $k$. Since
$b_{n'-3}(P)=b_{n'-3}(O_6\# B_n)=1$,
we have $P\in\mathcal{C}(P_1\#P_2)$ for some
irreducible polytopes $P_1$ and $P_2$. Let $n_1$ and $n_2$ be the number of
vertices of $P_1$ and $P_2$ respectively. Then $n_1+n_2=n+6$. Assume $n_1\geq
n_2$. Since $b_{n'-4}(P)=b_{n'-4}(O_6\# B_n)$, by Lemma~\ref{thm:mainlemma},
we have $b_{n_1-4}(P_1)+b_{n_2-4}(P_2)=b_{6-4}(O_6)+b_{n-4}(B_n)$. Since
$b_{6-4}(O_6)+b_{n-4}(B_n) = f(6)+f(n)$ and $b_{n_1-4}(P_1)+b_{n_2-4}(P_2)
\leq f(n_1)+f(n_2)$, by Lemma~\ref{thm:flemma}, we have $n_2\leq 6$.
If $n_2=6$, then the equality holds and we get that $P_1=B_n$ and $P_2=O_6$.
Since there is no irreducible simplicial polytope with $5$ vertices, we have
$n_2\ne 5$. Now assume $n_2=4$. Then $P_2=T_4$ and $n_1=n+2$. Since
$b_{n_1-4}(P_1)+b_{n_2-4}(P_2)=b_{6-4}(O_6)+b_{n-4}(B_n)=3+
\binom{n-3}{2}-1+\delta_{n,6}$ and $b_{n_2-4}(P_2)=-1$, we get
\begin{equation}\label{eq:cube pf2}
b_{n_1-4}(P_1)=\binom{n-3}{2}+\delta_{n,6}+3.
\end{equation}
\begin{description}
\item[Case 1] $P_1$ is a bipyramid. Then
$$\binom{n-1}{2}-1+\delta_{n+2,6}=\binom{n-3}{2}+\delta_{n,6}+3.$$ Thus
we have $2n-5=4+\delta_{n,6}-\delta_{n,4}$, which does not have
an integer solution for $n\geq6$.
\item[Case 2] $P_1$ is not a bipyramid. Then by Theorem~\ref{thm:second
max}, we get
$b_{n_1-4}(P_1)\leq \binom{n-3}2 + 2$, which is a contradiction to
\eqref{eq:cube pf2}.
\end{description}
Thus we can only have $P_1=B_n$ and $P_2=O_6$, thus $P=O_6\#B_n$ and $O_6\#B_n$
is rigid.
\end{proof}
\begin{prop}
The connected sum $T_4\# I_{12} $ is rigid.
\end{prop}
\begin{proof}
Let $b_k(P)=b_k(T_4\#I_{12})$ for all $k$. Using the same argument in the
proof of Proposition~\ref{thm:cube}, we have $P=P_1\# P_2$ for some
irreducible polytopes $P_1$ and $P_2$ with $n_1$ and $n_2$ vertices
respectively, and
$b_{n_1-4}(P_1)+b_{n_2-4}(P_2)=b_{4-4}(T_4)+b_{12-4}(I_{12})=-1+0$. Note that
if a polytope $Q$ with $n$ vertices satisfies $b_{n-4}(Q)<0$, then
$Q=T_4$. Moreover, $I_{12}$ is the only polytope with $12$ vertices without
$4$-belts. Thus $\{P_1,P_2\}=\{T_4,I_{12}\}$ and $P=T_4\# I_{12}$.
\end{proof}
\section{Further study}
In this paper, we show that for a $3$-dimensional irreducible simplicial
polytope $P$ with $n$ vertices, $b_{n-4}(P)$ has the maximum when $P=B_n$.
\begin{prob}
Find the maximum of $b_k(P)$ for a $3$-dimensional irreducible simplicial
polytope $P$ with $n$ vertices.
\end{prob}
Since the polytopes in Table~\ref{tab:rigid} are the only possible
$3$-dimensional reducible combinatorially rigid polytopes, the following problem
is solved if we prove the rigidity of those polytopes. Note that except
$I_{12}\# B_n$, the remaining polytopes are finite.
\begin{prob}
Classify all $3$-dimensional reducible combinatorially rigid polytopes.
\end{prob}
\begin{prob}
Generalize the arguments in this paper to arbitrary dimensional simplicial
polytopes.
\end{prob}
\section*{acknowledgement}
The authors would like to thank Young Soo Kwon for suggesting us
Lemma~\ref{lem:face-transitive}.
\bibliographystyle{amsplain}
|
\section{First-Principles Calculations}
Taking the 50/50 (Eu,Ba)TiO$_3$ ordered alloy as our starting point
(Fig.~\ref{th_phonons} inset),
we next calculate its
properties using first-principles. For details of the computations
see the Methods section.
We began by calculating the phonon dispersion for the high symmetry, cubic perovskite
reference structure at a lattice constant of 3.95 \AA\ (chosen, somewhat arbitrarily,
for this first step because it is the average
of the experimental BaTiO$_3$ and EuTiO$_3$ lattice constants), with the magnetic spins
aligned ferromagnetically; our results are shown in Fig.~\ref{th_phonons}, plotted
along the high symmetry lines of the Brillouin zone.
Importantly we find a polar $\Gamma$-point instability with an imaginary frequency of
103$i$~cm$^{-1}$
which is dominated by relative oxygen -- Ti/Eu displacements (the eigenmode displacements for
Eu, Ba, Ti, O$_{\parallel}$ and O$_{\perp}$ are 0.234, -0.059, 0.394, -0.360 and -0.303
respectively); such polar instabilities are indicative
of a tendency to ferroelectricity. The zone boundary rotational instabilities that often occur in
perovskite oxides and lead to non-polar, antiferrodistortive ground states are notably absent
(in fact the flat bands at $\sim$60 cm$^{-1}$ are stable rotational vibrations).
Interestingly we find
that the Eu ions have a significant amplitude in the soft-mode eigenvector, in contrast
to the Ba ions both here and in the parent BaTiO$_3$.
\begin{figure}
\begin{center}
\resizebox{0.9\columnwidth}{!}{\includegraphics{phonon_spectrum.pdf} }
\end{center}
\caption{
Calculated phonon dispersion of ferromagnetic Eu$_{0.5}$Ba$_{0.5}$TiO$_3$
in its high symmetry reference structure with pseudo-cubic lattice constant $a_0=3.95$ \AA.
The imaginary frequency polar phonon at $\Gamma$ indicates a structural instability
to a ferroelectric phase.
The inset shows the supercell of the ferromagnetic Eu$_{0.5}$Ba$_{0.5}$TiO$_3$ ordered alloy used
in our calculations. The Ti and O ions are omitted for clarity; arrows represent the Eu magnetic
moments.
\label{th_phonons}}
\end{figure}
Next we performed a structural optimization of both the unit cell shape and the
ionic positions of our Eu$_{0.5}$Ba$_{0.5}$TiO$_3$ alloy with the total volume constrained
to that of the ideal cubic structure studied above (3.95$^3$ \AA\,$^3$ per formula unit).
Our main finding is that the
Eu$_{0.5}$Ba$_{0.5}$TiO$_3$ alloy is polar with large relative displacements of oxygen
and both Ti and Eu relative to the high symmetry reference structure. Using the Berry
phase method we obtain a ferroelectric polarization value of $P = 23$ $\mu$C/cm$^2$.
Our
calculated ground state is orthorhombic with the polarization oriented along a [011]
direction and lattice parameters $a=3.94$ \AA, $b=5.60$ \AA\ and $c=5.59$ \AA. As
expected from our analysis of the soft mode, the calculated ground state is characterized
by large oxygen -- Ti/Eu displacements, and the absence of rotations or tilts of the
oxygen octahedra. Importantly, the large Eu amplitude in the soft mode manifests as a
large off-centering of the Eu from the center of its oxygen coordination polyhedron in
the ground state structure. The origin of the large Eu displacement lies
in its small ionic radius compared with that of divalent Ba$^{2+}$:
The large coordination cage around the Eu ion which is imposed by the large lattice
constant of the alloy results in under-bonding of the Eu that can be relieved by
off-centering. Indeed, we find that in calculations for fully relaxed single phase
EuTiO$_3$, the oxygen octahedra tilt to reduce the volume of the A site in a similar
manner to those known to occur in SrTiO$_3$, in which the A cation size is almost
identical. This Eu off-centering is
desirable for the EDM experiment because the change in local environment at the magnetic
ions on ferroelectric switching determines the sensitivity of the EDM measurement.
\begin{table}[htbp]
\centering
\begin{tabular}{l l c}\hline
\multicolumn{2}{c}{volume (\AA\,$^3$)} & P ($\mu$C/cm$^2$)\\ \hline
61.63 & (constrained) & 23 \\
62.30 & (experimental) & 28 \\
64.63 & (relaxed) & 44 \\
\hline\end{tabular}
\caption{Calculated ferroelectric polarizations, P, of Eu$_{0.5}$Ba$_{0.5}$TiO$_3$ at
three different volumes.}
\label{PversusV}
\end{table}
We note that the magnitude of the polarization is strongly dependent on the volume
used in the calculation (Table~\ref{PversusV}).
At the experimental volume (reported in the next section),
which is only slightly larger than our constrained volume of $3.95^3$ \AA\,$^3$,
we obtain a polarization of 28 $\mu$C/cm$^2$.
At full relaxation, where we find a larger volume close to that
of BaTiO$_3$, we obtain a polarization of 44 $\mu$C/cm$^2$, almost certainly a
substantial over-estimate.
This volume dependence suggests that the use
of pressure to reduce the lattice parameters and suppress the ferroelectric
polarization could be a viable tool for reducing the coercivity at low temperatures. Indeed
our computations show that, at a pressure corresponding to 2.8 GPa applied to the experimental volume
the theoretical structure is cubic, with both the polarization and coercive field reduced
to zero.
Finally, to investigate the likelihood of magnetic ordering,
we calculated the relative energies of the ferromagnetic state discussed above
and of two antiferromagnetic arrangements: planes
of ferromagnetically ordered spins coupled antiferromagnetically along either the
pseudo-cubic $z$ axis or the $x$ or $y$ axes.
(Note that these are degenerate in the high-symmetry cubic structure).
For each magnetic arrangement we re-relaxed the lattice parameters and atomic
positions.
As expected for the highly localized Eu $4f$ electrons on their diluted sublattice,
the energy differences between the different configurations are small
-- around 1 meV per 40 atom supercell -- suggesting an absence of magnetic ordering
down to low temperatures. While our calculations find the ferromagnetic state to
be the lowest energy, this is likely a consequence of our A-site ordering
and should not lead us to anticipate ferromagnetism at low temperature
(Note that, after completing our study, we found a report of an early
effort to synthesize (Eu,Ba)TiO$_3$\cite{Janes/Bodnar/Taylor:1978} in which a
large magnetization, attributed to A-site ordering and ferromagnetism, was
reported. A-site ordering is now known to be difficult to achieve in perovskite-structure
oxides, however, and we find no evidence of it in our samples. Moreover the earlier work
determined a tetragonal crystal structure in contrast to our refined orthorhombic structure.)
In summary, our predicted properties of the (Eu,Ba)TiO$_3$ alloy -- large ferroelectric
polarization, reducible with pressure, with large Eu displacements, and strongly
suppressed magnetic ordering -- meet the criteria for the electron electric dipole
moment search and motivate the synthesis and characterization of the compound,
described next.
\section{Synthesis}
Eu$_{0.5}$Ba$_{0.5}$TiO$_3$ was synthesized by solid-state reaction using mechanochemical
activation before calcination. For details see the Methods section.
The density of the sintered pellets was 86-88\% of the
theoretical density. X-ray diffraction at room temperature revealed the cubic perovskite
$Pm\bar{3}m$ structure with a=3.9642(1)\,\AA. At 100\,K we obtain an orthorhombic ground
state with space group $Amm2$, in agreement with the GGA$+U$ prediction, and lattice
parameters 3.9563(1), 5.6069(2) and 5.5998(2) \AA.
\section{Characterization}
The final step in our study is the characterization of the samples, to
check that the measured properties are indeed the same as those that we predicted and
desired.
Figure~\ref{Fig3} shows the temperature dependence of the complex permittivity between
1\,Hz and 1\,MHz, measured using an impedance analyzer ALPHA-AN (Novocontrol). The
low-frequency data below 100\,kHz are affected above 150\ensuremath{\,\mbox{K}}\, by a small defect-induced
conductivity and related Maxwell-Wagner polarization; the high-frequency data clearly
show a maximum in the permittivity near $T_c$=213\ensuremath{\,\mbox{K}}\ indicating the ferroelectric phase transition.
Two regions of dielectric dispersion -- near 100\ensuremath{\,\mbox{K}}\ and below 75\ensuremath{\,\mbox{K}}\ -- are seen in
tan$\delta(T)$; these could originate from oxygen defects or from ferroelectric domain
wall motion.
\begin{figure}
\begin{center}
\resizebox{0.9\columnwidth}{!}{\includegraphics{eps_No3.pdf} }
\end{center}
\caption{Temperature dependence of permittivity and dielectric loss in
Eu$_{0.5}$Ba$_{0.5}$TiO$_3$ ceramics. The arrows indicate the direction
of increasing frequency and the colors are for clarity to assist the eye
in distinguishing the
lines. The inset shows ferroelectric hysteresis loops measured
at three temperatures and 50\,Hz.} \label{Fig3}
\end{figure}
Measurement of the polarization was adversely affected by the sample conductivity above
150\ensuremath{\,\mbox{K}}, but at lower temperatures good quality ferroelectric hysteresis loops were obtained
(Fig.~\ref{Fig3}, inset). At 135\ensuremath{\,\mbox{K}}\, we obtain a saturation polarization of $\sim$8 $\mu$C/cm$^2$.
The deviation from the predicted value could be the result of incomplete saturation as
well as the strong volume dependence of the polarization combined with the well-known
inaccuracies in GGA$+U$ volumes.
As expected, at lower temperatures the coercive field strongly increases, and only
partial polarization switching was possible even with an applied electric field of
18\,kV/cm (at higher electric field dielectric breakdown was imminent). The partial
switching is responsible for the apparent decrease in saturation polarization below 40\,K.
\begin{figure}[h]
\begin{center}
\resizebox{0.9\columnwidth}{!}{\includegraphics{magChi_new.pdf} }
\end{center}
\caption{Temperature dependence of ac magnetic susceptibility, $\chi$, at various static magnetic
fields and frequency of 214\,Hz. Inset shows magnetization curves at various
temperatures. We note that no hysteresis in magnetization was observed.} \label{Fig4}
\end{figure}
Time-domain THz transmission and infrared reflectivity spectra (not shown here) reveal
a softening of the polar phonon from $\sim$40\ensuremath{\,\mbox{cm}^{-1}}\, at 300\ensuremath{\,\mbox{K}}\ to $\sim$15\ensuremath{\,\mbox{cm}^{-1}}\, at $T_c$,
and then its splitting into two components in the ferroelectric phase. Both components
harden on cooling below $T_c$, with the lower frequency component remaining below 20\ensuremath{\,\mbox{cm}^{-1}}\,
down to 10\ensuremath{\,\mbox{K}}, and the higher-frequency branch saturating near 70\ensuremath{\,\mbox{cm}^{-1}}\, at 10\ensuremath{\,\mbox{K}}. This
behavior is reminiscent of the soft-mode behavior in BaTiO$_{3}$\cite{Hlinka:2008}.
However, when we extract the contribution to the static permittivity that comes from
the polar phonon, we find that it is
considerably smaller than our measured value (Fig.~\ref{Fig3}) indicating an additional
contribution to the dielectric relaxation. Our observations suggest that the phase transition
is primarily soft-mode driven, but also exhibits some order-disorder character.
Finally, we measured the magnetic susceptibility $\chi$ at various static magnetic
fields as a function of temperature down to 0.4\,K. (For details see the Methods section.)
Our results are shown in Fig.~\ref{Fig4}. $\chi(T)$ peaks at $T\sim$1.9\,K indicating an
absence of magnetic ordering above this temperature.
The $\chi(T)$ data up to 300\,K show Curie-Weiss behavior
$\chi(T)=\frac{C}{T+\theta}$ with
$\theta$=-1.63\,K and $C = 0.017$ emuK/(gOe).
The peak in susceptibility at 1.9K is frequency independent and not influenced by
zero field heating measurements after field cooling, confirming antiferromagnetic
order below $T_N = 1.9$\,K.
As in pure EuTiO$_3$, the $\chi(T)$ peak is suppressed by a
static external magnetic field, indicating stabilization of the paramagnetic phase
\cite{Katsufuji/Takagi:2001}.
Magnetization curves (Fig.~\ref{Fig4} inset) show saturation above $2\times10^4$ Oe
at temperatures below $T_{N}$ and slower saturation at 5\,K. No open magnetic hysteresis
loops were observed.
In summary, we have
designed a new material -- Eu$_{0.5}$Ba$_{0.5}$TiO$_3$ -- with the properties required
to enable a measurement of the EDM to a higher accuracy than can currently be
realized. Subsequent synthesis of Eu$_{0.5}$Ba$_{0.5}$TiO$_3$ ceramics confirmed
their desirable ferroelectric polarization and absence of magnetic ordering above
1.9\,K.
The search for the permanent dipole moment of the electron using Eu$_{0.5}$Ba$_{0.5}$TiO$_3$
is now underway.
Initial measurements have already achieved an EDM upper limit of 5 $\times
10^{-23}$ e.cm, which is within a factor of 10 of the current record with a
solid-state-based EDM search \cite{Heidenreich2005}.
We are currently studying a number of systematic effects that may mask the
EDM signal. The primary error originates
from ferroelectric hysteresis-induced heating of the samples during
polarization reversal.
This heating gives rise to a change in magnetic susceptibility,
which, in a non-zero external magnetic field, leads to an undesirable sample
magnetization response. We are working to control the absolute magnetic
field at the location of the samples to the 0.1 $\mu$G level.
Our projected sensitivity of 10$^{-28}$ e.cm should then be achievable.
\section{Acknowledgments}
This work was supported by the US National Science Foundation under award number
DMR-0940420 (NAS), by Yale University, by the Czech Science Foundation (Project Nos.
202/09/0682 and AVOZ10100520) and the
Young Investigators Group Programme of Helmholtz Association, Germany, contract VH-NG-409.
We thank O. Pacherova, R. Krupkova and G. Urbanova for technical assistance and Oleg
Sushkov for invaluable discussions.
\section{Author contributions }
SKL supervised the EDM measurement effort at Yale. AOS and SE performed the analysis and made
preliminary measurements, showing that these materials could be useful in an EDM experiment.
ML and NAS selected (Eu,Ba)TiO$_3$ as the candidate material according to the experimental
requirements and supervised the ab-initio calculations. KZR performed the ab-initio calculations.
ML, NAS and KZR analysed the ab-initio results and wrote the theoretical component of the paper.
Ceramics were prepared by PV. Crystal structure was determined by KK and FL. Dielectric measurements
were performed by MS. JP investigated magnetic properties of ceramics. VG performed infrared
reflectivity studies. DN investigated THz spectra. SK coordinated all experimental studies and
wrote the synthesis and characterization part of manuscript. NAS coordinated the preparation of
the manuscript.
\section{Methods}
\subsection{Computational details}
We performed first-principles density-functional calculations within the spin-polarized generalized
gradient approximation (GGA) \cite{PBE:1996}. The strong on-site correlations of the Eu $4f$
electrons were treated using the GGA+$U$ method \cite{Anisimov/Aryasetiawan/Liechtenstein:1997} with the double
counting treated within the Dudarev approach \cite{Dudarev_et_al:1998} and parameters
$U=5.7$~eV and $J=1.0$~eV. For structural relaxation and lattice
dynamics we used the Vienna \textit{Ab Initio} Simulation Package (VASP)
\cite{VASP_Kresse:1996} with the default
projector augmented-wave (PAW) potentials~\cite{Bloechl:1994} (valence-electron configurations
Eu: $5s^2 5p^6 4f^{7}6s^{2}$, Ba: $5s^{2}5p^{6}6s^{2}$, Ti: $3s^{2}3p^{6}3d^{2}4s^{2}$ and O: $2s^{2}2p^{4}$.)
Spin-orbit interaction was not included.
The 50/50 (Eu,Ba)TiO$_3$ alloy was represented by an ordered A-site structure with the
Eu and Ba ions alternating in a checkerboard pattern (Fig.~\ref{th_phonons}, inset). Structural
relaxations and total
energy calculations were performed for a 40-atom supercell (consisting of two 5-atom perovskite
unit cells in each cartesian direction) using a $4\times4\times4$ $\Gamma$-centered $k$-point
mesh and a plane-wave cutoff of 500 eV.
Ferroelectric polarizations and Born effective charges were calculated using the Berry phase method
\cite{King-Smith:1993}.
Lattice instabilities were investigated in the frozen-phonon scheme \cite{Kunc:1982, Alfe:2009}
for an 80 atom supercell using a $\Gamma$-centered $2\times2\times2$ $k$-point mesh
and 0.0056~\AA\ atomic displacements to extract the Hellman-Feynman forces.
\subsection{Synthesis}
Eu$_2$O$_3$, TiO$_2$ (anatase) and BaTiO$_3$ powders (all
from Sigma-Aldrich) were mixed in stoichiometric ratio then milled intensively in a
planetary ball micro mill Fritsch Pulverisette 7 for 120\,min. in a dry environment
followed by
20 min. in suspension with n-heptane. ZrO$_2$ grinding bowls (25\,ml) and balls (12\,mm
diameter, acceleration 14\,g) were used. The suspension was dried under an IR lamp and the
dried
powder was pressed in a uniaxial press (330\,MPa, 3\,min.) into 13\,mm diameter
pellets. The
pellets were calcined in pure H$_2$ atmosphere at 1200\ensuremath{\,{}^\circ}\!C\ for 24\,hr (to reduce
Eu$^{3+}$ to Eu$^{2+}$), then milled and pressed by the same procedure as above and
sintered at 1300\ensuremath{\,{}^\circ}\!C\ for 24\,hr in Ar\,+\,10\%\,H$_2$ atmosphere. Note that pure
H$_2$ can not be used for sintering without adversely increasing the conductivity of
the sample.
\subsection{Characterization}
Magnetic susceptibility was measured using a Quantum Design PPMS9 and a
He$^3$ insert equipped with a home-made induction coil that allows measurement
of ac magnetic susceptibility, $\chi$ from 0.1 to 214\,Hz.
|
\section{Introduction}
Speed of growth of molecular electronics is being accelerated more
and more as it has brought together scientists and engineers from
various disciplines. The reason behind this attraction is inscribed into
its smallness of size with wonderful electronic properties. In addition,
several other properties such as magnetic, optical, etc., have been
recognized in different molecules, which may be utilized in artificially
tailored devices that are not possible with conventional
materials~\cite{tao}. The concept of electron transport which emerged
first in the theoretical work of Aviram and Ratner in $1974$~\cite{aviram}
has opened a new era in the field of nanoscience. But at that time any type
of measurement in such a small scale was a long-sought goal. Study at
molecular scale level is not a simple one as we cannot avoid the effect
of interface to the external electrodes. However, the progress in the
theoretical works~\cite{nardelli} was continuing, which bestowed
inspirations to the experimentalists to take such task as a challenge.
Now with the advancement in nanotechnology, it is possible to investigate
several transport properties not only through a group of
molecules~\cite{dadosh} but also through a single molecule~\cite{reed}.
This single molecular electronics may play a key role in designing
nanoelectronic circuits. For this we have to have a thorough understanding
of the electronic transport processes at this molecular scale
level~\cite{xue1,baer1,baer2,baer3,walc1,walc2}. Many problems are
yet to be solved to make this field much more reliable. Therefore, the
electron transport in molecular systems is an open area and detailed
investigations of molecular transport are still needed.
All these works we have referred above are related to two-terminal
electron transports. We can also analyze various transport phenomena
of a multi-terminal system, which was first addressed by
B\"{u}ttiker~\cite{buttiker}. The B\"{u}ttiker formalism, which is an
extension of the Landauer two-terminal conductance formula, is a very
simple and elegant way to divulge the transport mechanism in terms of
various transmission probabilities. There are several pioneering
works~\cite{xu1,stafford,sun,zhao,emberly} based on this formalism,
which are very interesting from the theoretical as well as experimental
point of view.
Several {\em ab initio} methods~\cite{damle, derosa, taylor, xue2} are
there which may be utilized to study electron transport properties through
molecular junctions. At the same time, tight-binding model has been
extended to density functional theory (DFT) for transport
calculations~\cite{tagami1}. But in case of molecular systems, the
investigations based on this theory (DFT) have some quantitative
discrepancies compared to the experimental predictions. More over, these
{\em ab initio} theories are computationally very expensive. To avoid
this we do model calculations by using a simple tight-binding framework.
In the present article we do a theoretical study of multi-terminal
electron transport through a single phenalenyl molecule~\cite{tagami2,
tagami3} attached to semi-infinite one-dimensional ($1$D) metallic leads.
We do exact numerical calculations based on single particle Green's
function formalism~\cite{san1,san2} to evaluate conductance, reflection
probability and current-voltage characteristics. Quite interestingly, we
show that the positions where the leads are connected to the molecule as
well as the presence of other leads have eloquent effects on these
transport properties. More over, these characteristics are also
influenced significantly by the molecule-to-lead coupling strengths.
These aspects can be utilized in designing nano-electronic devices.
We organize the paper as follows. With a brief introduction (Section
$1$), in Section $2$ we describe our model and the theoretical
background. Results are analyzed in Section $3$. Finally we conclude
our results in Section $4$.
\section{Model and a view of theoretical formulation}
In this section we focus our attention on the systems where a single
phenalenyl molecule is attached symmetrically or asymmetrically to
semi-infinite $1$D metallic leads through thiol (SH bond) groups. The
models are shown schematically in Figs.~\ref{two}, \ref{three} and
\ref{four} where, the number of leads attached to the molecule is $2$,
$3$ and $4$, respectively. To evaluate the conductance ($g$) and current
($I$) through this single molecular system we adopt the Green's function
technique. For this, first we define the Green's function for the whole
system as,
\begin{equation}
G=\left(E - H \right)^{-1}
\label{green}
\end{equation}
where, $E = \epsilon +i \eta$ with $\eta$ arbitrarily very small number
which can be set as zero in the limiting approximation. $\epsilon$ is the
injecting electron energy. $H$ is the Hamiltonian of the entire system
\begin{figure}[ht]
{\centering \resizebox*{7.3cm}{3cm}{\includegraphics{two.eps}}\par}
\caption{(Color online). Two-terminal quantum system. A phenalenyl molecule
is attached symmetrically to two semi-infinite $1$D metallic leads, viz,
Lead-1 and Lead-2 through thiol (SH bond) groups in the chemisorption
technique where sulfur (S) atoms reside and hydrogen (H) atoms remove.
The filled yellow circles correspond to the location of S atoms.}
\label{two}
\end{figure}
which is of infinite dimension. So, the above equation deals with the
inversion of an infinite dimensional matrix corresponding to the system
consisting of a finite size molecule and semi-infinite leads. However,
the full Hamiltonian can be partitioned into sub-matrices that correspond
to the individual sub-systems like,
\begin{equation}
H = H_M + \sum \limits_{p=1}^N \left(H_p + H_{pM} + H_{pM}^{\dag}\right)
\end{equation}
where, $H_M$ and $H_p$ are the Hamiltonians of the molecule and lead-p,
respectively. $N$ is the number of leads to which the molecule is attached.
$H_{pM}$ represents the coupling matrix that will be non-zero only for the
adjacent points in the molecular system (molecule with sulfur atoms) and
the lead-p. Here all the leads are treated on an equal footing. Within the
non-interacting picture, the tight-binding Hamiltonian of the molecular
system can be manifested as,
\begin{equation}
H_M = \sum_i \epsilon_i c_i^{\dag} c_i + \sum_{<ij>} t
\left(c_i^{\dag} c_j + c_j^{\dag} c_i \right)
\label{hamil}
\end{equation}
where, $\epsilon_i$ is the on-site energy, $t$ is the nearest-neighbor
hopping integral and $c_i^{\dag}$ $(c_i)$ is the creation (annihilation)
operator of an electron at the site $i$. Each lead can be described by
\begin{figure}[ht]
{\centering \resizebox*{7.2cm}{5.3cm}{\includegraphics{three.eps}}\par}
\caption{(Color online). Three-terminal quantum system. A phenalenyl
molecule is attached asymmetrically to three semi-infinite $1$D metallic
leads, namely, Lead-1, Lead-2 and Lead-3 through sulfur (S) atoms.}
\label{three}
\end{figure}
using a similar kind of tight-binding Hamiltonian, as given in
Eq.~\ref{hamil}, characterized by two parameters $\epsilon_0$, the
on-site potential and $t_0$, the nearest-neighbor hopping integral.
Following the partition of the Hamiltonian, the Green's function can also
be partitioned into sub-matrices and the effective Green's function for
the molecular system can be indited (using Lowdin's partitioning
technique~\cite{lowdin1,lowdin2}) as,
\begin{equation}
G_M=\left(E - H_M -\sum\limits_{p=1}^N \Sigma_p \right)^{-1}
\label{greenmolecule}
\end{equation}
where, $\Sigma_p$ is the self-energy due to the coupling of the molecular
system to the lead-p. It is straightforward to obtain an explicit
expression for self-energy corresponding to lead-p,
\begin{equation}
\Sigma_p = H^{\dag}_{pM} G_p H_{pM}
\label{sigma}
\end{equation}
where, $G_p=\left(E-H_p\right)^{-1}$ is the Green's function of lead-p. All
the coupling information are inscribed into this self-energy expression.
Once the Green's function is established, the coupling function $\Gamma_p$
can be easily obtained from the equation~\cite{datta1,datta2},
\begin{equation}
\Gamma_p(E) = i \left[\Sigma^r_p(E)-\Sigma^a_p(E)\right]
\label{gamma}
\end{equation}
where, the advanced self-energy $\Sigma^a_p$ is the Hermitian conjugate of
the retarded self-energy $\Sigma^r_p$. Thus, we can write,
\begin{equation}
\Gamma_p= -2 \mbox{Im} \left(\Sigma^r_p\right)
\label{gammasd}
\end{equation}
In order to evaluate the conductance for the multi-terminal quantum
system, we use the B\"{u}ttiker formalism~\cite{datta1}, valid
at much low temperature and bias voltage, in the form,
\begin{equation}
g_{pq} = \frac{2 e^2} {h} T_{pq}
\label{conduc}
\end{equation}
where, $T_{pq}$ is the transmission probability of an electron across the
molecular system from the lead-p to lead-q and it is related to the
reflection probability by the equation,
\begin{equation}
R_{pp} + \sum \limits_{q(\neq p)}T_{qp}=1
\label{reflec}
\end{equation}
which is obtained from the condition of current conservation~\cite{xu2}.
Now, this transmission probability can be expressed in terms of the
effective Green's function of the molecular system and molecule-to-lead
coupling as,
\begin{equation}
T_{pq} = {\mbox{Tr}} \left[\Gamma_p G^r_M \Gamma_q G^a_M \right]
\label{trans}
\end{equation}
where, $G^r_M$ and $G^a_M$ are the retarded and advanced Green's functions
of the molecular system, respectively. $\Gamma_p$ and $\Gamma_q$
represent the couplings of the molecule to the lead-p and lead-q,
respectively. Since the coupling matrix $H_{pM}$ is non-zero only for
the adjacent points, $n$ and $m$, the transmission probability
becomes~\cite{mujica},
\begin{equation}
T_{pq} = 4~ \Delta_p(nn)~ \Delta_q(mm) \mid{G_M(nm)}\mid^2
\end{equation}
where, $\Delta_p(nn)=\langle n | \Delta_p | n \rangle$,
$\Delta_q(mm)=\langle m | \Delta_q | m \rangle$,
\begin{figure}[ht]
{\centering \resizebox*{5.1cm}{7.8cm}{\includegraphics{four.eps}}\par}
\caption{(Color online). Four-terminal quantum system. A phenalenyl
molecule is attached asymmetrically to four semi-infinite $1$D metallic
leads, viz, Lead-1, Lead-2, Lead-3 and Lead-4 through sulfur (S) atoms.}
\label{four}
\end{figure}
$G_M(nm)=\langle n | G_M | m \rangle$ and $\Delta_p$, $\Delta_q$ are
the imaginary parts of $\Sigma_p$ and $\Sigma_q$, respectively.
In case of two-terminal system~\cite{maiti}, Eq.~\ref{conduc} becomes
quite simpler like,
\begin{equation}
g = \frac{2 e^2} {h} T
\label{land}
\end{equation}
and accordingly, the reflection probability becomes $R=1-T$. For the
two-terminal quantum system the above expression (Eq.~\ref{land}) is the
so-called Landauer conductance formula.
The current $I_p$ passing through the lead-p can be obtained from the
following expression~\cite{datta1},
\begin{equation}
I_p=\frac{2 e} {h} \sum\limits_q \int\limits_{-\infty}^{\infty} T_{pq}(E)
\left [f_p(E)- f_q(E) \right]~dE
\label{current}
\end{equation}
where, $f_{p(q)}=f\left(E-\mu_{p(q)}\right)$ is the Fermi distribution
function with the chemical potential $\mu_{p(q)}=E_F \pm e V_{p(q)}/2$.
$E_F$ is the equilibrium Fermi energy. Throughout this calculation, we
assume that the entire voltage is dropped across the molecule-lead
interfaces as this assumption introduces a minimal effect on the
behavior of the $I$-$V$ characteristics. In this article we set
$c=h=e=1$ for the sake of simplicity.
\section{Numerical results and discussion}
In order to illustrate the results, let us first mention the values of
different parameters used in our numerical calculations. All the on-site
energies of molecule, sulfur atom and leads are set to zero, while the
nearest-neighbor hopping strengths are fixed at 3 for both the molecule
($t$) and leads ($t_0$). But the values of molecule-to-lead coupling
strengths ($\tau_p$ for lead-p) are different from the value assigned
for $t$ and $t_0$. Based on the molecular coupling strength, we analyze
our results in two distinct regimes. One is the weak-coupling regime and
the other is the strong-coupling regime. In the first case, $\tau_p<<t$
and we set $\tau_p=0.5$. In the second case, $\tau_p \sim t$ and for this
regime we fix $\tau_p=2.5$. Here we consider that $\tau_p$'s are identical
for all the leads $p$ and set the equilibrium Fermi energy $E_F$ at
$0$.
\subsection{Conductance-energy characteristics}
In the forthcoming sub-sections we present the characteristic properties
of electron transport for two-, three- and four-terminal molecular systems
where the molecules are attached to leads via thiol-linking groups (SH bond).
In experiments, leads are generally designed from gold (Au) and the thiol
groups are linked by using chemisorption technique~\cite{holleitner} where
hydrogen (H) atoms remove and sulfur (S) atoms survive.
\subsubsection{Two-terminal conductance}
In Fig.~\ref{condtwo} we present the variation of two-terminal conductance
$g$ (red curves) and reflection probability $R$ (green curves) with
injecting electron energy $E$. The results in the weak molecule-to-lead
coupling limit are shown in (a) and (c), while (b) and (d) represent the
same for the strong molecule-to-lead coupling limit.
\begin{figure}[ht]
{\centering \resizebox*{8.2cm}{6.9cm}{\includegraphics{condtwo.eps}}\par}
\caption{(Color online). Conductance $g$ (red lines) and reflection
probability $R$ (green lines) as functions of energy $E$ for two-terminal
molecular system. (a) and (c) represent the results for the weak-coupling
limit, while (b) and (d) correspond to the same for the strong-coupling
limit.}
\label{condtwo}
\end{figure}
In the weak-coupling regime, the presence of sharp resonant peaks indicates
that electron transmission occurs at some typical energy values, while
for all other energies conductance vanishes (see Fig.~\ref{condtwo}(a)).
All these resonant peaks are associated with the energy eigenvalues of
the molecular system, and therefore, we predict that conductance spectrum
is a fingerprint of the electronic structure of the system. Most of the
resonant peaks approach the value $2$, the maximum value of $g$ following
the Landauer conductance formula (see Eq.~\ref{land}) and hence $T$
goes to unity at these resonances indicating ballistic transmission through
the molecular wire. But the behavior of conductance spectrum changes in
case of the strong molecule-to-lead coupling limit. Width of each resonant
peak becomes larger and larger as we increase gradually the molecule-to-lead
coupling strength and for a large molecular coupling, we have the situation
where electron transmission takes place for the entire energy range (for
illustration, see Fig.~\ref{condtwo}(b)). The effect of such broadening
comes from the imaginary parts of the self-energies~\cite{datta1}. All
these phenomena emphasize that fine tuning in the energy scale is necessary
as long as the coupling strength is much weak, while it is not required in
case of the strong molecule-to-lead coupling limit.
This scenario is just inverted in case of reflection probability $R$.
In the weak-coupling regime, sharp dips appear (Fig.~\ref{condtwo}(c)) for
some particular energy values where conductance shows resonant peaks, as
$R$ follows the simple relation $R=1-T$ for the two-terminal molecular
system. For all other energy values, $R=1$ indicating no transmission of
electron across the molecule. The effect of molecule-to-lead coupling
is exactly similar to that for conductance spectrum. In the strong-coupling
regime (see Fig.~\ref{condtwo}(d)), the reflection probability no longer
reach the maximum value ($1$) for the entire energy range.
\subsubsection{Three-terminal conductance}
To illustrate the results for the three-terminal quantum system, constructed
by attaching three leads to the molecule (see Fig.~\ref{three}), let us
start by referring to Fig.~\ref{condthree} where the first column shows
the conductance spectra $g_{pq}$ (from lead-p to lead-q) and the second
column presents the nature of reflection probability $R_{pp}$. Similar
to the two-terminal molecular system, some sharp peaks appear in the
conductance spectra. But the point is that for the three-terminal system
most of the resonant peaks do not reach the value $2$. From the conductance
spectra it is clear that the heights are much reduced compared to the
two-terminal case. This is solely due to the effect of quantum
\begin{figure}[ht]
{\centering \resizebox*{8.3cm}{10.8cm}{\includegraphics{condthree.eps}}\par}
\caption{(Color online). Conductance $g_{pq}$ (red curves) and reflection
probability $R_{pp}$ (green curves) as functions of energy $E$ for
three-terminal molecular system. All the results are presented only
for the weak-coupling regime.}
\label{condthree}
\end{figure}
interference of the electronic waves passing through different arms of
the molecular rings. In this three-terminal system, three leads are
attached asymmetrically to the molecule at three different
locations, which provide different path ways for electron transmission
between the leads. This introduces anomalous features in the conductance
spectra as illustrated in Figs.~\ref{condthree}(a), (b) and (c). More
over, in this three-terminal case, nature of variations of reflection
probabilities is not so simple as we get in the case of two-terminal
molecular system. Here, it
is not necessary that $R_{pp}$ shows dips or peaks where $g_{pq}$ has
peaks or dips since it ($R_{pp}$) depends on the combined effect of
$T_{pq}$'s obeying the expression given in Eq.~\ref{reflec}. Another
important point we like to mention here is that, one can easily find
$T_{qp}$ for any two leads $q$ and $p$ if $T_{pq}$ is known since
the relation $T_{pq}=T_{qp}$ holds true following the time-reversal
symmetry. The effect of molecule-to-lead coupling strength is
identical to the case of two-terminal conductance and therefore, we do
not show those results further.
\subsubsection{Four-terminal conductance}
The transport properties of the four-terminal molecular system is also
\begin{figure}[ht]
{\centering \resizebox*{8.3cm}{11cm}{\includegraphics{condfour.eps}}\par}
\caption{(Color online). Four-terminal conductance $g_{pq}$ as functions
of energy $E$ in the limit of weak molecular coupling.}
\label{condfour}
\end{figure}
described by investigating
various conductances $g_{pq}$ and reflection probabilities $R_{pp}$, which
are obtained from the Eqs.~\ref{conduc} and~\ref{reflec}, respectively.
The results are plotted in Fig.~\ref{condfour} considering the weak
molecule-to-lead coupling. In each figure $g_{pq}$ and $g_{qp}$ are
plotted and they are superposed to each other due to their symmetry.
Sharp resonant peaks of different heights for some particular energies
appear in the conductance spectra similar to the two- or three-terminal
conductance spectra. From the conductance spectra, the effect of quantum
interference associated with the molecule-to-lead interface geometry is
well understood.
\begin{figure}[ht]
{\centering \resizebox*{8.3cm}{8cm}{\includegraphics{reflection.eps}}\par}
\caption{(Color online). Reflection probability $R_{pp}$ as a function of
energy $E$ for four-terminal quantum system. These results correspond to
the weak-coupling regime.}
\label{reflection}
\end{figure}
Also the dependence of electron transport on molecular coupling strength
is exactly similar to that in Fig.~\ref{condtwo} and accordingly, the
results are not given further. Comparing the results given in
Figs.~\ref{condthree} and \ref{condfour} we observe that, the conductances
for two particular leads located at the same positions of the molecule
exhibit completely different features in presence of the other leads. For
instance, $g_{13}$ of the four-terminal system and $g_{12}$ of the
three-terminal system show different spectra though these two conductances
are evaluated for the two leads located at the same positions of the
molecule.
Similarly, careful observation depicts that pathways between the lead-$2$
and lead-$3$ of the three-terminal system are exactly similar to that
between lead-$2$ and lead-$4$ of the four-terminal system. In spite of
this the corresponding spectra i.e., $g_{23}$ of Fig.~\ref{condthree}
and $g_{24}$ of Fig.~\ref{condfour} are different from each other due to
the presence of the fourth lead. The effect of the leads are incorporated
into the self-energies which lead to the change in conductance spectra.
Thus, we can say that the presence of additional leads has a pronounced
effect on the electron transport properties. In $R$-$E$ spectra given
in Fig.~\ref{reflection}, we get various dips at different energies
depending on the combined effect of the transmission probabilities
(Eq.~\ref{reflec}), similar to the case of three-terminal system.
All the reflection probabilities are calculated only for the limit
of weak molecular coupling. Exactly similar feature, except the
broadening, will be observed in the case of strong-coupling.
\subsection{Current-voltage characteristics}
The scenario of electron transport through molecular junction becomes
much more transparent when we discuss the current-voltage ($I$-$V$)
characteristics, where the current is evaluated by integrating the
transmission probability $T$ using Eq.~\ref{current}. The nature of
the variation of transmission probability is exactly similar to that
of conductance spectra except the factor $2$ as we have assumed $e=h=1$
in the Landauer conductance formula (Eq.~\ref{land}). Here, we discuss
the $I$-$V$ characteristics for two-, three- and four-terminal molecular
systems separately in the following sub-sections.
\subsubsection{Two-terminal molecular system}
As an illustration, we display the $I$-$V$ characteristics for the
two-terminal system in Fig.~\ref{currtwo}, where (a) and (b) correspond
to the results for the weak and strong molecule-to-lead coupling limits,
respectively. For this two-terminal case, current can be expressed
mathematically as follows,
\begin{eqnarray}
I &=& g~ (V_1-V_2) \nonumber \\
&=& g~ V_{12}
\end{eqnarray}
where, $V_{12}$ is the voltage difference between the lead-1 and lead-2.
In the case of two-terminal molecular system, we have attached two leads
to the molecule and their chemical potentials are changed as the bias
voltage is applied. With the increase of the voltage, the gap increases
more and more and eventually crosses molecular energy levels
\begin{figure}[ht]
{\centering \resizebox*{7.8cm}{10cm}{\includegraphics{currtwo.eps}}\par}
\caption{(Color online). Current $I$ as a function of applied bias voltage
$V$ for two-terminal molecular system. (a) Weak-coupling limit and (b)
strong-coupling limit.}
\label{currtwo}
\end{figure}
one after another. Accordingly, current channels are opened up and jumps
in the $I$-$V$ curve appear. This provides staircase-like structure in
the current-voltage spectrum and only for the case of weak molecular
coupling these sharp steps appear. But this feature i.e., step-like changes
gradually towards continuous nature with the increase of molecule-to-lead
coupling strength. In addition to that, current amplitude becomes much
higher compared to the case of weak-coupling limit. By noting the area
under the curve of Fig.~\ref{condtwo}(b) the reason behind this enhancement
of current amplitude is clearly understood. Thus, it can be manifested that
molecule-to-lead coupling strength has a significant influence on molecular
transport.
\subsubsection{Three-terminal molecular system}
Now, we describe the current-voltage characteristics for the three-terminal
system where we find the current $I_p$ in lead-p by integrating the
\begin{figure}[ht]
{\centering \resizebox*{7.4cm}{4.75cm}{\includegraphics{3i1v13.eps}}\par}
\caption{(Color online). $I_1$ as a function of $V_{13}$ ($=V_1-V_3$)
for the three-terminal molecular system in the limit of strong-coupling,
keeping $V_{12}$ ($=V_1-V_2$) as a constant. The orange, green and magenta
colors correspond to $V_{12}=1$, $2$ and $3$, respectively.}
\label{3curri1}
\end{figure}
transmission function $T_{pq}$ considering the effects of other terminals.
To be more precise, we write the current expressions for three different
leads as follows,
\begin{eqnarray}
I_1 & = & g_{12} \left(V_1-V_2\right)+g_{13}\left(V_1-V_3\right)\nonumber\\
& = & g_{12} V_{12}+g_{13} V_{13} \label{threecurr1} \\
I_2 & = & g_{21} \left(V_2-V_1\right)+g_{23}\left(V_2-V_3\right)\nonumber\\
& = & g_{21} V_{21}+g_{23} V_{23} \label{threecurr2} \\
I_3 & = & g_{31} \left(V_3-V_1\right)+g_{32}\left(V_3-V_2\right)\nonumber\\
& = & g_{31} V_{31}+g_{32} V_{32}
\label{threecurr3}
\end{eqnarray}
where, $V_{pq}=\left(V_p-V_q\right)$ is the voltage difference between the
two leads named as lead-p and lead-q.
In Fig.~\ref{3curri1}, we plot $I_{1}$ for the lead-1 as a function of
$V_{13}$ for different fixed values of $V_{12}$ in the strong
molecule-to-lead coupling limit. The orange, green and magenta curves
represent the currents for $V_{12}=1$, $2$ and $3$, respectively. It is
clear from the figure that for a constant value of $V_{12}$, the moment
we switch on the bias voltage between lead-1 and lead-3, current rises to
a large value. Then, for a wide range of $V_{13}$, it ($I_1$) slowly
increases with the rise of $V_{13}$ and finally the rate of increment
\begin{figure}[ht]
{\centering \resizebox*{7.4cm}{4.75cm}{\includegraphics{3i2v21.eps}}\par}
\caption{(Color online). $I_2$ as a function of $V_{21}$ ($=V_2-V_1$) for
the three-terminal molecular system in the strong-coupling regime,
keeping $V_{23}$ ($=V_2-V_3$) as a constant. The blue, pink and green
lines correspond to $V_{23}=1$, $2$ and $3$, respectively.}
\label{3curri2}
\end{figure}
of the current gets enhanced when $V_{13}$ is quite high. This behavior
i.e., the rate of increment of the current with bias voltage solely depends
on the positions of resonant peaks in the $g$-$E$ spectrum. For a particular
value of $V_{13}$, $I_1$ increases as we increase $V_{12}$ which is clearly
visible from three different curves plotted in Fig.~\ref{3curri1}.
The variation of current $I_2$ through lead-2 as a function of $V_{21}$
is shown in Fig.~\ref{3curri2}, keeping the voltage $V_{23}$ as a constant.
The currents are evaluated in the strong-coupling limit, where the blue,
pink and green lines correspond to $V_{23}=1$, $2$ and $3$, respectively.
In a similar fashion in Fig.~\ref{3curri3} we display $I_3$-$V_{31}$
characteristics for different fixed values of $V_{32}$ considering the
case of strong-coupling limit. The magenta, green and orange curves
correspond to $V_{32}=1$, $3$ and $5$, respectively.
The characteristic features of the currents $I_1$, $I_2$ and $I_3$ passing
through three different leads are quite analogous to each other. Depending
on the conductance-energy spectra, we get different current amplitudes for
three different leads which are clearly observed from the results presented
in Figs.~\ref{3curri1}, \ref{3curri2} and \ref{3curri3}. All these currents
\begin{figure}[ht]
{\centering \resizebox*{7.4cm}{4.75cm}{\includegraphics{3i3v31.eps}}\par}
\caption{(Color online). $I_3$ as a function of $V_{31}$ ($=V_3-V_1$) for
the three-terminal molecular system in the strong molecule-to-lead
coupling limit, keeping $V_{32}$ as a constant. The magenta, green and
orange colors correspond to $V_{32}=1$, $3$ and $5$, respectively.}
\label{3curri3}
\end{figure}
are computed only for the strong-coupling limit. We can also determine
the currents for the limit of weak-coupling and in that case we will get
sharp step-like features as a function of bias voltage with much reduced
amplitude compared to the strong-coupling case. The origin of step-like
behavior in current is clearly mentioned in the case of two-terminal
molecular system (Sec. 3.2.1).
From the above current expressions (Eqs.~\ref{threecurr1}, \ref{threecurr2}
and \ref{threecurr3}) we see that the current in anyone lead is related
to two potential functions. For instance in Eq.~\ref{threecurr1}, there
are two parameters like $V_{12}$ and $V_{13}$. Keeping $V_{12}$ as a
constant we plot the current $I_1$ in terms of $V_{13}$ (see
Fig.~\ref{3curri1}). At the same time, we can also draw the $I_1$-$V_{12}$
characteristics, considering $V_{13}$ as a constant. Both for these two
cases, the characteristic features are quite similar. This argument is
also valid for the other two current expressions (Eqs.~\ref{threecurr2}
and \ref{threecurr3}).
All the above features of current-voltage characteristics in this
three-terminal molecular system clearly support the basic features
of a traditional macroscopic transistor. Thus we can predict that
the three-terminal molecular system may be utilized to design a
nano-scale molecular transistor.
\subsubsection{Four-terminal molecular system}
Finally, we focus our attention on the current-voltage characteristics for
\begin{figure}[ht]
{\centering \resizebox*{7.4cm}{4.75cm}{\includegraphics{4i1v13.eps}}\par}
\caption{(Color online). $I_1$ as a function of $V_{13}$ for the
four-terminal molecular system in the strong-coupling limit, keeping
$V_{12}$ and $V_{14}$ as constant. The magenta, green and orange colors
correspond to $V_{12}=V_{14}=1$, $3$ and $5$, respectively.}
\label{4curri1}
\end{figure}
the four-terminal molecular system. In this molecular system, the current
expressions for the four different leads are as follows,
\begin{eqnarray}
I_1 & = & g_{12} \left(V_1-V_2\right)+g_{13}\left(V_1-V_3\right)+
g_{14}\left(V_1-V_4\right) \nonumber\\
& = & g_{12} V_{12}+g_{13} V_{13}+g_{14} V_{14} \label{fourcurr1} \\
I_2 & = & g_{21} \left(V_2-V_1\right)+g_{23}\left(V_2-V_3\right)+
g_{24}\left(V_2-V_4\right) \nonumber\\
& = & g_{21} V_{21}+g_{23} V_{23}+g_{24} V_{24} \label{fourcurr2} \\
I_3 & = & g_{31} \left(V_3-V_1\right)+g_{32}\left(V_3-V_2\right)+
g_{34}\left(V_3-V_4\right) \nonumber\\
& = & g_{31} V_{31}+g_{32} V_{32}+g_{34} V_{34} \label{fourcurr3} \\
I_4 & = & g_{41} \left(V_4-V_1\right)+g_{42}\left(V_4-V_2\right)+
g_{43}\left(V_4-V_3\right) \nonumber\\
& = & g_{41} V_{41}+g_{42} V_{42}+g_{43} V_{43}
\label{fourcurr4}
\end{eqnarray}
where, $V_{pq}$ is the voltage difference between the lead-p and lead-q.
As representative examples, in Fig.~\ref{4curri1} we plot the current in
lead-1 ($I_1$) as a function of $V_{13}$ keeping $V_{12}$ and $V_{14}$ as
constant. The results are computed for the strong-coupling limit, where
\begin{figure}[ht]
{\centering \resizebox*{7.4cm}{4.75cm}{\includegraphics{4i2v24.eps}}\par}
\caption{(Color online). $I_2$ as a function of $V_{24}$ ($=V_2-V_4$) for
the four-terminal molecular system in the limit of strong-coupling, keeping
$V_{21}$ and $V_{23}$ as constant. The green, pink and dark-blue lines
correspond to $V_{21}=V_{23}=2$, $4$ and $6$, respectively.}
\label{4curri2}
\end{figure}
the magenta, green and orange curves correspond to $V_{12}=V_{14}=1$, $3$
and $5$, respectively.
The variation of current $I_1$ as a function of $V_{13}$ in this
four-terminal molecular system is quite similar to that as presented in
\begin{figure}[ht]
{\centering \resizebox*{7.4cm}{4.75cm}{\includegraphics{4i3v31.eps}}\par}
\caption{(Color online). $I_3$ as a function of $V_{31}$ for the
four-terminal molecular system for the case of strong-coupling limit,
considering $V_{32}$ and $V_{34}$ as constant. The green, magenta and
dark-blue curves correspond to $V_{32}=V_{34}=1$, $3$ and $5$, respectively.}
\label{4curri3}
\end{figure}
the case of three-terminal system (Fig.~\ref{3curri1}). For a particular
value of $V_{13}$, here also the current amplitude gets increased with
the rise of $V_{12}$ and $V_{14}$. But quite significantly we observe that,
for a particular value of $V_{13}$, the current $I_1$ passing through the
lead-1 of four-terminal system acquires much higher amplitude compared to
the three-terminal system (see Figs.~\ref{3curri1} and \ref{4curri1}). The
reason behind this enhancement of current amplitude is explained as follows.
From Eq.~\ref{fourcurr1} we see that $I_1$ contains three additive terms
where the contributions come from other leads, while in Eq.~\ref{threecurr1}
there are two additive terms. The additional term appears in
Eq.~\ref{fourcurr1} is due to the presence of fourth terminal which is
responsible for the larger current in four-terminal system compared to
the three-terminal one.
In case of the current $I_2$ through lead-2 we show the variation with
respect to $V_{24}$, keeping $V_{21}$ and $V_{23}$ fixed to a particular
\begin{figure}[ht]
{\centering \resizebox*{7.4cm}{4.75cm}{\includegraphics{4i4v43.eps}}\par}
\caption{(Color online). $I_4$ as a function of $V_{43}$ ($=V_4-V_3$) for
the four-terminal molecular system, keeping $V_{41}$ and $V_{42}$ as
constant. The currents are evaluated in the strong-coupling limit,
where the green, dark-blue and pink lines correspond to $V_{41}=V_{42}=1$,
$5$ and $9$, respectively.}
\label{4curri4}
\end{figure}
value as presented in Fig.~\ref{4curri2}. The currents are determined in
the strong-coupling regime, where the green, pink and dark-blue lines
correspond to $V_{21}=V_{23}=2$, $4$ and $6$, respectively.
Similarly, in Fig.~\ref{4curri3} we display $I_3$-$V_{31}$ characteristics
considering $V_{32}$ and $V_{34}$ as constants in the limit of
strong-coupling. The green, magenta and dark-blue lines represent the
currents for $V_{32}=V_{34}=1$, $3$ and $5$, respectively.
At the end, in Fig.~\ref{4curri4} we show the variation of current $I_4$
passing through lead-4 as a function of $V_{43}$ in the limit of strong
molecular coupling, when $V_{41}$ and $V_{42}$ are kept as constants. The
green, dark-blue and pink curves correspond to the currents for
$V_{41}=V_{42}=1$, $5$ and $9$, respectively.
Similar to the case of three-terminal molecular system, for this
four-terminal case, we can also plot the current through any lead-p in
aspect of the other potential functions as given in Eqs.~\ref{fourcurr1},
\ref{fourcurr2}, \ref{fourcurr3} and \ref{fourcurr4}. For all these
cases, the characteristic features are very much similar those are
presented in Figs.~\ref{4curri1}, \ref{4curri2}, \ref{4curri3} and
\ref{4curri4}, and hence, we do not re-plot the results further.
\section{Closing remarks}
In the present paper we have used a parametric approach to study
multi-terminal electron transport through a single phenalenyl molecule.
Using a simple tight-binding framework we have performed all the
numerical calculations through single particle Green's function
formalism. The basic features of electron transport in this molecular
system are explored by investigating the multi-terminal conductance,
reflection probability and current. Following a detailed description
of electron transport in two-terminal quantum system, we have revealed
the essential features of electron transport in the three- and
four-terminal quantum systems separately.
Our clear investigation predicts that the electron transport in
multi-terminal molecular system significantly depends on (a) the
molecule-lead interface geometry, (b) the presence of other leads and
(c) the strength of molecular coupling to the side attached leads. The
unique characteristics of this phenalenyl molecule with a very small
size has enhanced the importance of the present article. Our parametric
study provides several significant features to reveal electron transport
through any complicated multi-terminal quantum system.
In the present paper we have done all the calculations by ignoring
the effects of temperature, electron-electron correlation, etc. We need
further study by incorporating all these effects.
|
\section{Introduction}
Understanding the origin and evolution of galaxies, in particular
the most massive, is one of the major challenges in modern astrophysics.
Many massive galaxies today are giant early-type systems;
hence the formation of spheroids should proceed to a certain extent
in locked step with the mass assembly. The compelling theory of
hierarchical galaxy formation predicts that galaxies are assembled
through successive mergers of smaller systems in overdensities,
or haloes, of hypothetical cold dark matter (CDM) \citep{white78}.
Massive galaxies therefore emerge in the last phase of the formation
history. Alternatively, massive galaxies could form through
the rapid collapse of gas followed by a single prominent starburst at
high redshifts \citep{eggen62, larson75}.
This 'monolithic' scenario is supported by, for example, the tight
colour--luminosity relation of early-type galaxies found in galaxy
clusters \citep[e.g.,][]{bower92, ellis97}.
While different evolution in different models makes distant massive
galaxies a unique test-bed for galaxy formation scenarios, observations
have not yet provided evidence for the evolutionary path
of those galaxies. The major obstacle in observations originates
from the scarcity of galaxies at the high end of the galaxy mass
function; it means that not only it is hard to find the population
but also cosmic variance, the field-to-field variation of observed
volume density arising from large-scale structure, is significant. In
the last decade, many large programmes of optical-band imaging
have been carried out, providing excellent data sets with which
to investigate distant red old galaxies in wide fields of sky exceeding
a whole deg$^2$
\citep[e.g.,][]{bell04, borch06, cimatti06, willmer06, brown07, faber07}.
They consistently suggest that the total
stellar mass locked in red galaxies with luminosities around
$L \sim L^*$, where $L^*$ is the characteristic luminosity of the luminosity
function, has at least doubled since $z \sim 1$. Some of them also claim
little growth in the number of very luminous galaxies well above $L \sim L^*$.
However, while luminous red galaxies roughly correspond
to massive galaxies, it is not clear how well the evolution in the
number of galaxies at the steep high end of the mass function is
understood from these results, since the much more numerous, less
massive galaxies with mass-to-luminosity ratios slightly less than
average could easily dominate the observed numbers of luminous
galaxies. The above authors also reveal that a field of view of the
order of a whole deg$^2$ is still not sufficient to conquer the uncertainty
arising from cosmic variance for the high-end populations of the
galaxy mass function.
The advent of the United Kingdom Infrared Telescope (UKIRT)
InfraredDeep Sky Survey \citep[UKIDSS;][]{lawrence07} provides
a unique opportunity to produce an ideal sample to trace various
aspects of massive galaxies in the distant Universe. Here we report
the results of a $K$-band survey with optical ($u$, $g$, $r$, $i$, $z$ band)
and near-infrared ($Y$, $J$, $H$ band) photometry and optical spectra,
focusing on massive ($M_{\star} > 10^{11} M_{\odot}$) galaxies out to $z = 1$ in an unprecedented
large area covering 55.2 deg$^2$. The $K$-band photometry
provides robust estimates of galaxy stellar masses \citep[e.g.,][]{matsuoka08}
while the very large field of view significantly suppresses
cosmic variance, which allow us to conduct a unique analysis of the
mean properties of distant massive galaxies.
This paper is organized as follows. In Section 2 we describe the
data sources and reduction process to extract the galaxy sample from
the available data. Photometric redshifts and stellar masses are measured
for each galaxy in Section 3. In Section 4, the number-density
evolution of massive galaxies and the associated uncertainties are
explored. We then discuss the star-forming properties of galaxies
and the compatibility of the present results with previous measurements
in Section 5. A summary follows in Section 6. Throughout
this paper, we adopt the concordance cosmology of $H_0$ = 70 km s$^{-1}$ Mpc$^{-1}$, $\Omega_{\rm M} = 0.3$,
and $\Omega_{\Lambda} = 0.7$.
Magnitudes are expressed
in the Vega magnitude system for the UKIDSS near-infrared
bands and in the AB magnitude system for the optical bands.
\section{Data Sources and Reduction}
\subsection{Near-Infrared Photometry}
We extract from the Data Release 3 (DR3; Warren et al., in preparation)
of the UKIDSS/Large Area Survey (LAS) the $K$-band sources
with right ascensions from 1$^{\rm h}$ 15$^{\rm m}$ to 3$^{\rm h}$ 6$^{\rm m}$
on the Sloan Digital
Sky Survey \citep[SDSS;][]{york00} southern equatorial stripe (see
Section 2.2). The range in right ascension is chosen so that the $K$-band
observations are fairly complete within the sample area. More
than half of the $K$-band sources have also been observed in the $Y$,
$J$ and/or $H$ bands. We exclude the sources assigned with serious
quality flags corresponding to \texttt{ppErrBits} attributes larger than 31,
or near (within 20 arcsec of) the detector edges of any exposure. We
also exclude those sources near bright sources. This is achieved by
searching for bright point and extended sources in the Two-Micron
All-Sky Survey \citep[2MASS;][]{skrutskie06} catalogues and rejecting
all the LAS sources within sufficiently large distances of the
bright 2MASS sources. The total effective area of the observations
defining our sample is 55.2 deg$^2$.
We retrieved all the images in our sample area from the DR3 data
base, and used the \texttt{SOURCE EXTRACTOR}, version 2.5 \citep{bertin96}
for magnitude measurements. The total magnitudes ($m_{\rm tot}$) of
the sources are measured with the \texttt{SOURCE EXTRACTOR} total magnitude
algorithm \texttt{MAG\_AUTO}. We measure the aperture magnitudes
($m_{\rm ap}$) with circular apertures of several diameters, 4.5, 3.3, 2.8 and
2.6 arcsec, which correspond to 20 kpc at $z$ = 0.3, 0.5, 0.7 and 0.9,
respectively. Since the seeing condition is generally superior in the
UKIDSS LAS ($\sim$0.8 arcsec) to that in the optical observations by
the SDSS ($\sim$1.5 arcsec), the near-infrared images were smoothed
with the Gaussian kernel in such a way that the resultant full widths
at half-maximum (FWHMs) of stellar profiles are similar to those in
the optical images of the same field. The seeing measurements and
smoothing were performed in each of the small rectangular subareas
of approximately 9 $\times$ 13 arcsec$^2$. Then we ran the \texttt{SOURCE EXTRACTOR}
on the smoothed images to measure the aperture magnitudes of the
sources.
The detection completeness of the $K$-band sources can be estimated
by comparing the numbers of the LAS detections with those
of the much deeper UKIDSS Deep Extragalactic Survey \citep[DXS;][]{lawrence07}
in an overlapping field. This 0.6-deg$^2$ field is
a part of the DXS VIMOS 4 field, which is centred at RA 22$^{\rm h}$ 17$^{\rm m}$,
Dec. $+$00$^{\circ}$ 24' on the celestial equator. While the field is outside
the RA range of our LAS fields, we confirmed that the evaluation
sample of this field has a similar distribution of magnitudes and
their errors to those for our actual sample. Below we show that
the derived detection-completeness function reproduces the galaxy
number counts from our sample, in excellent agreement with previous
measurements, while it gives the lower limit of the detection
completeness for our massive galaxies. We define our sample as
being brighter than the limiting total magnitude $K_{\rm tot}$ = 17.9 mag,
where the detection completeness is higher than 0.5.
\subsection{Optical Photometry \label{subsec:optphot}}
We use the optical $u$, $g$, $r$, $i$ and $z$-band images on the SDSS southern
equatorial stripe. The stripe has been observed repeatedly in
the SDSS-II Supernova Survey \citep{frieman08} during 2005--
2007, as well as in the original SDSS. We retrieved all the available
images taken on the stripe from the SDSS Data Release 6 Supplemental
and Supernova Survey data bases \citep{adelman08}, and stacked them
in each of the five bands. The images
observed in runs 2738 and 3325 are set apart from others, since they
are taken with the standard survey conditions of the original SDSS
and can work as the reference frames for stellar photometry. We
measure, for each retrieved image, the mean and root-mean-square
(rms) of the sky counts and the sky transparency at the observation
by comparing the stellar photometry of the relevant frame with that
of the reference frames. After discarding the worst 5 per cent of the
retrieved images with the largest rms of sky counts, which we find
is sufficient to reject apparently flawed exposures, the images are
zero-shifted and scaled according to the sky-count statistics and then
stacked by the inverse-of-variance weighted average using \texttt{IRAF}. The
photometric calibration of the stacked images is achieved by comparing
the stellar photometry with that of the reference frames. We
find that the stellar magnitudes on the stacked and reference frames
are in excellent agreement, with rms less than 0.05 mag (Fig. 1).
Fig. 2 shows the comparison of the original and stacked r-band
images of the same field. More than 100 original SDSS frames contribute
to each of the stacked frames, and the latter images are on
average $>$2mag deeper than the former images.
\begin{figure}
\includegraphics[width=84mm]{f1_lores.eps
\caption{Comparisons of stellar magnitudes in the stacked and reference
frames in $u$ (top left), $g$ (top right), $r$ (middle left), $i$ (middle right) and $z$
(bottom left) bands. The dashed lines represent the locus where the two
measurements are identical. The rms errors of the photometric calibration
are calculated in the magnitude ranges shown by the dotted lines, where the
stellar photometry is most reliable.}
\label{ref_vs_stack_phot}
\end{figure}
\begin{figure}
\includegraphics[width=84mm]{f2_lores.eps
\caption{Original (top) and stacked (bottom) SDSS $r$-band images of the
same field.}
\label{ref_vs_stack_image}
\end{figure}
We run the \texttt{SOURCE EXTRACTOR} on the stacked images to extract
detected sources. The groups with four or more pixels whose counts
are 1.5$\sigma$ above the local background level are identified as sources.
For every detection, we measure the aperture magnitudes in the
4.5-, 3.3-, 2.8- and 2.6-arcsec diameter apertures as we did for the
near-infrared images. The photometry errors are calculated from
the source photon counts and the background noise. The extracted
sources are cross-identified with the $K$-band sources within the
maximum paring tolerance of 1.0 arcsec. Owing to the deep stacking
of the SDSS images, nearly 90 per cent of the $K$-band sources have
counterparts in the $g$, $r$, $i$ and $z$ bands. Nearly 40 per cent of the
$K$-band sources also have counterparts in the $u$ band.
\subsection{Optical Spectroscopy}
We exploit the two redshift surveys carried out on the SDSS southern
equatorial stripe; the VIMOS-VLT Deep Survey \citep[VVDS;][]{lefevre05}
and the DEEP2 Redshift Survey \citep{davis03}.
Among the four fields of the VVDS 'Wide' survey, we use
the 4-deg$^2$ field of the F22 (2217$+$00), which lies on the celestial
equator, centred at RA 22$^{\rm h}$ 17$^{\rm m}$ 50$^{\rm s}$.4 and
Dec. $+$00$^{\circ}$ 24$^{\rm m}$ 00$^{\rm s}$. This field
coincides with the UKIDSS DXS VIMOS 4, and is observed in both
the LAS and the DXS. We use Field 4 (RA 02$^{\rm h}$ 30$^{\rm m}$, Dec. $+$00$^{\circ}$ 00$^{\rm m}$)
of the DEEP2, one of the '1-h survey' fields placed on our
LAS field. The VVDS adopts a pure $I$-band flux-limited selection
of the sample while the DEEP2 imposes strict colour pre-selection
on the spectroscopic targets to favour galaxies at $z > 0.7$; thus the
two surveys are complementary in terms of the sample selection.
We use only the spectroscopic sample with high redshift-quality
flags, \texttt{zflag}/\texttt{zQ} = 3 or 4 for the VVDS/DEEP2. As a result, we obtain
253 LAS $K$--VVDS and 375 LAS $K$--DEEP2 galaxies, as well
as 1084 DXS $K$--VVDS galaxies. We note that the redshift surveys
are deep enough that essentially all the LAS $K$ sources could be
sampled.
\subsection{Star/Galaxy Classification}
The $K$-band sources separate clearly into stars and galaxies on
the $r - z$ versus $z - K$ diagram as shown in Fig. 3. The colours
are measured with the 2.8-arcsec diameter aperture magnitudes.
We define the demarcation between stars and galaxies along the
minimum surface density on this diagram, i.e. the sources redder
than $z - K = 0.52 (r - z) + 1.74$ are classified as galaxies. The
additional criterion of $r - z < 4$ is set for galaxies to exclude cool
dwarf stars from the sample. We obtain 259 082 galaxies with these
classification criteria. The VVDS classification based on spectra
confirms that the above scheme works very well, yielding a rate
of misclassification (stars classified as galaxies and vice versa) of
less than $\sim$1 per cent. The $K$-band sources without $r$- and/or $z$-band
detections are excluded from the sample, since their extremely red
colours suggest that they are mostly galaxies beyond $z = 1$. Actually,
we find that more than 95 per cent of the LAS $K$--VVDS and LAS
$K$--DEEP2 galaxies are detected in both $r$ and $z$ bands in any redshift
bins at $z \le 1$.
\begin{figure}
\includegraphics[width=84mm]{f3_lores.eps
\caption{The $r - z$ versus $z - K$ diagram for a subset of $K$-band sources
(black), on which the VVDS stars (blue) and galaxies (red) are superimposed.
The green line represents the adopted star/galaxy classification
criteria.}
\label{star_gal}
\end{figure}
We show the $K$-band differential number counts of the extracted
galaxies in Fig. 4. They are in excellent agreement with previous
measurements \citep{daddi00, huang01} down
to the limiting magnitude of $K_{\rm tot}$ = 17.9 mag after the detection-completeness
correction is applied. This suggests that we have successfully
constructed a well-defined sample of galaxies through the
above processes.
\begin{figure}
\includegraphics[width=84mm]{f4.eps
\caption{The $K$-band differential number counts of the extracted galaxies
with (squares) and without (diamonds) detection-completeness correction.
The error bars denote Poisson noise. The blue circles and the red triangles
represent the measurements of \citet{daddi00} and \citet{huang01},
respectively.}
\label{num_counts}
\end{figure}
\section{Redshift and Stellar mass Measurements}
\subsection{Photometric Redshift \label{subsec:photoz}}
We estimate the redshifts of galaxies from the observed broad-band
colours, measured in the 2.8-arcsec diameter aperture, with
the optimized template-fitting method following \citet{ilbert06}.
We present a short summary of the method below, while the full
description of the concept and details can be found in the above
reference.
We choose the DXS $K$--VVDS galaxies to optimize the spectral
templates, leaving the LAS $K$--VVDS and LAS $K$--DEEP2 galaxies
as the evaluation sets of the redshift measurements (Table 1). First,
the DXS $K$--VVDS galaxies are classified into five spectral types,
Ell, Sbc, Scd, Irr and starburst (SB) by least-$\chi^2$ fitting with the
appropriate amounts of dust extinction assuming the extinction laws
of the Small Magellanic Cloud (SMC) by \citet{pei92} for Scd and Irr
and that of starburst galaxies by \citet{calzetti00} for SB. The
initial spectral templates are taken from \citet{cww80} for Ell, Sbc, Scd
and Irr and from \citet{kinney96} for
SB. Then, for each filter $f$, we minimize the sum
\begin{equation}
\psi^2 = \sum_{\rm galaxy}^{}
\biggl( \frac{F_{\rm obs}^f - A \times F_{\rm model}^f - s^f}{\sigma_{\rm obs}^f} \biggr)^2,
\end{equation}
where $F_{\rm obs}^f$ and $\sigma_{\rm obs}^f$ are the observed flux and its error in the filter $f$.
The sum is taken over all the sample galaxies. The parameters
$F_{\rm model}^f$ and $A$
represent the best-fitting template flux and its normalization
factor taken from the initial least-$\chi^2$ fitting. The last term $s^f$ is
a free parameter. While this term should be zero in the case of
a completely random uncertainty in the photometry, we find that
it has a non-zero value in every filter. These values are at most
0.05 mag and are comparable with the expected uncertainty in the
photometric zero-point calibration.
\begin{table}
\caption{Summary of the spectroscopic sample.}
\label{tab:spec_sample}
\begin{tabular}{@{}lcc}
\hline
Sample & Number & Use$^{*1}$\\
\hline
DXS $K$ -- VVDS & 1084 & training\\
LAS $K$ -- VVDS$^{*2}$ & 253 & evaluation\\
LAS $K$ -- DEEP2 & 375 & evaluation\\
\hline
\end{tabular}
\medskip
$^{*1}$Use in the photometric-redshift measurements.\\
$^{*2}$Approximately 70\% of the LAS $K$ -- VVDS galaxies are also the members of the DXS $K$ -- VVDS galaxies.
\end{table}
The DXS $K$--VVDS galaxies are re-classified into five spectral-type
groups with the terms $s^f$ considered. In each group, the observed
broad-band fluxes of galaxies are converted to the rest frame
according to the spectroscopic redshifts, after being normalized and
de-reddened by the best-fitting normalization factor $A$ and the dust
extinction. Since the galaxies have various redshifts, this conversion
generates a continuous spectral energy distribution for each
spectral type of galaxies over the relevant rest-frame wavelength
range. We sort the rest-frame fluxes according to their wavelengths
and bin them by groups of points, and connect the median flux in
each bin to produce the optimized templates. We keep the extrapolations
provided by the initial templates in ultraviolet and infrared
wavelengths where no broad-band data are available. The starburst
template is not optimized, in order to retain the emission lines in
the template. Finally, these optimized templates are interpolated to
produce a total of 62 templates, the first being Ell and the last being
SB, to improve the sampling of the redshift--colour space. Below
we define the spectral type of each galaxy using the best fits from
among these 62 templates.
The created spectral templates are fitted to the observed colours
of the actual sample to measure their redshifts. We evaluate the
measurement accuracy by applying the same procedure to the LAS
$K$--VVDS and LAS $K$--DEEP2 galaxies, as well as the DXS $K$--VVDS
galaxies. The DXS $K$--VVDS galaxies are reduced in number
according to the LAS $K$-band detection completeness and are
given additional random photometry errors in order to simulate the
LAS $K$-band galaxies.We find that the photometric redshifts ($z_{\rm phot}$)
are well correlated with the spectroscopic redshifts ($z_{\rm spec}$) as shown
in Fig. 5, thanks to our wide and relatively fine wavelength coverage
in the $u$ through $K$ bands. The deviation between the two (photometric
and spectroscopic) measurements closely follows a Gaussian
distribution with standard deviation $\sigma_{{\Delta}z/(1+z)} \sim 0.04$
(${\Delta}z = z_{\rm spec} - z_{\rm phot}$) for all three sets of the evaluation sample. Note that
$\sim$70 per cent of the LAS $K$--VVDS galaxies are also members of
the DXS $K$--VVDS galaxies and account for about a quarter of the
sample used to build up the spectral templates, so that the LAS
$K$--VVDS galaxies do not provide a completely independent test of
the photometric-redshift accuracy. As a further test, we created another
template set from the DXS $K$--VVDS galaxies omitting these
LAS $K$--VVDS galaxies and repeated the photometric-redshift measurement.
This test again gives $\sigma_{{\Delta}z/(1+z)} \sim 0.04$, which indicates that
it is a robust estimate of the photometric-redshift uncertainty. We
show the uncertainty as a function of redshift and stellar mass (as
determined below) in Table 2. They are relatively large in the lowest
and highest redshift bins for larger stellar mass classes, for which
relatively small numbers of sample contribute to the spectral templates.
We also show the uncertainty as a function of spectral type
in Table 3, which suggests there is little variation of uncertainty
among the different spectral types.
\begin{figure}
\includegraphics[width=84mm]{f5.eps
\caption{Comparison of the photometric and spectroscopic redshift measurements
of LAS $K$--VVDS (light blue), DXS $K$--VVDS (blue) and LAS
$K$--DEEP2 (red) galaxies. The solid line shows the locus where the two
measurements are identical.}
\label{photz_vs_specz}
\end{figure}
\begin{table}
\caption{Uncertainty ($\sigma_{{\Delta}z/(1+z)}$) in photometric redshift.}
\label{tab:photo-z_err}
\begin{tabular}{@{}lcccc}
\hline
& & log $M_{\star}$ & \\
Redshift & 10.0 -- 10.5 & 10.5 -- 11.0 & 11.0 -- 11.5 & 11.5 -- 12.0\\
\hline
0.2 -- 0.4 & 0.037 (50) & 0.046 (51) & 0.043 ( 6) & --- ( 0) \\
0.4 -- 0.6 & 0.040 (35) & 0.037 (55) & 0.039 (25) & --- ( 0) \\
0.6 -- 0.8 & 0.021 (35) & 0.030 (73) & 0.032 (93) & 0.034 (20) \\
0.8 -- 1.0 & 0.033 (17) & 0.038 (71) & 0.046 (55) & 0.047 ( 7) \\
\hline
\end{tabular}
\medskip
Note --- Numbers in parentheses represent the number in the evaluation samples.
\end{table}
\begin{table}
\caption{Uncertainty ($\sigma_{{\Delta}z/(1+z)}$) in photometric redshift.}
\label{tab:photo-z_err2}
\begin{tabular}{@{}cc}
\hline
Spectral type & $\sigma_{{\Delta}z/(1+z)}$\\
\hline
Ell -- Sbc & 0.035 (386)\\
Sbc -- Scd & 0.037 (171)\\
Scd -- Irr & 0.036 ( 35)\\
Irr -- SB & --- ( 0)\\
\hline
\end{tabular}
\medskip
Note --- Numbers in parentheses represent the number in the evaluation samples.
\end{table}
\subsection{Stellar Mass \label{subsec:stellarmass}}
The stellar masses ($M_{\star}$) of galaxies are determined by fitting to the
observed colours the stellar population synthesis models of \citet{bc03}.
The aperture magnitudes ($m_{\rm ap}$) of the 4.5-, 3.3-,
2.8- and 2.6-arcsec apertures are used for the fitting of galaxies with
photometric redshifts $z$ = 0.2--0.4, 0.4--0.6, 0.6--0.8, and 0.8--1.0,
respectively, so that we sample the stellar populations consistently
within the central $\sim$20 kpc of all the galaxies. The resultant stellar
mass is then scaled by $10^{0.4 (K_{\rm ap}-K_{\rm tot})}$ to correct for aperture loss.
We adopt the standard configurations with the Padova 1994 stellar
evolutionary tracks and BaSel 3.1 spectral library for the \citet{bc03} models.
We assume three values of metallicity: 0.2,
1 and 2.5 $Z_{\odot}$, where $Z_{\odot}$ is the solar metallicity. The star-formation
history is assumed to take the exponentially declining form
$\tau^{-1}$ exp($-t/\tau$), where the $e$-folding time $\tau$ is a free parameter, with
the \citet{salpeter55} initial mass function (IMF). Our stellar mass
estimates can be approximately converted to those with another
commonly used IMF, that of \citet{chabrier03}, by adding $\sim$0.25 dex.
Other free parameters are the age $t$ of the stellar population and the
colour excess $E(B-V)$ due to the dust extinction of the stellar
radiation, following the SMC extinction curve of \citet{pei92}. These
parameters are varied over the plausible ranges of
10 Myr $\le \tau \le$ 10 Gyr ($\Delta$log $\tau$ = 0.2),
10 Myr $\le t \le$ 10 Gyr ($\Delta$log $t$ = 0.1), and
$0.0 \le E(B-V) \le 0.5$ mag ($\Delta$$E(B-V)$ = 0.05),
and the best-fitting
parameter set is searched for by the least-$\chi^2$ method for
each galaxy (the values in parentheses represent the grid intervals).
The additional error of 0.05 mag is added in quadrature to all band
magnitudes in the fitting in order to take into account the uncertainty
in the photometric zero-point calibration.
We derive two kinds of stellar mass for the spectroscopic sample,
i.e. the stellar mass with spectroscopic redshifts ($M_{\star, spec}$) and
the stellar mass with photometric redshifts ($M_{\star, phot}$). The difference
between the two measures, $\Delta {\rm log} M_{\star} = {\rm log} M_{\star, spec} - {\rm log} M_{\star, phot}$, is
found to be clearly correlated with the photometric redshift deviation
${\Delta}z = z_{\rm spec} - z_{\rm phot}$. Such a correlation is expected, since
a larger $z_{\rm phot}$ leads to a larger estimate of the galaxy luminosity,
which then leads to a larger estimate of stellar mass $M_{\star, phot}$.
The observed relation between ${\Delta}z$ and $\Delta {\rm log} M_{\star}$ is actually quite
consistent with this expected correlation. Another expected cause
of the correlation between ${\Delta}z$ and $\Delta {\rm log} M_{\star}$ comes from the fact
that larger estimates of $z_{\rm phot}$ lead to systematically shorter rest-frame
wavelengths to which each of the observing wavebands
corresponds. After removal of the above first component of the
systematic correlation, we found marginal evidence for the second
correlation in our sample, which is $\Delta {\rm log} M_{\star}$ = $+/-$ ($0.04 \pm 0.07$)
when ${\Delta}z$ is negative/positive. In addition, we consider the uncertainty
associated with the least-$\chi^2$ model fitting. It is evaluated
by the 1$\sigma$ confidence surface of the $\chi^2$ distributions in the fitting
parameter space.
We show the total amplitudes of the stellar mass uncertainty
($\sigma_{\Delta {\rm log} M_{\star}}$) as a function of redshift and stellar mass in Table 4. Those
as a function of spectral type are shown in Table 5. The above estimates
of error amplitudes and the correlations between ${\Delta}z$ and $\Delta {\rm log} M_{\star}$
are taken into account in the Monte Carlo simulation presented
below. Note that further different assumptions on the stellar
mass estimation, such as different stellar population synthesis
models and different IMF, can cause additional uncertainty in the
derived properties of galaxies. We will address this issue in Section
5.
\begin{table}
\caption{Uncertainty ($\sigma_{\Delta {\rm log} M_{\star}}$) in stellar mass.}
\label{tab:logM_err}
\begin{tabular}{@{}lcccc}
\hline
& & log $M_{\star}$ & \\
Redshift & 10.0 -- 10.5 & 10.5 -- 11.0 & 11.0 -- 11.5 & 11.5 -- 12.0\\
\hline
0.2 -- 0.4 & 0.19 (50) & 0.20 (51) & 0.22 ( 6) & --- ( 0) \\
0.4 -- 0.6 & 0.19 (35) & 0.17 (55) & 0.16 (25) & --- ( 0) \\
0.6 -- 0.8 & 0.12 (35) & 0.16 (73) & 0.17 (93) & 0.19 (20) \\
0.8 -- 1.0 & 0.14 (17) & 0.17 (71) & 0.19 (55) & 0.25 ( 7) \\
\hline
\end{tabular}
\medskip
Note --- Numbers in parentheses represent the number in the evaluation samples.
\end{table}
\begin{table}
\caption{Uncertainty ($\sigma_{\Delta {\rm log} M_{\star}}$) in stellar mass}
\label{tab:logM_err2}
\begin{tabular}{@{}cc}
\hline
Spectral type & $\sigma_{\Delta {\rm log} M_{\star}}$\\
\hline
Ell -- Sbc & 0.16 (386)\\
Sbc -- Scd & 0.17 (171)\\
Scd -- Irr & 0.17 ( 35)\\
Irr -- SB & --- ( 0)\\
\hline
\end{tabular}
\medskip
Note --- Numbers in parentheses represent the number in the evaluation samples.
\end{table}
\section{Results}
We define our massive galaxy sample using two stellar mass classes,
i.e. $10^{11.0-11.5} M_{\odot}$ galaxies with
$10^{11.0} M_{\odot} < M_{\star} < 10^{11.5} M_{\odot}$ and $10^{11.5-12.0} M_{\odot}$ galaxies with
$10^{11.5} M_{\odot} < M_{\star} < 10^{12.0} M_{\odot}$.
The
galaxies are grouped into four redshift bins, $z$ = 0.2 -- 0.4, 0.4 -- 0.6, 0.6 -- 0.8, and 0.8 -- 1.0.
Total numbers included in the sample are
summarized in Table 6. The median photometry errors in the $r$-, $z$-, and
$K$-band aperture magnitudes and in the $K$-band total magnitudes
($r_{\rm ap}$, $z_{\rm ap}$, $K_{\rm ap}$, $K_{\rm tot}$) are also listed.
The aperture magnitude errors
are generally smaller than the typical uncertainty in the photometric
zero-point calibration ($\sim$0.05 mag).
\begin{table*}
\caption{Summary of the massive galaxy sample.}
\label{tab:sample_num}
\begin{tabular}{@{}cclcl}
\hline
& log $M_{\star}$ = & 11.0 -- 11.5 & log $M_{\star}$ = & 11.5 -- 12.0\\
Redshift & Number & Photometry Error$^*$ & Number & Photometry Error$^*$ \\
\hline
0.2 -- 0.4 & 9,720 & ($<$0.01, $<$0.01, 0.01, 0.05) & 1,408 & ($<$0.01, $<$0.01, $<$0.01, 0.03)\\
0.4 -- 0.6 & 15,300 & ($<$0.01, $<$0.01, 0.02, 0.08) & 572 & ($<$0.01, $<$0.01, 0.01, 0.04)\\
0.6 -- 0.8 & 18,582 & (0.02, 0.02, 0.03, 0.13) & 815 & (0.01, 0.01, 0.02, 0.09)\\
0.8 -- 1.0 & 12,371 & (0.04, 0.02, 0.04, 0.16) & 613 & (0.03, 0.02, 0.03, 0.12)\\
\hline
\end{tabular}
\medskip
Note ($^*$) --- Median photometry errors in the r-, z- and K-band aperture magnitudes and in the K-band total magnitudes
($r_{\rm ap}$, $z_{\rm ap}$, $K_{\rm ap}$, $K_{\rm tot}$).
\end{table*}
We show the differential number counts of the massive galaxies
in Fig. 6. It shows that the number-count distributions of
the $10^{11.5-12.0} M_{\odot}$ galaxies have faint-end drop-offs at magnitudes
brighter than the limiting magnitude, which assures the near-complete
detection of this population. On the other hand, the faintest
of the $10^{11.0-11.5} M_{\odot}$ galaxies at high redshifts ($z > 0.6$) fall below
the limiting magnitude, and are thus left uncounted. In order
to estimate the lost fraction of $10^{11.0-11.5} M_{\odot}$ galaxies at these redshifts,
we derive the detection completeness specifically for these
galaxies as follows. We take each of the $10^{11.0-11.5} M_{\odot}$ galaxies in
the lowest redshift bin and assign random redshifts in the $0.6 < z < 0.8$ and
$0.8 < z < 1.0$ ranges. The galaxies are dimmed and
reduced in apparent size according to the assigned redshifts, placed
on random positions of the LAS $K$-band images, and then extracted
by \texttt{SOURCE EXTRACTOR} in the same way as the actual sample sources
are detected. The recovery rate of the embedded objects as a function
of their magnitudes gives the detection completeness of the
galaxies, which we find is significantly better than that of the whole
sample derived before. The 50 per cent detection completeness is
actually achieved at $K_{\rm tot}$ = 18.6 mag instead of $K_{\rm tot}$ = 17.9 mag,
and almost all galaxies brighter than $K_{\rm tot}$ = 17.9 mag are detected.
With the new detection-completeness function taken into account,
the fractions of $10^{11.0-11.5} M_{\odot}$ galaxies fainter than the formal limiting
magnitude ($K$ = 17.9 mag, thus uncounted) are 11 and 16 per
cent at $z$ = 0.6--0.8 and 0.8--1.0, respectively
\begin{figure}
\includegraphics[width=84mm]{f6.eps
\caption{The differential number counts of $10^{11.0-11.5} M_{\odot}$ (green) and
$10^{11.5-12.0} M_{\odot}$ (orange) galaxies at $z$ = 0.2--0.4 (top left), 0.4--0.6 (top
right), 0.6--0.8 (bottom left) and 0.8--1.0 (bottom right). The black dashed
lines represent our formal limiting magnitude of $K_{\rm tot}$ = 17.9 mag.}
\label{z_num2_4}
\end{figure}
We estimate the uncertainty in the measured numbers of massive
galaxies by a Monte Carlo simulation, as follows. First, we generate
a mock galaxy catalogue containing 10 galaxies for each stellar
mass ($\Delta$log$M_{\star} = 0.1$) and redshift ($\Delta{z} = 0.02$) bin in the ranges
$8.0 <$ log $M_{\star}$ $< 13.0$ and $0.0 < z < 1.4$, where the numbers in
parentheses represent the bin widths. Each galaxy is assigned a
weighting factor corresponding to the number density of galaxies,
following the galaxy mass function of \citet{cole01}. Next the
redshifts of galaxies are given perturbations following the measured
uncertainty of the actual sample. The stellar masses are also given
perturbations, correlated with the photometric redshift perturbations
as explored in Section 3.2. Then the mock galaxies are weighted by
their weighting factors and redistributed, and counted in the redshift
and stellar mass bins to obtain the output mass function. We repeat
the calculation 100 times, varying the random components.
Fig. 7 shows the results of the simulation. We observe both systematic
and random components in the resultant error estimates.
The systematic component is evident in $M_{\star} > 10^{12.0} M_{\odot}$ bins. This
is the so-called Eddington bias \citep{eddington13}, caused by the
steep slope of the high end of the mass function; simply put, a small
portion of the less massive, much more numerous galaxies could
contaminate the more massive classes owing to measurement errors,
which significantly alters the steep part of the mass function.
This is why we limit our massive galaxy sample to those with $M_{\star} < 10^{12.0} M_{\odot}$;
the systematic increases in number are found to be insignificant
for our mass ranges, i.e. negligible for the $10^{11.0-11.5} M_{\odot}$
galaxies and 40--60 per cent for the $10^{11.5-12.0} M_{\odot}$ galaxies. The
output number densities from the 100 repeated calculations scatter
around the systematic components, which yields the random
components of the measurement error. The standard deviations of
the scatter are $\sim$5 per cent and $\sim$12 per cent of the numbers of
$10^{11.0-11.5} M_{\odot}$ and $10^{11.5-12.0} M_{\odot}$ galaxies, respectively. Note that
the current estimate is not perfect, since we assume a non-evolving
galaxy stellar mass function at $0 < z < 1$, while our results show
clear signs of its evolution (see below). We will discuss this issue
further in the following section.
\begin{figure}
\includegraphics[width=84mm]{f7.eps
\caption{The input (black line) and output (100 red lines) galaxy mass
functions of the Monte Carlo simulation at $z$ = 0.2--0.4 (top left), 0.4--0.6
(top right), 0.6--0.8 (bottom left) and 0.8--1.0 (bottom right). The dashed
lines show the mass ranges in which our massive galaxy sample is defined.}
\label{montecarlo}
\end{figure}
Another source of uncertainty comes from cosmic variance. We
estimate cosmic variance by dividing our sample into five subfields
along the right ascension, $\sim$11 deg$^2$ each, and then calculating
the fractional variation of the measured numbers of massive
galaxies in these subfields. We find that the fractional variations
are $<$10 per cent and $<$20 per cent for the $10^{11.0-11.5} M_{\odot}$ and $10^{11.5-12.0} M_{\odot}$
galaxies, respectively, in each of the four redshift
bins. Considering that the total fields are five times larger than the
subfields, we conclude that cosmic variance could affect the measured
number of massive galaxies in each redshift bin by up to 5 per
cent and 10 per cent for the two classes of galaxies, respectively.
The above estimates are roughly consistent with the theoretical
predictions provided by \citet{somerville04}.
We show our results with regard to the number-density measurements
in Table 7, and plot them along with the measurements for
the local Universe \citep{cole01} in Fig. 8. The error bars take
into account all uncertainties considered above, as well as the Poisson
noises. The number densities of the $10^{11.5-12.0} M_{\odot}$ galaxies
are corrected for Eddington bias, although this correction has little
significance for our final conclusions. The local densities were normalized
to take into account the assumptions of \citet{cole01}
with regard to stellar population synthesis that differ from ours; the
major difference is that they assume a constant formation redshift
of galaxies at $z_f = 20$, while we vary the age of the stellar population
as a free parameter. The normalization factor, $\sim0.4$, is derived
by applying their assumption to our $10^{11.0-11.5} M_{\odot}$ galaxies in the
lowest redshift bin $z$ = 0.2--0.4. As seen in the figure, we find that
the most massive ($10^{11.5-12.0} M_{\odot}$) galaxies have experienced rapid
evolution in number since $z = 1$. On the other hand, the number
densities of the less massive ($10^{11.0-11.5} M_{\odot}$) systems show a rather
mild evolution during the same period.
\begin{table}
\caption{Number densities of the massive galaxies.}
\label{tab:number_densities}
\begin{tabular}{@{}lcc}
\hline
& log $M_{\star}$ & \\
Redshift & 11.0 -- 11.5 & 11.5 -- 12.0\\
\hline
0.2 -- 0.4 & 10.9 $\pm$ 0.8 & 0.98 $\pm$ 0.15 \\
0.4 -- 0.6 & 7.8 $\pm$ 0.6 & 0.19 $\pm$ 0.03 \\
0.6 -- 0.8 & 6.8 $\pm$ 0.5 & 0.17 $\pm$ 0.03 \\
0.8 -- 1.0 & 3.7 $\pm$ 0.3 & 0.11 $\pm$ 0.02 \\
\hline
\end{tabular}
\medskip
\begin{flushleft}
Note --- Number densities are given in units of $10^{-4}$ Mpc$^{-3}$ log $M_{\star}^{-1}$.
\end{flushleft}
\end{table}
\begin{figure}
\includegraphics[width=84mm]{f8.eps
\caption{The number densities of massive galaxies versus the age of the
Universe or redshift. The green and orange symbols and lines represent the
$10^{11.0-11.5} M_{\odot}$ and $10^{11.5-12.0} M_{\odot}$ galaxies, respectively. The filled circles
represent our measurements while the open circles represent measurements
in the local Universe \citep{cole01}. The solid lines represent the predictions
of the Millennium Simulation with the semi-analytic galaxy formation
model (see text).}
\label{number_densities}
\end{figure}
\section{Discussion \label{sec:discuss}}
The measured number density evolution of massive galaxies shows
clear signs of the hierarchical evolution of these systems. Such a
galaxy evolution scenario is predicted in the latest galaxy formation
models based on $\Lambda$CDM theory. In Fig. 8 we overlay the predictions
of the Millennium Simulation (Lemson et al. 2006), the largest
numerical simulation to date based on $\Lambda$CDM theory, with the semianalytic
galaxy formation model of \citet{delucia07},
scaled to fit to the local observations (by a factor of 0.4). The
stellar mass of the Millennium model has been shifted by $+$0.25 dex
in order to correct for the different IMFs adopted (the model adopts
the \citet{chabrier03} IMF). The observed hierarchical pattern of
evolution is consistent with the prediction of the model, while we
find some discrepancies between the observation and the model
(e.g. the $10^{11.0-11.5} M_{\odot}$ galaxies at $z$ = 0.8--1.0).
This indicates that
the basic idea of the bottom-up construction of galaxy systems is
valid at least for the most massive galaxies with $M_{\star} > 10^{11} M_{\odot}$.
Meanwhile, we point out that our measurements plotted in Fig. 8
seem to have an apparently unnatural feature at $z$ = 0.2--0.4, where
the number densities are significantly high relative to the overall
trend considering their estimated errors, implying residual systematic
effects from the Eddington bias. We check this by investigating
the systematic increases in the number of massive galaxies in the
spectroscopic sample, from those obtained with the spectroscopic
redshifts to those with photometric redshifts (we consider only those
redshift and stellar mass classes with more than five spectroscopic
sources). The above Monte Carlo simulation shows that the use of
photometric redshifts (and the consequent stellar mass fluctuations)
is the dominant source of the Eddington bias. We list the measured
systematic increases in Table 8 separately for the VVDS and DEEP2
galaxies, since the amplitudes of the Eddington bias are subject to
the redshift distribution of sources. The DXS $K$--VVDS galaxies
have been given additional photometry errors and detection incompleteness
in order to simulate the LAS $K$ galaxies. The systematic
increases estimated in the Monte Carlo simulation are also listed in
Table 8. The table shows that the systematic increases measured in
the spectroscopic sample are mostly less than the estimates from the
Monte Carlo simulation, while the one for the VVDS $10^{11.0-11.5} M_{\odot}$ galaxies at $z = 0.2 - 0.4$
is exceptionally high. While it is based on
a small sample, it provides marginal evidence that the Eddington
bias is unexpectedly significant in this lowest redshift bin, due to
some unquantified sources of systematic uncertainty. In that case,
the number densities measured at $z$ = 0.2--0.4 should be regarded
as the upper limits.
We also note that we adopt a non-evolving galaxy stellar mass
function in the Monte Carlo simulation, while our results suggest
the steepening of the high end of the mass function toward high
redshifts; thus the estimated amount of Eddington bias should be
regarded as a lower limit. When such a steepening of the mass function
is taken into account in the correction of the Eddington bias, we
obtain even lower numbers of galaxies in the more massive classes
at higher redshifts, which would further strengthen our conclusion.
\begin{table*}
\caption{Estimates of the Eddington bias$^{*1}$.}
\label{tab:eddington_bias}
\begin{tabular}{@{}ccccccc}
\hline
& log $M_{\star}$ = & 11.0 -- 11.5 & & log $M_{\star}$ = & 11.5 -- 12.0 & \\
Redshift & MCS$^{*2}$ & VVDS$^{*3}$ & DEEP2$^{*3}$ & MCS$^{*2}$ & VVDS$^{*3}$ & DEEP2$^{*3}$ \\
\hline
0.2 -- 0.4 & $<$ 0.1 & 0.7 (6) & - (0) & 0.6 & - (0) & - (0) \\
0.4 -- 0.6 & $<$ 0.1 & $<$ 0.1 (23) & - (2) & 0.5 & - (0) & - (0) \\
0.6 -- 0.8 & 0.1 & 0.1 (21) & $<$ 0.1 (72) & 0.6 & 0.3 (7) & 0.3 (13) \\
0.8 -- 1.0 & $<$ 0.1 & - (2) & $<$ 0.1 (53) & 0.4 & - (0) & 0.1 (7) \\
\hline
\end{tabular}
\medskip
\begin{flushleft}
$^{*1}$Rates of increase (increased amounts divided by the original values) of the number densities are listed.\\
$^{*2}$Estimates from the Monte Carlo simulation.\\
$^{*3}$Estimates from the VVDS and DEEP2 spectroscopic samples (the number in the samples is shown in the parentheses).\\
\end{flushleft}
\end{table*}
Recently, \citet{marchesini09} provided a detailed analysis
of random and systematic uncertainties affecting the galaxy stellar
mass function. They adopt 14 different stellar mass estimations
with different combinations of metallicity, dust extinction law, stellar
population synthesis models and IMF, and show that the derived
number density of galaxies in a given stellar mass bin could be
altered by up to 1 dex. The 'bottom-light' IMFs in particular, with
a deficit of low-mass stars relative to a standard \citet{chabrier03}
IMF, give significantly different results for the stellar mass function
from other classical IMFs. A more top-heavy IMF at higher
redshifts is actually suggested by, for example, \citet{dave08} and
\citet{vandokkum08}. Thus we point out that our results are subject
to a systematic change of these stellar-population properties during
$0 < z< 1$. Future improvements in stellar population models based
on new observations are eagerly awaited to overcome these large
uncertainties inherent in stellar mass measurements.
In order to probe the star-forming properties of massive galaxies,
here we investigate their rest-frame optical colours. Nearby galaxies
are known to show a clear bimodality in the optical colour distribution,
in which early-type galaxies form a narrow red sequence that
is separated from blue star-forming populations by a valley of the
galaxy distribution \citep[e.g.,][]{strateva01, hogg03}. A
similar bimodality is observed out to $z > 1$ \citep[e.g.,][]{lin99, im02, bell04}.
We calculate the rest-frame $U$- and
$V$-band magnitudes of massive galaxies by $k$-correcting the
nearest observed $r$, $i$ or $z$-band magnitudes, where the amounts of
the $k$-corrections are estimated from the best-fitting spectral population
synthesis models. We show the resultant colour distributions
in Fig. 9.
\begin{figure}
\includegraphics[width=84mm]{f9.eps
\caption{$U - V$ colour distributions of the massive galaxies at $z$ =
0.2--0.4 (top left), 0.4--0.6 (top right), 0.6--0.8 (bottom left) and 0.8--1.0 (bottom right).
The grey, green and orange lines represent the
$< 10^{11.0} M_{\odot}$, $10^{11.0-11.5} M_{\odot}$ and $10^{11.5-12.0} M_{\odot}$ galaxies, respectively.
The dotted lines show the demarcation between the blue and red populations.}
\label{cmd}
\end{figure}
As Fig. 9 shows, the less massive ($< 10^{11.0} M_{\odot}$) galaxies show
a clear colour bimodality, as expected, with the peak colour of the
red sequence at $U - V \sim 1.2$ and a valley of galaxy distributions
at $U - V \sim 1.0$ in all redshift bins. Compared with these
galaxies, massive ($> 10^{11.0} M_{\odot}$) galaxies are apparently dominated
by the red population, with conspicuous peaks at $U - V \sim 1.2$.
We divide the galaxies into blue and red populations at $U - V = 1.0$
and calculate the blue fractions (fractions of the blue population).
The associated errors are estimated by repeatedly giving
random fluctuations to the $U - V$ colours, taking into account the
uncertainties in the photometry and the amount of $k$-correction, and
then re-measuring the blue fractions. We find that the fluctuated blue
fractions are systematically larger than the original values, since
there is a greater red population than blue around the demarcation
$U - V = 1.0$, and correct for the effect (the correction amounts are
included in the final errors). We plot the measured blue fractions as a
function of redshift and stellar mass in Fig. 10, and also list them in
Table 9. One can see that the blue fractions are significantly lower in
more massive galaxies, and that the fractions in massive systems decrease
toward the local Universe. The blue fractions in $< 10^{11.0} M_{\odot}$
galaxies increase from z = 0.4--0.6 to 0.2--0.4 because the dominant
population within the sample shifts to bluer, less massive galaxies
toward the local Universe due to the fainter detection limit. In fact
we observe a decreasing blue fraction toward the local Universe, as
seen in massive galaxies, if we take the subsample with stellar mass
$10^{10.5} M_{\odot}< M_{\star} < 10^{11.0} M_{\odot}$ (see Table 9). As discussed above,
the $10^{11.0-11.5} M_{\odot}$ galaxies at $z$ = 0.2--0.4 and the $10^{11.5-12.0} M_{\odot}$
galaxies in all redshift bins could have considerable fractions of
contamination from less massive galaxies, which likely have bluer
$U - V$ colours. Actually, investigating the spectroscopic sample
shows that the contamination makes the mean $U - V$ colours bluer
by up to 0.1 mag. Thus the true blue fractions in the above classes
of galaxies could be even smaller than the present measurements.
The lower blue fractions in more massive galaxies and the decreasing
trend toward the local Universe implies major star formation
at higher redshifts, which is in line with 'downsizing' of the star
formation \citep[e.g.,][]{cowie96}.
The above measurements suggest that the majority of massive
galaxies are fairly quiescent, while of the rest a considerable fraction
of galaxies are experiencing active star formation, especially at
higher redshifts ($\sim$30 per cent at $z \sim 1$). Such active star formation
in massive galaxies is also reported in \citet{conselice07}, who
find that nearly half of their massive ($M_{\star} > 10^{11} M_{\odot}$) galaxies at
$0.4 < z < 1.4$ are detected in the {\it Spitzer Space Telescope}/MIPS
24-$\mu$m band and the average star-formation rate amounts to
$\sim50 M_{\odot}$ yr$^{-1}$. Our measurements of blue fractions indicate that the
star-formation activity in massive galaxies is gradually quenched toward
the local Universe, leaving the most massive galaxies on the
red sequence. Star-formation quenching processes above a certain
stellar mass limit are actually proposed, such as the internal feedback
of mass assembly caused by active galactic nuclei \citep[e.g.,][]{silk98, granato04, springel05}.
The above scenario is consistent with our primary results
that the active bottom-up formation of massive galaxies is going on
during $0 < z < 1$.
\begin{table}
\caption{The blue population fraction.}
\label{tab:frac_blue}
\begin{tabular}{@{}lccc}
\hline
& & log $M_{\star}$ & \\
Redshift & $<$ 11.0 (10.5 -- 11.0) & 11.0 -- 11.5 & 11.5 -- 12.0\\
\hline
0.2 -- 0.4 & 0.54 $\pm$ 0.01 (0.36 $\pm$ 0.01) & 0.12 $\pm$ 0.05 & $<$ 0.06\\
0.4 -- 0.6 & 0.47 $\pm$ 0.01 (0.43 $\pm$ 0.01) & 0.25 $\pm$ 0.03 & 0.18 $\pm$ 0.03\\
0.6 -- 0.8 & 0.51 $\pm$ 0.01 (0.43 $\pm$ 0.01) & 0.28 $\pm$ 0.02 & 0.19 $\pm$ 0.06\\
0.8 -- 1.0 & 0.59 $\pm$ 0.01 (0.56 $\pm$ 0.01) & 0.29 $\pm$ 0.02 & 0.21 $\pm$ 0.03 \\
\hline
\end{tabular}
\end{table}
\begin{figure}
\includegraphics[width=84mm]{f10.eps
\caption{The fraction of the blue population in $< 10^{11.0} M_{\odot}$ (grey solid),
$10^{11.0-11.5} M_{\odot}$ (green) and $10^{11.5-12.0} M_{\odot}$ (orange) galaxies, respectively,
as a function of redshift.}
\label{cmd2}
\end{figure}
Finally, we comment on the compatibility of the present results
with previous studies. There are a number of studies of luminous
red galaxies (LRGs) at $z < 1$ \citep[e.g.,][]{brown07, brown08, cool08}
covering up to $\sim$10 deg$^2$. Authors consistently suggest
that the LRGs show little evolution in number density since $z \sim 1$.
However, the mass-to-optical luminosity ratio of galaxies has a significant
scatter even for the massive systems, so that galaxies with a
certain stellar mass are not quite the equivalent population of galaxies
with a certain optical luminosity. This leads to a consequence of
most significance for the most massive galaxies: a small portion of
the less massive, much more numerous galaxies could contaminate
the luminous class of galaxies if their mass-to-luminosity ratios
were slightly less than the average, and thus could easily dominate
the luminous population. Therefore a subtle (in absolute amplitude)
change in the number of most massive galaxies could be drowned
out in the measured evolution of the LRGs.
Studies also exist of massive galaxies at $z < 1$ with stellar mass
measurements based on infrared photometry \citep[e.g.,][]{conselice07, ilbert09},
although these studies cover a much smaller
field of view ($\la$ 1.5 deg$^2$) than ours. In contrast to the present results,
they report little evolution in number of the most massive
($> 10^{11.5} M_{\odot}$) galaxies. At least part of the discrepancy could be
due to small-number statistics and cosmic variance. Actually, while
we find $\sim$1000 samples of the most massive galaxies in each redshift
bin from our 55.2 deg$^2$, the number of samples observed over
$\sim$1.5 deg$^2$ should be only $\sim$30. We estimate the effects of cosmic
variance by dividing our total field into small subfields, each
covering $\sim$ 1.5 deg$^2$, and measure the number-density fluctuations
of massive galaxies among the subfields. As a result, we find that
the number densities of the most massive galaxies measured over
$\sim$ 1.5 deg$^2$ can fluctuate by up to a factor of a few. However, we
are not sure whether the above uncertainties alone can account
for the discrepancy between the present results and previous ones.
The number-density measurements at the steep high end of the
galaxy stellar mass function could be heavily affected by contamination
arising from less massive galaxies, thus quite accurate analysis
is required to unveil the subtle evolution of the most massive
galaxies.
In essence, our measurements provide a unique opportunity to
investigate the mean properties and evolution of the most massive
galaxies, owing to the reliable estimates of photometric redshifts
and stellar masses conducted over an unprecedentedly large field
of view. What is observationally clear is that we have discovered
a substantial deficit of the most massive galaxies out to $z = 1$
compared with the local Universe. The analysis of the rest-frame
$U - V$ colour distributions indicates that star-formation activity
might be responsible for the active build-up of these systems, while
it is possible that a so-called dry merger \citep[e.g.,][]{bell04, vandokkum05}
is the main driver of the evolution. Actually, some
observations suggest that a substantial fraction of massive early-type
galaxies go through active evolution in terms of the galaxy structure
as well as the star formation since $z \sim$ 1 \citep[e.g.,][]{treu05, vanderwel08}.
The present results provide crucial evidence
of hierarchical galaxy formation, the missing piece of observation
required to chart a course for future theoretical models based on
$\Lambda$CDM theory.
\section{Summary}
We present an analysis of $\sim$60 000 massive galaxies with stellar
masses $10^{11} M_{\odot} < M_{\star} < 10^{12} M_{\odot}$ in an unprecedentedly large
field of view of 55.2 deg$^2$. The galaxies are drawn from the UKIDSS
Large Area Survey K-band images on the SDSS southern equatorial
stripe. We have created deep-stacked $u$, $g$, $r$, $i$ and $z$-band images
from the SDSS Supplemental and Supernova Survey image frames,
which results in $\sim$90 per cent counterparts of the $K$-band sources in
the $g$, $r$, $i$ and $z$ bands. We also exploit the redshift surveys conducted
on the SDSS southern equatorial stripe, namely the VIMOS-VLT
Deep Survey and the DEEP2 Redshift Survey, in order to obtain
accurate photometric redshifts and associated uncertainties for the
galaxies. Stellar masses are estimated by comparing the observed
broad-band colours with stellar population synthesis models.
In each of the redshift bins $z$ = 0.2--0.4, 0.4--0.6, 0.6--0.8 and
0.8--1.0, we obtain $\sim$10 000 and $\sim$1 000 galaxies with stellar masses
$10^{11.0} M_{\odot} < M_{\star} < 10^{11.5} M_{\odot}$ and $10^{11.5} M_{\odot} < M_{\star} < 10^{12.0} M_{\odot}$,
respectively. The galaxies are almost completely detected out to
$z = 1$, and form by far the largest sample of massive galaxies
reaching to the Universe at about half its present age. We find that
the most massive ($10^{11.5} M_{\odot} < M_{\star} < 10^{12.0} M_{\odot}$) galaxies have
experienced rapid growth in number since $z = 1$, while the number
densities of less massive systems show rather mild evolution.
Such a hierarchical trend of evolution is consistent with the predictions
of the current semi-analytic galaxy formation model based on
$\Lambda$CDM theory. While the majority of the massive galaxies are red-sequence
populations, we find that a considerable fraction are blue
star-forming galaxies. The blue fraction is less in more massive systems
and decreases toward the local Universe, leaving the red, most
massive galaxies at low redshifts, which further supports the idea of
active bottom-up formation of these populations during $0 < z <1$.
The present results provide strong evidence that galaxy formation
proceeds in a hierarchical way, and place stringent observational
constraints on future theoretical models.
\section*{Acknowledgments}
We are grateful to K. Shimasaku, K. Kohno, J. Makino, N. Yasuda
and N.Yoshida for insightful discussions and suggestions. We thank
the referee for many useful comments that have helped to improve
this paper. YM acknowledges Grant-in-Aid from the Research Fellowships
of the Japan Society for the Promotion of Science (JSPS)
for Young Scientists. This work was supported by Grants-in-Aid
for Scientific Research (17104002, 21840027), Specially Promoted
Research (20001003) and the Global COE Program of Nagoya University
'Quest for Fundamental Principles in the Universe (QFPU)'
from JSPS and MEXT of Japan.
This publication makes use of data products from the Two-
Micron All-Sky Survey, which is a joint project of the University
of Massachusetts and the Infrared Processing and Analysis
Center/California Institute of Technology, funded by the National
Aeronautics and Space Administration and the National Science
Foundation. IRAF is distributed by the National Optical Astronomy
Observatories, which are operated by the Association of Universities
for Research in Astronomy, Inc., under cooperative agreement
with the National Science Foundation. Funding for the SDSS and
SDSS-II has been provided by the Alfred P. Sloan Foundation, the
Participating Institutions, the National Science Foundation, the US
Department of Energy, the National Aeronautics and Space Administration,
the Japanese Monbukagakusho, the Max Planck Society
and the Higher Education Funding Council for England. The SDSS
Web Site is http://www.sdss.org/. The SDSS is managed by the Astrophysical
Research Consortium for the Participating Institutions.
The Participating Institutions are the American Museum of Natural
History, Astrophysical Institute Potsdam, University of Basel, University
of Cambridge, CaseWestern Reserve University, University
of Chicago, Drexel University, Fermilab, the Institute for Advanced
Study, the Japan Participation Group, Johns Hopkins University, the
Joint Institute for Nuclear Astrophysics, the Kavli Institute for Particle
Astrophysics and Cosmology, the Korean Scientist Group, the
Chinese Academy of Sciences (LAMOST), Los Alamos National
Laboratory, the Max-Planck-Institute for Astronomy (MPIA), the
Max-Planck-Institute for Astrophysics (MPA), New Mexico State
University, Ohio State University, University of Pittsburgh, University
of Portsmouth, Princeton University, the United States Naval
Observatory and the University of Washington.
|
\section{Generalized polynomial functions\\
and generalized semialgebraic sets}
\label{signomials}
We write $\mathbb R_+=[0,\infty)$ and $\mathbb R_{++}=(0,\infty)$,
endowed with the usual, order topology. And the Cartesian product,
$\mathbb R_{++}^2:=\mathbb R_{++}\times\mathbb R_{++}$ will
be endowed with the usual, Euclidean topology.
\begin{definition}
\label{gpf}
A {\em generalized polynomial function\/} $a(x,y)$ of two
variables is a function $a:\mathbb R_{++}^2\to\mathbb R$
of the form
\begin{equation}
\label{signomial}
a:=a(x,y):=
c_1x^{\alpha_{1,1}}y^{\alpha_{1,2}}+
c_2x^{\alpha_{2,1}}y^{\alpha_{2,2}}+\dotsb+
c_mx^{\alpha_{m,1}}y^{\alpha_{m,2}},
\end{equation}
where $m\in\mathbb N:=\{0,1,2,\ldots\}$, the ``coefficients"
$c_i$ of $a$ are nonzero elements of $\mathbb R$,
and the (binary) ``exponents" $\alpha_i:=
(\alpha_{i,1},\alpha_{i,2})$
of $a$ are distinct elements of $\mathbb R^2$.
We write $\mathbb R[\mathbb R^2]$ for the ring (actually,
it is a group ring) of all generalized polynomial
functions $a:\mathbb R_{++}^2\to\mathbb R$.
\end{definition}
Thus, generalized polynomial functions (sometimes called
``signomial'' functions) of two variables can be defined, roughly,
as ``real polynomial functions on $\mathbb R_{++}^2$
with arbitrary real exponents.'' A simple example is
$a(x,y)=y-x^\pi$.
Generalized polynomial functions of two variables are
clearly real analytic on $\mathbb R_{++}^2$.
See
\cite{Delzell 2008}
for background on the general properties
and the history of generalized polynomials (in any number
of variables), and some motivation for studying them.
\begin{definition}
\label{sss}
We call a subset $A\subseteq\mathbb R_{++}^2$
a {\it generalized semialgebraic set\/}, or a {\it
semisignomial set\/}, if it is of the form
$\bigcup_{j=1}^JS_j$, where $J\in\mathbb N$ and each $S_j$
is a ``basic semisignomial'' set, i.e., one of the form
\begin{equation}
S_j=\{\,(x,y)\in\mathbb R_{++}^2\mid
f_j(x,y)=0,\ g_{j,1}(x,y)>0,\ldots,g_{j,K_j}(x,y)>0\,\},
\label{gss}
\end{equation}
where each $K_j\in\mathbb N$ and the $f_j$ and $g_{jk}$
are generalized polynomials.
\end{definition}
(Recall that ordinary semialgebraic subsets of $\mathbb R^2$
or $\mathbb R^n$ are defined analogously, but with the
$f_j$ and $g_{jk}$ being (ordinary) polynomials.)
\section{Piecewise generalized polynomial functions}
\begin{definition}
\label{pgp}
We call a function $h(x,y):\mathbb R_{++}^2\to\mathbb R$ a
{\it piecewise generalized polynomial function\/} of two
variables if there exist $g_1,\ldots,g_l\in\mathbb R
[\mathbb R^2]$ \eqref{gpf} such that the subsets
\begin{equation}
\label{A_i}
A_i:=\{\,(x,y)\in\mathbb R_{++}^2\mid h(x,y)=g_i(x,y)\,\}
\end{equation}
are generalized semialgebraic and cover $\mathbb R_{++}^2$, i.e.,
$\mathbb R_{++}^2=\bigcup_iA_i$.
We may, and shall, assume that the $g_i$ are distinct.
\end{definition}
\begin{example}
\label{example of h}
~\nopagebreak
\begin{picture}(110,110)(-180,-10)
\put(-180,50)
{$\displaystyle
h(x,y):=\begin{cases}
y-x^\pi&\hbox{if }y\ge x^\pi,\\
0 &\hbox{if }y< x^\pi.
\end{cases}$}
{\thicklines
\put(-5,0){\vector(1,0){105}}
\put(0,-5){\vector(0,1){105}}}
\put(-7,-10)0
\put(-10,95){$y$}
\put(95,-10){$x$}
\qbezier( 0, 0 )(13.633, 0 )(20 , 1.027)
\qbezier(20, 1.027)(31.265, 2.845)(40 , 9.065)
\qbezier(40, 9.065)(50.758,16.724)(60 ,32.403)
\qbezier(60,32.403)(70.543,50.290)(80 ,80 )
\qbezier(80,80 )(82.836,88.910)(85.6,98.947)
\put(80,80){\circle*4}
\put(55,80){$(1,1)$}
\put(77,60){$y=x^\pi$}
\put(10,50){$h=y-x^\pi$}
\put(65,15){$h=0$}
\end{picture}
\end{example}
The following, technical lemma will not be needed until
Proposition~\ref{one variable} and Lemma~\ref{finer partition}
below, and can be skipped on a first reading.
In it, for any set $A$ in $\mathbb R_{++}^2$, we shall write $A^\circ$
for the interior of $A$.
\begin{lemma}
\label{union}
Let $A_1,\ldots,A_l$ be as in \eqref{pgp}.
$(1)$ \vrule width0pt depth15pt
$\displaystyle\bigcup_{i=1}^lA_i^\circ$
is dense in $\mathbb R_{++}^2$.
$(2)$ \vrule width0pt depth5pt
$A_i^\circ\cap A_j^\circ=\emptyset$ for $i\ne j$.
$(3)$ If $h$ is continuous, then each $A_i$ is closed, whence
$\overline{A_i^\circ}\subseteq A_i$.
$(4)$ If $h$ is continuous, then
$\displaystyle
\bigcup_{i=1}^lA_i^\circ=\mathbb R_{++}^2\setminus
\!\!\bigcup_{1\le i<j\le l}\!\!\bigl(\overline{A_i^\circ}
\cap\overline{A_j^\circ}\bigr)$.
$(5)$ Suppose $h$ is continuous, and $E$ is a connected subset
of $\mathbb R_{++}^2$ such that for each $(x,y)\in E$, the $l$
values $g_1(x,y),g_2(x,y),\ldots,g_l(x,y)$ are distinct.
Then there exists an $i\in\{1,2,\ldots,l\}$
such that $E\subseteq A_i^\circ$ $($in particular, such that $h=g_i$
throughout $E)$. This $i$ is unique in case $E\ne\emptyset$.
\end{lemma}
\begin{proof}
(1) By \eqref{sss}, $\bigcup_iA_i$ is a combined, but still finite,
union of suitable basic semisignomial sets $S_j$ as in \eqref{gss}.
Let $T$ be the union of those $S_j$ for which $f_j\not\equiv0$;
thus, $T\subseteq Z(F):=\{\,(x,y)\in\mathbb R_{++}^2\mid F(x,y)=0\,\}$,
where $F$ is the product of those $f_j$'s.
$\mathbb R_{++}^2\setminus Z(F)$ is dense in $\mathbb R_{++}^2$,
by the identity theorem for real analytic functions.
{\it A fortiori\/}, $\mathbb R_{++}^2\setminus T$ is also
dense in $\mathbb R_{++}^2$. The union $U$ of the other $S_j$'s
(viz., those for which $f_j\equiv0$) must contain
$\mathbb R_{++}^2\setminus T$ (since $T\cup U=\bigcup_iA_i=\mathbb R_{++}^2$
\eqref{pgp}), and so $U$ is also dense in $\mathbb R_{++}^2$.
But $\bigcup_iA_i^\circ\supseteq U$.\footnote
{In fact, $\bigcup_iA_i^\circ=U$. But we don't need this.}
(2) If $A_i^\circ\cap A_j^\circ\ne\emptyset$, then $g_i$ would
agree with $g_j$ on a nonempty open set (by \eqref{A_i}),
and hence on all of $\mathbb R_{++}^2$ (again by the identity
theorem), contradicting the distinctness
of the $g_i$ in \eqref{pgp}.\footnote
{And if $g_i$ agrees with $g_j$ on all of $\mathbb R_{++}^2$,
then the coefficients of $g_i$ and $g_j$ (i.e., the $c$'s in
\eqref{signomial} above) would agree, too, by
\cite[Remark~4.3]{Delzell 2008}.}
(3) Obvious.
(4) $\subseteq$. Let $(x,y)\in A_i^\circ$ and suppose $j\ne i$.
It is enough to show that $(x,y)\notin\overline{A_j^\circ}$.
There exists an open disk in $A_i$ about $(x,y)$. In fact,
this disk is in $A_i^\circ$, and hence is disjoint from
$A_j^\circ$, by (2) above. Therefore
$(x,y)\notin\overline{A_j^\circ}$.\thinspace\footnote
{This half of the proof of (4) does not require the hypothesis
that $h$ be continuous.}
$\supseteq$. Suppose $(x,y)\in\mathbb R_{++}^2\setminus
\bigcup_iA_i^\circ$. For $r\in\mathbb R_{++}$ with
$r\le\min\{x,y\}$, let $B_r$ denote the open disk in
$\mathbb R_{++}^2$ of radius $r>0$ about $(x,y)$, and let
$I(r)=\{\,i\in\{1,2,\ldots,l\}\mid B_r\cap A_i^\circ\ne\emptyset\,\}$.
Then for every $r$, $|I(r)|\ge1$, by (1) above.
In fact, $I(r)>1$. Otherwise, for some $i$, $A_i^\circ\cap B_r$
would be dense in $B_r$ (by (1) again), whence $B_r=
\overline{A_i^\circ}\cap B_r\subseteq A_i\cap B_r$ (by (3)),\footnote
{In fact, this inclusion is actually an equality.}
i.e., $B_r\subseteq A_i$, whence $(x,y)\in A_i^\circ$,
contradiction. Now, for any $s\in\mathbb R_{++}$ with $s<r$,
$I(s)\subseteq I(r)$; i.e., the finite set $I(r)$ decreases
monotonically with $r$, and yet always has cardinality $\ge2$.
Thus, there exist at least two indices $i<j$ such that for
every $r\in(0,\min\{x,y\})$, $B_r$ meets $A_i^\circ$ and
$A_j^\circ$. Therefore $(x,y)\in\overline{A_i^\circ}\cap\overline
{A_j^\circ}$.
(5) The distinctness hypothesis of (5) can be rephrased as
\begin{equation*}
E\cap\bigcup_{i<j}(A_i\cap A_j)=\emptyset.
\end{equation*}
{\it A fortiori\/},
$E\cap\bigcup_{i<j}\bigl(\overline{A_i^\circ}\cap\overline{A_j^\circ})
=\emptyset$, using (3). By (4), $E\subseteq\bigcup_iA_i^\circ$.
The existence of the desired $i$ now follows from (2) and the
hypotheses that $E$ is connected. The uniqueness of $i$ in case
$E\ne\emptyset$ also follows from (2).
\end{proof}
\begin{remark}
\label{announcement of Remark "aside"}
In Remark~\ref{aside} below, we shall use \eqref{union} above
to see that when a piecewise
generalized polynomial function $h$ is continuous, each $A_i$
in \eqref{pgp} can automatically be taken to be a generalized
semialgebraic set; it is not necessary to include that
condition as a hypothesis in \eqref{pgp}.
\end{remark}
The set of piecewise generalized polynomial functions is closed
under differences and products, and so forms a ring;
it is also closed under pointwise suprema and infima,
and so forms an $l$-ring under those lattice operations.
(This ring is, of course, even an $f$-ring.) The continuous
functions in this $f$-ring comprise a sub-$f$-ring.
(See, e.g., \cite{Birkhoff et al. 1956} or
\cite{Henriksen et al. 1962} for background
on $l$-rings and $f$-rings.)
\section{Statement and discussion of the main result}
\begin{theorem}[Main Theorem: The Pierce-Birkhoff conjecture for generalized
polynomials in two variables]
\label{PBCSIG}
If $h:\mathbb R_{++}^2\to\mathbb R$ is continuous and piecewise
generalized polynomial, then $h$ is a $($pointwise\/$)$ sup of
infs of finitely many generalized polynomial functions; i.e.,
\begin{equation}
\label{SIP}
h(x,y)=\sup_j\inf_kf_{jk}(x,y)\text{ on }\mathbb R_{++}^2,
\end{equation}
for some finite number of generalized polynomials
$f_{jk}$. \rm(The converse is easy.)
\end{theorem}
\begin{example}
For the $h$ in Example~\ref{example of h} above,
$h(x,y)=\sup\{0,\,y-x^\pi\}$.
\end{example}
The representation of $h$ in the form \eqref{SIP}
makes both the continuity and the piecewise generalized
polynomial character of $h$ {\it obvious\/}.
For ordinary polynomials in $\mathbb R[X,Y]$
and ordinary piecewise polynomial functions on $\mathbb R^2$,
the analog of Theorem~\ref{PBCSIG} above was first proved by
L.~Mah\'e \citey{Mahe 1984} and Efroymson (unpublished),
independently. The statement and proofs of the
Mah\'e-Efroymson theorem generalize easily to the situation
where $\mathbb R$ is replaced by an arbitrary real closed
field $R$ (furnished with the topology induced by the unique
ordering on $R$). But the fact that then the coefficients
of the $f_{jk}$ in the Mah\'e-Efroymson theorem may be
taken to lie in the subfield of $R$ generated by the
coefficients of the $g_i$ defining $h$ (in the analog
of \eqref{pgp}), was not trivial, and was proved in
\cite{Delzell 1989}.
The extension of the Mah\'e-Efroymson
theorem to functions of three or more variables (like the
extension of \eqref{PBCSIG} above) remains unproved
and unrefuted; it is known as the Pierce-Birkhoff Conjecture
(first formulated in \cite{Birkhoff et al. 1956}).
In our proof of Theorem~\ref{PBCSIG} below, we shall make
no attempt to indicate which steps generalize easily to the
case where $n>2$ (though many of those steps do).
The first reason for this is that the
notation is often simpler when $n=2$. The second reason is that,
considering the many mathematicians who have tried to prove the
Pierce-Birkhoff Conjecture for $n>2$, we now lean toward the
opinion that it and Theorem~\ref{PBCSIG} are false for $n>2$.
In 1987 we proved that for all $n\ge1$ and every real
closed field $R$, if $h:R^n\to R$ is ``piecewise-rational''
(i.e., if there are rational functions $g_1,\ldots,g_l\in R(X)$
such that the sets $A_i:=\{\,x\in R^n\mid g_i(x)
\text{ is defined and }h(x)=g_i(x)\,\}$
are s.a.\ and cover $R^n$), then there are finitely many
$f_{jk}\in R(X)$ and there is a $k\in R[X_1,\ldots,X_n]
\setminus\{0\}$ such that for all $x\in R^n$ where
$k(x)\ne0$ (i.e., for ``almost all'' $x\in R^n$),
each $f_{jk}(x)$ is defined and $h(x)=\sup_j\inf_kf_{jk}(x)$;
this is true even if $h$ is not continuous.
This result was announced in
\cite[p.~659]{Delzell 1989}, and proved in \citey{Delzell 1990}.
Madden gave an ``abstract'' version of this result
that applies to arbitrary fields (and not just
$R(X)$); see \cite{Madden 1989}. In \citey{Delzell 2005}
we proved an analog of our 1987 result, for
``generalized piecewise-rational functions''
(i.e., functions that are, piecewise,
quotients of generalized polynomial functions).
The rest of this paper will be devoted to the proof
of Theorem~\ref{PBCSIG}. In \S4 we shall develop the
necessary one-variable machinery; in \S5 we shall
deal with the additional difficulties arising in the
two-variable situation.
\section{One-variable methods}
We imitate Mah\'e's proof as much as possible.
We are given a continuous function
\begin{equation}
h(x,y)=
\begin{cases}
g_1(x,y)&\text{if \ }(x,y)\in A_1\\
\kern14pt\vdots&\kern25pt\vdots\\
g_l(x,y)&\text{if \ }(x,y)\in A_l,
\end{cases}
\label{h}
\end{equation}
where, as in \eqref{pgp}, the $g_i$ are generalized polynomials
and the $A_i$ cover $\mathbb R_{++}^2$.
(Recall from Remark~\ref{announcement of Remark "aside"}
above that the $A_i$ are also, automatically,
generalized semialgebraic; but we don't use this.)
As before, we assume the $g_i$ are distinct.
Write each $a(x,y)\in\mathbb R[\mathbb R^2]\setminus\{0\}$
\eqref{gpf} in the form
\begin{equation}
a_1(x)y^{\beta_1}+a_2(x)y^{\beta_2}+\dotsb+
a_K(x)y^{\beta_K},
\label{K}
\end{equation}
where $K\ge1$, \ $\beta_1<\dotsb<\beta_K\in\mathbb R$,
and each $a_i$ is a nonzero generalized polynomial in $x$.
This representation is unique.
Let ${\mathcal A}=\{\,g_i-g_j\mid1\le i<j\le l\,\}$.
Let $\mathcal B$ be the smallest subset of $\mathbb R[\mathbb R^2]$
containing $\mathcal A$ and closed under the following two
operations, for each $a(x,y)\in\mathcal B$
for which $K>1$ in \eqref{K}:
\begin{align}
a\vrule width0pt depth20pt
&\mapsto
\begin{cases}
\displaystyle a':=\frac{\partial a}{\partial y}
&\text{if }\beta_1=0\text{, and}\\
\displaystyle y^{-\beta_1}a(x,y)&\text{if }\beta_1\ne0\
\footnotemark;\text{ and}
\end{cases}
\label{closed under partial}\\
a&\mapsto
\begin{cases}
\displaystyle r:=r_a(x,y)=a(x,y)-\frac y{\beta_K}\cdot a'(x,y)
&\text{if }\beta_1=0,\footnotemark\text{ and}\\
a&\text{if }\beta_1\ne0.
\end{cases}
\label{r}
\end{align}
\addtocounter{footnote}{-1}
\footnotetext{This trick (of dividing by $y^{\beta_1}$)
was first used by Sturm
\citey{Sturm 1829
.}
\stepcounter{footnote}
\footnotetext{Here we use $\beta_K\ne0$, which follows from
$\beta_1=0$ and $K>1$.}
\begin{remark}
\label{no x involved}
Suppose no $g_i$ involves the variable $x$;
i.e., each $g_i$ is a function of $y$ alone, and is constant in $x$.
Then the same is, of course, true for each $a\in\mathcal A$;
in fact, the same is true even for each $a\in\mathcal B$, in view
of \eqref{closed under partial} and \eqref{r}.
\end{remark}
\begin{lemma}
\label{K-1 terms}
For each $a\in\mathcal B$ for which $K>1$ and $\beta_1=0$,
$a'(x,y)$ and $r_a$ each have exactly $K-1$ $y$-terms.
Consequently, $\mathcal B$ is finite.
\end{lemma}
\begin{proof}
This is clear for $a'(x,y)$.
For $r_a$, observe (a)~that the $K^{\rm th}$ $y$-term
$a_K(x,y)y^{\beta_K}$ in $a$ \eqref{K} is cancelled
out by the $y$-term
\begin{equation*}
\frac{y}{\beta_K}
\bigl(\beta_K\,a_K(x,y)\,y^{\beta_K-1}\bigr)
\end{equation*}
in
\begin{equation}
\label{partial}
\frac{y}{\beta_K}\cdot a'(x,y),
\end{equation}
and
(b)~that the other $y$-terms of
\eqref{partial} involve the $y$-exponents
$\beta_1,\ldots,\beta_{k-1}$, but with coefficients
different from those of the corresponding $y$-terms
of $a$ (since for each $i<K$, $\beta_i/\beta_K\ne1$).
\end{proof}
\begin{lemma}
\label{projection}
There exist $L\in\mathbb N$ and
$\gamma_1<\gamma_2<\dotsb<\gamma_L\in\mathbb R_{++}$ such that,
writing $\gamma_0=0$ and $\gamma_{L+1}=\infty$,
for each $a\in\mathcal B$ and for each $p\in\{0,1,\ldots,L\}$,
the zeros of $a(x,y)$ in the $p$th vertical half
strip $H_p:=(\gamma_p,\gamma_{p+1})\times\mathbb R_{++}$
are the graphs of
continuous, monotonic\footnote{We do not need the monotonicity
of the $\xi_{a,p,j}$ in this paper.\label{monotonic}} ``generalized
semialgebraic''\footnote
{\label{generalized semialgebraic function}We say that
a function is {\it generalized semialgebraic\/}
if its graph, in the product space, is a generalized semialgebraic
set.} functions $y=\xi_{a,p,j}(x)$, $j=1,2,\ldots,s$
$($where $s:=s(a,p)$ satisfies $0\le s\le K\>$\footnote
{Here, $K$ is as in \eqref{K};
in fact, $s$ is even bounded by the number of {\it
alternations in sign\/} in the sequence $a_0(x),\ldots,a_K (x)$,
by Sturm's generalization \citey{Sturm 1829},
to one-variable generalized polynomials, of the Fourier-Budan theorem
(which contains Descartes' rule of signs as a special case).}$)$
with
\begin{equation*}
(0\mathrel{<)}\xi_{a,p,1}<\cdots<\xi_{a,p,s}
\text{ on }(\gamma_p,\gamma_{p+1}).
\end{equation*}
Moreover,
$\forall a_1,a_2\in\mathcal B$, \ $\forall p\le L$, \
$\forall j_1\le s(a_1,p)$, \
$\forall j_2\le s(a_2,p)$, throughout $(\gamma_p,\gamma_{p+1})
\subseteq\mathbb R_{++}$, only one of the following
three relations holds:
\begin{align}
\begin{split}
\xi_{a_1,p,j_1}&<\xi_{a_2,p,j_2},\\
\xi_{a_1,p,j_1}&=\xi_{a_2,p,j_2}\text{, or}\\
\xi_{a_1,p,j_1}&>\xi_{a_2,p,j_2}.
\label{uniform trichotomy}
\end{split}
\end{align}
\end{lemma}
Lemma~\ref{projection} and its Corollary~\ref{corollary}
are illustrated in Figure~\ref{cylindrical figure},
which also shows the stack of open connected sets $D_{2,1},
D_{2,2},D_{2,3}$ whose union is a dense open subset of $H_2$
(looking ahead to \eqref{corollary} below).
\bigskip
\begin{figure}[htb]
\setlength\parindent{0pt}
\begin{picture}(327,200)(-143.5,-20
{\thicklines
\put(-110, 0){\vector(1,0){280}}
\put(-100,-15){\vector(0,1){200}}
}
\put(-108,-10)0
\put(-113,175){$y$}
\put(130,-12){$x$}
\qbezier(-30,100)(-30,140)( 40,140)
\qbezier( 40,140)(110,140)(110,100)
\qbezier(-30,100)(-30, 60)( 40, 60)
\qbezier( 40, 60)(110, 60)(110,100)
\qbezier(20,100)(70,100)(120,160)
\qbezier(20,100)(70,100)(120, 40)
\put(-60, -5){\line(0,1){10}}
\put(-60, 15){\line(0,1){10}}
\put(-60, 35){\line(0,1){10}}
\put(-60, 55){\line(0,1){10}}
\put(-60, 75){\line(0,1){10}}
\put(-60, 95){\line(0,1){10}}
\put(-60,115){\line(0,1){10}}
\put(-60,135){\line(0,1){10}}
\put(-60,155){\line(0,1){10}}
\put(-30, -5){\line(0,1){10}}
\put(-30, 15){\line(0,1){10}}
\put(-30, 35){\line(0,1){10}}
\put(-30, 55){\line(0,1){10}}
\put(-30, 75){\line(0,1){10}}
\put(-30, 95){\line(0,1){10}}
\put(-30,115){\line(0,1){10}}
\put(-30,135){\line(0,1){10}}
\put(-30,155){\line(0,1){10}}
\put( 20, -5){\line(0,1){10}}
\put( 20, 15){\line(0,1){10}}
\put( 20, 35){\line(0,1){10}}
\put( 20, 55){\line(0,1){10}}
\put( 20, 75){\line(0,1){10}}
\put( 20, 95){\line(0,1){10}}
\put( 20,115){\line(0,1){10}}
\put( 20,135){\line(0,1){10}}
\put( 20,155){\line(0,1){10}}
\put(91.5, -5){\line(0,1){10}}
\put(91.5, 15){\line(0,1){10}}
\put(91.5, 35){\line(0,1){10}}
\put(91.5, 55){\line(0,1){10}}
\put(91.5, 75){\line(0,1){10}}
\put(91.5, 95){\line(0,1){10}}
\put(91.5,115){\line(0,1){10}}
\put(91.5,135){\line(0,1){10}}
\put(91.5,155){\line(0,1){10}}
\put(110, -5){\line(0,1){10}}
\put(110, 15){\line(0,1){10}}
\put(110, 35){\line(0,1){10}}
\put(110, 55){\line(0,1){10}}
\put(110, 75){\line(0,1){10}}
\put(110.5, 95){\line(0,1){10}}
\put(110,115){\line(0,1){10}}
\put(110,135){\line(0,1){10}}
\put(110,155){\line(0,1){10}}
\put(-60,31){\circle*{3}}
\put(-36, 7){$a(x,y)=0$}
\put(-39,10){\line(-1,0){10}}
\put(-49,10){\vector(-1,2){9}}
\put(-28,135){$\xi_{b,2,2}$}
\put(- 5,125){($=\xi^{2,2}$)}
\put(-28, 64){$\xi_{b,2,1}$}
\put(- 5, 68){($=\xi^{2,1}$)}
\put( 25,144){$\xi_{b,3,2}$ ($=\xi^{3,4}$)}
\put( 38,111){$\xi_{c,3,2}$}
\put( 67,105){($=\xi^{3,3}$)}
\put( 38, 88){$\xi_{c,3,1}$}
\put( 67, 88){($=\xi^{3,2}$)}
\put( 25, 52){$\xi_{b,3,1}$ ($=\xi^{3,1}$)}
\put( 88,149){$\xi_{c,4,2}$}
\put(125,131){$\xi_{b,4,2}$}
\put(117,120){($=\xi^{4,3}$)}
\put(122,133){\line(-1,0){10}}
\put(112,133){\vector(-1,-1){10}}
\put(125, 63){$\xi_{b,4,1}$}
\put(117, 52){($=\xi^{4,2}$)}
\put(122, 66){\line(-1,0){10}}
\put(112, 66){\vector(-1,1){10}}
\put( 90, 47){$\xi_{c,4,1}$}
\put(120,150){$\xi_{c,5,2}$ ($=\xi^{5,2}$)}
\put(115, 33){$\xi_{c,5,1}$ ($=\xi^{5,1}$)}
\put(-57,172){$s(1)$}
\put( 90,171){$s(4)$}
\put(-97,170){$s(0)=0$}
\put(-45,165){$=0$}
\put(-20,170){$s(2)=3$}
\put( 40,170){$s(3)=5$}
\put( 94.5,163){$=5$}
\put(125,170){$s(5)=3$}
\put(-83,17){$H_0$}
\put(-51,17){$H_1$}
\put(-11,17){$H_2$}
\put( 48,17){$H_3$}
\put( 95,17){$H_4$}
\put(130,17){$H_5$}
\put(-63,-12){$\gamma_1$}
\put(-33,-12){$\gamma_2$}
\put( 16,-12){$\gamma_3$}
\put( 88,-12){$\gamma_4$}
\put(106,-12){$\gamma_5$}
\put(-16, 37){$D_{2,1}$}
\put(-16, 97){$D_{2,2}$}
\put(-16,155){$D_{2,3}$}
\end{picture}
\caption{Illustrating Lemma~\ref{projection} and Corollary~\ref
{corollary} by showing the zeros in $\mathbb R_{++}^2$
of $a,b,c\in\mathcal B$: the isolated zero
of $a(x,y)$, and the graphs of $y=\xi_{b,p,j}(x)$
and $y=\xi_{c,p,j}(x)$ (which are also the graphs of
$y=\xi^{p,k}(x)$, for suitable $k$).
Here, $L=5$ (the number of $\gamma$'s).}
\label{cylindrical figure}
\end{figure}
\begin{proof}
Miller \citey{Miller 1994} considered a class of
functions $f:\mathbb R^n\to\mathbb R$ that properly contains
the class of (extensions by 0 to $\mathbb R^n$ of)
generalized polynomial functions. Specifically, he
considered terms built up (in a formal language) from
variable symbols $x_1,x_2,\dotsc$ and from constants in
$\mathbb R$ by the usual operation symbols $+$, $-$, and
$\cdot$~, together with the class of operation symbols
$\{\,x_i^r\mid i\ge1,\ r\in\mathbb R\,\}$;
the symbol $x_i^r$ indicates the function
$\mathbb R\to\mathbb R$ defined by
\begin{equation*}
x_i\mapsto
\begin{cases}
x_i^r&\text{if }x_i>0\\
0 &\text{if }x_i\le0.
\end{cases}
\end{equation*}
He considered the structure
\begin{equation*}
\mathbb R_{\text{an}}^{\mathbb R}:=
\bigl(\mathbb R,<,+,-,\,\cdot\,,0,1,(x_i^r)_{r\in\mathbb R,\>i\ge1},\,
\bigl(\tilde f\bigr)_{f\in\mathbb R\{X,n\},n\in\mathbb N}\bigr),
\end{equation*}
where $\bigl(\tilde f\bigr)_{f\in\mathbb R\{X,n\},n\in\mathbb N}$
denotes a certain class of functions $\tilde f:
\mathbb R^n\to\mathbb R$
that are analytic on $[-1,1]^n$. He proved that
the theory of $\mathbb R_{\text{an}}^{\mathbb R}$ admits
quantifier-elimination and analytic cell-decomposition,
and is universally axiomatizable, o-minimal, and
polynomially bounded.
The standard properties of o-minimal theories (cf., e.g.,
\cite{Dries 1998} or \cite{Miller 1994}) imply
that the zeros in $\mathbb R_{++}^2$ of all the various
$a\in\mathcal B$ consist of finitely many
isolated points together with the graphs of finitely
many continuous, monotonic functions $\xi_{a,p,j}:
(\gamma_p,\gamma_{p+1})\to\mathbb R_{++}$ (on suitable
intervals $(\gamma_p,\gamma_{p+1})\subseteq\mathbb R_{++}$)
satisfying \eqref{uniform trichotomy}, as stated in the lemma.
(That the $\xi_{a,p,j}$ are generalized semialgebraic is
just the definition of that term (footnote~\ref{generalized
semialgebraic function} above), since
the $a(x,y)$ are generalized polynomials.)
\end{proof}
\begin{notation}
It will be helpful in \eqref{increasing order} below
if we agree that $\xi_{a,p,0}(x)=0$ and
$\xi_{a,p,s+1}(x)=+\infty$ for all $x\in(\gamma_p,\gamma_{p+1})$,
where $p\in\{0,1,\ldots,L\}$ and $s=s(a,p)$ is as in
Lemma~\ref{projection}.
\label{xi_0}
\end{notation}
\begin{corollary}
Let $L$, $\gamma_0,\ldots,\gamma_{L+1}$, and $H_p$
be as in \eqref{projection
, for some fixed $p\in\{0,1,\ldots,L\}$.
Then the zeros in $H_p$ of all the $a\in\mathcal B$ are
the graphs of continuous, monotonic,
generalized semialgebraic functions $y=\xi^{p,k}(x)$,
$k=1,2,\ldots,s(p)$, where $s(p)$ satisfies
$0\le s(p)\le\sum_{a\in{\mathcal B}}s(a,p)$ $($where $s(a,p)$
is as in \eqref{projection}$)$, and where,
for each $x\in(\gamma_p,\gamma_{p+1})$,
\begin{equation}
0=:\xi^{p,0}(x)<\xi^{p,1}(x)<\cdots<\xi^{p,s(p)}(x)<
\xi^{p,s(p)+1}(x):=\infty.
\label{increasing order}
\end{equation}
Consequently, the sets
\begin{equation*}
D_{p,k}:=\{\,(x,y)\mid
\gamma_p<x<\gamma_{p+1},\>\xi^{p,k}(x)<y<\xi^{p,k+1}(x)\,\},
\end{equation*}
for $k\in\{0,1,\ldots,s(p)\}$, are nonempty, pairwise-disjoint,
generalized semialgebraic cells
$($in particular, they are open and $($pathwise$)$ connected$)$,
and their union is a dense open subset of $H_p$. Moreover,
the $D_{p,k}$ are ``stacked'' one upon the other in the $y$-direction,
so that for any $x\in(\gamma_p,\gamma_{p+1})$ and for any
$(s(p)+1)$-tuple $y_0,y_1,\ldots,y_{s(p)}\in\mathbb R_{++}$
for which each $(x,y_k)\in D_{p,k}$, $y_0<y_1<\cdots<y_{s(p)}$.
\label{corollary}
\end{corollary}
\begin{proof}
The required sequence $\xi^{p,1},\xi^{p,2},\ldots,\xi^{p,s(p)}$ of
functions is just a suitable permutation and relabelling of the set of
functions $\{\,\xi_{a,p,j}\mid a\in{\mathcal B},\>1\le j\le s(a,p)\,\}$.
That a permutation of the $\xi$'s satisfying \eqref{increasing order}
exists follows from \eqref{uniform trichotomy}.
\end{proof}
\begin{proposition}
\label{closed}
The set of suprema of infima of finitely many generalized polynomial
functions is closed under subtraction and multiplication,
and so is a ring.
\end{proposition}
\begin{proof}
This is a special case of a result of
Henriksen and Isbell \citey[Corollary~3.4]{Henriksen et al. 1962}:
If $S$ is a ring of real-valued functions on a set,
then the least lattice of functions that contains $S$
is also a ring. Here we may take $S=\mathbb R[\mathbb R^2]$
\eqref{gpf}.
For the proof of this corollary,
Henriksen and Isbell gave some $f$-ring identities which,
they said, reduce the proof to an exercise; they omitted
the details. \cite{Delzell 1989} gave a sketch of a proof.
The first complete proof of this fact to appear in print was
that of \cite[Theorem~1(B)]{Hager et al. 2010}; their proof incorporates
some simplifications due to Madden, and their statement is a little
more general than the Henriksen-Isbell statement above,
in that now $S$ may be an arbitrary subring of an arbitrary
$f$-ring.
\end{proof}
In the next lemma it will helpful to use the abbreviation
$a^+=\sup\{0,a\}$, for any real-valued function $a$.
\begin{lemma}[Generalized Mah\'e lemma]
\label{Mahe's lemma}
Using the notation of Lemma~\ref{projection} above,
for each $p\in\{0,1,\ldots,L\}$, each $a(x,y)\in\mathcal B$,
and each $j\in\{0,1,\ldots,s\}$ $($where $s=s(a,p)$
as in \eqref{projection}$)$,
there exists a function $c_{a,p,j}(x,y)$
that is a sup of infs of finitely many generalized
polynomials, such that for all $x\in(\gamma_p,\gamma_{p+1})$
and for all $y\in\mathbb R_{++}$,
\begin{equation}
\label{c}
c_{a,p,j}(x,y)=
\begin{cases}
a(x,y)&\text{if }y>\xi_{a,p,j}(x)\text{, and}\\
0 &\text{otherwise.}
\end{cases}
\end{equation}
\end{lemma}
\begin{proof}
Fix any $p\le L$.
We use induction on $K\ge1$,
the number of distinct $y$-exponents occurring in $a$
(recall \eqref{K}). Note that for any $K\ge1$,
we may (in fact, we must) take $c_{a,p,0}=a$;
this handles the case $K=1$,
i.e., the case where $a(x,y)$ is of the form
$a_1(x)y^{\beta_1}$ (which implies $s(a,p)=0$ for each
$p\le L$).
Now assume $K>1$.
We claim that we may assume
\begin{equation}
\label{beta_1=0}
\beta_1=0.
\end{equation}
If not, then write $b(x,y)=y^{-\beta_1}a(x,y)$.
Thus $b\in\mathcal B$, by \eqref{closed under partial}.
Note that $b(x,y)$
has the same positive $y$-roots $\xi$ as $a(x,y)$ has;
thus $s(a,p)=s(b,p)$.
Therefore, if for each $j\le s(b,p)$ we can construct
$c_{b,p,j}$ such that
\begin{equation*}
c_{b,p,j}(x,y)=
\begin{cases}
b(x,y)&\text{if }y>\xi_{b,p,j}(x)\text{, and}\\
0 &\text{otherwise,}
\end{cases}
\end{equation*}
then we may, for each $j\le s(a,p)$ ($=s(b,p)$),
take $c_{a,p,j}(x,y)=y^{\beta_1}c_{b,p,j}(x,y)$; the latter
product is a sup of infs of finitely many generalized
polynomials, since $c_{b,p,j}$ is, and since $y^{\beta_1}>0$
for all $y>0$ (or use \eqref{closed}).
Next, recall that $a'$ \eqref{closed under partial} and
$r_a$ \eqref{r} each have exactly $K-1$ $y$-terms,
by \eqref{K-1 terms} and \eqref{beta_1=0}.
Thus we assume, by the inductive hypothesis,
that for every $k\le s(a',p)$ and $l\le s(r_a,p)$,
we can construct $c_{a',p,k}$ and $c_{r_a,p,l}$
satisfying the appropriate analogs of \eqref{c}.
Note that $c_{a',p,k}$ and $c_{r_a,p,l}$ are, in
particular, continuous (either by their form as in
\eqref{c}, or by the fact that they are sups of
infs of finitely many generalized polynomial functions).
Finally, in order to construct $c_{a,p,j}$, we now use
induction on $j\in\{0,1,2,\ldots,\linebreak[0]
s(a,p)\}$.
We have already constructed $c_{a,p,0}$,
so now we assume that $j\in\{1,2,\ldots,\linebreak[0]
s(a,p)\}$ and that $c_{a,p,j-1}$ has
already been constructed with the properties stated in
Lemma~\ref{Mahe's lemma}.
Throughout the rest of this proof, $x$ will range over
$(\gamma_p,\gamma_{p+1})$.
By the uniform trichotomy in \eqref{uniform trichotomy},
all order relations involving the various $\xi$'s
below will hold uniformly for such $x$;
thus we
usually write, e.g., $\xi_{a,p,j}$
instead of $\xi_{a,p,j}(x)$.\linebreak
Let $k$ be the smallest index such
that $\xi_{a,p,j}\le\xi_{a',p,k}$ (then $1\le k\le
1+s(a',p)$).\linebreak
Let $l$ be the smallest index such
that $\xi_{a',p,k}\le\xi_{r_a,p,l}$ (then $1\le l\le
1+s(r_a,p)$).
Then
\begin{equation}
\label{Rolle}
\xi_{a',p,k}<\xi_{a,p,j+1}\text{\quad
(unless $\xi_{a',p,k}=\infty$), by Rolle's theorem, and}
\end{equation}
\begin{align}
g(x,y):&=\frac y{\beta_K}c_{a',p,k}(x,y)+c_{r_a,p,l}(x,y)\notag\\
&=
\begin{cases}
\ 0&\text{if }0<y<\xi_{a',p,k},\\
\displaystyle\frac y{\beta_K}a'(x,y)=a(x,y)-r_a(x,y)
&\text{if }\xi_{a',p,k}<y<\xi_{r_a,p,l},\\
\vrule width0pt height13pt
\displaystyle\frac y{\beta_K}a'(x,y)+r_a(x,y)
=a(x,y)&\text{if }\xi_{r_a,p,l}<y,
\end{cases}
\label{g}
\end{align}
where \eqref{g} follows from \eqref{r} and from the
definitions of $c_{a',p,k}$ and $c_{r_a,p,l}$.\footnote
{In \eqref{g}, the inequalities in the case-distinctions
$y<\xi_{a',p,k}$, $\xi_{a',p,k}<y<\xi_{r_a,p,l}$, and
$\xi_{r_a,p,l}<y$ are all strict (i.e., they are all $<$,
and not $\le$).
This strictness is necessary because $\xi_{a',p,k}$
and/or $\xi_{r_a,p,l}$ could be $\infty$.
If either or both of the $\xi$'s are finite,
the corresponding inequalities could be relaxed to
nonstrict inequalities (with $\le$). But even without
such a relaxation, \eqref{g} still uniquely determines $g$
even when $y$ is $\xi_{a',p,k}$ or $\xi_{r_a,p,l}$, since
$g$ is continuous for all $y>0$.}
This function $g$ is a supremum of infima of finitely many
generalized polynomial functions, by \eqref{closed}.
If $a'(x,\xi_{a,p,j})=0$, then
\begin{alignat*}2
\xi_{a',p,k}&=\xi_{a,p,j}&&
\text{by the minimality of $k$, and}\\
\xi_{r_a,p,l}&=\xi_{a',p,k}\quad&&
\text{by \eqref{r} and the minimality of $l$.}
\end{alignat*}
Thus we may take $c_{a,p,j}=g$, by \eqref{g}.
Now suppose, on the other hand, that
\begin{equation}
\label{a' ne 0}
a'(x,\xi_{a,p,j})\ne0
\end{equation}
(recall \eqref{uniform trichotomy}). (Then
\begin{equation}
\xi_{a,p,j}<\xi_{a',p,k}.)
\label{xi inequality}
\end{equation}
We may assume that in fact
\begin{equation}
\label{a'>0}
a'(x,\xi_{a,p,j})>0,
\end{equation}
by \eqref{uniform trichotomy}, by replacing $a$
with $-a$, and by the fact that $-c_{-a,p,j}$
($=c_{a,p,j}$) will still be a supremum of infima
of finitely many generalized polynomial functions
if $c_{-a,p,j}$ is, by \eqref{closed}.
Then
\begin{alignat}2
\label{a<0}
a(x,y)&<0\quad&&\text{for}\quad\xi_{a,p,j-1}<y<a_{a,p,j}
\quad\text{and}\\
\label{a>0}
a(x,y)&>0\quad&&\text{for}\quad\xi_{a,p,j}<y<a_{a,p,j+1},
\end{alignat}
by \eqref{a'>0}.
First suppose $\xi_{a',p,k}=\infty$ (i.e.,
$k=1+s(a',p)$). Then $a'(x,y)>0$ for all $y>\xi_{a,p,j}$,
whence $a(x,y)>0$ for all $y>\xi_{a,p,j}$.
Hence we may take $c_{a,p,j}=\inf\{c_{a,p,j-1}^+,a^+\}$,
using also \eqref{a<0}.
Second, suppose $\xi_{a',p,k}<\infty$ (i.e.,
$k\le s(a',p)$). Then
\begin{align
r_a(x,\xi_{a',p,k})
&=a(x,\xi_{a',p,k})-
\frac{\xi_{a',p,k}}{\beta_K}a'(x,\xi_{a',p,k})\ \
\text{(by \eqref{r})}\notag\\
&=a(x,\xi_{a',p,k})-\frac{\xi_{a',p,k}}{\beta_K}\cdot0
\notag\\
&=a(x,\xi_{a',p,k})>0,\quad
\text{by \eqref{a>0},
\eqref{Rolle}, and \eqref{xi inequality}}.
\label{r>0}
\end{align
Then for $\xi_{a',p,k}\le y<\xi_{r_a,p,l}$:
\begin{alignat}2
r_a(x,y)&>0&&\text{by \eqref{r>0} and the choice of $l$,
and}\label{r(x,y)>0}\\
g(x,y)&=a(x,y)-r_a(x,y)\quad&&\text{by \eqref{g}}\notag\\
&<a(x,y)&&\text{by \eqref{r(x,y)>0}.}
\label{1}
\end{alignat}
Then
\begin{equation*}
\sup\{a,g\}=
\left\lbrace
\begin{alignedat}3
&a^+&\quad&\text{if }0<y\le\xi_{a,p,j}&\ &\text{ by \eqref{g}, and}\\
&a & &\text{if }y\ge\xi_{a,p,j} &&
\text{ by \eqref{g}, \eqref{1}, \eqref{Rolle},
and \eqref{a>0}.}
\end{alignedat}
\right.
\end{equation*}
Therefore, we may take $c_{a,p,j}=
\inf\{c_{a,p,j-1}^+,\,\sup\{a,g\}\}$, by \eqref{a<0}.
\end{proof}
\begin{proposition}
\label{one variable}
Let $h$, $\mathcal A$, and $\mathcal B$ be as before Lemma~\ref{K-1 terms},
and let $L$ and $H_p$ be as in Lemma~\ref{projection}, for some fixed
$p\in\{0,1,\ldots,L\}$. Then there is a function $d_p:\mathbb
R_{++}^2\to\mathbb R$ that $(1)$~is a supremum of infima of
finitely many generalized polynomial functions $\in\mathbb
R[\mathbb R^2]$ and $(2)$~coincides
with $h(x,y)$ on $H_p$.
\end{proposition}
\begin{proof}
Let $\gamma_p$ and $\gamma_{p+1}$ be as in Lemma~\ref{projection},
and let $s(p)$, $\xi^{p,0},\ldots,\xi^{p,s(p)+1}$,
and $D_{p,0},\ldots,D_{p,s(p)}$ be as in Corollary~\ref{corollary}.
For each $k=0,1,\ldots,s(p)$ there exists a unique
$\mu:=\mu(p,k)\in\{1,2,\ldots,l\}$ such that
$D_{p,k}\subseteq A_\mu$ (hence $h=g_\mu$ on $D_{p,k}$,
by \eqref{h}), using Lemma~\ref{union}(5) and the fact
that each $g_i-g_j$ is nonzero throughout $D_{p,k}$.
If $s(p)=0$, we may define the required $d_p$ to be
$g_{\mu(p,0)}\in\mathbb R[\mathbb R^2]$. If $s(p)>0$, then we
shall define $d_p$ as follows. For $k=0,1,\ldots,s(p)-1$, let
$v_{p,k}:=g_{\mu(p,k+1)}-g_{\mu(p,k)}$. We have $v_{p,k}=0$ on
$\overline{D_{p,k}}\cap\overline{D_{p,k+1}}$, since $h$ is
continuous. We extend the notation $c_{a,p,j}$ of
Lemma~\ref{Mahe's lemma} from the case where $a\in\mathcal B$ to the
case where $a=0$: for $j=0,1,\ldots$, we define the function
$c_{0,p,j}$ by $c_{0,p,j}(x,y)=0$ $\forall(x,y)\in\mathbb R_{++}^2$.
If $v_{p,k}\ne0$, then $v_{p,k}\in{\mathcal A}\subset\mathcal B$, so by
\eqref{projection} and \eqref{corollary} there exists
a unique $j(p,k)\in\{1,2,\ldots,s(v_k,p)\}$ such that
the graph of $y=\xi_{v_k,p,j}(x)$ over $(\gamma_p,\gamma_{p+1})$
separates $D_{p,k}$ from $D_{p,k+1}$. We may now take
\begin{equation*}
d_p=g_{\mu(p,0)}+\sum_{k=0}^{s(p)-1}c(v_{p,k},p,j(p,k)),
\end{equation*}
by \eqref{Mahe's lemma} and \eqref{closed}.
\end{proof}
\begin{remark}
The above proposition proves the one-variable analog of
Theorem~\ref{PBCSIG}. For if the given function $h$ does not
involve one of the two variables (say, $x$), then by
Remark~\ref{no x involved} above, none of the functions
that we constructed in the sets $\mathcal A$ and $\mathcal B$
will involve $x$, either, whence we would be able to take
$L=0$ (which would mean that $H_0$ equals all of $\mathbb R_{++}^2$)
in \eqref{projection}--\eqref{corollary}, \eqref{Mahe's lemma},
and \eqref{one variable} above.
\end{remark}
\section{Conclusion of the proof of Theorem~\ref{PBCSIG}}
\label{proof}
Recall, after \eqref{h} we defined ${\mathcal A}=
\{\,g_i-g_j\mid i<j\,\}$, and we defined $\mathcal B$ to be the
set obtained from $\mathcal A$ by closing under the
operations~\eqref{closed under partial} and \eqref{r} with
respect to $y$. We got an $L\ge0$ and certain
$\gamma_p$ on the $x$-axis such that
$0=\gamma_0<\gamma_1<\cdots<\gamma_L<\gamma_{L+1}=\infty$,
and for each $p\in\{0,1,,\ldots,L\}$ we got \eqref{one variable} a
function $d_p(x,y):\mathbb R_{++}^2\to\mathbb R$ that
(1)~is a supremum of infima of finitely many generalized polynomial
functions and
(2)~agrees with $h$ on $H_p$ ($=(\gamma_p,\gamma_{p+1})\times
\mathbb R_{++}$).
Now let $\mathcal C$ be the subset of $\mathbb R[\mathbb R^2]$
obtained from ${\mathcal B}\cup\{\,x-\gamma_p\mid1\le p\le L\,\}$
by closing under the ``$x$-analogs''
of the operations \eqref{closed under partial} and \eqref{r};
i.e., interchanging $x$ and $y$ in \eqref{K}, \eqref{closed
under partial}, and \eqref{r}. Then we immediately obtain,
first, the following $x$-analog of Lemma~\ref{projection}
and its Corollary \ref{corollary}:
\begin{lemma}
There exist $M\in\mathbb N$ and
$\eta_1<\eta_2<\cdots<\eta_M\in\mathbb R_{++}$ such that,
writing $\eta_0=0$ and $\eta_{M+1}=\infty$,
and fixing any $q\in\{0,1,\ldots,M\}$,
the zeros, in the $q$th horizontal half-strip
$I_q:=\mathbb R_{++}\times(\eta_q,\eta_{q+1})$,
of all the $a\in\mathcal C$, are the graphs of continuous,
monotonic,$^{\ref{monotonic}}$
generalized semialgebraic functions $x=\zeta^{q,k}(y)$,
$k=1,2,\ldots,t(q)$ $($for a suitable $t(q)\in\mathbb N)$.
Moreover, for each $y\in(\eta_q,\eta_{q+1})$,
\begin{equation}
0=:\zeta^{q,0}(y)<\zeta^{q,1}(y)<\cdots<\zeta^{q,t(q)}(y)<
\zeta^{q,t(q)+1}(y):=\infty.
\label{x-increasing order}
\end{equation}
Consequently, the sets
\begin{equation*}
E_{q,k}:=\{\,(x,y)\mid
\eta_q<y<\eta_{q+1},\>\zeta^{q,k}(y)<x<\zeta^{q,k+1}(y)\,\},
\end{equation*}
for $k\in\{0,1,\ldots,t(q)\}$, are nonempty, pairwise-disjoint,
generalized semialgebraic cells
$($in particular, they are open and $($pathwise$)$ connected$)$,
and their union is a dense open subset of $I_q$. Moreover,
the $E_{q,k}$ are ``stacked'' one to the right of the other
in the $x$-direction,
so that for any $y\in(\eta_q,\eta_{q+1})$ and for any
$(t(q)+1)$-tuple $x_0,x_1,\ldots,x_{t(q)}\in\mathbb R_{++}$ for which
each $(x_k,y)\in E_{q,k}$, $x_0<x_1<\cdots<x_{t(q)}$.
Finally, for each $k$, there is a $p\in\{0,1,\ldots,L\}$
such that $E_{q,k}\subseteq H_p$ $($since the functions
$x-\gamma_1,\ldots,x-\gamma_L$ belong to $\mathcal C)$.\qed
\label{x-lemma}
\end{lemma}
The second immediate consequence of our choice of $\mathcal C$
is the following $x$-analog of Proposition~\ref{one variable}:
\begin{proposition}
\label{the other variable}
Let $h$, $\mathcal A$, $\mathcal C$, $M$,
$\eta_0,\eta_1,\ldots,\eta_{M+1}$,
$q$, and $I_q$ be as above.
There is a function $e_q:\mathbb
R_{++}^2\to\mathbb R$ that $(1)$~is a supremum of infima of
finitely many generalized polynomial functions $\in\mathbb
R[\mathbb R^2]$ and $(2)$~coincides
with $h(x,y)$ on $I_q$.\qed
\end{proposition}
Let
\begin{equation*}
Q=\{\,(q,k)\mid q\in\{0,1,\ldots,M\},\
k\in\{0,1,\ldots,t(q)\}\,\},
\end{equation*}
where $M$ and $t(q)$ are as in \eqref{x-lemma}.
Then
\begin{equation}
\bigcup_{(q,k)\in Q}E_{q,k}\text{
is a dense open subset of }\mathbb R_{++}^2,
\label{dense}
\end{equation}
by \eqref{x-lemma}.
\begin{lemma}
\label{finer partition}
There is a function $\nu:Q\to\{1,\ldots,l\}$ such that
$\forall(q,k)\in Q$, $E_{q,k}\subseteq A_{\nu(q,k)}^\circ$
$($in particular, $h=g_{\nu(q,k)}$ on $E_{q,k})$.
\end{lemma}
\begin{proof}
This follows from Lemma~\ref{union}(5) and Lemma~\ref{x-lemma}.
\end{proof}
\begin{remark}[on Definition~\ref{pgp}]
\label{aside}
We can now substantiate the statement in Remark~\ref
{announcement of Remark "aside"} above, viz.,
that in the definition of ``piecewise generalized polynomial function''
\eqref{pgp}, it was not necessary to require each $A_i$ to be a
generalized semialgebraic set in the case where $h$ is continuous,
since in that case we may (by \eqref{finer partition} and
\eqref{union}(3)) take each $A_i$ to be the closure of the
union of certain $E_{q,k}$, which is automatically generalized
semialgebraic.
\end{remark}
\begin{notation}
\label{Delta}
For $a,b\in\mathbb R\cup\{\pm\infty\}$ with $a<b$, let
\begin{equation*}
\Delta(a,b)=\{\,(x,y)\in\mathbb R^2\mid xy>0\ \&\ a<x+y<b\,\}.
\end{equation*}
(See Figure~\ref{DeltaFigure}.)
\end{notation}
\begin{figure}[htb]
\setlength\parindent{0pt}
\begin{picture}(327,180)(-143.5,-0
{\thicklines
\put(-90, 80){\vector( 1,0){220}}
\put(130, 80){\vector(-1,0){220}}
\put( 0,-0){\vector(0, 1){185}}
\put( 0,185){\vector(0,-1){185}}
}
\put(3,70)0
\put(-10,178){$y$}
\put(115,72){$x$}
\put(-60,80){\line(1,-1){60}}
\put(10,102){$\Delta(a,b)$}
\put(0,160){\line(1,-1){80}}
\put(-34, 59){$\Delta(a,b)$}
\put( 78, 70){$b$}
\put(-62, 83){$a$}
\put( -7,157){$b$}
\put( 4, 18){$a$}
\qbezier(0,150,)(0,150)(5,155)
\put( 0,140){\line(1,1){10}}
\put( 0,130){\line(1,1){15}}
\put( 0,120){\line(1,1){20}}
\put( 0,110){\line(1,1){25}}
\put( 0,100){\line(1,1){30}}
\put( 0, 90){\line(1,1){10}}
\put(21,111){\line(1,1){14}}
\put( 0, 80){\line(1,1){18}}
\put(30,110){\line(1,1){10}}
\put(10, 80){\line(1,1){18}}
\qbezier(39,109)(39,109)(45,115)
\put(20, 80){\line(1,1){30}}
\put(30, 80){\line(1,1){25}}
\put(40, 80){\line(1,1){20}}
\put(50, 80){\line(1,1){15}}
\put(60, 80){\line(1,1){10}}
\qbezier(70,80)(70,80)(75,85)
\put( 0,80){\line(-1,-1){11}}
\qbezier(-25,55)(-25,55)(-30,50)
\put(-10,80){\line(-1,-1){11}}
\put(-20,80){\line(-1,-1){20}}
\put(-30,80){\line(-1,-1){15}}
\put(-40,80){\line(-1,-1){10}}
\qbezier(-50,80)(-50,80)(-55,75)
\put(-15,55){\line(-1,-1){10}}
\put( 0,60){\line(-1,-1){20}}
\put( 0,50){\line(-1,-1){15}}
\put( 0,40){\line(-1,-1){10}}
\qbezier(0,30)(0,30)(-5,25)
\end{picture}
\caption{The ``double-triangular'' region $\Delta(a,b)$
\eqref{Delta}. In this figure, $a<0<b$.}
\label{DeltaFigure}
\end{figure}
\begin{lemma}
\label{analytic}
Let $f(x,y)$ be a real-valued function that is analytic on a
neighborhood of $(0,0)$ in $\mathbb R^2$.
Write $f_x$ and $f_y$ for $\partial f/\partial x$ and
$\partial f/\partial y$, respectively.
Suppose $f(0,0)=0$, $f_x(0,0)>0$, and $f_y(0,0)>0$.
Then there is an $\epsilon>0$ such that for all
$(x,y)\in\Delta(0,\epsilon)$, $f(x,y)>0$.
\end{lemma}
\begin{proof}
By the Weierstrass Preparation Theorem and the theory of
Puiseux series (see, e.g., \cite[Propositions~3.3 and
4.4, respectively]{Ruiz 1993}),
the germ at $(0,0)$ of the zero-set of $f$ consists
of finitely many curve germs $(\alpha_1(t),\beta_1(t))$,
$(\alpha_2(t),\beta_2(t))$, \dots, where for each $i$:
$\alpha_i$ and $\beta_i$ are analytic for $0\le t<\delta$
(some $\delta>0$); $\alpha_i(0)=\beta_i(0)=0$; and
\begin{align}
\label{t^m}
\begin{split}
\text{either }
\alpha_i(t)&=t^{m_i}\text{ and }\beta'_i(0)\ne0,\\
\text{or }\beta_i(t)&=t^{m_i}\text{ and }\alpha'_i(0)\ne0,
\end{split}
\end{align}
for some $m_i\in\{1,2,\ldots\}$. By the chain rule,
\begin{equation}
\label{chain}
0=\frac d{dt}\,0=\frac d{dt}\,f(\alpha_i(t),\beta_i(t))\bigr|_0
=f_x(0,0)\alpha'_i(0)+f_x(0,0)\beta'_i(0).
\end{equation}
Now we see that we cannot have both
$\alpha'_i(0)\ge0$ and $\beta'_i(0)\ge0$,
for this, together with \eqref{t^m} and the hypothesis of the lemma,
would make the right hand side of \eqref{chain} positive.
Thus there is an $\epsilon>0$ such that for all
$(x,y)\in\Delta(\epsilon)$, $f(x,y)\ne0$.
Since $\Delta(\epsilon)$ is connected and $f$ is
continuous and nonzero there, $f$ has constant sign
(positive or negative) throughout $\Delta(\epsilon)$.
This sign must, in fact, be positive, since
$\frac d{dt}f(t,t)\bigr|_0=f_x(0,0)+f_y(0,0)>0$ and $f(0,0)=0$.
\end{proof}
{\it Conclusion of the proof of Theorem~\ref{PBCSIG}\/}.
As in \cite{Mahe 1984}, the idea now is to construct, for each
two ordered pairs $(q,k)$ and $(r,m)\in Q$, a function
$u_{(q,k),(r,m)}$ that is the supremum of infima of finitely many
generalized polynomial functions, and is such that
\begin{equation}
\label{u}
u_{(q,k),(r,m)}
\begin{cases}
\le g_{\nu(q,k)}&\text{on }E_{q,k}\text{ and}\\
\ge g_{\nu(r,m)}&\text{on }E_{r,m}.
\end{cases}
\end{equation}
Then we shall be done, since the function
\begin{equation*}
u_{(r,m)}:=
\inf\bigl(\{g_{\nu(r,m)}\}\cup
\{\,u_{(q,k),(r,m)}\mid(q,k)\in Q\,\}\bigr)
\end{equation*}
will satisfy
\begin{alignat*}2
u_{(r,m)}& = g_{\nu(r,m)}&&\text{ on }E_{r,m}\text{, and,}\\
\text{ for each }(q,k)\in Q,\quad
u_{(r,m)}&\le g_{\nu(q,k)}&&\text{ on }E_{q,k};
\end{alignat*}
then $h=\sup_{(r,m)\in Q}u_{(r,m)}$ throughout
$\bigcup_{(q,k)\in Q}E_{q,k}$, and hence (by \eqref{dense}
and the continuity of $h$) throughout $\mathbb R_{++}^2$,
as required.
So suppose $(q,k)$ and $(r,m)\in Q$,
and let us prepare to construct a $u_{(q,k),(r,m)}$
satisfying \eqref{u}. If $E_{\nu(q,k)}$ and
$E_{\nu(r,m)}$ are both subsets of the same horizontal
half-strip $I_q$ \eqref{x-lemma},\footnote{This will occur
if and only if $q=r$.} or of the same vertical half-strip $H_p$
(for some $p\in\{1,2,\ldots,L\}$, using the last sentence of
\eqref{x-lemma}), then we may take $u_{(q,k),(r,m)}$
to be either $e_q$ or $d_p$, respectively,
by \eqref{the other variable} or \eqref{one variable}.
The case that makes the proof for two variables harder
than the proof for one variable is the case when $E_{\nu(q,k)}$
and $E_{\nu(r,m)}$ do {\it not\/} lie in a common
half-strip (either horizontal or vertical). We may
assume, without loss of generality, that $E_{\nu(q,k)}$
is below and to the left of $E_{\nu(r,m)}$
(i.e., that points in $E_{\nu(q,k)}$ have $x$- and
$y$-coordinates less than the $x$- and $y$-coordinates
of points in $E_{\nu(r,m)}$, respectively);
the other three possibilities could be handled similarly.
$E_{\nu(q,k)}$ lies in the horizontal half-strip
$I_q:=\mathbb R_{++}\times(\eta_q,\eta_{q+1})$,
and in a unique vertical half-strip $H_p:=
(\xi_p,\xi_{p+1})\times\mathbb R_{++}$,
for some $p$. $E_{\nu(r,m)}$ lies in exactly
one of the horizontal half-strips $I_{q+1},I_{q+2},\dots$,
and in exactly one of the vertical half-strips $H_{p+1},
H_{p+2},\,$\dots. (See Figure~\ref{E},
where, for simplicity, $E_{\nu(r,m)}$ is shown lying in
$I_{q+1}$ and $H_{p+1}$.)
\begin{figure}[htb]
\setlength\parindent{0pt}
\begin{picture}(327,190)(-143.5,-10
{\thicklines
\put(-110, 0){\vector(1,0){280}}
\put(-90,-15){\vector(0,1){200}}
}
\put(-98,-10)0
\put(-103,175){$y$}
\put(155,-12){$x$}
\put(-30, 0){\line(0,1){10}}
\put(-30, 20){\line(0,1){10}}
\put(-30, 40){\line(0,1){10}}
\put(-30, 60){\line(0,1){10}}
\put(-30, 80){\line(0,1){10}}
\put(-30,100){\line(0,1){10}}
\put(-30,120){\line(0,1){10}}
\put(-30,140){\line(0,1){10}}
\put(-30,160){\line(0,1){10}}
\put( 40, 0){\line(0,1){10}}
\put( 40, 20){\line(0,1){10}}
\put( 40, 40){\line(0,1){10}}
\put( 40, 60){\line(0,1){10}}
\put( 40, 80){\line(0,1){10}}
\put( 40,100){\line(0,1){10}}
\put( 40,120){\line(0,1){10}}
\put( 40,140){\line(0,1){10}}
\put( 40,160){\line(0,1){10}}
\put(-90,60){\line(1,0){10}}
\put(-70,60){\line(1,0){10}}
\put(-50,60){\line(1,0){10}}
\put(-30,60){\line(1,0){10}}
\put(-10,60){\line(1,0){10}}
\put( 10,60){\line(1,0){10}}
\put( 30,60){\line(1,0){10}}
\put( 50,60){\line(1,0){10}}
\put( 70,60){\line(1,0){10}}
\put( 90,60){\line(1,0){10}}
\put(110,60){\line(1,0){10}}
\put(130,60){\line(1,0){10}}
\put(150,60){\line(1,0){10}}
\put(-90,100){\line(1,0){10}}
\put(-70,100){\line(1,0){10}}
\put(-50,100){\line(1,0){10}}
\put(-30,100){\line(1,0){10}}
\put(-10,100){\line(1,0){10}}
\put( 10,100){\line(1,0){10}}
\put( 30,100){\line(1,0){10}}
\put( 50,100){\line(1,0){10}}
\put( 70,100){\line(1,0){10}}
\put( 90,100){\line(1,0){10}}
\put(110,100){\line(1,0){10}}
\put(130,100){\line(1,0){10}}
\put(150,100){\line(1,0){10}}
\put(40,100){\circle*3}
\qbezier(40,140)(70,110)(110,100)
\put(80,100){\circle*3}
\put(80,-3){\line(0,1)6}
\put(67,-10){$\xi_{p+1}+b^*$}
\put(40,140){\circle*3}
\put(-93,140){\line(1,0)6}
\put(-131,137){$\eta_{q+1}+b^*$}
\put( -8,49){$x=\zeta^{q,k}(y)$}
\put(-10,51){\line(-1,0){10}}
\put(-20,51){\vector(1,2){9.7}}
\qbezier(-28,100)(-25,87)(-20,80)
\qbezier(-20, 80)(-10,66)( 25,60)
\put(53,71){$\Delta(a^*,b^*)+(\xi_{p+1},\eta_{q+1})$}
\put(51,73){\line(-1,0){10}}
\put(41,73){\vector(-1,2){8.5}}
\put(41,73){\vector(1,4){8.5}}
\put(-88,112){$x=\zeta^{q,k+1}(y)$}
\put(-29,114){\line(1,0){11}}
\put(-18,114){\vector(1,-3){7.5}}
\qbezier(-24,100)(-10,87)(10,85)
\qbezier( 10, 85)( 30,83)(40,60)
\put(40,140){\line(1,-1){40}}
\qbezier(40,130)(40,130)(45,135)
\put(40,120){\line(1,1){10}}
\put(40,110){\line(1,1){15}}
\put(40,100){\line(1,1){20}}
\put(50,100){\line(1,1){15}}
\put(60,100){\line(1,1){10}}
\qbezier(70,100)(70,100)(75,105)
\put(6.1,100){\line(1,-1){33.9}}
\qbezier(20,100)(20,100)(13.05,93.05)
\put(30,100){\line(-1,-1){11.95}}
\put(40,100){\line(-1,-1){16.95}}
\qbezier(40,90)(40,90)(37,87)
\qbezier(33,83)(33,83)(28.05,78.05)
\qbezier(37,77)(37,77)(33.05,73.05)
\put(6.1,100){\circle*3}
\put(6.1,-3){\line(0,1)6}
\put(-11,-10){$\xi_{p+1}-a^*$}
\put(40,66.1){\circle*3}
\put(-93,66.1){\line(1,0)6}
\put(-132,64){$\eta_{q+1}-a^*$}
\put( -32,-10){$\xi_p$}
\put( 38,-10){$\xi_{p+1}$}
\put(-100, 56){$\eta_q$}
\put(-109, 97){$\eta_{q+1}$}
\put(5,15){$H_p$}
\put(-76,77){$I_q$}
\put(78,130){$E_{\nu(r,m)}$}
\put(-3,73){$E_{\nu(q,k)}$}
\end{picture}
\caption{The case where $E_{\nu(q,k)}$ and $E_{\nu(r,m)}$
do not lie in a common half-strip.
(In this illustration, $E_{\nu(r,m)}$ lies in $I_{q+1}$
and $H_{p+1}$).)}
\label{E}
\end{figure}
For any $a,b\in\mathbb R\cup\{\pm\infty\}$ with $a<b$,
write
\begin{equation*}
\Delta(a,b)+(\xi_{p+1},\eta_{q+1})=
\{\,(x+\xi_{p+1},\,y+\eta_{q+1})\mid(x,y)\in\Delta(a,b)\,\}.
\end{equation*}
Now let
\begin{align}
\label{a^*,b^*}
\begin{split}
a^*&=\min\{\,s\in\mathbb R\mid
(\Delta(s,0)+(\xi_{p+1},\eta_{q+1}))\cap E_{\nu(q,k)}
=\emptyset\,\}\text{ and}\\
b^*&=\max\{\,t\in\mathbb R\mid
(\Delta(0,t)+(\xi_{p+1},\eta_{q+1}))\cap E_{\nu(r,m)}
=\emptyset\,\}.
\end{split}
\end{align}
(Thus, $a^*\le0\le b^*$, by the assumptions on $E_{\nu(q,k)}$
and $E_{\nu(r,m)}$ made in the previous paragraph.)
To simplify notation, let
\begin{equation}
\label{g_nu(r,m)-g_nu(q,k)}
g(x,y)=g_{\nu(r,m)}(x,y)-g_{\nu(q,k)}(x,y).
\end{equation}
Pick any $e\in\mathbb N$ greater than every $x$- and
$y$-exponent ($\in\mathbb R$) occurring in (the unique
representation as in \eqref{signomial} of) $g(x,y)$.
There is a $T\ge a^*$ such that for all $(x,y)
\in\mathbb R_{++}^2$ with $x+y-\xi_{p+1}-\eta_{q+1}\ge T$,\footnote
{In particular, for all
$(x,y)\in\Delta(T,\infty)+(\xi_{p+1},\eta_{q+1})$.}
\begin{equation}
\label{(x+y-xi_{p+1}-eta_{q+1}-a^*)^e ge g(x,y)}
(x+y-\xi_{p+1}-\eta_{q+1}-a^*)^e\ge g(x,y).\footnote
{If we had allowed $e$ to be an arbitrary real number
(as opposed to an element of $e\in\mathbb N$),
then $(x+y-\xi_{p+1}-\eta_{q+1}-a^*)^e$ would not necessarily
be a signomial function (see \cite[Example~4.7]{Delzell 2008}).
Since, in fact, $e\in\mathbb N$, $(x+y-\xi_{p+1}-\eta_{q+1}-a^*)^e$
is a signomial function (it is even an ordinary polynomial).
We shall need this below.}
\end{equation}
We may assume that $T>b^*$ (in particular, $T>0$).
{\it Case 1\/}: $b^*-a^*>0$.
In this case, there is a $C\in\mathbb R$ such that
for all $(x,y)\in\Delta(b^*,T)+(\xi_{p+1},\eta_{q+1})$,
\begin{equation}
\label{b^*,T}
C\cdot(x+y-\xi_{p+1}-\eta_{q+1}-a^*)^e\ge g(x,y).\footnote
{Specifically, we may take
$C=(\max g(x,y))/\min((x+y-\xi_{p+1}-\eta_{q+1}-a^*)^e)$,
where the max and min are taken as $(x,y)$ ranges over the compact
set $\overline{\Delta(b^*,T)+(\xi_{p+1},\eta_{q+1})}$.
(Here we need $\min(x+y-\xi_{p+1}-\eta_{q+1}-a^*)>0$,
which follows from our assumption (here in case~1) that $b^*-a^*>0$.)}
\end{equation}
We may assume that $C\ge1$. Then we may take
\begin{equation*}
u_{(q,k),(r,m)}=
g_{q,k}(x,y)+C\cdot((x+y-\xi_{p+1}-\eta_{q+1}-a^*)^+)^e,
\end{equation*}
which satisfies \eqref{u} (using \eqref{a^*,b^*},
\eqref{(x+y-xi_{p+1}-eta_{q+1}-a^*)^e ge g(x,y)},
\eqref{b^*,T}, and \eqref{g_nu(r,m)-g_nu(q,k)}),
and which is a supremum of infima of finitely many
generalized polynomial functions (using Proposition~\ref
{closed}).
{\it Case 2\/}: $b^*-a^*=0$ (whence $a^*=0=b^*$).
In this case, let
\begin{equation*}
f(x,y)=g(x+\xi_{p+1},\,y+\eta_{q+1}).\footnote
{In general, $f$ is not a signomial function
(again, see \cite[Example~4.7]{Delzell 2008}),
but it is, at least, real analytic (for $x>-\xi_{p+1}$
and $y>-\eta_{q+1}$), and this is all we shall need.}
\end{equation*}
Pick any $D\in\mathbb R_{++}$ greater than
$\max\{f_x(0,0),\>f_y(0,0)\}$.
By Lemma~\ref{analytic}, there is an $\epsilon>0$ such that
$D\cdot(x+y)>f(x,y)$ for all $(x,y)\in\Delta(0,\epsilon)$;
equivalently,
\begin{equation}
\label{D}
D\cdot(x+y-\xi_{p+1}-\eta_{q+1})>g(x,y)
\end{equation}
for all $(x,y)\in\Delta(0,\epsilon)+(\xi_{p+1},\eta_{q+1})$.
We may assume that $\epsilon\le T$.
There is a $C\in\mathbb R$ such that
for all $(x,y)\in\Delta(\epsilon,T)+(\xi_{p+1},\eta_{q+1})$,
\begin{equation}
\label{epsilon,T}
C\cdot(x+y-\xi_{p+1}-\eta_{q+1})^e\ge g(x,y).\footnote
{Specifically, we may take
$C=(\max g(x,y))/\min((x+y-\xi_{p+1}-\eta_{q+1})^e)$,
where the max and min are taken as $(x,y)$ ranges over the compact
set $\overline{\Delta(\epsilon,T)+(\xi_{p+1},\eta_{q+1})}$.
(Here we need $\min(x+y-\xi_{p+1}-\eta_{q+1})>0$,
which follows from $\epsilon>0$.)}
\end{equation}
We may assume that $C\ge1$.
Then we may take
\begin{equation*}
u_{(q,k),(r,m)}=g_{q,k}(x,y)+\sup\{
D(x+y-\xi_{p+1}-\eta_{q+1})^+,
C((x+y-\xi_{p+1}-\eta_{q+1})^+)^e\},
\end{equation*}
which satisfies \eqref{u} (using \eqref{D},
\eqref{epsilon,T},
\eqref{(x+y-xi_{p+1}-eta_{q+1}-a^*)^e ge g(x,y)} (with $a^*=0$),
and \eqref{g_nu(r,m)-g_nu(q,k)}),
and which is a supremum of infima of finitely many
generalized polynomial functions (using Proposition~\ref
{closed}).
\qed
|
\section{Introduction}
In 1995 Kashaev introduced a complex valued link invariant for an integer $N\ge 2$ by using the quantum dilogarithm \cite{Kashaev:MODPLA95} and then he observed that his invariant grows exponentially with growth rate proportional to the volume of the knot complement for several hyperbolic knots \cite{Kashaev:LETMP97}.
He also conjectured that this also holds for any hyperbolic knot, where a knot in the three-sphere is called hyperbolic if its complement possesses a complete hyperbolic structure with finite volume.
\par
In 2001 J.~Murakami and the author proved that Kashaev's invariant turns out to be a special case of the colored Jones polynomial.
More precisely Kashaev's invariant is equal to $J_N\bigl(K;\exp(2\pi\sqrt{-1}/N)\bigr)$, where $J_N(K;q)$ is the $N$-dimensional colored Jones polynomial associated with the $N$-dimensional irreducible representation of the Lie algebra $\mathfrak{sl}(2;\mathbb{C})$ and $K$ is a knot (\S~\ref{sec:YB}).
We also generalized Kashaev's conjecture to any knot (Volume Conjecture) by using the Gromov norm, which can be regarded as a natural generalization of the hyperbolic volume (\S~\ref{sec:Volume_Conjecture}).
If it is true it would give interesting relations between quantum topology and hyperbolic geometry.
So far the conjecture is proved only for several knots and some links but we have supporting evidence which is described in \S~\ref{sec:VC_proof}.
\par
In the Volume Conjecture we study the colored Jones polynomial at the $N$-th root of unity $\exp(2\pi\sqrt{-1}/N)$.
What happens if we replace $2\pi\sqrt{-1}$ with another complex number?
Recalling that the complete hyperbolic structure of a hyperbolic knot complement can be deformed by using a complex parameter \cite{Thurston:GT3M}, we expect that we can also relate the colored Jones polynomial evaluated at $\exp\bigl((2\pi\sqrt{-1}+u)/N\bigr)$ to the volume of the deformed hyperbolic structure.
At least for the figure-eight knot this is true if $u$ is small \cite{Murakami/Yokota:JREIA2007}.
It is also true (for the figure-eight knot) that we can also get the Chern--Simons invariant, which can be regarded as the imaginary part of the volume, from the colored Jones polynomial (\S~\ref{sec:generalization}).
\par
In general we conjecture that this is also true, that is, for any knot the asymptotic behavior of the colored Jones polynomial would determine the volume of a three-manifold obtained as the deformation associated with the parameter $u$.
\par
The aim of this article is to give an elementary introduction to these conjectures including many examples so that nonexperts can easily understand.
I hope you will join us.
\begin{ack}
The author would like to thank the organizers of the workshop and conference ``Interactions Between Hyperbolic Geometry, Quantum Topology and Number Theory'' held at Columbia University, New York in June 2009.
\par
Thanks are also due to an immigration officer at J.~F.~Kennedy Airport, who knows me by papers, for interesting and exciting discussion about quantum topology.
\end{ack}
\section{Link invariant from a Yang--Baxter operator}
\label{sec:YB}
In this section I describe how we can define a link invariant by using a Yang--Baxter operator.
\subsection{Braid presentation of a link}
An $n$-braid is a collection of $n$ strands that go downwards monotonically from a set of fixed $n$ points to another set of fixed $n$ points as shown in Figure~\ref{fig:braid}.
\begin{figure}[h]
\includegraphics[scale=0.3]{braid_small.eps}
\caption{braid}
\label{fig:braid}
\end{figure}
The set of all $n$-braids makes a group $B_n$ with product of braids $\beta_1$ and $\beta_2$ given by putting $\beta_2$ below $\beta_1$.
It is well known (see for example \cite{Birman:1974}) that $B_n$ is generated by $\sigma_1, \sigma_2,\dots,\sigma_{n-1}$ (Figure~\ref{fig:s_i}) with relations $\sigma_i\sigma_j=\sigma_j\sigma_i$ ($|i-j|>1$) and $\sigma_k\sigma_{k+1}\sigma_k=\sigma_{k+1}\sigma_k\sigma_{k+1}$.
See Figure~\ref{fig:braid_relation} for the latter relation, which is called the braid relation.
\begin{figure}[h]
\includegraphics[scale=0.3]{s_i_small.eps}
\caption{$i$th generator of the $n$-braid group $B_n$}
\label{fig:s_i}
\end{figure}
\begin{figure}[h]
\includegraphics[scale=0.3]{braid_relation1_small.eps}
\quad\raisebox{12mm}{$=$}\quad
\includegraphics[scale=0.3]{braid_relation2_small.eps}
\caption{braid relation}
\label{fig:braid_relation}
\end{figure}
So we have the following group presentation of $B_n$.
\begin{equation}\label{eq:braid_presentation}
B_n
=
\langle
\sigma_1,\sigma_2,\dots,\sigma_{n-1}
\mid
\sigma_i\sigma_j=\sigma_j\sigma_i\,\text{($|i-j|>1$)},
\sigma_k\sigma_{k+1}\sigma_k=\sigma_{k+1}\sigma_k\sigma_{k+1}
\rangle
\end{equation}
\par
It is known that any knot or link can be presented as the closure of a braid.
\begin{theorem}[Alexander \cite{Alexander:PRONA1923}]\label{thm:Alexander}
Any knot or link can be presented as the closure of a braid.
\end{theorem}
\begin{figure}[h]
\includegraphics[scale=0.3]{braid_small.eps}
\quad\raisebox{20mm}{$\xrightarrow{\text{closure}}$}\quad
\includegraphics[scale=0.3]{closure_small.eps}
\quad\raisebox{20mm}{$=$}\quad
\raisebox{5mm}{\includegraphics[scale=0.3]{fig8_thin_small.eps}}
\caption{the figure-eight knot is presented as the closure of a braid}
\label{fig:braid_presentation}
\end{figure}
Here the closure of an $n$-braid is obtained by connecting the $n$ points on the top with the $n$ points on the bottom without entanglement as shown in the middle picture of Figure~\ref{fig:braid_presentation}.
\par
There are many braids that present a knot or link but if two braids present the same knot or link, they are related by a finite sequence of conjugations and (de-)stabilizations.
In fact we have the following theorem.
\begin{theorem}[Markov \cite{Markov:1936}]\label{thm:Markov}
If two braids $\beta$ and $\beta'$ give equivalent links, then they are related by a finite sequence of conjugations, stabilizations, and destabilizations.
Here a conjugation replaces $\alpha\beta$ with $\beta\alpha$, or equivalently $\beta$ with $\alpha^{-1}\beta\alpha$ {\rm(}Figure~$\ref{fig:conjugate}${\rm)}, a stabilization replaces $\beta\in B_n$ with $\beta\sigma_n^{\pm1}\in B_{n+1}$ {\rm(}Figure~$\ref{fig:stabilization}${\rm)}, a destabilization replaces $\beta\sigma_n^{\pm1}\in B_{n+1}$ with $\beta\in B_n$.
\begin{figure}[h]
\includegraphics[scale=0.2]{conjugate_ab_small.eps}
\raisebox{13mm}{\quad$\xrightarrow{\text{\rm conjugation}}$\quad}
\includegraphics[scale=0.2]{conjugate_ba_small.eps}
\caption{$\alpha\beta$ is conjugate to $\beta$.}
\label{fig:conjugate}
\end{figure}
\begin{figure}[h]
\includegraphics[scale=0.2]{stabilize_1_small.eps}
\raisebox{13mm}{
$\begin{array}{c}
\xrightarrow{\text{\rm stabilization}}\\[2mm]
\xleftarrow{\text{\rm destabilization}}
\end{array}$
}
\includegraphics[scale=0.2]{stabilize_2_small.eps}
\caption{$\beta$ ($\beta\sigma_{n}^{\pm}$, respectively) is stabilized (destabilized, respectively) to $\beta\sigma_{n}^{\pm}$ ($\beta$, respectively).}
\label{fig:stabilization}
\end{figure}
\end{theorem}
\subsection{Yang--Baxter operator}
Alexander's theorem (Theorem~\ref{thm:Alexander}) and Markov's theorem (Theorem~\ref{thm:Markov}) can be used to define link invariants.
I will follow Turaev \cite{Turaev:INVEM88} to introduce a link invariant derived from a Yang-Baxter operator.
\par
Let $V$ be an $N$-dimensional vector space over $\mathbb{C}$, $R$ an isomorphism from $V\otimes V$ to itself, $m$ an isomorphism from $V$ to itself, and $a$ and $b$ non-zero complex numbers.
\begin{definition}\label{def:YB}
A quadruple $(R,\mu,a,b)$ is called an enhanced Yang--Baxter operator if it satisfies the following:
\begin{enumerate}
\item $(R\otimes\operatorname{Id}_V)(\operatorname{Id}_V\otimes R)(R\otimes\operatorname{Id}_V)
=(\operatorname{Id}_V\otimes R)(R\otimes\operatorname{Id}_V)(\operatorname{Id}_V\otimes R)$,
\item $R(\mu\otimes\mu)=(\mu\otimes\mu)R$,
\item $\operatorname{Tr}_2\bigl(R^{\pm1}(\operatorname{Id}_V\otimes\mu)\bigr)=a^{\pm1}b\operatorname{Id}_V$.
\end{enumerate}
Here $\operatorname{Tr}_k\colon\operatorname{End}(V^{\otimes k})\to\operatorname{End}(V^{\otimes(k-1)})$ is defined by
\begin{equation*}
\operatorname{Tr}_k(f)(e_{i_1}\otimes e_{i_2}\dots\otimes e_{i_{k-1}})
:=
\sum_{j_1,j_2,\dots,j_{k-1},j=0}^{N-1}
f_{i_1,i_2,\dots,i_{k-1},j}^{j_1,j_2,\dots,j_{k-1},j}
(e_{j_1}\otimes e_{j_2}\otimes\dots\otimes e_{j_{k-1}}\otimes e_{j}),
\end{equation*}
where $f\in\operatorname{End}(V^{\otimes k})$ is given by
\begin{equation*}
f(e_{i_1}\otimes e_{i_2}\otimes\dots\otimes e_{i_k})
=
\sum_{j_1,j_2,\dots,j_k=0}^{N-1}
f_{i_1,i_2,\dots,i_k}^{j_1,j_2,\dots,j_k}
(e_{j_1}\otimes e_{j_2}\otimes\dots\otimes e_{j_k})
\end{equation*}
and $\{e_0,e_1,\dots,e_{N-1}\}$ is a basis of $V$.
\end{definition}
\begin{remark}
The isomorphism $R$ is often called an $R$-matrix, and the equation {\rm(1)} is known as the Yang--Baxter equation.
\end{remark}
\par
Given an $n$-braid $\beta$, we can construct a homomorphism $\Phi(\beta)\colon V^{\otimes n}\to V^{\otimes n}$ by replacing a generator $\sigma_i$ with $\operatorname{Id}_V^{\otimes(i-1)}\otimes R\otimes\operatorname{Id}_V^{\otimes(n-i-1)}$, and its inverse $\sigma_i^{-1}$ with $\operatorname{Id}_V^{\otimes(i-1)}\otimes R^{-1}\otimes\operatorname{Id}_V^{\otimes(n-i-1)}$ (Figure~\ref{fig:braid_homomorphism}).
\begin{figure}[h]
\raisebox{-4mm}{\includegraphics[scale=0.3]{crossing_pos_small.eps}}
$\Rightarrow$
\raisebox{-8mm}{\includegraphics[scale=0.3]{R_pos_small.eps}},
\quad\quad
\raisebox{-4mm}{\includegraphics[scale=0.3]{crossing_neg_small.eps}}
$\Rightarrow$
\raisebox{-8mm}{\includegraphics[scale=0.3]{R_neg_small.eps}}
\caption{Replace a generator with the $R$-matrix.}
\label{fig:braid_homomorphism}
\end{figure}
\begin{example}
For the braid $\sigma_1\sigma_2^{-1}\sigma_1\sigma_2^{-1}$ the corresponding homomorphism is give as follows (Figure~\ref{fig:fig_8_homomorphism}):
\begin{equation*}
\Phi(\sigma_1\sigma_2^{-1}\sigma_1\sigma_2^{-1})
=
(R\otimes\operatorname{Id}_V)(\operatorname{Id}_V\otimes R^{-1})(R\otimes\operatorname{Id}_V)(\operatorname{Id}_V\otimes R^{-1}).
\end{equation*}
\begin{figure}[h]
\includegraphics[scale=0.3]{braid_small.eps}
\quad\raisebox{20mm}{$\xrightarrow{\Phi}$}\quad
\includegraphics[scale=0.3]{braid_R_small.eps}
\caption{A braid and the corresponding homomorhism}
\label{fig:fig_8_homomorphism}
\end{figure}
\end{example}
\subsection{Invariant}
Let $(R,\mu,a,b)$ be an enhanced Yang--Baxter operator on an $N$-dimensional vector space $V$.
\begin{definition}
For an $n$-braid $\beta$, we define $T_{(R,\mu,a,b)}(\beta)\in\mathbb{C}$ by the following formula.
\begin{equation*}
T_{(R,\mu,a,b)}(\beta)
:=
a^{-w(\beta)}
b^{-n}
\operatorname{Tr}_1
\Bigl(
\operatorname{Tr}_2
\bigl(
\cdots
\left(
\operatorname{Tr}_n\left(\Phi(\beta)\mu^{\otimes n}\right)
\right)
\cdots
\bigr)
\Bigr),
\end{equation*}
where $w(\beta)$ is the sum of the exponents in $\beta$.
Note that $\operatorname{Tr}_1\colon\operatorname{End}(V)\to\mathbb{C}$ is the usual trace.
\end{definition}
\begin{example}
For the braid $\sigma_1\sigma_2^{-1}\sigma_1\sigma_2^{-1}$, we have
\begin{multline*}
T_{(R,\mu,a,b)}(\sigma_1\sigma_2^{-1}\sigma_1\sigma_2^{-1})
\\
=
b^{-2}
\operatorname{Tr}_1\bigl(\operatorname{Tr}_2(\operatorname{Tr}_3(
(R\otimes\operatorname{Id}_V)(\operatorname{Id}_V\otimes R^{-1})(R\otimes\operatorname{Id}_V)(\operatorname{Id}_V\otimes R^{-1})
(\mu\otimes\mu\otimes\mu)))\bigr)
\end{multline*}
since $w(\sigma_1\sigma_2^{-1}\sigma_1\sigma_2^{-1})=+1-1+1-1=0$ (Figure~\ref{fig:fig_8_invariant}).
\begin{figure}[h]
\raisebox{-18mm}{\includegraphics[scale=0.3]{closure_small.eps}}
\quad$\Rightarrow$\quad$a^{-w(\beta)}b^{-n}\times$
\raisebox{-21mm}{\includegraphics[scale=0.3]{closure_R_small.eps}}
\caption{A braid and its invariant}
\label{fig:fig_8_invariant}
\end{figure}
\end{example}
\par
We can show that $T_{(R,\mu,a,b)}$ gives a link invariant.
\begin{theorem}[Turaev \cite{Turaev:INVEM88}]
If $\beta$ and $\beta'$ present the same link, then $T_{R,\mu,a,b}(\beta)=T_{R,\mu,a,b}(\beta')$.
\end{theorem}
\begin{proof}[Sketch of a proof]
By Markov's theorem (Theorem~\ref{thm:Markov}) it is sufficient to prove that $T_{R,\mu,a,b}$ is invariant under a braid relation, a conjugation and a stabilization.
\par
The invariance under a braid relation $\sigma_i\sigma_{i+1}\sigma_i=\sigma_{i+1}\sigma_i\sigma_{i+1}$ follows from Figure~\ref{fig:braid_relation_YB}.
Note that the left hand side depicts a braid relation \eqref{eq:braid_presentation} and the right hand side depicts the corresponding Yang--Baxter equation (Definition~\ref{def:YB} (1)).
\begin{figure}[h]
\raisebox{4mm}{$\underset{\text{\raisebox{-5mm}{braid relation}}}
{\includegraphics[scale=0.3]{braid_relation1_small.eps}
\quad\raisebox{12mm}{=}\quad
\includegraphics[scale=0.3]{braid_relation2_small.eps}}$}
\raisebox{16mm}{\quad$\xrightarrow{\Phi}$\quad}
$\underset{\text{\raisebox{-1mm}{Yang--Baxter equation}}}
{\includegraphics[scale=0.3]{YB1_small.eps}
\quad\raisebox{16mm}{=}\quad
\includegraphics[scale=0.3]{YB2_small.eps}}$
\caption{braid relation corresponds to the Yang--Baxter equation}
\label{fig:braid_relation_YB}
\end{figure}
\par
The invariance under a conjugation follows from Figure~\ref{fig:invariance_conjugation}.
The first equality follows since $\operatorname{Tr}_k$ is invariant under a conjugation.
The second equality follows since $(\mu\otimes\mu)R=R(\mu\otimes\mu)$ (Definition~\ref{def:YB} (2)).
Note that the equality $(\mu\otimes\mu)R=R(\mu\otimes\mu)$ means that a pair $\mu\otimes\mu$ can pass through a crossing.
\begin{figure}[h]
\includegraphics[scale=0.3]{conjugate_R_abm_small.eps}
\quad\raisebox{24mm}{$=$}\quad
\includegraphics[scale=0.3]{conjugate_R_bma_small.eps}
\quad\raisebox{24mm}{$=$}\quad
\includegraphics[scale=0.3]{conjugate_R_bam_small.eps}
\caption{invariance under a conjugation}
\label{fig:invariance_conjugation}
\end{figure}
\par
To prove the invariance under a stabilization, we first note that if a homomorphism $f\colon V^{\otimes n}\to V^{\otimes n}$ given by
\begin{equation*}
f(e_{i_1}\otimes\dots\otimes e_{i_n})
=
\sum_{j_1,\dots,j_n=0}^{N-1}
f_{i_1,\dots,i_n}^{j_1,\dots,j_n}(e_{j_1}\otimes\dots\otimes e_{j_n}),
\end{equation*}
then its $n$-fold trace is given by
\begin{equation*}
\operatorname{Tr}_1(\cdots(\operatorname{Tr}_{n}(f))\cdots)
=
\sum_{j_1,\dots,j_n}f_{j_1,\dots,j_n}^{j_1,\dots,j_n}.
\end{equation*}
Therefore if $g$ is a homomorphism $g\colon V\otimes V\to V\otimes V$ given by $g_{l_1,l_2}^{k_1,k_2}$, then we have
\begin{multline*}
\operatorname{Tr}_1
\left(\cdots
\left(\operatorname{Tr}_n
\left(\operatorname{Tr}_{n+1}
\bigl(
(f\otimes\operatorname{Id}_V)(\operatorname{Id}_V^{\otimes(n-1)}\otimes g)
\bigr)
\right)
\right)
\right)
\\
=
\sum_{j_1,\dots,j_n,k_n,k_{n+1}}
f_{j_1,\dots,j_{n-1},j_n}^{j_1,\dots,j_{n-1},k_n}
g_{k_n,k_{n+1}}^{j_n,k_{n+1}},
\end{multline*}
which coincides with the $n$-fold trace of the homomorphism $f\bigl(\operatorname{Id}_V^{\otimes(n-1)}\otimes\operatorname{Tr}_2(g)\bigr)\colon V^{\otimes n}\to V^{\otimes n}$.
\par
Therefore for $\beta\in B_n$ we have
\begin{equation*}
\begin{split}
&\operatorname{Tr}_1
\Biggl(
\operatorname{Tr}_2
\biggl(
\cdots
\Bigl(
\operatorname{Tr}_n
\left(
\operatorname{Tr}_{n+1}\left(\Phi(\beta\sigma_n^{\pm1})\mu^{\otimes(n+1)}\right)
\right)
\Bigr)
\cdots
\biggr)
\Biggr)
\\
=&
\operatorname{Tr}_1
\Biggl(
\operatorname{Tr}_2
\biggl(
\cdots
\Bigl(
\operatorname{Tr}_n
\left(
\operatorname{Tr}_{n+1}
\left(
(\mu^{\otimes n}\Phi(\beta)\otimes\operatorname{Id}_V)
(\operatorname{Id}_V^{\otimes(n-1)}\otimes R^{\pm1}(\operatorname{Id}_V\otimes\mu))
\right)
\right)
\Bigr)
\cdots
\biggr)
\Biggr)
\\
=&
\operatorname{Tr}_1
\Biggl(
\operatorname{Tr}_2
\biggl(
\cdots
\Bigl(
\operatorname{Tr}_n
\left(
(\mu^{\otimes n}\Phi(\beta))
(\operatorname{Id}_V^{\otimes(n-1)}\otimes\operatorname{Tr}_2(R^{\pm1}(\operatorname{Id}_V\otimes\mu)))
\right)
\Bigr)
\cdots
\biggr)
\Biggr)
\\
=&
a^{\pm1}b
\operatorname{Tr}_1
\Biggl(
\operatorname{Tr}_2
\biggl(
\cdots
\Bigl(
\operatorname{Tr}_n
\left(
(\mu^{\otimes n}\Phi(\beta))
\right)
\Bigr)
\cdots
\biggr)
\Biggr)
\\
=&
a^{\pm1}b
\operatorname{Tr}_1
\Biggl(
\operatorname{Tr}_2
\biggl(
\cdots
\Bigl(
\operatorname{Tr}_n
\left(
(\Phi(\beta)\mu^{\otimes n})
\right)
\Bigr)
\cdots
\biggr)
\Biggr),
\end{split}
\end{equation*}
since $\operatorname{Tr}_2\bigl(R^{\pm1}(\operatorname{Id}_V\otimes\mu\bigr)=a^{\pm1}b\operatorname{Id}_V$ (Definition~\ref{def:YB} (3)) as depicted in Figure~\ref{fig:invariance_stabilization}.
\begin{figure}[h]
\includegraphics[scale=0.3]{stabilize_R2_small.eps}
\quad\raisebox{24mm}{$=\quad a^{\pm1}b\,$}
\includegraphics[scale=0.3]{stabilize_R3_small.eps}
\caption{invariance under a stabilization}
\label{fig:invariance_stabilization}
\end{figure}
Since $w(\beta\sigma_n^{\pm1})=w(\beta)\pm1$, the invariance under a stabilization follows.
\end{proof}
Therefore we can define a link invariant $T_{R,\mu,a,b}(L)$ to be $T_{R,\mu,a,b}(\beta)$ if $L$ is the closure of $\beta$.
\subsection{Quantum $(\mathfrak{g},V)$ invariant}
One of the important ways to construct an enhanced Yang--Baxter operator is to use a quantum group, which is a deformation of a Lie algebra.
\par
Let $\mathfrak{g}$ be a Lie algebra.
Then one can define a quantum group $U_q(\mathfrak{g})$ as a deformation of $\mathfrak{g}$ with $q$ a complex parameter (\cite{Drinfeld:ICM86}, \cite{Jimbo:LETMP1985}).
Given a representation $\rho\colon\mathfrak{g}\to\mathfrak{gl}(V)$ of $\mathfrak{g}$ one can construct an enhanced Yang--Baxter operator.
The corresponding invariant is called the quantum $(\mathfrak{g},V)$ invariant.
For details see \cite{Turaev:INVEM88}.
\par
To define the colored Jones polynomial we need the Lie algebra $\mathfrak{sl}_2(\mathbb{C})$ and its $N$-dimensional irreducible representation $\rho_N\colon\mathfrak{sl}_2(\mathbb{C})\to\mathfrak{gl}(V_N)$.
The quantum $(\mathfrak{sl}_2(\mathbb{C}),V_N)$ invariant is called the $N$-dimensional colored Jones polynomial $J_N(L;q)$.
\par
A precise definition is as follows.
\par
Put $V:=\mathbb{C}^N$ and define the $R$-matrix $R\colon V\otimes V\to V\otimes V$ by
\begin{equation*}
R(e_k\otimes e_l)
:=
\sum_{i,j=0}^{N-1}R_{kl}^{ij}e_{i}\otimes e_{j},
\end{equation*}
where
\begin{equation}\label{eq:R}
\begin{split}
R^{ij}_{kl}
:=&
\sum_{m=0}^{\min(N-1-i,j)}
\delta_{l,i+m}\delta_{k,j-m}
\frac{\{l\}!\{N-1-k\}!}{\{i\}!\{m\}!\{N-1-j\}!}
\\
&
\times q^{\bigl(i-(N-1)/2\bigr)\bigl(j-(N-1)/2\bigr)-m(i-j)/2-m(m+1)/4},
\end{split}
\end{equation}
with $\{e_0,e_1,\dots,e_{N-1}\}$ is the standard basis of $V$, $\{m\}:=q^{m/2}-q^{-m/2}$ and $\{m\}!:=\{1\}\{2\}\cdots\{m\}$.
Here $q$ is a complex parameter.
A homomorphism $\mu\colon V\to V$ is given by
\begin{equation*}
\mu(e_j)
:=
\sum_{i=0}^{N-1}\mu^i_j e_i
\end{equation*}
with
\begin{equation*}
\mu^i_j
:=
\delta_{i,j}q^{(2i-N+1)/2}.
\end{equation*}
\par
Then it can be shown that $(R,\mu,q^{(N^2-1)/4},1)$ gives an enhanced Yang--Baxter operator.
\begin{definition}[colored Jones polynomial]
For an integer $N\ge1$, put $V:=\mathbb{C}^N$ and define $R$ and $\mu$ as above.
The $N$-dimensional colored Jones polynomial $J_N(L;q)$ for a link $L$ is defined as
\begin{equation*}
J_N(L;q):=T_{(R,\mu,q^{(N^2-1)/4},1)}(\beta)\times\frac{\{1\}}{\{N\}},
\end{equation*}
where $\beta$ is a braid presenting the link $L$.
\end{definition}
\begin{remark}
Note that $J_N(\text{unknot};q)=1$ since
\begin{equation*}
\operatorname{Tr}_1(\mu)
=
\sum_{i=0}^{N-1}
q^{(2i-N+1)/2}
=
\frac{\{N\}}{\{1\}}.
\end{equation*}
\end{remark}
The two-dimensional colored Jones polynomial $J_2(L;q)$ is (a version) the original Jones polynomial \cite{Jones:BULAM385} as shown below.
\begin{lemma}
Let $L_+$, $L_-$, and $L_0$ be a skein triple, that is, they are the same links except for a small disk as shown in Figure~$\ref{fig:skein_triple}$.
\begin{figure}[h]
\raisebox{4mm}{$L_+:$}
\includegraphics[scale=0.3]{crossing_pos_small.eps}\quad,\quad
\raisebox{4mm}{$L_-:$}
\includegraphics[scale=0.3]{crossing_neg_small.eps}\quad,\quad
\raisebox{4mm}{$L_0:$}
\includegraphics[scale=0.3]{crossing_null_small.eps}
\caption{skein triple}
\label{fig:skein_triple}
\end{figure}
Then we have the following skein relation:
\begin{equation*}
q J_2(L_+;q)-q^{-1}J_2(L_-;q)=(q^{1/2}-q^{-1/2})J_2(L_0;q).
\end{equation*}
\end{lemma}
\begin{proof}
By the definition, the $R$-matrix is given by
\begin{equation*}
R
=
\begin{pmatrix}
q^{1/4}&0 &0 &0 \\
0 &q^{1/4}-q^{-3/4}&q^{-1/4}&0 \\
0 &q^{-1/4} &0 &0 \\
0 & &0 &q^{1/4}
\end{pmatrix}
\end{equation*}
with respect to the basis $\{e_0\otimes e_0,e_0\otimes e_1,e_1\otimes e_0,e_1\otimes e_1\}$ of $V\otimes V$, and $\mu$ is given by
\begin{equation*}
\mu
=
\begin{pmatrix}
q^{-1/2}&0 \\
0 &q^{1/2}
\end{pmatrix}
\end{equation*}
with respect to the basis $\{e_0,e_1\}$ of $V$.
\par
Therefore we can easily see that
\begin{equation}\label{eq:skein_R}
q^{1/4}R-q^{-1/4}R^{-1}
=
(q^{1/2}-q^{-1/2})\operatorname{Id}_V\otimes\operatorname{Id}_V.
\end{equation}
\par
Since $L_+$, $L_-$, and $L_0$ can be presented by $n$-braids $\beta\sigma_i\beta'$, $\beta\sigma_i^{-1}\beta'$, and $\beta\beta'$ respectively, we have
\begin{equation*}
\begin{split}
&
\left(
qJ_2(L_+;q)
-
q^{-1}J_2(L_-;q)
\right)
\times\frac{\{2\}}{\{1\}}
\\
=&
q\times
q^{-3(w(\beta\beta')+1)/4}
\operatorname{Tr}_1(\operatorname{Tr}_2(\cdots(\operatorname{Tr}_n(\Phi(\beta\sigma_i\beta')\mu^{\otimes n}))))
\\
&\quad
-
q^{-1}\times
q^{-3(w(\beta\beta')-1)/4}
\operatorname{Tr}_1(\operatorname{Tr}_2(\cdots(\operatorname{Tr}_n(\Phi(\beta\sigma_i^{-1}\beta')\mu^{\otimes n}))))
\\
=&
q^{-3w(\beta\beta')/4}
\\
&\times
\left(
q^{1/4}
\operatorname{Tr}_1(\operatorname{Tr}_2(\cdots(\operatorname{Tr}_n(\Phi(\beta\sigma_i\beta')\mu^{\otimes n}))))
-
q^{-1/4}
\operatorname{Tr}_1(\operatorname{Tr}_2(\cdots(\operatorname{Tr}_n(\Phi(\beta\sigma_i^{-1}\beta')\mu^{\otimes n}))))
\right)
\\
=&
q^{-3w(\beta\beta')/4}
\\
&\times
\left\{
\operatorname{Tr}_1(\operatorname{Tr}_2(\cdots(\operatorname{Tr}_n(
\Phi(\beta)
(\operatorname{Id}_V^{\otimes(i-1)}\otimes q^{1/4}R\otimes\operatorname{Id}_V^{\otimes(n-i-1)})
\Phi(\beta')
\mu^{\otimes n}))))
\right.
\\
&\left.
\qquad
-
\operatorname{Tr}_1(\operatorname{Tr}_2(\cdots(\operatorname{Tr}_n(
\Phi(\beta)
(\operatorname{Id}_V^{\otimes(i-1)}\otimes q^{-1/4}R^{-1}\otimes\operatorname{Id}_V^{\otimes(n-i-1)})
\Phi(\beta')\mu^{\otimes n}))))
\right\}
\\
\intertext{(from \eqref{eq:skein_R})}
=&
q^{-3w(\beta\beta')/4}
(q^{1/2}-q^{-1/2})
\operatorname{Tr}_1(\operatorname{Tr}_2(\cdots(\operatorname{Tr}_n(\Phi(\beta\beta')\mu^{\otimes n}))))
\\
=&
(q^{1/2}-q^{-1/2})J_2(L_0;q)\times\frac{\{2\}}{\{1\}},
\end{split}
\end{equation*}
completing the proof.
\end{proof}
\begin{remark}
The original Jones polynomial $V(L;q)$ satisfies
\begin{equation*}
q^{-1}V(L_+;q)-qV(L_-;q)
=
(q^{1/2}-q^{-1/2})V(L_0;q)
\end{equation*}
\cite[Theorem~12]{Jones:BULAM385}.
So we have $J_2(L;q)=(-1)^{\sharp(L)-1}V(L;q^{-1})$, where $\sharp(L)$ denotes the number of components of $L$.
\end{remark}
\subsection{Example of calculation}\label{subsec:calculation}
Put $\beta:=\sigma_1\sigma_2^{-1}\sigma_1\sigma_2^{-1}$.
Its closure $E$ is a knot called the figure-eight knot (Figure~\ref{fig:braid_presentation}).
We will calculate $J_N(E;q)$.
\par
Instead of calculating $\operatorname{Tr}_1(\operatorname{Tr}_2(\operatorname{Tr}_3(\Phi(\beta)\mu^{\otimes3})))\in\mathbb{C}$, we will calculate $\operatorname{Tr}_2(\operatorname{Tr}_3(\Phi(\beta)(\operatorname{Id}\otimes\mu\otimes\mu)))\in\operatorname{End}(V)$, which is a scalar multiple by Schur's lemma (for a proof see \cite[Lemma~3.9]{Kirby/Melvin:INVEM1991}).
See Figure~\ref{fig:tangle}.
\begin{figure}[h]
\includegraphics[scale=0.3]{tangle_small.eps}
\quad\raisebox{19mm}{$\Rightarrow$}\quad
\includegraphics[scale=0.3]{tangle_R_small.eps}
\caption{We close all the strings except for the first one.}
\label{fig:tangle}
\end{figure}
\par
Then $\operatorname{Tr}_1(\operatorname{Tr}_2(\operatorname{Tr}_3(\Phi(\beta)\mu^{\otimes3})))$ coincides with the trace of $\mu$ times the scalar $S$.
Since
\begin{equation*}
\begin{split}
T_{(R,\mu,q^{(N^2-1)/4},1)}(\beta)
&=
q^{-w(\beta)(N^2-1)/4}\operatorname{Tr}_1(\operatorname{Tr}_2(\operatorname{Tr}_3(\Phi(\beta)\mu^{\otimes3})))
\\
&=
q^{-w(\beta)(N^2-1)/4}\operatorname{Tr}_1(S\operatorname{Id}_V)
\\
&=
q^{-w(\beta)(N^2-1)/4}\sum_{i=0}^{N-1}S\, q^{(2i-N+1)/2}
\\
&=
q^{-w(\beta)(N^2-1)/4}\frac{\{N\}}{\{1\}}S,
\end{split}
\end{equation*}
we have $J_N(L;q)=q^{-w(\beta)(N^2-1)/4}S=S$.
\par
We need an explicit formula for the inverse of the $R$-matrix, which is given by
\begin{equation}\label{eq:R_inverse}
\begin{split}
(R^{-1})^{ij}_{kl}
=&
\sum_{m=0}^{\min(N-1-i,j)}
\delta_{l,i-m}\delta_{k,j+m}
\frac{\{k\}!\{N-1-l\}!}{\{j\}!\{m\}!\{N-1-i\}!}
\\
&\times
(-1)^{m}q^{-\bigl(i-(N-1)/2\bigr)\bigl(j-(N-1)/2\bigr)-m(i-j)/2+m(m+1)/4}.
\end{split}
\end{equation}
\par
To calculate the scalar $S$, draw a diagram for the braid $\beta$ and close it except for the first string (Figure~\ref{fig:R0}).
\begin{figure}[h]
\includegraphics[scale=0.3]{fig8_R0_small.eps}
\caption{Draw the braid $\sigma_1\sigma_2^{-1}\sigma_1\sigma_2^{-1}$
and close it except for the left-most one.}
\label{fig:R0}
\end{figure}
Fix a basis $\{e_0,e_1,\dots,e_{N-1}\}$ of $\mathbb{C}^N$.
\par
Label each arc with a non-negative integer $i$ less than $N$, which corresponds to a basis element $e_i$, where our braid diagram is divided into arcs by crossings so that at each crossing four arcs meet.
Since the homomorphism $\operatorname{Tr}_2(\operatorname{Tr}_1(\Phi(\beta)\otimes2))$ is a scalar multiple, we choose any basis for the first (top-left) arc of Figure~\ref{fig:R0} and calculate the scalar.
For simplicity we choose $e_{N-1}$ and so we label the first arc with $N-1$ (Figure~\ref{fig:R1}).
\begin{figure}[h]
\includegraphics[scale=0.3]{fig8_R1_small.eps}
\caption{Label the first (top-left) arc with $N-1$.}
\label{fig:R1}
\end{figure}
\par
Recall that we will associate the $R$-matrix or its inverse with each crossing as follows.
\begin{center}
\raisebox{-9mm}{\includegraphics[scale=0.3]{crossing_pos_ijkl_small.eps}}
\quad$\Rightarrow$\quad
$R^{ij}_{kl}$\, ,\quad
\raisebox{-9mm}{\includegraphics[scale=0.3]{crossing_neg_ijkl_small.eps}}
\quad$\Rightarrow$\quad
$(R^{-1})^{ij}_{kl}$
\end{center}
\par
Therefore we will label the other arcs following the following two rules:
\begin{enumerate}
\item[(i).]
At a positive crossing, the top-left label is less than or equal to the bottom-right label, the top-right label is greater than or equal to the bottom-left label, and their differences coincide (see \eqref{eq:R}).
\begin{center}
\raisebox{-9mm}{\includegraphics[scale=0.3]{crossing_pos_ijkl_small.eps}}
$: i+j=k+l$, $l\ge i$, $k\le j$,
\end{center}
\item[(ii).]
At a negative crossing, the top-left label is greater than or equal to the bottm-right label, the top-right label is less than or equal to the bottm-left label, and their differences coincide (see \eqref{eq:R_inverse}).
\begin{center}
\raisebox{-9mm}{\includegraphics[scale=0.3]{crossing_neg_ijkl_small.eps}}
$: i+j=k+l$, $l\le i$, $k\ge j$.
\end{center}
\end{enumerate}
\par
From Rule (i), the next arc should be labeled with $N-1$, and the difference at the top crossing is $0$ (Figure~\ref{fig:R2}).
This is why we chose $N-1$ for the label of the first arc.
\begin{figure}[h]
\includegraphics[scale=0.3]{fig8_R2_small.eps}
\caption{The label of the next arc should be $N-1$, and the difference
at the top crossing is $0$}
\label{fig:R2}
\end{figure}
\par
Label the top-right arc with $i$ with $0\le i\le N-1$ (Figure~\ref{fig:R3}).
\begin{figure}[h]
\includegraphics[scale=0.3]{fig8_R3_small.eps}
\caption{Label the top-right arc with $i$ ($0\le i\le N-1$).}
\label{fig:R3}
\end{figure}
\par
The label of the left-middle arc should be $i$ since the difference at the top crossing is $0$.
\begin{figure}[h]
\includegraphics[scale=0.3]{fig8_R4_small.eps}
\caption{The label of the left-middle arc should be $i$
since the difference at the top crossing is $0$.}
\label{fig:R4}
\end{figure}
\par
Label the arcs indicated in Figure~\ref{fig:R6} with $j$ and $k$ with $0\le j\le N-1$ and $0\le k\le N-1$.
Then the difference at the second top crossing is $N-k-1$.
Therefore the arc between the second top crossing and the second bottom crossing should be labeled with $N-k+j-1$ from Rule (ii) (Figure~\ref{fig:R7}).
Since the label should be less than $N$, we have $j-k\le0$.
\begin{figure}[h]
\includegraphics[scale=0.3]{fig8_R6_small.eps}
\caption{Choose $j$ and $k$.
Then the difference at the second top crossing is $N-k-1$.}
\label{fig:R6}
\end{figure}
\begin{figure}[h]
\includegraphics[scale=0.3]{fig8_R7_small.eps}
\caption{The label of the arc between the second top arc and
the second bottom arc should be $N-k+j-1$.
Note that $j-k\le0$.}
\label{fig:R7}
\end{figure}
\par
Look at the bottom-most crossing and apply Rule (ii).
We see that $i\ge k$ and the difference at the bottom-most crossing is $i-k$.
So the label of the arc between the second bottom crossing and the bottom-most crossing is $i+j-k$ (Figure~\ref{fig:R8}).
We also see that the label of the bottom-most arc is $N-1$ as we expected.
\begin{figure}[h]
\includegraphics[scale=0.3]{fig8_R8_small.eps}
\caption{The label of the arc between the second bottom crossing and
the bottom-most crossing should be $i+j-k$.
Note that $j-k\ge0$}
\label{fig:R8}
\end{figure}
\par
However from Rule (i) we have $j-k\ge0$ and so $k=j$.
Therefore we finally have the labeling as indicated in Figure~\ref{fig:R9}.
\begin{figure}[h]
\includegraphics[scale=0.3]{fig8_R9_small.eps}
\caption{The label $k$ should coincide with $i$.}
\label{fig:R9}
\end{figure}
\par
Now we can calculate the colored Jones polynomial.
We have
\begin{equation*}
\begin{split}
J_N(E;q)
&=
\sum_{i\ge j}
R^{N-1,i}_{i,N-1}\,
(R^{-1})^{N-1,j}_{N-1,j}\,
R^{i,N-1}_{N-1,i}\,
(R^ {-1})^{i,j}_{i,j}\,
\mu^{j}_{j}\,
\mu^{i}_{i}
\\
&=
\sum_{i\ge j}
(-1)^{N-1+i}
\frac{\{N-1\}!\{i\}!\{N-1-j\}!}{(\{j\}!)^2\{i-j\}!\{N-1-i\}!}
\\
&\phantom{=\sum_{i\ge j}}\times q^{(-i-i^2-2ij-2j^2+3N+6Ni+2Nj-3N^2)/4}.
\end{split}
\end{equation*}
\par
In this formula we need two summations.
To get a formula involving only one summation we regard the figure-eight knot $E$ as the closure of a tangle as shown in Figure~\ref{fig:fig8_calc}.
\begin{figure}[h]
\includegraphics[scale=0.3]{fig8_calc_small.eps}
\caption{The figure-eight knot can also regarded as the closure of a $(1,1)$-tangle.}
\label{fig:fig8_calc}
\end{figure}
In this case we need to put $\mu$ at each local minimum where the arc goes from left to right, and $\mu^{-1}$ at each local maximum where the arc goes from right to left.
See \cite[Theorem~3.6]{Kirby/Melvin:INVEM1991} for details.
\par
From Figure~\ref{fig:tangle}, we have
\begin{equation*}
\begin{split}
&J_N(E;q)
\\
=&
\sum_{\stackrel{0\le i\le N-1,0\le j\le N-1}{0\le i+j\le N-1}}
R^{i,0}_{0,i}\,
(R^{-1})^{i,j}_{i+j,0}\,
R^{0,i+j}_{i,j}\,
(R^ {-1})^{j,0}_{0,j}\,
(\mu^{-1})^{i}_{i}\,
\mu^{j}_{j}
\\
=&
\sum_{\stackrel{0\le i\le N-1,0\le j\le N-1}{0\le i+j\le N-1}}
(-1)^i
\frac{\{i+j\}!\{N-1\}!}{\{i\}!\{j\}!\{N-1-i-j\}!}
q^{-(N-1)i/2+(N-1)j/2-i^2/4+j^2/4-3i/4+3j/4}.
\end{split}
\end{equation*}
Putting $k:=i+j$, we have
\begin{equation*}
J_N(E;q)
=
\sum_{k=0}^{N-1}
\frac{\{N-1\}!}{\{N-1-k\}!}q^{k^2/4+Nk/2+k/4}
\left(
\sum_{i=0}^{k}(-1)^i
\frac{\{k\}!}{\{i\}!\{k-i\}!}
q^{-Ni-ik/2-i/2}
\right).
\end{equation*}
Using the formula (see \cite[Lemma~3.2]{Murakami/Murakami:ACTAM12001})
\begin{equation*}
\sum_{i=0}^{k}(-1)^iq^{li/2}\frac{\{k\}!}{\{i\}!\{k-i\}!}
=
\prod_{g=1}^{k}(1-q^{(l+k+1)/2-g}),
\end{equation*}
we have the following simple formula with only one summand, which is due to Habiro and L{\^e} (I learned this method from L{\^e}).
\begin{equation}\label{eq:fig8_single}
J_N(E;q)
=
\frac{1}{\{N\}}
\sum_{k=0}^{N-1}\frac{\{N+k\}!}{\{N-1-k\}!}.
\end{equation}
\section{Volume conjecture}\label{sec:Volume_Conjecture}
In this section we state the Volume Conjecture and then prove it for the figure-eight knot.
We also give supporting evidence for the conjecture.
\subsection{Statement of the Volume Conjecture}
In \cite{Kashaev:MODPLA95} Kashaev introduced a link invariant $\langle L\rangle_N\in\mathbb{C}$ for an integer $N$ greater than one and a link $L$ by using the quantum dilogarithm.
Then he observed in \cite{Kashaev:LETMP97} that the limit $2\pi\lim_{N\to\infty}\log|\langle K\rangle_N|/N$ is equal to the hyperbolic volume of the knot complement if a knot $K$ is hyperbolic.
Here a knot in the three-sphere $S^3$ is called hyperbolic if its complement possesses a complete hyperbolic structure with finite volume.
He also conjectured this would also hold for any hyperbolic knot.
\par
In \cite{Murakami/Murakami:ACTAM12001} J.~Murakami and I proved that Kashaev's invariant equals the $N$-dimensional colored Jones polynomial evaluated at the $N$-th root of unity, that is, $\langle L\rangle_N=J_N(L;\exp(2\pi\sqrt{-1}/N))$ and proposed that Kashaev's conjecture would hold for any knot by using the simplicial volume.
\begin{conjecture}[Volume Conjecture \cite{Kashaev:LETMP97,Murakami/Murakami:ACTAM12001}]\label{conj:VC}
The following equality would hold for any knot $K$.
\begin{equation*}
2\pi\lim_{N\to\infty}\frac{\log|J_N(K;\exp(2\pi\sqrt{-1}/N))|}{N}
=\operatorname{Vol}(S^3\setminus{K}).
\end{equation*}
\end{conjecture}
To define the simplicial volume (or Gromov norm), we introduce the Jaco--Shalen--Johannson (JSJ) decomposition (or the torus decomposition) of a knot complement.
\begin{definition}[Jaco--Shalen--Johannson decomposition \cite{Jaco/Shalen:MEMAM79,Johannson:1979}]
Let $K$ be a knot.
Then its complement $S^3\setminus{K}$ can be uniquely decomposed into hyperbolic pieces and Seifert fibered pieces by a system of tori:
\begin{equation*}
S^3\setminus{K}
=
\left(\bigsqcup H_i\right)\sqcup\left(\bigsqcup E_j\right)
\end{equation*}
with $H_i$ hyperbolic and $E_j$ Seifert-fibered.
\end{definition}
Then the simplicial volume of the knot complement is defined to be the sum of the hyperbolic volumes of the hyperbolic pieces.
\begin{definition}[Simplicial volume (Gromov norm) \cite{Gromov:INSHE82}]
If a knot complement $S^3\setminus{K}$ is decomposed as above, then its simplicial volume $\operatorname{Vol}(S^3\setminus{K})$ is defined as
\begin{equation*}
\operatorname{Vol}(S^3\setminus{K})
:=
\sum_{H_i:\text{hyperbolic piece}}\text{Hyperbolic Volume of $H_i$}.
\end{equation*}
\end{definition}
\begin{example}
Let us consider the $(2,1)$-cable of the figure-eight knot as shown in figure~\ref{fig:2_1_fig_8}.
\begin{figure}[h]
\includegraphics[scale=0.3]{2_1_cable_small.eps}
\caption{$(2,1)$-cable of the figure-eight knot}
\label{fig:2_1_fig_8}
\end{figure}
Then its complement can be decomposed by a torus into two pieces (Figure~\ref{fig:2_1_fig_8_JSJ}), one hyperbolic and one Seifert fibered.
\begin{figure}[h]
\includegraphics[scale=0.3]{complement_2_1_cable_small.eps}
\quad\raisebox{15mm}{$=$}\quad
$\underset{\text{\raisebox{-2mm}{hyperbolic}}}
{\includegraphics[scale=0.3]{complement_fig8_small.eps}}$
\quad\raisebox{15mm}{$\sqcup$}\quad
\raisebox{1mm}{$\underset{\text{\raisebox{-3mm}{Seifert fibered}}}
{\includegraphics[scale=0.3]{complement_2_1_torus_small.eps}}$}
\caption{The JSJ decomposition of the $(2,1)$-cable of the figure-eight knot}
\label{fig:2_1_fig_8_JSJ}
\end{figure}
Therefore we have
\begin{equation*}
\operatorname{Vol}
\left(
\raisebox{-14mm}{\includegraphics[scale=0.3]{complement_2_1_cable_small.eps}}
\right)
=
\operatorname{Vol}
\left(
\raisebox{-14mm}{\includegraphics[scale=0.3]{complement_fig8_small.eps}}
\right).
\end{equation*}
\end{example}
\subsection{Proof of the Volume Conjecture for the figure-eight knot}
\label{subsec:proof_fig8}
We give a proof of the Volume Conjecture for the figure-eight knot due to T.~Ekholm.
\subsubsection{Calculation of the limit}
We use the formula \eqref{eq:fig8_single} of the colored Jones polynomial for the figure-eight knot $E$ due to Habiro and L\^e.
(See \cite{Habiro:SURIK2000,Masbaum:ALGGT12003} for Habiro's method.)
By a simple calculation using it, we have
\begin{equation}\label{eq:Jones_fig8}
J_N\left(E;q\right)
=
\sum_{j=0}^{N-1}
\prod_{k=1}^{j}
\left(q^{(N-k)/2}-q^{-(N-k)/2}\right)
\left(q^{(N+k)/2}-q^{-(N+k)/2}\right).
\end{equation}
Replacing $q$ with $\exp(2\pi\sqrt{-1}/N)$, we have
\begin{equation*}
J_N\left(E;\exp(2\pi\sqrt{-1}/N)\right)
=
\sum_{j=0}^{N-1}\prod_{k=1}^{j}f(N;k)
\end{equation*}
where we put $f(N;k):=4\sin^2(k\pi/N)$.
The graph of $f(N;k)$ is depicted in Figure~\ref{fig:graph_f}.
\begin{figure}[h]
\includegraphics[scale=0.8]{graph_f_small.eps}
\caption{Graph of $f(N;k)$}
\label{fig:graph_f}
\end{figure}
\par
Put $g(N;j):=\prod_{k=1}^{j}f(N;k)$ so that $J_N\left(E;\exp(2\pi\sqrt{-1}/N)\right)=\sum_{j=0}^{N-1}g(N;j)$.
Then $g(N;j)$ decreases when $0<j<N/6$ and $5N/6<j$, and increases when $N/6<j<5N/6$.
Therefore $g(N;j)$ takes its maximum at $j=5N/6$.
(To be precise we need to take the integer part of $5N/6$.)
See Table~\ref{tbl:f_g}.
\begin{table}[h]
\begin{tabular}{|c|c|c|c|c|c|c|c|}
\hline
$j$ &$0$& $\cdots$ &$N/6$& $\cdots$ &$5N/6$& $ \cdots$ &$1$\\
\hline
$f(N;k)$& & $<1$ & $1$ & $>1$ & $1$ & $<1$ & \\
\hline
$g(N;j)$&$1$&$\searrow$& &$\nearrow$&maximum&$\searrow$& \\
\hline
\end{tabular}
\medskip
\caption{table of $f(N;k)$ and $g(N;j)$}
\label{tbl:f_g}
\end{table}
Since there are $N$ positive terms in $J_N\left(E;\exp(2\pi\sqrt{-1}/N)\right)=\sum_{j=0}^{N-1}g(N;j)$ and $g(N;5N/6)$ is the maximum of these terms, we have
\begin{equation*}
g(N;5N/6)
\le
J_N\left(E;\exp(2\pi\sqrt{-1}/N)\right)
\le
N\times g(N;5N/6).
\end{equation*}
Noting that each side is positive, we take their logarithms and divide them by $N$.
\begin{equation*}
\frac{\log{g(N;5N/6)}}{N}
\le
\frac{\log J_N\left(E;\exp(2\pi\sqrt{-1}/N)\right)}{N}
\le
\frac{\log{N}}{N}+
\frac{\log{g(N;5N/6)}}{N}.
\end{equation*}
Since $\log_{N\to\infty}(\log{N})/N=0$, we have
\begin{equation*}
\lim_{N\to\infty}\frac{\log{g(N;5N/6)}}{N}
\le
\lim_{N\to\infty}
\frac{\log J_N\left(E;\exp(2\pi\sqrt{-1}/N)\right)}{N}
\le
\lim_{N\to\infty}\frac{\log{g(N;5N/6)}}{N}.
\end{equation*}
Therefore we have
\begin{equation*}
\lim_{N\to\infty}
\frac{\log{J_N\left(E;\exp(2\pi\sqrt{-1}/N)\right)}}{N}
=
\lim_{N\to\infty}\frac{\log{g(N;5N/6)}}{N}.
\end{equation*}
\par
We can calculate the limit $\lim_{N\to\infty}\bigl(\log{g(N;5N/6)}\bigr)/N$ by integration.
Since $g(N;j)=\prod_{k=1}^{j}f(N;k)$, we have
\begin{equation}\label{eq:Jones_integral}
\begin{split}
\lim_{N\to\infty}\frac{\log{g(N;5N/6)}}{N}
&=
\lim_{N\to\infty}
\frac{1}{N}
\sum_{k=1}^{5N/6}\log{f(N;k)}
\\
&=
2
\lim_{N\to\infty}
\frac{1}{N}
\sum_{k=1}^{5N/6}\log\bigl(2\sin(k\pi/N)\bigr)
\\
&=
\frac{2}{\pi}
\int_{0}^{5\pi/6}
\log\bigl(2\sin{x}\bigr)\,dx.
\end{split}
\end{equation}
What does this mean?
I will explain a geometric interpretation of this integral.
\subsubsection{Lobachevsky function}
The function defined by the integral in \eqref{eq:Jones_integral} is known as the Lobachevsky function.
More precisely, we define the Lobachevsky function $\Lambda(\theta)$ as
\begin{equation*}
\Lambda(\theta)
:=
-\int_{0}^{\theta}\log|2\sin{x}|\,dx
\end{equation*}
for $\theta\in\mathbb{R}$.
By using this function, we can express the limit of the colored Jones polynomial as
\begin{equation}\label{eq:Jones_Lobachevsky}
\lim_{N\to\infty}
\frac{\log{J_N\left(E;\exp(2\pi\sqrt{-1}/N)\right)}}{N}
=
-\frac{2}{\pi}\Lambda(5\pi/6)
\end{equation}
We show some properties of the Lobachevsky function (see for example \cite{Milnor:BULAM382}).
\begin{lemma}
The Lobachevsky function satisfies the following two properties.
\begin{enumerate}
\item The Lobachevsky function is an odd function and has period $\pi$.
\item We have $\Lambda(2\theta)=2\Lambda(\theta)+2\Lambda(\theta+\pi/2)$.
In general we have $\Lambda(n\theta)=n\sum_{k=1}^{n-1}\Lambda(\theta+k\pi/n)$.
\end{enumerate}
\end{lemma}
\begin{proof}
The first property is easily shown by the periodicity of the sine function.
\par
To prove the second, we use the double angle formula of the sine function: $\sin(2x)=2\sin{x}\cos{x}$.
We have
\begin{equation*}
\log|2\sin(2x)|
=
\log|2\sin{x}|
+
\log|2\cos{x}|
=
\log|2\sin x|
+
\log|2\sin(x+\pi/2)|,
\end{equation*}
completing the proof.
\end{proof}
From (1) we have
\begin{equation*}
\Lambda(5\pi/6)
=
\Lambda(-\pi/6)
=
-\Lambda(\pi/6).
\end{equation*}
From (2) and (1) we also have
\begin{equation*}
\Lambda(\pi/3)
=
2\Lambda(\pi/6)+2\Lambda(2\pi/3)
=
2\Lambda(\pi/6)-2\Lambda(\pi/3).
\end{equation*}
Therefore we have
\begin{equation*}
\Lambda(5\pi/6)=-\frac{3}{2}\Lambda(\pi/3).
\end{equation*}
\par
Returning to the limit of the Jones polynomial \eqref{eq:Jones_Lobachevsky}, we conclude that
\begin{equation*}
2\pi\lim_{N\to\infty}
\frac{\log{J_N\left(E;\exp(2\pi\sqrt{-1}/N)\right)}}{N}
=
6\Lambda(\pi/3).
\end{equation*}
\par
Next we show how the Lobachevsky function is related to hyperbolic geometry.
\subsubsection{Hyperbolic geometry}
\label{subsubsec:geometry}
It is well known that the complement of the figure-eight knot can be decomposed into two ideal hyperbolic regular tetrahedra.
\begin{theorem}[W.~Thurston \cite{Thurston:GT3M}]
The complement of the figure-eight knot can be obtained by gluing two ideal hyperbolic regular tetrahedra.
\end{theorem}
I will explain what is an ideal hyperbolic regular tetrahedron later.
Topologically, the theorem states that the complement of the figure-eight knot can be obtained by gluing two truncated tetrahedra as in Figure~\ref{fig:fig8_tetra} (see also \cite{Murakami:FUNDM2004}).
In Figure~\ref{fig:fig8_tetra} we identify $A$ with $A'$, $B$ with $B'$, $C$ with $C'$ and $D$ with $D'$.
Note that edges with single arrows and edges with double arrows are also identified respectively.
\begin{figure}[h]
\includegraphics[scale=0.25]{complement_fig8_small.eps}
\quad\raisebox{13mm}{$=$}\quad
\includegraphics[scale=0.4]{tetrahedron_z_small.eps}
\quad\raisebox{13mm}{$\underset{\substack{A=A',B=B',\\mathbb{C}=C',D=D'}}{\cup}$}\quad
\includegraphics[scale=0.4]{tetrahedron_w_small.eps}
\caption{The complement of the figure-eight knot is decomposed into two ideal hyperbolic regular tetrahedra.
Shadowed triangles make the boundary of the knot complement.}
\label{fig:fig8_tetra}
\end{figure}
\par
Here I give a short introduction to hyperbolic geometry.
\par
First consider the upper half space $\{(x,y,z)\in\mathbb{R}^3\mid z>0\}$ with hyperbolic metric $ds:=\sqrt{dx^2+dy^2+dz^2}/z$ and denote it by $\H^3$.
It is known that a geodesic line in $\H^3$ is a semicircle or a straight line perpendicular to the $xy$-plane, and that a geodesic plane is a hemisphere or a flat plane perpendicular to the $xy$-plane.
\par
An ideal hyperbolic tetrahedron is a tetrahedron in $\H^3$ with geodesic faces with four vertices at infinity, that is, on the $xy$-plain or at the point at infinity $\infty$.
By isometry we may assume that one vertex is $\infty$ and the other three are on the $xy$-plane.
So its faces consist of three perpendicular planes and a hemisphere as shown in Figure~\ref{fig:tetrahedron}.
\begin{figure}[h]
\includegraphics[scale=0.7]{tetrahedron_small.eps}
\caption{An ideal hyperbolic tetrahedron is the part above the hemisphere surrounded by three perpendicular planes.}
\label{fig:tetrahedron}
\end{figure}
If we see the tetrahedron from the top, it is a (Euclidean) triangle with angles $\alpha$, $\beta$, and $\gamma$.
\begin{figure}[h]
\includegraphics[scale=0.7]{tetrahedron_skeleton_small.eps}
\hspace{20mm}
\raisebox{7mm}{\includegraphics[scale=0.5]{triangle_small.eps}}
\caption{A top view (right) of the ideal hyperbolic tetrahedron (left)}
\label{fig:top_view}
\end{figure}
It is known that an ideal hyperbolic tetrahedron is defined (up to isometry) by the similarity class of this triangle.
Therefore we can parametrize an ideal hyperbolic tetrahedron by a triple of positive numbers $(\alpha,\beta,\gamma)$ with $\alpha+\beta+\gamma=\pi$.
We denote it by $\Delta(\alpha,\beta,\gamma)$.
\par
The hyperbolic volume $\operatorname{Vol}(\Delta(\alpha,\beta,\gamma))$ can be expressed by using the Lobachevsky function $\Lambda(\theta)$.
In fact it can be shown that
\begin{equation*}
\operatorname{Vol}(\Delta(\alpha,\beta,\gamma))
=
\Lambda(\alpha)+\Lambda(\beta)+\Lambda(\gamma).
\end{equation*}
For a proof, see for example \cite[Chapter~7]{Thurston:GT3M}.
\par
Now we return to the decomposition of the figure-eight knot complement.
Figure~\ref{fig:fig8_tetra} shows that after identification we have two edges, edge with single arrow and edge with double arrow, each of them is obtained by identifying six edges.
So if the ideal hyperbolic tetrahedra we are using are regular, that is, isometric to $\Delta(\pi/3,\pi/3,\pi/3)$ then the sum of dihedral angles around each edge becomes $2\pi$.
This means that if we use two ideal hyperbolic regular tetrahedra, our gluing is geometric, that is, the complement of the figure-eight knot is isometric to the union of two copies of $\Delta(\pi/3,\pi/3,\pi/3)$.
In particular its volume equals $2\operatorname{Vol}(\Delta(\pi/3,\pi/3,\pi/3))=6\Lambda(\pi/3)$.
\par
Thus we have proved
\begin{equation*}
2\pi\lim_{N\to\infty}
\frac{\log{J_N\left(E;\exp(2\pi\sqrt{-1}/N)\right)}}{N}
=
6\Lambda(\pi/3)
=
\operatorname{Vol}\left(S^3\setminus{E}\right),
\end{equation*}
which is the statement of the Volume Conjecture for the figure-eight knot.
\subsection{Knots and links for which the Volume Conjecture is proved}
As far as I know the Volume Conjecture is proved for
\begin{enumerate}
\item figure-eight knot by Ekholm,
\item $5_2$ knot by Kashaev and Yokota,
\item Whitehead doubles of torus knots by Zheng \cite{Zheng:CHIAM22007},
\item torus knots by Kashaev and Tirkkonen \cite{Kashaev/Tirkkonen:ZAPNS2000},
\item torus links of type $(2,2m)$ by Hikami \cite{Hikami:COMMP2004},
\item knots and links with volume zero by van der Veen \cite{van_der_veen:2008},
\item Borromean rings by Garoufalidis and L{\^e} \cite{Garoufalidis/Le:2005},
\item twisted Whitehead links by Zheng \cite{Zheng:CHIAM22007},
\item Whitehead chains by van der Veen \cite{van_der_Veen:ACTMV2009},
\item a satellite link around the figure-eight knot with pattern the Whitehead link by Yamazaki and Yokota \cite{Yamazaki/Yokota}.
\end{enumerate}
Note that (1) and (2) are for hyperbolic knots, (3) is for a knot whose JSJ decomposition consists of a hyperbolic piece and a Seifert fibered piece, (4)--(6) are for knots and links only with Seifert pieces, (7)--(9) are for hyperbolic links, and (10) is for a link whose JSJ decomposition consists of a hyperbolic piece and a Seifert fibered piece.
\section{Supporting evidence for the Volume Conjecture}
\label{sec:VC_proof}
The Volume Conjecture is proved only for several knots and links but I think it is true possibly with some modification; for example we may need to replace the limit with the limit superior (see \cite[Conjecture~2]{van_der_Veen:ACTMV2009}).
In this section I will explain why I think it is true.
\begin{remark}[Caution!]
Descriptions in this section are {\em not rigorous}.
\end{remark}
\subsection{Approximation of the colored Jones polynomial}
Put $\xi_N:=\exp(2\pi\sqrt{-1}/N)$ and I will give an interpretation of the $R$-matrix used to define the colored Jones polynomial.
From \eqref{eq:R} and \eqref{eq:R_inverse} we have
\begin{equation}\label{eq:R_N}
\begin{split}
R^{ij}_{kl}\bigr|_{q:=\xi_N}
&=
\sum_{m}
\text{(a power of $\xi_N$)}\times
\delta_{l,i+m}\delta_{k,j-m}\,
\frac{\{l\}!\{N-1-k\}!}{\{i\}!\{m\}!\{N-1-j\}!},
\\
(R^{-1})^{ij}_{kl}\bigr|_{q:=\xi_q}
&=
\sum_{m}
\text{(a power of $\xi_N$)}\times
\delta_{l,i-m}\delta_{k,j+m}\,
\frac{\pm\{k\}!\{N-1-l\}!}{\{j\}!\{m\}!\{N-1-i\}!}.
\end{split}
\end{equation}
Since
\begin{equation*}
\{k\}!\bigr|_{q:=\xi_N}
=
\left(2\sqrt{-1}\right)^{k}
\sin(\pi/N)
\sin(2\pi/N)
\cdots
\sin(k\pi/N),
\end{equation*}
we have
\begin{equation*}
\begin{split}
&
\{k\}!\{N-k-1\}!\bigr|_{q:=\xi_N}
\\
=&
\left(2\sqrt{-1}\right)^{N-1}
\sin(\pi/N)\sin(2\pi/N)\cdots\sin(k\pi/N)
\\
&\times
\sin\bigl((N-(k+1))\pi/N\bigr)
\sin\bigl((N-(k+2))\pi/N\bigr)
\sin\bigl((N-(N-1))\pi/N\bigr)
\\
=&
(-1)^{N-k-1}\left(\sqrt{-1}\right)^{N-1}
2^{N-1}
\sin(\pi/N)\sin(2\pi/N)\cdots\sin((N-1)\pi/N)
\\
=&
\varepsilon N
\end{split}
\end{equation*}
for $\varepsilon\in\{1,-1,\sqrt{-1},-\sqrt{-1}\}$.
For the last equality, see for example \cite{Murakami/Murakami:ACTAM12001}.
So if we put
\begin{align*}
(\xi_N)_{k^+}&:=(1-\xi_N)\cdots(1-\xi^k_N),
\\
(\xi_N)_{k^{-}}&:=(1-\xi_N)\cdots(1-\xi^{N-1-k}_N),
\end{align*}
we have
\begin{align*}
\{k\}!\bigr|_{q:=\xi_N}
&=
\text{(a power of $\xi_N$)}\times(\xi_N)_{k^+},
\\
\{N-1-k\}!\bigr|_{q:=\xi_N}
&=
\text{(a power of $\xi_N$)}\times(\xi_N)_{k^-}
\\
\intertext{and}
(\xi_N)_{k^+}(\xi_N)_{k^-}
&=
\text{(a power of $\xi_N$)}\times N.
\end{align*}
\par
Therefore from \eqref{eq:R_N} the $R$-matrix and its inverse can be written as
\begin{align*}
R^{ij}_{kl}\bigr|_{q:=\xi_N}
&=
\sum_{m}
\delta_{l,i+m}\delta_{k,j-m}\,
\frac{\text{(a power of $\xi_N$)}\times N^2}
{(\xi_N)_{m^+}(\xi_N)_{i^+}(\xi_N)_{k^+}
(\xi_N)_{j^-}(\xi_N)_{l^-}},
\\
(R^{-1})^{ij}_{kl}\bigr|_{q:=\xi_N}
&=
\sum_{m}
\delta_{l,i-m}\delta_{k,j+m}\,
\frac{\text{(a power of $\xi_N$)}\times N^{2}}
{(\xi_N)_{m^+}(\xi_N)_{i^-}(\xi_N)_{k^-}
(\xi_N)_{j^+}(\xi_N)_{l^+}}.
\end{align*}
\par
If we are given a knot $K$, we express it as a closed braid and calculate the colored Jones polynomial as described in \S\ref{subsec:calculation}.
Then we have
\begin{equation}\label{eq:Jones_R}
\begin{split}
J_N(K;\xi_N)
&=
\sum_{\text{labelings}}
\left(
\prod_{\text{crossings}}
(R^{\pm1})^{ij}_{kl}
\right)
\\
&=
\sum_{\text{labelings}}
\left(
\prod_{\text{crossings}}
\frac{\text{(a power of $\xi_N$)}\times N^{2}}
{(\xi_N)_{m^+}(\xi_N)_{i^{\pm}}(\xi_N)_{k^{\pm}}
(\xi_N)_{j^{\mp}}(\xi_N)_{l^{\mp}}}
\right),
\end{split}
\end{equation}
where the summation is over all the labelings with $\{0,1,\dots,N-1\}$ corresponding to the basis $\{e_0,e_1,\dots,e_{N-1}\}$ and for a fixed labeling the product is over all the crossings, each of which corresponds to an entry $R^{ij}_{kl}$ (or $(R^{-1})^{i,j}_{kl}$, respectively), determined by the four labeled arcs around the vertex, of the $R$-matrix (or its inverse, respectively).
\par
We will approximate $(\xi_N)_{k^{+}}$ for large $N$.
By taking the logarithm, we have
\begin{equation*}
\begin{split}
\log(\xi_N)_{k^{+}}
&=
\sum_{j=1}^{k}
\log(1-\xi^j_N)
\\
&=
\sum_{j=1}^{k}
\log\bigl(1-\exp(2\pi\sqrt{-1}j/N)\bigr).
\end{split}
\end{equation*}
Putting $x:=j/N$, we may replace the summation with the following integral for large $N$ (this is {\em not rigorous!}).
\begin{equation*}
\begin{split}
\log(\xi_N)_{k^{+}}
&\underset{N\to\infty}{\approx}
N\int_{0}^{k/N}\log\bigl(1-\exp(2\pi\sqrt{-1}x)\bigr)\,dx
\\
&=
\frac{N}{2\pi\sqrt{-1}}
\int_{1}^{\exp(2\pi\sqrt{-1}k/N)}\frac{\log(1-y)}{y}\,dy,
\end{split}
\end{equation*}
where we put $y:=\exp(2\pi\sqrt{-1}x)$ in the last equality and $\underset{N\to\infty}{\approx}$ means a very rough approximation (which may be not true at all) for large $N$.
\par
This integral is known as the dilogarithm function.
We put
\begin{equation*}
\operatorname{Li}_2(z)
:=
-\int_{0}^{z}\frac{\log(1-y)}{y}\,dy
\end{equation*}
for $z\in\mathbb{C}\setminus{[1,\infty)}$.
This is called dilogarithm since it has the Taylor expansion for $|z|<1$ as
\begin{equation*}
\sum_{n=1}^{\infty}\frac{z^n}{n^2}.
\end{equation*}
For more details about this function, see for example \cite{Zagier:2007}.
\par
So by using the dilogarithm function, we have the following approximation:
\begin{equation*}
\log(\xi_N)_{k^{+}}
\underset{N\to\infty}{\approx}
\frac{N}{2\pi\sqrt{-1}}
\left[
\operatorname{Li}_2(1)-\operatorname{Li}_2(\xi^k_N)
\right].
\end{equation*}
Since $\operatorname{Li}_2(1)$, which equals $\zeta(2)=\pi^2/6$ ($\zeta(z)$ is the Riemann zeta function), can be ignored for large $N$, we have
\begin{equation*}
(\xi_N)_{k^{+}}
\underset{N\to\infty}{\approx}
\exp
\left[
-\frac{N}{2\pi\sqrt{-1}}\operatorname{Li}_2(\xi^{k}_N)
\right].
\end{equation*}
Similarly we have
\begin{equation*}
(\xi_N)_{k^{-}}
\underset{N\to\infty}{\approx}
\exp
\left[
-\frac{N}{2\pi\sqrt{-1}}\operatorname{Li}_2(\xi^{-k}_N)
\right].
\end{equation*}
\par
Therefore from \eqref{eq:Jones_R} the colored Jones polynomial can be (roughly) approximated as follows.
\begin{multline}\label{eq:Jones_approx}
J_N(K;\xi_N)
\underset{N\to\infty}{\approx}
\sum_{\text{labelings}}
\exp
\left[\vphantom{\sum_{\text{crossings}}}
\frac{N}{2\pi\sqrt{-1}}
\right.
\\
\left.
\times
\left(
\sum_{\text{crossings}}
\left\{
\operatorname{Li}_2(\xi^m_N)
+
\operatorname{Li}_2(\xi^{\pm i}_N)
+
\operatorname{Li}_2(\xi^{\mp j}_N)
+
\operatorname{Li}_2(\xi^{\pm k}_N)
+
\operatorname{Li}_2(\xi^{\mp l}_N)
+
\text{$\log$ terms}
\right\}
\right)
\right],
\end{multline}
where the $\log$ terms come from powers of $\xi_N$ and $N$.
\par
Since the term
\begin{equation}\label{eq:dilog}
\sum_{\text{crossings}}
\left\{
\operatorname{Li}_2(\xi^m_N)
+
\operatorname{Li}_2(\xi^{\pm i}_N)
+
\operatorname{Li}_2(\xi^{\mp j}_N)
+
\operatorname{Li}_2(\xi^{\pm k}_N)
+
\operatorname{Li}_2(\xi^{\mp l}_N)
+
\text{$\log$ terms}
\right\}
\end{equation}
can be regarded as a function of $\xi_N^{i_1},\xi_N^{i_2},\dots,\xi_N^{i_c}$ with $i_1,i_2,\dots,i_c$ labelings of arcs, we can write it as $V(\xi_N^{i_1},\xi_N^{i_1}\dots,\xi_N^{i_c})$.
Note that $m$ can be expressed as a difference of two such labelings.
So we have
\begin{equation}\label{eq:Jones_sum}
J_N(K;\xi_N)
\underset{N\to\infty}{\approx}
\sum_{i_1,i_2,\dots,i_c}
\exp
\left[
\frac{N}{2\pi\sqrt{-1}}
V(\xi_N^{i_1},\xi_N^{i_2},\dots,\xi_N^{i_c})
\right].
\end{equation}
\par
We want to apply a method used in the proof for the figure-eight knot in \S\ref{subsec:proof_fig8}.
Recall that the point of the proof is to find the maximum term of the summand because the maximum dominates the asymptotic behavior.
We will seek for the ``maximum'' term of the summation in \eqref{eq:Jones_sum}.
\par
To do that we first approximate this with the following integral:
\begin{equation}\label{eq:Jones_integral2}
J_N(K;\xi_N)
\underset{N\to\infty}{\approx}
\int_{J_1}\int_{J_2}\cdots\int_{J_c}
\exp
\left[
\frac{N}{2\pi\sqrt{-1}}
V(z_1,z_2\dots,z_c)
\right]\,dz_1dz_2\cdots dz_c,
\end{equation}
where $z_d$ corresponds to $\xi_N^{i_d}$ and $J_1,J_2,\dots,J_c$ are certain contours.
(The argument here is not rigorous.
In particular I do not know how to choose the contours.)
\par
Then we will find the maximum of the absolute value of the integrand.
To be more precise we apply the steepest descent method.
For a precise statement, see for example \cite[Theorem~7.2.9]{Marsden/Hoffman:Complex_Analysis}).
\begin{theorem}[steepest descent method]\label{thm:steepest_descent_method}
Under certain conditions for the functions $f$, $g$, and a contour $C$, we have
\begin{equation*}
\begin{split}
\int_{C}g(z)\exp(N\,h(z))\,dz
&\underset{N\to\infty}{\sim}
\frac{\exp(N\,h(z_0))\sqrt{2\pi}g(z_0)}{\sqrt{N}\sqrt{-h''(z_0)}}
\\
&\underset{N\to\infty}{\approx}
\exp(N\,h(z_0)),
\end{split}
\end{equation*}
where $h'(z_0)=0$ and $\Re(h(z))$ takes its positive maximum at $z_0$.
\par
Note that the symbol $\underset{N\to\infty}{\sim}$ means that the ratio of both sides converges to $1$ when $N\to\infty$ and that we ignore the constant term and $\sqrt{N}$ in the rough approximation $\underset{N\to\infty}{\approx}$ because $\exp(N\,h(z_0))$ grows exponentially when $N\to\infty$.
$($Recall that we want to know the limit of $\log|J_N(K,\xi_N)|/N$ and so polynomial terms will not matter.$)$
\end{theorem}
\begin{remark}
In general, to apply the steepest descent method, we need to change the contour $C$ so that it passes through $z_0$.
\end{remark}
Now we apply (a multidimensional version of) this method to \eqref{eq:Jones_integral2} and we will find the the maximum of $\{\Im V(z_1,\dots,z_c)\mid{(z_1,z_2,\dots,z_c)\in J_1\times J_2\dots\times J_c}\}$.
Let $(x_1,x_2,\dots,x_c)$ be such a point.
Then we have
\begin{equation}\label{eq:steepest_descent}
J_N(K;\xi_N)
\underset{N\to\infty}{\approx}
\exp
\left[
\frac{N}{2\pi\sqrt{-1}}
V(x_1,x_2,\dots,x_c)
\right]
\end{equation}
and so we finally have
\begin{equation}\label{eq:limit}
2\pi\sqrt{-1}
\lim_{N\to\infty}
\frac{\log{J_N(K;\xi_N)}}{N}
=
V(x_1,x_2,\dots,x_c).
\end{equation}
Note that the point $(x_1,x_2,\dots,x_c)$ is a solution to the following equation:
\begin{equation}\label{eq:differential}
\frac{\partial\,V}{\partial\,z_d}(z_1,z_2,\dots,z_c)=0
\end{equation}
for $d=1,2,\dots,c$.
\par
Remember that our argument here is far from rigor!
Especially I am cheating on the following points:
\begin{itemize}
\item Replacing a summation with an integral in \eqref{eq:Jones_integral2}.
Here we do not know how to choose the multidimensional contour.
\item Applying the steepest descent method in \eqref{eq:steepest_descent}.
In general, we have many solutions to the system of equations \eqref{eq:differential} but we do not know which solution gives the maximum.
Moreover we may need to change the contour so that it passes through the solution that gives the maximum but we do not know whether this is possible or not.
\end{itemize}
\subsection{Geometric interpretation of the limit}
In this subsection I will give a geometric interpretation of the limit \eqref{eq:limit}.
\par
Since $V(z_1,z_2,\dots,z_c)$ is the sum of the terms as in \eqref{eq:dilog}, we first describe a geometric meaning of $\operatorname{Li}_2(\zeta_N^{\pm i})$.
\par
Recall that an ideal hyperbolic tetrahedron can be put in $\H^3$.
Regarding the $xy$-plane as the complex plain, we can assume that the three of the four (ideal) vertices are at $0$, $1$ and $z\in\mathbb{C}$ ($\Im{z}>0$), respectively (Figure~\ref{fig:tetrahedron_parameter}).
\begin{figure}[h]
\includegraphics[scale=0.7]{tetrahedron_triangle_small.eps}
\quad\raisebox{13mm}{$\Rightarrow$}\quad
\raisebox{3mm}{\includegraphics[scale=0.3]{triangle_z_small.eps}}
\caption{Parametrization of ideal hyperbolic tetrahedra}
\label{fig:tetrahedron_parameter}
\end{figure}
Thus the set $\{z\in\mathbb{C}\mid\Im{z}>0\}$ gives a parametrization of ideal hyperbolic tetrahedra.
We denote by $\Delta(z)$ the hyperbolic tetrahedron parametrized by $z$.
\par
The volume of $\Delta(z)$ is given as follows (see for example \cite[p.~324]{Neumann/Zagier:TOPOL85}):
\begin{equation}\label{eq:tetra_vol}
\operatorname{Vol}(\Delta(z))
=
\Im\operatorname{Li}_2(z)+\log|z|\arg(1-z).
\end{equation}
\par
So we expect $V(x_1,x_2,\dots,x_c)$ gives the sum of the volumes of certain tetrahedra related to the knot.
\par
In fact we can express the volume of the knot complement in terms of $V(z_1,z_2,\dots,z_c)$ \cite{D.Thurston:Grenoble,Yokota:GTM02}.
\par
We follow \cite{D.Thurston:Grenoble} to describe this.
\par
We decompose the knot complement into topological, truncated tetrahedra.
To do this we put an octahedron at each positive crossing as in Figure~\ref{fig:crossing_octa}, where $i$, $j$, $k$ and $l$ are labeling of the four arcs around the vertex.
\begin{figure}[h]
\includegraphics[scale=0.3]{crossing_pos_ijkl_big_small.eps}
\quad\raisebox{17mm}{$\Rightarrow$}\quad
\includegraphics[scale=0.3]{crossing_octa_small.eps}
\caption{An octahedron put at a crossing}
\label{fig:crossing_octa}
\end{figure}
Then decompose the octahedron into five tetrahedra as in Figure~\ref{fig:octa_tetra}, where the four of them are decorated with $\xi_N^i$, $\xi_N^{-j}$, $\xi_N^k$ and $\xi_N^{-l}$, respectively and the one in the center is decorated with $\xi_N^m$, where $m:=l-i=j-k$.
Here each truncated tetrahedron is just a topological one with some decoration.
\begin{figure}[h]
\includegraphics[scale=0.3]{octa_small.eps}
\caption{decomposition of the octahedron into five tetrahedra}
\label{fig:octa_tetra}
\end{figure}
\par
Now only two of the vertices are attached to the knot.
We pull the two of the remaining four vertices to the top ($+\infty$) and the other two to the bottom ($-\infty$) as shown in Figure~\ref{fig:octa_tetra_infinity}.
\begin{figure}[h]
\includegraphics[scale=0.3]{crossing_octa_infinity_small.eps}
\caption{Pull the vertices to $+\infty$ and $-\infty$.}
\label{fig:octa_tetra_infinity}
\end{figure}
We attach five tetrahedra to every crossing (if the crossing is negative, we change them appropriately) in this way.
At each arc two faces meet, and we paste them together.
Thus we have a decomposition of $S^3\setminus(K\cup\{+\infty,-\infty\})$.
By deforming this decomposition a little we get a decomposition of $S^3\setminus{K}$ by (topological) truncated tetrahedra, decorated with complex numbers $\xi^{\pm i_k}_N$ (k=1,2,\dots,c).
\par
Next we want to regard each tetrahedron as an ideal hyperbolic one.
\par
Recall that when we approximate the summation in \eqref{eq:Jones_sum} by the integral in \eqref{eq:Jones_integral2} we replace $\xi_N^{i_k}$ with a complex variable $z_k$.
Following this we replace the decoration $\xi_N^{i_k}$ for a tetrahedron with a complex number $z_k$.
Then regard the tetrahedron decorated with $z_k$ as an ideal hyperbolic tetrahedron parametrized by $z_k$.
\par
So far this is just formal parametrizations.
We need to choose appropriate values for parameters so that the tetrahedra fit together to provide a complete hyperbolic structure to $S^3\setminus{K}$.
To do this we choose $z_1,z_2,\dots,z_c$ so that:
\begin{itemize}
\item
Around each edge several tetrahedra meet.
To make the knot complement hyperbolic, the sum of these dihedral angles should be $2\pi$,
\item
Even if the knot complement is hyperbolic, the structure may not be complete.
To make it complete, the parameters should be chosen as follows.
\par
Since we truncate the vertices of the tetrahedra, four small triangles appear at the places where the vertices were (see Figure~\ref{fig:tetrahedron_parameter} for the triangle associated with the vertex at infinity).
After pasting these triangles make a torus which can be regarded as the boundary of the regular neighborhood of the knot $K$.
Each triangle has a similarity structure provided by the parameter $z_k$.
We need to make this boundary torus Euclidean.
\end{itemize}
See \cite[Chapter~4]{Thurston:GT3M},\cite[\S~2]{Neumann/Zagier:TOPOL85} for more details.
\par
Surprisingly these conditions are the same as the system of equations \eqref{eq:differential} that we used in the steepest descent method!
Therefore we can expect that a solution $(x_1,x_2,\dots,x_c)$ to \eqref{eq:differential} gives the complete hyperbolic structure.
\par
Then, what does $V(x_1,x_2,\dots,x_c)=2\pi\sqrt{-1}\lim_{N\to\infty}\log\bigl(J_N(K,\xi_N)\bigr)/N$ mean?
\par
Recall the formula \eqref{eq:tetra_vol} and that $V(x_1,x_2,\dots,x_c)$ is a sum of dilogarithm functions and logarithm functions.
Using these facts we can prove
\begin{equation*}
\Im{V(x_1,x_2,\dots,x_c)}
=
\operatorname{Vol}(S^3\setminus{K}),
\end{equation*}
that is,
\begin{equation*}
\Im
\left(
2\pi\sqrt{-1}\lim_{N\to\infty}\frac{\log\bigl(J_N(K,\xi_N)\bigr)}{N}
\right)
=
\operatorname{Vol}(S^3\setminus{K}).
\end{equation*}
So we have proved
\begin{equation*}
2\pi\lim_{N\to\infty}\frac{\log|J_N(K,\xi_N)|}{N}
=
\operatorname{Vol}(S^3\setminus{K}),
\end{equation*}
which is the Volume Conjecture (Conjecture~\ref{conj:VC}).
\section{Generalizations of the Volume Conjecture}
\label{sec:generalization}
In this section we consider generalizations of the Volume Conjecture.
\subsection{Complexification}
In \cite{Thurston:BULAM31982} W.~Thurston pointed out that the Chern--Simons invariant \cite{Chern/Simons:ANNMA21974} can be regarded as an imaginary part of the volume.
Neumann and Zagier gave a precise conjecture \cite[Conjecture, p.~309]{Neumann/Zagier:TOPOL85} which was proved to be true by Yoshida \cite{Yoshida:INVEM85}.
For combinatorial approaches to the Chern--Simons invariant, see \cite{Neumann1992} and \cite{Zickert:DUKMJ2009}.
\par
So it would be natural to drop the absolute value sign of the left hand side of the Volume Conjecture and add the Chern--Simons invariant to the right hand side.
\begin{conjecture}[Complexification of the Volume Conjecture
\cite{Murakami/Murakami/Okamoto/Takata/Yokota:EXPMA02}]
\label{conj:CVC}
If a knot $K$ is hyperbolic, that is, its complement possesses a complete hyperbolic structure, then
\begin{multline*}
2\pi\lim_{N\to\infty}\frac{\log J_N(K;\exp(2\pi\sqrt{-1}/N)) }{N}
\equiv
\operatorname{Vol}(S^3\setminus{K})+\sqrt{-1}\operatorname{CS}(S^3\setminus{K})
\\
\pmod{\pi^2\sqrt{-1}\mathbb{Z}},
\end{multline*}
where $\operatorname{CS}$ is the Chern--Simons invariant defined for a three-manifold with torus boundary by Meyerhoff \cite{Meyerhoff:density}.
\end{conjecture}
\begin{remark}
We may regard the left hand side as a definition of the Chern--Simons invariant for non-hyperbolic knots provided that the limit of Conjecture~\ref{conj:CVC} exists.
\end{remark}
\subsection{Deformation of the parameter}
In the Volume Conjecture (Conjecture~\ref{conj:VC}) and its complexification (Conjecture~\ref{conj:CVC}), the (possible) limit corresponds to the complete hyperbolic structure of $S^3\setminus{K}$ for a hyperbolic knot $K$.
As described in \cite[Chapter~4]{Thurston:GT3M} the complete structure can be deformed to incomplete ones.
\par
How can we perform this deformation in the colored Jones polynomial?
If we deform the parameter $2\pi\sqrt{-1}$ in the Volume Conjecture, is the corresponding limits related to incomplete hyperbolic structures?
\par
Let us consider the limit
\begin{equation*}
\lim_{N\to\infty}\frac{\log J_N\Bigl(K;\exp\bigl((u+2\pi\sqrt{-1})/N\bigr)\Bigr) }{N}.
\end{equation*}
Note that when $u=0$, this limit is considered in the (complexified) Volume Conjecture.
\par
\subsubsection{Figure-eight knot}
Before stating a conjecture for general knots, I will explain what happens in the case of the figure-eight knot.
\begin{theorem}[\cite{Murakami/Yokota:JREIA2007}]\label{thm:MY}
Let $E$ be the figure-eight knot.
There exists a neighborhood $\mathcal{O}\subset\mathbb{C}$ of $0$ such that if $u\in(\mathcal{O}\setminus\pi\sqrt{-1}\mathbb{Q})\cup\{0\}$, then the following limit exists:
\begin{equation*}
\lim_{N\to\infty}
\frac{\log J_N(E;\exp\bigl((u+2\pi\sqrt{-1})/N\bigr))}{N}.
\end{equation*}
Moreover if we put
\begin{align*}
H(u)
&:=
(u+2\pi\sqrt{-1})\times(\text{the limit above}),
\\
\intertext{and}
v
&:=
2\dfrac{d\,H(u)}{d\,u}-2\pi\sqrt{-1}
\end{align*}
then we have
\begin{multline*}
\operatorname{Vol}(E_{u})+\sqrt{-1}\operatorname{CS}(E_{u})
\\
\equiv
-\sqrt{-1}H(u)
-\pi{u}
+u\,v\sqrt{-1}/4
-\pi\kappa(\gamma_u)/2
\pmod{\pi^2\sqrt{-1}\mathbb{Z}}.
\end{multline*}
\par
Here $E_{u}$ is the closed hyperbolic three-manifold associated with the following representation of $\pi_1\left(S^3\setminus{E}\right)\to SL(2;\mathbb{C})$:
\begin{equation}\label{eq:meridian_longitude}
\begin{split}
\mu
&\mapsto
\begin{pmatrix}\exp(u/2)&\ast\\0&\exp(-u/2)\end{pmatrix},
\\[5mm]
\lambda
&\mapsto
\begin{pmatrix}\exp(v/2)&\ast\\0&\exp(-v/2)\end{pmatrix}.
\end{split}
\end{equation}
Here $\mu$ is the meridian of $E$ $($a loop in $S^3\setminus{E}$ that goes around the knot, which generates $H_1(S^3\setminus{E})\cong\mathbb{Z}$$)$, $\lambda$ is the longitude $($a loop in $S^3\setminus{E}$ that goes along the knot such that it is homologous to $0$ in $H_1(S^3\setminus{E})$$)$, and $\gamma_u$ is the loop attached to $S^3\setminus{E}$ when we complete the hyperbolic structure defined by $u$.
We also put $\kappa(\gamma_{u}):=\operatorname{length}(\gamma_u)+\sqrt{-1}\operatorname{torsion}(\gamma_u)$, where $\operatorname{length}(\gamma_u)$ is the length of the attached loop $\gamma_u$, and $\operatorname{torsion}$ is its torsion, which is defined modulo $2\pi$ as the rotation angle when one travels along $\gamma_u$.
See \cite{Neumann/Zagier:TOPOL85} for details $($see also \cite{Murakami:FUNDM2004}$)$.
\end{theorem}
We will give a sketch of the proof in the following two subsections.
\subsubsection{Calculation of the limit}
First we calculate the limit.
Note that here I give just a sketch of the calculation but it can be done rigorously.
For details see \cite{Murakami/Yokota:JREIA2007}.
\par
From \eqref{eq:Jones_fig8} we have
\begin{equation*}
J_N(E;q)
=
\sum_{j=0}^{N-1}
q^{jN}
\prod_{k=1}^{j}
\left(1-q^{-N-k}\right)
\left(1-q^{-N+k}\right).
\end{equation*}
Put $q:=\exp(\theta/N)$ for $\theta$ near $2\pi\sqrt{-1}$.
If $\theta$ is not a rational multiple of $\pi\sqrt{-1}$ (but it can be $2\pi\sqrt{-1}$), we have
\begin{equation*}
\begin{split}
&\log
\left(
\prod_{k=1}^{j}
\left(1-q^{-N\pm k}\right)
\right)
\\
=&
\sum_{k=1}^{j}
\log
\big(1-\exp(\pm k\theta/N-\theta)\bigr)
\\
\underset{N\to\infty}{\approx}&
N
\int_{0}^{j/N}\log(1-\exp(\pm\theta s-\theta))\,ds
\\
=&
\frac{\pm N}{\theta}
\int_{\exp(-\theta)}^{\exp(\pm j\theta/N-\theta)}
\frac{\log(1-t)}{t}dt
\\
=&
\frac{\pm N}{\theta}
\left(
\operatorname{Li}_2\bigl(\exp(-\theta)\bigr)
-
\operatorname{Li}_2\bigl(\exp(\pm j\theta/N-\theta)\bigr)
\right).
\end{split}
\end{equation*}
So we have
\begin{equation*}
\begin{split}
&J_N\bigl(E;\exp(\theta/N)\bigr)
\\
\underset{N\to\infty}{\approx}&
\sum_{j=0}^{N-1}
\exp(j\theta)
\exp
\left[
\frac{N}{\theta}
\left(
\operatorname{Li}_2\bigl(\exp(-j\theta/N-\theta)\bigr)
-
\operatorname{Li}_2\bigl(\exp(j\theta/N-\theta)\bigr)
\right)
\right]
\\
=&
\sum_{j=0}^{N-1}
\exp
\left[
\frac{N}{\theta}
H\bigl(\exp(j\theta/N),\exp(\theta)\bigr)
\right]
\\
\underset{N\to\infty}{\approx}&
\int_{C}
\exp
\left[
\frac{N}{\theta}
H\bigl(x,\exp(\theta)\bigr)
\right]
\,dx
\end{split}
\end{equation*}
for a suitable contour $C$.
Here we put
\begin{equation}\label{eq:H}
H(\zeta,\eta)
:=
\operatorname{Li}_2(1/(\zeta\eta))-\operatorname{Li}_2(\zeta/\eta)+\log{\zeta}\log{\eta}.
\end{equation}
\par
To apply the steepest descent method (Theorem~\ref{thm:steepest_descent_method}), we find the maximum of $\Re\left(H\bigl(x,\exp(\theta)\bigr)/\theta\right)$ over $x$.
To do that we will find a solution $y$ to the equation $d\,H\bigl(x,\exp(\theta)\bigr)/d\,x=0$, which is
\begin{equation*}
\frac{\log\left[\exp(\theta)+\exp(-\theta)-x-x^{-1}\right]}{x}
=0.
\end{equation*}
We can show that appropriately chosen $y$ gives the maximum and from the steepest descent method we have
\begin{equation*}
J_N\bigl(E;\exp(\theta/N)\bigr)
\underset{N\to\infty}{\approx}
\exp
\left[
\frac{N}{\theta}
H\bigl(y,\exp(\theta)\bigr)
\right],
\end{equation*}
that is,
\begin{equation}\label{eq:GVC_limit}
\theta
\lim_{N\to\infty}
\frac{J_N\bigl(E;\exp(\theta/N)\bigr)}{N}
=
H\bigl(y,\exp(\theta)\bigr),
\end{equation}
where $y$ satisfies
\begin{equation*}
y+y^{-1}
=
\exp(\theta)+\exp(-\theta)-1.
\end{equation*}
\subsubsection{Calculation of the volume}
Next we will relate the limit to the volume of a three-manifold obtained by $S^3\setminus{E}$.
\par
As described in \S~\ref{subsubsec:geometry}, $S^3\setminus{E}$ is obtained by gluing two ideal hyperbolic tetrahedra as in Figure~\ref{fig:fig8_tetra}.
Here we assume that they are parametrized by complex numbers $z$ and $w$.
When $z=w=(1+\sqrt{-3})/2$, $S^3\setminus{E}$ has a complete hyperbolic structure as described in \S~\ref{subsubsec:geometry}.
We assume that the left tetrahedron (with faces labeled with $A$, $B$, $C$ and $D$) and the right tetrahedron (with faces labeled with $A'$, $B'$, $C'$ and $D'$) in Figure~\ref{fig:fig8_tetra} are $\Delta(z)$ and $\Delta(w)$ respectively.
\par
The boundary torus, which is obtained from the shadowed triangles in Figure~\ref{fig:fig8_tetra}, looks like Figure~\ref{fig:torus}.
Here the leftmost triangle is the one in the center of the picture of $\Delta(z)$ and the second leftmost one is the one in the center of the picture of $\Delta(w)$.
Let $\alpha$, $\beta$ and $\gamma$ be the dihedral angle between $B$ and $C$, $A$ and $B$, and $C$ and $A$ respectively.
Let $\alpha'$, $\beta'$ and $\gamma'$ be the dihedral angle between $B'$ and $D'$, $A'$ and $B'$, and $D'$ and $A'$ respectively.
\begin{figure}[h]
\includegraphics[scale=0.4]{torus_small.eps}
\caption{Identifying the sides as indicated by the circles,
we get a triangulation of the boundary torus.
Here the single circles denote the arrow head of the single arrow,
the double circles denote the arrow head of the double arrow,
the circles with $-$ denote the arrow tail of the single arrow,
and the circles with $+$ denote the arrow tail of the double arrow
in Figure~\ref{fig:fig8_tetra}.
Note that we view this torus from outside of $S^3\setminus{E}$.}
\label{fig:torus}
\end{figure}
As described in \S~\ref{subsubsec:geometry}, the sum of the dihedral angles around each edge should be $2\pi$.
So from Figure~\ref{fig:torus}, we have
\begin{align*}
\beta+2\gamma+\beta'+2\gamma'&=2\pi\quad\text{from the single circles},
\\
2\alpha+\beta+2\alpha'+\beta'&=2\pi\quad\text{from the double circles},
\\
\beta+2\gamma+\beta'+2\gamma'&=2\pi\quad\text{from the circles with $-$},
\\
2\alpha+\beta+2\alpha'+\beta'&=2\pi\quad\text{from the circles with $+$},
\end{align*}
which is equivalent to a single equation
\begin{equation*}
2\alpha+\beta+2\alpha'+\beta'=2\pi
\end{equation*}
since $\alpha+\beta+\gamma=\alpha'+\beta'+\gamma'=\pi$.
\par
Assume that $\alpha=\arg{z}$ and $\alpha'=\arg{w}$.
Since $\beta=\arg(1-1/z)$ and $\beta'=\arg(1-1/w)$ (see Figures~\ref{fig:top_view} and \ref{fig:tetrahedron_parameter}) this turns out to be
\begin{equation*}
2\arg{z}+\arg(1-1/z)+2\arg{w}+\arg(1-1/w)=2\pi.
\end{equation*}
So we have
\begin{equation}\label{eq:hyperbolicity}
zw(z-1)(w-1)=1.
\end{equation}
\begin{remark}
This is just a condition that $S^3\setminus{E}$ is hyperbolic.
To make the metric complete we need to add the condition that the upper side and the lower side of the parallelogram in Figure~\ref{fig:torus} are parallel.
\end{remark}
Now we introduce parameters $x$ and $y$ as
\begin{equation}\label{eq:xyzw}
\begin{split}
x&:=w(1-z),
\\
y&:=-zw.
\end{split}
\end{equation}
Note that the following equality holds from \eqref{eq:hyperbolicity}:
\begin{equation}\label{eq:xy}
y+y^{-1}=x+x^{-1}-1.
\end{equation}
\par
Since $\Delta(z)$ and $\Delta(w)$ can also be parametrized as $\Delta(1-1/z)$ and $\Delta(1-1/w)$, we have
\begin{multline*}
\operatorname{Vol}(\Delta(z))+\operatorname{Vol}(\Delta(w))
\\
=
\Im\operatorname{Li}_2(w')+\Im\operatorname{Li}_2(z')+\log|w'|\arg(1-w')+\log|z'|\arg(1-z'),
\end{multline*}
from \eqref{eq:tetra_vol}, where we put $z':=1-1/z$ and $w':=1-1/w$.
Using the equation
\begin{equation*}
\operatorname{Li}_2(z')+\operatorname{Li}_2({z'}^{-1})
=
-\frac{\pi^2}{6}-\frac{1}{2}\bigl(\log(-z')\bigr)^2
\end{equation*}
(see for example \cite[\S~1.2]{Zagier:2007}), we have
\begin{equation}\label{eq:Vol_H1}
\begin{split}
&\operatorname{Vol}(\Delta(z))+\operatorname{Vol}(\Delta(w))
\\
=&
\Im\operatorname{Li}_2(w')-\Im\operatorname{Li}_2({z'}^{-1})
+\log|w'|\arg(1-w')-\log|{z'}^{-1}|\arg(1-{z'}^{-1})
\\
&\text{(since $w'=1/(yx)$ and ${z'}^{-1}=y/x$)}
\\
=&
\Im\operatorname{Li}_2(1/(yx))-\Im\operatorname{Li}_2(y/x)
+\log|1/(yx)|\arg(1-1/(yx))-\log|y/x|\arg(1-y/x)
\\
&\text{(from \eqref{eq:xy})}
\\
=&
\Im\operatorname{Li}_2(1/(yx))-\Im\operatorname{Li}_2(y/x)
+\log|y|\arg{x}+\log|x|\arg\frac{1-y/x}{1-1/(yx)}
\\
&\text{(from \eqref{eq:xyzw} and \eqref{eq:hyperbolicity})}
\\
=&
\Im\operatorname{Li}_2(1/(yx))-\Im\operatorname{Li}_2(y/x)
+\log|y|\arg{x}+\log|x|\arg\frac{y}{z(z-1)}
\\
&\text{(from \eqref{eq:H})}
\\
=&
\Im H(y,x)-\log|x|\arg(z(z-1)).
\end{split}
\end{equation}
\par
Putting
\begin{align*}
u&:=\log{x}=\log(w(1-z)),
\\
v&:=2\log(z(1-z)),
\\
H(u)&:=H(y,x),
\end{align*}
we have
\begin{equation}\label{eq:Vol_H2}
\operatorname{Vol}(\Delta(z))+\operatorname{Vol}(\Delta(w))
=
\Im H(u)-\pi\Re{u}-\frac{1}{2}\Re{u}\Im{v}.
\end{equation}
Moreover from \eqref{eq:GVC_limit}, we have
\begin{equation*}
(u+2\pi\sqrt{-1})
\lim_{N\to\infty}
\frac{\log{J_N\bigl(E;\exp((u+2\pi\sqrt{-1})/N)\bigr)}}{N}
=
H(u)
\end{equation*}
if we put $\theta=u+2\pi\sqrt{-1}$ since $x=\exp(u)=\exp(\theta)$.
Note that $z$, $w$, $x$, $y$ and $v$ are functions of $u$.
Note also that $v$ is given as
\begin{equation*}
v
=
2\frac{d\,H(u)}{d\,u}-2\pi\sqrt{-1}
\end{equation*}
since $\dfrac{d\,H(u)}{d\,u}=\log(z(z-1))$.
\begin{remark}
We need to be more careful about the arguments of variables.
For details see \cite{Murakami/Yokota:JREIA2007}.
\end{remark}
\par
I will give geometrical interpretation of $u$ and $v$ to relate the term $\Re{u}\Im{v}$ in \eqref{eq:Vol_H2} to the length of $\gamma_u$.
\par
We first calculate $H_1(S^3\setminus{E})=H_1(\Delta(z)\cup\Delta(w))$.
Since the interiors of three-simplices do not matter to the first homology, we can calculate it from the boundary torus, the edges of $\Delta(z)$ and $\Delta(w)$, and the faces $A=A'$, $B=B'$, $C=C'$, and $D=D'$ (see Figures~\ref{fig:fig8_tetra} and \ref{fig:torus}).
From Figures~\ref{fig:fig8_tetra}, \ref{fig:torus} and \ref{fig:torus_edge} one reads
\begin{align*}
\partial A=\partial A'
&=
\raisebox{-1mm}{\rotatebox{45}{$\rightarrow$}}-e_{7}
+\raisebox{-1mm}{\rotatebox{45}{$\twoheadrightarrow$}}+e_{10}
-\raisebox{-1mm}{\rotatebox{45}{$\twoheadrightarrow$}}-e_{5},
\\
\partial B=\partial B'
&=
-\raisebox{-1mm}{\rotatebox{45}{$\rightarrow$}}-e_{11}
-\raisebox{-1mm}{\rotatebox{45}{$\twoheadrightarrow$}}+e_{6}
+\raisebox{-1mm}{\rotatebox{45}{$\twoheadrightarrow$}}-e_{9},
\\
\partial C=\partial C'
&=
\raisebox{-1mm}{\rotatebox{45}{$\twoheadrightarrow$}}+e_{4}
+\raisebox{-1mm}{\rotatebox{45}{$\rightarrow$}}+e_{8}
-\raisebox{-1mm}{\rotatebox{45}{$\rightarrow$}}+e_{1},
\\
\partial D=\partial D'
&=
\raisebox{-1mm}{\rotatebox{45}{$\rightarrow$}}+e_{3}
-\raisebox{-1mm}{\rotatebox{45}{$\twoheadrightarrow$}}+e_{2}
-\raisebox{-1mm}{\rotatebox{45}{$\rightarrow$}}+e_{12},
\end{align*}
where \raisebox{-1mm}{\rotatebox{45}{$\rightarrow$}} and \raisebox{-1mm}{\rotatebox{45}{$\twoheadrightarrow$}} mean the single arrowed edge and the double arrowed edge in Figure~\ref{fig:fig8_tetra} respectively, and the $e_i$ are the edges of the boundary torus as indicated in Figure~\ref{fig:torus_edge}.
Since
\begin{align*}
e_{7}&=e_{6}+e_{2},
\\
e_{10}&=e_{6},
\\
e_{5}&=e_{1}+e_{6},
\\
e_{11}&=e_{4}+e_{12}=e_{4}+e_{6},
\\
e_{9}&=e_{3}+e_{10}=e_{3}+e_{6},
\\
e_{8}&=e_{6},
\\
e_{12}&=e_{6}
\end{align*}
in the first homology group, we have
\begin{align*}
\raisebox{-1mm}{\rotatebox{45}{$\rightarrow$}}-e_1-e_2-e_6&=0,
\\
-\raisebox{-1mm}{\rotatebox{45}{$\rightarrow$}}-e_3-e_4-e_6&=0,
\\
\raisebox{-1mm}{\rotatebox{45}{$\twoheadrightarrow$}}+e_1+e_4+e_6&=0,
\\
-\raisebox{-1mm}{\rotatebox{45}{$\twoheadrightarrow$}}+e_2+e_3+e_6&=0.
\end{align*}
So if we put $\mu:=e_6$, $\lambda:=e_1+e_2+e_3+e_4+2\mu$, then we see that the first homology group of the boundary torus is generated by $\mu$ and $\lambda$, that $H_1(S^3\setminus{E})\cong\mathbb{Z}$ is generated by $\mu$, and that $\lambda=0$ in $H_1(S^3\setminus{E})$.
Therefore $\mu$ is the meridian and $\lambda$ is the longitude.
\begin{figure}[h]
\includegraphics[scale=0.4]{torus_edge_small.eps}
\caption{The cycle $e_6$ is the meridian and the cycle
$e_1+e_6+e_2+e_8+e_3+e_4$ is the longitude.}
\label{fig:torus_edge}
\end{figure}
\par
Now let us consider the universal cover $\widetilde{S^3\setminus{E}}$ of $S^3\setminus{E}=\Delta(z)\cup\Delta(w)$ which is $\H^3$.
We can construct it by developing $\Delta(z)$ and $\Delta(w)$ isometrically in $\H^3$.
Then each loop in $S^3\setminus{E}$ is regarded as a covering translation of $\widetilde{S^3\setminus{E}}$ and it defines an isometric translation of $\H^3$.
This defines a representation (holonomy representation) of $\pi_1\bigl(S^3\setminus{E})$ at $PSL(2;\mathbb{C})$.
Taking a lift to $SL(2;\mathbb{C})$, we can define a representation $\rho\colon\pi_1\bigl(S^3\setminus{E})\to SL(2;\mathbb{C})$.
\par
We consider how $\rho(\mu)$ and $\rho(\lambda)$ act on $\partial\H^3=S^2=\mathbb{C}\cup\{\infty\}$.
The image of the meridian $\rho(\mu)$ sends the top side to the bottom side.
So it is the composition of a $-\alpha$-rotation around the circle with $+$ in the top (between $e_1$ and $e_2$) and a $\gamma'$-rotation around the single circle in the bottom (between $e_2$ and $e_3$), which means a multiplication by $1/z\times1/(1-w))=w(1-z)$ plus a translation from \eqref{eq:hyperbolicity}.
Similarly $\rho(\lambda)$ acts as a multiplication by $z^2(1-z)^2$ plus a translation.
\par
Therefore $u=\log(w(1-z))$ and $v=2\log(z(1-z))$ can be regarded as the logarithms of the actions by the meridian $\mu$ and the longitude $\lambda$, respectively.
\par
Since the meridian and the longitude commute in $\pi_1(S^3\setminus{E})$, their images can be simultaneously triangularizable.
Recalling that $\mu$ and $\lambda$ define multiplications by $\exp(u)$ and $\exp(v)$ plus translations on $\partial\H^3$, we may assume
\begin{align*}
\rho(\mu)
&=
\begin{pmatrix}
\exp(u/2)&\ast \\
0 &\exp(-u/2)
\end{pmatrix},
\\
\rho(\lambda)
&=
\begin{pmatrix}
\exp(v/2)&\ast \\
0 &\exp(-v/2)
\end{pmatrix},
\end{align*}
which is \eqref{eq:meridian_longitude}.
This is a geometric interpretation of $u$ and $v$.
\par
Since $u$ determines $z$ and $w$, it defines a hyperbolic structure of $S^3\setminus{E}$ as the union of $\Delta(z)$ and $\Delta(w)$.
When $u\ne0$ this hyperbolic structure is incomplete.
We can complete this incomplete structure by attaching either a point or a circle.
\par
Since $v$ is not a real multiple of $u$ when $u$ is small, there exists a pair $(p,q)\in\mathbb{R}^2$ such that $pu+qv=2\pi\sqrt{-1}$.
The pair $(p,q)$ is called the generalized Dehn surgery coefficient \cite{Thurston:GT3M}.
If $p$ and $q$ are coprime integers, then the completion is given by attaching a circle $\gamma_{u}$ and the result is a closed hyperbolic three-manifold which we denote by $E_u$.
(For other cases the completion is given by adding either a point or a circle.
In the former case the regular neighborhood of the attached point is a cone over a torus, and in the latter case the regular neighborhood of the attached circle is topologically a solid torus but geometrically the angle around the core is not $2\pi$.)
\par
If $p$ and $q$ are coprime integers, the completion is nothing but the $(p,q)$-Dehn surgery along the knot, that is, we attach a solid torus $D$ to $S^3\setminus\operatorname{Int}(N(E))$ so that the meridian of $D$ coincides with the loop on the boundary of the regular neighborhood $N(E)$ of $E\subset S^3$ presenting $p\mu+q\lambda\in H_1(S^3\setminus\operatorname{Int}(N(E)))$, where $\operatorname{Int}$ denotes the interior (Figure~\ref{fig:Dehn_surgery}).
\begin{figure}[h]
\includegraphics[scale=0.3]{fig8_surgery1_small.eps}
\raisebox{20mm}{\quad$\cup$\quad}
\raisebox{4mm}{\includegraphics[scale=0.3]{fig8_surgery2_small.eps}}
\caption{$(2,1)$-Dehn surgery along the figure-eight knot}
\label{fig:Dehn_surgery}
\end{figure}
Then the circle $\gamma_u$ can be regarded as the core of $D$.
\par
To complete the proof of Theorem~\ref{thm:MY}, we want to describe the length of the attached circle $\gamma_u$ in terms of $u$ and $v$.
We will show
\begin{equation}\label{eq:length}
\operatorname{length}{\gamma}_{u}
=
-\frac{1}{2\pi}\Im\left(u\overline{v}\right).
\end{equation}
\par
When $u$ is small and non-zero, we can assume that $\exp(u)\ne1$ and $\exp(v)\ne1$.
So we can also assume that $\rho(\mu)$ and $\rho(\lambda)$ are both diagonal.
This means that the image of $\mu$ is a multiplication by $\exp(u)$ and that the image of $\lambda$ is a multiplication by $\exp(v)$ (with no translations).
Note that now $\widetilde{S^3\setminus{E}}$ is identified with $\H^3$ minus the $z$-axis, and the completion is given by adding the $z$-axis.
\par
Since $p$ and $q$ are coprime, we can choose integers $r$ and $s$ so that $ps-qr=1$.
We push $\gamma_u\in D$ to the boundary of the solid torus $\partial D$ and denote the resulting circle by $\tilde{\gamma}_u$.
Then we see that $[\tilde{\gamma}_u]=r\mu+s\lambda\in H_1(\partial(S^3\setminus\operatorname{Int}(N(E)));\mathbb{Z})$ since the meridian of $D$ is identified with $p\mu+q\lambda$, and the images of the meridian and $\tilde{\gamma}_u$ make a basis of $H_1(\partial(S^3\setminus\operatorname{Int}(N(E)));\mathbb{Z})$.
\begin{remark}
Even if we use another pair $(r',s')$ such that $ps'-qr'=1$, we get the same manifold.
This is because changing $(r,s)$ corresponds to changing of $\tilde{\gamma}_u\in\partial D$.
Observe that ambiguity of the choice of $\tilde{\gamma}_u$ is given by a twist of $D$ and that it does not matter to the resulting manifold.
\end{remark}
\par
Therefore $\rho(\gamma_u)$ corresponds to a multiplication by $\exp(\pm(ru+sv))$.
This means that if we identify the completion of $\widetilde{S^3\setminus{E}}$ with $\H^3$, a fundamental domain of the lift of $\gamma_u$ is identified with the segment $[1,\exp\bigl(\pm\Re(ru+sv)\bigr)]$ in the $z$-axis.
Since the metric is given by $\sqrt{dx^2+dy^2+dz^2}/z$, the length of $\gamma_u$ is given by
\begin{equation*}
\begin{split}
\operatorname{length}(\gamma_u)
&=
\int_{1}^{\exp\bigl(\pm\Re(ru+sv)\bigr)}\frac{dz}{z}
\\
&=
\pm\Re(ru+sv)
\\
&=
\pm\left(\frac{ps-1}{q}\Re(u)+s\Re(v)\right)
\\
&=
\pm\left(\frac{s}{q}(p\Re(u)+q\Re(v))-\frac{1}{q}\Re(u)\right)
\\
&=
\mp\frac{\Re(u)}{q}
\\
&=
\frac{\mp1}{2\pi}\left(\Re{u}\Im{v}-\Im{u}\Re{v}\right)
\\
&=
\frac{\pm1}{2\pi}\Im(u\overline{v}),
\end{split}
\end{equation*}
where the fourth and the sixth equalities follow from
\begin{equation*}
\begin{pmatrix}
\Re(u)&\Re(v) \\
\Im(u)&\Im(v)
\end{pmatrix}
\begin{pmatrix}
p\\q
\end{pmatrix}
=
\begin{pmatrix}
0\\2\pi
\end{pmatrix}.
\end{equation*}
Since $v=u\times\dfrac{|v|^2}{u\overline{v}}$ and the orientation of $(u,v)$ should be positive on $\mathbb{C}$, we see that $\Im(u\overline{v})$ is negative (see \cite{Neumann/Zagier:TOPOL85} for details) and so \eqref{eq:length} follows.
\par
Therefore from \eqref{eq:Vol_H2} we finally have
\begin{equation*}
\operatorname{Vol}(E_u)
=
\Re
\left(
-\sqrt{-1}H(u)
-\pi u
+\frac{1}{4}uv\sqrt{-1}
-\frac{\pi}{2}\kappa(\gamma_u)
\right).
\end{equation*}
\par
The Chern--Simons invariant is obtained by Yoshida's formula \cite{Yoshida:INVEM85}.
See \cite{Murakami/Yokota:JREIA2007} for details.
\subsubsection{General knots}
Here I propose a generalization of the Volume Conjecture for general knots.
\begin{conjecture}[\cite{Murakami:ADVAM22007}]\label{conj:PVC}
For any knot $K$, there exists an open set $U\in\mathbb{C}$ such that if $u\in U$, then the following limit exists:
\begin{equation*}
\lim_{N\to\infty}
\frac{\log J_N(K;\exp\bigl((u+2\pi\sqrt{-1})/N\bigr))}{N}.
\end{equation*}
Moreover if we put
\begin{align*}
H(K;u)&:=(u+2\pi\sqrt{-1})\times(\text{the limit above}),
\\
\intertext{and}
v&:=2\dfrac{d\,H(K;u)}{d\,u}-2\pi\sqrt{-1},
\end{align*}
then we have
\begin{equation*}
\operatorname{Vol}(K;u)
=
\Im{H(K;u)}
-\pi\Re{u}
-\frac{1}{2}\Re{u}\Im{v}.
\end{equation*}
Here $\operatorname{Vol}(K;u)$ is the volume function corresponding to the representation of $\pi_1(S^3\setminus{K})$ to $SL(2;\mathbb{C})$ as in Theorem~$\ref{thm:MY}$.
\end{conjecture}
\begin{remark}
In the case of a hyperbolic knot, we can also propose a similar conjecture with the imaginary part as in the case of the figure-eight knot.
For a general knot, a relation to the Chern--Simons invariant is also expected by using a combinatorial description of the Chern--Simons invariant by Zickert \cite{Zickert:DUKMJ2009}.
\end{remark}
\begin{remark}
Conjecture~\ref{conj:PVC} is known to be true for the figure-eight knot \cite{Murakami/Yokota:JREIA2007} and for torus knots \cite{Murakami:ADVAM22007}.
See also \cite{Murakami:ACTMV2008} and \cite{Hikami/Murakami:Bonn} for a possible relation to the Chern--Simons invariant.
\end{remark}
Finally note that Garoufalidis and L{\^e} proved the following result, which should be compared with Conjecture~\ref{conj:PVC}.
(See also \cite{Murakami:JPJGT2007} for the case of the figure-eight knot.)
\begin{theorem}[S.~Garoufalidis and T.~L{\^e} \cite{Garoufalidis/Le:aMMR}]
For any $K$, there exists $\varepsilon>0$ such that if $|\theta|<\varepsilon$, we have
\begin{equation*}
\lim_{N\to\infty}
J_N(K;\exp(\theta/N))
=
\frac{1}{\Delta(K;\exp{\theta})},
\end{equation*}
where $\Delta(K;t)$ is the Alexander polynomial of $K$.
\end{theorem}
|
\section{Introduction}
The Milky Way contains a supermassive black hole at its very center,
Sagittarius A*, whose mass is $\approx4\times10^{6}$\,M$_{\odot}$ \citep[e.g.][]{mdavies_schodel03}.
A dense stellar cluster surrounds this black hole, with a central density at least comparable
to that seen in the cores of the densest globular clusters. At least one, and possibly two, discs
containing young stars at distances between $\sim0.04$ and $\sim0.3$\,pc from the black hole
are also seen \citep{mdavies_paumard06}. The latter stellar population is thought to have an unusually
flat IMF \citep{mdavies_paumard06}. It has long been known that the central 0.2\,pc or so of the Galactic
center is deficient in bright red giants \citep{mdavies_genzel96}. Since the center of the Galaxy
has a high number density of stars, it is natural to suggest that stellar collisions may explain
the observed depletion.
A recent crop of papers have studied the stellar population in the central
regions \citep{mdavies_bartko10, mdavies_buchholz09, mdavies_do09}. They report
that the early-type stars (bright main-sequence stars) follow a cusp-like
profile whilst the surface density of late-type (red-giant) stars that are
brighter than a K magnitude of 15.5 is rather flat out to 0.4\,pc or 10\,arcsec.
This surprising result implies that the red-giant population is depleted out to
about 0.4\,pc from the central supermassive black hole. We will consider here
whether stellar collisions could be responsible for this observed depletion. We
will consider two different cases: 1) where the stellar population in the
Galactic center is drawn from a Miller-Scalo IMF, and 2) where the IMF is much
flatter. In each case, we calculate collision probabilities between the various
stellar species, and produce, via Monte Carlo techniques, a stellar population.
From this we measure the surface density of the early and late-type stars (to
compare directly with what is observed) making some reasonable assumptions
concerning the effects of collisions on the population.
The masses of stars contributing to the observed early and late-type populations
are rather different, as shown in Fig.~\ref{mdavies_figure1}, where we have
produced a synthetic population with a flat IMF. We see that the early-type
stars are virtually all between 12 and 27\,M$_{\odot}$ whereas stars
contributing to the observed late-type population have much lower masses,
between 1 and 5\,M$_{\odot}$.
\begin{figure}[!ht]
\plotfiddle{mdavies_figure1.eps}{8truecm}{0.0}{70}{70}{-100}{20}
\caption{A plot of the distribution of stellar masses contributing to the
observed early-type (between 12 and 27\,M$_{\odot}$) and late-type stars
(between 1 and 5\,M$_{\odot}$) at the Galactic center, assuming the stellar
population is drawn from a flat IMF ($\Gamma= 1.0$).}
\label{mdavies_figure1}
\end{figure}
\section{The Effects of Crowdedness}
The Galactic center is a crowded environment, with number densities of stars probably
of order $10^6$\,stars\,pc$^{-3}$ or more. In such dense environments, collisions between
stars will be frequent, and thus affect the stellar population.
The cross section for two stars, having a relative velocity at infinity
of $V_\infty$, to pass within a distance $R_{\rm min}$ is given by
\begin{equation}
\sigma = \pi R_{\rm min}^2 \left( 1 + {V^2 \over V_\infty^2} \right)
\end{equation}
\noindent where $V$ is the relative velocity of the two
stars at closest approach in a parabolic encounter ({\it i.e.}\ $V^2 = 2 G (M_1 + M_2)/R_{\rm min}$, where $M_1$ and $M_2$ are the masses
of the two stars).
The second term is due to the attractive gravitational force between the
two stars, and
is referred to as the gravitational focusing term. In the very center of the Galaxy, where
the supermassive black hole dominates, and stars may be assumed to
move on Keplerian orbits around the black hole at speeds exceeding
1000\,km/s, $V \ll V_\infty$ and we recover
the result, $\sigma \propto R_{\rm min}^2$. In this regime collisions involving
larger red giants will be relatively more frequent despite their short
lifetimes compared to main-sequence stars. This is not the case in globular clusters where
$V \gg V_\infty$ and thus $\sigma \propto R_{\rm min}$.
One may estimate the timescale for a given star to undergo an encounter
with another star,
$\tau_{\rm coll} = 1/n \sigma v$. The collision timescale will therefore be a
function of both the number density of stars and the makeup of the stellar
population, {\it i.e.}\ the distribution of stellar masses and types.
The effects of collisions will depend on the types of stars involved and on the
relative speed of the two stars when they collide. Collisions involving two
main-sequence stars occurring at relatively low speed (less than the surface
escape speed of the stars) are likely to result in the merger of the two objects
with relatively low amounts of mass loss. Collisions occurring at much higher
speeds are likely to lead to significant mass loss and even to the destruction
of the stars involved.
\begin{figure}[!ht]
\plotone{mdavies_figure2.eps}
\caption{A series of snap-shots of a collision between a main-sequence star and a stellar-mass
black hole using our SPH code. The position of the black hole is given by the black filled circle
and we show only 1\% of the SPH particles.}
\label{mdavies_figure2}
\end{figure}
Collisions between main-sequence stars and compact objects ({\it i.e.}\ black
holes, white dwarfs or neutron stars) are likely to be destructive (Dale et al.,
in preparation). In the case of a black hole impactor, the main-sequence star
is often torn apart by tidal forces, with some material being accreted by the
black hole and rest being ejected. This is illustrated in
Fig.~\ref{mdavies_figure2} where we show snapshots of a collision between a
1\,M$_\odot$ main-sequence star and a 10\,M$_\odot$ black hole. The close
passage of the black hole results in the tidal disruption of the main-sequence
star. A small fraction of the material is accreted by the black hole and the
rest is dispersed. For neutron star and white dwarf impactors, a larger
fraction of the material from the main-sequence star may form an envelope around
the compact object. However such an object is likely to be relatively short
lived, perhaps appearing as a bright red supergiant, before the envelope is
ejected. Thus all collisions between main-sequence stars and compact objects
will act to reduce the population of luminous stars.
\begin{figure}[!ht]
\plotone{mdavies_figure3.eps}
\caption{A series of snap-shots of a collision between a red giant and a stellar-mass
black hole using our SPH code. The black hole is labelled with a black filled circle
and the red-giant core is labelled with an X in the center of an open circle. For clarity in the
plot, only 1\% of the SPH particles are shown.}
\label{mdavies_figure3}
\end{figure}
Even though they are very frequent, collisions
between red giants and main-sequence stars in fact have very little effect, as the main-sequence
star passes through the envelope and effects very little mass loss \citep{mdavies_bailey99}.
The same is also true
for white dwarf or neutron star impactors.
More interesting are encounters involving black holes.
In Fig.~\ref{mdavies_figure3} we show snapshots of an encounter between a 1\,M$_\odot$ red giant and
and 10\,M$_\odot$ black hole ($V_\infty = 800$ km/s, $r_{\rm min} = 10$ R$_\odot$). As the black hole
passes close to the red-giant core, it gives it a jerk and the core is ejected at high speed,
retaining only a small fraction of the envelope (in this case about 13\%). Such excessive mass
loss will prevent this red giant from becoming brighter: we have thus removed it from the pool
of brighter red giants.
Thus we see that in order to deplete the stellar population of red giants (as indicated
by the observations), we must consider collisions of three types: 1) encounters between
red giants and black holes, 2) encounters between two main-sequence stars, and 3)
encounters between main-sequence stars and compact objects
({\it i.e.}\ black holes, white dwarfs and neutron stars).
The relative frequencies of encounters will depend on the IMF of the underlying stellar
population. With a Miller-Scalo type IMF, where most stars are of low mass, the majority
of stars formed will still be on the main sequence today, with relatively few stars
having evolved off the main-sequence to ultimately form compact remnants.
Thus collisions involving two main-sequence
stars will be more frequent than collisions between main-sequence stars and compact
objects. Collisions between red giants and black holes may be relatively frequent, at least
in the very center, owing to the larger size (and thus collisional cross section) of red giants.
\section{Collisions for a Miller-Scalo IMF}
\begin{figure}[!ht]
\plotfiddle{mdavies_figure4.eps}{8.4truecm}{0.0}{70}{70}{-100}{20}
\caption{The collision rates between the various stellar species.
Here the collision probability is the expected collision rate integrated
over the entire lifetime of the main-sequence or red-giant phase. In other
words, the star has a good chance of being involved in a collision if the collision
probability is one.
The gray band
is for encounters between red giants and black holes, the vertically-shaded region
is for encounters between main-sequence stars and compact objects,
and the horizontally-shaded region is for encounters between two main-sequence stars.
In all cases, we consider stars between 1.0\,M$_\odot$ and 5.0\,M$_\odot$.
The stellar population has a Miller-Scalo IMF.}
\label{mdavies_figure4}
\end{figure}
We consider first that the stellar population in the Galactic center has been drawn from
a Miller-Scalo IMF \citep{mdavies_miller79} where we also assume
the stars have a uniform spread in ages from 0 to 14\,Gyr. Using a density distribution for the stars given
by $n_{\rm *}(r)\propto r^{-1.4}$ and a density distribution for 10\,M$_\odot$
black holes of $n_{\rm BH}\propto r^{-1.8}$ \citep{mdavies_freitag06}, we compute the collision probabilities between the various
stellar types as a function of Galactocentric radius. The results are shown in Fig.~\ref{mdavies_figure4}. We see that collisions
between red giants and black holes are the most frequent in the very central regions
whilst collisions between two main-sequence stars dominate further out.
For stars following a Miller-Scalo IMF,
collisions between main-sequence stars and black holes (or other compact remnants) are
relatively less frequent and are unlikely to have a significant effect on the stellar population
apart from perhaps in the very central regions.
Even though collisions between
main-sequence stars are relatively frequent, in the case of a Miller-Scalo IMF
with a low-mass cutoff of 0.2\,M$_\odot$ the number of main-sequence stars in the mass
range 1--4\,M$_\odot$ will not be reduced. The reason is that although some stars in this
mass range will be removed as they merge with other stars to produce more massive
stars, they will be replaced by stars produced by the merger of lower-mass stars. This process will be investigated in detail in Dale et al.~(in preparation). It could be that a stellar population
with a Miller-Scalo IMF but with a larger low-mass cutoff may lead to a depletion of the lower-mass
main-sequence stars which later evolve to produce the observed population of red giants.
Collisions between red giants and stellar-mass black holes
are able to deplete the red-giant population within
the inner 0.08\,pc or so but not further out. Collisions can plausibly explain the depletion of
red giants of middle brightness ($10.5 < K < 12$) but not those in the brightest band
($K > 10.5$) which are seen to be depleted to about 0.2\,pc \citep{mdavies_genzel03, mdavies_dale09}.
Collisions are not able to explain the observed flattening of the red-giant population
seen out to 0.4\,pc \citep{mdavies_bartko10,mdavies_buchholz09}.
\section{Collisions for a Flat IMF}
In this section we consider the case where the stellar population in the Galactic center
is drawn from a much flatter IMF than a Miller-Scalo IMF. Observations of the young disc
of stars in the Galactic center suggest that they may have been drawn from such
an IMF with a power-low slope as low as $\Gamma = 0.45$ \citep{mdavies_bartko10}. A stellar population
drawn from such a flat IMF would be quite different from one drawn from a Miller-Scalo IMF.
A much larger fraction of stars would be massive, and explode as core-collapse supernovae
producing either black holes or neutron stars. In addition, assuming stars have an approximately
uniform spread in ages, a very large fraction of all stars will have evolved to become compact
remnants. ({\it i.e.}~black holes, white dwarfs, or neutron stars). In other words, the majority of stellar
objects will in fact be compact objects. The most common flavor of collision will then be
collisions involving compact objects.
\begin{figure}[!ht]
\plotfiddle{mdavies_figure5.eps}{8.4truecm}{0.0}{70}{70}{-100}{20}
\caption{The collision rates between the various stellar species.
Here the collision probability is the expected collision rate integrated
over the entire lifetime of the main-sequence or red-giant phase. In other
words, the star has a good chance of being involved in a collision if the collision
probability is one.
The gray band
is for encounters between red giants and black holes, the vertically-shaded region
is for encounters between main-sequence stars and compact objects,
and the horizontally-shaded region is for encounters between two main-sequence stars.
In all cases, we consider stars between 1.0\,M$_\odot$ and 5.0\,M$_\odot$.
The stellar population has a flat IMF ($\Gamma = 1.0$).}
\label{mdavies_figure5}
\end{figure}
In order to produce a stellar population, we assume a model where a stellar disc, similar
to the one seen today, is produced about every $3 \times 10^{7}$ years up to 14\,Gyr ago, with the
most recent disc being produced only 6 million years ago to be consistent
with observations.
The disc mass is set to match the observations of the early-type stars which come entirely
from the most-recent disc, whilst the total number of discs produced is set to match
the observed late-type population 10\,arcsec from the supermassive black hole. One should
note that the formation mechanism for such discs of stars is unclear, one suggestion being that
gas clouds interact tidally with the supermassive black hole \citep{mdavies_bonnell08}.
We consider here for purposes
of illustration the case of $\Gamma = 1.0$. For this IMF (which is flat in log mass),
the mass is each disc is
$10^{4}$\,M$_\odot$, with the stars having a range of masses between 1.0 and 32.5\,M$_\odot$.
It is entirely possible that the star formation rate was higher in the past.
As a simple first model, we assume that in addition to the discs produced at a steady rate,
there was an excess of discs produced sufficiently early in the history of the Galaxy that essentially
all the stars from these earlier discs have evolved into compact remnants. The exact excess
of compact objects will depend on the history of the disc formation rate,
but here we assume that two thirds of all star formation occurred at this earlier epoch.
This will increase the number of compact objects present in the Galactic center population.
As collisions involving compact remnants destroy main-sequence stars,
increasing the compact object population will enhance the red-giant depletion.
The collision probability for collisions between stars of various types for the stellar
population produced as described above are shown in Fig.~\ref{mdavies_figure5}.
We see here that collisions between main-sequence
stars and compact objects are most likely (at least further out from the center).
Main-sequence stars are likely to undergo such a collision, and thus be destroyed,
out to a distance of about 0.2\,pc.
As discussed earlier, collisions between compact objects
and main-sequence stars are likely to be destructive.
Therefore with a flat IMF, and a suitably
dense stellar population, the low-mass main-sequence
stars which would ordinarily have evolved to become
the observed late-type population are depleted within a distance of about 0.2\,pc of the supermassive
black hole. Therefore a stellar population produced from a flat IMF may be able to explain
the flat surface density profile which has been observed for the late-type stars (red giants) as
stars in the mass range 1-5\,M$_\odot$ within 0.2\,pc have simply been destroyed before
they could evolve into red giants.
\section{Discussion}
Equipped with the collision probabilities as a function of radius, we are now able
to synthesize a stellar population for both Miller-Scalo and flat IMFs allowing for
the collisional depletion of the red-giant population. In Fig.~\ref{mdavies_figure6} we show the
calculated surface density profiles for early-type stars
and late-type stars having K magnitudes brighter than 15.5 for both IMFs.
In both cases, the stellar population has been normalised to the surface brightness
of late-type stars 10\,arcsec from the Galactic center, seen to be about 3 sources per square arcsec
(this normalization was also applied to the collision probability calculations and figures
shown earlier).
In both plots we do not include the effect of stellar collisions on the population of
early-type stars as the collisional depletion of these massive stars will be very small.
However it should be noted that the existence of the so-called S stars in the very center
of the galaxy (distances less than 1 arcsecond from the central black hole) may place
limits on the number density of compact objects.
The plots shown in Fig.~\ref{mdavies_figure6}
should be compared to Fig.~11 of \citet{mdavies_buchholz09}.
One can see from the plots that the early-type population only matches the observations
in the case of the flat IMF. A Miller-Scalo population is clearly not able to consistently match
both the observed early and late populations.
It is also clear
from these plots that in the case of the Miller-Scalo IMF, the effect of stellar collisions on the red-giant population
is negligible. However for the flat IMF, appreciable red-giant depletion occurs out to 10\,arcseconds.
Thus a flat IMF would seem to have the potential to explain the observed (flat) surface
density profile of the red giants.
\begin{figure}[!ht]
\plottwo{mdavies_figure6a.eps}{mdavies_figure6b.eps}
\caption{The surface density of early and late-type stars and the depletion due stellar collisions
for a) Miller-Scalo IMF, and b) a flat IMF ($\Gamma=1.0$). In both cases, the top line is for late-type stars (without
allowing for destructive collisions), the middle line is for late-type stars (allowing for destructive collisions), and the bottom line is for the early-type stars. Note that these plots are the results of single Monte Carlo realisations, so that each line has an associated uncertainty.}
\label{mdavies_figure6}
\end{figure}
One problem with the model used here is that the total
stellar mass out to 0.4\,pc is about $5 \times 10^6$\,M$_\odot$. This is much higher
than suggested by observations \citep{mdavies_schodel03}. However, in our calculations
using a flat IMF here
we have made the unrealistic assumption that stars remain at the radius at which they
were formed. In other words, we have not allowed for the effects of mass segregation.
In reality, the large population of
heavier stellar-mass black holes will sink in the potential,
forming a central sub-cluster. This sub-cluster may then collapse to form a supermassive black hole in the manner
envisaged by \cite{mdavies_quinlan87} if a supermassive black hole is not there already,
or the stellar-mass black holes may simply be fed into an existing supermassive black hole.
In either case, a large fraction of the stellar-mass black holes may end up inside the supermassive
black hole. It is interesting in this context to note that the total mass contained in stellar-mass
black holes for our flat IMF (about $3.5 \times 10^6$\,M$_\odot$) is close to the observed
value for the supermassive black-hole mass.
A larger population of black holes produced by a flat IMF would also lead
to a much higher rate of EMRI--type events, where stellar-mass black holes spiral in to the
central, supermassive black hole, emitting gravitational radiation in the process. EMRIs
could be an important source for LISA.
Our calculations here have been a simplification. We have made the unrealistic
assumption that stars remain at the radius they were formed. In addition we have not
allowed for the growth of the supermassive black hole. Early on, before the
supermassive black hole has grown, the disc-mode of star formation may be different or not
occur at all owing to a lack of a supermassive black hole. A natural next step
in these calculations will be inclusion of the dynamical evolution of the stellar
cluster and for
the growth of the central black hole (Nzoke et al., in preparation).
\acknowledgements RPC acknowledges support from the Wenner Gren Foundation.
JED is supported by a Marie Curie fellowship as part of the European Commission FP6 Research Training Network `Constellation' under contract MRTN--CT--2006--035890 and the Institutional Research Plan AV0Z10030501 of the Academy of Sciences of the Czech Republic and project LC06014--Centre for Theoretical Astrophysics of the Ministry of Education, Youth and Sports of the Czech Republic.
|
\section{Introduction}
Details of MHD wave propagation inside magnetized regions are very
important for understanding of interaction, scattering, and
conversion of seismic waves by sunspots in the Sun. The structure of
sunspots themselves is not well understood. \citet{Pizzo1986}
proposed a self consistent magnetostatic sunspot model. Later
\citet{Low1975} presented a self-similar model. These two models
(model of Pizzo at the top and model of Low at the bottom) were
joined by \citet{Khomenko2008}. It is well understood now that the
sunspot is dynamic, so several attempts have been made to build
numerically a stable sunspot model with the background flows
\citep{Hurlburt2000,Botha2008}. Recently \citet{Rempel2009} obtained
realistically looking penumbral structures with outflows.
Local helioseismology provides a tool for reconstruction of the
internal structure (both profiles of wave speed and flows) of
sunspots from Doppler observations. The magnetic field of sunspots
affect the result of the helioseismic inversion. For understanding
the helioseismic effects of the magnetic field it is important to
perform direct simulations of MHD waves in different models of
sunspots. The simulations are also used for producing artificial
data for calibration and testing of helioseismic measurements and
inversion algorithms. There is wide gallery of numerical simulations
of propagation of MHD waves inside sunspots illustrating various
observed phenomena: power deficit of $p$-modes, acoustic halos,
azimuthal phase shift variations of the signal in the sunspot
penumbra, and so on. Not all of these phenomena are caused by the
direct effects of magnetic fields. Indirect effects of magnetic
fields must be also taken into account. For instance,
\citet{Parchevsky2007b} showed that 50\% of the observed acoustic
power deficit in sunspots can be explained by the absence of the
acoustic sources inside sunspots, where strong magnetic field
inhibits convective motions, which are the primary source of solar
oscillations. The suppression of acoustic sources is one of the most
important indirect effects of the solar magnetoseismology. Another
example of the indirect effects are changes in the density and
temperature stratifications caused by magnetic fields. Among the
direct effects of magnetic field (effects caused by magnetic
stresses on wave perturbations) an important role is played by
conversion into different types of MHD waves. For instance, 2D
simulations of wave propagation in inclined magnetic fields (e.g
\citet{Cally1997,Cally2000,Spruit1992}) showed that fast MHD waves
can be converted into slow MHD waves, which can leave the
computational domain, and thus interpreted as an "absorption" of the
fast mode. Acoustic halos around sunspots were obtained in
simulations by \citet{Hanasoge2008} as a result of wave
transformations. However, \citet{Jacoutot2008} noted that this
effect can be explained by changes in the excitation properties of
solar convection in moderate field strength regions.
\citet{Cameron2008} obtained the peak restrictions of 3 kG for the
photospheric magnetic field strength from simulations of scattering
$f$-modes by the sunspot. Three dimensional simulations of linear
MHD wave propagation performed by \citet{Parchevsky2009} show that
the inclined magnetic field can be partly responsible for the
azimuthal variations of travel times around sunspots found by
\citet{Zhao2006} from helioseismic observations. Refraction of
upward propagating MHD waves in the solar atmosphere was simulated
by e.g. \citet{Khomenko2006,Cally2008}. Recently,
\citet{Khomenko2009} carried out 2D simulations of the interaction
of MHD waves with magnetostatic sunspot models. In this paper we
present initial results of our 3D modeling of this problem.
Results of linear numerical simulations of MHD waves propagation in
sunspots mainly depend on the setup of the problem: 2D or 3D
simulations, choice of the background model of the sunspot, the way
of excitation of the waves, and treating of the boundary conditions.
The goal of this paper is to study and compare properties of 3D
propagation of MHD waves in different magnetostatic models of
sunspots, and also describe how different types of waves are
affected by the sunspot. Waves are generated by localized
subphotospheric (depth of 100 km) sources of the vertical force. For
such sources simulated waves propagate along the same ray paths as
in the Sun. The numerical method and the background sunspot models
are described in \S2 and \S3. The results of simulations are
discussed in \S4.
\section{Governing equations}
Propagation of MHD waves inside the Sun is described by the
following system of linearized equations:
\begin{equation}\label{Eq:MHD_3D}
\begin{array}{l}
\displaystyle \ddt{\rho'} + \nabla\cdot \mi{m}'=0,\\[12pt]
\displaystyle \ddt{\mi{m}'} + \nabla p' - \frac{1}{4\pi}\left[
(\nabla\times\mi{B}')\times\mi{B}_0 +
(\nabla\times\mi{B}_0)\times\mi{B}'\right] = \rho'\mi{g}_0 + \mi{S}(r,t),\\[12pt]
\displaystyle
\ddt{\mi{B}'}=\nabla\times\left(\frac{\mi{m}'}{\rho_0}\times\mi{B}_0\right),\\[12pt]
\displaystyle \ddt{p'} + c_{s0}^2\nabla\cdot \mi{m}' +
c_{s0}^2\frac{\mathcal{N}^{\,2}_0}{g_0}m_z = 0,
\end{array}
\end{equation}
where $\mi{m}'=\rho_0\mi{v}'$ is the momentum perturbation,
$\mi{v}'$, $\rho'$, $p'$, and $\mi{B}'$ are the velocity, density,
pressure, and magnetic field perturbations respectively,
$\mi{S}(r,t)=(0,0,S_z(r,t))^T$ is the wave source function. The
quantities with subscript 0, such as gravity $\mi{g}_0$, sound speed
$c_{s0}$, and Br\"unt-V\"ais\"al\"a frequency $\mathcal{N}_0$
correspond to the background model, $\mi{B}_0$ is the background
magnetic field satisfying the usual magnetohydrostatic equilibrium
equation. The spatial and temporal behavior of the wave source is
modeled by function $S_z(r,t)= A H(r)F(t)$:
\begin{eqnarray}
H(r) &=& \left\{
\begin{array}{ll}
\displaystyle
\left(1-\frac{r^2}{R_{src}^2}\right)^2 & \mbox{if } r\leq R_{src}\\
\displaystyle 0 & \mbox{if } r>R_{src}
\end{array}
\right.\label{Eq:SourceXYZ}\\
F(t)&=& \left(1-2\tau^2\right)e^{-\tau^2}.\label{Eq:SourceT}
\end{eqnarray}
where $R_{src}$ is the source radius,
$r=\sqrt{(x-x_{src})^2+(y-y_{src})^2+(z-z_{src})^2}$ is the distance
from the source center, $\tau$ is given by equation
\begin{equation}
\tau=\frac{\omega_0 (t-t_0)}{2} - \pi, \qquad t_0\leq t\leq
t_0+\frac{4\pi}{\omega},
\end{equation}
where $\omega_0$ is the central source frequency, $t_0$ is the
moment of the source initiation. This source model provides the wave
spectrum, which closely resembles the solar spectrum. It has a peak
near the central frequency $\omega_0$ and spreads over a broad
frequency interval. The source spectrum is:
\begin{equation}
|\hat{F}(\omega)| \equiv \left|\int_{-\infty}^\infty F(t)e^{-i\omega
t} dt\right|
=4\sqrt{\pi}\;\frac{\omega^2}{\omega_0^3}\;e^{-\frac{\omega^2}{\omega_0^2}}.
\end{equation}
A superposition of such sources with uniform distribution of central
frequencies randomly distributed below the photosphere describes
very well the observed solar oscillation spectrum
\citep{Parchevsky2007a}.
For numerical solution of Eqs (\ref{Eq:MHD_3D}) a semi-discrete
finite difference scheme of high order was used. At the top and
bottom boundaries non-reflective boundary conditions based on the
Perfectly Matched Layer (PML) technique were set. Details of
numerical realization of the code can be found in
\citet{Parchevsky2007a}.
\section{Background model of the sunspot}
We used two types of axially symmetric magnetohydrostatic background
models of the sunspot described in \citet{Khomenko2008} ("shallow"
model) and \citet{Khomenko2005} ("deep" model). The "shallow"
sunspot model is obtained as a combination of the self-similar
solution \citep{Low1975} in deep layers with the solution of Pizzo
\citep{Pizzo1986} in upper layers. We calculated three instances
(a), (b), and (c) of the "shallow" model with the following
photospheric strengths of the magnetic field at the sunspot axis:
0.83~kG, 1.4~kG, and 2.2~kG respectively. Strengths of the magnetic
field at the bottom of the domain for these models are 5.0~kG,
8.0~kG, and 12.5~kG respectively. The depth of the domain (from the
photosphere) for the "shallow" models is 9.87~Mm. The position of
the photosphere for the "quiet" Sun (at the boundary of the sunspot)
coincides with the photospheric level in the standard model S
\citep{Christensen-Dalsgaard1996}. For comparison purposes we
calculated one instance of the "deep" sunspot model with the maximum
strength of the magnetic field of 843~G at the photosphere and 29 kG
at the bottom of the domain. The "deep" sunspot model is based on
solution of Pizzo everywhere in the domain. The depths of the deep
model is 7.5~Mm.
Maps of the relative sound speed perturbations for the "shallow"
(panel a) and "deep" (panel c) models with photospheric strength of
the magnetic field of 843 G and 836 G respectively are presented in
Figure \ref{Fig:csMap}. The black horizontal line marks the position
of the photosphere. The red curve shows position of $\beta$~=~1
level. Both types of the models were calculated under the assumption
that $\Gamma_1$~=~5/3. Panels b and d show variations of relative
speed of the fast MHD wave. The velocity of the fast MHD wave
depends on the angle between the wave vector and the vector of the
magnetic field. The maximum value is plotted. So, panels (a) and (b)
represent bottom and top limits (depending on the direction of
propagation) for the speed of the fast MHD wave for the "shallow"
model. Panels (c) and (d) represent speed limits for the "deep"
model.
There are three main differences between these models: (i) the
topology of the magnetic field, (ii) the strength of the magnetic
field near the bottom of the domain, and (iii) dependence of the
horizontal profile of the sound speed on the depth. The magnetic
field lines (shown by blue curves in both panels of Figure
\ref{Fig:csMap}) are convex near the photosphere for the "deep"
model while in the "shallow" model the field lines are concave near
the photosphere. One dimensional cuts of the horizontal profiles of
the sound speed for different depths are shown in Figure
\ref{Fig:csHorz}. Profiles are scaled by the sound speed $c_{quiet}$
taken at the same depth of the outer boundary of the sunspot. In the
"shallow" model sound speed perturbation $\delta c/c =
c/c_{quiet}-1$ is close to zero everywhere starting from the depth
of about 2~Mm. In the "deep" model the sound speed perturbation is
non zero even at the bottom of the domain at 7.5~Mm. This means,
that in the "shallow" model waves, propagating at distances greater
than 10 Mm, propagate mostly in the region with the sound speed
profile of the quiet Sun below the region with the perturbed sound
speed. The vertical sound speed profiles at the sunspot center and
at the boundary for both models are shown in
Figure~\ref{Fig:csVert}.
The sound speed profiles along with the magnetic field structure are
the most significant parameters affecting propagation of the MHD
waves through the sunspot.
\section{Results and discussion}
In this section we present results of simulation of MHD waves
generated by a single source in different magnetohydrostatic
self-consistent background models. Axially symmetric models of
sunspots were interpolated on Cartesian grid with $\Delta x = \Delta
y = 0.15$~Mm. Vertical grid is non-uniform. We have three instances
(a), (b), and (c) of the "shallow" model (discussed above) with the
grid size of 376$\times$376$\times$67 and depth of 9.87~Mm (from the
photosphere). The vertical z-grid step varies from $\Delta z =
0.05$~Mm at the level of the photosphere and above to $\Delta z =
0.52$~Mm at the bottom of the domain. Time step $\Delta t = 0.05$~s
is the same for simulations with all three instances of the
background "shallow" model. The "deep" model has the grid size of
184$\times$184$\times$62, depth of 7.5~Mm (from the photosphere),
and $\Delta t = 0.1$~s. The vertical z-grid step varies from $\Delta
z = 0.05$~Mm at the level of the photosphere and above to $\Delta z
= 0.4$~Mm at the bottom of the domain.
The source of the vertical component of force described above with
the spectrum given by Eq.~(5) is placed at the distance of 9~Mm from
the axis of the sunspot for the "shallow" models and 6~Mm for the
"deep" model. The depth of the source is 0.1~Mm for both models. The
top absorbing PML is placed at the height of 0.5~Mm and extends up
to 0.9~Mm. In this region the vertical profile of the sound speed at
the outer boundary of the sunspot is close to the profile in the
standard solar model with $\Gamma_1$~= 5/3. The lateral boundary
conditions are chosen to be periodic.
Snapshots of the simulated wavefield for the "deep" model are shown
in Figure~\ref{Fig:OldSpot_map}. Panels a, b, and c represent maps
of perturbations of density $\rho'$, z-momentum $\rho_0 w'$, and
vertical component of the magnetic field $B_z'$ respectively at
moment $t$~= 20 min. Panels d, e, and f represent maps of the same
variables at $t$~= 25 min. Each panel consists of horizontal
XY-slice at the photospheric level (top) and vertical XZ-slice
through the center of the sunspot (bottom). The white line in the
XZ-slice shows position of the photosphere. Solid black curves
represent the magnetic field lines. The dashed lines near the top
and bottom of the XZ-slice show the position of the damping layers.
For the reference we plotted a circle with radius of 5 Mm and origin
at the wave source location. The non-uniform background model causes
anisotropy of the amplitude of the wave front. When the wave front
reaches the center of the sunspot the amplitude of density and
momentum perturbations decrease. After passing the center of the
sunspot the amplitude restores its original value. At some moments,
the amplitude inside the sunspot can become greater than the
amplitude of wave propagating outside the sunspot. The shape of the
wave front is changed as well. This is more noticeable for inner
parts of the wave front, located inside the circle in panels a, b,
and c. Later we show that these parts of the wave front are formed
mostly by $f$-modes. The shape of the outer parts of the wave front
(formed mostly by $p$-modes) remains close to the circle. A wave,
propagating along the magnetic field lines is appeared at the source
location. It is clearly seen in vertical slices of $\rho_0w'$ and
$B_z'$, but absent in $\rho'$. This wave consists of a mixture of
the Alfven and slow MHD waves.
It is interesting to compare propagation of MHD waves in the "deep"
and "shallow" models with similar photospheric strength of the
magnetic field. Results of wave simulations in the "shallow" model
with the photospheric strength of the magnetic field $B_0$~= 0.83~kG
are shown in Figure~\ref{Fig:NewSpot_map}. There are common features
with simulations for the "deep" model and there are differences. In
general, in the "shallow" model waves show the same behavior as in
the "deep" model. The amplitude of $\rho'$ and $\rho_0w'$ decreases
when the wave front reaches the center of the sunspot, but the ratio
of amplitudes is closer to unity than in the "deep" model. After
passing the center of the sunspot the wave restores its amplitude.
The amplitude of momentum perturbations $\rho_0w'$ becomes slightly
bigger inside the sunspot than outside, but again, not as much as in
the "deep" model. Perturbation of density $\rho'$ remains smaller
inside the sunspot than outside for all moments of time. The wave
that contains a mixture of the Alfven and slow MHD waves is
generated near the source location, as in the "deep" model, but the
amplitude of this wave is much smaller and it is not seen in
Figure~\ref{Fig:NewSpot_map}. The shape of the wave front is more
circular than in simulations for the "deep" model. In general, we
can say, that the "shallow" model affects the waves less than the
"deep" model.
Comparison of the z-momentum maps at the same moments of time in
different instances of the "shallow" model with different strength
of the photospheric magnetic field is shown in
Figure~\ref{Fig:NewSpt_Bdep}. The wave amplitude inside the sunspot
grows with the strength of the magnetic field and for the surface
strength of 2.2~kG the wave amplitude inside the sunspot becomes
bigger than outside.
Behavior of MHD waves inside sunspots depends on the distance of the
wave source from the sunspot center. This happens because
magnetoacoustic waves in the Sun propagate through the solar
interior. This means that waves, generated by the source located
farther from the sunspot center, propagate deeper (in the region
with stronger magnetic field) than waves from the closer source. The
angle between the direction of wave propagation and the background
magnetic field is also different in these two cases. Snapshots of
the wave field for the "shallow" model with surface strength of the
magnetic field 2.2~kG are shown in Figure~\ref{Fig:NewSpt_9-12Mm}.
The left panel represents the wave generated by the source located
at 9~Mm from the sunspot axis. In the right panel the source is
located at 12~Mm from the sunspot center. The wave amplitude from
the close source at some moment of time becomes bigger inside the
sunspot than outside. The amplitude of the wave front of the distant
source remains smaller inside the sunspot than outside for all
moments of time.
The wave source generates a mixture of fast, slow, Alfven, and
magnetogravity waves. The spectrum ($k-\nu$ diagram) for this case
is shown in Figure~\ref{Fig:kw_OldSpt}. The solid black curve
represents the theoretical curve of the $f$-mode in absence of the
magnetic field. The fast MHD and magnetogravity modes can be
separated by filtering out the $f$-mode. Result of such separation
is shown in Figure~\ref{Fig:fp_kwseprt_OldSpt}. The top row
represents the original $k$-$\nu$ diagram (obtained form the
z-component of velocity at the photospheric level) and $k$-$\nu$
diagrams obtained by filtering out $f$- and $p$-modes respectively.
The white solid curve shows the theoretical position of the $f$-mode
ridge in the absence of the magnetic field. The middle row
represents maps of $V_z$ for $p$- and $f$-modes at the photospheric
level at the moment $t$=23 min. It was shown \citep{Parchevsky2009}
that in case of the horizontally uniform background model and
uniform inclined background magnetic field wavefronts of $p$- and
$f$-modes are spatially separated after 20-30~min. This does not
happen in the case of the non-uniform background magnitohydrostatic
model. The solid black circle marks the inner part of the wavefront
of $p$-modes. Although the amplitude of the $f$-modes (the right
panel) is mostly concentrated inside the circle, there is noticeable
non-zero amplitude in the region outside the circle where $p$-modes
present. It is clear, that the deformation of the wavefront (both
the shape and wave amplitude) due to the interaction with the
sunspot is much stronger for $f$-modes than for $p$-modes. The wave
amplitude of the $p$-mode wavefront decreases at the sunspot center
and quickly restores its value when the front passes through the
center. The amplitude of the $f$-mode wavefront remains perturbed at
all moments of time.
\section{Conclusion}
According to our 3D numerical simulations of MHD wave propagation in
two different models (referred as "shallow" and "deep") we can point
out to the following characteristic behavior of waves inside
sunspots. The interaction with the sunspot changes the shape of the
wave front and amplitude of the $f$-mode waves significantly
stronger than of the $p$-mode waves. The amplitude of the wave front
of the $p$-modes decreases when the wave reaches the sunspot center
and restores its original value when the wave passes the center of
the sunspot. The "shallow" model of sunspot affects waves less than
the "deep" model. This means that the horizontal inhomogeneity of
the sound speed profile inside the sunspot is mostly responsible for
perturbations of the wave front. In the "shallow" model the
horizontal distribution of the sound speed below 2 Mm is almost
uniform and the sound speed coincides with the value in the quiet
Sun at the same depth. Only magnetic field perturbs the wave front
in this region and the total perturbation of the wave front becomes
weaker than in the "deep" case.
Inside the sunspot magnetoacoustic and magnetogravity waves are not
spatially separated unlike the case of the uniform inclined magnetic
field. The wave amplitude inside sunspots depends on the strength of
the magnetic field and the distance of the wave source from the
sunspot axis. The stronger photospheric magnetic field, the bigger
wave amplitude inside the sunspot (if the source is located at the
same distance from the sunspot center). For the source located at
9~Mm from the sunspot axis the wave amplitude inside the sunspot at
some moment becomes bigger than the amplitude outside. For the
source located at 12 Mm the wave amplitude inside the sunspot
remains smaller that outside for all moments of time.
In this paper, we presented initial results of 3D simulations of
helioseismic MHD waves in magnetostatic sunspot models.Future work
includes simulations with multiple random sources for testing the
travel-time measurement procedures of time-distance helioseismology,
and also modeling wave propagation in the MHD models of sunspots,
including flows \citep{Botha2008, Hurlburt2000}.
|
\section{Introduction}\label{sec:introduction}
Knot Floer homology, primarily as an invariant for knots and links
inside $S^3$, was discovered by Peter Ozsv\'{a}th{} and Zolt\'{a}n{}
Szab\'{o}{} \cite{POZSzknotinvariants}, and independently by Jacob
Rasmussen \cite{JR}. Later, a related invariant for links, called link
Floer homology, was constructed by Peter Ozsv\'{a}th{} and Zolt\'{a}n{}
Szab\'{o}{} \cite{POZSzlinkinvariants}. However, due to certain
orientation issues, the link invariant was only constructed over
$\mathbb{F}_2$, instead of $\mathbb{Z}$. This short note is the author's effort to
understand the orientation issues that are known, and to resolve some
of the issues that are unknown.
Let us describe the algebraic structure of knot Floer homology in the
simplest case, as described in \cite{POZSzknotinvariants}. Let $K$ be
a null-homologous knot in $\#^{l-1}(S^1\times S^2)$. Then there are
$2^{l-1}$ bi-graded chain complexes over $\mathbb{Z}$, such that they all give
rise to the same complex when tensored with $\mathbb{F}_2$. The two
gradings are called the Maslov grading $M$ and the Alexander grading
$A$. The boundary maps preserve the Alexander grading, but lower the
Maslov grading by one. Therefore, the Maslov grading acts as the
homological grading while the Alexander grading acts as an extra
filtration. The homology of the chain complexes is called the hat
version of the knot Floer homology. Therefore, we get an $\mathbb{F}_2$-valued
bi-graded hat version of knot Floer homology and $2^{l-1}$ $\mathbb{Z}$-valued
bi-graded hat versions of knot Floer homology.
The reason for working with null-homologous knots in connected sums of
$S^1\times S^2$ is very simple. We want to work with links in $S^3$.
However, a link with $l$ components in $S^3$ very naturally gives rise
to a null-homologous knot in $\#^{l-1}(S^1\times S^2)$,
\cite{POZSzknotinvariants}. Therefore, what we have is the following.
Given a link $L\subset S^3$, with $l$ components, and after making
certain auxiliary choices, we get $2^{l-1}$ bi-graded chain complexes
over $\mathbb{Z}$, henceforth denoted by $\widehat{\mathit{CFK}}(L,\mathbb{Z},\mathfrak{o})$, where
$\mathfrak{o}$, called an orientation system, takes values in an indexing
set of $2^{l-1}$ elements, and records which of the $2^{l-1}$ chain
complexes is the one under consideration. All of the $2^{l-1}$ chain
complexes give rise the same bi-graded chain complex over $\mathbb{F}_2$,
$\widehat{\mathit{CFK}}(L,\mathbb{F}_2)=\widehat{\mathit{CFK}}(L,\mathbb{Z},\mathfrak{o})\otimes \mathbb{F}_2$. The reader
should be warned that these bi-graded chain complexes,
$\widehat{\mathit{CFK}}(L,\mathbb{Z},\mathfrak{o})$ and $\widehat{\mathit{CFK}}(L,\mathbb{F}_2)$, are not
link-invariants (they might depend on the auxiliary choices that we
did not specify, but simply alluded to), but their homologies are link
invariants. Therefore, we get one $\mathbb{F}_2$-valued bi-graded hat version
of knot Floer homology $\widehat{\mathit{HFK}}(L,\mathbb{F}_2)=H_*(\widehat{\mathit{CFK}}(L,\mathbb{F}_2))$, and
$2^{l-1}$ $\mathbb{Z}$-valued bi-graded hat versions of knot Floer homology
$\widehat{\mathit{HFK}}(L, \mathbb{Z}, \mathfrak{o})=H_*(\widehat{\mathit{CFK}}(L,\mathbb{Z},\mathfrak{o}))$. We often
let $\widehat{\mathit{HFK}}(L,\mathbb{Z})$ denote any one of the $2^{l-1}$ versions, or a canonical one, namely the one coming from the canonical choice of orientation systems in \cite{POZSzapplications}. However, to decide which of the $2^{l-1}$ groups $\widehat{\mathit{HFK}}(L,\mathbb{Z},\mathfrak{o})$ is the canonical one, one needs to understand some of the other versions of link Floer homology, most notably the infinity version. This seems to be a harder problem, for reasons that we will discuss shortly.
In \cite{POZSzlinkinvariants}, the story for links is treated
in a slightly different light, and a new definition of link Floer
homology is given. Given a link $L$ with $l$ components in $S^3$,
modulo certain choices, a chain complex $\widehat{\mathit{CFL}}(L,\mathbb{F}_2)$ over $\mathbb{F}_2$
is constructed. The chain complex carries $(l+1)$ gradings: a single
Maslov grading $M$, which is lowered by one by the boundary map, and
$l$ Alexander gradings $A_1, A_2, \ldots, A_l$, one for each link
component, each of which is preserved by the boundary map. The
homology of the chain complex $\widehat{\mathit{HFL}}(L,\mathbb{F}_2)=H_*(\widehat{\mathit{CFL}}(L,\mathbb{F}_2))$
is an $\mathbb{F}_2$-valued $(l+1)$-graded homology theory, called the link
Floer homology, and it is a link invariant. These two definitions,
\emph{a priori}, are different. Therefore, we have been careful
throughout; we have called the definition from
\cite{POZSzknotinvariants} the knot Floer homology (even when talking
about links), and denoted it by $\widehat{\mathit{HFK}}$, and we have called the
definition from \cite{POZSzlinkinvariants} the link Floer homology,
and denoted it by $\widehat{\mathit{HFL}}$. However, by a miraculous coincidence, it
turns out that if we condense the $l$ Alexander gradings in
$\widehat{\mathit{HFL}}(L,\mathbb{F}_2)$ into one single Alexander grading $A=\sum_i A_i$,
then the resulting $\mathbb{F}_2$-valued bi-graded homology group is
isomorphic to $\widehat{\mathit{HFK}}(L,\mathbb{F}_2)$.
In this note, we will complete the picture by constructing $2^{l-1}$
$\mathbb{Z}$-valued chain complexes, $\widehat{\mathit{CFL}}(L,\mathbb{Z},\mathfrak{o})$, each carrying a
Maslov grading $M$, and $l$ Alexander gradings $A_1,A_2,\ldots,A_l$,
such that the homologies
$\widehat{\mathit{HFL}}(L,\mathbb{Z},\mathfrak{o})=H_*(\widehat{\mathit{CFL}}(L,\mathbb{Z},\mathfrak{o}))$ are link
invariants, and on condensing the $l$ Alexander gradings into one
Alexander grading $A=\sum_i A_i$, we get the $2^{l-1}$ $\mathbb{Z}$-valued
bi-graded homology groups $\widehat{\mathit{HFK}}(L,\mathbb{Z},\mathfrak{o})$.
A similar story (except possibly the last bit of coincidence) holds
for the other versions of link Floer homologies, most notably the
minus, plus and infinity versions; however, the holomorphic
considerations and the orientation issues are significantly more
subtle. In particular, we will encounter boundary degenerations, and we will have to orient the relevant moduli spaces in a consistent fashion. We plan to address this problem in future work. Understanding the orientation issues for all versions of link Floer homology will help us understand which of the $2^{l-1}$ link Floer homology groups is the canonical one and whether it has some sort of a useful characterization.
For the second part of the discourse, we concentrate on the
computational aspects of the theory. Ever since knot Floer homology
saw the light of day \cite{POZSzknotinvariants}, \cite{JR},
\cite{POZSzlinkinvariants}, and some of its immense strengths were
discovered \cite{POZSzgenusbounds}, \cite{POZSzthurstonnorm}, \cite{YN}, people were
interested in algorithms to compute it. There have been several
recent developments towards computing various versions of link Floer
homology for links in $S^3$ \cite{CMPOSS}, \cite{SSJW}, \cite{POZSzcube}, \cite{POASZSz}.
We choose to concentrate on the algorithm from \cite{CMPOSS}: the link
$L$ in $S^3$ is represented by a toroidal grid diagram $G$, such that
the $i^{\text{th}}$ component is represented by $m_i$ vertical line segments
and $m_i$ horizontal line segments; an $\mathbb{F}_2$-valued $(l+1)$-graded
chain complex $C(G)$ is constructed such that its homology $H_*(C(G))$
is isomorphic to
$\widehat{\mathit{HFL}}(L,\mathbb{F}_2)\otimes_i(\otimes^{m_i-1}(\mathbb{F}_2\oplus\mathbb{F}_2))$, where,
in the $(\mathbb{F}_2\oplus\mathbb{F}_2)$ that is tensored with itself $(m_i-1)$
times, for one of the generators, all the $(l+1)$ gradings are zero,
and for the other generator, the Maslov grading $M=-1$, and the
Alexander grading $A_j=-\delta_{ij}$.
Very shortly thereafter, \cite{CMPOZSzDT} assigned signs of $\pm 1$ to each of the boundary maps in the chain complex $C(G)$ in a well defined way, such that it remains a chain complex and its homology (over $\mathbb{Z}$) is isomorphic to $\widehat{\mathit{HFG}}(L,\mathbb{Z})\otimes_i(\otimes^{m_i-1}(\mathbb{Z}\oplus\mathbb{Z}))$, for some $(l+1)$-graded group $\widehat{\mathit{HFG}}(L,\mathbb{Z})$, which is a link invariant.
A very natural question that arises is whether the new homology group $\widehat{\mathit{HFG}}(L,\mathbb{Z})$ is isomorphic to $\widehat{\mathit{HFL}}(L,\mathbb{Z},\mathfrak{o})$ for some $\mathfrak{o}$. We establish that the answer is in the affirmative, and indeed, we construct $2^{l-1}-1$ other sign assignments on the boundary maps of $C(G)$, such that the homologies of these $2^{l-1}$ sign refined grid chain complexes correspond precisely to the $2^{l-1}$ $\mathbb{Z}$-valued $(l+1)$-graded homology groups $\widehat{\mathit{HFL}}(L,\mathbb{Z})$. Once again, it is an interesting question whether $\widehat{\mathit{HFG}}(L,\mathbb{Z})$ is isomorphic to the canonical $\widehat{\mathit{HFL}}(L,\mathbb{Z})$, and once again, we are unable to answer it with our present methods. It is also an interesting endeavor to find two $l$-component links $L_1$ and $L_2$, such that $\widehat{\mathit{CFL}}(L_1,\mathbb{F}_2)$ is isomorphic to $\widehat{\mathit{CFL}}(L_2,\mathbb{F}_2)$ as $(l+1)$-graded $\mathbb{F}_2$-modules, there is a bijection between the set of $2^{l-1}$ groups $\widehat{\mathit{CFK}}(L_1,\mathbb{Z})$ and the set of $2^{l-1}$ groups $\widehat{\mathit{CFK}}(L_2,\mathbb{Z})$ such that the corresponding groups are isomorphic as bi-graded $\mathbb{Z}$-modules, $\widehat{\mathit{HFG}}(L_1,\mathbb{Z})$ is isomorphic to $\widehat{\mathit{HFG}}(L_2,\mathbb{Z})$ as $(l+1)$-graded $\mathbb{Z}$-modules, but there is no bijection between the set of $2^{l-1}$ groups $\widehat{\mathit{HFL}}(L_1,\mathbb{Z})$ and the set of $2^{l-1}$ groups $\widehat{\mathit{HFL}}(L_2,\mathbb{Z})$ such that the corresponding groups are isomorphic as $(l+1)$-graded $\mathbb{Z}$-modules.
This is a rather short paper. We expect the reader to be already
familiar with most of \cite{CMPOZSzDT}, \cite{POZSzknotinvariants},
\cite{POZSzlinkinvariants}. Despite trying our level best to be as
self-contained as possible, we will still be rather fast in our
exposition.
\subsection*{Acknowledgment}
The work was done when the author was supported by the Clay Research Fellowship. He would like to thank Robert Lipshitz, Peter Ozsv\'{a}th{} and Zolt\'{a}n{} Szab\'{o}{} for several helpful discussions. He would also like to thank the referee for providing useful comments and for pointing out the errors.
\section{Floer homology}\label{sec:floerhomology}
For the first part of the section, in the following few numbered paragraphs, we will briefly review the basics of Heegaard Floer homology. The interested reader is referred to \cite{POZSz}, \cite{POZSzapplications} for more details.
\subsection{} A \emph{Heegaard diagram} is an object $\mathcal{H}=(\Sigma_g, \alpha_1, \ldots,\alpha_{g+k-1},\beta_1,\ldots,\beta_{g+k-1},\allowbreak X_1,\ldots,X_k,O_1,\ldots,O_k)$, where: $\Sigma_g$ is a Riemann surface of genus $g$; $\alpha=(\alpha_1,\ldots,\allowbreak \alpha_{g+k-1})$ is $(g+k-1)$-tuple of disjoint simple closed curves such that $\Sigma_g\setminus\alpha$ has $k$ components; $\beta=(\beta_1,\ldots, \beta_{g+k-1})$ is $(g+k-1)$-tuple of disjoint simple closed curves such that $\Sigma_g\setminus\beta$ has $k$ components; the $\alpha$ circles are transverse to the $\beta$ circles; $X=(X_1,\ldots,X_k)$ is a $k$-tuple of points such that each component of $\Sigma_g\setminus\alpha$ has an $X$ marking, and each component of $\Sigma_g\setminus\beta$ has an $X$ marking; $O=(O_1,\ldots,O_k)$ is a $k$-tuple of points such that each component of $\Sigma_g\setminus\alpha$ has an $O$ marking, and each component of $\Sigma_g\setminus\beta$ has an $O$ marking; and the diagram is assumed to be \emph{admissible}, which is a technical condition that we will describe later.
\subsection{}A Heegaard diagram represents an oriented link $L$ inside a three-manifold $Y$ in the following way: the pair $(\Sigma_g,\alpha)$ represents genus $g$ handlebody $U_{\alpha}$; the pair $(\Sigma_g,\beta)$ represents genus $g$ handlebody $U_{\beta}$; the ambient three-manifold $Y$ is obtained by gluing $U_{\alpha}$ to $U_{\beta}$ along $\Sigma_g$; the $X$ markings are joined to the $O$ markings by $k$ simple oriented arcs in the complement of the $\alpha$ circles, and the interiors of the $k$ arcs are pushed slightly inside the handlebody $U_{\alpha}$; the $O$ markings are joined to the $X$ markings by $k$ simple oriented arcs in the complement of the $\beta$ circles, and the interiors of the $k$ arcs are pushed slightly inside the handlebody $U_{\beta}$; the union of these $2k$ oriented arcs is the oriented link $L$. Let the link have $l$ components, and let $2 m_i$ be the number of arcs that represent $L_i$, the $i^{\text{th}}$ component of the link $L$. Therefore, $k=\sum_i m_i \geq l$. In \cite{POZSzlinkinvariants}, the case $k=l$ is studied, and in \cite{POZSzknotinvariants}, the subcase $k=l=1$ is dealt with. We will always assume that $L_i$ is null-homologous in $Y$, for each $i$.
\subsection{}Consider $(g+k-1)$-tuples of points $x=(x_1,\ldots,x_{g+k-1})$, such that each $\alpha$ circle contains some $x_i$, and each $\beta$ circle contains some $x_j$. To each such tuple $x$, we can associate a $\text{Spin}^{\text{C}}$ structure $\mathfrak{s}_x$ on the ambient three-manifold $Y$. In all the three-manifolds that we will consider, we will be interested in a canonical torsion $\text{Spin}^{\text{C}}$ structure. In particular, for $Y=\#^n S^1\times S^2$, we will be interested in the unique torsion $\text{Spin}^{\text{C}}$ structure. A \emph{generator} is a $(g+k-1)$-tuple $x$ of the type described above, such that $\mathfrak{s}_x$ is the canonical $\text{Spin}^{\text{C}}$ structure. The set of all generators in a Heegaard diagram $\mathcal{H}$ is denoted by $\mathcal{G}_{\mathcal{H}}$. An \emph{elementary domain} is a component of $\Sigma_g\setminus(\alpha\cup\beta)$. A \emph{domain} $D$ joining a generator $x$ to a generator $y$, is a $2$-chain generated by elementary domains such that $\partial(\partial D|_{\alpha})=y-x$. The set of all domains joining $x$ to $y$ is denoted by $\mathcal{D}(x,y)$. A \emph{periodic domain} $P$ is a $2$-chain generated by elementary domains such that $\partial(\partial P|_{\alpha})=0$. The set of periodic domains is denoted by $\mathcal{P}_{\mathcal{H}}$, and there is a natural bijection between $\mathcal{P}_{\mathcal{H}}$ and $\mathcal{D}(x,x)$ for any generator $x$. If $D$ is a domain, and if $p$ is a point lying in an elementary domain, then $n_p(D)$ denotes the coefficient of the $2$-chain $D$ at that elementary domain. Let $n_X(D)=\sum_i n_{X_i}(D)$ and $n_O(D)=\sum n_{O_i}(D)$. Furthermore, let $n_{X,i}(D)$ denote the sum of $n_{X_j}(D)$ for all the $X_j$ markings that lie in $L_i$, and let $n_{O,i}(D)$ denote the sum of $n_{O_j}(D)$ for all the $O_j$ markings that lie in $L_i$. A domain is said to be \emph{non-negative} if it has non-negative coefficients in every elementary domain. A domain $D$ is said to be \emph{empty} if $n_{X_i}(D)=n_{O_i}(D)=0$ for all $i$. A Heegaard diagram is called \emph{admissible} if there are no non-negative, non-trivial empty periodic domains. The set of all empty domains in $\mathcal{D}(x,y)$ is denoted by $\mathcal{D}^0(x,y)$, and the set of all empty periodic domains is denoted by $\mathcal{P}^0_{\mathcal{H}}$. The set $\mathcal{P}^0_{\mathcal{H}}$ forms a free abelian group of rank $b_1(Y)+l-1$.
\subsection{}Every domain $D$ has an integer valued \emph{Maslov index} $\mu(D)$ associated to it, which satisfies certain properties that we will mention as we need them. In all the Heegaard diagrams that we will consider, the following additional restrictions will hold: if $P\in\mathcal{D}(x,x)$, then $\mu(P)=2 n_O(P)$ and, since $L_i$ is null-homologous in $Y$, $n_{X,i}(P)= n_{O,i}(P)$ for all $i$. This allows us to define $(l+1)$ relative gradings. Given two generators $x,y$, choose a domain $D\in\mathcal{D}(x,y)$ (since $\mathfrak{s}_x=\mathfrak{s}_y$, the set $\mathcal{D}(x,y)$ is non-empty), and let the \emph{relative Maslov grading} $M(x,y)=\mu(D)-2n_O(D)$, and let the \emph{relative Alexander grading} $A_i(x,y)=n_{X,i}(D)-n_{O,i}(D)$. In certain situations, with certain additional hypotheses, these gradings can be lifted to absolute gradings. However, for convenience, we will not work with absolute gradings right away. Therefore, until Lemma \ref{lem:absolutegrading}, whenever we talk about the Maslov grading $M$, or the Alexander grading $A_i$, we mean some affine lift of the corresponding relative grading, which is only well-defined up to a translation by $\mathbb{Z}$. Let $Q_i=\mathbb{Z}\oplus\mathbb{Z}$ be the $(l+1)$-graded group, with the two generators lying in gradings $(0,0,\ldots,0)$ and $(-1,-\delta_{i1},\ldots,-\delta_{il})$, where $\delta$ is the Kronecker delta function.
\subsection{}For the analytical aspects of the theory, which we are about to describe now, the reader is strongly advised to read Section 3 of \cite{POZSz}. Let $\text{Sym}^{g+k-1}(\Sigma_g)$ be $(g+k-1)$-fold symmetric product, and let $J_s$ be a path of nearly symmetric almost complex structures on it, obtained as a small perturbation of the constant path of nearly symmetric almost complex structure $\text{Sym}^{g+k-1}(\mathfrak{j})$, where $\mathfrak{j}$ is a fixed complex structure on $\Sigma_g$, such that $J_s$ achieves certain transversality that we will describe later. The subspaces $\mathbb{T}_{\alpha}=\alpha_1\times\cdots\times\alpha_{g+k-1}$ and $\mathbb{T}_{\beta}=\beta_1\times\cdots\times\beta_{g+k-1}$ are two totally real tori. Notice that $\mathcal{G}_{\mathcal{H}}$ is in a natural bijection with a subset of $\mathbb{T}_{\alpha}\cap\mathbb{T}_{\beta}$. Fix $\mathfrak{p}>2$. Given a domain $D\in\mathcal{D}(x,y)$, let $\mathcal{B}(D)$ be the space of all $L^{\mathfrak{p}}_1$ maps $u$ from $[0,1]\times\mathbb{R}\subset\mathbb{C}$ to $\text{Sym}^{g+k-1}(\Sigma_g)$, such that: $u$ maps $\{0\}\times\mathbb{R}$ to $\mathbb{T}_{\alpha}$; $u$ maps $\{1\}\times\mathbb{R}$ to $\mathbb{T}_{\beta}$; $\lim_{t\rightarrow\infty}u(s+it)=x$ with a certain pre-determined asymptotic behavior; $\lim_{t\rightarrow -\infty}u(s+it)=y$ with a certain pre-determined asymptotic behavior; for any point $p$ in any elementary domain, the algebraic intersection number between $u$ and $\{p\}\times \text{Sym}^{g+k-2}(\Sigma_g)$ is $n_p(D)$, or, as it is colloquially stated, the domain $D$ is the \emph{shadow} of $u$. Ozsv\'{a}th{} and Szab\'{o}{} define a vector bundle $\mathcal{L}$ over $\mathcal{B}(D)$, and a section $\xi$ of that bundle depending on $J_s$, such that the linearization of the section $D_u\xi$ is a Fredholm operator for every $u\in\mathcal{B}(D)$. The transversality of the path $J_s$ that we mentioned earlier, simply means that the Fredholm section $\xi$ is transverse to the $0$-section of $\mathcal{L}$. The intersection of $\xi$ and the $0$-section is denoted by $\mathcal{M}_{J_s}(D)$, and it consists precisely of the $J_s$-holomorphic maps. There is an $\mathbb{R}$ action on $\mathcal{M}_{J_s}(D)$ coming from the $\mathbb{R}$ action on $[0,1]\times\mathbb{R}$, and the \emph{unparametrized moduli space} is denoted by $\widehat{\mathcal{M}_{J_s}}(D)=\mathcal{M}_{J_s}(D)/\mathbb{R}$. The virtual index bundle of the linearization map $D_u$ gives an element of the $K$-theory of $\mathcal{B}(D)$. Its dimension is the expected dimension of the moduli space $\mathcal{M}_{J_s}(D)$, and this dimension is in fact the Maslov index $\mu(D)$, that we had mentioned earlier. The determinant line bundle of the index bundle, henceforth denoted by $\det(D)$, turns out to be a trivializable line bundle over $\mathcal{B}(D)$. Therefore, a choice of a nowhere vanishing section on the trivializable line bundle $\det(D)$, produces an orientation of the moduli space $\mathcal{M}_{J_s}(D)$, and hence an orientation of the unparametrized moduli space $\widehat{\mathcal{M}_{J_s}}(D)$.
\subsection{}\label{para:special}If $D_1\in\mathcal{D}(x,y)$ and $D_2\in\mathcal{D}(y,z)$ are domains, then the $2$-chain $D_1+D_2$ lies in $\pi_2(x,z)$. The asymototic behaviors that we had mentioned earlier, along with some globally pre-determined choices, allows us to get a pre-gluing map from $\mathcal{B}(D_1)\times\mathcal{B}(D_2)$ to $\mathcal{B}(D_1+D_2)$. The pullback of the line bundle $\det(D_1+D_2)$ over $\mathcal{B}(D_1+D_2)$ can be canonically identified with the line bundle $\det(D_1)\wedge \det(D_2)$ over $\mathcal{B}(D_1)\times\mathcal{B}(D_2)$ by linearized gluing. An \emph{orientation system} $\mathfrak{o}$ is a choice of a nowhere vanishing section $\mathfrak{o}(D)$ of the line bundle $\det(D)$ for every domain $D\in\mathcal{D}(x,y)$, and for every pair of generators $x,y\in\mathcal{G}_{\mathcal{H}}$, such that if $D_1\in\mathcal{D}(x,y)$ and $D_2\in\mathcal{D}(y,z)$, then $\mathfrak{o}(D_1)\wedge\mathfrak{o}(D_2)=\mathfrak{o}(D_1+D_2)$. Therefore, two orientation systems $\mathfrak{o}_1$ and $\mathfrak{o}_2$ disagree on $D_1+D_2$ if and only if they disagree on exactly one of the two domains $D_1$ and $D_2$.
\subsection{}The following describes a method to find all possible orientation systems. Fix a generator $x\in\mathcal{G}_{\mathcal{H}}$, and for every other generator $y$, choose a domain $D_y\in\mathcal{D}(x,y)$. Then choose a set of periodic domains $P_1,\ldots,P_m$, which freely generate $\mathcal{P}_{\mathcal{H}}$. Orient the determinant line bundles over the domains $D_y$ and $P_j$ arbitrarily. Since any domain $D\in\mathcal{D}(y,z)$ can be written uniquely as $D=\sum_j a_j P_j +D_z-D_y$, this choice uniquely specifies an orientation system.
Thus, an orientation system is specified by its values on certain domains $D_y$ and certain periodic domains $P_j$. This allows us to define a chain complex over $\mathbb{Z}$, and it will turn out that the gauge equivalence class of the sign assignment on the chain complex is independent of the orientations of the line bundles $\det(D_y)$. Therefore, declare two orientations systems to be \emph{strongly equivalent} if they agree on all the periodic domains in $\mathcal{P}_{\mathcal{H}}$ (or in other words, they agree on all the periodic domains $P_1,\ldots,P_m$). There is a second notion of equivalence, which is of some importance to us, whereby two orientation systems are declared to be \emph{weakly equivalent} if they agree on all the periodic domains in $\mathcal{P}^0_{\mathcal{H}}$. Let $\widehat{\mathcal{O}}_{\mathcal{H}}$ denote the set of weak equivalence classes of orientation systems. Then $\widehat{\mathcal{O}}_{\mathcal{H}}$ is a torseur over $\operatorname{Hom}(\mathcal{P}^0_{\mathcal{H}},\mathbb{Z}/2\mathbb{Z})$, so there are exactly $2^{b_1(Y)+l-1}$ weak equivalence classes of orientation systems.
\vspace{12 pt}
If $D\in\mathcal{D}(x,y)$ is a domain, its unparametrized moduli space
$\widehat{\mathcal{M}_{J_s}}(D)$ is a compact, $(\mu(D)-1)$-dimensional manifold
with corners by Gromov compactness and the fact that $J_s$ achieves
transversality; an orientation system $\mathfrak{o}$ determines an
orientation on $\widehat{\mathcal{M}_{J_s}}(D)$. Therefore, if $\mu(D)=1$, then
$\widehat{\mathcal{M}_{J_s}}(D)$ is a compact oriented zero-dimensional manifold with
corners, or in other words, it is a finite number of signed
points. Let $c(D)$ be the total number of points, counted with
sign. The cornerstone of Floer homology in the present setting, is the
following lemma.
\begin{lem}\cite{POZSz}\label{lem:index2}
If $D\in\mathcal{D}(x,y)$ is a domain with $\mu(D)=2$, then
$\widehat{\mathcal{M}_{J_s}}(D)$ is an oriented one-dimensional
manifold. Furthermore, if $D=D_1+D_2$, where $D_1\in\mathcal{D}(x,z)$ and
$D_2\in\mathcal{D}(z,y)$, with $\mu(D_1)=1$ and $\mu(D_2)=1$, then the
total number of points in the boundary of $\widehat{\mathcal{M}_{J_s}}(D)$ that
correspond to a decomposition of $D$ as $D_1+D_2$, when counted with
signs induced from the orientation of $\widehat{\mathcal{M}_{J_s}}(D)$, equals
$c(D_1)c(D_2)$.
\end{lem}
An immediate corollary is the following: if all the points in the
boundary of $\widehat{\mathcal{M}_{J_s}}(D)$ correspond to such a decomposition --- in other words, if bubbling and boundary degenerations can be ruled out --- then the sum $\sum c(D_1)c(D_2)$ over all such possible decompositions
is zero. This allows us to define the following $(l+1)$-graded chain
complex over $\mathbb{Z}$. This is a well-known chain complex, and it was
first defined by Ozsv\'{a}th{} and Szab\'{o}{} for $k=1$. However, for a
general value of $k$, the chain complex was originally not defined
over $\mathbb{Z}$. There are certain subtleties that need to be resolved
before the minus version can be defined over $\mathbb{Z}$, namely, we have to
orient the boundary degenerations in a consistent manner such that the
proofs of Theorems \ref{thm:main}, \ref{thm:secondmain} and
\ref{thm:connectsum} go through; however, those
issues do not appear when we work only in the hat version.
\begin{defn}
Given an admissible Heegaard diagram $\mathcal{H}$ for $L$ and an orientation system $\mathfrak{o}\in\widehat{\mathcal{O}}_{\mathcal{H}}$, let $\widehat{\mathit{CFL}}_{\mathcal{H}}(L,\mathbb{Z},\mathfrak{o})$ be the chain complex freely generated over $\mathbb{Z}$ by the elements of $\mathcal{G}_{\mathcal{H}}$, with the $(l+1)$ gradings given by $M,A_1,\ldots,A_l$, and the boundary map given by $\partial x=\sum_{y\in\mathcal{G}_\mathcal{H}}\sum_{D\in\mathcal{D}^0(x,y),\mu(D)=1}c(D)y$.
\end{defn}
\begin{lem}
The map $\partial$ on $\widehat{\mathit{CFL}}_{\mathcal{H}}(L,\mathbb{Z},\mathfrak{o})$ reduces the Maslov grading by $1$, keeps all Alexander gradings fixed, and satisfies $\partial^2=0$.
\end{lem}
\begin{proof}
The claims regarding the gradings follow directly from the
definitions. To prove that $\partial^2=0$, by Lemma \ref{lem:index2}, we
only need to show that for any empty Maslov index $2$ domain $D$,
the boundary points of $\widehat{\mathcal{M}}(D)$ do not correspond to
bubbling or boundary degenerations. However, the shadow of a bubble
or a boundary degeneration is a $2$-chain in the Heegaard diagram,
whose boundary lies entirely within the $\alpha$ circles, or
entirely within the $\beta$ circles. Any such $2$-chain must have
non-zero coefficient at some $X$ marking, and therefore by
positivity of domains, the original domain must also have non-zero
coefficient at that $X$ marking, and therefore, could not have been
empty.
\end{proof}
Even though we did not specify in the notations, $\widehat{\mathit{CFL}}_{\mathcal{H}}(L,\mathbb{Z},\mathfrak{o})$ might also depend on the path of almost
complex structures $J_s$ on $\text{Sym}^{g+k-1}(\Sigma_g)$. However, the
homology $H_*(\widehat{\mathit{CFL}}_{\mathcal{H}}(L,\mathbb{Z},\mathfrak{o}))$, as an $(l+1)$-graded
object, depends only on the link $L$, the numbers of $X$ markings,
$m_i$, that lie on the $i^{\text{th}}$ link component for each $i$,
and the weak equivalence class of the orientation system $\mathfrak{o}$.
\begin{thm}\label{thm:main}
For a fixed Heegaard diagram $\mathcal{H}$ and a fixed path of almost
complex structures $J_s$, if $\mathfrak{o}_1$ and $\mathfrak{o}_2$ are weakly
equivalent, then the two chain complexes
$\widehat{\mathit{CFL}}_{\mathcal{H}}(L,\mathbb{Z},\mathfrak{o}_1)$ and
$\widehat{\mathit{CFL}}_{\mathcal{H}}(L,\mathbb{Z},\mathfrak{o}_2)$ are isomorphic. If $\mathcal{H}_1$ and
$\mathcal{H}_2$ are two different Heegaard diagrams for the same link
$L$, such that in both $\mathcal{H}_1$ and $\mathcal{H}_2$, the
$i^{\text{th}}$ link component $L_i$ is represented by $m_i$ $X$
markings and $m_i$ $O$ markings, and if $J_{s,1}$ and $J_{s,2}$ are
two paths of almost complex structures on the two symmetric
products, then there is a bijection $f$ between
$\widehat{\mathcal{O}}_{\mathcal{H}_1}$ and $\widehat{\mathcal{O}}_{\mathcal{H}_2}$, such that for
every $\mathfrak{o}\in\widehat{\mathcal{O}}_{\mathcal{H}_1}$, the
homology $H_*(\widehat{\mathit{CFL}}_{\mathcal{H}_1}(L,\mathbb{Z},\mathfrak{o}))$ is isomorphic to
the homology $H_*(\widehat{\mathit{CFL}}_{\mathcal{H}_2}(L,\mathbb{Z},f(\mathfrak{o})))$, as
$(l+1)$-graded groups.
\end{thm}
\begin{proof}
This is neither a new type of a theorem, nor a new idea of a proof.
For the first part, let $\mathfrak{o}_1$ and $\mathfrak{o}_2$ be two weakly
equivalent orientation systems. We are going to define a map
$t:\mathcal{G}_{\mathcal{H}}\rightarrow\{\pm 1\}$ in the following way. Call
two generators $x$ and $y$ to be connected if there is an empty
domain $D\in\mathcal{D}^0(x,y)$. For each connected component of
$\mathcal{G}_{\mathcal{H}}$, choose a generator $x$ in that connected
component, and declare $t(x)=1$. For every other generator $y$ in
that connected component, choose an empty domain
$D_y\in\mathcal{D}^0(x,y)$, and declare $t(y)=1$ if $\mathfrak{o}_1(D_y)$
agrees with $\mathfrak{o}_2(D_y)$, and $t(y)=-1$ otherwise. Since
$\mathfrak{o}_1$ and $\mathfrak{o}_2$ agree on all the empty periodic domains,
$t$ is a well-defined function. Furthermore, for any empty Maslov
index $1$ domain $D\in\mathcal{D}^0(x,y)$, the contribution
$c_{\mathfrak{o}_1}(D)$ coming from $\mathfrak{o}_1$ is related to the
contribution $c_{\mathfrak{o}_2}(D)$ coming from $\mathfrak{o}_2$ by the
equation $c_{\mathfrak{o}_1}(D)=t(x)t(y)c_{\mathfrak{o}_2}(D)$. That shows that
the two chain complexes are isomorphic via the map $x\mapsto t(x)x$.
For the second part of the theorem, recall the well known fact that
if two Heegaard diagrams $\mathcal{H}_1$ and $\mathcal{H}_2$ represent the
same link $L$, such that each component of the link has the same
number of $X$ and $O$ markings in both the Heegaard diagrams, then
they can be related to one another by a sequence of isotopies,
handleslides, and stabilizations. This essentially follows from
\cite[Proposistion 7.1]{POZSz} and \cite[Lemma
2.4]{CMPOSS}. However, during the isotopies, we do not require the
$\alpha$ circles to remain transverse to the $\beta$ circles.
Therefore, we can assume that $\mathcal{H}_1$ and $\mathcal{H}_2$ are related
by one of the following elementary moves: changing the path of
almost complex structures $J_s$ by an isotopy $J_{s,t}$; a
stabilization in a neighborhood of a marked point; a sequence of
isotopies and handleslides of the $\alpha$ circles in the complement of the
marked points; or a sequence of isotopies and handleslides of the
$\beta$ circles in the complement of the
marked points.
For the case of a stabilization, or an isotopy of the
path of almost complex structures, there is a natural identification
between $\mathcal{P}^0_{\mathcal{H}_1}$ and $\mathcal{P}^0_{\mathcal{H}_2}$, and a natural
identification of the determinant line bundles over the corresponding
empty periodic domains. Since a weak equivalence class of an
orientation system is determined by its values on the empty periodic
domains, this produces a natural identification between
$\widehat{\mathcal{O}}_{\mathcal{H}_1}$ and $\widehat{\mathcal{O}}_{\mathcal{H}_2}$. The proof that
the two homologies are isomorphic for the corresponding orientation
systems is immediate for the case of a stabilization, and follows from
the usual arguments of \cite{POZSz} for the other cases. We
do not encounter any new problems, since boundary degenerations are
still ruled out by the marked points.
For the remaining cases, namely, the case of isotopies and
handleslides of $\alpha$ circles or $\beta$ circles, the isomorphism is
established by counting holomorphic triangles. Let us assume that the
$\alpha$ circles are changed to the $\gamma$ circles by a sequence of
isotopies and handleslides in the complement of the marked points. Out
of the $2^{g+k-1}$ weak equivalence classes of orientation systems in
the Heegaard diagram $\mathcal{H}_3=(\Sigma,\gamma,\alpha,z,w)$, there is a
unique one $\mathfrak{o}_3$, for which the homology of $\mathcal{H}_3$ is
torsion-free. Each empty periodic domain in $\mathcal{H}_2$ can be written
uniquely as a sum of empty periodic domains in $\mathcal{H}_1$ and
$\mathcal{H}_3$. Therefore, we have a natural bijection between
$\widehat{\mathcal{O}}_{\mathcal{H}_1}$ and $\widehat{\mathcal{O}}_{\mathcal{H}_2}$: given an
orientation system $\mathfrak{o}\in\widehat{\mathcal{O}}_{\mathcal{H}_1}$, we can patch it
with $\mathfrak{o}_3$, to get an orientation system
$f(\mathfrak{o})\in\widehat{\mathcal{O}}_{\mathcal{H}_2}$. The triangle map, evaluated on
the top generator of the homology of $\mathcal{H}_3$, provides the required
isomorphism between the homology of $\mathcal{H}_1$ and the homology of
$\mathcal{H}_2$, for the corresponding orienation systems. The same proof
from \cite{POZSz} goes through without any problems since we do not encounter any boundary degenerations.
\end{proof}
Let $\vec{m}=(m_1,\ldots,m_l)$. The above theorem shows that
$H_*(\widehat{\mathit{CFL}}_{\mathcal{H}}(L,\mathbb{Z},\mathfrak{o}))$ is an invariant of the link $L$
inside the three-manifold, a choice of a weak equivalence class of an
orientation system $\mathfrak{o}$, and the vector $\vec{m}$. Let us
henceforth denote the homology as $\widehat{\mathit{HFL}}_{\vec{m}}(L,\mathbb{Z},\mathfrak{o})$. We
now investigate the dependence of
$\widehat{\mathit{HFL}}_{\vec{m}}(L,\mathbb{Z},\mathfrak{o})$ on $\vec{m}$.
\begin{thm}\label{thm:secondmain}
Let $\mathcal{H}$ be a Heegaard diagram for a link $L$, where the $i^{\text{th}}$
component $L_i$ is represented by $m_i$ $X$ markings and $m_i$ $O$
markings, and let $\mathcal{H}'$ be a Heegaard diagram for the same link,
where $L_i$ is represented by $m'_i=(m_i+\delta_{i_0 i})$ $X$
markings and $m'_i$ $O$ markings, for some fixed $i_0$. Then
there is a bijection $f$ between $\widehat{\mathcal{O}}_{\mathcal{H}}$ and
$\widehat{\mathcal{O}}_{\mathcal{H}'}$ such that for every weak equivalence class
of orientation system $\mathfrak{o}$, $\widehat{\mathit{HFL}}_{\vec{m}'}(L,\mathbb{Z},\mathfrak{o})$ is
isomorphic to $\widehat{\mathit{HFL}}_{\vec{m}}(L,\mathbb{Z},f(\mathfrak{o}))\otimes Q_{i_0}$ as
$(l+1)$-graded groups.
\end{thm}
\begin{proof}
Consider the Riemann sphere $S$ with one $\alpha$ circle and one
$\beta$ circle, intersecting each other at two points $p$ and $q$.
Put two $X$ markings, one $O$ marking and one $W$ marking, one in
each of the four elementary domains of
$S\setminus(\alpha\cup\beta)$, such that the boundary of either of
the two elementary domains that contain an $X$ marking runs from
$p$ to $q$ along the $\alpha$ circle, and from $q$ to $p$ along the
$\beta$ circle. Remove a small disk in the neighborhood of the point
$W$. In the Heegaard diagram $\mathcal{H}$, choose an $X$ marking that
lies in $L_{i_0}$, and remove a small disk in the neighborhood of
that point. Then connect the diagram $\mathcal{H}$ to the sphere $S$ via
the `neck' $S^1\times [0,T]$ to get a new Heegaard diagram for the
same link, where $L_i$ is represented by $m'_i$ $X$
markings, and $m'_i$ $O$ markings. This process is
shown in Figure \ref{fig:stabilization}. By Theorem \ref{thm:main},
we can assume that the new Heegaard diagram is $\mathcal{H}'$. There is a
natural correspondance between $\mathcal{P}^0_{\mathcal{H}}$ and
$\mathcal{P}^0_{\mathcal{H}'}$, and this induces the bijection $f$ between
$\widehat{\mathcal{O}}_{\mathcal{H}}$, and $\widehat{\mathcal{O}}_{\mathcal{H}'}$.
\begin{figure}
\psfrag{z}{$X$}
\psfrag{w}{$O$}
\psfrag{p}{$p$}
\psfrag{q}{$q$}
\psfrag{a}{$\alpha$}
\psfrag{b}{$\beta$}
\begin{center}
\includegraphics[width=\textwidth]{stabilization}
\end{center}
\caption{The Heegaard diagrams $\mathcal{H}$ and $\mathcal{H}'$.}\label{fig:stabilization}
\end{figure}
Fix $\mathfrak{o}\in\widehat{\mathcal{O}}_{\mathcal{H}}$. As $(l+1)$-graded groups,
$\widehat{\mathit{CFL}}_{\mathcal{H}'}(L,\mathbb{Z},\mathfrak{o})=\widehat{\mathit{CFL}}_{\mathcal{H}}(L,\mathbb{Z},f(\mathfrak{o}))\otimes(\mathbb{Z}\oplus\mathbb{Z})$,
where one $\mathbb{Z}$ corresponds to all the generators that contain the
point $p$, and has $(M,A_1,\ldots,A_l)$ multi-grading
$(0,0,\ldots,0)$, and the other $\mathbb{Z}$ corresponds to all the generators
that contain the point $q$, and has $(M,A_1,\ldots,A_l)$ multi-grading
$(-1,-\delta_{i_0 1},\ldots,-\delta_{i_0 l})$. We simply need to show
that the same identity holds as chain complexes. For this, it is
enough to show that there are no boundary maps from the generators
that contain the point $p$ to the generators that contain the point
$q$.
Following the arguments from \cite{POZSzlinkinvariants}, we extend the
`neck length' $T$, and move the point $W$ close to the $\alpha$ circle
in $S$. After choosing $T$ sufficiently large and $W$ sufficiently
close to the $\alpha$ circle, if there is an empty positive Maslov
index $1$ domain $D$, joining a generator containing $p$ to a
generator containing $q$, such that $c(D)\neq 0$, then $D$ must
correspond to a positive, Maslov index $2$ domain in $\mathcal{H}$ that
avoids all the $O$ markings and whose boundary lies entirely on the
$\alpha$ circles. However, any non-trivial domain in $\mathcal{H}$ whose
boundary lies entirely on the $\alpha$ circles must have non-zero
coefficients at some $O$ marking, thus producing a contradiction, and
thereby finishing the proof.
\end{proof}
Henceforth, denote $\widehat{\mathit{HFL}}_{(1,\ldots,1)}(L,\mathbb{Z},\mathfrak{o})$ by $\widehat{\mathit{HFL}}(L,\mathbb{Z},\mathfrak{o})$.
Theorems \ref{thm:main} and \ref{thm:secondmain} imply:
\begin{thm}\label{thm:fourthmain}
Let $\mathcal{H}$ be a Heegaard diagram for a link $L\subset S^3$ with
$l$ components, such that the $i^{\text{th}}$ component $L_i$ is represented
by exactly $m_i$ $X$ markings, and exactly $m_i$ $O$ markings. Then
the $2^{l-1}$ homology groups $\widehat{\mathit{HFL}}_{\vec{m}}(L,\mathbb{Z},\mathfrak{o})$ are isomorphic to
the $2^{l-1}$ groups
$\widehat{\mathit{HFL}}(L,\mathbb{Z},\mathfrak{o})\otimes_i(\otimes^{m_i-1}Q_i)$.
\end{thm}
We are almost done with the construction that we had set out to do. Given a link $L\subset S^3$ with $l$ components, we have produced $2^{l-1}$ $\mathbb{Z}$-valued $(l+1)$-graded homology groups $\widehat{\mathit{HFL}}(L,\mathbb{Z},\mathfrak{o})$. We would like to finish this section by showing that when we combine the $l$ Alexander gradings into one, then we get the $2^{l-1}$ $\mathbb{Z}$-valued bi-graded homology groups $\widehat{\mathit{HFK}}(L,\mathbb{Z},\mathfrak{o})$. Recall that the groups $\widehat{\mathit{HFK}}(L,\mathbb{Z},\mathfrak{o})$ are constructed by viewing the link $L\subset Y$ as a knot in $Y\#^{l-1}(S^1\times S^2)$, and then looking at the knot Floer homology. Therefore, the following lemma is all that we need.
\begin{thm}\label{thm:connectsum}
Let $\mathcal{H}$ be a Heegaard diagram for a link $L\subset Y$ with
$(l+1)$ components, such that each component is represented by one
$X$ and one $O$ marking. Let $\widetilde{L}$ be the link with $l$
components in $Y\#(S^1\times S^2)$, whose $l^{\text{th}}$ component
$\widetilde{L}_l$ is obtained by connect summing $L_{l+1}$ and $L_l$
through the one-handle, and let $\widetilde{\mathcal{H}}$ be a Heegaard diagram
for $\widetilde{L}$, where $\widetilde{L}_i$ is represented by $(1+\delta_{il})$
$X$ markings and $(1+\delta_{il})$ $O$ markings. Then, there is a
bijection $f$ between $\widehat{\mathcal{O}}_{\mathcal{H}}$ and
$\widehat{\mathcal{O}}_{\widetilde{\mathcal{H}}}$, such that for all
$\mathfrak{o}\in\widehat{\mathcal{O}}_{\mathcal{H}}$,
$H_*(\widehat{\mathit{CFL}}_{\widetilde{\mathcal{H}}}(\widetilde{L},\mathbb{Z},f(\mathfrak{o})))=\widehat{\mathit{HFL}}(L,\mathbb{Z},\mathfrak{o})\otimes
Q_l$ as $(l+1)$-graded groups, where the $(l+1)$ gradings on the
left hand side are $(M,A_1,\ldots,A_{l-1},A_l+A_{l+1})$.
\end{thm}
\begin{proof}
This proof is very similar to the proof of Theorem
\ref{thm:secondmain}. Once more, consider the Riemann sphere $S$
with one $\alpha$ circle and one $\beta$ circle, intersecting each
other at two points $p$ and $q$. Put two $X$ markings and two $W$
marking, one in each of the four elementary domains of
$S\setminus(\alpha\cup\beta)$, such that the boundary of either of
the two elementary domains that contain an $X$ marking runs from
$p$ to $q$ along the $\alpha$ circle, and from $q$ to $p$ along the
$\beta$ circle. Remove two small disks in the neighborhoods of the
$W$ markings. In the Heegaard diagram $\mathcal{H}$, remove two small
disks in the neighborhoods of the the two $X$ markings
that lie in $L_l$ and $L_{l+1}$. Then connect $\mathcal{H}$ to the sphere
$S$ via the two `necks,' $S^1\times[0,T_1]$ and $S^1\times[0,T_2]$,
as shown in Figure \ref{fig:connectsum}. The resulting picture is a
Heegaard diagram for the link $\widetilde{L}\subset Y\#(S^1\times S^2)$,
where the $i^{\text{th}}$ component $\widetilde{L}_i$ is represented by
$(1+\delta_{il})$ $X$ markings and $(1+\delta_{il})$ $O$ markings.
By the virtue of Theorem \ref{thm:main}, we can assume that this
Heegaard diagram is $\widetilde{\mathcal{H}}$.
\begin{figure}
\psfrag{z}{$X$}
\psfrag{p}{$p$}
\psfrag{q}{$q$}
\psfrag{a}{$\alpha$}
\psfrag{b}{$\beta$}
\begin{center}
\includegraphics[width=\textwidth]{connectsum}
\end{center}
\caption{The Heegaard diagrams $\mathcal{H}$ and $\widetilde{\mathcal{H}}$.}\label{fig:connectsum}
\end{figure}
An empty periodic domain in $\mathcal{H}$ gives rise to an empty periodic
domain in $\widetilde{\mathcal{H}}$. In the other direction, an empty periodic
domain in $\widetilde{\mathcal{H}}$ gives rise to a periodic domain in $\mathcal{H}$
which does not pass through any of the $O$ markings. Since each
component of the link $L$ is null-homologous in $Y$, such a periodic
domain is an empty periodic domain. Therefore, there is a natural
correspondance between the empty periodic domains of $\mathcal{H}$ and
$\widetilde{\mathcal{H}}$, and this induces the bijection $f$ between
$\widehat{\mathcal{O}}_{\mathcal{H}}$ and $\widehat{\mathcal{O}}_{\widetilde{\mathcal{H}}}$.
Fix $\mathfrak{o}\in\widehat{\mathcal{O}}_{\mathcal{H}}$. It
is immediate that as $(l+1)$-graded groups,
$\widehat{\mathit{CFL}}_{\widetilde{\mathcal{H}}}(\widetilde{L},\mathbb{Z},f(\mathfrak{o}))=\widehat{\mathit{CFL}}_{\mathcal{H}}(L,\mathbb{Z},\mathfrak{o})\otimes Q_l$.
However, quite like the case of Theorem \ref{thm:secondmain}, for
sufficiently large `neck lengths' $T_1$ and $T_2$, and with the two
$W$ markings sufficiently close to the $\alpha$ circle on $S$, the
above identity holds even as chain complexes.
\end{proof}
Before we conclude this section, a note regarding absolute gradings is
due. So far, we have worked with relative Maslov grading and relative
Alexander gradings. However, for links in $S^3$, and for links in
$\#^m (S^1\times S^2)$ that we obtain from links in $S^3$ by the
connect sum process described in Theorem \ref{thm:connectsum}, there
is a well-defined way to lift these gradings to absolute gradings, as
defined in \cite[Theorem 7.1]{POZSz4manifolds}, \cite[Subsections 3.3
and 3.4]{POZSzknotinvariants} and \cite[Lemma 4.6 and Equation
24]{POZSzlinkinvariants}. Since this is an oft-studied scenario, for
such links, let us improve the earlier theorems, and
henceforth work with absolute gradings.
\begin{lem}\label{lem:absolutegrading}
For links in $\#^m(S^1\times S^2)$ that come from links in $S^3$ by
the connect sum operation as decribed in Theorem
\ref{thm:connectsum}, the isomorphisms in Theorems \ref{thm:main},
\ref{thm:secondmain}, \ref{thm:fourthmain} and \ref{thm:connectsum}
preserve the absolute gradings.
\end{lem}
\begin{proof}
Recall that the isomorphisms in question come from chain maps that
preserve the relative gradings. Therefore, each such chain map must
shift each absolute grading by a fixed integer on the entire
chain complex. We want to show that each of these shifts is zero.
Since the absolute gradings are defined on the generators
themselves, this shift is unchanged if instead of working over $\mathbb{Z}$,
we tensor everything with $\mathbb{F}_2$ and work over $\mathbb{F}_2$. However,
since the Heegaard Floer homology of $\#^m(S^1\times S^2)$ is
non-trivial over $\mathbb{F}_2$, in each case, the homology of the entire
chain complex is non-trivial over $\mathbb{F}_2$. Furthermore, the
maps induced on the homology over $\mathbb{F}_2$ preserve the absolute gradings
\cite{POZSz4manifolds}, \cite{POZSzknotinvariants},
\cite{POZSzlinkinvariants}. Therefore, all the shifts are zero, and each
of the chain maps preserves all the gradings.
\end{proof}
\section{Grid diagrams}\label{sec:griddiagrams}
A \emph{planar grid diagram of index $N$} is the square
$S=[0,N]\times[0,N]\subset\mathbb{R}^2$, with the following additional
structures: if $1\leq i\leq N$, the horizontal line $y=(i-1)$ is
called $\alpha_i$, the $i^{\text{th}}$ $\alpha$ arc, and the vertical line
$x=(i-1)$ is called $\beta_i$, the $i^{\text{th}}$ $\beta$ arc; there are $2N$
markings, denoted by $X_1,\ldots,X_N,O_1,\ldots,O_N$, such that each
component of $S\setminus(\bigcup_i\alpha_i)$ contains one $X$ marking and
one $O$ marking, and each component of $S\setminus(\bigcup_i\beta_i)$
contains one $X$ marking and one $O$ marking.
A \emph{toroidal grid diagram of index $N$} is obtained from a planar
grid diagram of the same index by identifying the opposite sides of
the square $S$ to form a torus $T$. A careful reader will immediately
observe that this creates a Heegaard diagram $\mathcal{H}$ for some link
$L$ in $S^3$, and for the rest of the section, we will work with this
Heegaard diagram. The $\alpha$ arcs and the $\beta$ arcs become full
circles, and they are the $\alpha$ circles and the $\beta$ circles
respectively; the $N$ components of $T\setminus(\bigcup_i\alpha_i)$ are called
the \emph{horizontal annuli}, and each of them contains one $X$
marking and one $O$ marking; the horizontal annulus with $\alpha_i$ as
the circle on the bottom is called the $i^{\text{th}}$ horizontal annulus, and
is denoted by $H_i$; the $N$ components of $T\setminus(\bigcup_i\beta_i)$ are
called the \emph{vertical annuli}, and each of them also contains one
$X$ marking and one $O$ marking; the vertical annulus with $\beta_i$
as the circle on the left is called the $i^{\text{th}}$ vertical annulus,
and is denoted by $V_i$; the $N^2$ components of
$T\setminus\bigcup_i(\alpha_i\cup\beta_i)$ are the elementary
domains. Therefore, the link $L$ that the toroidal grid diagram
represents, can be obtained in the following way. We assume that the
toroidal grid diagram comes from a planar grid diagram on the square
$S$. Then in each component of $S\setminus(\bigcup_i\alpha_i)$, we join the $X$
marking to the $O$ marking by an embedded arc, and in each component
of $S\setminus(\bigcup_i\beta_i)$, we join the $O$ marking to the $X$ marking
by an embedded arc, and at every crossing, we declare the arc that
joins $O$ to $X$ to be the overpass. Henceforth, we also assume that
the link $L$ has $l$ components, and the $i^{\text{th}}$ component $L_i$ is
represented by $m_i$ $X$ markings and $m_i$ $O$ markings, and $\sum_i
m_i =N$.
There is only one $Spin^C$ structure, so generators in $\mathcal{G}_{\mathcal{H}}$
correspond to the permuatations in $\mathfrak{S}_N$ as follows: a generator
$x=(x_1,\ldots,x_N)\in\mathcal{G}_{\mathcal{H}}$ comes from the permutation
$\sigma\in\mathfrak{S}_n$, where $x_i=\alpha_i\cap\beta_{\sigma(i)}$ for
each $1\leq i\leq N$. The $N$ points $x_1,\ldots,x_N$ are called the
\emph{coordinates} of the generator $x$.
Let $\mathfrak{j}$ be the complex structure on $T$ induced from the standard
complex structure on $S\subset\mathbb{C}$, and let $J_s$ be the constant path
of almost complex structure $\text{Sym}^N(\mathfrak{j})$ on $\text{Sym}^N(T)$. After a
slight perturbation of the $\alpha$ and the $\beta$ circles, we can ensure
that $J_s$ achieves transversality for all domains up to Maslov index
two \cite[Lemma 3.10]{RL}. Henceforth, we work with these perturbed
$\alpha$ and $\beta$ circles and this path of nearly symmetric almost
complex structure.
Consider the $2^{l-1}$ chain complexes $\widehat{\mathit{CFL}}_{\mathcal{H}}(L,\mathbb{Z},\mathfrak{o})$. The boundary maps in each of the chain complexes correspond to objects called \emph{rectangles}. A rectangle $R$ joining a generator $x$ to a generator $y$ is a $2$-chain generated by the elementary domains of $\mathcal{H}$, such that the following conditions are satisfied: $R$ only has coefficients $0$ and $1$; the closure of the union of the elementary domains where $R$ has coefficient $1$ is a disk embedded in $T$ with four corners, or in other words, it looks like a rectangle; the top-right corner and the bottom-left corner of $R$ are coordinates of $x$; the top-left corner and the bottom-right corner of $R$ are coordinates of $y$; the generators $x$ and $y$ share $(N-2)$ coordinates; and $R$ does not contain any coordinates of $x$ or any coordinates of $y$ in its interior. It is easy to check that the rectangles are precisely the positive Maslov index one domains. We denote the set of all rectangles joining $x$ to $y$ by $\mathcal{R}(x,y)\subset \mathcal{D}(x,y)$. The set $\mathcal{R}(x,y)$ is empty unless $x$ and $y$ differ in exactly two coordinates, and even then, $\left|\mathcal{R}(x,y)\right|\leq 2$.
\begin{lem}\cite[Theorem 1.1]{CMPOSS}\label{lem:maslovone}
If $D\in\mathcal{D}(x,y)$ is a domain with $\mu(D)\leq 0$, then the
unparametrized moduli space $\widehat{\mathcal{M}_{J_s}}(D)$ is empty. If
$D\in\mathcal{D}(x,y)$ is a Maslov index one domain such that
$\widehat{\mathcal{M}_{J_s}}(D)$ is non-empty, then $D$ is a
rectangle. Conversely, if $R\in\mathcal{R}(x,y)$ is a rectangle, then
$\widehat{\mathcal{M}_{J_s}}(R)$ consists of exactly one point, and hence
$\left|c(R)\right|=1$.
\end{lem}
If $D\in\mathcal{D}(x,y)$, we say that $D$ can be decomposed as a sum of
two rectangles if there exists a generator $z\in\mathcal{G}_{\mathcal{H}}$ and
rectangles $R_1\in\mathcal{R}(x,z)$ and $R_2\in\mathcal{R}(z,y)$ such that
$D=R_1+R_2$. It is easy to check that the domains that can
be decomposed as sum of two rectangles are precisely the
positive Maslov index two domains. For any generator $x\in\mathcal{G}_T$,
there are exactly $2N$ Maslov index two positive domains in
$\mathcal{D}(x,x)$, namely the ones coming from the horizontal annuli
$H_1,\ldots,H_N$ and the vertical annuli $V_1,\ldots,V_N$.
\begin{lem}\label{lem:maslovtwo}
If $D\in\mathcal{D}(x,y)$ is a Maslov index two domain such that
$\widehat{\mathcal{M}_{J_s}}(D)$ is non-empty, then $D$ can be decomposed as a
sum of two rectangles. Conversely, if $D\in\mathcal{D}(x,y)$ can be
decomposed as a sum of two rectangles, then $\widehat{\mathcal{M}_{J_s}}(D)$
is a compact $1$-dimensional manifold with exactly two
endpoints. Furthermore, if $x=y$ (i.e. if $D$ comes from a
horizontal or a vertical annulus), then one of the endpoints
corresponds to the unique way of decomposing $D$ as a sum of two
rectangles, while the other endpoint corresponds to an $\alpha$ or a
$\beta$ boundary degeneration; and if $x\neq y$, then $D$ can be
decomposed as a sum of two rectangles in exactly two ways, and the
two endpoints correspond to the two decompositions.
\end{lem}
Lemma \ref{lem:maslovone} implies that once we choose an orientation system $\mathfrak{o}$ (and not just a weak equivalence class of orientation systems), we get a function $c_{\mathfrak{o}}$ from the set of all rectangles to $\{-1,1\}$. Lemma \ref{lem:maslovtwo} in conjunction with Lemma \ref{lem:index2} implies that if a domain $D\in\mathcal{D}(x,y)$ can be decomposed as a sum of two rectangles in two different ways $D=R_1+R_2=R_3+R_4$, then $c_{\mathfrak{o}}(R_1)c_{\mathfrak{o}}(R_2)=-c_{\mathfrak{o}}(R_3)c_{\mathfrak{o}}(R_4)$. This naturally leads to the definition of a sign assignment.
\begin{defn}
A \emph{sign assignment} $s$ is a function from the set of all rectangles to the set $\{-1,1\}$, such that the following condition is satisfied: if $x,y,z,z'\in\mathcal{G}_{\mathcal{H}}$ are distinct generators, and if $R_1\in\mathcal{R}(x,z)$, $R_2\in\mathcal{R}(z,y)$, $R'_1\in\mathcal{R}(x,z')$, $R'_2\in\mathcal{R}(z',y)$ are rectangles with $R_1+R_2=R'_1+R'_2$, then $s(R_1)s(R_2)=-s(R'_1)s(R'_2)$. Two sign assignments $s_1$ and $s_2$ are said to be \emph{gauge equivalent} if there is a function $t:\mathcal{G}_{\mathcal{H}}\rightarrow\{-1,1\}$, such that $s_1(R)=t(x)t(y)s_2(R)$, for all $x,y\in\mathcal{G}_{\mathcal{H}}$ and for all $R\in\mathcal{R}(x,y)$.
\end{defn}
In particular, a true sign assignment, as defined in \cite[Definition 4.1]{CMPOZSzDT}, is a sign assignment. Let $f$ be the map from the set of all orientation systems to the set of all sign assignments such that for all rectangles $R$, $f(\mathfrak{o})(R)=c_{\mathfrak{o}}(R)$. In this section, we will show that there are exactly $2^{2N-1}$ gauge equivalence classes of sign assignments on the grid diagram. We will put a weak equivalence on the sign assignments, which is weaker than the gauge equivalence. We will prove that there are exactly $2^{l-1}$ weak equivalence classes of sign assignments, and the map $f$ induces a bijection $\widetilde{f}$ between the set of weak equivalence classes of orientation systems and the set of weak equivalence classes of sign assignments. This will allow us to combinatorially calculate $\widehat{\mathit{CFL}}_{\mathcal{H}}(L,\mathbb{Z},\mathfrak{o})$ for all $\mathfrak{o}\in\widehat{\mathcal{O}}_{\mathcal{H}}$, and thereby calculate $\widehat{\mathit{HFL}}(L,\mathbb{Z})$ in all the $2^{l-1}$ versions. As a corollary, this will also show that any sign assignment (in particular, the one constructed in \cite{CMPOZSzDT}) computes $\widehat{\mathit{HFL}}(L,\mathbb{Z},\mathfrak{o})$ for some orientation system $\mathfrak{o}$.
We have an explicit (although slightly artificial) correspondance between the generators in $\mathcal{G}_{\mathcal{H}}$ and the elements of the symmetric group $\mathfrak{S}_N$, whereby a permutation $\sigma\in\mathfrak{S}_N$ gives rise to the generator $x=(x_1,\ldots,x_N)$ with $x_i=\alpha_i\cap\beta_{\sigma(i)}$. There is the following very natural partial order on the permutations: a reduction of a permutation $\tau$ is a permuation obtained by pre-composing $\tau$ by some transposition $(i,j)$ where $i<j$ and $\tau(i)>\tau(j)$; the permutation $\sigma$ is declared to be smaller than the permutation $\tau$, if $\sigma$ can be obtained from $\tau$ by a sequence of reductions. This induces a partial order $\prec$ on the elements of $\mathcal{G}_{\mathcal{H}}$.
For $x,y\in\mathcal{G}_{\mathcal{H}}$, if $y\prec x$ and there does not exist any $z\in\mathcal{G}_{\mathcal{H}}$ such that $y\prec z\prec x$, then we say that $x$ \emph{covers} $y$, and write that as $y\leftarrow x$. If we view the toroidal grid diagram as one coming from a planar grid diagram on $S=[0,N]\times[0,N]$, then $y\leftarrow x$ precisely when there is a rectangle from $x$ to $y$ contained in the subsquare $S'=[0,N-1]\times[0,N-1]$.
The poset $(\mathcal{G}_{\mathcal{H}},\prec)$ is a well-understood object \cite{shellPHE}. There is a unique minimum $p\in\mathcal{G}_{\mathcal{H}}$, which corresponds to the identity permutation. In particular, the Hasse diagram of $(\mathcal{G}_{\mathcal{H}},\prec)$, viewed as an unoriented graph, is connected. There is a unique maximum $q\in\mathcal{G}_{\mathcal{H}}$, which corresponds to the permutation that maps $i$ to $(N+1-i)$. The poset is shellable, which means that there is a total ordering $<$ on the maximal chains, such that if $\mathfrak{m}_1$ and $\mathfrak{m}_2$ are two maximal chains with $\mathfrak{m}_1<\mathfrak{m}_2$, then there exists a maximal chain $\mathfrak{m}_3<\mathfrak{m}_2$ with $\mathfrak{m}_1\cap\mathfrak{m}_2\subseteq\mathfrak{m}_3\cap\mathfrak{m}_2=\mathfrak{m}_2\setminus\{z\}$ for some $z\in\mathfrak{m}_2$. This in particular implies that given any two maximal chains $\mathfrak{m}_1$ and $\mathfrak{m}_2$, we can get from $\mathfrak{m}_2$ to $\mathfrak{m}_1$ via a sequence of maximal chains, where we get from one maximal chain to the next by changing exactly one element.
Given a sign assignment $s$ and a generator $x\in\mathcal{G}_{\mathcal{H}}$, we define two functions $h_{s,x},v_{s,x}:\{1,\ldots,N\}\rightarrow\{-1,1\}$, called the \emph{horizontal function} and the \emph{vertical function}, as follows: let $D\in\mathcal{D}(x,x)$ be Maslov index two positive domain which corresponds to the horizontal annulus $H_i$; then, $D$ can be decomposed as a sum of two rectangles in a unique way, and define the horizontal function $h_{s,x}(i)$ as the product of the signs of the two rectangles. The vertical function $v_{s,x}(i)$ is constructed similarly by considering the vertical annulus $V_i$ instead. Clearly, the horizontal and the vertical functions depend only on the gauge equivalence class of the sign assignment. The following theorem shows that the functions do not depend on the choice of the generator $x$, and will henceforth be denoted by $h_s$ and $v_s$.
\begin{thm}\label{thm:projection}
For any sign assignment $s$, for any two generators $x,y\in\mathcal{G}_{\mathcal{H}}$, and for any $1\leq i\leq N$, the horizontal and the vertical functions satisfy $h_{s,x}(i)=h_{s,y}(i)$ and $v_{s,x}(i)=v_{s,y}(i)$.
\end{thm}
\begin{proof}
Fix a sign assignment $s$, and fix $i\in\{1,\ldots,N\}$. We will only prove the statement for the vertical function; the argument for the horizontal function is very similar. Given $z\in\mathcal{G}_{\mathcal{H}}$, let $(z',R_z,R'_z)$ be the unique triple with $z'\in\mathcal{G}_T$, $R_z\in\mathcal{R}(z,z')$ and $R'_z\in\mathcal{R}(z',z)$ such that $R_z+R'_z\in\mathcal{D}(z,z)$ comes from the vertical annulus $V_i$. We simply want to show that for any two generators $x,y\in\mathcal{G}_{\mathcal{H}}$, $s(R_x)s(R'_x)=s(R_y)s(R'_y)$. Recall the partial order on $\mathcal{G}_{\mathcal{H}}$. The corresponding Hasse diagram, when viewed as an unoriented graph, is connected; therefore, it is enough to prove the above statement when $y\leftarrow x$. Thus, we can assume that there exists a rectangle $R\in\mathcal{R}(x,y)$. We end the proof by considering the following two cases.
\begin{figure}
\begin{center}
\includegraphics[height=170pt]{projection}
\end{center}
\caption{The case when $y$ and $x'$ disagree in exactly $3$ or exactly $4$ coordinates. The coordinates of $x$, $y$, $x'$ and $y'$ are denoted by white circles, black circles, white squares and black squares, respectively.}\label{fig:projection}
\end{figure}
\emph{The generators $y$ and $x'$ disagree on none of the coordinates.} In this case, $y=x'$, $y'=x$, $R_x=R'_y$ and $R_y=R'_x$. The equality $s(R_x)s(R'_x)=s(R_y)s(R'_y)$ follows trivially.
\emph{The generators $y$ and $x'$ disagree on exactly three or exactly four coordinates.} In this case, there exists a rectangle $R'\in\mathcal{R}(x',y')$, such that $R_x+R'=R+R_y\in\mathcal{D}(x,y')$ and $R'_x+R=R'+R'_y\in\mathcal{D}(x',y)$. The three essentially different types of diagrams that might appear (up to a rotation by $180^{\circ}$) are illustrated in Figure \ref{fig:projection}. Therefore, $s(R_x)s(R')=-s(R)s(R_y)$ and $s(R'_x)s(R)=-s(R')s(R'_y)$. Multiplying, we get the required identity $s(R_x)s(R'_x)=s(R_y)s(R'_y)$.
\end{proof}
The following two theorems will establish that there are exactly $2^{2N-1}$ gauge equivalence classes of sign assignments. Let $\Phi$ be the map from the set of gauge equivalence classes of sign assignments to $\{-1,1\}^{2N-1}$ given by $s\rightarrow (h_s(1),\ldots,\allowbreak h_s(N),\allowbreak v_s(1),\ldots,v_s(N-1))$.
\begin{thm}\label{thm:equivalence}
Given functions $g_h,g_v:\{1,\ldots,N\}\rightarrow\{-1,1\}$, such that $\left|g^{-1}_v(1)\right|\equiv \left|g^{-1}_h(-1)\right|\pmod{2}$, there exists a sign assignment $s$, such that $g_h=h_s$ and $g_v=v_s$. Therefore, in particular, the function $\Phi$ from the set of gauge equivalence classes of sign assignments to $\{-1,1\}^{2N-1}$ is surjective.
\end{thm}
\begin{proof}
By \cite[Theorem 4.2]{CMPOZSzDT}, there exists a sign assignment $s_0$ such that $h_{s_0}(i)=1$ and $v_{s_0}(i)=-1$ for all $i\in\{1,\ldots,N\}$. Given $g_h,g_v:\{1,\ldots,N\}\rightarrow\{-1,1\}$ with $\left|g^{-1}_v(1)\right|\equiv \left|g^{-1}_h(-1)\right|\pmod{2}$, we would like to modify $s_0$ to get $s$, such that $g_h=h_s$ and $g_v=v_s$.
The general method that we employ to modify a sign assignment $s_1$ to get another sign assignment $s_2$, is the following: we start with a multiplicative $2$-cochain $m$ which assigns elements of $\{-1,1\}$ to the elementary domains; if $D$ is a $2$-chain generated by the elementary domains, then $\langle m,D\rangle$ is simply the evaluation of $m$ on $D$; then, for a rectangle $R\in\mathcal{R}(x,y)$, we define $s_2(R)$ to be $s_1(R)\langle m,R\rangle$. It is easy to see that $s_2$ is a sign assignment if and only if $s_1$ is a sign assignment.
We prove the statement by an induction on the number $n(g_v,g_h)=\frac{1}{2}(\left|g^{-1}_v(1)\right|+\left|g^{-1}_h(-1)\right|)$. For the base case, when $n(g_v,g_h)=0$, we can simply choose $s=s_0$.
Assuming that the induction hypothesis is proved for $n=k$, let $g_h,g_v:\{1,\ldots,N\}\allowbreak\rightarrow\{-1,1\}$ be functions with $n(g_v,g_h)=k+1$. Choose functions $\widetilde{g}_h,\widetilde{g}_v:\{1,\ldots,N\}\allowbreak\rightarrow\{-1,1\}$ such that $n(\widetilde{g}_v,\widetilde{g}_h)=k$ and $\left|\{i\mid g_v(i)\neq\widetilde{g}_v(i)\}\right|+\left|\{i\mid g_h(i)\neq\widetilde{g}_h(i)\}\right|=2$. By induction, there is a sign assignment $\widetilde{s}$ such that $\widetilde{g}_h=h_{\widetilde{s}}$ and $\widetilde{g}_v=v_{\widetilde{s}}$. If $\left|\{i\mid g_v(i)\neq\widetilde{g}_v(i)\}\right|=2$, consider the two vertical annuli corresponding to the two values where $g_v$ disagrees with $\widetilde{g}_v$, choose a horizontal annulus, and let $m$ be the $2$-cochain which assigns $(-1)$ to the two elementary domains where the horizontal annulus intersects the two vertical annuli, and $1$ to every other elementary domain. Similarly, if $\left|\{i\mid g_h(i)\neq\widetilde{g}_h(i)\}\right|=2$, consider the two horizontal annuli corresponding to the two values where $g_h$ disagrees with $\widetilde{g}_h$, choose a vertical annulus, and let $m$ be the $2$-cochain which assigns $(-1)$ to the two elementary domains where the vertical annulus intersects the two horizontal annuli, and $1$ to every other elementary domain. Finally, if $\left|\{i\mid g_v\neq\widetilde{g}_v(i)\}\right|=\left|\{i\mid g_h\neq\widetilde{g}_h(i)\}\right|=1$, consider the vertical annulus corresponding to the value where $g_v$ disagrees with $\widetilde{g}_v$, consider the horizontal annulus corresponding to the value where $g_h$ disagrees with $\widetilde{g}_h$, and let $m$ be the $2$-cochain which assigns $(-1)$ to the elementary domain where the vertical annulus intersects the horizontal annulus, and $1$ to every other elementary domain. Let $s$ be the sign assignment obtained from $\widetilde{s}$ by modifying it by the $2$-cochain $m$. It is fairly straightforward to check that $g_h=h_s$ and $g_v=v_s$.
\end{proof}
\begin{thm}\label{thm:uniqueness}
The function $\Phi$ from the set of gauge equivalence classes of sign assignments to $\{-1,1\}^{2N-1}$ is injective.
\end{thm}
\begin{proof}
For this proof, we will closely follow the corresponding proof from \cite{CMPOZSzDT}. However, that proof uses the permutahedron whose $1$-skeleton is the Cayley graph of the the symmetric group, where the generators are the adjacent transpositions. In our proof, we will use a different simplicial complex, which is the order complex of the partial order $\prec$ on $\mathcal{G}_{\mathcal{H}}$.
Recall that the poset has a unique minimum $p$, and a unique maximum $q$. View the Hasse diagram of the poset as an oriented graph $\mathfrak{g}$. Choose a maximal tree $\mathfrak{t}$ with $p$ as a root, i.e. given any vertex $x$, there is a (unique) oriented path from $p$ to $x$ in $\mathfrak{t}$. The edges of $\mathfrak{g}$ correspond to the rectangles that are supported in $[0,N-1]\times[0,N-1]$. A sign assignment endows the edges of $\mathfrak{g}$ with signs $\pm 1$.
Let us choose a $(2N-1)$-tuple in $\{-1,1\}^{2N-1}$, and let $s$ be a sign assignment such that the $(2N-1)$-tuple equals $\Phi(s)$. We would like to show that the gauge equivalence class of $s$ is determined. Since $\mathfrak{t}$ is a tree, by replacing the sign assignment $s$ by a gauge equivalent one if necessary, we can assume that $s$ labels all the edges of $\mathfrak{t}$ with $1$'s. We will show that the values of $s$ on all the other edges are now determined.
Now consider any other edge $y\leftarrow x$ in $\mathfrak{g}$. Let $\mathfrak{c}_1$ be the unique oriented path from $p$ to $x$ in $\mathfrak{t}$, and let $\mathfrak{c}_2$ be the unique oriented path from $p$ to $y$ in $\mathfrak{t}$. Choose an oriented path $\mathfrak{c}_0$ from $x$ to $q$ in $\mathfrak{g}$. Let $\mathfrak{m}_1$ be the union of $\mathfrak{c}_1$ and $\mathfrak{c}_0$, and let $\mathfrak{m}_2$ be the union of $\mathfrak{c}_2$, the edge from $y$ to $x$, and $\mathfrak{c}_0$; these can be seen as maximal chains in $(\mathcal{G}_{\mathcal{H}},\prec)$. Clearly, $($the product of the signs on the edges in $\mathfrak{m}_1)\cdot($the product of the signs on the edges in $\mathfrak{m}_2)=($the product of the signs on the edges in $\mathfrak{c}_1)\cdot($the product of the signs on the edges in $\mathfrak{c}_2)\cdot($the sign on the edge from $y$ to $x)$. Since $\mathfrak{c}_1\cup\mathfrak{c}_2\subseteq\mathfrak{t}$, the signs on the edges of $\mathfrak{c}_1$ and $\mathfrak{c}_2$ are all $1$, so the sign on the edge from $y$ to $x$ equals $($the product of the signs on the edges in $\mathfrak{m}_1)\cdot($the product of the signs on the edges in $\mathfrak{m}_2)$. Since $(\mathcal{G}_{\mathcal{H}},\prec)$ is shellable, $\mathfrak{m}_2$ can be turned into $\mathfrak{m}_1$ through maximal chains by modifying one element at a time. Changing exactly one element of exactly one of the maximal chains negates the above product, so the product depends only on the graph $\mathfrak{g}$. Thus, $s$ is determined on all the edges of $\mathfrak{g}$.
Therefore, we have shown that there exists at most one sign assignment, up to gauge equivalence, on the rectangles that lie in the subsquare $S'=[0,N-1]\times[0,N-1]$. In fact, shellability of our poset also implies that there exists a sign assignment, but we do not need it. The rest of the proof for uniqueness is very similar to the proof from \cite{CMPOZSzDT}, but for the reader's convenience, we repeat the argument. Let $S''\subset T$ be the annular subspace corresponding to the rectangle $[0,N-1]\times[0,N]$ in the planar grid diagram. Next, we show that the value of $s$ is determined on all the rectangles that lie in $S''$.
This is done by an induction on the (horizontal) width of the rectangles. For the base case, if $R\in\mathcal{R}(x,y)$ is a rectangle of width one which is not supported in $S'$, then let $R'\in\mathcal{R}(y,x)$ be the unique rectangle such that $R+R'$ is a vertical annulus. The vertical function $v_s$ determines the product of the signs $s(R)s(R')$, and thereby the sign $s(R)$.
\begin{figure}
\begin{center}
\includegraphics[height=170pt]{uniqueness}
\end{center}
\caption{The induction step. The coordinates of $x$, $y$,
$y'$ and $z$ are denoted by white circles, white squares, black
squares and black circles, respectively.}\label{fig:uniqueness}
\end{figure}
Assuming that we have proved the uniqueness of sign assignments for all
the rectangles up to width $k$, let $R\in\mathcal{R}(x,y)$ be a width
$(k+1)$ rectangle. Let $R_1\in\mathcal{R}(y,z)$ be the width one rectangle
such that the bottom-left corner of $R_1$ is the top-left corner of
$R$. Then there exists a generator $y'\neq y$, a width one rectangle
$R'\in\mathcal{R}(x,y')$ and a width $k$ rectangle $R'_1\in\mathcal{R}(y',z)$,
such that $R+R_1=R'+R'_1\in\mathcal{D}(x,z)$. The situation is illustrated
in Figure \ref{fig:uniqueness}. By induction, the value of
$s$ is determined on $R_1$, $R'$ and $R'_1$. However,
$s(R)s(R_1)=-s(R')s(R'_1)$, and this determines the sign $s(R)$. This
completes the induction and shows that the value of the sign
assignment $s$ is fixed on all the rectangles that are supported
in $S''$. A similar argument, but with the diagrams rotated by
$90^{\circ}$, shows that the value of $s$ is, in fact, determined on
all the rectangles. This completes the proof of uniqueness.
\end{proof}
\begin{lem}\label{lem:product}
For any sign assignment $s$, the product $\prod_{i=1}^N h_s(i)v_s(i)$ equals $(-1)^N$.
\end{lem}
\begin{proof}
By Theorem \ref{thm:equivalence}, there exists a sign assignment $s'$ such that $h_{s'}=h_{s}$, $v_{s'}(i)=v_s(i)$ for $i\in\{1,\ldots,N-1\}$ and $v_{s'}(N)= (-1)^N h_s(N)\prod_{i=1}^{N-1} h_s(i)v_s(i)$. Since $\Phi(s)=\Phi(s')$, by Theorem \ref{thm:uniqueness}, $s$ and $s'$ are gauge equivalent. Therefore, $\prod_{i=1}^N h_s(i)v_s(i)=\prod_{i=1}^N h_{s'}(i)v_{s'}(i)=(-1)^N$.
\end{proof}
Fix a sign assignment $s$ and fix a link component $L_i$. Let $V(L_i)=\{j\mid \text{the }X\text{ marking in }V_j\text{ is in }L_i\}$ and let $H(L_i)=\{j\mid \text{the }X\text{ marking in }H_j\text{ is in }L_i\}$. The product $(\prod_{j\in H(L_i)}h_s(j))(\prod_{j\in V(L_i)}(-v_s(j)))$ is defined to be the \emph{sign of the link component $L_i$} and is denoted by $r_s(L_i)$.
Call two sign assignments $s_1$ and $s_2$ \emph{weakly equivalent} if $r_{s_1}$ agrees with $r_{s_2}$ on each of the link components. Clearly, if two sign assignments are gauge equivalent, then they are weakly equivalent. Due to Lemma \ref{lem:product}, the product of the signs of all the link components is $1$, and this is the only restriction on these numbers $r_s(L_i)$. Therefore, there are exactly $2^{l-1}$ weak equivalence classes of sign assignments. The following observation yields a direct proof that the chain complex $\widehat{\mathit{CFL}}_{\mathcal{H}}(L,\mathbb{Z},\mathfrak{o})$ depends only on the weak equivalence class of the sign assignment $f(\mathfrak{o})$.
\begin{lem}
If two sign assignments $s_1$ and $s_2$ are weakly equivalent, then there exists a sign assignment $s'_2$, which is gauge equivalent to $s_2$, such that $s_1$ and $s'_2$ agree on all the rectangles that avoid the $X$ markings and the $O$ markings.
\end{lem}
\begin{proof}
Since $s_1$ and $s_2$ are weakly equivalent, a proof similar to the proof of Theorem \ref{thm:equivalence} shows that there exists a $2$-cochain $m$ which assigns $1$ to every elementary domain that does not contain any $X$ or $O$ markings, such that the sign assignment $s'_2$ obtained by modifying $s_1$ by the $2$-cochain $m$ satisfies
$h_{s_2}=h_{s'_2}$ and $v_{s_2}=v_{s'_2}$. Therefore, by Theorem \ref{thm:uniqueness}, $s'_2$ is gauge equivalent to $s_2$.
\end{proof}
\begin{thm}
The map $f$ from the set of orientation systems to the set of sign assignments induces a well-defined bijection $\widetilde{f}$ from the set of weak equivalence classes of orientation systems to the set of weak equivalence classes of sign assignments.
\end{thm}
\begin{proof}
Recall that two orientation systems $\mathfrak{o}_1$ and $\mathfrak{o}_2$ are weakly equivalent if and only if, for a fixed generator $x\in\mathcal{G}_{\mathcal{H}}$, $\mathfrak{o}_1$ agrees with $\mathfrak{o}_2$ on all the domains in $\mathcal{D}(x,x)$ that correspond to the empty periodic domains of $\mathcal{P}^0_{\mathcal{H}}$. Therefore, we need to find a basis for the empty periodic domains.
For each $i\in\{1,\ldots,l\}$, let $P_i=\sum_{j\in V(L_i)}V_j-\sum_{j\in H(L_i)}H_j$. These $l$ empty periodic domains generate $\mathcal{P}^0_{\mathcal{H}}$, and $\sum_i P_i=0$ is the only relation among these domains. Therefore, the domains $P_1,\ldots,P_{l-1}$ freely generate $\mathcal{P}^0_{\mathcal{H}}$.
If $D\in\mathcal{D}(x,x)$ is a domain which corresponds to a vertical annulus $V_i$, then we know from Paragraph \ref{para:special} that $\mathfrak{o}_1$ agrees with $\mathfrak{o}_2$ on $D$ if and only if $v_{f(\mathfrak{o}_1)}(i) = v_{f(\mathfrak{o}_2)}(i)$. A similar statement holds for the horizontal annuli. A repeated application of the same principle shows that if $D\in\mathcal{D}(x,x)$ corresponds to the empty periodic domain $P_i$, then $\mathfrak{o}_1$ agrees with $\mathfrak{o}_2$ on $D$ if and only if $r_{f(\mathfrak{o}_1)}(L_i)=r_{f(\mathfrak{o}_1)}(L_i)$. Therefore, the orientation systems $\mathfrak{o}_1$ and $\mathfrak{o}_2$ are weakly equivalent if and only if the sign assignments $f(\mathfrak{o}_1)$ and $f(\mathfrak{o}_2)$ are weakly equivalent. This shows that the map in question is well-defined and injective. As both sets have $2^{l-1}$ elements, it is a bijection.
\end{proof}
A consequence of the theorems in this section is the following.
\begin{thm}
There is a bijection $\widetilde{f}$ between the weak equivalence classes of orientation systems and the weak equivalence classes of sign assignments, such that for each of the $2^{l-1}$ weak equivalence classes of orientation systems $\mathfrak{o}$, the homology of the grid chain complex, evaluated with the sign assignment $f(\mathfrak{o})$, is isomorphic as an absolutely $(l+1)$-graded group to $\widehat{\mathit{HFL}}(L,\mathbb{Z},\mathfrak{o})\otimes_i(\otimes^{m_i-1}Q_i)$.
\end{thm}
\begin{figure}
\psfrag{x}{$X$}
\psfrag{o}{$O$}
\psfrag{xo}{$XO$}
\begin{center}
\includegraphics[width=0.8\textwidth]{grid}
\end{center}
\caption{Grid diagrams for the two-component unlink and the Hopf link.}\label{fig:grid}
\end{figure}
Let us conclude with a couple of examples. The first grid diagram in Figure \ref{fig:grid} represents the two-component unlink. There are exactly two generators and exactly two rectangles connecting the two generators. One weak equivalence class assigns the same sign to both the rectangles while the other weak equivalence class assigns opposite signs. Therefore, for one weak equivalence class of orientation systems, the homology is $\mathbb{Z}/2\mathbb{Z}$, while for the other other weak equivalence class of orientation systems, the homology is $\mathbb{Z}\oplus\mathbb{Z}$.
The second grid diagram in Figure \ref{fig:grid} represents the Hopf link. There are twenty-four generators and sixteen rectangles. It can be checked by direct computation that the homology is independent of the sign assignment. Therefore, the link Floer homology of the Hopf link is the same for both the weak equivalence classes of orientation systems.
\bibliographystyle{amsplain}
|
\section{Introduction}
This paper considers the growth rates for Hill's equation with
parameters that vary from cycle to cycle. In this context, Hill's
equation takes the form
\begin{equation}
{d^2 y \over dt^2} + [ \af_k + q_k \qhat (t) ] y = 0 \, ,
\label{basic}
\end{equation}
where the barrier shape function $\qhat(t)$ is periodic, so that
$\qhat (t + \period) = \qhat(t)$, where $\period$ is the period. Here
we take $\period = \pi$, and the function $\qhat$ is normalized so
that $\int_0^\period \qhat dt$ = 1. The forcing strength parameters
$q_k$ are a set of independent identically distributed (i.i.d.) random
variables that take on a new value every cycle (where the index $k$
labels the cycle). The parameters $\af_k$, which determine the
oscillation frequency in the absence of forcing, also vary from cycle
to cycle (and are i.i.d.). In principal, the cycle interval $\period$
could also vary; however, this generalized case can be reduced to the
problem of equation (\ref{basic}) through an appropriate re-scaling of
the other parameters (see Theorem 1 of [AB]).
Hill's equations [HI] with constant values of the parameters have been
well studied and arise in a wide variety of applications [MW]. The
introduction of parameters that sample a distribution of values is
thus a natural generalization of this classic problem. Here we refer
to the case with constant parameters as the ``classical regime'' of
the general case.
For this class of periodic differential equations, the transformation
that maps the coefficients of the principal solutions from one cycle
to the next takes the form
\begin{equation}
{\mfont M}_k = \left[
\matrix{h_k & (h_k^2 - 1)/g_k \cr g_k & h_k} \right] \, ,
\label{mapzero}
\end{equation}
where the subscript denotes the cycle. The matrix elements are defined
by $h_k = y_1 (\pi)$ and $g_k = {\dot y}_1 (\pi)$ for the $kth$ cycle,
where $y_1$ and $y_2$ are the principal solutions for that cycle. Note
that the matrix has only two independent elements rather than four:
Since the Wronskian of the original differential equation
(\ref{basic}) is unity, the determinant of the matrix map must be
unity, and this constraint eliminates one of the independent elements.
In addition, this paper specializes to the case where the periodic
functions $\qhat(t)$ are symmetric about the midpoint of the period,
so that $y_1(\pi) = {\dot y}_2 (\pi)$, which eliminates a second
independent element [MW]; this symmetry applies to the applications
that motivated this work.
For transformation matrices ${\mfont M}_k$ of the form
(\ref{mapzero}), the eigenvalues $\lambda_k$ can be used to classify
the matrix types [LR]. The characteristic polynomial has the form
\begin{equation}
\lambda_k^2 - 2 h_k \lambda_k + 1 = 0 \, .
\end{equation}
This equation allows for three classes of eigenvalues $\lambda_k$: For
$|h_k| > 1$, the eigenvalues are real and have the same sign, and the
transformation matrix is hyperbolic symplectic; we denote this regime
as classically unstable. When $|h_k| < 1$, the eigenvalues are
complex and the matrix is elliptic; this regime is denoted as
classically stable. The remaining possibility is for $|h_k| = 1$,
which leads to degenerate eigenvalues equal to either $+1$ or $-1$;
these matrices are parabolic and are stable under multiplication.
This paper studies the multiplication of infinite strings of random
matrices of the form (\ref{mapzero}), i.e., the product of $N$ such
matrices in the limit $N \rightarrow \infty$. The problem of finding
growth rates for infinite products of matrices with random elements
was formulated over four decades ago [FU, FK], where existence results
were given. We recall the key result here for convenience:
\noindent
For a $k\times k$ matrix $A$ with real or complex entries, let $||A||$
denote the Frobenius norm.
\noindent
{\bf Theorem (FK):}
Let $X^1, X^2, X^3,\dots$ form a metrically transitive stationary
stochastic process with values in the set of $k\times k$ matrices.
Suppose $\log^+||X^1||$ exists, where ${\log}^+ t=\max (\log t,0)$,
then the limit $\lim_{N\rightarrow\infty}||X^NX^{N-1}\cdots X^1||$ exists.
Determination of the growth rates are thus carried out in the limit of
large $N$, and all probabilistic limits given here are meant almost
surely.
A great deal of subsequent work has studied differential equations of
the form (\ref{basic}) and the growth rates of the corresponding
random matrices [CL, PF, LGP]. In spite of this progress, there are
relatively few examples that provide explicit expressions for the
growth rates. The goal of this paper is relatively modest: It provides
(what we believe to be) new analytic expressions for the growth rates
of random matrices of the form (\ref{mapzero}). These expressions are
derived for various regimes of parameter space, as described below.
The outline of this paper is as follows: Section 2 reviews the
astrophysical background that led us to this topic. Section 3
considers matrix multiplication for the case where the solutions are
unstable in the classical regime. Section 4 develops approximations
for this regime and provides some numerical verification. Section 5
considers matrix multiplication in the regime where the solutions are
classically stable. In this case, the transformation matrices
$\mfont{M}_k$ correspond to elliptical rotations and matrix
multiplication is stable in the absence of fluctuations; random
variations in the matrix elements render the solutions unstable. The
paper concludes (in Section 6) with a brief summary of the results.
\section{Astrophysical Background}
The motivation for considering random Hill's equations arose in
studies of orbit problems in astrophysics [AK]. When an orbit starts
in the principal plane of a triaxial, extended mass distribution (such
as a dark matter halo), the motion is unstable to perturbations in the
perpendicular direction. The development of the instability is
described by a random Hill's equation with the form given by equation
(\ref{basic}).
To illustrate this type of behavior, consider an extended mass
distribution with a density profile of the form
\begin{equation}
\rho = {\rho_0 \over m} \qquad {\rm with} \qquad
m^2 = {x^2 \over a^2} + {y^2 \over b^2} + {z^2 \over c^2} ,
\label{halodef}
\end{equation}
where $\rho_0$ is a density scale. This form arises in many different
astrophysical contexts, including dark matter halos, galactic bulges,
and young embedded star clusters. The density field is thus constant
on ellipsoids, where, without loss of generality, $a > b > c > 0$.
For this density profile, one can find analytic forms for both the
gravitational potential and the force terms [AK]. From these results,
one can determine the orbital motion for a test particle moving in the
potential resulting from the triaxial density distribution of equation
(\ref{halodef}). When the orbit begins in any of the three principal
planes, the motion is generally unstable to perturbations in the
perpendicular direction [AB, AK]. For example, for an orbit initially
confined to the $x$-$z$ plane, the amplitude of the $y$ coordinate
will (usually) grow exponentially with time. In the limit of small
$|y| \ll 1$, the equation of motion for the perpendicular coordinate
simplifies to the form
\begin{equation}
{d^2 y \over dt^2} + \omega_y^2 y = 0
\qquad {\rm where} \qquad \omega_y^2 =
{ 4/b \over \sqrt{c^2 x^2 + a^2 z^2} + b \sqrt{x^2 + z^2} } \ .
\label{omegay}
\end{equation}
The time evolution of the coordinates $(x,z)$ is determined by the
orbit in the original $x$-$z$ plane. Since the orbital motion is
nearly periodic, the $[x(t),z(t)]$ dependence of $\omega_y^2$
represents a nearly periodic forcing term. The forcing strengths, and
hence the parameters $q_k$ appearing in Hill's equation (\ref{basic}),
are determined by the inner turning points of the orbit (with appropriate
weighting from the axis parameters $[a,b,c]$). Since the orbits are
usually chaotic, the distance of closest approach, and hence the
strength $q_k$ of the forcing, varies from cycle to cycle. The outer
turning points of the orbit provide a minimum value of $\omega_y^2$,
which defines the unforced oscillation frequency $\af_k$ appearing in
Hill's equation. As a result, the quantity $\omega_y^2$ can be written
in the form
\begin{equation}
\omega_y^2 = \af_k + Q_k (t) \, ,
\end{equation}
where the index $k$ counts the number of orbit crossings. The shapes
of the functions $Q_k$ are nearly the same, so that one can write
$Q_k$ = $q_k {\hat Q}(t)$, where ${\hat Q}(t)$ is periodic. The
chaotic orbit in the original plane leads to different values of
$\af_k$ and $q_k$ for each crossing. The equation of motion
(\ref{omegay}) for the $y$ coordinate thus takes the form of Hill's
equation (\ref{basic}), where the period, forcing strength, and
oscillation frequency vary from cycle to cycle.
\section{Matrix Multiplication for the Classically Unstable Regime}
The goal of this work is to find growth rates for solutions of the
differential equation (\ref{basic}). These growth rates are determined
by multiplication of the random matrices ${\mfont M}_k$ (from equation
[\ref{mapzero}]) that connect solutions from cycle to cycle. These
transformation matrices can also be written in the form
\begin{equation}
{\mfont M}_k = h_k {\mfont B}_k \qquad {\rm where} \qquad
{\mfont B}_k = \left[ \matrix{1 & x_k \phi_k \cr {1/x_k} & 1} \right] \, ,
\label{mbdefine}
\end{equation}
where $x_k$ = $h_k/g_k$ and $\phi_k$ = $1 - 1/h_k^2$. By virtue of
our assumption on the variables ($q_k$, $\af_k$), the matrices
${\mfont M}_k$ form a sequence of i.i.d. matrices. In this section,
we consider the problem of matrix multiplication with matrices of the
form (\ref{mbdefine}). We specialize to the case where the solutions
are unstable in the classical regime so that $|h_k| \ge 1$ and to the
case where $x_k > 0$. We also assume that the $h_k$, $x_k$, and
$1/x_k$ have finite means. With the matrices written in the form
(\ref{mbdefine}), the highly unstable regime considered in [AB] can be
defined as follows:
\noindent
{\bf Definition:} Given that solutions to Hill's equation
(\ref{basic}) are determined by transformation matrices of the form
(\ref{mbdefine}), the {\it highly unstable regime} is defined by
setting $\phi_k = 1$. This specification thus defines a restricted problem.
We remark that the above regime applies when the matrix elements
$|h_k| \gg 1$, which occurs for forcing strength parameters
$q_k \gg 1$ [AB2].
The growth rates for Hill's equation (\ref{basic}) are determined
by the growth rates for matrix multiplication of the full set of
matrices ${\mfont M}_k$. For a given matrix product, denoted here
as ${\mfont M}^{(N)}$, the {\it growth rate} $\gamma$ is determined by
\begin{equation}
\gamma = \lim_{N \to \infty} {1 \over N}
\log || {\mfont M}^{(N)} || \, ,
\label{growbasic}
\end{equation}
where the result is independent of the choice of norm $|| \cdot ||$.
We note that the growth rate is called the {\it top} or
{\it largest Lyapunov exponent}.
Equation (\ref{mbdefine}) separates the growth rate for this problem
into two parts. Let the expectation value of a sequence $X_k$
be denoted by
$$\langle X_k\rangle=\lim_{N \to \infty}
{1 \over N} \sum_{k=1}^N X_k$$
Then the first part $\gamma_h$ of the growth rate is given
by
\begin{equation}
\gamma_h = \lim_{N \to \infty} {1 \over N} \sum_{k=1}^N \log |h_k|
=\langle\log |h_k|\rangle \, .
\end{equation}
We limit our discussion to distributions of the $h_k$ for which this
limit is finite. The remaining part of the growth rate is determined
by matrix multiplication of the ${\mfont B}_k$. Note that the
original differential equation (\ref{basic}) is defined on a time
interval $0 \le t \le \pi$, so that the definition of its growth rate
includes a factor of $\pi$ [MW], whereas the growth rate for matrix
multiplication (\ref{growbasic}) generally does not [FK]. Ignoring
these normalization issues, this paper focuses on the calculation of
the growth rates for the matrices ${\mfont M}_k$ and ${\mfont B}_k$.
The product of $N$ matrices of type ${\mfont B}_k$
can be written in the form
\begin{equation}
{\mfont B}^{(N)} \equiv
\prod_{k=1}^N {\mfont B}_k = \left[ \matrix{\siga & x_1 \sigb \cr
(1 / x_1) \sigc & \sigd } \right] \, ,
\label{product}
\end{equation}
where the first equality defines notation and where
$$
\siga = \sum_{j=1}^{2^{N-1}} r_j a_j \, , \qquad
\sigb = \sum_{j=1}^{2^{N-1}} r_j b_j \, ,
$$
\begin{equation}
\sigc = \sum_{j=1}^{2^{N-1}} {1 \over r_j} c_j \, , \qquad
\sigd = \sum_{j=1}^{2^{N-1}} {1 \over r_j} d_j \, .
\end{equation}
Here, the variables $r_j$ are products of ratios of the form
\begin{equation}
r_j = {x_{\mu_1} x_{\mu_2} \dots x_{\mu_n} \over
x_{\nu_1} x_{\nu_2} \dots x_{\nu_n} } \, .
\end{equation}
The indices are confined to the range $1 \le \mu_i, \nu_i \le N$.
The additional factors $a_j$, $b_j$, $c_j$, $d_j$ are products of
the variables $\phi_j$, and can be written in the form
\begin{equation}
a_j = \prod_{k=1}^N \phi_k^{p_k} \qquad {\rm where} \qquad
p_k = 0 \, \, \, {\rm or} \, \, \, 1 \, \, .
\end{equation}
\medskip
\noindent
{\bf Result 1:} For the case where $|h_k| > 1$ for all cycles,
and in the limit of large $N$, the eigenvalue of the product matrix
is given by the formula
\begin{equation}
\lambda = \siga + \sigd \, + {\cal O} \left( h^{-2N} \right) \, ,
\label{result1}
\end{equation}
where each of these quantities should be labeled at the
$Nth$ iteration.
\noindent
{\it Proof:} The characteristic equation of the product matrix
of equation (\ref{product}) takes the form
\begin{equation}
\lambda^2 - \lambda (\siga + \sigd) + \siga \sigd - \sigb \sigc = 0 \, .
\label{charact}
\end{equation}
The final term is the determinant of the product matrix, and this
determinant is given by the product of the individual matrices, so
that
\begin{equation}
\siga \sigd - \sigb \sigc = \prod_{k=1}^N (1 - \phi_k) =
\prod_{k=1}^N {1 \over h_k^2} \, .
\end{equation}
Given that $|h_k| > 1$ $\forall k$, this term vanishes in the limit
$N \to \infty$. As a result, the growing eigenvalue of the
characteristic equation (\ref{charact}) simplifies to the form
$\lambda = \siga + \sigd$. $\Box$
\medskip
\noindent
{\bf Result 2:} The four sums that specify the matrix elements of
the product matrix are not independent. In particular, for the case
where $|h_k| > 1$ and in the limit $N \to \infty$, the ratios of the
matrix elements approach the form
\begin{equation}
{\sigb \over \siga} = {\sigd \over \sigc} = {\rm constant} \equiv f \, .
\label{result2}
\end{equation}
\noindent
{\it Proof:} As shown above, the determinant of the product matrix vanishes
in the limit $N \to \infty$, so that in the limit
\begin{equation}
\siga \sigd = \sigb \sigc \, .
\end{equation}
The result implied by the first equality of equation (\ref{result2})
follows immediately.
Further, one can show by direct construction that if the relation of
equation (\ref{result2}) holds, then the relation is preserved under
matrix multiplication. Let the product matrix after $N$ cycles have
the form
\begin{equation}
{\mfont B}^{(N)} = \left[ \matrix{ { \Sigma_T } & f x_1 { \Sigma_T } \cr
(1/x_1) { \Sigma_B } & f { \Sigma_B }} \right] \, ,
\label{ncycle}
\end{equation}
where $f$ is the constant in equation (\ref{result2}). Then the
matrix takes the following from after the next cycle:
\begin{equation}
{\mfont B}^{(N+1)} =
\left[ \matrix{ { \Sigma_T } + (x/x_1) \phi { \Sigma_B } & x_1
f ( { \Sigma_T } + (x/x_1) \phi { \Sigma_B } ) \cr
(1/x_1) ({ \Sigma_B } + (x_1/x) { \Sigma_T } ) &
f ({ \Sigma_B } + (x_1/x) { \Sigma_T }) } \right] \, ,
\label{andone}
\end{equation}
so that the left-right symmetry relation is conserved. $\Box$
In the above proof we have adopted notation that is used throughout
this paper: The subscript `1' denotes the values of the parameters
(e.g., $x_1$) for the first cycle in the series. Since the results of
this problem can be written in terms of this starting value, these
initial values play a recurring role. The subscript `$N$' denotes the
values of the parameters (e.g, $x_N$) appropriate for the $N$th cycle
of the series. In iteration formulae, however, we use unsubscripted
variables (e.g., $x$) for the next ($N+1$)st cycle.
\medskip
\noindent
{\bf Result 3:} In the highly unstable regime, the ratio of
${ \Sigma_T }$ to ${ \Sigma_B }$ has the form:
\begin{equation}
{{ \Sigma_T } \over { \Sigma_B }} = {x \over x_1} \, .
\label{result3}
\end{equation}
\noindent
{\it Proof:} From our previous results (see equation [19] of [AB]),
the product matrix after $N$ cycles has the form given by equation
(\ref{ncycle}) with $f = 1$ (in the highly unstable regime). After one
additional multiplication, we obtain the form given by equation
(\ref{andone}) with $f$ = 1. We thus find
\begin{equation}
{{ \Sigma_T }^{(N+1)} \over { \Sigma_B }^{(N+1)} } =
{{ \Sigma_T }^{(N)} + (x/x_1) { \Sigma_B }^{(N)} \over { \Sigma_B }^{(N)} +
(x_1/x) { \Sigma_T }^{(N)} } = {x \over x_1} \, .
\end{equation}
For each cycle the ratio $x/x_1$ has a different value, so that no
limit is reached as $N \to \infty$. However, the ratio at any given
finite cycle obeys equation (\ref{result3}). $\Box$
To derive an expression for the growth rate for matrix
multiplication, we first define
\begin{equation}
S \equiv \siga + \sigd \, .
\end{equation}
As shown in the proof of Result 1, the eigenvalue of the
product matrix approaches $S$, as defined above, in the limit
$N \to \infty$. By construction, the iteration formula for $S$
takes the form
\begin{equation}
S^{(N+1)} = S^{(N)} \left[ 1 +
{ (x/x_1) \phi \sigc^{(N)} + (x_1/x) \sigb^{(N)} \over
\siga^{(N)} + \sigd^{(N)} } \right] \, .
\end{equation}
Using the definition of $f$, ${ \Sigma_T }$, and ${ \Sigma_B }$, this expression
can be simplified to the form
\begin{equation}
S^{(N+1)} = S^{(N)} \left[ 1 +
{ (x/x_1) \phi { \Sigma_B }^{(N)} + (x_1/x) f { \Sigma_T }^{(N)} \over
{ \Sigma_T }^{(N)} + f { \Sigma_B }^{(N)} } \right] \, .
\label{siterate}
\end{equation}
\medskip
\noindent
{\bf Result 4:} In the highly unstable regime the
iteration formula for the eigenvalue reduces to the form
\begin{equation}
S^{(N+1)} = S^{(N)} \left[ 1 + {x_N \over x} \right] \, .
\label{result4}
\end{equation}
This result agrees with that of Theorem 2 from [AB].
\noindent
{\it Proof:} In the highly unstable regime $\phi = 1$, $f = 1$, and
equation (\ref{result3}) holds for the ratio of ${ \Sigma_T }/{ \Sigma_B }$.
The iteration formula of equation (\ref{siterate}) thus reduces to
\begin{equation}
S^{(N+1)} = S^{(N)} \left[ 1 +
{ (x/x_1) + (x_N/x) \over 1 + {x_N/x_1} } \right]
= S^{(N)} \left[ 1 + {x_N \over x} \right]
\left[ { x_1 + x \over x_1 + x_N } \right] \, .
\label{iteratehu}
\end{equation}
Since the starting value $x_1$ is fixed, the second factor in
square brackets approaches unity in the limit $N \to \infty$, i.e.,
\begin{equation}
\lim_{N \to \infty} \prod_{k=1}^N
\left[ { x_1 + x_{k+1} \over x_1 + x_k } \right] = 1 \, .
\end{equation}
The expression of equation (\ref{iteratehu}) thus reduces to
that of equation (\ref{result4}). $\Box$
Motivated by the result of equation (\ref{result3}) for the highly
unstable regime, we write the ratio of matrix elements for the general
case in the form
\begin{equation}
{{ \Sigma_T }^{(N)} \over { \Sigma_B }^{(N)}} = {x_N \over x_1} \alpha_N \, ,
\label{alphadef}
\end{equation}
so that
\begin{equation}
S^{(N+1)} = S^{(N)} \left[ 1 +
{ (x/x_1) \phi + (x_N/x) f \alpha_N \over f + \alpha_N (x_N/x_1) }
\right] \equiv {\cal F}_N S^{(N)} \, ,
\label{sitfactor}
\end{equation}
where the second equality defines ${\cal F}_N$. The parameter
$\alpha_N$ incorporates the correction due to the matrices not being
in the highly unstable regime. Note that $f$ approaches a constant
value (from Result 2) and $x_1$ is a constant (by definition). The
iteration factor ${\cal F}_N$ can be rewritten in the form
\begin{equation}
{\cal F}_N = \left[ 1 +
{ x^2 \phi + b \alpha_N x_N \over x (b + \alpha_N x_N) } \right]
\qquad {\rm where} \qquad b \equiv f x_1 \, .
\label{itfactor}
\end{equation}
\noindent
{\bf Theorem 1:} The growth rate for matrix multiplication, with
products of the general form defined through equation (\ref{product}),
is given by
\begin{equation}
\gamma = \lim_{N \to \infty} {1 \over N} \sum_{k=1}^N \log \left[ 1 +
{ x_k^2 \phi_k + \alpha_{k-1} x_{k-1} \over x_k (1 + \alpha_{k-1} x_{k-1}) }
\right] \, ,
\label{gammafull}
\end{equation}
where the $\alpha_k$ are determined through the iteration formula
\begin{equation}
\alpha_{k} = {x_k \phi_k + x_{k-1} \alpha_{k-1} \over
x_k + x_{k-1} \alpha_{k-1}} \, .
\label{iteralpha}
\end{equation}
\noindent
{\it Proof:} Note that existence of the required limit holds by the
Theorem of FK. Equations (\ref{sitfactor} -- \ref{itfactor}) show
that the growth rate is given by
\begin{equation}
\gamma = \lim_{N \to \infty} {1 \over N} \sum_{k=1}^N \log {\cal F}_k =
\lim_{N \to \infty} {1 \over N} \sum_{k=1}^N \log \left[ 1 +
{ x_k^2 \phi_k + b \alpha_{k-1} x_{k-1} \over x_k (b + \alpha_{k-1} x_{k-1}) }
\right] \, ,
\end{equation}
where this form is exact, provided that the $\alpha_k$ are properly
specified. This issue is addressed below. To complete the proof, we
must also show that the growth rate is independent of the value of $b$,
so that we can set $b=1$ in the above formula. The derivative of the
growth rate with respect to the parameter $b$ takes the form
\begin{equation}
{d \gamma \over db} = \lim_{N \to \infty} {1 \over N}
\sum_{k=1}^N {1 \over {\cal F}_k} {d {\cal F}_k \over db} \, ,
\end{equation}
which can be evaluated to take the form
\begin{equation}
{d \gamma \over db} = \lim_{N \to \infty} {1 \over N} \sum_{k=1}^N
{ (\alpha_{k-1} x_{k-1})^2 - x_k^2 \phi_k \over (b + \alpha_{k-1} x_{k-1})
\left[ x_k (b + \alpha_{k-1} x_{k-1}) + x_k^2 \phi_k +
b \alpha_{k-1} x_{k-1} \right] } \, .
\end{equation}
This expression vanishes in the limit.
To show that the $\alpha_k$ are given by equation (\ref{iteralpha}),
we start with the result of matrix multiplication from equation
(\ref{andone}) and use the definition of $\alpha_k$ from equation
(\ref{alphadef}); these two results imply that
\begin{equation}
\alpha_{k+1} = {x_1 \over x_{k+1}} { { \Sigma_T }^{(k+1)} \over { \Sigma_B }^{(k+1)} }
= {x_1 \over x_{k+1}} { { \Sigma_T }^{(k)} + (x_{k+1}/x_1) \phi_{k+1} { \Sigma_B }^{(k)}
\over { \Sigma_B }^{(k)} + (x_1/x_{k+1}) { \Sigma_T }^{(k)} } \, .
\end{equation}
We can then eliminate the factors of ${ \Sigma_T }$ and ${ \Sigma_B }$ by again
using the definition of $\alpha_k$ from equation (\ref{alphadef}),
and thus obtain
\begin{equation}
\alpha_{k+1} = {x_1 \over x_{k+1}} { (x_k/x_1) \alpha_k + (x_{k+1}/x_1)
\phi_{k+1} \over 1 + (x_k/x_{k+1}) \alpha_k } \, = \,
{ x_k \alpha_k + x_{k+1} \phi_{k+1} \over x_{k+1} + x_k \alpha_k} \, .
\end{equation}
After re-labeling the indices, we obtain equation (\ref{iteralpha}).
$\Box$
\section{Approximations for the Classically Unstable Regime}
For classically unstable matrices with $|h_k| > 1$, Theorem 1 provides
an exact expression for the growth rate. Since the formulae are
complicated, this section presents simpler but approximate expressions
for the growth rates for the case where $\phi_k$ are small (Theorem 2)
and where the differences $1 - \phi_k$ are small (Theorem 3). We also
present two heuristic approximations for the growth rates for the
general problem.
\medskip
\noindent
{\bf Theorem 2:} In the regime where the variables $\phi_k$ are small,
$\phi_k x_k \ll 1$ $\forall k$, the growth rate for the matrix
${\mfont B}_k$ tends in the limit of large $N$ to the form:
\begin{equation}
\gamma = \log \left( 1 + \left[ \langle {1 / x_k} \rangle
\langle {x_k \phi_k} \rangle \right]^{1/2} \right) +
{\cal O} \left( \langle {x_k \phi_k} \rangle \right) \, .
\label{theorem2}
\end{equation}
\noindent
{\it Proof:} We first break up the matrix into two parts so that
${\mfont B}_k$ = ${\mfont I}$ + ${\mfont A}_k$,
where $\mfont{I}$ is the identity matrix and where
\begin{equation}
{\mfont A}_k = \left[ \matrix{0 & x_k \phi_k \cr 1/x_k & 0} \right] =
\left[ \matrix{0 & \eta_k \cr y_k & 0} \right] \, .
\label{defineak}
\end{equation}
Note that the second equality defines $\eta_k = x_k \phi_k$ and
$y_k = 1/x_k$. We first show (by induction) that repeated
multiplications of the matrices ${\mfont A}_k$ lead to products
with simple forms. The products of even numbers $N = 2 \ell$ of
matrices $\mfont{A}_k$ produce diagonal matrices of the form
\begin{equation}
{\mfont A}^{(N)} = {\mfont A}^{(2 \ell)} = \prod_{k=1}^N {\mfont A}_k =
\left[ \matrix{P_\ell^A & 0 \cr 0 & P_\ell^B} \right] \, ,
\label{aproducteven}
\end{equation}
where the products $P_\ell$ are defined by
\begin{equation}
P_\ell^A = \prod_{i=1}^\ell
\left( \eta_{2i} \right) \left( y_{2i-1} \right)
\qquad {\rm and} \qquad
P_\ell^B = \prod_{i=1}^\ell
\left( \eta_{2i-1} \right) \left( y_{2i} \right) \, .
\label{pells}
\end{equation}
Similarly, the product of odd numbers $N = 2 \ell + 1$ of matrices
$\mfont{A}_k$ produce off-diagonal matrices of the form
\begin{equation}
{\mfont A}^{(N)} = {\mfont A}^{(2 \ell+1)} = \prod_{k=1}^N {\mfont A}_k =
\left[ \matrix{0 & Q_\ell^B \eta_1 \cr Q_\ell^A y_1 & 0 } \right] \, ,
\label{aproductodd}
\end{equation}
where the products $Q_\ell$ are defined analogously to the $P_\ell$.
The product of $N$ matrices $\mfont{B}_k$ can then be written in the form
\begin{equation}
\mfont{B}^{(N)} = \prod_{k=1}^N \mfont{B}_k = \left[
\matrix{\siga & \sigb \eta_1 \cr \sigc y_1 & \sigd } \right] \, .
\end{equation}
Without loss of generality, let $N = 2 \ell$ be even.
Then the matrix elements are given by
$$
\siga = \sum_{\ell=0}^{N/2} \sum_{j=1}^{C^N_{2\ell}}
\left( P_\ell^A \right)_j \, , \qquad
\sigd = \sum_{\ell=0}^{N/2} \sum_{j=1}^{C^N_{2\ell}}
\left( P_\ell^B \right)_j \, , \qquad
$$
\begin{equation}
\sigb = \sum_{\ell=0}^{N/2-1} \sum_{j=1}^{C^N_{2\ell+1}}
\left( Q_\ell^B \right)_j \, , \qquad
\sigc = \sum_{\ell=0}^{N/2-1} \sum_{j=1}^{C^N_{2\ell+1}}
\left( Q_\ell^A \right)_j \, , \qquad
\end{equation}
where $C^N_\ell$ is the binomial coefficient and where the subscripts
on the $P_\ell$ and $Q_\ell$ denote different realizations of the
products.
The eigenvalue $\Lambda_N$ of the product matrix at the $Nth$ iteration
is given by its characteristic equation, which has the solution
\begin{equation}
\Lambda_N = {1 \over 2} \left\{ \siga + \sigd + \left[
(\siga - \sigd)^2 + 4 \sigb \sigc \eta_1 y_1 \right]^{1/2} \right\} \, .
\end{equation}
In the limit of large $N$, we can make the approximation that
$\siga \approx \sigd$ and $\sigb \approx \sigc$, so that the
expression for the eigenvalue takes the form
\begin{equation}
\Lambda_N = \siga + \sigb \left[\eta_1 y_1 \right]^{1/2}
= \sum_{\ell=0}^{N/2} \sum_{j=1}^{C^N_{2\ell}}
\left( P_\ell^A \right)_j +
\sum_{\ell=0}^{N/2-1} \sum_{j=1}^{C^N_{2\ell+1}}
\left( Q_\ell^B \right)_j \left[\eta_1 y_1 \right]^{1/2} .
\end{equation}
In the limit of large $N$, all the binomial coefficients are
large except for the first and last one. We can thus rewrite the
above equation in the form
\begin{equation}
\Lambda_N = \sum_{\ell=0}^{N/2} C^N_{2\ell}
\left( \left\langle P_\ell^A \right\rangle + \varepsilon_\ell \right)
+ \sum_{\ell=0}^{N/2-1} C^N_{2\ell+1}
\left( \left\langle Q_\ell^B \right\rangle + \varepsilon_\ell \right)
\left[\eta_1 y_1 \right]^{1/2} \, .
\end{equation}
If the realizations of the products $(P_\ell)_j$ were independent, the
error terms $\varepsilon_\ell$ would vanish in the limit. However, for a
given $N$, the sums contain $C_{2\ell}^N$ terms, and $C_{2\ell}^N > N$
in general, so all of the terms in the sum cannot be independent.
We then write the products $\left\langle P_\ell^A \right\rangle$ and
$\left\langle Q_\ell^B \right\rangle$ in the form
\begin{equation}
\left\langle P_\ell^A \right\rangle + \varepsilon_\ell =
\langle \eta_j \rangle^\ell \langle y_j \rangle^\ell
(1 + \epsilon_\ell)^\ell \, ,
\end{equation}
and similarly for $\left\langle Q_\ell^B \right\rangle$. This form is
exact if one uses the proper expressions for the $\epsilon_\ell$.
Using this result, the expression for the eigenvalue $\Lambda_N$ becomes
\begin{equation}
\Lambda_N = \sum_{\ell=0}^{N/2} C^N_{2\ell}
\langle \eta_j \rangle^\ell \langle y_j \rangle^\ell (1 + \epsilon_\ell)^\ell
+ \sum_{\ell=0}^{N/2-1} C^N_{2\ell+1}
\langle \eta_j \rangle^\ell \langle y_j \rangle^\ell (1 + \epsilon_\ell)^\ell
\left[\eta_1 y_1 \right]^{1/2} \, ,
\end{equation}
which takes the form
\begin{equation}
\Lambda_N = \sum_{k=0}^{N} C^N_k
\langle \eta_j \rangle^{k/2} \langle y_j \rangle^{k/2}
(1 + \epsilon_{k} )^{k/2} \, .
\end{equation}
If we expand this result, we find that
\begin{equation}
\Lambda_N = 1 + N \langle \eta_j \rangle^{1/2} \langle y_j \rangle^{1/2}
(1 + \epsilon_{1} )^{1/2} + C_2^N \langle \eta_j \rangle \langle y_j \rangle
(1 + \epsilon_{2} ) + \dots
\end{equation}
Further, by performing an exact treatment of the first order expansion
[AB2] we find that $\epsilon_1$ = 0. This finding allows us to write the
product in the form
\begin{equation}
\Lambda_N = \left[ 1 + \langle \eta_j \rangle^{1/2} \langle y_j \rangle^{1/2}
+ {\cal O} (\eta_j) \right]^N \, .
\end{equation}
The growth rate thus becomes
\begin{equation}
\gamma = \log \left[ 1 + \langle \eta_j \rangle^{1/2}
\langle y_j \rangle^{1/2} \right] + {\cal O} (\eta_j) \, .
\end{equation}
This last expression is valid provided that $\eta_j \ll 1$ $\forall j$.
$\Box$
Note that to consistent order, we can replace the limiting form of
equation (\ref{theorem2}) with the equivalent, simpler function
\begin{equation}
\gamma \to
\left[ \langle 1/x_k \rangle \langle \eta_k \rangle \right]^{1/2} \, .
\end{equation}
Figure \ref{fig:smallamp} illustrates how well the approximation of
Theorem 2 works. For the sake of definiteness, the variables $x_k$
are log-uniformly distributed with $\log_{10} x_k \in [-2,2]$. The
$\phi_k$ obey the relation $\phi_k = a_\phi \xi_k$, where $\xi_k$ is a
uniformly distributed random variable over the interval $[0,1]$. As
shown by the figure, the limiting form of equation (\ref{theorem2})
provides an excellent description of the calculated growth rate for
sufficiently small $\phi_k$.
\begin{figure}
\centering
\includegraphics[width=120mm]{smallamp.eps}
\caption{Growth rates for small $\phi_k$. The variables $\phi_k$ are
determined through the relation $\phi_k = \ampz \, \xi_k$, where
$\xi_k$ is uniformly distributed on [0,1]. The solid curve shows the
growth rate $\gamma$ calculated directly from matrix multiplication as
a function of the amplitude $\ampz$. The dashed curve shows the
estimate $\gamma_2$ for the growth rate from Theorem 2. The dotted
curve shows the difference $\Delta \gamma$ = $\gamma_2 - \gamma$.
Note that $\gamma \propto \sqrt{\ampz}$ whereas
$\Delta \gamma \propto \ampz$. }
\label{fig:smallamp}
\end{figure}
Next we consider the case where the correction factors $\phi_k$ are
close to unity. In this case the variables $(1-\phi_k) \ll 1$, and we
can expand to leading order in $(1-\phi_k)$. This procedure leads to
the following result:
\medskip
\noindent
{\bf Theorem 3:} Let $\gamma_0$ be the growth rate for the highly
unstable regime where $\phi_k = 1$. For small perturbations about
this limiting case, the growth rate takes the form $\gamma$ =
$\gamma_0 - \delta \gamma$, where
\begin{equation}
\delta \gamma = \lim_{N \to \infty} {1 \over N} \sum_{k=1}^N
{(1 - \phi_k) x_k^2 \over (x_{k+1} + x_k) (x_k + x_{k-1}) } \, +
{\cal O} \left( \langle x_k^2 (1 - \phi_k)^2 \rangle \right) \, .
\label{theorem3}
\end{equation}
\noindent
{\it Proof:} We again break up the matrix into two parts,
\begin{equation}
{\mfont B}_k = {\mfont C}_k - \epsilon_k {\mfont Z}
\qquad {\rm with} \qquad {\mfont Z} \equiv
\left[ \matrix{0 & 1 \cr 0 & 0} \right] \, ,
\label{decomp}
\end{equation}
where here $\epsilon_k \equiv x_k (1 - \phi_k)$ and ${\mfont C}_k$ is
the matrix appropriate for the highly unstable regime. Note that
${\mfont Z}$ does not depend on the index $k$. Here we work to
first order in the small parameter $\epsilon_k$. After $N$ cycles,
the product matrix takes the form
\begin{equation}
{\mfont B}_k^{(N)} = \prod_{k=1}^N {\mfont B}_k = {\mfont C}_k^{(N)} -
\sum_{k=1}^N \epsilon_k {\mfont P}_k^N \, + {\cal O}(\epsilon_k^2) \, ,
\label{productb}
\end{equation}
where the partial product matrices ${\mfont P}_k^N$ are given by
\begin{equation}
{\mfont P}_k^N = \left\{ \prod_{j=k+1}^N {\mfont C}_j \right\} \, {\mfont Z}
\left\{ \prod_{j=1}^{k-1} {\mfont C}_j \right\} \, .
\end{equation}
We ignore the case where the ${\mfont Z}$ factors appear on the
ends -- this effect is ${\cal O} (1/N)$ and vanishes in the limit.
The products of the ${\mfont C}_k$ matrices can be written in the form
\begin{equation}
{\mfont C}_k^{(N)} = \Sigma_T^N \left[ \matrix{1 & x_1 \cr 1/x_N & x_1/x_N}
\right] \qquad {\rm where} \qquad \Sigma_T^N = \prod_{j=2}^N
\left(1 + {x_j \over x_{j-1} } \right) \, ,
\label{productc}
\end{equation}
where these results follow from previous work [AB]. As a result, the
matrices ${\mfont P}_k^N$ can be evaluated:
\begin{equation}
{\mfont P}_k^N = {x_k \Sigma_T^N \over (x_k + x_{k+1}) (x_{k-1} + x_{k}) }
\left[ \matrix{1 & x_{1} \cr 1/x_N & x_{1}/x_N} \right] =
{x_k \over (x_k + x_{k+1}) (x_{k-1} + x_{k}) } {\mfont C}_k^{(N)} \, .
\end{equation}
The product matrix ${\mfont B}_k^{(N)}$, given by equation (\ref{productb})
to leading order, can now be written in the form
\begin{equation}
{\mfont B}_k^N = {\mfont C}_k^N \left[ 1 - \sum_{k=1}^N
{(1 - \phi_k) x_k^2 \over (x_k + x_{k+1}) (x_{k-1} + x_{k}) } \right] \, .
\end{equation}
The first factor is the product of the matrices for the highly
unstable regime. Since the second factor is a function (not a matrix)
its contribution to the growth rate is independent of the first factor
and represents a correction to the growth rate of the form
\begin{equation}
\delta \gamma = \lim_{N \to \infty} {1 \over N} \sum_{k=1}^N
{(1 - \phi_k) x_k^2 \over (x_k + x_{k+1}) (x_{k-1} + x_{k}) }
+ {\cal O} (\epsilon_k^2) \, ,
\end{equation}
where the equalities hold to leading order. This correction to the
growth rate has the form given by equation (\ref{theorem3}). $\Box$
Figure \ref{fig:largegam} shows the growth rate for small departures
from the highly unstable regime. The correction factors are taken to
have the form $\phi_k = 1 - \amp \xi_k$, where $\xi_k$ is a uniformly
distributed random variable over the interval $[0,1]$. The highly
unstable regime corresponds to $\amp \to 0$. The figure shows the
growth rate calculated from direct matrix multiplication (solid curve)
and the approximation from Theorem 3 (dashed curve) plotted as a
function of the amplitude $\amp$. Both curves plot the difference
$\gamma_0 - \gamma$, where $\gamma_0$ is the growth rate for the
highly unstable regime (where the $\phi_k$ = 1).
\begin{figure}
\centering
\includegraphics[width=120mm]{large.eps}
\caption{Growth rates for $\phi_k$ near unity. The variables $\phi_k$
are determined through the relation $\phi_k = 1 - \amp \, \xi_k$,
where $\xi_k$ is uniformly distributed on [0,1]. The solid curve
shows the quantity $\delta \gamma = \gamma_0 - \gamma$, where
$\gamma$ is the growth rate calculated from matrix multiplication and
$\gamma_0$ is the growth rate for the highly unstable regime
($\phi_k$ = 1 $\forall k$). The dashed curve shows the estimate
$(\delta \gamma)_3 = (\gamma_0 - \gamma)_3$ for the difference in
growth rate calculated from Theorem 3. The dotted curve shows the
error $\Delta$ = $(\delta \gamma)_3 - \delta \gamma$. Note that
$\delta \gamma \propto \amp$ whereas the error term
$\Delta \propto (\amp)^2$. }
\label{fig:largegam}
\end{figure}
Since the general case is quite complicated it is useful to have
a good working approximation for the
case where one is not in one of the two regimes
$\phi_k$ small or near unity.
Toward this end, we first show that the values of $\alpha_k$
have a limited range:
\medskip
\noindent
{\bf Result 5:} The variables $\alpha_k$ are confined to the range
$\phi_{\rm min} \le \alpha_k \le 1$, where $\phi_{\rm min}$ is the
minimum value of $\phi_k$.
\noindent
{\it Proof:} We can rewrite the iteration formula (\ref{iteralpha})
for $\alpha_k$ in the alternate form
\begin{equation}
\alpha_{k} = {\phi_k + \beta_k \over 1 + \beta_k } \, ,
\end{equation}
where we have defined the composite random variable
$\beta_{k} \equiv \alpha_{k-1} x_{k-1} / x_k$.
In the present context, $0 \le \beta_k < \infty$, and we can show that
\begin{equation}
{d \alpha_k \over d \beta_k} > 0
\end{equation}
for all values of $\beta_k$. In the limit $\beta_k \to \infty$,
$\alpha_{k} \to 1$, whereas in the limit $\beta_k \to 0$,
$\alpha_{k} \to \phi$. Hence $\phi \le \alpha_k \le 1$ for all cycles.
But $\phi \ge \phi_{\rm min}$, by definition, so that $\phi_{\rm min}
\le \alpha_k \le 1$. $\Box$
\medskip
\noindent
{\bf Approximation 1:} As a first heuristic approximation, we replace
the full iteration expression of equation (\ref{iteralpha}) for
$\alpha_k$ with the following simplified form
\begin{equation}
\alpha_{k+1} = {x \phi + x_k \over x + x_k} \, ,
\label{firstalf}
\end{equation}
i.e., we use $\alpha_k$ = 1 as an approximation for the previous value
[keep in mind that $x$ is the value at the ($k+1$)th cycle]. Using
equation (\ref{firstalf}) to evaluate $\alpha_k$ in the iteration
formula for ${\cal F}_k$, we obtain a working approximation for the
growth rate. Notice that $\alpha_k$ appears in the iteration formula
for ${\cal F}_k$, so that we must use equation (\ref{firstalf})
evaluated at $k$ rather than $k+1$. As a result, the iteration factor
${\cal F}_k$ involves the random variables $x_k$ from three cycles,
or, equivalently (since the $x_k$ are i.i.d.) three separate samplings
of the variables. We change notation so that $x_{j1}, x_{j2}, x_{j3}$
denote the three independent samplings of the random variables $x_k$.
Similarly, let $\phi_{j1}, \phi_{j2}$ denote two independent samplings
of the $\phi_k$. The iteration formula for this approximation can then
be written in the form
\begin{equation}
{\cal F}_j = 1 + { x_{j1}^2 \phi_{j1} (x_{j2} + x_{j3}) + x_{j2}
(x_{j2} \phi_{j2} + x_{j3} ) \over x_{j1} \left[ (x_{j2} + x_{j3}) +
x_{j2} (x_{j2} \phi_{j2} + x_{j3} ) \right] } \, .
\label{iterapproxone}
\end{equation}
The growth rate for matrix multiplication can then be approximated by
\begin{equation}
\gamma = \lim_{N \to \infty} {1 \over N} \sum_{j=1}^N \log {\cal F}_j \, ,
\label{approxgamma}
\end{equation}
where ${\cal F}_j$ is given by equation (\ref{iterapproxone}). As a
consistency check, for the restricted problem where the $\phi_{jn} = 1$,
the iteration factor ${\cal F}_j$ reduces to that appropriate for the
highly unstable regime (see equation [\ref{result4}]).
\medskip
\noindent
{\bf Approximation 2:} To derive a second approximation for the growth
rate, we need a better approximation for the $\alpha_k$. If the values
of $x_k$ and $\phi_k$ were constant, then the $\alpha_k$ would
approach a constant value given by
\begin{equation}
\alpha_k = {1 \over 2} \left\{ (1 - x_k/ x_{k-1}) + \left[
(1 - x_k/ x_{k-1})^2 + 4 (x_k/ x_{k-1}) \phi_k \right]^{1/2} \right\} \, .
\label{zeropoint}
\end{equation}
Even though the $x_k$ and $\phi_k$ are not constant, and the
$\alpha_k$ vary, we can use equation (\ref{zeropoint}) as an
approximation to specify the values of $\alpha_k$ appearing in the
exact formula of equation (\ref{gammafull}) for the growth rate.
After using this form to specify the $\alpha_k$, and relabeling the
indices, the iteration factor takes the form
\begin{equation}
{\cal F}_k = 1 +
{ x_{k1}^2 \phi_{k1} 2 x_{k3} + x_{k2} \left\{ (x_{k3} - x_{k2}) + \left[
(x_{k3} - x_{k2})^2 + 4 x_{k2} x_{k3} \phi_{k2} \right]^{1/2} \right\}
\over
x_{k1} \left( 2 x_{k3} + x_{k2} \left\{ (x_{k3} - x_{k2}) +
\left[ (x_{k3} - x_{k2})^2 + 4 x_{k2} x_{k3} \phi_{k2} \right]^{1/2}
\right\} \right) } \, .
\label{approxtwo}
\end{equation}
In the case $\phi_{jn} = 1$, the iteration factor of equation
(\ref{approxtwo}) reduces to the expression for the highly unstable
regime (Result 4).
Figure \ref{fig:gamamp} shows how well these two approximation schemes
work. The $\phi_k$ variables are chosen from the expression $\phi_k$ =
$1 - \amp \xi_k$, where $\xi_k$ is a random variable uniformly sampled
from the interval $0 \le \xi_k \le 1$ and where $\amp$ sets the
amplitude of the departures of the $\phi_k$ from unity. The growth
rate is shown as a function of the amplitude.
\begin{figure}
\centering
\includegraphics[width=120mm]{gamamp.eps}
\caption{Validity of approximations of equation (\ref{approxgamma})
and equation (\ref{approxtwo}) as a function of the deviation of
$\phi_k$ from unity. The upper solid line shows the growth rate for
matrix multiplication in the highly unstable regime where $\phi_k = 1$.
The lower solid curve shows the growth rate for the case where
$\phi_k = 1 - \amp \xi_k$, where $\xi_k$ is a uniformly distributed random
variable $0 \le \xi_k \le 1$. The dotted curve shows the estimate for
growth rate calculated from equation (\ref{approxgamma}) using the
same sampling of the $\phi_k$ variables; similarly, the dot-dashed curve
shows the approximation of equation (\ref{approxtwo}). Notice that both
of these approximations are almost identical to the actual result. The
dashed curve shows the lower limit to the growth rate derived in [AB]. }
\label{fig:gamamp}
\end{figure}
In [AB], we derived a bound on the difference between the growth rate
for the general case $\gamma$ (considered here) and the growth rate in the
highly unstable regime $\gamma_0$, i.e.,
\begin{equation}
\gamma_0 - \gamma \le {1 \over 2} \langle \log \phi_k \rangle \, .
\end{equation}
This bound is shown as the dashed curve in Figure \ref{fig:gamamp}.
The true growth rates fall comfortably between this lower bound and
the growth rate for the highly unstable regime (where the latter
provides an upper bound).
Thus far, this paper has focused on the regime where the transformation
matrices are classically unstable. Before considering classically
stable matrix multiplication in the next section, we note the
following result that applies at the transition between the two
regimes:
\medskip
\noindent
{\bf Result 6:} Consider the matrix transformation that maps the
principal solutions from one cycle to the next. When the matrix
elements $g_k = {\dot y}_1 (\pi)$ vanish, then the remaining matrix
elements are $h_k = y_1 (\pi) = \pm 1$. The transformation matrix
${\mfont M}_{g0}$ for this case is stable under multiplication.
(The proof is a simple explicit computation.)
\section{Elliptical Rotations and the Classically Stable Regime}
When the principal solutions $h_k$ appearing in the discrete map of
equation (\ref{mapzero}) are less than unity, matrix multiplication is
stable for the case of constant parameters. In the case of interest,
however, the parameters in Hill's equation (\ref{basic}) and the
matrices (\ref{mapzero}) vary from cycle to cycle. This section
considers the case where the $|h_k| \le 1$, but vary from cycle to
cycle, and show that instability results. In this regime, the
discrete map takes the form of an elliptical rotation matrix [LR] as
described below. We thus find the growth rates for elliptical
rotation matrices for the case where the matrix elements vary from
cycle to cycle.
\medskip
\noindent
{\bf Definition:} An {\it elliptical rotation matrix} is defined to be
\begin{equation}
{\mfont E} (\theta; L) \equiv \left[
\matrix{ \cos\theta & - L \sin\theta \cr
(1/L) \sin\theta & \cos \theta }
\right] \, .
\label{ellipdef}
\end{equation}
\noindent
These matrices have the following properties:
\noindent
The product of elliptical rotation matrices with the same value of $L$
produces another elliptical rotation matrix, also with the same $L$,
\begin{equation}
{\mfont E} (\theta_1; L) {\mfont E} (\theta_2; L) =
{\mfont E} \left( [\theta_1 + \theta_2]; L \right) \, .
\end{equation}
As a result, the elliptical rotation matrices form a group.
\noindent
For fixed $L$, matrix multiplication is stable. Specifically, the
eigenvalues of the product of $N$ matrices (with fixed $L$) have the
form
\begin{equation}
\lambda = \exp \left[ \pm i \sum_{j=1}^N \theta_j \right] \, =
\exp \left[ \pm i \theta_N \right] \, ,
\end{equation}
where $\theta_N$ is the angle corresponding to the group element
produced after $N$ matrix multiplications.
\medskip
\noindent
{\bf Result 7:} When an individual cycle of Hill's equation is
stable, specifically when $|h_k| \le 1$, the full transformation
matrix ${\mfont M}_k$ takes the form of an elliptical rotation.
\noindent
{\it Proof:} Since $|h_k| \le 1$, we can define an angle $\theta_k$ such
that $h_k = \cos \theta_k$. The full matrix ${\mfont M}_k$ given by
equation (\ref{mbdefine}) then takes the form
\begin{equation}
{\mfont M}_k = \left[ \matrix{ \cos \theta_k & - (\sin^2 \theta_k)/g_k \cr
g_k & \cos \theta_k } \right] \, =
\left[ \matrix{ \cos \theta_k & - L_k \sin \theta_k \cr
(1/L_k) \sin \theta_k & \cos \theta_k } \right] \, = {\mfont E}_k (\theta_k; L_k) \, ,
\end{equation}
where we have defined $L_k = (\sin\theta_k)/g_k$. As before, we can
factor out the $\cos\theta_k = h_k$ and write the matrix in the form
\begin{equation}
{\mfont M}_k =
\cos\theta_k \left[ \matrix{1 & x_k \phi_k \cr 1/x_k & 1} \right] \, = \,
\cos\theta_k {\mfont B}_k \, ,
\label{anotherbdef}
\end{equation}
where
\begin{equation}
x_k = L_k / \tan\theta_k \qquad {\rm and} \qquad \phi_k = - \tan^2\theta_k \, .
\label{transellipse}
\end{equation}
Equation (\ref{transellipse}) thus specifies the transformation
between the random variables $(x_k, \phi_k)$ appearing in the original
transformation matrix and the random variables $(\theta_k, L_k)$ in
the corresponding elliptical rotation matrix. Note that the values of
$\phi_k$ are strictly negative in this formulation. Otherwise, the
matrix ${\mfont B}_k$ has the same form as in equation
(\ref{mbdefine}). $\Box$
If we let $\gamma_B$ be the growth rate for matrix ${\mfont B}_k$,
then the growth rate $\gamma_M$ for the full matrix ${\mfont M}_k$
takes the form
\begin{equation}
\gamma_M = \gamma_B + \lim_{N \to \infty} {1 \over N} \sum_{k=1}^N
\log [ \cos \theta_k ] \, .
\label{gammasum}
\end{equation}
The exact growth rate for the matrix ${\mfont B}_k$ (see equation
[\ref{anotherbdef}]) is given by Theorem 1. In particular, equations
(\ref{gammafull}) and (\ref{iteralpha}) remain valid for negative
values of the $\phi_k$ and can be used to calculate the growth rate.
\medskip
\noindent
{\bf Result 8:} For an elliptical rotation matrix with constant
angle $\theta$ and random $L_k$, the growth rate for matrix
multiplication vanishes in the two limits $h = \cos\theta \to 0$ and
$h = \cos\theta \to 1$.
\noindent
{\it Proof:} In the limit $h \to 1$ we have $\sin\theta$ = 0, and the
elliptical rotation matrix becomes the identity matrix. As a result,
the growth rate vanishes.
\noindent
In the other case where $h \to 0$, $\sin\theta$ = 1, and the matrix
takes the form
\begin{equation}
{\mfont E}_k \to {\mfont E}_{0k} =
\left[ \matrix{0 & -L_k \cr 1/L_k & 0 } \right] \, .
\end{equation}
In this case, for even numbers of matrix multiplications,
say $N$ = $2n$, the product matrix takes the form
\begin{equation}
{\mfont E}_{0k}^{(N)} = \prod_{k=1}^N {\mfont E}_{0k} =
(-1)^n \left[ \matrix{P^A_n & 0 \cr 0 & P^B_n } \right] \, ,
\end{equation}
where the matrix elements are given by the products
\begin{equation}
P^A_n = \prod_{k=1}^{n} {L_{2k} \over L_{2k-1}}
\qquad {\rm and} \qquad
P^B_n = \prod_{k=1}^{n} {L_{2k-1} \over L_{2k}} \, .
\label{pndef}
\end{equation}
The eigenvalues of the product matrix are given by $\lambda = P^A_n$
and $\lambda = P^B_n$. For odd $N = 2n+1$, the eigenvalue $|\lambda|$
= $(P^A_n P^B_n)^{1/2}$. In either case, in the limit of large $N$,
the growth rate for matrix multiplication takes the form
\begin{equation}
\gamma = \lim_{N \to \infty} {1 \over N} \sum_{k=1}^N \log
\left[ {L_{2k} \over L_{2k-1} } \right] =
\left\langle \log L_{2k} \right\rangle -
\left\langle \log L_{2k-1} \right\rangle \, = 0 \, .
\end{equation}
The final equality holds because the $L_k$ are independent. $\Box$
Elliptical rotation matrices are unstable under multiplication when
their parameters vary from cycle to cycle:
\medskip
\noindent
{\bf Theorem 4:} Consider an elliptical rotation matrix with variable
angle $\theta_k$ and symmetric fluctuations of the $L_k$ parameter
about its mean value $L_0$. The variations are thus written in the
form $L_k$ = $L_0 (1 + \eta_k)$, where the odd moments
$\langle \eta_k^{2n+1} \rangle$ = 0 for all integers $n$. For small
fluctuations $|\eta_k| < 1$, the growth $\gamma$ rate for matrix
multiplication takes the form
\begin{equation}
\gamma = {1 \over 2} \lim_{N\to\infty} {1 \over N}
\sum_{k=1}^N \log \left[ \cos^2 \theta_k + \sin^2 \theta_k
\left\langle {1 \over 1 + \eta_{j}} \right\rangle \right] +
{\cal O} \left( \eta_k^4 \right) \, .
\label{quadellipse}
\end{equation}
\noindent
{\it Proof:} We first break up the matrix into two parts so that
\begin{equation}
{\mfont E}_k = {\mfont I} \cos \theta_k + \sin \theta_k {\mfont Z}_k \, ,
\end{equation}
where $\mfont{I}$ is the identity matrix and where
\begin{equation}
{\mfont Z}_k = \left[ \matrix{0 & -L_k \cr 1/L_k & 0} \right] \, .
\end{equation}
The product of $N$ matrices $\mfont{E}_k$ becomes
\begin{equation}
\mfont{E}^{(N)} = \prod_{k=1}^N \mfont{E}_k =
\sum_{\ell=0}^N \sum_{k=1}^{C_\ell^N}
\left( \prod_{i=1}^{N-\ell} \cos \theta_i \right)_k
\left( \prod_{j=1}^{\ell} {\mfont Z}_j \sin\theta_j \right)_k \, ,
\end{equation}
where the subscripts on the products denote different realizations.
The products of even numbers $\ell = 2 n$ of matrices $\mfont{Z}_k$
produce diagonal matrices of the form
\begin{equation}
{\mfont Z}^{(\ell)} = {\mfont Z}^{(2 n)} =
\prod_{k=1}^n {\mfont Z}_{2k} {\mfont Z}_{2k-1}
= (-1)^n \left[ \matrix{P_n^A & 0 \cr 0 & P_n^B} \right] \, ,
\end{equation}
where the matrix elements $P^A_n$ and $P^B_n$ are given by equation
(\ref{pndef}). Similarly, the product of odd numbers $\ell = 2n + 1$
of matrices $\mfont{Z}_k$ produce off-diagonal matrices of the form
\begin{equation}
{\mfont Z}^{(\ell)} = {\mfont Z}^{(2n+1)} =
\left\{ \prod_{k=1}^n {\mfont Z}_{2k+1} {\mfont Z}_{2k} \right\}
{\mfont Z}_1 = (-1)^n
\left[ \matrix{0 & - P_n^A L_1 \cr P_n^B / L_1 & 0 } \right] \, ,
\end{equation}
where the $P_n$ are defined previously. Next we write the expectation
values of these products in the form
\begin{equation}
\left\langle P_n \right\rangle = \left\langle
\prod_{j=1}^n {L_{2j} \over L_{2j-1}} \right\rangle = \left\langle
\prod_{j=1}^n {1 + \eta_{2j} \over 1 + \eta_{2j-1}} \right\rangle =
\left\langle {1 \over 1 + \eta_{j}} \right\rangle^n
\equiv { {\cal R} }^n \, .
\label{pbar}
\end{equation}
This expression holds because the odd powers of the $\eta_j$ vanish in
the mean, and the samples of the different $\eta$'s are independent.
The eigenvalue $\Lambda_N$ of the product matrix at the $Nth$
iteration can be written in terms of its matrix elements, i.e.,
\begin{equation}
\Lambda_N = \sigma_{11} + \sigma_{22} \, .
\end{equation}
Without loss of generality, let $N = 2 K$ be even. The matrix
elements $\sigma_{11} = \sigma_{22} = \sigma$ are given by
\begin{equation}
\sigma = \sum_{m=0}^{K} \, \sum_{k=1}^{C_{2m}^{2K}} \,
\left( \prod_{i=1}^{2K-2m} \cos \theta_i \right)_k
\left( \prod_{i=1}^{2m} \sin \theta_i \right)_k (-1)^m \,
{ {\cal R} }^m \, ,
\end{equation}
where $C^{2K}_{2m}$ is the binomial coefficient and where we have
used equation (\ref{pbar}). This expression for $\sigma$ contains
the even terms of a binomial expansion. We can thus write the
eigenvalue in the form
\begin{equation}
\Lambda_N = \prod_{k=1}^N \left[ \cos \theta_k + i \sin \theta_k
{ {\cal R} }^{1/2} \right]_k
+ \prod_{k=1}^N \left[ \cos \theta_k - i \sin \theta_k
{ {\cal R} }^{1/2} \right]_k \, .
\end{equation}
Next we define
\begin{equation}
A_k \equiv \left[ \cos^2 \theta_k + \sin^2 \theta_k { {\cal R} } \right]^{1/2}
\qquad {\rm and} \qquad \tan \alpha_k \equiv { {\cal R} }^{-1/2} \tan \theta_k \, .
\end{equation}
The eigenvalue takes the form
\begin{equation}
\Lambda_N = 2 \, \left( \prod_{k=1}^N A_k \right) \,
\cos \left( \sum_{k=1}^N \alpha_k \right) \, ,
\end{equation}
and the corresponding growth rate becomes
\begin{equation}
\gamma = {1 \over 2} \lim_{N\to\infty} {1 \over N}
\sum_{k=1}^N \log \left[ \cos^2 \theta_k +
\sin^2 \theta_k { {\cal R} } \right] \, .
\label{egrow}
\end{equation}
Using the definition of ${ {\cal R} }$, we obtain the result of Theorem 4.
The order of the error term follows by comparing equation
(\ref{egrow}) with the leading order expansion [AB2].
$\Box$
In the regime of small $\eta_k \ll 1$, the expression for the growth
rate reduces to the form
\begin{equation}
\gamma = {1 \over 2} \left\langle \sin^2 \theta_k \right\rangle
\left\langle \eta_k^2 \right\rangle \, .
\end{equation}
This section shows that instability does not require a finite
threshold for the amplitude of the fluctuations in $L_k$. Nonzero
amplitude leads to instability with growth rate $\gamma \propto
\langle \eta_k^2 \rangle$. Variations in the original parameters
$(\lambda_k, q_k)$ of Hill's equation lead to fluctuations in the
principal solutions $(h_k, g_k)$; fluctuations in the $(h_k, g_k)$
lead to variations in the $L_k$ and hence growth. As a result, Hill's
equation with random forcing terms is generically unstable. One
notable exception occurs when the $h_k$ = 0 or $h_k$ = 1 (Result 8).
\section{Conclusion}
This paper provides expressions for the growth rates for the random
$2 \times 2$ matrices that result from solutions to the random Hill's
equation (\ref{basic}). Theorem 1 gives an exact expression for the
growth rate. Theorems 2 and 3 provide approximate growth rates for the
regimes where the variables $\phi_k$ are small, and close to unity,
respectively. Additional approximations for are given in Section 4.
When Hill's equation is classically stable, the discrete map that
governs the solutions has the form of an elliptical rotation matrix
(equ. [\ref{ellipdef}]). With fixed elements, such matrices are
stable under multiplication; variations in the $L_k$ parameter lead to
instability. For small symmetric fluctuations of the length parameter
$L_k$, the growth rate is given by Theorem 4.
\medskip
\begin{acknowledgements}
We would like to thank Scott Watson and Michael Weinstein for useful
conversations and suggestions. The work of FCA and AMB is jointly
supported by NSF Grant DMS-0806756 from the Division of Applied
Mathematics, and by the University of Michigan through the Michigan
Center for Theoretical Physics. AMB is also supported by the NSF
through grants DMS-0604307 and DMS-0907949. FCA is also supported by
NASA through the Origins of Solar Systems Program via grant
NNX07AP17G.
\end{acknowledgements}
|
\section{Introduction}
\label{intr}
\setcounter{equation}{0}
One of the important assumptions of the
classical Black-Scholes theory is the assumptions that any
trading strategy of any trader on the market do not affect asset
prices. This assumption is failed in the presence of large
traders whose orders involve a significant part of the available
shares. Their trading strategy has a strong feedback effect on
the price of the asset, and from there back onto the price of
derivative products. The continuously increasing volumes of financial markets
as well as a significant amount of large traders acting on
these markets force us to develop and to
study new option pricing models.\\
There are a number of suggestions on how to incorporate in a
mathematical model the feedback effects which correspond to
different types of frictions on the market like illiquidity or
transaction costs. Most financial market models are characterized
by nonlinear partial differential equations (PDEs) of the
parabolic type. They contain usually a small perturbation
parameter $\rho$ which vanishes if the feedback effect is removed.
If $\rho$ tends to zero then the corresponding nonlinear PDE tends
to the Black-Scholes equation.
Some of the option pricing models in illiquid markets possess
complicated analytical and algebraic structures which are singular
perturbed. We deal with singular perturbed PDEs if one of the
nonlinear terms in the studied equation incorporates the highest
derivative multiplied by the small parameter $\rho$. It is a
demanding task to study such models. Solutions to a singular
perturbed equation may blow up in the case $\rho=0$ and may not
have any pendants in the linear
case.\\
An example of a singular perturbed model is the continuous-time model developed by Frey
\cite{bib:frey-98a}. He derived a PDE for perfect replication
trading strategies and option pricing for the large traders. An
option price $u(S,t)$ in this case is a solution to the nonlinear
PDE
\begin{equation} \label{frey}
u_t + \frac{1}{2} \frac{\sigma^2 S^2 u_{SS}}{\left(1 - \rho
\lambda(S) S u_{SS}\right)^2} =0 \,,
\end{equation}
where $t$ is time, $S$ denotes the price and $\sigma$ the
volatility of the underlying asset. The continuous function
$\lambda(S)$ included in the adjusted diffusion coefficient
depends on the payoff of the derivative product. The Lie group
analysis and properties of the invariant solutions to Eq.
(\ref{frey}) for different types of the function $\lambda(S)$ were
studied in \cite{Bordag}-\cite{BordagFrey}. The analytic form of
the invariant solutions to this model allow us to follow up the
behavior of these solutions.
Under the similar assumptions Cetin, Jarrow and Protter
\cite{Protter} developed a model which includes liquidity risk for
a large trader. Liquidity risk is the additional risk due to the
timing and size of trade. The value $u(S,t)$ of a self financing
trading strategy for the large trader in this setting is a
solution of the following nonlinear PDE
\begin{equation}
u_t+\frac{1}{2}\sigma^2 S^2 u_{SS}(1-\rho S u_{SS})^2=0.
\label{jarrowProtter}
\end{equation}
This equation seems to be simpler as the previous one but it is
still a singular perturbed PDE. The Lie group analysis and the
symmetry algebra admitted by this equation were studied in
\cite{Bobrov}.\\
Sircar and Papanicolaou in \cite{bib:papanicolaou-sircar-98}
present a class of nonlinear pricing models that account for the
feedback effect from the dynamic hedging strategies on the price
of asset using the idea of a demand function of the reference
traders relative to the supply. They obtain a nonlinear PDE of the
following type
\begin{equation}\label{sircpapgen}
u_{t}+\frac{1}{2}\left[\frac{U^{-1}(1-\rho u_{S})U'(U^{-1}(1-\rho
u_{S}))}{U^{-1}(1-\rho u_{S})U'(U^{-1}(1-\rho u_{S}))- \rho S
u_{SS}}\right]^2\sigma^2S^2 u_{SS}+r(S u_{S}-u)=0,
\end{equation}
where $t$ is time, $S$ and $\sigma$ is the price and the
volatility of the underlying asset respectively, and the parameter
$r$ is the risk-free interest rate. The value $u(S,t)$ is the
price of the derivative security and depends on the form of the
demand function $U(\cdot)$. The expression $U^{-1}(\cdot)$ denotes
the correspondingly inverse function, because of the strong
monotonicity of the demand function the existence of the inverse
function $U^{-1}(\cdot)$ is guaranteed. In the bulk of their paper
\cite{bib:papanicolaou-sircar-98} authors studied the particular
model arising from taking $U(\cdot)$ as linear, i.e. $U(z)=\beta
z, ~ \beta>0 $. The authors mainly focused on the numerical
solution and discuss the difference to the classical Black-Scholes
option pricing theory.
In the present paper we study a more general case in which the
demand function of the type $U(z)=\beta z^{\alpha}, ~\alpha,\beta
\ne 0 $ is incorporated. Consistency of (\ref{sircpapgen}) with
the Black-Scholes model characterizes the class of the admitted
demand functions and leads to the condition $U'(z)=\beta \alpha
z^{\alpha-1}>0$.
In this case the model (\ref{sircpapgen}) takes the form
\begin{equation} \label{sipar}
\frac{\partial u}{\partial t}+\frac{1}{2}\left[\frac{1-\rho
\frac{\partial u}{\partial S}}{1-\rho \frac{\partial u}{\partial
S}-\frac{\rho}{\alpha} S\frac {\partial^2 u}{\partial
S^2}}\right]^2\sigma^2S^2\frac{\partial^2 u}{\partial
S^2}+r\left(S\frac{\partial u}{\partial S}-u\right)=0.
\end{equation}
The diffusion coefficient in Eq. (\ref{sipar}) depends on both,
$u_S$ and $u_{SS}$ multiplied by the small perturbation parameter
$\rho$. It depends also on the parameter $ \alpha $ which
characterizes the type of the demand function and on the interest
rate $r$.
We give a short overview of analytical properties of
Eq. (\ref{sipar}) in the next Section 2. In Section 3 we provide
the Lie group analysis of this equation. Depending on whether the
interest rate $r=0$ or $r \ne 0$, we obtain different Lie
algebras admitted by the respectively equation. Then we provide
optimal systems of subalgebras in the both cases. The optimal
systems of subalgebras give us the possibility to describe the set
of independent reductions of these nonlinear PDEs to different
ordinary differential equations (ODEs). In some cases we found the
explicit solutions to these equations. We discuss the properties
of invariant solutions in Section 4.
\section{Basic analytical properties of Eq. (\ref{sipar})}
\label{analyt}
\setcounter{equation}{0}
In the model (\ref{sipar}) introduced by Sircar and Papanicolaou
in \cite{bib:papanicolaou-sircar-98} the diffusion coefficient has
a very complicated analytical and non trivial algebraic structure.
In particular the diffusion term is represented by a fraction
which contains derivatives $u_S(S,t), u_{SS}(S,t)$. The authors
analyzed some analytical properties of this equation in the case
$\alpha =1$ in the vicinity of the Black-Scholes equation and
considered a valuation of a European derivative security with a
convex payoff using their model. They give asymptotic results for
when the volume of assets traded by the large traders is small
compared to the total number of unites of
the asset.\\
In this Section we discuss some global analytical properties of
Eq. (\ref{sipar}) for $\alpha \ne 1$.
We pay the main attention to the second term in (\ref{sipar}).
We assume that the space variable $S \in \Omega \cup \{
0\}$, where $\Omega = \mathbb{ R}^+ $ and the time variable $t$
lie in $ {\cal T}\cup \{ 0\}$, where ${\cal T}= \mathbb{R}^+ $.
This term can vanish for some values of the variable $S$ or on
some set of smooth functions and then the equation may change the
type from the parabolic one to another one. Other hand the
fraction in the second term may became meaningless because of
vanishing of the denominator on some set of smooth functions. We
should exclude such functions from the domain
of definition of our model.\\
The classical linear diffusion equation of type $u_t= u_{SS}$ is
well defined on the space $D= C^{2,1}(\Omega \times {\cal T})
\bigcap C(\{\Omega \cup \{0\} \} \times \{ {\cal T}\cup \{0\}
\})$ and $u(S,t)$ map the space $D$ to a space of continuous
functions $M=C(\{\Omega \cup \{0\} \} \times \{ {\cal T}\cup
\{0\} \}).$
Let us check whether the expression for the diffusion coefficient
in Eq. (\ref{sipar}) vanishes or has
singularities.\\
Fist we study the case that the denominator of the
fraction in (\ref{sipar}) is equal to zero, i.e. we have to
solve the equation
\begin{equation} \label{sipaDenominator}
1-\rho u_S - \frac{ \rho}{\alpha} S u_{SS} =0.
\end{equation}
It is easy to see that this equation has the following solution
\begin{eqnarray}
u_{sing}(S,t)= \frac{S}{\rho} + c_1(t)\frac{\alpha}{\alpha-1}
S^{\frac{\alpha-1}{\alpha}} +c_2(t),~~~ \alpha \ne 1, \nonumber \\
u_{sing}(S,t)= \frac{S}{\rho} + c_1(t)\ln (S) +c_2(t),~~ \alpha
=1,\label{solsing}
\end{eqnarray}
where $c_1(t)$ and $c_2(t)$ are arbitrary functions of $t$. We can
rewrite the expressions in (\ref{solsing}) as one expression which
includes the case $\alpha =1$ as a limit case. Then we obtain
\begin{equation}
u_{sing}(S,t)= \frac{S}{\rho} + c_1(t)\frac{\alpha}{\alpha-1}
\left(S^{\frac{\alpha-1}{\alpha}} -1 \right) +c_2(t).
\label{singtogether}
\end{equation}
The numerator of the second term in (\ref{sipar}) is equal to zero
if one of the equations is satisfied
\begin{eqnarray}
S^2 u_{SS}=0, \nonumber \\
1-\rho u_S =0. \label{numer}
\end{eqnarray}
The first equation is satisfied on all linear functions of $S$ and
in the point $S=0$, the second equation has the following solution
\begin{equation}\label{numer0}
u_{0}(S,t)= \frac{S}{\rho} + c_2(t).
\end{equation}
We notice that in the case $c_1(t)=0$ the functions
$u_{sing}(S,t)$ and $u_{0}(S,t)$ coincide. It means in this case
the numerator and the denominator of the fraction in the equation
(\ref{sipar}) are simultaneously equal to zero.\\
In the second step we should define a limiting procedure to
explain what we means if we say that (\ref{numer0}) is a solution
to (\ref{sipar}).
We chose in the space $D$ a one-parametric family of functions
$u_{\epsilon}(S,t)$ of the following type
\begin{equation}
u_{\epsilon}(S,t)=d_1(t) S + d_2(t) +\epsilon v(S,t),
\label{oneparam}
\end{equation}
where $ \epsilon \in \mathbb{R}$ is a parameter, the functions
$d_1(t), d_2(t)$ are arbitrary functions of time and $v(S,t) \in
D$. If now the parameter $\epsilon \to 0$ then the family of
functions of the type (\ref{oneparam}) converges in the norm of
the space $D$ to a linear function of $S$, i.e. to
$u_{0}(S,t)=d_1(t) S + d_2(t)$. We apply to this family the
differential operator defined by (\ref{sipar})
\begin{eqnarray} \label{ner}
d_1^{'}(t) S + d_2^{'}(t) + \epsilon v_t (S,t)+ \frac{\sigma^2}{2}
\frac{\epsilon S^2 v_{SS}(1- \rho d_1(t)- \epsilon \rho v_S)^2
}{\left(1- \rho d_1(t)- \epsilon \rho v_S - \epsilon \beta \rho
S v_{SS}\right)^2} =0 \,,
\end{eqnarray}
here $d_1^{'}(t), d_2^{'}(t) $ denotes the first derivatives of
the corresponding functions. From (\ref{ner}) follows that any
linear function of $S$ with constant coefficients will be solution
to equation (\ref{sipar}) if the last term in (\ref{ner}) is a
bounded function in the norm of the space $M$. If we replace the
linear part in (\ref{ner}) by $u_0(S,t)$ we see that this function
(\ref{numer0}) is a solution to (\ref{sipar}) if $c_2(t)= const.$.
The fraction in (\ref{ner}) is not well defined just in one case
if $v(S,t)$ coincide with the second term in (\ref{singtogether}).
So far we use as the domain of definition for our model the space
$D$ the functions of type (\ref{singtogether}) with $c_1(t) \ne 0$
are excluded because they or their derivatives have singularities
in the point $S=0$ and consequently they do not belong to the
space $D$.
In the classical case of a linear parabolic diffusion equation
solutions of type (\ref{singtogether}) which do not belongs to the
set of classical solutions are called viscosity solutions
\cite{Barles}, \cite{Grandal} and these solutions are well
studied.
We proved that the functions of type (\ref{singtogether}) should
be excluded from the further investigation because the model
(\ref{sipar}) is not well defined on them. The linear function
(\ref{numer0}) with $c_2(t)=const.$ is a solution to (\ref{sipar})
because any one parametric family of functions $u_{\epsilon}(S,t)$
in the norm of the space $D$ convergent to $u_{0}$ is mapped by
the differential operator defined by (\ref{sipar}) to a
zero-convergent family of functions in the norm of the space $M$.
\section{Symmetry properties of the model}
\label{sym}
\setcounter{equation}{0}
We provide in this Section the Lie group analysis of Eq.
(\ref{sipar}) first for the case $r=0$ then for $r \ne 0$. In both
cases it is possible to find the non-trivial Lie algebras
admitted by the equation. We use the standard method to obtain the
symmetry group suggested by Sophus Lie and developed further in
\cite{{Ovsiannikov}}, \cite{{Olver}} and \cite{Ibragimov}. In the
case $r=0$ we obtain a four dimensional Lie algebra $L_4$ and by
$r \ne 0 $ Eq. (\ref{sipar}) admits a three dimensional algebra
$L_3$ defined in the subsection 3.2.
All three and
four dimensional real Lie algebras and their subalgebras were
classified by Pattera and Winternitzs in
\cite{PateraWinternitzs1977}. The authors looked for
classifications of the subalgebras into equivalence classes under
their group of inner automorphisms. They used also the idea of
normalization which guarantees that the constructed optimal system
of subalgebras is unique up to the isomorphisms.
The symmetry group $G_4$ related to the symmetry algebra $L_4$ is
generated by a usual exponential map. We use the similar procedure
to obtain to each subalgebra $h_i$ from the optimal system of
subalgebras the correspondingly subgroup $H_i$.
The optimal system of subalgebras allows us to divide the
invariant solutions into non-intersecting equivalence classes. In
this way it is possible to find the complete set of essential
different invariant solutions to the equation under consideration.
Using the invariants of these subgroups we reduce the studied PDE
to different ODEs. Solutions to these ODEs give us the invariant
solutions to the nonlinear PDE (\ref{sipar}) in an analytical
form. In the both cases whether by $r=0$ or by $r\ne0$ we skip the
study of invariant reductions to the two- and three- dimensional
subgroups because of they give trivial results for the studied
equation.
\subsection{Symmetry reductions in the case $r=0$}
In the first step we solve the Lie determining equations for the
equation
\begin{equation} \label{sipa}
u_t + \frac{1}{2} \frac{\sigma^2(1-\rho u_S)^2 }{\left(1-\rho u_S
- \frac{\rho}{\alpha}
S u_{SS}\right)^2} S^2 u_{SS} =0 \,, {\alpha} \ne 0,
\end{equation}
and obtain the Lie algebra admitted by this equations. We
formulate the results in the following theorem.
{\begin{theorem} Eq. (\ref{sipa}) admits a four dimensional Lie
algebra $L_4$ with the following infinitesimal generators
\begin{equation}
e_1=- \frac{S}{2}\frac{\partial}{\partial S} +\left(\frac{S}{2
\rho} - u \right)\frac{\partial}{\partial u},
~~e_2=\frac{\partial}{\partial u}, ~~e_3=\frac{\partial}{\partial
t},~~~ e_4=\rho S \frac{\partial}{\partial S}+ S
\frac{\partial}{\partial u}. \label{generatorsSiPa}
\end{equation}
The commutator relations are
\begin{eqnarray}
[e_1,e_3]=[e_1,e_4]=[e_2,e_3]=[e_2,e_4]=[e_3,e_4]=0,~
[e_1,e_2]=e_2,
\end{eqnarray}
\end{theorem}
\begin{remark} In the very short letter \cite{Bordag:2008} there is a misprint in the theorem
formulation. We apologize by readers for the inconvenience.
\end{remark}
The Lie algebra $L_4$ has a two-dimensional subalgebra
$L_2=<e_1,e_2>$ spanned by the generators $e_1,e_2$. The algebra
$L_4$ is a decomposable Lie algebra and can be represented as a
semi-direct sum
$L_4=L_2 \bigoplus e_3 \bigoplus e_4$.
The optimal system of subalgebras for $L_4$ were provided in
{\cite{PateraWinternitzs1977} and presented in Table \ref{optsev}.
\begin{table}
{\begin{tabular}{|c|c|}
\hline
{Dimension}&{Subalgebras}\\
\hline
$1$&$h_1=<e_2>,~~h_2=<e_3 \cos {(\phi)}+e_4\sin{(\phi)}>,$\\
&$h_3=<e_1+ x(e_3 \cos {(\phi)}+e_4\sin{(\phi)})>,$\\
&$h_4=<e_2 +\epsilon(e_3 \cos{(\phi)}+e_4\sin{(\phi)})>$\\
\hline
$2$&$h_5=<e_1+x(e_3 \cos {(\phi)}+e_4\sin{(\phi)}),e_2>,~~h_6=<e_3,e_4>,$\\
&$h_7=<e_1 +x(e_3 \cos {(\phi)}+e_4\sin{(\phi)}),e_3 \sin{(\phi)}-e_4\cos{(\phi)}>,$\\
&$h_8=<e_2 +\epsilon(e_3 \cos{(\phi)}+e_4\sin{(\phi)}),e_3 \sin{(\phi)}-e_4\cos{(\phi)}>,$\\
&$h_9=<e_2,e_3 \sin {(\phi)}-e_4\cos{(\phi)}>$\\
\hline
$3$&$h_{10}=<e_1,e_3,e_4>,~~h_{11}=<e_2,e_3,e_4>,$\\
&$h_{12}=<e_1 +x(e_3 \cos {(\phi)}+e_4\sin{(\phi)}),e_3 \sin {(\phi)}-e_4\cos{(\phi)},e_2>$\\
\hline
\end{tabular}}
\caption{\cite{PateraWinternitzs1977} The optimal system of
subalgebras $h_i$ of the algebra $L_4$ where $x\in {\mathbb
R},~~\epsilon=\pm1,~~\phi \in [0,\pi ] $.} \label{optsev}
\end{table}
The optimal system of the one-dimensional subalgebras involves
four subalgebras $h^0_i,~i=1,\dots,4$. We take step-by-step each
of these subalgebras $h^0_i$ and the corresponding symmetry
subgroup $H^0_i$ and study which invariant reductions of the studied PDE are possible.\\
{\bf Case $H^0_1$.} This one-dimensional subgroup $H^0_1 \subset
G_4$ is generated by the subalgebra $
h^0_1=<e_2>=<\frac{\partial}{\partial u}> $. It means that we deal
with a subgroup of translations in the $u$ - direction. Hence, to
each solution to Eq. (\ref{sipa}) we can add an arbitrary constant
without destroying the property of the function to be solution.
This subgroup does not provide any
reduction.\\
{\bf Case $H^0_2$.} The subalgebra $h^0_2$ is spanned by the
generator $e_3 cos(\phi) +e_4 sin(\phi)$. In terms of the
variables $S,t,u$ it takes the form
\begin{equation} \label{h2inr0}
h^0_2= <\frac{\partial}{\partial t}\cos (\phi)+\left( \rho S
\frac{\partial}{\partial S}+ S \frac{\partial}{\partial u}\right)
\sin (\phi) >.
\end{equation}
The invariants $z,w$ of the corresponding subgroup $H^0_2$ are
equal to
\begin{eqnarray} \label{invh2r0}
z=S\exp(-t \rho \tan(\phi) ),~~ w = u-\frac{1}{\rho}S
\end{eqnarray}
and we take them as the new dependent and independent variables,
respectively. Then the PDE (\ref{sipa}) is reduced to the ordinary
differential equation of the following form
\begin{equation} \label{redeqh2r0}
- \rho {\delta} zw_z + \frac{1}{2}\sigma^2w_{zz}z^2\left(\frac{\alpha w_z}{\alpha
w_z+zw_{zz}}\right)^2=0,~
\delta=\tan(\phi), \phi \in [0,\pi], \phi \ne \frac{\pi}{2}.
\end{equation}
This second order ODE is reduced to the first order equation by
the substitution $w_z=v(z)$ which takes the form
\begin{equation}\label{redFirsteqh2r0}
z v \left( - \rho \delta +\frac{\sigma^2}{2} z (\ln{v})_z \left(
\frac{\alpha}{z (\ln{v})_z +\alpha} \right)^2 \right)=0.
\end{equation}
Eq. (\ref{redFirsteqh2r0}) has two trivial solutions $z=0$ and
$v(z)=w_z=0$ which are not very interesting for applications. The
non trivial solutions we obtain if we set the last factor in Eq.
(\ref{redFirsteqh2r0}) equal to zero. We obtain the solution to
(\ref{redFirsteqh2r0}) in the form
\begin{equation}\label{h2solr01}
w(z)= c_1 z^p, ~~ p=1+\alpha -a \pm \sqrt{a(a-2 \alpha)},~
a=\frac{\sigma^2}{4 \rho \delta},
\end{equation}
where $c_1$ is an arbitrary constant. In terms of the variables
$S,t,u$ the solution (\ref{h2solr01}) is equivalent to the
following solution to Eq.(\ref{sipa})
\begin{equation}\label{h2solr0}
u(S,t)= c_1 S^p \exp(-p \rho \tan(\phi) ~t), ~~ p=1+\alpha - a \pm
\sqrt{a(a - 2 \alpha)},~ a=\frac{\sigma^2}{4 \rho \delta},
\end{equation}
where $ \delta=\tan(\phi), \phi \in [0,\pi], \phi \ne
\frac{\pi}{2} $.\\
{\bf Case $H^0_3$.} The subalgebra $h^0_3$ is spanned by
$$h^0_3=<x
\cos (\phi)\frac{\partial}{\partial t}+\left(x\rho\sin (\phi)-
\frac{1}{2}\right)S\frac{\partial}{\partial S}+
\left(\left(\frac{1}{2\rho}+x\sin(\phi)\right) S-u\right)\frac{\partial}{\partial
u}>.$$
The invariants $z,w$ of the corresponding subgroup $H^0_3$ are
given by the expressions
\begin{equation}\label{invh3r0}
z=Se^{-ct},~~ w=S^b u-\frac{1}{\rho}S^{1+b},
\end{equation}
where $b=(x\rho\sin (\phi)-\frac{1}{2})^{-1},$ $c=({b
x\cos(\phi)})^{-1}$, $x \ne 0$, $\phi \in [0,\pi], \phi \ne
\frac{\pi}{2}$ and $x\rho\sin (\phi)-\frac{1}{2}\ne 0$.\\
We use $z,w$ as the new invariant variables and reduce
Eq.(\ref{sipa}) to the ODE
\begin{equation}\label{redh3r0}
- c z w_z +\frac{\sigma^2 \alpha^2}{2} \frac{\left( z^2 w_{zz} -2
{b} z w_z +{b}\left( 1+{b}\right)w \right) \left( z w_z - {b}w
\right)^2 } {\left(\alpha \left( z w_z - {b}w \right)+ \left( z^2
w_{zz} -2 {b} z w_z +{b}\left( 1+ {b}\right)w \right)\right)^2}=0.
\end{equation}
Eq. (\ref{redh3r0}) admits a solution of the type
\begin{equation} \label{solal}
w(z)=c_1 z^q,
\end{equation}
where $q$ is a real root of the polynomial of the degree 5
\begin{eqnarray}\label{rootredh3r0}
- c q {\left(
q (q-1) + \left(\alpha-2 {b}\right) q +{b}\left( 1+
{b}-\alpha \right) \right)^2} \nonumber\\+\frac{\sigma^2
\alpha^2}{2} {\left( q (q-1) -2 {b}q +{b}\left( 1+{b}\right)
\right) \left( q - {b} \right)^2 } =0.\nonumber
\end{eqnarray}
Eq. (\ref{redh3r0}) has a complicate structure and is hardly
possible to solve it in the general form. But by a special values
of involved parameters we can simplify the equation and obtain
some particular classes of solutions.\\
We take the special case of Eq. (\ref{sipa}) with $\alpha=1$.\\
Under the special choice of the
parameters $\phi=0,\pi$ and $b=-2$ in Eq. (\ref{invh3r0}) and by
using the invariants $z,w$ in the form
$$z= S \exp
\left(\frac{1}{2 x} b t \right), x\ne 0, ~~w= S^{-2} u - (\rho
S)^{-1}, $$ we reduce Eq.(\ref{sipa}) to the ODE
\begin{equation}\label{redfi}
z w_z ((z(z^2 w)_{z})_z)^2 + \sigma^2 x~~(z^2 w )_{zz} ((z^2
w)_z)^2=0.
\end{equation}
The substitution (\ref{solal}) in this case leads to the second
order algebraic equation on the value of the parameter $q$
\begin{equation}\label{reh1}
q^2 + q \left(2 + \sigma^2 x \cos (\phi)\right) + \sigma^2 x \cos
\phi=0, ~~q \ne -2, \phi=0,\pi,
\end{equation}
which has two roots $q_1=- \sigma^2 x \cos (\phi)$ and $q_2=-2.$
For the future study we take just the first value $q_1$. The value
$q_2=-2$ leads to the know solution $u_0(S,t)$ (\ref{numer0}).
Since by $q_2=-2$ both, the numerator and the denominator in the
fraction in (\ref{sipa}) vanish, and the solution (\ref{solal})
coincide then with $u_{0}(S,t)$ by $c_2(t)=const.$ which we discussed in the previous section.\\
We notice that the solution (\ref{solal}) differs from the
function $u_{sing}(S,t)$ (\ref{solsing}) which involves the
logarithmic term in the case $\alpha=1$.
The solution to (\ref{redfi}) or respectively to Eq.(\ref{sipa})
in the form (\ref{solal}) in terms of $S,t,u$ variables is equal
to
\begin{equation} \label{solfirst}
u(S,t)= \frac{S}{\rho} + C_1 S^{2- \sigma^2 x }
e^{-\frac{\sigma^2}{2} t} + C_2 ,~~ \alpha \ne -2,
\end{equation}
where $C_1,C_2$ are arbitrary constants, $x \in \mathbb R$ and the
first term is the only term which contains the dependency on the
parameter $\rho$. We skip the factor $\cos (\phi)$ by $x$ in this
expression because $\cos (\phi)=\pm 1$ in the case $\phi = 0,\pi$
and $x\in \mathbb R$.
It is remarkable that the reduced equation (\ref{redfi}) does not
contain any more the parameter $\rho$. Hence all invariant
solutions of this class can be represented as a sum of two terms:
the first one is equal to $S/{\rho}$ and the second one depends on
$z$ only but not on the parameter $\rho$.
If we left the values of parameters like in the previous case, but take
the invariants in another form
$z=\ln{S}+t b/4 $ and $w(z)=(u/S - 1/{\rho})S^{\gamma}$
than we obtain the different form of the reduced ODE
\begin{eqnarray}\nonumber
w_{zz}^2 w_z +w_{zz}(w^2_{z}(4(1-\gamma)+\kappa)+w_{z} w (2
(1-\gamma)^2 +2 \kappa (1-\gamma))\\ \nonumber +w^2 \kappa
(1-\gamma)^2 )+ w_z^3 (4(1- \gamma )^2+\kappa(1-2 \gamma)) +w_z^2
w (1-\gamma)(4 (1-\gamma )^2 \\+ \kappa (2 - 5 \gamma))
+ w_z w^2 (1-\gamma)^2
((1-\gamma )^2+\kappa (1-4 \gamma)) -\kappa \gamma (1-\gamma)^3
w^3=0,\label{seconred}
\end{eqnarray}
where $\kappa =2 \sigma^2/b$. This second order equation can be
reduced in the case $w_{z}\ne 0 $ to a first order ODE. We
substitute $p(w)=w_z(z(w))$ and correspondingly $w_{zz}=p_w p$,
i.e., $w$ is the independent variable and $p$ is the dependent
variable in this case. Then we obtain the first order ODE
\begin{eqnarray} \nonumber
p_w^2 p^3+ p_w p(p^2(4(1-\gamma)+\kappa)+p w (2 (1-\gamma)^2 +2
\kappa (1-\gamma)) +w^2 \kappa (1-\gamma)^2 )\\\nonumber + p^3
(4(1- \gamma )^2+\kappa(1-2 \gamma)) +p^2 w (1-\gamma)(4 (1-\gamma
)^2 + \kappa
(2 - 5 \gamma))\\
+ p w^2 (1-\gamma)^2
((1-\gamma )^2+\kappa (1-4 \gamma)) -\kappa \gamma (1-\gamma)^3
w^3=0.\label{seconredfirstod}
\end{eqnarray}
This equation is quadratic in the first derivative $p_w$ and it is
equivalent to the system of two first order equations. For some
values of the constants $\gamma$ and $\kappa$ it can be explicitly
solved. The simplest case we obtain if we chose $\gamma =1$ then
the solution coincide with (\ref{solfirst}). In other cases the
equation can be studied using qualitative methods.
{\bf Case $H^0_4$.} We
consider subalgebra $h^0_4$ spanned by
$$h^0_4=<\epsilon \cos
(\phi)\frac{\partial}{\partial t}+\epsilon\rho S \sin (\phi)
\frac{\partial}{\partial S}+
(1+\epsilon S \sin (\phi))\frac{\partial}{\partial u}>.$$
The invariants $z,w$ are given by the expressions
\begin{eqnarray}
z&=&S\exp(-t \rho \tan(\phi) ),~~~ \phi \in (0,\pi), \phi \ne \frac{\pi}{2},\nonumber\\
w&=&\frac{\epsilon}{\rho\sin(\phi)}\ln S+\frac{1}{\rho}S -V,
~~\epsilon = \pm 1.\nonumber
\end{eqnarray}
Using these expressions as the independent and dependent variables
we reduce the original equation to the ODE
\begin{equation}\label{redh4r0}
- \rho\tan (\phi) w_z
z+\frac{1}{2}\sigma^2(w_{zz}z^2+a)\left[\frac{\alpha( w_z
z-a)}{\alpha( w_z z- a)+a+w_{zz}z^2}\right]^2=0,
\end{equation}
where $a=(\epsilon\rho\sin(\phi))^{-1}$ and $\phi \in (0,\pi),
(\phi) \ne \frac{\pi}{2}$. Eq. (\ref{redh4r0}) is possible to
reduce to the first order ODE
\begin{equation}\label{firstredh4r0}
(v+a)\left(z v_z -(1- \alpha) v + a \right)^2
-\frac{\sigma^2\alpha^2}{2 \rho\tan (\phi)} v^2 (z v_z -v +a)=0
\end{equation}
after substitution $v(z)=z w_z-a$. It is a quadratic algebraic
equation on the value $z v_z$ which roots depends on $v$ only. If
we denote the roots as $f_{\pm}(v)$ we represent the solutions to
Eq. (\ref{firstredh4r0}) in the parametric form
\begin{equation}\label{solfirstredh4r0}
\int{\frac{{\rm d} v}{f_{\pm}(v)}}=\ln{z} +c_1, ~~c_1 \in {\mathbb
R }.
\end{equation}
\subsection{Symmetry reductions in the case $r\ne 0$}
{\begin{theorem} Eq. (\ref{sipar}) admits a three dimensional Lie
algebra $L_3$ with the following operators
\begin{equation}\label{generatorsSiPar}
e_1=\frac{\partial}{\partial t},~~e_2=\left( S - \rho \right)
\frac{\partial}{\partial u},~~ e_3= S \frac{\partial}{\partial S}
+ u \frac{\partial}{\partial u}.
\end{equation}
The algebra $L_3$ is abelian.
\end{theorem}
Similar to the previous investigation for $r=0$ first we find an
optimal systems of subalgebras for the algebra $L_3$. Because we
are interested just in the one-dimensional subalgebras we provide
the optimal system for these subalgebras only
\begin{eqnarray} \label{optsyssipar}
h_1=<e_3>,\;\;h_2=<e_2 +x e_3>,\:\:h_3=<e_1+x e_2+y e_3 >.
\end{eqnarray}
First we provide for each of these three one-dimensional
subalgebras $h_i, ~i=1,2,3$ and the corresponding subgroup $H_i,
~i=1,2,3$ a set of invariants. Then we use the invariants as the
new independent and independent variables and reduce Eq.
(\ref{sipar}) to some ODEs. In the cases where it is possible we
solve the ODEs.\\
{\bf Case $H_1$}. The algebra $h_1$ is spanned by
\begin{equation}
h_1=<S \frac{\partial}{\partial S} + u \frac{\partial}{\partial
u}>.
\end{equation}
It describes scaling symmetry of Eq. (\ref{sipar}) and means that
if we multiply the variable $S$ and $u$ with one and the same non
vanishing constant the equation will be unaltered. Respectively
the invariants of this transformations are
\begin{equation}
z=t, ~~ w=\frac{u}{S}.
\end{equation}
If we use these expressions as the new invariant variables we
obtain a rather simple reduction of the original equation
\begin{equation}
w_z=0,
\end{equation}
with the trivial solution $w=c_1=const.$ It describes all
solutions to Eq. (\ref{sipar}) of the type
\begin{equation}
u(S,t)=Sw(z)=Sw(t)=S c_1.
\end{equation}
{\bf Case $H_2$}. The second subalgebra $h_2$ in the optimal
system of subalgebras (\ref{optsyssipar}) is given by
\begin{equation}
h_2=< x~ S\frac{\partial}{\partial
S}+\left(S+u(x-\rho)\right)\frac{\partial}{\partial u}
>, ~~x \in {\mathbb R}.
\end{equation}
The invariants of the subgroup $H_2$ are
\begin{eqnarray} \label{invh2sipar1}
z=t,~~
w=-\frac{S^{\frac{\rho}{x}}}{\rho}+S^{\frac{\rho}{x}-1}u,~x\ne 0.
\end{eqnarray}
If we use the invariants (\ref{invh2sipar1}) as the new dependent
and independent variables then the reduced Eq. (\ref{sipar})
takes the form
\begin{equation}
w_z- \rho \left( \frac{\sigma^2 \alpha^2 (x -\rho)}{2 (\alpha x -
\rho )^2} + \frac{r}{x}\right)w=0.
\end{equation}
It has the solution
\begin{equation}
w(z)=c_1 \exp { \left( \rho \left( \frac{\sigma^2 \alpha^2 (x
-\rho)}{2 (\alpha x - \rho )^2} +\frac{r}{x}\right)\right) t},
~~c_1,x \in {\mathbb R}, x\ne 0.
\end{equation}
The corresponding solution to Eq. (\ref{sipar}) in terms of
variables $S,t,u$ is given by
\begin{equation}
u(S,t)= c_1 S^{1-\frac{\rho}{x}}e^{\rho \gamma t} +\frac{1}{\rho}
S, ~~~~\gamma= \frac{\sigma^2 \alpha^2 (x -\rho)}{2 (\alpha x -
\rho )^2} +\frac{r}{x},
\end{equation}
where $c_1,x \in {\mathbb R}, x\ne 0$ are arbitrary constants.
{\bf Case $H_3$}. The last subalgebra $h_3$ from the optimal
system (\ref{optsyssipar}) is spanned by
\begin{equation}
h_3=<\frac{\partial}{\partial t}+y~S\frac{\partial}{\partial
S}+\left(x~S-x\rho u+y~u\right)\frac{\partial}{\partial u}>,~~ x,y
\in {\mathbb R} .
\end{equation}
We take the two invariants $z,w$
\begin{eqnarray}
z=Se^{-ty},~~w=uS^{\kappa \rho -1}-\frac{1}{\rho}S^{\kappa \rho},~
y\ne 0 ,~~\kappa=\frac{x}{y} .\nonumber
\end{eqnarray}
as the new independent and dependent variables reduce Eq.
(\ref{sipar}) to the ODE
\begin{eqnarray}
&{}&(r-y)zw_z -{r\kappa \rho} w\label{h3ne0}\\
&+&\frac{\alpha^2\sigma^2}{2}\frac{(z^2w_{zz}+2(1-\kappa\rho)zw_z-\kappa\rho(1-\kappa\rho)w)(zw_z+(1-\kappa
\rho)w)^2}{(z^2w_{zz}+(\alpha +2(1-\kappa \rho)zw_z+(1-\kappa
\rho)(\alpha-\kappa \rho)w)^2}=0. \nonumber
\end{eqnarray}
Eq. (\ref{h3ne0}) possesses a solution of the type
\begin{equation}
w(z)=c_1 z^{q}, ~~c_1 \in \mathbb R,
\end{equation}
where $q$ is a real root of the fifth order algebraic equation
\begin{eqnarray}
((r-y)q -{r\kappa \rho}){(q(q-1)+(\alpha +2(1-\kappa
\rho)q+(1-\kappa \rho)(\alpha-\kappa \rho))^2} \nonumber\\
+\frac{\alpha^2\sigma^2}{2}
{(q(q-1)+2(1-\kappa\rho)q-\kappa\rho(1-\kappa\rho))(q+(1-\kappa
\rho))^2} =0.\label{h3ne0q}
\end{eqnarray}
Respectively the solution to Eq. (\ref{sipar}) takes in this case
the form
\begin{equation}
w(z)=c_1 S^{q+1-\kappa \rho}e^{-y q t} +\frac{1}{\rho}S,
~~~c_1,y,\kappa \in \mathbb R, y\ne 0.
\end{equation}
\section{Conclusions}
\label{intr}
\setcounter{equation}{0}
In the previous sections we studied the Sircar-Papanicolaou model
(\ref{sircpapgen}) in the case of the nonlinear demand function
$U(z)=\beta z, ~ \beta>0 $. The model which includes the linear
demand function was studied in \cite{bib:papanicolaou-sircar-98}
with numerical methods. In this paper we use the methods of Lie
group analysis which gives us a general point of view on the
structure of this equation. We found the symmetry algebra admitted
by the nonlinear PDE (\ref{sipar}) for $r\ne 0$ and by Eq.
(\ref{sipa}) in the case $r=0$. We present in the both cases the
optimal systems of subalgebras. Using the optimal systems of
subalgebras we provide the complete set of non-equivalent
reductions. In most cases we solve the ODEs or present particular
solutions to them and respectively to Eqs. (\ref{sipar}) and
(\ref{sipa}). The explicit and parametric solutions can be used as
benchmarks for numerical methods.
|
\section{Model}
We investigate a model of the Galactic Centre which contains following
constituents:
{\bf{}Supermassive black hole (SMBH)} of mass $M_{\bullet} = 3.5\times10^6M_{\odot}$
dominates the gra\-vi\-ta\-tional field and it is approximated by a fixed
Keplerian potential.
{\bf{}Circum-nuclear Disc (CND)} is modelled by several tens of particles
orbiting the SMBH in a toroidal structure. Its total mass is not well
determined by observations; we assume $M_\mathrm{CND} \approx 0.2M_{\bullet}$
orbiting at a characteristic radius of $R_\mathrm{CND} \approx 1.5\mathrm{pc}$.
{\bf{}Stellar cusp} of mass $\gtrsim0.1M_{\bullet}$ within $1\mathrm{pc}$ is assumed to be
spherically symmetric. We incorporate it by means of a smooth fixed potential.
{\bf{}Young stellar disc} is a set of $\sim100$ equal-mass gravitating particles.
Initially, they are on circular orbits with nearly colinear angular momenta.
Semi-major axes have distribution $\propto a^{-1}$
within the interval $\langle 0.04\mathrm{pc}, 0.4\mathrm{pc} \rangle$.
Provided the CND would be the only perturbation to the SMBH's potential,
individual stellar orbits would undergo secular evolution.
In particular, orbital eccentricity and inclination would oscillate on the
time-scale comparable to the age of the young stars.
These oscillations are considerably damped if additional {\em spherically\/}
symmetric perturbation of the late type stellar cusp is present. This
component, however, does not suppress
differential precession of the stellar orbits around the symmetry axis of
the CND. The precession rate can be approximated as \cite{subr09}
$$
\frac{\mathrm{d}\Omega}{\mathrm{d} t} \approx -{\textstyle\frac{3}{4}}\,
\cos i \, \frac{M_\mathrm{CND}}{M_{\bullet}} \, \frac{a\sqrt{GM_{\bullet} a}}{R_\mathrm{CND}^3} \,
\frac{1+\frac{3}{2}e^2}{\sqrt{1-e^2}} \approx const\,.
$$
There is a strong dependence of $\mathrm{d}\Omega / \mathrm{d} t$ on the semi-major
axis and the orbit inclination $i$ with respect to the plane of the CND.
Stars at the outer part of the disc and/or stars that are less perpendicular
to the CND undergo faster precession, i.e. the disc becomes warped. The shape
of the stellar disc in terms of the normal
vectors of the orbits after several Myrs of such a differential precession
is shown in the right panel of Fig.~\ref{fig:dissolve}.
We have verified \cite{subr09} that such a model of a warped stellar disc is
compatible with the observational data available in the literature
\cite{paumard06,bartko09}.
\begin{figure}
\includegraphics[width=\textwidth]{subr-gc09-fig1.eps}
\caption{Dissolution of the coherently rotating structure due to the gravity
of stellar cusp and the CND in terms of the direction of the angular momenta.
Self-gravity of the stellar disc is considered
only in the right panel, while it is omitted in
the left one. Normal vectors of all orbits lie within the shaded area
at $T=0$; crosses represent their position at $T=7\mathrm{Myr}$. Coordinate system
is aligned with the axis of the CND.}
\label{fig:dissolve}
\end{figure}
The picture changes when self-gravity of the stellar disc is taken into
account (see Fig.~\ref{fig:dissolve}). It appears that mutual torques of the
stellar orbits tend to drive
the coherently rotating `core' of the disc towards the orientation perpendicular
to the CND. At the same time, two-body relaxation scatters some stars to orbits
that undergo fast precession and do not apparently belong to the parent structure
after few Myrs.
Another consequence of the gravitational interaction
of the stellar orbits with both the CND and stellar disc itself is an
enhanced growth of orbital eccentricities \cite{haas10}.
\section*{Conclusions}
We suggest that all young stars (probably except for S-stars) in the Galactic
Centre could have been formed in a single thin self-gravitating disc.
Perturbative influence of the CND then led to partial destruction of the
coherently rotating structure. In addition to our previous results, we have
found that the core of the disc has tendency to migrate towards orientation
perpendicular to the CND, which is in a striking accord with the data provided
by the observations. This result cannot be achieved if the gravity of either
the CND, the spherical cusp or the stellar disc is ignored.
\acknowledgements
This work was supported by the Czech Science Foundation (ref.\ 205/07/0052)
and the Centre for Theoretical Astrophysics in Prague.
|
\section{Introduction}\label{sec:intro}
\indent The Galactic Centre (GC) hosts at least one clockwise rotating disc of O and Wolf-Rayet stars, most likely a result of the gravitational fragmentation of a gaseous accretion disc \citep{Lev03}. These stars are located at projected distances of $0.05 - 0.5$ pc, and have an average age of $\sim 6$ Myr \citep{Gen03a, Pau06,Lu09,Bar09}. Recent hydrodynamical simulations suggest that young stars forming in such discs can retain similar initial orbital eccentricities, which are often non-zero \citep[e.g.][]{BoR08}. The dynamics of stars in such eccentric discs can be very different from circular structures. In particular there exists an instability \citep{Mad09} that occurs in coherently eccentric discs which are embedded in nuclear stellar clusters such as the one in our GC (see review by Sch\"odel in this volume). This instability can propel stars to very high, or low, orbital eccentricities within $\sim 1$ Myr, overcoming the long relaxational time scales associated with galactic nuclei.
The background stellar cluster (or cusp) is an integral part of the instability. In this paper we identify the reasons for this, vary the density profile of the cluster to see the effect it has on the instability, and address the possible implications of the observed flattened density profile of the cusp of late-type stars in the GC \citep{Buc09, Do09, Bar09}.
\section{The Eccentric Disc Instability}\label{sec:EDI}
To describe the physics of the instability we work with a simplified template of a galactic nucleus. We consider a disc of stars with mass $M_{\rm disc}$ orbiting a MBH of mass $M_{\bullet}$. Both are embedded in a power-law stellar cusp of mass $M_{\rm cusp}$.
We make the assumptions that (1) $M_{\rm disc}\ll M_{\rm cusp}$, such that the apisidal precession of stellar orbits within the disc is driven by the cusp, and (2) initially each stellar eccentricity vector (which points to periapse) is aligned and similar in magnitude to those of the other stars in the disc. The stellar orbits precess with retrograde motion due to the presence of the cusp, i.e., in the direction opposite to the orbital rotation of the stars. The precession time scales as
\begin{equation} \label{eqn:t_prec}
t_{\rm prec} \sim \frac{M_{\bullet}} {N( < a) m} P(a) f(e) \quad \propto a^{\gamma-3/2} f(e) ,
\end{equation}
\noindent where $P(a) = 2 \pi \sqrt{a^3/G M_{\bullet}}$ is the period of a star with
semi-major axis $a$, $N(<a)$ is the number of stars in the cusp within $a$, $m$ is the individual mass of the stars, and $\gamma$ is the power-law index for the space-density profile of the cusp $\rho(r) \propto r^{-\gamma}$. For simplicity let us assume a power-law index $\gamma = 3/2$, such that $t_{\rm prec}$ is constant for all values of $a$. The key element of the instability is that $f(e)$ in Equation (\ref{eqn:t_prec}) is an increasing function of eccentricity, $df/de > 0$.
\begin{figure}[!ht]
\begin{center}
\includegraphics[scale = 0.55]{f1.eps}
\caption{This schematic shows the stellar disc from above its orbital plane. The stars move counterclockwise on their orbits and precess in the opposite direction (dashed arrow). A ``test star" is also shown, with a more eccentric orbit, and lagging behind in precession. The direction of its velocity vector, ${\bf v}$, at apoapse is opposite to the coherent gravitational force, ${\bf F}$, it feels from the rest of the stars in the disc. This causes the torque, ${\bf \tau} = {\bf r} \times {\bf F}$, acting on the star from the disc to decrease its angular momentum, ${\bf J} = {\bf r} \times {\bf v}$.
\label{f:torque_diagram}}
\end{center}
\end{figure}
The instability works as follows: A star that has a slightly more eccentric orbit than the average star in the disc consequently has a larger $t_{\rm prec}$ and lags behind in precession; see Figure (\ref{f:torque_diagram}). Such a star feels a strong, coherent torque from the other stars in the disc, in the {\it opposite} direction of its angular momentum vector, $J = \sqrt{G M_{\bullet} a (1 - e^2)}$. As a result, its angular momentum decreases in magnitude, which is equivalent to saying that its eccentricity increases\footnote{Now, with an even higher eccentricity, the stellar orbit lags even further behind the bulk of the disc and the cycle repeats. We emphasize that the precession due to the cusp is a key element of this instability. Without it, stars in the eccentric disc would experience torques but in both directions and it would not be an unstable configuration.}. In this way, very high eccentricities can be achieved. Conversely, if a stellar orbit is slightly less eccentric than the average, it has a smaller $t_{\rm prec}$ and moves ahead in precession, thus experiencing a torque which decreases its eccentricity further below the average.
\subsection{Dependency on Background Stellar Cluster}
The stellar orbits at the innermost radii of the disc (i.e, with the smallest $a$) undergo the greatest fractional change in $J$ and hence are most likely to be pushed to extreme eccentricities. Consequently, the precession rates of these orbits are important.
Recalling that $t_{\rm prec} \propto a^{\gamma-3/2} f(e)$, we perform $N$-body simulations of an eccentric stellar disc to examine the effect of varying the index of the power-law slope of the cusp, $\gamma$, on the instability; see Figure (\ref{f:gamma}) for examples where $\gamma = 0.5$ (shallow cusp) and $\gamma = 1.75$ (steep cusp). The disc has a surface density profile $\Sigma \propto r^{-2}$, $M_{\rm disc} = 10^4 M_{\odot}$, a semi major axis distribution of $0.05 \leq a \leq 0.5$ pc and a smooth stellar cusp with $M_{\rm cusp}(<1 {\rm pc}) = 0.5 \times 10^6 M_{\odot}$ \citep[see][for details]{Mad09}. For $\gamma < 3/2$, due to the shallowness of the cusp, the innermost orbits lag behind in precession and are predisposed to being torqued to high $e$ orbits; the instability is not as effective at generating low $e$ orbits. Conversely, an increase in $\gamma$ $(> 3/2)$ results in the innermost orbits precessing faster than the bulk of the stars and hence suppresses the generation of the highest $e$ orbits.
\begin{figure}[!ht]
\begin{center}
\includegraphics[scale = 0.44, angle = 270]{f2.eps}
\caption{Evolution of the (x and y components of the) eccentricity vectors (EVs) of stars within an eccentric disc embedded in a background cusp with index $\gamma = 0.5$ (top) and $\gamma = 1.75$ (bottom). Time is in units of the initial period of the innermost orbit at $a = 0.05 {\rm pc}$; final time corresponds to $\sim \!0.8$ Myr. The innermost circle shows the initial value of the eccentricities $e = 0.6$; the outermost is $e = 1$. The colour of the EVs indicate the semi major axes of the stars in the disc, going from red ($\sim0.05 - 0.1$ pc) outwards to $0.5$ pc (green, cyan and dark blue). The EVs spread out as the stars complete more and more orbits, and precess with retrograde motion (clockwise here). Many more high eccentricity stars are produced in simulations with a flatter cusp ($\gamma = 0.5$) as the innermost stars lag behind in precession. In simulations with a steep cusp ($\gamma = 1.75$), the stars precess more rapidly ($t_{\rm prec} \sim 900$ orbits) and the innermost orbits have lower average eccentricities.
\label{f:gamma}}
\end{center}
\end{figure}
\subsection{Implications for the Galactic Centre}
A (so far unknown) percentage of stars in a disc should be formed in binary systems. These binaries can be propelled to high eccentricities by the eccentric disc instability, pass within the tidal radius of the MBH and be disrupted \citep{Hil88, YuQ03}. In the GC, this mechanism may be responsible for producing both hypervelocity stars \citep[HVS; see][and references therein]{Bro09} and the S-stars \citep{Sch02, Ghe05, Eis05, Gil09}. There are three lines of evidence to support this idea. (1) The cusp of late-type stars in the GC is observed to have a flat surface density profile \citep{Buc09, Do09, Bar09} which, if it traces the underlying mass distribution, is maximally effective in producing high eccentricity orbits with the instability. (2) \citet{Lu09} find that the observed HVS are consistent with two planar structures, one alined with the inner edge of the clockwise disc\footnote{The instability preferentially propels stars at the inner edge of the disc to extreme eccentricities. These stars are likely to have significant inclinations however which suggests that, in this scenario, the HVS will not be tightly confined to a plane.}. The HVS travel time from the GC to their current positions is too large (100 - 200 Myr) for them to have originated in the observed young disc but an older disc could potentially explain this feature. (3) \citet{Bar09} find a bimodal distribution of eccentricities in the young disc. The eccentric disc instability naturally produces a double-peaked eccentricity profile as the stars are pushed away from the average value. However, we must emphasize that a direct comparison to the observations is not possible at this time as the most eccentric stars in our simulations are the most highly inclined and would not be observed along the plane of the disc. \\
\acknowledgements
AM thanks Olivera Raki\'c for her helpful comments.
|
\section{Introduction}
The Chern-Simons-matter (CSM) theories \cite{Schwarz:2004yj,Ivanov:1991fn,Gaiotto:2007qi,Avdeev:1991za,Avdeev:1992jt,Chen:1992ee,Kapustin:1994mt,Gaiotto:2008sd,Hosomichi:2008jd,Aharony:2008ug,Benna:2008zy} provide a large class of (super)conformal field theories in three-dimensions. It was pointed out in \cite{Gaiotto:2007qi} that even with given gauge group and matter content, the ${\cal N}=2$ CSM theory admits a large number of exact infrared fixed points, at least in the perturbative regime. In this paper, we make a very simple extension of the argument of \cite{Gaiotto:2007qi} to show that
in fact an entire continuous family of exact conformal fixed points exist; they are given by ${\cal N}=2$ CSM theory with appropriate quartic superpotentials.
Our general argument will be based on holomorphy of the effective superpotential, and promoting superpotential coefficients to dynamical
chiral fields a la Seiberg \cite{Seiberg:1993vc}. An explicit 4-loop check will be performed. We find nontrivial cancelation of certain components of the 4-loop beta functions, consistent with the claim that the family of two-loop IR fixed points survive to all loop order (the precise RG fixed point locus may be deformed by higher loop effects).
While in a general ${\cal N}=2$ CSM theory, the $U(1)_R$ charge of the matter fields can be renormalized, there is no anomalous $U(1)_R$ charge along the continuous family of fixed points. With appropriate choices of matter content, one special point in this family is the ${\cal N}=3$ CSM theory.
One moves along the family by turning on quartic chiral primary deformations of the superpotential.
On this space of superconformal CSM theories, there is a natural notion of metric -- the Zamolodchikov metric \cite{Zamolodchikov:1986gt}. We will consider the example of ${\cal N}=2$ $U(N)$ CSM theory with $M$ adjoint matter fields. At the leading nontrivial order, the family of fixed points modulo the $U(M)$ flavor symmetry (the quotient space denoted by ${\cal M}$) is given by a symplectic quotient of the linear complex vector space $V$ of all quartic superpotential coefficients, and the metric is the natural one associated with the symplectic form. We will compute the next-to-leading order perturbative correction to this metric in the 't Hooft limit. It will turn out that the moduli space ${\cal M}$ is a symplectic quotient defined by a deformed symplectic form on $V$.
However, the corrected Zamolodchikov metric, while still K\"ahler, is different from the one induced from the symplectic quotient.
In the next section, we will present the non-renormalization argument. Section 3 discusses the check via the 4-loop beta function, with details of the computation in Appendix A and B. While one may perform the computation using supergraph techniques, we found it more convenient to work with ordinary Feynman diagrams in component fields, utilizing the ``graphical rules" described in Appendix B. Section 4 studies the perturbative Zamolodchikov metric on the space of fixed points. The details of the computation of the metric are given in Appendix C and D. We summarize the results and conclude in section 5.
\section{A non-renormalization theorem}
Let us start by considering the example of ${\cal N}=2$ Chern-Simons-matter theory with $U(N)$ gauge group and $M$ {\sl adjoint} flavors. $k$ will denote the Chern-Simons level. For convenience, we will be mostly working in the 't Hooft limit, i.e. $N,k\to \infty$ with $\lambda=N/k$ fixed and treated perturbatively. This is a natural limit to consider, having in mind the holographic dual. Most of our arguments here can be straightforwardly generalized to finite $N$. The chiral matter superfields are denoted $\Phi_i$, with the flavor index $i=1,\cdots,M$. We will consider a general quartic (single trace) superpotential,
\ie\label{supw}
W = \frac{1}{4}\sum_{i,j,k,l} \alpha_{ijkl} {\rm Tr} (\Phi_i \Phi_j \Phi_k \Phi_l).
\fe
As argued in \cite{Gaiotto:2007qi}, the theory with $W=0$ is a superconformal field theory, in which the matter field $\Phi_i$ acquires a quantum corrected $U(1)_R$ charge, $J_\Phi={1\over 2}+{\cal O}({1\over k^2})$. Our normalization convention for the $U(1)_R$ charge $J$ is such that the unitarity bound on the scaling dimension of an operator of charge $J$ is $\Delta\geq J$. This bound is saturated by chiral primaries.
In the $W=0$ theory, every chiral operator is also a chiral primary. Nevertheless, the chiral primaries still acquire anomalous dimensions, which are equal to their anomalous $U(1)_R$ charges. Therefore, the operator ${\rm Tr} (\Phi_i \Phi_j \Phi_k \Phi_l)$ has dimension $4J_\Phi$ at the origin of the space of $\alpha$'s.
It was argued and also shown in explicit computation in \cite{Gaiotto:2007qi} that $J_\Phi<{1\over 2}$, i.e. $W=0$ is an unstable fixed point. Further, the beta function for $\alpha_{ijkl}$ (defined by normalizing the kinetic term for $\Phi_i$'s) up to two-loop order takes the form
\ie
\mu{d\alpha_{ijkl}\over d\mu} = (4J_\Phi-2) \alpha_{ijkl} + {1\over 4\pi^2}B_{(\underline{i}}{}^r \alpha_{r\underline{jkl})} + {\rm higher~loop}
\fe
where $(\underline{ijkl})$ stands for cyclic symmetrization. $B_i{}^j={1\over 2}N^2 \alpha_{iklm} \overline\alpha^{jklm}$ comes from the two-loop wave function renormalization in the corresponding Wess-Zumino model (obtained by decoupling the Chern-Simons gauge field).
Starting with the superpotential (\ref{supw}) in the UV, we can consider the Wilsonian effective action, of the form
\ie
S_{CS}^{{\cal N}=2}(V)+\int d^3x\int d^4\theta K(\Phi_i,\overline\Phi_i,V)
+ \int d^3x d^2\theta \sum f_{ijkl}(k) \alpha_{ijkl} {\rm Tr} (\Phi_i \Phi_j \Phi_k \Phi_l) + c.c.
\fe
For now we are working in the 't Hooft limit, and hence the multi-trace operators are ignored in the effective action. Note however that our argument will go through even with multi-trace operators included. In particular, the effective superpotential only contains the quartic terms as in the classical superpotential. This follows from holomorphy of the effective superpotential in $\Phi_i$ and the $U(1)$ R-symmetry. Note that unlike in four-dimensional gauge theories \cite{Seiberg:1993vc, ArkaniHamed:1997mj}, here there is no anomaly in the global $U(1)$ symmetries, nor a dynamically generated scale, to allow for non-perturbatively generated superpotentials. Further, by promoting $\alpha_{ijkl}$ to dynamical chiral fields, one sees that the effective superpotential is also holomorphic in $\alpha_{ijkl}$.
By assigning an $R$-symmetry\footnote{This $R$-symmetry is not to be confused with the $U(1)_R$ of the superconformal algebra.} charge $2$ to $\alpha_{ijkl}$, $1$ to $\theta$ (and $-1$ to $\bar\theta$), and $0$ to $\Phi_i$'s,
we conclude that the effective superpotential must be linear in $\alpha_{ijkl}$, and that the superpotential coefficient can only be renormalized by the Chern-Simons coupling $1/k$.\footnote{One may worry about the linear mixing of $\alpha_{ijkl}$ with say $\alpha_{ilkj}$, which is consistent with the $U(M)$ flavor symmetry. However, this is not possible because ${\rm Tr}\Phi^4$ would be a chiral primary in {\sl the $W=0$ theory}, and therefore do not mix with one another at leading order in $\alpha$.}
As pointed out in \cite{Gaiotto:2007qi}, such corrections will occur in general, since one cannot promote the Chern-Simons level $k$ to a dynamical field without breaking gauge symmetry.
After normalizing the two-derivative kinetic term for $\Phi_i$ in the K\"ahler potential, we see that the quantum correction to $\alpha_{ijkl}$ amounts to an anomalous dimension for the operator $ {\rm Tr} (\Phi_i \Phi_j \Phi_k \Phi_l) $ together with a wave function renormalization. So, in fact, we expect
\ie
\mu{d\alpha_{ijkl}\over d\mu} = (4J_\Phi(k)-2) \alpha_{ijkl} + {1\over 4\pi^2}B_{(\underline{i}}{}^r \alpha_{r\underline{jkl})}
\fe
to hold exactly, for some $B_i{}^j(\alpha,\overline\alpha,k)={1\over 2}N^2 \alpha_{iklm} \overline\alpha^{jklm}+ ({\rm higher~ order~terms~in~}1/k,\alpha,\bar\alpha)$. Here $\alpha_{ijkl}$ are considered to be of the same order as $1/k$, as is the case along the $W\not=0$ two-loop fixed point loci. $J_\Phi(k)$ is the quantum corrected $U(1)_R$ charge of $\Phi_i$ in the $W=0$ theory, which is a function of $k$ only.\footnote{If we did not have the $U(M)$ flavor symmetry, of course, the $\Phi_i$'s may have different anomalous $U(1)_R$ charges depending on their representation content, in the $W=0$ theory.} We then conclude that the IR fixed points, up to the global flavor symmetry $U(M)$, is parameterized by the quotient space
\ie
{\cal M}=\left\{\alpha_{ijkl}:\,B_i{}^j = c\,\delta_i^j\right\}/U(M)
\fe
where $c=4\pi^2(2-4J_\Phi)>0$. Denote by $V$ the linear vector space of all $\alpha_{ijkl}$'s. ${\cal M}$ is generally a deformation of the standard symplectic quotient $V//U(M)$
by the 't Hooft coupling $\lambda=N/k$. We will revisit the geometry of ${\cal M}$ in section 4. When $M=2n$ is even, one point on ${\cal M}$ is given by the ${\cal N}=3$ CSM theory with $n$ adjoint hypermultiplets.
So in the perturbative regime, the IR fixed points of the ${\cal N}=2$ $U(N)$ CSM with $M$ adjoint matter fields are given by the $W=0$ fixed point together with the fixed point manifold ${\cal M}$ (up to the $U(M)$ flavor symmetry). The tangent directions of ${\cal M}$ are in 1-1 correspondence with quartic chiral primary operators.\footnote{More precisely, the tangent vectors along the fixed point loci in $V$ (before quotienting by $U(M)$) are a linear combination of the quartic chiral primaries and the scalar operators in the supermultiplet of the $U(M)$ flavor currents.} It then follows that the $U(1)_R$-charge of $\Phi_i$ along ${\cal M}$ is given exactly by $J={1\over 2}$, in contrast to $J_\Phi<{1\over 2}$ at $W=0$. When the number of flavors $M$ is even, the non-renormalization of $U(1)_R$ charge is well known at the ${\cal N}=3$ point on ${\cal M}$. Now we conclude that this property continues to hold even when one deforms marginally away from the ${\cal N}=3$ point. At a given point on ${\cal M}$ with superpotential $W$, the chiral operators of the form ${\rm Tr}(\Phi_i \partial_j W)$ are descendants, and are transverse to the IR fixed point manifold in $V$. On the other hand, the explicit expressions of the quartic chiral primaries are dependent on $k$, and can be determined by looking at the tangent directions of ${\cal M}$. We will return to this in section 4.
Let us comment that the holomorphy argument above applies only to the Wilsonian effective action and not to the 1PI effective action \cite{ArkaniHamed:1997mj, Weinberg:1998uv}. This is because in the 1PI effective action, where massless modes are integrated out, nonlocal terms may be generated in the K\"ahler potential such that when one replaces the spurious chiral fields by their expectation values $\alpha_{ijkl}$, the term
looks like a superpotential term with non-holomorphic dependence on $\alpha_{ijkl}$ (see \cite{Poppitz:1996na}). In computing higher loop contributions to the beta function, the result from 1PI RG and Wilsonian RG may differ, depending on renormalization schemes. Nevertheless, the dimensionality of the loci of IR fixed points in the space of couplings $\alpha_{ijkl}$ clearly should not depend on the choice of renormalization group. The Chern-Simons gauge field may appear subtle from the perspective of Wilsonian RG. On one hand, the gauge fields are effectively infinitely massive and have no propagating degrees of freedom; on the other hand, they give rise to long range interactions, which may be thought of as a non-abelian generalization of the anyon statistics of the matter fields.
In principle, they must be treated carefully, using the regularization of \cite{Hayashi:1998ca} which cuts off the momenta in a way that
preserves supersymmetry and gauge symmetry manifestly. This is achieved, for instance, by replacing the Chern-Simons gauge super-propagator in ${\cal N}=2$ superspace by
\ie
\int_0^\infty d\tau f(\Lambda\tau) e^{-\tau D\bar D}\delta^7(z)
\fe
for some function $f(\tau)$ that vanishes (as well as its derivatives to all orders) at $\tau=0$, and approaches 1 at $\tau=\infty$. Here $\delta^7(z)\equiv \delta^3(x)\delta^4(\theta)$, and we did not take into account gauge fixing. If we are to integrate out a momenta shell from $\Lambda$ to $\Lambda+\delta\Lambda$, we may replace the regularized propagator by
\ie
\delta \Lambda \int_0^\infty d\tau \,\tau f'(\Lambda\tau) e^{-\tau D\bar D}\delta^7(z)
=-{\delta \Lambda\over\Lambda} \int_0^\infty d\tau f(\Lambda\tau) \partial_\tau\left[\tau e^{-\tau D\bar D}\right] \delta^7(z)
\fe
With the choice $f(\tau)=\theta(\tau-1)$, this is simply ${\delta\Lambda\over\Lambda^2} e^{-{D\bar D\over\Lambda}}\delta^7(z)$.
It is straightforward to generalize the non-renormalization argument to ${\cal N}=2$ CSM theories with any gauge groups and any matter representation content. The renormalization of the quartic superpotential coefficients, or Yukawa couplings, can be entirely absorbed into wave function renormalization. In particular, the a priori nontrivial $k$-dependent renormalization of the superpotential coefficients are entirely due to the anomalous dimensions of the chiral matter fields in the $W=0$ theory. The ``generic branch" of conformal fixed points are described by
\ie\label{gb}
B_i{}^j(\alpha,\overline\alpha,k) = 16\pi^2\left[ {1\over 2}-J_i(k) \right] \delta_i^j
\fe
where $B_i{}^j(\alpha,\bar\alpha,k)=(B_{WZ})_i{}^j(\alpha,\bar\alpha)+{\cal O}(1/k^2)$. Here $B_{WZ}$ represents the wave function renormalization in the corresponding Wess-Zumino model, and $J_i(k)$ is the quantum corrected $U(1)_R$ charge of the field $\Phi_i$ in the $W=0$ theory. For general matter content, there may also be non-generic branches of IR fixed points, where some of the $\alpha$'s set to zero, and (\ref{gb}) only needs to be satisfied for a subset of $\Phi_i$'s.
\section{A 4-loop check}
In the previous section we have given a holomorphy argument that the manifold of two-loop IR fixed points survives to all-loop order. That is, while the loci of the family of two-loop fixed points in $V$ may be deformed by higher loop effects, the dimension of the family remains unchanged. This is not at all obvious from the perspective of 1PI RG. In the 1PI effective action, a priori, there are terms that could potentially contribute to the beta function of $\alpha_{ijkl}$ in the form
\ie\label{poss}
\beta_{ijkl} = \beta_{ijkl}^{2-loop} + {C\over k^2} \alpha_{(\underline{ij}mn}
\bar\alpha^{mnpq}\alpha_{pq\underline{kl})} + \cdots
\fe
where we exhibited one possible 4-loop contribution. $C$ is a constant coefficient that generally depends on $M$ and $N$. Such a 4-loop contribution cannot be absorbed into the wave function renormalization of the matter fields. If $C$ is nonzero, the family of two-loop fixed points will further flow to a submanifold of lower dimension (possibly discrete points).
While the higher loop beta function in 1PI RG may not agree with that of the Wilsonian RG in general, the dimensionality of the loci of IR fixed points should not depend on which RG we use. Therefore we expect the higher-than-two-loop contributions such as the second term on the RHS of (\ref{poss}) to vanish. We will now check this explicitly
at 4-loop order.
We will make a few simplifying assumptions. Firstly, we shall work at the planar level. We expect the same conclusion to hold with non-planar diagrams included as well, but the computation would be more complicated. (Having in mind the holographic dual, the planar limit is interesting on its own.) Secondly, we will take the number of flavors, $M$, to be parametrically large, and start by looking at the leading nontrivial contribution in the large $M$ limit. This reduces the number of diagrams drastically. The subleading $1/M$ contributions involve many more diagrams, whose explicit computations are not consider in the current paper.
At the planar level, the potential contributions to the effective superpotential that cannot be absorbed into wave function renormalization take the general form
\begin{equation}
(c_1 M + c_2) {4\pi^2\over k^2}N^4\alpha_{ijmn}\bar\alpha^{mnpq}\alpha_{pqkl}
{\rm Tr}\left(\Phi^i\Phi^j\Phi^k\Phi^l\right)
\label{effective_superpotential}
\end{equation}
where $c_1$ and $c_2$ are constants. To see this, let us examine the diagrams. While we will perform the computation using ordinary Feynman diagrams in component fields, it is convenient to organize them using ${\cal N}=2$ supergraphs. Our notation is explained in Appendix A. In a supergraph that contributes to
(\ref{effective_superpotential}), the F-term vertices are contracted according to the following structure:
\bigskip
\centerline{\begin{fmffile}{basic}
\begin{tabular}{c}
\begin{fmfgraph*}(45,25)
\fmfstraight
\fmfset{arrow_len}{.3cm}\fmfset{arrow_ang}{12}
\fmfleft{i1,i2}
\fmfright{o1,o2}
\fmf{fermion}{i2,v1}
\fmf{fermion}{i1,v1}
\fmf{fermion}{o1,v2}
\fmf{fermion}{o2,v2}
\fmf{fermion,right=.7,tension=.3}{v3,v1}
\fmf{fermion,right=.7,tension=.3}{v3,v2}
\fmf{fermion,left=.7,tension=.3}{v3,v1}
\fmf{fermion,left=.7,tension=.3}{v3,v2}
\end{fmfgraph*}
\end{tabular}
\end{fmffile}}
\noindent The non-abelian ${\cal N}=2$ CS action minimally coupled to chiral fields can be written in
${\cal N}=2$ superspace as
\ie
S_D = \int d^3 x \int d^4\theta \left\{ {k\over 2\pi}\int_0^1 dt {\rm Tr} \left[
V \bar D^\alpha \left( e^{-tV}D_\alpha e^{tV} \right) \right] + \sum_i \bar\Phi_i e^V \Phi_i \right\}
\fe
In Wess-Zumino gauge, the D-term supervertices involve the cubic interactions of the super gauge fields and the standard minimal coupling to matter fields, as described in Appendix A. These D-supervertices can be attached to the above graph to form a 4-loop diagram that contribute to the beta function. Some examples are
\begin{equation}\nonumber
\begin{array}{ccc}
\begin{fmffile}{ex1}
\begin{tabular}{c}
\begin{fmfgraph*}(45,25)
\fmfstraight
\fmfset{arrow_len}{.3cm}\fmfset{arrow_ang}{12}
\fmfleft{i1,i2}
\fmfright{o1,o2}
\fmf{fermion}{i2,v4}
\fmf{fermion}{v4,v1}
\fmf{fermion}{i1,v1}
\fmf{fermion}{o1,v2}
\fmf{fermion}{o2,v6}
\fmf{fermion}{v6,v2}
\fmffixed{(.6h,0)}{v1,v3}
\fmffixed{(.6h,0)}{v3,v2}
\fmffixed{(.3h,.25h)}{v3,v5}
\fmf{fermion,right=.7,tension=.3}{v3,v1}
\fmf{fermion,right=.7,tension=.3}{v3,v2}
\fmf{fermion,left=.7,tension=.3}{v3,v1}
\fmf{fermion,left=.3,tension=.3}{v3,v5,v2}
\fmf{photon,left=.35,tension=0}{v4,v5}
\fmf{photon,left=.7,tension=0}{v5,v6}
\end{fmfgraph*}
\end{tabular}
\end{fmffile}
&~~~~
\begin{fmffile}{ex2}
\begin{tabular}{c}
\begin{fmfgraph*}(45,25)
\fmfstraight
\fmfset{arrow_len}{.3cm}\fmfset{arrow_ang}{12}
\fmfleft{i1,i2}
\fmfright{o1,o2}
\fmf{fermion}{i2,v4}
\fmf{fermion}{v4,v1}
\fmf{fermion}{i1,v1}
\fmf{fermion}{o1,v2}
\fmf{fermion}{o2,v6}
\fmf{fermion}{v6,v2}
\fmffixed{(.6h,0)}{v1,v3}
\fmffixed{(.6h,0)}{v3,v2}
\fmffixed{(.3h,.25h)}{v3,v5}
\fmffixed{(-.3h,.5h)}{v3,v7}
\fmf{fermion,right=.7,tension=.3}{v3,v1}
\fmf{fermion,right=.7,tension=.3}{v3,v2}
\fmf{fermion,left=.7,tension=.3}{v3,v1}
\fmf{fermion,left=.3,tension=.3}{v3,v5,v2}
\fmf{photon,left=.17,tension=0}{v4,v7,v5}
\fmf{photon,left=.35,tension=0}{v7,v6}
\end{fmfgraph*}
\end{tabular}
\end{fmffile}
&~~~~
\begin{fmffile}{ex3}
\begin{tabular}{c}
\begin{fmfgraph*}(45,25)
\fmfstraight
\fmfset{arrow_len}{.3cm}\fmfset{arrow_ang}{12}
\fmfleft{i1,i2}
\fmfright{o1,o2}
\fmf{fermion}{i2,v4}
\fmf{fermion}{v4,v1}
\fmf{fermion}{i1,v1}
\fmf{fermion}{o1,v2}
\fmf{fermion}{o2,v2}
\fmffixed{(.6h,0)}{v1,v3}
\fmffixed{(.6h,0)}{v3,v2}
\fmffixed{(.3h,.25h)}{v3,v7}
\fmffixed{(-.4h,.5h)}{v3,v5}
\fmffixed{(-.1h,.5h)}{v3,v6}
\fmf{fermion,right=.7,tension=.3}{v3,v1}
\fmf{fermion,right=.7,tension=.3}{v3,v2}
\fmf{fermion,left=.7,tension=.3}{v3,v1}
\fmf{fermion,left=.3,tension=.3}{v3,v7,v2}
\fmf{photon,left=.1,tension=0}{v4,v5}
\fmf{fermion,left=.7,tension=0}{v5,v6,v5}
\fmf{photon,left=.1,tension=0}{v6,v7}
\end{fmfgraph*}
\end{tabular}
\end{fmffile}
\end{array}
\end{equation}
In the limit of large $M$, the third diagram dominates the first two, due to the factor $M$ coming from the matter-loop-corrected vector superfield propagator. For now, we will consider this limit and ignore diagrams such as the first two above. In other words, we will be computing $c_1$ but not $c_2$ in (\ref{effective_superpotential}).
There are only three types of planar 4-loop supergraphs that contribute to $c_1$,
given by
\begin{equation}
\begin{array}{ccc}
\begin{fmffile}{A}
\begin{tabular}{c}
\begin{fmfgraph*}(45,25)
\fmfstraight
\fmfset{arrow_len}{.3cm}\fmfset{arrow_ang}{12}
\fmfleft{i1,i2}
\fmfright{o1,o2}
\fmf{fermion}{i2,v1}
\fmf{fermion}{i1,v1}
\fmf{fermion}{o2,v2}
\fmf{fermion}{o1,v2}
\fmffixed{(.6h,0)}{v1,v3}
\fmffixed{(.6h,0)}{v3,v2}
\fmffixed{(-.3h,.25h)}{v3,v4}
\fmffixed{(.3h,.25h)}{v3,v7}
\fmffixed{(-.15h,.5h)}{v3,v5}
\fmffixed{(.15h,.5h)}{v3,v6}
\fmf{fermion,right=.3,tension=.3}{v3,v4,v1}
\fmf{fermion,right=.7,tension=.3}{v3,v2}
\fmf{fermion,left=.7,tension=.3}{v3,v1}
\fmf{fermion,left=.3,tension=.3}{v3,v7,v2}
\fmf{photon,left=.3,tension=0}{v4,v5}
\fmf{fermion,left=.7,tension=0}{v5,v6,v5}
\fmf{photon,left=.3,tension=0}{v6,v7}
\end{fmfgraph*}
\end{tabular}
\end{fmffile}
&~~~~
\begin{fmffile}{B}
\begin{tabular}{c}
\begin{fmfgraph*}(45,25)
\fmfstraight
\fmfset{arrow_len}{.3cm}\fmfset{arrow_ang}{12}
\fmfleft{i1,i2}
\fmfright{o1,o2}
\fmf{fermion}{i2,v4}
\fmf{fermion}{v4,v1}
\fmf{fermion}{i1,v1}
\fmf{fermion}{o1,v2}
\fmf{fermion}{o2,v2}
\fmffixed{(.6h,0)}{v1,v3}
\fmffixed{(.6h,0)}{v3,v2}
\fmffixed{(.3h,.25h)}{v3,v7}
\fmffixed{(-.1h,.5h)}{v3,v6}
\fmffixed{(-.4h,.5h)}{v3,v5}
\fmf{fermion,right=.7,tension=.3}{v3,v1}
\fmf{fermion,right=.7,tension=.3}{v3,v2}
\fmf{fermion,left=.7,tension=.3}{v3,v1}
\fmf{fermion,left=.3,tension=.3}{v3,v7,v2}
\fmf{photon,left=.1,tension=0}{v4,v5}
\fmf{fermion,left=.7,tension=0}{v5,v6,v5}
\fmf{photon,left=.1,tension=0}{v6,v7}
\end{fmfgraph*}
\end{tabular}
\end{fmffile}
&~~~~
\begin{fmffile}{C}
\begin{tabular}{c}
\begin{fmfgraph*}(45,25)
\fmfstraight
\fmfset{arrow_len}{.3cm}\fmfset{arrow_ang}{12}
\fmfleft{i1,i2}
\fmfright{o1,o2}
\fmf{fermion}{i2,v4}
\fmf{fermion}{v4,v1}
\fmf{fermion}{i1,v1}
\fmf{fermion}{v7,v2}
\fmf{fermion}{o2,v7}
\fmf{fermion}{o1,v2}
\fmffixed{(.6h,0)}{v1,v3}
\fmffixed{(.6h,0)}{v3,v2}
\fmffixed{(-.15h,.5h)}{v3,v5}
\fmffixed{(.15h,.5h)}{v3,v6}
\fmf{fermion,right=.7,tension=.3}{v3,v1}
\fmf{fermion,right=.7,tension=.3}{v3,v2}
\fmf{fermion,left=.7,tension=.3}{v3,v1}
\fmf{fermion,left=.7,tension=.3}{v3,v2}
\fmf{photon,left=.1,tension=0}{v4,v5}
\fmf{photon,left=.1,tension=0}{v6,v7}
\fmf{fermion,left=.7,tension=0}{v5,v6,v5}
\end{fmfgraph*}
\end{tabular}
\end{fmffile}
\label{ABC}
\\
(a) & (b) & (c)
\end{array}
\end{equation}
Let us note that planarity forbids contributions to terms proportional to
$\alpha_{ijmn}\bar\alpha^{mpnq}\alpha_{pqkl}$ or $\alpha_{ijmn}\bar\alpha^{mqnp}\alpha_{pqkl}$ in the beta function. Similarly, in the beta function for $\alpha_{ijkl}$, planarity only allows
the term $\alpha_{ijmn}\bar\alpha^{mnpq}\alpha_{pqkl}$ with the indices $\{i,j,k,l\}$ appearing in cyclic order.
In component fields, the super gauge field propagator involves the vector gauge field as well as the auxiliary fields $D,\sigma,\chi$. The latter can be integrated out to give quartic scalar-fermion vertices and sextic scalar vertices (not used here). These diagrams are computed explicitly in Appendix A. The coefficient $c_1$ is
\ie
c_1 = {1\over 2}(a+b+c)
\fe
where $a$, $b$, $c$ are constants computed from the three supergraphs $(a), (b), (c)$ listed above. We find
\ie
a=c={1\over 256\pi^2},~~~b=-{1\over 128\pi^2},
\fe
and they indeed sum up to zero. We note however that the individual supergraph contribution does not vanish, and so $c_1=0$ here is a consequence of cancellation among different supergraphs in the 1PI RG.
\section{The metric on ${\cal M}$}
The manifold ${\cal M}$, defined as the $W\not=0$ IR fixed points in $V$ modulo $U(M)$ flavor symmetry, is naturally equipped with a Zamolodchikov metric. The metric is defined by the coefficient of the two-point function of quartic chiral primaries and their conjugates that parameterize the tangent directions of ${\cal M}$. In particular, the geometry of ${\cal M}$ will generally depend on the 't Hooft coupling $\lambda$. We expect the generic CSM theory to have a holographic dual, which may or may not have a gravity limit at strong 't Hooft coupling.\footnote{The theories with a large number of adjoint flavors, in particular, are expected to only have a stringy holographic dual. This is because the number of chiral primaries grow exponentially with the dimension, which cannot happen in a supergravity theory compactified to $AdS_4$.} Heuristically, had there been a gravity dual say of the form M-theory on $AdS_4\times M_7$, $M_7$ being the base of a Calabi-Yau 4-fold cone, then the analog of the manifold ${\cal M}$ at
strong 't Hooft coupling would be the moduli space of this CY 4-fold cone.\footnote{See \cite{Martelli:2008a, Martelli:2008b, Martelli:2009, Hanany:2008, Franco:2008} for recent work on such theories.} The geometry of ${\cal M}$ at
strong 't Hooft coupling is difficult to understand from the field theory perspective. In this section, we will investigate the perturbative corrections to the Zamolodchikov metric on ${\cal M}$.
We may write the $W\not=0$ IR fixed point locus as
\ie\label{mpc}
\mu_i{}^j(\alpha,\overline\alpha,k) = r(k)\delta_i^j
\fe
Here $\mu_i{}^j$ is proportional to $B_i{}^j$, with a possibly $k$-dependent normalization factor for later convenience. So, $r(k)$ is not necessarily the same as ${1\over 2}-J(k)$. Up to two-loop contribution, and in the limit of large $M$, we have
$r(k)=M({4\pi N\over k})^2 +\cdots$. Its precise $k$ dependence is not important for our purpose.
The tangent directions $\delta\alpha_{ijkl}=c_{ijkl}$ are determined by
\ie
c_{mnpq} {\partial\over\partial\alpha_{mnpq}} \mu_i{}^j(\alpha,\overline\alpha,k)=0,~~~\forall i,j.
\fe
The quartic chiral primaries are then given by ${\cal O}_c=\sum c_{mnpq} {\rm Tr}(\phi_m \phi_n \phi_p \phi_q)$ for such $c$. The two-point functions of a quartic chiral primary and an anti-chiral primary, in the SCFT corresponding to a point on ${\cal M}$, take the form
\ie
\langle {\cal O}_c(x) \overline{\cal O}_{c'}(0)\rangle = {g(c,\bar c')\over |x|^4}.
\fe
where the coefficient $g(c,\bar c')$ is the Zamolodchikov metric.
To begin, let us examine the IR manifold ${\cal M}$ at two-loop order. The leading contribution to the Zamolodchikov metric is simply given by the free correlator,
\ie
g^{(0)}(c,\bar c') = \sum c_{mnpq}\bar c'^{mnpq}.
\fe
It is the standard Euclidean metric on the space of $\alpha_{ijkl}$'s, corresponding to the symplectic form
\ie
\omega^{(0)} = d\alpha_{mnpq}\wedge d\bar\alpha^{mnpq}.
\fe
The metric on ${\cal M}$ at the leading order is the one induced from the symplectic quotient by
the flavor symmetry $U(M)$.
At the next-to-leading order, we must consider the 4-loop-corrected IR manifold ${\cal M}$, as well as the two-loop contributions to the two-point functions of the quartic chiral primaries. With the 4-loop contributions taken into account, $\mu_i{}^j$ takes the form
\ie\label{mucorr}
\mu_i{}^j = N^2 \alpha_{imnp}\bar\alpha^{jmnp} + a_1 N^4 \left(\alpha_{imnk}\bar\alpha^{mnpq}\alpha_{pqrs}\bar\alpha^{rskj}+\alpha_{kmni}\bar\alpha^{mnpq}\alpha_{pqrs}\bar\alpha^{rsjk}\right) + {\rm higher~ order}.
\fe
Here the constant $a_1$ is simply given by the 4-loop wave function renormalization in the Wess-Zumino model. All other 4-loop corrections to $B_i{}^j$ will be proportional to the first term in (\ref{mucorr}) and are subleading in $1/k^2$; they can be absorbed by a rescaling of $\mu_i{}^j$ and $r(k)$ in (\ref{mpc}).
$\mu_i{}^j$ is in fact a moment map associated with the symplectic form
\ie
\omega = N^2 d\alpha_{mnpq}\wedge d\bar\alpha^{mnpq}
+4a_1 N^4 \alpha_{rsij}\bar\alpha^{pqij} d\alpha_{mnpq}\wedge d\bar\alpha^{mnrs} + {\rm higher~ order~ terms}.
\label{a1}
\fe
By definition, given the $U(M)$ action as a vector field on $V$
\ie
v_i{}^j = \alpha_{imnp} {\partial\over\partial\alpha_{jmnp}}
- \bar\alpha^{jmnp} {\partial\over\partial \bar\alpha^{imnp}},
\fe
we have $d\mu_i{}^j = \iota_{v_i{}^j}\omega$, where $\iota$ stands for the contraction with a vector field.
Generally, the two-loop correction to the two-point function of the quartic chiral primaries takes the form
\ie
g(c,\bar c') = f(k) c_{mnpq}\bar c'^{mnpq} + a_2 N^2 c_{ijmn} \bar\alpha^{mnpq}\alpha_{pqrs}\bar c'^{rsij}
+ a_3 N^2 c_{imnp} \bar\alpha^{mnpq}\alpha_{qrst}\bar c'^{rsti},\label{a2}
\fe
where $f(k) = 1+ {\cal O}(1/k^2)$, $a_2,a_3$ are constants. $c_{ijkl}$ is constrained to be tangent to the IR manifold, in particular,
$c_{imnp}\bar\alpha^{mnpq} = {\cal O}(c/k^3)$ from the two-loop constraints. So we can ignore $a_3$ in the expression for $g(c,\bar c')$. The coefficient $a_1$ is computed in Appendix C, and $a_2$ is computed in Appendix D. We find that
\ie
a_1 = -{1\over 128},~~~~a_2=-{1\over 16}.
\fe
Therefore, the Zamolodchikov metric with the next-to-leading correction included takes the form
\ie\label{gcck}
g(c,\bar c') = f(k) c_{mnpq}\bar c'^{mnpq} -{N^2\over 16} c_{ijmn} \bar\alpha^{mnpq}\alpha_{pqrs}\bar c'^{rsij} + \cdots
\fe
whereas $\alpha_{ijkl}$'s are constrained by
\ie\label{momt}
\mu_i{}^j &= N^2 \alpha_{imnp}\bar\alpha^{jmnp} -{N^4\over 128} \left(\alpha_{imnk}\bar\alpha^{mnpq}\alpha_{pqrs}\bar\alpha^{rskj}+\alpha_{kmni}\bar\alpha^{mnpq}\alpha_{pqrs}\bar\alpha^{rsjk}\right) +
\cdots
\\
&= r(k) \delta_i{}^j
\fe
We note that due to a factor of 2 difference in the second term, $g(c,\bar c')$ is not the same as the natural symplectic metric on the quotient space ${\cal M}=V//U(M)$ defined using the symplectic form $\omega$. Nevertheless, $g(c,\bar c')$ is the restriction of a K\"ahler metric on the ambient space $V$ to the level set $\mu^{-1}(r(k)\delta_i^j)$; it follows easily that the Zamolodchikov metric on ${\cal M}$ is also K\"ahler, at least to the order we have computed, even though it is not the same as the K\"ahler metric induced from the symplectic quotient. We will now sketch an argument that the metric on ${\cal M}$ is K\"ahler to all order.
The variation of the metric along a tangent direction corresponding to a chiral primary ${\cal O}_{c''}$ is
\ie\label{delg}
\delta_{c''} g(c,\bar c') = |x|^4 \left\langle {\cal O}_{c}(x) \left[ \int d^3y Q^2\cdot {\cal O}_{c''}(y)\right]\overline {\cal O}_{\bar c'}(0) \right\rangle
\fe
The statement that $g$ is K\"ahler amounts to $\delta_{c_1} g(c_2,\bar c_3)=\delta_{c_2} g(c_1,\bar c_3)$, for chiral primaries $c_1,c_2,c_3$. Let us study the correlation function
\ie
F(c_1,c_2,\bar c_3;x,y,z) = \langle {\cal O}_{c_1}(x) \left[ Q^2\cdot {\cal O}_{c_2}(y) \right] \overline {\cal O}_{c_3}(z) \rangle
\fe
$Q^2\cdot {\cal O}_{c_2}$ is a primary with respect to the bosonic conformal algebra. Apart from potential contact terms, the spatial dependence of the three-point function of the primaries as above are fixed by conformal symmetry, to be $|x-y|^{-3}|y-z|^{-3} |x-z|^{-1}$. On the other hand, by Ward identity, we can move $Q^2$ from acting on ${\cal O}_{c_2}(y)$ to acting on ${\cal O}_{c_1}(x)$, and conclude that $F(c_1,c_2,\bar c_3;x,y,z) = F(c_2,c_1,\bar c_3;y,x,z)$. This is inconsistent with the naive spatial dependence determined by conformal symmetry, which implies that $F$ must vanish up to contact terms.
Indeed, there are such contact terms. We can explicitly compute $F(c_1,c_2,\bar c_3;x,y,z)$ in perturbation theory. To leading nontrivial order, this is computed by the same diagrams as in Appendix D, interpreted as a three point function rather than a two-point function. It is given by
\ie
&F(c_1,c_2,\bar c_3;x,y,z)
\\&\sim (c_1)_{ijkl}(c_2)_{mnpq}\bar c_3^{ijmn}\bar\alpha^{klpq}
{1\over |x-z|^2|y-z|^2}\int d^3w {1\over |x-w|^2 |y-w|^4}
+ {\rm higher~order}.
\fe
The integration over $w$ naively gives zero by analytic continuation in the exponents of the propagators. If we first integrate the integrand multiplied by a generic function of $x$ over $x$, and then integrate over $w$, we see that
\ie
\int d^3w {1\over |x-w|^2 |y-w|^4} = -{\pi\over 4}\delta^3(x-y).
\fe
This gives the contact term in $F(c_1,c_2,\bar c_3;x,y,z)$. With higher order contributions included, on dimensional grounds we expect $F$ to take the form
\ie
F(c_1,c_2,\bar c_3;x,y,z) = {f(c_1,c_2,\bar c_3)\over |x-z|^4}\delta^3(x-y).
\fe
where $f(c_1,c_2,\bar c_3)$ is symmetric in $c_1$ and $c_2$ by the Ward identity argument above. The closure of the K\"ahler form associated with $g(c,\bar c')$ then follows.
Coming back to (\ref{momt}), we may also perform a nonlinear redefinition of the coupling $\alpha_{ijkl}$,
\ie
\tilde\alpha_{ijkl} &= \alpha_{ijkl} - {N^2\over 256}\left(\alpha_{limn}\bar\alpha^{mnpq}\alpha_{pqjk}
+ \alpha_{ijmn}\bar\alpha^{mnpq}\alpha_{pqkl} \right)
\\
&= \alpha_{ijkl} - {N^2\over 128}\alpha_{(\underline{ij}mn}\bar\alpha^{mnpq}\alpha_{pq\underline{kl})}
\fe
where $(\underline{ij}\cdots\underline{kl})$ stands for cyclic symmetrization on the indices $\{i,j,k,l\}$, such that the moment map reduces to the standard one
\ie
\mu_i{}^j(\tilde\alpha,\bar{\tilde\alpha}) = N^2\tilde\alpha_{imnp}\bar{\tilde \alpha}^{jmnp} + (\rm{6-loop ~and~higher~order})
\fe
The tangent space basis vectors are now modified to
\ie
\tilde c_{ijkl} = c_{ijkl} - {N^2\over 128}\left[2\alpha_{(\underline{ij}mn}\bar\alpha^{mnpq}c_{pq\underline{kl})} + \alpha_{(\underline{ij}mn}\bar c^{mnpq}\alpha_{pq\underline{kl})} \right]
\fe
They satisfy $\tilde c_{imnp} \bar{\tilde \alpha}^{jmnp} = 0$ up to 6-loop contributions. The metric in the new coordinate system is written
\ie
&g(\tilde c,\bar{\tilde c};\tilde c',\bar{\tilde c}')= f(k) \tilde c_{ijkl}\bar{\tilde c}'^{ijkl}\\
&~-{N^2\over 32}\left( \tilde c_{ijmn} \bar\alpha^{mnpq}\alpha_{pqrs}{\bar{\tilde c}}'^{rsij}
-{1\over 4} \alpha_{ijmn}\alpha_{pqkl} \bar{\tilde c}^{ijkl} \bar{\tilde c}'^{mnpq}
-{1\over 4} \bar\alpha^{ijmn}\bar\alpha^{pqkl} {\tilde c}_{ijkl} {\tilde c'}_{mnpq} \right)+\cdots.
\fe
Due to the non-holomorphic change of coordinates, the metric is not Hermitian in this new coordinate system on ${\cal M}$.
So far we have focused on the chiral primaries ${\cal O}_c = c_{ijkl}{\rm Tr} \phi_i \phi_j \phi_k \phi_l$
which give rise to deformations along the IR fixed point loci $\mu^{-1}(r(k) \delta_i^j)\subset V$. In addition, there are the $U(M)$ rotation on the $\alpha_{ijkl}$'s, corresponding to the tangent vectors $v_i{}^j$. As exactly marginal deformations of the action, they can be written in terms of the operators
\ie
{\cal V}_i{}^j &= i \int d^2\theta {\rm Tr}\left(\Phi^j {\partial W\over\partial\Phi^i}\right) + c.c. \\
&= i \int d^2\theta \sum_{m,n,p}\alpha_{imnp}{\rm Tr} (\Phi^j \Phi^m \Phi^n \Phi^p) + c.c.
\fe
The ${\cal V}_i{}^j$'s lie in the same supermultiplet as the $U(M)$ flavor current.
Let us examine the quartic non-primary chiral operators ${\cal O}_i{}^j ={\rm Tr}\left[\phi^j {\partial \over\partial\phi^i}W(\phi)\right]$ more closely. While classically it is a descendant, the precise form of the descendant operator receives quantum corrections. This can be seen as follows. In general, ${\cal O}_i{}^j$ may be the linear combination of a purely descendant operator $\tilde{\cal O}_i{}^j$ with a chiral primary. $\tilde{\cal O}_i{}^j$ is the one orthogonal to all quartic chiral primaries. Namely, the corresponding tangent vector of the IR manifold $\tilde v_i^j$ satisfies
\ie\label{gvc}
g(\tilde v_i{}^j, \bar c) = 0
\fe
for all chiral primaries ${\cal O}_c$. On the other hand, since ${\cal O}_c$ is tangent to the level set of $\mu_i{}^j$, we have
\ie
\omega(v_i{}^j,\bar c) = 0.
\fe
We have seen that the next-to-leading order perturbative correction to $g$ does not agree with that of $\omega$, which implies that $\tilde v_i{}^j$ is different from $v_i{}^j$. In fact, demanding (\ref{gvc}) gives
\ie\label{otild}
\tilde {\cal O}_i{}^j &= {\cal O}_i{}^j +{N^2\over 32} \delta_{(\underline{k}}{}^j \alpha_{i\underline{lmn})} \bar \alpha^{mnpq}\alpha_{pqrs} {\rm Tr} \left(\phi^k\phi^l\phi^r\phi^s \right) + {\rm higher~order}
\\
&= {\cal O}_i{}^j +{N^2\over 128} \left[\alpha_{i kmn} \bar \alpha^{mnpq}\alpha_{pqrs} {\rm Tr} \left(\phi^j\phi^k\phi^r\phi^s \right) + \alpha_{kimn} \bar \alpha^{mnpq}\alpha_{pqrs} {\rm Tr} \left(\phi^k\phi^j\phi^r\phi^s \right) \right.\\
&~~~\left.+\alpha_{i mkl} \bar \alpha^{jmpq}\alpha_{pqrs} {\rm Tr} \left(\phi^k\phi^l\phi^r\phi^s \right) + \alpha_{mikl} \bar \alpha^{mjpq}\alpha_{pqrs} {\rm Tr} \left(\phi^k\phi^l\phi^r\phi^s \right)\right] + {\rm higher~order}.
\fe
On the other hand, ${\cal V}_i{}^j = i Q^2 {\cal O}_i{}^j - i\overline Q^2 \overline{\cal O}_i{}^j$ is a descendant of the flavor current and is therefore orthogonal to all quartic chiral primaries. This is not in conflict with (\ref{otild}), since the descendant of the quartic chiral primary in $iQ^2{\cal O}_i{}^j$ may be canceled by a descendant of $\overline{\cal O}_i{}^j$. A more detailed investigation of the operator spectrum of this family of SCFTs is left to future work.
\section{Discussion}
We have argued that there are large classes of continuous families of three-dimensional ${\cal N}=2$ superconformal field theories, described by Chern-Simons-matter theories with appropriate quartic superpotentials. Such SCFTs in the large $N$ limit are expected to have holographic duals
as string theories in $AdS_4$. The best understood example is the theory of ABJM \cite{Aharony:2008ug}, which is dual to type IIA string theory on $AdS_4\times \mathbb {CP}^3$. However, all current examples of CSM theories with known or conjectured gravity duals have a parametrically small number of matter flavors (given the gauge group). As was pointed out in \cite{Gaiotto:2007qi}, this is because for a large number of flavors $M$, the number of chiral primaries will grow exponentially in their dimensions, which is faster than the growth of KK modes and is characteristic of string modes. For instance, suppose the superpotential $W$ is a homogeneous polynomial of degree $d$ in $M$ adjoint flavors $\Phi_i$, $i=1,\cdots, M$. The chiral operators that are descendants of primaries have the form ${\rm Tr} (f_i(\Phi) \partial_i W )$, where $\partial_i$ stands for the derivatives with respect to $\Phi_i$. For single trace operators of length $L$, the number of such descendants grow with $M$ like $M^{L-d+2}$, whereas the total number of chiral operators of length $L$ grow like
$M^L/L$ at large $M$. From the perspective of the $AdS_4$ dual, this is a Hagedorn-like growth of states with Hagedorn temperature $T\sim (R \log M)^{-1}$.
A nice feature of our family of CSM superconformal field theories is that, various features of SCFTs can be studied in perturbation theory.
In particular, we considered the perturbative corrections to the geometry of the moduli space\footnote{Here the moduli space refers to that of CFTs, not to be confused with the moduli space of vacua in a given CFT.} ${\cal M}$ of ${\cal N}=2$ CSM SCFTs, given by $U(N)$ Chern-Simons theory coupled to $M$ adjoint chiral matter fields and quartic superpotentials. At the leading nontrivial (two-loop) order, ${\cal M}$ is the symplectic quotient $V//U(M)$ defined by the standard symplectic form on $V$. This symplectic form, and hence the moduli space ${\cal M}=V//U(M)$, is deformed by 4-loop corrections. We also found a nontrivial 4-loop correction to the Zamolodchikov metric on ${\cal M}$, which is K\"ahler but is {\sl not} the same as the induced metric from the symplectic quotient. We gave a general argument that the metric on ${\cal M}$ is K\"ahler. It would be interesting to put further constraints on the general structure of the quantum corrected Zamolodchikov metric at finite 't Hooft coupling, which may guide us toward finding the stringy AdS dual of such SCFTs.
\subsection*{Acknowledgments}
We would like to thank F. Denef, D. Jafferis, S. Penati, and especially D. Gaiotto for discussions and correspondences.
X.Y. would like to thank S. Giombi for collaborations in unpublished work on closely related matters in CSM theories.
This work is supported in part by the Fundamental Laws Initiative Fund at Harvard University.
X.Y. is supported by NSF Award PHY-0847457.
|
\section{Motivation}
It is presumed that quantum gravitational effects would become
important at length scales of the order of the Planck length, $L_{_{\rm P}} =
(G\, \hbar /c^3)^{1/2}$. At these scales, it seems quite likely that
the description of the spacetime structure in terms of a metric, as
well as certain notions of standard quantum field theory, would have
to undergo drastic changes. Since any quantum field has virtual
excitations of arbitrary high energy, which probe arbitrarily small
scales, it follows that the conventional quantum field theory can only
be an approximate description that is valid at energies smaller than
the Planck energy. In particular, one hopes that, in a complete
theory, gravity would provide an effective cut off at the Planck
scale. Some very general considerations based on the principle of
equivalence and the uncertainty principle seem to strongly indicate
that it may not be possible to operationally define spacetime events
beyond an accuracy of the order of $L_{_{\rm P}}$\cite{1987-Padmanabhan-CQG}.
Therefore, one may consider $L_{_{\rm P}}$ as the `zero point length' of
spacetime intervals.
The existence of a fundamental length implies that processes involving
energies higher than the Planck energies will be suppressed, thereby
improving the ultra-violet behavior of the theory. However, according
to a theorem due to Weinberg, the momentum space propagator of any
Lorentz invariant and local field theory {\it has}\/ to behave
as~$p^{-2}$ in the ultra-violet limit~\cite{1979-Weinberg-Proc}.
Therefore, if the short-distance behavior of the propagators have to
be improved, one has to either break Lorentz invariance or include
non-local terms in the field theory. In this talk, we shall focus on
a latter approach that is based on the {\it hypothesis of
path-integral duality} \cite{1997-Padmanabhan-PRL}. In flat
space-time, it has been shown that the modified propagators obtained
using the duality principle are Lorentz invariant and ultra-violet
finite. We shall extend this analysis to space-times with constant
curvature and explicitly show that the two-point function of the
scalar field in these space-times are finite in the coincident
limit~\cite{2009-Kothawala.etal-PRD}.
\section{The hypothesis of path-integral duality}
The basic postulate is that the path integral amplitude of a point
relativistic particle is invariant under the `duality transformation'
${\cal R} \rightarrow (L_{_{\rm P}}^2/ {\cal R})$, where $\mathcal{R}$ is the
proper length of the path~\cite{1997-Padmanabhan-PRL}. The specific
prescription being that the original sum over the paths is modified to
\begin{equation}
G_{_{\rm PID}}(x,x')
=\sum_{\mathrm{paths}}
\exp{-m \left({\cal R}(x,x')
+ \frac{{L_{_{\rm P}}^2}}{{\cal R}(x,x')}\right)},
\end{equation}
where $m$ is the mass of the relativistic particle.
It can be shown that the resulting, modified momentum space propagator
has the following forms in the infra-red and the ultra-violet limits:
\begin{eqnarray}
G_{_{\rm PID}}(p)\propto \left\{\begin{array}{ll}
(p^{2} + m^2)^{-1},\;\;\;\; &{\rm for }\;\;\;\;
(L_{_{\rm P}}\, \sqrt{p^2 + m^2}) \ll 1,\\
\frac{\exp-\left(L_{_{\rm P}}\, \sqrt{p^2 + m^2}\right)}{L_{_{\rm P}}^{1/2}\, (p^2 + m^2)^{3/4}},
\;\;\;\; & {\rm for }\;\;\;\; (L_{_{\rm P}}\, \sqrt{p^2 + m^2}) \gg 1.\\
\end{array}\right.
\end{eqnarray}
It is interesting to note that the modified propagator probably
corresponds to a field theory with infinite derivatives and is
therefore likely to be highly non-local~\cite{1999-Padmanabhan-MPLA}.
\section{Modified propagator in constant curvature space-times}
As we mentioned, in flat space-time, the hypothesis of path-integral
duality renders the propagator ultra-violet finite. It is then
natural to enquire whether the hypothesis performs in a similar
fashion in an arbitrary curved space-time as well. It turns out to be
a formidable tasks to compute the propagator in an arbitrary
background. To make the calculations tractable, we focus on
space-times of constant curvature,
$R\propto\ell^{-2}$~\cite{2009-Kothawala.etal-PRD}. In particular, we
consider (i)~the Einstein static spacetime in $(3+1)$-dimensions [i.e.
${\rm R} \times {\rm S}^3$], (ii)~the de Sitter and the anti-de Sitter
spacetimes in $(3+1)$-dimensions [i.e. Euclidean ${\rm S}^4$ and
${\rm H}^4$], and (iii)~the anti-de Sitter spacetime in
$(2+1)$-dimensions [i.e. Euclidean ${\rm H}^3$].
Upon using the Schwinger's proper time representation, we find that
we can evaluate the modified propagator for a massive scalar field
in the above backgrounds (for details, see
Ref.~\cite{2009-Kothawala.etal-PRD}).
The main results can be summarized as follows $\left[b = \left(m^2
+ (\xi - \frac{1}{6})\, R \right);
\beta = (1 + (m\, \ell)^2 + \xi\, R\, \ell^2) \right]$:
\begin{enumerate}
\item In $(3+1)$-dimensions:
{\small
\begin{equation}
\!\!\!\!\!\!\!\!\!\!
G_{_{\rm PID}}(x, x')
= -\left(\frac{\sqrt{b}}{8 \pi}\right)\;
\frac{H_{1}^{(2)}\left(\sqrt{b\,
\left[u_{xx'}^2 -L_{_{\rm P}}^2\right]}\,\right)}{\sqrt{u_{xx'}^2
- L_{_{\rm P}}^2}}
\; \Delta_{xx'}^{1/2}
\underset{x \rightarrow x'}{\longrightarrow}
\frac{\sqrt{b}}{8\, \pi\, i\, L_{_{\rm P}}}
H_1^{(2)} \left(i\, \sqrt{b}\, L_{_{\rm P}}\right),\nonumber
\end{equation}
}
\item
In $(2+1)$-dimensions:
{\small
\begin{equation}
\!\!\!\!\!\!\!\!\!\!
G_{_{\rm PID}}(x,x')
= \left(\frac{1}{4\, \pi\,}\right)\;
\frac{\exp{-\sqrt{ (\beta / \ell^2) \, \left[u_{xx'}^2 + L_{_{\rm P}}^2\right]}}}{\sqrt{u_{xx'}^2
+ L_{_{\rm P}}^2}}
\Delta_{xx'}^{1/2}
\underset{x \rightarrow x'}{\longrightarrow}
\frac{1}{4\, \pi\, L_{_{\rm P}}}
\, \exp{-\left(\sqrt{\beta}\; \frac{L_{_{\rm P}}}{\ell} \right)},\nonumber
\end{equation}
}
\end{enumerate}
where $u_{xx'}$ represents the geodesic distance between the two
points $x$ and $x'$, and $\Delta_{xx'}$ is the Van-Vleck determinant.
We see that the duality hypothesis: (i) regulates the theory at Planck
scales, (ii) yields modifications which are non-perturbative in~$L_{_{\rm P}}$
and, (iii) most interestingly, the quantum gravitational effects, as
accounted for by the duality prescription, can be looked upon as
leading to addition of $L_{_{\rm P}}$ to {\it all}\/ spacetime (geodesic)
intervals in a (peculiar) Pythagorean way, i. e. , $\left\langle
\sigma^2(x,x')\right\rangle = [\sigma^2(x,x') + {\cal O}(1)\, L_{_{\rm P}}^2]$, as
is evident from the above expressions for the
propagators. (here, the angular brackets represent a suitable
path integral average over quantum fluctuations of the background metric.)
\section{Implications}
Current approaches to quantum gravity seem to face two main obstacles:
(i)~construct a consistent description of physics at Planck scales,
and (ii)~make robust predictions that can be tested against
experiments. The hypothesis of path-integral duality accepts our
ignorance of physics at the Planck scale and instead, based on an
underlying principle, provides a prescription for calculating
gravitationally smeared propagators which can be used to make testable
predictions. For instance, path-integral duality predicts that the
Planck scale corrections to the primordial perturbation spectrum in
exponential inflation will be of the order
of~$(H/M_{_{\rm P}})$\cite{2006-Sriramkumar.Shankaranarayanan-JHEP}.
Our results here seem to support the viewpoint that demanding the
duality invariance of the relativistic point particle path integral is
{\it equivalent}\/ to `adding' a zero-point length to spacetime
intervals. Such a result might be an outcome of {\it the generic
short distance behavior}\/ of the spacetime structure itself and
hence could be expected to naturally appear in the (effective) low
energy sector of the full theory of quantum gravity
\cite{2006-Fontanini.etal-PLB}.
|
\section{Introduction}
Parshin's conjecture states that higher algebraic $K$-groups of smooth
projective schemes over finite fields are torsion.
In \cite{ichparshin}, we studied the properties that Parshin's
conjecture would imply for rational higher Chow groups. We compared
higher Chow groups to weight homology $H_i^W(X,{{\mathbb Q}}(n))$,
defined by Jannsen \cite{jannsen}
based on the work of Gillet-Soule \cite{gilletsoule}, and obtained
a diagram
\begin{equation}\label{maina}
\begin{CD}
H_i^c(X,{{\mathbb Q}}(n)) @>\pi>> H_i^W(X,{{\mathbb Q}}(n))\\
@V\alpha VV @A\gamma AA \\
\tilde H_i^c(X,{{\mathbb Q}}(n))@>\beta >>\tilde H_i^W(X,{{\mathbb Q}}(n)).
\end{CD}\end{equation}
The terms with the tilde are the cohomology of the first
non-vanishing $E^1$-line of the niveau spectral sequence.
Parshin's conjecture in weight $n$ is equivalent to $\pi$
being an isomorphism for all $X$ and $i$. We showed that
$\pi$ is an isomorphism if and only if $\alpha$, $\beta$
and $\gamma $ are isomorphisms, and gave criteria for
this to happen.
In this article, we take the cohomological point of view and
examine the properties that Parshin's conjecture implies
for motivic cohomology with compact support. Surprisingly,
the properties obtained are not dual to the properties
for higher Chow groups, but have a different flavor.
The method to study motivic cohomology with compact support
is to use the coniveau filtration. To avoid the problems
arising from the covariance of motivic cohomology with
compact support for open embeddings (for example, one gets very large by
taking inverse limits, and has to deal with derived inverse
limits), we consider the dual groups $H^i_c(X,{{\mathbb Q}}(n))^*$. We obtain
a niveau spectral sequence, and compare it with the
spectral sequence for the dual of weight cohomology
$H^i_W(X,{{\mathbb Q}}(n))^*$ as in \cite{ichparshin} to obtain a diagram
\begin{equation}\label{mainb}
\begin{CD}
\tilde H^i_W(X,{{\mathbb Q}}(n))^* @= \tilde H^i_c(X,{{\mathbb Q}}(n))^*\\
@V\gamma^* VV @V\alpha^*VV \\
H^i_W(X,{{\mathbb Q}}(n))^*@>\pi^* >> H^i_c(X,{{\mathbb Q}}(n))^*.
\end{CD}\end{equation}
Again, the upper terms are given by the first non-vanishing row
of $E^1$-terms in the niveau spectral sequence.
The map $\pi^*$ is an isomorphism for all $X$ if and only if
Parshin's conjecture holds.
In contrast to the homological situation,
$\alpha^*$ being an isomorphism is stronger than Parshin's conjecture.
We go on to examine the relationship between diagrams
\eqref{maina} and \eqref{mainb}. Not surprisingly, this is related to
Beilinson's conjecture that rational and numerical equivalence agree up
to torsion over finite fields.
Finally we relate bounds for all four rational motivic theories
to Parshin's conjecture.
Since the purpose of this work is to understand interrelations between
certain conjectures, we assume the existence of resolution of
singularities. Its use in the results of Friendlander and Voevoesky
\cite{friedvoe} maybe be dispensable with more work because
we work with rational coefficients,
but occasionally we need a smooth and projective model for every
function field to do an induction process.
Throughout this paper, the cateogory of schemes over $k$, written
$Sch/k$ denotes the cateogory of separated schemes of finite
type over $k$, and $Sm/k$ the category of smooth schemes over $k$.
\medskip
{\it Acknowledgements:} This paper was inspired by the work of and
discussions with U.Jannsen and S.Saito.
\section{Motivic cohomology with compact support}
For a scheme $X$ over a field $k$, motivic cohomology with
compact support is defined as
$$ H^i_c(X,{{\mathbb Z}}(n))=\operatorname{Hom}_{DM^-}(M^c(X),{{\mathbb Z}}(n)[i]).$$
A concrete description is given as follows \cite[\S 3]{friedvoe}:
Let $\rho:(Sch/k)_{cdh}\to (Sm/k)_{Nis}$ be the map from the
large cdh-site of $k$ to the smooth site with the Nisnevich topology.
Let ${{\mathbb Z}}(n)$ be the motivic complex on $(Sm/k)_{Nis}$, and consider
an injective resolution $\rho^* {{\mathbb Z}}(n)\to I^\cdot$
on $(Sch/k)_{cdh}$ (we need resolution of singularites to
ensure that $\rho^*$ is exact). Let ${{\mathbb Z}}^c(X)$
be the cdh-sheafification of the presheaf which associates to
$U$ the free abelian group
generated by those subschemes $Z\subseteq X\times U$ whose
projection to $U$ induces an open embedding $Z\subseteq U$.
Then $H^i_c(X,{{\mathbb Z}}(n))=\operatorname{Hom}_{D(Shv_{cdh})}({{\mathbb Z}}^c(X),I^\cdot[i])$.
This satisfies the following properties:
\begin{enumerate}
\item Contravariance for proper maps.
\item Covariance for flat quasi-finite maps.
\item For a closed subscheme $Z$ of $X$ with open complement $U$,
there is a localization sequence
\begin{equation}\label{locali}
\cdots \to H^i_c(U,{{\mathbb Z}}(n))\to H^i_c(X,{{\mathbb Z}}(n))\to H^i_c(Z,{{\mathbb Z}}(n)) \to\cdots.
\end{equation}
\end{enumerate}
If $X$ is proper, then since ${{\mathbb Z}}^c(X)={{\mathbb Z}}(X)$, motivic cohomology with
compact support agrees with motivic cohomology
$H^i(X,{{\mathbb Z}}(n)):= H^i_{cdh}(X,{{\mathbb Z}}(n))$. Moreover,
under resolution of singularities, we get for smooth $X$
of dimension $d$ isomorphisms \cite{susvoe00, voevodskysmooth}
\begin{equation}\label{smooth}
H^i_{cdh}(X,{{\mathbb Z}}(n)) \cong H^i_{Nis}(X,{{\mathbb Z}}(n))\cong CH^n(X,2n-i).
\end{equation}
\begin{proposition}\label{vanbound}
a) We have $H^i_c(X,{{\mathbb Z}}(n))=0$ for $i>n+\dim X$.
b) If $k$ is finite, resolution of singularities exists,
and if $n>\dim X$, then $H^i_c(X,{{\mathbb Q}}(n))=0$ for $i\geq n+\dim X$.
c) If $k$ is finite and $X$ is smooth of dimension $d$, then
$H^{n+d}(X,{{\mathbb Q}}(n))=0$ unless $n=d$.
\end{proposition}
\noindent{\it Proof. }
a) Using the localization sequence and induction on the dimension,
the statement is easily reduced to the case where $X$ is proper.
Then we use that the complex ${{\mathbb Z}}(n)$ is concentrated in degrees at most
$n$, and $X$ has $cdh$-cohomological dimension $d$.
b) This was proved in \cite[Prop.6.3]{ichsuslin}. The idea is to
use induction on the dimension to reduce to $X$ smooth and proper,
and then use c).
c) If $n<d$ then this follows by comparing to higher Chow groups.
If $n>d$, consider the spectral sequence
\begin{equation}
E_1^{s,t}=\bigoplus_{x\in X^{(s)}}H^{t-s}(k(x),{{\mathbb Z}}(n-s))
\Rightarrow H^{s+t}(X,{{\mathbb Z}}(n)).
\end{equation}
In order for the $E_1^{s,t}$-terms not to vanish, we need $t\leq n$
and $s\leq d$, hence to have $s+t=n+d$ we need $s=d$ and $t=n$.
But $E_1^{d,n}$ is a sum of $H^{n-d}(k(x),{{\mathbb Z}}(n-d))$ for finite fields $k(x)$,
and higher Milnor K-theory of finite fields is torsion.
\hfill $\Box$ \\
\subsection{The niveau spectral sequence}
In order not to deal with derived inverse limits and to get smaller
groups, we work with the dual of motivic cohomology with compact
support
$$ H^i_c(X,{{\mathbb Q}}(n))^* := \operatorname{Hom}(H^i_c(X,{{\mathbb Z}}(n)),{{\mathbb Q}}).$$
These groups are covariant for proper maps and contravariant
for quasi-finite flat maps.
Let $Z_s$ be set of closed subschemes of dimension at most $s$
and let $Z_s/Z_{s-1}$ be the set of ordered pairs
$(Z,Z')\in Z_s\times Z_{s-1}$ such that $Z'\subseteq Z$.
Then $Z_s$ as well as $Z_s/Z_{s-1}$ are ordered by inclusion,
and we obtain a filtration $Z_0\subseteq Z_1\subseteq \cdots$.
We use covariance for proper maps to define
$$ H^i_c(Z_s,{{\mathbb Q}}(n))^*:=\operatornamewithlimits{colim}_{Z\in Z_s}H^i_c(Z,{{\mathbb Q}}(n))^*.$$
For a point $x\in X $ we write
$x\in Z_s$ if $\overline{\{x\}}\in Z_s$, and using contravariance
for open embeddings define
$$ H^i_c(k(x),{{\mathbb Q}}(n))^*:=\operatornamewithlimits{colim}_{U\cap\overline{\{x\}}\not=\emptyset}
H^i_c(U\cap\overline{\{x\}},{{\mathbb Q}}(n))^*.$$
Beware that this is typically not the dual of any group. For example,
for the function field $k(C)$ of a smooth and proper curve $C$
we have
$\lim_U H^1_c(U,{{\mathbb Q}}(0)) =(\prod_{C_{(0)}} {{\mathbb Q}}) /{{\mathbb Q}}$, whereas taking duals
allows us to work with the countable "predual" group
$H^1_c(k(C),{{\mathbb Q}}(0))^*=\operatornamewithlimits{colim}_U H^1_c(U,{{\mathbb Q}}(0))^*
=\ker\big( \oplus_{C_{(0)}} {{\mathbb Q}}\to {{\mathbb Q}}\big) $.
From the localization sequence we obtain
$$H^i_c(Z_s/Z_{s-1},{{\mathbb Q}}(n))^*:=\operatornamewithlimits{colim}_{(Z,Z')\in Z_s/Z_{s-1}}H^i_c(Z-Z',{{\mathbb Q}}(n)^*
=\bigoplus_{x\in Z_d}H^i_c(k(x),{{\mathbb Q}}(n))^*.$$
The usual yoga with exact couples gives
\begin{proposition}
There is a homological spectral sequence
\begin{equation}\label{sseq}
E^1_{s,t}=
\bigoplus_{x\in X_{(s)}}H^{s+t}_c(k(x),{{\mathbb Q}}(n))^*
\Rightarrow H^{s+t}_c(X,{{\mathbb Q}}(n))^*.
\end{equation}
\end{proposition}
The $d^1$-differential is given by
$$ H^{i+1}_c(Z_{s+1}/Z_s,{{\mathbb Q}}(n))^*\to H^i_c(Z_s,{{\mathbb Q}}(n))^*
\to H^i_c(Z_s/Z_{s-1},{{\mathbb Q}}(n))^*.$$
By Proposition \ref{vanbound}a), we obtain
$H^i_c(k,{{\mathbb Q}}(n))^*=0$ for $i>n+s$, i.e. $E^1_{s,t}$ vanishes
for $t>n$, so that the spectral sequence \eqref{sseq} is concentrated below
and on the line $t=n$.
On the line $t=n$, the terms $E^1_{s,n}$
vanish for $s<n$ by Proposition \ref{vanbound}b).
We define $\tilde H^j_c(X,{{\mathbb Q}}(n))^*$ to be the cohomology of the line
$E^1_{*,n}$
\begin{equation}\label{comp}
\bigoplus_{x\in X_{(n)}}H^{2n}_c(k(x),{{\mathbb Q}}(n))^* \leftarrow \cdots
\leftarrow\bigoplus_{x\in X_{(d)}}H^{n+d}_c(k(x),{{\mathbb Q}}(n))^*,
\end{equation}
where we put the term indexed by $X_{(i)}$ in degree $n+i$.
It is easy to check that we obtain canonical maps
\begin{equation}\label{canmap}
\tilde H^i_c(X,{{\mathbb Q}}(n))^*\stackrel{\alpha^*}{\to} H^i_c(X,{{\mathbb Q}}(n))^*
\end{equation}
\section{Parshin's conjecture}
Parshin's conjecture states that for all smooth and projective
$X$ over ${{\mathbb F}}_q$, the groups $K_i(X)_{{\mathbb Q}}$ are torsion for $i>0$.
In \cite{ichtate} we showed that it is implies by Tate's conjecture
and Beilinson's conjecture that rational and numerical equivalence
agree up to torsion.
Since $K_i(X)_{{\mathbb Q}}=\oplus_n H^{2n-i}(X,{{\mathbb Q}}(n))$, it follows that Parshin's
conjecture is equivalent to the following conjecture for all $n$.
\medskip
\noindent{\bf Conjecture $P^n$:}
{\it For all smooth and projective schemes $X$
over the finite field ${{\mathbb F}}_q$, and all $i\not=2n$, the group $H^i(X,{{\mathbb Z}}(n))$
is torsion.}
\medskip
Conjecture $P^n$ is known for $n=0,1$ and is trivial for $n<0$.
In \cite{ichparshin}, we considered the homological analog
(it was denoted $P(m)$ in loc.cit.):
\medskip
\noindent{\bf Conjecture $P_m$:}
{\it For all smooth and projective schemes $X$
over the finite field ${{\mathbb F}}_q$, and all $i\not=2m$, the group $H_i^c(X,{{\mathbb Z}}(m))$
is torsion.}
\medskip
This conjecture is not known for any $m$.
One can also consider the restrictions $P^n(d)$ and $P_m(d)$
of the above conjectures to varieties of dimension at most $d$.
By the projective bundle formula one gets $P^n(d)\Rightarrow P^{n-1}(d-1)$
and $P_m(d)\Rightarrow P_{m-1}(d-1)$, hence $P^n\Rightarrow P^{n-1}$
and $P_m\Rightarrow P_{m-1}$.
\begin{lemma}
We have $P^n(d)\Leftrightarrow P_{d-n}(d)$.
\end{lemma}
\noindent{\it Proof. }
Let $X$ be smooth and projective of dimension $e\leq d$. Then
conjecture $P^{n-d+e}$ holds for $X$, hence the formula
$H^i(X,{{\mathbb Z}}(a))\cong H_{2e-i}^c(X,{{\mathbb Z}}(e-a))$ implies conjecture $P_{d-n}$
for $X$. The converse is proved the same way.
\hfill $\Box$ \\
Since conjecture $P^{-1}$ is trivially true, the following Lemma
explains why the spectral sequence for homology with compact support
in \cite{ichparshin} is concentrated in degrees $s\geq n$,
whereas \eqref{sseq} a priori is not:
\begin{lemma}
If conjecture $P_{-1}$ holds, $H^i_c(X,{{\mathbb Q}}(n))=0$ for $n>d=\dim X$ and any $X$.
In particular, the terms $E^1_{s,t}$ vanish for $s<n$ in the spectral
sequence \eqref{sseq}.
\end{lemma}
\noindent{\it Proof. }
By induction on the dimension of $X$ and the sequence \eqref{locali} we can
assume that $X$ is smooth and proper. Then
$H^i_c(X,{{\mathbb Q}}(n))=H_{2d-i}^c(X,{{\mathbb Q}}(d-n))$ which vanishes
by conjecture $P_{-1}$.
\hfill $\Box$ \\
\begin{lemma}\label{Ha}
The following statements are equivalent:
\begin{enumerate}
\item Conjecture $P^n$.
\item For all schemes $X$ over ${{\mathbb F}}_q$, we have $H^i_c(X,{{\mathbb Q}}(n))=0$ for $i<2n$.
\item For all finitely generated fields $k/{{\mathbb F}}_q$,
we have $H^i_c(k,{{\mathbb Q}}(n))^*=0$ for $i<2n$.
\end{enumerate}
\end{lemma}
\noindent{\it Proof. }
a) $\Rightarrow$ b) follows by induction on the dimension and
localization to recude to the smooth and proper case,
b) $\Rightarrow$ c) follows by taking colimits, and
c) $\Rightarrow$ a) follows with the spectral sequence \eqref{sseq}.
\hfill $\Box$ \\
It is not a priori clear if the terms $H^i_c(k(x),{{\mathbb Q}}(n))$ with
$2n\leq i<\operatorname{trdeg} k(x)+n$ should vanish or not. Thus the following
statement is possibly stronger than Parshin's conjecture
(but see Proposition \ref{compareto}):
\begin{proposition}\label{themapg}
The following statements are equivalent:
\begin{enumerate}
\item Conjecture $P^n$ holds, and for smooth and projective $X$ we have
$$ \tilde H^i_c(X,{{\mathbb Q}}(n))^* \cong \begin{cases}
CH^n(X)^* &i=2n;\\
0 &\text{else}.
\end{cases}$$
\item The groups $H^i_c(k,{{\mathbb Q}}(n))^*$ vanish for $i\not=n+\operatorname{trdeg} k$.
\item The map $\alpha^*$ is an isomorphism for all $X$.
\end{enumerate}
\end{proposition}
\noindent{\it Proof. }
a) $\Rightarrow$ b): We proceed by induction on the transcendence
degree. Choose a smooth and projective model $X$ of $k$.
Since $H^i_c(X,{{\mathbb Q}}(n))$ is $CH^n(X)$ for $i=2n$ and
vanishies for $i\not=2n$, an inspection of
the spectral sequence \eqref{sseq} shows the vanishing.
b) $\Rightarrow$ c) is clear.
c) $\Rightarrow$ a): Conjecture $P^n$ follows because
$\tilde H^i_c(X,{{\mathbb Q}}(n))^*$ vanishes for $i< 2n$, and the sequence
is exact because for smooth and proper $X$,
$H^i_c(X,{{\mathbb Q}}(n))^*$ vanishes for $i>2n$
and is isomorphic to $CH^n(X)$ for $i=2n$.
\hfill $\Box$ \\
The statements of this Proposition are non-trivial
even in the case $n=0$ (but they can be proven with methods similar
to \cite[Thm.5.10]{jannsen} in this case).
\section{Weight cohomology}
Let ${\cal C}$ be category of correspondences with objects smooth projective
varieties $[X]$ over the field $k$,
$\operatorname{Hom}_{\cal C}([X],[Y])=\oplus CH_{\dim X_i}(X_i\times Y)_{{\mathbb Q}}$ for $X=\coprod X_i$
the decomposition into connected components, and the usual composition
of correspondenes. In \cite{gilletsoule},
Gillet and Soul\'e defined, for every separated scheme of finite type, a
weight complex $W(X)$ in the homotopy category of bounded
complexes in ${\cal C}$, satisfying the following properties
\cite[Thm. 2]{gilletsoule}:
\begin{enumerate}
\item $W(X)$ is represented by a bounded complex
$$ [X_0]\gets [X_1]\gets \dots \gets [X_k]$$
with $\dim X_i\leq \dim X-i$.
\item $W(-)$ is covariant functorial for proper maps.
\item $W(-)$ is contravariant functorial for open embeddings.
\item If $T\to X$ is a closed embedding with open complement $U$,
then there is a distinguished triangle
$$ W(T)\stackrel{i_*}{\longrightarrow} W(X) \stackrel{j^*}{\longrightarrow}
W(U).$$
\end{enumerate}
Our notation differs from loc.cit. in variance. In loc.cit.,
resolution of singularities is used to obtain an integral result,
but see \cite{gilletsoule2} for a rational result.
We define dual weight cohomology (with compact support)
$H_W^i(X,{{\mathbb Q}}(n))^*$
to be the $i$th cohomology of the complex
$$CH^n(X_0)^*\leftarrow CH^n(X_1)^*\leftarrow CH^n(X_2)^*\leftarrow \cdots,$$
induced by contravariance of $CH^n$, and with
$CH^n(X_i)^*$ placed in degree $2n+i$.
Note that this is the dual of the functor obtained via the
contravariant analog of \cite[Thm.5.13]{jannsen} from the (contravariant)
functor $CH^n(-)$ on the category ${\cal C}$.
We define dual weight cohomology of a field to be
$$H^i_W(K,{{\mathbb Q}}(n))^*:= \operatornamewithlimits{colim}_U H^i_W(U,{{\mathbb Q}}(n))^*,$$
where $U$ runs through smooth schemes with function field $K$.
\begin{lemma}\label{popoa}
We have $H^i_W(X,{{\mathbb Q}}(n))^*=0$ unless $2n\leq i\leq \dim X+n$. In particular,
$H^i_W(K,{{\mathbb Q}}(n))^*=0$ for every finitely generated field $K/k$ unless
$2n\leq i\leq \operatorname{trdeg}_k K+n$.
\end{lemma}
\noindent{\it Proof. } This follows from the first property of
weight complexes together with $CH^n(T)=0$ for $n>\dim T$.
\hfill $\Box$ \\
It follows from Lemma \ref{popoa} that the niveau spectral sequence
\begin{equation}\label{uuiioo}
E^1_{s,t}=\bigoplus_{x\in X_{(s)}}H^{s+t}_W(k(x),{{\mathbb Q}}(n))^*\Rightarrow
H^{s+t}_W(X,{{\mathbb Q}}(n))^*
\end{equation}
is concentrated on and below the line $t=n$ and on and above the
line $s+t=2n$. If we let
$\tilde H^i_W(X,{{\mathbb Q}}(n))^*=E^2_{i-n,n}(X)$ be the homology of the complex
\begin{equation}\label{jannsenco}
\bigoplus_{x\in X_{(n)}}H^{2n}_W(k(x),{{\mathbb Q}}(n))^*\leftarrow
\cdots \leftarrow \bigoplus_{x\in X_{(d)}}H^{n+d}_W(k(x),{{\mathbb Q}}(n))^*,
\end{equation}
then we obtain a canonical and natural map
$$\gamma^*:\tilde H^i_W(X,{{\mathbb Q}}(n))^*\to H^i_W(X,{{\mathbb Q}}(n))^* .$$
\subsection{Comparison}
We are going to check the hypothesis of \cite[Prop.5.16]{jannsen}
to construct a functor between motivic cohomology with compact support
and weight cohomology.
Recall that motivic cohomology with compact support is
defined as the cohomology of $C'(X)= \operatorname{Hom}_{D(Shv_{cdh})}({{\mathbb Z}}^c(X),I^\cdot)$,
where $\rho^* {{\mathbb Z}}(n)\to I^\cdot$ is an injective resolution
on the cdh-site. Then $C'$ is a covariant functor from the category of schemes
over $k$ with proper maps to the category of complexes with bounded above
cohomology, which is contravariant for open embeddings.
Moreover, for proper $X$ we have $C'(X)=I^\cdot(X)$, and
a closed embedding $i:Y\to X$ with
open complement $j:U\to X$ gives a short exact sequence
$$0\to C'(U)\to C'(X)\to C'(Y)\to 0.$$
Restricting $C'$ to smooth and proper $X$, we have
$H^iC'(X)=0$ for $i>2n$, and a functorial isomorphism
$$H^{2n}C'(X)=H^{2n}I^\cdot(X)\cong \tau_{\geq 2n}I^\cdot(X) \cong CH^n(X).$$
by \eqref{smooth}. We obtain a morphism of functors on the category
of smooth and proper schemes,
$$C' = I^\cdot \to \tau_{\geq 2n}I^\cdot
\stackrel{\sim}{\longleftarrow}H^{2n}(I^\cdot)[-2n] = CH^n(-)[-2n]$$
Reversing all the arrows induced by arrows between schemes, but not
by arrows between cohomology
theories in the proof of \cite[Prop.5.16]{jannsen} gives a natural
transformation $H^i_c(X,{{\mathbb Z}}(n))\to H^i_W(X,{{\mathbb Z}}(n))$,
hence a natural transformation
$$\pi^* :H^i_W(X,{{\mathbb Q}}(n))^*\to H^i_c(X,{{\mathbb Q}}(n))^* .$$
From now on we return to the situation $k$ finite.
\begin{proposition}\label{samefields}
Assume that every finitely generated field $K/k$ has a smooth and
projective model over $k$, and let $K$ be finitely generated
of transcendence degree $d$ over $k$.
a) The map $\pi^*$ induces isomorphisms
$$H^{n+d}_W(K,{{\mathbb Q}}(n))^*\stackrel{\sim}{\to} H^{n+d}_c(K,{{\mathbb Q}}(n))^*.$$
In particular, we have $\tilde H^i_W(X,{{\mathbb Q}}(n))^*\cong \tilde H^i_c(X,{{\mathbb Q}}(n))^*$.
b) If $d>n$, then $\pi^*$ induces isomorphisms
$$H^{n+d-1}_W(K,{{\mathbb Q}}(n))^*\stackrel{\sim}{\to} H^{n+d-1}_c(K,{{\mathbb Q}}(n))^*.$$
\end{proposition}
\noindent{\it Proof. }
We proceed by induction on $d$.
Given $K$ of transcendence degree $d$, choose a smooth and projective
model $X$ of $K$ and compare \eqref{sseq} and \eqref{uuiioo}.
a) If $d<n$, then both terms vanish by Proposition \ref{vanbound}b) and
Lemma \ref{popoa}. For $d=n$ we obtain
$CH^n(X)\cong H^{n+d}_c(K,{{\mathbb Q}}(n))^*\cong H^{n+d}_W(K,{{\mathbb Q}}(n))^*$.
For $d>n$, we obtain
from $H^{n+d}_c(X,{{\mathbb Q}}(n))= H^{n+d}_W(X,{{\mathbb Q}}(n))=0$ a commutative diagram
with exact rows
$$\begin{CD}
\cdots @<<< \bigoplus_{x\in X_{(d-1)}}H^{n+d-1}_W(k(x),{{\mathbb Q}}(n))^*@<<<
H^{n+d}_W(K,{{\mathbb Q}}(n))^*@<<< 0\\
@EEE @| @VVV\\
\cdots @<<< \bigoplus_{x\in X_{(d-1)}}H^{n+d-1}_c(k(x),{{\mathbb Q}}(n))^*@<<<
H^{n+d}_c(K,{{\mathbb Q}}(n))^*@<<< 0.
\end{CD}$$
b) follows by a similar argument, noting that the $d_2$-differentials
originating from the terms in question end in terms considered in a),
and there are no higher differentials.
\hfill $\Box$ \\
We obtain a commutative diagram
\begin{equation}\label{maind}
\begin{CD}
\tilde H^i_W(X,{{\mathbb Q}}(n))^* @= \tilde H^i_c(X,{{\mathbb Q}}(n))^*\\
@V\gamma^*VV @V\alpha^*VV \\
H^i_W(X,{{\mathbb Q}}(n))^*@>\pi^*>> H^i_c(X,{{\mathbb Q}}(n))^*.
\end{CD}
\end{equation}
\begin{proposition}
The following statements are equivalent:
\begin{enumerate}
\item Conjecture $P^n$.
\item The map $\pi^*$ is isomorphisms for all $X$.
\item We have $H^i_W(k,{{\mathbb Q}}(n))^*\cong H^i_c(k,{{\mathbb Q}}(n))^*$ for all $i$ and $k$.
\end{enumerate}
\end{proposition}
\noindent{\it Proof. }
a) $\Leftrightarrow$ b): For smooth and proper $X$ this is clear.
In general, one does induction on the dimension and uses localization
sequences.
b) $\Leftrightarrow$ c): One direction follows by taking colimits,
and the other by comparing the spectral sequences \eqref{sseq}
and \eqref{uuiioo}.
\hfill $\Box$ \\
The following Proposition is analog to Proposition \ref{themapg}
and dual to \cite[Prop.3.4]{ichparshin}:
\begin{proposition}
The following statements are equivalent and follow from $\alpha^*$ being
an isomorphism:
\begin{enumerate}
\item For smooth and projective $X$, we have
$$\tilde H^i_W(X,{{\mathbb Q}}(n))^* \cong \begin{cases}
CH^n(X)^* &i=2n;\\
0 &\text{else}.
\end{cases}$$
\item The groups $H^i_W(k,{{\mathbb Q}}(n))^*$ vanish for $i\not=n+\operatorname{trdeg} k$.
\item The map $\gamma^*$ is an isomorphism for all $X$ and $i$.
\end{enumerate}
\end{proposition}
\noindent{\it Proof. }
The proof is similar to Proposition \ref{themapg}.
a) $\Rightarrow$ b): We proceed by induction on the transcendence
degree. Choose a smooth and projective model $X$ of $k$.
Since $H^i_W(X,{{\mathbb Q}}(n))$ is $CH^n(X)$ for $i=2n$ and
vanishes for $i\not=2n$, an inspection of
the spectral sequence \eqref{sseq} gives the result.
b) $\Rightarrow$ c) $\Rightarrow$ a) are clear. If $\alpha^*$ is
an isomorphism, then so is $\pi^*$, and hence $\gamma^*$.
\hfill $\Box$ \\
\section{Beilinson's conjecture and duality}
Beilinson conjectured
that over a finite field, rational and numerical
equivalence agrees up to torsion. This can be reformulated
to the following:
\medskip
\noindent{\bf Conjecture $D(n)$:}
{\it For all smooth and projective schemes $X$
over the finite field ${{\mathbb F}}_q$, the intersection pairing
gives a functorial isomorphism
$$ CH^n(X)_{{\mathbb Q}}\cong \operatorname{Hom}(CH_n(X),{{\mathbb Q}}).$$}
\medskip
Note that since both sides are countable, this implies finite
dimensionality. By the projection formula, the pairing induces
a map of complexes
$$\begin{CD}
CH_n(X_0)@<<< CH_n(X_1)@<<< CH_n(X_2)@<<< \cdots\\
@VVV @VVV @VVV\\
CH^n(X_0)^*@<<< CH^n(X_1)^*@<<< CH^n(X_2)^* @<<< \cdots.
\end{CD}$$
Taking homology, we obtain a map
$$ \delta :H_i^W(X,{{\mathbb Q}}(n)) \to H^i_W(X,{{\mathbb Q}}(n))^*.$$
Taking the limit over decreasing open sets with function field $K$,
$\delta$ induces a map $H_i^W(K,{{\mathbb Q}}(n)) \to H^i_W(K,{{\mathbb Q}}(n))^*$.
This in turn induces a map of complexes
$$\begin{CD}
\bigoplus_{x\in X_{(n)}}H_{2n}^W(k(x),{{\mathbb Q}}(n))
@<<< \bigoplus_{x\in X_{(n+1)}}H_{2n+1}^W(k(x),{{\mathbb Q}}(n))@<<< \cdots\\
@VVV @VVV \\
\bigoplus_{x\in X_{(n)}}H^{2n}_W(k(x),{{\mathbb Q}}(n))^*
@<<< \bigoplus_{x\in X_{(n+1)}}H^{2n+1}_W(k(x),{{\mathbb Q}}(n))^* @<<< \cdots,
\end{CD}$$
which gives the map $\tau$ making the following diagram commutative
\begin{equation}\label{alldiagram} \begin{CD}
H_i^c(X,{{\mathbb Q}}(n))@>\pi>> H_i^W(X,{{\mathbb Q}}(n))@>\delta>> H^i_W(X,{{\mathbb Q}}(n))^*
@>\pi^* >>H^i_c(X,{{\mathbb Q}}(n))^*\\ \
@V\alpha VV @A\gamma AA @A\gamma^* AA @A\alpha^* AA \\
\tilde H_i^c(X,{{\mathbb Q}}(n))@>\beta >> \tilde H_i^W(X,{{\mathbb Q}}(n))@>\tau>>
\tilde H^i_W(X,{{\mathbb Q}}(n))^* @= \tilde H^i_c(X,{{\mathbb Q}}(n))^*.
\end{CD}\end{equation}
\begin{lemma}
Conjecture $D(n)$ is equivalent to $\delta$ being an isomorphism for all
$i$ and $X$, and implies that $\tau$ is an isomorphism for all $i$ and $X$.
\end{lemma}
\noindent{\it Proof. }
The equivalence follows from the definition of $\delta$, and
the statement about $\tau$ follows by a colimit argument.
\hfill $\Box$ \\
Parshin's conjecture and Beilinson's conjecture can be combined
into the following
\medskip
\noindent{\bf Conjecture $BP(n)$:}
{\it For all smooth and projective schemes $X$
over the finite field ${{\mathbb F}}_q$, the cup product pairing
$$ H^i(X,{{\mathbb Q}}(n))\times H^{2d-i}(X,{{\mathbb Q}}(d-n))\to {{\mathbb Q}} $$
is perfect.}
\medskip
\begin{proposition}\label{compareto}
For fixed $n$, the following statements are equivalent:
\begin{enumerate}
\item Conjecture $BP(n)$.
\item Conjectures $D(n)$, $P^n$ and $P_n$.
\item There are perfect pairings of finite
dimensional vector spaces
$$ H_i^c(X,{{\mathbb Q}}(n))\times H^i_c(X,{{\mathbb Q}}(n))\to {{\mathbb Q}} $$
for all $X$, respectively smooth projective $X$.
\item All maps in \eqref{alldiagram} are isomorphisms for all $X$,
respectively for all smooth and proper $X$.
\end{enumerate}
\end{proposition}
\noindent{\it Proof. }
a) $\Leftrightarrow$ b): If $i>2n$, then the left hand side
in $BP(n)$ vanishes, hence
perfectness is equivalent to the vanishing of
$H^{2d-i}(X,{{\mathbb Q}}(d-n))\cong H_i^c(X,{{\mathbb Q}}(n))$
for $i>2n$, i.e. conjecture $P_n$ of \cite{ichparshin}.
If $i<2n$, then the right hand side in $BP(n)$ vanishes, so perfectness
is equivalent to $P^n$.
For $i=2n$, we recover conjecture $D(n)$.
b) $\Leftrightarrow$ c):
Clearly conjecture $BP(n)$ is a special case of the assertion in c).
For the other direction, it suffices to construct a functorial map
$H_i^c(X,{{\mathbb Q}}(n))\to H^i_c(X,{{\mathbb Q}}(n))^*$ which is the intersection pairing
for smooth and projective $X$, and which is
compatible with localization sequences on both sides. Indeed having
such a map one can use the usual devissage to reduce to the case that $X$
is smooth and projective. One way to construct such a map is
to write $H_i^c(X,{{\mathbb Z}}(n))\cong \operatorname{Hom}_{DM^-}({{\mathbb Z}}(n)[i],M^c(X))$,
$H^i_c(X,{{\mathbb Z}}(n))\cong \operatorname{Hom}_{DM^-}(M^c(X),{{\mathbb Z}}(n)[i])$, where
$DM^-$ is Voevodsky's triangulated category of homotopy invariant
Nisnevich sheaves with transfers. Then the pairing is given by the
composition
\begin{multline*}
\operatorname{Hom}_{DM^-}({{\mathbb Z}}(n)[i],M^c(X))\times \operatorname{Hom}_{DM^-}(M^c(X),{{\mathbb Z}}(n)[i])
\to \\
\operatorname{Hom}_{DM^-}({{\mathbb Z}}(n),{{\mathbb Z}}(n)) \cong \operatorname{Hom}_{DM^-}({{\mathbb Z}},{{\mathbb Z}})\cong {{\mathbb Z}},
\end{multline*}
using the cancellation theorem.
b) $\Leftrightarrow$ d)
Conjecture $P_n$, $D(n)$ and $P^n$ imply that the left square, middle
horizontal maps, and right horizontal maps of \eqref{alldiagram} are
isomorphisms for all $X$.
Conversely, isomorphisms of the three upper maps of \eqref{alldiagram}
for smooth and proper $X$
imply that $P_n$, $D(n)$, and $P^n$ hold, respectively.
\hfill $\Box$ \\
\section{Parshin's conjecture and the four motivic theories}
Recall from \cite{friedvoe} that we have four motivic theories:
Motivic cohomology, motivic cohomology with compact support,
motivic homology and motivic homology with compact support.
All four theories are homotopy invariant and satisfy a
projective bundle formula.
Motivic cohomology is contravariant, has a Mayer-Vietoris long
exact sequence for Zarsiki covers, and a long exact sequence
for abstract blow-ups. Motivic cohomology is contravariant
for proper maps, covariant for quasi-finite flat maps, and
satisfies a localization long exact sequence (which implies
in particular Mayer-Vietoris and abstract blow-up
long exact sequences). Motivic homology and motivic homology
with compact support satisfy the dual properties.
The theories are related by the following diagram
$$ \begin{CD}
H^i_c(X,{{\mathbb Q}}(n))@>proper>\sim> H^i(X,{{\mathbb Q}}(n))\\
@VsmoothV\cong V @VsmoothV\cong V \\
H_j(X,{{\mathbb Q}}(m))@>proper>\sim> H_j^c(X,{{\mathbb Q}}(m))
\end{CD}$$
The horizontal maps are isomorphisms for proper $X$, and the
vertical maps are isomorphisms if $X$ is smooth of pure dimension $d$,
and $m+n=d$ and $j+i=2d$. The functorialities suggest
that groups diagonally opposite should be in some form of duality;
we saw that with rational coefficients, this is equivalent to deep
conjectures, for a result with torsion coefficients see \cite{ichdual}.
The following diagram describes the range where these groups
can be non-zero, where they can be non-zero
assuming Parshin's conjecture, where they can be non-zero
assuming Parshin's conjecture
plus smoothness of $X$, and where they can be non-zero
assuming Parshin's conjecture plus properness of $X$, respectively.
The bold faced inequalities indicate that they are strong enough
to recover Parshin's conjecture.
\medskip
\noindent
{\renewcommand\arraystretch{1.5}
\begin{tabular}{l|c|c|c|c}
&Coh compact sup&Mot Cohomology&Mot Homology&Borel-Moore hom\\
\hline
&$H^i_c(X,{{\mathbb Q}}(n))$&$ H^i(X,{{\mathbb Q}}(n))$ &$ H_j(X,{{\mathbb Q}}(m))$ &$ H_j^c(X,{{\mathbb Q}}(m))$ \\
\hline
always&$i\leq n+d$ &$i\leq n+d $&$j\geq m$&$j\geq 2m $\\
& &$i\leq 2n$ X smooth & $j\geq 2m$ X proper \\
Parshin $\Rightarrow$ &
${\bf 2n\leq i}\leq n+d$ &$ n\leq i\leq n+d $&$ m\leq j\leq m+d$
&$2m\leq j\leq m+d$ \\
P+smooth& &$n\leq i\leq 2n$&$m\leq {\bf j\leq 2m}$&\\
P+proper& &${\bf 2n\leq i}\leq n+d$&$2m\leq j\leq m+d$&
\end{tabular}}
\bigskip
\noindent{\it Proof. }
The first row follows from the definitions (and that the cdh-cohomological
dimension agress with the dimension).
Since Borel-Moore homology $H_j^c(X,{{\mathbb Q}}(m))$ is isomorphic to
higher Chow groups $CH_m(X,j-2m)$, they can only be non-zero for
$j\geq 2m$. The second row is the translation of this fact into a statement
for motivic cohomology for smooth $X$, and for motivic homology
for proper $X$.
The results under Parshin's conjecture for Borel-Moore homology
and motivic cohomology with compact support can be obtained by using
induction on the dimension and the localization sequences.
To obtain them for motivic homology and cohomology,
one uses the isomorphisms
$H_i(X,{{\mathbb Z}}(n))\cong H^{2d-i}_c(X,{{\mathbb Z}}(d-n))$
and $H^i(X,{{\mathbb Z}}(n))\cong H_{2d-i}^c(X,{{\mathbb Z}}(d-n))$
for a smooth scheme $X$ of dimension $d$ to obtain the result for
smooth schemes. Then induction on the dimension and the blow-up long exact
sequences gives results for all schemes.
The extra information for the smooth and proper case in case
of homology and cohomology is obtained by comparing to the other
theories.
\hfill $\Box$ \\
The bold faced inequalities were a
motivation to write this paper: It might be difficult to prove
a statement which only holds for smooth and proper $X$, as in
the case of higher Chow groups. It might be easier to prove
a statement which holds for all smooth schemes (motivic homology), or
all proper schemes (motivic cohomology), or all schemes
(motivic cohomology with compact support).
|
\section{Introduction}
\label{sec: intro}
Fluctuations are known to play a key role in sufficiently
low-dimensional systems, whether classical or quantum, as they
can preempt spontaneous symmetry breaking.
When the symmetry is both global and continuous,
the tool of choice to address the role
of fluctuations in low-dimensional systems is the non-linear sigma model
(NL$\sigma$M). However, the usefulness of NL$\sigma$Ms
has come to transcend situations in which
a pattern of symmetry breaking is immediately obvious.
For example, NL$\sigma$Ms
have been used with success in the context of Anderson localization
(see Ref.~\onlinecite{Efetov97} for a review)
to access the transition from a metallic to an insulating phase
induced by weak disorder or to compute probability distributions
of spectral,\cite{Mirlin00} wavefunction,\cite{Foster09}
and transport characteristics in chaotic metallic grains and
disordered electronic systems.\cite{Efetov97}
Quite generally, the construction of a generic NL$\sigma$M on
a connected Riemannian manifold $\mathfrak{M}$
of finite dimension $\mathfrak{n}$,
the ``target manifold'',
can proceed in the following way.%
\cite{Friedan85}
One assigns to any point from Euclidean space in $d$ dimensions,
specified by coordinates $x^{\mu}$ ($\mu=1,\cdots,d$),
a point in the manifold $\mathfrak{M}$ with the coordinates $\phi^{i}(x)$
($i=1,\cdots,\mathfrak{n}$).
The simplest action $S$, which is
made of two derivatives of the coordinates $\phi^i$,
and is invariant under both the rotations of Euclidean space
and reparametrization of the target manifold, is
\begin{eqnarray}
\label{eq: construction NLSM on rieman manifold}
S=
\frac{1}{4\pi t}
\int \frac{d^{d}x}{\mathfrak{a}^{d-2}}
G^{\ }_{ij}\big[\phi(x)\big]
\partial^{\ }_{\mu}\phi^{i}(x)
\partial^{\ }_{\mu}\phi^{j}(x)
\end{eqnarray}
where $G^{ }_{ij}\left[\phi\right]$
is a component of the metric tensor on $\mathfrak{M}$,
$t$ is the coupling constant,
and $\mathfrak{a}$ is the short-distance cutoff.
The target manifold can be either
compact or non-compact.
An example of a NL$\sigma$M on a compact target manifold
is the O($N$)/O($N-1$) NL$\sigma$M
with $2<N=3,4,5,\cdots$ when
the target manifold is the unit sphere $S^{N-1}$
in $N$-dimensional Euclidean space.
When $N=3$ it describes spontaneous symmetry breaking
in a classical ferromagnet.
Non-compact target manifolds
are of relevance to the problem of Anderson localization
in the bosonic ``replica limit'' $N\to 0$ or
when the manifold is generalized to a supermanifold.%
\cite{Efetov97}
In Anderson localization the coupling constant $t$ has the meaning
of the inverse of the mean dimensionless conductance.%
\cite{Lee85}
The implicit assumption made in the construction%
~(\ref{eq: construction NLSM on rieman manifold})
is that all the invariant
scalars that contain $2s$ ($1<s=2,3,\cdots$)
derivatives of the field can be ignored.
The standard justification for this assumption is that
their ``engineering dimension'' $2s$ is
much larger than the spatial dimension $d=2+\epsilon$,
i.e., they are irrelevant in the renormalization group (RG)
sense, and this is expected to remain so after renormalization in
$d=2+\epsilon$ dimensions
for small $\epsilon$, and thus small $t$.
This assumption was called into question in Refs.%
~\onlinecite{Kravtsov88}--\onlinecite{Ryu07a},
for which the main results can be illustrated
most simply by the example
of the O($N$)/O($N-1$) NL$\sigma$M. We recall that the
O($N$)/O($N-1$) NL$\sigma$M has
an
infra-red
unstable
fixed point
located, to one loop order,
at $t^*=\epsilon/(N-2)$,
from which emerges a renormalization group (RG) flow
to strong and weak coupling.
In Ref.\ \onlinecite{Wegner90},
a family of perturbations of the
O($N$)/O($N-1$) NL$\sigma$M action
(\ref{eq: construction NLSM on rieman manifold}),
which we shall call high-gradient operators, was considered.
A high-gradient operator of order $s$ is
a homogeneous polynomial of order $2s$ in the derivatives of the fields
(all located at the same point) which is a scalar with respect
to both the symmetry group of the NL$\sigma$M [i.e., O($N$)]
and the rotation group of Euclidean space.
The minimum (i.e., dominant, or ``leading'')
value of the one-loop scaling dimensions%
~\cite{footnote: our conventions for dimensions}
of the high-gradient operators of order $s$
at the fixed point $t^{*}$ is found\cite{Wegner90} to be
\begin{eqnarray}
x^{(s)}&=&
2s
-
s(s-1)t^*
+
\mathcal{O}(\epsilon^2).
\label{eq: anomalous scaling dimensions for O(N) high-gradient operators}
\end{eqnarray}
Although strongly irrelevant by power counting
(i.e., in the absence of fluctuation corrections, $t^* \to 0$),
high-gradient operators of
order $s$ thus acquire a one-loop scaling dimension smaller
than two when the order $2s$ of derivatives
is large enough so that $s t^* \approx s \epsilon/(N-2) \sim 2$,
and thus would appear to become relevant,
based on the one-loop result.
In $d=2$ dimensions,
the lowest one-loop scaling dimensio
\cite{footnote: on anomalous dimensions}
for all high-gradient operators of order $s$ is
\begin{eqnarray}
x^{(s)}&=&
2s
-
s(s-1)t
+
\mathcal{O}(t^2)
\label{eq: anomalous dimensions for O(N) high-gradient operators}
\end{eqnarray}
along the trajectory to strong coupling
away from the infra-red unstable fixed point $t=0$.
Two-loop counterparts to Eqs.%
~(\ref{eq: anomalous scaling dimensions for O(N) high-gradient operators})
and (\ref{eq: anomalous dimensions for O(N) high-gradient operators})
yield the same conclusion:~\cite{Castilla97}
high-gradient operators of sufficiently
high-order $s$ appear to be relevant for any given
dimension $d=2+\epsilon$ at the non-trivial fixed point.
Similar results hold for the NL$\sigma$Ms defined on
the compact target manifolds
($M$ and $N$ are positive integers)
$\mathrm{Sp}(M+N)/\mathrm{Sp}(M)\times \mathrm{Sp}(N)$,%
~\cite{Kravtsov88,Kravtsov89}
$\mathrm{U}(M+N)/\mathrm{U}(M)\times \mathrm{U}(N)$,%
~\cite{Lerner90,Wegner91}
$\mathrm{O}(M+N)/\mathrm{O}(M)\times \mathrm{O}(N)$,%
~\cite{Mall93}
and on families of compact K\"ahler
(and super) manifolds.%
~\cite{Ryu07a}
Generalizations to the non-compact target manifolds
$\mathrm{Sp}(M,N)/\mathrm{Sp}(M)\times\mathrm{Sp}(N)$,
$\mathrm{U}(M,N)/\mathrm{U}(M)\times\mathrm{U}(N)$,
and
$\mathrm{O}(M,N)/\mathrm{O}(M)\times\mathrm{O}(N)$
follow from the rule that the coupling $t$ of the
compact NL$\sigma$M entering in one-loop anomalous dimensions
must be replaced by $-t$ in the corresponding non-compact
NL$\sigma$M.
In Anderson localization, compact target manifolds arise
when using fermionic replicas for disorder averaging,
whereas non-compact target manifolds arise when using
the bosonic replicas for disorder averaging.
If one uses supersymmetric disorder averaging,
the resulting NL$\sigma$M has both compact
and non-compact sectors.~\cite{Efetov97}
The high-gradient operators in the NL$\sigma$M defined on
$\mathrm{AdS}_5\times S^5$
($\mathrm{AdS}_5$
is non-compact
whereas $S^5$
is compact)
have also been discussed
in the context of
the AdS/CFT correspondence.
(See, for example, Refs.%
~\onlinecite{Polyakov01}--\onlinecite{Polyakov05}.)
The substitution $t\to-t$ does not affect
the value of the minimal
(i.e., dominant, or ``leading'')
one-loop scaling dimensions, when the
spectrum of anomalous one-loop
dimensions\cite{footnote: our conventions for dimensions}
of all high gradient operators of order $s$ is
distributed symmetrically about zero.
This turns out to be the case
whenever $m,n>1$ in the above examples.
On the other hand, there are some target manifolds, the simplest
examples being
$S^{N-1}=\mathrm{O}(N)/\mathrm{O}(N-1)$
and
$
\mathbb{C}P^{N-1}=
\mathrm{U}(N)/\mathrm{U}(N-1)\times \mathrm{U}(1)
$,
for which the full spectrum of one-loop anomalous dimensions
of order $s$ turns out to be not symmetric about zero,
in which case the substitution $t\to-t$ matters.
For example, high-gradient operators are made more irrelevant by one-loop
renormalization effects
in the non-compact NL$\sigma$M on
$
\mathrm{U}(N-1,1)/\mathrm{U}(N-1)\times \mathrm{U}(1)
$.
(We refer the reader to
Appendix~\ref{app: HGO on projective superspaces}
for a more detailed discussion of
``one-sided'' versus ``two-sided''
spectra of one-loop anomalous scaling dimensions
for high-gradient operators in NL$\sigma$Ms.)
Of course, one can only conclude that high-gradient operators
become relevant for sufficiently large values of $s$,
if the strong relevance seen in the one-loop expressions for their
scaling dimensions persists when all higher loop contributions
(not computed here or in other works on this subject)
have been taking into account.
For example, the one-loop expressions may not be
characteristic in the large-$s$ limit, if the actual
expansion parameter is not $\epsilon$ but $s\epsilon$.%
\cite{%
Ludwig90,%
Brezin97%
}
As any insight for resolving the nature of the $\epsilon$
expansion for high-gradient operators in NL$\sigma$Ms
must come from outside the $\epsilon$ expansion itself,
progress has stalled since the early 1990's.
The aim of this paper is to study
the operators that play the role of the high-gradient operators
in field theories which are two-dimensional Wess-Zumino-Witten
(WZW) theories~\cite{Wess71}$^{-}$\cite{Bocquet00}
on a Lie group $G$,
perturbed by
an interaction quadratic in the Noether currents
(``current-current interaction'').
Such theories are often also referred to as
``two-dimensional non-Abelian Thirring
(or Gross-Neveu) models''.
Any WZW theory,
which is a Principal-Chiral-Non-Linear-sigma model
supplemented by a WZW term at level $k$,
gives a prescription to construct high-gradient operators
in terms of powers of Noether currents.
Because it is possible to represent the Noether currents in WZW theories
in terms of free fermions,\cite{Witten84}
one might be inclined to think
that such operators are perhaps not capable of
displaying a ``pathological''
spectrum of scaling dimension
as in Eq.%
~(\ref{eq: anomalous scaling dimensions for O(N) high-gradient operators}).
However, as we demonstrate in this paper,
the situation is more interesting. Indeed, we will see
that under conditions specified below,
the one-loop spectra of the form
(\ref{eq: anomalous dimensions for O(N) high-gradient operators})
and
(\ref{eq: anomalous scaling dimensions for O(N) high-gradient operators})
can be realized by perturbing a WZW critical point by
a current-current perturbation.
We also want to investigate if there is a difference
between the properties of high-gradient operators
in unitary and non-unitary non-Abelian Thirring models.
This is important because
NL$\sigma$Ms describing the physics of
Anderson localization are non-unitary field theories.
Moreover, high-gradient operators in these theories
have been previously related to the statistical fluctuations
of the conductance of a disordered metal.%
~\cite{Altshuler86a}$^{-}$\cite{Altshuler91}
In this context, an appealing physical interpretation of the spectra
(\ref{eq: anomalous dimensions for O(N) high-gradient operators})
and
(\ref{eq: anomalous scaling dimensions for O(N) high-gradient operators})
has been proposed, attributing them to a broad tail
in the probability distribution of the conductance.%
~\cite{Altshuler86a}$^{-}$\cite{Altshuler91}
However, given that this interpretation
depends crucially on the ability to invert
the $s\to\infty$ and $\epsilon\to0$
limits in spectra which are analogous to those in Eqs
~(\ref{eq: anomalous dimensions for O(N) high-gradient operators})
and
~(\ref{eq: anomalous scaling dimensions for O(N) high-gradient
operators}),
it would be useful to have
an example of a critical field theory
describing a problem of Anderson localization
for which one can study the RG-relevance of high-gradient operators
without resorting to the $\epsilon$-expansion,
and for which one can reasonably expect a broad distribution
of the conductance.
We now provide an outline of the article and a summary of our results.
It is shown in Sec.\ \ref{sec: su2} that high-gradient operators
in the (unitary) $\widehat{\mathrm{su}}(2)^{\ }_{k}$ Thirring model
with strength $g$ of the ``current-current interaction'',
are made more irrelevant by the presence of these interactions
when the latter are (marginally) irrelevant
($g<0$, in our conventions).
On the other hand,
along the renormalization group (RG) flow driven
by a (marginally) relevant current-current interaction
($g>0$, in our conventions),
a one-loop spectrum of the form%
~(\ref{eq: anomalous dimensions for O(N) high-gradient operators})
is recovered
in the ``classical'' limit $1/k\to0$.
The inverse level $1/k$ plays here
the role of a ``quantum'' parameter.
Indeed, for any finite $k$, we find that
the quadratic growth in $s$ in the
unbounded one-loop spectrum
~(\ref{eq: anomalous dimensions for O(N) high-gradient operators})
does not persist for values of $s$ larger than $k$.
In effect, $1/k$ determines the efficiency in ``taming'' the
strong RG-relevance of high-gradient operators
seen at one-loop order,
which is related to the fact that there exists a representation
of the current algebra of the level-$k$
WZW theory in terms of free fermions.
Section~\ref{sec: gl(M|N) level k=1}
is devoted to high-gradient operators in
what we will call the $\widehat{\mathrm{gl}}(M|M)^{\ }_{k}$ Thirring
(or Gross-Neveu) model which was discussed in Ref.~\onlinecite{Guruswamy00}.
This is the $\widehat{\mathrm{gl}}(M|M)^{\ }_{k}$
WZW theory
on the Lie Supergroup $\mathrm{GL}(M|M)$,
perturbed by
\textit{two} current-current perturbations,
one which we call $g^{\ }_{\mathrm{M}}$, which
is exactly marginal, and
another
which we call $g^{\ }_{\mathrm{A}}$,
which flows logarithmically under the RG at a rate dependent on
$g^{\ }_{\mathrm{M}}$.
In spite of the presence of an RG
flow of the coupling $g^{\ }_{\mathrm{A}}$
there exists a sector of the theory, the
so-called $\mathrm{PSL}(M|M)$ sector,
which is scale (conformally) invariant throughout.%
~\cite{Guruswamy00}
The high-gradient operators turn out to reside
in this conformally invariant sector,
and are unaffected by the presence of the coupling $g^{\ }_{\mathrm{A}}$.
We will show that, for $k=1$, the spectrum of one-loop anomalous
dimensions of high-gradient operators is fundamentally different
for positive and negative
values of the coupling constant
$g^{\ }_{\mathrm{M}}$.
In particular, when $g^{\ }_{\mathrm{M}}>0$
all high-gradient operators are made more irrelevant by the
current-current perturbations,
whereas they are made more relevant when
$g^{\ }_{\mathrm{M}}<0$.
We close Sec.~\ref{sec: gl(M|N) level k=1}
by comparing the anomalous scaling dimensions
of high-gradient operators in
the $\widehat{\mathrm{gl}}(M|M)^{\ }_{k}$ Thirring
(or Gross-Neveu) models and those in the
$\mathrm{GL}(2N|2N)/\mathrm{OSp}(2N|2N)$
NL$\sigma$Ms, observing that they behave in the same way.
The result that the spectrum of one-loop anomalous dimensions of
high-gradient operators is strongly dependent on
the sign of $g^{\ }_{\mathrm{M}}$
has important implications
in the context of Anderson localization because, as discussed in Ref
~\onlinecite{Guruswamy00},
the $\widehat{\mathrm{gl}}(M|M)^{\ }_{k}$
Thirring (or Gross-Neveu) model at $k=1$
describes a disordered electronic system, where
$g^{\ }_{\mathrm{A}} >0$ and $g^{\ }_{\mathrm{M}}>0$
correspond to the strengths of disorder potentials.
The theory with $g^{\ }_{\mathrm{M}}>0$ thus offers
an example of a critical theory for Anderson localization with no
relevant high-gradient operator.
-- For example, this field theory describes~\cite{Guruswamy00}
a tight-binding model of electrons on the honeycomb lattice
with (real-valued) random hopping matrix elements
which are non-vanishing only between the two sublattices of the
bipartite honeycomb lattice (see also Ref.~\onlinecite{Foster06}).
Versions of the honeycomb tight-binding model provide the basic electronic
structure of graphene.
In the classification scheme of Zirnbauer, and Altland and Zirnbauer,%
~\cite{Verbaarschot94}$^{-}$\cite{Heinzner05}
this model belongs to
the ``chiral-orthogonal'' symmetry class (class BDI).
(Another example of a problem of Anderson localization
in the same symmetry class
is provided by a random tight-binding model on a square lattice
with $\pi$-flux through every plaquette.\cite{Hatsugai97})
By contrast, when $g^{\ }_{\mathrm{M}}<0$,
the spectrum of one-loop scaling dimensions is unbounded
from below for any $k$ as is the case in Eq.%
~(\ref{eq: anomalous scaling dimensions for O(N) high-gradient operators}).
The full spectrum of one-loop anomalous scaling dimensions of
high gradient operators as it appears, e.g.,
in the Grassmanian NL$\sigma$Ms with target manifolds
$\mathrm{Sp}(M+N)/\mathrm{Sp}(M)\times \mathrm{Sp}(N)$,
$\mathrm{U}(M+N)/\mathrm{U}(M)\times \mathrm{U}(N)$,
$\mathrm{O}(M+N)/\mathrm{O}(M)\times \mathrm{O}(N)$,
which is symmetric about zero,
is only recovered in the extreme ``classical'' limit
$M,k\to\infty$. In the context of Anderson localization,
the case with $g^{\ }_{\mathrm{M}}<0$
describes the surface state of a
three-dimensional topological insulator
in the chiral-symplectic class (symmetry class CII)
of Anderson localization.%
~\cite{Schnyder08,Ryu09,Hosur09}
After concluding in Sec.~\ref{sec: conclusion},
we review in Appendix~\ref{sec: HWK model}
the realization of the
$\widehat{\mathrm{gl}}(2N|2N)^{\ }_{k=1}$
Thirring (or Gross-Neveu) model as a problem
of Anderson localization in two dimensions
in symmetry class BDI,
which was established in
Ref.~\onlinecite{Guruswamy00}.
\section{
High-gradient operators and
$\widehat{\mathrm{su}}(2)^{\ }_{k}$ WZW theories
}
\label{sec: su2}
The O(3)/O(2) NL$\sigma$M with coupling constant $t$
is the simplest example of a NL$\sigma$M
containing \textit{infinitely many} high-gradient operators
all of which would appear to become
relevant based on one-loop results.
This happens at the infra-red unstable
fixed point $t^{*}=\epsilon$
in $d=2+\epsilon >2 $ dimensions within the one-loop approximation
as long as the order $s$ of these high-gradient operators is
large enough. A precursor to this perturbative property
also occurs in $d=2$ dimensions close to the infra-red
unstable fixed point $t=0$
as the NL$\sigma$M flows to strong coupling. Along this flow,
the spectrum of one-loop dimensions%
~\cite{footnote: on anomalous dimensions}
for the high-gradient operators
is unbounded from below.
In two dimensions,
the O(3)/O(2) NL$\sigma$M, supplemented by a topological
theta-term at $\theta=\pi$,
flows to a critical field theory,
the SU(2) WZW theory with $\widehat{\mathrm{su}}(2)^{\ }_{k=1}$
current algebra,
$\widehat{\mathrm{su}}(2)^{\ }_{k=1}$
WZW theory.%
~\cite{Affleck87}$^{-}$\cite{Zamolodchikov92}
The strongly relevant high-gradient operators
near the infra-red unstable fixed point $t=0$
must become
irrelevant at the WZW critical point,
because the full operator content of the
$\widehat{\mathrm{su}}(2)^{\ }_{k=1}$ WZW theory
is known to contain only a finite number
of relevant
fields (with scaling dimensions bounded from below
and above by zero and two, respectively).
The purpose of this section is to perturb the
$\widehat{\mathrm{su}}(2)^{\ }_{k}$ WZW theory
with a current-current perturbation
and to examine the fate of those operators in the
$\widehat{\mathrm{su}}(2)^{\ }_{k}$ WZW theory
which correspond to the
high-gradient operators in the O(3)/O(2) NL$\sigma$M.
We will refer to these operators still as ``high-gradient operators''.
We are going to argue that the spectrum of
one-loop scaling dimensions
associated with
all high-gradient operators is bounded from below by the lowest
one-loop scaling dimension corresponding to
high-gradient operators of order $k$.
This result is very different from the unbounded
spectrum~(\ref{eq: anomalous dimensions for O(N) high-gradient operators})
of one-loop scaling dimensions associated with
high-gradient operators in the two-dimensional
O(3)/O(2) NL$\sigma$M.
In the following sections, we first review the
$\widehat{\mathrm{su}}(2)^{\ }_{k}$ WZW theory
perturbed by a current-current interaction.
Second, we identify high-gradient operators of order $s$.
Finally, we compute the leading one-loop dimensions of
high-gradient operators of order $s$ up to one loop.
\subsection{
Definitions
}
The most fundamental property
of the $\widehat{\mathrm{su}}(2)^{\ }_{k}$ WZW theory
is the existence of
a pair of holomorphic and antiholomorphic SU(2) Noether
currents,
$J ^{\ }_1$,
$J^{\ }_2$,
$J^{\ }_3$,
and
$\bar J^{\ }_1$,
$\bar J^{\ }_2$,
$\bar J^{\ }_3$,
respectively,
which satisfy the affine (Kac-Moody) current algebra
\begin{subequations}
\label{eq: su(2) level k current algebra}
\begin{equation}
\begin{split}
&
J^{\ }_{\alpha}(z) J^{\ }_{\beta}(0)=
\frac{k C^{\ }_{\alpha\beta} }{z^{2}}
+
\frac{{i}}{z}
f^{\ }_{\alpha\beta}{}^{\gamma}
J^{\ }_{\gamma}(0)
+\cdots,
\\
&
\bar J^{\ }_{\alpha}(\bar z) \bar J^{\ }_{\beta}(0)=
\frac{kC^{\ }_{\alpha\beta}}{\bar z^{2}}
+
\frac{{i}}{\bar z}
f^{\ }_{\alpha\beta}{}^{\gamma}
\bar J^{\ }_{\gamma}(0)
+\cdots,
\\
&
J^{\ }_{\alpha}(z)\bar J^{\ }_{\beta}(0)=
0,
\end{split}
\label{eq: su(2) level k current algebra a}
\end{equation}
at level $k=1,2,3,\cdots,$
where the invariant (Casimir) tensor
of rank 2 in $\mathrm{su}(2)$
has the contravariant and covariant representations
(in our conventions)
\begin{equation}
C^{\ }_{\alpha\beta}=
\frac{1}{2}\delta^{\ }_{\alpha\beta},
\qquad
C^{\alpha\beta}=
2\delta^{\ }_{\alpha\beta},
\label{eq: invariant Casimir}
\end{equation}
respectively, while the structure constant of su(2)
is
the fully antisymmetric Levi-Civita tensor of rank 3,
\begin{eqnarray}
f^{\ }_{\alpha\beta}{}^{\gamma}=
\epsilon^{\ }_{\alpha\beta\gamma},
\qquad
\alpha,\beta,\gamma=1,2,3.
\label{eq: structure constants su(2)}
\end{eqnarray}
\end{subequations}
The dots in Eq.~(\ref{eq: su(2) level k current algebra a})
are terms of order zero and higher in powers of $z$ ($\bar z$)
with $z=x+{i}y$
($\bar z=x-{i}y$)
the holomorphic (antiholomorphic) coordinates
of the Euclidean plane. We shall also refer to the (anti) holomorphic sector
of the theory as the (right-) left-moving sector.
The $\widehat{\mathrm{su}}(2)^{\ }_{k}$ current algebra%
~(\ref{eq: su(2) level k current algebra})
has a representation
in terms of free-fermions.
More precisely,
it is obtained from the action%
~\cite{footnote: Einstein convention}
\begin{subequations}
\label{eq: def free fermion rep of su(2) level k}
\begin{equation}
S^{\ }_{*}:=
\sum_{\iota=1}^{k}
\int \frac{d\bar zd z}{2\pi{i}}
\left(
\psi^{c\dag}_{\ \iota}
\bar \partial\,
\psi^{\ }_{c\iota}
+
\bar \psi^{c\dag}_{\ \iota}
\partial\,
\bar \psi^{\ }_{c\iota}
\right)
\label{eq: def S*}
\end{equation}
constructed from $k$-independent flavors of
left ($\psi$) and right ($\bar\psi$) moving Dirac fermions,
whereby each one transforms
in the fundamental representation of SU(2)$\times$SU($k$),
with the partition function
\begin{equation}
Z^{\ }_{*}:=
\int\mathcal{D}[\psi^{\dag},\psi,\bar\psi^{\dag},\bar\psi]\,
\exp\left(-S^{\ }_{*}\right).
\label{eq: def Z*}
\end{equation}
\end{subequations}
One has the operator product expansions (OPE)
\begin{equation}
\begin{split}
&
\psi^{\ }_{c\iota}(z)
\psi^{d\dag}_{\ \iota'}(0)=
\psi^{d\dag}_{\ \iota'}(z)
\psi^{\ }_{c\iota}(0)
\sim
\frac{\delta^{\ }_{\iota\iota'}\delta^{\ }_{cd}}{ z},
\\
&
\bar\psi^{\ }_{c\iota}(z)
\bar\psi^{d\dag}_{\ \iota'}(0)=
\bar\psi^{d\dag}_{\ \iota'}(z)
\bar\psi^{\ }_{c\iota}(0)
\sim
\frac{\delta^{\ }_{\iota\iota'}\delta^{\ }_{cd}}{\bar z},
\\
&
\psi^{\ }_{c\iota}(z)
\bar \psi^{d\dag}_{\,\iota'}(0)
\sim
0,
\\
\end{split}
\label{eq: OPE free left and right movers}
\end{equation}
for $\iota,\iota'=1,\cdots,k$ and $c,d=1,2$.
In turn, the OPE~(\ref{eq: OPE free left and right movers})
imply that the left and right Noether currents
\begin{equation}
\begin{split}
J^{\ }_{\alpha}:=
\sum_{\iota=1}^{k}
\psi^{c\dag}_{\ \iota}
\frac{(\sigma^{\ }_{\alpha})^{\ }_{c}{}^{d}}{2}
\psi^{\ }_{d\iota},
\quad
\bar J^{\ }_{\alpha}:=
\sum_{\iota=1}^{k}
\bar \psi^{c\dag}_{\ \iota}
\frac{(\sigma^{\ }_{\alpha})^{\ }_{c}{}^{d}}{2}
\bar\psi^{\ }_{d\iota},
\end{split}
\label{eq: fermionic oscillator rep of su(2)}
\end{equation}
with $\alpha=1,2,3$
obey the $\mathrm{SU}(2)^{\ }_{k}$
current algebra~(\ref{eq: su(2) level k current algebra}).
The field theory defined by
Eq.~(\ref{eq: def free fermion rep of su(2) level k})
is a free-fermion field theory.
The content of local operators is thus known.
It contains a finite number of fields
whose scaling dimensions are bounded between 0 and 2
and are thus relevant,
as it should be for a field theory defined on a Hilbert space with
a positive definite inner product
and with a spectrum bounded from below which is built on
the Dirac-Fermi sea, in short a unitary field theory.
Clearly, within the set of powers of the Noether currents
(\ref{eq: fermionic oscillator rep of su(2)}) there is thus
no room for an infinite family of relevant operators.
We perturb the free-fermion field theory
by a current-current interaction
$\mathcal{O}^{\ }_I$
of the $\mathrm{SU}(2)^{\ }_{k}$ currents
(\ref{eq: fermionic oscillator rep of su(2)}).
\begin{equation}
\begin{split}
&
Z:=
\int \mathcal{D}[\psi^{\dag},\psi,\bar\psi^{\dag},\bar\psi]\,
\exp\left(-S\right),
\\
&
S:=
S^{\ }_{*}
+
g
\int\frac{d\bar zdz}{2\pi{i}}
\mathcal{O}^{\ }_I(\bar z,z),
\\
&
\mathcal{O}^{\ }_I(\bar z,z):=
C^{\alpha\beta}J^{\ }_{\alpha}(z)\bar J^{\ }_{\beta}(\bar z)\equiv
2 J^{\ }_{\alpha}(z)\bar J^{\ }_{\alpha}(\bar z).
\end{split}
\label{eq: def current current pert to su(2)k}
\end{equation}
We take the coupling constant $g$ to be real.
The (unitary) field theory%
~(\ref{eq: def current current pert to su(2)k})
is often
referred to as a non-Abelian Thirring (Gross-Neveu) model.
Suitable non-unitary generalizations
of the field theory%
~(\ref{eq: def current current pert to su(2)k})
compute
(disorder average) moments of Green's functions
in a class of problems of Anderson localization in
$d=2$ dimensions that we will investigate later on in this paper.
The one-loop beta function,
\begin{eqnarray}
\beta^{\ }_{g}=
\frac{{d}g}{{d}l}=
4 g^{2},
\end{eqnarray}
encodes the change in the coupling constant caused by
the infinitesimal rescaling $\mathfrak{a}\to(1+{d}l)\mathfrak{a}$
of the short-distance cutoff $\mathfrak{a}$.
Thus, the current-current interaction
is (marginally) irrelevant (relevant)
for $g<0$ ($g>0$)
with the free-fermion fixed point at $g=0$.
The $\mathrm{SU}(2)^{\ }_{k}$ Noether currents
(\ref{eq: fermionic oscillator rep of su(2)})
are those appearing at
the non-trivial fixed point of the
Principal Chiral NL$\sigma$M on the SU(2) group manifold
with a Wess-Zumino term.%
~\cite{Novikov82}$^{-}$\cite{Polyakov83}
This has, the well-known (Euclidean) action
\begin{subequations}
\label{eq: SU(N) WZW action}
\begin{equation}
S=
\frac{k}{16\pi}
\int d^2 x\,
\mathrm{tr}
\left(
\partial^{\ }_{\mu} G^{-1}\partial^{\ }_{\mu} G
\right)
+
k\Gamma[G],
\end{equation}
where $G\in \mathrm{SU}(2)$
is a group element, and
the integral $\Gamma[G]$
over a three-dimensional ball
$B$ with coordinates $r^{\ }_{\mu}$ and
whose boundary $\partial B$
is $d=2$-dimensional Euclidean space,
\begin{equation}
\Gamma[g]:= \frac{1}{24\pi}
\int\limits_{B} d^{3}r\,
\epsilon^{\ }_{\mu\nu\lambda}
\mathrm{tr}\,
\left(
G^{-1}\partial^{\ }_{\mu}G\,
G^{-1}\partial^{\ }_{\nu}G\,
G^{-1}\partial^{\ }_{\lambda}G
\right)
\end{equation}
\end{subequations}
is the Wess-Zumino term.%
~\cite{Wess71}
The Noether currents which generate the
SU(2)${\ }_{\mathrm{left}}$$\,\times\,$SU(2)${\ }_{\mathrm{right}}$
symmetry at the critical point of the WZW theory
can be fully represented by the fermionic expressions in
Eq.~(\ref{eq: fermionic oscillator rep of su(2)}).
In the bosonic
(i.e., NL$\sigma$M)
representation, these currents are built out of
first-order derivatives of the bosonic fields,
\begin{equation}
J^{\ }_{\alpha}\propto
k\,
\mathrm{tr}
\left[
(\partial G)G^{-1}\sigma^{\ }_{\alpha}
\right],
\quad
\bar{J}^{\ }_{\alpha}\propto
k\,
\mathrm{tr}
\left[
G^{-1}(\bar\partial G)\sigma^{\ }_{\alpha}
\right].
\label{eq: bosonized currents}
\end{equation}
The relationship~(\ref{eq: bosonized currents})
suggests that composite operators
built out of monomials in the currents%
~(\ref{eq: fermionic oscillator rep of su(2)})
in the WZW theory are the counterparts of
the high-gradient operators in NL$\sigma$M. For this reason,
we shall still call the former family of composite operators
high-gradient operators.
The ``classical'' counterparts of the
high-gradient operators of order $s$
in the NL$\sigma$M are thus the homogeneous polynomials
\begin{eqnarray}
T^{ \alpha^{\ }_{1}\cdots \alpha^{\ }_{s}
\bar\alpha^{\ }_{1}\cdots\bar\alpha^{\ }_{s}}
J^{\ }_{\alpha^{\ }_{1}}
\cdots
J^{\ }_{\alpha^{\ }_{s}}
\bar J^{\ }_{\bar \alpha^{\ }_{1}}
\cdots
\bar J^{\ }_{\bar \alpha^{\ }_{s}}
\label{eq: classical hgo su(2)k}
\end{eqnarray}
of the left and right currents that are invariant under the diagonal
SU(2) symmetry group of the interacting theory.%
~\cite{footnote: invariant tensor}
The generating set of classical high-gradient operators of order $s$
is specified once all the linearly
independent rank $2s$ tensors
$
T^{\alpha\beta\cdots\gamma\delta\cdots}
$
in the adjoint representation of SU(2)
that are invariant under
SU(2) transformations
can be fully enumerated.
In turn, the most general
SU(2) invariant tensor of even rank
in the adjoint representation
is the product of the Casimir tensor of rank 2.%
~\cite{Dittner72}
The high-gradient operators in Eq.~(\ref{eq: classical hgo su(2)k})
are classical in the sense that quantum fluctuations encoded
through the Pauli principle (or, equivalently,
through the underlying Dirac-Fermi sea)
in the free-fermion representatio
~(\ref{eq: fermionic oscillator rep of su(2)})
of the current algebra, have not yet been accounted for.
To account for these quantum fluctuations,
one needs to introduce a point-splitting
procedure that allows for the proper normal ordering,
i.e., the correct subtraction of all short-distance singularitie
~\cite{DiFrancesco97}
\begin{equation}
\begin{split}
&
:
T^{ \alpha^{\ }_{1}\cdots\bar\alpha^{\ }_{s}}
J^{\ }_{ \alpha^{\ }_{1}}
\cdots
\bar J^{\ }_{\bar\alpha^{\ }_{s}}
:(z,\bar z)\equiv
\\
&
\qquad
\lim_{z^{\ }_{i}\to z}
\Big[
T^{ \alpha^{\ }_{1}\cdots\bar\alpha^{\ }_{s}}
J^{\ }_{ \alpha^{\ }_{1}}( z^{\ }_{1})
\cdots
\bar J^{\ }_{\bar\alpha^{\ }_{s}}(\bar z^{\ }_{s})
\\
&
\qquad\qquad\quad
-\mbox{
(all short-distance singularities)
}
\Big].
\end{split}
\label{eq: quantum hgo su(2)k}
\end{equation}
Two objects of the form~(\ref{eq: classical hgo su(2)k})
that are linearly independent classically
might not survive as a pair of distinct
quantum operators of the form%
~(\ref{eq: quantum hgo su(2)k})
after normal ordering has been implemented.
More precisely, one might anticipate that
the underlying free-fermion representation
of the current algebra must manifest itself
as soon as the order $s$
becomes larger than the number $k$ of fermionic flavors
by changing the book-keeping relating classical expressions labeled by
SU(2) tensors of rank $2s$ and quantum operators.
Indeed, we are going to show that this is the mechanism that prevents
high-gradient operators of order $s>k$
from acquiring one-loop scaling dimensions smaller
than the smallest one-loop scaling dimensions associated with the set of
all high-gradient operators of order $s\leq k$.
In other words, the smallest
one-loop dimension associated with the set of \textit{all}
high-gradient operators is reached within the set of all
high-gradient operators of order $s\leq k$ when $g>0$.
It is thus bounded from below when $g>0$.
Had we ignored the underlying free-fermion representation
of the current algebra altogether,
we would have wrongly predicted that, when $g>0$,
the one-loop dimensions associated with the
classical objects~(\ref{eq: classical hgo su(2)k})
are of a form similar to the ones in Eq.%
~(\ref{eq: anomalous scaling dimensions for O(N) high-gradient operators})
i.e.,
that the set of one-loop dimensions of high-gradient operators
is unbounded from below. On the other hand,
this classical prediction is recovered
in the limit $k\to\infty$ with $s/k\to0$.
For this reason we shall separate the computation of
the most relevant one-loop dimension
associated with high-gradient operators
of order $s$ into the case when $s\leq k$ and the case when $k< s$.
In this context, we would like to remind the reader that
the $\widehat{\mathrm{su}}(2)^{\ }_{k}$ WZW theories are known to
describe quantum critical points in the parameter space of
quantum spin-$S$ antiferromagnetic chains when $k=2S$.
Here, we observe that both,
the number of relevant perturbations
and the number of independent local composite operators
built out of the generators of SU(2)
which are SU(2) singlets, grows with $S$.
[For $S=1/2$ the algebra obeyed by the Pauli
matrices only allows one invariant SU(2) tensor of rank 2,
the $2\times2$ unit matrix.]
On the other hand,
the strength of quantum fluctuations in
$\widehat{\mathrm{su}}(2)^{\ }_{k}$ WZW theories
decreases with increasing $k=2S$
for the same reason as the role of
quantum fluctuations decreases with increasing
$S$ in quantum spin chains.
\subsection{
Anomalous dimensions of high-gradient operators
}
\label{subsec: High-gradient operators when 1<s<k}
As the most general
SU(2) invariant tensor of even rank
in the adjoint representation
is the product of the Casimir tensor of rank 2,\cite{Dittner72}
we define the three diagonal SU(2) invariants
out of the three current bilinears
\begin{subequations}
\begin{equation}
\begin{split}
H
:=
C^{\alpha\beta}
J^{\ }_{\alpha}
\bar J^{\ }_{\beta},
\quad
A
:=
C^{\alpha\beta}
J^{\ }_{\alpha}
J^{\ }_{\beta},
\quad
B
:=
C^{\alpha\beta}
\bar J^{\ }_{\alpha}
\bar J^{\ }_{\beta},
\end{split}
\end{equation}
together with the
SU(2) invariant
\begin{equation}
C:=AB.
\end{equation}
\end{subequations}
The space of the high-gradient operators
is then spanned by the family
\begin{eqnarray}
\left\{
H^s,
H^{s-2}C,\cdots,
H^{2} C^{[s/2]-1},
C^{[s/2]}
\right\}
\label{eq: def high gradiant order s if s<k}
\end{eqnarray}
made of $[s/2]+1$ ``classical'' operators.%
~\cite{footnote: choice HGO for s<k}
We call these operators ``classical'' because we have
not yet taken into account the short distance singularities
associated with the
definition of composite operators
(i.e., the ``Pauli principle'' discussed above).
As announced below Eq.~(\ref{eq: quantum hgo su(2)k}),
these singularities need to be subtracted
from the ``classical'' expressions%
~(\ref{eq: def high gradiant order s if s<k})
upon normal ordering.
We shall nevertheless ignore the issue of
normal ordering at first and compute the one-loop RG equation
for these ``un-regularized'' (or un-normal-ordered) operators,
a step of no consequence in the (``classical'') limit $s/k\to0$.
We shall then contrast this un-regularized calculation with
the full quantum calculation for the special case of $k=1$,
i.e., when the proper normal ordering procedure has been accounted for.
We shall see that the calculation without normal ordering
gives an infinite tower of high-gradient operators
that are all relevant to one-loop order
for sufficiently large $s$ and for $g>0$.
The one-loop spectrum of anomalous dimensions is identical to
the one in the $\mathrm{O}(N)/\mathrm{O}(N-1)$ NL$\sigma$M
when $N=3$.
Indeed, once the normal ordering procedure is ignored,
the high-gradient operators
(\ref{eq: def high gradiant order s if s<k})
are analogous to those discussed in Ref.\ \onlinecite{Castilla97},
and the calculations of the anomalous dimensions
in the $\hat{\mathrm{su}}(2)^{\ }_{k}$ WZW model
and
in the $\mathrm{O}(N)/\mathrm{O}(N-1)$ NL$\sigma$M
run along parallel tracks.
The effect of normal ordering is weaker the smaller $s/k$ is, i.e.,
the closer proximity to the semi-classical limit
of the WZW theory.
To see this, consider the case when $k>s$.
The ``classical'' expression
$
T^{ \alpha^{\ }_{1}\cdots\bar\alpha^{\ }_{s}}
J^{\ }_{ \alpha^{\ }_{1}}
\cdots
\bar J^{\ }_{\bar\alpha^{\ }_{s}}
(z,\bar z)$
for the composite operator made of a local product of
holomorphic and antiholomorphic currents
is modified upon normal ordering.
To leading order in a short-distance expansion,
this classical expression is replaced by
\begin{eqnarray}
&& \!\!
\sum_{
\iota^{\ }_{1}\neq\cdots\neq \iota^{\ }_{s}\neq
\bar\iota^{\ }_{1}\neq\cdots\neq\bar\iota^{\ }_{s}=1
}^{k}
T^{ \alpha^{\ }_{1}\cdots \alpha^{\ }_{s}
\bar\alpha^{\ }_{1}\cdots\bar\alpha^{\ }_{s}}
:\!
J^{\ }_{\alpha^{\ }_{1}\iota^{\ }_{1}}
\cdots
J^{\ }_{\alpha^{\ }_{s}\iota^{\ }_{s}}
\!:\!(z)
\nonumber\\
&&\hspace*{27mm}{}\times
:\!
\bar J^{\ }_{\bar\alpha^{\ }_{1}\bar\iota^{\ }_{1}}
\cdots
\bar J^{\ }_{\bar\alpha^{\ }_{s}\bar\iota^{\ }_{s}} \!
:\!(\bar{z})
+\cdots.
\end{eqnarray}
Here, the terms included in the $\cdots$ arise
from the OPE for the product
$J^{\ }_{\alpha^{\ }_{i}\iota^{\ }_{i}}(z)
J^{\ }_{\alpha^{\ }_{j}\iota^{\ }_{j}}(0)$
when any two flavor indices $\iota^{\ }_{i}$ and $\iota^{\ }_{j}$
are identical. Evidently, normal ordering (or the Pauli principle)
has a much more potent effect when $k<s$,
for the condition
$\iota^{\ }_{1}\neq\cdots\neq \iota^{\ }_{s}\neq
\bar\iota^{\ }_{1}\neq\cdots\neq\bar\iota^{\ }_{s}$
can then never be met so that the
leading order term above is absent.
The operator contents with and without normal ordering
thus look very different.
When $k<s$, some operators in
the set (\ref{eq: def high gradiant order s if s<k})
completely disappear to leading order because of
Fermi statistics. This will be demonstrated explicitly for the case of $k=1$
[see Eqs.\ (\ref{eq: normal ordered JJ})
and (\ref{eq: normal ordered JJ second})
below],
for which we will show, after correctly taking into account
normal ordering, that all high-gradient operators
which would be relevant classically (when $g > 0$)
disappear from the operator content.
\subsubsection{RG equation for un-regularized high-gradient operators}
To compute the \textit{leading} one-loop scaling dimensions
for the high-gradient operators%
~(\ref{eq: def high gradiant order s if s<k}),
we start from the field theory%
~(\ref{eq: def current current pert to su(2)k})
in which we substitute the action by
\begin{equation}
\begin{split}
S:=& \,
S^{\ }_{*}
+
g
\int\frac{d\bar zdz}{2\pi{i}}\,
\mathcal{O}^{\ }_{I}
\\
&
-
\sum_{
m,n=0
}^{2m+n=s}
Z^{(s)}_{m,n}
\mathfrak{a}^{2s-2}
\int\frac{d\bar zdz}{2\pi{i}}\,
C^{m}H^n.
\end{split}
\label{eq: def pert action by g and Z's}
\end{equation}
To determine the one-loop dimensions
of the couplings
$\{Z^{(s)}_{m,n}|2m+n=s\}$,
we do not need the full one-loop RG flows,
i.e., the RG equations for
the coupling constants
up to and including order $Z^{(s)}_{m,n}Z^{(s)}_{p,q}$,
but
only the linear in $Z^{(s)}_{m,n}$
contributions to the one-loop RG flows.
Thus, all we need are the OPE of
$
C^{m}H^{n}(z,\bar{z})$
with
$\mathcal{O}^{\ }_I(0)
$,
where the
integers $m$ and $n$ satisfy $1<2m+n=s\leq k$.
Furthermore, we shall introduce the short-hand notation
\begin{equation}
\mathcal{A}
\times
\mathcal{B}=
\mathcal{C}
\Longleftrightarrow
\mathcal{A}(z,\bar z)
\mathcal{B}(0)=
\frac{1}{z\bar z}
\mathcal{C}(z,\bar z)
+
\cdots
\label{eq: def short hand OPE}
\end{equation}
for the OPE relating the operators
$\mathcal{A}$,
$\mathcal{B}$,
and
$\mathcal{C}$.
Here,
the dots are meant to contain not only regular terms of
zeroth and higher order
in $z$ or $\bar z$ but also second and higher order poles in $z$ or $\bar z$.
As an intermediary step, one verifies that the
OPE (\ref{eq: def short hand OPE})
between the building blocks $H$ and $C$
with $\mathcal{O}^{\ }_{I}$
(observe that $\mathcal{O}^{\ }_{I}=H$)
are
\begin{eqnarray}
\Wick{7mm}
\Wickunder{7mm}H
{\times}
\mathcal{O}^{\ }_{I}
&=&
-4
H,
\qquad
\Wick{7mm}
\Wickunder{7mm}
C{\times}
\mathcal{O}^{\ }_{I}=
0.
\label{eq: Step 1}
\end{eqnarray}
Here, we introduced yet another short-hand notation
$\Wick{8mm}\mathcal{A} \cdots \mathcal{B}$
or
$\Wickunder{8mm}\mathcal{A} \cdots \mathcal{B}$,
by which we mean that one current in $\mathcal{A}$
and one current in $\mathcal{B}$ are contracted
with the rule
\begin{equation}
\begin{split}
&
J^{\ }_{\alpha\iota}(z)
J^{\ }_{\beta \iota'}(0)=
\delta^{\ }_{\iota\iota'}\!\!
\left(
\frac{C^{\ }_{\alpha\beta}}{z^{2}}
\!+\!
\frac{{i}}{z}
f^{\ }_{\alpha\beta}{}^{\gamma}
J^{\ }_{\gamma\iota}(0)
\!+\!
\cdots
\right),
\\
&
\bar J^{\ }_{\alpha\iota }(\bar z)
\bar J^{\ }_{\beta \iota'}( 0)=
\delta^{\ }_{\iota\iota'}\!\!
\left(
\frac{C^{\ }_{\alpha\beta} }{\bar z^{2}}
\!+\!
\frac{{i}}{\bar z}
f^{\ }_{\alpha\beta}{}^{\gamma}
\bar J^{\ }_{\gamma\iota}(0)
\!+\!
\cdots
\right),
\\
&
J^{\ }_{\alpha\iota }(z)
\bar J^{\ }_{\beta \iota'}(0)=
0,
\end{split}
\label{eq: OPE for the currents for each species}
\end{equation}
for any
$
\alpha,\beta=1,2,3
$
and
$
\iota,\iota'=1,\cdots,k
$
at the free-fermion fixed point $g=0$, where
\begin{equation}
\begin{split}
J^{\ }_{\alpha\iota}:=
\psi^{a \dag}_{\ \iota}
\frac{\left(\sigma^{\ }_{\alpha}\right)^{\ }_{a}{}^{b}}{2}
\psi^{\ }_{b \iota},
\quad
\bar J^{\ }_{\alpha\iota}:=
\bar \psi^{a \dag}_{\ \iota}
\frac{\left(\sigma^{\ }_{\alpha}\right)^{\ }_{a}{}^{b}}{2}
\bar \psi^{\ }_{b \iota}.
\end{split}
\label{eq: flavor currents}
\end{equation}
When $\mathcal{A}$ and $\mathcal{B}$ consist of
more than one $J^{\ }_{\alpha}$ or $\bar{J}^{\ }_{\alpha}$,
and when there are many possible Wick contractions
between $\mathcal{A}$ and $\mathcal{B}$,
the short-hand notations
$\Wick{8mm}\mathcal{A} \cdots \mathcal{B}$,
$\Wickunder{8mm}\mathcal{A} \cdots \mathcal{B}$
and
$\Wick{8mm}\Wickunder{8mm}\mathcal{A} \cdots \mathcal{B}$
mean the resulting operator
obtained by taking all possible such Wick contractions.
One also verifies that the
OPE (\ref{eq: def short hand OPE})
between the building blocks $HH$, $CH$, and $CC$
with $\mathcal{O}^{\ }_{I}$
are
\begin{equation}
\begin{split}
\Wick{10mm}H
\Wickunder{7mm}H
{\times}
\mathcal{O}^{\ }_I
&=
-4 H^{2}
+ 4 C,
\\
\Wick{10mm}
C
\Wickunder{7mm}
H
{\times}
\mathcal{O}^{\ }_I
&=
\Wick{10mm}
C
\Wickunder{7mm}
C
{\times}
\mathcal{O}^{\ }_I
=
0.
\end{split}
\label{eq: Step 2}
\end{equation}
We then infer that, for any pair $(m,n)$ of
positive integer
that satisfies $1<2m+n=s\leq k$,
\begin{equation}
\begin{split}
C^{m}H^{n}
{\times}
\mathcal{O}^{\ }_{I}
=&
\hphantom{+}
\Wick{7mm}\Wickunder{7mm}C
{\times}
\mathcal{O}^{\ }_I
\times
m C^{m-1}H^{n}
\\
&
+
\Wick{7mm}\Wickunder{7mm}H
{\times}
\mathcal{O}^{\ }_I
\times
n C^{m}H^{n-1}
\\
&
+
\Wick{10mm}C\Wickunder{7mm} H
{\times}
\mathcal{O}^{\ }_{I}
\times
mn C^{m-1}H^{n-1}
\\
&
+
\Wick{10mm}C\Wickunder{7mm} C
{\times}
\mathcal{O}^{\ }_{I}
\times
\frac{m(m-1)}{2} C^{m-2}H^{n}
\\
&
+
\Wick{10mm}H\Wickunder{7mm} H
{\times}
\mathcal{O}^{\ }_{I}
\times
\frac{n(n-1)}{2} C^{m}H^{n-2}
\\
=&
-2
n\left(n+1\right)
C^{m} H^{n}
\\
&
+
2
n(n-1)
C^{m+1}H^{n-2}.
\end{split}
\label{eq: OPE's between H and CmHN}
\end{equation}
The contributions to the RG equations obeyed by the couplings
$Z^{(s)}_{m,n}$ where $1<2m+n=s\leq k$
needed to extract the spectrum of one-loop dimensions are
\begin{equation}
\begin{split}
\frac{{d}Z^{(s)}_{m,n}}{{d}l}=& \,
\big(
2
-
2s
\big)
Z^{(s)}_{m,n}
+
4gn(n+1)
Z^{(s)}_{m,n}
\\
&
-
4g
(n+2)(n+1)
Z^{(s)}_{m-1,n+2}
+
\cdots.
\end{split}
\label{eq: linearized RG flows for Z(m,n)}
\end{equation}
Here, the dots include non-linear contributions of
second order in $g$ or $Z^{(s)}_{m,n}$.
The linearized RG flows
(\ref{eq: linearized RG flows for Z(m,n)})
are closed. This is a justification a posteriori
for neglecting the RG effects of current monomials with
repeating flavor indices.
The linearized RG flows
(\ref{eq: linearized RG flows for Z(m,n)})
have a lower triangular structure,
i.e., there is no feedback effect on the flow of
a high-gradient operator of order $s$
from lower-order high-gradient operators.
Thus, we conclude that the leading
$[s/2]+1$ one-loop scaling dimensions associated
with the family of high-gradient operators%
~({\ref{eq: def high gradiant order s if s<k})
when $k\geq s=2m+n$ are given by
\begin{eqnarray}
x^{(s)}_{m,n}&=&
2(2m+n)
-
4 g n(n+1).
\end{eqnarray}
Observe that the spectrum of
anomalous dimensions%
~\cite{footnote: our conventions for dimensions}
\begin{equation}
\gamma^{(s)}_{m,n}:=
-
4 g n(n+1),
\qquad
2m+n=s
\label{eq: gamma(s)m,n is one sided for su(2) level k<s}
\end{equation}
is one sided with respect to 0.
When $g\leq0$ these anomalous dimensions are positive, i.e.,
the scaling dimensions are larger than their engineering value.
The opposite happens when $g\geq0$, i.e., when the current-current
perturbation is (marginally) relevant.
When $g>0$ and for a given $1<s\leq k$,
the smallest one-loop anomalous dimension
occurs for the pair $(m,n)=(0,s)$,
\begin{eqnarray}
\gamma^{(s)}_{\mathrm{min}}&:=&
\min_{2m+n=s}
\gamma^{(s)}_{m,n}
=
-
4gs(s+1).
\label{eq: max scaling dimension when g>0}
\end{eqnarray}
For $g>0$,
the quadratic dependence on $s$ can overcome the linear dependence on $s$
in the one-loop dimension
$x^{(s)}_{\mathrm{min}}:=2s+\gamma^{(s)}_{\mathrm{min}}$.
If the order $s$ ($1<s\leq k$) is allowed to be sufficiently large,
the one-loop dimension $x^{(s)}_{\mathrm{min}}$
decreases past the value 2 and eventually becomes negative.
The quadratic dependence on $s$ is reminiscent of that for the
one-loop dimensions%
~(\ref{eq: anomalous scaling dimensions for O(N) high-gradient operators})
in the $\mathrm{O}(N)/\mathrm{O}(N-1)$ NL$\sigma$M.
However, in contrast to the
$(2+\epsilon)$-dimensional $\mathrm{O}(N)/\mathrm{O}(N-1)$ NL$\sigma$M
at its non-trivial fixed point $t^{*}$,
a value smaller than 2 for the one-loop dimensions
$x^{(s)}_{\mathrm{min}}$
is not a threat to the internal stability of the WZW fixed point $g=0$
since it occurs along a flow to strong coupling.
Moreover, it is known that in $d=2$ dimensions the
$\mathrm{O}(3)/\mathrm{O}(2)$ NL$\sigma$M
with theta term at $\theta =\pi$ flows
in the infrared into
the level $k=1$ SU(2) WZW fixed point.
While the spectrum of one-loop dimensions
of high-gradient operators at the WZW fixed point is
bounded from below (as we will recall below),
the spectrum of these operators
is unbounded from below in the
weakly coupled $2$-dimensional
$\mathrm{O}(3)/\mathrm{O}(2)$ NL$\sigma$M
(the presence of the theta term does not affect this result).
\subsubsection{
Normal ordering revisited
}
\label{subsec: high-gradient operators when s>k }
We shall illustrate the effects of the Fermi statistics
for the family of high-gradient operators%
~({\ref{eq: def high gradiant order s if s<k})
when $s=2$ for the case of a (marginally) relevant ($g>0$)
current-current interaction.
We shall then show for the special case of $k=1$ and $s=2$
that the two one-loop dimensions associated with
the family of high-gradient operators%
~({\ref{eq: def high gradiant order s if s<k})
are unchanged, to one loop order, i.e.,
\begin{equation}
x^{(s)}_{0,2}=
x^{(s)}_{1,0}=4.
\end{equation}
We start from the family of high-gradient operators%
~({\ref{eq: def high gradiant order s if s<k})
with $s=2$. For clarity of presentation,
we rename the two members
of this family,
\begin{equation}
\mathcal{O}^{\ }_{1}\equiv
C^{\alpha\beta}J^{\ }_{\alpha}J^{\ }_{\beta}
C^{\gamma\delta}\bar J^{\ }_{\gamma}\bar J^{\ }_{\delta},
\quad
\mathcal{O}^{\ }_{2}\equiv
C^{\alpha\beta}J^{\ }_{\alpha}\bar J^{\ }_{\beta}
C^{\gamma\delta}J^{\ }_{\gamma}\bar J^{\ }_{\delta}.
\label{eq: classical O1 and O2 HGO}
\end{equation}
As implied by Eq.~(\ref{eq: quantum hgo su(2)k})
these are two classical expressions.
The two quantum expressions involve
point splitting and normal ordering
as in Eq.\ (\ref{eq: quantum hgo su(2)k}).
Without loss of generality, we consider only the left current sector.
Normal ordering of
\begin{equation}
\begin{split}
J^{\ }_{\alpha}(z)
J^{\ }_{\beta }(0)=&
\frac{k C^{\ }_{\alpha\beta}}{z^{2}}
+
\frac{{i}}{z}
\epsilon^{\ }_{\alpha\beta\gamma}J^{\ }_{\gamma}(0)
+
\frac{{i}}{2}
\epsilon^{\ }_{\alpha\beta\gamma}
\partial
J^{\ }_{\gamma}
(0)
\\
&+
\frac{\delta^{\ }_{\alpha\beta}}{4}
\sum_{\iota=1}^{k}
:
\left(
\psi^{a\dag}_{\ \iota}
\partial
\psi^{\ }_{a\iota}
-
\partial
\psi^{a\dag}_{\ \iota}
\psi^{\ }_{a\iota}
\right)
:
(0)
\\
&
+
\sum_{\iota,\iota'=1}^{k}
:
\psi^{a\dag}_{\ \iota}
\frac{(\sigma^{\ }_{\alpha})^{\ }_{a}{}^{b}}{2}
\psi^{\ }_{b\iota}
\psi^{c\dag}_{\ \iota'}
\frac{(\sigma^{\ }_{\beta})^{\ }_{c}{}^{d}}{2}
\psi^{\ }_{d\iota'}
:(0)
\\
&
+\cdots
\end{split}
\label{eq: OPE two currents made of fermions}
\end{equation}
amounts to the subtraction
from Eq.~(\ref{eq: OPE two currents made of fermions})
of the terms singular in the limit $z\to0$,
\begin{equation}
\begin{split}
\label{eq: normal ordered JJ}
&
:J^{\ }_{\alpha}J^{\ }_{\beta}:(0) =
\sum_{\iota\neq\iota'=1}^{k}
J^{\ }_{\alpha \iota}
J^{\ }_{\beta \iota'}
(0)
+
\frac{{i}}{2}
\epsilon^{\ }_{\alpha\beta\gamma}
\partial J^{\ }_{\gamma}(0)
\\
&
+
\frac{\delta_{\alpha\beta}}{4}\sum_{\iota=1}^{k}
:
\big(
\partial
\psi^{a\dag}_{\ \iota}
\psi^{\ }_{a \iota}
-
\psi^{a\dag}_{\ \iota}
\partial
\psi^{\ }_{a \iota}
-
\psi^{a\dag}_{\ \iota}
\psi^{\ }_{a \iota}
\psi^{b\dag}_{\ \iota}
\psi^{\ }_{b \iota}
\big)
:(0)
\end{split}
\end{equation}
for $\alpha,\beta=1,2,3$.
The proper quantum interpretation of the
classical currents~(\ref{eq: classical O1 and O2 HGO})
is then
\begin{equation}
\begin{split}
&
\label{eq: normal ordered JJ second}
:\mathcal{O}^{\ }_{1}:(z,\bar z)=
4
\sum_{\alpha,\beta=1}^{3}
: J^{\ }_{\alpha} J^{\ }_{\alpha}:(z)
:\bar J^{\ }_{\beta }\bar J^{\ }_{\beta }:(\bar z),
\\
&
:\mathcal{O}^{\ }_{2}:(z,\bar z)=
4\sum_{\alpha,\beta=1}^{3}
: J^{\ }_{\alpha} J^{\ }_{\beta}:(z)
:\bar J^{\ }_{\alpha}\bar J^{\ }_{\beta}:(\bar z).
\end{split}
\end{equation}
\subsubsection{
High-gradient operators when $k=1$
}
When $k=1$,
the summation over unequal flavors disappears
in Eq.~(\ref{eq: normal ordered JJ}).
(Observe in passing that
$:\!\mathcal{O}^{\ }_{1}\!:$
is then proportional to one component of the energy-momentum
stress tensor.)
One then verifies the OPE
\label{eq: correct OPE}
\begin{equation}
\begin{split}
&
:\!\mathcal{O}^{\ }_{1}\!:
\times \mathcal{O}^{\ }_{I}=
3:\!\mathcal{O}^{\ }_{1}\!:
- \,
9
:\!\mathcal{O}^{\ }_{2}\!: \,
,
\\
&
:\!\mathcal{O}^{\ }_{2}\!:
\times
\mathcal{O}^{\ }_{I}= \,
:\!\mathcal{O}^{\ }_{1}\!:
- \,
3:\!\mathcal{O}^{\ }_{2}\!:.
\end{split}
\end{equation}
If we diagonalize the linearized one-loop RG flows for
the coupling $Z^{(2)}_{1,0}$
associated with
$:\!\mathcal{O}^{\ }_{1}\!:$
and
the coupling $Z^{(2)}_{0,2}$
associated with
$:\!\mathcal{O}^{\ }_{2}\!:$,
we find that their one-loop dimensions
remain equal to their engineering dimensions,
\begin{equation}
x^{(2)}_{1,0}=
x^{(2)}_{0,2}=
4.
\end{equation}
The lesson that we draw from the example $s=2$ and $k=1$
is that it is necessary to use normal ordering
to properly define composite operators.
Had we not used normal ordering, we would have incorrectly
predicted that there are infinitely many high-gradient operators
which become relevant, at one-loop order, for large
enough $s$ and for $g>0$.
We believe that for a generic value of $k$,
there is no infinity of
one-loop relevant high-gradient operators.
Only a finite number of high-gradient operators become
relevant, at one-loop order,
for large enough $s$ and for $g>0$ when $k>1$.
In the next section, we turn attention
to a non-unitary WZW model of relevance to the problem
of Anderson localization to investigate whether
the loss of unitarity opens the door to an infinity
of relevant high-gradient operators.
\section{
High-gradient operators and $\widehat{\mathrm{gl}}(M|M)_{k}$ WZW theories
}
\label{sec: gl(M|N) level k=1}
An interesting example of a problem of Anderson localization
in two dimensions which possesses a special so-called sublattice
(or chiral) symmetry (SLS) and TRS (thus belonging to
the ``chiral-orthogonal'' symmetry class BDI
in the classification scheme of Zirnbauer, and Altland and
Zirnbauer%
~\cite{Zirnbauer96}$^{-}$\cite{Heinzner05})
is as follows.
Consider a tight-binding model of fermions on a honeycomb lattice
with random real-valued hopping matrix elements of non-vanishing mean,
which do not connect the same sublattice
(so that SLS is preserved).
[A related realization of the same problem of Anderson localization
is provided by a random tight-binding model on a square lattice with flux-$\pi$
through every plaquette, introduced in Ref.~\onlinecite{Hatsugai97}.]
In the absence of disorder this band structure is known to exhibit the
energy-momentum dispersion law of two species of (relativistic) Dirac
fermions at two points in the Brillouin zone at low energy
near the Fermi level (at zero energy).
It was shown in Ref. \onlinecite{Guruswamy00} that the
SLS-preserving disorder discussed above
leads to a theory for the disorder averages which, in the
supersymmetric formulation,\cite{Efetov97}
is a ${\mathrm{GL}}(2N|2N)$ Thirring (Gross-Neveu) model.
In other words, the problem of two-dimensional
Anderson localization
on the honeycomb lattice preserving SLS and TRS, is described by a set
of Dirac fermions (and SUSY boson partners) perturbed by a
current-current interaction of
the Noether currents
of its underlying ${\mathrm{GL}}(2N|2N)$ (super) symmetry.
The interaction strength corresponds to the strength of the disorder.
The system of free Dirac fermions (and SUSY boson partners)
is well known~\cite{Witten84,Bocquet00}
to be described by a WZW model on the supergroup
${\mathrm{GL}}(2N|2N)$
with $\widehat{\mathrm{gl}}(2N|2N)^{\ }_{k}$
conformal Kac-Moody current algebra symmetry at level $k=1$.
This section is devoted to the one-loop RG analysis of
high-gradient operators
in the perturbed $\widehat{\mathrm{gl}}(M|M)_{k}$ WZW theory.
The main result of this section and of this article
applies to the case of level $k=1$ of relevance
for the random tight-binding models discussed above.
In order to state this result,
we first need to
recall from Ref. \onlinecite{Guruswamy00} that
the ${\mathrm{GL}}(2N|2N)$ Thirring (Gross-Neveu) models
possess two coupling constants;
one, $g^{\ }_{\mathrm{M}}$,
which does not flow under the renormalization group (RG) and another,
$g^{\ }_{\mathrm{A}}$, which flows logarithmically under the RG
and a rate dependent on $g^{\ }_{\mathrm{M}}$.
Our main result then suggests that
all higher-order gradient operators are more irrelevant
in the presence of the current-current interaction
with $g^{\ }_{\mathrm{M}}>0$
than at zero coupling $g^{\ }_{\mathrm{M}}=0$.
A positive $g^{\ }_{\mathrm{M}}$
can be interpreted~\cite{Guruswamy00}
as the variance of the disorder strength
in the random tight-binding model in symmetry class BDI.
For the opposite sign of the coupling constant
$g^{\ }_{\mathrm{M}}<0$,
on the other hand,
higher-order gradient operators have a spectrum of one-loop dimensions
that is unbounded from below very much as in Eq.%
~(\ref{eq: anomalous scaling dimensions for O(N) high-gradient operators}).
In the context of Anderson localization,
the case with $g^{\ }_{\mathrm{M}}<0$
describes the surface state of a
three-dimensional topological insulator
in the chiral-symplectic class (symmetry class CII)
of Anderson localization.%
\cite{Schnyder08,Ryu09}
As in Sec.~\ref{sec: su2},
we are going to distinguish two limits.
In the first (classical) limit,
\begin{equation}
M\to\infty,
\qquad
k\to\infty,
\label{eq: extreme classical limit}
\end{equation}
OPEs between the high-gradient operators can be obtained without
any reference to the composite nature of the currents.
One then recovers a spectrum of
one-loop scaling dimensions
for high-gradient operators that mimics closely
that of the NL$\sigma$Ms discussed above.
The second limit,
\begin{equation}
M=1,2,3,\cdots,
\qquad
k=1,
\label{eq: extreme quantum limit}
\end{equation}
is the opposite extreme to the first one in that the
normal ordering of the currents and thus of the high-gradient operators
is essential and changes dramatically the spectrum of
one-loop scaling dimensions from the ``classical'' limit%
~(\ref{eq: extreme classical limit}).
\subsection{
Definitions
}
Our starting point is a two-dimensional conformal
field theory characterized by the current
algebra~\cite{Guruswamy00}
\begin{subequations}
\label{eq: def gl(M,N) current algebra of level k}
\begin{equation}
\begin{split}
J^{\,B}_{A}{ }(z)
J^{\,D}_{C}(0)=&
\frac{k\mathfrak{c}^{BD}_{AC}}{z^{2}}
+
\frac{1}{z}
\left[
\mathfrak{d}^{B}_{C}
J^{\,D}_{A}(0)
+
\mathfrak{e}^{BD}_{AC}
J^{\,B}_{C}(0)
\right]
\\
&
+
\cdots,
\\
\bar J^{\,B}_A(\bar z)
\bar J^{\,D}_C(0)=&
\frac{k\mathfrak{c}^{BD}_{AC}}{\bar z^{2}}
+
\frac{1}{\bar z}
\left[
\mathfrak{d}^{B}_{C}
\bar J^{\,D}_{A}(0)
+
\mathfrak{e}^{BD}_{AC}
\bar J^{\,B}_{C}(0)
\right]
\\
&
+
\cdots,
\\
J^{\,B}_A(z)
\bar J^{\,D}_C(0)=&
0,
\end{split}
\end{equation}
where
\begin{equation}
\begin{split}
\mathfrak{c}^{BD}_{AC}:=
(-)^{B+1}
\delta^{D}_{A}
\delta^{B}_{C},
\end{split}
\end{equation}
and
\begin{equation}
\mathfrak{d}^{B}_{C}=
-
(-)^{BC}\delta^{B}_{C},
\quad
\mathfrak{e}^{BD}_{AC}=
(-)^{BC+D(B+C)}\delta^{D}_{A},
\label{eq: structure constants gl(M|N)}
\end{equation}
\end{subequations}
with the indices $A,B,C,D=1,\cdots,M+N$,
where
$\delta^{B}_{C}$ denotes the Kronecker delta.
The capitalized indices
$A$,
$B$,
$C$,
and
$D$
also carry a grade which is either 0 for $M$ out of the $M+N$
values that they take or 1 for the remaining $N$ values.
It is the grade of the indices $A$ and $B$ that enters
expressions such as $(-)^{A}$ or $(-)^{AB}$.
The grade $0$ ($1$) will shortly be associated with
bosons (fermions).
The positive integer $k$ is the level of the current algebra%
~(\ref{eq: def gl(M,N) current algebra of level k}).
The current algebra%
~(\ref{eq: def gl(M,N) current algebra of level k})
is associated with the Lie superalgebra
$\mathrm{gl}(M|N)$
defined by the structure constants
Eq.\ (\ref{eq: structure constants gl(M|N)})
for $A,B,C,D=1,\cdots,M+N$.
When $N=0$, the structure constants%
~(\ref{eq: structure constants gl(M|N)})
reduce to
\begin{equation}
\mathfrak{d}^{B}_{C}=
-
\delta^{B}_{C},
\qquad
\mathfrak{e}^{BD}_{AC}=
+\delta^{D}_{A},
\label{eq: structure constants gl(M|0)}
\end{equation}
for $A,B,C,D=1,\cdots,M$. These define
the Lie algebra gl($M$) of the non-compact Lie group GL$($M$)$.
When $M=0$, the structure constants%
~(\ref{eq: structure constants gl(M|N)})
reduce to
\begin{equation}
\mathfrak{d}^{B}_{C}=
+\delta^{B}_{C},
\qquad
\mathfrak{e}^{BD}_{AC}=
-\delta^{D}_{A},
\label{eq: structure constants gl(0|N)}
\end{equation}
for $A,B,C,D=1,\cdots,N$. These define
the Lie algebra u($N$) of the compact Lie group U($N$).
There exists a free-fermion and free-boson realization of
the current algebra%
~(\ref{eq: def gl(M,N) current algebra of level k})
defined by the action%
~\cite{footnote: Einstein convention}
\begin{subequations}
\label{eq: free fermion and free boson rep gl(M|N) level k}
\begin{equation}
S^{\ }_{*}:=
\sum_{\iota=1}^{k}
\int \frac{d\bar zd z}{2\pi{i}}
\left(
\psi^{A\dag}_{\ \ \iota}
\bar \partial\,
\psi^{\ }_{A\iota}
+
\bar \psi^{A\dag}_{\ \ \iota}
\partial\,
\bar \psi^{\ }_{A\iota}
\right)
\label{eq: def S* gl(M|N)}
\end{equation}
with the partition function
\begin{eqnarray}
Z^{\ }_{*}:=
\int\mathcal{D}[\psi^{\dag},\psi,
\bar{\psi}^{\dag},\bar{\psi}
]
\exp(-S^{\ }_{*}),
\end{eqnarray}
\end{subequations}
where it is understood that
$\psi^{\ }_{A\iota}$ and $\bar\psi^{\ }_{A\iota}$ are
complex-valued integration variables
for the $M$ values of $A$ with grade 0
while $\psi^{\ }_{A\iota}$ and $\bar\psi^{\ }_{A\iota}$
are Grassmann-valued integration variables
for the $N$ values of $A$ with grade 1, regardless
of the value taken by the flavor index $\iota=1,\cdots,k$.
The current algebra (\ref{eq: def gl(M,N) current algebra of level k})
is then realized by the representation
\begin{equation}
J^{\,B}_{A}:=
\sum_{\iota=1}^{k}
:\!
\psi^{\ }_{A\iota}
\psi^{B\dag}_{\ \ \iota}
\!:\,
,
\qquad
\bar J^{\,B}_{A}:=
\sum_{\iota=1}^{k}
:\!
\bar \psi^{\ }_{A\iota}
\bar \psi^{B\dag}_{\ \ \iota}
\!:,
\label{eq: def GL(M|N) currents as free spinors}
\end{equation}
as follows from the OPE
\begin{equation}
\begin{split}
&
\psi^{\ }_{A\iota}(z)
\psi^{B\dag}_{\iota'}(0)=
(-1)^{AB+1}
\psi^{B\dag}_{\iota'}(z)
\psi^{\ }_{A\iota}(0)=
\frac{\delta^{\ }_{\iota\iota'}\delta^{B}_{A}}{z},
\\
&
\bar{\psi}^{\ }_{A\iota}(\bar{z})
\bar{\psi}^{B\dag}_{\iota'}(0)=
(-1)^{AB+1}
\bar{\psi}^{B\dag}_{\iota'}(\bar{z})
\bar{\psi}^{\ }_{A\iota}(0)=
\frac{\delta^{\ }_{\iota\iota'}\delta^{B}_{A}}{\bar z},
\\
&
\psi^{\ }_{A \iota }(z)
\bar\psi^{B\dag}_{\ \ \iota'}( 0)=0,
\end{split}
\end{equation}
with $A,B=1,\cdots,M+N$ and $\iota=1,\cdots,k$.
The expressions in
Eq.~(\ref{eq: def GL(M|N) currents as free spinors})
form a representation
of the $\widehat{\mathrm{gl}}(M|N)^{\ }_{k}$
current algebra in terms of free fermions.
There are two Casimir invariants of rank 2 in $\mathrm{gl}(M|N)$
that we use to perturb the free field theory
(\ref{eq: free fermion and free boson rep gl(M|N) level k})
with two types of current-current interactions,
both of which are invariant under the global $\mathrm{GL}(M|N)$
symmetry,\cite{Guruswamy00}
\begin{equation}
\begin{split}
&
Z:=
\int\mathcal{D}[\psi^{\dag},\psi,
\bar{\psi}^\dag,\bar{\psi}]
\exp(-S),
\\
&
S:=
S^{\ }_{*}
+
\int \frac{d\bar zdz}{2\pi{i}}
\left(
\frac{g^{\ }_{\mathrm{A}}}{2\pi}
\mathcal{O}^{\ }_{\mathrm{A}}
+
\frac{g^{\ }_{\mathrm{M}}}{2\pi}
\mathcal{O}^{\ }_{\mathrm{M}}
\right),
\\
&
\mathcal{O}^{\ }_{\mathrm{A}}:=
-
J^{\,A}_{A}\,
(-1)^A\,
\bar J^{\,B}_{B}\,
(-1)^B
\equiv
-
\mathrm{str}\, J\,
\mathrm{str}\,\bar J,
\\
&
\mathcal{O}^{\ }_{\mathrm{M}}:=
-
J^{\,B}_{A}
\bar J^{\,A}_{B}
(-1)^A
\equiv
-
\mathrm{str}\,
\left(
J\bar J
\right).
\label{eq: glMN perturbed by two current-current int.}
\end{split}
\end{equation}
Formally, one may allow the coupling constants
$g^{\ }_{\mathrm{A}}$
and
$g^{\ }_{\mathrm{M}}$
to take on any real (i.e., positive or \textit{negative})
values. However, to make connection with the
above mentioned two-dimensional
tight-binding models
in symmetry class BDI
of Anderson localization,
we must demand that
$g^{\ }_{\mathrm{A}}$
and
$g^{\ }_{\mathrm{M}}$
be positive.
(See Appendix~\ref{sec: HWK model}.)
The ``classical'' counterparts to the high-gradient operators of order $s$
in Eq.~(\ref{eq: classical hgo su(2)k})
are the homogeneous polynomials
\begin{equation}
T^{\,A^{\ }_{1}\cdots\,A^{\ }_{s}\,\bar A^{\ }_{1} \cdots\,\bar A^{\ }_{s}}
_{B^{\ }_{1}\,\cdots B^{\ }_{s} \bar B^{\ }_{1}\,\cdots \bar B^{\ }_{s}}
J^{\,B^{\ }_{1}}_{A^{\ }_{1}}
\cdots
J^{\,B^{\ }_{s}}_{A^{\ }_{s}}
\bar J^{\,\bar B^{\ }_{1}}_{\bar A^{\ }_{1}}
\cdots
\bar J^{\,\bar B^{\ }_{s}}_{\bar A^{\ }_{s}}
\label{eq: classical hgo gl(M|N)k}
\end{equation}
of the left and right currents that are invariant under the diagonal
GL($M|N$) symmetry group of the interacting theory.%
~\cite{footnote: invariant tensor}
The set of ``classical'' high-gradient operators
of order $s$ is specified once all the linearly independent
rank $2s$ invariant tensors
$
T^{\,A^{\ }_{1}\cdots\,A^{\ }_{s}\,\bar A^{\ }_{1} \cdots\,\bar A^{\ }_{s}}
_{B^{\ }_{1}\,\cdots B^{\ }_{s} \bar B^{\ }_{1}\,\cdots \bar B^{\ }_{s}}
$
in the adjoint representation of GL($M|N$)
which are invariant under GL($M|N$) transformations
have been enumerated.
At the quantum level, normal ordering
defines the quantum high-gradient operators of order $s$
as in Eq.\ (\ref{eq: quantum hgo su(2)k}).
We are now going to specialize to the case $M=N$
where the beta function for the coupling constant
$g^{\ }_{\mathrm{M}}$ vanishes identically,
an exact result.~\cite{Guruswamy00}
(As already mentioned, the other coupling constant
$g^{\ }_{\mathrm{A}}$ flows logarithmically
at a rate set by
$g^{\ }_{\mathrm{M}}$.)
The sector which we loosely denote by
\begin{equation}
\mathrm{PSL}(M|M)\sim
\mathrm{GL}(M|M)/\mathrm{U}(1)\times
\mathrm{U}(1)
\label{eq: PSL and GL}
\end{equation}
remains scale (conformally) invariant
for any value of $g^{\ }_{\mathrm{M}}$.
More specifically, $\mathrm{PSL}(M|M)$ is obtained
by first factoring out the U(1) subgroup
thereby obtaining the subgroup
$\mathrm{SL}(M|M)$ of $\mathrm{GL}(M|M)$,
followed in a second step by the ``gauging away''
of the states carrying the $\mathrm{U}(1)$
charges under
$
j:=
J^{\ A}_{A}
$
and
$
\bar{j}:=
\bar{J}^{\ A}_{A}
$.%
~\cite{Guruswamy00,Bershadsky99,Berkovits99}
This turns out to realize a line of
RG fixed points (and conformal field theories)
labeled by the coupling constant $g^{\ }_{\mathrm{M}}$.%
~\cite{Guruswamy00}
\subsection{
High-gradient operators
when $M,k\to\infty$
}
\label{subsec: largest HGO order s in GL(N|N) at N=k=infty}
We are going to show that, when $k$ and $M$ are very large, the spectrum
for the one-loop scaling dimensions of high-gradient operators
shares the same structure as that in Eq.%
~(\ref{eq: anomalous scaling dimensions for O(N) high-gradient operators}).
It will become clear by comparison to the case of $k=1$ that
the limit $M,k\to\infty$ is the extreme ``classical'' limit
whereas the limit $k=1$ is the extreme ``quantum'' limit.
We restrict the family of ``classical'' high-gradient operators to
objects of the form
\begin{eqnarray}
\mathrm{str}\,
\left(
J\bar{J}JJ\bar{J}J
\right)
\mathrm{str}\,
\left(
J\bar{J}\bar{J}
\right)
\cdots,
\label{eq: def HGO if M,k to infty}
\end{eqnarray}
i.e., to diagonal GL($M|M$)-invariant
monomials of order $s$ in both the holomorphic and antiholomorphic
currents. For any given order $s$,
the engineering dimensions are all equal and given by $2s$.
This degeneracy is lifted to first order in the coupling constant
$g^{\ }_{\mathrm{M}}$.
The task of enumerating
all linearly-independent high-gradient operators%
~(\ref{eq: def HGO if M,k to infty})
of order $s$ is greatly simplified by the assumption
$M,k\to\infty$.
We can rule out the scenario by which
it is a finite set of independent Casimir operators
of gl($M|M$) that fixes all the linearly independent classical
high-gradient operators of order $s$ once the limit $M\to\infty$
has been taken.
We can also rule out the scenario by which normal
ordering changes the book-keeping between classical and
quantum high-gradient operators of order $s$ once the limit $k\to\infty$
has been taken.
For high-gradient operators of
type Eq.\ (\ref{eq: classical hgo gl(M|N)k})
or (\ref{eq: def HGO if M,k to infty}),
the coupling $g^{\ }_{\mathrm{A}}$ does not
renormalize their scaling dimensions,
since $g^{\ }_{\mathrm{A}}$
(or $\mathcal{O}^{\ }_{\mathrm{A}}$)
can be removed from
the action (\ref{eq: glMN perturbed by two current-current int.})
by chiral transformation.
All that therefore is needed
to compute their one-loop scaling dimensions are their
OPE with the quadratic Casimir operator
$\mathcal{O}^{\ }_{\mathrm{M}}$.
We will write the following expressions for the general case of
$\mathrm{GL}(M|N)$, and will set $M=N$
(i.e., the case of interest)
only in Eqs.%
~(\ref{eq: result in large M,N for max and min lambda}),
(\ref{eq: result in large M, k for min x(s) max}),
and
(\ref{eq: result in large M, k for min x(s) min}).
The required OPEs follow from
(a) the intra-trace formula
\begin{subequations}
\begin{equation}
\begin{split}
&
\mathrm{str}\,
\big[
\Wick{24mm}J\mathcal{M}
\Wickunder{20mm}\bar{J}\mathcal{N}
\big]
\times \mathrm{str}\,
\big[
J\bar{J}
\big]
\\
&
\quad
=
\mathrm{str}\,\left(J\mathcal{N}\right)
\mathrm{str}\,\left(\mathcal{M}\bar{J}\right)
-
\mathrm{str}\,\left(
\mathcal{M}
\right)
\mathrm{str}\,\left(
J\bar{J}\mathcal{N}
\right)
\\
&
\qquad
-
\mathrm{str}\,\left(
J\mathcal{M}\bar{J}
\right)
\mathrm{str}\,\left(
\mathcal{N}
\right)
+
\mathrm{str}\,\left(J\mathcal{M}\right)
\mathrm{str}\,\left(\bar{J}\mathcal{N}\right)
\end{split}
\label{eq: glMN large MN intra trace formula}
\end{equation}
and (b) the inter-trace formula
\begin{equation}
\begin{split}
&
\mathrm{str}\,
\big[
\Wick{32mm}J\mathcal{M}
\big]
\mathrm{str}\,
\big[
\Wickunder{20mm}\bar{J}\mathcal{N}
\big]
\times \mathrm{str}\,
\big[
J\bar{J}
\big]
\\
&
\quad
=
\mathrm{str}\,\left(
J \mathcal{N}\bar{J}\mathcal{M}
\right)
-
\mathrm{str}\,
\left(
J\bar{J}
\mathcal{N}
\mathcal{M}
\right)
\\
&
\qquad\
-
\mathrm{str}\,\left(
J\mathcal{M}\mathcal{N}\bar{J}
\right)
+
\mathrm{str}\,
\left(
J
\mathcal{M}
\bar{J}
\mathcal{N}
\right)
\end{split}
\label{eq: glMN large MN inter trace formula}
\end{equation}
\end{subequations}
with $\mathcal{M}$ and $\mathcal{N}$ arbitrary operators.
Here we have used the short-hand notation of
Eq.\ (\ref{eq: def short hand OPE}).
To proceed we also need to distinguish linearly independent
high-gradient operators of order $s$. To this end,
a ``quantum number'', the number of switches, is introduced.%
~\cite{Lerner90}$^{-}$\cite{Mall93}
The number of switches of type
$n^{\ }_{\uparrow}$
and of type
$n^{\ }_{\downarrow}$
in a single trace are defined as follows. Consider the trace
\begin{subequations}
\begin{equation}
\mathrm{str}\,
\big(
J^{\ }_{\mu^{\ }_{1}}
J^{\ }_{\mu^{\ }_{2}}
J^{\ }_{\mu^{\ }_{3}}
\cdots
J^{\ }_{\mu^{\ }_{2n}}
\big)
\label{eq: str needed to define number switches}
\end{equation}
where $\mu^{\ }_{1},\cdots,\mu^{\ }_{2n}=\pm$
while $J^{\ }_{-}=J$
and $J^{\ }_{+}=\bar{J}$.
Write the sequence of ``conformal'' indices
\begin{equation}
\mu^{\ }_{1},\cdots,\mu^{\ }_{2n},\mu^{\ }_{2n+1}
\end{equation}
\end{subequations}
where $\mu^{\ }_{2n+1}=\mu^{\ }_{1}$
by cyclicity of the trace.
The number
$n^{\ }_{\uparrow}$
of switches of type $\uparrow$
is the number of sign changes from
$+\to -$
in two consecutive conformal indices when reading the sequence
$\mu^{\ }_{1},\cdots,\mu^{\ }_{2n},\mu^{\ }_{2n+1}$
from left to right.
The number
$n^{\ }_{\downarrow}$
of switches of type $\downarrow$
is the number of sign changes from
$-\to +$
in two consecutive conformal indices when reading the sequence
$\mu^{\ }_{1},\cdots,\mu^{\ }_{2n},\mu^{\ }_{2n+1}$
from left to right.
These quantum numbers are useful as it can be shown that
there is no contribution in the one-loop RG
of supertraces made out
of $2n$ currents as in
Eq.~(\ref{eq: str needed to define number switches})
from the subspace
with $n^{\ }_{\uparrow}$ and $n^{\ }_{\downarrow}$
to the one with
at least $n^{\ }_{\uparrow}+1$ and $n^{\ }_{\downarrow}+1$.
This implies a lower triangular structure for the linearized RG equations
obeyed by all supertraces of order $2n$ as in
Eq.~(\ref{eq: str needed to define number switches})
which allows to treat separately each sector defined by
a given number of switches.
We shall assume that the strongest renormalization of the
engineering scaling dimensions occurs within the sector made
of the maximum number of switches.
Within the subspace of maximal switches it is sufficient to introduce
\begin{equation}
\omega:=
J \bar{J}\equiv
J^{\ }_{-}J^{\ }_{+},
\qquad
\Omega^{\ }_{m}:=
\mathrm{str}\,\big(\omega^{m}\big),
\end{equation}
for any $m=1,2,3,\cdots$.
\begin{subequations}
With the help of the OPE
(\ref{eq: glMN large MN intra trace formula})
and
(\ref{eq: glMN large MN inter trace formula})
one verifies the OPE
\begin{equation}
\begin{split}
&
\mathrm{str}\,(\Wick{20mm}\omega \omega^m \Wickunder{13mm}\omega \omega^n)
\times \mathcal{O}^{\ }_{\mathrm{M}}=
-\Omega^{\ }_{m+2}\Omega^{\ }_n
-\Omega^{\ }_{n+2}\Omega^{\ }_m
\\
&
\hphantom{
\mathrm{str}\,(\Wick{20mm}\omega \omega^m \Wickunder{13mm}\omega \omega^n)
\times \mathcal{O}^{\ }_{\mathrm{M}}=
}
-2\Omega^{\ }_{m+1}\Omega^{\ }_{n+1},
\\
&
\mathrm{str}\,(\Wick{28mm}\omega \omega^m)
\mathrm{str}\,(\Wickunder{13mm}\omega \omega^n)
\times
\mathcal{O}^{\ }_{\mathrm{M}}=
-4\Omega^{\ }_{m+n+2},
\\
&
\mathrm{str}\,
(\Wick{13mm}\Wickunder{13mm}\omega \omega^m)
\times
\mathcal{O}^{\ }_{\mathrm{M}}=
-\Omega^{\ }_1 \Omega^{\ }_m
-(N-M)\Omega^{\ }_{m+1},
\end{split}
\label{eq: OPE within maximal switch subspace}
\end{equation}
and
\begin{equation}
\begin{split}
&
\Wick{10mm}\Wickunder{10mm}\Omega^{\ }_m
\times \mathcal{O}_{\mathrm{M}}
=
-2m \sum_{k,l=1}^{k+l=m}\Omega^{\ }_{k}\Omega^{\ }_{l}
-2m (N-M) \Omega^{\ }_m,
\\
&
\Wick{17mm}\Omega^{r_m}_m \Wickunder{11mm}
\Omega^{r_n}_n
\times
\mathcal{O}_{\mathrm{M}}
=
-4 r_m r_n mn \Omega^{\ }_{m+n}
\Omega^{r_m-1}_m \Omega^{r_n-1}_n,
\label{eq: OPE within maximal switch subspace 2}
\end{split}
\end{equation}
for any
$m,n,r^{\ }_m,r^{\ }_n=1,2,3,\cdots$.
\label{eq: OPE within maximal switch subspace all}
\end{subequations}
The action of the linearized one-loop RG flow
on the space of composite operators
in the subspace of maximal switches
spanned by
\begin{eqnarray}
\Omega^{r^{\ }_{1}}_{1}
\Omega^{r^{\ }_{2}}_{2}
\cdots
\Omega^{r^{\ }_{L}}_{L},
\quad
\sum_{p=1}^{L} p \ r^{\ }_p=2s,
\end{eqnarray}
is encoded by the operator
\begin{eqnarray}
\hat{R}
\!\!&:=&\!\!\!
-2
\left(N-M\right)
\sum_{k}
k \Omega^{\ }_{k}\frac{\partial}{\partial \Omega^{\ }_{k}}
\nonumber\\
&&\!\!\!{}
-
2
\sum_{l,n}
\left[
(l+n)
\Omega^{\ }_{l} \Omega^{\ }_{n}
\frac{\partial}{\partial \Omega^{\ }_{l+n}}
+
ln\,
\Omega^{\ }_{l+n}
\frac{\partial}{\partial \Omega^{\ }_{l}}
\frac{\partial}{\partial \Omega^{\ }_{n}}
\right].
\nonumber\\&&
\label{eq: RG equation}
\end{eqnarray}
It is instructive to compare the OPE
(\ref{eq: OPE within maximal switch subspace all})
and the RG equation
(\ref{eq: RG equation})
with the corresponding result in the weakly coupled
NL$\sigma$M on
the symmetric space
$\mathrm{U}(P+Q)/\mathrm{U}(P)\times \mathrm{U}(Q)$
with $P,Q>1$:%
~\cite{Lerner90,Wegner91}
They
are essentially identical to the corresponding result for the
$\mathrm{U}(P+Q)/\mathrm{U}(P)\times \mathrm{U}(Q)$
NL$\sigma$M.
Now we return to the case $M=N$.
The diagonalization of $\hat{R}$
gives the largest and smallest eigenvalue
~\cite{Lerner90, Wegner91}
\begin{eqnarray}
\lambda^{(s)}_{\mathrm{max}}=
+2 s(s-1)=
-\lambda^{(s)}_{\mathrm{min}}.
\label{eq: result in large M,N for max and min lambda}
\end{eqnarray}
Thus, both largest and smallest eigenvalues
depend quadratically on $s$.
In turn, one obtains a spectrum of one-loop scaling
dimensions with the upper and lower bounds
\begin{eqnarray}
&&
x^{(s)}_{\mathrm{max}}=
2s
+
\frac{g^{\ }_{\mathrm{M}}}{\pi}s(s-1),
\label{eq: result in large M, k for min x(s) max}
\\
&&
x^{(s)}_{\mathrm{min}}=
2s
-
\frac{g^{\ }_{\mathrm{M}}}{\pi}s(s-1),
\label{eq: result in large M, k for min x(s) min}
\end{eqnarray}
for any given $1<s=2,3,\cdots$.
Observe that these bounds
are interchanged when
$g^{\ }_{\mathrm{M}}\to-g^{\ }_{\mathrm{M}}$.
\subsection{
High-gradient operators
when $k=1$
}
\label{subsec: largest HGO order s in GL(N|N) at k=1}
Having dealt with the extreme ``classical'' limit,
we turn our attention to the extreme ``quantum'' limit
$M=1,2,3,\cdots$ and $k=1$ for which
the interacting field theory
(\ref{eq: glMN perturbed by two current-current int.})
describes a problem of Anderson localization in $d=2$ dimensions
reviewed in Appendix~\ref{sec: HWK model}.
The classification of all independent
high-gradient operators in GL($M|M$)
or in PSL($M|M$)
is more involved than in SU(2)
because the problem of listing all invariants
is more complex.\cite{Bershadsky99}
An increase of complexity can already be
seen at the level of SU($N$)
for which the invariant tensors of rank $2s$ are obtained
from all possible products of one rank 2 tensor and two rank 3 tensors.%
\cite{Dittner72}
Instead of considering the most generic family of ``classical''
high-gradient operators%
~(\ref{eq: classical hgo gl(M|N)k}),
we consider
the GL($M|M$) invariant family of ``classical'' objects
\begin{equation}
\left\{
\left.
\mathcal{O}^{m}_{\mathrm{M}}
\mathcal{O}^{n}_{\mathrm{A}}
\right|
m,n=0,1,2,3,\cdots,
\quad
m+n=s
\right\},
\label{eq: hgo glMNk=1}
\end{equation}
which must then be normal ordered.
We are going to prove that the coupling constant $Z^{(s)}_{m,n}$
of the high-gradient operator
$\mathcal{O}^{m}_{\mathrm{M}}\mathcal{O}^{n}_{\mathrm{A}}$
in the action
\begin{equation}
\begin{split}
S=&
S^{\ }_{*}
+
\int \frac{d\bar zdz}{2\pi{i}}
\left(
\frac{g^{\ }_{\mathrm{A}}}{2\pi}
\mathcal{O}^{\ }_{\mathrm{A}}
+
\frac{g^{\ }_\mathrm{M}}{2\pi}
\mathcal{O}^{\ }_{\mathrm{M}}
\right)
\\
&
-
\sum_{m,n = 0}^{m+n=s}
Z^{(s)}_{m,n}\mathfrak{a}^{2s-2}
\int \frac{d\bar zdz}{2\pi{i}}
\mathcal{O}^{m}_{\mathrm{M}}
\mathcal{O}^{n}_{\mathrm{A}}
\end{split}
\end{equation}
obeys the linearized one-loop RG equation
\begin{equation}
\begin{split}
\frac{d Z^{(s)}_{m,n}}{dl}=&
\left(
2
-
2s
\right)
Z^{(s)}_{m,n}
\\
&
-
4\frac{g^{\ }_{\mathrm{M}}}{2\pi}
m(m-1)Z^{(s)}_{m,n}
\\
&
+4 \frac{g^{\ }_{\mathrm{M}}}{2\pi}
(m+1)^2
Z^{(s)}_{m+1,n-1}
\label{eq: RG equations of Zmn in gl(MM)k=1}
\end{split}
\end{equation}
for any $m,n=0,1,2,3,\cdots$ with $m+n=s>1$.
For the $PSL(M|M)$ theory the operators $\mathcal{O}^{n}_{A}$
are all absent.
The RG equation (\ref{eq: RG equations of Zmn in gl(MM)k=1})
shows that there is no feedback from
high-gradient operators containing a factor
$\mathcal{O}^{n}_{\mathrm{A}}$
to those containing a factor
$\mathcal{O}^{n'}_{\mathrm{A}}$
provided $n'<n$.
Diagonalization of the RG equation gives
the set of one-loop scaling dimensions
\begin{equation}
\begin{split}
&
x^{(s)}_{m,n}=
2s
+
\frac{2g^{\ }_{\mathrm{M}}}{\pi}
m(m-1)
\end{split}
\label{eq: scaling dimension OM raised to the power s}
\end{equation}
for all $m,n=0,1,2,3,\cdots$ such that $m+n=s$.
For a positive
$g^{\ }_{\mathrm{M}}$ we get the lower and upper bounds
\begin{equation}
x^{(s)}_{\mathrm{min}}=
2s,
\qquad
x^{(s)}_{\mathrm{max}}=
2s
+
\frac{2g^{\ }_{\mathrm{M}}}{\pi}
s(s-1),
\label{eq: scaling dimension OM raised to the power s if positive}
\end{equation}
respectively, i.e.,
$x^{(s)}_{m,n}$ with $m+n=s$
is always much larger than the engineering dimension
$2s$ so that the high-gradient operator
$\mathcal{O}^{m}_{\mathrm{M}}\mathcal{O}^{n}_{\mathrm{A}}$
is irrelevant.
For a negative
$g^{\ }_{\mathrm{M}}$,
the spectrum of lower bounds on
$x^{(s)}_{m,n}$ with $m+n=s$
is unbounded from below when $s\to\infty$,
i.e.,
\begin{equation}
x^{(s)}_{\mathrm{min}}=
2s
-
\frac{2|g^{\ }_{\mathrm{M}}|}{\pi}
s(s-1),
\qquad
x^{(s)}_{\mathrm{max}}=
2s.
\label{eq: scaling dimension OM raised to the power s if negative}
\end{equation}
\textit{Proof:}
Having made the simplification
$g^{\ }_{\mathrm{A}}=0$
we only need to compute the OPE
$
\mathcal{O}^{m}_{\mathrm{M}}
\mathcal{O}^{n}_{\mathrm{A}}
\times
\mathcal{O}_{\mathrm{M}}
$,
where $1\le m+n=s$, to justify
Eqs.%
~(\ref{eq: RG equations of Zmn in gl(MM)k=1})
and%
~(\ref{eq: scaling dimension OM raised to the power s}).
Each operator in Eq.%
~(\ref{eq: hgo glMNk=1})
contains terms with
$4s$ bosons,
$4s-2$ bosons and $2$ different fermions,
$4s-4$ bosons and $4$ different fermions,
...,
$4s-2M$ bosons and $2M$ different fermions,
and so on.
The terms that contain identical
fermions have short-distance singularities
and hence they should be interpreted as
operators that involve gradients over fermion fields
after normal ordering.
It is understood from now on that the OPE
$
\mathcal{O}^{m}_{\mathrm{M}}
\mathcal{O}^{n}_{\mathrm{A}}
\times
\mathcal{O}_{\mathrm{M}}
$
is only over the terms in the expansion
$
\mathcal{O}^{m}_{\mathrm{M}}
\mathcal{O}^{n}_{\mathrm{A}}
$
involving different fermions, i.e., the OPE we present are
``accurate'' up to terms involving gradients over fermionic spinors.
Neglecting the OPE between derivatives
of the fermionic spinors and
$
\mathcal{O}^{\ }_{\mathrm{M}}
$
is harmless insofar as these OPE cannot feedback into the RG
flows of those contributions that we keep.
Let
\begin{equation}
(\chi \xi):=
\sum_{A=1}^{2M}
\chi^A \xi^{\ }_A=
\sum_{A=1}^{2M}
(-)^{A}
\xi^{\ }_A\chi^A
\end{equation}
and remember that
$
\mathcal{O}_{\mathrm{A}}
= -\left(\psi^{\dag}\psi\right)
\left(\bar{\psi}^{\dag}\bar{\psi}\right)
$
while
$
\mathcal{O}_{\mathrm{M}}
= -\left(\psi^{\dag}\bar{\psi}\right)
\left(\bar{\psi}^{\dag}\psi\right)
$.}
The OPE that involve
$\big(\psi^{\dag}\bar{\psi}\big)$
and
$\big(\bar{\psi}^{\dag}\psi\big)$
are
\begin{equation}
\begin{split}
&
\big(\Wick{18mm}\psi^{\dag}\Wickunder{10mm}\bar{\psi}\big)
\times
\big(\bar{\psi}^{\dag}\psi\big)=
0,
\\
&
(\Wick{26mm}\psi^{\dag}\bar{\psi})(\psi^{\dag}\Wickunder{9mm}\bar{\psi})
\times
(\bar{\psi}^{\dag}\psi)=
-
(\psi^{\dag} \bar{\psi}),
\\
&
(\Wick{35mm}\psi^{\dag}\bar{\psi})(\Wickunder{17mm}\bar{\psi}^{\dag} \psi)
\times
(\psi^{\dag}\bar{\psi})(\bar{\psi}^{\dag}\psi)=
-
\mathcal{O}^{\ }_{\mathrm{A}}.
\end{split}
\label{eq: OPE within the O_M sector}
\end{equation}
On the other hand,
the OPE that involve
$\big(\psi^{\dag} \psi \big)$,
$\big(\bar{\psi}^{\dag}\bar{\psi} \big)$,
$\big(\psi^{\dag}\bar{\psi}\big)$,
and
$\big(\bar{\psi}^{\dag}\psi \big)$ are given by
\begin{subequations}
\begin{eqnarray}
&&
(\Wick{35mm}\psi^{\dag}\psi)
(\Wickunder{17mm}\bar{\psi}^{\dag}\bar{\psi})
\times
(\psi^{\dag}\bar{\psi})(\bar{\psi}^{\dag}\psi)
\label{eq: OPE outside the O_M sector a}
\\
&&
\qquad
=
-
(\Wick{35mm}\psi^{\dag}\psi)
(\bar{\psi}^{\dag}\Wickunder{18mm}\bar{\psi})
\times
(\psi^{\dag}\bar{\psi})(\bar{\psi}^{\dag}\psi)
=
-
\mathcal{O}^{\ }_{\mathrm{M}},
\nonumber\\
&&
(\Wick{35mm}\psi^{\dag}\psi)
(\Wickunder{17mm}\bar{\psi}^{\dag} \psi)
\times
(\psi^{\dag}\bar{\psi})(\bar{\psi}^{\dag}\psi)
\label{eq: OPE outside the O_M sector b}
\\
&&
\qquad
=
-(\psi^{\dag}\Wick{18mm}\psi)
(\Wickunder{17mm}\bar{\psi}^{\dag} \psi)
\times
(\psi^{\dag}\bar{\psi})(\bar{\psi}^{\dag}\psi)
=
(\psi^{\dag}\psi)(\bar{\psi}^{\dag}\psi).
\nonumber
\end{eqnarray}
\label{eq: OPE outside the O_M sector}
\end{subequations}
Both $\mathcal{O}^{\ }_{\mathrm{M}}$ and $\mathcal{O}^{\ }_{\mathrm{A}}$
are generated
through the OPE
(\ref{eq: OPE within the O_M sector})
and
(\ref{eq: OPE outside the O_M sector}),
respectively.
However,
two OPE
in
Eq.\ (\ref{eq: OPE outside the O_M sector a})
always appear in a pairwise fashion
and cancel each other,
\begin{equation}
(\Wick{25mm}\psi^{\dag}\psi)
(\Wickunder{16mm}\bar{\psi}^{\dag}\bar{\psi})
\mathcal{A}
\times
\mathcal{O}^{\ }_{\mathrm{M}}
+
(\Wick{25mm}\psi^{\dag}\psi)
(\bar{\psi}^{\dag}\Wickunder{12mm}\bar{\psi})
\mathcal{A}
\times
\mathcal{O}_{\mathrm{M}}=0,
\end{equation}
where $\mathcal{A}$ is some operator.
Hence, the total number of $\mathcal{O}^{\ }_\mathrm{M}$
contained in a high-gradient operator
never increases under the linearized RG flow.
{}From the OPE (\ref{eq: OPE within the O_M sector})
and (\ref{eq: OPE outside the O_M sector})
one deduces the OPE
\begin{eqnarray}
&&
\mathcal{O}^{m}_{\mathrm{M}}
\mathcal{O}^{n}_{\mathrm{A}}
\times
\mathcal{O}^{\ }_{\mathrm{M}}
\nonumber\\
&&\quad
=
m
\mathcal{O}^{m-1}_{\mathrm{M}}
\mathcal{O}^{n}_{\mathrm{A}}
\Wick{10mm}
\Wickunder{10mm}
\mathcal{O}_{\mathrm{M}}
\times
\mathcal{O}_{\mathrm{M}}
+
n
\mathcal{O}^{m}_{\mathrm{M}}
\mathcal{O}^{n-1}_{\mathrm{A}}
\Wick{10mm}
\Wickunder{10mm}
\mathcal{O}_{\mathrm{A}}
\times
\mathcal{O}_{\mathrm{M}}
\nonumber\\
&&\qquad
+
mn
\mathcal{O}^{m-1}_{\mathrm{M}}
\mathcal{O}^{n-1}_{\mathrm{A}}
\Wick{15mm}
\mathcal{O}_{\mathrm{M}}
\Wickunder{10mm}
\mathcal{O}_{\mathrm{A}}
\times
\mathcal{O}_{\mathrm{M}}
\nonumber\\
&&\qquad
+
m(m-1)
\mathcal{O}^{m-2}_{\mathrm{M}}
\mathcal{O}^{n}_{\mathrm{A}}
\Wick{16mm}
\mathcal{O}_{\mathrm{M}}
\Wickunder{10mm}
\mathcal{O}_{\mathrm{M}}
\times
\mathcal{O}_{\mathrm{M}}
\nonumber\\
&&\qquad
+
n(n-1)
\mathcal{O}^{m}_{\mathrm{M}}
\mathcal{O}^{n-2}_{\mathrm{A}}
\Wick{15mm}
\mathcal{O}_{\mathrm{A}}
\Wickunder{10mm}
\mathcal{O}_{\mathrm{A}}
\times
\mathcal{O}_{\mathrm{M}}
\nonumber\\
&&\quad
=
2m(m-1)
\mathcal{O}^{m}_{\mathrm{M}}
\mathcal{O}^{n}_{\mathrm{A}}
-
2m^2
\mathcal{O}^{m-1}_{\mathrm{M}}
\mathcal{O}^{n+1}_{\mathrm{A}}.
\label{eq: full OPE hgo}
\end{eqnarray}
(When $m=0$, the term with
$\mathcal{O}^{m-1}_{\mathrm{M}}$
is absent from the last line.)
The linearized one-loop RG equation%
~(\ref{eq: RG equations of Zmn in gl(MM)k=1})
thus follows
from the OPE (\ref{eq: full OPE hgo}).
$\square$
Had we assumed the level $k$ to be larger than $k=1$,
the family~(\ref{eq: hgo glMNk=1})
would not have been closed under the OPE with
$\mathcal{O}^{\ }_{\mathrm{M}}$.
For example, in the extreme
classical limit $M,k\to\infty$
the family of high-gradient operators is given by the much
larger family~(\ref{eq: def HGO if M,k to infty}).
We close by pointing out that we could have reached the same conclusions
on the spectrum of one-loop scaling dimensions of high-gradient operators
had we used, instead of the effective action with diagonal GL($M|M)$
symmetry, an action built out of fermionic replicas or an action built out
of bosonic replicas and taken the number of replicas to zero at the end of
the day. Using bosonic replicas mimics very closely the line
of argument presented here. Using fermionic replicas singles out
high-gradient operators made of fermionic spinors that are all
distinct through their replica index and then taking this replica index
to zero, very much in the same way as replicated
vortices in certain classes of classical random
two-dimensional Coulomb gases.%
~\cite{Korshunov93}$^{-}$\cite{Fukui02}
We would like to stress that our
results depend crucially on the
continuous symmetry GL($2N|2N)$
of the $\widehat{\mathrm{gl}}(2N|2N)^{\ }_{k=1}$ Thirring model.
(From the point of view of Anderson localization,
it is the existence of a continuous symmetry
not the symmetry group per se that matters since
the symmetry group changes depending on the choice made to
represent single-particle Green's functions, say
a supersymmetric, bosonic replicas, fermionic replicas, or Keldysh
path integral.)
If we consider local perturbations (local operators)
that break the GL$(2N|2N)$ symmetry,
an infinite set of local operators with relevant (negative)
scaling dimensions can appear.
This alternative set of local operators may be related to
the situation recently considered by Le Doussal and Schehr.%
\cite{LeDoussal06}
The microscopic starting point of Ref.~\onlinecite{LeDoussal06}
is a class of classical random $XY$ models in two dimensions.
These models can also be viewed as
interacting models of Dirac fermions subjected to disorder,
by the magic of the boson-fermion duality in
$d=(1+1)$ dimensions.%
\cite{Mudry99,Guruswamy00}
The difference with our paper is that
their model is not invariant under a
\textit{continuous} symmetry group, but only under
the discretey symmetry group which permutes
the replica indices.
It is then necessary to use the full machinery of functional RG
to account for the one-loop relevance of high-gradient operators.
\subsection{Comparison with the
$\mathbb{C}P^{1|2}$ NL$\sigma$M}
The perturbed
$\widehat{\mathrm{gl}}(2N|2N)^{\ }_{k=1}$
WZW model with $g^{\ }_{\mathrm{M}}>0$ (Thirring model)
describes a problem of Anderson localization in
two dimensions.
As briefly reviewed in
Appendix~\ref{sec: HWK model},
this problem of Anderson localization
arises as the long-wavelength description of
a tight-binding model on a two-dimensional
bipartite lattice with a form of disorder that
preserves sublattice and time-reversal symmetries.
The long-wavelength theory is a $(2+1)$-dimensional
Dirac equation subject to disorder potentials
consistent with these symmetries.
In terms of the symmetry-based classification of Anderson
localization, the relevant symmetry class is the class BDI
(chiral-orthogonal symmetry class).%
~\cite{Zirnbauer96}$^{-}$\cite{Heinzner05}
It is possible to use a different
representation of this Anderson localization problem, in terms of
a NL$\sigma$M
whose target space is the \textit{non-compact} supermanifold
\begin{equation}
\mathrm{GL}(2N|2N)/\mathrm{OSp}(2N|2N).
\label{eq: NLSM for BDI}
\end{equation}
(A suitable analytical continuation in the boson-boson sector is needed
to implement the non-compactness.~\cite{Zirnbauer96})
These two descriptions, one in terms of the
Thirring model and the other in terms of the
NL$\sigma$M,
are complementary to each other in
that when one of the models is strongly coupled,
the other is weakly coupled.
A reflection of this appears in the conductivity. The
coupling constant of the
NL$\sigma$M is inversely proportional to the conductivity.
In the clean limit $g^{\ }_{\mathrm{M}}=0$ of the Thirring model
the conductivity is of order unity (in units of $e^2/h$),
consistent with the strongly coupled regime of the NL$\sigma$M.
The conductivity increases with $g^{\ }_{\mathrm{M}}>0$
as seen in perturbation theory.\cite{Ostrovsky06}
Furthermore, both $g^{\ }_{\mathrm{M}}$ and the conductivity
are exactly marginal.
This suggests a deeper relationship between
the Thirring model and the NL$\sigma$M,
and indeed (following Ref.~\onlinecite{Ryu09}),
one can turn the Thirring model into the NL$\sigma$M continuously
by tuning $g^{\ }_{\mathrm{M}}$ (or equivalently the conductivity) continuously.
We consider the case $N=1$ for which
the \textit{non-compact} target supermanifold
(\ref{eq: NLSM for BDI})
is isomorphic to
$\mathrm{U}(1)\times\mathrm{U}(1)\times\mathbb{C}P^{1|2}$,
where again a suitable analytical continuation is understood for
$\mathbb{C}P^{1|2}$, i.e.,
we need to consider the \textit{non-compact} counterpart to
$\mathbb{C}P^{1|2}$
as defined in Appendix~\ref{app: HGO on projective superspaces}.
Obtaining the non-compact $\mathbb{C}P^{1|2}$ target supermanifold
of the NL$\sigma$M from
$\mathrm{U}(1)\times \mathrm{U}(1)\times\mathbb{C}P^{1|2}$
corresponds in the Thirring model to the reduction of
the $\mathrm{GL}(2|2)$ to the $\mathrm{PSL}(2|2)$ current algebra
in Eq.\ (\ref{eq: PSL and GL}).
It is explicitly shown in Appendix%
~\ref{app: HGO on projective superspaces}
that all high-gradient operators are made more irrelevant
at one-loop order by fluctuations
in any non-compact $\mathbb{C}P^{N+M-1|N}$ NL$\sigma$M
labeled by the non-negative integers $M$ and $N$.
To be more precise,
we find that the largest and smallest
one-loop scaling dimensions
for the high-gradient operators of type%
~(\ref{eq: set of HGO}),
for a given $s$, are
\begin{equation}
\begin{split}
x^{(s)}_{\mathrm{max}}=&\,
2s
+
2|t|s(s-1),
\\
x^{(s)}_{\mathrm{min}}=&\,
2s
+
2|t|\times0,
\end{split}
\end{equation}
where $|t|>0$ is the coupling constant of the
non-compact $\mathbb{C}P^{N-1|N}$ NL$\sigma$M.
This is fully consistent with our finding
(\ref{eq: scaling dimension OM raised to the power s if positive})
in the Thirring model.
We conclude that, in symmetry class BDI,
high-gradient operators
in the Thirring model with $g^{\ }_{\mathrm{M}}>0$
behave in the same way as
in the corresponding NL$\sigma$M (i.e., the one that belongs
to the symmetry class BDI).
The sign of $g^{\ }_{\mathrm{M}}$
in the perturbed $\widehat{\mathrm{gl}}(2N|2N)^{\ }_{k=1}$
WZW model can be chosen to be negative,
$g^{\ }_{\mathrm{M}}<0$.
If so, this field theory does not represent anymore
the moments of the single-particle Green's function
in a problem of Anderson localization in (bulk) two dimensions.
Nevertheless, this field theory does describe a problem of
Anderson localization
which, however, now belongs to the \textit{different}
symmetry class CII (chiral-symplectic symmetry class)
describing the effect of disorder on the Dirac
fermions which are known to appear
at the two-dimensional boundary of a three-dimensional
topological band insulator in the same symmetry class.%
~\cite{Schnyder08,Ryu09,Hosur09}
Equation~(\ref{eq: scaling dimension OM raised to the power s if negative})
implies that high-gradient operators are now made more relevant
by the current-current perturbation $g^{\ }_{\mathrm{M}}<0$
to one-loop order.
As for the case with $g^{\ }_{\mathrm{M}}>0$,
a problem of Anderson localization in the symmetry class CII
is characterized by a NL$\sigma$M with a corresponding target
manifold. As before, the beta function of the coupling constant
$g^{\ }_{\mathrm{M}}<0$
of the Thirring model as well as that of the coupling
constant of the corresponding
NL$\sigma$M vanish, and
one can interpolate\cite{Ryu09}
between the weak coupling limit of
the Thirring model and the strong coupling limit
of the NL$\sigma$M
and conversely, by tuning $g^{\ }_{\mathrm{M}}$ continuously.
The target supermanifold in symmetry class CII is
the \textit{compact} supermanifold~(\ref{eq: NLSM for BDI}),
from which one extracts when $N=1$ the NL$\sigma$M
on the
\textit{compact} supermanifold $\mathbb{C}P^{1|2}$.
\cite{CommentSchomerusSaleur2009}
It is explicitly shown in Appendix%
~\ref{app: HGO on projective superspaces}
that all high-gradient operators are made more relevant
at one-loop order by fluctuations
in any compact $\mathbb{C}P^{N+M-1|N}$ NL$\sigma$M
labeled by the non-negative integers $M$ and $N$.
In particular, for $M=0$,
we find that the largest and smallest
one-loop scaling dimensions
for the high-gradient operators of type%
~(\ref{eq: set of HGO}),
for a given $s$, are
\begin{equation}
\begin{split}
x^{(s)}_{\mathrm{max}}=&\,
2s
-
2ts(s-1),
\\
x^{(s)}_{\mathrm{min}}=&\,
2s
+
2t \times0,
\end{split}
\end{equation}
where $t>0$ is the coupling constant of the
compact $\mathbb{C}P^{N-1|N}$ NL$\sigma$M.
Once again, we conclude that
high-gradient operators behave in the same way
in the Thirring model with $g^{\ }_{\mathrm{M}}<0$
and in the corresponding NL$\sigma$M that belongs to the
symmetry class CII.
\section{Conclusions}
\label{sec: conclusion}
More than twenty years after their discovery,
the role of high-gradient operators, which appear
to be highly relevant in one-loop computations of anomalous
dimensions in a great variety of NL$\sigma$Ms,
still remains a puzzle.
Indeed, this perturbative property is rather general
as it can apply to both compact and non-compact target manifolds.
In the absence of an exact calculation of observables that
would be sensitive to high-gradient operators,
it is still an outstanding question
whether the extreme
RG-relevance of these operators is an artifact of the
one-loop calculation (e.g., in the $2+\epsilon$-expansion),
or is a feature that is generally valid.
(For an attempt to compare the $\epsilon$
expansion in $d=2-\epsilon$ dimensions
with exact results obtained for $d=1$, see Ref.~\onlinecite{Ryu07a}.)
In order to shed some light on these issues
we have asked in this paper the following question.
Can high-gradient operators become relevant in the family of
two-dimensional
$\widehat{\mathrm{gl}}(M|M)^{\ }_{k}$ Thirring models
with $M$ and $k$ positive integers?
The strategy that we followed has three steps.
The first step consists of identifying all
the independent ``classical'' high-gradient operators of order $s$.
This is a problem of group theory that involves the enumeration of
all distinct GL($M|M$) singlets
in the direct product
of $2s$ adjoint representations
of GL($M|M$).
The second step consists of normal-ordering all independent
classical high-gradient operators of order $s$. This step depends
crucially on the level $k$ of the non-Abelian Thirring model.
The inverse level $1/k$ plays the role of a quantum parameter that
vanishes in the limit $k\to\infty$. The level $k=1$ is thus the most
``quantum''.
The computation of the linearized RG flows for the high-gradient
operators is the final step.
We could not solve the first step in its full generality.
We were nevertheless able to construct two sets
of high-gradient operators in the
extreme ``classical'' limit
$\widehat{\mathrm{gl}}(M|M)^{\ }_{k}$ with $M,k\to\infty$
and the extreme ``quantum'' case
$\widehat{\mathrm{gl}}(M|M)^{\ }_{k}$ with $M$ a positive integer and $k=1$,
respectively,
and carry out the second and third steps consistently,
i.e., show that each family
of normal-ordered high-gradient operators is closed under
the linearized RG flow equations.
The set of high-gradient operators that we considered in the extreme
``quantum'' limit is much smaller than the set of
high-gradient operators for the extreme ``classical'' case.
This is to be expected as normal ordering is extremely sensitive
to the free-field fermionic representation of the
$\widehat{\mathrm{gl}}(M|M)^{\ }_{k}$
current algebra at the unperturbed} WZW critical point.
This difference has dramatic consequences for the
spectrum of one-loop anomalous scaling dimensions
in the extreme ``classical'' and ``quantum'' cases.%
~\cite{Gepner86}
In the extreme ``classical'' case,
anomalous one-loop scaling dimensions
for high-gradient operators of order $s$ are
distributed in a symmetric fashion about zero with
the minimum and the maximum both depending
quadratically on the order $s$,
very much like for the family
of NL$\sigma$Ms on the target spaces
$\mathrm{U}(M+N)/\mathrm{U}(M)\times\mathrm{U}(N)$
with $M$ and $N$ positive integers.%
~\cite{Ryu07a,Altshuler86a}$^{-}$\cite{Altshuler91}
Hence, high-gradient operators
must become (one-loop) relevant for both signs of the current-current
interaction with increasing order $s$
very much in the same way as their cousins do in both
the compact family
$\mathrm{U}(M+N)/\mathrm{U}(M)\times\mathrm{U}(N)$
and the non-compact family
$\mathrm{U}(M,N)/\mathrm{U}(M)\times\mathrm{U}(N)$
with $M,N>1$.
In the extreme quantum case $k=1$, the spectrum of
anomalous one-loop scaling dimensions
of order $s$ is always one-sided,
i.e., positive for one sign of the current-current interaction.
For $\widehat{\mathrm{gl}}(M|M)^{\ }_{k=1}$ with $M$ a positive integer
the sign of the current-current interaction for which high-gradient
operators are always irrelevant corresponds to the interpretation
of the $\widehat{\mathrm{gl}}(2N|2N)^{\ }_{k=1}$ Thirring model as
a problem of Anderson localization
in random tight-binding models on two-dimensional
bipartite lattices (symmetry class BDI). We have shown in this paper that
the high-gradient operators in these random tight-binding models
are irrelevant at one-loop order.
High-gradient operators in those NL$\sigma$Ms
of relevance to the physics of Anderson localization
are related to the moments of the $dc$ conductance.%
~\cite{Altshuler86a}$^{-}$\cite{Altshuler91}
Their perturbative one-loop relevance has been interpreted
as the signature of broad tails in the probability distribution
of the conductance in Refs.~%
\onlinecite{Altshuler86a}--\onlinecite{Altshuler91}.
(One should, however, bear in mind that
the current-current correlation function entering
the Kubo formula for the conductance
looks rather different from a simple
$\mathrm{GL}(2N|2N)$ current-current
correlation function.%
~\cite{Ryu07b})
It would thus be very interesting to study the
probability distribution of
the $dc$ conductance in the relevant random tight-binding model
using nonperturbative techniques
(this may include, e.g., also numerical approaches)
in order to establish if it is broad or not.
\section*{Acknowledgments}
CM would like to thank Eduardo Fradkin for important comments.
This research was supported in part by the National Science Foundation
under Grant No.\ PHY05-51164 and under Grant No.\ DMR-0706140 (AWWL).
SR thanks the Center for Condensed Matter Theory at University of California,
Berkeley for its support.
|
\section{Introduction and the Model}
We are interested in the conserved Penrose-Fife type
equations
\begin{equation}\begin{split}\label{PF0}\partial_t\psi=\Delta\mu,\quad \mu=-\Delta\psi+\Phi'(\psi)-\lambda'(\psi)\vartheta,&\quad t\in J,\ x\in\Omega,\\
\partial_t\left(b(\vartheta)+\lambda(\psi)\right)-\Delta\vartheta=0,&\quad t\in J,\
x\in\Omega,
\end{split}\end{equation}
where $\vartheta=1/\theta$ and $\theta$ denotes the absolute temperature of the system, $\psi$ is the order parameter and $\Omega\subset\mathbb{R}^n$ is a bounded domain with boundary $\partial\Omega\in C^4$. The function $\Phi'$ is the derivative of the physical potential, which characterizes the different phases of the system. A typical example is the \emph{double well} potential $\Phi(s)=(s^2-1)^2$ with the two distinct minima $s=\pm1$. Typically, the nonlinear function $\lambda$ is a polynomial of second order.
For an explanation of \eqref{PF0} we will follow the lines of \textsc{Alt \& Pawlow} \cite{AltPaw} (see also \textsc{Brokate} \& \textsc{Sprekels} \cite[Section 4.4]{BroSpr}). We start with the rescaled Landau-Ginzburg functional (total Helmholtz free energy)
$$\mathcal{F}(\psi,\theta)=\int_\Omega\left(\frac{\gamma(\theta)}{2\theta}|\nabla\psi|^2+\frac{f(\psi,\theta)}{\theta}\right)\
dx,$$
where the free energy density $F(\psi,\theta):=\frac{\gamma(\theta)}{2}|\nabla\psi|^2+f(\psi,\theta)$ is rescaled by $1/\theta$. The reduced chemical potential $\mu$ is given by the variational derivative of $\mathcal{F}$ with respect to $\psi$, i.e.
$$\mu=\frac{\delta \mathcal{F}}{\delta\psi}(\psi,\theta)=\frac{1}{\theta}\left(-\gamma(\theta)\Delta\psi+\frac{\partial f(\psi,\theta)}{\partial \psi}\right).$$
Assuming that $\psi$ is a conserved quantity, we have the conservation law
$$\partial_t\psi+{\rm div} j=0.$$
Here $j$ is the flux of the order parameter $\psi$, for which we choose the well accepted constitutive law $j=-\nabla\mu$, i.e.\ the phase transition is driven by the chemical potential $\mu$ (see \cite[(4.4)]{BroSpr}). The kinetic equation for $\psi$ thus reads
$$\partial_t\psi=\Delta\mu,\quad \mu=\frac{1}{\theta}\left(-\gamma(\theta)\Delta\psi+\frac{\partial f(\psi,\theta)}{\partial \psi}\right).$$
If the volume of the system is preserved, the internal energy $e$ is given by the variational derivative
$$e=\frac{\delta\mathcal{F}(\psi,\theta)}{\delta (1/\theta)}.$$
This yields the expression
$$e(\psi,\theta)=
f(\psi,\theta)-\theta\frac{\partial f(\psi,\theta)}{\partial\theta}+\frac{1}{2}\left(\gamma(\theta)-
\theta\frac{\partial\gamma(\theta)}{\partial\theta}\right)|\nabla\psi|^2.$$
It can be readily checked that the \textsc{Gibbs} relation
$$e(\psi,\theta)=F(\psi,\theta)-\theta\frac{\partial F(\psi,\theta)}{\partial\theta}.$$
holds. If we assume that no mechanical stresses are active, the internal energy $e$ satisfies the conservation law
$$\partial_t e+{\rm div}q=0,$$
where $q$ denotes the heat flux of the system. Following \textsc{Alt \& Pawlow} \cite{AltPaw}, we assume that $q=\nabla\left(\frac{1}{\theta}\right)$, so that the kinetic equation for $e$ reads
$$\partial_t e+\Delta\left(\frac{1}{\theta}\right)=0.$$
Let us now assume that $\gamma(\theta)=\theta$ and $f(\psi,\theta)=\theta\Phi(\psi)-\lambda(\psi)-\theta\log\theta$. In this case we obtain
$e=\theta-\lambda(\psi)$ and
$$\mu=-\Delta\psi+\Phi'(\psi)-\lambda'(\psi)\frac{1}{\theta},$$
hence system \eqref{PF0} for $\vartheta=1/\theta$ and $b(s)=-1/s$, $s>0$. Suppose $(j|\nu)=(q|\nu)=0$ on $\partial\Omega$ with $\nu=\nu(x)$ being the outer unit normal in $x\in\partial\Omega$. This yields the boundary conditions
$\partial_\nu\mu=0$ and $\partial_\nu\vartheta=0$ for the chemical potential $\mu$
and the function $\vartheta$, respectively. Since \eqref{PF0}
is of fourth order with respect to the function $\psi$ we need an
additional boundary condition. An appropriate and classical one from
a variational point of view is $\partial_\nu\psi=0$. Finally, this yields
the initial-boundary value problem
\begin{equation}\label{PF}\begin{split}\partial_t\psi-\Delta\mu=f_1,\quad \mu=-\Delta\psi+\Phi'(\psi)-\lambda'(\psi)\vartheta,&\quad t\in J,\ x\in\Omega,\\
\partial_t\left(b(\vartheta)+\lambda(\psi)\right)-\Delta\vartheta=f_2,&\quad t\in J,\ x\in\Omega,\\
\partial_\nu\mu=g_1,\ \partial_\nu\psi=g_2,\ \partial_\nu\vartheta=g_3,&\quad t\in J,\ x\in\partial\Omega,\\
\psi(0)=\psi_0,\ \vartheta(0)=\vartheta_0,&\quad t=0,\ x\in\Omega,
\end{split}\end{equation}
The functions $f_j, g_j, \psi_0, \vartheta_0, \Phi, \lambda$ and $b$ are
given. Note that if $\theta$ has only a small deviation from a constant value $\theta_*>0$, then the term $1/\theta$ can be linearized around $\theta_*$ and \eqref{PF} turns into the nonisothermal Cahn-Hilliard equation for the order parameter $\psi$ and the relative temperature $\theta-\theta_*$, provided $b(s)=-1/s$.
In the case of the Penrose-Fife equations, \textsc{Brokate \& Sprekels}
\cite{BroSpr} and \textsc{Zheng} \cite{Zhe2} proved global
well-posedness in an $L_2$-setting if the spatial dimension is
equal to 1. \textsc{Sprekels \& Zheng} showed global well-posedness
of the non-conserved equations (that is $\partial_t\psi=-\mu$) in
higher space dimensions in \cite{SprZhe}, a similar result can be
found in the article of \textsc{Laurencot} \cite{Lau}. Concerning
asymptotic behavior we refer to the articles of \textsc{Kubo, Ito
\& Kenmochi} \cite{KubItoKen}, \textsc{Shen \& Zheng}
\cite{SheZhe}, \textsc{Feireisl \& Schimperna} \cite{FeiSchim} and
\textsc{Rocca \& Schimperna} \cite{Rocca1}. The last two authors
studied well-posedness and qualitative behavior of solutions to
the non-conserved Penrose-Fife equations. To be precise, they
proved that each solution converges to a steady state, as time
tends to infinity. \textsc{Shen \& Zheng} \cite{SheZhe}
established the existence of attractors for the non-conserved
equations, whereas \textsc{Kubo, Ito \& Kenmochi} \cite{KubItoKen}
studied the non-conserved as well as the conserved Penrose-Fife
equations. Beside the proof of global well-posedness in the sense
of weak solutions they also showed the existence of a global
attractor. Finally, we want to mention that the physical potential $\Phi$ may also be of logarithmic type, such that $\Phi'(s)$ has singularities at $s=\pm1$. This forces the order parameter to stay in the physically reasonable interval $(-1,1)$, provided that the initial value $\psi(0)=\psi_0\in (-1,1)$. In general, such a result cannot be obtained in the case of the double well potential, since there is no maximum principle available for the fourth order equation $\eqref{PF}_1$. For a result on global existence, uniqueness and asymptotic behaviour of solutions of the \emph{Cahn-Hilliard equation} in case of a logarithmic potential, we refer the reader to \textsc{Abels} \& \textsc{Wilke} \cite{AbWi}. However, in this paper we will only deal with smooth potentials.
In the following sections we will prove well-posedness of
\eqref{PF} for solutions in the maximal $L_p$-regularity classes
$$\psi\in H_p^1(J;L_p(\Omega))\cap L_p(J;H_p^4(\Omega)),$$
$$\vartheta\in H_p^1(J;L_p(\Omega))\cap L_p(J;H_p^2(\Omega)),$$
where $J=[0,T]$, $T>0$.
In Section 2 we investigate a linearized version of \eqref{PF} and prove maximal $L_p$-regularity. Section 3 is devoted to local well-posedness of \eqref{PF}. To this end we apply the contraction mapping principle. In Section 4, we show that the solution exists globally in time, provided that the absolute temperature $\vartheta$ is uniformly bounded from below and above. Finally, in Section 5, we study the asymptotic behavior of the solution to \eqref{PF} as $t\to\infty$. The Lojasiewicz-Simon inequality will play an important role in the analysis.
\section{The Linear Problem}\label{LinProblem}
In this section we deal with a linearized version of \eqref{PF}.
\begin{equation}\begin{split}\label{linPF}
\partial_tu+\Delta^2u+\Delta(\eta_1 v)=f_1,&\quad t\in J,\ x\in\Omega,\\
\partial_tv-a_0\Delta v+\eta_2\partial_tu=f_2,&\quad t\in J,\ x\in\Omega,\\
\partial_\nu\Delta u+\partial_\nu(\eta_1 v)=g_1,&\quad t\in J,\ x\in\partial\Omega,\\
\partial_\nu u=g_2,\ \partial_\nu v=g_3,&\quad t\in J,\ x\in\partial\Omega,\\
u(0)=u_0,\ v(0)=v_0,&\quad t=0,\ x\in\Omega.\end{split}\end{equation} Here
$\eta_1=\eta_1(x),\eta_2=\eta_2(x),a_0=a_0(x)$ are given functions
such that \begin{equation}\label{condeta}\eta_1\in B_{pp}^{4-4/p}(\Omega),\ \eta_2\in B_{pp}^{2-2/p}(\Omega)\quad\text{and}\quad a_0\in C(\overline{\Omega}).\end{equation} We
assume furthermore that $a_0(x)\ge \sigma>0$ for all
$x\in\overline{\Omega}$ and some constant $\sigma>0$. Hence equation
$\eqref{linPF}_2$ does not degenerate. We are interested in
solutions
$$u\in H_p^1(J;L_p(\Omega))\cap L_p(J;H_p^4(\Omega))=:E_1(T)$$
and
$$v\in H_p^1(J;L_p(\Omega))\cap L_p(J;H_p^2(\Omega))=:E_2(T)$$
of \eqref{linPF}. By the well-known trace theorems (cf.
\cite[Theorem 4.10.2]{Ama}) \begin{equation}\label{trace}E_1(T)\hookrightarrow
C(J;B_{pp}^{4-4/p}(\Omega))\quad\text{and}\quad E_2(T)\hookrightarrow
C(J;B_{pp}^{2-2/p}(\Omega)),\end{equation} we necessarily have $u_0\in
B_{pp}^{4-4/p}(\Omega)=:X_{\gamma}^1$, $v_0\in B_{pp}^{2-2/p}(\Omega)=:X_\gamma^2$
and the compatibility conditions
$$\partial_\nu\Delta u_0+\partial_\nu(\eta_1 v_0)=g_1|_{t=0},\quad \partial_\nu u_0=g_2|_{t=0},\quad\text{as well
as}\quad \partial_\nu v_0=g_3|_{t=0},$$ whenever $p>5$, $p>5/3$ and $p>3$, respectively (cf. \cite[Theorem 2.1]{DHP07}). In the sequel we will assume that $p>(n+2)/2$ and $p\ge 2$. This yields the embeddings
$$B_{pp}^{4-4/p}(\Omega)\hookrightarrow H_p^2(\Omega)\cap C^1(\bar{\Omega})\ \mbox{and}\ B_{pp}^{2-2/p}(\Omega)\hookrightarrow H_p^1(\Omega)\cap C(\bar{\Omega}).$$
We are going to prove the following theorem.
\begin{thm}\label{linthm}
Let $n\in\mathbb N$, $\Omega\subset\mathbb R^n$ a bounded domain with boundary $\partial\Omega\in C^4$ and let $p>(n+2)/2$, $p\ge 2$, $p\neq 3,5$. Assume in addition that $\eta_1\in B_{pp}^{4-4/p}(\Omega)$, $\eta_2\in B_{pp}^{2-2/p}(\Omega)$ and $a_0\in C(\bar{\Omega})$, $a_0(x)\ge \sigma>0$ for all $x\in\bar{\Omega}$. Then the linear problem \eqref{linPF} admits a unique solution
$$(u,v)\in H_p^1(J_0;L_p(\Omega)^2)\cap L_p(J_0;(H_p^4(\Omega)\times H_p^2(\Omega))),$$
if and only if the data are subject to the following conditions.
\begin{enumerate}
\item $f_1,f_2\in L_p(J_0;L_p(\Omega))=X(J_0)$,
\item $g_1\in W_p^{1/4-1/4p}(J_0;L_p(\partial\Omega))\cap L_p(J_0;W_p^{1-1/p}(\partial\Omega))=Y_1(J_0)$,
\item $g_2\in W_p^{3/4-1/4p}(J_0;L_p(\partial\Omega))\cap L_p(J_0;W_p^{3-1/p}(\partial\Omega))=Y_2(J_0)$,
\item $g_3\in W_p^{1/2-1/2p}(J_0;L_p(\partial\Omega))\cap L_p(J_0;W_p^{1-1/p}(\partial\Omega))=Y_3(J_0)$,
\item $u_0\in B_{pp}^{4-4/p}(\Omega)=X_{\gamma}^1$, $v_0\in B_{pp}^{2-2/p}(\Omega)=X_\gamma^2$,
\item $\partial_\nu\Delta u_0+\partial_\nu(\eta_1 v_0)=g_1|_{t=0},\ p>5$,
\item $\partial_\nu u_0=g_2|_{t=0}$, $\partial_\nu v_0=g_3|_{t=0},\ p>3$.
\end{enumerate}
\end{thm}
\begin{proof}
Suppose that the function $u\in E_1(T)$ in \eqref{linPF} is already
known. Then in a first step we will solve the linear heat equation
\begin{equation}\label{heateq} \partial_tv-a_0\Delta v=f_2-\eta_2\partial_tu,\end{equation} subject to
the boundary and initial conditions $\partial_\nu v=g_3$ and $v(0)=v_0$.
By the properties of the function $a_0$ we may apply \cite[Theorem 2.1]{DHP07} to obtain a unique solution $v\in E_2(T)$ of \eqref{heateq},
provided that $f_2\in L_p(J\times\Omega)$, $v_0\in
B_{pp}^{2-2/p}(\Omega)$,
$$g_3\in W_p^{1/2-1/2p}(J;L_p(\partial\Omega))\cap L_p(J;W_p^{1-1/p}(\partial\Omega))=:Y_3(J),$$ and
the compatibility condition $\partial_\nu v_0=g_3|_{t=0}$ if $p>3$ is
valid. The solution may then be represented by the variation of
parameters formula
\begin{align}\label{solvth}
v(t)&=v_1(t)-\int_0^te^{-A(t-s)}\eta_2\partial_tu(s)\ ds,
\end{align}
where $A$ denotes the $L_p$-realization of the differential
operator $\mathcal A(x)=-a_0(x)\Delta_N$, $\Delta_N$ means the
Neumann-Laplacian and $e^{-At}$ stands for the bounded analytic
semigroup, which is generated by $-A$ in $L_p(\Omega)$. Furthermore
the function $v_1\in E_2(T)$ solves the linear problem
$$\partial_tv_1-a_0\Delta v_1=f_2,\quad \partial_\nu v_1=g_3,\quad v_1(0)=v_0.$$
We fix a function $w^*\in E_1(T)$ such that $w^*|_{t=0}=u_0$ and make
use of \eqref{solvth} and the fact that $(u-w^*)|_{t=0}=0$ to
obtain
\begin{align*}
v(t)=v_1(t)+v_2(t)-(\partial_t+A)^{-1}\eta_2\partial_t(u-w^*)
\end{align*}
with $v_2(t):=-\int_0^te^{-A(t-s)}\eta_2\partial_t w^*$. Set
$v^*=v_1+v_2\in E_2(T)$ and
$$F(u)=-(\partial_t+A)^{-1}\eta_2\partial_t(u-w^*).$$
Then we may reduce \eqref{linPF} to the problem
\begin{equation}\begin{split}\label{linu}
\partial_tu+\Delta^2u=\Delta G(u)+f_1,&\quad t\in J,\ x\in\Omega,\\
\partial_\nu\Delta u=\partial_\nu G(u)+g_1,&\quad t\in J,\ x\in\partial\Omega,\\
\partial_\nu u=g_2&\quad t\in J,\
x\in\partial\Omega,\\
u(0)=u_0,&\quad t=0,\ x\in\Omega,
\end{split}\end{equation}
where $G(u):=-\eta_1(F(u)+v^*)$. For a given $T\in (0,T_0]$ we set
$$_0E_1(T)=\{u\in E_1(T):u|_{t=0}=0\}$$
and
$$E_0(T):=X(T)\times Y_1(T)\times Y_2(T)$$
$$_0E_0(T):=\{(f,g,h)\in E_0(T):g|_{t=0}=h|_{t=0}=0\},$$
where $X(T):=L_p((0,T)\times\Omega)$,
$$Y_1(T):=W_p^{1/4-1/4p}(0,T;L_p(\partial\Omega))\cap L_p(0,T;W_p^{1-1/p}(\partial\Omega)),$$
and
$$Y_2(T):=W_p^{3/4-1/4p}(0,T;L_p(\partial\Omega))\cap L_p(0,T;W_p^{3-1/p}(\partial\Omega)).$$
The spaces $E_1(T)$ and $E_0(T)$ are endowed with the canonical norms
$|\cdot|_1$ and $|\cdot|_0$, respectively. We introduce the new function $\tilde{u}:=u-w^*\in\! _0E_1(T)$ and we set
$$\tilde{F}(\tilde{u}):=-(\partial_t+A)^{-1}\eta_2\partial_t\tilde{u}$$
as well as $\tilde{G}(\tilde{u}):=-\eta_1\tilde{F}(\tilde{u})$. If $u\in E_1(T)$ is a solution of \eqref{linu}, then the function $\tilde{u}\in\! _0E_1(T)$ solves the problem
\begin{equation}\begin{split}\label{linutilde}
\partial_t\tilde{u}+\Delta^2\tilde{u}=\Delta \tilde{G}(\tilde{u})+\tilde{f}_1,&\quad t\in J,\ x\in\Omega,\\
\partial_\nu\Delta \tilde{u}=\partial_\nu \tilde{G}(\tilde{u})+\tilde{g}_1,&\quad t\in J,\ x\in\partial\Omega,\\
\partial_\nu \tilde{u}=\tilde{g}_2&\quad t\in J,\
x\in\partial\Omega,\\
\tilde{u}(0)=0,&\quad t=0,\ x\in\Omega,
\end{split}\end{equation}
with the modified data
$$\tilde{f}_1:=f_1-\Delta(\eta_1 v^*)-\partial_t w^*-\Delta^2 w^*\in X(T),$$
$$\tilde{g}_1:=g_1-\partial_\nu (\eta v^*)-\partial_\nu\Delta w^*\in\! _0Y_1(T),$$
and
$$\tilde{g}_2:=g_2-\partial_\nu w^*\in\! _0Y_2(T).$$
Observe that by construction we have $\tilde{g}_1|_{t=0}=0$ and $\tilde{g}_2|_{t=0}=0$ if $p>5$ and $p>5/3$, respectively.
Let us estimate the term $\Delta \tilde{G}(u)$ in $L_p(J;L_p(\Omega))$, where $u\in\! _0E_1(T)$. We compute
\begin{multline*}
|\Delta\tilde{G}(u)|_{L_p(J;L_p(\Omega))}\le |\tilde{F}(u)\Delta\eta_1|_{L_p(J;L_p(\Omega))}\\
+2|(\nabla \tilde{F}(u)|\nabla\eta_1)|_{L_p(J;L_p(\Omega))}+|\eta_1\Delta \tilde{F}(u)|_{L_p(J;L_p(\Omega))}.
\end{multline*}
Since $\eta_1\in B_{pp}^{4-4/p}(\Omega)$ does not depend on the variable $t$, we obtain
$$|\tilde{F}(u)\Delta\eta_1|_{L_p(J;L_p(\Omega))}\le |\Delta\eta_1|_{L_p(\Omega)}|\tilde{F}(u)|_{L_p(J;L_\infty(\Omega))},$$
$$|(\nabla \tilde{F}(u)|\nabla\eta_1)|_{L_p(J;L_p(\Omega))}\le |\nabla\eta_1|_{L_\infty(\Omega)}|\nabla\tilde{F}(u)|_{L_p(J;L_p(\Omega))},$$
and
$$|\eta_1\Delta \tilde{F}(u)|_{L_p(J;L_p(\Omega))}\le |\eta_1|_{L_\infty(\Omega)}|\Delta \tilde{F}(u)|_{L_p(J;L_p(\Omega))}.$$
Therefore we have to estimate $\tilde{F}(u)$ for each $u\in\! _0E_1(T)$ in the topology of the spaces $L_p(J;L_\infty(\Omega))$ and $L_p(J;H_p^2(\Omega))$. Let $u\in\! _0E_1$ and recall that $\tilde{F}(u)$ is defined by $\tilde{F}(u)=-(\partial_t+A)^{-1}\eta_2 \partial_t u$. The operator $(\partial_t+A)^{-1}$ is a bounded linear operator from $L_p(J;L_p(\Omega))$ to $_0H_p^1(J;L_p(\Omega))\cap L_p(J;H_p^2(\Omega))=\! _0E_2(T)$. Moreover, by the trace theorem and by Sobolev embedding, it holds that
$$_0H_p^1(J;L_p(\Omega))\cap L_p(J;H_p^2(\Omega))\hookrightarrow C(J;B_{pp}^{2-2/p}(\Omega))\hookrightarrow C(J;C(\bar{\Omega})).$$
Note that the bound of $(\partial_t+A)^{-1}$ as well as the embedding constant do not depend on the length of the interval $J=[0,T]\subset [0,T_0]=J_0$, since the time trace at $t=0$ vanishes. With these facts, we obtain
\begin{align*}
|(\partial_t+A)^{-1}\eta_2 \partial_t u|_{L_p(J;L_\infty(\Omega))}&\le T^{1/p}|(\partial_t+A)^{-1}\eta_2 \partial_tu|_{L_\infty(J;L_\infty(\Omega))}\\
&\le T^{1/p}C|(\partial_t+A)^{-1}\eta_2 \partial_tu|_{E_2(T)}\\
&\le T^{1/p}C|\eta_2\partial_tu|_{L_p(J;L_p(\Omega))}\\
&\le T^{1/p}C|\eta_2|_{L_{\infty}(\Omega)}|u|_{E_1(T)}.
\end{align*}
To estimate $\tilde{F}(u)$ in $L_p(J;H_p^2(\Omega))$ we need another representation of $\tilde{F}(u)$. To be precise, we rewrite $\tilde{F}(u)$ as follows
$$\tilde{F}(u)=-(\partial_t+A)^{-1}\eta_2 \partial_tu=-\partial_t^{1/2}(\partial_t+A)^{-1}\partial_t^{1/2}(\eta_2 u).$$
This is possible, since $u\in\! _0E_1(T)$. Now observe that for each $u\in\! _0E_1$ it holds that $\eta_2 u\in\! _0H_p^{3/4}(J;H_p^1(\Omega))$. This can be seen as follows. First of all, it suffices to show that $\eta_2 u\in L_p(J;H_p^1(\Omega))$, since $\eta_2$ does not depend on the variable $t$. But
\begin{align*}
|\eta_2 u|_{L_p(J;H_p^1(\Omega))}&\le |\eta_2\nabla u|_{L_p(J;L_p(\Omega))}+|u\nabla\eta_2|_{L_p(J;L_p(\Omega))}\\
&\le C\left(|\eta_2|_{L_\infty(\Omega)}|u|_{E_1(T)}+|u|_{L_p(J;L_\infty(\Omega))}|\eta_2|_{H_p^1(\Omega)}\right)\\
&\le C|u|_{E_1(T)}|\eta_2|_{B_{pp}^{2-2/p}(\Omega)},
\end{align*}
and this yields the claim, since
$$u\in\! _0H_p^1(J;L_p(\Omega))\cap L_p(J;H_p^4(\Omega))\hookrightarrow\! _0H_p^{3/4}(J;H_p^1(\Omega)),$$
by the mixed derivative theorem. It follows readily that $\partial_t^{1/2}(\eta_2 u)\in\! _0H_p^{1/4}(J;H_p^1(\Omega))$ and
$$(\partial_t+A)^{-1}(I+A)^{1/2}\partial_t^{1/2}(\eta_2 u)\in\! _0H_p^{5/4}(J;L_p(\Omega))\cap\! _0H_p^{1/4}(J;H_p^2(\Omega)).$$
Since the operator $(I+A)^{1/2}$ with domain $D((I+A)^{1/2})=H_p^1(\Omega)$ commutes with the operator $(\partial_t+A)^{-1}$,
this yields
$$(\partial_t+A)^{-1}\partial_t^{1/2}(\eta_2 u)\in\! _0H_p^{5/4}(J;H_p^1(\Omega))\cap\! _0H_p^{1/4}(J;H_p^3(\Omega))$$
for each fixed $u\in\! _0E_1(T)$. By the mixed derivative theorem we obtain furthermore
$$_0H_p^{5/4}(J;H_p^1(\Omega))\cap\! _0H_p^{1/4}(J;H_p^3(\Omega))\hookrightarrow\! _0H_p^{3/4}(J;H_p^2(\Omega)).$$
Therefore
$$\tilde{F}(u)=-\partial_t^{1/2}(\partial_t+A)^{-1}\partial_t^{1/2}(\eta_2 u)\in\! _0H_p^{1/4}(J;H_p^2(\Omega)),$$
and there exists a constant $C>0$ being independent of $T>0$ and $u\in\! _0E_1(T)$ such that
$$|\tilde{F}(u)|_{H_p^{1/4}(J;H_p^2(\Omega))}\le C|u|_{E_1(T)},$$
for each $u\in\! _0E_1(T)$. In particular this yields the estimate
\begin{align*}
|\tilde{F}(u)|_{L_p(J;H_p^2(\Omega))}&\le T^{1/2p}|\tilde{F}(u)|_{L_{2p}(J;H_p^2(\Omega))}\\
&\le T^{1/2p}|\tilde{F}(u)|_{H_p^{1/4}(J;H_p^2(\Omega))}\le T^{1/2p}C|u|_{E_1(T)},
\end{align*}
by H\"{o}lders inequality and $C>0$ does not depend on the length $T$ of the interval $J$. We have thus shown that
$$|\Delta\tilde{G}(u)|_{L_p(J;L_p(\Omega))}\le \mu_1(T)C|u|_{E_1(T)},$$
where we have set $\mu_1(T):=T^{1/2p}(1+T^{1/2p})$. Observe that $\mu_1(T)\to 0_+$ as $T\to 0_+$. The next step consists of estimating the term $\partial_\nu \tilde{G}(u)$ in $_0W_p^{1/4-1/4p}(J;L_p(\partial\Omega))\cap L_p(J;W_p^{1-1/p}(\partial\Omega))$. To this end, we recall the trace map
$$_0H_p^{1/2}(J;L_p(\Omega))\cap L_p(J;H_p^2(\Omega))\hookrightarrow\! _0W_p^{1/4-1/4p}(J;L_p(\partial\Omega))\cap L_p(J;W_p^{1-1/p}(\partial\Omega))$$
for the Neumann derivative on $\partial\Omega$. Therefore, by the results above, it remains to estimate $\tilde{G}(u)$ in $_0H_p^{1/2}(J;L_p(\Omega))$. By the complex interpolation method we have
$$|w|_{H_p^{1/2}(J;L_p(\Omega))}\le C|w|_{L_p(J;L_p(\Omega))}^{1/2}|w|_{H_p^1(J;L_p(\Omega))}^{1/2}$$
for each $w\in\! _0H_p^1(J;L_p(\Omega))$, and $C>0$ does not depend on $T>0$. Using H\"{o}lders inequality, this yields
\begin{align*}
|w|_{H_p^{1/2}(J;L_p(\Omega))}&\le T^{1/4p}C|w|_{L_{2p}(J;L_p(\Omega))}^{1/2}|w|_{H_p^1(J;L_p(\Omega))}^{1/2}\\
&\le T^{1/4p}C|w|_{H_p^1(J;L_p(\Omega))}.
\end{align*}
Finally we obtain the estimate
$$|\tilde{G}(u)|_{H_p^{1/2}(J;L_p(\Omega))}\le T^{1/2p}|\eta_1|_{L_\infty(\Omega)}C|u|_{\mathbb E_1(T)},$$
which in turn implies
\begin{align*}
|\partial_\nu\tilde{G}(u)|_{Y_1(J)}\le |\tilde{G}(u)|_{H_p^{1/2}(J;L_p(\Omega))}+|\tilde{G(u)}|_{L_p(J;H_p^2(\Omega))}\le \mu_2(T)C|u|_{E_1(T)},
\end{align*}
where $\mu_2(T):=T^{1/4p}(1+T^{1/4p})$ and $\mu_2(T)\to 0_+$ as $T\to 0_+$. Define two operators $L,B:\!_0E_1(T)\to\! _0E_0(T)$ by means of
$$Lu:=\begin{bmatrix}\partial_t u+\Delta^2 u\\\partial_\nu\Delta u\\\partial_\nu u\end{bmatrix}\ \mbox{and}\
Bu:=\begin{bmatrix}\Delta\tilde{G}(u)\\\partial_\nu\tilde{G}(u)\\0\end{bmatrix}.$$
With these definitions, we may rewrite \eqref{linutilde} in the abstract form
$$Lu=Bu+f,\quad f:=(\tilde{f}_1,\tilde{g}_1,\tilde{g}_2)\in\! _0E_0(T).$$
By \cite[Theorem 2.1]{DHP07}, the operator $L$ is bijective with bounded inverse $L^{-1}$, hence $u\in\! _0E_1(T)$ is a solution of \eqref{linutilde} if and only if $(I-L^{-1}B)u=L^{-1}f$. Observe that $L^{-1}B$ is a bounded linear operator from $_0E_1(T)$ to $_0E_1(T)$ and
$$|L^{-1}Bu|_{E_1(T)}\le |L^{-1}|_{\mathcal{B}(E_0(T),E_1(T))}|Bu|_{E_0(T)}\le (\mu_1(T)+\mu_2(T))C|u|_{E_1(T)}.$$
Here the constant $C>0$ as well as the bound of $L^{-1}$ are independent of $T>0$. This shows that choosing $T>0$ sufficiently small, we may apply a Neumann series argument to conclude that \eqref{linutilde} has a unique solution $u\in\! _0E_1(T)$ on a possibly small time interval $J=[0,T]$. Since the linear system \eqref{linutilde} is invariant with respect to time shifts, we may set $J=J_0$.
\end{proof}
\section{Local Well-Posedness}\label{LWP}
In this section we will use the following setting. For $T_0>0$, to
be fixed later, and a given $T\in (0,T_0]$ we define
$$\mathbb E_1(T):=E_1(T)\times E_2(T),\qquad\hspace{0.05cm}_0\mathbb E_1(T):=\{(u,v)\in\mathbb E_1(T):(u,v)|_{t=0}=0\}$$
and
$$\mathbb E_0(T):=X(T)\times X(T)\times Y_1(T)\times Y_2(T)\times Y_3(T),$$
as well as $$_0\mathbb E_0(T):=\{(f_1,f_2,g_1,g_2,g_3)\in\mathbb E_0(T):
g_1|_{t=0}=g_2|_{t=0}=g_3|_{t=0}=0\},$$ with canonical norms
$|\cdot|_1$ and $|\cdot|_0$, respectively. The aim of this section
is to find a local solution $(\psi,\vartheta)\in \mathbb E_1(T)$ of the quasilinear
system
\begin{equation}\label{LWPPF}\begin{split}\partial_t\psi-\Delta\mu=f_1,\quad \mu=-\Delta\psi+\Phi'(\psi)-\lambda'(\psi)\vartheta,&\quad t\in J,\ x\in\Omega,\\
\partial_t\left(b(\vartheta)+\lambda(\psi)\right)-\Delta\vartheta=f_2,&\quad t\in J,\ x\in\Omega,\\
\partial_\nu\mu=g_1,\ \partial_\nu\psi=g_2,\ \partial_\nu\vartheta=g_3,&\quad t\in J,\ x\in\partial\Omega,\\
\psi(0)=\psi_0,\ \vartheta(0)=\vartheta_0,&\quad t=0,\ x\in\Omega.
\end{split}\end{equation}
To this end, we will apply Banach's fixed
point theorem. For this purpose let $p>(n+2)/2$, $p\ge 2$, $f_1,f_2\in X(T_0)$, $g_j\in
Y_j(0,T_0),\ j=1,2,3$, $\psi_0\in X_{\gamma}^1$ and $\vartheta_0\in X_\gamma^2$ be
given such that the compatibility conditions
$$\partial_\nu\Delta\psi_0-\partial_\nu\Phi'(\psi_0)+\partial_\nu(\lambda'(\psi_0)\vartheta_0)=-g_1|_{t=0},\
\partial_\nu\psi_0=g_2|_{t=0}\quad\text{and}\quad
\partial_\nu\vartheta_0=g_3|_{t=0}$$ are satisfied, whenever $p>5$, $p>5/3$ and $p>3$,
respectively. In the sequel we will assume that $\lambda,\phi\in C^{4-}(\mathbb R)$, $b\in C^{3-}(0,\infty)$ and $b'(s)>0$ for all $s>0$. Note that by the Sobolev embedding theorem we have $\vartheta_0\in C(\bar{\Omega})$ as well as $b'(\vartheta_0)\in C(\bar{\Omega})$. Since $\vartheta$ represents the inverse absolute temperature of the system, it is reasonable to assume $\vartheta_0(x)>0$ for all $x\in\bar{\Omega}$. Therefore, there exists a constant $\sigma>0$ such that $\vartheta_0(x),b'(\vartheta_0(x))\ge \sigma>0$ for all $x\in\bar{\Omega}$. We define $a_0(x):=1/b'(\vartheta_0(x))$, $\eta_1(x)=\lambda'(\psi_0(x))$ and $\eta_2(x)=a_0(x)\eta_1(x)$. By assumption, it holds that $a_0\in B_{pp}^{2-2/p}(\Omega)$, $\eta_1\in B_{pp}^{4-4/p}(\Omega)$ and $\eta_2\in B_{pp}^{2-2/p}(\Omega)$, cf. \cite[Section 4.6 \& Section 5.3.4]{RuSi96}.
Thanks to Theorem \ref{linthm} we
may define a pair of functions $(u^*,v^*)\in \mathbb E_1(T_0)$ as the solution
of the problem \begin{equation}\begin{split}\label{linfix}
\partial_tu^*+\Delta^2u^*+\Delta(\eta_1v^*)=f_1,&\quad t\in [0,T_0],\ x\in\Omega,\\
\partial_tv^*-a_0\Delta v^*+\eta_2\partial_tu^*=a_0f_2,&\quad t\in [0,T_0],\ x\in\Omega,\\
\partial_\nu\Delta u^*+\partial_\nu(\eta_1 v^*)=-g_1-e^{-B^2t}g_0,&\quad t\in [0,T_0],\ x\in\partial\Omega,\\
\partial_\nu u^*=g_2,&\quad t\in [0,T_0],\ x\in\partial\Omega,\\
\partial_\nu v^*=g_3,&\quad t\in [0,T_0],\ x\in\partial\Omega,\\
u^*(0)=\psi_0,\ v^*(0)=\vartheta_0,&\quad t=0,\ x\in\Omega,
\end{split}\end{equation}
where $B=-\Delta_{\partial\Omega}$ is the Laplace-Beltrami operator on $\partial\Omega$ and
$e^{-B^2t}$ is the analytic semigroup which is generated by $-B^2$.
Furthermore $g_0=0$ if $p<5$ and
$g_0=-g_1|_{t=0}-(\partial_\nu\Delta\psi_0+\partial_\nu(\eta_1\vartheta_0))$ if $p>5$.
Define a linear operator $\mathbb L:\!_0\mathbb E_1(T_0)\to\!_0\mathbb E_0(T_0)$ by
$$\mathbb L(u,v)=\vectfuenf{\partial_tu+\Delta^2u+\eta_1\Delta v}{\partial_tv-a_0\Delta v+\eta_2\partial_tu}{\partial_\nu\Delta u+\partial_\nu(\eta_1 v)}{\partial_\nu u}{\partial_\nu v}.$$
Then, by Theorem \ref{linthm}, the
operator $\mathbb L:\!_0\mathbb E_1(T_0)\to\!_0\mathbb E_0(T_0)$ is bounded and
bijective, hence an isomorphism with bounded inverse $\mathbb L^{-1}$.
For all $(u,v)\in\!_0\mathbb E_1(T)$ we set
$$G_1(u,v)=(\lambda'(\psi_0)-\lambda'(u)) v+\Phi'(u),$$
$$G_2(u,v)=(a_0\lambda'(\psi_0)-a(v)\lambda'(u))\partial_tu-(a_0-a(v))\Delta v-(a_0-a(v))f_2,$$
where $a(v(t,x))=1/b'(v(t,x))$ and $a_0=a(\vartheta_0)$. Lastly we
define a nonlinear mapping
$G:\mathbb E_1(T)\times\!_0\mathbb E_1(T)\to\!_0\mathbb E_0(T)$ by
$$G((u^*,v^*);(u,v))=\vectfuenf{\Delta G_1(u+u^*,v+v^*)}{G_2(u+u^*,v+v^*)}{\partial_\nu G_1(u+u^*,v+v^*)-\tilde{g}_0}{0}{0},$$
where $\tilde{g}_0=0$ if $p<5$ and $\tilde{g}_0=e^{-B^2 t}\partial_\nu
G_1(\psi_0,\vartheta_0)$ if $p>5$. Then it is easy to see that
$\psi=u+u^*\in E_1(T)$ and $\vartheta=v+v^*\in E_2(T)$ is a solution of \eqref{PF} if and
only if
$$\mathbb L(u,v)=G((u^*,v^*);(u,v))$$
or equivalently
$$(u,v)=\mathbb L^{-1}G((u^*,v^*);(u,v)).$$
In order to apply the contraction mapping principle we consider a
ball $\mathbb B_R=\mathbb B_R^1\times\mathbb B_R^2\subset\!_0\mathbb E_1(T)$, where
$R\in (0,1]$. Furthermore we define a mapping
$\mathcal T:\mathbb B_R\to\!_0\mathbb E_1(T)$ by
$\mathcal T(u,v)=\mathbb L^{-1}G((u^*,v^*);(u,v))$. We shall prove that
$\mathcal T\mathbb B_R\subset\mathbb B_R$ and that $\mathcal T$ defines a strict
contraction on $\mathbb B_R$. To this end we define the shifted ball
$\mathbb B_R(u^*,v^*)=\mathbb B_R^1(u^*)\times\mathbb B_R^2(v^*)\subset\mathbb E_1(T)$ by
$$\mathbb B_R(u^*,v^*)=\{(u,v)\in\mathbb E_1(T):(u,v)=(\tilde{u},\tilde{v})+(u^*,v^*),\ (\tilde{u},\tilde{v})\in\mathbb B_R\}.$$
To ensure that the mapping $G_2$ is well defined, we choose
$T_0>0$ and $R>0$ sufficiently small. This yields that all
functions $v\in\mathbb B_R^2(v^*)$ have only a small deviation from the
initial value $\vartheta_0$. To see this, write
$$|\vartheta_0(x)-v(t,x)|\le |\vartheta_0(x)-v^*(t,x)|+|v^*(t,x)-v(t,x)|\le
\mu(T)+R,$$ for all functions $v\in\mathbb B_R^2(v^*)$, where
$\mu=\mu(T)$ is defined by
$$\mu(T)=\max_{(t,x)\in[0,T]\times\Omega}|v^*(t,x)-\vartheta_0(x)|.$$
Observe that $\mu(T)\to 0$ as $T\to 0$, by the continuity of $v^*$
and $\vartheta_0$. This in turn implies that $v(t,x)\ge\sigma/2>0$ and $b'(v(t,x))\ge\sigma/2>0$ for $(t,x)\in [0,T]\times\bar{\Omega}$ and all $v\in\mathbb B_R^2(v^*)$, with $T_0>0$, $R>0$ being sufficiently small.
Moreover, for all $v,\bar{v}\in\mathbb B_R^2(v^*)$ we obtain the
estimates \begin{equation}\label{lipscha1}|a(\vartheta_0(x))-a(v(t,x))|\le
C|\vartheta_0(x)-v(t,x)|\end{equation} and
\begin{equation}\label{lipscha2}|a(\bar{v}(t,x))-a(v(t,x))|\le
C|\bar{v}(t,x)-v(t,x)|,\end{equation} valid for all $(t,x)\in
[0,T]\times\bar{\Omega}$, with some constant $C>0$, since $b'$ is locally
Lipschitz continuous.
The next proposition provides all the facts to show the
desired properties of the operator $\mathcal T$.
\begin{prop}\label{fixpointest}
Let $n\in\mathbb N$ and $p>(n+2)/2$, $p\ge 2$, $b\in C^{2-}(0,\infty)$, $b'(s)>0$ for all $s>0$, $\lambda,\Phi\in C^{4-}(\mathbb R)$ and $\vartheta_0(x)>0$ for all $x\in\bar{\Omega}$. Then there
exists a constant $C>0$, independent of $T$, and functions
$\mu_j=\mu_j(T)$ with $\mu_j(T)\to 0$ as $T\to 0$, such that for
all $(u,v),(\bar{u},\bar{v})\in\mathbb B_R(u^*,v^*)$ the following
statements hold.
\begin{enumerate}
\setlength{\itemsep}{1ex}
\item $|\Delta G_1(u,v)-\Delta G_1(\bar{u},\bar{v})|_{X(T)}\le (\mu_1(T)+R)|(u,v)-(\bar{u},\bar{v})|_{\mathbb E_1(T)}$,
\item $|G_2(u,v)-G_2(\bar{u},\bar{v})|_{X(T)}\le C(\mu_2(T)+R)|(u,v)-(\bar{u},\bar{v})|_{\mathbb E_1(T)}$,
\item $|\partial_\nu G_1(u,v)-\partial_\nu G_1(\bar{u},\bar{v})|_{Y_1(T)}\le
C(\mu_3(T)+R)|(u,v)-(\bar{u},\bar{v})|_{\mathbb E_1(T)}$.
\end{enumerate}
\end{prop}
The proof is given in the Appendix.\vspace{0.2cm}
\noindent It is now easy to verify the self-mapping property of
$\mathcal T$. Let $(u,v)\in \mathbb B_R$. By Proposition \ref{fixpointest}
there exists a function $\mu=\mu(T)$ with $\mu(T)\to 0$ as $T\to
0$ such that
\begin{align*}
|\mathcal T(u,v)|_1&=|\mathbb L^{-1}G((u^*,v^*),(u,v))|_1\le |\mathbb L^{-1}||G((u^*,v^*),(u,v))|_0\\
&\le C(|G((u^*,v^*),(u,v))-G((u^*,v^*),(0,0))|_0+|G((u^*,v^*),(0,0))|_0)\\
&\le
C(|\Delta G_1(u+u^*,v+v^*)-\Delta G_1(u^*,v^*)|_{X(T)}\\
&\hspace{1cm}+|G_2(u+u^*,v+v^*)-G_2(u^*,v^*)|_{X(T)}\\
&\hspace{1cm}+|\partial_\nu G_1(u+u^*,v+v^*)-\partial_\nu G_1(u^*,v^*)|_{Y_1(T)}\\
&\hspace{1cm}+|G((u^*,v^*),(0,0))|_0)\\
&\le C(\mu(T)+R)|(u,v)|_1+|G((u^*,v^*),(0,0))|_0\\
&\le C(\mu(T)+R)R+|G((u^*,v^*),(0,0))|_0.
\end{align*}
Hence we see that $\mathcal T\mathbb B_R\subset\mathbb B_R$ if $T$ and $R$ are
sufficiently small, since $G((u^*,v^*),(0,0))$ is a fixed
function. Furthermore for all $(u,v),(\bar{u},\bar{v})\in\mathbb B_R$ we
have
\begin{align*}
|\mathcal T(u,v)-\mathcal T(\bar{u},\bar{v})|_1&=|\mathbb L^{-1}(G((u^*,v^*),(u,v))-G((u^*,v^*),(\bar{u},\bar{v})))|_1\\
&\le |\mathbb L^{-1}||G((u^*,v^*),(u,v))-G((u^*,v^*),(\bar{u},\bar{v}))|_0\\
&\le
C(|\Delta G_1(u+u^*,v+v^*)-\Delta G_1(\bar{u}+u^*,\bar{v}+v^*)|_{X(T)}\\
&\hspace{1cm}+|\partial_\nu G_1(u+u^*,v+v^*)-\partial_\nu G_1(\bar{u}+u^*,\bar{v}+v^*)|_{Y_1(T)}\\
&\hspace{1cm}+|G_2(u+u^*,v+v^*)-G_2(\bar{u}+u^*,\bar{v}+v^*)|_{X(T)})\\
&\le C(\mu(T)+R)|(u,v)-(\bar{u},\bar{v})|_1.
\end{align*}
Thus $\mathcal T$ is a strict contraction on $\mathbb B_R$, if $T$ and $R$ are
again small enough. Therefore we may apply the contraction mapping
principle to obtain a unique fixed point
$(\tilde{u},\tilde{v})\in\mathbb B_R$ of $\mathcal T$. In other words the
pair $(\psi,\vartheta)=(\tilde{u}+u^*,\tilde{v}+v^*)\in\mathbb E_1(T)$ is
the unique local solution of \eqref{PF}. We summarize the
preceding calculations in
\begin{thm}\label{LWPsolPF}
Let $n\in\mathbb N$, $p>(n+2)/2$, $p\ge 2$, $p\neq 3,5$, $b\in C^{3-}(0,\infty)$, $b'(s)>0$ for all $s>0$ and let $\lambda,\Phi\in C^{4-}(\mathbb R)$. Then there exists an interval
$J=[0,T]\subset [0,T_0]=J_0$ and a unique solution $(\psi,\vartheta)$ of
\eqref{PF} on $J$, with
$$\psi\in H_p^1(J;L_p(\Omega))\cap L_p(J;H_p^4(\Omega))$$
and
$$\vartheta\in H_p^1(J;L_p(\Omega))\cap L_p(J;H_p^2(\Omega)),\quad \vartheta(t,x)>0\ \mbox{for all}\ (t,x)\in J\times\bar{\Omega},$$
provided the data are subject to the following conditions.
\begin{enumerate}
\item $f_1,f_2\in L_p(J_0\times\Omega)$,
\item $g_1\in W_p^{1/4-1/4p}(J_0;L_p(\partial\Omega))\cap
L_p(J_0;W_p^{1-1/p}(\partial\Omega))$,
\item $g_2\in W_p^{3/4-1/4p}(J_0;L_p(\partial\Omega))\cap
L_p(J_0;W_p^{3-1/p}(\partial\Omega))$,
\item $g_3\in W_p^{1/2-1/2p}(J_0;L_p(\partial\Omega))\cap
L_p(J_0;W_p^{1-1/p}(\partial\Omega))$,
\item $\psi_0\in B_{pp}^{4-4/p}(\Omega)$, $\vartheta_0\in B_{pp}^{2-2/p}(\Omega)$,
\item $\partial_\nu\Delta \psi_0-\partial_\nu\Phi'(\psi_0)+\partial_\nu(\lambda'(\psi_0) \vartheta_0)=-g_1|_{t=0},$ if
$p>5$,
\item $\partial_\nu\psi_0=g_2|_{t=0}$, $\partial_\nu \vartheta_0=g_3|_{t=0},$ if $p>3$,
\item $\vartheta_0(x)>0$ for all $x\in\bar{\Omega}$.
\end{enumerate}
The solution depends continuously on the given data and if the
data are independent of $t$, the map $(\psi_0,\vartheta_0)\mapsto
(\psi,\vartheta)$ defines a local semiflow on the natural (nonlinear) phase
manifold
$$\mathcal M_p:=\{(\psi_0,\vartheta_0)\in B_{pp}^{4-4/p}(\Omega)\times B_{pp}^{2-2/p}(\Omega): \psi_0\ \text{and}\ \vartheta_0\ \text{satisfy}\ 6.-8.\}.$$
\end{thm}
\section{Global Well-Posedness}\label{SECPFGWP}
In this section we will investigate the global existence of the
solution to the conserved Penrose-Fife type system
\begin{equation}\label{PFGWP}\begin{split}\partial_t\psi-\Delta\mu=0,\quad \mu=-\Delta\psi+\Phi'(\psi)-\lambda'(\psi)\vartheta,&\quad t>0,\ x\in\Omega,\\
\partial_t\left(b(\vartheta)+\lambda(\psi)\right)-\Delta\vartheta=0,&\quad t>0,\ x\in\Omega,\\
\partial_\nu\mu=0,\ \partial_\nu\psi=0,\ \partial_\nu\vartheta=0,&\quad t>0,\ x\in\partial\Omega,\\
\psi(0)=\psi_0,\ \vartheta(0)=\vartheta_0,&\quad t=0,\ x\in\Omega,
\end{split}\end{equation}
with respect to time if the spatial dimension $n$ is less or equal to 3.
Note that the boundary conditions are equivalent to
$\partial_\nu\vartheta=\partial_\nu\psi=\partial_\nu\Delta\psi=0$. A successive application of Theorem \ref{LWPsolPF} yields
a maximal interval of existence $J_{\max}=[0,T_{\max})$ for the
solution $(\psi,\vartheta)\in E_1(T)\times E_2(T)$ of \eqref{PFGWP}, where $T\in(0,T_{\max})$. In the
sequel we will make use of the following assumptions.
\begin{itemize}
\setlength{\itemsep}{0.5ex}
\item[\textbf{(H1)}] $\Phi\in C^{4-}(\mathbb{R})$ and there exist some constants $c_j>0$, $\gamma\ge 0$
such that
$$\Phi(s)\ge-\frac{\eta}{2}s^2-c_1,\ |\Phi'''(s)|\le c_2(1+|s|^{\gamma}),$$
for all $s\in\mathbb R$, where $\eta<\lambda_1$ with $\lambda_1$ being the
smallest nontrivial eigenvalue of the negative Laplacian on $\Omega$
with Neumann boundary conditions and $\gamma<3$ if $n=3$.
\item[\textbf{(H2)}] $\lambda\in C^{4-}(\mathbb{R})$ and $\lambda'',\lambda'''\in L_\infty(\mathbb R)$. In particular, there is a constant $c>0$ such that
$|\lambda'(s)|\le c(1+|s|)$ for all $s\in\mathbb R$.
\item[\textbf{(H3)}] $b\in C^{3-}((0,\infty))$, $b'(s)>0$ on $(0,\infty)$ and there is
a constant $\kappa>1$ such that $$\frac{1}{\kappa}\le\vartheta(t,x)\le\kappa$$ on
$J_{\max}\times\Omega$. In particular, there exists $\sigma>1$ such that
$$\frac{1}{\sigma}\le b'(\vartheta(t,x))\le\sigma,$$
on $J_{\max}\times\Omega$.
\end{itemize}
\emph{Remark:} Condition (H1) is certainly fulfilled, if $\Phi$ is
a polynomial of degree $2m$, $m<3$.\vspace{0.25cm}
\noindent We prove global well-posedness with respect to time by contradiction. For
this purpose, assume that $T_{\max}<\infty$. Multiply
$\partial_t\psi=\Delta\mu$ by $\mu$ and integrate by parts to the result
\begin{equation}\label{eqpsi}\frac{d}{dt}\left(\frac{1}{2}|\nabla\psi|_{2}^2+\int_\Omega\Phi(\psi)\
dx\right)+|\nabla\mu|_{2}^2-\int_\Omega\lambda'(\psi)\vartheta\partial_t\psi\
dx=0.\end{equation} Next we multiply $\eqref{PFGWP}_2$ by $\vartheta$ and
integrate by parts. This yields \begin{equation}\label{eqtheta}\int_\Omega\vartheta
b'(\vartheta)\partial_t\vartheta\
dx+|\nabla\vartheta|_{2}^2+\int_\Omega\lambda'(\psi)\vartheta\partial_t\psi\ dx=0.\end{equation}
Set $\beta'(s)=sb'(s)$ and add \eqref{eqpsi} to \eqref{eqtheta} to
obtain the equation
\begin{align}\label{eneq}\begin{split}\frac{d}{dt}\Big(\frac{1}{2}|\nabla\psi|_{2}^2+
\int_\Omega\Phi(\psi)\ dx+\int_\Omega\beta(\vartheta)\
dx\Big)+|\nabla\mu|_{2}^2+|\nabla\vartheta|_{2}^2=0.\end{split}\end{align}
Integrating \eqref{eneq} with respect to $t$, we obtain
\begin{equation}\label{PFenineq2}
E(\psi(t),\vartheta(t))+\int_0^t\left(|\nabla\mu(s)|_{2}^2+|\nabla\vartheta(s)|_{2}^2\right)\ dt=E(\psi_0,\vartheta_0),\end{equation}
for all $t\in J_{\max}$, where
$$E(u,v):=\frac{1}{2}|\nabla u|_{2}^2+\int_\Omega\Phi(u)\
dx+\int_\Omega\beta(v)\ dx.$$ It follows from (H1) and the
Poincar\'{e}-Wirtinger inequality that
\begin{multline*}\frac{\varepsilon}{2}\int_\Omega|\nabla\psi(t)|^2\
dx+\frac{1-\varepsilon}{2}\int_\Omega|\nabla\psi(t)|^2\
dx+\int_\Omega\Phi(\psi(t))\ dx\\
\ge\frac{\varepsilon}{2}\int_\Omega|\nabla\psi(t)|^2\ dx+
\frac{(1-\varepsilon)\lambda_1-\eta}{2}|\psi(t)|_2^2-c_1|\Omega|-\frac{\lambda_1}{2|\Omega|}\left(\int_{\Omega}\psi_0\
dx\right),\end{multline*} since by equation $\partial_t\psi=\Delta\mu$ and
the boundary condition $\partial_\nu\mu=0$, it holds that
$$\int_\Omega\psi(t,x)\ dx\equiv \int_\Omega\psi_0(x)\ dx,\quad t\in J_{\max}.$$
Hence for a sufficiently small $\varepsilon>0$ we obtain the a priori
estimates \begin{equation}\label{enest2}\psi\in
L_\infty(J_{\max};H_2^1(\Omega))\quad\text{and}\quad
|\nabla\mu|,|\nabla\vartheta|\in L_2(J_{\max};L_2(\Omega)),\end{equation} since
$\beta(\vartheta(t,x))$ is uniformly bounded on $J_{\max}\times\Omega$, by
(H3). However, things are more involved for higher order estimates.
Here we have the following result.
\begin{prop}\label{estpsit}
Let $n\le 3$, $p>(n+2)/2$, $p\ge 2$ and let $(\psi,\vartheta)$ be the
maximal solution of \eqref{PFGWP} with initial value
$\psi_0\in B_{pp}^{4-4/p}(\Omega)$ and $\vartheta_0\in B_{pp}^{2-2/p}(\Omega)$. Suppose furthermore $b\in C^{3-}(0,\infty)$, $b'(s)>0$ for all $s>0$, $\lambda,\Phi\in C^{4-}(\mathbb R)$ and let (H1)-(H3) hold.
Then $\psi\in L_\infty(J_{\max}\times\Omega)$ and $\vartheta\in
H_2^1(J_{\max};L_2(\Omega))\cap L_\infty(J_{\max};H_2^1(\Omega))$.
Moreover, it holds that $\partial_t\psi\in L_r(J_{\max}\times\Omega)$,
where $r:=\min\{p,2(n+4)/n\}$.
\end{prop}
\bpr The proof is given in the Appendix.
\end{proof} \noindent Define the new function $u=b(\vartheta)$. Then u
satisfies the nonautonomous linear differential equation in divergence form
\begin{equation}\label{eqhoelder}\partial_t u-\diver (a(t,x)\nabla u)=f,\end{equation}
subject to the boundary and initial conditions $\partial_\nu u=0$ and
$u(0)=b(\vartheta_0)=:u_0$, where $a(t,x):=1/b'(\vartheta(t,x))$ and
$f:=-\lambda'(\psi)\partial_t\psi$. With (H3), the regularity of $\vartheta$ from
Proposition \ref{estpsit} carries over to the function $u$; in particular $u_0\in B_{pp}^{2-2/p}(\Omega)$. This yields, that $u$ is a
\emph{weak solution} of \eqref{eqhoelder} in the sense of
\textsc{Lieberman} \cite{Lieb96} \& \textsc{DiBenedetto} \cite{DiBen93}, and $u$ is bounded by (H3).
Furthermore, by (H3)
$$0<\frac{1}{\sigma}\le a(t,x)\le \sigma<\infty,$$
for all $(t,x)\in J_{\max}\times\Omega$. Note that by Proposition \ref{estpsit} it holds that $f=-\lambda'(\psi)\partial_t\psi\in
L_r(J_{\max}\times\Omega)$, $r:=\min\{p,2(n+4)/n\}$. Consider the case
$r=2(n+4)/n$. Then it can be readily checked that
$$\frac{n+2}{2}<\frac{2(n+4)}{n}=r$$
provided $n\le 5$. It follows from \textsc{Lieberman} \cite{Lieb96} \& \textsc{DiBenedetto} \cite{DiBen93} that there exists a real number $\alpha\in
(0,1/2)$ such that $u\in C^{\alpha,2\alpha}(\overline{\Omega_{T_{\max}}})$,
provided $f\in L_p(J_{\max}\times\Omega)$ and $p>(n+2)/2$. Here
$C^{\alpha,2\alpha}(\overline{\Omega_{T_{\max}}})$ is defined as
$$C^{\alpha,2\alpha}(\overline{\Omega_{T_{\max}}}):=\{v\in
C(\overline{\Omega_{T_{\max}}}):\sup_{(t,x),(s,y)\in
\overline{\Omega_{T_{\max}}}}\frac{|v(t,x)-v(s,y)|}{|t-s|^\alpha+|x-y|^{2\alpha}}<\infty\}.$$
and we have set $\Omega_{T_{\max}}=(0,T_{\max})\times\Omega$. The properties of the function $b$
then yield that $\vartheta=b^{-1}(u)\in C^{\alpha,2\alpha}(\overline{\Omega_{T_{\max}}})$.
In a next step we solve the
initial-boundary value problem
\begin{align}\label{PFGWPEQVTH}\begin{split}
\partial_t\vartheta-a(t,x)\Delta\vartheta=g,&\quad
t\in J_{\max},\ x\in\Omega,\\
\partial_\nu\vartheta=0,&\quad t\in J_{\max},\ x\in\partial\Omega,\\
\vartheta(0)=\vartheta_0,&\quad t=0,\ x\in\Omega, \end{split}\end{align} with
$g:=-a(t,x)\lambda'(\psi)\partial_t\psi\in L_r(J_{\max}\times\Omega)$ and
$r=2(n+4)/n>(n+2)/2$. By \cite[Theorem 2.1]{DHP07} we obtain
$$\vartheta\in H_r^1(J_{\max};L_r(\Omega))\cap L_r(J_{\max};H_r^2(\Omega)),$$
of \eqref{PFGWPEQVTH}, since
$$\vartheta_0\in
B_{pp}^{2-2/p}(\Omega)\hookrightarrow B_{rr}^{2-2/r}(\Omega),\quad p\ge
r.$$ At this point we use equation \eqref{psit} from the proof of
Proposition \ref{estpsit} to conclude $\partial_t\psi\in
L_s(J_{\max}\times\Omega)$, with $s=\min\{p,q\}$ where $q$ is
restricted by
$$\frac{1}{q}\ge\frac{1}{r}-\frac{2}{n+4}.$$
For the case $r=2(n+4)/n$, this yields
$$\frac{1}{q}\ge\frac{n-4}{2(n+4)},$$
i.e. $q$ may be arbitrarily large in case $n\le 3$ and we may set
$s=p$. Now we solve \eqref{PFGWPEQVTH} again, this time with $g\in
L_p(J_{\max}\times\Omega)$, to obtain
$$\vartheta\in H_p^1(J_{\max};L_p(\Omega))\cap L_p(J_{\max};H_p^2(\Omega))$$
and therefore $\vartheta(T_{\max})\in B_{pp}^{2-2/p}(\Omega)$ is well
defined. Next, consider the equation
$$\partial_t\psi+\Delta^2\psi=\Delta\Phi'(\psi)-\Delta(\lambda'(\psi)\vartheta),$$
subject to the initial and boundary conditions $\psi(0)=\psi_0$
and $\partial_\nu\psi=\partial_\nu\Delta\psi=0$. By maximal $L_p$-regularity there
exists a constant $M=M(J_{\max})>0$ such that
\begin{equation}\label{MRestPF}|\psi|_{E_1(T)}\le
M(1+|\Delta\Phi'(\psi)|_{X(T)}+|\Delta(\lambda'(\psi)\vartheta)|_{X(T)}).
\end{equation}
for each $T\in J_{\max}$. Since $\vartheta\in E_2(T_{\max})$ we may apply \cite[Lemma 4.1]{PrWi06} to the result
\begin{equation}\label{MRestPF2}
|\Delta\Phi'(\psi)|_{X(T)}+|\Delta(\lambda'(\psi)\vartheta)|_{X(T)}\le C(1+|\psi|_{E_1(T)}^\delta),
\end{equation}
with some $\delta\in (0,1)$ and $C>0$ being independent of $T\in J_{\max}$. Combining \eqref{MRestPF} with \eqref{MRestPF2}, we obtain the estimate
$$|\psi|_{E_1(T)}\le C(1+|\psi|_{E_1(T)}^\delta),$$
which in turn yields that $|\psi|_{E_1(T)}$ is bounded as $T\nearrow T_{\max}$, since $\delta\in (0,1)$. Therefore the value $\psi(T_{\max})\in B_{pp}^{4-4/p}(\Omega)$ is well defined and we may continue the solution $(\psi,\vartheta)$ beyond the point $T_{\max}$, contradicting the assumption that $J_{\max}=[0,T_{\max})$ is the maximal interval of existence.
We summarize these considerations in
\begin{thm}\label{GWPsolPF}
Let $n\le 3$, $p>(n+2)/2$, $p\ge 2$ and $p\neq 3,5$. Assume that (H1)-(H3) hold. Then for each $T_0>0$ there exists a unique
solution
$$\psi\in H_p^1(J_0;L_p(\Omega))\cap L_p(J_0;H_p^4(\Omega))=E_1(T_0)$$
and
$$\vartheta\in H_p^1(J_0;L_p(\Omega))\cap L_p(J_0;H_p^2(\Omega))=E_2(T_0),$$
of \eqref{PF}, provided the data are subject to the following conditions.
\be
\item $\psi_0\in B_{pp}^{4-4/p}(\Omega)$, $\vartheta_0\in B_{pp}^{2-2/p}(\Omega)$;
\item $\partial_\nu\Delta \psi_0=0,$ if
$p>5$, $\partial_\nu\psi_0=0$;
\item $\partial_\nu \vartheta_0=0,$ if $p>3$, $\vartheta_0(x)>0$ for all $x\in\bar{\Omega}$.
\end{enumerate} The solution depends continuously on the given data and the
map $(\psi_0,\vartheta_0)\mapsto (\psi,\vartheta)$ defines a semiflow
on the natural phase manifold
$$\mathcal M_p:=\{(\psi_0,\vartheta_0)\in B_{pp}^{4-4/p}(\Omega)\times B_{pp}^{2-2/p}(\Omega): \psi_0\ \text{and}\ \vartheta_0\ \text{satisfy}\ 2.\ \&\ 3.\}.$$
\end{thm}
\section{Asymptotic Behavior}\label{asymbeh}
Let $n\le3$. In the following we will investigate the asymptotic
behavior of global solutions of the homogeneous system
\begin{equation}\label{PFasym}\begin{split}\partial_t\psi-\Delta\mu=0,\quad \mu=-\Delta\psi+\Phi'(\psi)-\lambda'(\psi)\vartheta,&\quad t>0,\ x\in\Omega,\\
\partial_t\left(b(\vartheta)+\lambda(\psi)\right)-\Delta\vartheta=0,&\quad t>0,\ x\in\Omega,\\
\partial_\nu\mu=0,&\quad t>0,\ x\in\partial\Omega,\\
\partial_\nu\psi=0,&\quad t>0,\ x\in\partial\Omega,\\
\partial_\nu\vartheta=0,&\quad t>0,\ x\in\partial\Omega,\\
\psi(0)=\psi_0,\ \vartheta(0)=\vartheta_0,&\quad t=0,\ x\in\Omega,
\end{split}\end{equation}
as $t\to \infty$. To this end let $(\psi_0,\vartheta_0)\in\mathcal M_p$, $p>(n+2)/2$, $p\ge 2$ and denote
by $(\psi(t),\vartheta(t))$ the unique global solution of
\eqref{PFasym}. In the sequel we will make use of the following
assumptions.
\begin{itemize}
\setlength{\itemsep}{0.5ex}
\item[\textbf{(H4)}]$b\in C^{3-}((0,\infty))$, $b'(s)>0$ on $(0,\infty)$ and there is
a constant $\kappa>1$ such that $$\frac{1}{\kappa}\le\vartheta(t,x)\le\kappa$$ on
$J_{\max}\times\Omega$. In particular, there exists $\sigma>1$ such that
$$\frac{1}{\sigma}\le b'(\vartheta(t,x))\le\sigma,$$
on $J_{\max}\times\Omega$.
\item[\textbf{(H5)}] The functions $\Phi$, $\lambda$ and $b$ are real
analytic on $\mathbb R$.
\end{itemize}
We remark that assumption (H4) is identical to (H3) for a global solution. We stated it here for the sake of readability.
Note that the boundary conditions $\eqref{PFasym}_{3,5}$ yield
$$\int_\Omega \psi(t,x)\ dx\equiv\int_\Omega \psi_0(x)\
dx,$$ and $$\int_\Omega(b(\vartheta(t,x))+\lambda(\psi(t,x)))\ dx\equiv
\int_\Omega(b(\vartheta_0(x))+\lambda(\psi_0(x)))\ dx.$$ Replacing $\psi$ by
$\tilde{\psi}=\psi-c$, where $c:=\frac{1}{|\Omega|}\int_\Omega
\psi_0(x)\ dx$ we see that $\int_\Omega\tilde{\psi}\ dx \equiv 0$, if
$\Phi(s)$ and $\lambda(s)$ are replaced by $\tilde{\Phi}(s)=\Phi(s+c)$
and $\tilde{\lambda}(s)=\lambda(s+c)$, respectively. Similarly we can
achieve that
$$\int_\Omega(b(\vartheta(t,x))+\lambda(\psi(t,x)))\ dx\equiv 0,$$
by a shift of $\lambda$, to be precise $\bar{\lambda}(s):=\lambda(s)-d$,
where
$$d:=\frac{1}{|\Omega|}\int_\Omega(b(\vartheta_0(x))+\lambda(\psi_0(x)))\ dx.$$
With these modifications of the data we obtain the constraints
\begin{equation}\label{sidecond}\int_\Omega \psi(t,x)\
dx\equiv0\quad\text{and}\quad
\int_\Omega(b(\vartheta(t,x))+\lambda(\psi(t,x)))\ dx\equiv 0.\end{equation} Recall from
Section \ref{SECPFGWP} the energy functional
$$E(u,v)=\frac{1}{2}|\nabla u|_{2}^2+\int_\Omega\Phi(u)\
dx+\int_\Omega\beta(v)\ dx,$$ defined on the energy space
$V=V_1\times V_2$, where
$$V_1:=\left\{u\in H_2^1(\Omega):\int_\Omega u\
dx=0\right\},\qquad V_2:=H_2^r(\Omega),\ r\in (n/4,1).$$ and $V$ is
equipped with the canonical norm
$|(u,v)|_V:=|u|_{H_2^1(\Omega)}+|v|_{H_2^r( \Omega)}$. It is convenient
to embed $V$ into a Hilbert space $H=H_1\times H_2$ where
$$H_1:=\left\{u\in L_2(\Omega):\int_\Omega u\ dx=0\right\}\quad\text{and}\quad H_2:=L_2(\Omega).$$
\begin{prop}\label{relcomp}
Let $(\psi,\vartheta)\in E_1\times E_2$ be a global solution of \eqref{PFasym} and
assume (H1)-(H4). Then
\begin{enumerate}
\item $\psi\in L_\infty(\mathbb R_+;H_p^{2s}(\Omega)),\ s\in [0,1),\ p\in (1,\infty),\ \partial_t\psi\in
L_2(\mathbb R_+\times\Omega)$;
\item $\vartheta\in L_\infty(\mathbb R_+;H_2^1(\Omega)),\ \partial_t\vartheta\in
L_2(\mathbb R_+\times\Omega)$.
\end{enumerate}
In particular the orbits $\psi(\mathbb R_+)$ and $\vartheta(\mathbb R_+)$ are
relatively compact in $H_2^1(\Omega)$ and $H_2^{r}(\Omega)$,
respectively, where $r\in [0,1)$.\end{prop} \bpr Assertions 1 \& 2
follow directly from (H1)-(H4) and the proof of Proposition
\ref{estpsit}, which is given in the Appendix. Indeed, one may
replace the interval $J_{\max}$ by $\mathbb R_+$, since the operator
$-A^2=-\Delta_N^2$ generates an exponentially stable, analytic semigroup
$e^{-A^2 t}$ in the space
$$\mathbb X_p:=\{u\in L_p(\Omega):\int_\Omega u\ dx=0\}$$
with domain
$$D(A^2)=\{u\in H_p^4(\Omega)\cap\mathbb X_p:\partial_\nu u=\partial_\nu\Delta u=0\ \text{on}\
\partial\Omega\}.$$
\end{proof}
By Assumption (H4), there exists some bounded interval
$J_\vartheta\subset\mathbb R_+$ with $\vartheta(t,x)\in J_\vartheta$ for all $t\ge 0,\
x\in\Omega$. Therefore we may modify the nonlinearities $b$ and
$\beta$ outside $J_\vartheta$ in such a way that $b,\beta\in
C_b^{3-}(\mathbb R)$.
Unfortunately the energy functional $E$ is not yet the right one for our purpose,
since we have to include the nonlinear constraint
$$\int_\Omega(\lambda(\psi)+b(\vartheta))\ dx=0,$$
into our considerations. The linear constraint $\int_\Omega\psi\
dx=0$ is part of the definition of the space $H_1$. For the
nonlinear constraint we use a functional of Lagrangian type
which is given by
$$L(u,v)=E(u,v)-\overline{v} F(u,v),$$
defined on $V$, where $F(u,v):=\int_\Omega(\lambda(u)+b(v))\ dx$ and $\bar{w}=\frac{1}{|\Omega|}\int_\Omega w\ dx$ for a
function $w\in L_1(\Omega)$. Concerning the differentiability of $L$ we
have the following result.
\begin{prop}\label{diffL}
Under the conditions (H1)-(H4), the functional $L$ is twice continuously Fr\'{e}chet
differentiable on $V$ and the derivatives are given by
\begin{multline}\label{1stder}
\langle L'(u,v),(h,k)\rangle_{V^*,V}=\\
\langle E'(u,v),(h,k)\rangle_{V^*,V}-\overline{k}F(u,v)-\overline{v}\langle
F'(u,v),(h,k)\rangle_{V^*,V}
\end{multline}
and
\begin{multline}\label{2ndder}
\langle L''(u,v)(h_1,k_1),(h_2,k_2)\rangle_{V^*,V}=
\langle E''(u,v)(h_1,k_1),(h_2,k_2)\rangle_{V^*,V}-\\
\overline{k_1}\langle F'(u,v),(h_2,k_2)\rangle_{V^*,V}-\overline{k_2}\langle
F'(u,v),(h_1,k_1)\rangle_{V^*,V}-\\
\overline{v}\langle
F''(u,v)(h_1,k_1),(h_2,k_2)\rangle_{V^*,V},
\end{multline}
where $(h,k),(h_j,k_j)\in V,\ j=1,2$, and
\begin{align*}
\langle E'(u,v),(h,k)\rangle_{V^*,V}=\int_\Omega\nabla u\nabla h\
dx+\int_\Omega\Phi'(u) h\ dx+\int_\Omega\beta'(v) k\ dx,
\end{align*}
\begin{multline*}
\langle E''(u,v)(h_1,k_1),(h_2,k_2)\rangle_{V^*,V}=\\
\int_\Omega\nabla
h_1\nabla h_2\ dx+\int_\Omega\Phi''(u) h_1h_2\ dx+\int_\Omega\beta''(v)
k_1k_2\ dx,
\end{multline*}
\begin{align*}
\langle F'(u,v),(h,k)\rangle_{V^*,V}=\int_\Omega\lambda'(u) h\
dx+\int_\Omega b'(v) k\ dx
\end{align*}
and
\begin{align*}
\langle
F''(u,v)(h_1,k_1),(h_2,k_2)\rangle_{V^*,V}=\int_\Omega\lambda''(u)
h_1h_2\ dx+\int_\Omega b''(v) k_1k_2\ dx.
\end{align*}
\end{prop}
\bpr We only consider the first derivative, the second one is
treated in a similar way. Since the bilinear form
\begin{equation}\label{bilinform}a(u,v):=\int_\Omega\nabla u(x)\nabla v(x)\
dx\end{equation} defined on $V_1\times V_1$ is bounded and symmetric, the
first term in $E$ is twice continuously Fr\'{e}chet
differentiable. For the functional
$$G_1(u):=\int_\Omega\Phi(u)\ dx,\quad u\in V_1,$$
we argue as follows. With $u,h\in V_1$ it holds that
\begin{align*}
\Phi(u(x)+h(x))-\Phi(u(x))&-\Phi'(u(x))h(x)\\
&=\int_0^1\frac{d}{dt}\ \Phi(u(x)+th(x))\ dt-\int_0^1\Phi'(u(x))h(x)\ dt\\
&=\int_0^1\Big(\Phi'(u(x)+th(x))-\Phi'(u(x))\Big)h(x)\ dt\\
&=\int_0^1\int_0^t\frac{d}{ds}\ \Phi'(u(x)+sh(x))h(x)\ ds\ dt\\
&=\int_0^1\int_0^t\Phi''(u(x)+sh(x))h(x)^2\ ds\ dt\\
&=\int_0^1\Phi''(u(x)+sh(x))h(x)^2(1-s)\ ds.
\end{align*}
From the growth condition (H1), H\"{o}lder's inequality and the
Sobolev embedding theorem it follows that
\begin{align*}
\Big|\int_\Omega\Big(\Phi(u(x)+h(x))&-\Phi(u(x))-\Phi'(u(x))h(x)\Big)\ dx\Big|\\
&\le C\int_\Omega(1+|u(x)|^4+|h(x)|^4)|h(x)|^2\ dx\\
&\le C(1+|u|_{6}^4+|h|_{6}^4)|h|_{6}^2\\
&\le C(1+|u|_{V_1}^4+|h|_{V_1}^4)|h|_{V_1}^2.
\end{align*}
This proves that $G_1$ is Fr\'{e}chet differentiable and also
$G_1'(u)=\Phi'(u)\in L_{6/5}(\Omega)\hookrightarrow V_1^*$. The next
step is the proof of the continuity of $G_1':V_1\to V_1^*$. We
make again use of (H1), the H\"{o}lder inequality and the Sobolev
embedding theorem to obtain
\begin{align*}|G_1'(u)&-G_1'(\bar{u})|_{V_1^*}\\
&\le
C\left(\int_\Omega|\Phi'(u(x))-\Phi'(\bar{u}(x))|^{\frac{6}{5}}\
dx\right)^\frac{5}{6}\\
&\le
C\left(\int_\Omega\int_0^1|\Phi''(tu(x)+(1-t)\bar{u}(x))|^\frac{6}{5}|u(x)-\bar{u}(x)|^\frac{6}{5}\
dt\ dx\right)^\frac{5}{6}\\
&\le C
\left(\int_\Omega(1+|u(x)|^\frac{24}{5}+|\bar{u}(x)|^\frac{24}{5})|u(x)-\bar{u}(x)|^\frac{6}{5}\
dx\right)^\frac{5}{6}\\
&\le C\left(\int_\Omega(1+|u(x)|^6+|\bar{u}(x)|^6)\
dx\right)^\frac{2}{3}\left(\int_\Omega|u(x)-\bar{u}(x)|^{6}\right)^{\frac{1}{6}}\\
&\le C(1+|u|_{V_1}^4+|\bar{u}|_{V_1}^4)|u-\bar{u}|_{V_1}.
\end{align*}
Actually this proves that $G_1'$ is even locally Lipschitz
continuous on $V_1$. The Fr\'{e}chet differentiability of $G_1'$
and the continuity of $G_1''$ can be proved in an analogue way.
The fundamental theorem of differential calculus and the Sobolev
embedding theorem yield the estimate
\begin{multline*}
|\Phi'(u+h)-\Phi'(u)-\Phi''(u)h|_{V_1^*}\\
\le C\left(\int_\Omega\int_0^1|\Phi'''(u(x)+sh(x))|^{\frac{6}{5}}|h(x)|^{\frac{12}{5}}\
ds\ dx\right)^{\frac{5}{6}}.
\end{multline*}
We apply Assumption (H1) and H\"{o}lder's inequality to the result
\begin{align*}
|\Phi'(u+h)-\Phi'(u)&-\Phi''(u)h|_{V_1^*}\\
&\le C\left(\int_\Omega(1+|u(x)|^\frac{18}{5}+|h(x)|^\frac{18}{5})|h(x)|^\frac{12}{5}\
dx\right)^{\frac{5}{6}}\\
&\le C\left(\int_\Omega(1+|u(x)|^6+|h(x)|^6)\
dx\right)^{\frac{1}{2}}\left(\int_\Omega|h(x)|^6\
dx\right)^{\frac{1}{3}}\\
&=C(1+|u|_{V_1}^3+|h|_{V_1}^3)|h|_{V_1}^2.
\end{align*}
Hence the Fr\'{e}chet derivative is given by the multiplication
operator $G_1''(u)$ defined by $G_1''(u)v=\Phi''(u)v$ for all
$v\in V_1$ and $\Phi''(u)\in L_{3/2}(\Omega)$. We will omit the proof
of continuity of $G_1''$. The way to show the $C^2$-property of
the functional
$$G_2(u):=\int_\Omega\lambda(u(x))\ dx,\quad u\in V_1,$$
is identical to the one above, by Assumption (H2). Concerning the
$C^2$-differentiability of the functionals
$$G_3(v):=\int_\Omega\beta(v(x))\ dx\quad\text{and}\quad
G_4(v):=\int_\Omega b(v(x))\ dx,\quad v\in V_2,$$ one may adopt the
proof for $G_1$ and $G_2$. In fact, this time it is easier, since
$\beta$ and $b$ are assumed to be elements of the space
$C_b^{3-}(\mathbb R)$, however one needs the assumption $r\in (n/4,1)$. We will skip the details.
Finally the product rule of differentiation yields that $L$ is
twice continuously Fr\'{e}chet differentiable on $V_1\times V_2$.
\end{proof}
The corresponding stationary system to
\eqref{PFasym} will be of importance for the forthcoming
calculations. Setting all time-derivatives in \eqref{PFasym} equal
to 0 yields
$$\Delta\mu=0\quad\text{and}\quad\Delta\vartheta=0,$$
subject to the boundary conditions $\partial_\nu\mu=\partial_\nu\vartheta=0$.
Thus we have $\mu\equiv\mu_\infty=const$,
$\vartheta\equiv\vartheta_\infty=const$ and there remains the nonlinear
elliptic problem of second order \begin{equation}\label{statsys}
\begin{cases}
-\Delta\psi_\infty+\Phi'(\psi_\infty)-\lambda'(\psi_\infty)\vartheta_\infty=\mu_\infty,\quad
x\in\Omega,\\
\partial_\nu\psi_\infty=0,\quad x\in\partial\Omega,
\end{cases}
\end{equation} with the constraints \eqref{sidecond} for the unknowns
$\psi_\infty$ and $\vartheta_\infty$. The following proposition
collects some properties of the functional $L$ and the $\omega$-limit
set
\begin{multline*}
\omega(\psi,\vartheta):=\{(\varphi,\theta)\in V_1\times V_2:\exists\ (t_n)\nearrow \infty\ \mbox{s.t.}\\
(\psi(t_n),\vartheta(t_n))\to (\varphi,\theta)\ \mbox{in}\ V_1\times V_2\}.
\end{multline*}
\begin{prop}\label{propomlim}
Under Hypotheses (H1)-(H4) the following assertions are true.
\begin{enumerate}
\item The $\omega$-limit set is nonempty, connected and compact.
\item Each point $(\psi_\infty,\vartheta_\infty)\in\omega(\psi,\vartheta)$ is a strong solution of the
stationary problem \eqref{statsys}, where
$\vartheta_\infty,\mu_\infty=const$ and $(\psi_\infty,\vartheta_\infty)$
satisfies the constraints \eqref{sidecond} for the unknowns
$\vartheta_\infty,\mu_\infty$.
\item The functional $L$ is constant on $\omega(\psi,\vartheta)$ and each
point $(\psi_\infty,\vartheta_\infty)\in\omega(\psi,\vartheta)$ is a critical
point of $L$, i.e. $L'(\psi_\infty,\vartheta_\infty)=0$ in $V^*$.
\end{enumerate}
\end{prop}
\bpr The fact that $\omega(\psi,\vartheta)$ is nonempty, connected and
compact follows from Proposition \ref{relcomp} and some well-known
facts in the theory of dynamical systems.
Now we turn to 2. Let $(\psi_\infty,\vartheta_\infty)\in
\omega(\psi,\vartheta)$. Then there exists a sequence
$(t_n)\nearrow+\infty$ such that
$(\psi(t_n),\vartheta(t_n))\to(\psi_\infty,\vartheta_\infty)$ in $V$ as
$n\to\infty$. Since $\partial_t\psi,\partial_t\vartheta\in L_2(\mathbb R_+\times\Omega)$ it
follows that $\psi(t_n+s)\to\psi_\infty$ and
$\vartheta(t_n+s)\to\vartheta_\infty$ in $L_2(\Omega)$ for all $s\in [0,1]$ and
by relative compactness also in $V$. This can be seen as follows.
\begin{align*}
|\psi(t_n+s)-\psi_\infty|_{2}&\le
|\psi(t_n+s)-\psi(t_n)|_{2}+|\psi(t_n)-\psi_\infty|_{2}\\
&\le \int_{t_n}^{t_n+s}|\partial_t\psi(t)|_{2}\
dt+|\psi(t_n)-\psi_\infty|_{2}\\
&\le s^{1/2}\left(\int_{t_n}^{t_n+s}|\partial_t\psi(t)|_{2}^2\
dt\right)^{1/2}+|\psi(t_n)-\psi_\infty|_{2}.
\end{align*}
Then, for $t_n\to\infty$ this yields $\psi(t_n+s)\to\psi_\infty$ for
all $s\in[0,1]$. The proof for $\vartheta$ is the same. Integrating
\eqref{eneq} with $f_1=f_2=0$ from $t_n$ to $t_n+1$ we obtain
\begin{multline*}
E(\psi(t_n+1),\vartheta(t_n+1))-E(\psi(t_n),\vartheta(t_n))\\
+\int_0^1\int_\Omega\left(|\nabla\mu(t_n+s,x)|^2+|\nabla\vartheta(t_n+s,x)|^2\right)\
dx\ ds=0.
\end{multline*}
Letting $t_n\to+\infty$ yields
$$|\nabla\mu(t_n+\cdot,\cdot)|,|\nabla\vartheta(t_n+\cdot,\cdot)|\to 0\quad\text{in $L_2([0,1]\times\Omega)$}.$$
This in turn yields a subsequence $(t_{n_k})$ such that
$\nabla\mu(t_{n_k}+s),\nabla\vartheta(t_{n_k}+s)\to 0$ in
$L_2(\Omega;\mathbb R^n)$ for a.e. $s\in [0,1]$. Hence
$\nabla\vartheta_\infty=0$, since the gradient is a closed operator in
$L_2(\Omega;\mathbb R^n)$. This in turn yields that $\vartheta_\infty$ is a
constant.
\noindent Furthermore the Poincar\'{e}-Wirtinger inequality
implies that
\begin{multline*}|\mu(t_{n_k}+s^*)-\mu(t_{n_l}+s^*)|_2\\\le
C_p\Big(|\nabla\mu(t_{n_k}+s^*)-\nabla\mu(t_{n_l}+s^*)|_2+\int_\Omega|\Phi'(\psi(t_{n_k}+s^*))-\Phi'(\psi(t_{n_l}+s^*))|\
dx\\
+\int_\Omega|\lambda'(\psi(t_{n_k}+s^*))\vartheta(t_{n_k}+s^*)-\lambda'(\psi(t_{n_l}+s^*))\vartheta(t_{n_l}+s^*)|\
dx,\end{multline*} for some $s^*\in [0,1]$. Taking the limit
$k,l\to\infty$ we see that $\mu(t_{n_k}+s^*)$ is a Cauchy sequence
in $L_2(\Omega)$, hence it admits a limit, which we denote by
$\mu_\infty$. In the same manner as for $\vartheta_\infty$ we therefore
obtain $\nabla\mu_\infty=0$, hence $\mu_\infty$ is a constant.
Observe that the relation
$$\mu_\infty=\frac{1}{|\Omega|}\left(\int_\Omega(\Phi'(\psi_\infty)-\lambda'(\psi_\infty)\vartheta_\infty)\
dx\right)$$ is valid. Multiplying $\eqref{PFasym}_1$ by a function
$\varphi\in H_2^1(\Omega)$ and integrating by parts we obtain
\begin{multline*}
(\mu(t_{n_k}+s^*),\varphi)_2=(\nabla\psi(t_{n_k}+s^*),\nabla\varphi)_2+\\
(\Phi'(\psi(t_{n_k}+s^*)),\varphi)_2-(\lambda'(\psi(t_{n_k}+s^*))\vartheta(t_{n_k}+s^*),\varphi)_2.
\end{multline*}
As $t_{n_k}\to\infty$ it follows that \begin{equation}\label{RegPsiInftyPF}
(\mu_\infty,\varphi)_2=(\nabla\psi_\infty,\nabla\varphi)_2+(\Phi'(\psi_\infty),\varphi)_2-\vartheta_\infty(\lambda'(\psi_\infty),\varphi)_2.
\end{equation} By the Lax-Milgram theorem the bounded, symmetric and
elliptic form
$$a(u,v):=\int_\Omega\nabla u\nabla v\
dx,$$ defined on the space $V_1\times V_1$ induces a bounded
operator $A:V_1\to V_1^*$ with nonempty resolvent, such that
$$a(u,v)=\langle Au,v\rangle_{V_1^*,V_1},$$
for all $(u,v)\in V_1\times V_1$. It is well-known that the domain
of the part $A_p$ of the operator $A$ in
$$\mathbb X_p=\{u\in L_p(\Omega):\int_\Omega u\
dx=0\}$$ is given by
$$D(A_p)=\{u\in \mathbb X_p\cap H_p^2(\Omega),\ \partial_\nu u=0\}.$$ Going back to
\eqref{RegPsiInftyPF} we obtain from (H1) and (H2) that
$\psi_\infty\in D(A_q)$, where $q=6/(\beta+2)$. Since $q>6/5$ we
may apply a bootstrap argument to conclude $\psi_\infty\in
D(A_2)$. Integrating \eqref{RegPsiInftyPF} by parts, assertion
2 follows.
In order to prove 3.\ , we make use of \eqref{1stder} to obtain
\begin{align*}\langle
L'(\psi_\infty,\vartheta_\infty)&,(h,k)\rangle_{V^*,V}\\
&=\langle
E'(\psi_\infty,\vartheta_\infty),(h,k)\rangle_{V^*,V}-\vartheta_\infty\langle
F'(\psi_\infty,\vartheta_\infty),(h,k)\rangle_{V^*,V}\\
&=\int_\Omega(-\Delta\psi_\infty+\Phi'(\psi_\infty))h\
dx+\int_\Omega\beta'(\vartheta_\infty)k\
dx\\
&\hspace{3cm}-\vartheta_\infty\int_\Omega(\lambda'(\psi_\infty)h+b'(\vartheta_\infty)k)\
dx\\
&=\int_\Omega\mu_\infty h\ dx=0,
\end{align*}
for all $(h,k)\in V$, since $\mu_\infty$ and $\vartheta_\infty$ are
constant. A continuity argument finally yields the last statement
of the proposition.
\end{proof} \noindent The following result is crucial for the proof of
convergence.
\begin{prop}[Lojasiewicz-Simon inequality]\label{LS}
Let $(\psi_\infty,\vartheta_\infty)\in\omega(\psi,\vartheta)$ and assume
(H1)-(H5). Then there exist constants $s\in (0,\frac{1}{2}],
C,\delta>0$ such that
$$|L(u,v)-L(\psi_\infty,\vartheta_\infty)|^{1-s}\le C|L'(u,v)|_{V^*},$$
whenever $|(u,v)-(\psi_\infty,\vartheta_\infty)|_V\le \delta$.
\end{prop}
\bpr We show first that $\dim
N(L''(\psi_\infty,\vartheta_\infty))<\infty$. By \eqref{2ndder} we obtain
\begin{align*}
\langle
L''(\psi_\infty,\vartheta_\infty)(h_1,k_1),&(h_2,k_2)\rangle_{V^*,V}\\
&=\int_\Omega\nabla h_1\nabla h_2\
dx+\int_\Omega\Phi''(\psi_\infty)h_1h_2\
dx+\int_\Omega\beta''(\vartheta_\infty)k_1k_2\ dx\\
&-\overline{k_1}\int_\Omega(\lambda'(\psi_\infty)h_2+b'(\vartheta_\infty)k_2)\
dx\\
&-\overline{k_2}\int_\Omega(\lambda'(\psi_\infty)h_1+b'(\vartheta_\infty)k_1)\
dx\\
&-\overline{\vartheta_\infty}\int_\Omega(\lambda''(\psi_\infty)h_1h_2+b''(\vartheta_\infty)k_1k_2)\
dx.
\end{align*}
Since $\beta''(s)=b'(s)+s b''(s)$ and $\vartheta_\infty\equiv const$ we
have
\begin{align*}
\langle
L''(\psi_\infty&,\vartheta_\infty)(h_1,k_1),(h_2,k_2)\rangle_{V^*,V}\\
&=\int_\Omega\nabla h_1\nabla h_2\
dx+\int_\Omega\left(\Phi''(\psi_\infty)h_1-\overline{k_1}\lambda'(\psi_\infty)-\vartheta_\infty\lambda''(\psi_\infty)h_1\right)h_2\
dx\\
&+\int_\Omega(b'(\vartheta_\infty)(k_1-2\overline{k_1})-\overline{\lambda'(\psi_\infty)h_1})k_2\
dx
\end{align*}
for all $(h_j,k_j)\in V$. If $(h_1,k_1)\in
N(L''(\psi_\infty,\vartheta_\infty))$, it follows that
$$b'(\vartheta_\infty)(k_1-2\overline{k_1})-\overline{\lambda'(\psi_\infty)h_1}=0.$$
It is obvious that a solution $k_1$ to this equation must be
constant, hence it is given by
\begin{equation}\label{k1}k_1=-(b'(\vartheta_\infty))^{-1}\overline{\lambda'(\psi_\infty)h_1},\end{equation}
where we also made use of (H4). Concerning $h_1$ we have
\begin{equation}\label{eqh1}\langle
Ah_1,h_2\rangle_{V_1^*,V_1}=\int_\Omega(k_1\lambda'(\psi_\infty)+\vartheta_\infty\lambda''(\psi_\infty)h_1-\Phi''(\psi_\infty)h_1)h_2\
dx,\end{equation} since $k_1$ is constant. By Proposition \ref{propomlim} it
holds that $\psi_\infty\in D(A_2)\hookrightarrow L_\infty(\Omega)$,
hence $Ah_1\in H_1$, which means that $h_1\in D(A_2)$ and from
\eqref{eqh1} we obtain
$$A_2
h_1+P(\Phi''(\psi_\infty)h_1-\vartheta_\infty\lambda''(\psi_\infty)h_1-k_1\lambda'(\psi_\infty))=0,$$
where $P$ denotes the projection $P:H_2\to H_1$, defined by
$Pu=u-\overline{u}$. It is an easy consequence of the embedding
$D(A_2)\hookrightarrow L_\infty(\Omega)$ that the linear operator
$B:H_1\to H_1$ given by
$$Bh_1=P(\Phi''(\psi_\infty)h_1-\vartheta_\infty\lambda''(\psi_\infty)h_1-k_1\lambda'(\psi_\infty))$$
is bounded, where $k_1$ is given by \eqref{k1}. Furthermore the
operator $A_2$ defined in the proof of Proposition \ref{propomlim}
is invertible, hence $A_2^{-1}B:H_1\to D(A_2)$ is a compact operator
by compact embedding and this in turn yields that $(I+A_2^{-1}B)$ is
a Fredholm operator. In particular it holds that $\dim
N(I+A_2^{-1}B)<\infty$, whence $N(L''(\psi_\infty,\vartheta_\infty))$ is
finite dimensional and moreover
$$N(L''(\psi_\infty,\vartheta_\infty))\subset
D(A_2)\times (H_2^r(\Omega)\cap L_\infty(\Omega))\hookrightarrow
L_\infty(\Omega)\times L_\infty(\Omega).$$ By Hypothesis
(H5), the restriction of $L'$ to the space $D(A_2)\times
(H_2^r(\Omega)\cap L_\infty(\Omega))$ is analytic in a neighbourhood of $(\psi_\infty,\theta_\infty)$. For the definition of analyticity in
Banach spaces we refer to \cite[Section 3]{Chi}. Now the claim
follows from \cite[Theorem 3.10 \& Corollary 3.11]{Chi}.
\end{proof}
\noindent Let us now state the main result of this section.
\begin{thm}
Assume (H1)-(H5) and let $(\psi,\vartheta)$ be a global solution of \eqref{PFasym}. Then the limits
$$\lim_{t\to\infty}\psi(t)=:\psi_\infty,\quad\text{and}\quad
\lim_{t\to\infty}\vartheta(t)=:\vartheta_\infty=const$$ exist in
$H_2^1(\Omega)$ and $H_2^r(\Omega),\ r\in (0,1)$, respectively, and
$(\psi_\infty,\vartheta_\infty)$ is a strong solution of the stationary
problem \eqref{statsys}.
\end{thm}
\bpr Since by Proposition \ref{propomlim} the $\omega$-limit set is
compact, we may cover it by a union of \emph{finitely} many balls
with center $(\varphi_i,\theta_i)\in\omega(\psi,\vartheta)$ and radius $\delta_i>0$,
$i=1,\ldots,N$. Since $L(u,v)\equiv L_\infty$ on $\omega(\psi,\vartheta)$
and each $(\varphi_i,\theta_i)$ is a critical point of $L$, there are
\emph{uniform} constants $s\in(0,\frac{1}{2}]$, $C>0$ and an open
set $U\supset \omega(\psi,\vartheta)$, such that
\begin{equation}\label{CHGLS}|L(u,v)-L_\infty|^{1-s}\le C|L'(u,v)|_{V^*},\end{equation}
for all $(u,v)\in U$. Define
$H:\mathbb R_+\to\mathbb R_+$ by
$$H(t):=(L(\psi(t),\vartheta(t))-L_\infty)^s.$$
The function $H$ is nonincreasing and $\lim_{t\to\infty}H(t)=0$, since $L(\psi(t),\vartheta(t))=E(\psi(t),\vartheta(t))$ and since $E$ is a strict Lyapunov functional for \eqref{PFasym}, which follows from \eqref{eneq}. Furthermore we have
$\lim_{t\to\infty}\dist((\psi(t),\vartheta(t)),\omega(\psi,\vartheta))=0$, i.e.
there exists $t^*\ge 0$, such that $(\psi(t),\vartheta(t))\in U$,
for all $t\ge t^*$. Next, we compute and estimate the time
derivative of $H$. By \eqref{eneq} and Proposition \ref{LS} we
obtain
\begin{align}\label{PFA6}
-\frac{d}{dt}\ H(t)&=s\left(-\frac{d}{dt}\
L(\psi(t),\vartheta(t))\right)|L(\psi(t),\vartheta(t))-L_\infty|^{s-1}\nonumber\\
&\ge
C\frac{|\nabla\mu(t)|_2^2+|\nabla\vartheta(t)|_2^2}{|L'(\psi(t),\vartheta(t))|_{V^*}}
\end{align}
So have to estimate the term $|L'(\psi(t),\vartheta(t))|_{V^*}$.
For convenience we will write $\psi=\psi(t)$ and $\vartheta=\vartheta(t)$.
From \eqref{1stder} we obtain with $\bar{h}=0$
\begin{equation}\label{1}\begin{split} \langle
L'(\psi,\vartheta)&,(h,k)\rangle_{V^*,V}\\
&=\int_\Omega(-\Delta\psi+\Phi'(\psi))h\
dx+\int_\Omega\vartheta b'(\vartheta)k\
dx-\overline{\vartheta}\int_\Omega(\lambda'(\psi)h+b'(\vartheta)k)\
dx\\
&=\int_\Omega(\mu-\overline{\mu}) h\
dx+\int_\Omega(\vartheta-\overline{\vartheta})\lambda'(\psi)h\
dx+\int_\Omega(\vartheta-\overline{\vartheta})b'(\vartheta)k\ dx\end{split}\end{equation} An
application of the H\"{o}lder and Poincar\'{e}-Wirtinger inequality yields
the estimates
\begin{align}\label{2}
|\int_\Omega(\vartheta-\overline{\vartheta})\lambda'(\psi)h\ dx|&\le
|\lambda'(\psi)|_{\infty}|\vartheta-\overline{\vartheta}|_2|h|_2\le
c|\nabla\vartheta|_2|h|_2,
\end{align}
\begin{align}\label{3}
|\int_\Omega(\vartheta-\overline{\vartheta})b'(\vartheta)k\ dx|&\le
|b'(\vartheta)|_{\infty}|\vartheta-\overline{\vartheta}|_2|k|_2\le
c|\nabla\vartheta|_2|k|_2
\end{align}
and
\begin{align}\label{4}
|\int_\Omega(\mu-\overline{\mu})h\ dx|&\le c|\nabla\mu|_2|h|_2,
\end{align}
whence we obtain
$$|L'(\psi(t),\vartheta(t))|_{V^*}\le C(|\nabla\mu(t)|_2+|\nabla\vartheta(t)|_2),$$
by taking the supremum over all functions $(h,k)\in V$ with norm
less than 1 in \eqref{1}-\eqref{4}. This in connection with
\eqref{PFA6} yields
$$-\frac{d}{dt} H(t)\ge C(|\nabla\mu(t)|_2+|\nabla\vartheta(t)|_2),$$
hence $|\nabla\mu|,|\nabla\vartheta|\in L_1([t^*,\infty),L_2(\Omega))$. Using
the equation $\partial_t\psi=\Delta\mu$ we see that $\partial_t\psi\in
L_1([t^*,\infty),H_2^1(\Omega)^*)$, hence the limit
$$\lim_{t\to\infty}\psi(t)=:\psi_\infty$$ exists in $H_2^1(\Omega)^*$ and even in $H_2^1(\Omega)$ thanks to Proposition \ref{relcomp}.
From equation $\eqref{PFasym}_2$ it follows that $\partial_t e\in
L_1([t^*,\infty);H_2^1(\Omega)^*)$, where $e:=b(\vartheta)+\lambda(\psi)$,
i.e. the limit $\lim_{t\to\infty}e(t)$ exists in $H_2^1(\Omega)^*$.
This in turn yields that the limit
$$\lim_{t\to\infty}b(\vartheta(t))=:b_\infty$$
exists in $L_2(\Omega)$, by relative compactness, cf.
Proposition \ref{relcomp}. By the monotonicity
assumption (H3) we obtain $\vartheta(t)=b^{-1}(b(\vartheta(t)))$ and thus
the limit of $\vartheta(t)$ as $t$ tends to infinity exists in
$L_2(\Omega)$. From the relative compactness of the
orbit $\vartheta(\mathbb R_+)$ it follows that the limit
$$\lim_{t\to\infty}\vartheta(t)=:\vartheta_\infty$$
also exists in $H_2^r(\Omega),\ r\in [0,1)$. Finally Proposition
\ref{propomlim} yields the last statement of the theorem.
\end{proof}
\section{Appendix}
\emph{Proof of Proposition \ref{fixpointest}}
Let $(u,v),(\bar{u},\bar{v})\in \mathbb B_R(u^*,v^*)$. By Sobolev embedding it
holds that $u,\bar{u}$ and $v,\bar{v}$ are uniformly bounded in
$C^1(\overline{\Omega})$ and $C(\overline{\Omega})$, respectively.
Furthermore, we will use the following inequality, which has been
proven in \cite[Lemma 6.2.3]{Zach}. \begin{equation}\label{zachineqPF}
|f(w)-f(\bar{w})|_{H_p^s(L_p)}\le
\mu(T)(|w-\bar{w}|_{H_p^{s_0}(L_p)}+|w-\bar{w}|_{\infty,\infty}),\quad
0<s<s_0<1,\end{equation} valid for every $f\in C^{2-}(\mathbb R)$ and all
$w,\bar{w}\in\mathbb B_R^1(u^*)\cup\mathbb B_R^2(v^*)$. Here $\mu=\mu(T)$ denotes
a function, with the property $\mu(T)\to 0$ as $T\to 0$. The proof consists of several steps
(i) By H\"{o}lders inequality it holds that
\begin{align*}|\Delta&\Phi'(u)-\Delta\Phi'(\bar{u})|_{X(T)}\\
&\le|\Delta u\Phi''(u)-\Delta \bar{u}\Phi''(\bar{u})|_{X(T)}+
||\nabla u|^2\Phi'''(u)-|\nabla \bar{u}|^2\Phi'''(\bar{u})|_{X(T)}\\
&\le |\Delta u|_{rp,rp}|\Phi''(u)-\Phi''(\bar{u})|_{r'p,r'p}+|\Delta u-\Delta \bar{u}|_{rp,rp}|\Phi''(\bar{u})|_{r'p,r'p}\\
&\hspace{0.3cm}+T^{1/p}\left(|\nabla
u|_{\infty,\infty}^2|\Phi'''(u)-\Phi'''(\bar{u})|_{\infty,\infty}+|\nabla
u-\nabla \bar{u}|_{\infty,\infty}|\Phi'''(\bar{u})|_{\infty,\infty}\right)\\
&\le T^{1/r'p}\left(|\Delta
u|_{rp,rp}|\Phi''(u)-\Phi''(\bar{u})|_{\infty,\infty}+|\Delta u-\Delta
\bar{u}|_{rp,rp}|\Phi''(\bar{u})|_{\infty,\infty}\right)\\
&\hspace{0.3cm}+T^{1/p}\left(|\nabla
u|_{\infty,\infty}^2|\Phi'''(u)-\Phi'''(\bar{u})|_{\infty,\infty}+|\nabla
u-\nabla
\bar{u}|_{\infty,\infty}|\Phi'''(\bar{u})|_{\infty,\infty}\right),
\end{align*}
since $u,\bar{u}\in C(J;C^1(\overline{\Omega}))$. We have
$$\Delta w\in H_p^{\theta_2/2}(J;H_p^{2(1-\theta_2)}(\Omega))\hookrightarrow L_{r p}(J\times\Omega),\quad
\theta_2\in[0,1],$$ for every function $w\in E_1(T)$, since $r>1$ may
be chosen close to 1. Therefore we obtain
$$|\Delta\Phi'(u)-\Delta\Phi'(\bar{u})|_{X(T)}\le \mu(T)\left(R+|u^*|_1\right)|u-\bar{u}|_1,$$
due to the assumption $\Phi\in C^{4-}(\mathbb R)$.
(ii) Consider the term $(\lambda'(\psi_0)-\lambda'(u))\Delta
v-(\lambda'(\psi_0)-\lambda'(\bar{u}))\Delta\bar{v}$.
\begin{align*}
|(\lambda'(\psi_0)-\lambda'(u))&\Delta v-(\lambda'(\psi_0)-\lambda'(\bar{u}))\Delta\bar{v}|_{X(T)}\\
&\le |(\lambda'(\psi_0)-\lambda'(u))\Delta(v-\bar{v})|_{X(T)}+|(\lambda'(u)-\lambda'(\bar{u}))\Delta\bar{v}|_{X(T)}\\
&\le |\psi_0-u|_{\infty,\infty}|v-\bar{v}|_{E_2(T)}+|u-\bar{u}|_{\infty,\infty}|\bar{v}|_{E_2(T)}\\
&\le
(|\psi_0-u^*|_{\infty,\infty}+|u^*-u|_{\infty,\infty})|v-\bar{v}|_{E_2(T)}\\
&\hspace{1cm}+|u-\bar{u}|_{E_1(T)}(|\bar{v}-v^*|_{E_2(T)}+|v^*|_{E_2(T)})\\
&\le C(\mu(T)+R)|(u,v)-(\bar{u},\bar{v})|_1,
\end{align*}
since $\lambda\in C^{4-}(\mathbb R)$. Next, we consider the term
$\nabla(\lambda'(\psi_0)-\lambda'(u))\nabla
v-\nabla(\lambda'(\psi_0)-\lambda'(\bar{u}))\nabla\bar{v}$. We obtain
\begin{multline*}
|\nabla(\lambda'(\psi_0)-\lambda'(u))\nabla
v-\nabla(\lambda'(\psi_0)-\lambda'(\bar{u}))\nabla\bar{v}|_{X(T)}\\
\le
|\nabla(\lambda'(\psi_0)-\lambda'(u))|_\infty|\nabla(v-\bar{v})|_{X(T)}+
|\nabla(\lambda'(u)-\lambda'(\bar{u}))|_\infty|\nabla\bar{v}|_{X(T)}.
\end{multline*}
Since
$$\nabla(\lambda'(\psi_0)-\lambda'(u))=\nabla\psi_0(\lambda''(\psi_0)-\lambda''(u))+\lambda''(u)(\nabla\psi_0-\nabla
u),$$ and the same for $\nabla(\lambda'(u)-\lambda'(\bar{u}))$, we may
argue as above, to conclude
\begin{multline*}
|\nabla(\lambda'(\psi_0)-\lambda'(u))|_{\infty,\infty}|\nabla(v-\bar{v})|_{X(T)}+
|\nabla(\lambda'(u)-\lambda'(\bar{u}))|_{\infty,\infty}|\nabla\bar{v}|_{X(T)}\\
\le (\mu(T)+R)|(u,v)-(\bar{u},\bar{v})|_{1}.
\end{multline*}
Finally, we estimate the remaining part with H\"{o}lder's
inequality to the result
\begin{multline}\label{locpropPF1}
|v\Delta(\lambda'(\psi_0)-\lambda'(u))-\bar{v}\Delta(\lambda'(\psi_0)-\lambda'(\bar{u}))|_{X(T)}\\
\le|v-\bar{v}|_{\infty,\infty}|\Delta(\lambda'(\psi_0)-\lambda'(u))|_{X(T)}+|\bar{v}|_{r'p,r'p}|\Delta(\lambda'(u)-\lambda'(\bar{u}))|_{rp,rp},
\end{multline} where $1/r+1/r'=1$. For the first part, we obtain
\begin{align*}
|\Delta&(\lambda'(\psi_0)-\lambda'(u))|_{X(T)}\\
&\le|\Delta\psi_0|_p|\lambda''(\psi_0)-\lambda''(u)|_{\infty,\infty}+|\Delta\psi_0-\Delta
u
|_p|\lambda''(u)|_{\infty,\infty}\\
&\hspace{1cm}+|\nabla\psi_0|_{\infty,\infty}^2|\lambda'''(\psi_0)-\lambda'''(u)|_{\infty,\infty}+|\lambda'''(u)|_{\infty,\infty}|\nabla\psi_0-\nabla
u|_{\infty,\infty}\\
&\le C(|\psi_0-u|_{\infty,\infty}+|\nabla\psi_0-\nabla
u|_{\infty,\infty}+|\Delta\psi_0-\Delta u|_{p,p})\\
&\le C(\mu(T)+R),
\end{align*}
since $\psi_0\in H_p^2(\Omega)\cap C^1(\overline{\Omega})$ and $\lambda\in
C^{4-}(\mathbb R)$. For the second term in \eqref{locpropPF1} we obtain
\begin{align*}
|\Delta(\lambda'(u)&-\lambda'(\bar{u}))|_{rp,rp}\\
&\le |\Delta u|_{rp,rp}|\lambda''(u)-\lambda''(\bar{u})|_{\infty,\infty}+|\lambda''(\bar{u})|_{\infty,\infty}|\Delta
u-\Delta\bar{u}|_{rp,rp}\\
&\hspace{1cm}+|\nabla
u|_{\infty,\infty}^2|\lambda'''(u)-\lambda'''(\bar{u})|_{\infty,\infty}+|\lambda'''(\bar{u})|_{\infty,\infty}|\nabla
u-\nabla \bar{u}|_{\infty,\infty}\\
&\le C|u-\bar{u}|_{E_1(T)},
\end{align*}
since $u,\bar{u}\in C(J;C^1(\overline{\Omega}))$ and $r>1$ can be
chosen close enough to 1, due to the fact that $\bar{v}\in
C(J;C(\overline{\Omega}))$. Finally, we observe
$$|\bar{v}|_{r'p,r'p}\le |\bar{v}-v^*|_{r'p,r'p}+|v^*|_{r'p,r'p}\le
\mu(T)+R.$$
\noindent
(iii) For simplicity we set $f(u,v)=a_0\lambda'(\psi_0)-a(v)\lambda'(u)$.
Then we compute
\begin{align}\label{pro(v)}
|f(u,v)\partial_t u&-f(\bar{u},\bar{v})\partial_t \bar{u}|_{X(T)}\nonumber\\
&\le |\partial_t u(f(u,v)-f(\bar{u},\bar{v}))|_{X(T)}+|f(\bar{u},\bar{v})(\partial_tu-\partial_t\bar{u})|_{X(T)}\\
&\le (|\partial_t u-\partial_t u^*|_{X(T)}+|\partial_t
u^*|_{X(T)})|f(u,v)-f(\bar{u},\bar{v})|_{\infty,\infty}\nonumber\\
&\hspace{2cm}+|f(\bar{u},\bar{v})|_{\infty,\infty}|\partial_tu-\partial_t\bar{u}|_{X(T)}\nonumber\\
&\le
C(\mu_3(T)+R)|f(u,v)-f(\bar{u},\bar{v})|_{\infty,\infty}\nonumber\\
&\hspace{2cm}+|f(\bar{u},\bar{v})|_{\infty,\infty}|\partial_tu-\partial_t\bar{u}|_{X(T)}\nonumber.
\end{align}
Next we estimate
\begin{align*}
|f(u,v)&-f(\bar{u},\bar{v})|_{\infty,\infty}\\
&\le |a(v)(\lambda'(u)-\lambda'(\bar{u}))|_{\infty,\infty}+|\lambda'(\bar{u})(a(v)-a(\bar{v}))|_{\infty,\infty}\\
&\le |a(v)|_{\infty,\infty}|\lambda'(u)-\lambda'(\bar{u})|_{\infty,\infty}+|\lambda'(\bar{u})|_{\infty,\infty}|a(v)-a(\bar{v})|_{\infty,\infty}\\
&\le C(|u-\bar{u}|_{\infty,\infty}+|v-\bar{v}|_{\infty,\infty})\le
C|(u,v)-(\bar{u},\bar{v})|_1.
\end{align*}
Furthermore, we have
\begin{align*}
|f(\bar{u},\bar{v})|_{\infty,\infty}&\le
|a_0|_{\infty,\infty}|\lambda'(\psi_0)-\lambda'(\bar{u})|_{\infty,\infty}+|\lambda'(\bar{u})|_{\infty,\infty}|a_0-a(\bar{v})|_{\infty,\infty}\\
&\le C(|\psi_0-\bar{u}|_{\infty,\infty}+|\vartheta_0-\bar{v}|_{\infty,\infty})\\
&\le
C(|\psi_0-u^*|_{\infty,\infty}+|u^*-\bar{u}|_{\infty,\infty}+|\vartheta_0-v^*|_{\infty,\infty}+|v^*-\bar{v}|_{\infty,\infty})\\
&\le C(\mu(T)+R).
\end{align*}
The estimate of $(a_0-a(v))\Delta v-(a_0-a(\bar{v}))\Delta\bar{v}$ in $L_p(J;L_p(\Omega))$ can be carried out in a similar way.
(iv) We compute \begin{multline*}|(a(v)-a(\bar{v})f_2|_{X(T)}\le
|a(v)-a(\bar{v})|_{\infty,\infty}|f_2|_{X(T)}\le
|v-\bar{v}|_{\infty,\infty}|f_2|_{X(T)}\\\le\mu(T)|v-\bar{v}|_{E_2(T)}\le\mu(T)|(u,v)-(\bar{u},\bar{v})|_{1},
\end{multline*}
since $f_2\in X(T)$ is a fixed function, hence $|f_2|_{X(T)}\to 0$
as $T\to 0$.
(v) By trace theory, we obtain
\begin{multline*}
|\partial_\nu(\Phi'(u)-\Phi'(\bar{u}))|_{Y_1(T)}\\
\le C|\Phi'(u)-\Phi'(\bar{u})|_{H_p^{1/2}(J;L_p(\Omega))}+|\Phi'(u)-\Phi'(\bar{u})|_{L_p(J;H_p^2(\Omega))}.
\end{multline*}
The second norm has already been estimated in (i), so it remains
to estimate $\Phi'(u)-\Phi'(\bar{u})$ in $H_p^{1/2}(J;L_p(\Omega))$.
Here we will use \eqref{zachineqPF}, to obtain
\begin{align*}|\Phi'(u)-\Phi'(\bar{u})|_{H_p^{1/2}(L_p)}&\le
\mu(T)(|u-\bar{u}|_{H_p^{s_0}(L_p)}+|u-\bar{u}|_{\infty,\infty})\\
&\le \mu(T)C|u-\bar{u}|_{E_1(T)}\le
\mu(T)C|(u,v)-(\bar{u},\bar{v})|_{1},
\end{align*}
since $s_0<1$.
(vi) We may apply (ii) and trace theory, to conclude that it
suffices to estimate
\begin{multline*}
(\lambda'(\psi_0)-\lambda'(u))v-(\lambda'(\psi_0)-\lambda'(\bar{u}))\bar{v}\\
=(\lambda'(\psi_0)-\lambda'(u))(v-\bar{v})-(\lambda'(u)-\lambda'(\bar{u}))\bar{v}
\end{multline*}
in $H_p^{1/2}(J;L_p(\Omega))$. This yields
\begin{align*}
|(\lambda'&(\psi_0)-\lambda'(u))(v-\bar{v})|_{H_p^{1/2}(L_p)}\\
&\le|\lambda'(\psi_0)-\lambda'(u)|_{H_p^{1/2}(L_p)}|v-\bar{v}|_{\infty,\infty}+
|\lambda'(\psi_0)-\lambda'(u)|_{\infty,\infty}|v-\bar{v}|_{H_p^{1/2}(L_p)}\\
&\le
(|\lambda'(\psi_0)-\lambda'(u^*)|_{H_p^{1/2}(L_p)}+|\lambda'(u^*)-\lambda'(u)|_{H_p^{1/2}(L_p)})|v-\bar{v}|_{E_2(T)}\\
&\hspace{1cm}+
(|\psi_0-u^*|_{\infty,\infty}+|u^*-u|_{\infty,\infty})|v-\bar{v}|_{E_2(T)}\\
&\le
\left(|\lambda'(\psi_0)-\lambda'(u^*)|_{H_p^{1/2}(L_p)}+\mu(T)R+(\mu(T)+R)\right)|v-\bar{v}|_{E_2(T)}.
\end{align*}
Clearly
$\lambda'(\psi_0)-\lambda'(u^*)\in\, _0H_p^{1/2}(J;L_p(\Omega))$, since
$\psi_0$ does not depend on $t$ and since $\lambda\in C^{4-}(\mathbb R)$.
Therefore it holds that
$$|\lambda'(\psi_0)-\lambda'(u^*)|_{H_p^{1/2}(L_p)}\to 0$$
as $T\to 0$. The second part $(\lambda'(u)-\lambda'(\bar{u}))\bar{v}$ can
be treated as follows.
\begin{align*}
|(\lambda'(u)&-\lambda'(\bar{u}))\bar{v}|_{H_p^{1/2}(L_p)}\\
&\le|\lambda'(u)-\lambda'(\bar{u})|_{H_p^{1/2}(L_p)}|\bar{v}|_{\infty,\infty}+|\lambda'(u)-\lambda'(\bar{u})|_{\infty,\infty}|\bar{v}|_{H_p^{1/2}(L_p)}\\
&\le C(\mu(T)+R+\mu(T))|u-\bar{u}|_{E_1(T)},
\end{align*}
where we applied again \eqref{zachineqPF}. This completes the
proof of the proposition.\vspace{0.2cm}
\noindent\emph{Proof of Proposition \ref{estpsit}}
\noindent Let $J_{\max}^\delta:=[\delta,T_{\max}]$ for some small $\delta>0$.
Setting $A^2=\Delta_N^2$ with domain
$$D(A^2)=\{u\in H_p^4(\Omega):\partial_\nu u=\partial_\nu\Delta u=0\ \text{on}\
\partial\Omega\},$$ the solution $\psi(t)$ of equation $\eqref{PFGWP}_1$ may
be represented by the variation of parameters formula
\begin{equation}\label{varpar}\psi(t)=e^{-A^2t}\psi_0+\int_0^tAe^{-A^2(t-s)}\Big(\lambda'(\psi(s))\vartheta(s)-\Phi'(\psi(s))\Big)\
ds,\quad t\in J_{\max},\end{equation} where $e^{-A^2t}$ denotes the analytic
semigroup, generated by $-A^2=-\Delta_N^2$ in $L_p(\Omega)$. By (H1), (H2)
and \eqref{enest2} it holds that
$$\Phi'(\psi)\in L_\infty(J_{\max};L_{q_0}(\Omega))\quad\text{and}\quad
\lambda'(\psi)\in L_\infty(J_{\max};L_6(\Omega)),$$ with $q_0=6/(\gamma+2)$.
We then apply $A^{r},\ r\in (0,1)$, to \eqref{varpar} and make use
of semigroup theory to obtain \begin{equation}\label{estpsi1}\psi\in
L_\infty(J_{\max}^\delta;H_{q_0}^{2r}(\Omega)), \end{equation} valid for all $r\in
(0,1)$, since $q_0<6$. It follows from \eqref{estpsi1} that
$\psi\in L_\infty(J_{\max}^\delta;L_{p_1}(\Omega))$ if
$2r-3/{q_0}\ge-3/p_1$, and
$$\Phi'(\psi)\in L_\infty(J_{\max}^\delta;L_{q_1}(\Omega))\quad\text{as well as}\quad
\lambda'(\psi)\in L_\infty(J_{\max}^\delta;L_{p_1}(\Omega)),$$ with
$q_1=p_1/(\gamma+2)$. Hence we have this time
$$\psi\in L_\infty(J_{\max}^\delta;H_{q_1}^{2r}(\Omega)),\quad r\in (0,1).$$
Iteratively we obtain a sequence $(p_n)_{n\in\mathbb N_0}$ such that
$$2r-\frac{3}{q_n}\ge-\frac{3}{p_{n+1}},\quad n\in\mathbb N_0$$
with $q_n=p_n/(\gamma+2)$ and $p_0=6$. Thus the sequence
$(p_n)_{n\in\mathbb N_0}$ may be recursively estimated by
$$\frac{1}{p_{n+1}}\ge\frac{\gamma+2}{p_n}-\frac{2r}{3},$$
for all $n\in\mathbb N_0$ and $r\in (0,1)$. From this definition it is
not difficult to obtain the following estimate for $1/p_{n+1}$.
\begin{align}\label{bootstrap}
\frac{1}{p_{n+1}}&\ge\frac{(\gamma+2)^{n+1}}{p_0}-\frac{2r}{3}\sum_{k=0}^{n}(\gamma+2)^k\nonumber\\
&=\frac{(\gamma+2)^{n+1}}{p_0}-\frac{2r}{3}\left(\frac{(\gamma+2)^{n+1}-1}{\gamma_1+1}\right)\nonumber\\
&=(\gamma+2)^{n+1}\left(\frac{1}{p_0}-\frac{2r}{3\gamma+3}\right)+\frac{2r}{3\gamma+3},\quad
n\in\mathbb N_0.
\end{align}
By the assumption (H1) on $\gamma$ we see that the term in brackets
is negative if $r\in (0,1)$ is sufficiently close to 1 and
therefore, after finitely many steps the entire right side of
\eqref{bootstrap} is negative as well, whence we may choose $p_n$
arbitrarily large or we may even set $p_n=\infty$ for $n\ge N$ and
a certain $N\in\mathbb N_0$. In other words this means that for those
$r\in (0,1)$ we have \begin{equation}\label{highordest1}\psi\in
L_\infty(J_{\max}^\delta;H_p^{2r}(\Omega)),\end{equation} for all $p\in
[1,\infty]$. It is important, that we can achieve this result in
\emph{finitely} many steps!
Next we will derive an estimate for $\partial_t\psi$. For all
forthcoming calculations we will use the abbreviation
$\psi=\psi(t)$ and $\vartheta=\vartheta(t)$. Since we only have estimates on
the interval $J_{\max}^\delta$, we will use the following solution
formula.
$$\psi(t)=e^{-A^2(t-\delta)}\psi_\delta+\int_0^{t-\delta}Ae^{-A^2 s}\Big(\lambda'(\psi)\vartheta-\Phi'(\psi)\Big)(t-s)\
ds,\quad t\in J_{\max}^\delta$$ where $\psi_\delta:=\psi(\delta)$.
Differentiating with respect to $t$, we obtain
\begin{multline}\label{psit}
\partial_t\psi(t)=
A\int_0^{t-\delta}e^{-A^2s}(\lambda''(\psi)\vartheta\partial_t\psi+\lambda'(\psi)\partial_t\vartheta-\Phi''(\psi)\partial_t\psi)(t-s)\ ds\\
+F(t,\psi_\delta,\vartheta_\delta),
\end{multline}
for all $t\ge \delta$ and with
$$F(t,\psi_\delta,\vartheta_\delta):=Ae^{-A^2(t-\delta)}(\lambda'(\psi_\delta)\vartheta_\delta-\Phi'(\psi_\delta))-A^2e^{-A^2 (t-\delta)}\psi_\delta.$$
Let us discuss the function $F$ in detail. By the trace theorem we have
$\psi_\delta\in B_{pp}^{4-4/p}(\Omega)$ and $\vartheta_{\delta}\in
B_{pp}^{2-2/p}(\Omega)$. Since we assume $p>(n+2)/2$, it holds that
$\psi_\delta,\vartheta_\delta\in L_\infty(\Omega)$. Furthermore, the semigroup
$e^{-A^2 t}$ is analytic. Therefore there exist some constants
$C>0$ and $\omega\in\mathbb R$ such that
$$|F(t,\psi_\delta,\vartheta_\delta)|_{L_p(\Omega)}\le
C\left(\frac{1}{(t-\delta)^{1/2}}+\frac{1}{t-\delta}\right)e^{\omega t},$$
for all $t>\delta$. This in turn implies that
$$F(\cdot,\psi_\delta,\vartheta_\delta)\in L_p(J_{\max}^{\delta'}\times\Omega)$$
for all $p\in (1,\infty)$, where $0<\delta<\delta'<T_{\max}$. We will
now use equations $\eqref{PFasym}_{1,2}$ to rewrite the integrand
in \eqref{psit} in the following way.
\begin{align}\label{est2}
(\lambda''(\psi)\vartheta&-\Phi''(\psi))\partial_t\psi+\lambda'(\psi)\partial_t\vartheta\nonumber\\
&=(\lambda''(\psi)\vartheta-\Phi''(\psi))\Delta\mu+\frac{\lambda'(\psi)}{b'(\vartheta)}\Delta\vartheta
-\frac{\lambda'(\psi)^2}{b'(\vartheta)}\Delta\mu\nonumber\\
&=\diver\left[\left(\lambda''(\psi)\vartheta-\frac{\lambda'(\psi)^2}{b'(\vartheta)}-\Phi''(\psi)\right)\nabla\mu\right]+
\diver\left[\frac{\lambda'(\psi)}{b'(\vartheta)}\nabla\vartheta\right]\\
&-\nabla\left(\lambda''(\psi)\vartheta-\frac{\lambda'(\psi)^2}{b'(\vartheta)}-\Phi''(\psi)\right)\cdot\nabla\mu-\nabla\frac{\lambda'(\psi)}{b'(\vartheta)}\cdot\nabla\vartheta.\nonumber
\end{align}
Thus we obtain a decomposition of the following form
\begin{multline*}
(\lambda''(\psi)\vartheta-\Phi''(\psi))\partial_t\psi+\lambda'(\psi)\partial_t\vartheta\\
=\diver(f_\mu\nabla\mu+f_\vartheta\nabla\vartheta)+
g_\mu\nabla\mu+g_\vartheta\nabla\vartheta+h_\mu\nabla\vartheta\nabla\mu+h_\vartheta|\nabla\vartheta|^2,
\end{multline*}
with
\begin{align*}
f_\mu:=\lambda''(\psi)\vartheta-\frac{\lambda'(\psi)^2}{b'(\vartheta)}-\Phi''(\psi),&\quad
f_\vartheta:=\frac{\lambda'(\psi)}{b'(\vartheta)},\\
g_\mu:=-\left(\lambda'''(\psi)\vartheta-2\frac{\lambda'(\psi)\lambda''(\psi)}{b'(\vartheta)}-\Phi''(\psi)\right)\nabla\psi,&\quad
g_\vartheta:=-\frac{\lambda''(\psi)}{b'(\vartheta)}\nabla\psi,\\
h_\mu:=\lambda''(\psi)-\frac{b''(\vartheta)\lambda'(\psi)^2}{b'(\vartheta)^2},&\quad
h_\vartheta:=\frac{b''(\vartheta)\lambda'(\psi)}{b'(\vartheta)^2}.
\end{align*}
By Assumption (H3) and the first part of the proof it holds that
$f_j,g_j,h_j\in L_\infty(J_{\max}^\delta\times\Omega)$ for each
$j\in\{\mu,\vartheta\}$ and this in turn yields that
\begin{align*}\diver(f_\mu\nabla\mu+f_\vartheta\nabla\vartheta)&\in
L_2(J_{\max}^\delta;H_2^1(\Omega)^*),\\
g_\mu\cdot\nabla\mu+g_\vartheta\cdot\nabla\vartheta&\in L_2(J_{\max}^\delta\times\Omega),\\
h_\mu\nabla\vartheta\cdot\nabla\mu+h_\vartheta|\nabla\vartheta|^2&\in
L_1(J_{\max}^\delta\times\Omega),
\end{align*}
where we also made use of \eqref{enest2}. Setting
$$
T_1=Ae^{-A^2t}\ast \diver(f_\mu\nabla\mu+f_\vartheta\nabla\vartheta),\quad
T_2=Ae^{-A^2t}\ast (g_\mu\cdot\nabla\mu+g_\vartheta\cdot\nabla\vartheta)
$$
and
$$T_3=Ae^{-A^2t}\ast
(h_\mu\nabla\vartheta\cdot\nabla\mu+h_\vartheta|\nabla\vartheta|^2),$$ we may
rewrite \eqref{psit} as
$$\partial_t\psi=T_1+T_2+T_3+F(t,\psi_0,\vartheta_0).$$
Going back to \eqref{psit} we obtain
\begin{align*}&T_1\in
H_2^{1/4}(J_{\max}^\delta;L_2(\Omega))\cap L_2(J_{\max}^\delta;H_2^1(\Omega))\hookrightarrow L_2(J_{\max}^\delta\times\Omega),\\
&T_2\in H_2^{1/2}(J_{\max}^\delta;L_2(\Omega))\cap L_2(J_{\max}^\delta;H_2^2(\Omega))\hookrightarrow L_2(J_{\max}^\delta\times\Omega),\quad\text{and}\\
&F(\cdot,\psi_\delta,\vartheta_\delta)\in L_2(J_{\max}^{\delta'}\times\Omega).
\end{align*}
Observe that we do not have full regularity for $T_3$ since $A$
has no maximal regularity in $L_1(\Omega)$, but nevertheless we
obtain
$$T_3\in H_1^{1/2-}(J_{\max}^\delta;L_1(\Omega))\cap
L_1(J_{\max}^\delta;H_1^{2-}(\Omega)).$$ Here we used the notation
$H_p^{s-}:=H_p^{s-\varepsilon}$ and $\varepsilon>0$ is sufficiently small. An
application of the mixed derivative theorem then yields
$$H_1^{1/2-}(J_{\max}^\delta;L_1(\Omega))\cap
L_1(J_{\max}^\delta;H_1^{2-}(\Omega))\hookrightarrow
L_p(J_{\max}^\delta;L_2(\Omega)),$$ if $p\in (1,8/7)$, whence
$$\partial_t\psi\in L_2(J_{\max}^{\delta'}\times\Omega)+L_p(J_{\max}^{\delta'};L_2(\Omega))$$
for some $1<p<8/7$. Now we go back to \eqref{est2} where we
replace this time only $\partial_t\vartheta$ by the differential equation
$\eqref{PFasym}_2$ to obtain
\begin{align*}\label{est3}
(\lambda''(\psi)\vartheta&-\Phi''(\psi))\partial_t\psi+\lambda'(\psi)\partial_t\vartheta\\
&=\left(\lambda''(\psi)\vartheta-\Phi''(\psi)-\frac{\lambda'(\psi)^2}{b'(\vartheta)}\right)\partial_t\psi\\
&+\diver\left[\frac{\lambda'(\psi)}{b'(\vartheta)}\nabla\vartheta\right]-\frac{\lambda''(\psi)}{b'(\vartheta)}\nabla\psi\cdot\nabla\vartheta
+\frac{\lambda'(\psi)b''(\vartheta)}{b'(\vartheta)^2}|\nabla\vartheta|^2\\
&=f\partial_t\psi+\diver\left[g\nabla\vartheta\right]+h\cdot\nabla\vartheta+k|\nabla\vartheta|^2.
\end{align*}
Rewrite \eqref{psit} in the following way
\begin{equation}\label{S}\partial_t\psi=S_1+S_2+S_3+S_4+F(t,\psi_0,\vartheta_0),\end{equation}
where the functions $S_j$ are defined in the same manner as $T_j$.
Since $f,g,h\in L_\infty(J_{\max}^\delta\times\Omega)$ it follows again
from regularity theory that
\begin{multline*}
S_1\in H_2^{1/2}(J_{\max}^{\delta'};L_2(\Omega))\cap
L_2(J_{\max}^{\delta'};H_2^2(\Omega))\\
+H_p^{1/2}(J_{\max}^{\delta'};L_2(\Omega))\cap L_p(J_{\max}^{\delta'};H_2^2(\Omega)),
\end{multline*}
$$S_2\in H_2^{1/4}(J_{\max}^{\delta'};L_2(\Omega))\cap L_2(J_{\max}^{\delta'};H_2^1(\Omega)),$$
$$S_3\in H_2^{1/2}(J_{\max}^{\delta'};L_2(\Omega))\cap L_2(J_{\max}^{\delta'};H_2^2(\Omega)),$$
and it can be readily verified that
$$H_p^{1/2}(J_{\max}^{\delta'};L_2(\Omega))\cap L_p(J_{\max}^{\delta'};H_2^2(\Omega))\hookrightarrow
L_2(J_{\max}^{\delta'}\times\Omega),$$ whenever $p\in[1,2]$. Now we turn
our attention to the term $S_4=Ae^{-A^2t}\ast k|\nabla\vartheta|^2$.
First we observe that by the mixed derivative theorem the
embedding
$$Z_q:=H_q^{1/2-}(J_{\max}^{\delta'};L_1(\Omega))\cap
L_q(J_{\max}^{\delta'};H_1^{2-}(\Omega))\hookrightarrow
L_2(J_{\max}^{\delta'}\times\Omega)$$ is valid, provided that
$q\in(8/5,2]$. Hence it holds that
$$|S_4|_{2,2}\le C|S_4|_{Z_q}\le C|k|\nabla\vartheta|^2|_{q,1}\le
C|\nabla\vartheta|_{2q,2}^2,$$ with some constant $C>0$. Taking the
norm of $\partial_t\psi$ in $L_2(J_{\max}^{\delta'}\times\Omega)$ we obtain
from \eqref{S}
$$|\partial_t\psi|_{2,2}\le
C\left(\sum_{j=1}^3|S_j|_{2,2}+|\nabla\vartheta|^2_{2q,2}+|F(\cdot,\psi_\delta,\vartheta_\delta)|_{2,2}\right).$$
The Gagliardo-Nirenberg inequality in connection with
\eqref{enest2} yields the estimate
$$|\nabla\vartheta|_{2q,2}^2\le
c|\nabla\vartheta|_{2,2}^{2a}|\nabla\vartheta|_{\infty,2}^{2(1-a)}\le
c|\nabla\vartheta|_{\infty,2}^{2(1-a)},$$ provided that $a=1/q$.
Multiply $\eqref{PFGWP}_2$ by $\partial_t\vartheta$ and integrate by parts
to the result
\begin{multline*}
\int_\Omega b'(\vartheta(t,x))|\partial_t\vartheta(t,x)|^2\
dx+\frac{1}{2}\frac{d}{dt}|\nabla\vartheta(t)|_{2}^2=-\int_\Omega
\lambda'(\psi(t,x))\partial_t\psi(t,x)\partial_t\vartheta(t,x)\ dx.
\end{multline*}
Making use of (H3) and Young's inequality we obtain
\begin{equation}\label{enest4}C_1|\partial_t\vartheta|_{2,2}^2+\frac{1}{2}|\nabla\vartheta(t)|_2^2\le
C_2(|\partial_t\psi|_{2,2}^2+|\nabla\vartheta_0|_2^2),\end{equation} after integrating
w.r.t. $t$. This in turn yields the estimate
$$|\nabla\vartheta|_{2q,2}^2\le
c|\nabla\vartheta|_{\infty,2}^{2(1-a)}\le
c(1+|\partial_t\psi|_{2,2}^{2(1-a)}).$$ In order to gain something from
this inequality we require that $2(1-a)<1$, i.e. $q$ is restricted
by $1<q<2$. Finally, if we choose $q\in (8/5,2)$ and use the
uniform boundedness of the $L_2$ norms of $S_j,\ j\in\{1,2,3\}$ we
obtain
$$|\partial_t\psi|_{2,2}\le C(1+|\partial_t\psi|_{2,2}^{2(1-a)}).$$
Since by construction $2(1-a)<1$, it follows that the $L_2$-norm
of $\partial_t\psi$ is bounded on $J_{\max}^{\delta'}\times\Omega$. In
particular, this yields the statement for $\vartheta$ by equation
\eqref{enest4}.
Now we go back to \eqref{psit} with $\delta$ replaced by $\delta'$. By
Assumption (H5), by the bounds $\partial_t\vartheta,\partial_t\psi\in
L_2(J_{\max}^{\delta'};L_2(\Omega))$ and by the first part of the proof
we obtain
$$\lambda''(\psi)\vartheta\partial_t\psi+\lambda'(\psi)\partial_t\vartheta-\Phi''(\psi)\partial_t\psi\in
L_2(J_{\max}^{\delta'};L_2(\Omega)).$$ Since the operator $A^2=\Delta^2$
with domain
$$D(A^2)=\{u\in H_p^4(\Omega):\partial_\nu u=\partial_\nu\Delta u=0\}$$
has the property of maximal $L_p$-regularity (cf. \cite[Theorem 2.1]{DHP07}), we obtain from \eqref{psit}
$$\partial_t\psi-F(\cdot,\psi_{\delta'},\vartheta_{\delta'})\in
H_2^{1/2}(J_{\max}^{\delta'};L_2(\Omega))\cap
L_2(J_{\max}^{\delta'};H_2^2(\Omega))\hookrightarrow
L_r(J_{\max}^{\delta'};L_r(\Omega)),$$ and the last embedding is valid
for all $r\le 2(n+4)/n$. By the properties of the function $F$ it
follows
$$\partial_t\psi\in L_r(J_{\max}^{\delta''};L_r(\Omega)),$$
for all $r\le 2(n+4)/n$ and some $0<\delta''<T_{\max}$. To obtain an
estimate for the whole interval $J_{\max}$, we use the fact that
we already have a local strong solution, i.e. $\partial_t\psi\in
L_p(0,\delta'';L_p(\Omega))$, $p>(n+2)/2$. The proof is complete.
\bibliographystyle{plain}
|
\section{Introduction}
From the last century, many black hole candidates are observationally
discovered and it is believed that the spacetime around the object
is well described by a Kerr metric. In the Kerr metric, when a central
object is a black hole, it is required that its specific angular momentum
should not be larger than its mass \cite{mtw73,fn98}, i.e. $|a|\le M$.
Here, $a$ denotes a specific angular momentum of a black hole
and is defined as $a=J/M$ where $M$ and $J$ are the mass and
the angular momentum of the black hole, respectively.
In this case, an event horizon exists around the black hole. We call this bound
a Kerr bound \cite{gh09,bf09,bft09}. The observational confirmation of
this bound leads to the confirmation of the existence of a black hole.
In case that the specific angular momentum of a central object is larger than
its mass, curvature singularity where spacetime curvature
diverges is not surrounded by an event horizon. To avoid this, the
cosmic censorship conjecture in which spacetime singularity should
be concealed from our world is proposed \cite{p69,p79,w97,p98}.
In some numerical simulations which start from configurations similar
to a star before gravitational collapse, an apparent horizon
forms before the formation of spacetime curvature singularity and
a final object inevitably becomes a black hole \cite{ss04}.
On the other hand, the appearance of curvature singularity
without being surrounded by an apparent horizon has been suggested
in the numerical simulation of axisymmetric gravitational collapse of collisionless
particles \cite{st91, st92}. It has been also revealed that
curvature singularity not surrounded by a horizon appears in the
spherically symmetric collapse in the critical and super-critical
cases (\cite{harada04} and references therein).
Of course, it is also well known that in cylindrically symmetric systems
a black hole cannot be formed after gravitational collapse
\cite{t65,t72}, and it is not fully
excluded that the spacetime singularity which is not surrounded
by a horizon forms when the gravitational
collapse starts from nearly cylindrically symmetric systems
\cite{b02}. However, it is questionable that such
highly symmetric configuration forms in any astrophysical situations.
We can also see that most examples of spacetime singularities
not surrounded by a horizon can be regarded as precursory or transient
singularities followed by the formation of event and/or apparent horizons. It
is also discussed that spacetime singularity not surrounded by a
horizon is dynamically unstable\cite{cpcc08a,cpcc08b,pccc09}.
Moreover, the object with its specific angular
momentum larger than its mass evolves to a black hole through
accretion processes of mass and angular momentum from a rotating disk around
the object \cite{mtw73,t74}. That is, it would be a physically reasonable
assumption that the black hole candidates with mass accretion observed so far are
really black holes (although recently, another scenario has been discussed, see \cite{js09}).
Recently, the numerical simulations of hydrodynamic accretion onto a super-spinning
object have suggested that for the object with
its spin $a/M$ slightly larger than unity the mass accretion is prohibited
because of the repulsive force near the central object due to
the spacetime geometrical effects \cite{bfh09}.
On the other hand, in past studies many attempts were performed to obtain
direct evidence of the existence of a black hole from observational data
(see \S\ref{sec:dis}). Especially, thanks to the developments of radio interferometers
achieving highest spatial resolution in existing telescopes, the black hole in the
Galactic Center Sgr A* with a large apparent size begins to be spatially resolved
\cite{s05,d08}. The spatial resolution of these radio
interferometers is comparable to the size of the apparent size of a black
hole shadow in the nearby galactic centers such as Sgr A*
\cite{b09,y09,h09} and M87 \cite{bl09}. The data recently obtained
by sub-millimeter interferometers \cite{d08} contain information of the size of
luminous matters whose size is comparable to the size of the event horizon of the
black hole in Sgr A* \cite{b09,y09,h09}. Although as described above the
Kerr bound is assumed to be valid by many authors, we have not yet obtained the final
confirmation of the bound from observational data. Even for the black hole in Sgr A*
in which there are plenty of observational data such as energy spectrum, linear
and circular polarization, radio visibility and light curves in the wide range of
observed frequencies, we have not yet obtained the final value of the spin of the black hole
in Sgr A* \cite{b09,y09,h09} and the Kerr bound is not also observationally
confirmed.
Recently, several theoretical studies relating to the confirmation of the Kerr
bound are performed for a black hole which we can potentially directly image
in the near future such as the case of Sgr A* and in these studies the apparent size
of a central object is estimated \cite{bf09,bft09,hm09}. The strategy they adopted
is as follows. They first assume a central object which is described by the Kerr metric
but have specific angular momentum larger than its mass. Next, they calculate
the observational signatures of the assumed object. Finally, they compare the observational
signatures with those of a black hole. As observational signatures, the apparent shapes
and sizes of the assumed objects are calculated. From these calculations,
the very large values of the spin of the central object in Sgr A* are ruled out
from its large size \cite{bf09,hm09}. They assume that general relativity is not valid at
a central region near the curvature singularity but replaced by an alternative theory such as quantum gravity theory (see also \cite{gh09}).
In order to spatially resolve the apparent image of a black hole by future interferometers,
its apparent angular size should be larger than several micro-arcseconds, which correspond
to the spatial resolution of radio interferometers in the near future \cite{d08}.
Even with these highest spatial resolution, most of the black hole candidates discovered
in X-ray observations can not be spatially resolved because of their too small angular sizes.
For most of the black holes discovered in X-ray, energy spectrum and light curves
are observationally obtained. For some of these black holes observed in X-ray,
the parameters of the black hole such as the mass and the spin are determined by
the spectral fittings by assuming that the object is a black hole
\cite{zcc97,gmne01,dbht05,msnrdl06,smndlr06,rm06,nms08,mns08}.
For these black hole candidates, it is open to question whether it is possible to confirm the Kerr
bound from the observational data. The main purpose of the present study is to answer this.
In order to do this, we take the same strategy as described in the last paragraph. That is,
we first consider the object which violates the Kerr bound, i.e. with $a/M>1$. In this paper,
we call this object a super-spinar \cite{gh09,bfh09}.
Next, the energy spectrum of the assumed object is calculated and finally compare the
spectrum with that of a black hole. The X-ray spectrum of a black hole candidate generally
consists of a thermal component originating from the accretion disc around the black hole
and a non-thermal component originating from high energy photons which are up-scattered
in e.g. corona above or in the accretion disc. In this study, we only assume a thermal
component for simplicity. The observed energy spectrum is calculated by solving the general
relativistic radiative transfer including usual special and general relativistic effects such as
Doppler boosting, gravitational redshift, light bending and frame-dragging.
In this paper, we assume no emission from a central object.
The present paper is organized as follows. In \S\ref{sec:disc}, physical assumptions and
disc structure are given. In \S\ref{sec:spec}, we calculate the local radiation flux, the radial
temperature profile and the energy spectrum of the disc. We give discussion in \S\ref{sec:dis}
and conclusions are presented in \S\ref{sec:con}.
Throughout this paper, we use the geometrical units $c=G=1$.
\section{Structure of an Accretion disc}
\label{sec:disc}
In this section, we describe the basic disc structure used in the calculations in the next
section. Although some part of this section was already investigated in the past studies
\cite{df74, df78, cn79, rt79, s80, s81a, s81b, bsb89},
we describe these for the completeness of the description
and the preliminaries for the next section. In the process of the mass accretion onto the
central compact object such as a black hole or a super-spinar, when the accreting
matters have some angular momentum, accreting fluid usually forms disc-like structure
around a compact object. This astrophysical object is called an accretion disc. In the
accretion disc, the angular momentum of the fluid is transferred outward due to the
viscous stress caused by magnetic and/or turbulent effects. Then, matters are allowed
to gradually and spirally accrete inward. On the other hand, the viscous stress converts
the gravitational energy of the accreting matters into other forms of energy such as the
thermal, radiation and/or magnetic energies. In the present study,
we consider the geometrically thin and optically thick accretion disc,
where the gravitational energy is effectively released as the radiation energy, and then
produce the significant radiation which can be observed. According to the accretion
disc theory, this type of disc structure is achieved for the accretion disc whose mass
accretion rate is nearly sub-Eddington, i.e. $\dot{M}< L_{\rm Edd}/c^2$. For the
geometrically thin disc, the disc thickness $H$ in the vertical direction at some radius
$r$ is much smaller than the radius $r$, i.e. $H\ll r$. In terms of this type of the accretion
disc, the general relativistic disc around the Kerr black hole is given in \cite{nt73} and
\cite{pt74}. In the present study, by adopting the basically same calculation methods and
assumptions in \cite{pt74}, we calculate the disc structure around the super-spinar.
We neglect the self-gravity of the accretion disc. As a background
geometry, we consider the stationary, axisymmetric and asymptotically flat spacetime
described as
\begin{equation}
ds^2=-e^{2\nu}dt^2+e^{2\psi}(d\varphi-\omega dt)^2+e^{2\mu_1}dr^2+e^{2\mu_2}d\theta^2,
\label{eq:metric}
\end{equation}
where
\begin{eqnarray}
e^{2\nu}=\Sigma\Delta/A,~~e^{2\psi}=\sin^2\theta A/\Sigma,~~e^{2\mu_1}
=\Sigma/\Delta,~~e^{2\mu_2}=\Sigma,~~\omega=2Mar/A. \nonumber \\
\end{eqnarray}
Here, $M$ is the mass of the central object, $a$ is its angular momentum
per unit mass, $\omega$ is the angular velocity of the frame-dragging
around the central object, and the functions $\Delta$, $\Sigma$ and $A$
are defined as $\Delta=r^2-2Mr+a^2$, $\Sigma=r^2+a^2\cos^2\theta$, and
$A=(r^2+a^2)^2-a^2\Delta\sin^2\theta$, respectively. The horizon radius
$r_{\rm H}$ of the black hole is given by $r_{\rm H}/M=1+\sqrt{1-a_*^2}$
where $a_*\equiv a/M$ for $-1\le a_* \le 1$.
While for $-1\le a_* \le 1$ the central object is a black hole,
for $a_* < -1$ or $1< a_*$ it is a super-spinar. In this study, we
mainly focus on the super-spinar with $1<a_{*}$.
We assume that the matters in the disc move in equatorial circular
geodesic orbits around the central object as in \cite{pt74} and \cite{bpt72}. Such
orbits of the particles in the Kerr geometry are described by the three
constants of motion; the total energy $E$, the angular momentum parallel to
symmetry axis $L$ and the Carter constant $Q$ \cite{bpt72, c68}. The
radial component of the equations governing the orbital trajectory is
given by $\Sigma dr/d\lambda=\pm [V_r(r)]^{1/2}$, where $\lambda$ is the
parameter along the trajectory and
$V_r(r)=[E(r^2+a^2)-La]^2-\Delta[r^2+(L-aE)^2+Q]$ in \cite{bpt72}. The
energy and the angular momentum of the particle with the circular orbit
on the equatorial plane are calculated from the conditions $V_r(r)=0$ and
$V_r^\prime(r)=0$ and are given by
\begin{eqnarray}
E&=&s_1[r^{1/2}(r-2M)+s_2 aM^{1/2}]/p, \label{eq:E}\\
L&=&s_1 s_2 M^{1/2}(r^2+a^2-s_2 2M^{1/2}ar^{1/2})/p, \label{eq:L}
\end{eqnarray}
where $p=r^{3/4}(r^{3/2}-3Mr^{1/2}+s_2 2aM^{1/2})^{1/2}$, $s_1=\pm 1$
and $s_2=\pm 1$.
In the limit of $r\to \infty$, the sign of the energy becomes
positive (negative) for $s_1=1$ ($-1$).
It is noted that the sign $s_1 = -1$ corresponds to negative energy as measured
by local observers and gives unphysical solutions in the context of the present
paper (for details, see \cite{bsb89}).
On the other hand,
$s_2=1$ and $-1$ respectively correspond to equatorial circular orbits of
the 1st family and the 2nd family \cite{s80}. For the orbits with $s_1=1$,
the 2nd family orbit always corresponds to a retrograde orbit.
In the case of a black hole ($a_*\le 1$), the 1st family orbit always corresponds
to a prograde orbit outside the horizon \cite{bpt72}.
On the other hand, in the case of a super-spinar ($a_*>1$) the 1st family orbit
can be a retrograde orbit near the super-spinar, while far from the super-spinar
the 1st family orbit always corresponds to a prograde orbit. These features are
investigated in the past studies \cite{df74,s80}.
From the energy $E(=-u_t)$ and the angular momentum $L(=u_\phi)$
given above, all components of the four-velocity $u^\mu$ of the particle
with a circular orbit in the equatorial plane can be calculated. For
this orbit, the angular velocity $\Omega(=u^\phi/u^t)$ is given by
$\Omega=s_2M^{1/2}/[r^{3/2}+s_2 aM^{1/2}]$. This is the Keplerian
angular velocity in the Kerr geometry. By using these $E$ and $L$,
the marginally bound circular orbit $r_{\rm mb}$ and the ISCO $r_{\rm ISCO}$ are obtained
from $E/\mu=1$ and $V_r^{\prime\prime}(r)=0$, respectively, as in \cite{bpt72}.
The circular orbit exists for the case that the denominator of Eqs. (\ref{eq:E}) and
(\ref{eq:L}) is real, i.e. $p\ge 0$. The limiting case, $p=0$, gives an orbit with infinite
energy per unit rest mass, and hence
the radius of the photon circular orbit $r_{\rm ph}$ \cite{bpt72}.
\begin{figure}
\begin{center}
\vspace{+0mm}
\hspace*{-0mm}
\includegraphics[width=120mm]{f1.eps}
\caption{\label{fig:superKerr_radius}
The radii of the event horizon $r_{\rm H}$ ({\it dotted line}) and the innermost stable
circular orbit (ISCO) $r_{\rm ISCO}$ ({\it solid line}) for a black hole ($a_*\leq 1$) and
a super-spinar ($a_*>1$) as a function of the Kerr parameter $a_*$. Only the
quantities corresponding to the direct motions are shown.
}
\end{center}
\end{figure}
For the accretion disc consisting of the materials with the circular
orbit, the inner boundary with no torque is usually assumed. While
outside the inner boundary the gravitational energy is effectively released as the
radiation, inside the inner boundary the matters fall freely onto the central object with the
energy and the angular momentum with the values at the inner
boundary. That is, inside this radius which is called the plunging
region, there is no radiation. The recent three-dimensional
magnetohydrodynamic simulations around the rotating black holes support
this assumption \cite{s08}. In this study, we assume that the viscous torque vanishes
at the ISCO, $r_{\rm ISCO}$, where $dE/dr=dL/dr=0$. For the rotating black hole, the
analytic expression for $r_{\rm ISCO}$ is given by Eq. (2.21) in \cite{bpt72}. For any
values of the spin $a_*$, the radius of ISCO is analytically given by
\begin{eqnarray}
r_{\rm ISCO}/M&=&3+Z_2-{\rm sgn}_2\left[(3-Z_1)(3+Z_1+2Z_2)\right]^{1/2},\\
Z_1&=&1+\left|1-a_*^2\right|^{1/3}\left[\left|1-|a_*|\right|^{1/3}+{\rm sgn}_1(1+|a_*|)^{1/3}\right],\nonumber\\
Z_2&=&\left(3a_*^2+Z_1^2\right)^{1/2},\nonumber\\
{\rm sgn}_1&=&\left\{
\begin{array}{ll}
+1 & ~{\rm for}~~a_*^2\leq 1~~({\rm black~hole})\\
-1 & ~{\rm for}~~a_*^2> 1~~({\rm naked~singularity})\\
\end{array}
\right.
\nonumber\\
{\rm sgn}_2&=&\left\{
\begin{array}{ll}
+1 & ~{\rm for}~~a_*\geq 0 \\
-1 & ~{\rm for}~~a_*<0.\\
\end{array}
\right.
\nonumber
\end{eqnarray}
In the limit of $a_*\to +\infty$, $r_{\rm ISCO}/M\to \sqrt{3}a_*$.
In Fig. \ref{fig:superKerr_radius}, we show the event horizon $r_{\rm H}$ ({\it dotted line})
and the innermost stable circular orbit $r_{\rm ISCO}$ ({\it solid line}) as a function of the
spin parameter $a_*$. For $a_*=8\sqrt{6}/3~(\sim 6.532)$, the radius of the ISCO becomes
$6M$ which is the same value as that for the non-rotating black hole (i.e. $a_*=0$). The
minimum value of $r_{\rm ISCO}=(2/3)M~(\sim 0.667 M)$ which is achieved at
$a_*=a_{\rm cr}\equiv 4\sqrt{2}/(3\sqrt{3})~(\sim 1.089)$.
\begin{figure}
\begin{center}
\vspace{+0mm}
\hspace*{-0mm}
\includegraphics[width=70mm]{f2a.eps}
\hspace{+5mm}
\includegraphics[width=71mm]{f2b.eps}
\vspace{+0mm}\\
\hspace*{-5mm}
\includegraphics[width=72mm]{f2c.eps}
\hspace{+0mm}
\includegraphics[width=72mm]{f2d.eps}
\caption{\label{fig:superKerr_ELOm}
The total energy $E~(=-u_t)$ ({\it top left}), the angular momentum $L~(=u_\phi)$
({\it top right}), the angular velocity $\Omega~(=u^\phi/u^t)$ ({\it bottom left}) and the
angular velocity of the frame-dragging $\omega$ ({\it bottom right}) of the orbiting
particles around a super-spinar ({\it solid lines}) and
a black hole (BH) ({\it dotted lines}).
The values of $a_*$ are selected as 0, 0.9 and 1 for black holes and 1.001,
$a_{\rm cr}=4\sqrt{2}/(3\sqrt{3})~(\sim 1.089)$,1.5, 2 and 5 for super-spinars.
The radius of ISCO is denoted by the filled circles.
}
\end{center}
\end{figure}
In Fig \ref{fig:superKerr_ELOm}, we plot the total energy $E~(=-p_t)$
({\it top left}), the angular momentum $L~(=p_\phi)$ ({\it top right}),
the angular velocity $\Omega~(=u^\phi/u^t)$ ({\it bottom left}) and the
angular velocity of the frame-dragging $\omega$ ({\it bottom right}) of
the orbiting particles around a super-spinar ({\it solid
lines}) and a black hole (BH) ({\it dotted lines}). Here, we have assumed
$s_1=s_2=1$ (positive energy at a large radius, and the prograde orbit). The values
of $a_*$ are selected as 0, 0.9 and 1 for black holes and 1.001,
$a_{\rm cr}=4\sqrt{2}/(3\sqrt{3})~(\sim 1.089)$, 1.5, 2 and 5 for super-spinars.
The radius of ISCO is denoted by the filled circles. For the spin in the range of
$1<a_*<a_{\rm cr}$, the energy at the ISCO becomes negative.
\begin{figure}
\begin{center}
\vspace{+0mm}
\hspace*{-0mm}
\includegraphics[width=140mm]{f3.eps}
\caption{\label{fig:efficiency}
The radiation efficiency $\epsilon~(=1-E_{\rm ISCO})$ with which the gravitational
binding energy converts to the radiation energy when all the photons are escaping
from the disc for the super-spinars with negative ({\it solid line}) and positive ({\it dashed line})
energy at ISCO and a black hole ({\it dotted line}). The locations of the maximally
rotating black hole ({\it filled circle}) and the limit of maximum radiation efficiency
for the super-spinar with the negative energy at ISCO ({\it blank circle}) are also shown.
}
\end{center}
\end{figure}
The total radiation energy of the matter with the circular orbit in the equatorial plane
is equal to the gravitational binding energy of the matter when it is at the ISCO. The
efficiency $\epsilon$ with which the rest mass energy converts to the radiation energy
of photons escaping from the disc is defined as the ratio of the rate of the radiation
energy (=the rate of the gravitational binding energy) and the transportation rate of
mass energy onto the central object. When all the emitted photons
escape from the disc,
for the matter at the ISCO, this efficiency is calculated by using the total energy $E$
at the ISCO as $\epsilon=1-E$. As is well known, for the non-rotating and the
maximally rotating black holes these efficiencies become about 6\% and 42\%,
respectively \cite{mtw73,fn98}. In Fig. \ref{fig:efficiency}, we plot this
efficiency, $\epsilon~(=1-E_{\rm ISCO})$, for the super-spinars with negative ({\it solid line})
and positive ({\it dashed line}) energy at ISCO and a black hole ({\it dotted line})
for the case that all the emitted photons escape from the disc. The locations of the
maximally rotating black hole ({\it filled circle}) and the limit of maximum radiation
efficiency for the super-spinar with the negative energy at ISCO ({\it blank circle}) are also
shown in Fig \ref{fig:efficiency}. For the super-spinar (i.e., $1<a_*$), the efficiency decreases
as $a_*$ increases. For the super-spinar with the spin within $1<a_*<a_{\rm cr}$,
the radiation efficiency is over 100\% (i.e., $1-E_{\rm ISCO}>1$ ). This is
because the energy at ISCO is negative and then we can interpret that positive
energy is extracted from the super-spinar \cite{df78, cn79, rt79, s81b}. It should be noted
that the solution denoted by the dashed line in Fig \ref{fig:efficiency} for a super-spinar is
not considered in this study. This is because this solution have negative energy at
infinity and this is not the case for the accretion disc considered here. So, for a super-spinar,
only the solution denoted by the solid line in Fig \ref{fig:efficiency} is considered in
this study. The upper limit of $1-E_{\rm ISCO}$ is
$(1-E_{\rm ISCO})_{\rm max}=1+1/\sqrt{3}~(\sim 1.577)$, i.e., the upper limit of the
radiation efficiency is about 157.7\% as shown by the blank circle in Fig.
\ref{fig:efficiency}. This is achieved for the super-spinar with the spin which is just
above 1. The efficiency becomes 100\% at the spin $a_*=a_{\rm cr}$, where the
minimum value of the radius of ISCO is achieved as $r_{\rm ISCO}=(2/3)M$.
At the spin of $a_*=5/3~(\sim 1.667)$, the efficiency becomes $\sim 42\%$ which
is the same value as for the maximally rotating black hole. At the spin of
$a_*=8\sqrt{6}/3~(\sim 6.532)$, the efficiency becomes the same values as for the
non-rotating black hole, i.e. 6\%. For the super-spinar with $a_*>8\sqrt{6}/3$, the radiation
efficiency becomes smaller than that of the non-rotating black hole. It is noted that
in these calculations we have assumed that all photons escape from the disc which
is not completely realistic, i.e. in reality some photons emitted from the disc should
be absorbed by the central object, not escaping into infinity \cite{t74}.
Actually, in the vicinity of the black hole and the super-spinar, a large part of
photons are trapped by the strong gravitational field of the central object.
The analysis about this problem is presented in the past studies \cite{b73,s78,s81a}.
However, it can be expected that the efficiency for the super-spinar with the spin of
$1<a_*< 1.667$ is significantly larger than that of the
maximally rotating black hole.
\section{Radiation Flux and Energy Spectrum of an Accretion disc}
\label{sec:spec}
The stress-energy tensor $T^{\mu\nu}$ is given by
$T^{\mu\nu}=\rho_0 h u^\mu u^\nu+u^\mu q^\nu+u^\nu q^\mu+t^{\mu\nu}$,
where $\rho_0$, $h$, $q^\mu$ and $t^{\mu\nu}$ are the rest-mass density,
the relativistic enthalpy, the heat-flux tensor and the viscous tensor, respectively.
For $q^\mu$ and $t^{\mu\nu}$, we have the orthogonality relations as
$u_\mu q^\mu=0$ and $u_\mu t^{\mu\nu}=0$. For the stationary disc, the radiation
flux $\mathcal{F}$ at the disc surface is calculated as $\mathcal{F}=q^\theta$.
From the rest-mass conservation $\nabla_\mu (\rho_0 u^\mu)=0$, where
$\nabla_\mu$ is the covariant derivative. The mass accretion rate $\dot{M}_0$
of the disc is calculated as $\dot{M}_0=-2\pi r \Sigma_0 u^r (={\rm constant})$,
where $\Sigma_0$ is the surface density of the disc which is obtained by the
integration of the rest-mass density along the disc thickness $2H$ as
$\Sigma_0=\int^{H_\theta}_{-H_\theta}\rho_0~rd\theta$, where $H_\theta$ is
the angular thickness of the disc, i.e. $H=r H_\theta$. Here, the radial
component of the four-velocity is set to be negative, i.e. $u^r<0$. From
the energy conservation $\nabla_\mu T^{\mu\nu}h_\nu^{~t}=0$ and the
angular momentum conservation $\nabla_\mu T^{\mu\nu}h_\nu^{~\phi}=0$,
where $h^{\mu\nu}=g^\mu\nu+u\mu u^\nu$ is the projection tensor, we can
obtain $\partial_r [\dot{M}_0 E + 2\pi r W^r_{~t}]=4\pi r \mathcal{F} E$ and
$\partial_r [\dot{M}_0 L - 2\pi r W^r_{~\phi}]=4\pi r \mathcal{F} L$ where
$W^{\mu\nu}$ is defined by the integration of the viscous tensor along the
disc thickness as $W^{\mu\nu}=\int^{H_\theta}_{-H_\theta} t^{\mu\nu}~rd\theta$.
From the orthogonality condition $u_\mu t^{\mu\nu}=0$, we have the relation
$W^r_{~t}=-\Omega W^r_{~\phi}$.
\begin{figure}
\begin{center}
\vspace{+0mm}
\hspace*{-0mm}
\includegraphics[width=140mm]{f4.eps}
\caption{\label{fig:PTflux}
The flux $\mathcal{F}$ [erg/s/cm$^2$] emitted from the geometrically thin and
optically thick disc with the mass accretion rate $\dot{M}=0.1 \dot{M}_{\rm Edd}$
around a super-spinar ({\it solid lines}) and a black hole ({\it dotted lines})
with a mass $M=10M_\odot$. The values of $a_*$ are selected as 0, 0.9 and 1
for black holes and 1.00001, 1.01, $a_{\rm cr}=4\sqrt{2}/(3\sqrt{3})~(\sim 1.089)$,
2, 5 and 9 for super-spinars.
}
\end{center}
\end{figure}
By using these relations,
the local flux at the disc $\mathcal{F}(r)$ [erg cm$^{-2}$ s$^{-1}$] is calculated
as \cite{pt74}
\begin{equation}
\mathcal{F}(r)=\frac{\dot{M}_0}{4\pi r}f(r), \label{eq:flux}
\end{equation}
where $f(r)$ is given by
\begin{equation}
f(r)=\frac{-\partial_r \Omega}{(E-\Omega L)^2}\int^r_{r_{\rm ms}} (E-\Omega L)~\partial_r L~dr.
\end{equation}
Here, the inner boundary of the disc is set to be the radius of the ISCO
inside which there is no torque and no radiation. For a Kerr black hole with
$a_*<1$, the analytic expression for $f(r)$ is given by the Eq. (15n) in \cite{pt74},
which cannot be used for $a_*=1$. For a black hole with $a_*=1$, we can calculate $f(r)$ as
\begin{equation}
f(r)=\frac{3}{2M}\frac{1}{x^2(x+2)(x-1)^2}\left[x-1-\frac{3}{2}\ln x+\frac{3}{2}\ln\left(\frac{x+2}{3}\right)\right]~~~({\rm for}~~~x>1),~~~
\label{eq:fa1}
\end{equation}
where $x=(r/M)^{1/2}$,
and for $x=1$, $f(r)=1/(3M)$. For a super-spinar, i.e. $a_*>1$, we can obtain the analytic form of $f(r)$ as
\begin{eqnarray}
f(r)&=&\frac{3}{2M}\frac{1}{x^2(x^3-3x+2a_*)}
\left\{
x-x_0
+\frac{3a_*^2}{x_*(x_*^2-3)}
\ln(x/x_0)
\right. \nonumber \\
&&\left.
-\frac{(x_*-a_*)^2}{x_*(x_*^2-1)}
\ln\left(\frac{x-x_*}{x_0-x_*}\right)
\right. \nonumber \\
&&\left.
+\frac{1}{2(x_*^2-1)}
\left(
x_*-2a_*-\frac{2a_*x_*}{x_*^2-3}
\right)
\ln\left(\frac{x^2+x_*x+x_*^2-3}{x_0^2+x_*x_0+x_*^2-3}\right)
\right. \nonumber \\
&&\left.
-\frac{6}{(x_*^2-1)\sqrt{3(x_*^2-4)}}
\left(
x_*^2/2-1+a_*x_1+\frac{a_*^2}{x_*^3-3}
\right)
\right.\nonumber\\
&&\left.
\left[
\tan^{-1}\left(\frac{2x+x_*}{\sqrt{3(x_*^2-4)}}\right)-
\tan^{-1}\left(\frac{2x_0+x_*}{\sqrt{3(x_*^2-4)}}\right)
\right]
\right\}
\label{eq:f}
\end{eqnarray}
where
\begin{eqnarray}
x_0&=&(r_{\rm ISCO}/M)^{1/2},\nonumber\\
x_*&=&-(a_*-\sqrt{a_*^2-1})^{1/3}-(a_*+\sqrt{a_*^2-1})^{1/3}.
\end{eqnarray}
Based on Eqs. (\ref{eq:flux}), (\ref{eq:fa1}) and (\ref{eq:f}), we can analytically
calculate the flux $\mathcal{F}$ [erg/s/cm$^2$] emitted from the disc. In
Fig \ref{fig:PTflux}, we plot the flux emitted from the geometrically thin and
optically thick disc with the mass accretion rate $\dot{M}=0.1 \dot{M}_{\rm Edd}$ around
a super-spinar ({\it solid lines}) and a black hole ({\it dotted lines}) with a mass $M=10M_\odot$. The
values of $a_*$ are selected as 0, 0.9 and 1 for black holes and 1.00001, 1.01,
$a_{\rm cr}=4\sqrt{2}/(3\sqrt{3})~(\sim 1.089)$, 2, 5 and 9 for the super-spinars.
For all the calculations presented in Fig. \ref{fig:PTflux}, the radiation flux becomes
zero within the radius of ISCO within which there is no torque. Then, the peak flux is
achieved at a radius which is slightly larger than the radius of ISCO.
As denoted in the previous section, the radiation efficiency for the super-spinar with the spin in the range of $1<a_*<5/3$ is larger than the radiation efficiency of the maximally rotating black hole (see Fig \ref{fig:efficiency}). In Fig \ref{fig:PTflux}, we can see that in the cases when a super-spinar with its spin in the range of $1<a_*<5/3$, for any radius of the disc around such a super-spinar, the radiation flux is larger than that for the maximally rotating black hole. Then, for such a super-spinar, the total flux integrated along the disc surface also becomes larger than that of the maximally rotating black hole. For a super-spinar with the spin in the range $a_*>8\sqrt{6}/3$, the radiation efficiency becomes smaller than that of the non-rotating black hole as seen in the previous section. In the case of $a_*=9$ in Fig. \ref{fig:PTflux}, for any radius of the disc, the radiation flux is
smaller than that for the non-rotating black hole ($a_*=0$). It is interesting that the amount of the local radiation flux for the super-spinar with the spin in $1<a_*<5/3$ can be larger than that of the maximally rotating black hole ($a_*= 1$) by a few orders of magnitude.
\subsection{Temperature of the Accretion disc}
\label{sec:temp}
\begin{figure}
\begin{center}
\vspace{+0mm}
\hspace*{-0mm}
\includegraphics[width=140mm]{f5.eps}
\caption{\label{fig:T}
The radial profile of the effective temperature $T$ [K] [$=(F/\sigma)^{1/4}$] of the geometrically thin and optically thick disc for the same parameters used in Fig. \ref{fig:PTflux}.
}
\end{center}
\end{figure}
By assuming that the local radiation spectrum of the disc surface obeys the blackbody spectrum, it is possible to calculate the effective temperature of the accreting matter. In this case, the effective temperature $T$ of the disc is related to the blackbody flux $\mathcal{F}$ with the Stefan-Boltzmann equation, $\mathcal{F}=\sigma T^4$, where $\sigma$ is the Stefan-Boltzmann constant ($\sigma=5.670\times10^{-5}$ erg s$^{-1}$ cm$^2$ K$^{-4}$). By using the local flux $\mathcal{F}$ calculated from Eqs. (\ref{eq:flux}), (\ref{eq:fa1}) and (\ref{eq:f}) and the Stefan-Boltzmann relation, the effective temperature of the disc can be calculated. In Fig \ref{fig:T}, we plot the radial profile of the effective temperature $T$ [K] [$=(F/\sigma)^{1/4}$] of the geometrically thin and optically thick disc around super-spinars ({\it solid lines}) and black holes ({\it dotted lines}) with the same parameters used in Fig. \ref{fig:PTflux}.
In a similar manner as radiation flux of the disc in Fig \ref{fig:PTflux}, we can see in Fig \ref{fig:T} that in the cases when a super-spinar with its spin in the range of $1<a_*<5/3$, for any radius of the disc around such a super-spinar, the temperature of the accreting matter is larger than that for the maximally rotating black hole. Also, for a super-spinar with the spin in the range $a_*>8\sqrt{6}/3$, the temperature of the accreting matters becomes smaller than that of the non-rotating black hole as seen in the case of $a_*=9$ in Fig. \ref{fig:T}. It is noted that the temperature of the accreting matters for the super-spinar with the spin in $1<a_*<5/3$ becomes larger than that of the maximally rotating black hole ($a_*\sim 1$) by at most a several factor.
The radiation efficiency of the particle at the radius of ISCO for $a_*=1.00001$ is larger than
that for $a_*=1.01$ by about a factor 2 (see Fig. \ref{fig:efficiency}). In addition, the disc
for $a_*=1.00001$ achieves much higher temperature than for the case of $a_*=1.01$ as shown
in Fig. \ref{fig:T}.
\subsection{Observed Energy Spectrum}
\begin{figure}
\begin{center}
\vspace{+0mm}
\hspace*{-0mm}\includegraphics[width=140mm]{f6a.eps}\vspace{3mm}\\
\hspace*{-0mm}\includegraphics[width=140mm]{f6b.eps}
\caption{\label{fig:SED}
The observed spectral energy distribution $\nu L_\nu$ [erg/s] of the geometrically thin and
optically thick disc for the same parameters used in Fig. \ref{fig:PTflux} for the viewing angles
$i=0^\circ$ and $85^\circ$.
}
\end{center}
\end{figure}
Here, we calculate the energy spectrum of the disc observed by the distant observer which is the
function of the mass of the central object $M$, the mass accretion rate $\dot{M}$, the spin of the
central object $a_*$ and the viewing angle between the direction of the observer and the rotation
axis of the disc $i$. The calculations are performed by solving general relativistic radiative
transfer in the Kerr spacetime by including the usual special and general relativistic effects such
as the Doppler boosting, the gravitational redshifts, the light bending and the frame-dragging.
The calculation method of the observed energy spectrum $L_\nu$ [erg/s] is given in
\ref{app:cm}, and this method is basically same as that used in \cite{tw07}.
In Fig \ref{fig:SED}, we plot the spectral energy distribution $\nu L_\nu$ [erg/s] of the geometrically
thin and optically thick disc around super-spinars ({\it solid lines}) and black holes ({\it dotted lines}) for the
viewing angles $i=0^\circ$ ({\it top}) and $85^\circ$ ({\it bottom}). The other parameters are same
as those used in Fig \ref{fig:PTflux}. Here, we assume the outer radius of the disc as $r_{\rm
max}=10^4 M$.
As expected from the calculations given in Figs \ref{fig:PTflux} and \ref{fig:T}, the disc with a larger
maximum temperature produces the energy spectrum extending to higher energy. The part of the
lowest photon energy in the energy spectrum corresponds to the outer region (lowest temperature
region) of the disc. So, for all the cases in Fig \ref{fig:SED}, we have the same energy spectrum in
the part of the lowest photon energy \cite{kfm08}. For the case of $i=85^\circ$ in Fig \ref{fig:SED},
in the part of the middle photon energy in the energy spectrum, for a super-spinar with the spin around
$1.01 \lesssim a_* \lesssim a_{\rm cr} $ the slope of the energy spectrum becomes slightly steeper
than that of others. The slope in the middle part reflects the radial profile of the temperature (see,
Fig \ref{fig:T}). In Fig \ref{fig:T} we can see that for super-spinars with spins of $a_*=1.01$ and $a_{\rm cr}$
have more increasing temperature profiles than other cases. These signatures can be seen in the
middle part of the energy spectrum in the bottom panel of Fig \ref{fig:SED}. As a result, for a super-spinar
with the spin of $1.01\lesssim a* \lesssim a_{\rm cr}$ and the large viewing angle, the slope of the
middle part of the energy spectrum gives the characteristic signature of such a super-spinar. On the other
hand, for other super-spinars (i.e. $1<a_*\lesssim 1.01$ and $a_{\rm cr}\lesssim a_*$) this signature
cannot be seen clearly.
In the same way as the radiation flux and the temperature of the disc in Fig \ref{fig:SED}, we can
see that in the cases when a super-spinar with its spin in the range of $1<a_*<5/3$, for any energy of the
photon emitted from the disc around such a super-spinar, the emitted energy becomes larger than that for
a maximally rotating black hole. For a super-spinar with the spin in the range $a_*>8\sqrt{6}/3$, for any
energy band the emitted energy becomes smaller than that of the non-rotating black hole.
The disc for $a_*=1.00001$ produces photons with
much higher energy. This is because the disc for $a_*=1.00001$ achieves much higher
temperature than for the case of $a_*=1.01$.
\subsection{Contributions of Energy Extracted from a Central Onject in Energy Spectrum}
\begin{figure}
\begin{center}
\vspace{+0mm}
\hspace*{-0mm}
\includegraphics[width=140mm]{f7.eps}
\caption{\label{fig:SEDtpn}
The energy spectrum $\nu L_\nu$ [erg/s] for a super-spinar
with the spin of $a_*=1.01$; the energy spectrum contributed
from the whole region of the disc ({\it solid line}) and the energy spectra
contributed from the positive energy region ({\it dashed line}) and
the negative energy region ({\it dotted line}). The total energy
spectrum is the sum of the energy spectra from the positive and the
negative energy regions.
}
\end{center}
\end{figure}
As already stated, the negative energy region appears in the accretion
disc around a super-spinar with the spin in the range of
$1<a_*<a_{\rm cr}=4\sqrt{2}/(3\sqrt{3})~(\sim 1.089)$. We next investigate the effects of
photons originated from the negative energy region in the energy
spectrum and the total radiation energy. In Fig \ref{fig:SEDtpn}, we
give the energy spectrum $\nu L_\nu$ [erg/s] for a super-spinar
with the spin of $a_*=1.01$; the total energy spectrum
({\it solid line}), the energy spectra contributed from the positive
energy region ({\it dashed line}) and the negative energy region ({\it
dotted line}). The total energy spectrum is the sum of the energy
spectra of positive and negative energy regions. In Fig
\ref{fig:SEDtpn}, we can see that the contributions to the energy
spectrum from the emission originated from the negative energy region is
a relatively minor component. The total bolometric luminosity
$L_B^{\rm total}$ is calculated by the integration of the energy flux as
$L_B^{\rm total}=\int L_\nu~d\nu$. Separately, we can also calculate the
luminosity $L_B^-$ by the integration of the energy spectrum $L^-_\nu$
contributed by the negative energy region as $L^-_B=\int L^-_\nu~d\nu$,
where $L_\nu^-$ is the luminosity calculated from the photons emitted in
the negative energy region as shown by the dotted line in Fig
\ref{fig:SEDtpn}.
\subsection{Can we confirm the Kerr bound from the thermal X-ray spectrum?}
\begin{figure}
\begin{center}
\vspace{+0mm}
\hspace*{-0mm}
\includegraphics[width=140mm]{f8.eps}
\caption{\label{fig:similarSED}
The similar pairs of the energy spectra $\nu L_\nu$ [erg/s] of
the geometrically thin and optically thick disc
for a black hole ({\it dotted lines}) and a super-spinar ({\it solid lines}).
The pairs of the spins for a black hole $a_*^{\rm BH}$ and its
counterpart super-spinar $a_*^{\rm super-spinar}$ are
$(a_*^{\rm BH},~a_*^{\rm super-spinar})=(0,~6.56)$, (0.5, 4.8) and (0.9,~3) ({\it left to right}).
}
\end{center}
\end{figure}
We finally give the examples of the very similar pairs of the energy
spectra of a black hole and a super-spinar. In Fig
\ref{fig:similarSED}, we plot the similar energy spectrum $\nu F_\nu$
[erg/s] of the geometrically thin and optically thick disc for
a black hole ({\it dotted lines})
and
a super-spinar ({\it solid lines}).
The pairs of the spins for a black hole $a_*^{\rm BH}$
and a super-spinar $a_*^{\rm super-spinar}$ are $(a_*^{\rm BH},~a_*^{\rm super-spinar})=(0,~6.53)$, (0.5, 5) and (0.9,~3.35)
({\it left to right}). As shown in this figure, surprisingly, for given black holes with some value of the spin $a_*(\le 1)$, we can always find its counterpart objects with the spin $a_*$ larger than the unity whose observed spectrum is very similar to and practically indistinguishable from that of the black hole. As a results, we can not confirm the Kerr bound only by using the X-ray thermal spectrum of the black hole candidates.
\section{Discussion}
\label{sec:dis}
Here, we discuss about important topics relating to this study.
After giving the discussion mainly in the research of the gravitational physics,
we discuss about astrophysical topics.
In this paper, we consider the object which is described by the Kerr metric but
have specific angular momentum larger than its mass. For such object,
no emission from the object is assumed and we implicitly assume the general
relativity near curvature singularity is replaced by other gravitational theory
described in the Introduction. These assumptions have the background as follows.
Since the last century, extensive investigations have been performed on
the research of black holes and naked singularities, around which the strong-field
of gravity is achieved and the nonlinearity of gravitation is prominent. In particular,
the cosmic censorship conjecture \cite{p69,p79,w97,p98} has been proposed but
its proof is yet very limited. On the other hand, it has been revealed that there are
many examples of solutions to the Einstein field equations which have naked
singularities and contain physically reasonable matter fields,although no such
example has been proven to be completely generic (see, e.g. \cite{harada04,joshi00}).
This conjecture is often useful to deduce the properties of spacetimes and
black holes (see, e.g. \cite{he73}). and hence many researchers on
classical general relativity tend to assume it. However, from a quantum gravity
point of view, the motivation is not so clear to believe that the cosmic
censorship must hold within classical general relativity and, in fact, it has been
pointed out that naked singularities in classical theory can be viewed as
a window into new physics including quantum gravity~\cite{hn04}.
If the specific angular momentum is greater than its mass in the Kerr solution,
which is a stationary, axisymmetric and vacuum solution to the Einstein field
equations, there is no horizon but naked singularity. However, it might be
reasonable to consider that around the singularity classical general relativity
is broken down and actually there is no singularity at the center \cite{bf09}.
In such cases, some physical mechanism such as quantum gravity effects
replace the singularity with some finite radius $R$. One can imagine
the radius $R$ is very small but do not know how. Even within classical
theory, the supercritically rotating Kerr solution might approximately describe
the geometry exterior to a rapidly rotating compact object. In the past studies,
based on quantum field theory in curved spacetime, the explosive emission
from a forming naked singularity in gravitational collapse has been argued as
a possible observational signature of naked singularities
\cite{hwe82,bsvw98a, bsvw98b,hin00a,hin00b,hin02}. Although
these explosive signatures might appear in the observational signatures,
in this study we assume no emission from the central objects.
We do not know whether these assumptions are reasonable or not, but the similar
assumptions are used in the past studies (see the references in the following
paragraphs).
Before this study, there are a lot of past studies about the observational
feasibility of the super-spinning objects and/or naked singularity in a
variety of astrophysical contexts such as; direct radio observations of
a super-spinning Kerr object~\cite{bf09, hm09}, accretion disc around
a super-spinning Kerr object~\cite{ts05}, gravitational lensing
phenomena by a super-spinning Kerr object~\cite{ve02,vk08,gy08,wp07},
light rays from a forming naked singularity~\cite{d98,nki03},
particle creation (emission) from a forming naked singularity~
\cite{bsvw98a,bsvw98b,hin00a,hin00b,vw98,sv00},
gravitational radiation from a forming naked singularity~
\cite{nsn93,ihn99a,ihn99b,ihn00}, physical processes in naked
singularity formation~\cite{hin02, jdm02,jgd04}, connection to
gamma-ray bursts (GRBs)~\cite{jdm00,a99}.
Most of the past calculations related to general relativistic
accretion disc models assume a black hole as the central compact object
in the center of the accretion disc. If the observational data contain
the physical information in the strong-field of gravity, these data can
be used to test the assumption of the black hole as the central object
and/or test the theory of gravitation including general relativity
in the strong-field regime. Based on these motivations, recently several authors
considered and/or calculated the accretion disc model around the central objects
except the black hole in general relativity \cite{btb01,t02,ynr04,g06,pkh08,hkl08,phk08}.
These calculations give the emissivity profiles of the discs surrounding the central objects
including quark, boson, or fermion stars, wormholes and brane-world
black holes, or discs in $f(R)$ modified gravity models.
It is widely accepted that most of the astrophysical black hole candidates discovered
so far consist of the central object and the viscous accretion disc system.
Therefore, in order to identify the objects with the super-spinning objects,
it is essential to study the observational signatures of the viscous accretion
disc around the assumed objects. Since the last century,
a number of black hole candidates have been discovered by astronomical
observations through the electromagnetic signatures (from radio to
X-ray/gamma-ray) from the accreting matters plunging onto the central objects.
Such radiation from the vicinity of the central objects contains the information
about the space-time structure and the plasma in the strong gravitational field.
Strictly speaking, to identify the central object with a black hole, we need to
show not only that the observation is explained by the assumption that
the central object is a black hole but also that it cannot be explained by
the assumption that the central object is anything else. Although the test
of gravitational theory is out of scope in the present paper, even general
relativity is not a trivial assumption because it has never been so
accurately tested in such a strong-field regime as around black holes.
In this context, it is essentially important to clarify the relationships between
the physics in the strong gravity and the observational features of
the accreting plasma such as electromagnetic energy spectrum.
In the context of the cosmic censorship, the observational identification
of the central object provides a rather direct astrophysical test.
For this purpose, it is at least required to find the distinguishable
observational features between black holes and super-spinning objects
(or naked singularities)~\cite{bf09,ve02,wp07}.
As denoted in the Introduction, the X-ray spectrum of the black hole
candidate generally consists of the thermal component originating
from the accretion disc around the black hole and the non-thermal
component originating from the high energy photons which are
up-scattered in e.g. corona above or in the accretion disc.
In this study, we only assumed the thermal component for simplicity.
However, this is not valid especially for the hard X-ray
spectrum. We know that there are many astrophysical objects with
high energy radiation which can not be naturally explained by
assuming the accretion disc used in this study. Especially, hard
X-ray and gamma-ray radiation from the observed objects can
not be simply explained by the standard disc around a black hole.
For the explanation of such observed high energy radiation, past
researchers proposed many physical processes such as the
inverse-Comptonization of the corona near the central object for
the hard X-ray emission (e.g. \cite{st80,t94,ps96,mlm00,lms02, c09}),
dark matter annihilation for X-ray and gamma-ray radiation
(e.g. \cite{ubel02,cflm04,bfp06,hfd07,dhs08}), and other
many processes \cite{an05}. Although these emissions are
also expected around a super-spinning object, if exists, these are the topics for
the future studies. In addition to the geometrically thin accretion
discs considered in this paper, many kinds of the discs are
proposed. Among the variety of types of the accretion discs/flows,
most basic one is the so-called standard disc or Shakura-Sunyaev
disc \cite{ss73}, whose state is achieved when the mass accretion
rate $\dot{M}$ is sub-Eddington, i.e. $\dot{M}< L_{\rm Edd}/c^2$,
where $L_{\rm Edd}$ is the Eddington luminosity given by
$L_{\rm Edd}=1.25\times10^{39}(M/10M_\odot)$ erg/s. Here,
$M$ is the mass of the central object. For this mass accretion
rate, the disc becomes geometrically thin and optically thick.
This standard disc can be applied to black hole candidates in
black hole binaries and active galactic nuclei. The general
relativistic version of the standard disc is given for the first time
in \cite{nt73} and \cite{pt74}. The accretion discs assumed in this
study belongs to this type. Based on the accretion disc theory,
for the other range of the mass accretion rate, different forms of
the disc structures are realized; for example, radiatively
inefficient accretion flow or advection-dominated accretion
flow for much smaller mass accretion rate, supercritical accretion
disc or slim disc for super-critical mass accretion rate,
hypercritical accretion disc or neutrino-dominated accretion
flow for hyper-critical mass accretion rate (for review,
see e.g. \cite{kfm08}). For such accretion discs with a different mass
accretion rate, since the physical processes in the disc and
the equation of state are different, the resultant energy spectrum
also becomes different from the results given in this paper.
Especially, it is important to investigate the observational
signatures of the radiatively inefficient accretion flow in the
Galactic Center, where the direct imaging observations by
the radio interferometers will be performed in the near future.
It is expected that the direct imaging observations will
determine the background spacetime geometry such as
the spin parameter \cite{d08, b73,fma00,t04,t05,bl06,m07}.
These studies will be performed in the future.
\section{Conclusions}
\label{sec:con}
The observational confirmation of the Kerr bound directly suggests the existence of a black hole.
In this study, in order to investigate testability of this bound by using
the observed X-ray energy
spectrum of black hole candidates, we first calculate the energy spectrum for the object whose
spacetime geometry is described by the Kerr metric but
whose specific angular momentum is larger than its mass, and then compare
the results with that of a black hole. We call this object a super-spinar in this study.
The optically thick and geometrically thin disc is assumed and only
the thermal energy spectrum seen by the distant observer is calculated
by general relativistic radiative transfer calculations including
usual special and general relativistic effects such as Doppler boosting,
gravitational redshift, light bending and frame-dragging.
After calculating a disc structure such as velocity fields (Fig\ref{fig:superKerr_ELOm})
and radiation efficiency at ISCO (Fig\ref{fig:efficiency}), we have calculated
energy flux radiated from the disc (Fig \ref{fig:PTflux}), disc temperature (Fig{\ref{fig:T}}) and
observed energy spectrum (Fig \ref{fig:SED}). We use the new analytic formula
for the radiation flux of a disc. As known in past studies, some energy is extracted from the central
objects whose specific angular momentum is larger than its mass. We have investigated
the influence of the extracted energy on the energy spectrum of a disc. Finally, we compare the
energy spectra of a super-spinar and that of a black hole.
In terms of the energy spectrum observed by a distant observer, we have obtained the
following results:
\begin{itemize}
\item For the super-spinar with $1<a_*<5/3\simeq 1.667$,
higher energy photons are emitted from the disc than those from the disc
around the maximally rotating black hole (see Fig\ref{fig:SED}). This
signature can be seen especially for the cases with large viewing angles,
e.g. the case with $i=85^\circ$ in Fig \ref{fig:SED}.
\item For the super-spinar with $1.01\lesssim a_* \lesssim 1.1$ and its large viewing
angle, the slope of the middle energy part of the energy spectrum becomes
slightly steeper than that of the case of the black hole (see, e.g., the case
for $i=85^\circ$ in Fig \ref{fig:SED}).
\item The influence of the extracted energy from a super-spinar on energy spectrum is negligible
(see, Fig \ref{fig:SEDtpn}). That is, most of the radiation energy comes from
the accreting matters with positive energy even when the energy is maximally
extracted from the super-spinar.
\item For a given black hole, we can always find its super-spinning counterpart
in the range $5/3<a_{*}<8\sqrt{6}/3$ whose observed spectrum is very similar
to and practically indistinguishable from that of the black hole
(see Fig \ref{fig:similarSED}).
As a result, we conclude that to confirm the Kerr bound we need more than
the X-ray thermal spectrum of the black hole candidates.
Although in principle black holes and super-spinars can be distinguished
by the detailed observations of the energy spectrum,
the distinction between the black
holes and the super-spinars only by the steady-state emergent spectrum
becomes a severe challenge to the future observational facilities.
\item For the super-spinar with $a_{*}>8\sqrt{6}/3 \simeq 6.532$, the total
radiation energy of the disc is lower than the disc around the non-rotating
black hole.
\end{itemize}
As a result of this study, we found, surprisingly, that for a given black hole
we can always find its super-spinning counterpart whose observed
spectrum is very similar to and practically indistinguishable from that of
the black hole. Then, in order to confirm the Kerr bound
we need more than the X-ray thermal spectrum of the black hole candidates.
\section*{Acknowledgments}
The author would like to thank Akira Tomimatsu, Masaaki Takahashi,
Cosimo Bambi, Naoki Isobe, Hitoshi Negoro, Makoto Miyoshi and Mareki Honma
for valuable comments and discussion. The author also thank anonymous referees
for useful comments and suggestions which improve the original manuscript.
The authors are supported by the Grant-in-Aid for Scientific
Research Fund of the Ministry of Education, Culture, Sports, Science
and Technology, Japan [Young Scientists (B) 18740144 and 21740190 (TH);
21740149 (RT)].
\section*{References}
|
\section{Introduction}
At present, it is believed that the high field and low temperature (HFLT) phase of the heavy fermion superconductor CeCoIn5 with a $d$-wave pairing symmetry is a Fulde-Ferrell-Larkin-Ovchinnikov (FFLO) vortex lattice with a paramagnetically induced modulation {\it parallel} to the applied magnetic field \cite{B1,RI1}. This identification is based on several key observations \cite{W,M,T} and their consistency with theoretical results \cite{RI1}. For instance, the instability of HFLT phase via Cd or Hg doping \cite{T}, which induces a localized \cite{P} antiferromagnetic (AFM) order, seems to imply a destruction of long range order of the FFLO modulation and is well understood if the HFLT phase includes an {\it inhomogeneous} AFM order \cite{RI1}. In this sense, the FFLO picture of the HFLT phase of CeCoIn$_5$ is {\it compatible} with the recent observation of AFM order there \cite{K}. Nevertheless, a relation between the HFLT phase and the AFM {\it fluctuation}, which is believed to be enhanced near $H_{c2}(0)$ \cite{R}, has not been well understood yet. Here, we will report on some results of our study on the vortex state taking account of both the strong paramagnetic depairing {\it and} the AFM quantum fluctuation enhanced near $H_{c2}(0)$.
\section{Phase diagram in ${\bf H} \perp c$}
As mentioned earlier \cite{RI2}, stability of the FFLO state with a modulation parallel to the field against those with lateral modulations is ensured by going {\it beyond} the weak-coupling BCS approach and including some quasiparticle damping. In CeCoIn$_5$ with strong AFM fluctuation {\it near} $H_{c2}(0)$, this fluctuation-induced damping seems to be a main origin of making the states with {\it lateral} modulations relatively unstable, although the AFM fluctuation tends to suppress even the FFLO state with the longitudinal modulation (see Fig.1). On the other hand, the AFM fluctuation does not necessarily suppress the first order $H_{c2}$-transition, which is another peculiar feature in CeCoIn$_5$ in high fields \cite{B1,W}, and seems to also have a partial role of inducing the first order $H_{c2}$ transition. An example of the $H$-$T$ phase diagram in ${\bf H} \perp c$ we obtain in terms of a microscopically-derived Ginzburg-Landau (GL) free energy functional including the AFM fluctuation is shown in Fig.1. Since an applied pressure in real experiments corresponds to a reduction of AFM fluctuations, the dependences of the two (mean-field) transition curves and the onset of the first order $H_{c2}$-transition on the AFM fluctuation strength are consistent with the features found in experiments \cite{M}.
\begin{figure}[t]
\begin{center}
\includegraphics[scale=0.43]{af005m2s.eps}
\caption{The resulting $H_{c2}(T)$ (dashed curves) and the FFLO transition curves (solid ones) in ${\bf H} \perp c$ for a smaller (red) and larger (black) strength of the AFM fluctuation. Each arrow indicates the onset of the first order $H_{c2}$-transition. The upturn of $H_{c2}(T)$ below $t=0.2$ is an artifact of the expansion on the wavenumber of the FFLO modulation and can be improved. Here, $h$ is the reduced magnetic field normalized by the orbital-limiting field in 2D case, and $t=T/T_{c0}$.
. \label{fig:ph}}
\end{center}
\end{figure}
\section{Flux distribution in ${\bf H} \parallel c$}
Neutron scattering data on CeCoIn$_5$ in ${\bf H} \parallel c$ show \cite{B2} the vortex lattice form factor enhanced with increasing the field, which is an opposite trend to the conventional one in type II superconductors. Although such an anomalous behavior has been explained in intermediate fields as a result of strong Pauli-paramagnetic effect \cite{IM}, this interpretation is insufficient in higher fields where the AFM fluctuation is not negligible. If the damping effect suppressing the paramagnetic depairing is the main consequence of the AFM fluctuation, it seems difficult that the above-mentioned anomaly is attributed to the AFM fluctuation. However, the AFM fluctuation also leads to an effective increase of the electron correlation and thus, of the paramagnetic depairing, implying that the field dependence of roles of the AFM fluctuation is not intuitively clear. For this reason, we have investigated the flux density distribution of the vortex lattice based on the microscopically derived GL free energy. The flux distribution follows from the Maxwell equation. Expanding up to the lowest order in the squared pair field, $|\Delta|^2$, and also in a local value of AFM fluctuation correlator $\chi$ ($> 0$), we obtain the superconducting contribution to the longitudinal flux density parallel to the applied field in the form
\begin{equation}
B^{(s)}({\bf r}) = 4 \pi C_{\rm G} \, |\Delta|^2 \, [ \, \tilde{B}({\bf r}) + \chi \, \tilde{B}^{\rm AF}({\bf r}) \, ],
\end{equation}
where $C_{\rm G}$ is a scale factor measuring the magnitude of the local flux density in the familiar Abrikosov lattice near $T_c$. Figure 2 shows spacial dependences of $\tilde{B}({\bf r})$ and $\tilde{B}^{\rm AF}({\bf r})$ on the coordinate between neighboring two vortices for the two cases of a small (dotted lines) and large (solid lines) Maki parameter $\alpha_M$, which measures the Pauli-paramagnetic effect. The feature in $\tilde{B}$ that the magnetic flux is accumurated near the vortex center with increasing the paramagnetic depairing is qualitatively the same as that in Ref.\cite{IM}. A crucial result in this study is that, as the red curves show, the fluctuation contribution $\tilde{B}^{\rm AF}$ is competitive with $\tilde{B}$ for small Maki parameters, while it assists and {\it enhances} the flux accumuration in the vortex core induced by the paramagnetic effect for large enough $\alpha_M$. It appears that, among roles of the AFM fluctuation, the mass enhancement of quasiparticles will overcome an increase of the quasiparticle damping. This enhanced flux accumuration due to the AFM fluctuation should be a direct origin of the unresolved enhancement \cite{B2} of the structure factor in higher fields and seems to imply an apparent presence of an AFM quantum critical point near $H_{c2}(0)$.
\begin{figure}[t]
\begin{center}
\includegraphics[scale=0.6]{VCpaulih01t01J02M001-65n.eps}
\caption{Dependences of ${\tilde B}({\tilde {\bf r}})$ (black) and ${\tilde B}^{\rm AF}({\tilde {\bf r}})$ (red) at $t=0.1$ and $h=0.1$ on the coordinates ranging from a vortex center (${\tilde r}=0$) to the midpoint (${\tilde r}=1.4$) between the two vortices. Here, a triangular lattice was assumed in a d-wave pairing. The dotted and solid curves correspond to those for $\alpha_{M}=0.01$ and $6.5$, respectively. \label{fig:vortexcore}}
\end{center}
\end{figure}
\section{Acknowledgement}
The work is supported by the Japan Society for the Promotion of Science.
|
\section{Introduction}
It might be useful, in order to account for the complexity of the world, to embed the observed world in a higher-dimensional space. This allows, by proper engineering, to give a geometric origin for the low energy features of models. Extra dimensions have thus been introduced about a century ago by Nordstr\"{o}m, Kaluza and Klein \cite{Nordstrom:1988fi}. Such a program has then been pursued with a renewal of interest in the last decades (see for example \cite{Antoniadis:1990ew,ADD,DDG,Intermediate,RS,PQ}). Of particular interest for us, there are many attempts to engineer supersymmetry (SUSY) breaking mechanisms, classify them, and obtain experimental predictions. In these studies, the importance of extra dimensions depends on the relative sizes of the supersymmetry breaking scale ($M_{SUSY}$) and the compactification one ($M_C$). For $M_{SUSY}\ll M_C$, one can restrict the analysis to the four-dimensional effective theory (for a review, see for example \cite{Kitano:2010fa}). There, a knowledge of the data of the extra dimensions allows to "understand" the fields content and their interactions. In contrast, for $M_{SUSY}\gtrsim M_C$, the analysis should be performed in the higher dimensional theory. We are interested in the latter case.
A way of breaking supersymmetry with extra-dimensions is the Scherk-Schwarz mechanism \cite{Scherk:1978ta}, where different
higher dimensional supersymmetries are conserved at different points. For instance, in five dimensional models
compactified on an interval, $\mathcal{N}=2$ SUSY is preserved in the bulk, while at each of the two boundaries a different $\mathcal{N}=1$
supercharge survives, leading to supersymmetry fully broken in the four-dimensional effective theory (see for example \cite{Antoniadis:1990ew,PQ,ADPQ,DPQ,Barbieri:2000vh,MP,Diego:2006py}). Here, we shall be interested in a straightforward generalization of such a scenario, where instead of two branes at the boundaries, one deals with many branes at different points of the extra dimension(s).
Assumptions on the size, shape, and content of extra dimensions, turn in fact into assumptions on the very early history of the universe.
As the extra dimensions are expected to have a small volume, their evolution can be thought to be short, ending before nucleosynthesis to avoid unobserved variations of fundamental constants. Even if, for such a small size, all parts are causally related, we want to argue that in the presence of branes the homogenization of this space might not have been efficient enough. In an alternative to the usual scenario of a symmetric internal space, we assume that the cooling has been very fast, may be in a non-adiabatic way, and as a consequence, the internal space was not driven all to a single ground state. Instead, we will assume that a situation similar to the domains of ferromagnets might arise.
In section 2, we describe the scenario of branes represented as a disordered system of spins living in the internal dimensional space. This is to be contrasted with the usual wisdom of considering symmetric internal spaces where, in a brane-world framework, the disorder consists in the presence of a single anti-brane as the source of SUSY breaking. In the disordered extra dimension set-up discussed here, the "defects" separating different supersymmetric sub-spaces play a major role, as they are responsible for the breaking of (super)symmetry. Section 3, discusses the case of one extra-dimension, and illustrates how a "domain wall" can be approximated as a localization of a gravitino mass making further computations simpler. In fact, effects of localized gravitino masses are well studied in one extra dimension for the case of an interval with two boundaries\cite{Bagger:2001qi,Meissner:2002dg,Delgado:2002xf}, or including many branes\cite{Benakli:2007zza}. In contrast, the case of localized mass in six dimensions has not yet been discussed to our knowledge, and it will be treated in section 4. Most of the content of these two sections is original material that can be read independently of the rest. We end the paper by some conclusions.
\section{The world as a lattice and the branes as spins}
As for the observable ones, properties of the extra dimensions, such as geometry and topology, could be determined by the very early history of the universe. Contrary to the large observed space-time, very little is known about the evolution of the (small volume) internal space; some geometrical parameters have to be frozen before nucleosynthesis as their values are associated with those of fundamental constants, whose variations are very constrained. The interaction between different parts of the extra dimensions can be achieved very quickly ($ t< 10^{-13}$s), as all parts can be causally connected. The possibility that some degrees of freedom in the internal space are at a finite and sizable temperature is not excluded.
The history of the internal space is assumed to proceed through three steps (i) space-time is nucleated (ii) diverse bubbles are created in the extra-dimensions where the branes are in excited states preserving no supersymmetry (iii) while the "temperature" decreases, the branes move to minimize their energy and different patches have branes that are ``pointing'' towards a different supersymmetry. The internal space finishes frozen in a (meta-stable) state made of "domains" described by different ground states.
The branes can be located at arbitrary points in the extra-dimensions. Patches in well defined supersymmetric ground states have branes positioned as long molecules of a liquid crystal in a nematic phase. One can also consider the branes on a regular lattice, the inter-branes spacing is then fixed by some kind of Van der Waals forces \cite{ArkaniHamed:1998kx}, or by the equilibrium between forces due to a combination of electric and dyonic charged branes \cite{Corley:2001rt,Denef:2000nb}. Such mechanisms appeal to physics at scales of order or smaller than the inter-branes separation, which needs the knowledge of the field content of the fundamental theory. The existence of such a possibility to locate the branes at fixed position would imply the non-supersymmetric vacuum to be (meta)stable, and also stabilizes the size of the extra-dimension \cite{ArkaniHamed:1998kx}. As we stress again in the conclusions, achieving the stability of non-supersymmetric configurations is not obvious, and remains an open issue in string theory.
For each brane $i$, we associate a vector, we will denote as the spin $\vec{S_i}$, corresponding to the central charge in the $\mathcal{N}=2$ super-algebra which describes the direction of the supersymmetry preserved by the brane. The spins $\vec{S_i}$, with unit norm $|\vec{S_i}|^2=1$, live in a two-dimensional space spanned by the unitary orthogonal basis vectors $(\vec{e}_X, \vec{e}_Y)$:
\begin{equation}
\vec{S}_i = \cos{\theta_{i}}\ \vec{e}_X + \sin{\theta_{i}} \ \vec{e}_Y
\label{Hamilton1}
\end{equation}
and we will define $\vec{e}_Z=\vec{e}_X \wedge \vec{e}_Y$.
While, in the extra-dimensions discussed here, the branes appear as localized spins at some points, in other smaller dimensions they are wrapping cycles intersecting at angles that define the associated spin. In treating the branes as a spin system, one can have: i) long range interactions (This is the case for toroidal constructions in string models) ii) the values for the spins to take discrete values. Because of these, there would be no fluctuation destruction of long-range order following the Mermin-Wagner-Berezinskii theorem, and supersymmetry ordering is expected in all dimensions. Moreover, the spins system is supposed to be nowadays at zero temperature. Unless stated otherwise,\textit{ we will always implicitly assume the spins to take discrete values} but we shall consider finite range spin-exchange interactions.
Working in an effective description, we can take for the Hamiltonian of this system the very simple form:
\begin{eqnarray}
\mathcal{H}&=& \sum_{i,j} J_{ij} |\vec{S}_i \wedge \vec{S}_j|^p+ \sum_{i}a_i |\vec{S}_i \wedge \partial_t \vec{S}_i|^q - \sum_{i} \vec{H}_i \cdot \vec{S}_i \\
&=& \sum_{i,j} J_{ij} |\sin{(\theta_{j}-\theta_{i})}|^p+ \sum_{i} a_i |\partial_t {\theta}_i|^q -\sum_{i} b_i \cos({\theta}_i-{\theta}^{(H)}_i)
\label{Hamilton2}
\end{eqnarray}
where we have used $\vec{H}_i = b_i (\cos{\theta^{(H)}_{i}}\ \vec{e}_X + \sin{\theta^{(H)}_{i}} \ \vec{e}_Y)$. Here, $\{p,q\}$ are exponents that depend on the microscopic theory, we take to be $p=q=2$ .
Let us discuss each term:
\begin{itemize}
\item The first term tends to align all the spins, giving a supersymmetric ground state with all the branes preserving the same $\mathcal{N}=1$ supersymmetry. The interaction strength $J_{ij}$ are positive with strength and action range depending on the microscopic model.
\item The second term is included to take into account that the rotation of the spin requires a rotation of the brane which costs a non-zero amount of energy. This is because the supersymmetry preserved by the brane is associated to its orientation in other smaller extra dimensions. The rigidity $a_i$ could be related to the brane tension, and it does not need to be the same at each point.
\item The third term describes the effect of the background fields. It will determine the non-trivial vacuum structure as it leads to constraints that pull the branes towards a specific direction. This can contain, for example, the effect of the presence of specific boundary conditions in the internal space (as orientifold objects).
\end{itemize}
It is the last term of Eq. (\ref{Hamilton1}) that parametrizes the source of supersymmetry breaking. Our aim then would be to find simple forms of $ \vec{H}_i$ that, on one side, lead to a minimum of the Hamiltonian with broken supersymmetry, and on the other side, can {\it a posteriori} be understood at the level of a microscopic fundamental theory.
The system of $N$ spins with constrained boundaries can be thought to evolve during the early time from the initial conditions as
\begin{equation}
\frac{\partial \vec{H}_i}{\partial t}=\vec{F}_i(\vec{H}_j, \vec{S}_j ) \qquad \mbox{with } \qquad \vec{H}_0(t)=\vec{H}_0, \qquad \vec{H}_N(t)=\vec{H}_N
\label{b2g}
\end{equation}
where the exact form of $\vec{F}_i$ requires the knowledge of the microscopic theory. A simple phenomenological approximate of such equation could take the form
\begin{equation}
\vec{F}_i(\vec{H}_j, \vec{S}_j )=\sum_{j}\alpha_{ij} \vec{H}_j\wedge \vec{H}_i + \sum_{j}\beta_{ij} \vec{S}_j \wedge \vec{H}_i
\label{b2f}
\end{equation}
that tend to bring the vacuum to aligned spins but for the imposed non-supersymmetric boundary conditions. With appropriate $\alpha_{ij},\beta_{ij}$, one can achieve a background made of domains preserving different supersymmetries.
As stated above, the problem of supersymmetry breaking is now formulated as the problem of obtaining appropriate forms of the constraints $ \vec{H}_i$ and study their properties as the phase space and the correlation function in the corresponding spin model.
Examples of constraints that allow obtaining ground states with domains where the spins point towards different directions can be constructed either by having a strong localized force, or a weak long range force, in both cases opposing the effect of the first term of (\ref{Hamilton1}) . We will illustrate this through examples for one extra dimension.
The first option can be realized as:
\begin{equation}
\vec{H}_i=\sum_I b_I \exp[{-\frac{(y_i-Y_I)^2}{\Delta}}](f_I(y)\vec{e}_X + g_I(y)\vec{e}_Y); \qquad b_I \gg J_{ij}
\label{b22}
\end{equation}
where $f_I,g_I$ are slowly varying functions that force the spins to change directions around the point $Y_I$. For example, if the points $Y_I =0, \pi R$ are the boundaries of a compact dimension, we can take:
\begin{equation}
\vec{H}_i= b_0 \exp[{-\frac{(y_i)^2}{\Delta}}]\vec{e}_X + b_\pi \exp[{-\frac{(y_i-\pi R)^2}{\Delta}}] \vec{e}_Y
\label{b2b}
\end{equation}
Another example is used in the next section.
The second option can also be obtained from the interplay of weak and long range forces. For instance, the aligning force is taken to be effective only between nearest neighbors, while parametrizes an anti-alignment force at long range:
\begin{equation}
\vec{H}_i=\sum_{j\neq i} \frac{b}{(y_j-y_i)^\alpha} \vec{S}_j
\label{b2c}
\end{equation}
where $\alpha$ is a positive integer. We take $J_{ij}$ to a nearest neighbor interaction, and $b$ positive and small $b\ll J$, we see that the energy of the system is lowered for spins aligned in domains, but a large amount of such spins generate an interaction that tends to flip the spins. The net result is the formation of domains with different alignments separated by transition regions.
In this picture of disordered extra dimension, the supersymmetry breaking is localized in the "defects" separating different supersymmetric sub-spaces. It is the subject of the next sections. We will assume that the fundamental length scale $\kappa$, the inter-brane separation $d_i$ and the compactification scale $R$ are well separated $\kappa \ll d_i\ll R$. The first inequality allows to neglect quantum geometry effects, the second to get a large number of branes. When needed for the purpose of explicit computations, we will also use a flat metric.
\section{A defect in one extra dimension}
We consider the space-time extended with a fifth dimension with coordinate $y \in [0, \pi R]$. Along this direction $N + 1$ branes are located at the points $y = y_{i}$, $i = 0 \cdots N$ with $y_{0} = 0$, $y_{N} = \pi R$, and $y_{n} < y_{n+1} $.
We assume that the evolution of the universe has ended in a non-supersymmetric vacuum. The branes orientation varies when going from one to the opposite boundary, such that the distribution of the associated spins has the form of a localized kink. Obviously, non-trivial boundary conditions are needed such that at different regions the (spin) branes have to point to different directions. An abrupt separation of domains would be costly in spin-exchange energy. In fact, with the combination of boundary conditions, spin-exchange, and long range interactions, the minimization of the total energy will induce a kink with a certain thickness $\Delta'$ localized around a point $Y$, as for ferromagnets. As translation invariance is broken by both the presence of boundaries and the conditions imposed to break of supersymmetry, the minimization of the total energy will fix the value of $Y$ (i.e. gives masses to the collective modes of the soliton). Because, we are not interested here in this issue of fixing the moduli, we choose a final configuration, and illustrate how it can be parametrized as the result of an applied ``magnetic field'' $H_i$ on the spin system. Our aim will be to show how our order parameter, the gravitino mass, is related to the final spin configuration.
We will first show how a kink localized at a single position $Y$ can be described as if the branes are embedded in a background field $\vec{H}_i$. We consider the constraint:
\begin{equation}
\vec{H}_i=b \exp[{-\frac{(y_i-Y)^2}{\Delta'}}]( \cos{(\frac{y_i}{\Delta})} \vec{e}_X + \sin{(\frac{y_i}{\Delta})} \vec{e}_Y)
\label{b2d}
\end{equation}
which means that at the boundary $y=0$, the spins at forced to point in the direction $\vec{e}_X $ while at $y=\pi R$, the spins at forced to point along $\vec{e}_Y $, as well as
\begin{equation}
J_{i j} = J[ \delta_{j,i+1}+ \delta_{j,i-1}]; \qquad a_i=a
\label{J2}
\end{equation}
The Hamiltonian is given by:
\begin{equation}
\mathcal{H}= \sum_{i} J \sin^2{(\theta_{i}-\theta_{i+1})}+ \sum_{i} a |\partial_t {\theta}_i|^2 -\sum_{i} b \exp[{-\frac{(y_i-Y)^2}{\Delta'}}] \cos{( \theta_i-\frac{y_i}{\Delta})}
\label{H2}
\end{equation}
The minimum of the potential is obtained for:
\begin{equation}
\sin{2(\theta_{i}-\theta_{i+1})}= -\frac{b}{J} \exp[{-\frac{(y_i-Y)^2}{\Delta'}}] \sin{( \theta_i-\frac{y_i}{\Delta})}
\label{min1}
\end{equation}
Taking $b\gg J$, we obtain the following limits:
In the vicinity of Y, $y_i \sim Y$, we can approximate
\begin{equation}
\sin{( \theta_i-\frac{y_i}{\Delta})} = -\frac{J}{b} \sin{2(\theta_{i}-\theta_{i+1})} \rightarrow 0 \qquad \mbox{i.e. } \qquad \theta_i \rightarrow \frac{y_i}{\Delta}
\label{min2}
\end{equation}
while elsewhere:
\begin{equation}
\sin{2(\theta_{i}-\theta_{i+1})} \rightarrow 0 \qquad \mbox{i.e. } \qquad \theta_{i} \rightarrow \theta_{i+1}
\label{min3}
\end{equation}
This means we have two patches of aligned spins, separated by a region of size of order $2 \Delta'$. In his transition region, the spins rotate by an angle of order $2 \Delta'/ \Delta$, which if not a multiple of $\pi$, implies that the two patches preserve two different supersymmetries. It is only this quantity that will be relevant for our purpose. Of course, one can compute the energy carried by the interface between the two domains, and study the process of bubbles nucleation, the size of the interface, or the processes of homogenization, all well known issues.
We are interested to describe here the effect of the simplest such defect in a five-dimensional supergravity. The total action is given by the sum of a bulk and brane components:
\begin{equation}
S = \int^{2 \pi R}_{0} dy \int d^{4}x \left[ \frac{1}{2} {\cal L}_{BULK} +
\sum_{i = 0}^{N} {\cal L}_{i} \delta({y - y_{i}}) \right] .
\label{ActionN}
\end{equation}
The brane $n$ will be characterized by the supersymmetry it preserves, which is correlated with the choice of the couplings to the bulk operators, in particular the gravitino. The non-vanishing set of such operators $\Phi_{even}$ are determined as those being even under a $\mathbb{Z}_2$
action at the point $y = y_{i}$:
\begin{equation}
\Phi_{even}(y_{i} + y) = {\cal P}_{i} \Phi_{even}(y_{i} - y)=
\Phi_{even}(y_{i} - y).
\label{ParityN}
\end{equation}
The operators might be themselves made of products of even numbers of odd fields. The supersymmetry preserved by the brane $i$, associated with the ``spin'' $\vec{S_i}$ can be read from the gravitino components $\psi_{\mu +}^{~i}$ which couples to it, while it breaks the orthogonal supersymmetry direction associated to $\psi_{\mu -}^{~i}$. We can choose a basis $(\psi_{\mu 1} , \psi_{\mu 2})$ for the $\mathcal{N}=2$ gravitino in term of two-components spinors, and
define:
\begin{eqnarray}
\psi_{\mu +}^{~i} & = &
\cos{ 2\theta_{i} } \ \psi_{\mu 1} + \sin{ 2\theta_{i} } \ \psi_{\mu 2}
\nonumber \\
\psi_{\mu -}^{~i} & = &
\! \! \! \! \! - \sin{2 \theta_{i} }\ \psi_{\mu 1} + \cos{2 \theta_{i} } \ \psi_{\mu 2}
\nonumber \\
\psi_{5 +}^{~i} & = &
\sin{2 \theta_{i} }\ \psi_{5 1} + \cos{ 2\theta_{i} } \ \psi_{5 2}
\nonumber \\
\psi_{5 -}^{~i} & = &
\cos{2 \theta_{i} }\ \psi_{5 1} - \sin{2 \theta_{i} }\ \psi_{5 2}
\label{ParityEigenvectorsN}
\end{eqnarray}
Without lost of generality, we will take $\theta_{0} = 0$. The spin $\vec{S_i}$ of the chain makes an angle $\theta_i$ with $\vec{S_0}$.
We shall now consider the case of a single domain wall separating two phases. the generalization to more than one domain is straightforward. At leading order, we shall consider the branes world-volumes to be supersymmetric. The whole supersymmetry breaking is then concentrated in the transition interval $\left[ y_n, y_{n+1}\right]$ of length $d_n$. It is encoded in the wave function of the gravitino which interpolates between the two values $\psi_{\mu +}^{~n}$ and $\psi_{\mu +}^{~n+1}$. This variation is associated with a gravitino mass $M_{n} (\theta_n, d_n)$. We take a configuration where:
\begin{itemize}
\item all spins $\vec{S_i}$, $0\leqslant i \leqslant n$ are aligned with $\vec{S_0}$, thus $\theta_{0} = ...= \theta_n=0$. We will denote this as the phase $A$ .
\item all spins $\vec{S_i}$, $n+1\leqslant i \leqslant N$ are aligned with $\vec{S_N}$, thus $\theta_{n+1} = ...= \theta_N=2 \theta$. We will denote this as the phase $B$.
\end{itemize}
For the purpose of the illustration, we can take the extra dimension to be flat, thus $M_{n} (\theta, d_n) = \theta/ d_n$. The gravitino wave function associated to the supersymmetry preserved on the left side of the defect is given by:
\begin{eqnarray}
\psi_{\mu 1}(y) = \left\{
\begin{array}{lll}
1 & \mbox{for } y \in [0, y_{n}] \\
\cos{ ( \frac{y-y_n}{d_n}2 \theta)} & \mbox{for } y \in [y_{n}, y_{n+1}] \\
\cos{2 \theta} & \mbox{for} y \in [y_{n+1}, \pi R]
\end{array}
\right.
\end{eqnarray}
while the orthogonal one is given by
\begin{eqnarray}
\psi_{\mu 2}(y) = \left\{
\begin{array}{lll}
0 & \mbox{for } y \in [0, y_{n}] \\
\sin{ ( \frac{y-y_n}{d_n} 2\theta)} & \mbox{for } y \in [y_{n}, y_{n+1}] \\
\sin{2 \theta} & \mbox{for } y \in [y_{n+1}, \pi R]
\end{array}
\right.
\end{eqnarray}
such that the right side of the defect preserves the combination: $ \cos{ 2 \theta } \psi_{\mu 1} + \sin{2 \theta } \psi_{\mu 2} $.
For a bulk observer outside the domain $\left[ y_n, y_{n+1}\right]$, in the limit $d_n\ll \pi R$ the breaking of supersymmetry can be accounted a variation of the bulk field (here the gravitino's) wave function between the points $y_n$ and $y_{n+1}$. We would like to describe the wave function outside the defect in the limit where the latter can be considered as point-like, i.e $d_n \rightarrow 0$.
In this limit, we describe the five-dimensional gravitino by two wave functions: a continuous one $\Psi_{C}$ that couples to the defect with a mass $M_n$, and a discontinuous one, $\Psi_{D}$, that does not couple. We can find the respective values building these functions in the interval $\left[ y_n, y_{n+1}\right]$:
\begin{eqnarray}
\Psi^{C n}_{\mu}(y) & = & c_{\theta} \psi^{n}_{\mu 1}(y)+ s_{\theta} \psi^{n}_{\mu 2}(y)\\
\Psi^{D n}_{\mu}(y) & = & s_{\theta} \psi^{n}_{\mu 1}(y)- c_{\theta} \psi^{n}_{\mu 2}(y)
\end{eqnarray}
where:
\begin{eqnarray}
c_{\theta}= \cos{ \theta}, && s_{\theta}= \sin{ \theta}
\end{eqnarray}
In the limit $d_n \rightarrow 0$, $y_n = y_{n+1}= Y_n$, the gravitino component that couples to the defect domain wall at $y_n = y_{n+1}= Y_n$ is given by the even wave function value:
\begin{eqnarray}
\Psi^{C n}_{\mu}(Y^<_n)= \Psi^{C n}_{\mu}(y_n) = c_{\theta}= \Psi^{C n}_{\mu}(y_{n+1}) = \Psi^{C n}_{\mu}(Y^>_n)
\end{eqnarray}
while the orthogonal component
\begin{eqnarray}
\Psi^{D n}_{\mu}(y_n) & = &s_{\theta}= - \Psi^{D n}_{\mu}(y_{n+1})
\end{eqnarray}
is odd and corresponds to the a gravitino component that does not couple to the defect.
We can now build, in this limit, an effective wave function:
\begin{eqnarray}
\Psi^{C }_{\mu}(y) = \left\{
\begin{array}{llll}
c_{\theta} & \mbox{for } y \in [0, Y_n] \\
s_{\theta} & \mbox{for } y \in [Y_{n}, \pi R]
\end{array}
\right.
\end{eqnarray}
orthogonal to
\begin{eqnarray}
\Psi^{D }_{\mu}(y) = \left\{
\begin{array}{lll}
c_{\theta}& \mbox{for } y \in [0, Y_{n}] \\
- s_{\theta} & \mbox{for } y \in [Y_{n}, \pi R]
\end{array}
\right.
\end{eqnarray}
The breaking of supersymmetry by defect can now be described as due to a localized mass term $M_n$. Such a localized mass gives rise to the equations of motion for the gravitinos $\Psi^{I n}_{\mu}$(we assume $e^{\hat{5}}_{5} =
1$):
\begin{eqnarray}
\partial_{5}\Psi^{D }_{\mu}
+ m_{3/2} \Psi^{C }_{\mu }
&=& 2 M_n \Psi^{C }_{\mu} \delta(Y_n)
\nonumber \\
\partial_{5} \Psi^{C}_{\mu}
- m_{3/2}\Psi^{D}_{\mu}
&=& 0
\label{GravitinosEOM2}
\end{eqnarray}
where we have used the four-dimensional equation of
motion for gravitinos of mass $m_{3/2}$:
\begin{equation}
\epsilon^{\mu \nu \rho \lambda}
\sigma_{\nu}\partial_{\rho}\overline{\Psi}^{I}_{\lambda } =
- 2 m_{3/2} \sigma^{\mu \nu} \Psi^{I }_{\nu } \qquad I=C, D
\label{MassiveGravitinoEOM}
\end{equation}
It can be clearly seen from equations (\ref{GravitinosEOM2}) that while $\Psi^{C }_{\mu}$ is a
continuous field, $\Psi^{D i}_{\mu }$ has a
jump at the point $y = Y_n$, its first derivative being proportional to a
Dirac $\delta$ distribution. We can then identify the gravitino mass as:
\begin{equation}
M_n = \kappa^{-1} \tan{ \theta_{n+1}} = \kappa^{-1} \frac {(\vec{S}_{n+1}\wedge \vec{S_{n}}) \cdot \vec{e}_Z }{(\vec{S}_{n+1}\cdot \vec{S_{n}})}
\label{MassiveGravitino}
\end{equation}
Given the knowledge of the localized gravitino mass in 5d, we can use this result to derive the four-dimensional gravitino mass.
\begin{equation}
m_{3/2} = \frac{1}{\pi R} \arctan{( \frac {(\vec{S}_{n+1}\wedge \vec{S_{n}}) \cdot \vec{e}_Z }{(\vec{S}_{n+1}\cdot \vec{S_{n}})}})
\label{MassiveGravitino2}
\end{equation}
As an illustration of this simple case, let us consider $\kappa \lesssim d_n \sim$ TeV$^{-1}$. The observable sector lives on a 4-brane extended between the two points $y_n$ and $y_{n+1}$, that is part of a large extra dimension of size $\pi R$ responsible for the hierarchy between the string and the Planck length. We can use the previous example to see that the resulting gravitino mass is $ \theta_{n+1}/\pi R$. In the case of a system of brane-anti-brane, $ \theta_{n+1}= \pi/2 $ leads to $m_{3/2} = 1/2R$. As explained in \cite{Benakli:2007zza} for the case of an explicit localized F-term, the breaking can not be compensated by opposite twists in other parts of the extra dimension.
\section{A localized defect in two dimensions }
\label{secSixDSugra}
In this section, we will illustrate the case of supergravity with two extra dimensions, i.e. in six dimensions. The defects can be either a one-dimensional curve or a point. The latter can appear in the spin system as the zero size limit of a vortex. The gravitino wave functions can be taken in the absence of supersymmetry breaking as holomorphic (or anti-holomorphic) function of the complex coordinate $z= x^5+ix^6$ describing the two extra dimensions. When a gravitino mass $m_0$ localized at the point $z_0$ is included for the component ${\psi}_{\mu 1}$, it appears as a flux in the circulation of the gravitno wave function ${\psi}_{\mu 2}$, of the form
\begin {equation}
\oint_{\partial S} {\psi}_{\mu 2} dz = -i m_1 \int_S {\psi}_{\mu 1} dx^5 dx^6 = - 2 im_0 {\psi}_{\mu 1}(z_0)
\label {GravitinoKineticTerm1}
\end {equation}
where $S$ is a surface containing the point $z_0$ and having as boundary $\partial S$, while $m_1$ is the bulk mass appearing in the equation of motion of ${\psi}_{\mu 1}$. In this section, we will derive the resulting lightest four-dimensional gravitino mass.
The two extra dimensions are taken compactified on the orbifold $ T^2 / \mathbb { Z }_2 $ parametrized by the coordinates $(x^5 , x^6)$. The torus $T^2$ coordinates obey $(x^5 , x^6) \equiv (x^5 + 2 \pi m R_5, x^6 + 2 \pi n R_6)$, $(m,n)\in \mathbb{Z}$, and the orbifold is obtained through the identification ${(x^5 , x^6)} \equiv {-(x^5 , x^6)}$. There are four fixed points of this action at $(0,0)$, $(\pi R_5,0)$, $(0,\pi R_6)$ and $(\pi R_5, \pi R_6)$. We will consider the simplest case with a single defect located at the origin $(x^5 , x^6) = (0,0)$.
The bulk Lagrangian volume must describe the six-dimensional supergravity. The supermultiplets of supergravity in six dimensions in its minimal form are the sechsbein $ e^a_m$, the gravitino $ \Psi_m$, a real scalar field $ \Phi $, the dilaton; a fermion $ X $ the dilatino; and the Kalb-Ramond two-form denoted by $ B_{MN} $ which gives rise to the three-form $ H = 3 \partial_{[B_M {NP}]} $. The action of supergravity in the volume is $ N = 2 $ supersymmetric as it preserves eight supercharges. Our study focuses on the gravitino. Its standard kinetic term reads:
\begin {equation}
{\cal L}_{kin} = - i E_6 M^2_6 \overline {\Psi}_M \Gamma^{MNP} D_{N} \Psi_{P}
\label {GravitinoKineticTerm}
\end {equation}
where $M_6 = \kappa^{-1} $ is the fundamental Planck mass in six dimensions, $E_6 $ the sechsbein determinant. It is useful to express this in two-components spinor notation:
\begin{eqnarray}
{\cal L}_{kin} &=& \kappa^{-2} e_6 \bigg[
\frac{1}{2} \epsilon^{\mu \nu \rho
\lambda} \left(
\overline{\psi}_{\mu 1}\overline{\sigma}_{\nu}D_{\rho}\psi_{\lambda 1}
+ \overline{\psi}_{\mu 2}\overline{\sigma}_{\nu}D_{\rho}\psi_{\lambda 2}
\right)
\nonumber \\
&&
+ \psi_{\mu 1} \sigma^{\mu \nu} \left( D_{\hat{5}} + i D_{\hat{6}} \right) \psi_{\nu 2}
- \psi_{\mu2} \sigma^{\mu \nu} \left( D_{\hat{5}} + i D_{\hat{6}} \right) \psi_{\nu1}
\nonumber \\
&&
- \left( \psi_{\hat{5}1} + i \psi_{\hat{6}1} \right) \sigma^{\mu \nu}D_{\mu}\psi_{\nu2}
+ \left( \psi_{\hat{5}2} + i \psi_{\hat{6}2} \right) \sigma^{\mu \nu}D_{\mu}\psi_{\nu 1}
\nonumber \\
&&
- \psi_{\mu 1} \sigma^{\mu \nu}D_{\nu} \left( \psi_{\hat{5}2} + i \psi_{\hat{6}2} \right)
+ \psi_{\mu 2} \sigma^{\mu \nu}D_{\nu} \left( \psi_{\hat{5}1} + i \psi_{\hat{6}1} \right)
\bigg]+ h.c.
\label{KineticTerm2}
\end{eqnarray}
To define this theory in the orbifold $ T^2 / \mathbb {Z}_2 $ we must impose parity of various fields under the action of the symmetry $ \mathbb{Z}_2 $ in a manner consistent with the action of supergravity and supersymmetry transformations. Expressing $ \Psi_m $ and $ X $ in two-components spinor notation, (i) the fields $ e^a_{\mu} $, $ e^{i_{j}} $, $ B_{\mu \nu} $, $ B_{ij} $, $ \Phi $, $ \psi_{\mu 1} $, $ \psi_{i 2} $ and $ \chi_1$ are taken even under $ \mathbb{Z}_2 $ (ii) the fields $ e^i_{\mu} $, $ e^{a}_{i} $, $ B_{\mu i} $ $ \psi_{\mu 2} $, $ \psi_{i 1} $ and $ \chi_2$ are odd under the $ \mathbb {Z}_2 $ action. Here the indices $ i, j$ denote coordinates of the extra dimensions: $ i, j \in \{5,6\} $.
The defect is located at the fixed point of the orbifold $ (x^5, x^6) = {(0,0)} $, therefore only the operators even under the $ \mathbb {Z }_2 $ action couple to it. We are interested in the case where a constant localized four-dimensional gravitino mass:
\begin {equation}
{\cal L}_{mass} = - e_4 \delta {(x^5)} \delta {(x^6)}
\left (M_0 \psi_{\mu 1} \sigma^{\mu \nu} \psi_{\nu 1}+ hc \right)
\label {BraneMassTerm}
\end {equation}
is present, as well as new bi-linear terms which mix the four-dimensional gravitino $ \psi_{\mu 1} $ with the internal dimensional components $ \psi_{5 2} $ and $ \psi_{6 2} $. The constant $ M_0 $ is proportional to the value $ W_0 $ of the localized superpotential: $ M_0 = \sqrt {g_{\hat {5} \hat {5}}} \, W_0 $. A necessary step is gauge fixing. A possible choice is the unitary gauge where the terms bi-linear mixing the four-dimensional gravitino $ \psi_{\mu } $ fields with $ \psi_{5} $ and $ \psi_{6} $ are absent, so that the part of the Lagrangian which describes the bi-linear terms for the gravitino is given by:
\begin{eqnarray}
{\cal L}_{k + m} &=& \kappa^{-2} \left[
\frac{1}{2} \epsilon^{\mu \nu \rho
\lambda} \left(
\overline{\psi}_{\mu 1}\overline{\sigma}_{\nu} \partial_{\rho}\psi_{\lambda 1}
+ \overline{\psi}_{\mu 2}\overline{\sigma}_{\nu} \partial_{\rho}\psi_{\lambda 2}
\right)
+ 2 \psi_{\mu 1} \sigma^{\mu \nu} \left( \partial_{5} + i \partial_{6} \right) \psi_{\nu 2}
\right]
\nonumber \\
&&
- \delta(x^5) \delta(x^6) M_0 \psi_{\mu 1} \sigma^{\mu \nu} \psi_{\nu 1}
+ h.c.
\label{KineticAndMassTerms}
\end{eqnarray}
To study the properties of the gravitino there are two approaches: one can study its equations of motion and boundary conditions as done in the one-dimensional case, or we can study the theory reduced to four dimensions. In this section we follow the second method.
First, we Fourier expand the gravitinos $ \psi_{\mu 1} (x^{\mu}, x^5, x^6) $ and $ \psi_{\mu 2} (x^{\mu}, x^5, x^6)$, taking into account their parities under the $ \mathbb {Z}_2 $ action:
\begin{eqnarray}
\psi_{\mu 1}(x^{\mu},x^5, x^6) &=& \frac{\kappa}{\sqrt{\pi^2 R_5 R_6 }} \left[
\frac{1}{\sqrt{2}}\psi_{\mu 1}^{0}(x^{\mu})
+ \sum_{p,q \in Y} \psi_{\mu 1}^{p,q}(x^{\mu})
\cos \left( \frac{p x^5}{R_5} + \frac{q x^6}{R_6} \right) \right]
\nonumber \\
\psi_{\mu 2}(x^{\mu},x^5, x^6) &=& \frac{\kappa}{\sqrt{\pi^2 R_5 R_6 }}
\sum_{p,q \in Y} \psi_{\mu 2}^{p,q}(x^{\mu})
\sin \left( \frac{p x^5}{R_5} + \frac{q x^6}{R_6} \right)
\label{FourrierExpansion}
\end{eqnarray}
with the sum over $Y$ is defined as: $\sum_{p,q \in Y} = \sum^{p = + \infty}_{p = 1}\sum^{q = + \infty}_{q = - \infty} + \left[ \sum^{q = + \infty}_{q = 1} \right]_{{p = 0}} $ . When plugged in (\ref {KineticAndMassTerms}), it gives:
\begin{eqnarray}
{\cal L}_{k + m} &=& \frac{1}{2} \epsilon^{\mu \nu \rho \lambda} \left[
\overline{\psi}^0_{\mu 1}\overline{\sigma}_{\nu}\partial_{\rho}\psi^0_{\lambda 1}
+ \sum_{p,q \in Y} \overline{\psi}^{p,q}_{\mu 1}\overline{\sigma}_{\nu}\partial_{\rho}\psi^{p,q}_{\lambda 1}
+ \sum_{p,q \in Y} \overline{\psi}^{p,q}_{\mu 2}\overline{\sigma}_{\nu}\partial_{\rho}\psi^{p,q}_{\lambda 2}
\right]
\nonumber \\
&&
- \frac{M_0 \kappa^2}{\pi^2 R_5 R_6} \left[
\frac{1}{\sqrt{2}} \psi_{\mu 1}^0 + \sum_{k,l \in Y} \psi_{\mu 1}^{k,l} \right]
\sigma^{\mu \nu} \left[
\frac{1}{\sqrt{2}} \psi_{\nu 1}^0 + \sum_{p,q \in Y} \psi_{\nu 1}^{p,q} \right]
\nonumber \\
&&
+ 2 \sum_{p,q \in Y} \psi^{p,q}_{\mu 1} \sigma^{\mu \nu}
\left( \frac{p}{R_5} + i \frac{q}{R_6} \right) \psi^{p,q}_{\nu 2}
+ h.c.
\label{4dLagrangian}
\end{eqnarray}
Note that the phases in the masses which appear in the Lagrangian have no physical consequences, because the phases of the masses of Kaluza-Klein $ \frac {p} {R_5} + i \frac {q} {R_6}$ can eliminated by the redefinition of fields $\psi^{p, q}_{\nu 2} $. A phase in localized masses $ M_0 $ may also be eliminated by a redefinition of the fields $ \psi_{\mu 1}^0 $ and $ \psi_{\nu 1}^{p, q} $. We can then do so the following substitutions:
\begin{equation}
\frac{p}{R_5} + i \frac{q}{R_6} \rightarrow
\sqrt{\frac{p^2}{R_5^2} +\frac{q^2}{R_6^2}} = m_{p,q}.
\label{KKMass}
\end{equation}
with a change of basis $\psi_{\mu +}^{p,q} = \frac{1}{\sqrt{2}} \left[ \psi_{\mu 1}^{p,q} + \psi_{\mu 2}^{p,q} \right]$ and $\psi_{\mu -}^{p,q} = \frac{1}{\sqrt{2}} \left[ \psi_{\mu 1}^{p,q} - \psi_{\mu 2}^{p,q} \right]$.
In the new basis $\psi_{\mu}^{\lambda} = {\psi_{\mu 1}^0, \psi_{\mu +}^{p,q}, \psi_{\mu -}^{p,q}}$ the bi-linear terms can be expressed as:
\begin{equation}
{\cal L}_{k + m} = \frac{1}{2} \sum_{i} \epsilon^{\mu \nu \rho \lambda}
\overline{\psi}^i_{\mu}\overline{\sigma}_{\nu}\partial_{\rho}\psi^i_{\lambda}
- \sum_{i,j} \psi^i_{\mu} M_{3/2~ij} \sigma^{\mu \nu} \psi^j_{\nu}
+ h.c.
\label{4dLagrangian2}
\end{equation}
and the gravitinos mass matrix takes the form
\begin{equation}
M_{3/2} =
\left(\begin{matrix}
m_0 & m_0 & m_0 \\
m_0 & m_0 -\delta_{kp} \delta_{lq} m_{p,q} & m_0 \\
m_0 & m_0 & m_0 +\delta_{kp} \delta_{lq} m_{p,q} \\
\end{matrix}\right) \qquad m_0 = \frac{M_0 \kappa^2}{2 \pi^2 R_5 R_6}
\label{GravitinoMassMatrix}
\end{equation}
We will now diagonalize the mass matrix and obtain the eigenvalues and eigenvectors. We denote by $ \Psi_m$ the eigenvector associated with the eigenvalue $ m $. It can be written in the above basis (\ref {4dLagrangian2}) as $ \Psi_m = (\Psi_m^0, \psi_{+ m}^{p , q}, \psi_{m -}^{p, q}) $. With these notations, the equations that define the vectors and eigenvalues of the mass matrix is $ M_{3/2} \psi_{m} = m \psi_{m} $ takes the form:
\begin{eqnarray}
m_0 \left[ \psi_m^0 + \sum_{p,q \in Y} \psi^{p,q}_{m +}
+ \sum_{p,q \in Y} \psi^{p,q}_{m -} \right] &=& m \psi_m^0
\nonumber \\
m_0 \left[ \psi_m^0 + \sum_{p,q \in Y} \psi^{p,q}_{m +}
+ \sum_{p,q \in Y} \psi^{p,q}_{m -} \right]
- m_{p,q} \psi^{k,l}_{m +}&=& m \psi^{k,l}_{m +}
\nonumber \\
m_0 \left[ \psi_m^0 + \sum_{p,q \in Y} \psi^{p,q}_{m +}
+ \sum_{p,q \in Y} \psi^{p,q}_{m -} \right]
+ m_{p,q} \psi^{k,l}_{m -} &=& m \psi^{k,l}_{m -} .
\label{EigenEq}
\end{eqnarray}
Some straightforward algebra leads then to the eigenvalues equation:
\begin{equation}
\sum_{p,q = - \infty}^{+ \infty} \frac{m_0}{m^2 - \frac{p^2}{R_5^2} -\frac{q^2}{R_6^2} } = \frac{1}{m}.
\label{MassEq2}
\end{equation}
We note that the double infinite sum in this equation has a logarithmic divergence. A regularization procedure is needed and leads to a result dependent on the ultraviolet cutoff. A "truncation" of the sum leads for the lowest eigenvalue gravitino mass:
\begin{equation}
\frac{1}{m_0 m} \simeq - \pi R_5 R_6.\ln \left( \Lambda^2 R^2 \right) +\frac{1}{m^2}.
\label{MassEqAprox}
\end{equation}
when taking $R \approx R_5 \approx R_6$. Retaining only the dominant terms in $M_0^2 \kappa^4 \ln \left( \Lambda R\right) /{R_5 R_6}$, we get:
\begin{equation}
\frac{1}{m} \simeq \frac{1}{m_0} + \frac{M_0 \kappa^2}{\pi} \ln\left(\Lambda R\right) .
\label{MassAprox}
\end{equation}
We see that for a small size of the extra dimensions ($R \Lambda \sim 1$), we recover the effective four-dimensional result $M_{3/2} \sim m_0$. On the other hand, for very large radius (typically $ \ln\left( {\Lambda R}\right) > 2 \pi$), we can instead have
\begin{equation}
M_{3/2} \simeq \frac{m_0}{ 2 \pi m_0^2 R^2 \ln\left( {\Lambda R}\right)}
\label{MassAprox2}
\end{equation}
which is reduced compared to $m_0$.
We describe now the wave functions for eigenstates of the gravitinos. According to (\ref {FourrierExpansion}) the eigenvectors of the mass matrix $ M_{3 / 2} $ (we have denoted $ \psi_{m \, \mu} $) can be written as:
\begin{equation}
\psi_{m \, \mu 1}(x^{\mu},x^5, x^6) = \frac{\kappa N e^{i \beta}}
{\sqrt{\pi^2 R_5 R_6 }} \left[ \frac{1}{\sqrt{2}}
+ \sum_{p,q \in Y} \frac{\sqrt{2} m^2
\cos \left( \frac{p x^5}{R_5} + \frac{q x^6}{R_6} \right)
}{m^2 - \frac{p^2}{R_5^2} -\frac{q^2}{R_6^2}}
\right] \chi_{m \, \mu}(x^{\mu})
\label{GravitinoEigenstate1}
\end{equation}
\begin{equation}
\psi_{m \, \mu 2}(x^{\mu},x^5, x^6) = - \frac{\kappa N m \sqrt{2} }{\sqrt{\pi^2 R_5 R_6 }}
\sum_{p,q \in Y}
\frac{ \sqrt{ \frac{p^2}{R_5^2} + \frac{q^2}{R_6^2} }}
{m^2 - \frac{p^2}{R_5^2} -\frac{q^2}{R_6^2}}
e^{i \alpha_{p,q}}
\sin \left( \frac{p x^5}{R_5} + \frac{q x^6}{R_6} \right) \chi_{m \, \mu}(x^{\mu}).
\label{GravitinoEigenstate2}
\end{equation}
In these expressions the field $ \chi_{m \, \mu} $ is a spinor that does not depend on extra dimensions $ (x^5, x^6) $. It is a massive spin $ 3/2$ state with mass $ m $ as given by the equation (\ref {MassEq2}). The phases $ e^{i \beta} $ and $ e^{i \alpha_{p, q}} $ are given by:
\begin{equation}
e^{i \beta} = \sqrt{\frac{\left| M_0 \right|}{M_0}}
,\quad
e^{i \alpha_{p,q}} =
\frac{\sqrt{ \frac{p^2}{R_5^2} + \frac{q^2}{R_6^2} }}{ \frac{p}{R_5} + i \frac{q}{R_6} }
e^{- i \beta}.
\label{Phases}
\end{equation}
The normalization constant $ N $ can be determined by imposing unitary standard eigenvector $ \Psi_m = {\Psi_m^0, \psi_{m 1}^{p, q}, \psi_{m 2 }^{p, q}} $:
\begin{equation}
N = \left[
\sum_{p,q = - \infty}^{+ \infty} \frac{2 m^2}
{\left(m^2 - \frac{p^2}{R_5^2} -\frac{q^2}{R_6^2}\right)^2}
- \frac{m}{m_0}
\right]^{-1} .
\label{NormConst}
\end{equation}
One can check explicitly that the wave functions (\ref {GravitinoEigenstate1}) and (\ref {GravitinoEigenstate2}) are solutions of the equations of motion of the gravitinos in six dimensions.
The divergence in the tree level computation (\ref{MassRunningSolut}) of the gravitino brane mass
arises here because of the $\delta(x_5)\delta(x_6)$ singularity due to the zero brane thickness limit. It
shows that the field theory considered here is not a valid description of the physics in the UV
as the internal structure of the brane cannot be neglected. This behavior is well known, it has been encountered in \cite{Antoniadis:1993jp} ( see also \cite{Giudice:1998ck}), and it was shown that in a fundamental theory, as in string models, it is finite, regularized by an effective thickness \cite{Antoniadis:1993jp,Antoniadis:2000jv}.
The sensitivity to cut-off scale of the theory can be interpreted as a classical running of the mass parameter between the cut-off and the compactification scales, and it can be re-summed. It was studied in six-dimensional models with
orbifold compactifications as a tree level renormalization of brane coupling constants \cite{Goldberger:2001tn}. While these properties have been discussed
for particles of spin $0$ and $1/2$, here we can generalize this phenomenon for a spin $3/2$ gravitino with brane localized masses. The logarithmic divergence in (\ref{MassAprox}) can be absorbed by defining a bare coupling $M_0$ is replaced by the renormalized coupling $M_{0}^{ren}$:
\begin{equation}
\frac{1}{M_0} = \frac{1}{M_{0}^{ren}} - \frac{M_0 \kappa^4}{2 \pi^3 R_5 R_6} \ln\left(\Lambda R\right).
\label{MassRen}
\end{equation}
which implies the following running for the gravitino brane mass:
\begin{equation}
\mu \frac{d}{d \mu} M_0(\mu) = \frac{\kappa^4}{2 \pi^3 R_5 R_6}
\left[ M_0(\mu) \right] ^3.
\label{MassRunning}
\end{equation}
This has the solution:
\begin{equation}
M^2_0(\mu) = \frac{ M^2_0(\mu_0)}
{1 - \frac{\kappa^4}{\pi^3 R_5 R_6} M_0^2(\mu_0)
\ln\left( \frac{\mu}{\mu_0} \right) } .
\label{MassRunningSolut}
\end{equation}
If $M_0$ is positive then equation (\ref{MassRunningSolut}) shows that
the mass $M_0$ increases in the UV and would reach a Landau singularity at
$ \mu = \mu_0 \exp \left[ \frac{\pi^3 R_5 R_6}{\kappa^4 M_0^2(\mu_0)} \right]$.
\section{Conclusions}
Our interest in this work, is a situation where spontaneous supersymmetry breaking is described in a higher dimensional theory. The presence of many localized objects (branes) coupled to bulk fields forces on the latter specific boundary conditions. When the bulk wave functions have to interpolate between the different boundary conditions, as in the Scherk-Schwarz mechanism, this leads to supersymmetry breaking. We try to formulate these multiple boundary conditions as a system of spins, forced to point in peculiar directions by a constraining field $ \vec{H}_i$. The amount of supersymmetry breaking is measured by the departure from the total alignment of all the spins. The problem of building a particular configuration comes back to find the appropriate field $ \vec{H}_i$, and then, a microscopic realization of it.
The relevance of this picture requires that the fundamental scale, $\kappa \sim M_s^{-1}$, the branes separation distance $d_i$, and the compactification radius $R \sim M_c^{-1}$ to be separated as $\kappa \ll d_i\ll R$. A non-exhaustive set of examples is given by:
\begin{itemize}
\item the very large extra dimensions as introduced by \cite{ADD} responsible of the weakness of the strength of four dimensional gravitational interactions; these have compactification scales as low as $M_c \sim 10^{-4} -10^{7}$ eV. Along these, we can suppose the presence of 3-branes separated by typically $d_i\sim$ TeV$^{-1}$ distances. The observable world lives on some of these 3-branes or on higher branes stretched between them.
\item The so-called "large volume" scenario \cite{Intermediate}, with a fundamental scale in the intermediate energies $M_s\sim 10^{11}$GeV, with electroweak scale compactification radius $M_c \sim M_w$ , where the branes are separated by distances smaller than TeV$^{-1}$.
\end{itemize}
Our formulation aims to study the phase space of a number of branes located at well separated points in extra-dimensions, to discuss the "landscape" of configurations that break supersymmetry, with estimate of its size. Given such a configuration, one needs to explain the origin of the constraint $\vec{H}_i$, as well as information about the microscopic details at scales below the inter-branes distance which is relevant for many phenomenological issues.
Finally, we would like to comment on the possibility of embedding such a scenario, with discrete values for the spins $\vec{S_i}$, in a string theory framework. In the very early history of the internal space, one can imagine that some cycles shrink to a very small size. If a cycle carries some (quantized) Ramond-Ramond flux, this might give birth to a (stack of) D-brane(s) at this point (see for example \cite{Heckman:2007ub}). Different shrinking cycles can be located at points separated by potential barrier (due to wrapping factors for example) which make them potential wells where D-brane are located. The latter are driven there in order to minimize their energy contribution through the wrapping effect \cite{Giddings:2001yu}. A construction of compactifications with both branes and anti-branes exist, see for example \cite{Dabholkar:2001gz}. However, there is an issue of the stability of the configuration, either by annihilation between the brane and anti-brane, or due to decompactification pushing them infinitely away from each other. Often, studies of non-supersymmetric branes in toroidal compactifications ignore the instability problem, having another aim as to try to find exact conformal field theory descriptions of systems of brane-anti-branes. Most recent studies have concentrates on trying to find a meta-stable configuration of a single anti-brane (negatively charged) in a background with positive charge Ramond-Ramond flux (see \cite{Klebanov:2000hb}). Assuming that one can explain why the anti-brane appears at that point, and then accounts correctly for the back reaction on the backgrounds, it remains to check that the antibrane will not annihilate too quickly with part of the fluxes such as discussed in \cite{Kachru:2002gs}. Building meta-stable string backgrounds with broken supersymmetry, even with a minimal number of branes, remains an intersecting open issue.
The main purpose here is to try instead to have a description where the whole complex system is parametrized by an effective Hamiltonian as a spin system in a bottom-up approach to the brane models constructions. The branes play here a role similar to the one of atoms in solid state physics, and some of the machinery of spin systems could be applied.
\section*{Acknowledgments}
I thank my former student C. Moura for collaboration on some parts of this work in its early stage, A. Dabholkar, B. Dou\c{c}ot and Y. Oz for useful discussions. This work is supported in parts by the European contract "UNILHC" PITN-GA-2009-237920.
|
\section{Introduction}
The zeros of the derivative $\zeta'(s)$ of the Riemann
zeta-function are intimately connected with the behavior of the
zeros of $\zeta(s)$ itself. Indeed, a theorem by
Speiser~\cite{Spe34} states that the Riemann Hypothesis (RH) is
equivalent to $\zeta'(s)$ having no zeros to the left of the
critical line. Thus, understanding of the properties the zeros of
$\zeta'(s)$ can provide important tools and insight into the study
of RH. After Speiser's article this idea was explored by
Berndt~\cite{Ber70} and Spira~\cite{Spi65}, but not much progress
was achieved until the work of Levinson and
Montgomery~\cite{LM74}, who proved a quantitative refinement of
Speiser's theorem. They showed that $\zeta(s)$ and $\zeta'(s)$
have essentially the same number of zeros to the left of the
critical line $\sigma = \Re(s)=\frac12$, and proved that as $T
\rightarrow \infty$, where $T$ is the height on the critical line,
a positive proportion of the zeros of $\zeta'(s)$ lie in the
region
\begin{equation}
\label{eq:lm_res}
\sigma < \frac12 + (1 + \epsilon) \frac{\log \log T}{\log T},
\quad \epsilon > 0.
\end{equation}
Consider the group of unitary matrices $\mathrm{U}(N)$ with probably
distribution given by Haar measure, which is the unique measure
invariant under the left and right action of $\mathrm{U}(N)$ on itself.
Such a probability space is often known as the \textit{Circular
Unitary Ensemble} (CUE). Let $\Lambda(z)$ be the characteristic
polynomial of a matrix in the CUE. In recent years evidence has
been accumulated suggesting that, in the limit as $T \to
\infty$, the local statistical properties of $\zeta(s)$ can be
modeled by the characteristic polynomials of matrices in the
CUE where $N\approx \log(T/2\pi)$. The connection between the
Riemann zeta-function and characteristic polynomials is
extensive; examples include the distribution of the zeros of
$\zeta(s)$, its value distribution and its moments. (For a
series of review articles on the subject see~\cite{MS05} and
references therein.)
Assuming that random matrix theory (RMT) provides an accurate
description of $\zeta(s)$, the horizontal distribution of the
zeros of $\zeta'(s)$ in proximity of the critical line should
be the same as the radial distribution of the roots of
$\Lambda'(z)$ close to the unit circle. This idea was first
developed by Mezzadri~\cite{Mez03}, who determined the
distribution of the zeros of $\Lambda'(z)$ that are very far
from the unit circle and conjectured the leading order term of
the distribution very close to the unit circle. In this paper
we prove his conjecture. We also perform an analogous
calculation for the Riemann zeta-function and conjecturally
find that the result agrees with the RMT model. In addition, we
do numerical computations in both cases and find a surprising
feature in the distribution of zeros of the derivative, namely
that the probability distribution is bimodal.
\section{The zeros of $\zeta'(s)$.}
We have mentioned that the main motivation for studying the zeros
of the derivative of the Riemann zeta-function is its connection
with RH. Indeed, Levinson and Montgomery's result is the basis for
Levinson's method~\cite{Lev74}, which Conrey~\cite{Con89} used to
prove that at least 40\% of the zeros of the zeta-function are on
the line $\sigma = \frac12$.
Levinson's method involves estimating a weighted average of the zeros
of $\zeta'(s)$ to the left of $\frac12+a/\log T$ for some fixed~$a>0$.
Thus, zeros of $\zeta'(s)$ in the region $\frac12 \le \sigma <
\frac12+a/\log T$ are an inherent loss in Levinson's method. It would
be useful to understand the magnitude of this loss. Alternatively, if
we could find a lower bound for the number of zeros of $\zeta'$ in this region we
could improve the estimate for the number of zeros on the critical
line.
\begin{figure}[htp]
\begin{center}
\scalebox{1.00}[1.00]{\includegraphics{zzpzpp}}
\caption{\sf Zeros of $\zeta(s)$ (dot), $\zeta'(s)$ (triangle),
$\zeta''(s)$ (square), and $(\zeta'/\zeta)'(s)$ (star), with
imaginary parts in the range $1015<T<1040$.
\label{fig:zzpzpp}
}
\end{center}
\end{figure}
Figure~\ref{fig:zzpzpp} gives a representative example of the
location of zeros of $\zeta'(s)$ and the relationship to the zeros
of $\zeta(s)$ and various other derivatives of the zeta-function.
It illustrates that zeros of $\zeta'(s)$ close to the
critical-line correspond to closely spaced zeros of~$\zeta(s)$. We
make this statement precise in Section~\ref{sec:riemann}. Also,
a zero of $\zeta'(s)$ seems to be ``missing'' when zeros of
$\zeta$ are particularly far apart or when there are two
successive large gaps. Indeed, there can't be a zero of
$\zeta'(s)$ between every pair of zeros of $\zeta(s)$ because the
density of zeros of $\zeta(s)$ is $\frac{1}{2\pi}\log(T/2\pi)$
while the density of zeros of $\zeta'(s)$ is
$\frac{1}{2\pi}\log(T/4\pi)$. So on average there is a
``missing'' zero of $\zeta'(s)$ in each $T$ interval of
width~$2\pi/\log 2 \approx 9.06$.
Conrey and Ghosh~\cite{CG90} and subsequently Guo~\cite{Guo95}
improved Levinson and Montgomery's result~\eqref{eq:lm_res} and showed
that a positive proportion of the zeros of $\zeta'(s)$ are much closer
to the line $\sigma=\frac12$. Indeed, for any fixed $a>0$, the region
\begin{equation}
\label{eq:cg_res}
\sigma - \frac12 \ge \frac{a}{\log T}
\end{equation}
contains a positive proportion of the zeros.
Soundararajan~\cite{Sou98} made further progress and introduced the
functions
\begin{subequations}
\label{eq:sound_fns}
\begin{align}
\label{eq:sound_fn-}
m^-(a)& := \liminf_{T \to \infty} \frac{1}{N_1(T)}
\sum_{\substack{\beta' \le \frac12 + \frac{a}{\log T}\\
0 < \gamma' \le T}} 1,\\
\label{eq:sound_fn+}
m^+(a)& := \limsup_{T \to \infty} \frac{1}{N_1(T)}
\sum_{\substack{\beta' \le \frac12 + \frac{a}{\log T}\\
0 < \gamma' \le T}} 1,
\end{align}
\end{subequations}
where $N_1(T)$ is the number of zeros $\beta' + i\gamma'$ of
$\zeta'(s)$ with $0 < \gamma' \le T$. Soundararajan proved
that $m^-(a)>0$ for $a > 2.6$, and conjectured that
\begin{equation}
\label{eq:sound_conj}
m(a) = m^-(a)=m^+(a).
\end{equation}
He also conjectured that $m(a)$ is continuous, that $m(a) > 0$
for all $a >0$, and that $m(a) \to 1$ as $a \to \infty$.
Zhang~\cite{Zha01}, Feng~\cite{Fen05}, Garaev and
Y{\i}ld{\i}r{\i}m~\cite{GY06}, and Ki~\cite{Has06} proved
refinements of Soundararajan's results. In particular
Feng~\cite{Fen05} showed that $m^-(a)>0$ for all $a>0$
unconditionally of RH but assuming a conjecture on the
frequency of small gaps between consecutive critical zeros of
$\zeta(s)$.
\section{Zeros of derivatives of polynomials
and statement of results}
Suppose $f(z)$ is a polynomial with all zeros on the unit
circle. (Eventually, $f$ will be a random polynomial obtained as the
characteristic polynomial of a random unitary matrix.) The
Gauss-Lucas theorem assures that all the roots of $f'(z)$ lie on or
inside the unit circle (zeros of $f'$ on the circle occur only if $f$
has multiple zeros). If $f(z)$ has two zeros which are very close
together, then $f'(z)$ will have a zero close by. (This is a
consequence of the continuous dependence of the zeros of $f'$ on those
of~$f$. This dependence is actually piecewise analytic as will be
described below.) The specific location of the nearby zero of $f'(z)$
will depend primarily on how close those two zeros of $f(z)$ are, and
on the general position of the remaining zeros of~$f(z)$. Thus, to
leading order (with respect to the size of the gap), the distribution
of zeros of $f'(z)$ near $|z|=1$ should largely depend on the
distribution of small gaps between zeros of $f(z)$, that is, on the
tail of the nearest-neighbor spacing of zeros of~$f(z)$. We will make
this idea precise and treat in detail the case where $f(z)=\Lambda(z)$
is the characteristic polynomial of a CUE matrix (a matrix chosen from
the unitary group~$\mathrm{U}(N)$, uniformly with respect to Haar measure).
Let $z'$ be a root of $\Lambda'(z)$ and define the random variable
\begin{equation}
\label{eq:dis_un_c}
S := N(1 - \abs{z'}).
\end{equation}
Denote by $Q(s;N)$ the probability density
function (p.d.f.) of $S$. Mezzadri~\cite{Mez03} showed that the
limit
\begin{equation}
\label{eq:limit_pdf}
Q(s) := \lim_{N \to \infty} Q(s;N)
\end{equation}
exists, and proved that
\begin{equation}
\label{eq:m_r1}
Q(s;N) \sim \frac{1}{s^2}, \quad N \to \infty, \quad s \to \infty,
\end{equation}
with $s=o\left(N\right)$. He also conjectured that
\begin{equation}
\label{eq:m_con}
Q(s) \sim \frac{4}{3 \pi}s^{1/2}, \quad s \to 0.
\end{equation}
Formula~\eqref{eq:m_r1} can be interpreted as the RMT counterpart of
the Levinson-Montgomery bound~\eqref{eq:lm_res} for the roots of
$\zeta'(s)$. The RMT model of the Riemann-zeta function is based on
the observation that the local correlations of the non-trivial zeros
of $\zeta(s)$ coincide with those of the eigenvalues of matrices
in the CUE. In order to make this correspondence quantitative, the
densities of the eigenvalues and of the zeros of $\zeta(s)$ must be
made (asymptotically) equal, i.e.,
\begin{equation}
\label{eq:dens}
\frac{N}{2\pi} = \frac{1}{2\pi} \log \frac{T}{2\pi}.
\end{equation}
It follows from~\eqref{eq:m_r1} that the expected value of $S$
does not exist---its p.d.f.\ does not decay sufficiently
rapidly. On the other hand, using~(\ref{eq:m_r1}), the average
of the values of $S$ not exceeding $N$ is
\begin{equation}
\label{eq:rmeqlm}
\sim\int_1^N s\cdot\frac{ds}{s^2} = \log N, \qquad N \to \infty.
\end{equation}
Recalling the relation~(\ref{eq:dis_un_c}) between $S$ and $|z'|$, we
conclude that a positive proportion of the roots of $\Lambda'(z)$ must
lie within a distance from the unit circle bounded from above by
\begin{equation}
\label{eq:rmeqlm2}
(1+\epsilon)\frac{\log N}{N}
\end{equation}
from the unit circle. Because of~\eqref{eq:dens},
formula~\eqref{eq:rmeqlm2} corresponds to Levinson and
Montgomery's result~\eqref{eq:lm_res}.
Now consider the roots $e^{it_1}, \ldots, e^{it_N}$ (with $-\pi
< t_i \leq \pi$) of the characteristic polynomial~$\Lambda(z)$
of a random unitary matrix distributed with Haar measure. It is
convenient for our purposes to define
\begin{equation}
\label{eq:resc_eig}
x_j = \frac{Nt_j}{2\pi}, \qquad j=1,\ldots,N,
\end{equation}
so that, on average, the distance between two consecutive
$x_j$'s is one. The joint probability density function
(j.p.d.f.) of the eigenvalues is given in terms of
$x_1,\ldots,x_N$ as
\begin{equation}
\label{eq:weyl_int}
p_2(x_1,\ldots,x_N) := \frac{1}{N^N N!} \prod_{1 \le j < k \le
N} \abs{e_N(x_k) - e_N(x_j)}^2,
\end{equation}
where we have used the notation $e_N(x):=\exp(2\pi ix/N)$.
Relabeling the indexes $j=1,\ldots,N$, if necessary, we assume
that
\begin{equation}
\label{eq:asc_ord}
x_1 \le \ldots \le x_N < x_1+N.
\end{equation}
We also extend the sequence $\{x_j\}$ to be periodic by setting
$x_{N+1}=x_1$, $x_{N+2}=x_2$, etc. Fix integers $n$ and $j$
with $0\leq n\leq N-2$. Let us denote by $p_2(n;s)$ the
probability density function of $x_{j+n+1}-x_j$. Since the
j.p.d.f.~(\ref{eq:weyl_int}) is invariant under translations,
which means
\begin{equation}
\label{eq:tran_inv}
p_2(x_1 + \alpha,\ldots,x_N + \alpha)= p_2(x_1,\ldots,x_N)
\quad \text{for all
$\alpha \in \mathbb{R}$},
\end{equation}
it follows that $p_2(n;s)$ does not depend on~$j$. If $n=0$,
$p_2(s):=p_2(0;s)$ is known as the \textit{spacing distribution}. It
has an asymptotic expansion in powers of $s$:
\begin{align}
\label{eq:sp_as}
p_2(s) = & \left(\frac{1}{3}-\frac{1}{3 N^2}\right) \pi^2 s^2-
\left(\frac{2}{45}-\frac{1}{9 N^2}
+\frac{1}{15 N^4}\right) \pi^4 s^4\cr
& \quad + \left(\frac{1}{315}-\frac{2}{135 N^2}+
\frac{1}{45 N^4}-\frac{2}{189 N^6}\right)\pi^6 s^6 + O(s^7).
\end{align}
This expansion follows from the pair correlation function
for $U(N)$ (see \cite{MS05}),
\begin{equation}
\frac{\sin^2(\pi y)}{N^2 \sin^2(\pi y/N)},
\end{equation}
and the fact that the pair correlation function and the nearest neighbor
spacing for $U(N)$ agree to order~6.
To describe our main result, suppose that a root of
$\Lambda(z)$ is degenerate (which means $x_{j+1} = x_j$) so
that $z'= \exp\left(2\pi i x_j/N\right)$ will also be a root of
$\Lambda'(z)$. Simple considerations of continuity show that,
if $x_j$ and $x_{j+1}$ are slightly moved apart (say, while
keeping the remaining roots of $f$ fixed), then $z'$ will also
move, but will still be close to the midpoint of the segment
joining $\exp(2\pi i x_j /N)$ to $\exp(2\pi i x_{j+1}/N)$. In
Proposition~\ref{prop:doma-analyt-root} we make this precise,
showing that $z'$ stays close to that midpoint
provided~$x_{j+1}-x_j<1/\pi$. Henceforth we assume that the
rescaled distance
\begin{equation}
\label{eq:def_theta}
\theta: = x_{j+1}-x_j
\end{equation}
is small.
By the translation invariance of the j.p.d.f.\ of the $x_j$'s,
we may assume without loss of generality that
\begin{equation}
\label{eq:xjxjpu}
x_{j+1}=\frac{\theta}{2} \quad \text{and} \quad x_j =
-\frac{\theta}{2}.
\end{equation}
Define
\begin{equation}
\label{eq:delta_def}
\delta := N(1 - z').
\end{equation}
In section~\ref{sec:ch_pol_proof} we shall show that
\begin{equation}
\label{eq:ant_res}
\delta = b_1 \pi^2 \theta^2 + b_2 \pi^4 \theta^4 + O(\theta^6), \quad \text{ as } \theta \to 0,
\end{equation}
where $b_1$ and $b_2$ are explicit functions of the zeros $x_k$ for $k\not= j, j+1$.
By combining the distribution of $\theta$ given in~\eqref{eq:sp_as} with information
we will determine about $b_1$ and $b_2$ in Section~\ref{sec:ch_pol_proof}, we will prove
\begin{theorem}\label{thm:delta}
Let $\Lambda(z)$ be the characteristic polynomial of a random matrix
in $\mathrm{U}(N)$ distributed with respect to Haar measure.
The distribution of $\delta=N(1-|z'|)$ arising from closely spaced zeros of
$\Lambda(z)$ is given by
\begin{equation}
\frac{4}{3 \pi} s^{1/2}
- \frac{82}{45\pi}s^{3/2} + O\left(s^{5/2}\right),
\end{equation}
as $N\to\infty$.
\end{theorem}
Note that Theorem~\ref{thm:delta} refers to the small values of $\delta$ that
arise from closely spaced zeros of the polynomial.
The theorem does \emph{not} account for all small values of~$\delta$.
That distinction is often missed, because
examples such as shown in Figure~\ref{fig:zzpzpp} give the mistaken impression
that zeros of the derivative very close to the unit
circle can only arise from closely spaced zeros of the polynomial.
Farmer and Ki~\cite{FK} give examples of families of polynomials whose
(rescaled) zeros are bounded away from each other,
but for which the density function of $\delta$ vanishes like $C\cdot s$ as $s\to 0$.
They also argue that
any larger density of zeros of $z'$ near the unit circle must arise from
closely spaced zeros of the polynomial. Therefore we have the following
corollary of Theorem~\ref{thm:delta}, which proves Mezzadri's conjecture
about the distribution of~$|z'|$.
\begin{corollary}
\label{thm:unitary}
Let $\Lambda(z)$ be the characteristic polynomial of a random matrix
in $\mathrm{U}(N)$ distributed with respect to Haar measure, and let $Q(s;N)$
be the \textit{p.d.f.} of $S= N(1-|z'|)$ for $z'$ a root of
$\Lambda'(z)$. Then
\begin{equation}
Q(s) = \lim_{N \to \infty} Q(s;N) = \frac{4}{3 \pi} s^{1/2} +O(s).
\end{equation}
\end{corollary}
Note that it remains an unsolved problem to show that $Q(s)$ is
a proper probability distribution. That is, to show
$\int_0^\infty Q(s)ds=1$. This is the random matrix analogue of
Soundararajan's conjecture~$m(a)\to 1$ as $a\to\infty$.
\section{Comparison with data}
\label{sec:data}
We compare our formulas with numerical data.
We generated Haar-random
matrices in $\mathrm{U}(N)$ using the simple algorithm
described in \cite{Mezz}.
Figure~\ref{fig:rmtderiv} shows the empirical distribution of the rescaled
zeros of~$\Lambda'$ for various size matrices.
Figure~\ref{fig:Ip40} shows the empirical cumulative distribution
function $Ip(x)=\int_0^x Q(s,40)ds$ for $\mathrm{U}(40)$ and
a comparison with the tail of the empirical cumulative distribution function
with our results, showing good agreement.
\begin{figure}[htp]
\scalebox{0.7}[0.7]{\includegraphics{rmtderiv15}}
\hskip 0.1in
\scalebox{0.7}[0.7]{\includegraphics{rmtderiv40}}
\vskip 0.1in
\scalebox{0.7}[0.7]{\includegraphics{rmtderiv100}}
\hskip 0.1in
\scalebox{0.7}[0.7]{\includegraphics{rmtderiv15_40_100}}
\caption{\sf
Distribution of the (normalized) distance from the unit circle of zeros of $\Lambda'(z)$
for $\Lambda$ the characteristic polynomial of a random matrix
from $\mathrm{U}(N)$. Top row: $N=15$, $40$. Bottom row: $N=100$
and all three plots together.
} \label{fig:rmtderiv}
\end{figure}
\begin{figure}[htp]
\begin{center}
\scalebox{0.7}[0.7]{\includegraphics{Ip40}}
\hskip 0.1in
\scalebox{0.7}[0.7]{\includegraphics{bestguessplot40}}
\caption{\sf
Experimental cumulative distribution function of $\delta$ for random matrices from $\mathrm{U}(40)$,
and a comparison of the tail of the distribution with the
main term in Theorem~\ref{thm:unitary}.
} \label{fig:Ip40}
\end{center}
\end{figure}
We would like to know the underlying cause of the curious ``second
bump'' in the distribution of zeros of derivatives. This seems to be
a completely general phenomenon. In Figure~\ref{fig:coepoiss40} we
show the analogous distributions for characteristic polynomials of
matrices from $\mathrm{COE}(40)$ and for degree-$40$ polynomials whose
roots are independently and uniformly distributed on the unit circle.
Both cases show the ``second bump'', although not quite at the same
location. In Figure~\ref{fig:rezetaprime} we find a similar shape for
the distribution of zeros of~$\zeta'$.
\begin{figure}[htp]
\scalebox{0.7}[0.7]{\includegraphics{coederiv40}}
\hskip 0.1in
\scalebox{0.7}[0.7]{\includegraphics{poissderiv40}}
\caption{\sf
Distribution of the (normalized) absolute value of zeros of $\Lambda'(z)$
for $\Lambda$ the characteristic polynomial of a random matrix
from $\mathrm{COE}(40)$ (left) and
$\Lambda$ a degree-$40$ polynomial with independent uniformly distributed
(Poisson)
roots on the unit circle (right). For the Poisson case, the plot is truncated
to suppress the large contribution from the small values.
} \label{fig:coepoiss40}
\end{figure}
\section{Proof of theorem~\ref{thm:delta}}
\label{sec:ch_pol_proof}
Suppose $f(z)$ is a degree-$N$ polynomial having all zeros on the unit
circle, for which two zeros $z_1$, $z_2$ are very close together.
Then the derivative $f'(z)$ will have a zero close to the midpoint
$(z_1+z_2)/2$. This follows because, if $z_1=z_2$ (that is, if $f(z)$
has a multiple root at $z_1$), then $z_1$ is also a root of the
derivative $f'(z)$, and the roots of $f'$ are continuous functions of
the roots of~$f$. By a rotation we can assume that
\begin{equation}\label{eq:2}
f(z)=F(z)(z-e^{-i\Theta/2})(z-e^{i \Theta/2}),
\end{equation}
where $F(z)$ does not have any zeros $e^{i t}$ with $-\Theta/2\le t\le
\Theta/2$. The root of $f'$ near~$1$ is the root near~$1$ of
\begin{equation}\label{eqn:fprimeoverf}
\frac{f'}{f}(z)=\frac{F'}{F}(z)
+ \frac{1}{z-e^{-i\Theta/2}} + \frac{1}{z-e^{i \Theta/2}} ,
\end{equation}
and we denote that root by $z'=1-\Delta$.
We are concerned with the case when $f$ is the characteristic
polynomial of a random CUE matrix, and we want to understand the
distribution of the zeros of~$f'$. Since the CUE measure (i.~e.,
normalized Haar measure on $\mathrm{U}(N)$) is invariant under
rotation, the distribution depends only on the absolute value of the
roots of~$f'$. Those roots will accumulate near the unit circle as
$N$ grows, so we must rescale them suitably in order to get a
meaningful result in the limit $N\to\infty$. We let
\begin{equation}\label{eq:1}
\begin{aligned}
\Theta &= 2 \pi\theta/N \\
\Delta &= \delta/N .
\end{aligned}
\end{equation}
Note that this rescaling gives $\langle \theta \rangle =1$.
The rescaling of~$\delta$ is more subtle, and it will
be found in equation~\eqref{eqn:deltathetarelationN} that this is the
correct rescaling. Note that, while $\theta$ is a real number,
$\delta$ is usually complex (but typically with small imaginary part,
as we shall see).
We will determine the leading order behavior
of the roots which are close to the unit circle, so in the above
notation we are interested in
\begin{align}\label{eqn:normalized}
N(1-|z'|) =\mathstrut & N( 1- |1-\Delta| ) \cr
= \mathstrut & \Re(\delta) - \frac{\Im(\delta)^2}{2N} + O\left(\frac{\delta^3}{N^2}\right) .
\end{align}
\subsection{Expansion for the roots}
Exploiting the symmetry of the \textit{j.p.d.f.}~(\ref{eq:weyl_int})
under arbitrary relabellings of the variables $x_j$, as well as its
translation invariance (cf., Section~\ref{thm:unitary}), we will
assume (without loss of generality) that $F(z)$ from
Equation~(\ref{eq:2}) is given in the form (recall that
$e_N(x):=\exp(2\pi ix/N)$)
\begin{equation}
F(z) = \prod_{n=1}^{N-2} \left(z-e_N(x_n)\right),\qquad x_n \in
\left(-\frac N2,-\frac\theta2\right]\cup\left[\frac\theta2,\frac N2\right).
\label{eq:3}
\end{equation}
(We are not excluding the possibility that $F$ has
roots at $e_N(\pm\theta/2)$.) Therefore,
\[
\frac{F'}{F}(z) = \sum_{n=1}^{N-2} \frac{1}{z-e_N(x_n)}.
\]
Now let $z=z'$ be a root of $f'$ (say, that root which is closest to
$z=1$). The Implicit Function Theorem shows that $z'$ is an analytic
function of $\theta$ (at least for $\theta$ sufficiently
small). Indeed, fixing $F$ and regarding $f'$ as a function
$f'(\theta;z)$ of both $\theta$ and~$z$, we have $f'(0;1)=0$; by the
assumption that $F$ has no root at $z=1$ all that remains to observe
is that
\begin{equation}
\label{eq:not_inline}
0 \neq 2F(1) = \frac{\partial}{\partial z}f'(0;1)=f''(0;1).
\end{equation}
if $N$ is sufficiently large, by
Proposition~\ref{prop:doma-analyt-root} $z'$ is defined uniquely and
analytically as a function of $\theta$ in the domain
\begin{equation}
|\theta| < \min\{ \tfrac{1}{\pi}, |x_1|, \ldots, |x_{N-2}|\}.
\end{equation}
We write $z'=1-\delta/N$ and wish to expand
$(1-\delta/N-e_N(x_n))^{-1}$ as a Taylor series in~$\delta$. This
will be justified when $|\delta| < N |1-e_N(x_n)|$. Hence, we have
\begin{equation}\label{eqn:FprimeoverF}
\frac{F'}{F}\(1-\frac{\delta}{N}\) = N \sum_{j=0}^\infty A_j
\delta^j
\end{equation}
for $\delta$ sufficiently small, where
\begin{align}\label{eqn:Ajdef}
A_j &= (-1)^j\frac{1}{j! N^{j+1}} \left(\frac{F'}{F}\right)^{(j)}(1) \cr
&=\frac{1}{N^{j+1}} \sum_{n=1}^{N-2} \frac{1}{(1-e_N(x_n))^{j+1}}.
\end{align}
We will see later that the prefactor of $N$ is the right choice to
make the coefficients $A_j$ approximately bounded. Note that, for
$|\theta| < 1$ (equivalently, for $|\Theta|<2\pi/N$), any $\delta$
such that $|\delta|<|\theta|=N|\Theta|/(2\pi)$ makes the expansion in
\eqref{eqn:FprimeoverF} valid (independently of the exact location of
the zeros of~$F$), since
\begin{equation}
|\delta| < \frac{N \Theta}{2\pi} \leq
N \left| 1-e_N(\theta/2) \right| \leq
N \left| 1-e_N(x_n) \right|,\qquad 1\leq n\leq N-2.
\end{equation}
We also wish to expand the other terms in \eqref{eqn:fprimeoverf}
as a series in $\delta$. We have
\begin{multline}
\frac{1}{z'-e^{-i\Theta/2}} + \frac{1}{z'-e^{i \Theta/2}}
=\frac{2-\frac{2\delta}{N} - 2\cos\left(\frac{\pi\theta}{N}\right)}
{2-\frac{2\delta}{N} +\frac{\delta^2}{N^2}
-2\left(1-\frac{\delta}{N}\right) \cos\frac{\pi\theta}{N}} \cr
=\frac{-2\delta+\pi^2 \theta^2 N^{-1} - \frac{1}{12} \theta^4 \pi^4 N^{-3} + O(\theta^6 N^{-5})}
{\delta^2 + \pi^2\theta^2 - \pi^2 \delta \theta^2 N^{-1} - \frac{1}{12}\pi^4\theta^4 N^{-2} + \frac{1}{12}\pi^4\delta \theta^4 N^{-3} + O(\theta^6 N^{-4})}.
\end{multline}
Combining this with equations \eqref{eqn:fprimeoverf} and
\eqref{eqn:FprimeoverF}, putting all terms over a common denominator,
and using the fact that $f'(z')=0$, we have
\begin{multline}\label{eqn:deltathetarelationN}
0= \sum_{j=0}^\infty A_j \delta^j
\left(\delta^2+ \pi^2 \theta^2 - \pi^2 \delta \theta^2 N^{-1} - \tfrac{1}{12}\pi^4\theta^4 N^{-2} + \tfrac{1}{12}\pi^4\delta \theta^4 N^{-3} + O(\theta^6 N^{-4})\right) \cr
-2\delta + \pi^2 \theta^2 N^{-1} - \tfrac{1}{12} \pi^4 \theta^4 N^{-3} + O(\theta^6 N^{-5}).
\end{multline}
Note that a global factor of $N$ canceled to give the above
equation, suggesting that we have chosen the correct scaling for~$\delta$.
Equation~\eqref{eqn:deltathetarelationN} is simply a more explicit and
manageable form of the equation $f'(\theta;z')=0$ defining $z'$
implicitly as a function of~$\theta$. Noting that $f'(0;1)=0$,
together with the functional equation $f'(\theta;z)=f'(-\theta;z)$, it
follows that $\delta=\delta(\theta)$ has an expansion in powers
of~$\theta^2$, with no constant term, of the form
\begin{equation}\label{eqn:deltaapprox1}
\delta = b_1 \pi^2\theta^2 + b_2 \pi^4 \theta^4 + O(\theta^6).
\end{equation}
From~\eqref{eqn:deltathetarelationN}, we obtain
\begin{multline}\label{eqn:subdelta}
0= \left( A_0 - 2 b_1 + \frac{1}{N} \right)\pi^2 \theta^2 \cr
+\left(A_1 b_1 + A_0 b_1^2 - 2 b_2 - \frac{A_0 b_1}{N} -
\frac{A_0}{12N^2} - \frac{1}{12N^3} \right)\pi^4 \theta^4
+O(\theta^6).
\end{multline}
Setting each term in~\eqref{eqn:subdelta} equal to 0 and solving for
$b_1$ and $b_2$ we have the following:
\begin{proposition}\label{prop:delta} In the notation above, if $0\le\theta<1/\pi$
and $N$ is sufficiently large then
$\delta= b_1 \pi^2\theta^2 + b_2 \pi^4 \theta^4 + O(\theta^6)$ where
\begin{align}
\label{eqn:bAN}
b_1 &= \frac{A_0}{2} + \frac{1}{2N} \\
b_2 &=\frac{1}{8}\left(A_0^3 +2 A_0 A_1\right) + \frac{A_1}{4N} -
\frac{A_0}{6N^2} - \frac{1}{24 N^3} ,
\end{align}
with $A_j$ given in~\eqref{eqn:Ajdef}.
\end{proposition}
Note that in this analysis we have treated $F$ (and hence $A_j$) as
being fixed, in the sense that we assume its zeros do not vary with
$\theta$. In the next section we will show that
when $f(z)$ is the characteristic polynomial of a
matrix drawn from the CUE, this can be justified up to
$O(\theta^7)$.
Using \eqref{eqn:deltaapprox1} and \eqref{eqn:bAN}, we can determine the
distribution of $\delta$ from the distributions of $\theta$
and the~$A_j$. For small $\theta$, this comes from the tail of the
nearest-neighbor spacing.
\subsection{Nearest-neighbor spacing}
\label{nnsp}
Proposition~\ref{prop:delta}
provides a formula for $\delta=N(1-z')$, but what we really want is
the distribution of $\delta^*:=N(1-|z'|)$.
So by \eqref{eqn:normalized} and \eqref{eqn:deltaapprox1}, and writing
$B_j=\Re(b_j)$, we have
\begin{align}\label{eqn:delta*}
\delta^* =\mathstrut& B_1 \pi^2\theta^2 + B_2 \pi^4 \theta^4
+ O\left(\theta^6 + \frac{|\delta|^2}{N} + \frac{|\delta|^3}{N^2} \) \\
=\mathstrut& B_1 \pi^2\theta^2 + B_2 \pi^4 \theta^4
+ O\left(\theta^6 + \frac{\theta^2}{N}\).
\end{align}
The second line is a corollary of Proposition~\ref{prop:doma-analyt-root},
because $\delta\ll\theta$ if $\theta<1/\pi$.
Suppose for the moment that
$B_1$ and $B_2$ were
constants (instead of being random).
We would have
$\delta^*=g(\theta)=B_1\pi^2\theta^2 + B_2\pi^4\theta^4 +O_N(\theta^6)$
where $\theta$ is random
with p.d.f.~\eqref{eq:sp_as}
given by the nearest neighbor spacing of eigenvalues of unitary matrices.
Then the distribution function of
$ \delta^*$ would be given by
\begin{align}\label{eqn:deldist}
\frac{p_2(g^{-1}(s))}{g'(g^{-1}(s))}
=\mathstrut&
\frac{B_1^{-3/2}}{6 \pi} \left(1-\frac{1}{N^2}\right) s^{1/2} \cr
&-
\frac{1}{\pi}
\left(\frac{ B_1^{-5/2}}{45} \left(1-\frac{5}{2N^2} +\frac{3}{2N^4}\right)
+ \frac{5 B_1^{-7/2} B_2}{12}\left(1-\frac{1}{N^2}\right)
\right) s^{ 3/2} \cr
&+ O_N(s^{5/2}).
\end{align}
It turns out that $B_1$ actually is a constant: in
Section~\ref{sec:Aj} we show that $B_1=\frac{1}{4}$. It is fortunate
that $B_1$ is a constant, otherwise it could be difficult to determine
the expected value of quantities like~$B_1^{-3/2}$.
The contribution of $B_2$ takes a bit more work. If $B_2$ was
independent of $\theta$ then we could just average over the possible
contributions of $B_2$ to~\eqref{eqn:deldist}. That is, in
\eqref{eqn:deldist} replace $B_2$ by its expected value. This can be
computed from the expected values of various combinations of $A_0$ and
$A_1$. But $B_2$ is not independent of $\theta$. However, it is
independent of $\theta$ to leading order. Our specific concern is the
distribution of the other roots when $\theta$ is very small. This
approximates the polynomial having a double zero. Since the
next-nearest-neighbor spacing of $\mathrm{U}(N)$ eigenvalues vanishes to
order~$7$, the dependence of $B_2$ on $\theta$ is only to order
$O(\theta^7)$. Thus, for the terms we are computing for the
distribution of $\delta^*$ we can treat $B_2$ as independent
of~$\theta$. The expected value of $B_2$ is calculated in
Section~\ref{sec:Aj}.
\subsection{Expected value of $A_j$}\label{sec:Aj}
We have
\begin{align}\label{eqn:Ajsum}
A_j :=\mathstrut &\frac{1}{N^{j+1}j!} \left(\frac{F'}{F}\right)^{(j)}(1) \cr
=\mathstrut &(-1)^j N^{-j-1} \sum_{n=1}^{N-2} \frac{1}{\left(1-e^{i t_n} \right)^{j+1} },
\end{align}
where $t_1,t_2,\ldots$ are the arguments of the zeros of $F(z)$.
Since
\begin{equation}
\frac{1}{1-e^{i t}} = \frac12 + \frac{i}{2} \cot\(\frac{t}{2}\)
\end{equation}
we see that
\begin{equation}
\Re(A_0)= \frac{N-2}{2 N},
\end{equation}
so
\begin{equation}
B_1=\Re(b_1)=\frac14 ,
\end{equation}
as claimed.
Note also that
\begin{equation}
\langle A_0\rangle= \frac{1}{2}-\frac{1}{N},
\end{equation}
because the imaginary part of the summand is odd.
For $A_j$ with $j\ge 1$, we require a random matrix calculation.
The sum in \eqref{eqn:Ajsum} is dominated by the terms
where $e^{i t_n}$ is close to~$1$, so one possibility is to determine
the level densities of the~$t_n$. We will find the expected value
of $A_j$ by appealing to prior results on averages of ratios
of characteristic polynomials~\cite{CFZ}.
We assume that $f(z)$ is the characteristic polynomial of a matrix
chosen uniformly with respect to Haar measure from the unitary group~$\mathrm{U}(N)$.
We restrict to those matrices which
have two eigenvalues very close to~$1$, and we wish to determine
the joint distribution of the remaining eigenvalues. This is very
similar to the calculations of Due\~nez~\cite{D} and Snaith~\cite{Sna05}
for the orthogonal group~$\mathrm{SO}(N)$.
First we restrict the measure on the entire ensemble to determine the
measure on the remaining eigenvalues. Haar measure on $\mathrm{U}(N)$
is given by
\begin{equation}
d\mu = C \prod_{1\le n<m\le N}
\left|e^{i t_n} - e^{i t_m}\right|^2 dt_1 \cdots dt_N .
\end{equation}
Here and following, $C$ is a normalization constant which may
vary from line to line, chosen so that the measure has total mass~$1$.
Restricting to those matrices which have \emph{one} eigenvalue
at~$1$ is equivalent to rotating (changing variables) to move
an eigenvalue to~$1$. So we can also write the measure as
\begin{equation}\label{eqn:measure1}
d\mu_{1} = C
\prod_{1\le n<m\le N-1}
\left|e^{i t_n} - e^{i t_m}\right|^2
\prod_{1\le n\le N-1}
\left|e^{i t_n} -1\right|^2
dt_1 \cdots dt_{N-1} .
\end{equation}
The set of matrices which have a repeated eigenvalue at~$1$ has measure zero,
so there is no canonical way to restrict the measure. However, we are
interested in the limiting case of two eigenvalues which are very
close together, so we determine the measure by restricting the
measure~\eqref{eqn:measure1} to have $|t_{N-1}|\le t$, and then
let $t\to 0$. The resulting measure is
\begin{equation}\label{eqn:measure2}
d\mu_{2} = C
\prod_{1\le n<m\le N-2}
\left|e^{i t_n} - e^{i t_m}\right|^2
\prod_{1\le n\le N-2}
\left|e^{i t_n} -1\right|^{4}
dt_1 \cdots dt_{N-2} .
\end{equation}
Let $U_2(N-2)$ denote the ensemble of unitary matrices with
joint eigenvalue
measure~$\mu_2$. Then if $g=g(e^{i t_1},\ldots,e^{i t_{n-2}})$ we have
\begin{equation}
\langle g\rangle_{U_2(N-2)} =
C_N
\langle g |\Lambda(1)|^4 \rangle_{\mathrm{U}(N-2)}
\end{equation}
where the right side is a Haar measure average, $\Lambda$
is the characteristic polynomial, and
\begin{equation}
C_N=\langle |\Lambda(1)|^4 \rangle_{\mathrm{U}(N-2)}^{-1} .
\end{equation}
In other words, an expectation involving repeated eigenvalues on $\mathrm{U}(N)$
is equivalent to an expectation on $\mathrm{U}(N-2)$ with an extra factor of the
$4^{th}$ power of the characteristic polynomial.
This is the key observation for
computing the expected values of the $A_j$ because it reduces it
to the evaluation of known quantities.
Specifically,
let
\begin{equation}
G(\alpha_1,\alpha_2,\alpha_3,\alpha_4; \beta_1,\beta_2; \gamma_1,\gamma_2)
=
\frac{
\Lambda(e^{-\alpha_1})
\Lambda(e^{-\alpha_2})
\Lambda(e^{-\alpha_3})
\Lambda(e^{-\alpha_4})
\overline{\Lambda}(e^{-\beta_1})
\overline{\Lambda}(e^{-\beta_2})
}
{\Lambda(e^{-\gamma_1})
\Lambda(e^{-\gamma_2})
} .
\end{equation}
Theorem 4.1 of \cite{CFZ} provides an explicit formula for the expected
value of
$G$ for $\Lambda$ the characteristic polynomial of
Haar distributed matrices on~$\mathrm{U}(N-2)$.
The formula is complicated so we do not reproduce it here.
This is sufficient to determine
the expected values of all the quantities in~\eqref{eqn:bAN}.
The calculation requires the assistance of a computer algebra
package. We now present the answers, which we determined with
the help of Mathematica.
The normalization constant for the measure $\mu_2$ is (the reciprocal of)
\begin{align}
\langle G(0,0,0,0;0,0;0,0)\rangle_{\mathrm{U}(N-2)}
&= \langle |\Lambda(1)|^4 \rangle_{\mathrm{U}(N-2)}\cr
&=
\frac{N^4}{12}-\frac{N^2}{12} \cr
&= C_N^{-1},
\end{align}
say. As $\theta\to 0$ we have
\begin{align}
\langle A_0^3\rangle =\mathstrut&
\left\langle \left(\frac{\Lambda'}{\Lambda}\right)^3 (1)\right\rangle_{U_2(N-2)} \cr
=\mathstrut& - C_N
\frac{\partial^3}{\partial_{\alpha_1,\alpha_2,\gamma_1}}
\bigg|_{(\alpha_1,\alpha_2,\gamma_1) =(0,0,0)}
\left\langle
G(\alpha_1,\alpha_2,0,0; 0,0; \gamma_1,0) \right\rangle_{\mathrm{U}(N-2)}
\cr
=\mathstrut&
\frac{1}{10} N^3-\frac{7}{10} N^2+\frac{8}{5 }N-\frac{6}{5}\cr
&\cr
\langle A_1 \rangle =\mathstrut&
\left\langle \left(\frac{\Lambda'}{\Lambda}\right)' (1)\right\rangle_{U_2(N-2)}\cr
=\mathstrut& C_N
\left(1+\frac{d}{d\alpha_1}\right)\bigg|_{\alpha_1=0}
\
\frac{\partial}{\partial \alpha_1} \bigg|_{\gamma_1=\alpha_1}
\left\langle
G(\alpha_1,0,0,0; 0,0; \gamma_1,0)
\right\rangle_{\mathrm{U}(N-2)}\cr
=\mathstrut&
\frac{1}{15}N^2-\frac{1}{2} N+\frac{11}{15} \cr
&\cr
\langle A_0 A_1 \rangle =\mathstrut&
\left\langle \left(\frac{\Lambda'}{\Lambda}\right)(1) \left(\frac{\Lambda'}{\Lambda}\right)'(1) \right\rangle_{U_2(N-2)} \cr
=\mathstrut& -C_N
\left(1+\frac{d}{d\alpha_1}\right)\bigg|_{\alpha_1=0}
\
\frac{\partial}{\partial \alpha_1} \bigg|_{\gamma_1=\alpha_1}
\
\frac{\partial}{\partial \alpha_2} \bigg|_{(\alpha_2,\gamma_2)=(0,0)} \cr
&\phantom{XXXXX}
\left\langle
G(\alpha_1,\alpha_2,0,0; 0,0; \gamma_1,\gamma_2)
\right\rangle_{\mathrm{U}(N-2)}\cr
=\mathstrut&
\frac{1}{30} N^3-\frac{3}{10}N^2+\frac{13}{15}N-\frac{4}{5}.
\end{align}
Thus,
\begin{equation}\label{eqn:b2}
\langle B_2 \rangle=\Re \langle b_2 \rangle =\frac{1}{48}-\frac{7}{48}N^{-1}+ O(N^{-2}).
\end{equation}
Inserting this into
\eqref{eqn:deldist} gives the expansion claimed in Theorem~\ref{thm:delta}.
\section{The Riemann zeta-function}\label{sec:riemann}
We do analogous calculations for derivatives of the Riemann zeta-function
and compare our results with data.
We start with
\begin{equation}
\frac{\zeta'(s)}{\zeta(s)} = b-\frac{1}{s-1}-\frac{1}{2}\frac{\Gamma'(\frac{1}{2}s+1)}{\Gamma(\frac{1}{2}s+1)}+\sum_\rho\left(\frac{1}{s-\rho}+\frac{1}{\rho}\right)
\end{equation}
where $b= \log2\pi-1-\frac{1}{2}\gamma$.
As in the polynomial case, we assume there are two very closely spaced zeros
of the $\zeta$-function and look for the nearby zero of $\zeta'$.
Suppose the closely spaced zeros are
\begin{equation}
\rho_\pm = \frac12 + i (t\pm \tfrac12 \Theta)
\end{equation}
with
\begin{equation}
s'= \frac12 + X + i t
\end{equation}
a zero of~$\zeta'$.
Using
\begin{equation}
\frac{\Gamma'}{\Gamma}(s) = \log(s) + O(1/s)
\end{equation}
we have
\begin{equation}
0= b^* -\frac12 \log t + \frac{1}{s'-\rho_-}+\frac{1}{s'-\rho_+}
+ \sum_{\rho\not= \rho_\pm} \left(\frac{1}{s'-\rho}+\frac{1}{\rho}\right),
\end{equation}
where
\begin{equation}
b^* = b + \frac12\log(2) - i\frac{\pi}{4} + O(1/t).
\end{equation}
Note that
\begin{equation}
\frac{1}{s'-\rho_-}+\frac{1}{s'-\rho_+}= \frac{8 X}{4 X^2 +\Theta^2}
\end{equation}
and
\begin{equation}
\frac{1}{s'-\rho}+\frac{1}{\rho}
= \frac{X-i(t-\gamma)}{X^2+(t-\gamma)^2} + \frac{\frac12 -i \gamma}{\frac14 + \gamma^2} ,
\end{equation}
which has real part
\begin{equation}
\frac{X}{X^2+(t-\gamma)^2} + \frac{\frac12 }{\frac14 + \gamma^2}.
\end{equation}
As in the polynomial case, we rescale:
\begin{align}
X=& x /\log t \cr
\Theta=&2\pi \theta /\log t .
\end{align}
Note that this is analogous to the rescaling in the
unitary case because $N\approx\log(t/2\pi)$.
We have
\begin{equation}\label{eqn:scaledx}
0= b^{**} +I -\frac12 \log t +\frac{2 x\log t}{x^2+\pi^2\theta^2}
+ x \log t \sum_{\gamma_*} \frac{1}{x^2+ 4\pi^2 \gamma_*^2 },
\end{equation}
where $b^{**}$ is a real constant, $I$ is purely imaginary, and
the sum is over the rescaled zeros $\gamma_* = \log t (t-\gamma)/2\pi$.
We follow the same procedure as in the $\mathrm{U}(N)$ case.
First multiply through~\eqref{eqn:scaledx} by $(x^2+\pi^2\theta^2)/\log t$ and
expand the final summand as a series in~$x$,
giving
\begin{equation}\label{eqn:xthetaalpha}
0=2x +(x^2+\pi^2 \theta^2)\left(-\frac12 + x\alpha_1-x^3\alpha_2+O(x^5)\right)+\text{ smaller terms},
\end{equation}
where
\begin{equation}
\alpha_j=\sum_{\gamma_*} \frac{1}{(4\pi^2 \gamma_*^2)^j }.
\end{equation}
Note that \eqref{eqn:xthetaalpha} has the same form as~\eqref{eqn:deltathetarelationN}.
Now write
$x=\beta_1 \pi^2 \theta^2 + \beta_2\pi^4\theta^4+O(\theta^6)$ and gather terms to
get
\begin{equation}
\left(2\beta_1-\frac12\right)\pi^2\theta^2
+
\left(2\beta_2-\frac12\beta_1^2+\beta_1\alpha\right) \pi^2\theta^4 + \text{ smaller terms} .
\end{equation}
Thus,
\begin{align}
\beta_1\sim\mathstrut& \frac14 \\
\beta_2 \sim\mathstrut& \frac{1}{64} -\frac18 \alpha_1,
\end{align}
which exactly corresponds to the $\mathrm{U}(N)$ case~in
Proposition~\ref{prop:delta}.
The sum over zeros $\alpha_1$ is similar to the expression for $A_1$ in the
unitary case. We can determine the expected value of the sum
if we assume $\mathrm{CUE}$ statistics for the zeros of the Riemann zeta-function,
restrict to having two closely spaced zeros, and find the one-level
density of the remaining zeros. That calculation is in Section~\ref{sec:1level}.
In the notation of Lemma~\ref{lem:1level} we have
\begin{equation}
\langle \alpha_1 \rangle = \frac{1}{4\pi^2}
\int_{-\infty}^\infty \frac{1}{t^2} W_1^{(2,0)}(t)\, dt = \frac{1}{15}.
\end{equation}
Note that this is the same as the expected value of $A_1$.
Thus, assuming that the spacing of zeros of the Riemann zeta function has the
same distribution as the spacing of eigenvalues of random unitary matrices,
we find that the leading order behavior of zeros of $\zeta'$ near the
$\tfrac12$-line is the same as that of zeros of $\Lambda'$ near the unit circle.
In Figure~\ref{fig:rezetaprime} we show the empirical distribution of
the zeros of $\zeta'$ for $10^6 < t < 10^6 + 60000$. The general shape
of the distribution shows a striking similarity with the zeros of
the derivative of characteristic polynomials of unitary matrices.
\begin{figure}[htp]
\begin{center}
\scalebox{1.0}[1.0]{\includegraphics{zetaderiv12}}
\caption{\sf
Normalized distribution of the real part of the zeros of $\zeta'(s)$.
Data is for the approximately 100000 zeros
with imaginary part in~$[10^6,10^6+60000]$.
} \label{fig:rezetaprime}
\end{center}
\end{figure}
\section{Calculation of the 1-level density}\label{sec:1level}
We prove the following
\begin{lemma}\label{lem:1level}
Fix $a,b>-1/2$. If $t_1,t_2,\ldots,t_M$ are independently
distributed with respect to the probability measure
\begin{equation}
\label{eqn:measurek}
d\mu = d\mu^{(a,b)} =
\frac1{C_{M,a,b}}\prod_{1\le n<m\le M}
\left|e^{i t_n} - e^{i t_m}\right|^2
\prod_{1\le n\le M}
\left|e^{i t_n} -1\right|^{2a}\left|e^{it_n}+1\right|^{2b}
dt_1 \cdots dt_{M} ,
\end{equation}
then the large-$M$ limiting (rescaled) 1-level density of the
normalized values $\tilde t_j = t_j M/2\pi $ is given by
\begin{equation}
\label{eq:W-1}
W_1^{(a,b)}(t)
=t \frac{\pi^2}{2} \left(J_{a-\frac12}(\pi t)^2+J_{a+\frac12}(\pi t)^2\right)
-a \pi J_{a-\frac12}(\pi t) J_{a+\frac12}(\pi t).
\end{equation}
\end{lemma}
The measure $d\mu^{(a,b)}$ is Haar measure on $U(M+a+b)$ restricted to those
matrices which have $a$ eigenvalues equal to~$1$ and $b$ eigenvalues equal to~$-1$. Note that
$W_1^{(a,b)}$ is independent of~$b$.
\begin{proof}
For fixed $a,b>-\frac12$ define $\omega(z)=|z-1|^{2a}|z+1|^{2b}$.
Let $\{\phi_n\}_{n=0}^\infty$ be the sequence in $\mathbb C[x]$ uniquely
determined by the following requirements:
\begin{enumerate}
\item For all $n\geq0$, $\phi_n$ is of degree~$n$ and has positive
leading coefficient.
\item For all $m,n\geq0$,
\begin{equation}
\label{eq:IPC}
\langle\phi_m,\phi_n\rangle := \frac1{2\pi}
\int_{-\pi}^{\pi}\phi_m(e^{it})\overline{\phi_n(e^{it})}
w(e^{it})dt = \delta_{mn},
\end{equation}
where $\delta_{mn}$ is the Kronecker delta.
\end{enumerate}
Then $\{\phi_n\}$ is the sequence of normalized orthogonal polynomials on the
unit circle with respect to the measure $d\nu(z)=(2\pi)^{-1}\omega(z)d\ell(z)$
(where $d\ell$ is the arc-length element.)
Let
\begin{equation}
\label{eq:K_M}
K_M(z,w) := \sum_{n=0}^{M-1}\overline{\phi_n(z)}\phi_n(w)
\end{equation}
be the projection kernel onto polynomials of degree less than $M$
with respect to the inner product~\eqref{eq:IPC}. By the
Gaudin-Mehta method, the probability measure~\eqref{eqn:measurek},
when regarded as a measure on the unit circle, can be rewritten~as
\begin{equation}
\label{eq:dmu_k-GM}
d\mu_k =\frac1{M!}
\det_{1\leq j,k\leq M}(K_M(z_j,z_k))\prod_{j=1}^Md\nu(z_j)
\end{equation}
(note that the normalization constant $C_{M,k}$ is no longer needed).
Then the $1$-level measure is $W_1^{(M)}(z)d\nu(z)$, where
\begin{equation}
\label{eq:W1_M}
W_1^{(M)}(z) = K_M(z,z).
\end{equation}
(The normalization above is such that the total mass of the $1$-level
measure is equal to $M$.)
Let the ``dual'' $\phi^*$ of a polynomial
$\phi(z)=c_nz^n+c_{n-1}z^{n-1}+\dots+c_1z+c_0$ of degree~$n$ be the
polynomial
\begin{equation}
\label{eq:dual}
\phi^*(z) = z^n\bar\phi(z^{-1})
= \bar c_n + \bar c_{n-1}+\dots+\bar c_1z^{n-1}+\bar c_0z^n.
\end{equation}
Then we have the following formula of Szeg\H{o} for the projection
kernel
(\cite{Sze39}, Theorem~11.4.2):
\begin{equation}
\label{eq:Szego-kernel}
K_M(z,w) =
\frac{\overline{\phi^*_M(z)}\phi^*_M(w) -
\overline{\phi_M(z)}\phi_M(w)}
{1-\bar z w}.
\end{equation}
(This formula is analogous to the classical one of
Christoffel and Darboux for the projection kernel of orthogonal
polynomials on the line.)
In view of~\eqref{eq:W1_M} and~\eqref{eq:Szego-kernel}, in order to
find the rescaled limit of the $1$-level measure as $M\to\infty$ it
will suffice to derive the asymptotic behavior of the orthogonal
polynomials $\phi_n$ as $n\to\infty$ near the point $z=+1$.
Theorem~\ref{thm:phi-vs-Jacobi} and formula~\eqref{eq:Hilb} below are
the key ingredients, but first we need to introduce some notation.
Denote by $P_n^{(a,b)}$ the classical Jacobi polynomials: they are
orthogonal in the interval $[-1,1]$ with respect to the measure
\begin{equation}
\label{eq:w}
w^{(a,b)}(x):=(1-x)^a(1+x)^b
\end{equation}
and are normalized as follows (\cite{Sze39}, Equation~4.3.3):
\begin{equation}
\label{eq:h_n}
h_n^{(a,b)} := \int_{-1}^1|P_n^{(a,b)}(x)|^2w^{(a,b)}(x)\,dx =
\frac{2^{a+b+1}}{2n+a+b+1}\,
\frac{\Gamma(n+a+1)\Gamma(n+b+1)}{\Gamma(n+1)\Gamma(n+a+b+1)}.
\end{equation}
Let also
\begin{align}
\label{eq:hpm}
h^+_n &= 2^{a+b}h_n^{(a+1/2,b+1/2)}, &
h^-_n &= 2^{a+b}h_n^{(a-1/2,b-1/2)}.
\end{align}
\begin{theorem}\cite{Sze39}
\label{thm:phi-vs-Jacobi}
\begin{align}
\notag
z^{-n}\phi_{2n}(z) &=
AP_n^{(a-1/2,b-1/2)}\left(\frac{z+z^{-1}}2\right)
+B(z-z^{-1})P_{n-1}^{(a+1/2,b+1/2)}\left(\frac{z+z^{-1}}2\right) \\
\label{eq:phi_e/o}
z^{-n+1}\phi_{2n-1}(z) &=
CP_n^{(a-1/2,b-1/2)}\left(\frac{z+z^{-1}}2\right)
+D(z-z^{-1})P_{n-1}^{(a+1/2,b+1/2)}\left(\frac{z+z^{-1}}2\right).
\end{align}
Letting $c_n=(a+b)/(n+a+b)$, we have
\begin{align*}
A &= \sqrt{\frac\pi2\,\frac{1+c_n}{h_n^-}} &
B &= \frac12\sqrt{\frac\pi2\,\frac{1-c_n}{h_{n-1}^+}} \\
C &= \sqrt{\frac\pi2\,\frac{1-c_n}{h_n^-}} &
D &= \frac12\sqrt{\frac\pi2\,\frac{1+c_n}{h_{n-1}^+}}.
\end{align*}
\end{theorem}
Equation~\eqref{eq:phi_e/o} appears as~(11.5.4) in Szeg\H{o}'s book
(except for an obvious typographical mistake therein.) The constants
$A,B,C,D$ can be easily found using Szeg\H{o}'s equation~(11.5.2)
together with the fact that $\phi_{2n-1}$ is a polynomial, which
forces the coefficient of $z^{-n}$ on the right-hand side of
equation~\eqref{eq:phi_e/o} to vanish. We omit the details.
A meaningful rescaling of the $1$-level measure is achieved via the
change of variables
\begin{align}
\label{eq:xi}
t&=\frac{2\pi\xi}M.
\end{align}
Indeed, one finds that the limiting $1$-level measure (as
$M\to\infty$) is $W_1(\xi)d\xi$, where
\begin{align}
\label{eq:K_inf}
W_1(\xi)&=K_\infty(\xi,\xi), \\
K_\infty(\xi,\eta) &= \lim_{M\to\infty}\frac1M
K_M(e^{2\pi\xi/M},e^{2\pi\eta/M})\sqrt{\omega(e^{2\pi\xi/M})\omega(e^{2\pi\eta/M})}.
\end{align}
(Actually, all limiting local correlations can be expressed
in terms of the limiting kernel $K_\infty$, not just the $1$-level
density.)
It remains to compute $K_\infty(\xi,\eta)$. It suffices to use
equation~\eqref{eq:phi_e/o} in formula~\eqref{eq:Szego-kernel} and
an asymptotic formula by Szeg\H{o}'s (see~\cite{Sze39},
equation~(8.21.17)):
\begin{equation}
\label{eq:Hilb}
\left(\sin\frac t2\right)^a\left(\cos\frac t2\right)^bP_n^{(a,b)}(\cos t)
= N^{-a}\frac{\Gamma(n+a+1)}{n!}\sqrt{\frac t{\sin t}}J_a(Nt)
+ t^{a+2}O(n^a),
\end{equation}
valid for fixed $a>-1$, $b\in\mathbb R$, in the range $0<t\leq c/n$ for any
fixed constant $c>0$, where $N=n+(a+b+1)/2$ and $J_a$ is a Bessel
function of the first kind.
A tedious but straightforward computation finally gives:
\begin{align}
\label{eq:K_infty}
K_\infty(\xi,\eta) &=
\frac\pi2 e^{i\pi(\eta-\xi)}\frac{\sqrt{\xi\eta}}{\xi-\eta}
\left(J_{a+1/2}(\pi\xi)J_{a-1/2}(\pi\eta)
- J_{a-1/2}(\pi\xi)J_{a+1/2}(\pi\eta)\right)\\
K_\infty(\xi,\xi) &=
\frac\pi2\left\{ \pi\xi[J_{a+1/2}(\pi\xi)^2+J_{a-1/2}(\pi\xi)^2]
- 2a J_{a+1/2}(\pi\xi)J_{a-1/2}(\pi\xi) \right\}.
\end{align}
\begin{remark}
The kernel $K_{\infty}(\xi,\eta)$ can be used to compute $n$-level
correlations and spacing statistics through (matrix or operator)
determinants (see~\cite{TW98} for an explanation and proof of these
applications). Incidentally, for the purposes of evaluating such
determinants, the factor $e^{i\pi(\eta-\xi)}$ may be suppressed
in~\eqref{eq:K_infty} (this is tantamount to conjugating the
corresponding integral operator by a unitary transformation).
\end{remark}
\end{proof}
\section{Appendix: The Domain of Analyticity of the Root of the Derivative}
\label{sec:dom-analyt-root}
\begin{center}
{by Eduardo Due\~nez}
\end{center}
Define as before $e(x):=e^{2\pi ix}$ for $x$ real. For any fixed
$N\geq3$ define $e_N(x):=e(x/N)=e^{2\pi ix/N}$. Consider $N$ unit
complex numbers
\begin{equation}
e_N(\theta_0),e_N(-\theta_0),
e_N(\theta_1),e_N(\theta_2),\dots,e_N(\theta_{N-2}),
\label{eq:roots-f}
\end{equation}
where
\begin{equation}\label{eq:pyram}
0\leq\theta_0\leq\theta_j\leq N-\theta_0,\qquad j=1,2,\dots,N-2.
\end{equation}
(I.~e., the open arc centered at $1$ of the unit circle $|z|=1$ having
endpoints $e_N(\pm\theta_0)$ contains none of the remaining $N-2$
numbers). The inequalities~(\ref{eq:pyram}) define a ``pyramid''
$\mathcal{P}_N$ contained in the cube $[0,N]^{N-1}$. For fixed
$0<T\leq N/2$, denote by $\mathcal{P}_N^{(T)}$ the closed truncation
of $\mathcal{P}_N$ at height~$T$. Then $\mathcal{P}_N^{(T)}$ is defined by
the inequalities
\begin{align}
\theta_0&\leq\theta_j\leq N-\theta_0,\qquad j=1,2,\dots,N-2;\notag\\
\label{eq:pyram-T}0&\leq\theta_0\leq T.
\end{align}
For notational convenience we will set
$\Theta=(\theta_1,\theta_2,\dots,\theta_{N-2})$ and
$\Theta_0=(\theta_0;\Theta)$. Finally, let
\begin{equation}\label{eq:9}
f(z)=f(\Theta_0;z)=
(z-e_N(\theta_0))(z-e_N(-\theta_0))\prod_{j=1}^{N-2}(z-e_N(\theta_j))
\end{equation}
be the monic polynomial of degree~$N$ with roots~(\ref{eq:roots-f}).
In this section we prove the following result.
\begin{proposition}\label{prop:doma-analyt-root}
With the above notation, for every
$T<\frac1\pi$ there exists $N(T)$ such that for all $N\geq
N(T)$ and for each $\Theta_0$ in $\mathcal{P}_N^{(T)}$, the
derivative $f'(z)=\frac{\partial}{\partial z}f(\Theta_0;z)$ of $f$
has a unique root $z'=\varsigma(\Theta_0)$ in the open disk with
diameter $[e_N(-\theta_0),e_N(\theta_0)]$. Moreover, $\varsigma$ is an
analytic function of $\Theta_0$ in the interior of
$\mathcal{P}_N^{(T)}$.
\end{proposition}
We remark that, by the Gauss-Lucas theorem, the root $z'$ alluded to
in Proposition~\ref{prop:doma-analyt-root} must lie on or to the
left of the vertical diameter~$[e_N(-\theta_0),e_N(\theta_0)]$.
\begin{proof}
Fix $T<1/\pi$ and
consider $N$ as a parameter ($N\geq3$) for the time being. Let
\begin{equation}
\label{eq:4}
F(z)=F(\Theta;z):=\prod_{j=1}^{N-2}(z-e_N(\theta_j)),
\end{equation}
so
\begin{equation}
f(z)=(z-e_N(\theta_0))(z-e_N(-\theta_0))F(z)\label{eq:6}
\end{equation}
and
\begin{equation}
\label{eq:7}
\frac{f'}{f}(z) = g(z) + L(\Theta;z),
\end{equation}
where
\begin{align}
g(z)&:=\frac1{z-e_N(\theta_0)}+\frac1{z-e_N(-\theta_0)},\label{eq:12}\\
L(\Theta;z) &:= \frac{F'}{F}(z)
= \sum_{j=1}^{N-2}\frac1{z-e_N(\theta_j)}. \label{eq:8}
\end{align}
Define $c_N(x):=\cos(2\pi x/N)$ and $s_N(x):=\sin(2\pi x/N)$, so
$e_N(x)=c_N(x)+is_N(x)$. Parametrize the boundary of the disk with
diameter $e_N(\pm\theta_0)$ (i.~e., the disk
$|z-c_N(\theta_0)|\leq s_N(\theta_0)$) as:
\begin{equation}
z(\phi)=c_N(\theta_0)+ie^{i\phi}s_N(\theta_0).
\label{eq:10}
\end{equation}
Since roots of $f$ occur at $e_N(\pm\theta_0)$ whenever $f$ has
multiple roots there, it is best to work instead with a
slightly deformed contour $\mathcal{C}$ obtained as the boundary of
the ``twice bitten'' disk
\begin{align*}
|z-c_N(\theta_0)|&\leq s_N(\theta_0)\\
|z-e_N(\theta_0)|&\geq \epsilon \\
|z-e_N(-\theta_0)|&\geq \epsilon.
\end{align*}
missing tiny $\epsilon$-neighborhoods of the points
$e_N(\pm\theta_0)$. We still assume that $\mathcal{C}$ is parametrized
by~(\ref{eq:10}), except for $\phi$ in a small $\delta$-neighborhood
of any multiple of~$\pi$. (We will not write down the exact
parametrization of $\mathcal{C}$ for such values of $\phi$ since
its precise form will not be needed.)
Write $\mathcal{C}$ as a union of four pieces: $\mathcal{C}
=\mathcal{C}_\text{d}\cup\mathcal{C}_\text{r}\cup\mathcal{C}_\text{u}
\cup\mathcal{C}_\ell$ (down, right, up, left), where
\begin{align*}
\mathcal{C}_\text{d}&=\{z(\phi):\pi-\delta\le\phi\le\pi+\delta\},\\
\mathcal{C}_\text{r}&=\{z(\phi):-\pi+\delta\le\phi\le-\delta\},\\
\mathcal{C}_\text{u}&=\{z(\phi):-\delta\le\phi\le\delta\},\\
\mathcal{C}_\ell&=\{z(\phi):\delta\le\phi\le\pi-\delta\}.
\end{align*}
The assumed inequalities~(\ref{eq:pyram-T}) ensure that
there are no zeros of $f$ anywhere on $\mathcal{C}$ (and,
\textit{a fortiori,} no multiple zeros); hence,
every
zero of $f'$ inside $\mathcal{C}$ is a zero of $f'/f$ with the
same multiplicity. It suffices to ensure that
that (for all sufficiently small $\epsilon$)
the contour $\mathcal{C}$ encloses exactly one zero of~$f'/f$. Note
that, since $f(\Theta_0;z)$ has no zeros on or inside $\mathcal{C}$,
$f'/f$ has no poles there either. By the Argument Principle, the
claim in Proposition~\ref{prop:doma-analyt-root} regarding the
uniqueness
of the zero $z'$ of $f'$ is equivalent to showing that the image
$\mathcal{D}$ of
$\mathcal{C}$ under
$f'/f$ has index $1$ about the origin for large $N$ and all
sufficiently small~$\epsilon$.
We let
$w(\phi):=\frac{f'}{f}(z(\phi))$ and
denote the images of the four pieces $\mathcal{C}_\text{d},
\mathcal{C}_\text{r},\mathcal{C}_\text{u},\mathcal{C}_\ell$ of
$\mathcal{C}$ under $f'/f$ by
$\mathcal{D}_\text{d},\mathcal{D}_\text{r},\mathcal{D}_\text{u},
\mathcal{D}_\ell$.
The idea of the proof is very simple. As we shall show, the dominant
part of the logarithmic derivative $f'/f(z)$ (for $z$ in
$\mathcal{C}$) is $g(z)$.
The image of $\mathcal{C}$ under $g(z)$ is easy to describe
explicitly; indeed, it is a curve $\mathcal{E}$ having index $1$
about the
origin. The curve $\mathcal{D}$ can be regarded as a perturbation of
$\mathcal{E}$. We show that, if $H$ and $\epsilon$ are sufficiently
small, then the remaining terms $1/(z-e_N(\theta_j))$,
$j=1,2,\dots,N-2$, are small enough to keep the index of
$\mathcal{D}$ equal to that of~$\mathcal{E}$.
\begin{figure}[h]
\centering
\scalebox{.4}[.4]{\includegraphics{contourE}}
\caption{The curve $\mathcal{E}$.}
\label{fig:contourE}
\end{figure}
We denote by
$\mathcal{E}_{\text{d}},\mathcal{E}_{\text{r}},\mathcal{E}_{\text{u}},
\mathcal{E}_\ell$ the images under $g$ of the respective parts of
$\mathcal{C}$ (figure~\ref{fig:contourE}).
\begin{itemize}
\item $\mathcal{E}_{\text{r}}$ is parametrized as
\begin{equation}
g(z(\phi))=\frac1{2s_N(\theta_0)\sin\phi},\qquad
-\pi+\delta\leq\phi\leq-\delta.\label{eq:11}
\end{equation}
Thus, $\mathcal{E}_{\text{r}}$ is a twice-traversed straight
segment starting at the faraway point
$H=g(z(-\pi+\delta))=1/(2s(\theta_0)\sin\delta)$, moving
leftward to the point $A=g(z(0))=1/(2s(\theta_0))$ and retracing
itself back to~$B=g(z(-\delta))$ (here $B=H$).\vspace{1ex}
\item $\mathcal{E}_{\text{u}}$ is a large arc $\overarc{BCD}$ on the
upper half-plane. ($\mathcal{E}_{\text{u}}$ is essentially a
semicircle, for $\epsilon$ small. This is easily seen from the
fact that the dominant term in $f'/f(z)$ for $z$ very close to
$e_N(\theta_0)$ is $1/(z-e_N(\theta_0))$, namely an inversion with
center
$e_N(\theta_0)$, taking the tiny (almost) semicircle
$\mathcal{C}_{\text{u}}$ to a huge (almost)
semicircle~$\mathcal{E}_{\text{u}}$.) $\mathcal{E}_{\text{u}}$
escapes any bounded region of the plane as $\epsilon$ approaches
zero.
\item $\mathcal{E}_{\ell}$ and $\mathcal{E}_{\text{d}}$ are obtained
from $\mathcal{E}_{\text{r}}$ and $\mathcal{E}_{\text{u}}$ through
central symmetry with respect to the origin.
\end{itemize}
It is clear that $\mathcal{E}$ has index $1$ about the origin for
all sufficiently small $\epsilon$.
For notational convenience, denote by $\tilde A=w(-\pi/2),\tilde
B=w(-\delta),\dots,\tilde H=w(-\pi+\delta)$ the points on
$\mathcal{D}$ analogous to $A,B,\dots,H$ on~$\mathcal{E}$.
We claim that if $N$ is large enough and
$\Theta_0\in\mathcal{P}_N^{(T)}$, then
the index of $\mathcal{D}$ about the origin is also~$1$ for all
sufficiently small~$\epsilon$ and all $N\geq N(T)$. Note that
$\epsilon$ is allowed to depend on $\Theta_0$, whereas the lower
bound $N(T)$ for $N$ depends only on~$T$.
For notational simplicity, we set $\theta_{N-1}:=-\theta_0$. Write
\begin{equation*}
\frac{f'}{f}(z) = \frac{m}{z-e_N(\theta_0)}
+
\underbrace{\sum_{j=m}^{N-1}\frac1{z-e_N(\theta_j)}}_{\displaystyle
R(\Theta_0;z)},
\end{equation*}
where $m\geq1$ is the multiplicity of the zero $e_N(\theta_0)$ of
$f(z)$ and $\theta_j$, $j=m,\dots,N-1$, are the remaining zeros
of~$f(z)$.
Let $\epsilon_0>0$ be the minimum of the distances from
$e_N(\theta_0)$
to the other $e_N(\theta_j)$, $m\leq j\leq N-1$.
Now let
$\epsilon=\epsilon_0/N$. If $|z-e_N(\theta_0)|=\epsilon$, then
$|z-e_N(\theta_j)|\geq\epsilon_0$ ($m\leq j\leq N-1$), so
$|R(\Theta_0;z)|\leq(N-m)/\epsilon_0$. Moreover,
$m/(z-e_N(\theta_0))$ maps $\mathcal{C}_{\text{u}}$ into (part of) the
upper half-circle $\mathcal{S}$: $|w|=m/\epsilon$, $\Re
w>0$. Therefore, $\mathcal{D}_{\text{u}}$ will be supported on the
$\frac{N-1}{\epsilon_0}$-neighborhood $\mathcal{T}$
of~$\mathcal{S}$. Since $\mathcal{T}$ intersects the imaginary axis
$\Re w=0$ on the interval
$i\left[\frac m\epsilon-\frac{N-m}{\epsilon_0},
\frac m\epsilon+\frac{N-m}{\epsilon_0}\right]
=i\left[\frac{m+N(m-1)}{\epsilon_0},\frac{N+m(N-1)}{\epsilon_0}\right]$,
our claim
that $\mathcal{D}_{\text{u}}$ only intersects the positive imaginary
axis is proved. The claim that $\mathcal{D}_\ell$ only intersects
the negative imaginary axis for suitably small $\epsilon$ is proved
along identical lines.
In order to prove the claim that $\mathcal{D}$ has index~$1$ about
the origin, it suffices now to show that $\mathcal{D}_{\text{r}}$ is
contained in the right half-plane $\Re(w)>0$ and
$\mathcal{D}_{\ell}$ in the left half-plane $\Re(w)<0$. The truth of
said statement for $\mathcal{D}_{\text{r}}$ is obvious, since each
of the terms $1/(z-e_N(\theta_j))$ ($j=0,\dots,N-1$) in $f'/f(z)$
(cf., equations~(\ref{eq:7})--(\ref{eq:8})) has positive real part
for $z$ on $\mathcal{C}_{\text{r}}$.
Now note that
\begin{equation}
\label{eq:13}
-g(z(\phi))=\frac1{2s_N(\theta_0)\sin\phi},\qquad
\delta\leq\phi\leq\pi-\delta,
\end{equation}
is real and positive; we anticipate it to be the
dominant term of the logarithmic derivative $-f'/f(z)$. We will show
that the real parts of each of the terms
$\frac1{z(\phi)-e_N(\theta_j)}$ (for
$\delta\leq\phi\leq\pi-\delta$ and $j=1,\dots,N-2$) are bounded above
by a suitably small multiple of $-g(z(\phi))$.
\begin{lemma}\label{lem:bound-reciprocals}
For all $z$ with $|z|<1$ we have
\begin{equation}
\label{eq:17}
\max_{|\zeta|=1}\Re\frac{1}{z-\zeta} = \frac{1-\Re(z)}{1-|z|^2}.
\end{equation}
\end{lemma}
This lemma can be proved by rewriting the equation $|\zeta|=1$
in the form
\begin{equation*}
\left|w+\frac{\bar z}{1-|z|^2}\right| = \frac1{1-|z|^2},
\end{equation*}
in terms of the variable
\begin{equation*}
w = \frac{1}{z-\zeta},
\end{equation*}
whence the result follows trivially.
\begin{lemma}\label{lem:bound-eta}
For $N\geq3$, $0\leq\theta_0\leq1$, $0<\phi<\pi$ and all $\psi$ we
have
\begin{equation}
\label{eq:18}
\Re\frac{1}{z(\phi)-e_N(\psi)}
\leq -\left(\frac{\pi\theta_0}N
+ 2\pi^2(7+4\sqrt3)\frac{\theta_0^2}{N^2}\right)g(z(\phi)),
\end{equation}
\end{lemma}
First, since $-g(z(\phi))>0$ for $0<\phi<\pi$ (cf.,
equation~(\ref{eq:13})), it suffices to prove
the upper bound
\begin{equation}\label{eq:20}
\left(\frac{\pi\theta_0}N
+ 2\pi^2(7+4\sqrt3)\frac{\theta_0^2}{N^2}\right)
\end{equation}
for the quantity
\begin{equation}
\label{eq:19}
h(\theta_0;\phi,\psi)
:= \frac{-1}{g(z(\phi))}\Re\frac{1}{z(\phi)-e_N(\psi)}
= s_N(\theta_0)\sin\phi\Re\frac{1}{z(\phi)-e_N(\psi)}.
\end{equation}
By Lemma~\ref{lem:bound-reciprocals},
\begin{equation*}
h(\theta_0;\phi,\psi) \leq
\left(\frac{1-\Re(z(\phi))}{1-|z(\phi)|^2}\right) s_N(\theta_0)\sin\phi.
\end{equation*}
From Equation~(\ref{eq:10}) it quickly follows that
\begin{align*}
1-|z(\phi)|^2 &= s_N(2\theta_0)\sin\phi,\quad\text{and}\\
\Re(z(\phi)) &= c_N(\theta_0)-s_N(\theta_0)\sin\phi.
\end{align*}
Thus,
\begin{align}\label{eq:21}
h(\theta_0;\phi,\psi)
&\leq \left(\frac{1-c_N(\theta_0)+s_N(\theta_0)\sin\phi}
{s_N(2\theta_0)\sin\phi}\right)s_N(\theta_0)\sin\phi
=\frac{1-c_N(\theta_0)+s_N(\theta_0)\sin\phi}{2c_N(\theta_0)}\notag\\
&= \frac12\tan\left(\frac{2\pi\theta_0}N\right)\sin\phi
+\frac12\left(\sec\left(\frac{2\pi\theta_0}N\right)-1\right)\notag\\
&\leq \frac12\tan\left(\frac{2\pi\theta_0}N\right)
+\frac12\left(\sec\left(\frac{2\pi\theta_0}N\right)-1\right).
\end{align}
Let us now use the Taylor formul\ae\ with remainder
\begin{align*}
|\tan\theta-\theta| &\leq \frac{\theta^2}2 \sup_{|\vartheta|\leq\theta}|\tan''\vartheta|\\
|\sec\theta-1| &\leq \frac{\theta^2}2 \sup_{|\vartheta|\leq\theta}|\sec''\vartheta|
\end{align*}
with $\theta:=2\pi\theta_0/N$.
Both $|\tan''\vartheta|$ and
$|\sec''\vartheta|$ are even functions of $\vartheta$, increasing
with $|\vartheta|$,
so their respective suprema are bounded by $|\tan''(2\pi/3)|=14$ and
$|\sec''(2\pi/3)|=8\sqrt3$ (since
$\vartheta=\theta=2\pi\theta_0/N\leq2\pi/3$, by the assumptions
$\theta_0\leq1$ and $N\geq 3$). The inequalities
\begin{align*}
\tan\theta &\leq \theta + 7\theta^2,\quad\text{and}\\
\sec\theta-1 &\leq 4\sqrt3\,\theta^2,
\end{align*}
follow immediately. These inequalities together with~(\ref{eq:21})
complete the proof of Lemma~\ref{lem:bound-eta} upon setting
$\theta=2\pi\theta_0/N$.
To complete the proof that $\mathcal{D}_\ell$ is contained in $\Re
w<0$, it remains to show that
\begin{equation}
\label{eq:22}
-\Re\left(\frac{f'}{f}(z(\phi))\right)>0
\qquad\text{for $\delta\leq\phi\leq\pi-\delta$.}
\end{equation}
But
\begin{align*}
-\Re\left(\frac{f'}{f}(z(\phi))\right) &=
-g(z(\phi)) -
\sum_{j=1}^{N-2}\Re\frac1{z(\phi)-e_N(\theta_j)}
\qquad\text{since $-g(z(\phi))>0$} \\
&\geq -g(z(\phi)) \left[1-(N-2)\left(\frac{\pi\theta_0}N
+ 2\pi^2(7+4\sqrt3)\frac{\theta_0^2}{N^2}\right)\right]
\qquad\text{by Lemma~\ref{lem:bound-eta}}\\
&= -g(z(\phi))
\left[1-\pi\theta_0 + U\frac{\theta_0}{N} + V\frac{\theta_0^2}{N} + W\frac{\theta_0^2}{N^2}\right],
\end{align*}
say, for suitable absolute constants $U,V,W$. As $N\to\infty$, the last
bracketed quantity above has the limit $1-\pi\theta_0$. As long as
$|\theta_0|\leq T<\frac1\pi$, there will exist $N(T)$ such that said quantity
is positive for $N\geq N(T)$. This completes the proof of~(\ref{eq:22}) and of
the uniqueness of the desired root $z'=\varsigma(\Theta_0)$ of~$f'$. The
analyticity of $\varsigma$ as a function of $\Theta_0$ follows from the joint
analyticity of the function $f(\Theta_0;z)$ in the variables $\Theta_0$ and
$z$ together with the formula
\begin{equation}
\label{eq:23}
\varsigma(\Theta_0) =
\frac1{2\pi i}\oint_{\mathcal{C}}
\frac{z f''(\Theta_0;z)}{f'(\Theta_0;z)}dz.
\end{equation}
While $\mathcal{C}$ depends on the choice of a fixed $\epsilon$, it
is clear that $\epsilon$ can be chosen so that
$z'=\varsigma(\Theta_0)$ is uniquely
defined by the integral~(\ref{eq:23}) for all $\Theta_0$ in any
desired
compact subset of the interior of $\mathcal{P}_N^{(T)}$. This is
enough to ensure that $\varsigma$ is analytic in the whole interior
of $\mathcal{P}_N^{(T)}$ and concludes the proof of Proposition~8.1.
\end{proof}
|
\section{Introduction}
Subdivision schemes have been the subject of active research in recent years.
In such algorithms, discrete data are recursively generated from coarse to fine by
means of local rules. When the local rules are independent of the data, the underlying
refinement process is linear. This case is extensively studied in literature.
The convergence of this process and the existence of the limit function
was studied in \cite{CDM} and \cite{Dyn} when the
scales are dyadic. When the scales are related to a dilation
matrix $M$, the convergence to a limit function in $L^p$ was studied in
\cite{HanJia} and
generalized to Sobolev spaces in \cite{Jia1} and \cite{Jia2}.
In the linear case, the stability is a consequence of
the smoothness of the limit function.
The nonlinearity arises naturally when one needs to adapt locally the refinement
rules to the data such as in image or geometry processing.
Nonlinear subdivision schemes based
on dyadic scales were originally introduced by Harten
\cite{harten}\cite{hartenENO} through the so-called essentially non-oscillatory (ENO)
methods. These methods have recently been
adapted to image processing into essentially non-oscillatory
edge adapted (ENO-EA) methods. Different versions of ENO methods exist
either based on
polynomial interpolation as in \cite{Co}\cite{Ar} or in a wavelet
framework \cite{Chan}, corresponding to interpolatory or non-interpolatory
subdivision schemes respectively.
In the present paper, we study nonlinear subdivision schemes
associated to dilation matrix $ M$. After recalling the definitions
on nonlinear subdivision schemes in that context, we give sufficient
conditions for convergence in Sobolev and $L^p$ spaces.
\section{General Setting}
\subsection{Notations}
Before we start, let us introduce some notations that will be used throughout the paper.
We denote $\# Q$ the cardinal of the set $Q$.
For a multi-index $\mu=(\mu_1,\mu_2,\cdots,\mu_d) \in \mathbb{N}^{d}$ and a vector
$x=(x_1,x_2,\cdots,x_d)\in \mathbb{R}^{d}$ we define
$|\mu|=\sum\limits_{i=1}^d{\mu_i}$, $\mu !=\prod\limits_{i=1}^d{\mu_i!}$ and $x^{\mu}=\prod\limits_{i=1}^d{{ x_i }^{\mu_i}}.$
For two multi-index $m,\mu\in \mathbb{N}^{d}$ we also define
$$\left (\begin{array}{c}
\mu\\
m
\end{array}
\right )=\left (\begin{array}{c}
\mu_1\\
m_1
\end{array}
\right )
\cdots
\left ( \begin{array}{c}
\mu_d\\
m_d
\end{array}
\right ).
$$
Let $\ell(\mathbb{Z}^{d})$ be the space of all sequences indexed by $\mathbb{Z}^{d}$.
The subspace of bounded sequences is denoted by
$\ell^{\infty}(\mathbb{Z}^{d})$ and $\|u\|_{\ell^{\infty}(\mathbb{Z}^{d})}$ is the supremum of $\{ |u_k|:k \in \mathbb{Z}^{d}\}$.
We denote $\ell^0(\mathbb{Z}^{d})$ the subspace of all sequences with finite
support (i.e. the number of non-zero components of a sequence is finite).
As usual, let $\ell^p(\mathbb{Z}^{d})$ be
the Banach space of sequences
$u$ on $\mathbb{Z}^{d}$ such that $\|u\|_{\ell^p(\mathbb{Z}^{d})} < \infty$, where
$$
\|u\|_{\ell^p(\mathbb{Z}^{d})} := \left ( \sum\limits_{k \in \mathbb{Z}^{d}} |u_k|^p \right
)^{\frac{1}{p}} \textrm { for }1\leq p < \infty.
$$
As in the discrete case, we denote by $L^p(\mathbb{R}^{d})$ the space of all
measurable functions $f$ such that $\|f\|_{L^p(\mathbb{R}^{d})} < \infty$, where
$$
\|f\|_{L^p(\mathbb{R}^{d})} := \left ( \int_{\mathbb{R}^{d}} |f(x)|^p dx \right )^\frac{1}{p}
\textrm{ for } 1 \leq p < \infty
$$
and $\|f\|_{L^{\infty}(\mathbb{R}^{d})}$ is the essential supremum of $|f|$ on $\mathbb{R}^{d}$.
Let $\mu\in \mathbb{N}^{d}$ be a multi-index, we define $\nabla^{\mu}$ the
difference operator $\nabla_1^{\mu_1}\cdots \nabla_d^{\mu_d}$, where
$\nabla_j^{\mu_j}$ is the $\mu_j$th
difference operator with respect to the $j$th coordinate of the canonical basis.
We define $D^{\mu}$ as $D_1^{\mu_1}\cdots D_d^{\mu_d}$, where $D_j$ is the
differential operator with respect to the $j$th coordinate of the canonical basis.
Similarly, for a vector $x \in \Bbb R^{d}$ the differential
operator with respect to $x$ is denoted by $D_x$.
A matrix $M$ is called a dilation matrix if it has integer entries and if
$\lim\limits_{n\rightarrow \infty}M^{-n} =0$.
In the following, the invertible dilation matrix is always denoted by $M$
and $m$ stands for $|det(M)|$.
For a dilation matrix $M$ and any arbitrary function $\Phi$ we put
$\Phi_{j,k}(x) = \Phi(M^{j}x -k)$.
We also recall that a compactly supported function
$\Phi$ is called $L^p$-stable if there exist two constants $C_1, C_2 >0$
satisfying
$$
\quad C_1 \|c\|_{\ell^p(\mathbb{Z}^{d})} \leq \|\sum\limits_{k \in \Bbb Z^{d}}
c_k \Phi(x - k)\|_{L^p(\mathbb{R}^{d})} \leq C_2 \|c\|_{\ell^p(\mathbb{Z}^{d})}.
$$
Finally, for two positive quantities $A$ and $B$ depending on a set of parameters,
the relation $A\raisebox{-1ex}{$~\stackrel{\textstyle<}{\sim}~$} B$ implies the existence of a positive constant $C$,
independent of the parameters, such that $A\leq C B$. Also $A\sim B$
means $A\raisebox{-1ex}{$~\stackrel{\textstyle<}{\sim}~$} B$ and $B\raisebox{-1ex}{$~\stackrel{\textstyle<}{\sim}~$} A$.
\subsection{Local, Bounded and Data Dependent Subdivision Operators, Uniform Convergence
Definition}
In the sequel, we will consider the general class of local, bounded and data dependent subdivision operators which are defined as follows:
\newtheorem{LocalPred}{Definition}
\begin{LocalPred}
\label{locpred}
For $v \in \ell^{\infty}(\mathbb{Z}^{d})$,
a local, bounded and data dependent subdivision operator is defined by
\begin{eqnarray}\label{subdi}
S(v) w_k =
\sum\limits_{l \in \mathbb{Z}^{d}} a_{k-M l} (v)
w_l,
\end{eqnarray}
for any $w$ in $\ell^{\infty}(\mathbb{Z}^{d})$ and where the real coefficients $a_{k- M l}(v) \in \Bbb R$
are such that
\begin{eqnarray}
\label{localdef}
a_{k- M l}(v) = 0, \quad if \quad \|k - M l \|_{\ell^{\infty}(\mathbb{Z}^{d})} > K
\end{eqnarray}
for a fixed constant $K$.
The coefficients $a_{k}(v)$ are assumed to be
uniformly bounded by a constant $C$, i.e. there is $C>0$ independent of $v$
such that:
$$
|a_k(v)| \leq C.
$$
Note that the definition of the coefficients depends on some sequence
$v$, while $S(v)$ acts on the sequence $w$.
Note also that, from \iref{subdi} and \iref{localdef}
the new defined value $S(v) w_k$ depends only on those values $l$
satisfying $\|k - M l \|_{\ell^{\infty}(\mathbb{Z}^{d})} > K$. The subdivision operator in this sense is
{\it local}.
\end{LocalPred}
To simplify, in what follows a data dependent subdivision operator
is an operator in the sense of Definition \ref{locpred}.
With this definition, the associated subdivision scheme
is the recursive action
of the data dependent rule $Sv = S(v)v$ on an initial set of data $v^0$,
according to:
\begin{eqnarray}
v^j = Sv^{j-1} = S(v^{j-1})v^{j-1}, \ j \geq 1.
\end{eqnarray}
\subsection{Polynomial Reproduction for Data Dependent Subdivision Operators}
The study of the convergence of data dependent subdivision operators
will involve the polynomial reproduction property. We recall the definition of
the space $\mathbb{P}_N$ of polynomials of total
degree $N$:
$$
\mathbb{P}_N:= \{ P; P(x) = \sum\limits_{|{\bf \mu}|\leq N} a_\mu
x^{\mu} \}.
$$
With these notations, the polynomial reproduction properties read:
\newtheorem{Exactness}[LocalPred]{Definition}
\begin{Exactness}
Let $N \geq 0$ be a fixed integer.
\begin{enumerate}
\item The data dependent subdivision operator $S$ has
the property of reproduction of polynomials of total degree $N$ if for all $u \in
\ell^{\infty}(\mathbb{Z}^{d})$ and $P\in\mathbb{P}_N$ there exists $\tilde P \in \mathbb{P}_N$
with $P-\tilde P \in \mathbb{P}_{N-1}$ such that $S(u)p=\tilde p$ where $p$
and $\tilde p$ are defined by $p_k=P(k)$ and $\tilde p_k
= \tilde P (M^{-1}k)$.
\item The data dependent subdivision operator $S$ has
the property of exact reproduction of polynomials of total degree $N$ if for all $u \in
\ell^{\infty}(\mathbb{Z}^{d})$ and $P\in\mathbb{P}_N$, $S(u)p=\tilde p$ where $p$
and $\tilde p$ are defined by $p_k=P(k)$ and $\tilde p_k
= P (M^{-1}k)$.
\end{enumerate}
\end{Exactness}
\underline{Remark:}
The case $N=0$ is the so-called "constant reproduction property".
For a data dependent subdivision operator defined as in \iref{subdi},
the constant reproduction property reads $\sum\limits_{k\in\mathbb{Z}^{d}} a_{k-M l}(v) = 1,$ for all ${v} \in \ell^{\infty}(\mathbb{Z}^{d})$.
\section{Definition of Schemes for the Differences}
Another ingredient for our study is the schemes for the differences associated
to the data dependent subdivision operator.
The existence of schemes for the differences is obtained by using the
polynomial reproduction property of the data dependent subdivision operator.
Let us denote $\Delta^{l} = (\nabla^\mu,|\mu|=l)$ and then state
the following result on the existence of schemes for the differences:
\newtheorem{DataDependent}{Proposition}
\begin{DataDependent}
Let $S$ be a data dependent subdivision operator which reproduces polynomials up
to total degree $N$. Then for $1\leq l\leq N+1$ there exists a data dependent
subdivision rule $S_l$ with the property that for all $v$,$w$ in $\ell^{\infty}(\mathbb{Z}^{d})$,
$$
\Delta^l S(v) w := S_l(v) \Delta^l w
$$
\end{DataDependent}
\textsc{Proof:}
Let $l$ be an integer such that $1 \leq l \leq N+1$.
By using the definition of $\nabla^{\mu}$ with $|\mu|=l$, we write:
\begin{eqnarray}
\nonumber
\nabla^{\mu} S(v) w_k =
\nabla^{\mu_1}_1
\cdots
\nabla^{\mu_d}_d S(v) w_k.
\end{eqnarray}
From the definition of $S(v) w$ we infer that
\begin{eqnarray}
\nonumber
\nabla^{\mu} S(v) w_k = \sum\limits_{m_1,\cdots,m_d=0}^{\max (\mu_1,\cdots,\mu_d)}
(-1)^{l} \left
(\begin{array}{c}
\mu\\
m
\end{array}
\right )
\sum_{p \in \mathbb{Z}^{d}}
a_{k-m\cdot e-Mp}(v)
w_p,
\end{eqnarray}
where we have used the notation $m\cdot e= m_1e_1+\cdots+m_d e_d$.
Straightforward computations give
\begin{eqnarray}
\nonumber
\label{vjj1}
\nabla^{\mu} S(v) w_k &=& \sum\limits_{p \in \mathbb{Z}^{d}}
w_p
\sum\limits_{m_1,\cdots,m_d=0}^{\max (\mu_1,\cdots,\mu_d)}
(-1)^{l} \left (\begin{array}{c}
\mu\\
m
\end{array}
\right )
a_{k - m\cdot e -Mp}(v)\ \\
&=&\sum\limits_{p \in \mathbb{Z}^{d}}
w_p f_{k,p}(v,\mu).
\end{eqnarray}
Let us clarify the definition of $f_{k,p}(v,\mu).$
Since the data dependent subdivision operator is local we have $a_{k-Mp}(v) = 0$ for any data $v\in\ell^{\infty}(\mathbb{Z}^{d})$ and any index
$k$ such that $\|k-Mp\|_{\ell^{\infty}(\mathbb{Z}^{d})} > K$. Now by putting
$k = \varepsilon + Mn$, we get that $f_{k,p}(v,\mu)$ is defined for $p$ in the set
$$
V^{\mu}(k) :=
\left \{ p : \|n-p+{M}^{-1}(\varepsilon- m\cdot e)
\|_{\infty} \leq K \|M^{-1}\|_{\infty},\; 0 \leq m_i \leq \mu_i \forall i \right \}
$$
Then, we define
$V(k) := \left \{ p : \|k-Mp\|_{\infty} \leq K \right \} $.
Since the data dependent subdivision scheme reproduces polynomials up
to total degree $N$, we have for any $|\nu|= r \leq N$:
\begin{eqnarray}
\label{poly}
\sum\limits_{p \in V (k)}
a_{k-Mp}(v) p^{\nu} = P_{\nu}(k) \quad \mbox{ for all }k \in \Bbb Z^d,
\end{eqnarray}
where $P_{\nu}$ is a polynomial of total degree $r$.
By tacking the differences of order $|\nu'| = r+1$ in \iref{poly} we get
\begin{eqnarray*}
\sum\limits_{p \in V^{\nu'} (k)}
f_{k,p}(v,\nu') p^{\nu} &=& 0.
\end{eqnarray*}
Note that the above equality is true for any $\nu$
such that $|\nu|=r$. We deduce
that $\left ( f_{k,p}(v,\nu') \right )_{k} \in \Bbb Z^d$
is orthogonal to $\left ( p^q \right )_{p \in V^{\nu'}(k)}$ where
$|q| \leq r$. Note that
$
\left \{
\left ( \nabla^{\nu} \delta_{n-\beta}
\right )_{n \in V^{\nu'}(k)},
|\nu|= r+1,
\beta \in \Bbb Z^d
\right \}
$ spans $\left ( p^q \right )_{p \in V^{\nu'}(k)}$
and we may thus write for any $p \in V^{\nu'}(k)$:
\begin{eqnarray*}
f_{k,p}(v,\nu')=
\sum\limits_{|{\nu}| = r+1}
\sum\limits_{r \in \mathbb{Z}^{d}}
c^\mu_{k,r}(v) \nabla^{\mu} \delta_{p-r}.
\end{eqnarray*}
Now, by using \iref{vjj1} we obtain for any $|\mu| \leq N+1$:
\begin{eqnarray*}
\nabla^{\mu} S(v)w_k &=&
\sum\limits_{p \in V^{\mu}(k)}
w_p
\sum\limits_{|\nu| = l}
\sum\limits_{r \in \Bbb Z^d}
c^\nu_{k,r}(v) \nabla^{\nu} \delta_{p-r}\\
&=&\sum\limits_{ p \in V^{\mu}(k)}
\sum\limits_{|{\nu}| = l}
c^\nu_{k,r}(v)
\nabla^\nu w_p
\end{eqnarray*}
If we now make $\mu$ vary, we obtain the desired relation.
\fin \\
Now that we have proved the existence of schemes for the differences,
we introduce the notion of joint spectral radius for these schemes,
which is a generalization of the one dimensional case which
can be found in \cite{Ost}.
\newtheorem{SpectralRad}[LocalPred]{Definition}
\begin{SpectralRad}
Let $S(v): \ell^p(\mathbb{Z}^{d}) \rightarrow \ell^p(\mathbb{Z}^{d}) $ be a
data dependent subdivision operator such that the difference operators
$S_l(v): \left ( \ell^p(\mathbb{Z}^{d}) \right )^{q_l} \rightarrow \left (\ell^p(\mathbb{Z}^{d}) \right)^{q_l} $,
with $q_l =\#\{\mu,|\mu|=l\}$
exists for $l \leq N+1$.
Then, to each operator $S_l$, $l=0,\cdots,N+1$ (putting $S_0 = S$)
we can associate the joint spectral radius given by
$$
\rho_{p,l}(S) := \inf\limits_{j \geq 1}
\|(S_l)^j\|_{\ell^p(\mathbb{Z}^{d})^{q_l}}^{\frac{1}{j}}.
$$
\end{SpectralRad}
In other words, $\rho_p(S)$ is the infimum of all $\rho >0$
such that for all $v \in \ell^p(\mathbb{Z}^{d}) $, one has
\begin{eqnarray}
\label{spectprat}
\|\Delta^l S^j v \|_{\ell^p(\mathbb{Z}^{d})^{q_l}} \raisebox{-1ex}{$~\stackrel{\textstyle<}{\sim}~$} \rho^j \|\Delta^l v\|_{\ell^p(\mathbb{Z}^{d})^{q_l}},
\end{eqnarray}
for all $j \geq 0$.\\
\noindent \underline{Remark:} Let us define a set of vectors
$\{x_{\bf 1},\cdots,x_{\bf n}\}$ such that
$[x_{\bf 1},\cdots,x_{\bf n}]\mathbb{Z}^n=\mathbb{Z}^{d}$,
$n \geq d$ (i.e. a set such that
the linear combinations of its elements with coefficients in $\mathbb{Z}$
spans $\mathbb{Z}^{d}$). We use the bold notation in the definition of the set so as
to avoid the confusion with the coordinates of vector $x$. Then, consider the differences in the directions $x_{\bf 1},\cdots,x_{\bf n}$. One can show that there exists a scheme for that
differences which we call $\tilde S_l$ for $l \leq N+1$ provided the data dependent
subdivision operator reproduces polynomials up
to degree $N$ (the proof is similar to that using the canonical directions).
If we denote by $\tilde \Delta^l$ the difference operator of order $l$ in the
directions $x_{\bf 1},\cdots,x_{\bf n}$, one can see that
$\|\tilde \Delta^l v\|_{\ell^p(\mathbb{Z}^{d})^{\tilde q_l}} \sim \|\Delta^l v\|_{\ell^p(\mathbb{Z}^{d})^{q_l}}$
for all $v$ in $\ell^p(\mathbb{Z}^{d})$ and where
$\tilde q_l = \# \{\mu, |\mu|=l, \mu =(\mu_i)_{i=1,\cdots,n} \}$.Then,
following (\ref{spectprat}), one can deduce that the joint spectral radius
of $\tilde S_l$ is the same as that of $S_l$.
\section{Convergence in $L^p$ spaces}
In the following, we study the convergence of data dependent
subdivision schemes in $L^p$ which corresponds to the following
definition:
\newtheorem{Converg}[LocalPred]{Definition}
\begin{Converg}
The subdivision scheme $v^j =Sv^{j-1}$
converges in $L^p(\mathbb{R}^{d})$, if for every set of initial control points
$v^0 \in\ell^p(\mathbb{Z}^{d})$, there exists a non-trivial function $v$ in
$L^p(\mathbb{R}^{d})$, called the limit function, such that
$$
\lim\limits_{j \rightarrow \infty} \|v_j-v\|_{L^p(\mathbb{R}^{d})}=0.
$$
\end{Converg}
where $v_j(x) = \sum\limits_{k \in \mathbb{Z}^{d}}
v^j_k \phi_{j,k}(x)$ with
$\phi(x) = \prod\limits_{i=1}^{d} \max(0,1-|x_i|)$.
\subsection{Convergence in the Linear Case}
When $S$ is independent of $v$, the rule (\ref{subdi}) defines a linear subdivision
scheme:
$$
Sv_k =
\sum\limits_{l\in \mathbb{Z}^{d}} a_{k-M l}
v_l.
$$
If the linear subdivision scheme converges for any $v \in
\ell^p(\mathbb{Z}^{d})$ to some function in $L^p(\mathbb{R}^{d})$ and if there exists
$v^0$ such that $\lim\limits_{j\rightarrow +\infty} v^j \neq 0$,
then $\{a_k, k \in \mathbb{Z}^{d}\}$ determines a unique continuous compactly supported
function $\Phi$ satisfying
\begin{eqnarray*}
\Phi (x) =
\sum\limits_{k \in \mathbb{Z}^{d}} a_k \Phi(Mx-k) \textrm{ and }
\sum\limits_{k \in \mathbb{Z}^{d}} \Phi (x -k) = 1.
\end{eqnarray*}
Moreover, $v(x)= \sum\limits_{k \in \mathbb{Z}^{d}}
v^0_k \Phi (x-k)$.
\subsection{Convergence of Nonlinear Subdivision Schemes in $L^p$ Spaces}
In the sequel, we give a sufficient
condition for the convergence of nonlinear subdivision
schemes in $L^p(\Bbb R^{d})$. This result will be a generalization of the
existing result in the
linear context established in \cite{HanJia} and only uses the operator $S_1$.
\newtheorem{ConvContrac}{Theorem}
\begin{ConvContrac}
\label{ConvContract1}
Let $S$ be a data dependent subdivision operator that reproduces the
constants. If $\rho_{p,1}(S) < m^{\frac{1}{p}}$, then
$Sv^j$ converges to a $L^p$ limit function.
\end{ConvContrac}
\textsc{Proof:}
Let us consider
\begin{equation}
\label{fjdef}
v_j(x) := \sum\limits_{k \in \mathbb{Z}^{d}} v^{j}_k
\phi_{j,k}(x),
\end{equation}
where
$\phi(x) = \prod\limits_{i=1}^{d} \max(0,1-|x_i|)$ is the hat function.
With this choice, one can
easily check that $\sum\limits_{k \in \mathbb{Z}^{d}}
\phi(x-k) = 1$.
Let $\rho_{p,1} (S) < \rho < m^{\frac{1}{p}}$,
it follows that
$$
\|\Delta^1 v^j\|_{(\ell^p(\mathbb{Z}^{d}))^d} \raisebox{-1ex}{$~\stackrel{\textstyle<}{\sim}~$} \rho^j \|\Delta^1 v^0\|_{(\ell^p(\mathbb{Z}^{d}))^d},
$$
since $q_1=d$.
We now show that the sequence $v_j$ is a Cauchy sequence in $L^p$:
\begin{eqnarray*}
v_{j+1}(x) - v_j(x)
=
\sum\limits_{k \in \mathbb{Z}^{d}}
v^{j+1}_k
\phi_{j+1,k}(x) -
\sum\limits_{p \in \mathbb{Z}^{d}}
v^{j}_p
\phi_{j,p}(x)
\\
=
\sum\limits_{k \in \mathbb{Z}^{d}}
\sum\limits_{p \in \mathbb{Z}^{d}}
(v^{j+1}_k-v^{j}_p)
\phi_{j+1,k}(x)
\phi_{j,p}(x)
\end{eqnarray*}
where we have used $\sum\limits_{k \in \mathbb{Z}^{d}} \phi(\cdot-k)= 1 .$
Now, since the subdivision operator reproduces the constants:
\begin{eqnarray*}
v_{j+1}(x) - v_j(x)
=\sum\limits_{p \in \mathbb{Z}^{d}}
\sum\limits_{k \in \mathbb{Z}^{d}}
\sum\limits_{l \in \mathbb{Z}^{d}}
a_{k -M l}(v^j) (v^{j}_l-v^{j}_p)
\phi_{j+1,k}(x)
\phi_{j,p}(x).
\end{eqnarray*}
Note that
\begin{eqnarray*}
\sum\limits_{l \in \mathbb{Z}^{d}}
a_{k -M l}(v^j) (v^{j}_l-v^{j}_p) =
\sum\limits_{l \in \mathbb{Z}^{d}} a_{k -M l}(v^j) v^{j}_ l - v^j_p = \sum\limits_{l \in \mathbb{Z}^{d}} (a_{k -M l}(v^j) -\delta_{p-l}) v^{j}_ l
.\end{eqnarray*}
Since $\sum\limits_{l \in \mathbb{Z}^{d}} a_{k -M l}(v^j) -\delta_{p-l} = 0$,
$
\left \{ \nabla_i \delta_{l-\beta}, l \in \left \{ F(k)
\cup \{p\}\right \}, \beta \in \mathbb{Z}^{d}, i=1,\cdots,d \right \}
$ spans $(a_{k-Ml}-\delta_{p-l})_{ l \in \left \{ F(k)
\cup \{p\}\right \}}$.
This enables us to write:
\begin{eqnarray*}
v_{j+1}(x) - v_j(x)
=
\sum\limits_{p \in \mathbb{Z}^{d}}
\sum\limits_{k \in \mathbb{Z}^{d}}
\sum\limits_{l \in V(k)\bigcup \{p\}}
\sum\limits_{i=1}^d d^i_{k,p,l} \nabla_i v^j_l
\phi_{j+1,k}(x)
\phi_{j,p}(x).
\end{eqnarray*}
Since $|\sum\limits_{k \in \mathbb{Z}^{d}}
\phi_{j+1,k}(x)| = 1$ following the same argument
as in Theorem 3.2 of \cite{HanJia}, we may write:
\begin{eqnarray}
\label{diffj}
\|v_{j+1} - v_j\|_{L^p(\mathbb{R}^{d})}
&\raisebox{-1ex}{$~\stackrel{\textstyle<}{\sim}~$}&
m^{-\frac{j}{p}} \max\limits_{1 \leq i \leq d}
\| \nabla_i v^j\|_{\ell^p(\mathbb{Z}^{d})}\nonumber\\
&\raisebox{-1ex}{$~\stackrel{\textstyle<}{\sim}~$}& m^{-\frac{j}{p}}\|\Delta^1 v^j\|_{\ell^p(\mathbb{Z}^{d})}\nonumber\\
&\sim& (\frac{\rho}{m^{\frac{1}{p}}})^j \|\Delta^1 v^0\|_{\ell^p(\mathbb{Z}^{d})}
\end{eqnarray}
which proves that $v_j$ converges in $L^p$, since $\rho < m^{\frac{1}{p}}$.
Note that, for $p=\infty$, we obtain that the limit function is continuous.
\fin
\\
Furthermore, the above proof is valid for any function $\Phi_0$ satisfying the property of partition of unity when $p=\infty$.
In general, we could show, following Theorem 3.4 of \cite{HanJia},
that the limit function in $L^p$ is independent of the choice
of a continuous and compactly supported $\Phi_0$.
\subsection{Uniform Convergence of the Subdivision Schemes to $C^s$ functions
$(s < 1)$}
We are now ready to establish a sufficient condition for the $C^s$ smoothness
of the limit function with $s<1$.
\newtheorem{ConvContracS}[ConvContrac]{Theorem}
\begin{ConvContracS}
\label{ConvContrac2}
Let $S(v)$ be a data dependent subdivision operator which reproduces the constants.
If the scheme for the differences satisfies
$\rho_{p,1} (S)< m^{-s+\frac{1}{p}}$, for some $0< s < 1$ then
$Sv^j$ is convergent in $L^p$ and the limit
function is $C^{s}$ .
\end{ConvContracS}
\textsc{Proof:}
First, the convergence in $L^p$ is a consequence of
$\rho_{p,1} (S)< m^{\frac{1}{p}}$.
In order to prove that the limit function $v$ be in $C^s$, it suffices to
evaluate $|v(x)-v(y)|$ for $\| x - y\|_{\infty}\leq 1$. Let $j$ be
such that $m^{-j-1} \leq \| x-y\|_{\infty} \leq m^{-j}$.
We then write :
\begin{eqnarray*}
|v(x)-v(y)|
&\leq&|v(x)-v_j(x)|+|v(y)-v_j( y)|+
|v_j(x)-v_j(y)| \\
&\leq& 2\|v-v_j\|_{L^{\infty}(\mathbb{R}^{d})}+|v_j(x)-v_j(y)|
\end{eqnarray*}
Note that (\ref{diffj}) implies that
$\|v-v_j\|_{L^{\infty}(\mathbb{R}^{d})}\raisebox{-1ex}{$~\stackrel{\textstyle<}{\sim}~$} \rho^j \|\Delta^1 v^0\|_{\ell^{\infty}(\mathbb{Z}^{d})}$.
Since $v_j$ is absolutely continuous, it is almost everywhere
differentiable, so putting $y = x+M^{-j}h$,
with $h = (h_i)_{i=1,\cdots,d}$ satisfying $\|h\|_\infty \leq 1$ we get:
\begin{eqnarray*}
|v_j(x+ M^{-j}h)-v_j(x)|
&\leq&
|v_j(x+M^{-j}h)-v_j(x+M^{-j}(h-h_d e_d))| \\
&&+|v_j(x+ M^{-j}(h-h_d e_d))-
v_j(x+M^{-j}(h-h_d e_d-
h_{d-1} e_{d-1}))|\\
&&+\cdots+
|v_j(x+M^{-j}(h_1 e_1))-
v_j(x)|
\end{eqnarray*}
Then, using a Taylor expansion we remark that, there exists $\theta \in
]-h_d,h_d[$ such that:
\begin{eqnarray*}
|v_j(x+M^{-j}h)-v_j(x+M^{-j}(h-h_d e_d))| &=&
\sum_{k \in \mathbb{Z}^{d}}
v_k^j h_d D_d\phi(M^{j}x+h-k+\theta_d e_d)
\end{eqnarray*}
If we denote $\Psi_d(x) = \Phi(x_1)\cdots \Phi(x_{d-1})\Psi(x_d)$, where
$\Psi$ is the characteristic function of $[0,1]$ and $\Phi(x_i) =
\max(0,1-|x_i|)$, we may write:
\begin{eqnarray*}
|v_j(x+M^{-j}h)-v_j(x+M^{-j}(h-h_d e_d))| &\sim&
\sum_{k \in \mathbb{Z}^{d}} \nabla_d v_k^j h_d \Psi_d(y)
\end{eqnarray*}
where $y_i=(M^{-j}x+h-k+\theta_d e_d)_i$ if $i <d$ and $y_d = 2(M^{-j}x+h+
\theta_d e_d)_d-k_d$ (we have used the fact that the differential of the hat
function $\Phi$ is the Haar wavelet).
Iterating the procedure for other differences in the sum, we get:
\begin{eqnarray*}
|v_j(x+ M^{-j}h)-v_j(x)|
&\raisebox{-1ex}{$~\stackrel{\textstyle<}{\sim}~$}& \sum_{i=1}^d \|\nabla_i v^{j}\|_{\ell^{\infty}(\mathbb{Z}^{d})}
\raisebox{-1ex}{$~\stackrel{\textstyle<}{\sim}~$} \|\Delta^1 v^j\|_{\ell^{\infty}(\mathbb{Z}^{d})}.
\end{eqnarray*}
Combining these results we may finally write:
\begin{eqnarray*}
|v(x) -v(y)|&\leq& |v(x)-v_j(x)|
+|v(y)-v_j(y)|
+|v_j(x)-v_j(y)|\\
&\leq&2\|v-v_j\|_{L^{\infty}(\mathbb{R}^{d})} + |v_j(x) -v_j(y)|\\
&\raisebox{-1ex}{$~\stackrel{\textstyle<}{\sim}~$}&\rho^j\|\Delta^1 v^0 \|_{\ell^{\infty}(\mathbb{Z}^{d})}+ \|\Delta^1 v^j\|_{\ell^{\infty}(\mathbb{Z}^{d})}\\
&\raisebox{-1ex}{$~\stackrel{\textstyle<}{\sim}~$}&\rho^j \|\Delta^1 v^0 \|_{\ell^{\infty}(\mathbb{Z}^{d})}\raisebox{-1ex}{$~\stackrel{\textstyle<}{\sim}~$} \|x-y\|^s_{\infty}
\end{eqnarray*}
with $s < -\log(\rho_{\infty,1} )/\log m$.
\fin
\section{Examples of Bidimensional Subdivision Schemes }
In the first part of this section, we construct an interpolatory subdivision scheme having the dilation matrix the hexagonal matrix which is:
\begin{eqnarray*}
{ M} &=& \left (
\begin{array}{c c}
2&1\\
0&-2
\end{array}
\right ),
\end{eqnarray*}
For the hexagonal dilation matrix, the coset vectors are $\varepsilon_0=(0,0)^T
,\varepsilon_1=(1,0)^T,\varepsilon_2=(1,-1)^T,\varepsilon_3=(2,-1)^T$.
The coset vector $\varepsilon_i$,$i=0,\cdots,3$ of $M$ defines a partition of $\mathbb{Z}^2$ as
follows:
$$
\mathbb{Z}^2 = \bigcup\limits_{i=0}^{3} \left
\{ Mk+\varepsilon_i, k \in \mathbb{Z}^2 \right \}.
$$
The discrete data at the level $j$, $v^j$ is defined on the grid $\Gamma^{j}= M^{-j}\mathbb{Z}^2$, the value $v_k^j$ is then associated to the location
$M^{-j}k$.
We now define our bi-dimensional
interpolatory subdivision scheme based on
the data dependent subdivision operator which acts from the coarse
grid $\Gamma^{j-1}$ to the fine grid grid $\Gamma^{j}$.
To this end, we will compute $v^j$ at the different coset points on the fine grid $\Gamma^{j}$ using the existing values $v^{j-1}$ of the coarse grid $\Gamma^{j-1}$,
as follows: for the first coset vector
$\varepsilon_0=(0,0)^T$ we simply put
$v^j_{Mk +\varepsilon_0 } = v^{j-1}_k$,
for the coset vectors ${\varepsilon_i}$, $i=1,\cdots,3$.
the value $v^j_{Mk+\varepsilon_i}$, $i=1,\cdots,3$
is defined by affine interpolation of the values on the coarse grid.
To do so,
we define four different stencils on $\Gamma^{j-1}$ as follows:
\begin{eqnarray*}
V_k^{j,1} &=& \{M^{-j+1} k, M^{-j+1}(k+e_1),M^{-j+1}(k+e_2)\},\\
V_k^{j,2} &=& \{M^{-j+1} k, M^{-j+1}(k+e_2),M^{-j+1}(k+e_1+e_2)\},\\
W_k^{j,1} &=& \{M^{-j+1} (k+e_1),M^{-j+1}(k+e_2),M^{-j+1}(k+e_1+e_2)\},\\
W_k^{2} &=& \{M^{-j+1} k,M^{-j+1}(k+e_1),M^{-j+1}(k+e_1+e_2)\}.
\end{eqnarray*}
We determine to which stencils each point of $\Gamma^j$ belongs to, and we then define the prediction as its barycentric coordinates. Since we use an affine interpolant we have:
\begin{eqnarray}
\label{lb1}
v^{j}_{Mk} = v^{j-1}_k \textrm{ and }
v^{j}_{Mk + \varepsilon_1} = \frac{1}{2} v^{j-1}_{k} +
\frac{1}{2} v^{j-1}_{k+e_1}.
\end{eqnarray}
To compute the rules for the coset point $\varepsilon_2$,
$V_k^{1}$ or $V_k^{2}$ can be used leading respectively to:
\begin{eqnarray}
\label{lb2}
v^{j,1}_{Mk+\varepsilon_2} &=& \frac{1}{4} v^{j-1}_{k+e_1} +
\frac{1}{2} v^{j-1}_{k+e_2}
+\frac{1}{4} v^{j-1}_{k} \nonumber\\
v^{j,2}_{Mk+\varepsilon_2} &=& \frac{1}{2} v^{j-1}_k +
\frac{1}{4} v^{j-1}_{k+e_2} +
\frac{1}{4} v^{j-1}_{k+e_1+e_2}.
\end{eqnarray}
When one considers the rules for the coset point $\varepsilon_3$,
$W_k^{1}$ or $W_k^{2}$ can be used leading respectively to:
\begin{eqnarray}
\label{lb3}
v^{j,1}_{Mk+\varepsilon_3} &=& \frac{1}{4} v^{j-1}_{k+e_2} +
\frac{1}{4} v^{j-1}_{k+e_1+e_2}
+\frac{1}{2} v^{j-1}_{k+e_1} \nonumber \\
v^{j,2}_{Mk+\varepsilon_3} &=& \frac{1}{4} v^{j-1}_k +
\frac{1}{4} v^{j-1}_{k+e_1} +
\frac{1}{2} v^{j-1}_{k+e_1+e_2}.
\end{eqnarray}
This nonlinear scheme is converging in $L^\infty$ since we have the following result:
\newtheorem{conver}[DataDependent]{Proposition}
\begin{conver}
The prediction defined by (\ref{lb1}), (\ref{lb2}), (\ref{lb3}) satisfies:
$$
\|\Delta^1 v^j_{M.+\varepsilon_i}\|_{\ell^{\infty}(\mathbb{Z}^{d})} \leq \frac{3}{4}
\|\Delta^1 v^{j-1}\|_{\ell^{\infty}(\mathbb{Z}^{d})}
$$
\end{conver}
We do not detail the proof here but the result is obtained by computing the differences in the canonical directions
at each coset points.
If we then use Theorem \ref{ConvContrac2} we can find the regularity of the
corresponding limit function
is $C^s$ with $s< -\frac{\log (3/4)}{\log(4)} \approx 0.207$.
In the second pat of the section, we build an
example of bidimensional subdivision scheme based on the same
philosophy but this time using as dilation matrix the quincunx matrix
defined by:
\begin{eqnarray*}
M&=&\left (
\begin{array}{cc}-1 & 1\\
1&1
\end{array}
\right ),
\end{eqnarray*}
whose coset vectors are $\varepsilon_0=(0,0)^T$ and
$\varepsilon_1=(0,1)^T$.
Note that $a_{0,0}=1$ and since the nonlinear subdivision operator
reproduces the constants we have
$\sum\limits_{i} a_{Mi+\varepsilon}=1$ for all coset vectors
$\varepsilon$. To build the subdivision operator, we consider the subdivision rules based on interpolation
by of first degree polynomials on the grid $\Gamma^{j-1}$.
$v^j_{Mk+\epsilon_1}$ corresponds to a point
inside the cell delimited by
$M^{-j+1} \{ k,k+e_1,k+e_2,k+e_1+e_2\}$.
There are four potential stencils, leading in this case only to two subdivision rules:
\begin{eqnarray}
\label{quin1}
\hat v^{j,1}_{Mk+\epsilon_1}&=\frac 1 2 (v_{k}^{j-1}+v_{k+e_1+e_2}^{j-1})
\end{eqnarray}
\begin{eqnarray}
\label{quin2}
\hat v^{j,2}_{Mk+\epsilon_1}&=\frac 1 2 (v_{k+e_1}^{j-1}+v_{k+e_2}^{j-1})
\end{eqnarray}
Note also, that since the scheme is interpolatory we have the
relation: $v^j_{Mk} = v^{j-1}_k$. Let us
now prove a contraction property for the above scheme.
\newtheorem{convquin}[DataDependent]{Proposition}
\begin{convquin}
\label{convquin1}
The nonlinear subdivision scheme defined by (\ref{quin1}) and (\ref{quin2}) satisfies the following property:
\begin{enumerate}
\item when $k=Mk'$:
\begin{eqnarray*}
\|v^{j,1}_{M.+\varepsilon_1}-v^j_{M.}\|_{\ell^{\infty}(\mathbb{Z}^{d})}&\leq&
\frac{1}{2}\|\Delta^1 v_.^{j-2}\|_{\ell^{\infty}(\mathbb{Z}^{d})}\\
\|v^{j,2}_{M.+\varepsilon_1}-v^j_{M.}\|_{\ell^{\infty}(\mathbb{Z}^{d})}&\leq&
\|\Delta^1 v_.^{j-2}\|_{\ell^{\infty}(\mathbb{Z}^{d})}
\end{eqnarray*}
\item when $k=Mk'+\varepsilon_1$, we can show that:
\begin{eqnarray*}
\|v^{j,2}_{M.+\varepsilon_1}-v^j_{M.}\|_{\ell^{\infty}(\mathbb{Z}^{d})}&\leq& \frac{1}{2}
\|\Delta^1 v_.^{j-2}\|_{\ell^{\infty}(\mathbb{Z}^{d})}\\
\|v^{j,1}_{M.+\varepsilon_1}-v^j_{M.}\|_{\ell^{\infty}(\mathbb{Z}^{d})}&\leq&
\|\Delta^1 v_.^{j-2}\|_{\ell^{\infty}(\mathbb{Z}^{d})}
\end{eqnarray*}
\end{enumerate}
\end{convquin}
The proof of this theorem is obtained computing all the potential differences.
This theorem shows that the nonlinear subdivision scheme converges in
$L^\infty$ since $\rho_{1,\infty}(S) < 1$.
\section{Convergence in Sobolev Spaces}
In this section, we extend the result established in \cite{Jia2}
on the convergence of linear subdivision scheme to
our nonlinear setting. We will first recall the notion of
convergence in Sobolev spaces in the linear case.
Following \cite{Jia} Theorem 4.2, when
$\Phi_0(x) = \sum\limits_{k\in \mathbb{Z}^{d}} a_k \Phi_0(Mx-k)$ is $L^p$-stable, the
so-called "moment condition of order $k+1$ for $a$" is equivalent to the
polynomial reproduction property of polynomial of total degree $k$
for the subdivision scheme associated to $a$. In what follows, we will say
that $\Phi_0$ reproduces polynomial of total degree $k$. When the subdivision
associated to $a$ exactly reproduces polynomials, we will say that $\Phi_0$
exactly reproduces polynomials.
We then have the following definition for the convergence of subdivision
schemes in Sobolev spaces in the linear case \cite{Jia2}:
\newtheorem{LinearSobo}[LocalPred]{Definition}
\begin{LinearSobo}
We say that $v^j=Sv^{j-1}$ converges in the Sobolev space $W_N^k(\mathbb{R}^{d})$ if there exists
a function $v$ in $W_N^k(\mathbb{R}^{d})$ satisfying:
$$
\lim\limits_{j\rightarrow +\infty} \|v_j-v\|_{W_N^k(\mathbb{R}^{d})} = 0
$$
where $v$ is in $W_N^k(\mathbb{R}^{d})$, and $v_j=\sum\limits_{k \in \mathbb{Z}^{d}} v_k^j \Phi_0(M^jx-k)$
for any $\Phi_0$ reproducing polynomials of total degree $k$.
\end{LinearSobo}
We are going to see that in the nonlinear case, to ensure the convergence
we are obliged to make a restriction on the choice of $\Phi_0$.
We will first give some results when the matrix $M$ is an isotropic dilation
matrix, we will also emphasize a particular class of isotropic matrices,
very useful in image processing.
\subsection{Definitions and Preliminary Results}
\newtheorem{isotropic}[LocalPred]{Definition}
\begin{isotropic}
\label{isotropic1}
We say that a matrix $M$ is isotropic if it is similar to the diagonal matrix $\textrm{diag}(\sigma_1, \ldots, \sigma_d)$, i.e. there exists an invertible matrix $\Lambda$ such that
\begin{equation*}
M=\Lambda^{-1} \textrm{diag}(\sigma_1, \ldots, \sigma_d) \Lambda,
\end{equation*}
with $|\sigma_1| = \ldots = |\sigma_d|$ being the eigenvalues of matrix $M$.
\end{isotropic}
Evidently, for an isotropic matrix holds $|\sigma_1| = \ldots = |\sigma_d|=\sigma=m^{\frac{1}{d}}$.
Moreover, for any given norm in $\mathbb{R}^{d}$, any integer
$n$ and any $v\in \mathbb{R}^{d}$ we have
$$
\sigma^{n}\|u\| \raisebox{-1ex}{$~\stackrel{\textstyle<}{\sim}~$} \|M^n u\| \raisebox{-1ex}{$~\stackrel{\textstyle<}{\sim}~$} \sigma^{n} \|u\|.
$$
A particular class of isotropic matrices is when there exists a set
$\tilde e_1, \tilde e_2,\cdots,\tilde e_q$ such that:
\begin{eqnarray}
\label{caspart}
M \tilde e_i = \lambda_i \tilde e_{\gamma(i)}
\end{eqnarray}
where $\gamma$ is a permutation of $\{1,\cdots,q\}$. Such matrices are
particular cases of isotropic matrices
since $M^q=\lambda I$ where $I$ is the identity matrix and where $\lambda = \prod\limits_{i=1}^d \lambda_i$.
For instance, when $d=2$, the quincunx (resp. hexagonal) matrix satisfies $M^2=2I$ (resp. $M^2=4I$).
We establish the following property on joint spectral radii that will be useful when dealing with the convergence in Sobolev spaces.
\newtheorem{spectral1}[DataDependent]{Proposition}
\begin{spectral1}
\label{spectral}
Assume that $S$ reproduces polynomials up to total
degree $N$. Then,
$$
\rho_{p,n+1}(S)\geq
\frac {1} {\|M\|_{\infty}} \rho_{p,n}(S),
$$
for all $n=0,\ldots,N$.
\end{spectral1}
\noindent \underline{Remark:}
If $M$ is an isotropic matrix and $S$ reproduces polynomials
up to total degree $N$, then
\begin{equation*}
\rho_{p,n+1}(S)\geq
\sigma^{-1} \rho_{p,n}(S),
\end{equation*}
for all $n=0,\ldots,N$.\\
\textsc{Proof:} It is enough to prove
\begin{equation*}
\rho_{p,1}( S)\geq
\frac {1} {\|M\|_{\infty}} \rho_p(S).
\end{equation*}
According to the definition of spectral radius there exists
$\rho>\rho_{p,1}(S)$ such that for any $u^0$
\begin{equation*}
\|S_{1}(u^{j-1}) \ldots S_{1}(u^0){\nabla} u\|_{\ell^p(\mathbb{Z}^{d})}
\raisebox{-1ex}{$~\stackrel{\textstyle<}{\sim}~$} \rho^j \|{\nabla} u\|_{(\ell^p(\mathbb{Z}^{d})}.
\end{equation*}
Using the notation $\omega^j := S(u^{j-1})\cdot \ldots \cdot S(u^0) u$
we obtain
\begin{equation*}
\|{\nabla} \omega^j\|_{\ell^p(\mathbb{Z}^{d})} \raisebox{-1ex}{$~\stackrel{\textstyle<}{\sim}~$} \rho^j \|{\nabla}
u\|_{\ell^p(\mathbb{Z}^{d})}.
\end{equation*}
Since
\begin{equation*}
\omega^j_l = \sum_n A^j_{l,n} u_n,
\end{equation*}
where
\begin{equation*}
A^j_{l,n} = \sum_{l_1, \ldots, l_{j-1}}a_{l-Ml_{j-1}}(u^{j-1})
a_{l_{j-1}-Ml_{j-2}}\cdot \ldots \cdot a_{l_1-Mn}(u^0).
\end{equation*}
We can write down the $\ell^p$-norm as follows:
\begin{equation*}
\|\omega^j\|^p_{\ell^p(\mathbb{Z}^{d})} =
\sum_{k\in \mathbb{Z}^d}\sum_{i=1}^{m^j}|\omega^j_{M^jk+\varepsilon^j_i}|^p,
\end{equation*}
where $\{\varepsilon^j_i\}_{i=1}^{m^j}$ are the representatives of cosets of $M^j$.
First note that:
\begin{equation*}
\|k-n\|_{\infty} \leq \|k-n+M^{-j}\varepsilon^j_i\|_{\infty}+
\|M^{-j}\varepsilon^j_i\|_{\infty}.
\end{equation*}
Note that $M^{-j} \varepsilon_i^j$ belongs to the unit square so that
$\|M^{-j}\varepsilon^j_i\|_{\infty} \leq K_1$.
When $A^j_{M^jk+\varepsilon^j_i,
n} \neq 0$, one can prove that there exists $K_2>0$ such that
\begin{equation*}
\|k-n+M^{-j}\varepsilon^j_i\|_{\infty} \leq K_2,
\end{equation*}
the proof being similar to that of Lemma 2 in \cite{Goodman}.
From these inequalities it follows that if
$A^j_{M^jk+\varepsilon^j_i, n} \neq 0$ there exists $K_3>0$ such that
\begin{equation*}
\|k-n\|_{\infty} \leq K_3,
\end{equation*}
that is, for a fixed $k$, the values of $\omega^j_l$ for
$l \in \{M^jk+\varepsilon^j_i\}^{m^j}_{i=1}$ depend only on
$u_n$ with $n : \{\|k-n\|_{\infty} \leq K_3\}$.
\noindent Let us now fix $k$ and define $\tilde{u}$ such that
$$
\tilde{u}_l=\begin{cases}
u_l,&\text{if $\|k-l\|_{\infty} \leq K_3$;}\\
0,&\text{otherwise.}
\end{cases}
$$
Let $\tilde{\omega}^j := S(u^{j-1})\cdot \ldots \cdot S(u^0) \tilde{u}$, then
$$
\tilde{\omega}^j_l=\begin{cases}
\omega^j_l,&\text{ if $l \in \{M^jk+\varepsilon^j_i\}^{m^j}_{i=1}$;}\\
0,&\text{ if $\|k-M^{-j}l\|_{\infty} \geq K_4$,}
\end{cases}
$$
since if $A^j_{l, n}\neq 0$, then
\begin{equation*}
\|k-M^{-j}l\|_{\infty} \leq \|k-n\|_{\infty} + \|n-M^{-j}l\|_{\infty} \leq K_3 +
K_2 := K_4.
\end{equation*}
Moreover, from $\|k-M^{-j} l\|_{\infty} \leq K_4$, it
follows that $\|M^{j}k-l\|_{\infty} \leq K_4 \|M^{j}\|_{\infty}$.
Taking all this into account, we get
\begin{eqnarray*}
\sum\limits_{k \in \Bbb Z^d}
\sum_{l \in \{M^jk+\varepsilon^j_i\}}
|\omega^j_l|^p &=&
\sum\limits_{k \in \Bbb Z^d} \sum_{l \in \{M^jk+\varepsilon^j_i\}}|
\tilde{\omega}^j_l|^p \leq
\sum\limits_{k \in \Bbb Z^d}
\sum_{\|M^{j}k-l\|_{\infty} \leq K_4 \|M^{j}\|_{\infty}}
|\tilde{\omega}^j_l|^p \\
&\raisebox{-1ex}{$~\stackrel{\textstyle<}{\sim}~$}& \|M\|_{\infty}^j \|\Delta^1 \tilde{\omega}^j_l\|_{\ell^p(\mathbb{Z}^{d})}
\raisebox{-1ex}{$~\stackrel{\textstyle<}{\sim}~$} (\|M\|_{\infty} \rho)^j \|\Delta^1 \tilde{u}\|_{\ell^p(\mathbb{Z}^{d})}.
\end{eqnarray*}
That is, $\|\omega^j\|_{\ell^p(\mathbb{Z}^{d})} \raisebox{-1ex}{$~\stackrel{\textstyle<}{\sim}~$} (\|M\|_{\infty} \rho)^j
\|u\|_{\ell^p(\mathbb{Z}^{d})}$, consequently $\rho_p(S) \raisebox{-1ex}{$~\stackrel{\textstyle<}{\sim}~$} \|M\|_{\infty} \rho$.
Now, if $\rho \rightarrow \rho_{p,1}(S)$ we get
$\rho_p(S) \leq \|M\|_{\infty} \rho_{p,1}( S)$.
\fin
\subsection{Convergence in Sobolev Spaces When $M$ is Isotropic}
First, Let us recall that the Sobolev norm on $W_N^p(\mathbb{R}^{d})$ is defined by:
\begin{eqnarray}
\label{defSobo}
\|f\|_{W_N^p(\mathbb{R}^{d})} = \|f\|_{L^p(\mathbb{R}^{d})} + \sum\limits_{|\mu| \leq N} \|D^\mu f\|_{L^p(\mathbb{R}^{d})}.
\end{eqnarray}
If one considers a set $x_{\bf 1},\cdots,x_{\bf n}$ such that $
[x_{\bf 1},\cdots,x_{\bf n}]\mathbb{Z}^n=\mathbb{Z}^d$, an equivalent norm
is given by:
\begin{eqnarray}
\label{defSobo1}
\|f\|_{W_N^p(\mathbb{R}^{d})} = \|f\|_{L^p(\mathbb{R}^{d})} + \sum\limits_{|\mu| \leq N} \|\tilde D^\mu f\|_{L^p(\mathbb{R}^{d})}.
\end{eqnarray}
where
$\tilde D^{\mu} = D^{\mu_1}_{x_{\bf 1}} \cdots D^{\mu_n}_{x_{\bf n}}$.
\\
We then enounce a convergence theorem for general isotropic matrix $M$:
\newtheorem{segaln}[ConvContrac]{Theorem}
\begin{segaln}
\label{segaln}
Let $S$ be a data dependent nonlinear subdivision scheme which exactly
reproduces polynomials up to total degree $N-1$, then the subdivision scheme
$Sv^j$ converges in $W_N^p(\Bbb R^{d})$, provided
$\Phi_0$ is compactly supported and exactly reproduces polynomials up to total degree $N-1$ and
\begin{eqnarray}
\label{rspect1}
\rho_{p,N}(S) < m^{\frac{1}{p}-\frac{s}{d}} \textrm{ for some }s > N.
\end{eqnarray}
\end{segaln}
\textsc{ Proof:}
Note that because of Proposition \ref{spectral}, the hypotheses of Theorem
\ref{segaln} imply
that
$
\rho_{p,k}(S) < m^\frac{1}{d} \rho_{p,k+1}(S) < m^{\frac{1}{p}-\frac{s-1}{d}},
$
which means that (\ref{rspect1}) is also true for $k < N$.
Let us now show that $v_j$ is a Cauchy sequence in $L^p$. To do so, let us define
$$
q_j(x)=\sum_{l=1}^d \lambda_{j,l}x_l,
$$
where $\Lambda = (\lambda_{j,l})$ is defined in (\ref{isotropic1}).
For a multi-index $\mu=(\mu_1, \ldots, \mu_d) \in \mathbb Z^d$ let
$$
q_\mu (x) = q_1^{\mu_1}(x) \ldots q_d^{\mu_d}(x).
$$
Since $\Lambda$ is invertible, the set $\{q_{\mu}: |\mu| = N\}$
forms a basis of the space of all polynomials of exact degree $N$,
which proves that
$$
\|D^\mu (v_{j+1}-v_j)\|_{L^p(\mathbb{R}^{d})} \sim
\|q_{\mu}(D)(v_{j+1}-v_j)\|_{L^p(\mathbb{R}^{d})}
$$
Now, we use the fact that, since $M$ is isotropic,
$
q_\mu(D) (f(M^jx))=\sigma^{j\mu}(q_\mu(D) f)(M^jx)
$
where $\tilde \sigma^\mu = \prod\limits_{i=1}^d \sigma_i^{\mu_i}$
(\cite{Jia1}). We can thus write:
\begin{eqnarray*}
q_\mu(D)(v_{j+1}-v_j)=
q_\mu(D) \left(
\sum_{l \in \mathbb{Z}^{d}} v^{j+1}_l \Phi_0 (M^{j+1}x-l) -
\sum_{l \in \mathbb{Z}^{d}}
v^j_l \Phi_0 (M^jx-l)\right ).
\end{eqnarray*}
We use now the scaling equation of $\Phi_0$ to get
\begin{eqnarray*}
q_\mu(D)(v_{j+1}-v_j)
&=&q_\mu(D)
\left(
\sum_{l \in \mathbb{Z}^{d}}v^{j+1}_l \Phi_0 (M^{j+1}x-l) -
\sum_{l \in \mathbb{Z}^{d}}
\sum_{r \in \mathbb{Z}^{d}} v^j_r g_{l-Mr} \Phi_0 (M^{j+1} x- l) \right )\\
&=& \sum_{l \in \mathbb{Z}^{d}} (v^{j+1}_l - \sum_{r \in \mathbb{Z}^{d}} v^j_r g_{k-Mr} )
q_\mu (D) \left ( \Phi_0 (M^{j+1}x-l) \right )\\
&=& \sum_{l \in \mathbb{Z}^{d}} \sum_{r \in \mathbb{Z}^{d}} (a_{l-Mr}(v^j)-g_{l-Mr}) v^j_r
\tilde \sigma^{\mu(j+1)} (q_{\mu}(D) \Phi_0) (M^{j+1}x-l).
\end{eqnarray*}
Since $S$ and $\Phi_0$ exactly reproduce polynomials up to total degree
$N-1$, we have for $|\mu| \leq N-1$:
$$
\sum_{r \in \mathbb{Z}^{d}} (a_{l-Mr}(v^j)-g_{l-Mr}) r^{\mu} = 0.
$$
Remark that $g_{l-Mr}= 0$ for $\|l-Mr\| > \tilde K$ since $\Phi_0$ is compactly
supported.
Since
$
\left \{
\nabla^{\nu} \delta_{l-\beta}, |\nu| = N, r
\in F(l)=\left \{ \|l-Mr \| \leq \max(K,\tilde K) \right \}, \beta \in \mathbb{Z}^{d}
\right \}
$ spans $(a_{l-Mr}(v^j)-g_{l-Mr})_{r \in F(l)}$,
we deduce:
\begin{eqnarray*}
q_\mu(D)(v_{j+1}-v_j) &=&
\sum_{l \in \mathbb{Z}^{d}} \sum_{r \in F(l)} \sum_{|\nu|=N}c^{\nu}_{r}(v^{j})
\nabla^{\nu} v^{j}_r \tilde \sigma^{\mu (j+1)} (q_{\mu}(D) \Phi_0) (M^{j+1}x-l),
\end{eqnarray*}
Consequently,
\begin{eqnarray*}
\|q_{\mu}(D)(v_{j+1}-v_j)\|_{L^p(\mathbb{R}^{d})}
&\raisebox{-1ex}{$~\stackrel{\textstyle<}{\sim}~$}& \sigma^{(j+1)N} m^{-(j+1)/p} (\rho_{p,N}(S))^j
\|\Delta^{N} v^{0}\|_{(\ell^p(\mathbb{Z}^{d}))^{q_{N}}}
\end{eqnarray*}
Since $\rho_{p,N}(S) < m^{1/p-s/d}$, with $s > N$ we obtain
\begin{eqnarray*}
\|q_{\mu}(D)(v_{j+1}-v_j)\|_{L^p(\mathbb{R}^{d})} &\raisebox{-1ex}{$~\stackrel{\textstyle<}{\sim}~$}& \sigma^{j(N-s)}
\|\Delta^{N} v^0\|_{(\ell^p(\mathbb{Z}^{d}))^{q_{N}}}.
\end{eqnarray*}
From this we deduce that $\|q_{\mu}(D)(v_{j+1}-v_j)\|_{L^p(\mathbb{R}^{d})}$ tends to $0$ with $j$. Making $\mu$ vary, we deduce the
convergence in $W_{N}^p(\mathbb{R}^{d})$ \fin
We now show that when the matrix $M$ satisfies (\ref{caspart}) and when $\Phi_0$ is a box spline
satisfying certain properties, the limit function is in $W_N^p(\mathbb{R}^{d})$. Before that, we need to recall
the definition of box splines and some properties that we will use.
Let us define a set of $n$ vectors, not necessarily distinct:
$$
X_n = \{ x_{\bf 1}, \cdots,x_{\bf n} \}
\subset \mathbb{Z}^{d} \setminus \{ 0 \}.
$$
We assume that $d$ vectors of $ X_n$ are linearly independent.
Let us rearrange the family $X_n$ such that
$X_d = \{ x_{\bf 1},\cdots,x_{\bf d} \}$ are linearly
independent. We denote by
$[x_{\bf 1},\cdots,x_{\bf d}][0,1[^{d}$ the collection of linear combinations
$\sum\limits_{i=1}^{d} \lambda_i x_{\bf i}$ with $\lambda_i \in [0,1[$.
Then, we define multivariate box splines as follows \cite{Ch}\cite{Pr}:
\begin{eqnarray}\label{defSpli}
\beta_0(x,X_d ) &=&\left \{ \begin{array}{c}
\frac{1}{|\det(x_{\bf 1},\cdots,x_{\bf d})|} \textrm{ if }
x \in [x_{\bf 1},\cdots,x_{\bf d}][0,1[^{d}\\
0 \textrm{ otherwise}
\end{array}
\right . \nonumber \\
\beta_0(x,X_k) &=& \int_0^1 \beta_0 (x-tx_{\bf k},X_{k-1}) dt,
\quad n \geq k > d.
\end{eqnarray}
One can check by induction that the support of $\beta_0(x,X_n)$ is
$[x_{\bf 1},x_{\bf 2},\cdots,x_{\bf n}][0,1]^n$. The regularity of box
splines is then given by the following theorem \cite{Pr}:
\newtheorem{Differentiability}[DataDependent]{Proposition}
\begin{Differentiability}
\label{Different}
$\beta_0(x,X_n)$ is $r$ times continuously differentiable if all
subsets of $X_n$ obtained by deleting $r+1$ vectors spans
$\mathbb{R}^{d}$.
\end{Differentiability}
We recall a property on the directional derivatives
of box splines, which we use in the convergence theorem that follows:
\newtheorem{Differentiation}[DataDependent]{Proposition}
\begin{Differentiation}
\label{Different1}
Assume that $X_n \setminus x_{\bf r}$ spans $\mathbb{R}^{d}$, and
consider the following box spline function $s(x) = \sum\limits_{k \in \mathbb{Z}^{d}}
c_k \beta_0(x-k,X_n)$
then the directional derivative of $s$ in the direction $x_{\bf r}$
reads:
$$
D_{x_{\bf r}} s(x) = \sum\limits_{k \in \mathbb{Z}^{d}}
\nabla_{x_{\bf r}} c_k
\beta_0(x-k,X_n \setminus x_{\bf r}).
$$
\end{Differentiation}
We will also need the property of polynomial reproduction which is \cite{Pr}:
\newtheorem{Polrep}[DataDependent]{Proposition}
\begin{Polrep}
\label{Polrep1}
If $\beta_0(x,X_n)$ is r times continuously differentiable then, for any polynomial $c(x)$ of total degree $d \leq r+1$,
\begin{eqnarray}
\label{defv}
p(x) = \sum_{i \in \mathbb{Z}^{d}} c(i) \beta_0(x-i,X_n)
\end{eqnarray}
is a polynomial with total degree $d$, with the same leading coefficients
(i.e. the coefficients corresponding to degree $d$). Conversely, for any
polynomial $p$, it satisfies (\ref{defv}) with
$c$ being a polynomial having the same leading coefficients as $p$.
\end{Polrep}
\newtheorem{segaln2}[ConvContrac]{Theorem}
\begin{segaln2}
\label{segaln1}
Let $S$ be a data dependent nonlinear subdivision scheme which reproduces
polynomials up to total degree $N-1$ and assume that $M$ satisfies
relation (\ref{caspart}), then the subdivision scheme $Sv^j$ converges
in $W_N^p(\Bbb R^{d})$, if when $N \geq 2$, $\Phi_0$ is a
$C^{N-2}$ box spline generated by $x_{\bf 1},\cdots,x_{\bf n}$
satisfying $\Phi_0(x) =\sum\limits_{k} g_k \Phi_0(Mx-k)$ and if
$N =1$ $\Phi_0(x) =\sum\limits_{k} g_k \Phi_0(Mx-k)$ and
$\sum\limits_{k \in \mathbb{Z}^{d}} \Phi_0(x-k)=1$ and if
\begin{eqnarray}
\label{rspect}
\rho_{p,N}(S) < m^{\frac{1}{p}-\frac{s}{d}} \textrm{ for some } s > N.
\end{eqnarray}
\end{segaln2}
\textsc{Proof:} We here prove the case $N \geq 2$, the case $N=1$ can be proved
similarly. First note that since $\Phi_0(x)$ is a $C^{N-2}$ box spline,
we can write for any polynomial $p$ of total degree $N-1$ at most:
\begin{eqnarray*}
p(M^{-1}x) &=& \sum\limits_{i \in \mathbb{Z}^{d}} \tilde p(i) \Phi_0(M^{-1}x-i,X_n)\\
&=&\sum\limits_{q\in \mathbb{Z}^{d}} \sum\limits_{i \in \mathbb{Z}^{d}} g_{q-Mi} \tilde p(i)
\Phi_0(x-q,X_n)
\end{eqnarray*}
Using Proposition \ref{Polrep1} we get $p$ and $\tilde p$ have the same leading coefficients, and that
$\sum\limits_{i \in \mathbb{Z}^{d}} g_{q-Mi} \tilde p(i)$ is a polynomial evaluated in
$M^{-1}i$ having the same leading
coefficients as $p$. That is to say the subdivision scheme
$(Sv^j)_q = \sum\limits_{i \in \mathbb{Z}^{d}} g_{q-Mi} v^j_i$ reproduces
polynomials up to degree $N-1$.
As already noticed, the joint spectral radius of difference operator is independent of the choice of the
directions $x_{\bf 1},\cdots,x_{\bf n}$ that spans $\mathbb{Z}^{d}$.
Furthermore, it is shown in \cite{Me}, that the existence of a scaling
equation for $\Phi_0$ implies that the vectors $x_{\bf i}$, ${\bf i} =
{\bf 1},\cdots,{\bf n}$ satisfy a relation of type (\ref{caspart}).
We consider such a set $\{ x_{\bf i} \}_{{\bf i}={\bf 1},\cdots,{\bf n}}$
and then define $\Phi_0(x)=\beta_0(x,Y_N)$ the box spline associated
to the set
$$
Y_N:=\left \{ \overbrace{x_{\bf 1},\cdots,x_{\bf 1}}^{N},\cdots,
\overbrace{x_{\bf n},\cdots,x_{\bf n}}^{N} \right \}.
$$
which is $C^{N-2}$ by definition. We then define the differential operator
$ \tilde D_{M^{-j}}^\mu :=\tilde D^{\mu_1}_{M^{-j}
x_{\bf 1}}\cdots \tilde D^{\mu_n}_{M^{-j} x_{\bf n}}$. We will use the
characterization (\ref{defSobo1}) of Sobolev spaces therefore $\mu =
(\mu_i)_{i=1,\cdots,n}$. For any $|\mu| \leq N$ we may write:
\begin{eqnarray*}
\tilde D^{\mu}_{M^{-j-1}} (v_{j+1}(x) - v_j(x))
&=& \sum\limits_{k \in \mathbb{Z}^{d}}
v^{j+1}_k
(\tilde D^{\mu} \beta_0)(M^{j+1}x-k,Y_N)\\
&&-
\sum\limits_{p \in \mathbb{Z}^{d}} \sum_{i \in \mathbb{Z}^{d}}
v^{j}_i g_{p-Mi} (\tilde D^{\mu} \beta_0)(M^{j+1}x - p,Y_N),\\
\end{eqnarray*}
using the scaling property satisfied by $\beta_0$.
Then, we get:
\begin{eqnarray*}
\tilde D^{\mu}_{M^{-j-1}} (v_{j+1}(x) - v_j(x))
&=& \sum\limits_{k \in \mathbb{Z}^{d}} \sum_{i \in \mathbb{Z}^{d}}
(a_{k-Mi}(v^j)-g_{k-Mi}) v^j_i (\tilde D^{\mu} \beta_0)(M^{j+1}x-k,Y_N) \\
&=&\sum\limits_{k \in \mathbb{Z}^{d}} \tilde \nabla^{\mu} ( \sum_{i \in \mathbb{Z}^{d}}
(a_{k-Mi}(v^j)-g_{k-Mi}) v^j_i) \beta_0(M^{j+1}x-k,Y_N^\mu)
\end{eqnarray*}
where $Y_N^\mu$ is obtained by removing $\mu_i$ vector $x_{\bf i}$,
$i=1,\cdots,d$ to $Y_N$ and $\tilde \nabla^{\mu} =
\left (\nabla^{\mu_i}_{x_{\bf i}} \right )_{i=1,\cdots,n}$.
As both $a_{k-M.}(v^j)$ and $g_{k-M.}$ reproduce polynomials up to total
degree $N-1$, there exist a finite sequence $c_{k,p}$ such that:
\begin{eqnarray*}
\tilde \nabla^{\mu} ( \sum_{i \in \mathbb{Z}^{d}} (a_{k-Mi}(v^j)-g_{k-Mi}) v^j_i)
&=&\sum\limits_{p \in V(k)\bigcup \tilde V (k)}
\sum\limits_{|\nu|=|\mu|} c_{k,p}(\nu) \tilde \nabla^\nu v^{j}_p,
\end{eqnarray*}
where $\tilde V (k) = \{i,\|k-Mi\| \leq \tilde K\}$, where $g_{k-Mi} = 0$
if $\|k-Mi\| > \tilde K$.
We finally deduce that:
\begin{eqnarray*}
\tilde D_{M^{-j-1}}^{\mu} (v_{j+1}(x) -v_j(x))=
\sum\limits_{k \in \Bbb Z^{d}} \sum\limits_{p \in V(k)\bigcup \tilde V(k)}
\sum\limits_{|\nu|=|\mu|}
c_{k,p}( \nu ) \tilde \nabla^{\nu} v_p^j \beta_0(M^{j+1}x-k,Y_N^\mu).
\end{eqnarray*}
From this, we conclude that:
\begin{eqnarray*}
\| \tilde D_{M^{-j-1}}^{\mu} (v_{j+1}(x) -v_j(x)) \|_{L^p(\mathbb{R}^{d})} \raisebox{-1ex}{$~\stackrel{\textstyle<}{\sim}~$}
\rho_{p,|\mu|}(S)^j m^{-\frac{j+1}{p}} \|\tilde \Delta^{|\mu|} v^{j}_0\|_
{(\ell^p(\mathbb{Z}^{d}))^{\tilde q_{|\mu|}}}.
\end{eqnarray*}
Now, consider a sufficiently differentiable function
$f$ and remark that $D_{M^{-j-1}x_{\bf 1}} f(x) = (D f)(x).M^{-j-1}x_{\bf 1}$,
where $Df$ is the differential of the function $f$.
We also note that $M^q = \lambda I$ which implies that
$\lambda = \sigma^q$ and we then put $j+1 = q \times\lfloor \frac{j+1}{q}
\rfloor + r$ with $r < q$ and where $\lfloor . \rfloor$ denotes the integer part. From
this we may write:
$$
D_{M^{-j-1}x_{\bf 1}} f(x) = \sigma^{-q \lfloor \frac{j+1}{q} \rfloor}
(D f)(x).M^{-r}x_{\bf 1}
$$
and then
$$
D_{M^{-j-1}x_{\bf 1}} f(x) \sim \sigma^{-q\lfloor \frac{j+1}{q}
\rfloor}
(D f)(x).x_{r_j}
$$
where $r_j$ depends on $j$. Making the same reasoning for any order $\mu$ of
differentiation and any direction $x_{\bf i}$, we get, in $L^p$:
$$
\|(\tilde D_{M^{-j-1}}^{\mu} f)(x)\|_{L^p(\mathbb{R}^{d})} \sim \sigma^{-q|\mu|
\lfloor \frac{j+1}{q} \rfloor }
\| (\tilde D^{\mu} f)(x)\|_{L^p(\mathbb{R}^{d})}.
$$
We may thus conclude that
\begin{eqnarray*}
\|\tilde D^{\mu} (v_{j+1}(x)-v_j(x)) \|_{L^p(\mathbb{R}^{d})} &\sim&
\|\tilde D_{M^{-j-1}}^{\mu} (v_{j+1}(x)-v_j(x)) \|_{L^p(\mathbb{R}^{d})}
\sigma^{q|\mu|
(\times\lfloor \frac{j+1}{q} \rfloor)|}\\
&\raisebox{-1ex}{$~\stackrel{\textstyle<}{\sim}~$}&
\rho_{p,|\mu|}(S)^j m^{-\frac{j+1}{p}} \sigma^{q|\mu|
\lfloor \frac{j+1}{q} \rfloor} \|\tilde \Delta^{|\mu|} v^0\|_
{(\ell^p(\mathbb{Z}^{d}))^{\tilde q_{|\mu|}}}.
\end{eqnarray*}
To state the above result, we have used the fact that the joint spectral radius is
independent of the directions used for its computation.
Since we have the hypothesis that $\rho_{p,|\mu|}(S)
\leq m^{\frac{1}{p}-\frac{s}{d}}$ for $s > |\mu|$, we get that
\begin{eqnarray*}
\|\tilde D^{\mu} (v_{j+1}(x)-v_j(x)) \|_{L^p(\mathbb{R}^{d})} \raisebox{-1ex}{$~\stackrel{\textstyle<}{\sim}~$}
\sigma^{(|\mu|-s)j} \|\tilde \Delta^{|\mu|} v^0\|_
{(\ell^p(\mathbb{Z}^{d}))^{\tilde q_{|\mu|}}},
\end{eqnarray*}
which tends to zero with $j$, and thus the limit function is in $W_N^p(\Bbb R^{d})$
$\blacksquare$.\\
A comparison between Theorem \ref{segaln} and \ref{segaln1} shows
that when the subdivision scheme reproduces exactly polynomials, which is the
case of interpolatory subdivision schemes, the convergence is ensured provided
$\Phi_0$ also exactly reproduces polynomials. When the subdivision scheme only
reproduces polynomial the convergence is ensured provided that $\Phi_0$ is a
box spline. Note also that the condition on the joint spectral radius is the
same. We are currently investigating illustrative examples which involve the
adaptation of the local averaging subdivision scheme proposed in \cite{Co} to
our non-separable context.
\section{Conclusion}
We have addressed the issue of the definition of nonlinear
subdivision schemes associated to isotropic dilation matrix $M$.
After the definition of the convergence concept of such operators,
we have studied the convergence of these subdivision schemes in $L^p$ and in
Sobolev spaces.
Based on the study of the joint spectral radius of these operators,
we have exhibited sufficient conditions for the convergence of the proposed
subdivision schemes. This study has also brought into light the
importance of an appropriate choice of $\Phi_0$ to define the limit
function. In that context, box splines functions have shown to be a very
interesting tool.
|
\section{Introduction}\label{Intro}
The exponential stability plays a central role in the theory of
asymptotic behaviors for dynamical systems. In this paper we
consider the more general concepts of nonuniform exponential
stability for skew-evolution semiflows on Banach spaces. These
seem to be more appropriate for the study of evolution equations
in the nonuniform case, because of the fact that they depend on
three variables, contrary to a skew-product semiflow or an
evolution operator, which depend only on two, and, hence, the
study of asymptotic behaviors for skew-evolution semiflows in the
nonuniform setting arises as natural, relative to the third
variable.
Our main objectives are to establish relations between these
concepts and to give some integral characterizations for them. We
also remark that we use the concept of nonuniform exponential
stability, given and studied in the papers of L. Barreira and C.
Valls, as for example \cite{BaVa_LNM}, \cite{BaVa_NA09} or
\cite{BaVa_NA10}, and which we call \textit{"Barreira-Valls
exponential stability"}.
The paper presents some generalizations for the results obtained
in the uniform case in our paper \cite{StMe_NA}.
We remark that Theorems \ref{Datko} and \ref{Datko_BV} are
generalizations of Datko type for the nonuniform exponential
stability in the case of skew-evolution semiflows. The uniform
exponential stability was characterized by R. Datko in
\cite{Da_JMA}. The particular case of evolution operators was
considered by C. Bu\c{s}e in \cite{Bu_RSMUPT} and by M. Megan,
A.L. Sasu and B. Sasu in \cite{MeSaSa_MIR}. Theorem \ref{Rolewicz}
is the nonuniform variant for skew-evolution semiflows of the
known result of S. Rolewicz in \cite{Ro_JMAA}. Theorem
\ref{Barbashin} is a generalization of a result proved by E.A.
Barbashin in \cite{Ba_NAU}. A similar result was obtained
Bu\c{s}e, M. Megan, M. Prajea and P. Preda for the uniform
exponential stability in \cite{BuMePrPr_IEOT}.
Some illustrating examples clarify the connections between the
stability concepts considered in this paper.
\section{Skew-evolution semiflows}\label{Def_ses}
Let $( X,d)$ be a metric space, $ V$ a Banach space and $ V^{*}$
its topological dual. Let $\mathcal{B}( V)$ be the space of all
$V$-valued bounded operators defined on $ V$. The norm of vectors
on $ V$ and on $ V^{*}$ and of operators on $\mathcal{B}( V)$ is
denoted by $\left\Vert \cdot \right\Vert$. Let us consider $Y=
X\times V$ and $T=\left\{ (t,t_{0})\in \mathbb{R}_{+}^{2}:t\geq
t_{0}\right\}$. $I$ is the identity operator.
\begin{definition}
A mapping $\varphi: T\times X\rightarrow X$ is called
\emph{evolution semiflow} on $ X$ if the following
propositionerties are satisfied:
$(es_{1})$ $\varphi(t,t,x)=x, \ \forall (t,x)\in
\mathbb{R}_{+}\times X$;
$(es_{2})$ $\varphi(t,s,\varphi(s,t_{0},x))=\varphi(t,t_{0},x), \
\forall (t,s),(s,t_{0})\in T, \ \forall x\in
X$.
\end{definition}
\begin{definition}
A mapping $\Phi: T\times X\rightarrow \mathcal{B}( V)$ is called
\emph{evolution cocycle} over an evolution semiflow $\varphi$ if
it satisfies following propositionerties:
$(ec_{1})$ $\Phi(t,t,x)=I, \ \forall t\geq0,\ \forall x\in
X$;
$(ec_{2})$
$\Phi(t,s,\varphi(s,t_{0},x))\Phi(s,t_{0},x)=\Phi(t,t_{0},x),
\forall (t,s), (s,t_{0})\in T,\forall x\in X$.
\end{definition}
If $\Phi$ is an evolution cocycle over an evolution semiflow
$\varphi$, then the mapping
\begin{equation}
C: T\times Y\rightarrow Y, \
C(t,s,x,v)=(\varphi(t,s,x),\Phi(t,s,x)v)
\end{equation}
is called \emph{skew-evolution semiflow} on $ Y$.
\begin{remark}\rm
The concept of skew-evolution semiflow generalizes the notion of
skew-product semiflow, considered and studied by M. Megan, A.L.
Sasu and B. Sasu in \cite{MeSaSa_BBMS} and \cite{MeSaSa_MB}, where
the mappings $\varphi$ and $\Phi$ do not depend on the variables
$t$ and $x$.
\end{remark}
\begin{example}\rm
Let $ X=\mathbb{R}_{+}$. The mapping $\varphi:
T\times\mathbb{R}_{+} \rightarrow \mathbb{R}_{+}, \
\varphi(t,s,x)=t-s+x$ is an evolution semiflow on
$\mathbb{R}_{+}$. For every evolution operator $E: T\rightarrow
\mathcal{B}(V)$ (i.e. $E(t,t)=I$, $ \forall t\in \mathbb{R}_{+}$
and $E(t,s)E(s,t_{0})=E(t,t_{0})$, $\forall (t,s),(s,t_{0})\in T$)
we obtain that $\Phi_{E}: T\times \mathbb{R}_{+}\rightarrow
\mathcal{B}(V), \ \Phi_{E}(t,s,x)=E(t-s+x,x)$ \noindent is an
evolution cocycle on $V$ over the evolution semiflow $\varphi$.
Hence, an evolution operator on $V$ is generating a skew-evolution
semiflow on $Y$.
\end{example}
\begin{example}\rm\label{shift}
If $C=(\varphi, \Phi)$ denotes a skew-evolution semiflow and
$\alpha \in \mathbb{R}$ a parameter, then $C_{\alpha}=(\varphi,
\Phi_{\alpha})$, where
\begin{equation}
\Phi_{\alpha}: T\times X\rightarrow \mathcal{B}(V), \
\Phi_{\alpha}(t,t_{0},x)=e^{\alpha(t-t_{0})}\Phi(t,t_{0},x),
\end{equation}
is also a skew-evolution semiflow, being the
\emph{$\alpha$-shifted skew-evolution semiflow}.
\end{example}
Other examples of skew-evolution semiflows are given in
\cite{StMe_NA}.
\section{Nonuniform exponential stability}\label{Def_stab}
In this section we define five concepts of exponential stability
for skew-evolution semiflows. For each, an equivalent definition
is given. Also, we will establish some connections between these
concepts and we will emphasize that they are not equivalent.
We will begin by considering the notion of uniform exponential
stability for skew-evolution semiflows, as given in \cite{StMe_NA}
and which was characterized for evolution operators in
\cite{MeSt_IEOT}.
\begin{definition} \label{ues}
A skew-evolution semiflow $C =(\varphi,\Phi)$ is \emph{uniformly
exponentially stable} $(u.e.s.)$ if there exist some constants
$N\geq 1$ and $\alpha>0$ such that, for all $(t,s),(s,t_{0})\in
T$, following relation holds:
\begin{equation}
\left\Vert \Phi(t,t_{0},x)v\right\Vert \leq
Ne^{-(t-s)\alpha}\left\Vert \Phi(s,t_{0},x)v\right\Vert, \ \forall
\ (x,v)\in Y.
\end{equation}%
\end{definition}
An equivalent definition is given by
\begin{remark}\rm
The skew-evolution semiflow $C =(\varphi,\Phi)$ is uniformly
exponentially stable iff there exist $N\geq 1$ and $\alpha>0$ such
that, for all $(t,s)\in T$, the relation holds:
\begin{equation}
\left\Vert \Phi(t,s,x)v\right\Vert \leq Ne^{-
(t-s)\alpha}\left\Vert v\right\Vert, \ \forall \ (x,v)\in Y.
\end{equation}%
\end{remark}
The nonuniform exponential stability is defined by
\begin{definition} \label{es}
A skew-evolution semiflow $C =(\varphi,\Phi)$ is
\emph{exponentially stable} $(e.s.)$ if there exist a mapping
$N:\mathbb{R}_{+}\rightarrow[1,\infty)$ and a constant $\alpha>0$
such that, for all $(t,s)\in T$, following relation takes place:
\begin{equation}
\left\Vert \Phi(t,t_{0},x)v\right\Vert \leq N(s)e^{-\alpha
t}\left\Vert \Phi(s,t_{0},x)v\right\Vert,\ \forall (x,v)\in Y.
\end{equation}%
\end{definition}
Instead of the previous definition we have
\begin{remark}\rm
The skew-evolution semiflow $C =(\varphi,\Phi)$ is exponentially
stable iff there exist $N\geq 1$ and $\alpha>0$ such that, for all
$(t,s)\in T$, we have:
\begin{equation}
\left\Vert \Phi(t,s,x)v\right\Vert \leq N(s)e^{-\alpha
t}\left\Vert v\right\Vert,\ \forall \ (x,v)\in Y.
\end{equation}%
\end{remark}
A concept of nonuniform exponential stability for evolution
equations is given by L. Barreira and C. Valls in \cite{BaVa_LNM},
which we will generalize for skew-evolution semiflows. In what
follows, allow us to name this asymptotic propositionerty
\emph{"Barreira-Valls exponential stability".}
\begin{definition}\label{BVes}
A skew-evolution semiflow $C =(\varphi,\Phi)$ is
\emph{Barreira-Valls exponentially stable} $(BV.e.s.)$ if there
exist some constants $N\geq 1$, $\alpha>0$ and $\beta\geq\alpha$
such that, for all $(t,s),(s,t_{0})\in T$, the relation holds:
\begin{equation}
\left\Vert \Phi(t,t_{0},x)v\right\Vert \leq Ne^{-\alpha t}e^{\beta
s}\left\Vert \Phi(s,t_{0},x)v\right\Vert, \ \forall \ (x,v)\in Y.
\end{equation}%
\end{definition}
We also have, as an equivalent definition, the next
\begin{remark}\rm
A skew-evolution semiflow $C =(\varphi,\Phi)$ is Barreira-Valls
exponentially stable iff there some constants $N\geq 1$,
$\alpha>0$ and $\beta\geq\alpha$ such that, for all $(t,s)\in T$,
following relation is verified:
\begin{equation}
\left\Vert \Phi(t,s,x)v\right\Vert \leq Ne^{-\alpha t}e^{\beta
s}\left\Vert v\right\Vert, \ \forall \ (x,v)\in Y.
\end{equation}%
\end{remark}
The asymptotic propositionerty of nonuniform stability is
considered in
\begin{definition} \label{s}
A skew-evolution semiflow $C =(\varphi,\Phi)$ is \emph{stable}
$(s.)$ if there exists a mapping
$N:\mathbb{R}_{+}\rightarrow[1,\infty)$ such that, for all
$(t,s),(s,t_{0})\in T$, the relation is true:
\begin{equation}
\left\Vert \Phi(t,t_{0},x)v\right\Vert \leq N(s)\left\Vert
\Phi(s,t_{0},x)v\right\Vert, \ \forall \ (x,v)\in Y.
\end{equation}%
\end{definition}
We also have
\begin{remark}\rm
The skew-evolution semiflow $C =(\varphi,\Phi)$ is stable iff
there exists a mapping $N:\mathbb{R}_{+}\rightarrow[1,\infty)$
such that, for all $(t,s)\in T$, the relation is verified:
\begin{equation}
\left\Vert \Phi(t,s,x)v\right\Vert \leq N(s)\left\Vert
v\right\Vert, \ \forall \ (x,v)\in Y.
\end{equation}%
\end{remark}
Let us remind the propositionerty of exponential growth for
skew-evolution semiflows, given by
\begin{definition} \label{eg}
A skew-evolution semiflow $C =(\varphi,\Phi)$ has
\emph{exponential growth} $(e.g.)$ if there exist two
nondecreasing mappings
$M,\omega:\mathbb{R}_{+}\rightarrow[1,\infty)$ such that, for all
$(t,s),(s,t_{0})\in T$, we have:
\begin{equation}
\left\Vert \Phi(t,t_{0},x)v\right\Vert \leq
M(s)e^{\omega(t-s)}\left\Vert \Phi(s,t_{0},x)v\right\Vert,\
\forall \ (x,v)\in Y.
\end{equation}%
\end{definition}
Similarly, we have
\begin{remark}\rm
The skew-evolution semiflow $C =(\varphi,\Phi)$ has exponential
growth iff there exist two nondecreasing mappings
$M,\omega:\mathbb{R}_{+}\rightarrow[1,\infty)$ such that, for all
$(t,s)\in T$, the relation holds:
\begin{equation}
\left\Vert \Phi(t,s,x)v\right\Vert \leq
M(s)e^{\omega(t-s)}\left\Vert v\right\Vert,\ \forall \ (x,v)\in Y.
\end{equation}%
\end{remark}
We obtain following relations concerning the previously defined
asymptotic propositionerties for skew-evolution semiflows.
\begin{remark}\rm From the previous definitions it follows that:
\begin{equation}
(u.e.s.)\Longrightarrow (BV.e.s.)\Longrightarrow
(e.s.)\Longrightarrow(s.)\Longrightarrow (e.g.)
\end{equation}
\end{remark}
The reciprocal statements are not true, as shown in what follows.
\vspace{3mm}
The next example emphasizes a skew-evolution semiflow which is
Barreira-Valls exponentially stable but is not uniformly
exponentially stable.
\begin{example}\rm
Let $ X=\mathbb{R}_{+}$ and $V=\mathbb{R}$. The mapping $\varphi:
T\times\mathbb{R}_{+} \rightarrow \mathbb{R}_{+}$, where $
\varphi(t,s,x)=t-s+x$ \noindent is an evolution semiflow on
$\mathbb{R}_{+}$.
We will consider the function
$u:\mathbb{R}_{+}\rightarrow\mathbb{R}$, given by
$u(t)=e^{2t-t\sin t}$. We define
$$\Phi_{u}(t,s,x)v=\frac{u(s)}{u(t)}v, \ \textrm{with} \ (t,s)\in T,\ (x,v)\in Y.$$
As we have
$$\left| \Phi_{u}(t,s,x)v\right| \leq |v|\cdot e^{t\sin t-s\sin s+2s-2t}\leq |v|e^{3s-2t}=e^{-2t}e^{3t}|v|,$$
for all $(t,s,x,v)\in T\times Y$. It follows that
$C_{u}=(\varphi,\Phi_{u})$ is Barreira-Valls exponentially stable.
Let us suppose now that the skew-evolution semiflow
$C_{u}=(\varphi,\Phi_{u})$ is uniformly exponentially stable.
According to Definition \ref{ues}, there exist $N\geq 1$,
$\alpha>0$ and $t_{1}>0$ such that
$$e^{t\sin t-s\sin s+2s-2t}\leq Ne^{\alpha(s-t)}, \ \forall t\geq s\geq t_{1}.$$
If we consider $t=2n\pi+\frac{\pi}{2}$ and $s=2n\pi$, we have that
$$\exp\left(2n\pi-\frac{3\pi}{2}\right)\leq N\exp\left(-\frac{\pi}{2}\right),$$
which, for $n\rightarrow \infty$, leads to a contradiction, which
proves that $C_{u}$ is not uniformly exponentially stable.
\end{example}
The following example presents a skew-evolution semiflow which is
exponentially stable but not Barreira-Valls exponentially stable.
\begin{example}\rm
Let $ X=\mathbb{R}_{+}$. The mapping $\varphi:
T\times\mathbb{R}_{+} \rightarrow \mathbb{R}_{+}, \
\varphi(t,s,x)=x$ \noindent is an evolution semiflow on
$\mathbb{R}_{+}$.
Let us consider a continuous function
$u:\mathbb{R}_{+}\rightarrow[1,\infty)$ with
$$u(n)=n\cdot 2^{2n}\ \textrm{and} \ u\left(n+\frac{1}{2^{2n}}\right)=1.$$
We define
$$\Phi_{u}(t,s,x)v=\frac{u(s)e^{s}}{u(t)e^{t}}v, \ \textrm{where} \ (t,s)\in T,\ (x,v)\in Y.$$
As following relation
$$\left\Vert \Phi_{u}(t,s,x)v\right\Vert \leq u(s)e^{s}e^{-t}\left\Vert v\right\Vert$$
holds for all $(t,s,x,v)\in T\times Y$, it results that the
skew-evolution semiflow $C_{u}=(\varphi,\Phi_{u})$ is
exponentially stable.
Let us now suppose that the skew-evolution semiflow
$C_{u}=(\varphi,\Phi_{u})$ is Barreira-Valls exponentially stable.
Then, according to Definition \ref{BVes}, there exist $N\geq 1$,
$\alpha>0$, $\beta>0$ and $t_{1}>0$ such that
$$\frac{u(s)e^{s}}{u(t)e^{t}}\leq Ne^{-\alpha t}e^{\beta s}, \ \forall t\geq s\geq t_{1}.$$
For $t=n+\frac{1}{2^{2n}}$ and $s=n$ it follows that
$$e^{n\left(2^{2n}+1\right)}\leq Ne^{n+\frac{1}{2^{2n}}}e^{-\alpha\left(n+\frac{1}{2^{2n}}\right)}e^{\beta n},$$
which is equivalent with $$e^{n\left(2^{2n}-\beta\right)}\leq
Ne^{\frac{1}{2^{2n}}-\alpha\left(n+\frac{1}{2^{2n}}\right)}.$$ For
$n\rightarrow \infty$, a contradiction is obtained, which proves
that $C_{u}$ is not Barreira-Valls exponentially stable.
\end{example}
There exist skew-evolution semiflows that are stable but not
exponentially stable, as results from the following
\begin{example}\rm
Let us consider $ X=\mathbb{R}_{+}$, $V=\mathbb{R}$ and
$$u:\mathbb{R}_{+}\rightarrow[1,\infty)\ \textrm{with the propositionerty}
\ \underset{t\rightarrow\infty}\lim\frac{u(t)}{e^{t}}=0.$$ The
mapping
\[
\Phi_{u}: T\times \mathbb{R}_{+} \rightarrow
\mathcal{B}(\mathbb{R}), \ \Phi_{u}(t,s,x)v=\frac{u(s)}{u(t)}v
\]
is an evolution cocycle. As $|\Phi(t,s,x)v|\leq u(s)|v|$, $\forall
(t,s,x,v)\in T\times Y$, it follows that $C_{u}=(\varphi,
\Phi_{u})$ is a stable skew-evolution semiflow, for every
evolution semiflow $\varphi$ on $\mathbb{R}_{+}$.
On the other hand, if we suppose that $C_{u}$ is exponentially
stable, according to Definition \ref{es}, there exist a mapping
$N:\mathbb{R}_{+}\rightarrow[1,\infty)$ and a constant $\alpha>0$
such that, for all $(t,s),(s,t_{0})\in T$, we have
$$
\left\Vert \Phi(t,t_{0},x)v\right\Vert \leq N(s)e^{-\alpha
t}\left\Vert \Phi(s,t_{0},x)v\right\Vert,\ \forall \ (x,v)\in Y.
$$
It follows that
\[
\frac{u(s)}{N(s)}\leq \frac{u(t)}{e^{\alpha t}}.
\]
For $t\rightarrow\infty$ we obtain a contradiction, and, hence,
$C_{u}$ is not exponentially stable.
\end{example}
Following example gives a skew-evolution semiflow that has
exponential growth but is not stable.
\begin{example}\rm
We consider $ X=\mathbb{R}_{+}$, $V=\mathbb{R}$ and
$$u:\mathbb{R}_{+}\rightarrow[1,\infty)\ \textrm{with the propositionerty}
\ \underset{t\rightarrow\infty}\lim\frac{e^{t}}{u(t)}=\infty.$$
The mapping
\[
\Phi_{u}: T\times \mathbb{R}_{+} \rightarrow
\mathcal{B}(\mathbb{R}), \
\Phi_{u}(t,s,x)v=\frac{u(s)e^{t}}{u(t)e^{s}}v
\]
is an evolution cocycle. We have $|\Phi(t,s,x)v|\leq
u(s)e^{t-s}|v|$, $\forall (t,s,x,v)\in T\times Y$. Hence,
$C_{u}=(\varphi, \Phi_{u})$ is a skew-evolution semiflow, over
every evolution semiflow $\varphi$, and has exponential growth.
Let us suppose that $C_{u}$ is stable. According to Definition
\ref{s}, there exists a mapping
$N:\mathbb{R}_{+}\rightarrow[1,\infty)$ such that $u(s)e^{t}\leq
N(s)u(t)e^{s}$, for all $(t,s)\in T$. If $t\rightarrow\infty$, a
contradiction is obtained. Hence, $C_{u}$ is not stable.
\end{example}
\section{Datko type theorems for the nonuniform exponential
stability}\label{Th_D_R}
A different type of stability for skew-evolution semiflows in the
nonuniform setting is presented in this section, as well a
particular class of skew-evolution semiflows, which allows
connections between various stability types.
\begin{definition}\label{is}
A skew-evolution semiflow $C=(\varphi,\Phi)$ is called
\emph{integrally stable} $(i.s.)$ if there exists a mapping
$D:\mathbb{R}_{+}\rightarrow\mathbb{R}_{+}^{*}$ such that:
\begin{equation}
\int^{\infty}_{s}\left\Vert \Phi(t,t_{0},x)v\right\Vert dt\leq
D(s)\left\Vert \Phi(s,t_{0},x)v\right\Vert,
\end{equation}
for all $(s,t_{0})\in T$ and all $(x,v)\in Y$.
\end{definition}
An equivalent definition can be considered the next
\begin{remark}\rm\label{rem_is}
A skew-evolution semiflow $C=(\varphi,\Phi)$ is integrally stable
iff there exists a mapping
$D:\mathbb{R}_{+}\rightarrow\mathbb{R}_{+}^{*}$ such that:
\begin{equation}
\int^{\infty}_{s}\left\Vert \Phi(t,s,x)v\right\Vert dt\leq
D(s)\left\Vert v\right\Vert,
\end{equation}
for all $s\in \mathbb{R}_{+}$ and all $(x,v)\in Y$.
\end{remark}
\begin{definition}
A skew-evolution semiflow $C=(\varphi,\Phi)$ has \emph{bounded
exponential growth} if $C$ has exponential growth and function $M$
from Definition \ref{eg} is bounded.
\end{definition}
\begin{proposition}\label{proposition_stab}
An integrally stable skew-evolution semiflow $C=(\varphi,\Phi)$
with bounded exponential growth is stable.
\end{proposition}
\begin{proof}
Let us denote $M=\underset{t\geq 0}\sup M(t)$ and
$c=\int_{0}^{1}e^{-\omega(t)}$, where functions $M$ and $\omega$
are given by Definition \ref{eg}.
We observe that for $t\geq s+1$ we have
$$c\leq \int_{0}^{t-s}e^{-\omega(r)dr}=\int_{s}^{t}e^{-\omega(t-\tau)}d\tau$$
and, further,
$$c|<v^{*},\Phi(t,s,x)v>|\leq \int_{s}^{t}e^{-\omega(t-\tau)d\tau}\left\Vert \Phi(t,\tau,\varphi(\tau,s,x))^{*}v^{*}\right\Vert
\left\Vert \Phi(\tau,s,x)v\right\Vert d\tau\leq$$
$$\leq M\int_{s}^{t}\left\Vert \Phi(\tau,s,x)v\right\Vert
d\tau\leq MD(s)\left\Vert v\right\Vert\left\Vert
v\right\Vert^{*},$$ for all $(t,t_{0})\in T$, all $(x,v)\in Y$ and
all $v^{*}\in V^{*}$, function $D$ being given by Remark
\ref{rem_is}.
By taking supremum relative to $\left\Vert v\right\Vert^{*}\leq
1$, we obtain $$\left\Vert \Phi(t,s,x)v\right\Vert\leq
\frac{MD(s)}{c}, \ \forall t\geq s+1, \ (x,v)\in Y.$$ Finally, it
follows that $$\left\Vert \Phi(t,s,x)v\right\Vert \leq
N(s)\left\Vert v\right\Vert, \ \forall (t,s)\in T, \ \forall
(x,v)\in Y,$$ where we have denoted
$$N(s)=M\left[\frac{D(s)}{c}+e^{\omega(s)}\right],$$ and which proves that $C$ is
stable.
\end{proof}
\begin{definition}\label{eis}
A skew-evolution semiflow $C=(\varphi,\Phi)$ is said to be
\emph{exponentially integrally stable} $(e.i.s.)$ if there exist a
mapping $D:\mathbb{R}_{+}\rightarrow\mathbb{R}_{+}^{*}$ and a
constant $d>0$ such that following relation:
\begin{equation}
\int^{\infty}_{s}e^{(t-s)d}\left\Vert \Phi(t,s,x)v\right\Vert
dt\leq D(s)\left\Vert \Phi(s,t_{0},x)v\right\Vert,
\end{equation}
holds for all $(s,t_{0})\in T$ and all $(x,v)\in Y$.
\end{definition}
We also have
\begin{remark}\rm
A skew-evolution semiflow $C=(\varphi,\Phi)$ is exponentially
integrally stable iff there exist a mapping
$D:\mathbb{R}_{+}\rightarrow\mathbb{R}_{+}^{*}$ and a constant
$d>0$ such that:
\begin{equation}
\int^{\infty}_{s}e^{(t-s)d}\left\Vert \Phi(t,s,x)v\right\Vert
dt\leq D(s)\left\Vert v\right\Vert,
\end{equation}
for all $(t,s)\in T$ and all $(x,v)\in Y$.
\end{remark}
\begin{remark}\rm As a connection between the presented asymptotic propositionerties, we
have:
\begin{equation}
(e.i.s.)\Longrightarrow (i.s.)
\end{equation}
\end{remark}
In what follows, we will emphasize some characterizations of the
various types of nonuniform stability considered in Section
\ref{Def_stab}. We will begin this section by considering a
particular class of skew-evolution semiflows, given in
\begin{definition}
A skew-evolution semiflow $C=(\varphi,\Phi)$ is said to be
\emph{strongly measurable} $(s.m.)$ if for all
$(t_{0},x,v)\in\mathbb{R}_{+}\times Y$ the mapping
$s\mapsto\left\Vert\Phi(s,t_{0},x)v\right\Vert$ is measurable on
$[t_{0},\infty)$.
\end{definition}
\begin{theorem}\label{Datko}
A strongly measurable skew-evolution semiflow $C=(\varphi,\Phi)$
with bounded exponential growth is exponentially stable if and
only if it is exponentially integrally stable.
\end{theorem}
\begin{proof}
\emph{Necessity.} It is a simple verification for
$$d=\frac{\alpha}{2} \ \textrm{and} \ D(t)=\frac{N(t)}{\alpha}, \
t\geq 0.$$ \emph{Sufficiency.} If $C=(\varphi,\Phi)$ is
exponentially integrally stable, then there exists a constant
$d>0$ such that the $d$-shifted skew-evolution semiflow $C_{d}$,
given as in Example \ref{shift}, is integrally stable with bounded
exponential growth.
According to Proposition \ref{proposition_stab}, it follows that
$C_{d}$ is stable, which assures the existence of a mapping
$N:\mathbb{R}_{+}\rightarrow[1,\infty)$ with
$$
\left\Vert \Phi(t,s,x)v\right\Vert \leq N(s)e^{-(t-s)d}\left\Vert
v\right\Vert,\ \forall (t,s)\in T, \ \forall (x,v)\in Y,$$ which
proves that $C$ is exponentially stable.
\end{proof}
\begin{remark}\rm
Theorem \ref{Datko} can be viewed as a Datko type theorem for the
propositionerty of nonuniform exponential stability for
skew-evolution semiflows. The case of uniform stability was
considered in \cite{StMe_NA}. For the particular case of evolution
operators, this result was proved by R. Datko in \cite{Da_JMA} in
the uniform setting and by C. Bu\c{s}e in \cite{Bu_RSMUPT} for the
nonuniform case.
\end{remark}
Let us denote by $\mathcal{F}$ the set of all nondecreasing
functions $F:\mathbb{R}_{+}\rightarrow\mathbb{R}_{+}$ with the
propositionerties $F(0)=0$ and $F(t)>0$, $\forall t>0$.
\begin{remark}\rm
Analogously to the uniform case studied in \cite{StMe_NA}, the
proof of Theorem \ref{Datko} can be easily adapted to prove a
variant of Rolewicz type for the propositionerty of exponential
stability of skew-evolution semiflows in the nonuniform setting,
as given by
\end{remark}
\begin{theorem}\label{Rolewicz}
Let $C=(\varphi,\Phi)$ be a strongly measurable skew-evolution
semiflow with exponential growth. Then $C$ is exponentially stable
if and only if there exist two mappings
$F,R:\mathbb{R}_{+}\rightarrow\mathbb{R}_{+}$ and a constant $d>0$
with $F\in\mathcal{F}$ and:
\begin{equation}
\int^{\infty}_{s}F\left(e^{(t-s)d}\left\Vert
\Phi(t,s,x)v\right\Vert dt\right)\leq R(s)F\left(\left\Vert
v\right\Vert\right),
\end{equation}
for all $(s,x,v)\in\mathbb{R}_{+}\times Y$.
\end{theorem}
\begin{remark}\rm
For the particular case of evolution operators, Theorem
\ref{Rolewicz} was proved by S. Rolewicz in \cite{Ro_JMAA} for the
propositionerty of uniform exponential stability.
\end{remark}
\begin{remark}\rm
By means of the methods used in the proofs of Proposition
\ref{proposition_stab} and of Theorem \ref{Datko}, one can obtain
a Datko type theorem for the exponential stability of
Barreira-Valls type, in the case of skew-evolution semiflows in
the nonuniform setting, as shown by
\end{remark}
\begin{theorem}\label{Datko_BV}
Let $C=(\varphi,\Phi)$ be a strongly measurable skew-evolution
semiflow with exponential growth. Then $C$ is Barreira-Valls
exponentially stable if and only if there exist some constants
$N\geq 1$, $a>0$ and $b\geq a$ such that:
\begin{equation}
\int^{\infty}_{s}e^{at}\left\Vert \Phi(t,s,x)v\right\Vert dt\leq
Ne^{bs}\left\Vert v\right\Vert,
\end{equation}
for all $(s,x,v)\in\mathbb{R}_{+}\times Y$.
\end{theorem}
\begin{remark}\rm
Analogously, a Rolewicz type theorem can be given for the
propositionerty of Barreira-Valls exponential stability, in the
case of skew-evolution semiflows.
\end{remark}
\section{A Barbashin type theorem for the nonuniform exponential
stability}\label{Th_B}
In this section let us consider a particular class of
skew-evolution semiflows, given by
\begin{definition}
A skew-evolution semiflow $C=(\varphi,\Phi)$ is said to be
\emph{$*$-strongly measurable} $(\ast-s.m.)$ if for every
$(t,t_{0},x,v^{*})\in T\times X\times V^{*}$ the mapping defined
by
$s\mapsto\left\Vert\Phi(t,s,\varphi(s,t_{0},x))^{*}v^{*}\right\Vert$
is measurable on $[t_{0},t]$.
\end{definition}
The main result of this section is
\begin{theorem}\label{Barbashin}
Let $C=(\varphi,\Phi)$ be a $*$-strongly measurable skew-evolution
semiflow with exponential growth. If there exist a constant $b>0$
and a mapping $B:\mathbb{R}_{+}\rightarrow [1,\infty)$ such that:
\begin{equation}
\int^{t}_{s}e^{(t-\tau)b}\left\Vert
\Phi(t,\tau,\varphi(\tau,s,x))^{*}v^{*}\right\Vert d\tau\leq
B(t)\left\Vert v^{*}\right\Vert,
\end{equation}
for all $(t,s)\in T$ and all $(x,v^{*})\in X\times V^{*}$, then
$C$ is exponentially stable.
\end{theorem}
\begin{proof}
For $t\geq s\geq 0$ we will denote
$$f_{s}(t)=M(s)B(t)e^{tb}e^{\omega(t)} \ \textrm{and} \ K(s)=\int_{0}^{1}\frac{du}{f_{s}(u)},$$ where the functions $M$ and
$\omega$ are given by Definition \ref{eg}.
We remark that, if $t\geq s+1$, then $$K(s)\leq
\int^{t-s}_{0}\frac{du}{f_{s}(u)}=\int^{t}_{s}\frac{d\tau}{f_{s}(\tau-s)}.$$
It follows that $$B(t)e^{(t-s)b}K(s)|<v^{*},\Phi(t,s,x)v>|\leq$$
$$\leq
\int^{t}_{s}\frac{e^{(t-s)b}|<\Phi(t,\tau,\varphi(\tau,s,x))^{*}v^{*},
\Phi(\tau,s,x)v>|}{M(s)e^{(\tau-s)b}e^{\omega(\tau-s)}}d\tau\leq$$
$$\leq \int^{t}_{s}e^{(t-\tau)b}\left\Vert \Phi(t,\tau,\varphi(\tau,s,x))^{*}v^{*}\right\Vert
\left\Vert v\right\Vert d\tau \leq B(t)\left\Vert
v\right\Vert\left\Vert v^{*}\right\Vert,$$ which implies
$$\left\Vert \Phi(t,s,x)v\right\Vert\leq
\frac{e^{-(t-s)b}}{K(s)}\left\Vert v\right\Vert$$ for all $t \geq
s+1$ and all $(x,v)\in Y$.
Now, if we consider $t\in [s,s+1)$, we have $$\left\Vert
\Phi(t,s,x)v\right\Vert\leq M(s)e^{\omega(t-s)}\left\Vert
v\right\Vert\leq M(s)e^{\omega(1)}\left\Vert v\right\Vert\leq
M(s)e^{b+\omega(1)}e^{-b(t-s)}\left\Vert v\right\Vert.$$ Finally,
we obtain, $$\left\Vert \Phi(t,s,x)v\right\Vert\leq
N(s)e^{-(t-s)b}\left\Vert v\right\Vert,$$ for all $(t,s)\in T$ and
all $(x,v)\in X\times V$, where we have denoted
$$N(s)=M(s)e^{b+\omega(1)}+\frac{1}{K(s)},$$ and which proves the
exponential stability of the skew-evolution semiflow $C$.
\end{proof}
\begin{remark}\rm
Theorem \ref{Barbashin} is a generalization of a known result of
E.A. Barbashin emphasized in \cite{Ba_NAU}. A similar result was
obtained by C. Bu\c{s}e, M. Megan, M. Prajea and P. Preda for the
uniform exponential stability of evolution operators in
\cite{BuMePrPr_IEOT}.
\end{remark}
\begin{remark}\rm
Analogously as in the proof of Theorem \ref{Barbashin}, one can
prove a Barbashin type theorem for the propositionerty of
Barreira-Valls exponential stability, in the case of
skew-evolution semiflows.
\end{remark}
\textbf{Acknowledgments.} This work is financially supported from
the Exploratory Research Grant CNCSIS PN II ID 1080 No. 508/2009
of the Romanian Ministry of Education, Research and Innovation.
{\footnotesize
|
\section{\label{intro}Introduction}
In a $\Lambda$CDM cosmology galaxies acquire mass mostly through
minor merger events, where one galaxy has 0.3 times or less the
mass of its collision partner. Only the most luminous elliptical
galaxies experience a major merger event in their history
\citep{pef09}. The growth of large elliptical systems is
facilitated particularly well in the low-velocity environments of galaxy
groups where dynamical friction \citep{cha43,nus99} is very efficient.
This effect increases with the mass of the infalling galaxy and is
higher for lower velocities. In this way galaxies cool down into the
group or cluster core, losing their gas through ram-pressure stripping
along the way \citep{qmb00}. At the same time they undergo slower
morphological transformations \citep{pef09,sws09}, leading to the
formation of the red sequence in the inner region. The time
scale for dynamical friction depends on
the mass of the infalling galaxy and its distance from the core. For
the most massive galaxies ($\sim 10^{12}{\rm M_\odot}$) it is as short as a
few Gyrs \citep{nat08,bmq08}, implying that large elliptical galaxies
in groups can already accur at early times.
Several mechanisms for the formation of BCGs (the brightest cluster
galaxies) have been suggested, ranging from galactic cannibalism and
cooling flows to merger processes during cluster collapse
\citep[see][and references therein]{lbk07}. Elliptical galaxies
growing in this fashion should be located at the centre of the
gravitational potential, and their recession velocity should match the
mean of the radial velocities of the other cluster members for
virialised systems. Recently, \cite{sby10} have shown that in
$\sim40\%$ of all haloes of mass $\sim5\times10^{13}\,h_{100}\,{\rm M_\odot}$
the BCG is not the central galaxy, falsifying this paradigm. This was
also demonstrated for clusters with higher masses
\cite[][]{oeh01,lbk07}. Most of these analyses have in
common that the centre of the halo is identified by the distribution
centre of elliptical galaxies or, more rarely, by the X-ray centroid
or weak gravitational lensing. Either of these methods has advantages
and disadvantages, for instance, they can be hampered by small numbers
of galaxies, low X-ray S/N or projection effects. In this paper we are
in the lucky situation that a strongly lensed galaxy let us put tight
constraints on the dark matter halo centre, and in this way show
that the BCG is not located at the minimum of the
potential. We conclude that the BCG was formed outside the cluster in
a nearby group, which is now falling into the cluster.
\subsection{Fossil groups}
Contrary to the quick dynamical collapse of galaxy groups, the
cooling times for their X-ray haloes are comparable to
one or several Hubble times \citep{sar88}. This can lead to isolated
giant elliptical galaxies, embedded in X-ray haloes with luminosities
characteristic for entire galaxy groups. Such objects exist in
the Universe \citep{vnh99}, either isolated \citep{yft04} or
surrounded by groups of less luminous satellite galaxies
\citep{jph03,kpj06,bcr09}. One of the first systems has been reported
by \cite{paj94} coining the term `fossil group', and \cite{jph03}
have introduced general selection criteria. Accordingly, the galaxies
must be embedded in an extended X-ray halo with
$L_X>10^{42}\,h_{50}^{-2}\rm ~erg~s^{-1}$, integrated over the $0.5-2.0$ keV
range. In addition, the central elliptical galaxy must be
$\Delta m_{12}^{\rm min}\geq 2$ mag brighter in $R$-band than the
second brightest galaxy (independent of morphology) within half
the virial radius. This magnitude gap is motivated by the accretion of
$L_*$-galaxies in the inner volume, which are then absent in the
group's luminosity function. Current observational samples
\citep[e.g.][]{kpj07,bcr09,vbh10} are largely based on this
definition, and so are simulations \citep{bog08}.
The selection criteria by \cite{jph03} have been relaxed in the course
of systematic searches. \cite{sms07} have favoured a fixed
radius of $0.5\,h_{50}$ Mpc within which the magnitude
gap must hold, independent of the cluster's virial state. \cite{vbh10}
adopt $0.7\,r_{500}\sim0.4\,r_{\rm vir}$, with $r_{500}$ being
calculated from the group's X-ray luminosity. Similar relaxations have
been adopted for the magnitude gap. \cite{mmf06} and \cite{bcr09} show
that there is no sharp transition in the magnitude gap of galaxy
clusters, hence there is no physical motivation for a particular
numeric value. \cite{vbh10} and \cite{bcr09} favour smaller gap sizes
of $\Delta m_{12}^{\rm min}=1.7$ and 1.75 mag, respectively. According
to \cite{vbh10} the gap should not be too strict a requirement, as the
determination of the total magnitude of very extended galaxies is not
trivial.
As for the formation of the magnitude gap, \cite{bog08} have found in
simulations that it usually arises at redshifts $0<z<0.7$, after the
haloes assembled half of their final mass at $0.8<z<1.2$. This is
significantly earlier than the formation of normal groups
\citep{dkp07} and leads to increased NFW \citep{nfw97} concentration
parameters. Accordingly, the last major merger in the simulated fossil
groups took place more than 6 Gyrs ago for more than 50\% of the
galaxies. Most of the magnitude gaps are closed at later (current)
times when more infall of satellite galaxies occurs.
The exact formation process of fossils is not yet entirely understood.
For example, \cite{yft04} observe mass-to-light ratios as high as
$1000$, which are difficult to explain if these galaxies assembled
their mass only through dynamical friction. Another uncertainty lies
in the type of the galaxies from which the giant elliptical forms.
\cite{kpj06} argue that their disky isophotes indicate gas-rich
mergers, which would distinguish these galaxies from the BCGs in
normal clusters. These tend to show more boxy isophotes from gas-poor
mergers. However, \cite{bcr09} do not find a preference for either
disky or boxy shapes in their larger sample. They have argued
that fossils merely represent a transitional state in the
last stages of mass assembly than a class of their own.
While the formation process of fossils is still a matter of debate,
their occurrence is not. About $10\%-20\%$ of all X-ray luminous
groups and clusters have fossil character \citep[e.g.][]{jph03,bog08},
with typical masses of $1-10\times10^{13} {\rm M_\odot}$. However, they are
difficult to identify observationally. Only a few dozen systems
are known so far, mostly extracted from large-area
surveys such as SDSS \citep{sms07,bcr09} or the 400D cluster catalogue
\citep{vbh10}. The last authors discuss various difficulties in the
selection process, in particular completeness and
problems in the accurate determination of the magnitude of the
brightest galaxy. In general, the observationally determined
abundances of fossil groups agree with those predicted by
simulations. However, in terms of absolute numbers samples are
systematically incomplete since the second brightest galaxy can be at
a sufficiently large physical distance from the centre and still
appear projected onto the inner volume. Assuming that all galaxies
are within the virial radius and follow a radially symmetric
distribution, we estimate that $20\%$ $(25\%)$ of all fossils
are overlooked for $r_{\rm min}=0.4$ (0.5)
$r_{\rm vir}$ due to this effect.
Almost all of the few dozen fossil groups known were discovered and
analysed based upon comparatively shallow optical and/or X-ray survey
data. In general the observational data are poor compared to what are
available for normal clusters. Only a few fossils were investigated in
detail, such as ESO 3060170 \citep{sfv04}, RXJ1552.2+2013
\citep{mcs06}, RXJ1416.4+2315 \citep{jph03,cms06,kmp06}, CL0259+0013
\citep{vbh10} or UGC 842 \citep{lcm10}. For a comprehensive comparison
with normal groups and clusters a systematic deep survey of a larger
number is needed.
In this paper we present our analysis of J0454.0-0308 (hereafter:
J0454), a fossil group at $z=0.26$. It is projected
8\hbox{$^{\prime}\;$} south of the well-known cluster MS0451-0301
(hereafter: MS0451, $z=0.54$), thus a large amount of
archival data are available for our analysis. J0454 consists of at
least 60 galaxies and was identified by us in Subaru/Suprime-Cam
images. It is dominated by a giant elliptical galaxy (hereafter:
E0454), which strongly lenses a distant background source. We use
Subaru/Suprime-Cam and CFHT/MegaPrime for photometry, XMM-Newton
to study the intra-cluster gas and HST/ACS for the weak and strong
lensing analysis. The imaging data (see Fig. \ref{fields} for an
overview) are complemented by VLT and Keck spectroscopy.
\subsection{Terminology and assumptions}
In this work we present evidence that J0454 is composed of a poor
cluster and an infalling fossil group. We refer to the global
system as J0454, but also to the cluster without the fossil
group. The latter distinction is only made in Sect. 8 when
we discuss the results. E0454 is the brightest galaxy of the system.
\begin{figure}[t]
\includegraphics[width=1.0\hsize]{13810_f01.eps}
\caption{\label{fields}{Pointings of the imaging data sets.
The positions of the fossil group J0454 and the
background cluster MS0451 are shown as well.}}
\end{figure}
The paper is organised as follows: In Sect. 2 we describe the imaging
and spectroscopic data provided this was not done elsewhere. In
Sect. 3 we study foreground and background contamination and select a
red cluster sequence in colour-colour and colour-magnitude space. In
Sect. 4 we investigate the galactic content of the system, establish a
morphology-density relation and obtain the velocity dispersions of
early- and late-type galaxies. We use virial properties and the
size-richness relation for an estimate of $r_{200}$. In Sect. 5 we
present the X-ray results, followed by our weak and strong lensing
analysis in Sects. 6 and 7. We discuss our findings in Sect. 8 and
summarise in Sect. 9.
We assume a flat standard cosmology with $\Omega_m=0.27$,
$\Omega_\Lambda=0.73$ and $H_0=70\,h\rm ~km~s^{-1}\,{\rm Mpc}^{-1}$. On
occasion we refer to relations from the literature with
parameterisations $H_0=100\,h_{100}\rm ~km~s^{-1}\,{\rm Mpc}^{-1}$ or
$H_0=50\,h_{50}\rm ~km~s^{-1}\,{\rm Mpc}^{-1}$. To avoid confusion
we quote them as published originally, indexing $h$ accordingly. X-ray
luminosities are reported for the $0.5-2.0$ keV range, and optical
luminosities are given in solar units. The relation between angular
and physical scales at $z=0.26$ is $1^{\prime}=243\,h^{-1}$ kpc. All
numeric values quoted for physical distances in J0454 must be scaled
with $h^{-1}$. Magnitudes are reported for both the Johnson-Cousins and
the Sloan passbands and denoted with uppercase and lowercase
letters, respectively. All error bars represent the $1\sigma$
confidence level.
\section{Observations and data reduction}
\subsection{Subaru/Suprime-Cam and CFHT/Megaprime data reduction}
We serendipitously discovered J0454 in deep Subaru/Suprime-Cam
\citep{mks02} images of MS0451. The data were reduced with
THELI\footnote{Available at
http://www.astro.uni-bonn.de/$\sim$mischa/theli.html} \citep{esd05},
our pipeline for the reduction of wide-field optical and near-infrared
images. In the following we summarise those aspects where our
reduction scheme deviated significantly from the standard approach.
Images were taken in nine different nights during six periods
between 2001-01-22 and 2006-12-21 (PIs: H. Ebeling, N. Yasuda,
G. Kosugi). Suprime-Cam consists of 10 CCDs, covering
$34^{\prime}\times27^{\prime}$ with 0\myarcsec202 per pixel. In 2001
Suprime-Cam had one broken CCD and individual gain settings. The
defect CCD and three others were replaced, and the gains were
homogenised and refined once more another year later. We brought all
chips to the same gain and then performed the standard pre-processing
including debiasing, flatfielding, superflatting, defringing, and sky
subtraction. The data were astrometrically calibrated with
\textit{Scamp} \citep{ber06} and then stacked. Since images were taken
with two different sky position angles we could recover areas
initially lost due to blooming. The data did not allow for correction
of scattered light in the flat fields, for which extensive dithering
of photometric standard fields is required
\citep[see][]{mas01,mac04,kog04}.
We complemented the $BVRIz$ Subaru/Suprime-Cam data with $u^*griz$
CFHT/Megaprime images, which improves the photometric redshifts as a
result of the presence of $u^*$-band. The CFHT/Megaprime data were
pre-reduced using {\tt ELIXIR} \citep{mac04} at CFHT, including
corrections for scattered light of the order of 0.1 mag. The remaining
processing was done with THELI following \cite{ehl09}. The properties
of the coadded images are summarised in Table \ref{table_data_sup}.
\subsubsection{\label{catcreation}Catalogue creation}
Object detection and photometry was done using SExtractor
\citep{bea96} in double image mode. We stacked all exposures in all
filters of one camera with an image seeing of less than 1\myarcsec0,
obtaining a deep noise-normalised detection image. Coadded images in
the different filters were convolved to a common seeing of
0\myarcsec95, ensuring that the object flux in each waveband was
integrated over identical apertures. We kept objects with at least 5
connected pixels with $S/N\geq 2$ each.
The Subaru/Suprime-Cam data were only partially taken in photometric
conditions, with zeropoint variations of up to 0.1 mag in other
nights. We tied the photometric $z$-band image to CFHT/Megaprime data
taken in the same filter. The other Subaru/Suprime-Cam zeropoints were
inferred by comparing the fluxes from non-saturated stars, measured in
$3^{\prime\prime}$ wide apertures, against the \cite{pic98}
library \citep[for details see][]{ehl09}. We took into account filter
transmission, quantum efficiency, and the combined mirror reflectivity
and corrector throughput (Table \ref{subaru_transmission},
S. Miyazaki, priv. comm.). The photometric calibration of the
CFHT/Megaprime data was taken from the {\tt ELIXIR} headers.
The zeropoints of both data sets were ultimately fine-tuned during the
calculation of the photometric redshifts based on several hundred
calibration spectra (see Sect. \ref{photz}).
\begin{figure}[t]
\includegraphics[width=1.0\hsize]{13810_f02.ps}
\caption{\label{specphotz}{Comparison of photometric and
spectroscopic redshifts.}}
\end{figure}
\subsubsection{\label{photz}Photometric redshifts}
We need photometric redshifts for the weak gravitational lensing
analysis in Sect. \ref{weaklens}, mainly to distinguish between
lensed background and unlensed foreground galaxies. The photometric
redshifts were obtained as outlined in \cite{hpe09} for all objects in
the catalogues (see Sect. \ref{catcreation}) and calibrated
against 774 and 493 spectroscopic redshifts from \cite{met07}, for
CFHT and Subaru, respectively. Notice that \cite{met07} obtained
spectroscopic redshifts for a total of 1562 sources in the field of
view of MS0451. We performed the phot-z calibration using only those
spectra of sources with photometric errors smaller than 0.1 mag in all
bands. In detail, we fix the
redshifts of the corresponding galaxies to their spectroscopically
determined values. The magnitude differences between the best-fit
templates and the observed photometry then yield the zeropoint
corrections, in the range of $0.02-0.09$ mag for CFHT
and $0.04-0.18$ mag for Subaru. The correlation between
photometric and spectroscopic redshifts is shown in
Fig. \ref{specphotz} for both data sets, using a confidence limit
(ODDS parameter) higher than $0.8$. Due to the lack of $u$-band data
the Subaru photo-zs are highly unreliable for $z\lesssim0.3$, moving a
significant fraction of lensed galaxies into the unlensed foreground
sample. The CFHT data are much better in this respect, but shows $2-3$
times as much scatter for $z\gtrsim0.6$ and fails for fainter galaxies
due to inferior depth in $i$- and $z$-band. The accuracy of the
photo-zs is $\sigma\sim0.040$.
We run both data sets simultaneously through the photo-z process but
found the results to be significantly worse than the photo-zs obtained
separately for CFHT and Subaru. This is due to different PSF
characteristics of the two data sets which could not be homogenised
sufficiently, and in particular due to the fact that the Subaru data
could not be corrected for scattered light in the flats, resulting in
inconsistent magnitudes across similar passbands. We therefore created
a composite photo-z catalogue in the following manner. We took the CFHT
estimate if $z_{\rm phot}^{\rm CFHT}<=0.4$ and the average if both
estimates are between 0.4 and 0.7. The remaining galaxies were split
in two groups. The first is formed by galaxies for which the Subaru
redshift is higher than 0.7, and we assigned them this
estimate. Galaxies in the second group, with
$z_{\rm phot}^{\rm Subaru}<=0.7$ and $z_{\rm phot}^{\rm CFHT}>0.4$ got
either the CFHT or the Subaru redshift assigned, depending on which
one has higher confidence. Ultimately, the redshifts were transformed
into relative lensing strengths,
\begin{equation}
\beta=\frac{\mathrm{D}(z_{\mathrm{l}},z_{\mathrm{s}})}{\mathrm{D}(0,z_{\mathrm{s}})},
\label{beta}
\end{equation}
where $\mathrm{D}(z_1,z_2)$ is the angular diameter distance between two
sources at redshifts $z_1$ and $z_2$, and $z_{\mathrm{l}}$ and $z_{\mathrm{s}}$ are the lens
and source redshifts, respectively. See Sects. \ref{acs} and
\ref{weaklens} for more details.
\begin{table}
\caption{Summary of the Subaru/Suprime-Cam and CFHT/Megaprime
data. The limiting AB magnitudes (50\% completeness limit) are for
$10\sigma$ point sources, and are on average 0.8 mag brighter for
extended objects.}
\label{table_data_sup}
\begin{tabular}{l r r r r}
\hline
\hline
Telescope/Instrument & Filter (abbr.) & $t_{\rm exp}$ [s] & Seeing & $M_{\rm lim}$\\
\hline
\noalign{\smallskip}
Subaru/Suprime-Cam & WJB ($B$) & 12240 & 0\myarcsec82 & 26.7 \\
Subaru/Suprime-Cam & WJV ($V$) & 5040 & 0\myarcsec95 & 26.0 \\
Subaru/Suprime-Cam & WCRC ($R$) & 11400 & 0\myarcsec83 & 26.6 \\
Subaru/Suprime-Cam & WCIC ($I$) & 4920 & 0\myarcsec92 & 25.9 \\
Subaru/Suprime-Cam & WSZ ($z$) & 4380 & 0\myarcsec76 & 25.1 \\
\hline
CFHT/Megaprime & $u^*$ & 5220 & 0\myarcsec87 & 25.7\\
CFHT/Megaprime & $g$ & 3400 & 0\myarcsec85 & 26.0\\
CFHT/Megaprime & $r$ & 14850 & 0\myarcsec71 & 26.2\\
CFHT/Megaprime & $i$ & 1280 & 0\myarcsec71 & 23.7\\
CFHT/Megaprime & $z$ & 1440 & 0\myarcsec70 & 22.4\\
\hline
\end{tabular}
\end{table}
\begin{table}
\caption{Combined Subaru mirror reflectivity and corrector throughput}
\label{subaru_transmission}
\begin{tabular}{c c}
\hline
\hline
Wavelength [\AA]& Throughput\\
\hline
\noalign{\smallskip}
4450 & 0.774 \\
5500 & 0.828 \\
6590 & 0.828 \\
7710 & 0.791 \\
9220 & 0.765 \\
\hline
\end{tabular}
\end{table}
\subsection{\label{acs}HST/ACS imaging and shear catalogue}
For the weak lensing measurements and the strong lens modelling we
rely on wide-field imaging with HST/ACS through the F814W filter (PI:
R. Ellis). The data consist of 41 single orbit pointings of 2036s
each, covering a continuous area of $19^{\prime}\times19^{\prime}$,
and was reduced according to \cite{ses07,shj09}. An extensive
description of our shape measurement pipeline is given in
\cite{shj09}. In the following we summarise the main
characteristics.
The shear catalogue is based on SExtractor \citep{bea96} detections,
for which we required a minimum number of 8 connected pixels with
${\rm S/N}>1.4$ each after filtering with a $5\times5$ pixel wide
Gaussian kernel. The object catalogue created in this manner was then
fed into our implementation \citep[see][]{ewb01} of the KSB method
\citep{ksb95,luk97,hfk98} for the shape measurement, adapted for
HST/ACS as detailed in \citet{ses07,shj09}. We employed a
principal component interpolation for the variable HST/ACS
point-spread function and parametric corrections for charge-transfer
inefficiency for both stars and galaxies. In addition, we applied
weights $w_i$ to the individual shear estimates given by
\begin{equation}
w_i^{-1} = \left(\frac{2}{\mathrm{Tr}[P^g_i]}\right)^2
\sigma_{e_\mathrm{ani}}^2(\mathrm{mag}_i) + 0.25^2\,,
\end{equation}
where $\left(2/\mathrm{Tr}[P^g_i]\right)$ is the isotropic PSF
correction factor for galaxy $i$ and
$\sigma_{e_\mathrm{ani}}^2(\mathrm{mag})$ denotes the variance of the
PSF anisotropy corrected galaxy polarisations fitted as a function of
magnitude. We then selected galaxies with a minimum half light radius
of $r_h>1.2\,r_h^{*,{\rm max}}$, where $r_h^{*,{\rm max}}$ is the
maximum half light radius of the 0.25 pixel wide stellar locus in a
size-magnitude diagram. An explicit magnitude cut was not
performed. For more details we refer the reader to \cite{shj09}.
After all filtering, the shear catalogue contains 33500 galaxies with
redshift estimates $z>0.3$, corresponding to a number density of
$n=73$ arcmin$^{-2}$. 42\% of the galaxies have their redshifts
estimated photometrically as outlined in Sect. \ref{photz}.
For those galaxies without redshift estimate (median magnitude
$I_{\rm F814W}=26.0$) we used the mean magnitude-redshift relation
from \cite{shj09}. Thereto we split the galaxies into magnitude bins
of width 0.5 mag, starting from $I_{\rm F814W}=23.0$ down to
$I_{\rm F814W}=27.5$. For each bin we calculated the average lensing
strength $\langle \beta\rangle$ defined in eq. (\ref{beta}). Since the
lens is at a very low redshift of $z_{\mathrm{l}}=0.26$, it is insensitive
to the redshift distribution, in particular for galaxies with
redshifts $z\gtrsim0.7$. Essentially, $\langle \beta\rangle$ is
between 0.70 and 0.80 for 96\% of these galaxies, and we might as well
have assumed a constant redshift without affecting our results.
The median and mean redshift of all galaxies in the shear catalogue
are 1.39 and 1.16, respectively. Objects are evenly distributed over
the sky, and the area around J0454 has only very few masks for bright
stars, none of which is larger than $\sim20^{\prime\prime}$.
We estimate the 50\% completeness limiting AB magnitude of our shear
catalogue to $I_{\rm F814W}\sim26.1$ mag, consistent with the results
for the COSMOS field, which was observed with very similar strategies
\citep{saa07}. The depth matches the one for the ground-based data
(see Table \ref{table_data_sup}).
\begin{figure}
\includegraphics[width=1.0\hsize]{13810_f03.eps}
\caption{\label{cl0454_spectra}Redshifted FORS2 spectra of the lens
and the arc (binned 3 times). No significant features were found in
the spectrum of the arc. The noise level was offset by -0.5 for
better visibility.}
\end{figure}
\subsection{\label{vltfors2}VLT/FORS2 spectroscopy of the strong lens
system}
We used VLT/FORS2 to determine the redshifts of the fossil group's
brightest elliptical, E0454, and its arc system. Data were taken on
2009-03-23 in DDT time and in 1\myarcsec0
seeing, using the OG590 order sorting filter, GRIS\_300I grism and a
1\myarcsec0 long slit, resulting in a resolution of $R\sim660$
(1.68\AA$\;$ pixel$^{-1}$). The spectra were exposed for
$2\times600$s and their useable range extends over
$6250$\AA$-9300$\AA. The long slit covered the core of E0454 and the
bright northern arc.
We debiased, flat-fielded and sky-corrected the data using modified
THELI modules. A third-order polynomial was fit to the calibration
lamp emission lines for wavelength calibration, and a small residual
offset was corrected by comparison to sky lines. We obtained the
spectrum of E0454 by averaging 6 detector rows, yielding $S/N\sim20$
in the continuum (Fig. \ref{cl0454_spectra}). The spectrum of the arc
is strongly blended with that of E0454. To remove this contamination,
we exploited the symmetry of the lens and extracted a spectrum from
the opposite side of E0454 at the same distance as the arc. This
spectrum was subtracted from the arc's spectrum, which was then
averaged over 4 rows yielding $S/N\sim1-2$. The noise level was
determined from 200 nearby detector rows which only contained sky
background.
The lens redshift is $z=0.2594\pm0.0004$ and based on five absorption
features: MgI/MgH (5156/5196\AA), E-band (a blend
of Fe and Ca at 5269\AA), and NaD (5890/5896\AA). Thus E0454 is a
physical member of J0454, establishing the magnitude gap and thus the
fossil character. The redshift of the arc is more difficult to
infer. Our lens modelling (see Sect. \ref{stronglens}) yields a
magnification of $8-33$ for the arc, which allows us to resolve two
maxima in its light distribution. The colours of the object and the
morphology rule out an early-type galaxy. If the morphology is
indicative of star formation and if the redshift ($z_{\rm arc}$) of
the arc is less than about $1.0$, then there would be a chance to
detect the common set of nebular emission lines such as [OII]
(3728\AA), H$\beta$ (4863\AA), [OIII] (5008\AA) and H$\alpha$
(6565\AA) with the given exposure time. However, the spectrum
does not contain any significant features. There are two possible
explanations:
First, $z_{\rm arc}$ is lower than $\sim1.0$ and the morphology
observed is not indicative for star formation, or the star formation
rate (SFR) is low. In this case we can at least infer upper limits for
the SFR based on the non-detection of lines. For
$z_{\rm arc}=0.4$ the H$\alpha$ line would still be
accessible. Using \cite{ken98}, a presumed line width of
$30$\AA$\;$ and correcting for the strong lens magnification
(see Sect. \ref{stronglens}), we find
${\rm SFR}<0.15\,{\rm M_\odot}\,{\rm yr}^{-1}$. For
$z_{\rm arc}=0.7-1.0$ both [OII] and H$\beta$ are covered. Following
\cite{arl09} the upper limits for the SFR from these two lines are
${\rm SFR}<0.2-2.5\,{\rm M_\odot}\,{\rm yr}^{-1}$ for
$z_{\rm arc}=0.7$ and ${\rm SFR}<1-10\,{\rm M_\odot}\,{\rm yr}^{-1}$
for $z_{\rm arc}=1.0$, the uncertainties being due to the unknown
[OII]/H$\beta$ line ratio.
The second possibility is that the lensed source is at significantly
higher redshift, $z_{\rm arc}\gtrsim1.8$, such that possibly present
emission lines are redshifted beyond the spectral range covered by our
observations. The clear detection in $u^*$-band on the other hand
means $z_{\rm arc}<2.4$ and therefore $z_{\rm arc}=2.1\pm0.3$. Based
on strong-lensing properties and the stellar velocity dispersion of
E0454 we show in Sect. \ref{piemd} that this higher redshift is indeed
the most plausible assumption. The actual redshift of the arc is not
relevant for our main conclusions (see Sect. \ref{masscentroid}).
\begin{figure*}[t]
\includegraphics[width=1.0\hsize]{13810_f04.ps}
\caption{\label{galselection}{Target selection in colour-colour
and colour-magnitude space. The left and middle panels: galaxies
with $0.52<z<0.56$ and $0.28<z<0.5$ are shown as red and cyan
diamonds, respectively. Confirmed members of J0454 are coded
green, and blue triangles are objects with $0.1<z<0.24$. A
selection in colour-magnitude space leads to significant
contamination with objects at higher redshifts (left panel,
exemplary for $B-R$ vs. $R$). Instead, we selected galaxies in
$B-V$ vs. $V-I$ (middle). The right panel shows that the
galaxies selected in $B-V$ vs. $V-I$ form a well-defined red
sequence in $V-I$ vs. $I$, and the box indicates additional
selection criteria. Black points represent galaxies that were
kept based on this purely photometric selection, and grey ones
were excluded. Small corrections were made by means of available
spectra (see text for details).}}
\end{figure*}
\subsection{X-ray observations}
The field was observed for 42 ksec on 2004-09-17 with XMM-Newton (PI:
D. Warroll, observation ID 0205670101), covering a radius of $\sim14$
around MS0451. J0454 is contained in the 2XMM catalogue \citep{wsf09}
as source 2XMM J045400.6-030832:41489. We reduced the data using
XMM-SAS\footnote{http://xmm.esac.esa.int/sas/} v8.0.0. The maximum
flare level in the $10-15$ keV range is well below 0.35 counts
s$^{-1}$, and about half of the data were taken during completely
quiescent periods. Thus we did not reject any data due to high
background rates.
X-rays are particularly absorbed by neutral hydrogen,
\begin{equation}
I(E) = I_0\,e^{-\sigma_{\rm ph}(E)\, N_{\rm HI}}
\end{equation}
where $N_{\rm HI}=3.53\times10^{20} {\rm cm}^{-2}$ is the column
density along the line of sight \citep[taken from][]{kbh05}. Assuming
that all hydrogen atoms are in their ground state, we obtained the
quantum mechanical photon cross section as
\begin{equation}
\sigma_{\rm ph}(E) = 1.61\times10^{-23}\,{\rm cm}^2\,
\left(\frac{E}{\rm keV}\right)^{-3.5}\,.
\end{equation}
Accordingly, the absorption is significant for low X-ray energies
($0.2-0.5$ keV) and becomes low for energies higher than $0.5-1$
keV \citep[see also][]{moc83}. For the soft cluster spectrum of J0454
($T=1.1$ keV, see Sect. \ref{xray}), $2.5\%$ of the flux are absorbed
in the $0.5-2.0$ keV range. We also applied a k-correction factor of
1.06, interpolated from the values tabulated by \cite{bsg04}.
\section{Cluster members and field contamination}
\subsection{Object selection}
The MS0451 field had extensive wide-field spectroscopy with Keck by
\cite{met07}, who kindly made their redshift catalogue publicly
available. They randomly selected galaxies in a Subaru $I$-band
image (a subset of the data we use) from a sample with $I<21.5$ mag,
irrespective of morphology. Remaining spaces in the 14 slit masks were
then filled up with fainter objects. In total, redshifts
were obtained for 1562 galaxies in a $25^{\prime}\times20^{\prime}$
field that covers J0454 as well.
\subsubsection{\label{memberselection}Selection of J0454 member
galaxies}
The spectroscopic sampling of J0454 is complete to about 44\% for
$I<21.5$ mag (see Sect. \ref{contamination}). First, J0454 is at
significantly lower redshift than MS0451, and thus its brighter
galaxies were not observed as it is implausible that they are members
of MS0451. This holds in particular for the central elliptical with
$I=16.6$ mag. Second, its angular separation from MS0451 is
8\hbox{$^{\prime}\;$} and thus it was not at the centre of interest. Lastly,
slit masks cannot be configured arbitrarily due to source
clustering.
Our photometric redshifts ($\sigma_{\rm photz}=0.040$) do not offer
sufficient power to distinguish unambiguously between
structures in the range $z=0.24-0.32$, which are present along
the line of sight (Sect. \ref{contamination}). For a more complete
picture we therefore selected ellipticals in colour-colour and
colour-magnitude space using the red cluster sequence (RCS) method
from \cite{gly00}. The spectroscopic redshifts were used to identify
suitable areas. However, the large angular extent of MS0451 and its
significant content of blue galaxies leads to a high contamination
when using the red sequence alone (see e.g. Fig. \ref{galselection},
left panel). We investigated various colour-colour combinations and
found that in $B-V$ vs. $V-I$ (Fig. \ref{galselection}, middle panel)
the highest redshift differentiation is achieved. All galaxies at
$z\sim 0.26$ are cleanly separated from those at $z=0.54$, thus
removing the bulk of the contamination. There is also very little
overlap with galaxies at $z\geq 0.3$. Only bluer objects at $z=0.26$
cannot be separated from those at lower redshift. In a first pass, we
selected objects with
\begin{equation}
B-V>1.2
\end{equation}
\begin{equation}
B-V<1.7
\end{equation}
\begin{equation}
B-V > 1.286\,(V-I) + 0.07
\end{equation}
\begin{equation}
B-V < 1.286\,(V-I) + 0.59\,.
\end{equation}
These form a red sequence in $V-I$ vs. $I$ (diamonds in the right
panel of Fig. \ref{galselection}) with a typical width of
$\sigma=0.049$ \citep[see e.g.][]{hsw09}. Only objects within
$2\sigma$ of the red sequence and with $I\leq22$ are kept for later
analysis. The $I<22$ cut-off was chosen for two reasons. First, the
width of the red sequence increases significantly for fainter galaxies
(see right panel of Fig. \ref{galselection}), and thus the
contamination rate would increase as well. Second, the Keck
spectroscopic survey is limited by $I\lesssim21.5$. Pushing the
photometrically selected sample significantly beyond this limit would
mean that we could not quantify anymore the contamination rate by
structures with similar redshifts.
Taken all together, these selection criteria exclude all galaxies with
$z_{\rm spec}<0.24$, and all but two ellipticals with
$z_{\rm spec}>0.29$. A good fit to the red sequence formed by the
remaining galaxies is
\begin{equation}
V-I = -0.0430\,I + 1.768\,.
\end{equation}
From the sample of 55 galaxies selected in this manner (black
diamonds in the right panel of Fig. \ref{galselection}) we removed
three with higher and one with lower spectroscopic redshift, and
those where the photometric redshifts deviated by more than 0.1
from the cluster redshift ($2.5\sigma$ rejection, 4 objects). In
total, 47 galaxies remained to which we added 17 with confirmed
redshifts, most of them galaxies with blue colours. One galaxy
(object \#10) had colours redder than the red sequence (caused by
a prominent dust lane) and was added back to the sample.
In total, 15 of the red sequence galaxies have spectroscopic
redshifts, including the red galaxy that was added back to the
sample. Assuming that similar effects hold for the 32 red sequence
galaxies without spectra, we estimate that about 2 galaxies were
overlooked. Our sample of red sequence galaxies is then $95\%$
complete down to $i=22.0$ mag ($M_i=-18.6\pm0.05$).
We confined the galaxy sample to within 6\hbox{$^{\prime}\;$} of the
brightest elliptical galaxy, E0454, corresponding to $1.7\,\langle
r_{200}\rangle$ (see Sects.
\ref{r200galcounts} and \ref{tangshear}). Beyond this perimeter the
number density of red sequence galaxies is indistinguishable from
the density of field galaxies selected in the same manner ($n=0.09$
arcmin$^{-2}$, determined from a $10^{\prime}\times11^{\prime}$ wide
area where structures with $0.2<z<0.3$ are unknown).
After correcting for galactic extinction \citep{sfd98} we
determined the k-correction \citep{hbb02} using {\tt kcorrect}
\citep[v. 4.1.4,][]{blr07}. For better comparison with
other publications we report the rest-frame absolute magnitudes in the
Sloan $g$ and $i$ passbands. The errors for M$_g$ and
M$_i$ are 0.07 and 0.05 mag, respectively, based upon measurement
uncertainties and the internal error estimate of {\tt kcorrect}. The
spatial distribution of the member galaxies is shown in
Fig. \ref{gal_spatialdist}, and their properties are summarised in
Table \ref{galsample}.
\begin{figure}[t]
\includegraphics[width=1.0\hsize]{13810_f05.ps}
\caption{\label{gal_spatialdist}{Photometrically and
spectroscopically selected cluster galaxies. The circle
indicates $\langle r_{200}\rangle=830$ kpc, centred on E0454.}}
\end{figure}
\subsubsection{Magnitude gap}
Assuming a virial radius of $\langle r_{200}\rangle=840$ kpc
(3\myarcmin5) from our analysis presented below, we determined a
magnitude gap of $\Delta m_{12}=2.5$ mag in $I$-band for J0454 within
half the virial radius. The second-brightest galaxy is object \#20
from Table \ref{galsample}, an elliptical galaxy at a separation of
$0.46\,\langle r_{200}\rangle$ and with spectroscopic confirmation of
its redshift. Notice that within $0.5\,\langle r_{200}\rangle$
there is no other possible foreground or background galaxy brighter
than the second-brightest member galaxy, hence the fossil character of
J0454 is secured. The third- and fourth-brightest members within $0.5\,
\langle r_{200}\rangle$ are 2.8 mag fainter than E0454 and also
spectroscopically confirmed. Two brighter galaxies exist at larger
radii with $\Delta_m=1.8-1.9$ (objects \#34 and \#45), but they do
not have their redshifts measured. For a meaningful luminosity
function we need complete spectroscopic sampling, in particular
because the line of sight is contaminated by nearby structures in
redshift space (see Sect. \ref{contamination}).
\begin{figure*}[t]
\includegraphics[width=1.0\hsize]{13810_f06.ps}
\caption{\label{zslices}{Clustering for different spectroscopic
redshift bins and their width (in parentheses). J0454 is shown
in the upper right, MS0451 in the lower left. The circle is
centred on E0454 and traces $\langle r_{200}\rangle=830$ kpc at
$z=0.26$. North is up and East is left. The field is $25^\prime$
wide and centred on $\alpha=$ 04:54:06, $\delta=$ -03:02:06.}}
\end{figure*}
\subsubsection{\label{contamination}Structures along the line of sight}
Based on the Keck spectra we identified 16 structures between
$0.1<z<0.8$, consisting of at least 12 galaxies within
$\Delta z=0.01$. The spatial distributions of the 12 most significant
ones are shown in Fig. \ref{zslices}. The circle indicates
$\langle r_{200}\rangle=830$ kpc determined below from galaxy counts
(Sect. \ref{r200galcounts}) and weak gravitational lensing
(Sect. \ref{tangshear}). MS0451 overlaps significantly with J0454,
whereas other structures contribute fewer interlopers.
The distributions shown in Fig. \ref{zslices} are representative of
the actual galaxy distribution. This is not self-evident due to
the incomplete spectroscopic sampling with slit masks. However, the
main selection criterion of \cite{met07} was simply $I<21.5$ mag, with
a possible bias preferring galaxies closer to MS0451 over those with
larger separations. Thus the selection function is approximately
constant across the field and does not favour one particular structure
over another.
The line of sight towards J0454 is not only contaminated by MS0451
but also by structures at $z=0.240$, 0.246, 0.282, 0.293 and
0.325. Without spectra we cannot distinguish these from members at
$z=0.26$. We estimated the contamination assuming that the
interlopers had the same probability of being selected for
spectroscopy as the members of J0454. From the number of red
sequence galaxies with and without spectra we determined the
spectroscopic coverage to be 44\% complete for $I<21.5$. Five
interlopers were kept by the initial selection (see Sect.
\ref{memberselection}) and therefore we expect that about 10 of the 32
purely photometrically selected galaxies in Table \ref{galsample} are
not true members of J0454. We applied corrections for this where
necessary.
\section{\label{kinematics}Morphology-density relation, kinematics and
\boldmath${r_{200}}$}
In this section we show that J0454 has characteristics
typical for normal galaxy clusters, such as a distinct
morphology-density relation \citep[see e.g.][]{gyf03} and a significantly
lower velocity dispersion for the central population of elliptical
galaxies as compared to the population of spirals. Based upon general
cluster scaling relations, we obtain size and mass estimates.
\begin{figure}[t]
\includegraphics[width=1.0\hsize]{13810_f07.ps}
\caption{\label{radecz}{The kinematic structure of J0454 with
respect to E0454. Filled and open symbols mark red- and
blueshifted galaxies, respectively.}}
\end{figure}
\subsection{Cluster extent and mass: $r_{200}$ and $M_{200}$}
A characteristic key estimate of a cluster's linear extent is the
virial radius. It is often approximated by $r_{200}$, within which the
mean density is 200 times higher than the critical density
$\rho_{\rm c}$,
\begin{equation}
\label{rhocrit}
\rho_{\rm c}(z)= \frac{3}{8 \pi G} H^2(z), \;{\rm with}
\end{equation}
\begin{equation}
\label{hubble}
H^2(z) = H_0^2\;[\Omega_m(1+z)^3+\Omega_\Lambda]
\end{equation} being the Hubble function. The mass enclosed
within $r_{200}$ is
\begin{equation}
\label{m200}
M_{200} = 200\,\rho_{\rm c}(z)\,\frac{4 \pi}{3}\,r_{200}^3\,.
\end{equation}
A common estimator for the virial mass is
\begin{equation}
\label{m200vir}
M_{200}^{\rm dyn} \sim \frac{3\sigma_v^2}{G}\; r_{200}\,,
\end{equation}
which can be combined with (\ref{rhocrit}) and (\ref{m200}) yielding a
dynamic estimate for $r_{200}$,
\begin{equation}
\label{r200dyn}
r_{200}^{\rm dyn} = \frac{\sqrt{3}}{10}\frac{\sigma_v}{H(z)}\;.
\end{equation}
Estimating virial masses from galaxy dynamics is non-trivial
\citep[see e.g.][]{cye97}, in particular if the cluster under
investigation is poorly sampled with spectroscopic redshifts. Our
dynamic mass and size estimates for J0454 should therefore be viewed
with caution, and we complement them with more robust cluster scaling
relations, weak lensing and X-ray estimates.
\subsection{\label{velfield}Velocity field and virial estimate of $r_{200}$}
In Fig. \ref{radecz} we show the positions of all
galaxies around J0454 with spectroscopic redshifts in the range
$0.255<z<0.265$. The symbol size encodes the relative velocity with
respect to E0454, and open (filled) symbols denote blueshifted
(redshifted) motions. We notice two filaments extending up to
4.3 Mpc to the North and to the North-West. The former is on average
blue-shifted by $-595\rm ~km~s^{-1}$ compared to E0454, whereas the latter
does not show a significant motion. A photometric selection of more
member galaxies in these areas would result in significant
contamination as these filaments are projected onto four structures at
similar redshifts (Fig. \ref{zslices}). We thus confined our
subsequent analysis to the region within 6\hbox{$^{\prime}\;$} from E0454.
We compute the velocity dispersion $\sigma_v$ as
\begin{equation}
\sigma_v^2=\left(\frac{c}{1+\langle z \rangle}\right)^2\,
\left(\frac{1}{N-1}\sum_i\left[z_i-\langle
z\rangle\right]^2 - \langle\delta\rangle^2\right)\,,
\end{equation}
excluding E0454 and following the prescription of \cite{dzt80} and
\cite{har74}. Therein, $c$ is the speed of light, $\langle z\rangle$
the mean cluster redshift, and $\langle\delta\rangle$ the uncertainty
in the redshift measurement \citep[$50\rm ~km~s^{-1}$, from][]{met07}. The
factor $(1+\langle z\rangle)^{-1}$ cancels the stretching effect of
cosmic expansion. After the visual classification of the galaxies'
morphologies based on their appearance in the HST/ACS data, we
determined $\sigma_v$ for the red (E, S0) and the blue (Sa-Sc, Irr)
population and for all galaxies together (see left panel of
Fig. \ref{galtypes}). Including a correction for local peculiar
motions \citep{rrs06} we have
$\sigma_v^{\rm red}=480\pm20\rm ~km~s^{-1}$,
$\sigma_v^{\rm blue}=590\pm20\rm ~km~s^{-1}$, and
$\sigma_v^{\rm all}=570\pm20\rm ~km~s^{-1}$.
The errors were obtained from the propagated mean measurement error,
and include a conservative estimate for the uncertainty of the local
peculiar motion and a possible net motion of J0454.
The velocity dispersion of the red galaxies is significantly lower
than the one of the blue galaxies, which is expected from dynamical
friction and the morphology-density relation \citep[right panel of Fig.
\ref{galtypes}, consistent with the findings of][for a much larger
sample of clusters]{gyf03}. Their mean velocities are different too,
and offsets exist with respect to E0454 ($+240\rm ~km~s^{-1}$ for the red
population, significant on the $2.5\sigma$ level, and $+540\rm ~km~s^{-1}$
($5.7\sigma$) for the blue galaxies). For the red galaxies this could
still be an observational effect due to incomplete sampling, as within
1\hbox{$^{\prime}\;$} of E0454 only two of nine ellipticals have their
redshifts measured. If confirmed by future observations, these
features would indicate that these galaxies have a different origin
than those which already collapsed into E0454, and that significant
substructure exists in the entire system \citep[see also][]{oeh01}.
Using equations (\ref{m200vir}) and (\ref{r200dyn}) we obtained
$r_{200}^{\rm dyn}=1054\pm44$ kpc and
$M_{200}^{\rm dyn}=(1.69\pm0.14)\times10^{14}\,{\rm M_\odot}$ for the red
population, and $r_{200}^{\rm dyn}=1295\pm44$ kpc and
$M_{200}^{\rm dyn}=(3.14\pm0.21)\times10^{14}\,{\rm M_\odot}$ for the blue
population.
\subsection{\label{r200galcounts}Size-richness relation}
\cite{hmw05} and \cite{jsw07} have shown that $r_{200}$ and $M_{200}$
can be estimated starting from the number $N_{\rm gal}$ of galaxies
within a radius of $1\,h_{100}^{-1}$ Mpc of the BCG. Only galaxies in
the red sequence and with $i$-band luminosities $L>0.4 L_*$ are
considered. Based on $N_{\rm gal}$ one has
\begin{equation}
\label{hansen05}
r_{200}^{\rm gal}=0.156\, h_{100}^{-1}\; {\rm Mpc}\; N_{\rm gal}^{0.60}\,,
\end{equation}
a refined version of the original relation from \cite{hmw05}. Within
$r_{200}^{\rm gal}$ the luminosity is 200 times the mean luminosity of
the Universe. It must not be mistaken for $r_{200}$ which refers to
matter overdensity, yet the two are closely related \citep{jsw07}.
Based on weak lensing measurements and the number $N_{200}$ of
galaxies within $r_{200}^{\rm gal}$, \cite{jsw07} and \cite{hsw09}
obtain
\begin{equation}
r_{200}=0.182\,h_{100}^{-1}\,{\rm Mpc}\,N_{200}^{0.42}
\end{equation}
\begin{equation}
M_{200}=1.75\times10^{12}\,h_{100}^{-1}\,{\rm M_\odot}\,N_{200}^{1.25}\,.
\end{equation}
Using $M_{i,*}=-21.8$ mag from \cite{hsw09}, we counted
$N_{200}=15^{+1}_{-2}$ red sequence galaxies with $M_i<-20.8$ mag
(corresponding to $0.4\,L_*$). This richness estimate contains a
correction for field contamination, and the errors are due to an
uncertainty of $0.2$ mag which we allowed for $M_{i,*}$. As a result
we have $r_{200}=811\pm46$ kpc and
$M_{200}=(0.74\pm0.15)\times10^{14}\,{\rm M_\odot}$, including 13\%
intrinsic uncertainty for the mass-richness relation.
\begin{figure}[t]
\includegraphics[width=1.0\hsize]{13810_f08.ps}
\caption{\label{galtypes}{Left panel: Redshift distribution for
the red (E, S0) and blue (Sa-Sc, Irr) galaxy populations. Notice
that the spectroscopic sampling of ellipticals is complete to
only $\sim44\%$. The cross marks E0454. Right panel: Galaxy
types as a function of angular separation from E0454.}}
\end{figure}
\section{\label{xray}X-ray halo}
The XMM-Newton image of J0454 is shown in Fig. \ref{j0454_mass_xmm},
overlaid over the HST/ACS optical image, and in Fig.
\ref{j0454_label_cropped} in the appendix (overlaid over a colour
picture of the Subaru/Suprime-Cam data, online material). X-ray flux
is detected locally out to 1\hbox{$^{\prime}\;$} (240 kpc) from the core of
E0454, encompassing the 10 innermost galaxies. If azimuthally
averaged, we can trace the halo about twice as far. It is possible
that this very extended emission is not associated with E0454 anymore
but with the surrounding cluster of galaxies (see below). The offset
of the X-ray centroid with respect to E0454 is
$6^{\prime\prime}\pm4^{\prime\prime}$ (24 kpc). The luminosity profile
is described by an isothermal $\beta$-model with
$\beta=0.57\pm0.06$ and a core radius of $r_{\rm c}=120\pm17$ kpc
(Fig. \ref{cl0454_xray_profile}). The best-fit isothermal redshifted
bremsstrahlung model of the spectrum yields $T=1.1\pm0.1$ keV.
Assuming a mean particle mass of $\mu=0.6$ we find
$M_{200}=(0.34\pm0.10)\times10^{14}\,{\rm M_\odot}$ and $r_{200}=617\pm28$
kpc, respectively, and for the total luminosity within $r_{200}$ we
have $L_X=(1.4\pm0.2)\times10^{43}\,h^{-2}\rm ~erg~s^{-1}$.
A cooling flow is absent from the data as can be seen from the
luminosity profile. Consequently, we do not expect star formation in
the core of E0454. This is confirmed by our VLT/FORS2 spectrum
(Fig. \ref{cl0454_spectra}) which does not show any
H$\alpha$-emission, which would be a prime indicator for star
formation second to molecular CO emission \citep{edg01}.
The X-ray properties of J0454 agree with those of normal
groups and clusters. \cite{rmb08} find a tight correlation between
$\langle L_X\rangle$ and $\langle N_{200}\rangle$ of 17000 maxBCG
clusters, and this relation describes J0454 well. The $L_X-\sigma$
relation drawn from the same sample predicts $\sigma\sim480\pm30\rm ~km~s^{-1}$,
the same as we measured for the elliptical galaxy population. In the
compilation of \cite{mul00} J0454 falls comfortably within the
natural scatter of the $L_X-\sigma$ relation, resembling either a
rich group or a poor cluster.
\begin{figure}[t]
\includegraphics[width=1.0\hsize,angle=90]{13810_f09.ps}
\caption{\label{j0454_mass_xmm}HST/ACS image of J0454. The (jagged)
blue contours trace the S/N-ratio of the $0.5-2.0$ keV X-ray flux,
starting with $3\sigma$ and increasing in steps of $2\sigma$. A
6\hbox{$\;\!\!^{\prime\prime}\;$} wide kernel was used for smoothing. The (smooth)
black contours trace the S/N of the weak lensing mass reconstruction,
starting with $2\sigma$ and increasing in steps of $0.5\sigma$,
smoothed with a $40^{\prime\prime}$ wide kernel. The white square
outlines the area of the strong lensing system shown in
Fig. \ref{fig:SLImagePos}, and the red cross marks the centroid of
the distribution of elliptical galaxies within
$\langle r_{200}\rangle\sim830$ kpc.}
\end{figure}
Differences occur in temperature-based scaling relations. While no
deviation is found with respect to the $L_X-T$ relation from the
HIFLUGCS sample \citep{seb06,reb02}, J0454 appears cooler than
expected ($\sim2$ keV) when comparing it to the $L_X-T$ relations
presented by \cite{mul00} and \cite{rmb08}. A similar trend is
seen for $T-\sigma$ \citep{mul00}, i.e. for $\sigma=480\rm ~km~s^{-1}$ one would
expect $T\sim2.0$ keV (or $\sigma\sim330\rm ~km~s^{-1}$ for $T=1.1$ keV). These
deviations can be explained by the natural scatter seen in groups of
galaxies. A different explanation would be that we see a group-sized
substructure embedded in, but not yet fully merged with, a larger
sparse cluster. Extended and patchy X-ray emission exists on the
lowest levels and at radii $\gtrsim 1^{\prime}$. It is unclear whether
this emission is still part of the E0454 halo or if we see the
brightest emission features of the gas associated with
J0454. With deeper X-ray data we could look for temperature variations
or different chemical compositions to distinguish these two
components. We discuss these findings in Sect. \ref{interpretation}.
The inner, flat core of the X-ray halo is elongated, tracing the
optical ellipticity of E0454. These trends have been seen
previously for groups \citep[e.g.][]{muz98} and clusters
\citep[e.g.][]{hhb08}, and also for fossils
\citep{kjp04,sfv04,kmp06}. In general, the X-ray contours of the halo
analysed in this work are not as concentric and regular as e.g. those
for the fossil groups RX J1331.5+1108 and RX J1416.4+2315 from
\cite{kpj07}, yet they do not appear more disturbed than those of the
other three fossil groups presented by the same authors.
We mention here that the X-ray halo of E0454 was detected
previously and is listed as object \#6 in the Chandra cluster sample
of \cite{bos02}. The reported centroid of the X-ray flux is located
$\sim41\pm8$ kpc south-east of E0454, whereas the XMM-Newton data
reveals only a small offset of $24\pm16$ kpc to the North-West. We
explain this by the fact that XMM-Newton collected more than 10 times
as many photons as Chandra.
\begin{figure}[t]
\includegraphics[width=1.0\hsize]{13810_f10.ps}
\caption{\label{cl0454_xray_profile}Best-fit $\beta$-model for the
X-ray halo. A cooling flow is absent.}
\end{figure}
\section{\label{weaklens}Weak lensing analysis}
The strength of a gravitational lens scales with the ratio of the
angular diameter distances $D(z_1,z_2)$ between the lens and the source and
between the observer and the source. The more distant the source the
stronger the lensing effect, but for a lens redshift $z_{\mathrm{l}}=0.26$ and
sources at $z_{\mathrm{s}}>0.8$ it is effectively constant. One must project the
sources to some arbitrarily chosen reference redshift $(z_{\mathrm{r}}=1)$ and
rescale the shear estimator (the image ellipticities) accordingly to
obtain comparable shear values,
\begin{equation}
\varepsilon_{1/2}=\varepsilon_{1/2}^0
\frac{\mathrm{D}(z_{\mathrm{l}},z_{\mathrm{r}})}{\mathrm{D}(0,z_{\mathrm{r}})}\,
\frac{\mathrm{D}(0,z_{\mathrm{s}})}{\mathrm{D}(z_{\mathrm{l}},z_{\mathrm{s}})}\,.
\end{equation}
This rescaling decreases (enhances) the noise for $z_{\mathrm{s}}>z_{\mathrm{r}}$
($z_{\mathrm{s}}<z_{\mathrm{r}}$) and is taken into account by individual weighting factors
\begin{equation}
w = \left(\frac{\mathrm{D}(z_{\mathrm{l}},z_{\mathrm{s}})}{\mathrm{D}(0,z_{\mathrm{s}})} \,
\frac{\mathrm{D}(0,z_{\mathrm{r}})}{\mathrm{D}(z_{\mathrm{l}},z_{\mathrm{r}})}\right)^2\,.
\end{equation}
Before we could proceed on the weak lensing analysis of J0454 we had
to remove the lensing contribution of MS0451 from the data by
subtracting a singular isothermal sphere (SIS) tangential shear
profile parametrised with $\sigma_v=1354\rm ~km~s^{-1}$. This value was taken
from \cite{cye97}, who used an iterative outlier rejection process for
its determination. A concrete error estimate was not given, but by
comparing to other measurement methods in their work, we adopted
an uncertainty of $5\%$.
Other known structures apart from MS0451 (see Fig. \ref{zslices}) do
not need to be taken into account, as their angular separation is too
large and their velocity dispersion is too low to leave a measurable
footprint at the position of J0454. The X-ray data are consistent with
this picture, revealing no structures apart from MS0451 that could add
discernible lensing signals to J0454.
\begin{figure}[t]
\includegraphics[width=1.0\hsize]{13810_f11.ps}
\caption{\label{cl0454_tangshear}Tangential shear of J0454 and
best-fit SIS and NFW profiles. The shaded area shows the 68\%
confidence region of the SIS fit.}
\end{figure}
\subsection{\label{massrec}Mass reconstruction}
We use the finite-field method from \cite{ses01} to reconstruct the
projected surface mass density, $\kappa$, from the sheared images.
This method uses the field border as a boundary condition, which makes
reconstructions of non-rectangular areas difficult. We therefore work
on a 16\myarcmin8 wide rectangle inscribed into the HST/ACS
mosaic. Our code is freely
available\footnote{http://www.astro.uni-bonn.de/$\sim$mischa/download/massrec.tar}
and based on the original version from \cite{ses01}.
The convergence $\kappa$ is determined up to an additive constant, the
`mass-sheet' degeneracy, which is safely broken by assuming that
$\kappa$ vanishes on average along the border of the field. The
algorithm only works for under-critical regions with $\kappa<1$,
i.e. strong lensing areas are not reconstructed reliably. In the case
of J0454 this affects only the innermost 4\hbox{$\;\!\!^{\prime\prime}\;$} (see Sect.
\ref{stronglens}), which is well below the resolution limit and thus
of no concern.
The resulting density map must not be interpreted without a
corresponding noise map. For example, bright stars cause holes in the
data field, which locally increase the noise due to the reduced number
density of galaxies. In addition, the smoothing length for the shear
field must be larger than these holes. Otherwise, the boundary
condition of a rectangular data field is violated, resulting in a
corrupted solution. To obtain the noise map, we created 1000
realisations of randomised galaxy orientations keeping their positions
fixed, and obtained $\kappa$ for each. The two-dimensional rms of
these $\kappa$-maps yields the desired noise map. Since lensing
increases the ellipticities of galaxies, we removed the SIS shear
profile of J0454 (Sect. \ref{tangshear}) from the data prior to the
randomisations. Otherwise the noise at the cluster position would
be overestimated.
The S/N-level of the mass map is shown in Fig. \ref{j0454_mass_xmm}.
J0454 is detected on the $4.7\sigma$ level with a peak convergence of
$\kappa=0.20$. It is the only significant ($S/N>4$) mass peak besides
MS0451 (${\rm S/N}=7.7$), and located $12\pm5$\hbox{$\;\!\!^{\prime\prime}\;$} south of
E0454. The uncertainty in the position was determined from
boot-strapping the shear catalogue. The mass of J0454 within 182 kpc
(approximately tracing the ${\rm S/N}=1$ contour) is
$M=(0.38\pm0.09)\times10^{14}\,{\rm M_\odot}$. This is not comparable to
$M_{200}$ since it is integrated within a much smaller radius. A
determination of $M_{200}$ from the reconstructed density map is not
sensible as the noise entirely dominates the signal in the larger
aperture. However, we can infer a lower limit of $r_{200}\gtrsim
650\pm50$ kpc. One way to test the integrity of the detection is to
check for noise peaks in the 1000 randomisations with equal or higher
significance. No such peak is found, consistent with the expectation
(0.21 peaks) from idealised Gaussian noise. In reality the noise is
non-Gaussian as the dispersion of image ellipticities is
non-Gaussian. Probing the actual differences for $\sim5\sigma$
peaks would require many more randomisations, but would not change
our main conclusion here that is that we detected a real signal.
As mentioned previously, we removed the contribution of MS0451 by
subtracting a SIS profile with $\sigma_v=1354\rm ~km~s^{-1}$. Changing this
value by 5\% alters the mass estimate by 0.1\%, hence this measurement
is insensitive to the presence of MS0451. This is not unexpected as
the separation between J0454 and M0451 is large and $\kappa$ is a
local quantity, resulting in no overlap of the clusters' projected
surface mass densities.
\subsection{\label{tangshear}SIS and NFW fits to the tangential shear
profile}
We fit SIS and NFW profiles to the tangential shear
around J0454 (Fig. \ref{cl0454_tangshear}), assuming a spherical
symmetric density distribution. As compared to the mass
reconstruction, the results are not model-independent. For
the SIS we furthermore assumed that the system is in virial
equilibrium with isotropic distribution of the orbits, having a
density profile
\begin{equation}
\rho_{\rm SIS}(r)=\frac{\sigma_v^2}{2\pi\,G\,r^2}
\end{equation}
which yields, in analogy to the derivation of equation (\ref{r200dyn}),
\begin{equation}
r_{200}=\frac{\sqrt{2}}{10}\,\frac{\sigma_v}{H(z)}
\end{equation}
(note the different pre-factor).
The tangential shear is measured with respect to a reference point,
which should be near or at the centre of mass, depending on
substructure. We identify the position where the tangential shear
is maximised with a matched-filter technique \citep[the $S$-statistics
or peak finder, see][]{seh07}. The signal is maximised for a
5\myarcmin5 wide filter approximating the NFW shear profile, detecting
J0454 on the $5.2 \sigma$ level $12^{\prime\prime}\pm5^{\prime\prime}$
south-east of E0454 (position angle $162\pm2$ degrees, both error
estimates from bootstrapping). This is coincident with the peak of the
mass reconstruction and indicates a robust choice for the reference
point. In general, the two peaks would not necessarily coincide as
both methods compute very different quantities. Deviations can occur
in particular for clusters with significant substructure
\citep[see e.g.][]{hsd09}, provided that the $S/N$ is high enough to
resolve such features. With this reference point the SIS fit yields
$\sigma_v^{\rm wl}=476\pm46\rm ~km~s^{-1}$ and
$M_{200}=(0.90\pm0.26)\times10^{14} {\rm M_\odot}$, and from the NFW fit
we obtained a concentration parameter of $c=9.5\pm4.8$,
$r_{200}=834\pm219$ kpc and
$M_{200}=(0.84\pm0.66)\times10^{14}{\rm M_\odot}$. Contrary to the
convergence $\kappa$, the shear is a non-local quantity and therefore
more susceptible to changes in the velocity dispersion assumed for
MS0451. For example, decreasing (increasing) its $\sigma_v$ by 5\%
results in a 2\% (4\%) increase of the velocity dispersion for
J0454. These effects are included in our error budget.
To quantify the effect of possible errors in the choice of
the reference point, we repeated the analysis using the core of E0454
and the centroid of the distribution of elliptical galaxies within
$r_{200}=830$ kpc. This yielded $\sigma_v=462\pm49\rm ~km~s^{-1}$ and
$376\pm58\rm ~km~s^{-1}$, respectively. The first fit is qualitatively slightly
worse than the original fit but still acceptable, whereas the second
is significantly deteriorated. The centroid of the distribution of
elliptical galaxies can therefore be ruled out as the centre of
mass. We show below based on strong lensing that this also
applies to E0454.
\begin{figure}
\includegraphics[width=1.0\hsize]{13810_f12.ps}
\caption{\label{fig:SLImagePos} HST/ACS image of the strong
lens. The counter image and the arcs reveal two maxima in the source
intensity distribution, forming two sets of multiple image
systems. They are marked by circles and squares and are used for
the lens modelling.}
\end{figure}
\section{\label{stronglens}Strong lensing analysis}
\subsection{\label{lensmodel}Lens modelling}
We identified two sets of multiple image systems, corresponding to two
bright knots in the source intensity distribution and identified by
circles and squares in Figure \ref{fig:SLImagePos}. The lens is
modelled using a pseudo-isothermal elliptic mass distribution
\citep[PIEMD,][with zero core radius]{KassiolaKovner93},
\begin{equation}
\label{eq:piemd}
\kappa(\theta_1, \theta_2) = \frac{b}{1+q}
\left({\theta_1^2}+\frac{\theta_2^2}{q^2}\right)^{-1/2},
\end{equation}
where $b$ is the strength and $q$ is the axis ratio. The $1/(1+q)$
normalisation is needed to match the profile in
\citet{KassiolaKovner93} that was defined using ellipticities
($\epsilon \equiv (1-q)/(1+q)$) instead of axis ratios. The
distribution is translated by the centroid position and rotated by the
position angle, $PA_{\rm L}$. Furthermore, we allowed a constant
external shear with strength $\gamma_{\rm ext}$ and $PA_{\rm ext}$. In
total, there are 11 parameters: 4 for the two source positions, 5 for
the PIEMD, and 2 for the external shear. The two sets of multiple
images provide 16 constraints. Note that the modelling is independent
of lens and source redshifts.
\begin{figure}[t]
\begin{center}\includegraphics[]{13810_f13.ps}
\end{center}
\caption{\label{fig:SLCritCaus} The most probable critical curve
(dashed) and caustic curve of the lens. The caustic
consists of four folds (solid lines) joining at four cusps. The
modelled source and image positions for the two sets of
multiple-image systems are marked by squares and circles.}
\end{figure}
We used the strong lens modelling code (Halkola et al. 2010, in
preparation) based on \citet{HalkolaEtal06}, \citet{HalkolaEtal08},
\citet{SuyuEtal06} and \citet{DunkleyEtal05}. Markov chain Monte Carlo
(MCMC) methods were employed to obtain the posterior probability
distributions of the lens parameters. We placed Gaussian priors on the
centroid, $q$, and $PA_{\rm L}$ of the PIEMD (with Gaussian widths of
$0.05''$, $0.09$ and $10^{\circ}$, respectively) based on the observed
light distribution.
Table \ref{tab:SLparams} lists the results of the marginalised lens
parameters from a MCMC chain of length $10^5$ after the burn-in phase.
Typical predicted image positions agree with the observations within 1
pixel (rms $\sim0\myarcsec03$). Figure \ref{fig:SLCritCaus} shows the
critical and caustic curves of the most probable lens parameters.
The arc is in a fold configuration, close to being in a cusp
configuration (i.e., the positions of the bright knots in the source
lie next to a fold and are in the vicinity of a cusp of the caustic
curves). The two merging images form the northern half of the arc
with magnifications $\mu=8.2-33.8$, and the other image the southern
arc with $\mu=3.0-4.4$. The counter image has $\mu\sim2.3$.
The separations between the arc and the lens, and between the counter
image and the lens, are $1\myarcsec85-2\myarcsec18$ and 3\myarcsec42,
respectively. This asymmetry requires the presence of significant
external shear, $\gamma_{\rm ext}=0.12$. As gravitational lensing
is an achromatic process, all images should have similar colours,
which allowed us to test the counter image hypothesis. Since the lensed
images are very near the core of E0454 we subtracted a model for
the lens galaxy light before obtaining usable photometry. Thereto we
fit an elliptic Sersic model to the $u^*BVRIz$ data using GALFIT
\citep{phi02}. The resulting images and source fluxes are shown in
Figs. \ref{cl0454_galfit} and \ref{lens_colours}. We found good
agreement confirming the lens modelling. Only the $V$-band flux of the
southern arc appears too low, which is a consequence of the worse
seeing in this filter and the fact that this image is closest to the
lens making it very susceptible to over-subtraction effects.
\subsection{\label{piemd}PIEMD and stellar velocity dispersions}
The equivalent Einstein radius of the PIEMD reads
\begin{equation}
\theta_{\rm E}^{\rm\,PIEMD}= 2 b\, \frac{\sqrt{q}}{1+q}
\end{equation}
\citep[e.g.][]{KoopmansEtal06} and evaluates to
$2\myarcsec37\pm0\myarcsec04$. These authors have also shown that
$\theta_{\rm E}^{\rm PIEMD}$ corresponds to that of a classical
spherically symmetric SIS,
\begin{equation}
\theta_E=4 \pi \left(\frac{\sigma_v}{c}\right)^2 \frac{D_{\mathrm{ls}}}{D_{\mathrm{s}}}\,,
\end{equation}
yielding a PIEMD velocity dispersion of
$\sigma_v^{\rm sl}=319\pm4\rm ~km~s^{-1}$ for a source redshift
of $z_{\mathrm{s}}=2.1\pm0.3$.
\begin{figure}[t]
\includegraphics[width=1.0\hsize]{13810_f14.eps}
\caption{\label{cl0454_galfit}The arc system after subtracting an
elliptic Sersic model from the CFHT $u^*$-band and the Subaru
$BVRIz$-band images. The lens is not entirely removed. The counter
image is marked with a box. North is up and East is left,
the image width is 27$^{\prime\prime}$.}
\end{figure}
\begin{table}[t]
\caption{Marginalised strong lens parameters. The position angles (PA)
are counted from North to East. The uncertainties on the parameters
correspond to the 68\% posterior credible interval.}
\label{tab:SLparams}
\begin{center}
\begin{tabular}{l c}
\hline
\hline
$\gamma_{\rm ext}$ & $0.12\pm0.02$ \\
$PA_{\rm ext}$ & $91_{-3}^{+5}$ [deg] \\
\hline
$b$ & $2.38_{-0.05}^{+0.03}$ [arcsec]\\
$q$ & $0.80_{-0.08}^{+0.06}$\\
$PA_{\rm L}$ & $161_{-6}^{+9}$ [deg]\\
\hline
\noalign{\smallskip}
\end{tabular}\\
\end{center}
\end{table}
\cite{elr07} provide a recipe through which the PIEMD strong
lensing velocity dispersion can be linked to the observed stellar
velocity dispersion $\sigma^*$. Their mass model is a
parametrised truncated PIEMD (hereafter: dPIE) with core radius $a$
and scale radius $s$, which becomes identical to our model in the
limit of $a\rightarrow0$ and $s\rightarrow\infty$. We obtain
\begin{equation}
\sigma_{\rm dPIE}=\left(\frac{2}{3}\frac{c^2}{4 \pi}
\frac{D_{\mathrm{s}}}{D_{\mathrm{ls}}}\,\theta_{\rm E}^{\rm\,PIEMD}\right)^{1/2}= 260\pm4\rm ~km~s^{-1}\,.
\end{equation}
The relation between $\sigma_{\rm dPIE}$ and $\sigma^*$ is shown in
Fig. 20 of \cite{elr07}. Our PIEMD corresponds to their asymptotic
limit, approximated by the solid line in that figure. The radius $R$
($3$ FORS2 detector rows or 0\myarcsec6) within which we measured
$\sigma^*$ is significantly smaller than their effective radius $R_e$
(half mass radius), hence $R/R_e<<1$. We thus expect $\sigma^*$ to be
between $0.95\,\sigma_{\rm dPIE}$ and $1.15\,\sigma_{\rm dPIE}$ or
$250-300\rm ~km~s^{-1}$. A fit of the Doppler-broadened NaD absorption doublet
yields $\sigma^*=210\pm80\rm ~km~s^{-1}$, indicating that E0454 possibly does
not contain all the lensing mass. Up to 50\% could be located in
the halo of J0454 \citep[see also][for a similar example]{mak09}, but
the large error bars do not allow a definite conclusion.
These considerations depend on the arc redshift which
could not be determined unambiguously from spectroscopy (see
Sect. \ref{vltfors2}). For $z_{\rm arc}=0.4$ (1.0) we would have
$\sigma_{\rm dPIE}=515\,\rm ~km~s^{-1}$ ($345\,\rm ~km~s^{-1}$) and similarly high stellar
velocity dispersions. Thus $z_{\rm arc}=0.4$ is clearly ruled out by
the velocity dispersion measured ($\sigma^*=210\pm80\rm ~km~s^{-1}$),
and also by the fundamental plane properties of BCG galaxies
\citep{dqm07}. The latter predict $\sigma^*=280\pm35\rm ~km~s^{-1}$ for a BCG
with the absolute $I$-band luminosity of E0454 ($-24.1$ mag). A
redshift of $z_{\rm arc}\sim1.0$ would still be permitted within
$2\sigma$ errors, but the redshift most consistent with the data are
$z_{\rm arc}=2.1\pm0.3$, as assumed throughout our paper.
\begin{figure}[t]
\includegraphics[width=1.0\hsize]{13810_f15.ps}
\caption{\label{lens_colours}Magnitudes of the arc components and the
counter image, obtained from the images in
Fig. \ref{cl0454_galfit}. The uncertainties in the values are of the
order of $0.2$ mag due to the residuals from the removal of
E0454.}
\end{figure}
\section{\label{interpretation}Interpretation and discussion}
\subsection{\label{masscentroid}E0454 is not at the centre of the dark
matter potential}
Strong gravitational lensing by galaxy groups is very sensitive to the
local group environment, and in particular to the internal
distribution of dark matter \citep{mwk06}. During the build-up
of the morphology-density relation individual dark matter haloes get
partially stripped and integrated into the group halo. In these
systems, strong lensing can occur by individual galaxies with typical
Einstein radii of $\theta_E=1^{\prime\prime}-2^{\prime\prime}$, but
also by the common and more massive group halo with
$\theta_E=3^{\prime\prime}-8^{\prime\prime}$
\citep[see e.g.][]{fka08,lcg09}.
The Einstein radius for the $476\pm46 \rm ~km~s^{-1}$ weak lensing SIS
halo of J0454 and a source redshift of $z_s=2.1\pm0.3$ is
\begin{equation}
\theta_E=4 \pi \left(\frac{\sigma_v}{c}\right)^2\,\frac{D_{\mathrm{ls}}}{D_{\mathrm{s}}}
=5\myarcsec28\pm1\myarcsec02\,,
\end{equation}
more than twice as large as the observed arc radius of $2\myarcsec37$.
Hence the strong lens effect is caused by E0454 and not by the more
massive group halo. Given that all masses along the line of sight
contribute to the lensing potential, the immediate consequence of this
observation is that E0454 cannot be, on a $3\sigma$-level, located at
the centre of the gravitational potential of J0454. Otherwise the arc
radius would be significantly larger.
This argument would not hold anymore if the lensed source was at
very low redshift (we assume that it is at high redshift, see Sects.
\ref{vltfors2} and \ref{piemd}). The Einstein radius
of $2\myarcsec37$ would be reproduced for $z_s=0.42$, but already for
slightly higher redshifts such as 0.5 (0.6) it would increase to
$2\myarcsec93$ ($3\myarcsec45$). But even if we overestimated the
source redshift significantly, E0454 could still not be located at
J0454's halo centre as we show in the following using the lensed image
configuration.
The lens modelling (Table \ref{tab:SLparams}) requires a large amount
of external shear, $\gamma_{\rm ext}=0.12\pm0.02$, which can be caused
by a different lens along the line of sight, but also by an offset of
E0454 in the halo of J0454. The SIS model for the background cluster
MS0451 predicts $\gamma_{\rm ext}=0.021$ with ${\rm PA}=108$ degrees.
To obtain the external shear required, we need additional
components whose net shear is $\gamma_{\rm ext}=0.100\pm0.017$ and
${\rm PA}=86.5\pm0.6$ degrees. Since there is no evidence for other
suitable lenses in the Keck spectra and the XMM-Newton data, this
signal must come from J0454 alone. Its centre of mass must be located
$22^{\prime\prime}\pm4^{\prime\prime}$ ($89\pm16$ kpc) south of E0454
for a SIS profile, and $31^{+30}_{-12}$ arcseconds ($126^{+122}_{-49}$
kpc) for NFW (see Fig. \ref{j0454_centreofmass}). The same
external shear could also be caused if the halo was at identical
distances to the North of E0454, but this is ruled out at the
$4\sigma$ level by the mass reconstruction and the peak finder, both
of which locate the centre of mass
$12^{\prime\prime}\pm5^{\prime\prime}$ south of E0454. This is
unlikely to be the real centre, as neither the mass reconstruction nor
the peak finder are able to resolve such substructures in the halo
for the given lensing signal-to-noise ratio. A much larger number
density of lensed background galaxies than $n=73$ arcmin$^{-2}$ would
be required for this purpose.
\begin{figure}[t]
\includegraphics[width=1.0\hsize,angle=90]{13810_f16.ps}
\caption{\label{j0454_centreofmass}
The various estimates for the halo centre of J0454: distribution of
elliptical galaxies, mass reconstruction and peak finder, and from
the external shear coming from an SIS or NFW profile. The arrow
marks a tidal feature in the optical halo of E0454.}
\end{figure}
The different offsets predicted by the spherically symmetric NFW and
SIS density profiles are obviously model-dependent. Furthermore, an
elliptical halo would increase the offset if the halo's projected
major axis pointed to E0454. The reason for this is the increased
projected mass and thus shear seen by the strongly lensed light
bundle towards J0454, putting the halo at larger separation from E0454
to satisfy the external shear constraint. Likewise a smaller offset
would result if the halo minor axis pointed towards E0454. The first
scenario is more likely as the distribution of red sequence galaxies
is significantly elongated North-South within $r_{200}$ (see Fig.
\ref{gal_spatialdist}), which is expected if the galaxies are
virialised within an elliptic halo. In addition, E0454 is elongated
along the same direction, and \cite{ojl09} show that the central
luminous red galaxies in clusters are preferentially aligned within
$\sim35$ degrees with their host dark matter haloes. The separations
between E0454 and the halo centre should therefore be regarded as
lower limits.
\subsubsection{Effect of sub-haloes on the external shear}
Individual galaxies can significantly affect strong lensing systems in
cluster environments. We modelled the red sequence galaxies with SIS
profiles to estimate their contribution to the external shear, using
the velocity dispersions predicted by the Faber-Jackson relation
from \cite{dqm07}. Including successively more galaxy
haloes going from the strong lens outwards, we find that the external
shear increases gradually and stabilises at $\gamma_{\rm ext}=0.036\pm0.056$
with $PA=76$ degrees. Only 5 galaxies within 0\myarcmin51 contribute
to the signal. The uncertainty in the shear is large and based on the
intrinsic scattering of the Faber-Jackson relation. If we
systematically increase the predicted velocity dispersions of all 5
galaxies by the $1\sigma$ range allowed by Faber-Jackson, then the
entire external shear can be explained by these haloes. In the other
extreme, by lowering the velocity dispersion by $1\sigma$, the
contribution to the external shear becomes zero.
We therefore expect halo substructures to contribute about 30\% to the
total external shear. Since this lowers the shear
coming from a smooth common group or cluster halo with
$\sigma\sim480 \rm ~km~s^{-1}$, the offset from E0454 to the centre of this halo
becomes larger. For example, in case of a SIS halo the separation
would increase from 22\hbox{$\;\!\!^{\prime\prime}\;$} to 40\hbox{$\;\!\!^{\prime\prime}\;$}. The position
angle, i.e. the location of the halo centre south of E0454, remains
unchanged. We ignore the smaller effects of substructure for the rest
of the paper as the main result, i.e. the presence of an offset of
E0454 with respect to the centre of the projected mass distribution,
remains unaffected.
\subsubsection{Interpretation: A group falling into a cluster, or a
filament collapsing onto a group?}
How can it be explained that E0454, which is significantly more
luminous and massive than all other member galaxies, is not located
at the minimum of the gravitational potential? Such offsets are not
uncommon for normal groups and clusters \citep{oeh01,lbk07,sby10},
but for old and evolved fossils they are unusual
and have not been reported previously. If all galaxies in J0454
had the same origin, then the observations are difficult to reconcile.
A simple solution would be that E0454 formed outside of J0454 in
a separate small group which is now falling into J0454. This could
also explain the velocity offset of $+240\rm ~km~s^{-1}$ ($+540\rm ~km~s^{-1}$) observed
between the BCG and the population of elliptical (spiral) galaxies
\citep[see Sect. \ref{velfield} and][]{zam98,oeh01}. We present
more support for this scenario below based on the properties of the
X-ray halo.
There is one observational difficulty in this picture. If the halo
into which the group is falling represents a fully formed NFW mass
distribution, then one would expect a galaxy or concentration of
galaxies at its centre whose brightness is reasonably scaled to the
halo mass. However, no such galaxies are seen. From this we
conclude that the halo is not fully assembled yet, or
decomposed into several sub-haloes in a filament projected along the
line of sight, mimicking a spherical system. In the latter case,
galaxies could be streaming along the filament onto the denser fossil
group, presenting an alternative interpretation of the system
(B. Fort, private communication). Unfortunately neither the strong nor
the weak lensing data allows us to distinguish between a filament and
a more spherical cluster. We pool both scenarios under the term
`infall hypothesis', indicating a fossil group still forming an object
of its own in a larger system (J0454).
\subsection{X-ray halo properties support infall hypothesis}
Whereas the strong lensing data require an offset of $90-120$
kpc between E0454 and the halo of J0454, only a weakly
significant offset of $24\pm16$ kpc exists between E0454 and the X-ray
halo. The latter appears to be gravitationally bound by E0454 and not
by J0454. Even though the X-ray halo overlaps in projection with the
presumed core of J0454, no significant mass transfer has happened yet
as the X-ray halo appears undisturbed. This can be explained if E0454
still forms a local minimum and thus a system of its own in the larger
gravitational potential of J0454. The significantly lower X-ray mass of
$0.34\times10^{14}\,{\rm M_\odot}$ as compared to
$(0.75-0.90)\times10^{14}\,{\rm M_\odot}$ from galaxy counts and weak
lensing supports this interpretation. The same holds for the
$\beta$-model velocity dispersion ($316\pm26 \rm ~km~s^{-1}$) which matches the
strong lensing value ($319\pm4 \rm ~km~s^{-1}$) much better than the one derived
from weak lensing ($476\pm46 \rm ~km~s^{-1}$). The halo temperature of 1.1 keV
is also more characteristic for a group with $\sim330\rm ~km~s^{-1}$ than the
optically determined overall $480 \rm ~km~s^{-1}$. If E0454 indeed represents a
very evolved and virialised former group of galaxies that is now
falling into J0454 (Sect. \ref{masscentroid}), then one would expect
the X-ray properties to reflect a less massive and smaller system than
J0454. Given the undisturbed X-ray halo, the absence of shock fronts
and the low temperature, we presume that E0454 has not yet passed
through J0454 and is thus inbound for the first time. The present
data does not allow us to infer more information about the
three-dimensional orientation of the trajectory. To this
end we would need a better sampling of the peculiar motions of the
member galaxies in J0454.
\subsection{X-ray offsets in other strong lensing or fossil groups}
The offset of $24\pm16$ kpc between the X-ray centroid and E0454 is
consistent with those of other groups. Four of the strong lensing
selected systems by \cite{fka08} have X-ray haloes, coinciding within
$25-50\,h_{100}^{-1}$ kpc with the brightest group galaxy (BGG). Even
smaller offsets have been observed for the five fossil groups in
\cite{kpj06}, where the X-ray centroids of four systems match those of
the BGGs. For the fifth system, RXJ1552.2+2013, an offset of 12 kpc is
reported, but the authors argue that it is unlikely to be real. Fossil
samples that were cross-matched with the ROSAT All-Sky Survey
\citep{vab99} such as those by \cite{sms07} and \cite{bcr09} show
larger offsets of up to 50 and 90 kpc, respectively. These should be
interpreted with caution though due to the poor angular resolution of
ROSAT. In general, X-ray selected galaxy groups often show small
offsets, but in dynamically disturbed or merging systems they can
become larger than 100 kpc \citep{jml06,jml07}.
Whether small X-ray offsets are representative for strong lensing
X-ray groups is currently difficult to answer due to the small
sample size and possible selection effects. For example, the groups
from \cite{fka08} could be biased towards systems for which the X-ray
centroid and the BGG coincide, as this would boost the central density
giving rise to strong lensing. On the other hand, groups with more
complicated dynamical states (and larger X-ray offsets) have increased
lensing cross-sections and should therefore be selected as well.
\subsection{Comparison with other strong lensing groups}
\cite{lcg09} use an automated algorithm to detect strong
lensing features. They were looking for Einstein radii larger than
$3^{\prime\prime}$, targetting group-scale strong lenses, and found 13
such systems. The authors also report weak lensing measurements of the
velocity dispersions in the range of $500-800\rm ~km~s^{-1}$. A comparison of
the weak and strong lensing Einstein radii is not made, but the
according values have been tabulated. In general there seems to be
good agreement between the two estimates if the group halo is
responsible for the lensing. About half of the groups show
significantly larger strong lensing Einstein radii, most likely
systems where the lensing has been boosted by the potential of an
individual galaxy in addition to the group halo. Due to the lower
cut-off in $\theta_E$ objects like J0454 with large weak and small
strong lensing Einstein radii are filtered out. A survey aiming at
smaller strong lensing features near the core of the BGG could
identify systems similar to J0454. In combination with a high
external shear this would be a prime indicator for substructure and a
possible infall.
\subsection{Spatial and dynamic misalignment of E0454}
More evidence for the infall hypothesis arises when looking at the
sample of seven groups selected by \cite{fka08} for their strong
lensing effects. These authors found that the BGG almost always
coincides with the spatial and the dynamical group
centre. E0454 on the other hand is marginally consistent within
$1\sigma$ with the centre of the distribution of elliptical
galaxies. In addition, its velocity deviates by $2.5\sigma$
(half the velocity dispersion) from the mean recession velocity of
the ellipticals, and even more so from that of the spiral galaxies
(comparable to the velocity dispersion, see Sect. \ref{velfield}).
E0454 also contradicts the nine X-ray selected groups and poor
clusters of \cite{mlf06}, who found that BGGs coinciding with the
X-ray centroid have the same mean recession velocity as the
surrounding group. It would be worthwhile to look for similar
deviations in the currently existing samples of fossil groups.
With a more complete sampling of velocities of the elliptical
galaxies we could analyse these deviations in more detail, possibly
identifying a dynamic sub-population of galaxies belonging to the
fossil group (or, if we summon our alternative interpretation,
identify galaxies in the filament streaming towards the group).
With the data at hand we cannot estimate how many galaxies
comprise the fossil group. A tidal feature in E0454's
optical halo (see Fig. \ref{j0454_centreofmass}) indicates that the
accretion process in the fossil component of J0454 has not yet
finished, and therefore it is plausible that E0454 is not the only
galaxy belonging to that component.
\begin{table}[t]
\caption{Summary of the main results. Values in bold face are the
primary measurements, the others were derived from these.}
\label{resultstable}
\begin{tabular}{l l l l}
\hline
\hline
\tiny
Method & $\sigma_v$ [$\rm ~km~s^{-1}$]& $r_{200}$ [kpc] & $M_{200}$ [$10^{14}\,{\rm M_\odot}$]\\
\hline
\noalign{\smallskip}
Galaxy counts & -- & \boldmath$811\pm46$ & $0.74\pm0.15$\\
Spectr. (early type) & \boldmath$480\pm20$ & $1054\pm44$ & $1.69\pm0.14$\\
Spectr. (late type) & \boldmath$590\pm20$ & $1295\pm44$ & $3.14\pm0.21$\\
X-ray ($\beta$-model) & \boldmath$316\pm26$ & $617\pm28$ & $0.34\pm0.10$\\
Weak lens. (SIS) & \boldmath$476\pm46$ & $853\pm82$ & $0.90\pm0.26$\\
Weak lens. (NFW) & -- & \boldmath$834\pm219$ & $0.84\pm0.66$\\
Weak lens. (MR) & -- & $650\pm50\,^*$ & \boldmath$0.38\pm0.09\,^*$\\
Strong lensing & \boldmath$319\pm4$ & $700\pm9$ & $0.50\pm0.01$\\
\hline
\noalign{\smallskip}
\end{tabular}\\
$^*$ The mass estimate from the weak lensing mass reconstruction is
obtained within a radius of 182 kpc, significantly smaller than
$r_{200}$. The given values are therefore lower limits.
\end{table}
\subsection{Dynamic disturbances, X-ray offsets and cooling flows}
Dynamically disturbed haloes can suppress or reheat cooling cores, as
has been shown by \cite{ses09} for the 65 systems in the Local Cluster
Substructure Survey (LoCuSS, median redshift $z=0.23$). They
demonstrated that for clusters without cooling core or with inactive
BCGs the probability distribution function of the projected offset
between the X-ray centroid and the BCG peaks between 40 and 60
kpc. Conversely, cooling core clusters never showed
offsets larger than 15 kpc. With an offset of 24 kpc and no traces of
star formation in the FORS2/VLT spectrum, E0454 matches the LoCuSS
observations. In addition there are also hints for dynamic
disturbances. E0454 is embedded in an extended optical halo (see
Fig. \ref{j0454_centreofmass}), forming a tidal tail in the North
at a distance of about 90 kpc from the core. This
feature is too small and too localised to be attributed to a current
major merger event. It could be a residual of a recently disrupted
small companion galaxy, or caused by tidal disturbances of galaxies
orbiting very close to the halo (objects \#2, 3 and 4 from Table
\ref{galsample}). A similar feature is observed in the BCG of the
fossil cluster ESO 3060170 \citep{sfv04}.
E0454 is currently accreting mass, but the rate is too small to have
destroyed a previously existing cool core. \cite{bhg08} have
analysed the survival rate of cool cores in merger simulations and
found that non-cooling core clusters experienced a high accretion rate
with major mergers at $z>0.5$, destroying a potentially existing
cool core and also prevent their later reformation. Minor, and in
particular late accretion events such as the one observed in E0454
do not suppress a cool core. Thus, if a cooling core were present
in E0454, it must have been destroyed at early times. This is
consistent with E0454 being a giant elliptical galaxy, as only these
galaxies have experienced major mergers in their history \citep{pef09},
and fossil systems in particular form their halos at earlier times
than other groups \citep{dkp07,bog08}. A systematic survey of fossils
regarding cooling flows and cool cores would be helpful in testing
theoretical predictions. So far, the number of examined systems is
small \citep[see e.g.][]{sfv04,kjp04,kmp06} and the amount of
fossils with suitable X-ray data poor. This also applies to J0454, for
which the data are insufficient to derive a temperature map with
sufficient resolution to establish the non-existence of a cooling
core.
\subsection{Mass-to-light ratio}
\cite{sjm09} have obtained mass-to-light ratios for the maxBCG cluster
sample in SDSS, using weak lensing to estimate $M_{200}$. The total
luminosity within $r_{200}$ was inferred from red sequence galaxies
only. To minimise K-corrections, it was calculated for the
$i$-band bandpass shifted to the median cluster redshift of $0.25$. We
adopt their terminology and refer to the shifted bandpass as
$^{0.25}i$. A minimum $^{0.25}i$-band luminosity of
$10^{9.5}\,h_{100}^{-2}\,{\rm L_\odot}$ was required for each galaxy.
From the SIS and NFW fit to the tangential shear profile we obtained
an average $\langle r_{200}\rangle=843$ kpc, corresponding to
3\myarcmin47. Within this radius are 14 and 22 red sequence galaxies
with and without spectroscopic confirmation above the minimum
luminosity threshold. To correct for the field contamination
determined in Sect. \ref{contamination}, we randomly selected a
corresponding number of 5 galaxies from the sample without
spectroscopic redshifts and calculated their total flux
contribution. This was repeated 100 times to estimate the average
background correction. The total luminosity found is
$L_{^{0.25}i}^{\rm tot}=(6.9\pm0.6)\times10^{11}\,h^{-2}{\rm L_\odot}$, and
$M_{200}/L_{^{0.25}i}=130\pm39\,h$ for the SIS profile and
$M_{200}/L_{^{0.25}i}=122\pm96\,h$ for NFW. In the rest-frame bandpass
the $M/L$ ratios would be 8\% lower. For a cluster with the same
number of $N_{200}$ (Sect. \ref{r200galcounts}) galaxies as J0454,
\cite{sjm09} predict
$\langle M_{200}/L_{^{0.25}i}\rangle=200\pm30\,h$.
Given the scatter present in the luminosity and the mass of
a given $N_{200}$ bin \citep[see][]{sjs09,sjm09}, J0454
does not appear exceptionally over-luminous compared to non-fossil
systems. The contribution of the BGG to the total
luminosity in $i$-band within $r_{200}$ is 38\%.
For completeness we also report the corresponding result for
rest-frame $B$-band and the SIS mass, $M_{200}/L_B=115\pm34\,h$. If we
include also galaxies bluer than the red sequence, the ratio becomes
$101\pm30\,h$. The contribution of the BGG to $L_B^{\rm tot}$
is 34\% for red sequence galaxies alone, decreasing to 29\% if late
type galaxies are included. The latter is an upper limit as the sample
of late types is incomplete. For the two fossil clusters
RXJ1416.4+2315 and RXJ1552.2+2013 the non-brightest cluster members
contribute only $\sim55\%$ of the flux of the BCG, i.e. the BCG
provides about 2/3 of $L_B^{\rm tot}$
\citep[see][]{kpj07,cms06,mcs06}.
\section{Summary and conclusions}
In deep ground-based Subaru/Suprime-Cam data we discovered a
galaxy strongly lensed by a very bright elliptical galaxy
(E0454). Using VLT/FORS2 spectroscopy we confirmed that E0454
is a member of a larger association of galaxies (J0454) at
$z=0.26$. The system forms a fossil group with a gap of 2.5 mag in
$I$-band between the brightest and the second brightest galaxy within
half the virial radius. We have spectroscopically confirmed the
membership of 31 galaxies, and furthermore selected 33 objects based
on their photometric properties. Our catalogue is complete down to
$i\leq22$ ($M_i=-18.6$).
The data, being the deepest so far for a fossil group, show that J0454
is a complex system in various stages of mass assembly. Stripping away
the layers from outside to inside, we find two filaments extending 4
Mpc from J0454. Within a projected distance of 1.5 Mpc of the centre
is a population of spirals with $\sigma_v=590\rm ~km~s^{-1}$, surrounding
a more concentrated and dynamically cooler group of $\sim50$
galaxies ($\sigma_v=480\rm ~km~s^{-1}$). These form a red sequence with an
intrinsic width of $\sigma=0.049$.
Using HST/ACS and photometric redshifts we performed the
first weak lensing analysis for a fossil group. The tangential shear
profile yields $r_{200}\sim840$ kpc and
$M_{200}\sim0.85\times10^{14}\,{\rm M_\odot}$, fully consistent
with the predictions made by the cluster size-richness relation of
\cite{hsw09}. From this point of view J0454 is indistinguishable from
normal clusters, forming either a rich fossil group or a poor fossil
cluster. The X-ray halo can be described by a classic $\beta$-model
and is only marginally offset ($24$ kpc) from the brightest group
galaxy. However, the velocity dispersion of $316\rm ~km~s^{-1}$ is lower
than the one measured from weak lensing ($476\rm ~km~s^{-1}$) and
spectroscopy ($480\rm ~km~s^{-1}$), and so is
$M_{200}$ ($0.34\times10^{14}\,{\rm M_\odot}$). The low X-ray halo
temperature of $1.1$ keV also favours a smaller structure with
$\sim330 \rm ~km~s^{-1}$.
Peculiarities arise when analysing the brightest group galaxy with
respect to its environment. It not located at the spatial centre of
elliptical galaxies and shows a significantly different velocity than
the mean velocity of the ellipticals. This indicates a different
origin of E0454 from the surrounding galaxies. More
evidence for this hypothesis is provided by the strongly lensed galaxy
near the core of E0454. We constrained its redshift to $z=2.1\pm0.3$
and determined an Einstein radius of $2\myarcsec37$. The weak lensing
velocity dispersion of $476\rm ~km~s^{-1}$ corresponds to an Einstein radius of
$\theta_E=5\myarcsec28$, meaning that E0454 cannot be located at
the centre of the dark matter halo of J0454. Even stronger evidence
comes from the external shear required to fit the position
of the counter image. About 15\% of the shear can be attributed
to the background cluster MS0451, and about 30\% to individual
galaxies near E0454, but the dominant contribution must come from
J0454 itself. This can only be explained if E0454 is not at
the centre of the gravitational potential. If we describe the density
profile with NFW, then the projected distance between E0454 and the
halo centre must be at least $\sim120$ kpc. Whereas such offsets have
been shown to be common for other groups and clusters
\citep{oeh01,lbk07,sby10} this has not yet been reported for fossils.
An explanation that reconciles all observations is that E0454 is
currently infalling for the first time into the sparse cluster J0454,
seeding the brightest cluster galaxy. An alternative interpretation is
that J0454 is of filamentary nature, projected along the line
of sight, and galaxies therein stream towards the denser fossil core.
Both scenarios explain why the X-ray halo appears associated with E0454,
has undisturbed isophotes, no shock fronts, a low temperature
and a velocity dispersion and mass that fits a smaller group. This
hypothesis is only possible because of the presence and properties of
the strong lens, ruling out that E0454 is at the gravitational
centre. Without the lens all data would form a consistent picture.
Recently, \cite{lcm10} have demonstrated for the fossil UGC 842
that it segregates into two groups with $\sigma_v\sim220 \rm ~km~s^{-1}$ each,
separated by about $820 \rm ~km~s^{-1}$. Contrary to J0454 with a comparatively
low temperature of 1.1 keV, UGC 842 shows a, with respect to the
velocity dispersion, increased temperature of 1.9 keV, which has been
interpreted as a sign of an advanced interaction or merging state.
\subsection{Future observations}
Our Subaru/Suprime-Cam images are significantly deeper than those of
any other fossil system investigated so far, reaching
$10\sigma$-limits of 24.9 in $z$-band down to 26.6 in $B$-band. Hence
this data set could probe the luminosity function $\sim9$ magnitudes
below the BGG, provided deep spectroscopic data are available to
remove objects with very similar redshifts. The existing
spectra are limited to $I<21.5$ mag and
complete to only about 40\% at this depth. Hence we limited our
analysis to galaxies with $I<22$ mag. With several hours of exposure
time at $4-8$m telescopes we could push the spectroscopic limit by
$\sim2.5$ magnitudes, which would enable us to present an
uncontaminated luminosity function extending down to dwarf
ellipticals. Numerous of those are seen in the data with colours
matching that of the red sequence. With a better spectroscopic
sampling we could remove all line of sight contamination and construct
a complete red sequence down to much fainter magnitudes, and a fairly
complete sample of blue galaxies. There are also possibilities that we
could resolve J0454 from a dynamical point of view into members
belonging to the fossil component, and into galaxies belonging to the
sparse surrounding. If both components have indeed separate origins,
then one could attempt to identify stellar populations of different
age and composition.
Significantly deeper X-ray data could be used to better determine the
offset with respect to the BGG. We could also look for temperature
variations and changes in the chemical composition of the gas, which
would tell us more about the different origins of the sparse cluster
and the infalling group.
Lastly, deeper space-based observations could double the number
density of lensed galaxies and we could attempt to obtain direct
evidence for the separation of J0454 and E0454 in the mass
reconstructions. However, given the aged detectors of the HST/ACS
instrument this will be a difficult endeavour.
\begin{acknowledgements}
MS thanks Bodo Ziegler at ESO and the staff at Paranal for the
prompt execution of the DDT programme, and Helen Eckmiller, Bernard
Fort, Sarah Hansen, Stefan Hilbert, Satoshi Miyazaki and Achille
Nucita for their expertise and helpfulness concerning various aspects
of this work. Andrew Cardwell, Karianne Holhjem and
Peter Schneider provided very useful comments on the manuscript. We
thank the anonymous referee for very helpful suggestions that improved
the paper significantly. Author contributions: MS did the
scientific analysis, obtained the VLT spectrum, reduced
the Subaru, VLT and the XMM data, discovered the strong lens system
and wrote most parts of the manuscript. SS did the strong lens
modelling, based upon a code developed by herself and by AH. TS
reduced the HST/ACS data and provided the shear catalogue, while HH
complemented it with photometric redshifts. TE provided the stacks of
the {\tt Elixir} pre-processed CFHT $u^*griz$ images. Some figures in
this paper were made with the plotting tool WIP \citep{mor95}. The
authors wish to recognize and acknowledge the very significant
cultural role and reverence that the summit of Mauna Kea has always
had within the indigenous Hawaiian community. We are most fortunate
to have the opportunity to conduct observations from this mountain.
MS acknowledges support by the German Ministry for Science and
Education (BMBF) through DESY under the project 05AV5PDA/3 and the
Deutsche Forschungsgemeinschaft (DFG) in the frame of the
Schwerpunktprogramm SPP 1177 `Galaxy Evolution'. SS is supported in
part through the DFG under project SCHN 342/7-1. TS acknowledges
financial support from the Netherlands Organisation for Scientific
Research (NWO). HH is supported by DUEL-RTN, MRTN-CT-2006-036133, and
AH by the DFG cluster of excellence `Origin and Structure of the
Universe'.
\end{acknowledgements}
|
\section{Introduction}\label{intro}
The two-by-two $ABCD$ matrix plays a central role in optical
sciences. The four elements of this matrix are all real, and its
determinant is one. Thus, it has three independent parameters.
Yet, it has a rich mathematical content which could lead
interesting results in physics.
It is generally assumed that this matrix can be diagonalized by
a rotation, but this is not the case as shown in our previous
paper~\cite{bk09josa}. We have shown there that this matrix can
be brought to an equi-diagonal matrix by a rotation, and then by
a squeeze to one of the following four Wigner matrices.
\begin{equation}\label{wmat11}
\pmatrix{\cos\theta & -\sin\theta \cr \sin\theta & \cos\theta},
\quad
\pmatrix{\cosh\lambda & \sinh\lambda \cr
\sinh\lambda & \cosh\lambda}, \quad
\pmatrix{1 & -\gamma \cr 0 & 1}, \quad
\pmatrix{1 & 0 \cr \gamma & 1}.
\end{equation}
This squeeze portion of the similarity transformation is not yet
widely known. Thus the similarity transformation which brings the
$ABCD$ matrix to one of the Wigner matrices is a rotation followed
by a squeeze. Even though the two triangular matrices in
Eq.(\ref{wmat11}) can be similarity-transformed from each other, it
is convenient to work with the four branches of the $ABCD$ matrix.
The purpose of this paper is to reduce these four matrices into one
analytic matrix with four different branches. First of all, each
of the matrices in Eq.(\ref{wmat11}) is generated by
\begin{equation}\label{gen01}
\frac{1}{2} \pmatrix{0 & -i \cr i & 0}, \qquad
\frac{1}{2} \pmatrix{0 & i \cr i & 0}, \qquad
\frac{1}{2} \pmatrix{0 & -i \cr 0 & 0}, \qquad
\frac{1}{2} \pmatrix{0 & 0 \cr i & 0} ,
\end{equation}
respectively. The last two matrices can be obtained from a linear
combination of the first two, with two independent coefficients.
We can then study the general property of the $ABCD$ matrix by
exponentiating the linear combination of the two matrices
\begin{equation}\label{gen02}
\pmatrix{0 & -i \cr i & 0} \quad \mbox{and} \quad
\pmatrix{0 & i \cr i & 0} ,
\end{equation}
and making Taylor expansions.
One of the present authors noted this aspect of the $ABCD$ matrix
while studying optical activities~\cite{kim09jmo}. He then
concluded that the asymmetric optical activity can lead to the
study of the fundamental space-time symmetries of elementary
particles~\cite{wig39,knp86}.
In this paper, we study the resulting exponential form more
systematically. We first exponentiate the linear combination
of these two independent matrices. While the exponent is a fully
analytic function, the Taylor expansion of the exponential form
leads to complications, leading to four separate branches.
Again in this paper, we use the same optical activity to study
the origin of this branching property. We note then that the
exponential form is convenient for repeated applications of the
$ABCD$ matrix, such as periodic systems including laser cavities
and multi-layer optics.
In Sec.~\ref{branch}, we discuss how the $ABCD$ matrix can be
written as an exponential function of one analytic matrix, with
four branches. In Sec.~\ref{acti}, we use optical activities
to study the physics of the mathematics of Sec.~\ref{branch}.
Section~\ref{periodic} is devoted to application of this methods
to periodic systems. Laser cavities and multilayer optics
are discussed in detail.
\section{Exponential Form and Branches}\label{branch}
Let us start with the $ABCD$ matrix as a rotated equi-diagonal
$abcd$ matrix:
\begin{equation}\label{abcd11}
[ABCD] = R(\alpha) [abcd] R(-\alpha) ,
\end{equation}
where $R(\alpha)$ is a rotation matrix
\begin{equation}\label{rot11}
R(\alpha) = \pmatrix{\cos(\alpha/2) & -\sin(\alpha/2) \cr
\sin(\alpha/2) & \cos(\alpha/2)} ,
\end{equation}
and $[abcd]$ is an equi-diagonal matrix
\begin{equation}
[abcd] = \pmatrix{a & b \cr c & d} ,
\end{equation}
with
\begin{eqnarray}\label{alpha}
&{}& \tan\alpha = \frac{D - A}{B + C}, \nonumber\\[1ex]
&{}& a = d = \frac{A + B}{2} , \nonumber\\[1ex]
&{}& b = \frac{(B - C) + \sqrt{(A - D)^2 + (B + C)^2}}{2} ,
\nonumber\\[1ex]
&{}& c = \frac{(C - B) + \sqrt{(A - D)^2 + (B + C)^2}}{2} .
\end{eqnarray}
Since the determinant of the $ABCD$ matrix is assumed to be one,
this matrix has three independent parameters. One of those
parameters is the angle $\alpha$. Thus, the $abcd$ matrix has
two independent parameters. Since the two diagonal elements
of the $abcd$ matrix are the same, it can be exponentiated as
\begin{equation}\label{exp01}
[abcd] = \exp{\left\{r M(\theta)\right\}} ,
\end{equation}
with
\begin{equation}\label{m11}
M(\theta) = \pmatrix{0 & -\cos\theta +
\sin\theta \cr \cos\theta + \sin\theta & 0} .
\end{equation}
Here the two independent parameters are $r$ and $\theta$.
Thus, we are led to study in detail this $M(\theta)$ matrix
which can also be written as
\begin{equation}\label{gen03}
M(\theta) = \pmatrix{0 & -1\cr 1 & 0} \cos\theta +
\pmatrix{0 & 1\cr 1 & 0} \sin\theta .
\end{equation}
Other than the factor of $(i/2)$, this expression becomes the four
generators given in Eq.(\ref{gen01}) when $\theta = 0, \pi/2,
\pi/4, -\pi/4 $ respectively.
\begin{figure}[thb]
\centerline{\includegraphics[scale=0.5]{conic.eps}}
\vspace{6mm}
\caption{Forms of the $ABCD$ matrix depending on the angle
$\theta.$}\label{conic}
\end{figure}
In this way, we can combine the four Wigner matrices into an
exponential function of one analytic matrix. The problem is how
to compute the exponential form of Eq.(\ref{exp01}). Its Taylor
expansion is
\begin{equation}\label{taylor11}
[abcd] = \sum_{n} \frac{r^n}{n!} [M(\theta)]^n .
\end{equation}
This is an infinite series except at $\theta = \pm \pi/4$. If
$\theta = 45^o$, the $M$ matrix becomes
\begin{equation}
M = \pmatrix{0 & 0 \cr \sqrt{2} & 0} .
\end{equation}
Since $M^2 = 0$, the series truncates. The $abcd$ matrix becomes
\begin{equation}\label{trian11}
[abcd] = \pmatrix{1 & 0 \cr r\sqrt{2} & 1} .
\end{equation}
Likewise, when $\theta = -\pi/4$,
\begin{equation}\label{trian22}
[abcd] = \pmatrix{1 & -r\sqrt{2} \cr 0 & 1} .
\end{equation}
This aspect is illustrated in Fig.~\ref{conic}. The Taylor series
truncates at $\theta = \pm \pi/4$. In the circular regions 1 and
2, $(\sin\theta)^2$ is smaller than $(\cos\theta)^2$. In the
hyperbolic regions 1 and 2, $(\cos\theta)^2$ is smaller than
$(\sin\theta)^2$.
Also in Fig.~\ref{conic}, the symmetry of trigonometry tells us
the region $\pi/2 < \theta < 3\pi/2 $ is redundant if we allow
both positive and negative values of $r$ in Eq.(\ref{exp01}).
In this way, we restrict $\cos\theta$ to positive values. In
circular region 1, $\sin\theta$ can be both positive or negative.
In the region $(\cos\theta)^2 > (\sin\theta)^2,$ and $ |\theta| < \pi/4$,
if $\sin\theta$ is positive, we can write the $M$ matrix as
\begin{equation}
M(\theta) = \sqrt{\cos(2\theta)}
\pmatrix{0 & - \exp{(-\eta)} \cr \exp{(\eta)} & 0} .
\end{equation}
with
\begin{equation}\label{eta11}
\exp(-\eta) = \sqrt{\frac{\cos\theta - \sin\theta}
{\cos\theta + \sin\theta}} ,
\end{equation}
where $\eta$ is positive.
This formula is valid also when $\sin\theta$ is negative, but
$\eta$ is also negative.
We now write $M(\theta)$ as
\begin{equation}
M(\theta) = \sqrt{\cos(2\theta)}
\pmatrix{e^{-\eta/2} & 0 \cr 0 & e^{\eta/2}}
\pmatrix{0 & - 1 \cr 1 & 0}
\pmatrix{e^{\eta/2} & 0 \cr 0 & e^{-\eta/2}} .
\end{equation}
Then
\begin{equation}
\left(M(\theta)\right)^n = \left(\cos(2\theta)\right)^{n/2}
\pmatrix{e^{-\eta/2} & 0 \cr 0 & e^{\eta/2}}
\pmatrix{0 & - 1 \cr 1 & 0}^n
\pmatrix{e^{\eta/2} & 0 \cr 0 & e^{-\eta/2}} .
\end{equation}
Thus
\begin{equation}
[abcd] = \pmatrix{e^{-\eta/2} & 0 \cr 0 & e^{\eta/2}}
\pmatrix{\cos\phi & -\sin\phi \cr \sin\phi & \cos\phi}
\pmatrix{e^{\eta/2} & 0 \cr 0 & e^{-\eta/2}} ,
\end{equation}
which is
\begin{equation} \label{mat11}
[abcd] = \pmatrix{\cos\phi & -e^{-\eta}\sin\phi \cr
e^{\eta}\sin\phi & \cos\phi}
\end{equation}
with
\begin{equation}\label{phi11}
\phi = r\sqrt{\cos(2\theta)} .
\end{equation}
In terms of the four parameters of the $ABCD$ matrix,
\begin{eqnarray}\label{phi22}
&{}& \cos\phi = \frac{A + B}{2} , \nonumber\\[1ex]
&{}& e^{-2\eta} = \frac{-b}{c} =
\frac{C - B - \sqrt{(B + C)^2 + (A - D)^2}}
{C - B + \sqrt{(B + C)^2 + (A - D)^2}} .
\end{eqnarray}
where $\cos\phi$ is smaller than one, and $b$ is negative.
If $(\sin\theta)^2 >(\cos\theta)^2$, or
$ \pi/4 < |\theta| < \pi/2 $, we have to consider two separate
regions in Fig.~\ref{conic}, where $\cos\theta$ is positive, while
$\sin\theta$ can take both positive and negative signs.
$\cos(2\theta)$ is negative. The $M$ matrix should becomes
\begin{equation}\label{m22}
M(\theta) = \sqrt{-\cos(2\theta)}
\pmatrix{0 & \exp{(-\eta)} \cr \exp{(\eta)} & 0} .
\end{equation}
with
\begin{equation}\label{eta22}
\exp(-\eta) =\sqrt{\frac{\sin\theta - \cos\theta}
{\sin\theta + \cos\theta}} ,
\end{equation}
for both positive and negative values $\sin\theta$, but $\eta$ is
positive and is negative respectively.
Then the Taylor expansion leads to
\begin{equation}\label{mat22}
[abcd] = \pmatrix{\cosh\chi & e^{-\eta}\sinh\chi \cr
e^{\eta}\sinh\chi & \cosh\chi} ,
\end{equation}
with
\begin{equation}\label{chi11}
\chi = r \sqrt{-\cos(2\theta)} .
\end{equation}
In terms of the parameters of the original $ABCD$ matrix,
\begin{eqnarray}
&{}& \cosh\chi = \frac{A + B}{2} , \nonumber\\[1ex]
&{}& e^{-2\eta} = \frac{b}{c} = \frac{B - C + \sqrt{(B + C)^2 + (A - D)^2}}
{C - B + \sqrt{(B + C)^2 + (A - D)^2}} .
\end{eqnarray}
Here $\cosh\chi$ is greater than one, and both $b$ and $c$ are positive.
We can now go back to Eq.(\ref{eta11}) and Eq.(\ref{eta22}), and write
$\tan\theta$ in terms of $\eta$. Then $\tan\theta$ becomes
\begin{equation}
\tan\theta = \frac{b + c}{c - b} =
\frac{\sqrt{(A - D)^2 + (B + C)^2}}{C - B} ,
\end{equation}
for all values of $\theta$ between $-\pi/2$ and $\pi/2$.
The parameter $r$ is
\begin{eqnarray}
r = \left[\frac{b^2 + c^2}{-2bc}\right]^{1/2}\phi, \nonumber \\[1ex]
r = \left[\frac{b^2 + c^2}{2bc}\right]^{1/2} \chi ,
\end{eqnarray}
for $(\sin\theta)^2 < (\cos\theta)^2$ and
$(\sin\theta)^2 > (\cos\theta)^2$ respectively, with
\begin{equation}
\left[\frac{b^2 + c^2} {2bc}\right]^{1/2}
= \left[\frac{2\left(B^2 + C^2\right) + (A - D)^2}
{4BC + (A - D)^2} \right]^{1/2}.
\end{equation}
Let us now look at how the transition of the $abcd$ from
Eq.~(\ref{mat11}) to Eq.~(\ref{mat22}). This is a puzzling question
because the matrix $M(\theta)$ remains analytic in the neighborhood
of $\theta = \pi/4 $ (see Fig.~\ref{conic}. In order to tackle this
problem, we write $M(\theta)$ of Eq.(\ref{m11}) as
\begin{equation}
M(\theta) = (\cos\theta)
\pmatrix{0 & -(1 - \tan\theta) \cr 1 + \tan\theta & 0} .
\end{equation}
In the neighborhood of $\theta = \pi/4$, we can set $\cos\theta
= 1/\sqrt{2}$ and $(1 + \tan\theta) = 2$, and
\begin{equation}
M(\theta) = \frac{1}{\sqrt2}
\pmatrix{0 & -(1 - \tan\theta) \cr 2 & 0} .
\end{equation}
Then up to $r^2$, the Taylor leads to
\begin{equation}
[abcd] = = \pmatrix{1 - r^2(1 - \tan\theta)/2
& -r(1 - \tan\theta)/\sqrt{2}
\cr r\sqrt{2} & 1 - r^2(1 - \tan\theta)/2} .
\end{equation}
If $\theta$ is smaller than $\pi/4$, the diagonal elements of this
matrix are smaller than $1$, like $\cos\phi$ in Eq.(\ref{mat11}).
If $\theta$ becomes greater than $\pi/4$, the diagonal element
becomes greater than $1$ like $\cosh\chi$ in Eq.(\ref{mat22}).
If $\tan\theta = 1$, the result becomes that of
Eq.(\ref{trian11}).
We can give a similar reasoning for the neighborhood of
$\tan\theta = -1$. The Taylor expansion leads to
\begin{equation}
[abcd] =
\pmatrix{1 - r^2(1 + \tan\theta)/2 & -r\sqrt{2} \cr
r(1 + \tan\theta)/\sqrt{2} & 1 - r^2(1 + \tan\theta)/2} .
\end{equation}
leading to Eq.(\ref{trian22}) for $\theta= -\pi/4.$
The exponential form given in Eq.(\ref{exp01}) is very convenient
when we study periodic systems where the $ABCD$ matrix is applied
repeatedly. We shall return to this problem in
Sec.~\ref{periodic}.
\section{Optical Activities}\label{acti}
In his recent paper~\cite{kim09jmo}, one of the present authors used
the two-by-two matrix formulation of optical activities applicable
to the transverse electric field of an optical wave. The direction
of the electric component rotates as the optical wave propagates.
In the real world, the medium causes also an attenuation of the
transverse components. This does not interfere with the rotational
character. However, there is a problem if the dissipation
coefficients are different for two perpendicular directions.
Let us start from a circularly polarized light wave which can be
decomposed into the right-polarized and left polarized components.
If they have different indexes of refraction, we can write the light
wave as
\begin{equation}
\pmatrix{E_{x} \cr E_{y}} =
\frac{1}{2}\pmatrix{ 1 \cr i}
\exp{\left\{i \left(k_1 z - \omega t \right)\right\}}
+ \frac{1}{2}\pmatrix{ 1 \cr -i}
\exp{\left\{i \left(k_2 z - \omega t \right)\right\}}
\end{equation}
This two terms can be combined into
\begin{equation}\label{ray11}
\pmatrix{E_{x} \cr E_{y}} =
\pmatrix{ \cos(\gamma z) \cr \sin(\gamma z)}
\exp{\left\{i \left(k z - \omega t \right)\right\}},
\end{equation}
with
\begin{equation}
k = \frac{1}{2}\left(k_1 + k_2\right) , \qquad
\gamma = \frac{k_1 -k_2}{2} .
\end{equation}
If we start with a polarized light wave taking the form
\begin{equation}
\pmatrix{E_{x} \cr E_{y}} =
\pmatrix{A \exp{\left\{i(kz - \omega t)\right\}} \cr 0} ,
\end{equation}
the optical activity is carried out by the rotation matrix
\begin{equation}\label{rot22}
R(\gamma z) = \pmatrix{\cos(\gamma z) & -\sin(\gamma z) \cr
\sin(\gamma z) & \cos(\gamma z)} .
\end{equation}
The optical ray is expected to be attenuated due to absorption
by the medium. The attenuation coefficient in one transverse
direction could be different from the coefficient along the
other direction. Thus, if the rate of attenuation along the
$x$ direction is different from that along $y$ axis, this
asymmetric attenuation can be described by
\begin{equation}\label{atten}
\pmatrix{\exp{\left(-\mu_{1}z \right)} & 0 \cr
0 & \exp{\left(-\mu_{2}z \right)}}
= e^{-\lambda z} \pmatrix{\exp{(\mu z)} & 0 \cr
0 & \exp{(-\mu z)}} ,
\end{equation}
with
\begin{equation}
\lambda = \frac{\mu_{2} + \mu_{1}}{2} , \qquad
\mu = \frac{\mu_{2} - \mu_{1}}{2} .
\end{equation}
The exponential factor $\exp{(-\lambda z)}$ is for the overall
attenuation, and the matrix
\begin{equation} \label{sq01}
\pmatrix{\exp{(\mu z)} & 0 \cr 0 & \exp{(-\mu z)}}
\end{equation}
performs a squeeze transformation. This matrix expands the $x$
component of the polarization, while contracting the $y$ component.
We shall call this the squeeze along the $x$ direction.
The squeeze does not have to be along the $x$ and $y$directions
For convenience, let us rotate the squeeze axis by $45^o$. Then
the squeeze matrix becomes
\begin{equation}\label{sq02}
S(\mu z) = \pmatrix{\cosh(\mu z) & \sinh(\mu z) \cr
\sinh(\mu z) & \cosh(\mu z) } .
\end{equation}
If this squeeze is followed by the rotation of Eq.(\ref{rot22}),
the net effect is
\begin{equation}
e^{-\lambda z} \pmatrix{\cos(\gamma z) & -\sin(\gamma z) \cr
\sin(\gamma z) & \cos(\gamma z)}
\pmatrix{\cosh(\mu z) & \sinh(\mu z) \cr
\sinh(\mu z) & \cosh(\mu z) } ,
\end{equation}
where $z$ is in a macroscopic scale, perhaps measured in
centimeters. However, this is not an accurate description
of the optical process.
This happens in a microscopic scale of $z/N$, and becomes
accumulated into the macroscopic scale of $z$ after the $N$
repetitions, where $N$ is a very large number. We are thus
led to the transformation matrix of the form
\begin{equation}\label{trans}
Z(\gamma,\mu,z)= \left[e^{-\lambda z/N}S(\mu z/N)
R(\gamma z/N)\right]^N .
\end{equation}
In the limit of large $N$, this quantity becomes
\begin{equation}
e^{-\lambda z} \left[\pmatrix{1 & \mu z/N \cr \mu z/N & 1}
\pmatrix{1 & - \gamma z/N \cr \gamma z/N & 1}\right]^N .
\end{equation}
Since $\gamma z/N$ and $\mu z/N$ are very small,
\begin{equation}\label{z11}
Z(\gamma,\mu,z)= e^{-\lambda z} \left[\pmatrix{1 & 0 \cr 0 & 1}
+ \pmatrix{0 & - \gamma + \mu \cr
\gamma + \mu & 0}\frac{z}{N} \right]^N .
\end{equation}
For large $N$, we can write this matrix as~\cite{kim09jmo}
\begin{equation}\label{expo22}
Z(\gamma,\mu, z) = e^{-\lambda z} \exp{\left\{kz M(\theta) \right\}} ,
\end{equation}
where the $M$ matrix is
\begin{equation}\label{m33}
M(\theta) = \pmatrix{0 & -\cos\theta + \sin\theta
\cr \cos\theta + \sin\theta & 0} ,
\end{equation}
with
\begin{eqnarray}\label{gm11}
&{}& k ={\sqrt{\gamma^2 + \mu^2}}, \nonumber \\[1ex]
&{}& \cos\theta = \frac{\gamma}{\sqrt{\gamma^2 + \mu^2}}, \nonumber \\[1ex]
&{}& \sin\theta = \frac{\mu}{\sqrt{\gamma^2 + \mu^2}}.
\end{eqnarray}
We note here that the $M(\theta)$ matrix of Eq.(\ref{m33}) is the
same as that of Eq.(\ref{m11}) which determines the branch property
of the $ABCD$ matrix. At this point, it is more convenient to work
with $kM(\theta)$.
\begin{equation}
kM(\theta) = \pmatrix{0 & -\gamma + \mu \cr \gamma + \mu & 0} .
\end{equation}
If $\gamma > \mu$, the $rM$ matrix can then be written as
\begin{equation}
kM = \sqrt{\gamma^2 - \mu^2}\pmatrix{0 & -e^{-\eta}
\cr e^{\eta} & 0},
\end{equation}
where $\eta$ of Eq.(\ref{eta11}) becomes
\begin{equation}
e^{-2\eta} = \sqrt{\frac{\gamma - \mu}{\gamma + \mu}} =
\sqrt{\frac{\cos\theta - \sin\theta}{\cos\theta + \sin\theta}} .
\end{equation}
Thus, the exponential function in Eq.(\ref{expo22}) can be evaluated
according to the procedure defined in Sec.~\ref{branch}. This
expression is the same as that of Eq.(\ref{eta11}).
The exponential form $\exp{(kzM)}$ in of Eq.(\ref{expo22})
becomes
\begin{equation}
\pmatrix{e^{-\eta/2} & 0 \cr 0 & e^{\eta/2}}
\pmatrix{\cos(\gamma' z) & -\sin(\gamma' z) \cr
\sin(\gamma'z) & \cos(\gamma' z)}
\pmatrix{e^{\eta/2} & 0 \cr 0 & e^{-\eta/2}} ,
\end{equation}
where
\begin{equation}\label{gammak}
\gamma' = \sqrt{\gamma^2 - \mu^2} .
\end{equation}
The transformation matrix of Eq.(\ref{expo22}) takes the form
\begin{equation}
Z(\gamma,\mu,z) = e^{-\lambda z}
\pmatrix{\cos(\gamma' z) & - e^{-\eta} \sin(\gamma' z) \cr
e^{\eta} \sin(\gamma' z) & \cos(\gamma' z)} ,
\end{equation}
If $\mu > \gamma$, the $rM$ matrix becomes
\begin{equation}
kM = \sqrt{\mu^2 - \gamma^2}\pmatrix{0 & e^{-\eta} \cr e^{\eta} & 0},
\end{equation}
where $\eta$ of Eq.(\ref{eta22}) takes the form
\begin{equation}
e^{-2\eta} = \sqrt{\frac{\mu - \gamma}{\mu + \gamma}} =
\sqrt{\frac{\sin\theta - \cos\theta}
{\cos\theta + \sin\theta}},
\end{equation}
and $Z$ becomes
\begin{equation}
Z(\gamma,\mu,z) = e^{-\lambda z}
\pmatrix{\cosh(\mu'z) & e^{-\eta} \sinh(\mu'z) \cr
e^{\eta} \sinh(\mu' z) & \cosh(\mu' z)} ,
\end{equation}
where
\begin{equation}\label{muk}
\mu' = \sqrt{\mu^2 - \gamma^2}.
\end{equation}
In this section, we discussed a system of optical activities with
asymmetric dissipation as a physical illustration of the
mathematical procedure discussed in Sec.~\ref{branch}. We have
already seen that $M(\theta)$ of Eq.(\ref{m33}) has the same
form as that of Eq.(\ref{m11}), and that angle $\theta$ can be
defined in terms of the parameters $\gamma$ and $\mu$, as shown
in Eq.(\ref{gm11}). The parameter $\eta$ can also be defined
in terms of $\gamma$ and $\mu$, and its expression is the
same as the one given in terms of the angle $\theta$.
As for the branches, we note that both $\gamma$ and $\mu$ can be
negative and positive. Thus, the angle $\theta$ can cover the
entire range from zero to $2\pi$. We can write $\gamma'$ and
$\mu'$ as
\begin{equation}
\gamma' = k\sqrt{\cos^2\theta - \sin^2\theta}, \qquad
\mu' = k \sqrt{\sin^2\theta - \cos^2\theta} .
\end{equation}
Since $\cos^2\theta - \sin^2\theta = \cos(2\theta),$
$\gamma' z$ and $\mu' z$ correspond to $\phi$ and $\chi$ of
Eq.(\ref{phi11}) and Eq.(\ref{chi11}) respectively.
If we start with $\mu = 0$, it is simply a rotation of the
transverse component of the electric field and the overall
attenuation factor is $\exp{(-\lambda z)}$. The rate of this
rotation decreased as $\mu$ increases, and the rotation stops at
$\gamma = \mu$. For $\mu > \gamma$, there are no rotations. It
would be very interesting to test these effects experimentally.
We should not forget the fact that the equi-diagonal
$[abcd]$ matrix is a rotated $ABCD$ matrix. The rotation
matrix is given in Eq.(\ref{rot11}). This rotation changes
the optical ray of Eq.(\ref{ray11}) to
\begin{equation}\label{ray22}
\pmatrix{E_{x} \cr E_{y}} =
\pmatrix{ \cos(\gamma z + \alpha/2) \cr
\sin(\gamma z + \alpha/2)}
\exp{\left\{i \left(k z - \omega t \right)\right\}}.
\end{equation}
This is also an observable effect.
We have seen in this section that the asymmetric optical activity
can serve as an analog computer for the mathematical procedure
given in Sec.~\ref{branch} which is in fact an alternative to the
diagonalization of the $ABCD$ matrix.
\section{Periodic Systems in Optics}\label{periodic}
Let us summarize what we can do about the $ABCD$ matrix.
\begin{itemize}
\item[1.] We should first rootate to an equi-diagonal matrix [abcd].
\item[2.] If the diagonal elements of this equi-diagonal matrix are
smaller than one, it can be written as
\begin{equation}
\pmatrix{\cos\phi & -e^{-\eta} \sin\phi \cr
e^{\eta} \sin\phi & \cos\phi } ,
\end{equation}
with $\exp{(-\eta)} = -b/c$, which can also be written in
terms of the elements of the original $ABCD$ matrix, as shown
in Eq.~(\ref{phi11}).
\item[3.] If the diagonal elements of the equi-diagonal matrix are
greater than than one, the matrix can be written as
\begin{equation}
\pmatrix{\cosh\chi & e^{-\eta} \sinh\chi \cr
e^{\eta} \sinh\chi & \cosh\chi } ,
\end{equation}
with $\exp{(-\eta)} = b/c$, which takes the form of
Eq.(\ref{chi11}) in terms of the elements of the $ABCD$ Matrix.
\item[4.] If one of the off-diagonal elements vanish, the diagonal
elements have to be one.
\item[5.] It is possible to combine all these cases into one
exponential function of one analytic matrix. It can be
written as
\begin{equation}
[abcd] = \exp{ \left\{r\pmatrix{0 & -\cos\theta + \sin\theta \cr
\cos\theta + \sin\theta & 0} \right\}} .
\end{equation}
\item[6.] When $\theta = \pm\pi/4,$ the Taylor series truncates,
and
\begin{equation}
[abcd] = \pmatrix{1 & -r(1 \mp 1)/\sqrt{2} \cr
r(1 \pm 1)/\sqrt{2} & 1 } .
\end{equation}
\end{itemize}
\subsection{Laser Cavities}\label{cav}
\begin{figure}[thb]
\centerline{\includegraphics[scale=0.4]{cavity22.eps}}
\vspace{5mm}
\caption{Optical rays in a laser cavity.
(a) Multiple cycles in a laser cavity are equivalent to the beam
going through multiple lenses, for which one cavity cycle corresponds
to the propagation of light through a sub-system of two lenses. The
$ABCD$ matrix becomes equi-diagonal when the cycle begins at the midway
between the lenses. (b) A laser cavity consisting of two concave
mirrors with separation $s$.}\label{cavity22}
\end{figure}
A laser cavity consists of two concave mirrors separated by distance
$s$ as illustrated in Fig.~\ref{cavity22}. The mirror matrix takes
the form
\begin{equation}\label{lens01}
\pmatrix{1 & 0 \cr -2/R & 1} ,
\end{equation}
where $R$ is the radius of the concave mirror. The separation matrix
is
\begin{equation}\label{lens02}
\pmatrix{1 & s \cr 0 & 1} .
\end{equation}
If we start the cycle from one of the two mirrors
one complete cycle consists of
\begin{equation}\label{lens03}
\pmatrix{1 & 0 \cr -2/R & 1} \pmatrix{1 & s \cr 0 & 1}
\pmatrix{1 & 0 \cr -2/R & 1} \pmatrix{1 & s \cr 0 & 1} .
\end{equation}
If we start the beam at the position $x$ from the mirror, then one
complete cycle becomes
\begin{eqnarray}\label{lens04}
&{}&\pmatrix{1 & x \cr 0 & 1} \pmatrix{1 & 0 \cr -2/R & 1}
\pmatrix{1 & s -x \cr 0 & 1} \nonumber \\[1ex]
&{}& \hspace{20mm}\times \pmatrix{1 & x \cr 0 & 1}
\pmatrix{1 & 0 \cr -2/R & 1}\pmatrix{1 & s - x \cr 0 & 1} .
\end{eqnarray}
This cycle consist of two identical half cycles.
Thus, we shall use the half-cycle matrix as our starting point.
Then the half-cycle $ABCD$ matrix becomes
\begin{equation}\label{lens05}
[ABCD] = \pmatrix{1 & x \cr 0 & 1} \pmatrix{1 & 0 \cr -2/R & 1}
\pmatrix{1 & s - x \cr 0 & 1},
\end{equation}
It is now possible to replace replace $R$ and $x$ and by $R/s$
and $x/s$ respectively and set $s = 1$~\cite{bk09josa}. Then
\begin{equation}\label{lens07}
[ABCD] = \pmatrix{ 1 - 2xf & 1 - 2xf(1 - x) \cr -2f & 1 - 2f(1 - x)} ,
\end{equation}
where $f = s/R$, and is expected to be a small number because the
mirror radius $R$ is much larger than the separation of the mirrors.
It can be brought to an equi-diagonal form by a rotation as given in
Eq.(\ref{abcd11}). According to Eq.(\ref{alpha}), the rotation angle
is
\begin{equation}
\tan\alpha = \frac{2f(2x - 1)}{1 - 2f\left(1 + x - x^{2}\right)} .
\end{equation}
This angle is zero when $x = 1/2$. In this case, the laser cycle
starts at the midway between the lenses~\cite{bk09josa}. Then the
$ABCD$ matrix becomes
\begin{equation}\label{lens09}
[abcd] = \pmatrix{1 - f & 1 - f/2 \cr - 2f & 1 - f} ,
\end{equation}
This matrix can then be written as
\begin{equation}
[abcd] = \pmatrix{\cos\phi & e^{\eta}\sin\phi \cr
-e^{-\eta}\sin\phi & \cos\phi} .
\end{equation}
with
\begin{equation}\label{phieta}
\cos\phi = 1 - f, \qquad
e^{2\eta} = \frac{2 - f}{4f} .
\end{equation}
This is the result we obtained in our earlier paper on laser
cavities~\cite{bk02}, where the cycle starts from the midway
between the lenses. The signs of $\phi$ and $\eta$ are opposite
to those given in Eq.(\ref{mat11}), but this is purely for
convenience. There are no fundamental problems.
We can now write this expression in an exponential form
\begin{equation}
[abcd] = \exp{\left\{r\pmatrix{0 & \cos\theta + \sin\theta \cr
-\cos\theta + \sin\theta & 0 }\right\}},
\end{equation}
with
\begin{eqnarray}
&{}& \tan\theta = \frac{2 - 5f}{2 + 3f} , \nonumber \\[1ex]
&{}& r = \left[\frac{13f -17f^2}{8 - 17f^2}\right]^{1/2} \phi ,
\end{eqnarray}
where $\phi$ is given in Eq.(\ref{phieta}).
Since the radius of the mirror is much larger than the mirror
separation, $f$ is a small number, and $\tan\theta$ is close to
one and $r$ is a small number.
The $N$-cycle laser consists of $2N$ half-cycles, and its $abcd$
matrix is
\begin{equation}
[abcd]^{2N} = \exp{\left\{2Nr\pmatrix{0 & \cos\theta + \sin\theta \cr
-\sin\theta + \cos\theta & 0}\right\}} .
\end{equation}
This section is a straight-forward application of the procedure given
in Sec.~\ref{branch}. We know that $\sin(r\theta)^{2N}$ is not
$\sin(2Nr\theta)$, but the exponential form gives us the convenience of
$[\exp{(ir\theta)}]^{2N} = \exp{(i2Nr\theta)}.$ We have given a
two-by-two matrix formulation of this convenience applicable to the
$ABCD$ matrix.
\subsection{Multilayer Optics}\label{multi}
From the physical concept of Wigner's little group whose transformations
leave the four-momentum of a given particle invariant~\cite{wig39,knp86},
it has been established in the literature that~\cite{gk03}
\begin{equation}\label{lg11}
S(\eta) W S(-\eta) = R(\xi) B(-2\lambda) R(\xi) ,
\end{equation}
where $W$ is one of the four Wigner matrices given in Eq.(\ref{wmat11}),
$R(\xi)$ is the rotation matrix of the form of Eq.(\ref{rot11}),
and
\begin{eqnarray}\label{sqz11}
&{}& S(\eta) = \pmatrix{e^{\eta/2} & 0 \cr 0 & e^{-\eta/2}},
\nonumber\\[1ex]
&{}& B(\lambda) = \pmatrix{\cosh(\lambda/2) & \sinh(\lambda/2) \cr
\sinh(\lambda/2) & \cosh(\lambda/2)} ,
\end{eqnarray}
and the continuous parameters $\xi$ and $\lambda$ take care of the four
different Wigner parameters. These parameters can be written in terms
of $\eta$ and the parameter of the Wigner matrix, as shown in
Ref.~\cite{bk09josa,gk03}.
Since the left side of Eq.(\ref{lg11}) can be written as an exponential
form, we can write
\begin{equation}
R(\xi) B(-2\lambda)R(\xi) = \exp{\left\{ r \pmatrix{0 &
-\cos\theta + \sin\theta \cr \cos\theta + \sin\theta & 0}\right\}} ,
\end{equation}
where $r$ and $\theta$ are also continuous variables.
\begin{figure}[thb]
\centerline{\includegraphics[scale=0.5]{mlayer22.eps}}
\caption{Multilayer consisting of two different refractive indices.
One complete cycle starts at the boundary between medium 2 and
medium 1.}\label{mlayer22}
\end{figure}
With this mathematical preparation, let us study multilayer optics.
In this branch of optics, we have to consider the $ABCD$ matrix
applicable to two beams moving in opposite directions, one which is
the incident beam and the other is the reflected beam~\cite{azzam77}.
We can represent them as
a two component column matrix
\begin{equation}
\pmatrix{ E_{+}e^{ikz} \cr E_{-} e^{-ikz}} ,
\end{equation}
where the upper and lower components correspond to the incoming
and reflected beams respectively.
For a given frequency, the wave number depends on the index of the
refraction. Thus, if the beam travels along the distance $d$, the
column matrix should be multiplied by the two-by-two
matrix~\cite{azzam77}
\begin{equation}\label{ps11}
P(\beta_{j})=\pmatrix{e^{i\beta_{j}/2} & 0 \cr 0 & e^{-i\beta_{j}/2}} ,
\end{equation}
where $\beta_{j}/2 = k_{j}d $ and $j$ is denoting each different medium.
If the beam propagates along the first medium and meets the boundary
at the second medium, it will be partially reflected and partially
transmitted. The boundary matrix is~\cite{azzam77}
\begin{equation}\label{bn1}
B(\nu)=\pmatrix{\cosh(\nu/2) & \sinh(\nu/2) \cr
\sinh(\nu/2) & \cosh(\nu/2) } ,
\end{equation}
with
\begin{equation}
\cosh(\nu/2) = 1/t_{12}, \qquad \sinh(\nu/2) = r_{12}/t_{12} ,
\end{equation}
where $t_{12}$ and $r_{12}$ are the transmission and reflection
coefficients respectively, and they satisfy
$\left(r_{12}^2 + t_{12}^2\right) = 1.$
The boundary matrix for the second to first medium is the
inverse of the above matrix.
Therefore, one complete cycle, starting from the second medium,
consists of
\begin{equation}
B(\nu)P(\beta_{1})B(-\nu)P(\beta_{2}),
\end{equation}
as illustrated in Fig.~\ref{mlayer22}. This complex-valued
matrix can be cast into a real matrix by a similarity
transformation with the transformation matrix
\begin{equation} \label{cmatrix}
C=\frac{1}{\sqrt{2}}
\pmatrix{e^{i\pi/4} & e^{i\pi/4} \cr
- e^{-i\pi/4} & e^{-i\pi/4}} ,
\end{equation}
This transforms the boundary matrix $B(\nu)$ of Eq.(\ref{bn1}) to a
squeeze matrix $S(\nu)$ of Eq.(\ref{sqz11}), and the phase shift
matrices $P(\beta_{j})$ of Eq.(\ref{ps11}) to rotation matrices
$R(\beta_{j})$ of the form given in Eq.(\ref{rot11}). We are thus
led to consider the $ABCD$ matrix of the form
\begin{equation}\label{abcd44}
[ABCD] = S(\nu)R(\beta_{1})S(-\nu)R(\beta_{2}).
\end{equation}
If $W$ in Eq.(\ref{lg11}) is a rotation matrix, we can write
\begin{equation}\label{lg22}
S(\nu)R(\beta_{1})S(-\nu) = R(\xi_{1})B(-2\lambda)R(\xi_{1})
\end{equation}
where
\begin{eqnarray}
&{}&\cosh \lambda = (\cosh\nu) \sqrt{1-\cos^{2} (\beta_{1}/2)
\tanh^{2}\nu}, \\
&{}& \cos\xi_{1}= \frac{\cos(\beta_{1})}
{(\cosh\nu)\sqrt{1-\cos^{2}(\beta_{1}/2)\tanh^{2}\nu}}.
\end{eqnarray}
The $ABCD$ matrix can then be simplified to
\begin{equation}\label{bd22}
[ABCD] = R(\xi_{1})B(-2\lambda)R(\xi_{2})
\end{equation}
with
\begin{equation}
\xi_{2} = \xi_{1} + \beta_{2}
\end{equation}
It is now possible to write the above form as
\begin{equation}
[ABCD] = R(\alpha)[R(\xi)B(-2\lambda)R(\xi)]R(-\alpha) ,
\end{equation}
with
$$ \xi = \frac{1}{2} \left(\xi_1 + \xi_2\right), \qquad
\alpha = \frac{1}{2} \left(\xi_1 - \xi_2\right).
$$
The role of the rotation matrix $R(\alpha)$ matrix is clearly
defined in Sec.~\ref{branch}. Thus $R(\xi)B(-2\lambda)R(\xi)$
is the equi-diagonal matrix, and
\begin{equation}
R(\xi)B(-2\lambda)R(\xi) = \pmatrix{\cosh\lambda\cos\xi
& -(\sin\xi\cosh\lambda + \sinh\lambda) \cr
\sin\xi~\cosh\lambda - \sinh\lambda &
\cosh\lambda~\cos\xi} .
\end{equation}
Thus, if $(\cosh\lambda \cos\xi)$ is smaller than one, we can
write this matrix as
\begin{equation}
\pmatrix{\cos\phi & -e^{\eta}\sin\phi \cr
e^{-\eta}\sin\phi & \cos\phi} ,
\end{equation}
with
\begin{equation}
\cos\phi = (\cosh\lambda)\cos\xi, \qquad
e^{2\eta} = \frac{(\cosh\lambda)\sin\xi + \sinh\lambda}
{(\cosh\lambda)\sin\xi - \sinh\lambda} .
\end{equation}
Thus, if $(\cosh\lambda \cos\xi)$ is greater than one, we should
write the equi-diagonal matrix as
\begin{equation}
\pmatrix{\cosh\chi & -e^{\eta}\sinh\chi \cr
-e^{-\eta}\sinh\chi & \cosh\chi} ,
\end{equation}
with
\begin{equation}
\cosh\chi = (\cosh\lambda)\cos\xi, \qquad
e^{2\eta} = \frac{\sinh\lambda + (\cosh\lambda)\sin\xi}
{\sinh\lambda - (\cosh\lambda)\sin\xi} .
\end{equation}
We are now interested in the exponential form
\begin{equation}\label{expo66}
R(\xi)B(-2\lambda)R(\xi)
= \exp{\left\{r\pmatrix{0 & - (\cos\theta + \sin\theta) \cr
\cos\theta - \sin\theta & 0}\right\}} .
\end{equation}
with
\begin{equation}
\tan\theta = \frac{\tanh\lambda}{\sin\xi} .
\end{equation}
As for the $r$ parameter,
\begin{equation}
r = \left[\frac{\sin^2\xi + \tanh^2\lambda}
{\sin^2\xi - \tanh^2\lambda} \right]^{1/2} \phi , \quad
r = \left[\frac{\sin^2\xi + \tanh^2\lambda}
{\tanh^2\lambda - \sin^2\xi} \right]^{1/2} \chi ,
\end{equation}
for $(\cosh\lambda \cos\xi) < 1$, and for
$(\cosh\lambda \cos\xi) > 1$, respectively.
In this section, we started with two media with two different
indexes of refraction, corresponding to two rotation matrices
$R\left(\beta_{1}\right)$ and $R\left(\beta_{2}\right)$ given
in Eq.(\ref{abcd44}) respectively. However, the combined effect
in not necessarily a rotation matrix. It can be analytically
continued to the hyperbolic branch through the exponent of the
$ABCD$ matrix.
When $(\cosh\lambda \cos\xi)^2 = 1$, one of the off-diagonal
elements in Eq.(\ref{expo66}) vanishes, and this case was
repeatedly discussed in the literature~\cite{bk09josa,gk03},
also in the present paper.
\section*{Concluding Remarks}
In this paper, we noted first that the two-by-two $ABCD$ matrix
can be represented as a similarity transformation of one of the
four matrices which we choose to call the Wigner matrices. We
then combined these Wigner matrices into one exponential form of
an analytic matrix.
While $\cos\phi$ and $\cosh\chi$ correspond to a circle and a
hyperbola respectively, the lines in Fig.~\ref{conic} correspond
parabolas in the four-dimensional representation of the Lorentz
group~\cite{wig39,kiwi90jm}. Ancient Greeks used a circular cone
to combine these curves into one. This is the reason why we call
them conic sections. It is gratifying to note that the optical
devices we discussed in Secs.~\ref{acti} and~\ref{periodic} can
play the role of a conic section. Instead of three-dimensional
cone, we used a two-dimensional plane in Fig.~\ref{conic}.
We have seen in this paper that Taylor expansion of this analytic
form results in four branches. How does this happen? Let us go
to the Taylor expansion of Eq.(\ref{taylor11}). This infinite
series truncates at $(\sin\theta)^2 = (\cos\theta)^2$ or along
the two lines in Fig.~\ref{conic}. We are not familiar with
mathematical singularities resulting from the truncation of the
infinite Taylor series. This appears to be an interesting problem
in mathematics, but it is beyond the scope of this paper.
|
\section{Introduction}
One of the most important physical systems for both classical
and quantum mechanics is the harmonic oscillator. In
contrast to the nonrelativistic harmonic oscillator that is discussed in
most textbooks, the theory of relativistic harmonic oscillator is far
from complete. The reason for this is the complexity of the problem related to
the nonlinearity of differential equations of motion for the
classical relativistic oscillator. No wonder that even in the
simple case of the massive one-dimensional relativistic oscillator
there are problems with identification of periodic solutions to
equations of motion \cite{1}. The problem of a quantum relativistic harmonic
oscillator is usually formulated in one of three different frameworks:
the Klein-Gordon, Dirac or Salpeter equations. The first one uses the spinless
Klein-Gordon equation with a Lorentz invariant oscillatory potential \cite{2}.
However, the solutions of that equation are blamed
by pathologies such as the appearance of ghost states. The second
approach, referred to by Moshinsky \cite{3} as the ``Dirac oscillator" and describing
spin one-half particles utilizes the Dirac equation with an appropriate combination
of the scalar, vector and tensor couplings with an external field \cite{4}. It can
be successively applied to analysis of relativistic symmetries which recently were
recognized experimentally in both nuclear and hadron spectroscopy \cite{5}.
Unfortunately, this approach has no classical relativistic counterpart. Finally,
the third approach follows from the relativistic Hamiltonian dynamics for a scalar
particle and on the quantum level it is based on the spinless Salpeter equation \cite{6}.
The Salpeter equation \cite{6,7,8,9,10,11,12,13,14,15,16} is a "square root" of the
Klein-Gordon equation \cite{17} and can be regarded as its alternative \cite{16}.
The serious advantages of the Salpeter scheme are the lack of problems with probabilistic
interpretation on the quantum level as well as the classically well-defined physical
content of this theory. This last framework is frequently used as a phenomenological
description of the quark-antiquark-gluon system as a hadron model.
Surprisingly, to our best knowledge,
the simplest case of the massless relativistic harmonic oscillator
was not discussed in the literature. In this work we perform a
detailed analysis of the massless relativistic harmonic oscillator.
In particular we find the exact solutions to the classical Hamilton
equations as well as to the corresponding quantum Salpeter equation
and discuss their basic properties. The article is organized as follows.
In Sec.\ II, by integrating the corresponding Hamilton system we identify all
kinds of possible motion of the oscillator as well as find its
quantative characteristics. For an easy illustration of the
dynamics of the relativistic massless harmonic oscillator we also
provide a graphical presentation of numerical integration of
equations of motion. Section III is devoted to the
quantization of the massless relativistic harmonic oscillator.
\section{The analysis of the classical relativistic massless
harmonic oscillator}
The Hamiltonian of the relativistic massless particle subject to the
potential $\frac{1}{2}\kappa^2\bm{x}^2$ is given by
\begin{equation}
H = c|{\bm p}| + \frac{1}{2}\kappa^2{\bm x}^2,
\end{equation}
where ${\bm x}$ and ${\bm p}$ are the position and the momentum of a
particle, $|{\bm p}|=\sqrt{{\bm p}^2}$ is the norm of the vector
${\bm p}$ (so $c|{\bm p}|$ is the kinetic energy of the particle),
$\kappa$ is a constant and $c$ is the speed of light.
Therefore, the Hamilton's equations are
\begin{eqnarray}
\dot {\bm x} &=& c\frac{{\bm p}}{|{\bm p}|},\nonumber\\
\dot {\bm p} &=& -\kappa^2{\bm x}.\nonumber\\
\end{eqnarray}
We point out that an immediate consequence of Eq.\ (2.2) is ${\dot{\bm
x}}^2=c^2$, that is the length of velocity is $c$ as should be for a
massless particle. The familiar integrals of the motion in a
central field \cite{18} are the energy $E$ and the angular momentum ${\bm J}$:
\begin{eqnarray}
E &=& c|{\bm p}|+\frac{\kappa^2}{2}{\bm x}^2,\\
{\bm J} &=& {\bm x}\times{\bm p}.
\end{eqnarray}
As a result of the conservation of the angular momentum ${\bm J}$ the
motion is planar and we can restrict, without loss of
generality, to the case of a particle moving in the $(x^1,x^2)$ plane.
On passing to the polar coordinates ${\bm x}=(x^1,x^2)=(r\cos\varphi,
r\sin\varphi)$ and ${\bm p}=(p^1,p^2)=(p\cos\theta,p\sin\theta)$, where
$r=|{\bm x}|$, and $p=|{\bm p}|$, we obtain from Eq.\ (2.2) the following
system:
\begin{eqnarray}
\dot r &=& c\cos(\theta-\varphi),\nonumber\\
\dot\varphi &=& \frac{c}{r}\sin(\theta-\varphi),\nonumber\\
\dot p &=& -\kappa^2r\cos(\theta-\varphi),\nonumber\\
\dot\theta &=& \kappa^2\frac{r}{p}\sin(\theta-\varphi).\nonumber\\
\end{eqnarray}
The integrals of the motion take the form
\begin{eqnarray}
E &=& cp+\frac{1}{2}\kappa^2r^2,\\
J &\equiv& J_3 = rp\sin(\theta-\varphi).
\end{eqnarray}
From (2.5), (2.6) and (2.7) we find
\begin{equation}
r\sqrt{r^2\left(E-\frac{\kappa^2}{2}r^2\right)^2-(Jc)^2}\,\,d\varphi=\pm |J|c dr.
\end{equation}
We point out that the ``$+$'' and ``$-$'' signs correspond to the
two possible orientations of the angular momentum. We choose, without
loss of generality, the sign ``$+$'' and $J>0$ throughout this
work. Now, from Eq.\ (2.8) we find that the trajectories should satisfy
\begin{equation}
r\left(E-\frac{\kappa^2}{2}r^2\right)\ge Jc,
\end{equation}
and we can classify the types of motion as folows. The first
two possibilities refer to $J\ne 0$. Namely,\bigskip\\
\noindent 1)\quad For $r\left(E-\frac{\kappa^2}{2}r^2\right) = Jc$, we get
\begin{equation}
r=R=\sqrt{\frac{2E}{3\kappa^2}},\qquad \varphi=\omega
t+\varphi_0,\qquad p=p_0=\frac{2E}{3c},\qquad \theta =\omega
t+\varphi_0+\frac{\pi}{2},
\end{equation}
where $\omega
=\frac{c}{R}=\frac{\kappa^2R}{p_0}=c\sqrt{\frac{3\kappa^2}{2E}}$.
So in this case we have a uniform motion in a circle with the linear
speed $|{\bm v}|=\omega R=c$. This solution can also be obtained from
Eq.\ (2.2) by demanding ${\bm x}\mbox{\boldmath${\cdot}$}{\bm p}=0$.
Indeed, it can be easily checked that (2.2) and (2.3) imply the system
\begin{eqnarray}
&&\frac{d}{dt}{\bm x}\mbox{\boldmath${\cdot}$}{\bm p} =
E-\frac{3}{2}\kappa^2{\bm x}^2,\nonumber\\
&&\frac{d}{dt}{\bm p}^2 = -2\kappa^2{\bm
x}\mbox{\boldmath${\cdot}$}{\bm p},\nonumber\\
&&\frac{d}{dt}{\bm x}^2 = \frac{2c}{|{\bm p}|}{\bm
x}\mbox{\boldmath${\cdot}$}{\bm p}.\nonumber\\
\end{eqnarray}
From (2.11) it follows easily that the orthogonality of ${\bm x}$
and ${\bm p}$ refers to the motion of a particle in a circle with radius
$|{\bm x}|=R=\sqrt{\frac{2E}{3\kappa^2}}$.\bigskip\\
\noindent 2)\quad For $r\left(E-\frac{\kappa^2}{2}r^2\right)> Jc$,
the path lies entirely within the annulus bounded by the circles
$r=r_{\rm min}$ and $r=r_{\rm max}$, that is we have
\begin{equation}
r_{\rm min}\le r\le r_{\rm max},
\end{equation}
where $r_{\rm min}$ and $r_{\rm max}$ are the real positive
solutions of the equation
\begin{equation}
\frac{\kappa^2}{2}r^3 -rE+Jc=0.
\end{equation}
We find after some calculation
\begin{eqnarray}
r_{\rm min} &=& 2\sqrt{\frac{2E}{3\kappa^2}}\sin\frac{\alpha}{3},\\
r_{\rm max} &=&
\sqrt{\frac{2E}{3\kappa^2}}(\sqrt{3}\cos\frac{\alpha}{3}-\sin\frac{\alpha
}{3}),
\end{eqnarray}
where $\sin\alpha
=\frac{Jc}{\kappa^2}\left(\frac{3\kappa^2}{2E}\right)^\frac{3}{2}$,
and $0\le\alpha \le\frac{\pi}{2}$.
We now return to Eq.\ (2.8). An immediate consequence of integration of
Eq.\ (2.8) is the the relation
\begin{equation}
\varphi = \varphi_0 +\frac{Jc}{2}\int_{r_0^2}^{r^2}\frac{dx}
{x\sqrt{x(E-\frac{\kappa^2}{2}x)^2-(Jc)^2}}.
\end{equation}
In the particular case of $r_0=r_{\rm min}\ne 0$, and $r(t)>r_{\rm
min}$, $t>0$, the integral from the right-hand side of (2.16) can be
expressed by means of the elliptic integral of the third kind
$\Pi(\phi,n,k)$ (see Ref.\ \cite{19}, 3.137, Eq.\ 3), namely we have
\begin{equation}
\varphi=\varphi_0+\frac{2Jc}{\kappa^2r_{\rm min}^2\sqrt{r_-^2-r_{\rm min}^2}}
\Pi\left(\arcsin\sqrt{\frac{r^2-r_{\rm min}^2}{r_{\rm max}^2-r_{\rm
min}^2}},1-\frac{r_{\rm max}^2}{r_{\rm min}^2},\sqrt{\frac{r_{\rm max}^2
-r_{\rm min}^2}{r_-^2-r_{\rm min}^2}}\right),\quad r_0=r_{\rm min}.
\end{equation}
where $r_-$ is the negative root of the polynomial from the left-hand
side of Eq.\ (2.13) satisfying
\begin{equation}
r_{\rm min}+r_{\rm max}+r_-=0,\qquad r_-^2>r_{\rm max}^2>r_{\rm min}^2.
\end{equation}
Clearly, Eq.\ (2.17) defines $r$ as an implicit function of $\varphi$.
Furthermore, for $r_0=r_{\rm max}\ne 0$, and $r(t)<r_{\rm max}$,
$t>0$, the implicit equation for the trajectory can be obtained from
(2.16) with the help of the elliptic functions of the third kind
$\Pi(\phi,n,k)$ and first kind $F(\phi,k)$ (see \cite{19}, 3.137, Eq.\
4). It follows that
\begin{eqnarray}
&&\varphi = \varphi_0-\frac{2Jc}{\kappa^2r_-^2r_{\rm max}^2\sqrt{r_-^2-r_{\rm min}^2}}
\Bigg[(r_-^2-r_{\rm max}^2)\nonumber\\
&&{}\times\Pi\left(
\arcsin\sqrt{\frac{(r_-^2-r_{\rm min}^2)(r_{\rm max}^2-r^2)}{(r_{\rm
max}^2-r_{\rm min}^2)(r_-^2-r^2)}},
\frac{r_{\rm max}^2-r_{\rm min}^2}{r_-^2-r_{\rm min}^2}
\frac{r_-^2}{r_{\rm max}^2},
\sqrt{\frac{r_{\rm max}^2-r_{\rm min}^2}{r_-^2-r_{\rm min}^2}}\right)\nonumber\\
&&{}+r_{\rm max}^2F\left(\arcsin\sqrt{\frac{(r_-^2-r_{\rm min}^2)(r_{\rm max}^2-r^2)}
{(r_{\rm max}^2-r_{\rm min}^2)(r_-^2-r^2)}},\sqrt{\frac{r_{\rm max}^2-r_{\rm min}^2}
{r_-^2-r_{\rm min}^2}}\right)\Bigg],\quad r_0=r_{\rm max}.\nonumber\\
\end{eqnarray}
Now, the four-dimensional system (2.2), where ${\bm x}=(x^1,x^2)$,
and ${\bm p}=(p^1,p^2)$ is completely integrable. Indeed, it
possesses two integrals in involution $E$ and $J$. Therefore, the
motion between two circles with the radius $r_{\rm min}$ and $r_{\rm
max}$ can be only quasiperiodic and periodic. Of course the case
of the periodic motion refers to a closed path. This means that an
angle $\Delta\varphi$ given by (see formula (2.17))
\begin{eqnarray}
\Delta\varphi &=& \frac{Jc}{2}\int_{r_{\rm min}^2}^{r_{\rm max}^2}\frac{dx}
{x\sqrt{x(E-\frac{\kappa^2}{2}x)^2-(Jc)^2}}\nonumber\\
{} &=& \frac{2Jc}{\kappa^2r_{\rm min}^2\sqrt{r_-^2-r_{\rm min}^2}}
\Pi\left(\frac{\pi}{2},1-\frac{r_{\rm max}^2}{r_{\rm min}^2},\sqrt{\frac{r_{\rm max}^2
-r_{\rm min}^2}{r_-^2-r_{\rm min}^2}}\right),
\end{eqnarray}
should be a rational function of $\pi$, i.e.\ $\Delta\varphi=\pi
m/n$, where $m$ and $n$ are integers. An example of a periodic
motion between two circles is presented in Figs 1 and 2. It
should be noted that the length of momentum (kinetic energy of the
particle $c|{\bm p}|$) has maximum at $r=r_{\rm
min}$, decreases (increases) as $r$ approaches $r_{\rm max}$ ($r_{\rm min}$),
and for $r=r_{\rm max}$ has minimum. Clearly, such behavior of the momentum of a
massless particle is consistent with the form of (2.6). The values
of $r_{\rm min}$ and $r_{\rm max}$ as well as extrema of the length
of momentum can be expressed as a function of the energy by means of
the implicit formulas (2.14) and (2.15). We finally
remark that the case of the uniform motion in a circle discussed
earlier [type 1) of the motion] refers to the condition $r_{\rm
min}=r_{\rm max}=R=\sqrt{\frac{2E}{3\kappa ^2}}$.
\begin{figure*}
\centering
\includegraphics[scale=.8]{fig1.eps}
\caption{The periodic solution of the system (2.2) obtained
by numerical integration. The initial data are ${\bm x}_0=(0.479000,0.000000)$ m,
${\bm p}_0=(0.000000,1.290805)$ Js${\rm m}^{-1}$ , the parameter $\kappa^2=1$
J${\rm m}^{-2}$, and $c=1$ m${\rm s}^{-1}$.}
\end{figure*}
\begin{figure*}
\centering
\includegraphics[scale=.8]{fig2.eps}
\caption{The plot of the radius $r=|{\bm x}|$ (solid line) and the length of momentum
$p=|{\bm p}|$ (dotted line) vs time obtained by numerical integration of (2.11). The initial
condition is the same as in Fig.\ 1.}
\end{figure*}
The third type of the motion corresponds to $J=0$, so we
have\bigskip\\
\noindent 3)\quad $r(E-\frac{\kappa^2}{2}r^2)\ge 0$. From this
inequality we find $0\le r\le r_{\rm max}=\frac{\sqrt{2E}}{\kappa}$.
On the other hand, taking into account (2.7) we find that for $J=0$
the system (2.5) reduces to
\begin{eqnarray}
\dot r &=& \pm c,\nonumber\\
\dot\varphi &=& 0,\nonumber\\
\dot p &=& \mp\kappa^2r,\nonumber\\
\dot\theta &=& 0,\nonumber\\
\end{eqnarray}
where $\theta-\varphi=0$ or $|\theta-\varphi|=\pi$. Therefore a
particle motion is uniform in a segment
$[0,\frac{\sqrt{2E}}{\kappa}]$, more precisely, we have $r=\pm ct+r_0$,
where $0\le r\le \frac{\sqrt{2E}}{\kappa}$ and the two signs correspond
to two possible directions of motion. Assuming that a particle
moves in the $x$-coordinate line, (i.e.\ $x=x^1$), we get
\begin{equation}
x=\pm ct+x_0,\qquad -r_{\rm max}\le x\le r_{\rm max},
\end{equation}
where the turning points are $x=r_{\rm max}$ and $x=-r_{\rm
max}$. On setting $x_0=-r_{\rm max}$ we can write the trajectory
explicitly as
\begin{equation}
x(t) =
(-1)^{\left[\frac{2t}{T}\right]}c\left\{t-\left(2\left[\frac{2t}{T}\right]+
1\right)\frac{T}{4}\right\},
\end{equation}
where $T=\frac{4r_{\rm max}}{c}$ is the period of oscillations of a
massless particle between the turning points $x=r_{\rm max}$ and
$x=-r_{\rm max}$, and $[a]$ is the biggest integer in $a$. The
trajectory (2.23) is illustrated in Fig.\ 3.
Notice that at the turning points the momentum of a massless particle
vanishes [see Eq.\ (2.6) for $r=r_{\rm max}$] that is $p_{\rm min}=0$. The
maximum value of momentum $p_{\rm max}=\frac{E}{c}$ is reached for $x=0$.
The time development of the momentum for $p_0=0$ and $x_0=-r_{\rm max}$
can be written in the form
\begin{equation}
p=-\frac{\kappa^2}{2}\left(x^2(t)-c\frac{T}{4}\right),
\end{equation}
where $x(t)$ is given by (2.23). The plot of $p$ versus $t$ is shown in
Fig.\ 3. Because the momentum
of a massless particle tends to zero as its position approaches the turning
point we deal with a ``red shift'' similar to the gravitational
one.
\begin{figure*}
\centering
\includegraphics[scale=.8]{fig3.eps}
\caption{The plot of the coordinate (solid line) and the momentum (dotted line)
of a massless oscillating particle vs time given by (2.23) and (2.24), respectively,
where $c=1$ m${\rm s}^{-1}$, $\kappa^2=1$ J${\rm m}^{-2}$, $r_{\rm max}=1$ m,
and $T=4$ s.}
\end{figure*}
It should also be noted that the motion in the segment can be
easily obtained from (2.11) by setting $\frac{{\bm x}\mbox{\boldmath$
\scriptstyle{\cdot}$}{\bm p}}{|{\bm x}||{\bm p}|}=\pm 1$, that is
${\bm x}$ and ${\bm p}$ are parallel or antiparallel and therefore
satisfy ${\bm x}\times{\bm p} ={\bm 0}$.
Evidently, in the case of the system (2.2) this
condition is equivalent to $J=0$. We point out that the motion
in a segment corresponds to the condition $r_{\rm min}=0$
and $r_{\rm max}=\frac{\sqrt{2E}}{\kappa}$ for the nonnegative
solutions to (2.13). We finally remark that the type of motion
is completely determined by the values of the energy $E$ and the
angular momentum $J$. Namely, using the parametrization of $r_{\rm
min}$ and $r_{\rm max}$ defined by (2.14) and (2.15) we find
\begin{equation}
0\le \frac{Jc}{\kappa^2}\left(\frac{3\kappa^2}{2E}\right)^{\frac{3}{2}}
\le 1,
\end{equation}
where
$\frac{Jc}{\kappa^2}\left(\frac{3\kappa^2}{2E}\right)^{\frac{3}{2}}=1$
refers to the motion in a circle,
$0<\frac{Jc}{\kappa^2}\left(\frac{3\kappa^2}{2E}\right)^{\frac{3}{2}}<1$
corresponds to the motion between two circles, and
$\frac{Jc}{\kappa^2}\left(\frac{3\kappa^2}{2E}\right)^{\frac{3}{2}}=0$,
i.e.\ $J=0$ is the condition for the motion in the segment.
\section{Quantum mechanics of the relativistic massless harmonic
oscillator}
In relativistic quantum mechanics the massless harmonic oscillator
defined by the Hamiltonian (2.1) is described by a massless version of
the spinless Salpeter equation
\begin{equation}
{\rm i}\hbar\frac{\partial}{\partial t}\psi({\bm x},t)=
\left(c\hbar\sqrt{-\Delta_{\bm x}}+\frac{\kappa^2}{2}{\bm
x}^2\right)\psi ({\bm x},t),
\end{equation}
where $\Delta_{\bm x}=(\frac{\partial}{\partial{\bm x}})^2$.
Therefore the eigenvalue equation for the Hamiltonian
$\hat H\psi_E =E\psi_E$ takes the form of the pseudodifferential
equation
\begin{equation}
\left(c\hbar\sqrt{-\Delta_{\bm x}}+\frac{\kappa^2}{2}{\bm
x}^2\right)\psi_E({\bm x})=E\psi_E({\bm x}).
\end{equation}
Performing the Fourier transformation
\begin{equation}
\psi({\bm x)}=\frac{1}{(2\pi\hbar)^\frac{3}{2}}\int d^3{\bm
k}e^{{\rm i}\frac{{\bm k}\mbox{\boldmath$\scriptstyle{\cdot}$}
{\bm x}}{\hbar}}\tilde\psi({\bm k}),
\end{equation}
we get from (3.2) the following equation:
\begin{equation}
\left(-\Delta_{\bm k}+\frac{2c}{(\kappa\hbar)^2}|{\bm
k}|\right)\tilde\psi_E({\bm k})=\frac{2E}{(\kappa\hbar)^2}
\tilde\psi_E({\bm k}),
\end{equation}
where $\Delta_{\bm k}=(\frac{\partial}{\partial{\bm k}})^2$.
Finally, switching over to the spherical coordinates ${\bm
k}=(k\sin\alpha\cos\beta,k\sin\alpha\sin\beta,k\cos\alpha)$, where
$k=|{\bm k}|$, and making the ansatz
\begin{equation}
\tilde\psi_E({\bm k})=\frac{\chi(k)}{k}Y^m_l(\alpha,\beta),
\end{equation}
where $Y^m_l(\alpha,\beta)$ are the spherical functions, we obtain
the ``radial equation''
\begin{equation}
\left(-\frac{d^2}{dk^2}+\frac{l(l+1)}{k^2}+\frac{2c}{(\kappa\hbar)^2}k
\right)\chi(k)=\frac{2E}{(\kappa\hbar)^2}\chi(k).
\end{equation}
To our best knowledge in the case of $l\ne0$ the solution of (3.6) is
not known. For $l=0$ the solution to (3.6) can be expressed by
means of the Airy function ${\rm Ai}(x)$ \cite{20}, namely
\begin{equation}
\chi(k) =
C{\rm Ai}\left[\frac{2c}{(2c\kappa\hbar)^\frac{2}{3}}\left(k-\frac{E}{c}\right)\right],
\end{equation}
where $C$ is constant. We point out that $l=0$ was also the case
discussed in \cite{6}, where the recurrence was identified
satisfied by coefficients of the formal power series expansion for the
solution to the spinless Salpeter equation corresponding to the massive
relativistic harmonic oscillator. Clearly, $l=0$ refers to the vanishing
angular momentum, therefore we deal in this case with the
quantization of the motion of a massless particle in the segment
$0\le r\le r_{\rm max}=\frac{\sqrt{2E}}{\kappa}$ discussed in the
previous section corresponding to the condition $J=0$ (third type of
the motion). Furthermore, for $l=0$ the ansatz (3.5) takes the form
\begin{equation}
\tilde\psi_E({\bm k})=
\frac{\chi(k)}{k}Y^0_0(\alpha,\beta)=\frac{1}{\sqrt{4\pi}}\frac{\chi(k)}{k}.
\end{equation}
Demanding that $\tilde\psi_E({\bm k})$ is well defined for $k=0$
we find $\chi(0)=0$ (compare \cite{21} Eq.\ (32.11)), which leads to
${\rm Ai}\left(-\frac{2E}{(2c\kappa\hbar)^\frac{2}{3}}\right)=0$. This
quantization condition means that the values of the energy $E_n$,
$n=1,2,\ldots$, are given by zeros of the {\rm Ai}ry function $a_n$. We
have
\begin{equation}
E_n = -\frac{(2c\kappa\hbar)^\frac{2}{3}}{2}a_n,\qquad n=1,2,\ldots .
\end{equation}
Using the fact that the functions ${\rm Ai}(x+a_n)/{\rm Ai}'(a_n)$,
$n=1,2,\ldots$, where ${\rm Ai}'(x)$ designates the derivative of the
Airy function ${\rm Ai}(x)$, form an orthonormal basis on the interval
$[0,\infty)$ \cite{16}, we find that the normalized solutions (3.5)
to (3.4) in the Hilbert space $L^2({\Bbb R}^3,d^3{\bm k})$, with $l=0$
can be written as
\begin{equation}
\tilde\psi_n({\bm k})\equiv \tilde\psi_{E_n}({\bm k}) =
\sqrt{\frac{c}{2\pi}}\frac{1}{(2c\kappa\hbar)^\frac{1}{3}}\frac{1}
{{\rm Ai}'(a_n)}\frac{1}{k}{\rm
Ai}\left(\frac{2c}{(2c\kappa\hbar)^\frac{2}{3}}k+a_n\right).
\end{equation}
From (3.10) and (3.3) we finally obtain the normalized wave functions
such that
\begin{equation}
\psi_n({\bm x}) =
\sqrt{\frac{c}{\hbar}}\,\frac{1}{\pi}\frac{1}{(2c\kappa\hbar)^\frac{1}{3}}\frac{1}
{{\rm Ai}'(a_n)}\frac{1}{r}\int_{0}^{\infty}dk\sin\frac{kr}{\hbar}{\rm
Ai}\left(\frac{2c}{(2c\kappa\hbar)^\frac{2}{3}}k+a_n\right),
\end{equation}
where $r=|{\bm x}|$.
As in the case of the nonrelativistic harmonic
oscillator with the probability density different from zero outside
the turning points, the probability density $\rho_n(r)=|\psi_n({\bm
x})|^2$ does not vanish for $r>r_{\rm max}(E_n)$, where $r_{\rm
max}(E_n)=\frac{\sqrt{2E_n}}{\kappa}$, $n=1,2,\ldots,$. However, it
follows from the numerical calculation that $\rho_n(r)$ has no
maxima for $r>r_{\rm max}(E_n)$ (see Fig.\ 4). Furthermore, taking
into account all directions of the motion in the segment $[0,r_{\rm max}]$
(classical limit does not deal with a single classical orbit but an
ansamble of classical orbits \cite{22}) and taking into account that the
probability of finding a particle in the spherical layer $r$, $r+dr$
is inverse proportional to the surface of the sphere with radius
$r$, we find that the classical probability density is given by the
formula
\begin{equation}
\rho_{\rm cl}({\bm x}) \equiv \rho_{\rm cl}(r)= \frac{\theta(r_{\rm
max}-r)}{4\pi r_{\rm max}r^2},
\end{equation}
where $r_{\rm max}=\frac{\sqrt{2E}}{\kappa}$ and $\theta(x)$ is the
Heaviside step function. Clearly, the normalization
condition is of the form
\begin{equation}
\int d^3{\bm x}\rho_{\rm cl}({\bm x})=\int_0^\infty
\rho_{\rm cl}(r)d\mu(r) = 1,
\end{equation}
where $d\mu(r)=4\pi r^2dr$. The comparison of the quantum
probability density $\rho_n(r)$, and the classical one $\rho_{\rm
cl}(r)$ for $r_{\rm max}(E_n)$ is shown in Fig.\ 4. As expected the
differences between the quantum and the classical descriptions decrease
as the quantum number $n$ increases.
\begin{figure*}
\centering
\begin{tabular}{c@{}c}
\includegraphics[width =.5\textwidth]{fig4.eps}&
\includegraphics[width =.5\textwidth]{fig5.eps}\\
\includegraphics[width =.5\textwidth]{fig6.eps}&
\includegraphics[width =.5\textwidth]{fig7.eps}
\end{tabular}
\caption{The plot of quantum probability density
$\rho_n(r)=|\psi_n({\bm x})|^2$ (solid line), where $\psi_n({\bm x})$
is the wave function (3.12) and the classical probability density
$\rho_{\rm cl}(r)$ (dashed line) given by (3.13), where $r_{\rm
max}=r_{\rm max}(E_n)=\frac{\sqrt{2E_n}}{\kappa}$. We set
$c=1$ m${\rm s}^{-1}$, $\kappa^2=1$ J${\rm m}^{-2}$, and $\hbar=1$ Js.}
\end{figure*}
We now discuss the expectation values of both the kinetic and potential
energies. Using the identity \cite{23}
\begin{equation}
\frac{1}{[{\rm Ai}'(a_n)]^2}\int_0^\infty x{\rm Ai}^2(x+a_n)dx=
-\frac{2}{3}a_n,
\end{equation}
(3.10) and (3.9) we get
\begin{equation}
\langle \psi_n|c\hat p\psi_n\rangle = c\int d^3{\bm k}|{\bm k}||\psi_n(\bm k)|^2
=\frac{2}{3}E_n,
\end{equation}
where $\hat p=\sqrt{{\hat{\bm p}}^2}$. Hence, taking into account
the form of the Hamiltonian in Eq.\ (3.1) we find
\begin{equation}
\left\langle \psi_n\Bigg\vert\frac{\kappa^2}{2}{\hat r}^2\psi_n\right\rangle
= \frac{1}{3}E_n,
\end{equation}
where $\hat r=\sqrt{{\hat {\bm x}}^2}$. We conclude that the virial
theorem takes the nonstandard form in the case of the massless relativistic
harmonic oscillator. More precisely, the roles of the kinetic
energy and potential energies are exchanged. Interestingly, we have
the same formulas on average kinetic and potential energies in the
classical case. Indeed, from Eq.\ (3.12) it follows easily that
\begin{equation}
\left\langle\frac{\kappa^2}{2}r^2\right\rangle_{\rm cl}=\frac{\kappa^2}{2}
\int_0^{\infty}r^2\rho_{\rm cl}(r)d\mu(r)=\frac{1}{3}E.
\end{equation}
Therefore, by virtue of the first equation of Eq.\ (2.3) we have
\begin{equation}
\langle cp\rangle_{\rm cl} = \frac{2}{3}E.
\end{equation}
We finally write down the following approximate relation obtained
numerically:
\begin{equation}
\langle \psi_n|\hat r\psi_n\rangle\approx \frac{r_{\rm max}(E_n)}
{2}=\frac{\sqrt{2E_n}}{2\kappa},
\end{equation}
where the formula is exact in the limit $n\to\infty$. The
approximation in (3.19) is very good. The maximal relative error
$|(\langle \psi_n|\hat r\psi_n\rangle-r_{\rm max}(E_n)/2)
/\langle \psi_n|\hat r\psi_n\rangle|$ arising in the case with $n=1$
is about 5\% and is lesser than 1\% for $n=2$. The fact that the
limit $n\to\infty$ when we have the exact equality in (3.19), is the
classical limit is confirmed by the classical formula
\begin{equation}
\langle r\rangle_{\rm cl} = \int_0^{\infty}r\rho_{\rm cl}(r)d\mu(r)=
\frac{r_{\rm max}}{2}=\frac{\sqrt{2E}}{2\kappa},
\end{equation}
following directly from Eq.\ (3.12).
\section{Conclusion}
In this work we study the relativistic massless harmonic
oscillator in both classical and quantum cases. It seems that the
obtained results concerning the classical oscillator are of
importance not only from the physical point of view. Indeed, Eq.\ (2.2)
is one of the simplest examples of a nonlinear Hamiltonian system
with constant length of velocity. As far as we are aware such an
interesting class of nonlinear dynamical systems was not discussed
in the literature. Referring to the observations of this work
related to the quantum mechanics of the relativistic massless
harmonic oscillator we wish to point out that Eq.\ (3.11) is, to
our best knowledge, the first example of the nontrivial exact solution
to the Salpeter equation. We also stress the good behavior of
the corresponding probability density and expectation values of
observables which confirms the correctness of the quantization based
on the massless spinless Salpeter equation. Furthermore, we
obtain the exact formula (3.9) on the spectrum of the
Hamiltonian. It should be noted that for the Salpeter
equation only energy bounds were analyzed in the literature so far
(for the massive relativistic harmonic oscillator see Ref.\ \cite{10}).
Finally, we have obtained the interesting form of the virial theorem for the
massless relativistic harmonic oscillator with the exchanged roles of
the kinetic and potential energies.
|
\section{Introduction}
Tunable all-optical delay systems that dynamically manipulate the group velocity of light have received a great deal of attention for optical information processing applications such as data buffering and synchronization. Various slow-light devices, including those based on electromagnetically induced transparency (EIT) in atomic vapor, stimulated Brillouin and Raman scattering (SBS and SRS) in optical fiber, and photonic structures in dielectric material, have been explored as potential realizations of a practical delay system [1--10]
As for on-chip approaches, coupled resonators or photonic crystals are promising techniques that would allow easy integration with other electronics or optical components. Many recent demonstrations of coupled resonator optical waveguides (CROW) and side coupled integrated spaced sequence of resonators (SCISSOR) have been designed and fabricated in compact sizes ($\sim$ 10 $\mu\textrm{m}^2$ ) and with the possibility of dynamic delay control and large delay-bandwidth product [2--4].
A more recent analysis from Otey $\textit{et al}$ shows that cascaded resonators can even capture light pulses (i.e., stopped light) by completely compressing the system bandwidth and that the captured pulse can then be released \cite{otey}.
A large fractional delay (equivalent to the delay-bandwidth product) can be achieved by a chain of resonators. Unfortunately, these devices suffer a fundamental trade-off between transmission loss and delay, which potentially limits the use of large numbers of resonators. For example, a CROW consisting of 6 ring resonators demonstrated continuously controllable fractional delay up to 3 at a signal bit rate (BR) of 10 Gbps and a bit error rate (BER) of $10^{-9}$ \cite{mori08}. Its transmission loss, however, is 3 dB (i.e., 0.5 dB/ring) and therefore the use of any additional resonator will increase the BER higher than $10^{-9}$. For comparison, Xia $\textit{et al}$ have demonstrated a chain of 56 cascaded micro-ring resonators in a side-coupled configuration using a silicon-on-insulator waveguide. They achieved a large fractional delay ($\sim$ 5) at a BR of 10 Gbps and BER of $10^{-4}$ \cite{xia}. This high BER is a direct consequence of the 22 dB transmission loss of the device resulting from the use of the large number of rings.
One possible way of preserving acceptable output signal quality without sacrificing the delay performance is to use a Brillouin amplifier. An SBS gain-based delay system could provide significant signal amplification and its tunable gain bandwidth could be increased up to 25 GH
, which allows high speed data transmission \cite{song,Zhu}. In addition, an optimal SBS gain system would provide additional fractional delay of up to 3 \cite{hugo}. Therefore, combining this system with cascaded resonators or other photonic resonance structures seems like a promising method for compensating their respective disadvantages while increasing maximum fractional delay [14--16]
In general, large slow light delay is accompanied by substantial group velocity dispersion (GVD) that manifests itself as signal distortion. The presence of higher-order GVD terms lead to changes in the pulse shape. Under such a condition, we require a metric to quantitatively measure the output data $\emph{quality}$ along with the $\emph{delay}$.
A common measure of communications performance for the propagation of a pulse train is an $\emph{eye-diagram}$. Eye-diagrams are useful for estimating signal distortion via the maximum eye-opening; and its location represents the delay [16--19]. Neifeld and Lee have presented an alternative metric that uses Shannon information to estimate the information capacity and information delay in the presence of noise \cite{mark,mark1}. In this paper, we utilize these two metrics to evaluate SCISSOR, SBS, and SCISSOR + SBS under practical resource and fidelity constraints. By jointly optimizing the system parameters of the SCISSOR + SBS system, we determine the maximum fidelity-constrained fractional delay at a BR of 10 Gbps.
\section{Data fidelity metric}
The important quantities to consider for evaluating slow-light system performance are the fractional delay and received data fidelity. When a single pulse or a pulse sequence propagates through the dispersive media, it undergoes GVD. There are several metrics including the pulse broadening factor \cite{chin09,SHC2}, amplitude and phase distortion \cite{stenner05}, eye-opening [16--19], and mutual information \cite{mark} that have been introduced to quantify the slow-light performance. In what follows, we consider the eye-opening and information-theoretic metrics.
\subsection{Eye-opening metric }
An eye-diagram is used to visualize the shape of communications waveforms and it is generated by repetitively superimposing subsequent traces of a given data stream over a fixed time interval. The eye-opening (EO) is the maximum difference between the minimum value of ``ones" and the maximum value of ``zeros" at the bit center.
The data distortion (D) can be quantified using the eye-opening and it is defined as
\begin{equation}
D=1- \max(EO).
\end{equation}
If a pulse sequence passes through a dispersive medium, the output signal could be broadened or distorted, and then D will increase due to the increased intersymbol interference (ISI). Note that distortion has a monotonic relationship with BER and D = 0.35 indicates a corresponding BER $\simeq$ 10$^{-9}$, resulting in reliable communication \cite{Lee2,John}.
An eye-opening based delay can be calculated by the time difference $T_{EO}$ between the input and output eye center defined when the EO is maximal. The fractional eye-opening delay (EOD) is defined as the time delay divided by the input pulsewidth $T_{p}$, that is, EOD = $T_{EO}/T_{p}$.
\subsection{Information theoretic metric}
Information theory was first explored by Shannon and information rate has become a standard method to characterize the quality of a communication channel \cite{sh1,sh2}.
Recently, an information theoretic metric was introduced using the mutual information between the slow light input and output signals, in order to measure the information based delay (ID) and information throughput (IT) \cite{mark}.
The IT-metric in this paper is based on the channel model displayed in figure 1(a). Figure 1(b) shows the examples of 3 bits output signal propagated through an arbitrary slow light channel, which may include effects of delay, distortion, and noise.
The input X is a binary-valued sequence and it is modulated via on-off keying (OOK). The slow-light delay system is represented by the channel operator H$_S$$_L$, where H$_S$$_L$ could represent any kind of delay device. The mutual information (MI) represents the quantity of transmitted data, and estimates how much input information about X is known when the output Y is observed. Thus, the MI can be defined as $I(X;Y) = H(X)- H(X|Y)$,
where $H(X)$ is the entropy of the discrete input X, representing the a priori uncertainty, and $H(X|Y)$ is the conditional entropy after the output is observed \cite{sh1,sh2}.
We assume that the output signal Y is corrupted by additive white Gaussian noise (AWGN) with zero mean and variance $\sigma^{2}$. We also assume that the elements $x_{i}$ of
a specific n-bit input sequence X are independent and identically-distributed (IID), leading to a prior probability $p(x_{i})$ = (1/2)$^{n}$.
Under these assumptions, the MI can be written as:
\begin{equation}
I(X;Y) = n+ \int\sum_{i=1}^{M}p(x_{i})p(Y|x_{i})\log_{2}\frac{p(Y|x_{i})p(x_{i}) }{\sum_{j=1}^{M} p(x_{j})p(Y|x_{j}) }dY,
\end{equation}
where $\textit{n}$ is the number of input bits, $\textit{M} = 2^\textit{n}$ is the number of possible $\textit{n}$-bit input sequences, and $p(x_{i},Y)$ is the joint probability density function (PDF) of $x_{i}$ and Y.
The integral over Y in equation (2) can be solved by the Monte Carlo simulation with important sampling.
Here, $p(Y|x_{i})$ is the PDF of Y conditioned on $x_{i}$ that is expressed by the Gaussian PDF:
\begin{equation}
p(Y|x_{i}) \simeq \frac{1}{ (2\pi \sigma^{2}) ^{nL}} \exp(-\frac{1}{2\sigma^{2}}|Y-H_{SL}x_{i}|^{2}),
\end{equation}
where L is the number of simulation samples used to represent a single Gaussian pulse.
Note that the concept of delay is not easily captured within $I(X;Y)$. In order to apply I(X;Y) to the analysis of slow light systems, we impose a window structure, which confines an input pulse sequence within a finite duration window \cite{mark, stenner08}. With this approach, we can compute the MI between $\emph{X and only that part of Y}$ contained within the output window (OW) as a function of window offset \cite{mark, mark1}. Here we use a simple example to describe the IT-metric, let us first consider an ideal distortion free delay device with $\sigma^{2}$ = 0, as shown in figure 2.
The 3 bits of Gaussian pulses with a 50\% return-to-zero (RZ) modulation format serve as an input, and the Gaussian pulse is defined to have field amplitude $E(t)$ = exp$(- (t/T_{HW})^{2})$, where $T_{HW}$ = $T_{b}$/2 is the bit half-width at 1/$e^{2}$ intensity and $T_{b}$ is the bit period. The 50\% RZ modulation denotes that a logical one is represented by a half-bit wide pulse, therefore, $T_{p}$ = $T_{b}$/2. In order to compute the MI for this example of the three bit transmission, we consider all 8 possible states ($\textit{M}$ = 8), as shown in figure 2(a). We assume that the input bit period $T_{b}$ = 100 ps and the value of delay $T_{D}$ = 400 ps. The input signal is fitted within an input window (IW), and then we can compute the MI between input X within the IW and only part of Y contained in the OW for many different OW locations in figures 2(a) and (b).
We observe the values of MI = 3 bit and 1 bit for the two candidate output windows (OW1 and OW2) at two different values of window offset = 400 ps and 600 ps, respectively, as shown in figure 2(b). For this example, when the window offset is the same as the delay $T_{D}$, all the input signal information can be transferred without loss caused by distortion, noise, and energy leaking outside the window.
Thus, the peak value of I(X;Y) represents the amount of information that can be transmitted through the slow light channel; while the location of this peak provides an information-theoretic measure of delay. Therefore, we define the peak-height as the information throughput (IT) and the peak location as the information delay (ID) of the SL device, where the normalized IT is
\begin{equation}
IT = \frac{\max\{I(X;Y)\} }{\textmd{n-bits} },
\end{equation}
and this definition will be used throughout the remainder of the paper.
\section{Ring resonators}
\subsection{Single resonator}
For a coupled ring resonator, as shown in figure 3, the output fields can be related to the input fields through a complex amplitude transfer function
\begin{equation}
H_{\mathrm{Ring}}(\omega)=\frac{E_{2}(\omega)}{E_{1}(\omega)}=\frac{k-a\ \mathrm{exp}(i\phi(\omega))}{1-ka\ \mathrm{exp}(i\phi(\omega))},
\end{equation}
where $E_{i}$ is the complex field amplitudes, k is the self-coupling coefficient ($k^{2}=1-\rho^{2}$), $\rho$ is a cross-coupling coefficient,
$\textit{a}$ = exp(-$\alpha L_{R}$/2) is the round trip amplitude loss of the resonator, $L_{R}$ is the ring circumference, and $\alpha$ is the total attenuation coefficient which includes all sources of loss such as material absorption, bending loss, and scattering loss from waveguide roughness \cite{mario,xia}.
The round-trip phase shift $\phi(\omega)$ in the ring can be represented by $\phi(\omega)$ = 2$\pi n_{R} L_{R}$ $(\omega-\omega_0)$/c, where $n_{R}$ is the effective index of the ring, c is the speed of light, and $\omega_0$ is the resonance angular frequency.
The phase response $\Phi_{\mathrm{Ring}}(\omega)$ of the transfer function is obtained by the relation $H_{\mathrm{Ring}}(\omega)$ = $|H_{\mathrm{Ring}}(\omega)| $exp(j$\Phi_{\mathrm{Ring}}(\omega)$) and it is given in terms of $\textit{k}$, $\textit{a}$, and $\phi(\omega)$ as follows:
\begin{equation}
\Phi_{\mathrm{Ring}}(\omega) = \pi + \phi(\omega) + \tan^{-1}\Big[\frac{k\ \mathrm{sin}\phi(\omega)}{a-k\ \mathrm{cos}\phi(\omega)}\Big] + \tan^{-1}\Big[\frac{ka\ \mathrm{sin}\phi(\omega)}{1-ka\ \mathrm{cos}\phi(\omega)}\Big].
\end{equation}
Next, we consider the group delay which is a direct consequence of the amount of phase shift in equation (6) within the filter passband. It is defined as the negative derivative of the phase of the transfer function with respect to the angular frequency:
\begin{eqnarray}
\tau_{\mathrm{Ring}} = -\frac{d\Phi_{\mathrm{Ring}}(\omega)}{d\omega} \nonumber\\
= -\frac{n_{R}L_{R}}{c} + \frac{k(k-a\ \cos\phi(\omega)) }{a^{2}-2ka \ \cos\phi(\omega) + k^{2}}
+ \frac{ka(ka- \cos\phi(\omega)) }{1-2ka\ \cos\phi(\omega) + k^{2}a^{2}}.
\end{eqnarray}
Equations (6) and (7) explain the behavior of the propagated light through the resonator.
At resonance, for $\textit{k} < \textit{a}$, the ring and the waveguide are overcoupled and the phase shift increases rapidly as a function of angular frequency, leading to pulse delay \cite{heebner,heebner2,blair,Lenz}. On the other hand, for $\textit{k} > \textit{a}$, they are undercoupled and the phase shift decreases rapidly as a function of angular frequency, resulting in pulse advancement. Critical coupling occurs when $\textit{k} = \textit{a}$. Here, the transmission becomes zero at the resonance frequency as the round trip loss of the ring is exactly the same as the fractional loss through the resonance coupling \cite{O_S}. In our design study, we are particularly interested in pulse delay, and thus all candidate systems use overcoupled resonators.
In figure 4, we depict the resonator characteristics for four different values of $\textit{k}$ = 0.8, 0.9, 0.95, and 0.97 with a practical value of attenuation coefficient $\alpha$ = 1 cm$^{-1}$ and $L_{R}$ = 150 $\mu$m. Using the numerical simulations based on equations (5) - (7), we calculate and plot the transmission, phase shift, and group delay spectra in figures 4(a), 4(b), and 4(c), respectively. Corresponding Gaussian input and delayed output pulses are shown in figure 4(d), where the input pulsewidth is $T_{p}$ = 50 ps. A silicon waveguide is assumed, thus an effective refractive index $n_{R}$ = 3.0 is used. In figure 4(a), as $\textit{k}$ approaches the critical value of the round trip loss $\textit{a}$ from below, the full width at half depth (FWHD) of the resonator transmission function becomes narrower and deeper. This leads to the slope of the phase shift becoming larger, as shown in figure 4(b), and therefore a larger group delay is achieved for the larger value of $\textit{k}$ = 0.97 in figure 4(c). However, the maximum achievable pulse delay is limited by the tradeoff between the group delay and pulse distortion, causing oscillation at the pulse rising edge, as shown in figure 4(d).
Now, let us consider a pulse train at a BR of 10 Gbps rather than a single pulse, where BR = 1/$T_{b}$.
Figures 5(a), 5(b), and 5(c) present the EO-delay, distortion, and power throughput (PT), respectively, for a resonator as a function of $\textit{k}$ and $L_{R}$. We define the PT as the ratio of the propagated output signal power to the input signal power in the resonator. This numerical simulation is performed by propagating a 127-bit pseudo-random Gaussian pulse train with 50\% RZ modulation format at a BR of 10 Gbps. Our computation covers a range of $\textit{k}$ from 0.94 to 0.99 and $L_{R}$ from 10 $\mu$m to 250 $\mu$m and these ranges are chosen to observe the ring resonator characteristics. It is interesting to note that we observe both the slow and fast light regimes to the left and right sides, respectively, of the critical coupling line (green dashed), in figure 5(a). To increase the delay one can increase $\textit{k}$ or L$_{R}$, but both distortion and energy loss increase at the same time. For a given value of the maximum distortion constraint (e.g. D $\simeq$ 0.35) in figure 5(b), we can find many $\textit{k}$ and $L_{R}$ pairs that provide the same values of distortion-constrained EO-delay $\simeq$ 0.76 with corresponding PT $\simeq$ 0.67, as observed in black dotted lines in figures 5(a) and 5(c). Therefore, we will focus on varying $\textit{k}$ while keeping a fixed practical value of $L_{R}$ = 150 $\mu$m.
\subsection{SCISSOR}
We now consider a SCISSOR, as shown in figure 6. It is assumed that the SCISSOR has multiple identical rings and its transfer function
$E_{\mathrm{out}}(\omega)/E_{\mathrm{in}}(\omega) = H_{\mathrm{SCISSOR}}(\omega)=(H_{\mathrm{Ring}}(\omega))^{N}$,
where $\textit{N}$ is the number of resonators. Figure 7 presents the characteristics of the SCISSOR for four different numbers of rings ($\textit{N}$ = 1, 3, 5, and 8) with $L_{R}$ = 150 $\mu$m, $\alpha$ = 1 cm$^{-1}$, and $\textit{k}$ = 0.85. Because the phase shift at resonance for multiple resonators is additive, the magnitude of the total group delay from a summation of the delays of all individual ring resonators increases as a function of $\textit{N}$. The FWHD of SCISSOR transmission resonance also becomes wider than that of a single ring with the resonance transmission
approaching zero. As a result, output pulse power decreases and the output pulse shape becomes more distorted from the its original input shape as SCISSOR length $\textit{N}$ increases, shown in figure 7(d).
\section{Optimal System Design Study}
In this section, we explore optimal system designs for SCISSOR, SBS, and SCISSOR + SBS. Our approach is to maximize the delay performance under practical system resource constraints while maintaining constant data fidelity \cite{MLeeAO, ravi07, ravi08} .
\subsection{SCISSOR}
We use EO and IT metrics, as described in Section 2, to evaluate the SCISSOR structure. Figure 8 describes the results of the computations summarizing (1) the EO-delay with associated D and (2) the information theoretic delay with associated IT as a function of $\textit{N}$, where three different noise strengths of $\sigma^{2}$= 0.2, 0.3, and 0.4 are used for the IT computation.
The EO-based results presented in this paper are based on propagating a 127 bit pseudo-random pulse train with a RZ modulation format at a BR = 10 Gbps. For information-based results, we have utilized 8 bit input sequences with the same modulation format and BR, and therefore a total 256 ( $\textit{M}$ = 2$^{8}$ states in equation (2) ) possible bit patterns are considered. For each input pattern, we use 10$^{6}$ noise samples to obtain reliable results by using a Monte Carlo technique.
As expected we found that increasing $\textit{N}$ increases both the EOD and ID at the cost of increased distortion. As a result, the normalized IT values decrease. We observe both EOD and ID yield similar delay values, as shown in figure 8(a).
From figure 8(b), we see that IT decreases faster for higher noise strength with increasing $\textit{N}$, thus the fidelity of information transmission decreases with increasing $\sigma^{2}$ because the decreased signal to noise ratio (SNR) causes information to be lost. Based on D and IT results for the SCISSOR with $\textit{N}$ = 4, distortion of D = 0.342 is measured, while three different values of IT = 0.943, 0.873, and 0.812 are computed with corresponding AWGN levels of $\sigma^{2}$ = 0.2, 0.3, and 0.4, respectively, as shown in figure 8(b). For the given specific noise level of $\sigma^{2}$ = 0.3, one can use at most 4 cascaded resonators while simultaneously maintaining more than 87\% (i.e., IT $\geq$ 0.87 or approximately 7 out of 8 bits) of the transmitted information. Therefore, we take the IT constraint IT $\geq$ 0.87 for $\sigma^{2}$ = 0.3, to correspond with distortion constraint D $\leq$ 0.35.
Next, under the two signal quality constraints (IT $\geq$ 0.87 and D $\leq$ 0.35), we optimize $\textit{k}$ to maximize EOD and ID for three different attenuation coefficients $\alpha$ = 0, 1, and 3 cm$^{-1}$ as a function of $\textit{N}$ from 1 to 70. The optimal SCISSOR characteristics using both IT and EO metrics are presented in figure 9. We note that the results from the two different metrics provide similar trends.
When $\textit{N}$ = 70, the maximum fractional delays of approximately 10, 8, and 4 are achieved for $\alpha$ = 0, 1, and 3 cm$^{-1}$, respectively, at the BR = 10 Gbps.
As $\textit{N}$ increases, both fidelity-constrained-EOD and ID increases, however, early delay saturation for the highest attenuation value is observed in figure 9(a). As $\textit{N}$ increases, $\textit{k}$ must decrease in order to increase the effective FWHD so that the system can satisfy the D and IT constraints. For the same reason, optimal $\textit{k}$ at higher attenuation is smaller than that at smaller attenuation, as shown in figure 9(c). Note that these trends are also explained by the group delay relation of equation (7). The transferred energy in a lossy SCISSOR decreases exponentially because of the induced loss, shown in figure 9(b), and thus, the transmission losses of the $\textit{N}$ = 70 SCISSOR become around 22 dB and 43 dB for $\alpha$ = 1 and 3 cm$^{-1}$, respectively. This is what would limit the delay performance and reduce the data fidelity of such a system. Inevitably, we must conclude that an amplification process is required for the SCISSOR. As mentioned earlier, an SBS gain medium is a good choice for both increasing the delay performance and the signal amplification by combining it with the SCISSOR.
\subsection{Broadband SBS}
Slow light via the stimulated Brillouin scattering process has previously been demonstrated for tunable delay in optical fiber \cite{stenner05, SHC2,zhu2}. The SBS process is a nonlinear interaction between a strong pump wave and a weak probe wave that is mediated by an acoustic wave. The acoustic wave generated from this interaction scatters photons to the probe wave, shifting the scattered light downward to the Stokes frequency $\omega_{s}=\omega_{p}-\Omega_{B}$, where $\Omega_{B}$ is the Brillouin frequency shift in optical fiber. As a result, the Stokes field experiences strong gain at $\omega_{s}$. For a typical single mode fiber, the Brillouin frequency shift $\Omega_{B}$ is $\sim$ 10 GHz and the Brillouin linewidth $\Gamma$ is $\sim$ 40 MHz near the communication wavelength of 1550 nm. However, this narrow bandwidth limits the achievable data rate to only several megabits per second. Much of the recent research in the SBS slow-light community focuses on broadening the available SBS bandwidth, and several techniques have been experimentally demonstrated that accommodate a GHz data rate. A primary technique is direct modulation of a Gaussian noise source, generated by an arbitrary waveform generator. Gain bandwidths of up to 25 GHz have been experimentally demonstrated \cite{song, hugo, Zhu}.
Under the small signal approximation, the input field $E(0,\omega$) will be amplified at the fiber output according to $E(L_{f},\omega$) = $E(0,\omega)H_{\mathrm{SBS}}(\omega)$, with the SBS transfer function $H_{\mathrm{SBS}}(\omega)$ = exp($k(\omega)L_{f}$). Here, $L_{f}$ represents the fiber length and $k(\omega$) is the complex wave vector. For the pump broadened SBS, $k(\omega$) = $P_{p}(\omega)\otimes g_{B}(\omega)$ can be obtained by convolving the pump spectrum $P_{p}(\omega)$ with the Lorentzian gain profile $g_{B}(\omega)=g_{0}/ [1-j( (\omega-\omega_\mathrm{s})/(\Gamma/2) )]$, where $g_{0}$ is the line-center gain coefficient. Pant $\textit{et al}$ showed that a super Gaussian function provides a good approximation of the optimal pump profile
$P_{p}(\omega) = (x_{1}/x_{2})$exp$[-( \omega - ((\omega_{s}+\Omega_{B} ))/x_{2})^{2 x_{3}}]$, where the parameters $x_{1}$, $x_{2}$, and $x_{3}$ define pump peak power, pump width, and pump shape (i.e. $x_{3}$=1 is Gaussian and $x_{3} \gg$ 1 becomes nearly rectangular) \cite{ravi08}. In the next subsection, these three parameters will be optimized subject to the fidelity constraint and the maximum SBS gain constraint, $\textit{G}$ = max\{$k(\omega)L_{f}$\} $\leq$ 10.
When $\omega = \omega_{0}$, the line-center gain of the broadband SBS is defined by $\textit{G} \simeq g_{0} x_{1} \Gamma \pi L_{f} / 2A x_{2}$, where A is the mode area.
This gain constraint is imposed to avoid the nonlinear amplifier behavior and the maximum available gain $\textit{G}$ = 10 is conservative value as compare to the Brillouin gain threshold of $\sim$ 25 \cite{zhu2}.
\subsection{SBS + SCISSOR}
Recall the results presented in figure 9, from which we proposed the utility of a joint SCISSOR + SBS system. To demonstrate more explicitly the advantages of this combined slow-light device, we present a practical system design, analyzing its performance in terms of several important factors such as FD, PT, D, and IT in this section. The transfer function for such a device is given by $H(\omega)$ = $H_{\mathrm{SCISSOR}}(\omega) \times H_{\mathrm{SBS}}(\omega)$, and its real and imaginary parts at resonance are related to the gain and the refractive index profiles through the Kramers-Kronig relation. Figure 10 shows the normalized transmission spectra for individual and combined systems along with the spectrum of a 128 bit pseudo-random RZ sequence at BR = 10 Gbps. We assumed the signal carrier frequency and the SCISSOR resonance frequency $\omega_{0}$ are the same as the SBS Stokes frequency $\omega_{s}$.
To better understand the impact of using this combined system, we look at the input and output eye-diagrams after propagating through the combined transmission spectrum, as shown in figure 10, for several different SBS gain values of $\textit{G}$ = 0, 1, 5, and 10. These results are shown in figure 11. For simplicity, we first consider the SCISSOR with $\textit{N}$ = 1 and note that the combined system with SBS $\textit{G}$ = 0 is a resonator-only system.
It is known that the Gaussian pulse propagating through the ring resonator undergoes dispersion effects that can cause oscillations at the pulse rising and/or trailing edges mainly due to the cubic-GVD, as shown in figure 11(b) \cite{mad}. On the other hand, in the SBS system, the output pulse undergoes distortion in the form of pulse broadening mainly due to the quadratic-GVD. Note that distortion management techniques for the SBS system basically suppress the quadratic-GVD term as demonstrated by Stenner \textit{et al} \cite{stenner05}.
By comparing figures 11(b) and 11(d), the fractional EOD for the SCISSOR only and SCISSOR + SBS ($\textit{G}$ = 10) are 0.61 and 1.88 respectively, therefore, it is clearly observed that the SCISSOR + SBS combination not only improves delay performance, but also suppresses the pulse oscillation in the pulse trailing edge arising from the resonator. Although the SBS process also introduces the pulse broadening, it is not significant in this example. In addition, combining SBS + SCISSOR provides additional benefits in terms of delay and PT improvement. Therefore, the combined system provide $\sim$ 3.1 times larger delay with only a small sacrifice of
$\sim$ 1.2 times eye-closing when compared to the SCISSOR-only system.
Figure 12 shows the summary of the optimal results for the resonator loss $\textit{a}$ = 1 cm$^{-1}$ as a function of $\textit{N}$ = 1 - 70 for two candidate systems: SCISSOR-only and SCISSOR + SBS systems under data fidelity constraints (IT $\geq$ 0.87 and D $\leq$ 0.35). We observe that results via both metrics agree well. In general, the maximum fidelity-constrained delays gradually increase, while optimal SBS gain $\textit{G}$ and SCISSOR coupling coefficient $\textit{k}$ decrease. The gain and $\textit{k}$ must be chosen effectively to achieve the maximum delay performance under the IT and D limit. Therefore, for the region of N $<$ 7, the maximum gain can remain constant, whereas k decreases. However, any further increase in N requires a decrease in SBS gain as shown in figure 12(c). We know that the presence of loss causes a nonnegligible decrease in PT for increasing $\textit{N}$ as shown in figure 12(b). The results, however, indicate that the combined system can significantly improve the PT and the delay performance. Even for large number of rings ($\textit{N}$ = 70) the combined system can achieve unit power transmission ratio. The optimal design curves presented in figures 12(a) - (d) represent bounds on the performance of our proposed delay devices subject to real-world operating and fidelity constraints.
In summary, the proposed technique enables a maximum fractional delay of $\sim$ 17, which is $\sim$ 2.1 times the maximum SCISSOR-only delay, with unit power transmission using a cascade of 70 ring resonators combined with an SBS gain medium and can overcome Khurgin's fundamental limit for the fractional delay for the SCISSOR, in which $\textit{N}$ $>$ 100 resonators are required for fractional delay of 10 \cite{khurgin}.
\section{Conclusion}
We have presented a practical system design for increasing the fractional delay while maintaining high data fidelity by combining SBS and SCISSOR. We have employed two different fidelity metrics (EO-metric and IT-metric) to evaluate the slow-light system performance subject to real-world resource constraints. By jointly optimizing the system parameters, the combined SBS + SCISSOR system can provide larger delay and improved power throughput compare to the SCISSOR-only system. We have shown that the maximum fidelity constrained-delay of $\sim$ 17 for SBS + SCISSOR can be achieved with an unit power transmission at a bit rate of 10 Gbps.
\ack
We gratefully acknowledge the financial support of the DARPA/DSO Slow-Light Program.
\section*{References}
|
\section{Introduction}
The electron's spin degree of freedom plays a key role in the emerging area of
semiconductor spintronics \cite{Fert:2008,Grunberg:2008,Zutic04RMP}. A first
scheme for a semiconductor device is the spin field-effect Datta-Das transistor
(DDT). It was proposed 20 years ago \cite{dattadas} and implemented recently
\cite{Koo:2009Science}. Atomic and polaritonic analogs of the electron spin
transistor have also been suggested
\cite{Vaisnav08PRL-DDT,Johne09arxiv-Polariton-DDT}. An important ingredient of
the DDT is the spin-orbit coupling of the Rashba
\cite{Rashba60,Winkler03Review,Rashba-review08} or Dresselhaus
\cite{Dresselhaus55PR,Schliemann03PRL-DDT-Balanced} types. This
Rashba-Dresselhaus (RD) coupling scheme is described by a vector potential which can
be made proportional to the spin-$1/2$ operator of a particle within a plane
\cite{Schlieman06PRB}. It applies to electrons
\cite{Zutic04RMP,Winkler03Review,Rashba-review08} or atoms with two relevant
internal states
\cite{Dudarev04PRL,Ruseckas05PRL,Stanescu07PRL-Rashba,Jacob07APB,Juzeliunas08PRAR,Merkl09EPL-ZB,Vaishnav08PRL-ZB,Larson09PRA,Oh09-tripod}.
In the case of atoms, the spin-orbit coupling can be generated using two
counterpropagating light beams
\cite{Jacob07APB,Juzeliunas08PRAR,Merkl09EPL-ZB,Oh09-tripod} (or two standing
waves \cite{Stanescu07PRL-Rashba,Vaishnav08PRL-ZB,Larson09PRA}) and a third beam
propagating in an orthogonal direction, the beams being coupled to the atoms in
a tripod scheme \cite{Ruseckas05PRL,Unanyan98OC,Unanyan99pra}. The tripod atoms
have two degenerate internal dressed states known as \emph{dark states}, which
are immune to atom-light coupling. The center-of-mass motion of the dark-state
atoms is described by a two-component spinor and is equivalent to the motion of
a spin-$1/2$ particle with spin-orbit coupling
\cite{Stanescu07PRL-Rashba,Jacob07APB,Juzeliunas08PRAR,Merkl09EPL-ZB,Vaishnav08PRL-ZB,Larson09PRA,Oh09-tripod}
of the RD type.
\begin{figure}
\centering
\includegraphics[width=0.5\columnwidth]{N-pod}
\caption{$N$-pod configuration. An atomic state $|0\rangle$ is coupled to $N$
different atomic states $|j\rangle$, $j=1,...,N$ by $N$ resonant laser fields.}
\label{fig:figure-Npod}
\end{figure}
In the present article we investigate the possibility to generalize the
RD spin-orbit coupling scheme to spins larger than $1/2$. We
show that this can be achieved using cold atoms with more than two internal dark
states. We start our analysis with the general scheme in which $N$ laser beams
couple $N$ atomic internal ground states to a common excited state, thus forming
the $N$-pod setup shown in Fig.~\ref{fig:figure-Npod}. In the
$(N+1)$-dimensional Hilbert space, we identify $N-1$ dark states,
that is, zero-energy eigenstates of the atom-light Hamiltonian that are
superpositions of
the $N$ ground states and are immune to atom-light coupling.
Subsequently, we analyze the tetrapod case ($N=4$) for which the center-of-mass
motion of the dark-state atoms is described by a three-component spinor and thus
corresponds to the motion of a spin-$1$ particle. We show that the resulting
spin-orbit coupling can be made of the RD type and yields three
cylindrically symmetric dispersion branches. Two of them are similar to those
for the familiar RD spin-$1/2$ Hamiltonian, so the atom can exhibit the
well-known
quasirelativistic behavior \cite{Juzeliunas08PRAR,Vaishnav08PRL-ZB} for
small wave vectors. Furthermore there is an extra branch with a flat dispersion
around zero momentum. The formation of the latter branch leads to interesting
phenomena, such as a possibility to have a negative refraction at a potential
step, characterized by a larger amplitude as compared to the spin-$1/2$ case.
Finally we explore a possible implementation of the tetrapod scheme for
alkali-metal
atoms using Raman transitions. To avoid a strong heating due to spontaneous
emission, all the states forming the tetrapod scheme are chosen among the Zeeman
sublevels of the atomic ground state, and are coupled by far-detuned Raman
lasers beams.
\section{The $N$-pod scheme}
\subsection{Atomic Hamiltonian}
We are interested in the center-of-mass motion of atoms in the field of several
light beams. The atoms are characterized by $N$ internal ground states
$|1\rangle$, $|2\rangle$, $\ldots$ , $|N\rangle$, which are resonantly coupled
to an extra state $|0\rangle$ by laser beams. This provides the $N$-pod
configuration shown in Fig.~\ref{fig:figure-Npod}. Note that the state
$|0\rangle$ does not necessarily represent an electronic excited level; it can
be a sublevel of the atomic ground state coupled to the states $|1\rangle$,
$|2\rangle$, $\ldots$ ,$|N\rangle$ via stimulated Raman transitions. A more
detailed discussion on practical implementation is presented in the
Sec.~\ref{sec:Production-of-the}.
The Hamiltonian describing the motion of an atom in the presence of the light
beams is
\begin{equation}
H_0 =\frac{\mathbf{p}^2}{2m}+V_{0}+V_1
\label{}
\end{equation}
where $m$ is the atomic mass and $\mathbf{p}=-i\hbar \nabla$ the atomic
momentum operator. The terms $V_0$ and $V_1$ describe the atom-light
interaction in the $N$-pod configuration and a possible additional external
potential, respectively. We assume for simplicity that all couplings
$|j\rangle \leftrightarrow |0\rangle$, $j=1,\ldots,N$ are resonant, so that
$V_0$ reads using the interaction representation and the rotating wave
approximation:
\begin{equation}
V_0=\hbar\sum_{j=1}^N\Omega_j(\mathbf{r})\, |0\rangle\langle
j|+\mathrm{H.c.}\,,
\label{eq:H-0}
\end{equation}
where $\Omega_j$ is the Rabi frequency that couples the internal state
$|j\rangle$ to the common state $|0\rangle$, with $j=1,2,\ldots,N$. The
coupling $V_0$ can be rewritten as
\begin{equation}
V_0=\hbar\Omega(\mathbf{r})\Bigl[|0\rangle\langle
B(\mathbf{r})|+|B(\mathbf{r})\rangle\langle0|\Bigr]\,,
\label{eq:H-0-alternative}
\end{equation}
with
\begin{equation}
|B\rangle=\frac{1}{\Omega}\sum_{
j=1}^N\Omega_j^*|j\rangle\,,\qquad\Omega^2=\sum_{j=1}^N|\Omega_j|^2\,.
\label{eq:B}
\end{equation}
Here $|B\rangle$ is the so-called bright (coupled) state and $\Omega$ is the
total Rabi frequency.
The diagonalization of the atom-light interaction potential $V_0$ is straightforward:
(a) The coupling between the bright state $|B\rangle$ and the state $|0\rangle$ with
a strength equal to the Rabi frequency $\Omega$ in
Eq.~(\ref{eq:H-0-alternative}) gives rise to the two eigenstates
\begin{equation}
|\pm\rangle=\left(|B\rangle\pm|0\rangle\right)/\sqrt{2}\, ,
\label{eq:H-0-diagonal}
\end{equation}
with energies $\pm\hbar\Omega$.
(b) The remaining orthogonal $(N-1)$-dimensional subspace corresponds to dark states.
We denote $|D_n\rangle$, $n=1,\ldots, N-1$ an orthonormal basis of this
subspace. All dark states are eigenstates of the Hamiltonian $\hat{H}_0$ with
zero eigenenergy: $\hat{H}_0|D_n\rangle=0$ . They are orthogonal to the bright
state and to the state $|0\rangle$: $\langle B|D_n\rangle=\langle
0|D_n\rangle=0$.
Although the eigenenergies of the dark states are position-independent, the
states $|D_n\rangle$ depend on the atomic position through the spatial variation
of the Rabi frequencies $\Omega_j$. This leads to the appearance of the gauge
potentials to be considered next.
\subsection{Adiabatic motion of dark-state atoms}
We now suppose that the atoms are prepared in the dark-state subspace, and that
they move sufficiently slowly to remain in this manifold. This adiabatic
approximation is justified if the light fields are strong enough, so that the
energy difference $\pm \hbar\Omega$ between the dark-state manifold and the
other eigenstates $|\pm\rangle$ of $V_0$ is large compared to the detuning
due to Doppler shifts. The atomic state-vector $|\Phi\rangle$ can then be
expanded on the dark-state basis
\begin{equation}
|\Phi\rangle=\sum_{j=1}^{N-1}\Psi_j(\mathbf{r})|D_j(\mathbf{r})\rangle,
\end{equation}
where $\Psi_j(\mathbf{r})$ is the wave function for the center-of-mass motion of
the atom in the $j$th dark state. The atomic center-of-mass motion is described
by an $(N-1)$-component wave function
\begin{equation}
\Psi=\left(
\begin{array}{c}
\Psi_1\\\ldots\\\Psi_{N-1}\end{array}\right)
\label{eq:psi-D-original}
\end{equation}
obeying the Schr\"odinger equation
\begin{equation}
i\hbar\frac{\partial}{\partial t}\Psi=H\Psi,
\label{eq:SE-reduced}
\end{equation}
with the Hamiltonian
\begin{equation}
H=\frac{1}{2m}(-i\hbar\nabla-\mathbf{A})^2+\Phi+V\,.
\label{eq:H}
\end{equation}
The potentials governing the atomic center-of-mass motion $\mathbf{A}$, $\Phi$,
and $V$ are $(N-1)\times(N-1)$ matrices. Here $\mathbf{A}$ and $\Phi$ are the
geometric potentials that emerge due to the spatial dependence of the atomic
dark states
\cite{Ruseckas05PRL,Berry84PRSA,wilczek84PRL,mead91,Bohm03Book,Shapere1989}.
The matrix $\mathbf{A}(\mathbf{r})$ represents a non-Abelian vector potential,
with the matrix elements
\begin{equation}
\mathbf{A}_{n,m}=i\hbar\langle D_n(\mathbf{r})|\nabla
D_m(\mathbf{r})\rangle\,,\quad n,m=1,\ldots,N-1\,.
\label{eq:A-nm}
\end{equation}
The matrix $\Phi(\mathbf{r})$ is an effective scalar potential known as the
Born-Huang potential. It can be expressed through the matrix elements of the
vector potential between the dark states and the bright state
$|B\rangle\equiv|D_0\rangle$:
\begin{equation}
\Phi_{n,m}=\frac{1}{2m}\mathbf{A}_{n,0}\cdot\mathbf{A}_{0,m}\,,\qquad
n,m=1,\ldots,N-1\,.
\label{eq:Phi-nm}
\end{equation}
The matrix $V(\mathbf{r})$ represents the restriction of $V_1(\mathbf{r})$ to
the dark state subspace. For simplicity we assume in the following that (i) the
matrix elements of $V_1$ between the dark-state manifold and the states
$|B\rangle$ or $|0\rangle$ are negligible, so that $V_1$ cannot cause any
significant departure of atoms from the dark-state manifold; (ii) $V$ is
proportional to the identity matrix in the dark state subspace, so that it does
not break the gauge symmetry of $(\mathbf{A},\Phi)$. For the particular case of
alkali-metal
atoms, this occurs when the trapping is provided by far-detuned laser beams. The
confinement potential is then the same for all sublevels of the electronic
ground state, in particular for the states $|j \rangle$ ($j=1,\,\ldots,\,N$)
considered here.
The non-Abelian vector potential $\mathbf{A}$ provides a curvature
(or effective ``magnetic'' field)
\begin{equation}
\mathbf{B}=\nabla\times\mathbf{A}+\frac{1}{i\hbar}\mathbf{A}\times\mathbf{A}\ .
\label{eq:B-eff-initial}
\end{equation}
The first term represents the usual curl. Note that the second term
$\mathbf{A}\times\mathbf{A}$ does not vanish in general, since the Cartesian
components of the vector potential $\mathbf{A}$ do not necessarily commute
(i.e.\ the vector potential is non-Abelian). Therefore in contrast to the
Abelian case, even a constant vector potential can produce a nonzero curvature
and thus provide nontrivial topological effects, leading, for example, to unusual
dispersion curves.
\section{Effective fields generated by plane-wave laser beams }
\subsection{Dark states and gauge potentials}
From now on we focus on the case where the laser beams represent plane running
waves characterized by wave vectors $\mathbf{k}_j$, $j=1,\ldots,N$.
We suppose that the $N$
Rabi frequencies have equal amplitudes and read
\begin{equation}
\Omega_j=\frac{1}{\sqrt{N}}\Omega e^{i\mathbf{k}_j\cdot\mathbf{r}}\,,\qquad
j=1,2,\ldots,N\, .
\label{eq:Omega_j-plane-wave}
\end{equation}
At this stage the directions of the wave vectors $\mathbf{k}_j$ are still
arbitrary; we will address some specific geometries in Secs.~\ref{sec:planar}
and \ref{sec:tetrahedron}.
A convenient orthogonal set of $N-1$ normalized dark states is
\begin{equation}
|D_n\rangle=\frac{1}{\sqrt{N}}\sum_{j=1}^N|j\rangle e^{i2\pi
jn/N-i\mathbf{k}_j\cdot\mathbf{r}}\,,
\label{eq:D-n}
\end{equation}
with $n=1,2,\ldots,N-1$. Note that the bright state given by Eqs.~(\ref{eq:B})
and (\ref{eq:Omega_j-plane-wave}) has the form of Eq.~(\ref{eq:D-n}) with $n=0$,
so we will use in the following the notation $|D_0\rangle\equiv|B\rangle$.
Equations (\ref{eq:A-nm}) and (\ref{eq:D-n}) provide the matrix elements of the
vector potential
\begin{equation}
\mathbf{A}_{n,m}=\frac{\hbar}{N}\sum_{j=1}^N\mathbf{k}_je^{i2\pi
j(m-n)/N}\;.
\label{eq:A-nm-symmetric}
\end{equation}
It is evident that the vector potential $\mathbf{A}_{n,m}$ depends only on the
difference $n-m$, i.e.\ $\mathbf{A}_{n,m}=\mathbf{A}_{n-m,0}$.
Since the vector potential given by Eq.~(\ref{eq:A-nm-symmetric}) is constant in
space, the effective magnetic field (\ref{eq:B-eff-initial}) simplifies to
$i\hbar\mathbf{B}=\mathbf{A}\times\mathbf{A}$. Using
Eq.~(\ref{eq:A-nm-symmetric}), it can expressed in terms of the off-diagonal
matrix elements of the vector potential $\mathbf{A}_{n,0}$ and
$\mathbf{A}_{0,m}$ :
\begin{equation}
\mathbf{B}_{n,m}=\frac{i}{\hbar}\mathbf{A}_{n,0}\times\mathbf{A}_{0,m}\,.
\label{eq:B-eff-specific-nm}
\end{equation}
\subsection{Vector potential and angular momentum
\label{sub:Vector-potential-and}}
We now address the following question: Can the vector potential $\mathbf{A}$
be made proportional to a three-dimensional (3D) angular momentum operator
$\mathbf{J}$, that is,
$\mathbf{A}=\gamma\mathbf{J}$ , where $\gamma$ is a constant? If the answer
was positive, this would allow one to achieve a three-dimensional RD-type
coupling. This would be formally similar to the effective spin-orbit
interaction discussed in \cite{Zygelman:1990}, arising from non-Abelian gauge
fields in molecular physics. However as we see now, one cannot use the present
scheme to achieve $\mathbf{A}\propto \mathbf{J}$.
The angular momentum operator is known to obey the following relations:
\begin{equation}
\mathbf{J}\times\mathbf{J}=i\hbar\mathbf{J}\, .
\label{eq:J-J-cross-product-J-relationship}
\end{equation}
If $\mathbf{A}=\gamma\mathbf{J}\,,$ the cross product of the vector potential
should be proportional to the vector potential itself:
$\gamma\mathbf{A}\times\mathbf{A}=i\hbar\mathbf{A}$ or simply
$\gamma\mathbf{B}=\mathbf{A}$. Using Eq.~(\ref{eq:B-eff-specific-nm}), the last
relationship would lead to
\begin{equation}
\gamma\mathbf{A}_{n,0}\times\mathbf{A}_{m,0}^*=-i\hbar\mathbf{A}_{n
-m,0}\,,\quad n,m=1,\,\ldots,\,N-1\,.
\label{eq:C-cross-product}
\end{equation}
Multiplying Eq.~(\ref{eq:C-cross-product}) by $\mathbf{A}_{m,0}^*$, the
left-hand side of the resultant equation is zero. Thus one arrives at
\begin{equation}
\mathbf{A}_{m,0}^*\cdot\mathbf{A}_{n-m,0}=0.
\label{eq:C-d-relation}
\end{equation}
Equation (\ref{eq:C-d-relation}) should hold for all possible values of $n$ and
$m$. In particular, by taking $m=1$ and $n=2$, one finds
$\mathbf{A}_{1,0}^*\cdot\mathbf{A}_{1,0}=0$. This equation can be fulfilled only
if $\mathbf{A}_{1,0}=0$. Then by taking $m=1$, the relationship
(\ref{eq:C-cross-product}) yields that
$\mathbf{A}_{n-1,0}=\mathbf{A}_{n+p,p+1}=0$ for integer $n$ and $p$. This means
that the vector potential $\mathbf{A}=\gamma\mathbf{J}$ should be identically
equal to zero.
In this way, we have proved that when using the $N$-pod scheme with plane waves of
equal amplitudes it is not possible to generate a nonzero vector potential
which is proportional to the 3D angular momentum operator $\mathbf{J}$. In other
words, it is not possible to produce a 3D spin-orbit coupling of the RD type
using the $N$-pod scheme. Yet one can get a two-dimensional (2D) RD coupling
by means of the
$N$-pod scheme. This includes not only the usual spin-$1/2$ RD coupling but
also a generalized 2D RD coupling for the spin-$1$ case, as we shall see
later on.
\section{Plane matter-wave solutions}
We suppose in the following that the external potential $V$ is uniform in space.
In this case the Schr\"odinger equation (\ref{eq:SE-reduced}) has plane-wave
solutions:
\begin{equation}
\Phi_{\mathbf{k}}(\mathbf{r},t)=\Psi_{\mathbf{k}}e^{i\mathbf{k}\cdot\mathbf{r}
-\omega_{\mathbf{k}}t},
\label{eq:Phi-k-definition}
\end{equation}
where $\omega_{\mathbf{k}}$ is an eigenfrequency and $\Psi_{\mathbf{k}}$ is a
\textbf{$\mathbf{k}$}-dependent spinor:
\begin{equation}
\Psi_{\mathbf{k}}=\left(
\begin{array}{c}
\Psi_{1,\mathbf{k}}\\\ldots\\\Psi_{(N-1),\mathbf{k}}\end{array}\right)\ .
\end{equation}
Note that the direction of the wave vector \textbf{$\mathbf{k}$} is arbitrary
and it is not related to the wave vectors of the light beams
\textbf{$\mathbf{k}_j$}.
The \textbf{$\mathbf{k}$}-dependent spinor $\Psi_{\mathbf{k}}$ obeys the
stationary Schr\"odinger equation
\[
H_{\mathbf{k}}\Psi_{\mathbf{k}}=\hbar\omega_{\mathbf{k}}\Psi_{\mathbf{k}},
\]
with the $\mathbf{k}$-dependent Hamiltonian
\begin{equation}
H_{\mathbf{k}}=\frac{\hbar^2}{2m}k^2-\frac{\hbar}{m}\mathbf{A}\cdot\mathbf{k}
+\frac{1}{2m}\mathbf{A}^2+\Phi+V \ .
\label{eq:H-k}
\end{equation}
Exploiting Eqs.~(\ref{eq:Phi-nm}) and (\ref{eq:A-nm-symmetric}), the scalar term
$\mathbf{A}^2/2m+\Phi$ takes the form
\begin{equation}
\left(\frac{1}{2m}\mathbf{A}^2+\Phi\right)_{n,m}=\frac{\hbar^2}{2m}\frac{1}{
N}\sum_{j=1}^N\mathbf{k}_j^2e^{i\frac{2\pi}{N}(m-n)j}\,.
\label{eq:A-square
+ Phi-1}
\end{equation}
If the wave vectors of all the Rabi frequencies have the same modulus
$\mathbf{k}_j^2=2\kappa^2$, the term
\begin{equation}
\frac{1}{2m}\mathbf{A}^2+\Phi=\frac{\hbar^2\kappa^2}{m}\hat{I}
\end{equation}
is proportional to the unit matrix $\hat{I}$ for any arrangement of the
wave vectors (both planar and 3D). In this case the Hamiltonian (\ref{eq:H-k})
simplifies to
\begin{equation}
H_{\mathbf{k}}=\frac{\hbar}{2m}
\Bigl(\hbar k^2-2\mathbf{A}\cdot\mathbf{k}+2\hbar \kappa^2\Bigr)+V\,.
\end{equation}
If the external trapping potential $V$ is proportional to the unit matrix, the
eigenvectors $\Psi_{\mathbf{k}}^{\beta}$ of the Hamiltonian $H_{\mathbf{k}}$ are
also the eigenvectors of the operator
$A_{\mathbf{k}}=\mathbf{A}\cdot\mathbf{k}/k$ representing the projection of the
vector potential along the wave vector,
\begin{equation}
A_{\mathbf{k}}\Psi_{\mathbf{k}}^{\beta}=-\hbar\kappa\beta\Psi_{\mathbf{k}}^{
\beta}\,,
\label{eq:A-k-eigenstate}
\end{equation}
where the dimensionless parameter $\beta\equiv\beta_{\mathbf{k}}$ depends
generally on the wave-vector $\mathbf{k}$. The corresponding eigenvalues of the
Hamiltonian $H_{\mathbf{k}}$ are
\begin{equation}
\hbar\omega_{\mathbf{k}}^{\beta}=\frac{\hbar^2}{2m}\Bigl[(k+\beta\kappa)^2+(2
-\beta^2)\kappa^2\Bigr]+V.
\label{eq:omega-k-beta}
\end{equation}
For $\mathbf{k}=\mathbf{0}$ all the eigenenergies
$\hbar\omega_{\mathbf{k}}^{\beta}$ are
equal and do not depend on the branch parameter $\beta$. Consequently all
dispersion branches merge to $\omega_{0}^{\beta} \equiv \omega_0$ at the origin
where $k=0$. To find the eigenstates and the eigenenergies for $k\ne0$, one
needs to specify the arrangement of the wave vectors $\mathbf{k}_j$.
\section{Planar geometry}
\label{sec:planar}
\subsection{Wave vectors on a regular polygon}
Let us analyze a situation where the wave vectors $\mathbf{k}_j$ are situated in
a plane and form a regular polygon
\begin{eqnarray}
\mathbf{k}_j & = &
\sqrt{2}\kappa[-\cos\alpha_j\mathbf{e}_x+\sin\alpha_j\mathbf{e}_y]\\ &
= & -\kappa\Bigl(e^{i\alpha_j}\mathbf{e}_{+}+e^{-i\alpha_j}\mathbf{e}_{-}\Bigr)\,,
\end{eqnarray}
with $\mathbf{e}_{\pm}=\frac{1}{\sqrt{2}}(\mathbf{e}_x\pm i\mathbf{e}_y)$, where
$\alpha_j=2\pi j/N$ is the angle between the wave vector and the $x$ axis. The
scalar and vector potentials, Eqs.~(\ref{eq:Phi-nm}) and
(\ref{eq:A-nm-symmetric}), take the form
\begin{eqnarray}
\Phi_{n,m} & = & \frac{\hbar^2\kappa^2}{2m}(\delta_{m,1}\delta_{n,1}+\delta_{m,N
-1}\delta_{n,N-1})\,,
\label{eq:Phi-general-result} \\
\mathbf{A}_{n,m} & = & -\hbar\kappa\sum_{\pm}\mathbf{e}_{\pm}\delta_{n,m\pm1}\,.
\label{eq:A-nm-poligon}
\end{eqnarray}
The vector potential is thus a tridiagonal matrix whose elements are
proportional to $\mathbf{e}_x\pm i\mathbf{e}_y$, whereas the scalar potential
$\Phi_{n,m}$ is a diagonal matrix with nonzero elements only for $n=m=1$ or
$n=m=N-1$.
Note that the matrices $A_x$ and $A_y$, are proportional to the $x$ and $y$
components of the angular momentum operator $\mathbf{J}$ only for the tripod
($N=3$) and tetrapod ($N=4$) schemes. In these cases the scalar potential is
proportional to $J_z^2$.
The projection of $\mathbf{A}_{n,m}$ along the wave vector is
\begin{equation}
(A_{\mathbf{k}})_{n,m}=-\frac{\hbar\kappa}{\sqrt{2}}\Bigl(\delta_{n,m+1}e^{i\varphi}
+\delta_{n,m-1}e^{-i\varphi}\Bigr)\,,
\end{equation}
where $\varphi$ is the angle between the wave vector $\mathbf{k}$ and the $x$
axis. The eigenvectors of this operator are
\begin{equation}
\Psi_{\mathbf{k}}^{\beta}=\sqrt{\frac{2}{N}}\left(
\begin{array}{c}
\sin\left(\frac{\pi q}{N}\right)\\\sin\left(2\frac{\pi
q}{N}\right)e^{i\varphi}\\\cdots\\\sin\left((N-1)\frac{\pi
q}{N}\right)e^{i(N-2)\varphi}\end{array}\right)\,,
\label{eq:Phi-k-beta-N}
\end{equation}
with $q=1,\dots N-1$. The corresponding eigenvalues are given by
Eq.~(\ref{eq:A-k-eigenstate}) with
\begin{equation}
\beta=\sqrt{2}\cos\left(\frac{\pi q}{N}\right)\,.
\label{eq:beta}
\end{equation}
It is to be emphasized that the dimensionless parameter $\beta$ does not depend
on $\mathbf{k}$ for this particular geometry. The vectors
$\Psi_{\mathbf{k}}^{\beta}$ represent eigenstates of the Hamiltonian with
eigenenergies $\omega_k^{\beta}$ given by Eqs.~(\ref{eq:omega-k-beta}) and
(\ref{eq:beta}). This provides $N-1$ dispersion branches.
\subsection{Tripod setup}
\begin{figure}
\centering
\includegraphics[width=0.9\columnwidth]{fig-directions}
\caption{(Color online) Planar arrangement of laser beams for tripod (a) and
tetrapod (b) setups.}
\label{fig:tripod-directions}
\end{figure}
Consider first the tripod setup ($N=3$) in which the wave vectors $\mathbf{k}_j$
form an equilateral triangle [Fig.~\ref{fig:tripod-directions}(a)]. The parameter
$\beta$ featured in Eqs.~(\ref{eq:A-k-eigenstate}) and (\ref{eq:beta}) then
takes the values $\hbar\beta/\sqrt{2}=\pm\hbar/2$, representing the eigenvalues
of the projection of a spin-$1/2$ on a given axis. In such a situation the
operator $\mathbf{A}$ is related to the spin $1/2$ operator
$\hbar\boldsymbol{\sigma}_{\bot}$, providing the RD coupling along the $xy$
plane as in the previous studies
\cite{Stanescu07PRL-Rashba,Jacob07APB,Juzeliunas08PRAR,Merkl09EPL-ZB,Vaishnav08PRL-ZB,Larson09PRA,Oh09-tripod}:
\begin{equation}
\mathbf{A}=-\hbar\kappa\boldsymbol{\sigma}_{\bot}/\sqrt{2}\; .
\label{eq:A-tripod}
\end{equation}
It is noteworthy that the present setup produces a cylindrically symmetric
spin-orbit coupling in a more straightforward manner than the previously
suggested tripod schemes. Those schemes involved two counterpropagating light
beams \cite{Jacob07APB,Juzeliunas08PRAR,Merkl09EPL-ZB,Oh09-tripod} (or two
standing waves \cite{Stanescu07PRL-Rashba,Vaishnav08PRL-ZB,Larson09PRA}) and a
third beam propagating in an orthogonal direction. Consequently, one needed to
add a detuning potential and make the amplitudes of the Rabi frequencies
asymmetric in order to have dispersion curves of the RD-type, with the
proper cylindrical symmetry
\cite{Stanescu07PRL-Rashba,Jacob07APB,Juzeliunas08PRAR,Merkl09EPL-ZB,Vaishnav08PRL-ZB,Larson09PRA,Oh09-tripod}.
On the contrary, for the present regular polygon arrangement of wave vectors,
the dispersion relation is naturally symmetric as long as the amplitudes of all
four Rabi frequencies are equal.
\subsection{Tetrapod setup\label{sub:Tetrapod-seup}}
For $N=4$ one arrives at the tetrapod setup involving two pairs of
counterpropagating laser fields shown in Fig.~\ref{fig:tripod-directions}(b). In
this case the vector potential reads
\begin{equation}
\mathbf{A}=\frac{\hbar\kappa}{\sqrt{2}}\left(
\begin{array}{ccc}
0 & -\mathbf{e}_x+i\mathbf{e}_y & 0\\ -\mathbf{e}_x-i\mathbf{e}_y & 0 &
-\mathbf{e}_x+i\mathbf{e}_y\\ 0 & -\mathbf{e}_x-i\mathbf{e}_y & 0
\end{array}\right)\,.
\label{eq:A-tetrapod-matrix}
\end{equation}
The possible values for the parameter $\beta$ featured in
Eq.~(\ref{eq:A-k-eigenstate}) are $\hbar\beta=0,\,\pm\hbar$, representing the
eigenvalues of the component of a spin $1$ along a given axis. Consequently the
operator $\mathbf{A}$ is proportional to the projection $\mathbf{J}_{\bot}$ of a
spin $1$ operator along the $xy$ plane
\begin{equation}
\mathbf{A}=-\kappa\mathbf{\mathbf{J}_{\bot}}\,,\qquad\mathbf{J}_{
\bot}=J_x\mathbf{e}_x+J_y\mathbf{e}_y\, .
\label{eq:A-tetrapod}
\end{equation}
The scalar potential can be represented in terms of the $z$ component of the spin operator
\begin{equation}
\Phi=\frac{\hbar^2\kappa^2}{2m}J_z^2\,.
\label{eq:Phi-tetrapod}
\end{equation}
\begin{figure}
\centering
\includegraphics[width=0.9\columnwidth]{fig2}
\caption{(Color online) Dispersion curves for the tetrapod scheme calculated
using Eq.~(\ref{eq:omega-k-beta-quatro-pod}) for $V=0$. Here
$\omega_0=\hbar \kappa^2/m$.}
\label{fig:dispersion}
\end{figure}
The eigenstates and the eigenenergies of the Hamiltonian are now
\begin{equation}
\Psi_{\mathbf{k}}^{\pm1}=\frac{1}{2}\left(
\begin{array}{c}
1\\\pm\sqrt{2}e^{i\alpha}\\ e^{2i\alpha}
\end{array}\right)\,,\quad\Psi_{\mathbf{k}}^0=\frac{1}{\sqrt{2}}\left(
\begin{array}{c}
-1\\ 0\\ e^{2i\alpha}\end{array}\right)\,,
\label{eq:Psi-k-tetrapod}
\end{equation}
and
\begin{equation}
\hbar\omega_k^{\beta}=\frac{\hbar^2}{2m}\Bigl(k^2+2\kappa
k\beta+2\kappa^2\Bigr)+V\,,\quad\beta=0\,,\pm1.
\label{eq:omega-k-beta-quatro-pod}
\end{equation}
For $\beta=\pm1$ the dispersion curves shown in Fig.~\ref{fig:dispersion} are
analogous to those of the spin-$1/2$ RD model. An additional dispersion curve
with $\beta=0$ represents a parabola centered at $k=0$.
The dispersion curve with $\beta=-1$ has its minimum at
$\hbar\omega=\hbar^2\kappa^2/2m$, whereas the other two dispersion branches have
minima at the double energy $\hbar\omega=\hbar^2\kappa^2/m$ (for $V=0$).
Therefore, all dispersion curves have a strictly positive minimum energy. This
nonzero minimum originates from the micromotion of the atom in the light
field, caused by nonadiabatic transitions between the dark and bright states
\cite{Aharonov:1992,Cheneau08EPL}. The associated kinetic energy gives rise to
the scalar potential given by Eq.~(\ref{eq:Phi-general-result}), which has a
nonzero contribution even when acting on the dark states.
Finally, we note an important difference in the ``topology'' of the
eigenfunctions for the RD spin-$1/2$ and spin-$1$ problems, even thought the
$\beta=\pm1$ branches have the same dispersion in the two cases: The
wave functions $\Psi_{\mathbf{k}}^{\beta}$ exhibit a $\pi$ Berry's phase in $k$
space in the spin-$1/2$ case, whereas this Berry's phase is absent for the
spin-$1$.
\section{Tetrahedron geometry}
\label{sec:tetrahedron}
In this section we present an example of a nonplanar setup, which has some
advantages with respect to the planar configuration investigated in the previous
section, because it leads to a simpler scalar potential. We consider again the
tetrapod setup ($N=4$) with wave vectors $\mathbf{k}_j$ arranged in a regular
tetrahedron geometry:
\begin{equation}
\hat{\mathbf{k}}_j\cdot\hat{\mathbf{k}}_{j^{\prime}}=-\frac{1}{3}\,,\quad
j\ne j^{\prime}\,.
\label{eq:tetrahedron-condition}
\end{equation}
where $\hat{\mathbf{k}}_j=\mathbf{k}_j/k_j$ is a unit vector. More precisely,
we choose
\begin{equation}
\mathbf{k}_{1,3} = \kappa^{\prime}(\pm \mathbf{e}_y\sqrt{2}-\mathbf{e}_z)\;,\quad
\mathbf{k}_{2,4} = \kappa^{\prime}(\pm\mathbf{e}_x\sqrt{2}+\mathbf{e}_z)\;.
\label{}
\end{equation}
Using Eq.~(\ref{eq:A-nm-symmetric}), the vector potential then reads
\begin{equation}
\mathbf{A}=\frac{\hbar\kappa^{\prime}}{\sqrt{2}}\left(
\begin{array}{ccc}
0 & -\mathbf{e}_x+i\mathbf{e}_y &\sqrt{2}\mathbf{e}_z\\
-\mathbf{e}_x-i\mathbf{e}_y & 0 & -\mathbf{e}_x+i\mathbf{e}_y\\
\sqrt{2}\mathbf{e}_z & -\mathbf{e}_x-i\mathbf{e}_y & 0\end{array}\right)\;.
\label{eq:A-tetrahedron}
\end{equation}
For atoms moving in the $xy$ plane the vector potential can be expressed in
terms of a spin-$1$ operator in the $xy$ plane:
$\mathbf{A}_{\bot}=-\kappa^{\prime}\mathbf{J}_{\bot}$. Hence we obtain as before
a RD-type spin-orbit coupling for the atomic motion in the $xy$ plane,
characterized by the dispersion relation shown in Fig.~\ref{fig:dispersion}. Yet
we are now dealing with a 3D problem, so the same dispersion also characterizes
the atomic motion along two other planes perpendicular to the vectors
$\mathbf{e}_x+\mathbf{e}_y$ and $\mathbf{e}_x-\mathbf{e}_y$. By making an atomic
lattice along these directions, the atomic tunneling will be influenced by a
spin-$1$ RD coupling, thus extending the previous studies of spin-$1/2$
RD coupling in lattices \cite{Goldman09PRA}. This will be investigated in a
separate study.
A distinguished feature of the tetrahedron geometry is that the scalar potential
is proportional to the unit matrix $\hat{I}$:
\begin{equation}
\Phi=\frac{\hbar^2\kappa^{\prime2}}{2m}\hat{I}\;.
\label{eq:Phi-tetrahedron}
\end{equation}
Thus for atoms placed in a 3D lattice, there is no energy mismatch between
different dark states located in adjacent sites. This contrasts with the planar
tetrapod case (Eq.~(\ref{eq:Phi-tetrapod})), where the spin components are
likely to get frozen in the lattice because tunneling matrix elements are
normally much smaller than the atomic recoil energy, which gives the scale for
the scalar potential.
It is noteworthy that the $z$ component of the vector potential given by
Eq.~(\ref{eq:A-tetrahedron}) is not proportional to $J_z$. Hence, one cannot
generate a 3D Hamiltonian with RD-type spin-orbit coupling for all
directions of the atomic motion. This is a particular case of the general
conclusion reached in the Sec.~\ref{sub:Vector-potential-and}.
\section{Transmission by a potential step}
A spectacular consequence of spin-orbit RD coupling is the negative
refraction and reflection that occurs when a matter wave is incident on a
potential step. The problem was investigated for spin-$1/2$ atoms
\cite{Juzeliunas08PRAR,juz08-neg-refl} and electrons \cite{Winkler09PRB}. In
this case one can calculate relatively easily the transmission and reflection of
the atomic wave packet. For small wave vectors of the incident atoms,
$k\ll\kappa$, the transmission probability is close to unity at zero angle of
incidence. Here the parameter $\kappa$ characterizes the strength of the
spin-orbit interaction, see Eq.~(\ref{eq:A-tripod}). This nearly complete
transmission is a manifestation of the Klein paradox appearing also for electron
tunneling in graphene \cite{Katsnelson06NP}. For a nonzero angle of incidence,
the transmission probability is less than $1$ and decreases with increasing
angle. Furthermore the transmitted matter wave experiences negative refraction
\cite{Juzeliunas08PRAR}, similar to the case of electrons in graphene
\cite{Cheianov07Science}.
Particles with a spin larger than $1/2$ have additional degrees of freedom,
which modifies the continuity conditions at the potential step. This can lead to
a significant increase of the transmission probability of atoms, as we show now
for particles submitted to a spin-$1$ RD coupling.
\subsection{The Hamiltonian}
We consider in this section the motion of a particle in the $xy$ plane described
by the Hamiltonian
\begin{equation}
H=\frac{1}{2m}\left(\hat{\mathbf{p}}^2+2\hbar\kappa\hat{\mathbf{p}}\cdot\mathbf{
J}_{\bot}+2\hbar^2\kappa^2\right)+V(x)
\label{eq:H-Spin-1}
\end{equation}
where $\mathbf{J}_{\bot}=J_x\mathbf{e}_x+J_y\mathbf{e}_y$ is the projection of
spin-$1$ operator onto the $xy$ plane. Such a Hamiltonian can be obtained using
the tetrapod setups described in the Secs.~\ref{sub:Tetrapod-seup} and
\ref{sec:tetrahedron}. The external potential $V(x)$ is given by the step
function along $x$
\begin{equation}
V(x)=
\begin{cases}
0, & x\leq 0\\ V_0, & x>0\end{cases}
\label{eq:V(x)}
\end{equation}
with $V_0>0$. It is convenient to introduce the wave vector
$k_0=2mV_0/\hbar^2\kappa$ characterizing the height of the barrier.
For a constant potential the eigenvalue equation has plane-wave solutions
(\ref{eq:Phi-k-definition}) characterized by the spinor part
$\Psi_{\mathbf{k}}^{\beta}$ [Eq.~(\ref{eq:Psi-k-tetrapod})]. The corresponding
eigenvalues $\hbar\omega_k^{\beta}$ are given by
Eq.~(\ref{eq:omega-k-beta-quatro-pod}) with $k=\sqrt{k_x^2+k_y^2}$ and are
plotted in Fig.~\ref{fig:dispersion}. Additionally there can be evanescent wave
solutions localized in the vicinity of the potential step in the $x>0$ region.
In that case we have $k_x=iq$ with $q>0$, giving
\begin{equation}
k=\sqrt{k_y^2-q^2}\,,\qquad q^2<k_y^2\,.
\label{eq:k-evanesc}
\end{equation}
For the present problem, only the evanescent wave with $\beta=0$ will play a
role,
\begin{equation}
\Psi_{k_y,q}^0=c_{k_y,q}^0\left(
\begin{array}{c}
-1\\ 0\\\frac{q+k_y}{q-k_y}\end{array}\right)\,,
\label{eq:Evanescent-beta--0}
\end{equation}
where $c_{k_y,q}^{\beta}$ is the normalization factor.
\subsection{Incident waves with $\beta=1$}
\begin{figure}
\centering
\includegraphics[width=0.7\columnwidth]{fig-dispersion-barrier}
\caption{(Color online) Wave numbers of reflected and transmitted waves and
energy conservation at the potential step.}
\label{fig:dispersion-barrier}
\end{figure}
\begin{figure}
\centering
\includegraphics[width=0.7\columnwidth]{fig-geometry}
\caption{(Color online) Reflection and transmission of atoms at a potential
step. In addition there is an evanescent transmitted wave with
$\mathbf{k}_6=k_y\mathbf{e}_y+iq\mathbf{e}_x$.}
\label{fig:geomery}
\end{figure}
In this paragraph we restrict our analysis to the case where the incident atom
is prepared in the upper dispersion branch ($\beta=1$) in the region $x<0$.
Denoting its wave vector by $\mathbf{k}$, the incident wave is
\begin{equation}
\Psi_{\mathrm{in}}=\Psi_{\mathbf{k}}^1e^{i\mathbf{k}\cdot\mathbf{r}}\;.
\label{eq:Psi-in}
\end{equation}
The potential step is assumed to be high enough,
\begin{equation}
k_0>k^2/\kappa+2k\,,
\label{eq:V-0--Condition}
\end{equation}
so that there can be no propagating transmitted waves with chirality $\beta=0$
or $\beta=1$ (see Fig.~\ref{fig:dispersion-barrier}). At the same time, to allow
for propagation of plane waves in the region $x>0$ for the lower dispersion
branch $\beta=-1$, the step height should not be too large:
\begin{equation}
k_0<\kappa+k^2/\kappa+2k\,.
\end{equation}
The directions of reflected and transmitted waves are depicted in
Fig.~\ref{fig:geomery}. The reflected waves generally contain all three
components,
\begin{equation}
\Psi_{\mathrm{refl}}=r_1\Psi_{\mathbf{k}_1}^1e^{i\mathbf{k}_1\cdot\mathbf{r}}
+r_2\Psi_{\mathbf{k}_2}^0e^{i\mathbf{k}_2\cdot\mathbf{r}}+r_3\Psi_{\mathbf{
k}_3}^{-1}e^{i\mathbf{k}_3\cdot\mathbf{r}}\,,
\label{eq:Psi-refl}
\end{equation}
where $k_1=k$, $k_2=\sqrt{k^2+2\kappa k}$ and $k_3=k+2\kappa$. The reflection
angles are $\pi-\alpha$, $\pi-\alpha_2$, and $\pi-\alpha_3$, with
$\alpha_2=\arcsin[\sin(\alpha)k/k_2]$ and $\alpha_3=\arcsin[\sin(\alpha)k/k_3]$.
The transmitted waves are
\begin{equation}
\Psi_{\mathrm{tr}}=t_4\Psi_{\mathbf{k}_4}^{-1}e^{i\mathbf{k}_4\cdot\mathbf{r}}
+t_5\Psi_{\mathbf{k}_5}^{-1}e^{i\mathbf{k}_5\cdot\mathbf{r}}+t_6\Psi_{\mathbf{
k}_6}^0e^{i\mathbf{k}_6\cdot\mathbf{r}}\,,
\label{eq:Psi-tr}
\end{equation}
where $k_4=\kappa-\sqrt{(k+\kappa)^2-k_0\kappa}$,
$k_5=\kappa+\sqrt{(k+\kappa)^2-k_0\kappa}$ , and $k_6=k^2+2\kappa k-k_0\kappa$
(see Fig.~\ref{fig:dispersion-barrier}). The first and second transmitted waves
experience negative and positive refraction, respectively, and propagate at the
angles $\pi-\alpha_4$ and $\alpha_5$, where
$\alpha_4=\arcsin[\sin(\alpha)k/k_4]$ and $\alpha_5=\arcsin[\sin(\alpha)k/k_5]$.
On the other hand, due to the condition (\ref{eq:V-0--Condition}) the third
transmitted wave with the helicity $\beta=0$ is an evanescent one along the $x$
axis and thus is characterized by the wave vector
$\mathbf{k}_6=k_y\mathbf{e}_y+iq\mathbf{e}_x$, with $k_y=k\sin\alpha$ and
$q=\sqrt{k_y^2-k_6^2}$. Note that there is no evanescent transmitted wave in the
upper dispersion branch ($\beta=1$) because it cannot comply with the momentum
conservation along the interface in addition to the energy conservation.
The multicomponent wave function and its first derivative in the $x$ direction
are required to be continuous at the barrier ($x=0$), providing six equations
containing six unknown coefficients $r_1$, $r_2$, $r_3$, $t_4$, $t_5$, and $t_6$.
Of special interest is the situation where $k_0=4k$. In this case the
wave number of the first refracted wave coincides with the wave number of the
incident wave, $k_4=k$, so the angle of refraction is equal to the angle of
incidence for the first reflected wave. $\alpha_4=\alpha$.
The analytical solution for the six coefficients is generally complicated. It is
instructive to obtain approximate solutions for small wave vectors and small
angles of incidence, $k\ll\kappa$ and $\alpha\ll1$. In such a case one can
restrict to reflected (\ref{eq:Psi-refl}) and transmitted (\ref{eq:Psi-tr})
waves containing only the contributions of $\mathbf{k}_1$, $\mathbf{k}_2$,
$\mathbf{k}_4$, and $\mathbf{k}_6$. The transmitted wave with $\mathbf{k}_6$
represents a rapidly decaying evanescent wave characterized by a spinor
component given by Eq.~(\ref{eq:Evanescent-beta--0}) with $q\gg k_y$:
\begin{equation}
\Psi_{\mathbf{k}_6}^0\approx\frac{1}{\sqrt{2}}\left(
\begin{array}{c}
-1\\ 0\\ 1\end{array}\right)\;.
\label{eq:g-k6-0}
\end{equation}
The continuity of the wave function at $x=0$ gives
\begin{equation}
\Psi_{\mathbf{k}}^1+r_1\Psi_{\mathbf{k}_1}^1+r_2\Psi_{\mathbf{k}_2}^0e^{
i\mathbf{k}_2\cdot\mathbf{r}}=t_4\Psi_{\mathbf{k}_4}^{-1}+a\Psi_{\mathbf{k}_6}^0\;.
\label{eq:continuity-approx-2}
\end{equation}
In addition, we require continuity of the derivative in the $x$ direction for
the component with $\beta=0$, which is the largest:
\begin{equation}
k_2\cos(\pi-\alpha_2)r_2\Psi_{\mathbf{k}_2}^0=iqa\Psi_{\mathbf{k}_6}^0
\label{eq:continuity-derivatives-approx-2}
\end{equation}
Here $k_4\approx k_0/2-k$ and $\alpha_2\approx\sqrt{k/2\kappa}\sin(\alpha)$, with
$k_2=\sqrt{2\kappa k}$ and $q\approx\sqrt{\kappa(k_0-2k)}$. Using the spinors
(\ref{eq:Psi-k-tetrapod}) and (\ref{eq:g-k6-0}) one obtains the following
solution to Eqs.~(\ref{eq:continuity-approx-2}) and
(\ref{eq:continuity-derivatives-approx-2}):
\begin{eqnarray}
t_4 & = &
\frac{2\cos\alpha}{\cos\alpha+\cos\alpha_4}e^{i(\alpha+\alpha_4)}
\label{eq:t4-approx-2}
\\ r_1 & = &
\frac{\cos\alpha_4-\cos\alpha}{\cos\alpha+\cos\alpha_4}e^{i2\alpha}
\label{eq:r1-approx-2}
\\ r_2 & = &
\frac{\sqrt{k_0-2k}}{\sqrt{k}+i\sqrt{k_0/2-k}}\cos\alpha\tan\left(\frac{\alpha
+\alpha_4}{2}\right)e^{i\alpha}
\label{eq:r2-approx-2}
\end{eqnarray}
The calculated reflection and transmission coefficients $r_1$ and $t_4$ obey the
probability conservation up to terms of the order of $O(\alpha^2)$:
\begin{equation}
|r_1|^2+\frac{\cos\alpha_4}{\cos\alpha}|t_4|^2\approx1\,.
\label{eq:probability-conservation-1}
\end{equation}
If the barrier height is such that $k_0=4k$, we have $\alpha_4=\alpha$. In that
case $|t_4|\approx1$ and $|r_1|\approx0$ leading to an almost perfect negative
refraction, at the exactl opposite refraction angle, provided $k\ll\kappa$ and
the angles of incidence are not too large.
Figure~\ref{fig-compar-1-0.5} presents the comparison of the transmission
probabilities for the spin-$1$ and spin-$1/2$ RD coupling using the exact
numerical solutions of the continuity equations at the boundary $x=0$. For the
spin-$1/2$ case the incident wave is also prepared in the upper dispersion
branch. The figure shows a marked increase in the transmission probability for
small angles of incidence in the case of spin $1$. Note that the transmitted
waves experience negative refraction both for the spin-$1$ and the
spin-$1/2$ cases.
\begin{figure}
\centering
\includegraphics[width=0.9\columnwidth]{fig-comparison}
\caption{(Color online) Transmission probability of negatively refracted atoms
at a potential step as a function of the angle of incidence $\alpha$ for a
spin-$1$ (solid red) and spin-$1/2$ (dashed green) systems. The parameters used
for the calculation are $k/\kappa=0.1$ and $k_0/\kappa=0.4$ for both systems.}
\label{fig-compar-1-0.5}
\end{figure}
\section{Implementation of the tetrapod setup with alkali-metal atoms}
\label{sec:Production-of-the}
\begin{figure}
\centering
\includegraphics[width=0.6\columnwidth]{lasers}
\caption{Directions of the laser beams used for implementing the tetrapod
coupling scheme with alkali-metal atoms, via stimulated Raman transitions.}
\label{fig:exp-directions}
\end{figure}
\begin{figure}
\centering
\includegraphics[width=1\columnwidth]{levels}
\caption{Implementation of the tetrapod scheme for alkali-metal atoms with two
hyperfine levels of angular momentum $F=1$ and $F=2$. The laser couplings
involved in this scheme correspond to stimulated Raman transitions between
hyperfine states of the ground atomic level. We choose
$|0\rangle\equiv|F=1,m_F=0\rangle$. (a) The laser beams $A$, $A^{\prime}$, $B$
and $B^{\prime}$ induce the transitions
$|0\rangle\to|1\rangle\equiv|F=2,m_F=1\rangle$ and
$|0\rangle\to|3\rangle\equiv|F=2,m_F=-1\rangle$. (b) The laser beams
$A^{\prime\prime}$ and $B^{\prime\prime}$ induce the transitions
$|0\rangle\to|2\rangle\equiv|F=1,m_F=1\rangle$ and
$|0\rangle\to|4\rangle\equiv|F=1,m_F=-1\rangle$. }
\label{fig:exp_scheme}
\end{figure}
We now discuss a possible implementation of the tetrapod scheme. We consider the
case of alkali-metal atoms, which are the most frequently used in current experiments.
In order to avoid a strong heating due to spontaneous emission, we study the
case where the state $|0\rangle$ is actually one of the Zeeman sublevels of the
ground state. The states $|j\rangle$ (with $j=1,\ldots,4$) are also Zeeman
sublevels of the ground state, and the coupling between the state $|0\rangle$
and a state $|j\rangle$ is provided by a pair of laser beams that induce a Raman
transition under the condition of the two-photon resonance. The use of Raman
transitions in this context is an extension to the tetrapod case of a recent
proposal \cite{Spielman09PRL} to implement a $\Lambda$-type scheme for the
generation of an effective magnetic field by means of the counterpropagating
laser beams \cite{Cheneau08EPL,Juzeliunas06PRA}.
We recall that the electronic ground level $nS_{1/2}$ of alkali-meal atoms is split
by hyperfine interaction in two sublevels with angular momenta $F=I+1/2$ and
$F=I-1/2$, where $I$ is the nuclear spin. We consider in the following the case
$I=3/2$ that is relevant for lithium ($^7\mathrm{Li}$, $n=2$), sodium
($^{23}\mathrm{Na}$, $n=3$) or rubidium ($^{87}\mathrm{Rb}$, $n=5$). In order to
minimize the rate of spontaneous emission processes, we restrict to Raman
transitions that are far detuned from the resonance with the ``true'' excited
states $nP_{1/2}$ or $nP_{3/2}$ of the $D_1$ or $D_2$ transitions. More
precisely the typical one-photon detuning of the beams involved in the Raman
process is chosen much larger than the hyperfine structure of the excited level
$nP_{1/2}$ or $nP_{3/2}$ ($0.8\,\mathrm{GHz}$ for the hyperfine splitting of the
level $5P_{1/2}$ of $^{87}\mathrm{Rb}$). At the same time the one-photon
detuning should be smaller than the fine structure splitting, that is, the
difference between the energies of $nP_{1/2}$ and $nP_{3/2}$
($7000\,\mathrm{GHz}$ for $^{87}\mathrm{Rb}$). When the one-photon detuning
exceeds the hyperfine splitting, the nucleus angular momentum does not play any
role in the selection rules that determine the allowed transitions for photon
absorption or emission. For the $D_1$ ($D_2$) transition, the allowed
couplings are the same as between a spin-$1/2$ ground level and a spin-$1/2$
($3/2$) excited level. In particular, the only allowed Raman transitions
correspond to a change $\Delta m_J=0$ or $\Delta m_J=\pm1$ of the azimuthal
quantum number $m_J$.
A scheme that fulfills the aforementioned constraints is represented in
Figs.~\ref{fig:exp-directions} and \ref{fig:exp_scheme}. The atomic motion along
the $z$ direction is supposed to be frozen thanks to a trapping potential
$m\omega_z^2z^2/2$ such that $\hbar\omega_z$ is much larger than the atomic
kinetic energy. The atom is placed in a uniform magnetic field $B_0$ directed
along the $x$ direction. The role of this magnetic field is to allow for a
selective Raman excitation between two given Zeeman sublevels. More precisely
the Larmor frequency $\omega_{\mathrm{L}}=\mu_{{\rm B}}B_0/\hbar$ ($\mu_{{\rm
B}}$ is the Bohr magneton) is chosen much larger than the two-photon (Raman)
Rabi frequency $\Omega$. Typically we choose $\omega_{\mathrm{L}}/2\pi$ on the
order of a few MHz (i.e.\ $B_0$ on the order of a few Gauss) and $\Omega/2\pi$
in the range $10^5-10^6\,\mathrm{Hz}$. The latter choice is sufficient to ensure
that the splitting $\hbar\Omega$ between the dark-state manifold and the states
$|\pm\rangle$ is large compared to the two-photon Doppler shift, as required for
the adiabatic approximation to be valid. The state $|0\rangle$ is chosen equal
to the $|F=1,m_F=0\rangle$ sublevel and the states $|j\rangle$ with
$j=1,\ldots,4$ are the $|F=2,m_F=\pm1\rangle$ and $|F=1,m_F=\pm1\rangle$
sublevels. Here the quantization axis is the $x$ axis, parallel to the direction
of the magnetic field $\mathbf{B}_0$. As indicated in
Fig.~\ref{fig:exp_scheme}(a), the transition between $|0\rangle$ and
$|1\rangle\equiv|F=2,m_F=+1\rangle$ is driven by a pair of laser beams $(A,B)$
with a frequency difference equal to
$\omega_{\mathrm{hf}}+\omega_{\mathrm{L}}/2$, where $\omega_{\mathrm{hf}}$ is
the hyperfine splitting between the $F=1$ and $F=2$ manifolds
($\omega_{\mathrm{hf}}/2\pi$ is on the order of $7\,\mathrm{GHz}$ for
$^{87}\mathrm{Rb}$). The laser beam $A$ propagates along the $y$ axis (wave
vector $k{\bf\mathbf{e}_y}$, where ${\bf\mathbf{e}_y}$ is a unit vector). It is
linearly polarized along $x$, so that it carries no angular momentum along the
$x$ axis. The laser beam $B$ propagates along the $x$ axis (wave vector
$-k{\bf\mathbf{e}_x}$) and is circularly ($\sigma_{-}$) polarized. In the
transition $|0\rangle\to|1\rangle$ the momentum change of the atom is $\hbar{\bf
k_1}=\hbar k({\bf\mathbf{e}_x}+{\bf\mathbf{e}_y})$. One can readily check that
the transition $|0\rangle\to|1\rangle$ is the only one that is driven resonantly
by this pair of beams, thanks to the fact that the Land\'e factors are opposite
for the $F=1$ and $F=2$ manifolds, as one can see in Fig.~\ref{fig:exp_scheme}(a).
Similarly, the transition between $|0\rangle$ and
$|3\rangle\equiv|F=2,m_F=-1\rangle$ is driven by a pair of laser beams $(A',B')$
with a frequency difference equal to
$\omega_{\mathrm{hf}}-\omega_{\mathrm{L}}/2$. The beam $A'$ propagates along $y$
with wave vector $-k{\bf\mathbf{e}_y}$ and is linearly polarized along $x$. The
beam $B'$ propagates along $x$ with wave vector $k{\bf\mathbf{e}_x}$ and is
circularly ($\sigma_{+}$) polarized. The atomic momentum change in the
transition $|0\rangle\to|3\rangle$ is $\hbar{\bf k_3}=-\hbar{\bf k_1}$. The
difference in the frequencies of $A$ and $A'$ is chosen large enough so that no
transition is driven with a significant probability by the pairs of beams
$(A,B')$ and $(A',B)$.
The two remaining states of the tetrapod configuration are
$|2\rangle\equiv|F=1,m_F=+1\rangle$ and $|4\rangle\equiv|F=1,m_F=-1\rangle$. The
coupling between these states and the state $|0\rangle$ is provided by a single
pair of laser beams $(A'',B'')$, as in the recent experiment
\cite{Spielman09PRL}, in which the $\Lambda$ (ladder) type coupling was
generated within the Zeeman sublevels of the $F=1$ manifold. The wave vector of
$A''$ is $k{\bf\mathbf{e}_y}$ and this beam is linearly polarized along $x$. The
beam $B''$ propagates along $x$ with wave vector $k{\bf\mathbf{e}_x}$ and is
circularly ($\sigma_{+}$) polarized. The frequency difference between the beam
$A''$ and $B''$ is $\omega_{\mathrm{L}}/2$ so that the pair $(A'',B'')$
resonantly drives the transition $|0\rangle\to|2\rangle$ with a momentum
transfer $\hbar{\bf k_2}=\hbar k({\bf\mathbf{e}_x}-{\bf\mathbf{e}_y})$, and the
transition $|0\rangle\to|4\rangle$ with a momentum transfer $\hbar{\bf
k_4}=-\hbar{\bf k_2}$. Note that here again we take advantage of the different
signs of the Land\'e factors of the $F=1$ and $F=2$ manifolds: The pair of beams
$(A'',B'')$ cannot resonantly drive a transition between two sublevels of the
$F=2$ manifold []see Fig.~\ref{fig:exp_scheme}(b)]. Note also that another
consequence of two-photon processes is a modification of the energies of the
states $|j\rangle$, via the absorption and stimulated emission of photons in the
same laser beam. It can be accounted for by including these energy shifts in the
choice of the two-photon detunings and, for example, by taking advantage of the
(small) second-order Zeeman shift.
This configuration therefore constitutes a suitable implementation of the scheme
discussed in the first part of this article. The momentum transfers $\hbar{\bf
k_j}=\hbar k(\pm{\bf\mathbf{e}_x}\pm{\bf\mathbf{e}_y})$ form a square in the
$xy$ plane shown in Fig.~\ref{fig:tripod-directions}(b) (subject to the rotation
of the coordinate system by $45^{\circ}$). The intensities of the various beams can be
adjusted so that all Rabi frequencies $\Omega_j$ are equal, once the
Clebsch-Gordan coefficients associated to each Raman transition have been taken
into account (note that the two-photon Rabi frequencies for
$|0\rangle\to|2\rangle$ and $|0\rangle\to|4\rangle$ transitions are equal by
construction). With a one-photon detuning of $3\,\mathrm{nm}$ , which represents
$1/5$ of the fine structure splitting for rubidium atoms, the residual photon
scattering rate is below $1\,\mathrm{s}^{-1}$ for a two-photon Rabi frequency
$\Omega/(2\pi)=10^5\,\mathrm{Hz}$. The corresponding heating rate is thus small
enough to provide enough time for the investigation of the RD coupling
studied in this article.
\section{Conclusions}
In this article we have explained how to produce a spin-orbit coupling of the
RD type for a spin larger than $1/2$. Our scheme makes use of
cold atoms with three or more internal dark states so that their quasi-spin is
equal to or greater than unity. We have analyzed a general scheme in which $N$
laser beams couple $N$ atomic internal ground states to an extra state, thus
forming an $N$-pod setup of light-matter interaction. In this case the atoms
have $N-1$ dark states representing superpositions of the $N$ ground states that
are immune to the atom-light coupling.
We have analyzed in detail the particular case of the tetrapod setup ($N=4$), in
which the center of mass motion of the atoms in their dark state manifold is
described by a three-component spinor and thus corresponds to the motion of a
spin-$1$ particle. We have shown that the resulting spin-orbit coupling can be
made of the RD type and yields three cylindrically symmetric dispersion
branches. Two of them are similar to those known for the familiar RD spin-$1/2$ Hamiltonian, so the atom can exhibit a quasirelativistic behavior
\cite{Juzeliunas08PRAR,Merkl09EPL-ZB,Vaishnav08PRL-ZB} for small wave vectors.
Furthermore, we have shown that there exists an extra branch with a flat
dispersion around zero momentum. We have studied the modifications that this
extra branch brings to the problem of negative refraction of matter waves on a
potential step, and shown that it enhances the negative refraction probability.
Finally we have discussed a possible implementation of the tetrapod setup with
cold alkali-metal atoms. We have shown that in order to avoid heating due to
spontaneous emission, it is possible to choose all the states involved in this
tetrapod scheme among the various Zeeman sublevels of the ground atomic state.
All laser couplings are then provided by stimulated Raman transitions. For
rubidium atoms, realistic parameters yield a residual spontaneous emission rate
below $1\,\mathrm{s}^{-1}$, which makes the observation of this spin-orbit
coupling scheme experimentally feasible.
\begin{acknowledgments}
We thank M. Lewenstein and S. Das Sarma for helpful discussions. This work has
been supported by the Gilibert program, the Lithuanian Science and Studies
Foundation (Grant No.~V-34/2009), the Research Council of Lithuania, the R{\'e}gion
Ile de France IFRAF, the ANR (Grant No.~ANR-08-BLAN-65 BOFL), and the EU projects
SCALA and STREP NAMEQUAM. LKB is a mixed research unit No.8552 of CNRS, ENS, and Universit{\'e}
Pierre et Marie Curie.
\end{acknowledgments}
|
\section{Introduction}
It has been realised over the past decade that the supermassive black hole (SMBH) at the centre of a galaxy bulge plays a major role in galaxy evolution.
The process commonly considered is the negative feedback effect which may determine the final stellar mass of a galaxy so that it ends up with the locally observed universal bulge stellar mass over SMBH mass ratio (e.g. Cattaneo et al. 2009, Fabian 2009), M$_{\rm GAL}$/M$_{\rm BH}\sim$ 700 (500 in Marconi \& Hunt 2003, McLure \& Dunlop 2001 or Ferrarese et al. 2006, 700 in Kormendy \& Gebhardt 2001 and 830 in McLure \& Dunlop 2002). Indeed the correlation that exists between the mass of the central SMBH of local galaxies and their bulge luminosity (Kormendy \& Richstone 1995, Magorrian et al. 1998), stellar mass or velocity dispersion (Gebhardt et al. 2000, Ferrarese \& Merritt 2000), suggests the existence of a physical mechanism connecting the activity of a galaxy nucleus and the bulding of its host galaxy stellar bulge. The most massive galaxies exhibiting the largest over-abundance of $\alpha$-elements with respect to the solar value, this means that they formed their stars the fastest (Thomas et al. 2005) and indeed a downsizing of the typical star forming galaxy has been observed in deep surveys, i.e. the comoving star formation rate activity of the Universe is dominated by more and more massive galaxies going back in lookback time. If negative feedback may provide a way to "kill" galaxies to reproduce the downsizing of galaxies, it does not necessarily explain why most stars in massive ellipticals formed rapidly. Hence positive feedback must have been at play early on shortly after the formation of the proto-galaxy. Major mergers have long been thought to be the best candidates to explain those rapid processes by first triggering a starburst then feeding the central SMBH and activating the QSO phase (Sanders et al. 1988), but this scenario has been subject to debate, in particular since it was found that AGN activity in the nearby Universe was not shown to correlate with galaxy pairs (Li et al. 2008).
Since the masses of SMBH and of stars in galaxies are correlated, it is natural to consider the alternative possibility that both processes are related, with one possibly affecting the other. If radio jets are shown to be able to trigger star formation in galaxies, as suggested by early studies of distant galaxies but more or less abandoned since then, they may represent an interesting candidate process to understand the universal mass ratio since the most massive SMBH will have the largest accretion rates, hence also most powerful radio jets, possibly triggering more efficiently star formation. Because radio jets are difficult to identify, especially in the distant Universe, they have not yet been considered as a major actor in the process of galaxy formation. However, they must have been ubiquitous in the past to explain the ubiquity of SMBH in the center of local galaxies. Together with the following pieces of evidence, this suggests that their role may have been underestimated until now : (i) the discovery of a jet-induced ultra-luminous infrared galaxy (ULIRG), one of the strongest starbursts ever observed, in the system HE0450$-$2958, (ii) the presence of massive extended emission line regions (EELR) surrounding quasars and mostly radio galaxies, (iii) evidence that radio jets were more abundant in the past, (iv) the presence of large concentrations of molecular gas with an offset with respect to their neighboring QSO, possibly resulting from the impact of radio jets, (v) theoretical arguments suggesting that radio jets may trigger star formation efficiently.
Our aim here is not to firmly state that we are convinced that galaxy formation cannot be understood without accounting for the role of radio jets, since existing data and models are too sparse to either infer or reject such statement. However, in the near future, new observatories such as ALMA, eVLA, E-ELT or the JWST will provide crucial observations that should allow us to better address this issue, hence we wish here to bring back to the front this mechanism as a possible driver not only of star formation, but maybe even galaxy formation.
\section{The role of radio jets in star and galaxy formation}
\label{SEC:jetinduced}
Evidence for jet-induced star formation has been found in various objects and environments, either far away from the host galaxy, such as in the lobe of radio jets (e.g. van Breugel et al. 1985), or inside the host galaxy, resulting in the so-called radio-optical alignment (e.g. McCarthy et al. 1987, Rees 1989). McCarthy et al. (1987) noticed that both the stellar continuum and size of the emission line region of 3CR radio galaxies at $z\ge$0.6 were highly elongated in parallel with their radio jets. This so-called radio--optical alignment was interpreted as evidence that the radio jets interact with the interstellar medium and stimulate large-scale star formation in the host galaxy (see also Rees 1989, Rejkuba et al. 2002, Oosterloo \& Morganti 2005). Other mechanisms than jet-induced star formation have been suggested in the literature to explain the radio-optical alignment effect such as the scattering of light from the central AGN (e.g. Dey \& Spinrad 1996) or the nebular continuum emission from warm line-emitting regions (Dickson et al. 1995). However, many of these objects show clear evidence of star formation. This is, in particular, the case of 4C 41.17 ($z$=3.8), which rest-frame UV continuum emission is aligned with the radio axis of the galaxy, unpolarized and showing P Cygni-like features similar to those seen in star-forming galaxies (Dey et al. 1997). The most dramatic events, with star formation rates as high as 1000 M$_{\odot}$ yr$^{-1}$, are associated with very luminous radio galaxies at redshifts up to z $\sim$ 4 often located at the center of proto-clusters of galaxies (Dey et al. 1997, Bicknell et al. 2000, Zirm et al. 2005). Even in the closest radio galaxy Centaurus A, the reality of jet-induced star formation has been the center of a debate. Early on, Blanco et al. (1975) noticed the presence of extended gaseous filaments with bright knots exhibiting strong H$\alpha$ and UV emission or loose chains of blue compact objects aligned with the radio jets of the central active nucleus. The origin of these optical filaments has been subject to a long debate regarding the main source of their ionization until the young and massive stars inhabiting these optical knots were individually resolved showing that jet-induced star formation must indeed have taken place in this system (Rejkuba et al. 2002). If there is indeed a mechanism through which radio jets may induce star formation in quasars or radio galaxies, then this mechanism may also be at play inside their scaled-down version, i.e. microquasars. The relativistic jets of the microquasar GRS 1915$+$105, which show apparent superluminal motion, are aligned with two IRAS sources themselves associated with radio knots. One of the radio source was resolved with the VLA showing a non thermal extension pointing also in the direction of the microquasar on one side and the IRAS source on the other side (see Fig.~\ref{FIG:GRS1915}). Both IRAS sources appear to be H II regions ionized by late O or early B stars (Rodriguez \& Mirabel 1998).
\begin{figure
\begin{center}
\includegraphics[width=1.8in]{elbaz_fig1.eps}
\caption{VLA mosaic at 20-cm of the surroundings of GRS 1915+105 (Rodriguez \& Mirabel 1998). The red dashed line illustrates the direction of the radio jets of the microquasar. }
\label{FIG:GRS1915}
\end{center}
\end{figure}
Inversely, radio jets are known to be efficient in producing a negative feedback on the environment of the massive radio galaxies located at the center of the so-called "cooling flow" clusters (see Fabian 2009, Cattaneo 2009). There, even when its cooling time gets lower than 1 Gyr, the intracluster medium (ICM) temperature does not fall below a given threshold which implies that non gravitational heating is taking place. The presence of buoyant radio bubbles brings direct evidence that on these large scales, radio jets can prevent cooling from happening if they are long-lived. At the smaller scale of individual galaxies, there is no such strong direct evidence for negative feedback and the process must be different since radio jets are collimated, hence should escape galaxies without affecting the bulk of their ISM. For this reason, a second mode has been considered, called the radiative or wind mode (see discussion in Fabian 2009). Yet observationally, active nuclei are often found associated with star formation and both effects may not be exclusive in the sense that positive feedback may occur first at early stages in galaxy evolution followed by the quenching of star formation at a later stage.
The discovery of a case of a ULIRG in which the intense star formation activity appears to be triggered by the impact of a radio jet (Elbaz et al. 2009 and Sect.~\ref{SEC:HE0450}) suggests that, at least in some cases, positive feedback from radio jets can be a powerful mechanism. This is a scaled up version of the prototypical example for jet-induced star formation, Minkowski's Object, and in both cases, no evidence has been found for the presence of stars older than the starburst which suggests that the process may be responsible for the triggering of the formation of a whole galaxy. The fact that at $z$$\sim$2, the comoving space density of radio galaxies with powerful jets was 1000 times higher than at the present epoch (Willott et al. 2001), suggests that the process taking place in these locally rare cases may have been an important actor in the process of galaxy formation and evolution. Relic traces of this activity may be seen in the presence of the extended emission line regions (EELRs) surrounding QSOs, and radio galaxies, with masses which may be as large as that of a massive galaxy, i.e. 10$^{10}$ M$_{\odot}$ (Fu \& Stockton 2008). Those gas reservoirs may represent the seeds for new galaxies if they are shocked by a radio jet as it may have happened in HE0450$-$2958 and such events may have been more common in the past. Indeed, hydrodynamical simulations of radiative shock-cloud interactions indicate that for moderate gas cloud densities ($>$ 1 cm$^{-3}$) such as those observed in EELRs, cooling processes can be highly efficient and result in more than 50\,\% of the initial cloud mass cooling to below 100 K, hence leading to star formation (Fragile et al. 2004). The mere existence of these EELRs itself is suggestive of the impact of radio jets on galaxies since their abundance ratios are similar to those observed in the host galaxies of their neighboring AGN, which Fu \& Stockton (2008) interpret as evidence that they were expelled out of the host galaxy by the radio jets.
Because they are collimated, radio jets may not affect very strongly the host galaxy of the active nucleus that produces them but they may produce an impact on the local environment. We present here two elements favoring this interpretation:
- the pair of ULIRGs in the HE0450$-$2958 system, discussed in the next section, provide evidence for a major event in the formation process of a galaxy related to radio jets (see Sect.~\ref{SEC:HE0450}).
- several AGNs have been found with neighboring molecular gas clouds, which may be the result of the impact of a radio jet on its environment (see Sect.~\ref{SEC:CO}).
\section{The case of HE0450$-$2958}
\label{SEC:HE0450}
HE0450$-$2958 is a nearby luminous ($M_V$$=$$-25.8$) radio quiet quasar located at a redshift of $z$=0.2863 (Canalizo \& Stockton 2001) and presently the only known quasar for which no sign of a host galaxy has been found (Magain et al. 2005). This object has been the center of a debate and various scenarios have been proposed to explain the absence of detection of a host galaxy. Since dust extinction could potentially explain it, we imaged HE0450$-$2658 with VISIR, the Very Large Telescope Imager and Spectrometer in the Infrared (VISIR) at the ESO-VLT. Indeed, a bright source was detected by IRAS at the approximate location of HE0450$-$2958 within the position error bar of 5 arcmin which, at the luminosity distance of the QSO ($z$=0.2863), would translate into the total infrared luminosity of an ultra-luminous infrared galaxy (ULIRG, L$_{\rm IR}$=L(8--1000\,$\mu$m)$\geq$10$^{12}$ L$_{\odot}$, de Grijp et al. 1987, Low et al. 1988). In a first analysis, a unique mid-IR source was detected with VISIR associated with the QSO, but combined with HST--NICMOS imaging at 1.6\,$\mu$m, it was not possible to disentangle a local dust torus from a whole dust-obscured galaxy (Jahnke et al. 2009). However, a re-analysis of the VISIR data led to the detection of a second mid-IR source in the field, associated with the 7 kpc distant companion galaxy to the QSO (Elbaz et al. 2009).
\begin{figure}[h
\begin{center}
\includegraphics[width=2.2in]{elbaz_fig2a.ps}
\includegraphics[width=1.4in]{elbaz_fig2b.ps}
\includegraphics[width=1.4in]{elbaz_fig2c.ps}
\caption{\textbf{\textit{(a)}} Full spectrum of the companion galaxy of HE0450$-$2958 from 0.3 to 400\,$\mu$m in the rest-frame (red line). The plain green and blue lines represent the median and 68 percentile reddest SEDs from the Atlas of 47 quasar SEDs of Elvis et al. (1994), normalized to the VISIR measurement at 11.3\,$\mu$m (observed; 8.9\,$\mu$m rest-frame) associated to the QSO (filled blue circle). Filled blue star for the companion galaxy. \textbf{\textit{(b)}} VISIR image with contours from NICMOS-F160W at 1.6\,$\mu$m (in yellow, 1.2\,$\mu$m in the rest-frame, Jahnke et al. 2009) and ATCA CO 1--0 (dark blue lines, from Papadopoulos et al. 2008). \textbf{\textit{(c)}} ATCA 6208 MHz radio continuum (dark blue lines, from Feain et al. 2007) emission associated with the radio jets overlayed on the VISIR image.}
\label{FIG:im}
\end{center}
\end{figure}
The VISIR image provides the first direct evidence that the IRAS "source" is in fact made of two well separated objects, the QSO and its companion galaxy. The infrared SED of the "system" (Fig.~\ref{FIG:im}a) peaks around $\sim$50\,$\mu$m (rest-frame) which indicates that the bulk of the far infrared light arises from star formation (the QSO and companion galaxy are both contained in the single IRAS measurement) and not from and active nucleus that would drop beyond $\sim$20\,$\mu$m (Netzer et al. 2007, minimum temperature of $\sim$200 K typically). The IR SED of the companion galaxy, which produces the bulk of the far-IR light, is very well fitted with a standard star forming template SED (red plain line, from the library of Chary \& Elbaz 2001). The total infrared luminosity of the companion galaxy is L$_{\rm IR}$$\sim$2$\times$10$^{12}$ L$_{\odot}$, which corresponds to a SFR$\sim$340 M$_{\odot}$yr$^{-1}$ (conversion factor of Kennicutt 1998). Hence HE0450$-$2958 appears to be a composite system made of a pair of ULIRGs (both galaxies have similar L$_{\rm IR}$), where both mechanisms, star formation and QSO activity, are spatially separated by 7 kpc in two distinct sites. The fact that the starburst is associated with the companion galaxy is reinforced by the finding that the molecular gas traced by the CO molecule appears to avoid the QSO but to peak at the location of the companion galaxy (dark blue contours in Fig.~\ref{FIG:im}b) close to the peak mid-infrared emission measured with VISIR (Papadopoulos et al. 2008). This both suggests that there is a large amount of dust in this galaxy and that there is a large gas reservoir (M(H$_2$)$\simeq$2.3$\times$10$^{10}$ M$_{\odot}$) to fuel an intense star formation event. No evidence for the presence of an old stellar population is found in the VLT-FORS optical spectrum of the companion, although the radiation of such stars could be diluted in that of the youngest population. The age of the dominant stellar population is about 40--200 Myr suggesting that the galaxy was recently born. The stellar mass of the galaxy (M$_{\star}$$\simeq$[5--6]$\times$10$^{10}$ M$_{\odot}$) is consistent with this timescale for the SFR derived from its IR luminosity.
The companion galaxy coincides with one of the two radio lobes on both sides of the QSO HE0450$-$2958 (Fig.~\ref{FIG:im}c), suggesting that it is hit by a radio jet emitted by the QSO. Evidence for shock excitation of the ISM in the companion galaxy, as traced by an high [NII]/H$\alpha$ emission line ratio (Letawe et al. 2008b), confirms that the association of the jet and the galaxy is physical and not due to a projection effect.
The VISIR image itself shows the presence of a bridge of mid-IR emission between the QSO and the companion Fig.~\ref{FIG:jets}-left.
. Altogether, this suggests that it is the radio jet that is triggering the starburst in the companion. As a result, HE0450$-$2958 is comparable to Minkowski's Object, a proto-typical case for jet-induced star formation (van Breugel et al. 1985). Minkowski's Object is a newly formed galaxy, with a stellar population age of only 7.5 Myr and a stellar mass of 1.9$\times$10$^7$ M$_{\odot}$ (Croft et al. 2006), aligned with the radio jet of NGC 541, an FR I galaxy (Fanaroff \& Riley 1974) located in the cluster Abell 194. The companion galaxy of HE0450$-$2958 might be a scaled-up version of Minkowski's Object, i.e. a massive galaxy whose formation was jet-induced.
\begin{figure}[h
\begin{center}
\includegraphics[width=1.98in]{elbaz_fig3a.eps}
\includegraphics[width=3.02in]{elbaz_fig3b.eps}
\caption{\textbf{\textit{(left)}} composite image of the HE0450$-$2958 system with VISIR 11.3\,$\mu$m image (blue) and HST in the R-band (yellow). \textbf{\textit{(right)}} Composite image showing the details of Minkowski's Object (Croft et al. 2006). Optical image of the field of Abell 194, overlaid with radio continuum ( purple), HI data (dark blue), and H$\alpha$ data (light blue).}
\label{FIG:jets}
\end{center}
\end{figure}
Evidence for the presence of large amounts of matter in the surrounding of radio quasars, in the form of massive gas clouds exhibiting strong emission lines with high excitation levels, has been gathered since the early 60s. These extended emission lines regions (EELRs, see Stockton \& MacKenty 1983) can reach masses of several 10$^{10}$ M$_{\odot}$ (Fu \& Stockton 2008, Stockton et al. 2007). Spectro-imaging and narrow-band filter imaging in the fields of radio quasars demonstrated that EELRS were made of clouds of gas, whose strong emission lines were excited by the hard radiation of the quasar. Fu \& Stockton (2008) found that low metallicity quasar host galaxies happened to be surrounded by EELRs with similar abundance ratios, which they interpreted as evidence that the EELRs had been expelled from the host galaxy itself. The double-lobed morphologies of extragalactic radio sources shows that the relativistic jets not only couple strongly with the ISM of the host galaxies, but are capable of projecting their power on the scales of galaxy haloes and clusters of galaxies (up to $\sim$1 Mpc, Tadhunter 2007).
Several EELRs were found in the surroundings of HE0450$-$2958 following the axis of the radio jets. They have been detected up to 30 kpc away from the QSO (Letawe et al. 2008a). The N-W blob of gas, photoionized by the radiation of HE0450$-$2958, is itself located in the direction of the radio jet opposite to the one pointing at the companion galaxy. The thermodynamic state of EELRs is uncertain since the densities one can infer from their optical spectra are of the order of n$_{\rm e}$=100--300 cm$^{-3}$ and T$_{\rm e}$=10,000--15,000 K (Fu \& Stockton 2008), which imply that they should be gravitationally unstable and have already started forming stars. Hence Fu \& Stockton (2008) modeled these regions using two components, a low density component, where most of the mass is locked at densities of order n$_{\rm e}\sim$1 cm$^{-3}$, and a higher density component responsible for the observed emission lines. Hydrodynamical simulations of radiative shock-cloud interactions indicate that for moderate gas cloud densities ($>$ 1 cm$^{-3}$), cooling processes can be highly efficient and result in more than 50\,\% of the initial cloud mass cooling to below 100 K (Fragile et al. 2004). Hence, the companion galaxy of HE0450$-$2958 could have formed from a seed EELR that was hit by the radio jet. This process could explain why radio jets do not always produce the same effect.
\section{A test of jet-induced galaxy formation for future instrumentation: offset molecular gas}
\label{SEC:CO}
Here we search for other cases such as HE0450$-$2958, where a large mass of molecular gas is found offset with respect to the QSO and associated with newly formed stars which formation could be induced by a radio jet. Such an offset between molecular gas and QSOs may be a common feature in distant radio sources as suggested by Klamer et al. (2004) who produced a systematic search in $z$$>$3 CO emitters for an AGN offset with respect to the CO concentrations and for a connection with radio jets. Out of the 12 $z>$ 3 CO emitters that they studied, six showed evidence of jets aligned with either the CO or dust, five have radio luminosities above 10$^{27}$ W Hz$^{-1}$ and are clearly AGNs, and a further four have radio luminosities above 10$^{25}$ W Hz$^{-1}$ indicating either extreme starbursts or possible AGNs. In the following, we discuss three proto-typical cases classified with increasing redshift from $z$$=$2.6 to 6.4 that may be compared with HE0450$-$2958.
At $z$=1.574, 3C18 is radio loud quasar with recently formed radio jets, as inferred from their small physical size. A large mass of molecular gas ($M_{\rm H_2}$$\sim$(3.0$\pm$0.6)$\times$10$^{10}$ M$_{\odot}$), inferred from CO 2--1 emission, is found associated to the quasar with a positional offset of $\sim$20 kpc (Willott et al. 2007).
At $z$= 2.6, TXS0828$+$193 is a radio galaxy with a neighboring CO gas concentration of $\sim$1.4$\times$10$^{10}$ M$_{\odot}$ located 80 kpc away and shows no evidence for an underlying presence of stars or galaxy (Nesvadba et al. 2009). An upper limit of 0.1 mJy at 24\,$\mu$m, from the MIPS camera onboard Spitzer, was obtained by Nesvadba et al. (2009), who concluded from this limit that no major starburst with more than several hundred solar masses per year could be taking place associated with this CO concentration. We used the library of SED templates of local galaxies from Chary \& Elbaz (2001) to convert this mid infrared measurement into an upper limit for the total IR luminosity at this position of L$_{\rm IR}^{\rm max}$$\sim$7.8$\times$10$^{12}$ L$_{\odot}$, which would translate to a maximum SFR of $\sim$1340 M$_{\odot}$ yr$^{-1}$ using the Kennicutt (1998) conversion factor for a Salpeter IMF. Hence there is still room for a large amount of star formation in this object that may be in a similar stage than HE0450$-$2958 but at a much larger distance.
At $z$= 4.695, BR 1202$-$0725 is the first high-redshift quasar for which large amounts of molecular gas was detected (Omont et al. 1996).
The CO map presents two well separated emission peaks which coincide with radio continuum emission interpreted as evidence for the presence of radio jets by Carilli et al. (2002). The radio jets themselves would be too faint to be detected at the sensitivity level of the radio image but marginal evidence for variability suggest that the radio emission is not due to star formation. Omont et al. (1996) suggested that the double CO emission might be due to gravitational lensing, but no optical counterpart of the QSO is found associated with the second source and the CO (2--1) line profiles are different for the two components (Carilli et al. 2002). Klamer et al. (2004) suggested that stars may have formed along the radio jets, providing both the metals and dust for cooling and "conventional" star formation to take place afterwards.
In the case of HE0450$-$2958, the mid infrared emission is spatially associated with the offset CO concentration suggesting that similar systems might be found not only through CO imaging but also infrared imaging. SDSS160705+533558, located at a redshift of $z$=3.65, might be a distant analog of HE0450$-$2958 in that respect since it also presents a positional offset between the maximum sub-millimeter (from the submillimeter arra, SMA) and optical emission (Clements et al. 2009).
A key test for the role of radio jets in the formation of galaxies will be the detection of either molecular gas or mid infrared emission with a positional offset with respect to their neighboring QSO. This test will be fulfilled with the advent of ALMA, the JWST or project instruments such as METIS (mid infrared E-ELT Imager and Spectrograph) for the ELT (extremely large telescope).
|
\section{Introduction}
The existence of a thick disk in our Galaxy was revealed by \cite{gilmore1983},
who analyzed starcounts towards the South Galactic Pole. Thanks to the many studies carried out
since then, the main spatial, kinematic, and chemical features
of this population are well established.
Thick disks have been also observed in many disk galaxies \citep{yoachim2006}, and they represent
the frozen relics of the first phases of disk galaxy formation \citep{freeman2002}.
However, in spite of the many scenarios proposed until now, the origin of this component is still unclear.
In the context of CDM hierarchical galaxy formation models, it is possible that thick disks are formed by the heating of a pre-existing thin disk through a minor merger \citep[e.g.][]{villalobos2008}, by accretion of stars from disrupted satellites \citep{abadi2003}, or by the stars formed {\it in situ}
from gas-rich chaotic mergers at high redshift \citep{brook2005}. On the other hand, simulations suggest that
thick disks could simply be produced through secular radial migration of stars induced by the spiral arms \citep{roskar2008, schonrich2009}.
In any event, most astronomers agree that our thick disk is formed of an old stellar population
with an age of 8-12 Gyr \citep[e.g.][and references therein]{haywood2008}.
The bulk of the thick disk stars have metallicity in the range
$-1\la $[Fe/H]$\la -0.3$ ( [Fe/H]$\simeq -0.6$, on average) with enhanced [$\alpha$/Fe] \citep{bensby2005, reddy2006}, but note that tails with metal-poor stars down to [Fe/H]$\simeq -2$ \citep{chiba2000} and metal-rich stars up to [Fe/H]$\simeq 0$ \citep{bensby2007} have also been revealed.
Moreover, according to \citet{ivezic2008}, a mild vertical metallicity gradient shifts
the mean metallicity to [Fe/H]$\simeq -0.8$ beyond $|z|\ga 3$ kpc.
The spatial distribution is usually modeled with a symmetric exponential density distribution as a function of galactocentric coordinates $(R,z)$.
Its scale height spans a wide range of measurements, between $h_z=640$~pc and 1500~pc, while the local normalization varies beetween 13\% and 2\% in anticorrelation with $h_z$
\citep[see Fig. 3 of][]{arnadottir2008}.
The distribution above the galactic plane is supported by a vertical velocity dispersion, $\sigma_W\simeq 40$ km~s$^{-1}$, which is associated with an asymmetric drift of $\sim 50$ km~s$^{-1}$, relative to the local standard of rest.
Significant asymmetries have also been detected, such as the prominent Hercules thick disk cloud \citep{parker2003, juric2008}, which could correspond to a merger remnant or indicate a triaxial thick disk \citep{larsen2008}.
In this letter, we present new results regarding the vertical rotation gradient and, for the first time to our knowledge, evidence of a metallicity-rotation correlation in the thick disk stellar population.
\section{The SDSS -- GSC-II catalog}
This study is based on a new kinematic catalog derived by assembling the astrometric
parameters extracted from the database used for the construction of the Second Guide Star Catalog \citep[GSC-II; ][]{lasker2008}
with spectro-photometric data from the Seventh Data Release of the Sloan Digital Sky Survey \citep[SDSS DR7; e.g.\ ][]{abazajian2009, yanny2009}.
The SDSS--GSC-II catalog contains positions, proper motions, classification, and $ugriz$ photometry for 77 million sources down to $r\approx 20$, over 9000 square-degrees.
Proper motions are computed by combining multi-epoch positions from SDSS DR7 and the GSC-II database. Typically, 5-10 observations are available for each source, spanning $\sim50$ years.
Total errors are in the range 2-3 mas~yr$^{-1}$ for $16<r<18.5$, comparable with those of the SDSS proper motions \citep{munn2004}, as confirmed by external comparisons against QSOs.
The construction and properties of this catalog are described in detail by Smart et al. (2010, in preparation), while a concise description can be found in \citet{spagna2009}.
Radial velocities ($\sigma_{Vr}< 10$ km~s$^{-1}$) and astrophysical parameters ($\sigma_{\rm Teff}\simeq 150$ K, $\sigma_{\log g}\simeq 0.25$, $\sigma_{\rm [Fe/H]}\simeq 0.20$) are available for 151\,000 sources cross-matched with the SDSS spectroscopic catalog.
From this list, we select sources with $4500$ K$<T_{\rm eff}< 7500$ K and $\log g> 3.5$, corresponding to FGK (sub)dwarfs, and apply the color thresholds from \citet{klement2009}
in order to remove turn-off stars.
Spectro-photometric distances are computed by means of metallicity-dependent absolute magnitude relations, $M_r=f(g-i,{\rm [Fe/H]})$, from \citet{ivezic2008}.
Here, the observed magnitudes are corrected for interstellar absorption via the extinction maps of
\citet{schlegel1998}, while the spectroscopic [Fe/H] is used, instead of the photometric metallicity applied by \citet{ivezic2008}.
The mean distance of the sample is $\sim 2$ kpc, while most (92\%) of the sources are distributed between 0.5 kpc $< |z| < 3.5$ kpc and 6 kpc $<R<11$ kpc. The typical accuracy of the $M_r$ calibration is 0.3 mag (random) and 0.1 mag (systematic), which corresponds to distance errors of $\Delta d/d=15$\% and 5\%, respectively.
Finally, 3D velocities in the galactocentric reference frame, $(V_R, V_\phi, V_z)$, are derived by assuming $R_\odot=8$ kpc, solar motion $(U_\odot,V_\odot,W_\odot)$ from \citet{dehnen1998}, and local standard of rest velocity of 220 km~s$^{-1}$.
In order to produce an accurate sample, we select only stars with $(i)$ proper motion errors $<10$ mas~yr$^{-1}$ per component,
$(ii)$ errors on the velocity components $<50$ km~s$^{-1}$, $(iii)$ total velocity $< 600$ km~s$^{-1}$, $(iv)$ distance $<5$ kpc, and $(v)$ magnitude $13.5<g<20.5$.
Overall, the {\it kinematic} catalog contains 46\,000 stars; in the following sections a subsample of 27\,000 low metallicity dwarfs with $-3<$[Fe/H]$<-0.5$ will be used as {\it tracers} of the inner halo and thick disk and analyzed in details.
\section{Analysis and results}
\subsection{Vertical rotation gradient}
\label{sect:verticalGradient}
Figure \ref{spagna_fig1} shows the $V_\phi$ distribution of 6538 stars
with $1.0$ kpc $< \left|z\right|\le 1.5$ kpc and [Fe/H]$<-0.5$. In this sample, the contamination of thin disk stars is expected to be negligible\footnote
{Assuming a standard model with a thin disk and a thick disk having scale-heights of 300 pc and 900 pc, respectively, and a thick disk normalization of 10\% at $z$=0 pc, about half of the stars belongs to the thin disk for $1.0$ kpc $<\left|z\right|< 1.5$ kpc, but only a few percent of them with [Fe/H]$<-0.5$ \citep[cfr. e.g.][]{aumer2009}. Also, we estimate the contamination of metal poor thin disk stars does not exceed 10\%, even if we adopt a thick disk with a shorter $h_z=580$ pc and a local normalization of 13\% \citep{chen2001}. }, so that
we fit the distribution with only two gaussian populations, corresponding to the thick disk and halo.
The least-squares solution of the two-component model is good, although the counts at $V_\phi\approx$ 220 km~s$^{-1}$ are slightly underestimated ($\sim-16$\%) and the velocity peak is overestimated of about 7\%;
this explains a non-optimal $\chi^2_\nu=3.18$.
(If we force a third gaussian component corresponding to the thin disk, the formal goodness of fit improves significantly, $\chi^2_\nu=1.37$, but the solution becomes ill-conditioned with an inaccurate thin disk normalization of $(19\pm 6)$\%. )
\begin{figure}[]
\resizebox{\hsize}{!}{\includegraphics[trim=0mm 1mm 0mm 1mm clip=true]{13538f1.ps}}
\caption{\footnotesize
Histogram of the velocity distribution, $V_\phi$, of the kinematic sample with [Fe/H]$<-0.5$, between $\left| z\right|=$1.0 kpc and 1.5 kpc. The thick solid line shows the best fit of a two Gaussian component model, representing the thick disk and halo populations (thin lines).}
\label{spagna_fig1}
\end{figure}
\begin{table}
\caption{Parameters of a two-component Gaussian best fit (thick disk and halo) for six height intervals. }
\label{table:1}
\begin{center}
\begin{tabular}{crllllcc}
\hline\hline
& & \multicolumn{2}{c}{\sc Thick Disk} & \multicolumn{2}{c}{\sc Halo} & & \\
$\langle|z|\rangle$ & N & $\langle V_\phi\rangle$ & $\sigma_{V\phi}$ & $\langle V_\phi\rangle$ & $\sigma_{V\phi}$ & $\frac{\rho_{\rm TD}}{\rho_{\rm tot}}$ & $\chi^2_\nu$\\
(kpc) & & \multicolumn{2}{c}{(km s$^{-1}$)} & \multicolumn{2}{c}{(km s$^{-1}$)} & (\%) & \\
\hline
0.76 & 7022 & 186$\pm$1& 34$\pm$1 & 46$\pm$5 & 92$\pm$3 & 74$\pm$2 & 5.18\\
1.24 & 6538 & 173$\pm$1& 39$\pm$1 & 32$\pm$4 & 90$\pm$3 & 68$\pm$2 & 3.18\\
1.73 & 4753 & 163$\pm$1& 44$\pm$2 & 29$\pm$11 & 96$\pm$5 & 60$\pm$4 & 1.42\\
2.23 & 3044 & 155$\pm$2& 47$\pm$3 & 36$\pm$10 & 90$\pm$4 & 48$\pm$5 & 1.83\\
2.73 & 1637 & 144$\pm$4& 42$\pm$5 & 49$\pm$6 & 97$\pm$3 & 29$\pm$6 & 1.14\\
3.36 & 988 & 166$\pm$11 & 41$\pm$11& 44$\pm$8 & 90$\pm$4 & 13$\pm$7 & 0.80\\
\hline
\end{tabular}
\end{center}
\end{table}
\begin{figure*}[]
\begin{center}
\resizebox{0.90\hsize}{!}{\includegraphics[trim=0mm 0mm 0mm 0mm clip=true]{13538f2.ps}}
\caption{Velocity-metallicity distribution of 20\,251 stars
with $\left|z\right|=$1.0-3.0 kpc and [Fe/H]$<-0.3$.
The dashed line indicates the thick disk rotation, $V_\phi=173$ km~s$^{-1}$ at $\langle \left|z\right|\rangle = 1.24$ kpc (Table \ref{table:1}).
The box defines the region, shown in Fig. \ref{spagna_fig3}, in which the thick disk population dominates.}
\label{spagna_fig2}
\resizebox{0.80\hsize}{!}{\includegraphics[trim=0mm 0mm 0mm 0mm clip=true]{13538f3.ps}}
\caption{
\footnotesize
Iso-density contours of 13\,108 relatively metal-poor stars with $-1.0<$[Fe/H]$<-0.3$ and $\left|z\right|=$1.0-3.0 kpc. White crosses mark the ridge line of the maximum likelihood $V_\phi$ vs. [Fe/H].}
\label{spagna_fig3}
\end{center}
\end{figure*}
\begin{figure*}[]
\resizebox{0.90\hsize}{!}{\includegraphics[clip=false]{13538f4.ps}}
\caption{ $V_\phi$ vs.\ [Fe/H] distribution of stars with $-1.0<$[Fe/H]$<-0.5$.
Black crosses mark the rejected stars beyond 3$\sigma$ of the velocity ellipsoid of the thick disk to minimize the contamination from halo stars (see Sect. \ref{sect:correlation}). The solid line connects the mean velocities $\langle V_\phi\rangle$, which are plotted with 2$\sigma$ error bars. The dashed lines indicate the $\pm 1\sigma$ spread of the velocity distribution.
Top, middle, and bottom panels refer to $\left| z\right|=$1.0--1.5 kpc, 1.5--2.0 kpc, and 2.0--3.0 kpc, respectively.}
\label{spagna_fig4}
\end{figure*}
The same procedure is repeated for six height bins: $\Delta\left| z\right| =0.5$--1.0 kpc, 1.0--1.5 kpc, 1.5--2.0 kpc, 2.0--2.5 kpc, 2.5--3.0 kpc, and 3.0--4.0 kpc. The results are reported in Table \ref{table:1}, which lists mean height, number of stars, mean rotation velocities and dispersions, fraction of thick disk stars, and reduced $\chi^2_\nu$.
The halo parameters appear quite stable: on average, $V_{\phi}\simeq 37 \pm 3$ km~s$^{-1}$ ($1<\left| z\right|\le 4$ kpc), which indicates a slow prograde rotation of the inner halo, in agreement with some authors \citep{chiba2000,kepley2007} but different from others that favor a non-rotating inner halo \citep{vallenari2006, smith2009, bond2009}.
The halo velocity dispersion also appears rather constant up to $\left|z\right|\simeq 4$ kpc, with a mean value
of $\sigma_{V\phi}=93\pm 2$ km~s$^{-1}$ (uncorrected for the velocity errors).
Conversely, the thick disk shows a monotonic decreasing of the rotation velocity from
$V_\phi=186$ km~s$^{-1}$ to 146 km~s$^{-1}$, for height from 0.5 kpc to 3 kpc.
In the highest bin (3 kpc $\le \left|z\right|< 4$ kpc), $V_\phi$ increases to 166$\pm 11$ km~s$^{-1}$, but we think this is a spurious effect of both the larger velocity errors and the small fraction, $(13\pm 7)$\%, of thick disk stars that are strongly entangled with the halo population.
Similarly, in the same $z$-range, the velocity dispersion increases from $\sigma_{V\phi}=34$ km~s$^{-1}$ to $\sim$45 km~s$^{-1}$, in part because of the tangential velocity errors that scale with distance.
We exclude the highest bin and also the lowest, as it is probably contaminated by thin disk stars which are difficult to deconvolve from the thick disk population.
Thus, we estimate the gradient,
\begin{equation}
{\partial \langle V_{\phi}\rangle}/{\partial \left|z\right|} = -19 \pm 2 \hbox{\rm { } km~s$^{-1}$ kpc$^{-1}$}
\end{equation}
and the extrapolated intercept, $V_{\phi}(z=0)=196\pm 3$ km~s$^{-1}$.
Our result is significantly smaller than the value, $-30\pm 3$ km~s$^{-1}$~kpc$^{-1}$, measured by
\citet{chiba2000}, who analyzed stars with abundance in the range, $-0.8\le$[Fe/H]$\le -0.6$, where the thick disk dominates. A similar trend was estimated by \citet{girard2006}, \citet{carollo2009}, and by \citet{bond2009}, although they adopted a nonlinear function.
Instead, a shallower slope was found by \citet{majewski1992}, who derived a gradient of $-21\pm 1$ km~s$^{-1}$~kpc$^{-1}$ for $\left|z\right|<5$ kpc, after separating the halo population from that of the thick disk.
A low kinematical gradient was also found by \citet{spagna1996} and, more recently, by \citet{allende-prieto2006}, who estimated $-10$ km~s$^{-1}$~kpc$^{-1}$ and $-16$ km~s$^{-1}$~kpc$^{-1}$, respectively.
The difference between these results can be explained, at least in part, by thin disk and halo stars contamination, which tends to produce steeper velocity gradients.
\begin{table}
\caption{Kinematics-metallicity correlation of thick disk stars with $-1.0<$[Fe/H]$<-0.5$ for three height intervals.}
\label{table:2}
\begin{center}
\begin{tabular}{ccccc}
\hline\hline
$\langle \left|z\right| \rangle$ & N$_{\rm tot}$ & N$_{\rm used}$ & $\partial \langle V_\phi\rangle / \partial$[Fe/H] & $\rho_s$ \\
(kpc) & & & (km s$^{-1}$ dex$^{-1}$) & ($\times 10^{-2}$)\\
& & $3\sigma$ { } { } $2\sigma$ & $3\sigma$ { } { } { } $2\sigma$ & $3\sigma$ { } { } { } $2\sigma$\\
\hline
1.23 & 3994 & 3672 { } 2915 & 50$\pm$5 { } 39$\pm$5 & 17$\pm$2 { } 15$\pm$2 \\
1.73 & 2641 & 2348 { } 1715 & 54$\pm$6 { } 35$\pm$5 & 18$\pm$2 { } 16$\pm$2 \\
2.37 & 2194 & 1768 { } 1131 & 35$\pm$8 { } 33$\pm$5 & 10$\pm$2 { } 14$\pm$3\\
\hline
\end{tabular}
\end{center}
\end{table}
\begin{figure}[]
\resizebox{\hsize}{!}{\includegraphics[trim=0mm 3mm 0mm 2mm]{13538f5a.ps}}
\resizebox{\hsize}{!}{\includegraphics[trim=0mm 3mm 0mm 2mm]{13538f5b.ps}}
\resizebox{\hsize}{!}{\includegraphics[trim=0mm 3mm 0mm 2mm]{13538f5c.ps}}
\caption{$V_\phi$ distribution for [Fe/H]$<-0.5$, $-0.7$, and $-1.0$, in descending order. The solid lines show the best fit of two Gaussian-component models (thick disk and halo). Bottom, middle, and top panels refer to $\left| z\right|=$1.0--1.5 kpc, 1.5--2.0 kpc, and 2.0--3.0 kpc, respectively. The solid and dashed lines mark the values $V_\phi=0$ km~s$^{-1}$ and 170 km~s$^{-1}$. }
\label{spagna_fig5}
\end{figure}
\subsection{Rotation -- metallicity correlation}
\label{sect:correlation}
The {\it disk} and {\it halo} populations are apparent in the $V_\phi$ vs.\ [Fe/H] distribution (see Fig.\ \ref{spagna_fig2}).
In particular, the region $-1.0<$[Fe/H]$<-0.5$ and 0 km~s$^{-1}<V_\phi<300$ km~s$^{-1}$ does contain the bulk of the regular thick disk stars, besides a small number of stars belonging to the metal-poor tail of the thin disk and to the high-metallicity tail of the inner halo. Actually, a significant fraction of thin disk stars are expected for [Fe/H]$>-0.5$, while towards lower abundances, [Fe/H]$<-1$, the thick disk metal weak tail and the newly discovered flattened inner halo \citep{morrison2009} are also present.
Figure \ref{spagna_fig3} shows the iso-density contours of the velocity-metallicity distribution of stars with $\left|z\right|=1.0$-3.0 kpc and $-1.0<$[Fe/H]$<-0.3$.
As in \citet{ivezic2008} and \citet{bond2009}, no correlation appears in the transition region between the thin and thick disks ( [Fe/H]$\ga -0.5$).
Instead, we notice a shallow but clear slope for [Fe/H]$\la -0.5$, undetected by previous studies, which indicates that the metal-rich stars tend to rotate faster than the metal-poor ones.
In particular, the top-density ridge increases from $V_\phi\simeq 150$ km~s$^{-1}$ at [Fe/H]$\approx -1$ to
$V_\phi\simeq 170$ km~s$^{-1}$ at [Fe/H]$\approx -0.4$.
Inspection of Fig. \ref{spagna_fig3} also proves a bimodal distribution with a secondary maximum located at [Fe/H]$\approx -0.55$, close to the value of the mean metallicity of the thick disk, and the peak at [Fe/H]$\simeq -0.38$ due to thin disk stars.
\begin{table*}
\caption{Fitted parameters, as in Table \ref{table:1}, for different metallicity intervals, $-3.0<$ [Fe/H] $\le$ [Fe/H]$_\mathrm{max}$, where $-1.0\le$ [Fe/H]$_\mathrm{max} \le -0.5$. }
\label{table:3}
\begin{center}
\begin{tabular}{crllllcc}
\hline\hline
& & \multicolumn{2}{c}{\sc Thick Disk} & \multicolumn{2}{c}{\sc Halo} & & \\
$\left[\mathrm{Fe}/{\mathrm H}\right]_\mathrm{max}$ & N & $\langle V_\phi\rangle$ & $\sigma_{V\phi}$ & $\langle V_\phi\rangle$ & $\sigma_{V\phi}$ & $\frac{\rho_{\rm TD}}{\rho_{\rm tot}}$ & $\chi^2_\nu$\\
(dex) & & \multicolumn{2}{c}{(km s$^{-1}$)} & \multicolumn{2}{c}{(km s$^{-1}$)} & (\%) & \\
\hline
\multicolumn{8}{c}{$1.0$ kpc $<\left|z\right| \le 1.5$ kpc}\\
\hline
$-0.5$ & 6537 & 173 $\pm$ 1& 39 $\pm$ 1 & 33 $\pm$ 4 & 90 $\pm$ 3 & 68 $\pm$ 2 & 3.19\\
$-0.6$ & 5470 & 170 $\pm$ 1& 39 $\pm$ 1 & 35 $\pm$ 7 & 94 $\pm$ 4 & 61 $\pm$ 3 & 2.13\\
$-0.7$ & 4511 & 167 $\pm$ 1& 39 $\pm$ 2 & 33 $\pm$ 7 & 94 $\pm$ 4 & 55 $\pm$ 3 & 1.80\\
$-0.8$ & 3675 & 165 $\pm$ 1& 38 $\pm$ 3 & 31 $\pm$ 6 & 93 $\pm$ 3 & 46 $\pm$ 3 & 1.57\\
$-0.9$ & 3036 & 162 $\pm$ 2& 38 $\pm$ 2 & 29 $\pm$ 6 & 93 $\pm$ 3 & 38 $\pm$ 4 & 1.51\\
$-1.0$ & 2543 & 162 $\pm$ 2& 37 $\pm$ 2 & 24 $\pm$ 5 & 92 $\pm$ 3 & 30 $\pm$ 4 & 1.59\\
\hline
\multicolumn{8}{c}{$1.5$ kpc $<\left|z\right| \le 2.0$ kpc}\\
\hline
$-0.5$ & 4753 & 163 $\pm$ 1& 44 $\pm$ 2 & 29 $\pm$ 10 & 96 $\pm$ 5 & 60 $\pm$ 4 & 1.42\\
$-0.6$ & 4113 & 157 $\pm$ 2& 45 $\pm$ 2 & 23 $\pm$ 10 & 94 $\pm$ 5 & 56 $\pm$ 4 & 1.49\\
$-0.7$ & 3480 & 154 $\pm$ 2& 45 $\pm$ 2 & 23 $\pm$ 11 & 95 $\pm$ 5 & 49 $\pm$ 5 & 1.28\\
$-0.8$ & 2936 & 152 $\pm$ 2& 43 $\pm$ 3 & 22 $\pm$ 9 & 95 $\pm$ 4 & 41 $\pm$ 5 & 1.30\\
$-0.9$ & 2488 & 149 $\pm$ 3& 42 $\pm$ 3 & 19 $\pm$ 8 & 94 $\pm$ 4 & 33 $\pm$ 5 & 1.20\\
$-1.0$ & 2112 & 145 $\pm$ 4& 44 $\pm$ 5 & 18 $\pm$ 9 & 94 $\pm$ 4 & 25 $\pm$ 6 & 1.18\\
\hline
\multicolumn{8}{c}{$2.0$ kpc $<\left|z\right| \le 3.0$ kpc}\\
\hline
$-0.5$ & 4680 & 152 $\pm$ 2& 45 $\pm$ 2 & 43 $\pm$ 6 & 94 $\pm$ 2 & 40 $\pm$ 4 & 1.72\\
$-0.6$ & 4176 & 150 $\pm$ 3& 46 $\pm$ 3 & 42 $\pm$ 6 & 94 $\pm$ 2 & 34 $\pm$ 4 & 1.56\\
$-0.7$ & 3671 & 147 $\pm$ 3& 45 $\pm$ 3 & 37 $\pm$ 6 & 93 $\pm$ 2 & 30 $\pm$ 4 & 1.41\\
$-0.8$ & 3183 & 144 $\pm$ 4& 43 $\pm$ 4 & 33 $\pm$ 5 & 92 $\pm$ 2 & 25 $\pm$ 4 & 1.31\\
$-0.9$ & 2830 & 143 $\pm$ 4& 40 $\pm$ 5 & 32 $\pm$ 5 & 92 $\pm$ 2 & 19 $\pm$ 4 & 1.20\\
$-1.0$ & 2486 & 141 $\pm$ 5& 38 $\pm$ 6 & 30 $\pm$ 4 & 93 $\pm$ 2 & 14 $\pm$ 4 & 1.20\\
\hline
\end{tabular}
\end{center}
\end{table*}
To quantify the correlation, we first select the stars within $\Delta$[Fe/H]=0.05 bins
in the range $-1.0<$[Fe/H]$< -0.5$ and located at the different height intervals: $\Delta\left| z\right|= 1.0$--1.5 kpc, 1.5--2.0 kpc, and 2.0--3.0 kpc.
Then, the stars with velocities $(V_R, V_\phi, V_z)$ outside 3$\sigma$ from the thick disk velocity ellipsoid, corresponding to a confidence level of 97.1\%, were rejected to minimize the contamination from the halo stars.
We adopted $\langle V_\phi\rangle$ as a function of $z$ derived from Table \ref{table:1} and assumed constant dispersions: $\sigma_{V\phi}\equiv\sigma_{Vz}= 40$ km~s$^{-1}$ and $\sigma_{V_R}\equiv 1.5\, \sigma_{V\phi}= 60$ km~s$^{-1}$.
Finally, mean velocities were computed for the {\it bona fide} thick disk stars and the slope, $\partial \langle V_\phi\rangle / \partial$[Fe/H], is estimated by means of a linear fit for the height intervals $\Delta\left| z\right| =$1.0--1.5 kpc, 1.5--2.0 kpc, and 2.0--3.0 kpc. For each bin, mean height, total number of stars, number of stars used (after 3$\sigma$ and 2$\sigma$ rejection), slope, and Spearman's rank correlation coefficient are listed in Table \ref{table:2}, while the observed distributions are shown in Fig. \ref{spagna_fig4}.
Overall, a kinematic-metallicity correlation of about 50 km~s$^{-1}$~dex$^{-1}$ is detected up to $ \left| z\right|\simeq 2$ kpc, while a shallower slope ($\sim35$ km~s$^{-1}$~dex$^{-1}$) is present between $2 < \left| z\right|\le 3$ kpc.
It is possible that these values are affected by a residual contamination of halo stars, whose presence can be inferred by the number of rejected high velocity stars shown in Table \ref{table:2} being greater than the 3\% expected in the case of a pure Gaussian distribution. Nevertheless, even if we apply a conservative 2$\sigma$ selection (73.8\% confidence level), we still find a correlation at the level of 30$\div$40 km~s$^{-1}$~dex$^{-1}$, as reported in the last column of Table \ref{table:2}.
This conclusion is consistent with the systematic slowing down of the thick disk rotation, which results from fitting a {\it two} Gaussian-component model, representing the thick disk and halo populations, as more metal-poor thresholds are applied: [Fe/H]$_{\mathrm{max}}<-0.5, <-0.6,$ ... $<-1.0$ (see Table \ref{table:3}). This effect is depicted in Fig. \ref{spagna_fig5}, which shows how the thick disk component both decreases {\it and} shifts towards lower $V_\phi$ values, when different subsamples of metal poor stars are selected.
In addition, we estimate the rotation-metallicity correlation by fitting the thick disk $\langle V_\phi\rangle$ values from Table \ref{table:3} through the following integral linear model:
\begin{equation}
\langle V_\phi \rangle = V_\phi(\mathrm{[Fe/H]}_0) + a\cdot \left( \langle \mathrm{[Fe/H]}\rangle - \mathrm{[Fe/H]}_0 \right),
\label{eq:2}
\end{equation}
where $\langle \mathrm{[Fe/H]}\rangle$ is the average for the stars with $-3<$ [Fe/H]$\le $[Fe/H]$_{\rm max}$, $a=\partial \langle V_\phi\rangle/\partial\mathrm{[Fe/H]}$, and $V_\phi$([Fe/H]$_0$) is the mean velocity of the reference metallicity, which we set to [Fe/H]$_0=-0.6$.
In Figure \ref{spagna_fig6}, the lines connect the values from Eq. \ref{eq:2} at the different [Fe/H]$_{\mathrm{max}}$ thresholds. These results confirm both a vertical gradient consistent with the value derived in Sect.\ \ref{sect:verticalGradient} and a rotation-metallicity correlation in the range of 40$\div$50 km~s$^{-1}$~dex$^{-1}$ for the thick disk.
We also considered the hypothesis that a false trend $V_\phi$ vs.\ [Fe/H] might derive from the tangential velocity estimated through the metallicity-dependent photometric parallaxes. Actually, the correlation would still be significant even if the $M_r$-calibration were subjected to a systematic error up to 0.4 mag per dex.
Moreover, no kinematics-metallicity correlation is expected to arise because of the
color-selection criteria of the SDSS spectroscopic targets, which although they produce a bias towards metal poor stars, cannot affect the conditional $V_\phi$ probability distribution at a given metallicity, Pr($V_\phi \left|\right.$[Fe/H]), and no further kinematical selection is applied.\\
Thus, we conclude that the observed correlation is an intrinsic signature of our sample.
\begin{figure}[]
\resizebox{\hsize}{!}{\includegraphics[trim=0mm 3mm 0mm 3mm]{13538f6.ps}}
\caption{
\footnotesize
Mean $V_\phi$ vs. [Fe/H]$_{\rm max}$ for different abundance ranges,
$-3< $[Fe/H]$\le $[Fe/H]$_{\rm max}$, with $-1.0\le $[Fe/H]$_{\rm max} \le -0.5$. The three lines show the $\langle V_\phi\rangle$ computed from Eq. \ref{eq:2} with fitted slopes, $\partial \langle V_\phi\rangle / \partial$[Fe/H]=
$( 39 \pm 5, 56 \pm 9, 38\pm 13)$ km s$^{-1}$ dex$^{-1}$, and zero-points,
$V_\phi(\mathrm{[Fe/H]_0})=(178\pm 1, 170\pm 2, 158\pm 3)$ km s$^{-1}$ for $\left|z\right|=$(1.0-1.5, 1.5-2.0, 2.0-3.0) kpc, respectively.}
\label{spagna_fig6}
\end{figure}
\section{Discussion and conclusions}
The existence of a vertical velocity gradient and a rotation-metallicity correlation sets important constraints on the origin of the thick disk.
The estimated gradient of $-19\pm 2$ km~s$^{-1}$ kpc$^{-1}$ is consistent with Nbody simulations of disks thickened by a single minor merger with a low/intermediate orbital inclination \citep[e.g.][]{villalobos2008},
as well as by the interaction with numerous dark subhalos, as discussed by \citet{hayashi2006} and \citet{kazantzidis2008},
whose simulations show kinematic gradients of $-(10\div 30)$ km~s$^{-1}$~kpc$^{-1}$ and of $-20$ km~s$^{-1}$~kpc$^{-1}$, respectively, for 1 kpc $<\left|z\right|\le 3$ kpc.
A vertical rotation gradient of about $-20$ km~s$^{-1}$~kpc$^{-1}$ can also be inferred from Fig.\ 5 of \citet{abadi2003},
who investigated thick disks formed by accretion of both the stars of a pre-existing thin disk and the debris from disrupted satellites.
Unfortunately, we have not found any explicit kinematic prediction in the scenario of the chaotic gas-rich mergers described by \citet{brook2005},
although \citet{hayashi2006} state that a velocity shear {``may have difficulties in this regard''}.
Finally, to the best of our knowledge, explicit predictions of kinematics-metallicity correlations
are missing in the current CDM scenarios of satellite accretion or minor mergers. Hopefully, our results will motivate theoreticians to investigate this issue in their future models.
In the context of models based on disk secular processes of stellar migration driven by interactions with spiral arms, a vertical gradient of $\sim -15$ km~s$^{-1}$~kpc$^{-1}$ is reported by \citet{loebman2008}, who, conversely,
did not detected any $V_\phi$ vs.\ [Fe/H] correlation. The simulations carried out by \citet{schonrich2009} indicate a mild trend ($\sim10$ km~s$^{-1}$~dex$^{-1}$) at $z\approx 0$ kpc, which decreases with height and disappears for $\left| z\right|\ga 1$ kpc. Possibly, by adopting appropriate parameters, their inside-out disk formation model could reproduce the observed downtrend (Sch\"{o}nrich, 2009, private communication).
Thus, more attention should be devoted to this scenario as a possible theoretical framework to explain the rotation--metallicity relation in the thick disk of the Milky Way.
\begin{acknowledgements}
We are grateful to the anonymous referee for all the valuable comments.
A.S. thanks Beatrice Bucciarelli and Ralph Sch\"{o}nrich for helpful discussions.
We acknowledge B. McLean and the GSC-II team for supporting the data mining of the GSC-II database.
The authors acknowledge the financial support of INAF through the PRIN 2007 grant n. CRA 1.06.10.04 ``The local route to galaxy
formation''.
Support through the Marie Curie Research Training Network ELSA under contract MRTN-CT-2006-033481 to P.R.F. is also thankfully acknowledged.
Funding for the SDSS and SDSS-II has been provided by the Alfred P. Sloan Foundation, the Participating Institutions, the National Science Foundation, the U.S. Department of Energy, the National Aeronautics and Space Administration, the Japanese Monbukagakusho, the Max Planck Society, and the Higher Education Funding Council for England. The Guide Star Catalogue~II is a joint project of the Space Telescope Science Institute and the Osservatorio Astronomico di
Torino.
\end{acknowledgements}
\bibliographystyle{aa}
|
\section{Introduction
It is one of the most important issues for cosmologists and
particle physicists to understand the physical origin of the dark
energy (DE) which is responsible for an accelerated expansion of
the current Universe. Although the standard spatially flat ${\rm
\Lambda}$-Cold-Dark-Matter (${\rm \Lambda CDM}$) model is consistent with all
kinds of current observational data \cite{O01_WMAP7},
some tentative deviations from it have been reported recently
\cite{O01_Shafieloo:2009ti,O01_Bean:2009wj}. Furthermore, in the
${\rm \Lambda CDM}$ model, the cosmological term is regarded as a new
fundamental constant whose observed value is much smaller than any
other energy scale known in physics. Hence it is natural to seek
for non-stationary models of the current DE. Among them, $f(R)$
gravity which modifies and generalizes the Einstein gravity by
incorporating a new phenomenological function of the Ricci scalar
$R$, $f(R)$, can provide a self-consistent and non-trivial
alternative to the ${\rm \Lambda CDM}$
model\cite{O01_Hu:2007nk,O01_Starobinsky:2007hu}.
In the previous paper \cite{O01_Motohashi:2009qn}, we calculated
evolution of matter density fluctuations in viable $f(R)$ models
\cite{O01_Hu:2007nk,O01_Starobinsky:2007hu} for redshifts $z \gg
1$ during the matter-dominated stage and found an analytic
expression for them. In this paper we extend the previous analysis
and perform numerical calculations of the evolution of both
background space-time and density fluctuations for the particular
$f(R)$ model of Ref.~\cite{O01_Starobinsky:2007hu} without such a
restriction. As a result, we have found crossing of the phantom
boundary $w_{\rm DE}=-1$ at an intermediate redshift $z\lesssim 1$ for
the background space-time metric and an anomalous behavior of the
growth index of fluctuations.
\section{Background
We adopt the following action
of $f(R)$ models with model parameters $n,~\lambda$ and $R_s$
\cite{O01_Starobinsky:2007hu}:
\begin{equation} \label{O01_fR} S=
\frac{1}{16\pi G} \int d^4x \sqrt{-g} f(R) + S_m, \quad f(R)=R +
\lambda R_s \left[ \left( 1 +
\frac{R^2}{R_s^2}\right)^{-n}-1\right] , \end{equation}
where $S_m$ is the
action of the matter content which is assumed to be minimally
coupled to gravity. To make the late-time asymptotic de Sitter
regime where $R=$ constant stable, $\lambda$ has to satisfy
$f'(R)>Rf''(R)$. As a result, $\lambda$ has a lower limit
$\lambda_{\min}$ for each $n$. Numerically we find
$(n,\lambda_{\min})=$(2, 0.9440), (3, 0.7259), and (4, 0.6081).
From the action \eqref{O01_fR}, we obtain field equations as
\begin{align}
R^{\mu}_{\nu}-\frac{1}{2}\delta^{\mu}_{\nu}R&=
-8\pi G\mk{T^{\mu}_{\nu (m)}+T^\mu_{\nu ({\rm DE})}}, \\
\label{O01_EMtensor}
8\pi G T^\mu_{\nu ({\rm DE})}&\equiv
(F-1)R^\mu_\nu-\frac{1}{2}(f-R)\delta^{\mu}_{\nu}
+(\nabla^\mu\nabla_\nu-\delta^{\mu}_{\nu}\square)F,\quad
F(R)\equiv f'(R).
\end{align}
Working in the spatially flat Friedmann-Robertson-Walker (FRW)
space-time with a scale factor $a(t)$,
\begin{align}
\label{O01_hubble} 3H^2&=8\pi G\rho-3(F-1)H^2+\frac{1}{2}(FR-f)-3H\dot F, \\
\label{O01_hdot} 2\dot{H}&=-8\pi G\rho -2(F-1)\dot{H}-\ddot{F}+H\dot{F},
\end{align}
where $H$ is the Hubble parameter and $\rho$ is the energy density
of the material content which we assume to consist of
non-relativistic matter. From \eqref{O01_fR}, we can determine the
DE equation of state parameter $w_{\rm DE}$,
\begin{equation} \label{O01_wDE}
w_{\rm DE}\equiv\f{P_{\rm DE}}{\rho_{\rm DE}}
=-1+\f{2\dot{H}(F-1)-H\dot{F}+\ddot{F}}{-3H\dot{R}F'
+3(H^2+\dot{H})(F-1)-(f-R)/2}. \end{equation}
We solve the evolution equation \eqref{O01_hdot} numerically using
\eqref{O01_hubble} to check numerical accuracy. The moment when
the matter density parameter $\Omega(t)=16\pi G\rho/(16\pi
G\rho+\lambda R_s)$ equals to $0.998$ is chosen as the initial
time $t_i$. We determine the current epoch $t=t_0$ by the
requirement that the value of $\Omega$ takes the observed central
value $\Omega_0=0.27$. $R_s$ is fixed in such a way as to
reproduce the current Hubble parameter $H_0=72$km/s/Mpc. We find
the ratio $R_s/H_0^2$ is well fit by a simple power-law
$R_s/H_0^2=c_n\lambda^{-p_n}$ with $(n,c_n,p_n)=$(2, 4.16, 0.953),
(3, 4.12, 0.837), and (4, 4.74, 0.702), respectively, whereas in
the ${\rm \Lambda CDM}$ limit it would behave as
$R_s/H_0^2=6(1-\Omega_0)/\lambda\simeq 4.38\lambda^{-1}$.
\begin{figure}[t]
\centering
{
\includegraphics[width=80mm]{O01_w.eps}
\label{O01_fig:w-a}}
{
\includegraphics[width=80mm]{O01_w_min.eps}
\label{O01_fig:w-b}}
\caption{Evolution of the equation-of-state parameter of
effective dark energy.}
\label{O01_fig:w}
\end{figure}
Figures \ref{O01_fig:w} depict evolution of $w_{\rm DE}$ as a
function of redshift $z$ where phantom crossing is manifest. As
expected, it approaches $w_{\rm DE}=-1=\text{constant}$ as we
increase $\lambda$ for fixed $n$. For the minimal allowed values
of $\lambda$, deviations from $w_{\rm DE}=-1$ are observed at 5\%
level on both directions in $z \lesssim 2$ independently of $n$.
From \eqref{O01_wDE}, we can read off that this phantom crossing
behavior is not peculiar to the specific choice of the function
\eqref{O01_fR} but a generic one for models which satisfy the
stability condition $F'>0$.
\section{Perturbations
We now turn to the evolution of density fluctuations. In $f(R)$
gravity, the evolution equation of density fluctuations, $\delta$,
deeply in the sub-horizon regime is given
by\cite{O01_Tsujikawa:2007gd}
\begin{equation} \label{O01_de1} \ddot \delta +
2H\dot \delta - 4\pi G_{\text{eff}}\rho \delta = 0, \quad
G_{\text{eff}}=\frac{G}{F} \frac{1+4\frac{k^2}{a^2}\frac{F'}{F}}
{1+3\frac{k^2}{a^2}\frac{F'}{F}}. \end{equation}
In the previous
paper\cite{O01_Motohashi:2009qn} we obtained an analytic solution
for the high-curvature regime when the scale factor evolves as
$a(t)\propto t^{2/3}$ and $F$ takes an asymptotic form $F\simeq
1-2n\lambda \mk{R/R_s}^{-2n-1}$. In the present paper, we
numerically integrate \eqref{O01_de1} up to $z=0$ without using
the approximation $|F-1|\ll 1$.
The wavenumber of our particular interest is the scale
corresponding to $\sigma_8$ normalization, for which we find
$k_{\rm eff}(r=8h^{-1}{\rm Mpc})=0.174h{\rm Mpc}^{-1}$. Since the
standard ${\rm \Lambda CDM}$ model normalized by CMB data explains galaxy
clustering at small scales well, $\delta_{{\rm fRG}}$ should not be too
much larger than $\delta_{\rm \Lambda CDM}$ on these scales. We may typically
require $(\delta_{{\rm fRG}}/\delta_{{\rm \Lambda CDM}})^2\lesssim 1.1$. Although
we neglect non-linear effects here, the difference between linear
calculation and non-linear N-body simulation remained smaller than
5\% at wavenumber $0.174h{\rm Mpc}^{-1}$\cite{O01_Oyaizu:2008tb}.
\begin{figure}[tbp]
\centering
{
\includegraphics[width=80mm]{O01_rt.eps}
\label{O01_fig:rtnl-a}}
{
\includegraphics[width=80mm]{O01_n_lam.eps}
\label{O01_fig:rtnl-b}} \caption{Constraints from the power
spectrum.} \label{O01_fig:rtnl}
\end{figure}
The left panel of Fig.~\ref{O01_fig:rtnl} represents
$(\delta_{{\rm fRG}}/\delta_{{\rm \Lambda CDM}})^2$ as a function of $\lambda$ for
$n=2$ together with two fitting functions. The solid line is from
the analytic formula obtained in Ref.~\cite{O01_Motohashi:2009qn},
and the broken line is numerical fitting using an exponential
function $1+b_ne^{-q_n\lambda}$. From these analysis, we can
constrain the parameter space as the right panel of Fig.~\ref{O01_fig:rtnl}.
The region which satisfy $(\delta_{{\rm fRG}}/\delta_{{\rm \Lambda CDM}})^2 < 1.1$ lies
above the solid line. The region below the dotted line is
forbidden from instability of de Sitter regime.
Next we turn to another important quantity used to distinguish
different theories of gravity, namely, the gravitational growth
index, $\gamma(z)$, of density fluctuations. It is defined through
\begin{equation} \f{d\ln\delta}{d\ln a}=\Omega_m(z)^{\gamma(z)},~~~\text{or}~~~
\gamma(z)=\f{\log\mk{\f{\dot{\delta}}{H\delta}}}{\log\Omega_m}. \end{equation}
It takes a practically constant value $\gamma\cong 0.55$ in the
standard ${\rm \Lambda CDM}$ model,
but it evolves with time in modified
gravity theories in general. We also note that $\gamma(z)$ has a
nontrivial $k$-dependence in $f(R)$ gravity since density
fluctuations with different wavenumbers evolve differently.
Therefore, this quantity is a useful measure to distinguish
modified gravity from ${\rm \Lambda CDM}$ model in the Einstein gravity.
\begin{figure}[t]
\centering
{
\includegraphics[width=80mm]{O01_g.eps}
\label{O01_fig:gG-a}}
{
\includegraphics[width=80mm]{O01_Geff.eps}
\label{O01_fig:gG-b}}
\caption{Evolution of $\gamma(z)$ and $G_{\rm eff}/G$.}
\label{O01_fig:gG}
\end{figure}
Figures \ref{O01_fig:gG} show the evolution of $\gamma(z)$ together
with that of $G_{\rm eff}/G$ for different values of $k$. In the early
high-redshift regime, $\gamma(z)$ takes a constant value identical
to the ${\rm \Lambda CDM}$ model. It gradually decreases in time, reaches a
minimum which may be even negative, and then increase again
towards the present epoch. We note that recently Narikawa and
Yamamoto\cite{O01_Narikawa:2009ux} numerically calculated time
evolution of $\gamma(z)$ in a simplified model which we had used
in the previous paper and also obtained some analytic expansion,
which behaves qualitatively the same as the numerical result but
with much more exaggerated amplitudes. Our results, which satisfy
all viability conditions, exhibit milder deviation from ${\rm \Lambda CDM}$
model than they found. At present, the constraints for the growth
index is not so strict to distinguish the deviation from the
${\rm \Lambda CDM}$ model\cite{O01_Rapetti:2009ri}, but observations may
reveal its time and wave number dependence in future.
\section{Conclusion
In the present paper we have numerically calculated the evolution
of both homogeneous background and density fluctuations in a
viable $f(R)$ model of accelerated expansion based on the specific
functional form proposed in Ref.\ \cite{O01_Starobinsky:2007hu}.
We have found that viable $f(R)$ gravity models of accelerated
expansion generically exhibit phantom behavior during the
matter-dominated stage with crossing of the phantom boundary
$w_{\rm DE}=-1$ at redshifts $z\lesssim 1$. The predicted time
evolution of $w_{\rm DE}$ has qualitatively the same behaviour as
that was recently obtained from observational data in
\cite{O01_Shafieloo:2009ti}.
As for density fluctuations, we have numerically confirmed our
previous analytic results on a shift in the power spectrum index
for large wavenumbers which exceed the scalaron mass during the
matter dominated epoch\cite{O01_Motohashi:2009qn}, while for
smaller wavenumbers fluctuations have the same amplitude as in the
${\rm \Lambda CDM}$ model.
We have also investigated the growth index $\gamma(k,z)$ of
density fluctuations and have given an explanation of its
anomalous evolution in terms of time dependence of $G_{\rm eff}$. Since
$\gamma$ has characteristic time and wavenumber dependence, future
detailed observations may yield useful information on the validity
of $f(R)$ gravity through this quantity, although current
constraints have been obtained assuming that it is constant both
in time and in
wavenumber\cite{O01_Bean:2009wj,O01_Rapetti:2009ri}.
\acknowledgments
HM and JY are grateful to T.\ Narikawa and K.\ Yamamoto for useful
communications.
AS acknowledges RESCEU hospitality as a visiting professor. He was also
partially supported by the grant RFBR 08-02-00923 and by the Scientific
Programme ``Astronomy'' of the Russian Academy of Sciences.
This work was supported in part by
JSPS Grant-in-Aid for Scientific Research No.\ 19340054(JY),
JSPS Core-to-Core program ``International Research Network
on Dark Energy'', and
Global COE Program ``the Physical Sciences Frontier'', MEXT, Japan.
}
|
\section{Introduction}
\label{sect:intro}
The sensor network localization problem, \mbox{\boldmath$SNL$,}\,%
\index{\mbox{\boldmath$SNL$}\,, sensor network localization}%
\index{sensor network localization, \mbox{\boldmath$SNL$}\,}
consists in locating the positions
of $n$ wireless sensors, $p_i \in \mathbb{R}^r$, $i=1,\ldots,n$, given only the
(squared) Euclidean distances $D_{ij}=\|p_i-p_j\|_2^2$
between sensors that are within a given
radio range, $R>0$, and given the positions of a
\index{radio range, $R$}
subset of the sensors, $p_i$, $i=n-m+1,\ldots,n$
(called anchors); $r$ is the {\em embedding dimension} of the problem.
\index{embedding dimension ({\em fixed}), $r$}
Currently, many solution techniques for this problem use a relaxation to
a nearest, weighted, semidefinite approximation problem
\begin{equation}
\label{eq:sdprelax}
\min_{Y \succeq 0, \, Y \in \Omega} \left\| W\circ\left(\KK(Y) - D\right) \right\|,
\end{equation}
where $Y\succeq 0$ denotes positive semidefiniteness,
$Y \in \Omega$ denotes additional linear constraints, $\KK$ is a
specific linear mapping, and $\circ$ denotes the
{\em Hadamard (elementwise) product}.
\index{Hadamard product}
This approach requires
semidefinite programming, {\mbox{\boldmath$SDP$,}\,} primal-dual interior point (p-d i-p)
techniques; see, for example,
\cite{AlfakihAnjosKPW:08,AlKaWo:97,biswasphd07,BiswasYe:04,MR2191577,dattorro:05,pongtseng:09}.
This yields an expensive and inexact solution.
The \mbox{\boldmath$SNL$}\, problem is a special case of the Euclidean Distance Matrix,
\index{\mbox{\boldmath$EDM$}\,} \mbox{\boldmath$EDM$,}\, completion problem, \mbox{\boldmath$EDMC$}.\,\,
If $D$ is a {\em partial} \mbox{\boldmath$EDM$,}\, then the completion problem consists in finding
the missing elements (squared distances) of $D$.
It is shown in \cite{DiKrQiWo:06},
that there are advantages for handling the \mbox{\boldmath$SNL$}\, problem as an
\mbox{\boldmath$EDMC$},\, and ignoring the distinction between the
anchors and the other sensors until after the \mbox{\boldmath$EDMC$}\, is solved.
In this paper we use this framework and derive an algorithm that
locates the sensors by exploiting the structure and
implicit degeneracy in the \mbox{\boldmath$SNL$}\, problem. In particular, we solve the
\mbox{\boldmath$SDP$}\, problems {\em explicitly} (exactly)
without using any p-d i-p techniques.
We do so by repeatedly viewing \mbox{\boldmath$SNL$}\, in three equivalent forms:
as a graph realization problem, as a \mbox{\boldmath$EDMC$}\,, and as a rank restricted
\mbox{\boldmath$SDP$}\,.
A common approach to solving the \mbox{\boldmath$EDMC$}\, problem is to
relax the rank constraint and solve
a weighted, nearest, positive semidefinite
completion problem (like problem~\eqref{eq:sdprelax}) using semidefinite programming, \mbox{\boldmath$SDP$}\,.
The resulting \mbox{\boldmath$SDP$}\,
is, implicitly, highly degenerate in the
sense that the feasible semidefinite matrices have low rank. In particular,
\label{rankpage}
cliques in the graph of the \mbox{\boldmath$SNL$}\, problem reduce the ranks of these feasible
semidefinite matrices.
This means that the Slater constraint qualification (strict feasibility)
implicitly fails for the \mbox{\boldmath$SDP$}\,. Our algorithm is based on
exploiting this degeneracy.
We characterize the face of the \mbox{\boldmath$SDP$}\, cone that corresponds to a given
clique in the graph, thus reducing the size of the \mbox{\boldmath$SDP$}\, problem.
Then, we characterize the intersection
of two faces that correspond to overlapping cliques.
This allows us to explicitly {\em grow/increase} the size of the cliques by
repeatedly finding the intersection of subspaces that represent
the faces of the \mbox{\boldmath$SDP$}\, cone that correspond to these cliques.
Equivalently, this corresponds to completing overlapping blocks of the \mbox{\boldmath$EDM$}\,.
In this way, we further reduce the dimension of the faces until
we get a completion of the entire \mbox{\boldmath$EDM$}\,. The intersection of the
subspaces can be found using a singular value decomposition (SVD) or by
exploiting the special structure of the subspaces.
No \mbox{\boldmath$SDP$}\, solver is used.
\index{embedding dimension ({\em fixed}), $r$}
Thus we solve the \mbox{\boldmath$SDP$}\, problem in a finite number of steps, where the work
of each step is to find the intersection of two subspaces (or, equivalently,
each step is to find the intersection of two faces of the \mbox{\boldmath$SDP$}\, cone).
Though our results hold for general embedding dimension $r$,
\index{embedding dimension ({\em fixed}), $r$}
our preliminary numerical tests involve sensors with
embedding dimension $r=2$ and $r=3$. The sensors are in the region $[0,1]^r$.
There are $n$ sensors, $m$ of which are anchors. The radio range is
$R$ units.
\index{radio range, $R$}
\subsection{Related Work/Applications}
The number of applications for distance geometry problems is large and
increasing in number and importance. The particular case of \mbox{\boldmath$SNL$}\, has
applications to environmental monitoring of geographical regions, as
well as tracking of animals and machinery; see, for example,
\cite{biswasphd07,dattorro:05}.
There have been many algorithms published recently that solve the \mbox{\boldmath$SNL$}\,
problem. Many of these involve \mbox{\boldmath$SDP$}\, relaxations and use \mbox{\boldmath$SDP$}\, solvers;
see, for example,
\cite{biswasphd07,biswasliangtohwangye,biswasliangtohye:05,MR2398864,BiswasYe:04,
MR2191577,DiKrQiWo:06}
and more recently \cite{KimKojimaWaki:09,WangZhengBoydYe:06}.
Heuristics are presented in, for example, \cite{cassioli:08}.
\mbox{\boldmath$SNL$}\, is closely related to the \mbox{\boldmath$EDMC$}\, problem; see, for example,
\cite{AlKaWo:97,dattorro:05} and the survey \cite{AlfakihAnjosKPW:08}.
Jin et al \cite{MR2274505,Jin:05} propose the {SpaseLoc} heuristic.
It is limited to $r=2$ and uses an \mbox{\boldmath$SDP$}\, solver for small localized
subproblems. They then {\em sew} these subproblems together.
So \& Ye \cite{MR2295148} show that the problem of solving a noiseless
\mbox{\boldmath$SNL$}\, with a unique solution can be phrased as an \mbox{\boldmath$SDP$}\, and thus can be solved
in polynomial time. They also give an efficient criterion for checking
whether a given instance has a unique solution for $r=2$.
Two contributions of this paper are: we do not use
iterative p-d i-p techniques to solve the \mbox{\boldmath$SDP$}\,, but rather, we solve it
with a finite number of explicit solutions;
we start with local cliques and expand the cliques.
Our algorithm has four different basic steps. The first basic step
takes two cliques for which the intersection contains at least $r+1$ nodes
and implicitly completes the
corresponding \mbox{\boldmath$EDM$}\, to form the union of the cliques.
The second step does this when one of the cliques is a single element.
Therefore, this provides an extension of the algorithm in
\cite{egwymab04}, where Eren et al
have shown that the family of {\em trilateration graphs}
\index{trilateration graph} admit a polynomial time algorithm for computing
a realization in a required dimension.\footnote{A graph is a trilateration graph
in dimension $r$ if there exists an
ordering of the nodes $1, \ldots, r + 1, r + 2, \ldots, n$
such that: the first $r + 1$ nodes form a clique,
and each node $j > r + 1$ has at least $r + 1$ edges to nodes earlier
in the sequence.} Our first basic step also provides an explicit form for finding a
realization of a \index{uniquely localizable graph}
{\em uniquely localizable graph}\footnote{A graph is uniquely localizable in
dimension $r$ if it has a unique realization in $\mathbb{R}^r$ and it does not
have any realization whose affine span is $\mathbb{R}^h$, where $h>r$; see
\cite{MR2295148}.}.
Our algorithm repeatedly finds explicit solutions of an \mbox{\boldmath$SDP$}\,.
Other examples of finding explicit solutions of an \mbox{\boldmath$SDP$}\, are given in
\cite{MR97b:90077,Wolk:93}.
The \mbox{\boldmath$SNL$}\, problem with given embedding dimension $r$ is NP-hard
\index{embedding dimension, $r$}
\cite{hendrickson:91,MR92m:05182,sax79}.
However, from our numerical tests it appears that random problems that
have a unique solution can be solved very efficiently.
This phenomenon fits into the results in
\cite{AmVav:09,FeiKrau:00}.
\subsection{Outline}
We continue in Section \ref{sect:prels} to present notation and
results that will be used. The facial reduction process is based on the results in Section~\ref{sect:cliqred}.
The single clique facial reduction is given in Theorem~\ref{thm:onecliquered};
the reduction of two overlapping cliques in the
rigid and nonrigid cases is
presented in Theorem~\ref{thm:twocliquered} and
Theorem~\ref{thm:degcompl}, respectively;
absorbing nodes into cliques in the rigid and nonrigid
cases is given in Corollaries~\ref{cor:absorbsingle} and
\ref{cor:degabsorbsingle}, respectively.
These results are then used in our algorithm in
Section~\ref{sect:algor}. The numerical tests appear in
Section~\ref{sect:numerics} and Section~\ref{sect:noisy}.
Our concluding remarks are given in Section~\ref{sect:concl}.
\subsection{Preliminaries}
\label{sect:prels}
We work in the vector space of
{\em real symmetric $k \times k$ matrices}, ${\mathcal S} ^k$,
\index{symmetric $k \times k$ matrices, ${\mathcal S^{k}}\,$}
equipped with the
{\em trace inner product}, $\langle A,B\rangle = \trace AB$.
\index{trace inner product, $\langle A,B\rangle = \trace AB$}
We let ${\mathcal S^{k}_+}\,$ and ${\mathcal S^{k}_{++}}\,$
denote the cone of positive semidefinite and positive definite
\index{cone of positive semidefinite matrices, ${\mathcal S^{k}_+}\,$}
matrices, respectively;
\index{cone of positive definite matrices, ${\mathcal S^{k}_{++}}\,$}
$A \succeq B$ and $A \succ B$ denote the
L\"owner partial order, $A-B \in {\mathcal S^{k}_+}\,$ and
\index{L\"owner partial order, $A \succeq B$}
$A-B \in {\mathcal S^{k}_{++}}\,$, respectively;
$e$ denote the vector of ones of appropriate dimension;
\index{vector of ones, $e$}
${\mathcal R} ({\mathcal L} )$ and ${\mathcal N} ({\mathcal L} )$ denote the range space and null space of the linear transformation ${\mathcal L} $, respectively;
$\cone(S)$ denote the convex cone generated by the set $S$.
We use the {\sc Matlab} notation $1\!:\!n = \{1,\ldots,n\}$.
\index{$1:n = \{1,\ldots,n\}$}
\index{range space of ${\mathcal L} $, ${\mathcal R} ({\mathcal L} )$}
\index{null space of ${\mathcal L} $, ${\mathcal N} ({\mathcal L} )$}
\index{cone generated by $C$, $\cone (C)$}
A subset $F\subseteq K$ is a {\em face of the cone $K$}, denoted
$F \unlhd K$, if \index{face, $F \unlhd K$}
\[
\left( x, y \in K, \ \frac 12(x+y) \in F\right)
\implies \left(\cone \{x,y\} \subseteq F\right).
\]
If $F \unlhd K$, but is not equal to $K$, we write $F \lhd K$.
If $\{0\} \neq F \lhd K$, then $F$ is a {\em proper face} of $K$.
\index{proper face}
For $S \subseteq K$, we let $\face\!(S)$ denote the smallest face of $K$
that contains $S$.
A face $F \unlhd K$ is an {\em exposed face} if it is the
intersection of $K$ with a hyperplane.
\index{exposed face}
The cone $K$ is {\em facially exposed} if every face $F\unlhd K$ is exposed.
\index{facially exposed cone}
The cone ${\mathcal S^n_+\,}$ is facially exposed. Moreover, each face $F\unlhd {\mathcal S^n_+\,}$ is
determined by the range of any matrix $S$ in the
relative interior of the face, $S \in
\index{relative interior, $\relint \cdot$}
\relint F$: if $S=U\Gamma U^T$ is the compact spectral decomposition of $S$
with the diagonal matrix of eigenvalues $\Gamma \in {\mathcal S} _{++}^t$, then
\label{pagecite}(e.g., \cite{PatakiSVW:99})
\begin{equation}
\label{eq:facerepr}
F=U {\mathcal S} _+^t U^T.
\end{equation}
A matrix $D=(D_{ij})\in {\mathcal S^n}$
with nonnegative elements and zero diagonal is called
a {\em pre-distance matrix} \index{pre-distance matrix, $D$}.
In addition, if there exist points $p_1,\ldots,p_n \in \mathbb{R}^r$ such that
\begin{equation} \label{eq:dn}
D_{ij}=\|{p_i-p_j}\|_2^2, \quad i,j=1,\ldots,n,
\end{equation}
then $D$ is called a {\em Euclidean distance matrix}, denoted
\index{Euclidean distance matrix, \mbox{\boldmath$EDM$}\,}
\mbox{\boldmath$EDM$}\,\@. Note that we work with {\em squared distances}.
The smallest value of $r$ such that \eqref{eq:dn} holds
is called the {\em embedding dimension} of $D$.
Throughout the paper, we assume that $r$ is given and {\em fixed}.
\index{embedding dimension ({\em fixed}), $r$}
The set of \mbox{\boldmath$EDM$}\, matrices forms a closed convex cone in ${\mathcal S^n}$, denoted
${{\mathcal E}^n} $.
\index{cone of \mbox{\boldmath$EDM$}\,, ${{\mathcal E}^k} $}
If we are given an $n \times n$ partial \mbox{\boldmath$EDM$}\, $D_p$,
let ${\mathcal G} =(N,E,\omega)$ be the corresponding simple graph on the
\index{graph of the \mbox{\boldmath$EDM$}\,, ${\mathcal G} =(N,E,\omega)$}
nodes $N = 1\!:\!n$ whose edges $E$ correspond to the known
entries of $D_p$, with $(D_p)_{ij}=\omega_{ij}^2$, for all $(i,j) \in E$.
\begin{figure}[htb]
\epsfxsize=310pt
\centerline{\epsfbox{SNLfig.eps}}
\caption{Graph of partial \mbox{\boldmath$EDM$}\, with sensors
\textcolor{red}{$\circ$} and anchors $\blacksquare$}
\label{fig:fc1}
\end{figure}
\begin{defi}
For $Y\in {\mathcal S^n}$ and $\alpha \subseteq 1\!:\!n$, we let $Y[\alpha]$
denote the corresponding
\index{principal submatrix, $Y[\alpha]$}
{\em principal submatrix} formed from the rows and columns with indices
$\alpha$. If, in addition, $|\alpha |=k$ and
$\bar Y \in {\mathcal S} ^k$ is given, then we define
\[
{\mathcal S}^{n}(\alpha,\bar Y):= \left\{ Y \in {\mathcal S^n} : Y[\alpha] = \bar Y \right\}, \quad
{\mathcal S}_+^{n}(\alpha,\bar Y):= \left\{ Y \in {\mathcal S^n_+\,} : Y[\alpha] = \bar Y \right\}.
\]
That is, the subset of matrices $Y \in {\mathcal S^n}$ ($Y \in {\mathcal S^n_+\,}$)
\index{principal submatrix set, ${\mathcal S}^{n}(\alpha,\bar Y)$}
\index{principal submatrix positive semidefinite set, ${\mathcal S}_+^{n}(\alpha,\bar Y)$}
with principal submatrix $Y[\alpha]$ fixed to $\bar Y$.
\end{defi}
\noindent
For example, the subset of matrices
in ${\mathcal S^n}$ with the top left $k\times k$ block fixed is
\begin{equation}
\label{eq:setSblock}
{\mathcal S}^{n}(1\!:\!k,\bar Y) = \left\{Y \in {\mathcal S^n} : Y=
\left[
\begin{array}{c|c}
\bar Y & \cdot \cr
\hline
\cdot & \cdot
\end{array}
\right]
\right\}.
\index{top-left block fixed, ${\mathcal S}^{n}(1\!:\!k,\bar Y))$}
\index{${\mathcal S}^{n}(1\!:\!k,\bar Y))$, top-left block fixed}
\index{principal submatrix top-left block, ${\mathcal S}^{n}(1\!:\!k,\bar Y)$}
\index{${\mathcal S}^{n}(1\!:\!k,\bar Y)$, principal submatrix top-left block}
\end{equation}
A clique $\gamma \subseteq 1\!:\!n$
\index{clique}
in the graph ${\mathcal G} $ corresponds to a subset of sensors for
which the distances $\omega_{ij}=\|p_i-p_j\|_2$ are known, for all
$i,j \in \gamma$; equivalently, the clique corresponds
to the principal submatrix $D_p[\gamma]$ of the
partial \mbox{\boldmath$EDM$}\, matrix $D_p$, where all the elements of $D_p[\gamma]$ are known.
Suppose that we are given a subset of the
(squared) distances from \eqref{eq:dn} in the form of a partial \mbox{\boldmath$EDM$,}\, $D_p$.
The {\em \mbox{\boldmath$EDM$}\, completion problem} consists of finding the missing entries
of $D_p$ to complete the $\mbox{\boldmath$EDM$}\,$; see Figure \ref{fig:fc1}.
\index{\mbox{\boldmath$EDM$}\, completion problem}
This completion problem can be solved by finding a set of points $p_1,\ldots,p_n \in \mathbb{R}^r$ satisfying \eqref{eq:dn}, where $r$ is the embedding dimension of the partial \mbox{\boldmath$EDM$,}\, $D_p$.
\index{embedding dimension ({\em fixed}), $r$}
This problem corresponds to the graph
realizability problem with dimension $r$, which is the problem of finding
\index{graph realizability}
positions in $\mathbb{R}^r$ for the vertices of a graph such that the
inter-distances of these positions satisfy the
given edge lengths of the graph.
Let $Y \in {\mathcal M^n\,}$ be an $n \times n$ real matrix and $y \in \mathbb{R}^n$ a vector.
\index{$n\times n$ matrices, ${\mathcal M^n\,}$}
We let $\mbox{diag}(Y)$ denote the vector in $\mathbb{R}^n$ formed from the
diagonal of $Y$ and we let $\mbox{Diag}(y)$ denote the diagonal matrix in
${\mathcal M^n\,}$ with the vector $y$ along its diagonal.
Note that $\diag$ and $\Diag$ are the adjoint
linear transformations of each other: $\Diag = \mbox{diag}^*$.
The operator $\offDiag$ can then be defined as
$\offDiag (Y) := Y - \Diag ( \diag Y )$.
\index{offDiag operator of a matrix, $\offDiag M$}
\index{diagonal of a matrix, $\diag M$}
\index{diagonal matrix from a vector, $\Diag v$}
For
\[
P=\begin{bmatrix}
p_1^T\cr p_2^T\cr \vdots\cr p_n^T\cr
\end{bmatrix} \in {\mathcal M\,}^{n\times r},
\]
where $p_j$, $j=1,\ldots,n$, are the points used in \eqref{eq:dn},
let $Y:=PP^T$, and let $D$ be the corresponding \mbox{\boldmath$EDM$}\, satisfying \eqref{eq:dn}.
\index{matrix of points in space, $P$}
Defining the linear operators $\KK$ and ${\mathcal D} _e$ on ${\mathcal S^n}$ as follows, we see that
\begin{equation} \begin{array}{rcl} \label{KK}
\KK(Y)& := & {\mathcal D} _e(Y)-2Y \\
& := & \mbox{diag}(Y)\,e^T + e\,\mbox{diag}(Y)^T - 2Y \\
& = & \left( p_{i}^{T} p_{i} + p_{j}^{T} p_{j}
- 2 p_{i}^{T} p_{j} \right)_{i,j = 1}^{n} \\
& = & \left( \| p_{i} - p_{j} \|_{2}^{2} \right)_{i,j = 1}^{n} \\
& = & D.
\end{array} \end{equation}
That is, $\KK$ maps the positive semidefinite matrix $Y$ onto the \mbox{\boldmath$EDM$}\, D.
More generally, we can allow for a general vector $v$ to replace $e$,
and define ${\mathcal D} _v(Y) := \mbox{diag}(Y)\,v^T + v\,\mbox{diag}(Y)^T$.
By abuse of notation, we also allow ${\mathcal D} _v$ to act on a vector;
that is, ${\mathcal D} _v(y) := yv^T+vy^T$.
The adjoint of $\KK$ is
\begin{equation} \begin{array}{rcl} \label{KKs}
\KK^*(D) &=& 2(\Diag(De)-D).
\end{array} \end{equation}
The linear operator $\KK$ is one-one and onto between the
{\em centered} and {\em hollow} subspaces of ${\mathcal S^n}$, which are defined as
\begin{equation} \begin{array}{rcll}
{\mathcal S}_C &:=& \{ Y \in {\mathcal S^n} : Ye = 0 \} & \mbox{(zero row sums)}, \\
{\mathcal S}_H& := & \{ D \in {\mathcal S^n} : \diag(D) = 0 \} & = {\mathcal R} (\offDiag).
\end{array}
\end{equation}
Let $J:=I-\frac 1n ee^T$ denote the orthogonal projection onto the
\index{$J$, orthogonal projection onto $\{e\}^\perp$}
subspace $\{e\}^\perp$ and define the linear operator
${\mathcal T} (D) := -\frac 12 J \offDiag (D) J$.
\index{${\mathcal T} = \KK^\dagger$}
Then we have the following relationships.
\begin{prop}(\cite{homwolkA:04})
\label{prop:KTonetoone}
The linear operator ${\mathcal T} $ is the generalized inverse of
the linear operator $\KK$; that is, $\KK^\dagger ={\mathcal T} $.
Moreover:
\begin{equation} \label{eq:RN1}
\begin{array}{rcl}
{\mathcal R} (\KK) = {\mathcal S}_H; & \quad &{\mathcal N}(\KK) = {\mathcal R} ({\mathcal D} _e);\\
{\mathcal R} (\KK^*)={\mathcal R} ({\mathcal T} ) = {\mathcal S}_C ; & \quad & {\mathcal N}(\KK^*)= {\mathcal N}({\mathcal T} ) = {\mathcal R} (\Diag);
\end{array}
\end{equation}
\begin{equation} \label{eq:RN3}
\begin{array}{rcl}
{\mathcal S^n}= {\mathcal S}_H \oplus {\mathcal R} (\Diag) = {\mathcal S}_C \oplus {\mathcal R} ({\mathcal D} _e).
\end{array}
\end{equation}
\end{prop}
\begin{thm}(\cite{homwolkA:04})
\label{thm:KTonetoone}
The linear operators ${\mathcal T} $ and $\KK$ are one-to-one and onto mappings between the cone ${{\mathcal E}^n} \subset {\mathcal S}_H$ and the face of the semidefinite cone ${\mathcal S^n_+\,} \! \cap {\mathcal S}_C $. That is,
\[ {\mathcal T} ({{\mathcal E}^n} ) = {\mathcal S^n_+\,} \! \cap {\mathcal S}_C
\quad \mbox{and} \quad
\KK({\mathcal S^n_+\,} \! \cap {\mathcal S}_C ) = {{\mathcal E}^n} . \]
\end{thm}
\begin{rem}
$D \in {{\mathcal E}^n} $ has embedding dimension $r$ if and only if
\index{embedding dimension ({\em fixed}), $r$}
$\KK^\dagger (D)\succeq 0$ and
${\rm rank\,}\!(\KK^\dagger (D))=r$. In addition, we get $\KK^\dagger (D) e=0$.
Therefore, we can factor $\KK^\dagger (D) =PP^T$,
for some $P \in {\mathcal M\,}^{n\times r}$, to recover the
(centered) sensors in $\mathbb{R}^r$ from the rows in $P$.
Note that rotations of the points in the rows of $P$ do not change the
value $Y=PP^T$, since $PP^T=PQ^TQP$ if $Q$ is orthogonal.
However, the nullspace of $\KK$ is related to translations of the points in $P$.
Let $D \in {{\mathcal E}^n} $ with embedding dimension $r$ and
let $Y:=\KK^{\dagger}(D)$ have full rank factorization $Y=PP^T$, with $P
\in {\mathcal M\,}^{n \times r}$.
Then the translation of points in the rows of $P$ to $\bar P:=P+ew^T$, for
some $w\in \mathbb{R}^r$, results in $\bar Y := \bar P \bar P^T = Y+{\mathcal D} _e(y)$, with
$y:=Pw+\frac {w^Tw}2 e$, and
$\KK(\bar Y)=\KK(Y)=D$, since ${\mathcal D} _e(y)\in {\mathcal N}(\KK)$.
Note that ${\mathcal R} (Y)={\mathcal R} (P)$, therefore
$y=Pw+\frac {w^Tw}2 e \in {\mathcal R} (Y) + \cone \{e\}$,
as we will also see in more generality in Lemma~\ref{lem:psdDe} below.
\end{rem}
Let $D_p \in {\mathcal S} ^n$ be a \emph{partial} \mbox{\boldmath$EDM$}\, with embedding dimension $r$
and let $W \in {\mathcal S} ^n$ be the $0$--$1$ matrix corresponding to the known entries of $D_p$.
One can use the substitution $D=\KK(Y)$, where $Y\in {\mathcal S^n_+\,} \cap {\mathcal S} _C$,
in the \mbox{\boldmath$EDM$}\, completion problem
\[
\begin{array}{cc}
\mbox{Find} & D \in {\mathcal E} ^n \\
\mbox{s.t.} & W \circ D = W \circ D_p
\end{array}
\]
to obtain the \mbox{\boldmath$SDP$}\, relaxation
\[
\begin{array}{cc}
\mbox{Find} & Y \in {\mathcal S} ^n_+ \cap {\mathcal S} _C \\
\mbox{s.t.} & W \circ \KK(Y) = W \circ D_p
\end{array}.
\]
This relaxation does not restrict the rank of $Y$ and may
yield a solution with embedding dimension that is too large,
if ${\rm rank\,} (Y) > r$.
Moreover, solving \mbox{\boldmath$SDP$}\, problems with rank restrictions is NP-hard.
However, we work on faces of ${\mathcal S^n_+\,}$ described by $U{\mathcal S} _+^tU^T$, with $t \leq n$.
In order to find the face with the smallest dimension $t$, we must have the correct knowledge of the matrix $U$.
In this paper, we obtain information on $U$ using the cliques in the
graph of the partial \mbox{\boldmath$EDM$}\,.
\section{Semidefinite Facial Reduction}
\label{sect:cliqred}
We now present several techniques for reducing an \mbox{\boldmath$EDM$}\, completion problem when
one or more (possibly intersecting) cliques are known.
This extends the reduction using disjoint cliques presented in
\cite{DiKrQiWo:06,DiKrQiWo:08}. In each case, we take advantage of the
loss of Slater's constraint qualification and project the problem
to a lower dimensional \mbox{\boldmath$SDP$}\, cone.
We first need the following two technical lemmas
that exploit the structure of the \mbox{\boldmath$SDP$}\, cone.
\begin{lem}
\label{lem:psdDe}
Let $B \in {\mathcal S^n}$, $Bv=0$, $v \neq 0$, $y \in \mathbb{R}^n$ and $\bar Y := B+{\mathcal D} _v(y)$.
If $\bar Y \succeq 0$, then
\[
y \in {\mathcal R} (B) + \cone \{v\}.
\]
\end{lem}
\begin{proof}
First we will show that $y \in {\mathcal R} (B) + {\rm span}\,\{v\} = {\mathcal R} \left( \begin{bmatrix} B & v \end{bmatrix} \right)$. If this is not the case, then $y$ can be written as the orthogonal decomposition
\[
y = B u + \beta v + \bar y,
\]
where $0 \neq \bar y \in {\mathcal R} \left( \begin{bmatrix} B & v \end{bmatrix} \right)^\perp = {\mathcal N}\left( \begin{bmatrix} B & v \end{bmatrix}^{T} \right)$.
Note that $\bar y$ satisfies $B \bar y = 0$ and $v^{T} \bar y = 0$. To get a contradiction with the assumption that $\bar Y \succeq 0$, we let
\[
z := \frac{1}{2}\frac{v}{\|v\|^{2}} - (1+|\beta|)\frac{\bar y}{\|\bar y\|^{2}},
\]
and observe that $Bz = 0$ and $v^{T}z = 1/2$. Then,
\[
\begin{array}{rcl}
z^T\bar Y z
&=&
z^T {\mathcal D} _v(y) z
\\&=&
z^T \left( yv^T +vy^T \right) z
\\&=&
y^{T} z
\\&=&
\frac{1}{2}\beta + \bar y^{T} z
\\&<&
\frac{1}{2}(1+|\beta|) + \bar y^Tz
\\&=&
-\frac{1}{2}(1+|\beta|)
\\&<& 0,
\end{array}
\]
which gives us the desired contradiction. Therefore, $y \in {\mathcal R} (B) + {\rm span}\,\{v\}$, so to show that $y \in {\mathcal R} (B) + \cone\{v\}$, we only need to show that if $y = B u + \beta v$, then $\beta \geq 0$. First note that $v^{T}y = \beta v^{T}v$. Then,
\[
\begin{array}{rcl}
v^T \bar Y v
&=&
v^T \left( yv^T +vy^T \right) v
\\&=&
2 v^{T}y v^{T}v
\\&=&
2 \beta (v^{T}v)^{2}.
\end{array}
\]
Since $\bar Y \succeq 0$, we have $2 \beta (v^{T}v)^{2} \geq 0$. This
implies that $\beta \geq 0$, since $v \neq 0$.
\end{proof}
If $\bar Y \in {\mathcal S^{k}_+}\,$, then we can use the minimal face of ${\mathcal S^{k}_+}\,$
containing $\bar Y$ to find an expression for the minimal face of
${\mathcal S^n_+\,}$ that contains ${\mathcal S}_+^{n}(1\!:\!k,\bar Y)$.
\begin{lem}
\label{lem:rangeYbar}
Let $\bar U \in {\mathcal M\,}^{k\times t}$ with $\bar U^T \bar U=I_t$.
If $\face \{\bar Y\} \unlhd \bar U {\mathcal S^{t}_+}\, \bar U^{T}$, then
\begin{equation}
\label{eq:facereprYsub}
\face {\mathcal S}_+^{n}(1\!:\!k,\bar Y) \unlhd
\begin{bmatrix} \bar U & 0 \cr 0 & I_{n-k} \end{bmatrix}
{\mathcal S} ^{n-k+t}_+
\begin{bmatrix} \bar U & 0 \cr 0 & I_{n-k} \end{bmatrix}^T.
\end{equation}
Furthermore, if $ \face \{\bar Y\} = \bar U {\mathcal S^{t}_+}\, \bar U^{T}$,
then
\begin{equation}
\label{eq:facereprY}
\face {\mathcal S}_+^{n}(1\!:\!k,\bar Y) =
\begin{bmatrix} \bar U & 0 \cr 0 & I_{n-k} \end{bmatrix}
{\mathcal S} ^{n-k+t}_+
\begin{bmatrix} \bar U & 0 \cr 0 & I_{n-k} \end{bmatrix}^T.
\end{equation}
\end{lem}
\begin{proof}
Since $\bar Y \in \bar U {\mathcal S^{t}_+}\, \bar U^{T}$, then $\bar Y = \bar U S \bar U^T$,
for some $S \in {\mathcal S^{t}_+}\,$.
Let $Y \in {\mathcal S}_+^{n}(1\!:\!k,\bar Y)$ and choose $\bar V$ so that
$\begin{bmatrix}\bar U & \bar V\end{bmatrix}$ is an orthogonal matrix.
Then, with $Y$ blocked appropriately, we evaluate the congruence
\[
0\preceq \begin{bmatrix} \bar V & 0 \cr 0 &I_{n-k} \end{bmatrix}^T
Y \begin{bmatrix} \bar V & 0 \cr 0 & I_{n-k} \end{bmatrix} =
\begin{bmatrix} 0 & \bar V^T Y_{21}^T\cr Y_{21} \bar V & Y_{22} \end{bmatrix}=
\begin{bmatrix} 0 & 0 \cr 0 & Y_{22} \end{bmatrix}.
\]
Therefore, $Y\succeq 0$ implies that $\bar V^TY_{21}^T =0$.
Since ${\mathcal N}(\bar V^T)= {\mathcal R} (\bar U)$, we get $Y_{21}^T=\bar U X$, for some $X$.
Therefore, we can write
\[
Y=
\begin{bmatrix} \bar U & 0 \cr 0 & I_{n-k} \end{bmatrix}
\begin{bmatrix} S & X \cr X^T & Y_{22} \end{bmatrix}
\begin{bmatrix} \bar U & 0 \cr 0 & I_{n-k} \end{bmatrix}^T.
\]
This implies that $\face {\mathcal S}_+^{n}(1\!:\!k,\bar Y) \unlhd U {\mathcal S} ^{n-k+t}_+ U^T$, where
\[
U :=
\begin{bmatrix} \bar U & 0 \cr 0 & I_{n-k} \end{bmatrix}.
\]
This proves \eqref{eq:facereprYsub}. To prove \eqref{eq:facereprY},
note that if $ \face \{\bar Y\} = \bar U {\mathcal S^{t}_+}\, \bar U^{T}$ then
$ \bar Y \in \relint \left( \bar U {\mathcal S^{t}_+}\, \bar U^{T}\right)$, so
$\bar Y = \bar U S \bar U^{T}$, for some $S \in {\mathcal S^{t}_{++}}\,$. Letting
\[
\hat{Y} :=
\begin{bmatrix} \bar U & 0 \cr 0 & I_{n-k} \end{bmatrix}
\begin{bmatrix} S & 0 \cr 0 & I_{n-k} \end{bmatrix}
\begin{bmatrix} \bar U & 0 \cr 0 & I_{n-k} \end{bmatrix}^{T},
\]
we see that $\hat{Y} \in {\mathcal S}_+^{n}(1\!:\!k,\bar Y) \cap
\relint \left( U {\mathcal S} ^{n-k+t}_+ U^T \right)$.
This implies that there is no smaller face of ${\mathcal S^n_+\,}$ containing
${\mathcal S}_+^{n}(1\!:\!k,\bar Y)$, completing the proof.
\end{proof}
\subsection{Single Clique Facial Reduction}
\label{sect:singleclique}
If the principal submatrix $\bar D \in {\mathcal E} ^k$ is given,
for index set $\alpha \subseteq 1\!:\!n$, with $|\alpha| = k$,
we define
\begin{equation}
\label{eq:setS}
{\mathcal E}^{n}(\alpha,\bar D):= \left\{ D \in {{\mathcal E}^n} : D[\alpha] = \bar D \right\}.
\end{equation}
\index{${\mathcal E}^{n}(\alpha,\bar D)$, principal submatrix of \mbox{\boldmath$EDM$}\,}
\index{principal submatrix of \mbox{\boldmath$EDM$}\,, ${\mathcal E}^{n}(\alpha,\bar D)$}
Similarly, the subset of matrices
in ${{\mathcal E}^n} $ with the top left $k\times k$ block fixed is
\begin{equation}
\label{eq:setEblock}
{\mathcal E}^{n}(1\!:\!k,\bar D) = \left\{D \in {{\mathcal E}^n} : D =
\left[
\begin{array}{c|c}
\bar D & \cdot \cr
\hline
\cdot & \cdot
\end{array}
\right]
\right\}.
\index{top-left block fixed, ${\mathcal E}^{n}(1\!:\!k,\bar D)$}
\index{principal submatrix top-left block, ${\mathcal E}^{n}(1\!:\!k,\bar D)$}
\end{equation}
A fixed principal submatrix $\bar D$ in a partial \mbox{\boldmath$EDM$}\, $D$
corresponds to a clique $\alpha$ in the graph ${\mathcal G} $ of the
partial \mbox{\boldmath$EDM$}\, $D$.
Given such a fixed clique defined by the submatrix $\bar D$,
the following theorem shows that the following set, containing the feasible set
of the corresponding \mbox{\boldmath$SDP$}\, relaxation,
\[
\left\{
Y \in {\mathcal S^n_+\,} \cap {\mathcal S}_C :
\KK(Y[\alpha]) = \bar D
\right\}
=
\KK^\dagger \left({\mathcal E}^{n}(\alpha,\bar D) \right),
\]
is contained in a proper face of ${\mathcal S^n_+\,}$.
This means that the Slater
constraint qualification (strict feasibility) fails, and we can reduce
the size of the \mbox{\boldmath$SDP$}\, problem; see \cite{DiKrQiWo:06}.
We expand on this and find an explicit expression for
$\face \KK^\dagger \left({\mathcal E}^{n}(\alpha,\bar D) \right)$ in the following Theorem~\ref{thm:onecliquered}.
For simplicity, here and below, we often work with
ordered sets of integers for the two cliques. This simplification
can always be obtained by a permutation of the indices of the
sensors.
\begin{thm}
\label{thm:onecliquered}
Let $D \in {{\mathcal E}^n} $, with embedding dimension $r$.
Let $\bar D := D[1\!:\!k] \in {{\mathcal E}^k} $ with embedding dimension $t$,
and $B := \KK^\dagger(\bar D) = \bar U_B S \bar U_B^T$,
where $\bar U_B \in {\mathcal M\,}^{k\times t}$, $\bar U_B^T \bar U_B = I_t$,
and $S \in {\mathcal S} ^t_{++}$.
Furthermore, let $U_B :=
\begin{bmatrix}
\bar U_B & \frac 1{\sqrt{k}} e
\end{bmatrix}
\in {\mathcal M\,}^{k\times (t+1)}$,
$U := \begin{bmatrix} U_B & 0 \cr 0 & I_{n-k} \end{bmatrix}$,
and let $\begin{bmatrix} V & \frac{U^Te}{\|U^Te\|} \end{bmatrix}
\in {\mathcal M\,}^{n-k+t+1}$ be orthogonal.
Then
\begin{equation}
\label{eq:UBY}
\face \KK^\dagger\left({\mathcal E}^{n}(1\!:\!k,\bar D)\right)
=\left( U {\mathcal S} _+^{n-k+t+1} U^T \right) \cap {\mathcal S}_C
= (U V) {\mathcal S} _+^{n-k+t} (U V)^T.
\end{equation}
\end{thm}
\begin{proof}
Let $Y \in \KK^\dagger\left({\mathcal E}^{n}(1\!:\!k,\bar D)\right)$ and $\bar Y := Y[1\!:\!k]$.
Then there exists $D \in {\mathcal E}^{n}(1\!:\!k,\bar D)$ such that $Y = \KK^\dagger(D)$,
implying that $\KK(Y) = D$, and that $\KK(\bar Y) = \bar D = \KK(B)$.
Thus, $\bar Y \in B + {\mathcal N}(\KK) = B + {\mathcal R} ({\mathcal D} _e)$,
where the last equality follows from Proposition~\ref{prop:KTonetoone}.
This implies that $\bar Y = B + {\mathcal D} _{e}(y)$, for some $y \in \mathbb{R}^k$.
From Theorem~\ref{thm:KTonetoone}, we get $\bar Y \succeq 0$ and $Be=0$.
Therefore, Lemma~\ref{lem:psdDe} implies that $y = Bu + \beta e$,
for some $u \in \mathbb{R}^k$ and $\beta \geq 0$. This further implies
\[
\bar Y = B + Bue^T + eu^TB + 2\beta ee^T.
\]
From this expression for $\bar Y$, we can see that
${\mathcal R} (\bar Y) \subseteq
{\mathcal R} \left(\begin{bmatrix} B & e \end{bmatrix}\right)
= {\mathcal R} (U_B)$, where the last equality follows from the fact that $Be = 0$.
Therefore, $\bar Y \in U_{B} {\mathcal S} ^{t+1}_+ U_{B}^{T}$,
implying, by Lemma~\ref{lem:rangeYbar}, that
$\face {\mathcal S}_+^{n}(1\!:\!k,\bar Y) \unlhd U {\mathcal S} ^{n-k+t+1}_+ U^T$.
Since $Y \in {\mathcal S}_+^{n}(1\!:\!k,\bar Y)$ and $Ye = 0$,
we have that $Y \in \left( U {\mathcal S} _+^{n-k+t+1} U^T \right) \cap {\mathcal S}_C $.
Therefore, $\face \KK^\dagger\left({\mathcal E}^{n}(1\!:\!k,\bar D)\right)
\unlhd \left( U {\mathcal S} _+^{n-k+t+1} U^T \right) \cap {\mathcal S}_C $.
Since $V^T U^T e = 0$, we have that
\begin{equation}
\label{eq:UVdotUVT}
\left( U {\mathcal S} _+^{n-k+t+1} U^T \right) \cap {\mathcal S}_C
= U V {\mathcal S} _+^{n-k+t} V^T U^T.
\end{equation}
To show that $\face \KK^\dagger\left({\mathcal E}^{n}(1\!:\!k,\bar D)\right)
= \left( U {\mathcal S} _+^{n-k+t+1} U^T \right) \cap {\mathcal S}_C $, we need to find
\begin{equation}
\label{eq:Yhat}
\hat Y = UZU^T \in \KK^\dagger\left({\mathcal E}^{n}(1\!:\!k,\bar D)\right), \quad
\mbox{with} \ {\rm rank\,}(\hat Y) = n-k+t, \ \hat Y e = 0, \ Z \in {\mathcal S} _+^{n-k+t+1}.
\end{equation}
To accomplish this, we let $T_1= \begin{bmatrix} S & 0 \cr 0 & 1 \end{bmatrix}$.
Then $T_1 \succ 0$ and
\[
B+\frac 1k ee^T = U_B T_1 U_B^T = \bar P \bar P^T, \quad
\mbox{where} \ \bar P := U_B T_1^{1/2} \in {\mathcal M\,}^{k \times (t+1)}.
\]
Let
\[
P :=
\left[
\begin{array}{c|c}
\bar P & 0 \cr
\hline
0 & I_{n-k-1} \cr
-e^T \bar P & -e^T
\end{array}
\right]
\in {\mathcal M\,}^{n \times (n-k+t)}.
\]
Since $\bar P$ has full-column rank, we see that $P$ also has full-column rank. Moreover, $P^Te = 0$. Therefore,
\[
\hat Y := PP^T =
\left[
\begin{array}{c|cc}
\bar P \bar P^T & 0 & -e \cr
\hline
0 & I_{n-k-1} & -e \cr
-e^T & -e^T & n - 1
\end{array}
\right]
\in {\mathcal S^n_+\,},
\]
satisfies $\hat Y e = 0$ and ${\rm rank\,}(\hat Y) = n-k+t$.
Furthermore, we have that $\hat Y = U Z U^T$, where
\[
Z =
\left[
\begin{array}{cc|cc}
S & 0 & 0 & 0 \cr
0 & 1 & 0 & -\sqrt{k} \cr
\hline
0 & 0 & I_{n-k-1} & -e \cr
0 & -\sqrt{k} & -e^T & n - 1
\end{array}
\right]
\in {\mathcal S} ^{n-k+t+1}.
\]
Note that we can also write $Z$ as
\[
Z =
\begin{bmatrix}
S & 0 \cr
0 & T
\end{bmatrix}
\in {\mathcal S} ^{n-k+t+1},
\]
where
\[
T :=
\begin{bmatrix}
1 & 0 & -\sqrt{k} \cr
0 & I_{n-k-1} & -e \cr
-\sqrt{k} & -e^T & n - 1
\end{bmatrix}
\in {\mathcal S} ^{n-k+1}.
\]
The eigenvalues of $T$ are $0$, $1$, and $n$, with multiplicities $1$, $n-k-1$, and $1$, respectively. Therefore, ${\rm rank\,}(T) = n-k$, which implies that ${\rm rank\,}(Z) = n-k+t$
and $Z \succeq 0$.
Letting $\hat D := \KK(\hat Y)$, we have that $\hat D \in {\mathcal E}^{n}(1\!:\!k,\bar D)$, since
\[
\hat D[1\!:\!k] = \KK(\hat Y[1\!:\!k]) = \KK(\bar P \bar P^T)
= \KK\left(B + \frac{1}{k}ee^T\right) = \KK(B) = \bar D.
\]
Therefore, $\hat Y$ satisfies \eqref{eq:Yhat}, completing the proof.
\end{proof}
\begin{remark}
Theorem~\ref{thm:onecliquered} provides a reduction in the dimension of
the \mbox{\boldmath$EDM$}\, completion problem. Initially, our problem consists in finding
$Y\in {\mathcal S^n_+\,} \cap {\mathcal S}_C $ such that the constraint
\[
\KK(Y[\alpha])=D[\alpha], \quad \alpha = 1\!:\!k,
\]
holds.
After the reduction, we have the smaller dimensional variable
$Z \in {\mathcal S} _+^{n-k+t}$; by construction $Y := (UV)Z(UV)^T$ will automatically
satisfy the above constraints.
This is a reduction of $k-t-1=(n-1) -(n-k+t)$ in the dimension of the matrix variable.
The addition of the vector $e$ to the range of $B$,
$U_B := \begin{bmatrix} \bar U_B & \frac 1{\sqrt{k}} e \end{bmatrix}$,
has a geometric interpretation.
If $B=PP^T$, $P\in {\mathcal M\,}^{k \times t}$, then the rows of $P$ provide
{\em centered} positions for the $k$ sensors in the clique $\alpha$.
However, these sensors are not necessarily centered once they are combined
with the remaining $n-k$ sensors. Therefore, we have to
allow for translations, e.g.
to $P+ev^T$ for some $v$. The multiplication
$(P+ev^T)(P+ev^T)^T=PP^T +Pve^T+ev^TP^T+ev^Tve^T$ is included in the set of matrices that we get after adding $e$ to the range of $B$.
Note that $Pve^T+ev^TP^T+ev^Tve^T = {\mathcal D} _e(y)$, for $y = Pv + \frac12 ev^Tv$.
\end{remark}
The special case $k=1$ is of interest.
\begin{cor}
\label{cor:cliqueredk1}
Suppose that the hypotheses of Theorem~\ref{thm:onecliquered} hold but
that $k=1$ and $\bar D = 0$. Then $U_B=1$, $U=I_n$, and
\begin{equation}
\label{eq:UBYk1}
\face \KK^\dagger\left({\mathcal E}^{n}(1\!:\!k,\bar D)\right)
=\face \KK^\dagger\left({{\mathcal E}^n} \right)
= {\mathcal S} _+^{n} \cap {\mathcal S}_C = V {\mathcal S} _+^{n-1} V^T,
\end{equation}
where $\begin{bmatrix} V & \frac{1}{\sqrt{n}}e \end{bmatrix}
\in {\mathcal M\,}^{n}$ is orthogonal.
\end{cor}
\begin{proof}
Since $k=1$, necessarily we get $t=0$ and we can set $U_B=1$.
\end{proof}
\subsubsection{Disjoint Cliques Facial Reduction}
\label{sect:disjointcliques}
Theorem~\ref{thm:onecliquered} can be easily extended to two or more
disjoint cliques; see also \cite{DiKrQiWo:06}.
\begin{cor}
\label{cor:disjcliques}
Let $D \in {{\mathcal E}^n} $ with embedding dimension $r$.
Let $k_0:=1< k_1 < \ldots < k_l \leq n$.
For $i=1,\ldots, l$, let
$\bar D_i := D[k_{i-1}\!:\!k_i] \in {\mathcal E} ^{k_i-k_{i-1}+1}$
with embedding dimension $t_i$,
$B_i := \KK^\dagger(\bar D_i) = \bar U_{B_i} S \bar U_{B_i}^T$,
where $\bar U_{B_i} \in {\mathcal M\,}^{k\times t_i}$,
$\bar U_{B_i}^T \bar U_{B_i} = I_{t_i}$, $S_i \in {\mathcal S} ^{t_i}_{++}$,
and $U_{B_i} :=
\begin{bmatrix}
\bar U_{B_i} & \frac 1{\sqrt{k_i}} e
\end{bmatrix}
\in {\mathcal M\,}^{k\times (t_i+1)}$.
Let
\[
U := \begin{bmatrix} U_{B_1} & \ldots & 0 & 0 \cr
\vdots & \ddots & \vdots & \vdots \cr
0 & \ldots & U_{B_l} & 0 \cr
0 &\ldots & 0& I_{n-k_l} \end{bmatrix}
\]
and $\begin{bmatrix} V & \frac{U^Te}{\|U^Te\|} \end{bmatrix}
\in {\mathcal M\,}^{n-k_l+\sum_{i=1}^l t_i +l}$ be orthogonal.
Then
\begin{equation}
\label{eq:UBYdisjointcl}
\begin{array}{rcl}
\bigcap_{i=1}^l \face \KK^\dagger\left({{\mathcal E}^n} (k_{i-1}\!:\!k_i,\bar D_i) \right)
&=&
\left( U {\mathcal S} _+^{n-k_l+\sum_{i=1}^l t_i +l} U^T \right) \cap {\mathcal S}_C
\\&=&
(U V) {\mathcal S} _+^{n-k_l+\sum_{i=1}^l t_i+l-1} (U V)^T.
\end{array}
\end{equation}
\end{cor}
\begin{proof}
The result follows from noting that the range of $U$ is the intersection
of the ranges of the matrices $U_{B_i}$ with appropriate identity
blocks added.
\end{proof}
\subsection{Two (Intersecting) Clique Facial Reduction}
\label{sect:twoclique}
The construction \eqref{eq:UVdotUVT} illustrates how we can find the
intersection of two faces. Using this approach, we now extend
Theorem~\ref{thm:onecliquered} to two cliques that (possibly) intersect;
see the ordered indices in \eqref{eq:ordcliques}
and the corresponding Venn diagram in Figure~\ref{fig:alpha}.
We also find expressions for the intersection of
the corresponding faces in ${\mathcal S^n_+\,}$; see equation \eqref{eq:UBY2}.
The key is to find the intersection of the subspaces that
represent the faces, as in condition \eqref{eq:rangeU}.
\begin{figure}[htb]
\epsfxsize=180pt
\centerline{\epsfbox{alpha.eps}}
\caption{Venn diagram of the sets of ordered indices, $\alpha_1$ and $\alpha_2$,
in Theorem~\ref{thm:interclique}}
\label{fig:alpha}
\end{figure}
\begin{thm}
\label{thm:interclique}
Let $D \in {{\mathcal E}^n} $ with embedding dimension $r$ and, as in Figure~\ref{fig:alpha},
define the sets of positive integers
\begin{equation}
\label{eq:ordcliques}
\begin{array}{c}
\alpha_1 := 1\!:\!(\bar k_1 + \bar k_2), \quad
\alpha_2 := (\bar k_1 + 1)\!:\!(\bar k_1 + \bar k_2 + \bar k_3)
\subseteq 1\!:\!n, \\
k_1 := |\alpha_1| = \bar k_1 + \bar k_2, \quad
k_2 := |\alpha_2| = \bar k_2 + \bar k_3, \\
k := \bar k_1 + \bar k_2 + \bar k_3.
\end{array}
\end{equation}
For $i=1,2$, let $\bar D_i := D[\alpha_i] \in {\mathcal E} ^{k_i}$
with embedding dimension $t_i$,
and $B_i := \KK^\dagger(\bar D_i) = \bar U_i S_i \bar U_i^T$,
where $\bar U_i \in {\mathcal M\,}^{k_i\times t_i}$, $\bar U_i^T \bar U_i = I_{t_i}$,
$S_i \in {\mathcal S} ^{t_i}_{++}$, and
$U_i :=
\begin{bmatrix}
\bar U_i & \frac 1{\sqrt{k_i}} e
\end{bmatrix}
\in {\mathcal M\,}^{k_i \times (t_i+1)}$.
Let $t$ and $\bar U \in {\mathcal M\,}^{k \times (t+1)}$ satisfy
\begin{equation}
\label{eq:rangeU}
{\mathcal R} (\bar U) =
{\mathcal R} \left(
\begin{bmatrix} U_1 & 0 \cr 0 & I_{\bar k_3} \end{bmatrix}
\right)
\cap
{\mathcal R} \left(
\begin{bmatrix} I_{\bar k_1} & 0 \cr 0 & U_2 \end{bmatrix}
\right), \mbox{ with } \bar U^T \bar U = I_{t+1}.
\end{equation}
Let
$U := \begin{bmatrix} \bar U & 0 \cr 0 & I_{n-k} \end{bmatrix}
\in {\mathcal M\,}^{n \times (n-k+t+1)}$
and
$\begin{bmatrix} V & \frac{U^Te}{\|U^Te\|} \end{bmatrix}
\in {\mathcal M\,}^{n-k+t+1}$ be orthogonal. Then
\begin{equation}
\label{eq:UBY2}
\bigcap_{i=1}^2
\face \KK^\dagger\left({{\mathcal E}^n} (\alpha_i,\bar D_i)\right)
=\left( U {\mathcal S} _+^{n-k+t+1} U^T \right) \cap {\mathcal S}_C
= (U V) {\mathcal S} _+^{n-k+t} (U V)^T.
\end{equation}
\end{thm}
\begin{proof}
From Theorem~\ref{thm:onecliquered}, we have that
\[
\face \KK^\dagger\left({{\mathcal E}^n} (\alpha_1,\bar D_1)\right) =
\left(
\left[\begin{array}{cc|c}
U_1 & 0 & 0 \\
0 & I_{\bar k_3} & 0 \\
\hline
0 & 0 & I_{n-k}
\end{array}\right]
{\mathcal S} _+^{n-k_1+t_1+1}
\left[\begin{array}{cc|c}
U_1 & 0 & 0 \\
0 & I_{\bar k_3} & 0 \\
\hline
0 & 0 & I_{n-k}
\end{array}\right]^T
\right) \cap {\mathcal S}_C
\]
and, after a permutation of rows and columns in
Theorem~\ref{thm:onecliquered},
\[
\face \KK^\dagger\left({{\mathcal E}^n} (\alpha_2,\bar D_2)\right) =
\left(
\left[\begin{array}{cc|c}
I_{\bar k_1} & 0 & 0 \\
0 & U_2 & 0 \\
\hline
0 & 0 & I_{n-k}
\end{array}\right]
{\mathcal S} _+^{n-k_2+t_2+1}
\left[\begin{array}{cc|c}
I_{\bar k_1} & 0 & 0 \\
0 & U_2 & 0 \\
\hline
0 & 0 & I_{n-k}
\end{array}\right]^T
\right) \cap {\mathcal S}_C .
\]
The range space condition \eqref{eq:rangeU} then implies that
\[
{\mathcal R} (U) =
{\mathcal R} \left(
\left[\begin{array}{cc|c}
U_1 & 0 & 0 \\
0 & I_{\bar k_3} & 0 \\
\hline
0 & 0 & I_{n-k}
\end{array}\right]
\right)
\cap
{\mathcal R} \left(
\left[\begin{array}{cc|c}
I_{\bar k_1} & 0 & 0 \\
0 & U_2 & 0 \\
\hline
0 & 0 & I_{n-k}
\end{array}\right]
\right),
\]
giving us the result \eqref{eq:UBY2}.
\end{proof}
\begin{remark}
Theorem~\ref{thm:interclique} provides a reduction in the dimension of
the \mbox{\boldmath$EDM$}\, completion problem.
Initially, our problem consists in finding
$Y\in {\mathcal S^n_+\,} \cap {\mathcal S}_C $ such that the two constraints
\[
\KK(Y[\alpha_i])=D[\alpha_i],\quad i=1,2,
\]
hold. After the reduction, we want to find the smaller dimensional
$Z \in {\mathcal S} _+^{n-k+t}$; by construction $Y := (UV)Z(UV)^T$ will automatically
satisfy the above constraints.
\end{remark}
The explicit expression for the intersection of the two faces is
given in equation~\eqref{eq:UBY2} and uses the matrix $\bar U$
obtained from the intersection of the two ranges
in condition~\eqref{eq:rangeU}.
Finding a matrix whose range is the intersection
of two subspaces can be done using
\cite[Algorithm~12.4.3]{GoVan:79}. However, our subspaces have special
structure. We can exploit this structure to find the intersection;
see Lemma~\eqref{lem:twocliquered} and Lemma~\eqref{lem:twocliquereddeg} below.
The dimension of the face in \eqref{eq:UBY2} is reduced to $n-k+t$.
However, we can get a dramatic reduction if we have a common
block with embedding dimension $r$, and a reduction
in the case the common block has embedding dimension $r-1$ as well.
This provides an algebraic
proof using semidefinite programming of the rigidity of the union
\index{rigidity}
of the two cliques under this intersection assumption.
\subsubsection{Nonsingular Facial Reduction with Intersection Embedding Dimension $r$}
\label{sect:rigidreduction}
\index{nonsingular reduction}
We now consider the case when the intersection of the two cliques
results in $D[\alpha_1\cap \alpha_2]$ having embedding dimension $r$;
see Figure~\ref{fig:intersembedr}.
\begin{figure}[htb]
\epsfxsize=180pt
\centerline{\epsfbox{figclique_intersect.eps}}
\caption{Two clique reduction with intersection with embedding
dimension $r$}
\label{fig:intersembedr}
\end{figure}
We see that we can explicitly
find the completion of the \mbox{\boldmath$EDM$}\, $D[\alpha_1\cup \alpha_2]$.
We first need the following result on the
intersection of two structured subspaces.
\begin{lem}
\label{lem:twocliquered}
Let
\[
U_1:= \kbordermatrix{
& r+1 \cr
s_1 & U^{\prime}_1 \cr
k & U^{\prime\prime}_1
}, \quad
U_2:= \kbordermatrix{
& r+1 \cr
k & U^{\prime\prime}_2 \cr
s_2 & U^{\prime}_2
}, \quad
\hat U_1:= \kbordermatrix{
& r+1 & s_2 \cr
s_1 & U_1' & 0 \cr
k & U_1'' & 0 \cr
s_2 & 0 & I
}, \quad
\hat U_2:= \kbordermatrix{
& s_1 & r+1 \cr
s_1 & I & 0 \cr
k & 0 & U_2'' \cr
s_2 & 0 & U_2'
}
\]
be appropriately blocked with $U_1'', U_2'' \in {\mathcal M\,}^{k \times (r+1)}$
full column rank and ${\mathcal R} (U_1'') = {\mathcal R} (U_2'')$.
Furthermore, let
\begin{equation}
\label{eq:U1U2}
\bar U_1 :=
\kbordermatrix{
& r+1 \cr
s_1 & U_1^\prime\cr
k & U_1^{\prime\prime}\cr
s_2 & U_2^{\prime} (U_2^{\prime\prime})^\dagger U_1^{\prime\prime}
}, \quad
\bar U_2 :=
\kbordermatrix{
& r+1 \cr
s_1 & U_1^{\prime} (U_1^{\prime\prime})^\dagger U_2^{\prime\prime}\cr
k & U_2^{\prime\prime}\cr
s_2 & U_2^\prime
}.
\end{equation}
Then $\bar U_1$ and $\bar U_2$ are full column rank and satisfy
\[
{\mathcal R} (\hat U_1) \cap {\mathcal R} (\hat U_2) =
{\mathcal R} \left(\bar U_1 \right) = {\mathcal R} \left(\bar U_2 \right).
\]
Moreover, if $e_{r+1} \in \mathbb{R}^{r+1}$ is the $(r+1)^\mathrm{st}$ standard unit vector, and
$U_ie_{r+1} = \alpha_i e$, for some $\alpha_i \neq 0$, for $i=1,2$, then
$\bar U_ie_{r+1} = \alpha_i e$, for $i=1,2$.
\end{lem}
\begin{proof}
\label{pageproofblocks}
From the definitions, $x \in {\mathcal R} (\hat U_1) \cap {\mathcal R} (\hat U_2)$
if and only if
\[
x =
\begin{bmatrix}
x_1 \\
x_2 \\
x_3
\end{bmatrix}
=
\begin{bmatrix}
U^{\prime}_1 v_1 \\
U^{\prime\prime}_1 v_1 \\
v_2
\end{bmatrix}
=
\begin{bmatrix}
w_1 \\
U^{\prime\prime}_2 w_2 \\
U^{\prime}_2 w_2
\end{bmatrix}, \quad
\mbox{for some
$v = \begin{bmatrix} v_1 \cr v_2 \end{bmatrix}$,
$w = \begin{bmatrix} w_1 \cr w_2 \end{bmatrix}$. }
\]
Note that $U_1'' v_1 = U_2'' w_2$ if and only if $w_2 = (U_2'')^\dagger U_1'' v_1$;
this follows from the facts that
$U_2''$ full column rank implies $(U_2'')^\dagger U_2'' = I$,
and
${\mathcal R} (U_1'') = {\mathcal R} (U_2'')$ implies $U_2''(U_2'')^\dagger U_1'' = U_1''$.
Therefore, $x \in {\mathcal R} (\hat U_1) \cap {\mathcal R} (\hat U_2)$ if and only if
\[
x =
\begin{bmatrix}
x_1 \\
x_2 \\
x_3
\end{bmatrix}
=
\begin{bmatrix}
U_1' v_1 \\
U_1'' v_1 \\
U_2' (U_2'')^\dagger U_1'' v_1
\end{bmatrix}
=
\bar U_1 v_1,
\quad \mbox{for some $v_1$},
\]
with $v_2 := U_2' (U''_2)^\dagger U''_1 v_1$, $w_1 := U_1'v_1$, and
$w_2 := (U''_2)^\dagger U''_1 v_1$, implying that
${\mathcal R} (\hat U_1) \cap {\mathcal R} (\hat U_2) = {\mathcal R} (\bar U_1)$;
a similar argument shows that ${\mathcal R} (\hat U_1) \cap {\mathcal R} (\hat U_2) = {\mathcal R} (\bar U_2)$.
Now suppose, for $i=1,2$, that $U_i e_{r+1} = \alpha_i e$, for some $\alpha_i \neq 0$.
Then $e \in {\mathcal R} (\hat U_1) \cap {\mathcal R} (\hat U_2)$, so $e \in {\mathcal R} (\bar U_1)$,
implying that $\bar U_1 v = e$, for some vector $v$.
Since $\bar U_1 = \begin{bmatrix} U_1 \cr U_2' (U_2'')^\dagger U_1'' \end{bmatrix}$,
we have $U_1 v = e$.
Furthermore, since $U_1$ has full column rank, we conclude that
$v = \frac{1}{\alpha_1} e_{r+1}$, implying that $\bar U_1 e_{r+1} = \alpha_1 e$.
Similarly, we can show that $\bar U_2 e_{r+1} = \alpha_2 e$.
\end{proof}
We now state and prove a key result that shows we can complete the distances in the
union of two cliques provided that their intersection has embedding dimension equal to $r$.
\begin{thm}
\label{thm:twocliquered}
Let the hypotheses of Theorem~\ref{thm:interclique} hold.
Let
\[
\beta \subseteq \alpha_1 \cap \alpha_2, \quad
\bar D := D[\beta], \quad
B := \KK^\dagger(\bar D), \quad \bar U_\beta := \bar U[\beta,:],
\]
where
$\bar U \in {\mathcal M\,}^{k \times (t+1)}$
satisfies equation~\eqref{eq:rangeU}.
Let
$\begin{bmatrix} \bar V & \frac{\bar U^Te}{\|\bar U^Te\|} \end{bmatrix}
\in {\mathcal M\,}^{t+1}$ be orthogonal. Let
\begin{equation}
\label{eq:Zcalc1}
Z:= (J \bar U_\beta \bar V)^\dagger B ((J \bar U_\beta \bar V)^\dagger)^T.
\end{equation}
If the embedding dimension for $\bar D$ is $r$, then
$t = r$, $Z \in {\mathcal S} ^{r}_{++}$ is the unique
solution of the equation
\begin{equation}
\label{eq:Zbeta}
(J \bar U_\beta \bar V) Z (J \bar U_\beta \bar V)^T = B,
\index{$J$, orthogonal projection onto $\{e\}^\perp$}
\end{equation}
and
\begin{equation}
\label{eq:DKKUVZ}
D[\alpha_1 \cup \alpha_2] = \KK\left((\bar U \bar V)Z(\bar U \bar V)^T\right).
\end{equation}
\end{thm}
\begin{proof}
Since the embedding dimension of $\bar D$ is $r$, we have ${\rm rank\,}(B) = r$.
Furthermore, we have $Be = 0$ and $B \in {\mathcal S} ^{|\beta|}_+$, implying that
$|\beta| \geq r+1$. In addition, since the embedding dimension of $D$ is also
$r$, we conclude that the embedding dimension of $\bar D_i$ is $r$, for $i=1,2$.
Similarly, the embedding dimension of $D[\alpha_1 \cap \alpha_2]$ is also $r$.
Since $\bar U \in {\mathcal M\,}^{k \times (t+1)}$ satisfies equation~\eqref{eq:rangeU},
we have that
\[
{\mathcal R} (\bar U) =
{\mathcal R} \left(
\begin{bmatrix}
U^{\prime}_1 & 0 \cr
U^{\prime\prime}_1 & 0 \cr
0 & I_{\bar k_3}
\end{bmatrix}
\right)
\cap
{\mathcal R} \left(
\begin{bmatrix}
I_{\bar k_1} & 0 \cr
0 & U^{\prime\prime}_2 \cr
0 & U^{\prime}_2
\end{bmatrix}
\right).
\]
Note that we have partitioned $U_i =
\begin{bmatrix} \bar U_i & \frac{1}{\sqrt{k_i}} e \end{bmatrix}
\in {\mathcal M\,}^{k_i \times (r+1)}$ so that $U^{\prime\prime}_i =
\begin{bmatrix} \bar U''_i & \frac{1}{\sqrt{k_i}} e \end{bmatrix}
\in {\mathcal M\,}^{|\alpha_1 \cap \alpha_2| \times (r+1)}$, for $i=1,2$.
Moreover, we have used the fact that the embedding dimension of
$\bar D_i$ is $r$, so that $t_i=r$, for $i=1,2$.
We claim that $U_1''$ and $U_2''$ have full column rank and that
${\mathcal R} (U_1'') = {\mathcal R} (U_2'')$.
First we let $Y := \KK^\dagger(D[\alpha_1 \cup \alpha_2])$.
Then $Y \in \KK^\dagger\left({{\mathcal E}^k} (\alpha_1,\bar D_1)\right)$.
By Theorem~\ref{thm:onecliquered}, there exists $Z_1 \in {\mathcal S} ^{\bar k_3 + r + 1}_+$
such that
\[
Y =
\begin{bmatrix}
U^{\prime}_1 & 0 \cr
U^{\prime\prime}_1 & 0 \cr
0 & I_{\bar k_3}
\end{bmatrix}
Z_1
\begin{bmatrix}
U^{\prime}_1 & 0 \cr
U^{\prime\prime}_1 & 0 \cr
0 & I_{\bar k_3}
\end{bmatrix}^T.
\]
Therefore, $Y[\alpha_1 \cap \alpha_2] =
\begin{bmatrix} U^{\prime\prime}_1 & 0 \end{bmatrix}
Z_1
\begin{bmatrix} U^{\prime\prime}_1 & 0 \end{bmatrix}^T
\in
U_1'' {\mathcal S} ^{r+1}_+ (U_1'')^T$, so
\[
{\mathcal R} (Y[\alpha_1 \cap \alpha_2]) \subseteq {\mathcal R} (U_1'').
\]
Furthermore, since $\KK(Y) = D[\alpha_1 \cup \alpha_2]$, we have that
$\KK(Y[\alpha_1 \cap \alpha_2]) = D[\alpha_1 \cap \alpha_2]
= \KK \left( \KK^\dagger( D[\alpha_1 \cap \alpha_2] ) \right)$,
so $Y[\alpha_1 \cap \alpha_2] \in \KK^\dagger( D[\alpha_1 \cap \alpha_2] ) + {\mathcal N}(\KK)$.
Since ${\mathcal N}(\KK) = {\mathcal R} ({\mathcal D} _e)$, there exists a vector $y$ such that
\[
Y[\alpha_1 \cap \alpha_2] = \KK^\dagger( D[\alpha_1 \cap \alpha_2] ) + {\mathcal D} _e(y)
= \KK^\dagger( D[\alpha_1 \cap \alpha_2] ) + ye^T + ey^T.
\]
By Lemma~\ref{lem:psdDe},
$y \in
{\mathcal R} \left(
\begin{bmatrix} \KK^\dagger( D[\alpha_1 \cap \alpha_2] ) & e \end{bmatrix}
\right)$.
Therefore,
\[
{\mathcal R} (Y[\alpha_1 \cap \alpha_2])
=
{\mathcal R} \left(
\begin{bmatrix}
\KK^\dagger( D[\alpha_1 \cap \alpha_2] ) & e
\end{bmatrix}
\right).
\]
Moreover, ${\rm rank\,} \KK^\dagger( D[\alpha_1 \cap \alpha_2] ) = r$ and
$\KK^\dagger( D[\alpha_1 \cap \alpha_2] )e = 0$, so
\[
r+1 = \dim {\mathcal R} (Y[\alpha_1 \cap \alpha_2]) \leq \dim {\mathcal R} (U_1'') \leq r+1.
\]
Therefore, $U_1''$ has full column rank and
${\mathcal R} (U_1'') = {\mathcal R} (Y[\alpha_1 \cap \alpha_2])$.
Similarly, we can show that $U_2''$ has full column rank and
${\mathcal R} (U_2'') = {\mathcal R} (Y[\alpha_1 \cap \alpha_2])$, so we conclude that
${\mathcal R} (U_1'') = {\mathcal R} (U_2'')$.
We now claim that $t = r$, where $\bar U \in {\mathcal M\,}^{k \times (t+1)}$
satisfies equation~\eqref{eq:rangeU}.
Since $U_1'', U_2'' \in {\mathcal M\,}^{|\alpha_1 \cap \alpha_2| \times (r+1)}$ have full column rank
and ${\mathcal R} (U_1'') = {\mathcal R} (U_2'')$, we have by
Lemma~\ref{lem:twocliquered} that
${\mathcal R} (\bar U) = {\mathcal R} (\bar U_1) = {\mathcal R} (\bar U_2)$, where
\[
\bar U_1 :=
\begin{bmatrix}
U_1^\prime\cr
U_1^{\prime\prime}\cr
U_2^{\prime} (U_2^{\prime\prime})^\dagger U_1^{\prime\prime}
\end{bmatrix}
\quad \mbox{and} \quad
\bar U_2 :=
\begin{bmatrix}
U_1^{\prime} (U_1^{\prime\prime})^\dagger U_2^{\prime\prime}\cr
U_2^{\prime\prime}\cr
U_2^\prime
\end{bmatrix}.
\]
Therefore,
\[
t+1 = \dim {\mathcal R} (\bar U) = \dim {\mathcal R} (\bar U_1) = \dim {\mathcal R} (\bar U_2) = r+1,
\]
so we have $t = r$, as claimed.
Recall, $Y = \KK^\dagger(D[\alpha_1 \cup \alpha_2])$, so
$Y \in \cap_{i=1,2} \KK^\dagger\left({{\mathcal E}^k} (\alpha_i,\bar D_i)\right)$.
Thus, Theorem~\ref{thm:interclique} implies that there exists
$\bar Z \in {\mathcal S} ^{r}_+$ such that
$Y = (\bar U \bar V)\bar Z(\bar U \bar V)^T$.
Observe that $\KK(Y[\beta]) = D[\beta] = \bar D$.
Thus,
\[
\KK\left(
(\bar U_\beta \bar V)
\bar Z
(\bar U_\beta \bar V)^T
\right) = \bar D,
\]
implying that
\[
\KK^\dagger \KK\left(
(\bar U_\beta \bar V)
\bar Z
(\bar U_\beta \bar V)^T
\right) = B.
\]
Since $\KK^\dagger \KK$ is the projection onto ${\mathcal R} (\KK^*) = {\mathcal S}_C $,
we have that $\KK^\dagger \KK( \cdot ) = J(\cdot)J$.
Therefore, we have that $\bar Z$ satisfies equation~\eqref{eq:Zbeta}.
It remains to show that equation~\eqref{eq:Zbeta} has a unique solution.
Let $A := J \bar U_\beta \bar V \in {\mathcal M\,}^{|\beta| \times r}$.
Then $A \bar Z A^T = B$ and ${\rm rank\,}(B) = r$ implies that ${\rm rank\,}(A) \geq r$,
so $A$ has full column rank.
This implies that equation~\eqref{eq:Zbeta} has a unique solution,
and that $\bar Z = A^\dagger B (A^\dagger)^T = Z$.
Finally, since $Y = (\bar U \bar V)Z(\bar U \bar V)^T$ and
$D[\alpha_1 \cup \alpha_2] = \KK(Y)$,
we get equation~\eqref{eq:DKKUVZ}.
\end{proof}
The following result shows that if we know the minimal face of ${\mathcal S} ^n_+$ containing
$\KK^\dagger(D)$, and we know a small submatrix of $D$, then we can compute
a set of points in $\mathbb{R}^r$ that generate $D$ by solving a small equation.
\begin{cor}
\label{cor:finalUbar}
Let $D \in {{\mathcal E}^n} $ with embedding dimension $r$, and let $\beta \subseteq 1\!:\!n$.
Let $U \in {\mathcal M\,}^{n \times (r+1)}$ satisfy
\[
\face \KK^\dagger\left(D\right)
=
\left( U {\mathcal S} _+^{r+1} U^T \right) \cap {\mathcal S}_C ,
\]
let $U_\beta := U[\beta,:]$,
and let $\begin{bmatrix} V & \frac{U^Te}{\|U^Te\|} \end{bmatrix}
\in {\mathcal M\,}^{r+1}$ be orthogonal.
If $D[\beta]$ has embedding dimension $r$, then
\[
(J U_\beta V) Z (J U_\beta V)^T = \KK^\dagger(D[\beta])
\]
has a unique solution $Z \in {\mathcal S} ^r_{++}$, and $D = \KK(PP^T)$,
where $P := UVZ^{1/2} \in \mathbb{R}^{n \times r}$.
\end{cor}
\begin{proof}
Apply Theorem~\ref{thm:twocliquered} with $\alpha_1 = \alpha_2 = 1\!:\!n$.
\end{proof}
\begin{remark}
\label{rem:effZcalc}
A more efficient way to calculate $Z$ uses the full rank factorization
\[
B=QD^{1/2}\left(QD^{1/2}\right)^T, \quad Q^TQ=I_r, \quad D\in {\mathcal S} _{++}^r.
\]
Let $C = (J \bar U_\beta \bar V)^\dagger \left(QD^{1/2}\right)$. Then
$Z$ in \eqref{eq:Zcalc1} can be found from $Z=CC^T$.
Note that our algorithm postpones finding $Z$ until the end where we
can no longer perform any clique reductions. At each iteration, we
compute the matrix $\bar U$ that represents the face corresponding to
the union of two cliques; $\bar U$ is chosen from
one of $\bar U_i$, for $i=1,2$ in \eqref{eq:U1U2}. Moreover, for
stability, we maintain $\bar U^T \bar U =I$, $\bar Ue_{r+1}=\alpha e$.
For many of our test problems, we can repeatedly apply Theorem~\ref{thm:twocliquered}
until there is only one clique left. Since each
repetition reduces the number of cliques by one, this means that there
are at most $n$ such steps.
\end{remark}
\subsubsection{Singular Facial Reduction with Intersection Embedding Dimension $r-1$}
\label{sect:nonrigidreduction}
\index{singular reduction}
\begin{figure}[htb]
\epsfxsize=160pt
\centerline{\epsfbox{deg_clique_intersect.eps}}
\caption{Two clique reduction with intersection having embedding
dimension $<r$}
\label{fig:deg2cliqred}
\end{figure}
We now show that if the embedding dimension of
the intersection is $r-1$ (i.e., deficient),
then we can find at most two completions.
If exactly one of these two completions is feasible in the sense that it satisfies the related distance equality constraints and, if included, the related lower bound inequality constraints obtained from the radio range $R$, then we have identified the unique completion; see Figure~\ref{fig:deg2cliqred}.
We first need the following extension of Lemma~\ref{lem:twocliquered}
on the intersection of two structured
subspaces for the case where the common middle blocks are not full rank.
\index{singular intersection}
\begin{lem}
\label{lem:twocliquereddeg}
Let $U_i, \hat U_i, \bar U_i$, for $i=1,2$, be defined and appropriately blocked
as in Lemma~\ref{lem:twocliquered},
with $U^{\prime\prime}_i \in {\mathcal M\,}^{k \times (r+1)}$
having rank $r$, for $i=1,2$, and ${\mathcal R} (U_1'') = {\mathcal R} (U_2'')$.
Let $0\neq u_i \in {\mathcal N}(U_i^{\prime\prime})$, for $i=1,2$.
If $\bar U \in {\mathcal M\,}^{k \times (t+1)}$ satisfies
${\mathcal R} (\bar U) = {\mathcal R} (\hat U_1) \cap {\mathcal R} (\hat U_2)$, then $t=r+1$ and
\begin{equation}
\label{eq:rangeUsdeg}
\begin{array}{rcl}
{\mathcal R} (\bar U)
&=&
{\mathcal R} \left(\begin{bmatrix}
U_1^\prime& 0\cr
U_1^{\prime\prime}& 0\cr
U_2^{\prime} (U_2^{\prime\prime})^\dagger U_1^{\prime\prime} &
U_2^\prime u_2
\end{bmatrix} \right)
=
{\mathcal R} \left(
\begin{bmatrix}
\bar U_1
&
\begin{bmatrix}
0\cr
0\cr
U_2^\prime u_2
\end{bmatrix}
\end{bmatrix}
\right)
\vspace{.1in}
\\&=&
{\mathcal R} \left(\begin{bmatrix}
U_1^{\prime} (U_1^{\prime\prime})^\dagger U_2^{\prime\prime} &
U_1^{\prime} u_1\cr
U_2^{\prime\prime}& 0\cr
U_2^\prime & 0
\end{bmatrix} \right)
=
{\mathcal R} \left(
\begin{bmatrix}
\bar U_2
&
\begin{bmatrix}
U_1^\prime u_1 \cr
0\cr
0\cr
\end{bmatrix}
\end{bmatrix}
\right).
\end{array}
\end{equation}
Moreover, if $e_{r+1} \in \mathbb{R}^{r+1}$ is the $(r+1)^\mathrm{st}$ standard unit vector, and
$U_ie_{r+1} = \alpha_i e$, for some $\alpha_i \neq 0$, for $i=1,2$, then
$\bar U_ie_{r+1} = \alpha_i e$, for $i=1,2$.
\end{lem}
\begin{proof}
From the definitions, $x \in {\mathcal R} (\bar U)$ if and only if
\begin{equation}
\label{eq:xU12lemdeg}
x =
\begin{bmatrix}
x_1 \\
x_2 \\
x_3
\end{bmatrix}
=
\begin{bmatrix}
U^{\prime}_1 v_1 \\
U^{\prime\prime}_1 v_1 \\
v_2
\end{bmatrix}
=
\begin{bmatrix}
w_1 \\
U^{\prime\prime}_2 w_2 \\
U^{\prime}_2 w_2
\end{bmatrix},
\mbox{ for some }
v = \begin{bmatrix} v_1 \cr v_2 \end{bmatrix},
w = \begin{bmatrix} w_1 \cr w_2 \end{bmatrix}.
\end{equation}
Since ${\mathcal R} (U_1'') = {\mathcal R} (U_2'')$,
and $U''_i, i=1,2$, are both rank $r$, we conclude that
$x_2=U^{\prime\prime}_1 v_1 = U^{\prime\prime}_2 w_2$,
for some $v_1,w_2$ if and only if $x_2 \in {\mathcal R} ( U^{\prime\prime}_1)$, with
$v_1,w_2$ determined by
\[
v_1=(U''_1)^\dagger x_2 + \alpha_1u_1, \mbox{ for some } \alpha_1\in \mathbb{R},
\qquad w_2=(U''_2)^\dagger U''_1 v_1 + \alpha_2 u_2, \mbox{ for some }
\alpha_2 \in \mathbb{R}.
\]
In other words, we get
\begin{equation}
\label{eq:iffxvwdeg}
\begin{array}{c}
x_2=U^{\prime\prime}_1 v_1 = U^{\prime\prime}_2 w_2,
\mbox{ for some } v_1,w_2, \\
\mbox{ if and only if } \\
x_2=U_1''v_1, \mbox{ for some } v_1, \mbox{ with }
w_2=(U''_2)^\dagger U''_1 v_1+\alpha_2 u_2, \mbox{ for some } \alpha_2 \in \mathbb{R}.
\end{array}
\end{equation}
After substituting for $v_2$ with
$v_2=U_2' w_2 = U_2' \left( (U_2'')^\dagger U_1''v_1 +
\alpha_2 u_2\right)$,
we conclude that \eqref{eq:xU12lemdeg} holds if and only if
the first equality in \eqref{eq:rangeUsdeg} holds; i.e., \eqref{eq:xU12lemdeg} holds if and only if
\[
x =
\begin{bmatrix}
x_1 \\
x_2 \\
x_3
\end{bmatrix}
=
\begin{bmatrix}
U^{\prime}_1 v_1 \\
U^{\prime\prime}_1 v_1 \\
U_2^{\prime} (U_2^{\prime\prime})^\dagger U_1^{\prime\prime} v_1
+\alpha_2 U_2^{\prime} u_2
\end{bmatrix},
\mbox{ for some } v_1, \alpha_2,
\]
where
\[
v_2=U_2^\prime(U_2^{\prime\prime})^\dagger U_1^{\prime\prime} v_1
+\alpha_2 U_2^\prime u_2, \quad
w_1=U_1'v_1, \quad
w_2=(U_2^{\prime\prime})^\dagger U_1^{\prime\prime} v_1
+\alpha_2 u_2.
\]
The second equality in \eqref{eq:rangeUsdeg} follows similarly.
The last statements about $\bar U_ie_{r+1}$ follow as in the proof of
Lemma~\ref{lem:twocliquered}.
\end{proof}
In the rigid case in Theorem~\ref{thm:twocliquered},
we use the expression for $\bar U$
from Lemma~\ref{lem:twocliquered} to obtain a unique $Z$ in order to
get the completion of $D[\alpha_1 \cup \alpha_2]$. The $Z$ is unique because the
$r+1$ columns of $\bar U$ that represent the new clique
$\alpha_1\cup \alpha_2$ are linearly independent,
$e\in {\mathcal R} (\bar U)$, ${\rm rank\,}(B)=r$, and $Be=0$. This means that the solution
$C$ of $(J \bar U_\beta \bar V)C = QD^{1/2}$ in Remark~\ref{rem:effZcalc}
exists and is unique. (Recall that
$J \bar U_\beta \bar V$ is full column rank.)
This also means that the two
matrices, $U_1$ and $U_2$, that represent the cliques, $\alpha_1$ and $\alpha_2$,
respectively, can be replaced by the single matrix $\bar U$ without
actually calculating $C$; we can use $\bar U$ to represent
the clique $\alpha_1 \cup \alpha_2$ and complete all or part of the
partial \mbox{\boldmath$EDM$}\, $D[\alpha_1 \cup \alpha_2]$ only when needed.
We have a similar situation for the singular intersection case
following Lemma~\ref{lem:twocliquereddeg}.
We have the matrix $\bar U$ to represent the intersection of the two subspaces,
where each subspace represents one of the cliques, $\alpha_1$ and $\alpha_2$.
However, this is not
equivalent to {\em uniquely} representing the union of the
two cliques, $\alpha_1$ and $\alpha_2$,
since there is an extra column in $\bar U$ compared to the
nonsingular case.
In addition, since ${\rm rank\,}(B)=r-1$, then $J \bar U_\beta \bar V$ is not
necessarily full column rank.
Therefore, there may be infinite solutions for $C$ in
Remark~\ref{rem:effZcalc}; any
$C \in (J \bar U_\beta \bar V)^\dagger \left(QD^{1/2}\right) +
{\mathcal N}(J \bar U_\beta \bar V)$ will give us a solution.
Moreover, these solutions will not
necessarily satisfy $\KK\left((\bar U C) (\bar UC)^T\right) = D[\alpha_1 \cup \alpha_2]$.
We now see that we can continue and use the
$\bar U$ to represent a set of cliques rather than just $\alpha_1 \cup \alpha_2$.
Alternatively, we can use other relevant distance equality constraints or lower bound constraints from the radio range $R$ to determine the correct $C$ in order to get the
correct number of columns for $\bar U$; we can then get the correct
completion of $D[\alpha_1 \cup \alpha_2]$ if exactly one of the two possible completions with embedding dimension $r$ is feasible.
\begin{thm}
\label{thm:degcompl}
Let the hypotheses of Theorem~\ref{thm:twocliquered} hold with the
special case that
$U_i^TU_i=I$, $U_ie_{r+1}=\alpha_ie$, for $i=1,2$.
In addition, let
$\bar U$ be defined by one of the expressions in
\eqref{eq:rangeUsdeg} in Lemma~\ref{lem:twocliquereddeg}.
For $i=1,2$, let $\beta \subset \delta_i \subseteq \alpha_i$ and
$A_i := J \bar U_{\delta_i} \bar V$, where
$\bar U_{\delta_i} := \bar U(\delta_i,:)$. Furthermore,
let $B_i := \KK^\dagger(D[\delta_i])$, define the linear system
\begin{equation}
\label{eq:A1A2}
\begin{array}{rcl}
A_1 Z A_1^T & = & B_1 \\
A_2 Z A_2^T & = & B_2,
\end{array}
\end{equation}
and let $\bar Z \in {\mathcal S} ^t$ be a
particular solution of this system \eqref{eq:A1A2}.
If the embedding dimensions of $D[\delta_1]$ and $D[\delta_2]$
are both $r$, but the embedding dimension of $\bar D := D[\beta]$ is $r-1$,
then the following holds.
\begin{enumerate}
\item
\label{item:cornulls}
$\dim {\mathcal N}(A_i) = 1$, for $i=1,2$.
\item
\label{item:corsolsZ}
For $i=1,2$, let $n_i \in {\mathcal N}(A_i)$, $\|n_i\|_2 = 1$,
and $\Delta\!Z := n_1 n_2^T + n_2 n_1^T$. Then,
$Z$ is a solution of the linear system~\eqref{eq:A1A2} if and only if
\begin{equation}
\label{eq:solnZ}
Z = \bar Z + \tau \Delta\!Z,
\quad \mbox{for some $\tau \in \mathbb{R}$}.
\end{equation}
\item
\label{item:corcompletion}
There are at most two nonzero solutions, $\tau_1$ and $\tau_2$,
for the generalized eigenvalue
problem $-\Delta\!Z v = \tau \bar Z v$, $v \neq 0$.
Set $Z_i := \bar Z + \frac{1}{\tau_i} \Delta\!Z$, for $i=1,2$. Then
\[
D[\alpha_1 \cup \alpha_2] \in \left\{
\KK(\bar U \bar V Z_i \bar V^T \bar U^T) : i=1,2 \right\}.
\]
\end{enumerate}
\end{thm}
\begin{proof}
We follow a similar proof as in the nonsingular case.
For simplicity, we assume that $\delta_i=\alpha_i$, for $i=1,2$
(choosing smaller $\delta_i$ can reduce the cost of solving the linear
systems).
That a particular solution $\bar Z$
exists for the system \eqref{eq:A1A2}, follows
from the fact that $\bar U$ provides a representation for the
intersection of the two faces (or the union of the two cliques).
Since the embedding dimension of $\bar D$ is $r-1$, we have ${\rm rank\,}(B) = r-1$.
Furthermore, we have $Be = 0$ and $B \in {\mathcal S} ^{|\beta|}_+$, implying that
$|\beta| \geq r$. Without loss of generality, and for simplicity, we assume
that $|\beta| = r$. Therefore, there exists $0\neq u_i \in
{\mathcal N}(U_i^{\prime\prime})$, for $i=1,2$. From Lemma~\ref{lem:twocliquereddeg},
we can assume that we maintain $\bar U_i^T \bar U_i=I$, $\bar U_ie_{r+1}=\alpha_i
e$, for some $\alpha_i \neq 0$, for $i=1,2$. Therefore, the action of $\bar V$ is equivalent
to removing the $r+1$ column of $\bar U_i$. We can then explicitly use
$u_i$ to write down $n_i \in {\mathcal N}(A_i)$.
By construction, we now have $A_i(n_1 n_2^T + n_2 n_1^T)A_i^T=0$, for $i=1,2$.
From the first expression for $\bar U$
in \eqref{eq:rangeUsdeg}, we see that the choices for $n_1$ and $n_2$ in
Part~\ref{item:cornulls} are in the appropriate nullspaces. The
dimensions follow from the assumptions on the embedding dimensions.
Part~\ref{item:corsolsZ} now follows from the definition of the general
solution of a linear system of equations; i.e., the sum of a particular
solution with any solution of the homogeneous equation.
Part~\ref{item:corcompletion} now follows from the role that $\bar U$
plays as a representation for the union of the two cliques.
\end{proof}
\begin{remark}
\label{rem:effZcalcdeg}
As above in the nonsingular case,
a more efficient way to calculate $\bar Z$ uses the full rank factorization
\[
B_i=QD^{1/2}\left(Q_iD_i^{1/2}\right)^T, \quad
Q_i^TQ_i=I_r, \quad
D_i\in {\mathcal S} _{++}^{r}, \quad
i=1,2.
\]
(We have assumed that both have embedding dimension $r$, though we
only need that one does.)
We solve the equations $A_i C = \left(Q_iD_i^{1/2}\right)\bar Q_i$,
$\bar Q_i\bar Q_i^T=I$, for $i=1,2$,
for the unknowns $C$, and $\bar Q_i$, for $i=1,2$.
Then a particular solution
$\bar Z$ in \eqref{eq:A1A2} can be found from $\bar Z=CC^T$.
Note that the additional orthogonal matrices $\bar Q_i$, for $i=1,2$ are needed
since, they still allow $A_iC(A_iC)^T=B_i$, for $i=1,2$.
Also, without loss of generality, we can
assume $\bar Q_1=I$.
\end{remark}
\subsection{Clique Initialization and Node Absorption}
Using the above clique reductions,
we now consider techniques that allow one clique to grow/absorb other
cliques. This applies Theorem~\ref{thm:twocliquered}.
We first consider an elementary and fast technique to find some of the
existing cliques.
\begin{lem}
\label{lem:gencliques}
For each $i \in \{1,\ldots,n\}$, use half the radio range
and define the set
\[
C_i := \left\{ j \in \{1,\ldots,n\} : D_{ij} \leq (R/2)^2 \right\}.
\]
Then each $C_i$ corresponds to a clique of sensors that are
within radio range of each other.
\index{half radio range clique centered at node $i$, $C_i$}
\end{lem}
\begin{proof}
Let $j,k \in C_i$ for a given $i \in \{1,\ldots,n\}$.
An elementary application of the triangle inequality shows that
$\sqrt{(D_{jk})} \leq \sqrt{(D_{ji})} + \sqrt{(D_{ki})} \leq R$.
\end{proof}
We can now assume that we have a finite set of indices
${\mathcal C} \subseteq \mathbb{Z}_{+}$ corresponding to a family of cliques,
\index{clique index set, ${\mathcal C} $}
$\{C_i\}_{i\in {\mathcal C} }$. We can combine cliques using the reductions given in
Theorems~\ref{thm:twocliquered} and \ref{thm:degcompl}.
We now see how a clique can grow further by absorbing individual sensors;
see Figure \ref{fig:cliqabsorb}.
\begin{figure}[htb]
\epsfxsize=160pt
\centerline{\epsfbox{node_absorb.eps}}
\caption{Absorption with intersection having embedding
dimension $r$}
\label{fig:cliqabsorb}
\end{figure}
\begin{cor}
\label{cor:absorbsingle}
Let $C_k$, for $k\in {\mathcal C} $, be a given clique with node $l \notin C_k$,
$\beta := \left\{j_1,\ldots,j_{r+1}\right\} \subseteq C_k$,
such that the
distances $D_{lj_i}$, for $i=1,\ldots,r+1$ are known.
If
\begin{equation}
\label{eq:rankabsorb}
{\rm rank\,} \KK^\dagger (D[\beta]) = r,
\end{equation}
then $l$ can be absorbed by the clique $C_k$ and we can complete
the missing elements in column (row) $l$ of $D[C_k\cup \{l\}]$.
\end{cor}
\begin{proof}
Let $\alpha_1:=C_k$, $\alpha_2:=\{j_1,\ldots,j_{r+1},l\}$, and
$\beta := \alpha_1 \cap \alpha_2 = \{j_1,\ldots,j_{r+1}\}$. Then the
conditions in Theorem~\ref{thm:twocliquered} are satisfied and we
can recover all the missing elements in $D[C_k\cup \{l\}]$.
\end{proof}
\subsubsection{Node Absorption with Degenerate Intersection}
We can apply the same reasoning as for the clique reduction in the
nonsingular case, except now we apply Theorem~\ref{thm:degcompl}.
To obtain a unique completion, we test the feasibility of the two possible completions against any related distance equality constraints or, if included, any related lower bound inequality constraints.
See Figure~\ref{fig:degnodeabsorg}.
\begin{figure}[htb]
\epsfxsize=160pt
\centerline{\epsfbox{deg_node_absorb.eps}}
\caption{Degenerate absorption with intersection with embedding
dimension $<r$}
\label{fig:degnodeabsorg}
\end{figure}
\begin{cor}
\label{cor:degabsorbsingle}
Let $C_k$, for $k\in {\mathcal C} $, be a given clique with node $l \notin C_k$,
$\beta := \left\{j_1,\ldots j_{r}\right\} \subseteq C_k$
such that the distances $D_{lj_i}$, for $i=1,\ldots,r$ are known. If
\begin{equation}
\label{eq:rankabsorbdeg}
{\rm rank\,} \KK^\dagger (D[\beta]) = r-1,
\end{equation}
then we can determine two possible completions of the distances. If exactly one of these two completions is feasible, then $l$ can be absorbed by the clique $C_k$. We can also complete
the missing elements in column (row) $l$ of $D[C_k\cup \{l\}]$.
\end{cor}
\begin{proof}
Let $\alpha_1:=C_k$, $\alpha_2:=\{j_1,\ldots,j_{r},l\}$, and
$\beta := \alpha_1 \cap \alpha_2 = \{j_1,\ldots ,j_{r}\}$. Then the
conditions in Theorem~\ref{thm:degcompl} are satisfied and we
can recover all the missing elements in $D[C_k\cup \{l\}]$.
\end{proof}
\section{\texttt{SNLSDPclique} Facial Reduction Algorithm and Numerical Results}
\label{sect:algor}
Our \texttt{SNLSDPclique} algorithm starts by forming a clique $C_i$ around each
sensor $i$.
If and when we use this clique, we find a subspace representation
from the $r$ eigenvectors corresponding to the $r$ nonzero
eigenvalues of $B=\KK^\dagger(D[C_i])$.
The algorithm then grows and combines cliques using
Theorem~\ref{thm:twocliquered}, Theorem~\ref{thm:degcompl},
Corollary~\ref{cor:absorbsingle}, and Corollary~\ref{cor:degabsorbsingle}.
In particular, we do not complete the \mbox{\boldmath$EDM$}\, each time we combine or grow
cliques; i.e., we do not evaluate the missing distances. Instead, we use
the subspace representations of the corresponding faces of the \mbox{\boldmath$SDP$}\, cone
and then find the intersection of the subspaces that represent the faces.
This yields a subspace representation of the new smaller
face representing the union of two cliques. This is based on
Lemma~\ref{lem:twocliquered} and Lemma~\ref{lem:twocliquereddeg} and is therefore
inexpensive.
Once we cannot, or need not, grow cliques, we complete the distances
using Corollary~\ref{cor:finalUbar}. This is also inexpensive.
Finally, we rotate and translate the anchors to their original positions using
the approach outlined in \cite{DiKrQiWo:06}. We have provided an outline of our facial reduction algorithm \texttt{SNLSDPclique} in Algorithm~\ref{alg:SNLSDPclique}.
\begin{algorithm}
\caption{\texttt{SNLSDPclique} -- a facial reduction algorithm}
\label{alg:SNLSDPclique}
\SetAlgoLined
\LinesNumbered
\SetKwFunction{Face}{Face}
\SetKwFunction{RigidCliqueUnion}{RigidCliqueUnion}
\SetKwFunction{RigidNodeAbsorption}{RigidNodeAbsorption}
\SetKwFunction{NonRigidCliqueUnion}{NonRigidCliqueUnion}
\SetKwFunction{NonRigidNodeAbsorption}{NonRigidNodeAbsorption}
\SetKwInOut{Input}{input}
\SetKwInOut{Output}{output}
\Input{Partial $n \times n$ Euclidean Distance Matrix $D_p$ and anchors $A \in \mathbb{R}^{m \times r}$\;}
\Output{$X \in \mathbb{R}^{|C_i| \times r}$, where $C_i$ is the largest final clique that contains the anchors\;}
\BlankLine
Let $\mathcal{C} := \{1,\ldots,n+1\}$\;
Let $\{C_i\}_{i \in \mathcal{C}}$ be a family of cliques satisfying $i \in C_i$ for all $i = 1,\ldots,n$%
\tcc*{For example, by Lemma~\ref{lem:gencliques}, we could choose $C_i := \left\{ j : (D_p)_{ij} < (R/2)^2 \right\}$, for $i=1,\ldots,n.$ Alternatively, we could simply choose $C_i := \{i\}$, for $i=1,\ldots,n$.}
Let $C_{n+1} := \{n-m+1,\ldots,n\}$%
\tcc*{$C_{n+1}$ is the clique of anchors}
\BlankLine
\tcc{GrowCliques}
Choose $\mbox{\sc{MaxCliqueSize}} > r+1$%
\tcc*{For example, $\mbox{\sc{MaxCliqueSize}} := 3(r+1)$}
\For{$i \in \mathcal{C}$}{
\While{($|C_i| < \mbox{\sc{MaxCliqueSize}}$)
{\bf and} ($\exists$ a node $j$ adjacent to all nodes in $C_i$)}{
$C_i := C_i \cup \{j\}$\;
}
}
\BlankLine
\tcc{ComputeFaces}
\For{$i \in \mathcal{C}$}{
Compute $U_{B_i} \in \mathbb{R}^{|C_i| \times (r+1)}$ to represent face for clique $C_i$%
\tcc*{see Theorem~\ref{thm:onecliquered}}
\tcc{Alternatively, wait to compute $U_{B_i}$ when first needed. This can be more efficient since $U_{B_i}$ is not needed for every clique.}
}
\BlankLine
\Repeat{not possible to decrease $|\mathcal{C}|$ or increase $|C_i|$ for some $i \in \mathcal{C}$}{
\uIf{$|C_i \cap C_j| \geq r+1$, for some $i,j \in \mathcal{C}$}{
\RigidCliqueUnion{$C_i$,$C_j$}%
\tcc*{see Algorithm~\ref{alg:RigidCliqueUnion}}
}
\uElseIf{$|C_i \cap {\mathcal N}(j)| \geq r+1$, for some $i \in \mathcal{C}$ and node $j$}{
\RigidNodeAbsorption{$C_i$,$j$}%
\tcc*{see Algorithm~\ref{alg:RigidNodeAbsorption}}
}
\uElseIf{$|C_i \cap C_j|= r$, for some $i,j \in \mathcal{C}$}{
\NonRigidCliqueUnion{$C_i$,$C_j$}%
\tcc*{see Algorithm~\ref{alg:NonRigidCliqueUnion}}
}
\ElseIf{$|C_i \cap {\mathcal N}(j)| = r$, for some $i \in \mathcal{C}$ and node $j$}{
\NonRigidNodeAbsorption{$C_i$,$j$}%
\tcc*{see Algorithm~\ref{alg:NonRigidNodeAbsorption}}
}
}
\BlankLine
Let $C_i$ be the largest clique that contains the anchors\;
\uIf{clique $C_i$ contains some sensors}{
Compute a point representation $P \in \mathbb{R}^{|C_i| \times r}$ for the clique $C_i$%
\tcc*{see Cor.~\ref{cor:finalUbar}}
Compute positions of sensors $X \in \mathbb{R}^{(|C_i|-m)\times r}$ in clique $C_i$ by rotating $P$ to align with anchor positions $A \in \mathbb{R}^{m \times r}$%
\tcc*{see Ding et al.~\cite[Method~3.2]{DiKrQiWo:06}}
\Return $X$\;
}
\Else{
\Return $X := \emptyset$\;
}
\end{algorithm}
\begin{algorithm}
\caption{\texttt{RigidCliqueUnion}}
\label{alg:RigidCliqueUnion}
\SetAlgoLined
\LinesNumbered
\SetKwInOut{Input}{input}
\SetKwInOut{Output}{output}
\SetKwInOut{Empty}{}
\Input{Cliques $C_i$ and $C_j$ such that $|C_i \cap C_j| \geq r+1$\;}
\BlankLine
Load $U_{B_i} \in \mathbb{R}^{|C_i|\times(r+1)}$ and $U_{B_j} \in \mathbb{R}^{|C_j|\times(r+1)}$ representing the faces corresponding to the cliques $C_i$ and $C_j$, respectively\;
Compute $\bar U \in \mathbb{R}^{|C_i \cup C_j| \times (r+1)}$ using one of the two formulas in equation~\eqref{eq:U1U2} from Lemma~\ref{lem:twocliquered}, where $U_1 = U_{B_i}$, $U_2 = U_{B_j}$, and $k = |C_i \cap C_j|$%
\tcc*{see Theorem~\ref{thm:interclique}}
Update $C_i := C_i \cup C_j$\;
Update $U_{B_i} := \bar U$\;
Update $\mathcal{C} := \mathcal{C} \setminus\{j\}$\;
\end{algorithm}
\begin{algorithm}
\caption{\texttt{RigidNodeAbsorption}}
\label{alg:RigidNodeAbsorption}
\SetAlgoLined
\LinesNumbered
\SetKwInOut{Input}{input}
\SetKwInOut{Output}{output}
\SetKwInOut{Empty}{}
\Input{Clique $C_i$ and node $j$ such that $|C_i \cap {\mathcal N}(j)| \geq r+1$\;}
\BlankLine
Load $U_{B_i} \in \mathbb{R}^{|C_i|\times(r+1)}$ representing the face corresponding to clique $C_i$\;
\If{$C_i \cap {\mathcal N}(j)$ not a clique in the original graph}{
Use $U_{B_i}$ to compute a point representation $P_i \in \mathbb{R}^{|C_i| \times r}$ of the sensors in $C_i$\;
\tcc*[f]{see Cor.~\ref{cor:finalUbar}} \\
Use $P_i$ to compute the distances between the sensors in $C_i \cap {\mathcal N}(j)$\;
}
Use the distances between the sensors in $(C_i \cap {\mathcal N}(j))\cup\{j\}$ to compute the matrix $U_{B_j} \in \mathbb{R}^{(|C_i \cap {\mathcal N}(j)|+1)\times(r+1)}$ representing the face corresponding to the clique $(C_i \cap {\mathcal N}(j))\cup\{j\}$%
\tcc*{see Theorem~\ref{thm:onecliquered}}
Compute $\bar U \in \mathbb{R}^{(|C_i|+1) \times (r+1)}$ using one of the two formulas in equation~\eqref{eq:U1U2} from Lemma~\ref{lem:twocliquered}, where $U_1 = U_{B_i}$, $U_2 = U_{B_j}$, and $k = |C_i \cap {\mathcal N}(j)|$%
\tcc*{see Theorem~\ref{thm:interclique}}
Update $C_i := C_i \cup \{j\}$\;
Update $U_{B_i} := \bar U$\;
\end{algorithm}
\begin{algorithm}
\caption{\texttt{NonRigidCliqueUnion}}
\label{alg:NonRigidCliqueUnion}
\SetAlgoLined
\LinesNumbered
\SetKwInOut{Input}{input}
\SetKwInOut{Output}{output}
\SetKwInOut{Empty}{}
\Input{Cliques $C_i$ and $C_j$ such that $|C_i \cap C_j| = r$\;}
\BlankLine
Load $U_{B_i} \in \mathbb{R}^{|C_i|\times(r+1)}$ and $U_{B_j} \in \mathbb{R}^{|C_j|\times(r+1)}$ representing the faces corresponding to the cliques $C_i$ and $C_j$, respectively\;
Using $U_{B_i}$ and $U_{B_j}$, find the two point representations of the sensors in $C_i \cup C_j$\;
\tcc*[f]{see Theorem~\ref{thm:degcompl}} \\
\If{exactly one of these two point representations is feasible}{
Use the feasible point representation to compute $\bar U \in \mathbb{R}^{|C_i \cup C_j| \times (r+1)}$ representing the face corresponding to the clique $C_i \cup C_j$%
\tcc*{see Theorem~\ref{thm:onecliquered}}
Update $C_i := C_i \cup C_j$\;
Update $U_{B_i} := \bar U$\;
Update $\mathcal{C} := \mathcal{C} \setminus\{j\}$\;
}
\end{algorithm}
\begin{algorithm}
\caption{\texttt{NonRigidNodeAbsorption}}
\label{alg:NonRigidNodeAbsorption}
\SetAlgoLined
\LinesNumbered
\SetKwInOut{Input}{input}
\SetKwInOut{Output}{output}
\SetKwInOut{Empty}{}
\Input{Clique $C_i$ and node $j$ such that $|C_i \cap {\mathcal N}(j)| = r$\;}
\BlankLine
Load $U_{B_i} \in \mathbb{R}^{|C_i|\times(r+1)}$ representing the face corresponding to clique $C_i$\;
\If{$C_i \cap {\mathcal N}(j)$ not a clique in the original graph}{
Use $U_{B_i}$ to compute a point representation $P_i \in \mathbb{R}^{|C_i| \times r}$ of the sensors in $C_i$\;
\tcc*[f]{see Cor.~\ref{cor:finalUbar}} \\
Use $P_i$ to compute the distances between the sensors in $C_i \cap {\mathcal N}(j)$\;
}
Use the distances between the sensors in $(C_i \cap {\mathcal N}(j))\cup\{j\}$ to compute the matrix $U_{B_j} \in \mathbb{R}^{(|C_i \cap {\mathcal N}(j)|+1)\times(r+1)}$ representing the face corresponding to the clique $(C_i \cap {\mathcal N}(j))\cup\{j\}$%
\tcc*{see Theorem~\ref{thm:onecliquered}}
Using $U_{B_i}$ and $U_{B_j}$, find the two point representations of the sensors in $C_i \cup \{j\}$\;
\tcc*[f]{see Theorem~\ref{thm:degcompl}} \\
\If{exactly one of these two point representations is feasible}{
Use the feasible point representation to compute $\bar U \in \mathbb{R}^{|C_i \cup C_j| \times (r+1)}$ representing the face corresponding to the clique $C_i \cup \{j\}$%
\tcc*{see Theorem~\ref{thm:onecliquered}}
Update $C_i := C_i \cup \{j\}$\;
Update $U_{B_i} := \bar U$\;
}
\end{algorithm}
\subsection{Numerical Tests}
\label{sect:numerics}
Our tests are on problems with sensors and anchors randomly placed in the region $[0,1]^r$ by means of a uniform random distribution.
We vary the number of sensors from $2000$ to $10000$ in steps of
$2000$, and the radio range $R$ from $.07$ to $.04$ in steps of $-.01$.
We also include tests on very large problems with $20000$ to $100000$ sensors.
In our tests, we did not use the lower bound inequality constraints coming from the radio range; we only used the equality constraints coming from the partial Euclidean distance matrix.
Our tests were done using the 32-bit version of {\sc Matlab} R2009b on a laptop running Windows XP, with a 2.16 GHz Intel Core 2 Duo processor and with 2 GB of RAM. The source code used for running our tests has been released under a GNU General Public License, and has been made available from the authors' websites.
We in particular
emphasize the low CPU times and the high accuracy of the solutions we obtain.
Our algorithm compares well with the recent work in
\cite{pongtseng:09,WangZhengBoydYe:06}, where they use, for example,
$R = .06$ for $n = 1000, 2000$,
$R = .035$ for $n = 4000$,
$R = .02$ for $n = 10000$, and also
use $10$\% of the sensors as anchors and limit the degree for each
node in order to maintain a low sparsity for the graph.
Tables~\ref{table:RigidCliqueUnion}, \ref{table:RigidNodeAbsorb}, and \ref{table:NonRigidCliqueUnion} contain the results of our tests on noiseless problems. These tables contain the following information.
\begin{enumerate}
\item
{\bf \# sensors, $r$, \# anchors, and $R$:}
We use $m = (\# anchors)$, $n = (\# sensors) + (\# anchors)$, and $r$ to generate ten random instances of $p_1,\ldots,p_n \in \mathbb{R}^r$; the last $m$ points are taken to be the anchors. For each of these ten instances, and for each value of the radio range $R>0$, we generate the the $n \times n$ partial Euclidean distance matrix $D_p$ according to
\[
(D_p)_{ij} =
\begin{cases}
\|p_i-p_j\|^2, & \text{if $\|p_i-p_j\|<R$, or both $p_i$ and $p_j$ are anchors} \\
\text{unspecified}, & \text{otherwise}.
\end{cases}
\]
\item
{\bf \# Successful Instances:}
An instance was called \emph{successful} if at least some, if not all, of the sensors could be positioned. If, by the end of the algorithm, the largest clique containing the anchors did not contain any sensors, then none of the sensor positions could be determined, making such an instance unsuccessful.
\item
{\bf Average Degree:}
We have found that the average degree of the nodes of a graph is a good indicator of the percentage of sensors that can be positioned. In the results reported, we give the average of the average degree over all ten instances.
\item
{\bf \# Sensors Positioned:}
We give the average number of sensors that could be positioned over all ten instances. Note that below we indicate that the error measurements are computed only over the sensors that could be positioned.
\item
{\bf CPU Time:}
Indicates the average running time of {\texttt{SNLSDPclique}} over all ten instances. This time does not include the time to generate the random problems, but it does include all aspects of the Algorithm~\ref{alg:SNLSDPclique}, including the time for {\texttt{GrowCliques}} and {\texttt{ComputeFaces}} at the beginning of the algorithm.
\item
{\bf Max Error:}
The maximum distance between the positions of the sensors
found and the true positions of those sensors. This is defined as
\[
\mbox{Max Error} := \max_{\mbox{\tiny $i$ positioned} } \| p_i - p_i^{\mathrm true} \|_2.
\]
\item
{\bf RMSD:} The root-mean-square deviation of the positions of the sensors
found and the true positions of those sensors. This is defined as
\[
\mbox{RMSD} :=
\left(
\frac{1}{\mbox{\# positioned}}
\sum_{\mbox{\tiny $i$ positioned} } \| p_i - p_i^{\mathrm true} \|_2^2
\right)^\frac12.
\]
\end{enumerate}
We note that for each set of ten random instances, the Max Error and RMSD values reported are the average Max Error and average RMSD values over the successful instances only; this is due to the fact that an unsuccessful instance will have no computed sensor positions to compare with the true sensor positions.
We have three sets of tests on noiseless problems.
\begin{enumerate}
\item
In Table~\ref{table:RigidCliqueUnion} we report the results of using only the {\texttt{RigidCliqueUnion}} step (see Figure~\ref{fig:intersembedr}) to solve our random problems.
\begin{table}
\caption[\texttt{RigidCliqueUnion}]{Results of Algorithm~\ref{alg:SNLSDPclique}
on noiseless problems,
using step {\texttt{RigidCliqueUnion}}.
The values for Average Degree, \# Sensors Positioned, and CPU Time are averaged over ten random instances. The values for Max~Error and RMSD values are averaged over the successful instances.}
\label{table:RigidCliqueUnion}
\begin{center}
\begin{footnotesize}
\begin{tabular}{|c|c|c|c||c|c|c|c|c|c|}
\hline
& & & & \# Successful & Average & \# Sensors & & & \\
\# sensors & $r$ & \# anchors & $R$ & Instances & Degree & Positioned & CPU Time & Max Error & RMSD \\
\hline
2000 & 2 & 4 & .07 & 9/10 & 14.5 & 1632.3 & 1 s & 6e-13 & 2e-13 \\
2000 & 2 & 4 & .06 & 5/10 & 10.7 & 720.0 & 1 s & 1e-12 & 4e-13 \\
2000 & 2 & 4 & .05 & 0/10 & 7.5 & 0.0 & 1 s & - & - \\
2000 & 2 & 4 & .04 & 0/10 & 4.9 & 0.0 & 1 s & - & - \\
\hline
4000 & 2 & 4 & .07 & 10/10 & 29.0 & 3904.1 & 2 s & 2e-13 & 6e-14 \\
4000 & 2 & 4 & .06 & 10/10 & 21.5 & 3922.3 & 2 s & 6e-13 & 2e-13 \\
4000 & 2 & 4 & .05 & 10/10 & 15.1 & 3836.2 & 2 s & 4e-13 & 2e-13 \\
4000 & 2 & 4 & .04 & 1/10 & 9.7 & 237.8 & 2 s & 1e-13 & 4e-14 \\
\hline
6000 & 2 & 4 & .07 & 10/10 & 43.5 & 5966.9 & 4 s & 3e-13 & 8e-14 \\
6000 & 2 & 4 & .06 & 10/10 & 32.3 & 5964.4 & 4 s & 2e-13 & 7e-14 \\
6000 & 2 & 4 & .05 & 10/10 & 22.6 & 5894.8 & 3 s & 3e-13 & 1e-13 \\
6000 & 2 & 4 & .04 & 10/10 & 14.6 & 5776.9 & 3 s & 7e-13 & 2e-13 \\
\hline
8000 & 2 & 4 & .07 & 10/10 & 58.1 & 7969.8 & 6 s & 3e-13 & 8e-14 \\
8000 & 2 & 4 & .06 & 10/10 & 43.0 & 7980.9 & 6 s & 2e-13 & 8e-14 \\
8000 & 2 & 4 & .05 & 10/10 & 30.1 & 7953.1 & 5 s & 6e-13 & 2e-13 \\
8000 & 2 & 4 & .04 & 10/10 & 19.5 & 7891.0 & 5 s & 6e-13 & 2e-13 \\
\hline
10000 & 2 & 4 & .07 & 10/10 & 72.6 & 9974.6 & 9 s & 3e-13 & 7e-14 \\
10000 & 2 & 4 & .06 & 10/10 & 53.8 & 9969.1 & 8 s & 9e-13 & 1e-13 \\
10000 & 2 & 4 & .05 & 10/10 & 37.7 & 9925.4 & 7 s & 5e-13 & 2e-13 \\
10000 & 2 & 4 & .04 & 10/10 & 24.3 & 9907.2 & 7 s & 3e-13 & 1e-13 \\
\hline
\hline
20000 & 2 & 4 & .030 & 10/10 & 27.6 & 19853.3 & 17 s & 7e-13 & 2e-13 \\
40000 & 2 & 4 & .020 & 10/10 & 24.7 & 39725.2 & 50 s & 2e-12 & 6e-13 \\
60000 & 2 & 4 & .015 & 10/10 & 21.0 & 59461.1 & 1 m 52 s & 1e-11 & 8e-13 \\
80000 & 2 & 4 & .013 & 10/10 & 21.0 & 79314.1 & 3 m 24 s & 4e-12 & 1e-12 \\
100000 & 2 & 4 & .011 & 10/10 & 18.8 & 99174.4 & 5 m 42 s & 2e-10 & 9e-11 \\
\hline
\end{tabular}
\end{footnotesize}
\end{center}
\end{table}
\item
In Table~\ref{table:RigidNodeAbsorb} we report the results of increasing the level of our algorithm to use both the {\texttt{RigidCliqueUnion} and \texttt{RigidNodeAbsorb}} steps (see Figures~\ref{fig:intersembedr} and \ref{fig:cliqabsorb}) to solve the random problems.
We see that
the number of sensors localized has increased and
that there has been a small, almost insignificant, increase in the CPU time.
\begin{table}
\caption[{\texttt{RigidCliqueUnion}} and {\texttt{RigidNodeAbsorb}}]{Results of Algorithm~\ref{alg:SNLSDPclique}
on noiseless problems,
using steps {\texttt{RigidCliqueUnion}} and {\texttt{RigidNodeAbsorb}}.
The values for Average Degree, \# Sensors Positioned, and CPU Time are averaged over ten random instances. The values for Max~Error and RMSD values are averaged over the successful instances.}
\label{table:RigidNodeAbsorb}
\begin{center}
\begin{footnotesize}
\begin{tabular}{|c|c|c|c||c|c|c|c|c|c|}
\hline
& & & & \# Successful & Average & \# Sensors & & & \\
\# sensors & $r$ & \# anchors & $R$ & Instances & Degree & Positioned & CPU Time & Max Error & RMSD \\
\hline
2000 & 2 & 4 & .07 & 10/10 & 14.5 & 2000.0 & 1 s & 6e-13 & 2e-13 \\
2000 & 2 & 4 & .06 & 10/10 & 10.7 & 1999.9 & 1 s & 8e-13 & 3e-13 \\
2000 & 2 & 4 & .05 & 10/10 & 7.5 & 1996.7 & 1 s & 9e-13 & 2e-13 \\
2000 & 2 & 4 & .04 & 9/10 & 4.9 & 1273.8 & 3 s & 2e-11 & 4e-12 \\
\hline
4000 & 2 & 4 & .07 & 10/10 & 29.0 & 4000.0 & 2 s & 2e-13 & 6e-14 \\
4000 & 2 & 4 & .06 & 10/10 & 21.5 & 4000.0 & 2 s & 6e-13 & 2e-13 \\
4000 & 2 & 4 & .05 & 10/10 & 15.1 & 3999.9 & 2 s & 6e-13 & 3e-13 \\
4000 & 2 & 4 & .04 & 10/10 & 9.7 & 3998.2 & 2 s & 1e-12 & 5e-13 \\
\hline
6000 & 2 & 4 & .07 & 10/10 & 43.5 & 6000.0 & 4 s & 3e-13 & 8e-14 \\
6000 & 2 & 4 & .06 & 10/10 & 32.3 & 6000.0 & 4 s & 2e-13 & 7e-14 \\
6000 & 2 & 4 & .05 & 10/10 & 22.6 & 6000.0 & 3 s & 3e-13 & 1e-13 \\
6000 & 2 & 4 & .04 & 10/10 & 14.6 & 5999.4 & 3 s & 8e-13 & 3e-13 \\
\hline
8000 & 2 & 4 & .07 & 10/10 & 58.1 & 8000.0 & 6 s & 3e-13 & 7e-14 \\
8000 & 2 & 4 & .06 & 10/10 & 43.0 & 8000.0 & 5 s & 2e-13 & 8e-14 \\
8000 & 2 & 4 & .05 & 10/10 & 30.1 & 8000.0 & 5 s & 6e-13 & 2e-13 \\
8000 & 2 & 4 & .04 & 10/10 & 19.5 & 8000.0 & 4 s & 7e-13 & 2e-13 \\
\hline
10000 & 2 & 4 & .07 & 10/10 & 72.6 & 10000.0 & 9 s & 3e-13 & 7e-14 \\
10000 & 2 & 4 & .06 & 10/10 & 53.8 & 10000.0 & 8 s & 3e-13 & 1e-13 \\
10000 & 2 & 4 & .05 & 10/10 & 37.7 & 10000.0 & 7 s & 5e-13 & 2e-13 \\
10000 & 2 & 4 & .04 & 10/10 & 24.3 & 10000.0 & 6 s & 3e-13 & 1e-13 \\
\hline
\hline
20000 & 2 & 4 & .030 & 10/10 & 27.6 & 20000.0 & 17 s & 7e-13 & 2e-13 \\
40000 & 2 & 4 & .020 & 10/10 & 24.7 & 40000.0 & 51 s & 2e-12 & 6e-13 \\
60000 & 2 & 4 & .015 & 10/10 & 21.0 & 60000.0 & 1 m 53 s & 2e-12 & 7e-13 \\
80000 & 2 & 4 & .013 & 10/10 & 21.0 & 80000.0 & 3 m 21 s & 4e-12 & 1e-12 \\
100000 & 2 & 4 & .011 & 10/10 & 18.8 & 100000.0 & 5 m 46 s & 2e-10 & 9e-11 \\
\hline
\end{tabular}
\end{footnotesize}
\end{center}
\end{table}
\item
In Table~\ref{table:NonRigidCliqueUnion} we report the results of increasing the level of our algorithm to use steps \texttt{RigidCliqueUnion}, \texttt{RigidNodeAbsorb}, and \texttt{NonRigidCliqueUnion} (see Figures~\ref{fig:intersembedr}, \ref{fig:cliqabsorb}, and \ref{fig:deg2cliqred}) to solve the random problems, further increasing the class of problems that we can complete.
\end{enumerate}
Testing a version of our algorithm that uses all four steps is still ongoing. From the above results, we can see that our facial reduction technique works very well for solving many instances of the \mbox{\boldmath$SNL$}\, problem. We are confident that the results of our ongoing tests will continue to show that we are able to solve an even larger class of \mbox{\boldmath$SNL$}\, problems.
\begin{table}
\caption[{\texttt{RigidCliqueUnion}}, {\texttt{RigidNodeAbsorb}}, and {\texttt{NonRigidCliqueUnion}}]{Results of Algorithm~\ref{alg:SNLSDPclique}
on noiseless problems,
using steps {\texttt{RigidCliqueUnion}}, {\texttt{RigidNodeAbsorb}}, and {\texttt{NonRigidCliqueUnion}}.
The values for Average Degree, \# Sensors Positioned, and CPU Time are averaged over ten random instances. The values for Max~Error and RMSD values are averaged over the successful instances.
The results of the tests with more than $6000$ sensors remain the same as in Table~\ref{table:RigidNodeAbsorb}.}
\label{table:NonRigidCliqueUnion}
\begin{center}
\begin{footnotesize}
\begin{tabular}{|c|c|c|c||c|c|c|c|c|c|}
\hline
& & & & \# Successful & Average & \# Sensors & & & \\
\# sensors & $r$ & \# anchors & $R$ & Instances & Degree & Positioned & CPU Time & Max Error & RMSD \\
\hline
2000 & 2 & 4 & .07 & 10/10 & 14.5 & 2000.0 & 1 s & 6e-13 & 2e-13 \\
2000 & 2 & 4 & .06 & 10/10 & 10.7 & 1999.9 & 1 s & 8e-13 & 3e-13 \\
2000 & 2 & 4 & .05 & 10/10 & 7.5 & 1997.9 & 1 s & 9e-13 & 2e-13 \\
2000 & 2 & 4 & .04 & 10/10 & 4.9 & 1590.8 & 5 s & 2e-11 & 7e-12 \\
\hline
4000 & 2 & 4 & .07 & 10/10 & 29.0 & 4000.0 & 2 s & 2e-13 & 6e-14 \\
4000 & 2 & 4 & .06 & 10/10 & 21.5 & 4000.0 & 2 s & 6e-13 & 2e-13 \\
4000 & 2 & 4 & .05 & 10/10 & 15.1 & 3999.9 & 2 s & 6e-13 & 3e-13 \\
4000 & 2 & 4 & .04 & 10/10 & 9.7 & 3998.2 & 3 s & 1e-12 & 5e-13 \\
\hline
6000 & 2 & 4 & .07 & 10/10 & 43.5 & 6000.0 & 4 s & 3e-13 & 8e-14 \\
6000 & 2 & 4 & .06 & 10/10 & 32.3 & 6000.0 & 4 s & 2e-13 & 7e-14 \\
6000 & 2 & 4 & .05 & 10/10 & 22.6 & 6000.0 & 3 s & 3e-13 & 1e-13 \\
6000 & 2 & 4 & .04 & 10/10 & 14.6 & 5999.4 & 3 s & 8e-13 & 3e-13 \\
\hline
\end{tabular}
\end{footnotesize}
\end{center}
\end{table}
\subsection{Noisy Data and Higher Dimensional Problems}
\label{sect:noisy}
The above algorithm was derived based on the fact that the \mbox{\boldmath$SNL$}\, had exact data;
i.e., for a given clique $\alpha$ we had an exact correspondence between the
\mbox{\boldmath$EDM$}\, and the corresponding Gram matrix $B=\KK^\dagger (D[\alpha])$.
To extend this to the noisy case, we apply a naive, greedy approach. When
the Gram matrix $B$ is needed, then we use the best rank $r$ positive
semidefinite approximation to $B$ using the well-known Eckert-Young
result; see e.g., \cite[Cor.~2.3.3]{GoVan:79}.
\begin{lem}
Suppose that $B \in {\mathcal S^n}$ with spectral decomposition $B=\sum_{i=1}^n
\lambda_i u_iu_i^T, \lambda_1 \geq \ldots \geq \lambda_n$.
Then the best positive semidefinite approximation with at
most rank $r$ is $B_+= \sum_{i=1}^r (\lambda_i)_+ u_iu_i^T$, where
$(\lambda_i)_+ = \max \{0, \lambda_i\}$.
\qed
\end{lem}
We follow the multiplicative noise model in, e.g.,~\cite{BYIEEE:06,BY:04,KimKojimaWaki:09,pongtseng:09,tseng:07,WangZhengBoydYe:06};
i.e., the noisy (squared) distances $D_{ij}$ are given by
\[
D_{ij} = \left( \|p_i-p_j\|(1+\sigma\epsilon_{ij}) \right)^2,
\]
where $\sigma \geq 0$ is the noise factor and $\epsilon_{ij}$ is chosen
from the standard normal distribution ${\mathcal N\,}(0,1)$.
We include preliminary test results in Table~\ref{table:NoisyDim23} for problems with 0\%-1\% noise
with embedding dimension $r=2,3$. Note that we do not apply the noise to the distances between the anchors.
\begin{table}
\label{table:NoisyDim23}
\caption[Problems with noise and $r = 2,3$]{Results of Algorithm~\ref{alg:SNLSDPclique}
for problems with noise and $r = 2,3$, using {\texttt{RigidCliqueUnion}} and {\texttt{RigidNodeAbsorb}}. The values for Average Degree, \# Sensors Positioned, CPU Time, Max~Error and RMSD are averaged over ten random instances.}
\begin{center}
\begin{footnotesize}
\begin{tabular}{|c|c|c|c|c||c|c|c|c|c|}
\hline
& & & & & Average & \# Sensors & & & \\
$\sigma$ & \# sensors & $r$ & \# anchors & $R$ & Degree & Positioned & CPU Time & Max Error & RMSD \\
\hline
\hline
0 & 2000 & 2 & 4 & .08 & 18.8 & 2000.0 & 1 s & 1e-13 & 3e-14 \\
1e-6 & 2000 & 2 & 4 & .08 & 18.8 & 2000.0 & 1 s & 2e-04 & 4e-05 \\
1e-4 & 2000 & 2 & 4 & .08 & 18.8 & 2000.0 & 1 s & 2e-02 & 4e-03 \\
1e-2 & 2000 & 2 & 4 & .08 & 18.8 & 2000.0 & 1 s & 2e+01 & 3e+00 \\
\hline
0 & 6000 & 2 & 4 & .06 & 32.3 & 6000.0 & 4 s & 2e-13 & 7e-14 \\
1e-6 & 6000 & 2 & 4 & .06 & 32.3 & 6000.0 & 4 s & 8e-04 & 3e-04 \\
1e-4 & 6000 & 2 & 4 & .06 & 32.3 & 6000.0 & 4 s & 9e-02 & 3e-02 \\
1e-2 & 6000 & 2 & 4 & .06 & 32.3 & 6000.0 & 4 s & 2e+01 & 3e+00 \\
\hline
0 & 10000 & 2 & 4 & .04 & 24.3 & 10000.0 & 6 s & 3e-13 & 1e-13 \\
1e-6 & 10000 & 2 & 4 & .04 & 24.3 & 10000.0 & 6 s & 5e-04 & 2e-04 \\
1e-4 & 10000 & 2 & 4 & .04 & 24.3 & 10000.0 & 6 s & 5e-02 & 2e-02 \\
1e-2 & 10000 & 2 & 4 & .04 & 24.3 & 10000.0 & 7 s & 4e+02 & 1e+02 \\
\hline
\hline
0 & 2000 & 3 & 5 & .20 & 26.6 & 2000.0 & 1 s & 3e-13 & 8e-14 \\
1e-6 & 2000 & 3 & 5 & .20 & 26.6 & 2000.0 & 1 s & 7e-04 & 2e-04 \\
1e-4 & 2000 & 3 & 5 & .20 & 26.6 & 2000.0 & 1 s & 8e-02 & 2e-02 \\
1e-2 & 2000 & 3 & 5 & .20 & 26.6 & 2000.0 & 1 s & 2e+03 & 4e+02 \\
\hline
0 & 6000 & 3 & 5 & .15 & 35.6 & 6000.0 & 5 s & 3e-13 & 6e-14 \\
1e-6 & 6000 & 3 & 5 & .15 & 35.6 & 6000.0 & 5 s & 1e-03 & 2e-04 \\
1e-4 & 6000 & 3 & 5 & .15 & 35.6 & 6000.0 & 5 s & 1e-01 & 2e-02 \\
1e-2 & 6000 & 3 & 5 & .15 & 35.6 & 6000.0 & 6 s & 9e+01 & 9e+00 \\
\hline
0 & 10000 & 3 & 5 & .10 & 18.7 & 10000.0 & 9 s & 3e-12 & 2e-13 \\
1e-6 & 10000 & 3 & 5 & .10 & 18.7 & 10000.0 & 10 s & 4e-02 & 2e-03 \\
1e-4 & 10000 & 3 & 5 & .10 & 18.7 & 10000.0 & 10 s & 2e+00 & 8e-02 \\
1e-2 & 10000 & 3 & 5 & .10 & 18.7 & 10000.0 & 10 s & 4e+02 & 1e+01 \\
\hline
\end{tabular}
\end{footnotesize}
\end{center}
\end{table}
\section{Conclusion}
\label{sect:concl}
The \mbox{\boldmath$SDP$}\, relaxation of \mbox{\boldmath$SNL$}\, is highly (implicitly) degenerate, since
the feasible set of this \mbox{\boldmath$SDP$}\, is restricted to a low dimensional
face of the \mbox{\boldmath$SDP$}\, cone, resulting in the failure of the Slater constraint
qualification (strict feasibility).
We take advantage of this degeneracy by finding
explicit representations of intersections
of faces of the \mbox{\boldmath$SDP$}\, cone corresponding to unions of intersecting cliques.
In addition, from these representations we force further degeneracy in
order to find the minimal face that contains the optimal solution.
In many cases, we can efficiently compute the exact solution to the \mbox{\boldmath$SDP$}\, relaxation
\emph{without} using any \mbox{\boldmath$SDP$}\, solver.
In some cases it is not possible to reduce the problem down to a single clique. However, in these cases, the intersection of the remaining faces returned by {\texttt{SNLSDPclique}} will produce a face containing the feasible region of the original problem. This face can then be used to reduce the problem before passing the problem to an \mbox{\boldmath$SDP$}\, solver, where, for example, the trace of the semidefinite matrix can be maximized \cite{MR2398864} to try to keep the embedding dimension small. As an example, if the problem is composed of disjoint cliques, then Corollary~\ref{cor:disjcliques} can be used to significantly reduce the problem size. This reduction can transform a large intractable problem into a much smaller problem that can be solved efficiently via an \mbox{\boldmath$SDP$}\, solver.
\clearpage
\phantomsection
\addcontentsline{toc}{section}{References}
|
\section{Introduction}
The use of sub-picosecond laser pulses have enabled the observation of rapid processes in the time domain through the use of a pump-probe experiment. A typical experiment involves applying a ``pump'' pulse to a sample, for example a gas of atoms or molecules, to initiate a time dependent process. A second ``probe'' pulse is used after a controllable time delay to make a measurement of the state of the sample at this delay. Because the delay between the pulses may be set very precisely, a time dependent measurement may be made on the sample. Over the last decade, pump-probe experiments have been used for the observation of Rydberg electron wavepackets \cite{wolde_1988, noel_1996, yeazell_1990}, atomic spin-orbit precession \cite{christian_1996}, and molecular vibration \cite{khundkar_1990, baumert_1992, gruebele_1993}.
Recently, the observation of spin-orbit precession in rubidium atoms has been demonstrated \cite{zamith_2000}, and used as a substrate for the coherent control of the pump pulses \cite{chatel_2008}.
In this paper, the theoretical description of spin-orbit precession is extended to the photoionization of long range Rb$_2$ molecules. It is also shown that the same mechanism that enables the spin-orbit precession to be visible can provide a measurement of the separation of the atoms in the dimer.
Two results are obtained from the extension of the atomic orbital precession to molecular orbital precession. The molecular pump-probe signal contains a high frequency component and a low frequency component, in contrast to the atomic case, when only a high-frequency component is present, at the spin-orbit precession frequency. The high frequency component decays, as a result of the orbital precession rate being dependent on the internuclear separation. The low frequency component is due to a vibrational oscillation of the molecule.
Both components would be visible in a pump-probe ion signal. The reason is that the Hund's case (c) states change their admixtures of Hund's case (a) states as a function of internuclear separation, and the ionization process effectively measures the Hund's case (a) populations rather than the Hund's case (c) populations due to the contrasting timescales on which the ionization process and spin-orbit coupling operate. This gives an internuclear separation dependent measurement that may be used to observe coherent vibrational motion in the molecule, using a direct ionization scheme.
The calculations of the photoionization dynamics presented here are based on a simple model used to describe the rubidium atom. The model is optimized to recover the atomic energy levels and transition strengths of the rubidium atom. Ionization cross sections and the derived model of above threshold short pulse ionization are presented also.
Section \ref{model} details the atomic model, and gives the optimization process, section~\ref{cross_sections} give the ionization cross sections of various states as a function of wavelength, section~\ref{molecular_model} gives the molecular model, and sections~\ref{ftl} and \ref{cut_pulse} describe two experiments and their predicted outcomes that show the molecular orbital precession and the molecular vibration in the ion signal.
\section{Atomic model\label{model}}
The rubidium atom has a core containing 36 electrons and a single valence electron. An effective potential for the valence electron is employed with three parameters. These parameters are fit to several energy levels of the rubidium atom and to the transition dipole moments of several transitions. In calculating transition dipole moments, accuracy is improved by assuming that near the core, the electron is effectively shielded from any incident radiation, and a fourth parameter is used to describe the shielding effect.
The model Hamiltonian, in atomic units, is given by
\begin{eqnarray}
V(r) &=& -\frac{Z(r)}{r} - \left\lbrace1-e^{-\left(\frac{r}{\rho}\right)^2}\right\rbrace^2\frac{\alpha_D}{2r^4}.\\
Z(r) &=& 1+be^{-\left(\frac{r}{a}\right)^c}.\\
H(t) &=& V(r)+\frac{1}{2}\alpha^2\frac{1}{r}\frac{\partial V}{\partial r}\mathbf{l}\cdot\mathbf{s} + \frac{\nabla^2}{2} + D(\mathbf{r})\varepsilon(t). \label{H_e}\\
D(\mathbf{r}) &=& (1-e^{-\frac{r}{r_s}})\textbf{r}.
\end{eqnarray}
Here, the first term may be considered to be the Coulomb potential of an effective charge, $Z(r)$ of the core, and the second term is the core polarizability term. The second term is taken from reference \cite{szentpaly_1982}, which gives the parameters for the static electric core dipole polarizability, $\alpha_D$ and core size, $\rho$ the values \unit[8.67]{E$_ha_0^4$} and \unit[2.09]{$a_0$} respectively. The Hamiltonian contains a spin-orbit coupling term. $a$, $b$, $c$ and $r_s$ are the optimised parameters of the model. $r$ is the electron-nuclear separation. $\mathbf{l}$ is the electron orbital angular momentum operator, and $\mathbf{s}$ is the electron spin angular momentum operator. $\alpha$ is the fine-structure constant.
The dipole operator is defined as above in order to include a shielding effect. This can be explained physically by the fact that when the valence electron is in the core region of the atom, the core electrons will shift adiabatically in the presence of an electric field, thus shielding the valence electron from the electric field. It also contains any contribution from core rearrangement that might accompany a transition in the valence electron state that might change the transition strength.
The parameters given here were optimized to give a compromise between the energy levels and the transition strengths. It is possible to optimize the parameters to get more accurate energy levels, but at the expense of the transition dipole moments.
The values for $a$, $b$ and $c$ are chosen to minimize a fitness function, chosen to be
\begin{eqnarray}
\epsilon &=& \sum_j \left(\frac{E_j(\hbox{model})-E_j(\hbox{experimental})}{h\times\hbox{THz}}\right)^2 \nonumber\\
&&+ \sum_k 1000\log^2\left|\frac{\mu_k(\hbox{model})}{\mu_k(\hbox{experimental})}\right|,
\end{eqnarray}
where the sums are over just the energy levels and transitions selected for optimization. The fitness function contains a weighting between the energy level and the transition dipole accuracies. This value was chosen according to the particular requirements of the experiments presented here: in particular the accuracy of the energy levels had to be accurate to some hundreds of h$\times$GHz.
Values for the four parameters are given in table~\ref{table_params}, and the experimental observables that went into the optimization are given in table~\ref{table_opt}.
\begin{table}
\begin{center}
\begin{tabular}{ l c}
\hline
\hline
Parameter & Value\\
\hline
$a / a_0$&0.296760\\
$b$& 37.68652\\
$c$& 0.729772\\
$r_s / a_0$&1.790893\\
\hline
\end{tabular}
\caption{\label{table_params} Optimised parameters for the atomic model presented here.}
\end{center}
\end{table}
\begin{table}
\begin{tabular}{ l r r r l}
\hline
\hline
Observable & Exp.&Model&\\
& \cite{sansonetti_2006} &&\\
\hline
E(5s) &-1010.025&-1010.027&THz \\
E(6s) &-406.468&-406.276&THz \\
E(7s) &-221.228&-221.179&THz \\
E(8s) &-139.223&-139.206&THz \\
E(9s) &-95.687&-95.680&THz \\
E(5p$_{1/2}$) &-632.918&-632.932&THz \\
E(6p$_{1/2}$) &-299.065&-299.236&THz \\
E(7p$_{1/2}$) &-175.552&-175.663&THz \\
E(8p$_{1/2}$) &-115.596&-115.669&THz \\
E(9p$_{1/2}$) &-81.900& -81.947&THz \\
E(5p$_{3/2}$) &-625.795&-625.420&THz \\
E(6p$_{3/2}$) &-296.741&-296.862&THz \\
E(7p$_{3/2}$) &-174.500&-174.599&THz \\
E(8p$_{3/2}$) &-115.031&-115.101&THz \\
E(9p$_{3/2}$) &-81.562&-81.608&THz \\
E(4p$_{5/2}$) &-429.771&-431.418&THz \\
E(5p$_{5/2}$) &-239.454&-239.750&THz \\
E(6p$_{5/2}$) &-149.939&-150.004&THz \\
\hline
$\mu$ 5s$\longrightarrow$5P$_{1/2}$&2.99&3.016&$q_ea_0$ \\
$\mu$ 5S$\longrightarrow$6P$_{1/2}$&0.236&0.230&$q_ea_0$ \\
$\mu$ 5S$\longrightarrow$7P$_{1/2}$&0.0813&0.0762&$q_ea_0$ \\
$\mu$ 5S$\longrightarrow$8P$_{1/2}$&0.0407&0.0382&$q_ea_0$ \\
$\mu$ 5S$\longrightarrow$9P$_{1/2}$&0.0252&0.0234&$q_ea_0$ \\
$\mu$ 5S$\longrightarrow$5P$_{3/2}$&2.98&3.008&$q_ea_0$ \\
$\mu$ 5S$\longrightarrow$6P$_{3/2}$&0.255&0.265&$q_ea_0$ \\
$\mu$ 5S$\longrightarrow$7P$_{3/2}$&0.0950&0.0963&$q_ea_0$ \\
$\mu$ 5S$\longrightarrow$8P$_{3/2}$&0.0504&0.0517&$q_ea_0$ \\
$\mu$ 5S$\longrightarrow$9P$_{3/2}$&0.0326&0.0333&$q_ea_0$ \\
$\mu$ 5P$_{3/2}\longrightarrow$6d$_{5/2}$&0.677&0.631&$q_ea_0$ \\
\hline
$\Delta E_{SO}$&7.123&7.513&THz\\
\hline
\hline
\end{tabular}
\caption{\label{table_opt} The experimental observables that went into the optimizations. The energies of 18 levels were optimized, as were eleven transition dipole moments.}
\end{table}
The model is much simpler than multi-electron models which require configuration interaction calculations or self consistent methods to be applied, and allow calculations to be performed much more easily. When compared to other single-electron models, such as reference~\cite{marinescu_1994}, the model presented here recovers remarkably accurate transition strengths and ionization cross sections, as well as treating spin-orbit coupling accurately. These are essential for estimating the visibility of the spin-orbit precession in an ion signal, and so the model is well suited to the applications presented here.
The numerical approach to solving the Schr\"{o}dinger equation for Eqn.~\ref{H_e} is given in the Appendix.
\begin{figure}
\centering
\includegraphics[width=\columnwidth, angle=0]{level_diagram.eps}
\caption{\label{level_diagram}(Color online) The energy levels of the rubidium atom. The atoms in the experiments discussed here are initially in the ground state, 5s. A short pulse transfers them to 5p states. An ionizing pulse then transfers population to continuum states with either s or d symmetry. If both accessible 5p states are populated and the ionizing pulse is short enough, then interference effects cause the precession of the orbital angular momentum about the total angular momentum to be visible in the ion signal.}
\end{figure}
\section{Atomic photoionization cross section \label{cross_sections}}
Using this atomic model, the photon energy dependent photoionization cross section of the ground state is calculated for verification as described in the Appendix.
The continuous wave photoionization cross section as a function of photon energy for the ground state is shown in Fig.~\ref{5sPI}. The model agrees with experiment \cite{lowell_2002, suemitsu_1983} for the photoionization of the ground state, but overestimates cross sections for the photoionization cross section of the 5P state by around 20\% when compared to experiment \cite{dinneen_1992}. This could be for a variety of reasons. Firstly, the error bars on the experimental result are wide, and it is possible that the true value is quite close to the calculation presented here. Secondly, the photionization cross sections of the 5p states depend on the transition dipole moments from p orbitals to d orbitals. Due to the scarcity of experimental transition dipole moments from p to d orbitals, the model was only optimized to recover a single p to d transition. Therefore it is quite likely that although the s and h{p} electron wavefunctions are quite well recovered, the d orbital wavefunctions could be less accurate, and this affects the p orbital ionization cross sections.
\begin{figure}
\centering
\includegraphics[width=\columnwidth, angle=0]{ionization_cross_section_5s.eps}
\caption{\label{5sPI}The continuous-wave photoionization cross section as a function of wavelength for the ground state rubidium atom. The blue line is the model's predictions. The point (a) shows the result from reference~\cite{lowell_2002}, including error. The two results are in agreement. The points (b) shows the results from reference.~\cite{suemitsu_1983}. These are a relative magnitude measurement and so have been scaled appropriately.}
\end{figure}
For short-pulse photoionization, the ionization probability may be expressed as the magnitude of the final wavefunction when projected onto the scattering electronic states:
\begin{eqnarray}
P(\hbox{ion}) \simeq |\wp_{\hbox{ion}} U \psi_{\hbox{i}} |^2,
\end{eqnarray}
where $U$ is the propagator from before the start of the pulse to after the end of the pulse, $\wp_{\hbox{ion}}$ projects onto only ionised electronic states, and $\psi_{\hbox{i}}$ is the initial state. Equivalently, this can be expressed as the expectation value of an operator, $\hat{M}$.
The atom has six 5P excited states: two spin states multiplied by the three orbital angular momentum projection quantum numbers: -1, 0 and 1. The axis of projection will be chosen to be the electric field axis of the ionizing electric field. In the fine structure basis, the 5P$_{1/2}$ state has 2-fold degeneracy and the 5P$_{3/2}$ state has 4-fold degeneracy.
Any operator acting on this basis may be expressed as a six by six matrix, but the reversal symmetry ( $m_{\ell}\rightarrow-m_{\ell}$, $m_s\rightarrow-m_s$) means that the states with positive total angular momentum are never coupled to states with negative angular momentum, and so we can limit the discussion to the positive to positive matrix elements in the understanding that the negative to negative elements are the same. More generally, neither the electric field nor the spin-orbit coupling changes the total projected angular momentum of the atom, so the four possible values, -3/2, -1/2, +1/2 and +3/2 remain separate: Nothing in this model couple between values for this quantum number. The ionization operator will therefore be block diagonal with each block representing a different total angular momentum.
The three positive total angular momentum basis states are
\begin{eqnarray}
|j=1/2, m_j=1/2\rangle,\\
|j=3/2, m_j=1/2\rangle,\\
|j=3/2, m_j=3/2\rangle.
\end{eqnarray}
If an arbitrary 5P wavefunction $|\psi\rangle$ is represented by a vector $\mathbf{a}$:
\begin{eqnarray}
|\psi\rangle &=& a_1 |j=1/2, m_j=1/2\rangle \nonumber \\ \nonumber
&&+ a_2|j=3/2, m_j=1/2\rangle\\ \nonumber
&&+ a_3 |j=3/2, m_j=3/2\rangle, \nonumber
\end{eqnarray}
then the ionization operator, $\hat{M}$, whose expectation value gives the ionization probability is a three by three Hermitian matrix.
As an example, the ionization operator whose expectation value gives the ionization probability for a Gaussian probe pulse, centred at \unit[650]{THz}, with a full width at half maximum of \unit[18]{THz}, and with a total fluence of \unit[0.3]{Jm$^{-2}$} is given as
\begin{eqnarray}
\hat{M}= \left(\begin{array}{c c c}
0.997 & -0.223 & 0\\
-0.223 & 1.218 & 0\\
0 & 0 & 0.817\\
\end{array}\right)\times10^{-3}.
\end{eqnarray}
A basis change can be made to the $s,m_s, \ell,m_{\ell}$ basis, so that the 5P state is expressed as
\begin{eqnarray}
|\psi\rangle = b_1 |m_\ell=0, m_s=+1/2\rangle \nonumber \\ \nonumber
+ b_2|m_\ell=1, m_s=-1/2\rangle \\ \nonumber
+ b_3 |m_\ell=1, m_s=+1/2\rangle.
\end{eqnarray}
In this basis, the ionization matrix takes the form
\begin{eqnarray}
\hat{M}= \left(\begin{array}{c c c}
1.355 & 0.0297 & 0\\
0.0297 & 0.860 & 0\\
0 & 0 & 0.817\\
\end{array}\right)\times10^{-3}.
\end{eqnarray}
The small off-diagonal elements show that the ionization process effectively measures the populations in the $m_\ell$ basis. The physical reason for this is that the pulse is shorter than the spin-orbit interaction time, and so the electron spins may be neglected, leaving the differing $\ell,m_\ell$ quantum numbers as the elements that affect the ionization probability. The difference in ionization probability between the two $m_\ell=1$ states is due to the subtly different spatial wavefunction of the $j=3/2$ and $j=1/2$ states
The ionization matrix in this example, and in all the cases presented, was calculated by propagating the three initial states under the influence of the electric field and using their ionization probability and the inner product between the resulting ionized wavefunctions to infer matrix elements. Although the field used here is in the weak field regime, this method of parameterizing the effect of the field applies to any field.
The ionization process will be approximated in the molecular simulations as
\begin{eqnarray}
\hat{M}=&&|m_\ell = 0\rangle 1.36\cdot10^{-3}\langle m_\ell = 0 |\\
&+& |m_\ell = 1\rangle 0.84\cdot10^{-3}\langle m_\ell = 1 |.\label{M_ion}\\
\end{eqnarray}
The approximation is correct to within a few percent, which is also roughly the error in transition dipole moments of the atomic model.
\subsection{Ionization operator dependence}
The framework described above is used to find the ionization parameters as a function of pulse bandwidth and centre frequency. The two quantities that determine whether the ionization operator is selective for one orbital orientation are the angle between the operator's principle axis and the $|m_\ell=0\rangle$ axis, and the ratio of ionization probabilities for the $|m_\ell=1\rangle$ and $|m_\ell=0\rangle$ states. These two values are shown in Fig.~\ref{ionization_parameterised} as a function of probe wavelength and bandwidth. The ionization operator can be seen to vary slowly with laser parameters over these intervals, becoming more orbital-alignment selective at smaller probe bandwidths.
\begin{figure}
\begin{center}
\includegraphics[width=\columnwidth, angle=0]{ionization_parameter_plot.eps}
\caption{\label{ionization_parameterised}The ionization operator is a function of the probe pulse parameters. The operator has a principle axis defined as the state that is most easily ionized. The angle between this and the $|m_\ell=0\rangle$ state is shown in the first panel. The ratio of ionization probabilities for the $|m_\ell=1\rangle$ compared to the $|m_\ell=0\rangle$ state is shown in the second panel. The result is that both quantities have a weak dependence on the probe pulse, but narrower bandwidth probes are more sensitive to orbital alignment.}
\end{center}
\end{figure}
\begin{figure}
\centering
\includegraphics[width=\columnwidth, angle=0]{exit_electron_4b.eps}
\caption{\label{atomic_state_illustratio}The wavefunction of the valence electron at the peak of a short ionizing laser pulse. Contours of probability density multiplied by the electron-core separation squared are chosen at intervals to highlight the exiting electron's wavefunction, which is a superposition of s and d waves. Because of the short timescale, the electron has not had time to travel far from the core. The shelled structure is due to the electron being emitted at a specific phase of the incident radiation. The electric field axis is aligned vertically.}
\end{figure}
\section{Atomic spin-orbit dynamics}
The spin-orbit dynamics of the first excited state of the rubidium atom are of most interest here due to the ease of production of the state. The two levels in question are illustrated in Fig.~\ref{level_diagram}.
The atom has a single valence electron, giving it an electron spin of $\frac{1}{2}$. In the first excited state, the 5P state, the electron orbital has one unit of angular momentum. This gives the atom 6 allowed angular momentum states. Two pairs of these are coupled by the spin-orbit coupling: $|m_\ell = 0, m_s = \pm \frac{1}{2}\rangle$ is coupled to $|m_\ell = \pm1, m_s = \mp \frac{1}{2}\rangle$. Since the atom is transferred by a short laser pulse to the $|m_\ell = 0\rangle$ state, where the quantization axis is chosen along the electric field of the exciting laser pulse, the spin-orbit coupling causes an oscillation between the $|m_\ell = 0\rangle$ and $|m_\ell = 1\rangle$ states. Since the $|m_\ell = 0$ state is about 60\% easier to ionize than the $|m_\ell = 1\rangle$ state, this gives the pump-probe signal of the atom an oscillation at the spin-orbit frequency of \unit[7.123]{THz}.
The initial state and the $|m_\ell = 0\rangle$ and $|m_\ell = 1\rangle$ states are shown in Fig.~\ref{atomic_state_illustration}.
\begin{figure}
\centering
\includegraphics[width=\columnwidth, angle=0]{orbitals_1c.eps}
\caption{\label{atomic_state_illustration}The rubidium ground state and first excited states. In the experiments discussed, the atom starts off in the ground state (5s) and is transferred to the first excited state. If the transition is instantaneous, then the excited population will be in the $m_{\ell} = 0$ state (centre). This is not an eigenstate of the full Hamiltonian since it is spin orbit coupled to the $m_{\ell}=1$ state (right). The result is that the electron orbital oscillates between the $m_{\ell} = 0$ and $m_{\ell} = 1$ states. Since the two states have different ionization probabilities when an ionizing pulse is applied, the probability of ionizing the atom oscillates. The surfaces are a contour of $\rho(\mathbf{r})|r|^2$.}
\end{figure}
\section{Extension to molecular case\label{molecular_model}}
The calculations can be extended to the molecular case by increasing the 8 atomic states (2 ground, 6 excited) to 15 molecular states (3 ground, 12 excited), which make up all the accessible electronic states of the molecule, assuming it only absorbs a single photon and is initially in a triplet state. The internuclear separation is added as a continuous variable, and the angle the internuclear axis makes with the electric field of the laser pulses is added as a constant parameter since it does not change appreciably over the timescale of the experiment. The Schr\"{o}dinger equation is thus converted from a discrete 8 by 8 matrix equation into a one dimensional partial differential equation with 15 channels. The details of the model are given in a previous publication \cite{martay_2009}.
The ionization step is modelled by projecting the molecular state onto the separated atomic states and taking the expectation value of the ionization matrix (Eqn.~\ref{M_ion}) as the ionization probability.
\section{Experiments}
The application of the atomic model to the study of the direct photoionization of long range molecules is demonstrated by considering two experiments.
The two experiments outlined here are both possible, and an effort has been made to ensure they are feasible with current technology. Both have the same initial state which can be prepared using a ramped magnetic field over a Feshbach resonance. In both experiments, a pump pulse is applied, then after a delay, a probe pulse with the same polarization is applied. The probe pulses are identical, but the pump pulses are different. The polarization was chosen parallel to the magnetic field used to associate the molecules.
In the first experiment, the pump pulse is a Fourier limited Gaussian pulse whose bandwidth spans both the transition from the ground state, 5S, to the 5P$_{1/2}$ and to the 5P$_{3/2}$ state. This maximises the visibility of the spin-orbit precession. In the second experiment, the same Gaussian pump pulse is spectrally cut to remove intensity at the 5S --- 5P$_{3/2}$ transition. This prevents the preparation of a precessing orbital. As a result, the only mechanism by which the orbital can change orientation is through the interaction of the two atoms in the molecule. Since this depends on internuclear separation, the oscillation in the internuclear separation coordinate becomes apparent in the pump-probe signal.
\subsection{Initial state\label{initial}}
The initial state for each experiment is prepared using a magnetic field sweep across a Feshbach resonance in a cold $^{87}$Rb gas with a high phase space density in order to associate atom pairs to loosely bound molecules.
The ground electronic state rubidium dimer has a discrete number of bound states plus several continua of scattering states, each with a different electron and nuclear spin state. The energies of the bound and continuum states change as a function of magnetic field and spin state. The dynamics and molecular properties of long range dimers bound by less than a few h$\times$GHz is dictated by the hyperfine and Zeeman interactions. Several reviews \cite{kohler_2006} and other publications are concerned with the behaviour of such molecules, which constitute a major area of study in their own right.
The use of a Feshbach resonance to create molecules was demonstrated in 2002 \cite{donley_2002}. Work had already been done on using a magnetic field to control atom-atom interactions both theoretically and experimentally \cite{tiesinga_1993, inouye_1998}. This mechanism is the mechanism for creating the initial state in the work here. An essay by D. Kleppner presents a fuller picture of the use of Feshbach resonances in this context \cite{kleppner_2004}. Magnetic sweep association of the species under the conditions studied here has been demonstrated experimentally \cite{thalhammer_2006}.
Two interacting atoms are prepared in their high-field seeking $f = 1$ state. The spin-spin interaction between the two atoms is neglected, which constrains the atom pair to stay in the $m_f = 2$ state, since the remaining hyperfine, Zeeman, and electronic interactions do not break the rotational symmetry of the electron plus nuclear spins about the magnetic field axis. For numerical reasons, the internuclear separation is constrained to be in a box. The highest 9 vibrational states and lowest 191 box states for each of the 5 accessible spin states are taken as a basis set. The Hamiltonian, consisting of the vibrational energies, the electron and nuclear Zeeman energies, and the hyperfine interaction, was diagonalised for each magnetic field. The energy levels of the dimer as a function of magnetic field for this model are shown in Fig.~\ref{feshbach_diagram}.
\begin{figure}
\centering
\includegraphics[width=\columnwidth, angle=0]{feshbach_diagram_.eps}
\caption{\label{feshbach_diagram}The highest bound state energies of the rubidium dimer as a function of magnetic field. For the experiments outlined here, the atom pairs are initially unbound at a magnetic field of greater than \unit[1007]{G} (a). The magnetic field is swept down to \unit[940]{G} (b) for the first experiment, and to \unit[850]{G} (c) for the second experiment. This prepares a suitable initial state for each experiment.}
\end{figure}
By following a single state adiabatically from above \unit[1007]{G}, where it is a box state, to \unit[940]{G}, where it is bound by 24 h$\times$MHz, or to \unit[850]{G} for the second experiment, a suitable initial state is prepared.
For the \unit[940]{G} state, the state is well represented by the highest single channel vibrational state (whether singlet or triplet) multiplied by a spin state $|\psi\rangle$ which is 74\% $|S=1, M_S = -1, I=3, M_I=3\rangle$, \newline 12\% $|S=1, M_S = 0, I=3, M_I=2\rangle$, and 12\% $|S=0, M_S = 0, I=2, M_I=2\rangle$. The superposition is coherent, but the phases between these three states are not measured by the experiment, since the rest of the experiments take place on a timescale too short to change the nuclear spins and so the nuclear spins effectively measure the initial state of the electron spin. For this state, the singlet population is neglected. There are two justifications for this. The first is that it is only 12\% of the population, and the second is that it is expected to behave similarly to the triplet population: It will spin-orbit precess at first, but molecules at different internuclear separations will spin-orbit precess at different rates, so the visibility of the precession will wash out.
For the \unit[850]{G} state, the state is well represented by the fifth highest triplet state with 70\% in the \newline $|S=1, M_S = 1, I=3, M_I=1\rangle$ state, and 28\% in the $|S=1, M_S = -1, I=3, M_I=3\rangle$ state.
\begin{figure}
\centering
\includegraphics[width=\columnwidth, angle=0]{pump_probe_signal_uncut.eps}
\caption{\label{pump_probe_signa_uncut}The ionization probability of a long-range molecule prepared as described in section~\ref{initial} at \unit[940]{G} in a pump-probe experiment where the molecule is excited to the states associated with the 5S + 5P asymptote and then ionized directly. The pump and probe are sub-picosecond pulses. This is the calculated experimental outcome for experiment 1, where the molecules are put in a superposition of fine structure states in order to cause spin-orbit precession. The spin-orbit precession is clearly visible in the signal here.}
\end{figure}
\begin{figure}
\centering
\includegraphics[width=\columnwidth, angle=0]{signal_contribution_uncut.eps}
\caption{\label{pump_probe_signal_uncut}The relative number of molecular ions yielded as a function of internuclear separation for a long-range molecule prepared as described in section~\ref{initial} in a pump-probe experiment where the molecule is excited to the states associated with the 5S + 5P asymptote and then ionized directly. The angle between the internuclear axis and the electric field is 1.4 radians. The contribution is shown as a function of internuclear separation and time. This is the population density multiplied by ionization probability for experiment 1. It shows the spin orbit precession, and how the contributions from different internuclear separations become dephased leading to the signal washing out. The figure has been deliberately undersampled to increase the apparent oscillation period to around \unit[1]{ps} rather than \unit[1/7]{ps}.}
\end{figure}
\subsection{Experiment 1: Broadband uncut pump\label{ftl}}
In this experiment, we wish to observe spin-orbit precession in the long range Rb$_2$ dimer. The molecule is prepared as described in section~\ref{initial}. This initial state is very loosely bound, which increases the dephasing time for the spin-orbit precession, allowing a clearer signal. The molecules are excited with a Gaussian pump probe with a centre wavelength of \unit[375]{THz} and a full width at half maximum intensity of \unit[30]{THz}. After a variable time delay, the molecules are ionized by a second Gaussian laser pulse, centred at \unit[650]{THz}, with a full width at half maximum of \unit[18]{THz}, and with a total fluence of \unit[0.3]{Jm$^{-2}$}. The ionization probability as a function of delay is measured by counting the number of molecular ions produced.
The experiment is modelled by taking the initial state, propagating it under the influence of the first pulse. At each timestep, the expectation of the ionization operator can be taken, which gives the ionization probability of the molecule at the specified delay. The calculation is repeated for 21 different angles between the internuclear axis and the electric field and averaged. This ionization probability is given in Fig.~\ref{pump_probe_signa_uncut}. The contribution to this signal as a function of internuclear separations for one angle is given in Fig.~\ref{pump_probe_signal_uncut}.
The spin-orbit precession is initially visible as a \unit[7]{THz} oscillation, with a good signal to background, but then decays due to the interaction of the two atoms.
\begin{figure}
\centering
\includegraphics[width=\columnwidth, angle=0]{pump_probe_signal_cut.eps}
\caption{\label{pump_probe_signa_cut}The ionization probability of a long-range molecule prepared as described in section~\ref{initial} at \unit[850]{G} in a pump-probe experiment where the molecule is excited to the states associated with the 5S + 5P asymptote and then ionized directly. This is the signal for the second experiment. The pump pulse is spectrally cut to suppress the spin-orbit precession. As a result, and as a result of the different initial state, the molecular dynamics are visible as a long period contribution to the signal.}
\end{figure}
\begin{figure}
\centering
\includegraphics[width=\columnwidth, angle=0]{signal_contribution_cut.eps}
\caption{\label{pump_probe_signal_cut}The contribution to the total ion yield of a long-range molecule prepared as described in section~\ref{initial}, as for Fig.~\ref{pump_probe_signa_uncut}, but for experiment 2. It shows the internuclear separation oscillations clearly. The admixture of excited states that ionize easily change with internuclear separation, and this results in a time-dependent ion signal.}
\end{figure}
\subsection{Experiment 2: Spectrally cut pump\label{cut_pulse}}
In this experiment, the aim is to observe the internuclear oscillations rather than spin-orbit precession. As a result, a different initial state is chosen that has the fifth highest triplet state occupied rather than the highest. This shifts the molecular population to separations where the ionization process is more sensitive to separation, and where the population oscillates more uniformly. The pump pulse is spectrally cut to suppress the orbital precession. The probe is the same as for experiment 1. The experiment is modelled as before, and the expected ionization probability is given in Fig.~\ref{pump_probe_signa_cut}. The contributions to this signal are shown in Fig.~\ref{pump_probe_signal_cut}.
The spin-orbit precession is suppressed as hoped. A persistant long period oscillation may be observed mixed with a short period oscillation. The long period signal is due to the oscillation in the internuclear separation. The short period signal is due to spin-orbit precession.
\subsection{Discussion}
The first experiment would demonstrate that it is possible to observe spin-orbit precession in long-range molecules. One application for this is to verify which molecular vibrational state has been prepared by a magnetic field sweep, since the speed with which the oscillation in the ion signal washes out depends on the vibrational state or states that the molecules are in.
The second experiment is similar to the first experiment discussed in reference~\cite{martay_2009}. In the earlier work, it was shown that a coherently oscillating excited state wavefunction could be formed in long-range molecules. It was stated that if a position sensitive measurement could be made, then the molecular dynamics could be observed. In this paper, the result is extended by showing that a position dependent measurement can indeed be made, using orbital alignment. The initial state has to be modified from the earlier work, since it is difficult experimentally to use an above threshold ionization laser on a gas in a magneto-optical trap due to atomic ionization. In this way, the pump-probe experiment can now be modelled from initial state preparation to the creation of the molecular ions. The observation of wavepacket dynamics in the excited state of long range molecules would be a step towards the control and stabilization of such molecules, and would be a valuable contribution to the field of ultracold chemistry.
\section{Conclusions}
A simple atomic model is presented that is accurate enough to model spin-orbit precession and photoionization. A method for extending atomic spin-orbit precession calculations to the case of long range molecules is presented. This is demonstrated by the calculation of the ionization signal for two pump-probe experiments. A consequence of this is that it is shown that long range molecules can exhibit spin-orbit precession, but that it washes out. A second consequence is that it is shown that the mechanism that causes spin-orbit precession to affect the ionization signal also causes an internuclear separation oscillation to affect the ionization signal, allowing pump-probe experiments to observe coherent wavepacket behaviour in ultracold long range molecules.
\begin{acknowledgments}
The authors wish to acknowledge discussion and advice from Thorsten K\"{o}hler, Jordi Mur-Petit and Jovana Petrovi\u{c}.
Financial support is acknowledged from the EPSRC, grant number EP/D002842/1.
\end{acknowledgments}
\section*{APPENDIX}
The eigenstates and eigenenergies of the Schr\"{o}dinger equation are found in the basis defined by the quantum numbers of total angular momentum $j=l+s$, the projected total angular momentum $m_j = m_l + m_s$, and the electron orbital angular momentum $l$ --- a basis for which equation \ref{H_e} is diagonal. The lowest 100 energy levels for the s,p,d,f and g orbitals are used. The calculations are performed in a spherical box of \unit[1600]{$a_0$}. This ensures that the lowest 100 energy levels contain scattering states with an energy range that covers the possible free electron energies in the experiments discussed here. The finite spacing of the scattering states puts an upper bound on the duration of any ionizing pulse that can be accurately modelled. However, all the pulses used here are short enough that the scattering states included in the basis set provide an accurate representation of the continuum. The energy levels were obtained using fourth order Runge-Kutta.
The time-dependent Schr\"{o}dinger equation was solved using a matrix representation of the propagator for each timestep, chosen to be shorter than the time variation in the Hamiltonian. For weak fields, a 1st or 2nd order perturbative expansion may be used.
Continuous-wave photoionization cross sections can be calculated in this framework using the equation
\begin{eqnarray}
\sigma \simeq \frac{4\pi}{\sqrt{2}} LE^{-\frac{3}{2}}\omega \sum_j\mu_j^2,
\end{eqnarray}
where $\mu_j$ is the transition dipole moment from the initial state to the final scattering state at energy $E$ normalized to a box of length $L$, and $\omega$ is the angular frequency of the incident linearly polarized light. The sum is over the various electronic configurations of the scattering state: the different allowed orbital and spin angular momenta of the exiting electron. The equation becomes exact in the limit that $L\rightarrow\infty$.
|
\section{Introduction}
\label{s:intro}
Two broad finance theories make predictions about stock prices. They are the efficient market hypothesis (EMH) and the rational bubbles view (RBV). Both theories begin from the standpoint that an asset has a fundamental value, defined as the market's expected discounted present value of the future cash flows associated with the asset. Empirical tests of both the EMH and the RBV often fail to explain large market price falls or `crashes', as such crashes are not usually associated with any specific news item.\footnote{
Our definition of a stock market crash is similar to that of Hong and Stein \cite{HS:2003}, in that they represent unusual large market falls that are not followed by large public news events and such falls are market wide in nature.}
For example, Cutler et al. \cite{CPS:1989} find that of the 50 largest daily price falls in aggregate stock prices for the period 1946-1987, the majority are not accompanied by external news of specific importance.\footnote{
Recently, several specific theoretical models of stock market crashes have been put forward. In Romer's \cite{R:2001} symmetric rational asset-price model, neither rational behavior nor external news plays an important role in stock market crashes.
Also both the Hong and Stein \cite{HS:2003} and Barley and Veronesi \cite{BV:2003} models assume that economically significant differences in the views of investors can lead to stock market crashes when they are revealed.}
Empirical tests of the RBV have also had limited success in identifying price bubbles prior to large price falls (see, e.g., \cite{BW:1982} and \cite{W:1987}).
In particular, Donaldson and Kamstra \cite{DK:1996} estimate a non-linear ARMA-ARCH artificial neural network model which allows them to reject the claim that the 1929 stock market crash was the outcome of a bubble. One reason for the failure of tests of the RBV is the difficulty of explicitly isolating an asset's fundamental value from the component of the bubble tied to the asset's market price.
An alternative approach to modeling stock market crashes and their bubbles is to fit a Log-Periodic Power-Law (LPPL) to asset prices.\footnote{
Kumar et al. \cite{KMP:2003} also apply logit models to microeconomic and financial data and show that currency crashes can be predicted.}
The notion that statistical description of financial crashes as manifestations of power law accelerations essentially suggests that stock market crashes obey a particular power law with log-periodic fluctuations. Sornette et al. \cite{SJB:1996}, Sornette and Johansen \cite{SJ:1997} and Lillo and Mantegna \cite{LM:2004} show that this model is able to capture a shift over time in the log-periodic oscillations of financial prices that are associated with market crashes.
The analogy of financial crashes as being similar in their statistical signatures to critical points as depicted in natural phenomena has, however, been argued to be unrealistic. Laloux et al. \cite{LPC:1999} express doubts about the validity of a seven-parameter model fitted to highly noisy data. They argue that such a model would suffer from severe over-fitting. Also, some log-periodic precursors do not always lead to crashes but to a smooth draw-down or to an even greater draw-up. This suggests that there is no universality in the manner financial bubbles are manifested. Indeed, some evidence (see, e.g., \cite{F:2001}) shows that the predicted time of a crash is sensitive to the size of the event-window used to predict the crash.
Whilst the LPPL model is not perfect, it is empirically appealing as it provides a forecast of the date by which a financial crash will occur. This is an important attribute relative to other methods of financial risk assessment. For example, Novak and Beirlant \cite[p. 461]{NB:2006} argue Extreme Value Theory provides a means of predicting ``\ldots the magnitude of a market crash but not the day of the event." As such, the LPPL model appears to contain important statistical attributes that require serious empirical consideration.
For example, the LPPL model contains a component that captures the market's excessive volatility prior to a crash. This feature is consistent with several theoretical models of financial crashes as well as with empirical results \cite{L:2008,C:1996}.
There are several critical considerations associated with fitting an LPPL to financial data and we take issue with some of them as follows:
first, studies that support the LPPL (see e.g., \cite{JLS:2000}) show that the parameter estimates for this model are confined within certain ranges and that it is these that are the indicators of market crashes.
This view considerably restricts the number of classes of permissible models that fit bubbles preceding stock market crashes to just those LPPLs whose parameters fall within the specified ranges rather than to any 7 parameter fitted LPPL.
Second, the mechanism underlying the LPPL is such that prices must increase throughout the bubble, which is largely in line with the rational bubbles literature, but which is not what has been found in empirical fits of the LPPL (see Section \ref{ss:underlying mechanism}).
Finally, we do not feel that there has been sufficient critical analysis of the LPPL and its goodness-of-fit to market bubbles. In particular, a goodness-of-fit test is rarely applied and the sensitivity of the parameters of the fitted LPPL is usually not reported (see Section \ref{ss:sensitivity}).\footnote{
Laloux et al. \cite[p. 4]{LPC:1999} report two instances when financial crashes were predicted ex ante.
The prediction was correct in one case but not in the other despite both predictions being published prior to the expected crash date.
Indeed, they conclude that ``\ldots recent claims on the predictability of crashes are at this point not trust worthy".}
The remaining main sections of this paper are as follows: Section~\ref{s:LPPL} introduces the LPPL; Section~\ref{s:mechanism} describes the mechanism underlying the LPPL and tests it using already published results; Section~\ref{s:fitting} gives some details of the procedure used for finding the parameters of an LPPL so that it best fits the data; Section~\ref{s:results} fits the LPPL to pre-crash bubbles on the Hang Seng index, checks on the fits already published and tests whether the parameters of the fitted LPPLs do in fact predict crashes.
\section{The LPPL}
\label{s:LPPL}
The simplest form of the LPPL can be written as:
\begin{equation}
\label{lppl} y_t = A +B(t_c-t)^\beta\left\{1 + C\cos(\omega \log(t_c-t) + \phi)\right\},
\end{equation}
\noindent
where:\\
\begin{tabular}{ll}
$y_t > 0$ &is the price (index), or the log of the price, at time $t$; \\
$A > 0$ &is the value that $y_t$ would have if the bubble were to last until\\ &the critical time $t_c$; \\
$B < 0$ &is the increase in $y_t$ over the time unit before the crash, \\
&if C were to be close to zero; \\
$C$ &is the magnitude of the fluctuations around the exponential \\
&growth, as a proportion; \\
$t_c>0$ &is the critical time; \\
$t < t_c$ &is any time into the bubble, preceding $t_c$; \\
$\beta= 0.33 \pm 0.18$ &is the exponent of the power law growth; \\
$\omega = 6.36 \pm 1.56$ &is the frequency of the fluctuations during the bubble; \\
$0\le \phi \le 2\pi$ &is a shift parameter.
\end{tabular}
\noindent The ranges of values given for both $\beta$ and $\omega$ are based on the observed parameters of crashes for many stock markets \cite{J:2003}.
These ranges for $\beta$ and $\omega$, rather than any goodness-of-fit test, are used to identify the bubbles that precede crashes.
Empirical studies that fit the LPPL to financial data make a number of claims:
\begin{enumerate}
\item The mechanism that characterizes traders on financial markets is one in which they mutually influence each other within local neighborhoods.
This leads, in turn to co-ordinated behavior through a martingale condition, which in the extreme can lead to a bubble and then a crash (see e.g., \cite{JLS:2000}).
\item Financial crashes are preceded by bubbles with fluctuations. Both the bubble and the crash can be captured by the LPPL when specific bounds are imposed on the critical parameters $\beta$ and $\omega$ (see e.g., \cite{J:2003, JS:2001}).
\item The established parameters are sufficient to distinguish between LPPL fits
that precede a crash from those that do not (see e.g., \cite{SJ:2001}).\footnote{
Recently Lin et al. \cite{LRS:2009} carried out such an evaluation on a variant of the LPPL model.}
\end{enumerate}
\noindent In this paper, we examine the first two of the above claims and suggest a new approach for testing them. The third claim is more controversial; it only makes sense to evaluate it once we have a positive evaluation of the second claim.
\section{Is the Underlying Mechanism Correct?}
\label{s:mechanism}
\subsection{The underlying mechanism}
The mechanism driving the change in price during a bubble as posited in Johansen et al. \cite{JLS:2000} is based on rational expectations,
namely, that the expected price rise must compensate for the expected risk.
The mechanism is a stochastic process such that the conditional expected value of the asset at time $t+1$, given all previous data before and up to $t$, is equal to its price at time $t$. The martingale condition is formulated by Johansen et al. \cite{JLS:2000} as:
\begin{equation}
\label{dp1} dp \leftarrow \kappa p(t) h(t) dt,
\end{equation}
\noindent
\begin{tabular}{lll}
where: &$dp$ &is the expected change in price, conditional on no crash occurring \\
& & over the next time interval $dt$, at equilibrium; \\
&$p(t)$ & is the price at time $t$; \\
&$\kappa$&is the proportion by which the price is expected to drop during \\
& & a crash, if it were to occur; \\
&$h(t)$ &is the hazard rate at time $t$, i.e. the chance the crash will occur \\
& & in the next unit of time, given that it has not occurred already. \\
\end{tabular}
\noindent
Under this martingale condition, investors will buy shares at time $t$ if their expectation is that the price at time $t+1$ will exceed the price at $t$ by more than the associated risk; that is: $ E(p(t+1)) > p(t) + dp$. This buying would drive up today's price.
So the expected rise in price between today and tomorrow will be less (assuming that the expected price tomorrow remains constant); this buying will continue until the expected rise is in line with the perceived risk according to Eq. \ref{dp1}.
And, vice versa, if investors believe that the expected rise in price tomorrow will be insufficient to compensate for the risk, i.e. $ E(p(t+1)) < p(t) + dp$, then they will try to sell today, going short if necessary, thus driving today's price down.
Note that all the terms on the right side of Eq. \ref{dp1} are positive, so $dp > 0$,
i.e., the price must always be expected to be increasing during a bubble.
This condition was not treated as a constraint in early work (see, e.g., \cite{JLS:2000}) and as such gives us the opportunity of treating this requirement as a testable prediction.\footnote{
Sornette and Zhou \cite{SZ:2006}, a later work, does treat this condition as a constraint on the permissible parameter values.}
We now follow the consequences of Eq. \ref{dp1} for the behavior of prices.
Re-arranging Eq. \ref{dp1} gives us:
\begin{eqnarray}
\nonumber \frac{1}{p(t)} dp &=& \kappa h(t) dt, \\
\label{p=h} \log p(t) &=& \kappa \int_{t_0}^t h(t') dt' .
\end{eqnarray}
To capture the behavior of the price, the hazard rate, $h(t)$, needs to be specified.
Here, Johansen et al. \cite{JLS:2000} posit a model in which each trader $i$ is in one of two states, either bull (+1) or bear (-1). At the next time step, the position of trader $i$ is given by:
\begin{equation}
\label{trader}
\mathrm{sign} \left( K \displaystyle \sum_{j \in N(i)} s_j + \sigma \epsilon_i \right),
\end{equation}
\noindent
\begin{tabular}{lll}
where: &$K$ &is the coupling strength between traders; \\
&$N(i)$ &is the set of traders who influence trader $i$; \\
&$s_j$ &is the current state of trader $j$; \\
&$\sigma$ &is the tendency towards idiosyncratic behavior for all traders; \\
&$\epsilon_i$ &is a random draw from a zero mean unit variance normal \\
& & distribution.
\end{tabular}
\noindent
The relevant parameter determining the behavior of a collection of such traders is the ratio $K/\sigma$, which determines a critical value of $K$, say $K_c$.
If $K\ll K_c$ then the collection is in a disordered state.
However, as $K$ approaches $K_c$ order begins to appear in the collection, with a majority of traders having the same state.
As the value of $K$ approaches $K_c$ from below, the system becomes more sensitive to small initial perturbations.
At the critical value, $K_c$, all the traders will have the same state, either +1 or -1.
Johansen et al. \cite{JLS:2000} further assume that: i) the coupling strength $K$ increases smoothly over time up to $K_c$; and ii) the hazard rate is proportional to $K$. They do not justify these assumptions but the first one might be based on assuming that, as the frequency of fluctuations increases, traders become less sure of their own judgment and rely more on the judgment of their neighbors. In the next sections, we consider the evolution of $K$ over time.
\subsection{Simple power law hazard rate }
In the simplest scenario $K$ evolves linearly with time.
Assuming that each trader has four neighbors arranged in a regular two dimensional grid, then the {\em susceptibility} of the system near the critical value, $K_c$, can be shown to be given by the approximation:
\begin{equation}
\chi \approx B'' (K_c-K)^{-\gamma},
\end{equation}
\noindent
where $B''>0$ and $0 < \gamma < 1$ (see \cite{JLS:2000}).
The three assumptions taken together give:
\begin{equation}
\label{h1} h(t) \approx B' (t_c-t)^{-\alpha},
\end{equation}
\noindent where $0<\alpha<1$.
Substituting in Eq. \ref{p=h} for $h$ as given by Eq. \ref{h1} and integrating gives:
\begin{eqnarray}
\nonumber \log p(t) &=& \kappa \int_{t_0}^t B'(t_c-t')^{-\alpha} dt'
= \frac{-\kappa B' }{1-\alpha} \left[(t_c-t)^{1-\alpha}\right]^t_{t_0}\\
\nonumber &=&\frac{-\kappa B' }{1-\alpha} \left((t_c-t)^{1-\alpha} - (t_c-t_0)^{1-\alpha}\right).\\
\nonumber \mbox{At } t= t_c,\ \log p(t_c) &=&\frac{-\kappa B'} {1-\alpha} \left(0 - (t_c-t_0)^{1-\alpha}\right). \\
\nonumber \mbox{So } \log p(t) &=& \log p(t_c) + \frac{\kappa B'}{1-\alpha} (t_c-t)^{1-\alpha} \\
\label{p1} &=& A + B (t_c-t)^{\beta},
\end{eqnarray}
where:
$A = \log p(t_c),\
B = \kappa B'/(1-\alpha) $ and $\beta = 1-\alpha$.
This is a simple exponential growth.
\subsection{Log periodic hazard rate}
To introduce log periodic fluctuations into the exponential growth function we need a different form of interconnected structure.
Such a structure is assumed to be equivalent to one created by: i) starting with a pair of linked traders; ii) replacing each link in the current network by a diamond with four links and two new nodes diagonally opposite each other. This process continues until some stopping criterion is met. Then (see \cite{JLS:2000}):
\begin{eqnarray}
\nonumber \chi &\approx
&B'' (K_c-K)^{-\gamma} + C''(K_c-K)^{-\gamma} \cos(\omega \log(K_c-K) + \phi') + \ldots .\\
\label{h2} \mbox{So } h(t) &\approx
&B'(t_c-t)^{-\alpha}[1 + C' \cos(\omega \log(t_c-t) + \phi')]
\mbox{, from Eq. \ref{h1}.}
\end{eqnarray}
\noindent
Substituting for $h$ in Eq. \ref{p=h} from Eq. \ref{h2} gives:
\begin{equation}
\label{p2} \log p(t) = \kappa \int_{t_0}^t B'(t_c-t')^{-\alpha} \{1 + C' \cos(\omega \log(t_c-t') + \phi')\} dt'.
\end{equation}
\noindent Substituting $\beta =1-\alpha$ and
$\psi(t') = \omega \log(t_c-t') + \phi' $ in the integral
\begin{eqnarray}
\label{integral}
\int_{t_0}^t (t_c-t')^{-\alpha} \cos(\omega \log(t_c-t') + \phi') dt'
&=& \int_{t_0}^t (t_c-t')^{\beta -1} \cos \psi(t') dt' \\
\nonumber &= &\left[\frac{-(t_c-t')^{\beta}}{\omega^2 +\beta^2}
\{\omega \sin \psi(t') + \beta \cos \psi(t')\}\right]_{t_0}^t.
\end{eqnarray}
\noindent
Integrating Eq. \ref{p2} using Eq. \ref{integral} gives:
\begin{eqnarray}
\nonumber \log p(t) &=& \kappa \left[\frac{-B'}{1-\alpha} (t_c-t')^{1-\alpha}
- \frac{B'C'(t_c-t')^{\beta}}{\omega^2+\beta^2 }
\left(\omega\sin \psi(t') +\beta \cos \psi(t') \right)\right]^t_{t_0} \\
\nonumber &=& \kappa \left[\frac{-B'}{\beta} \left\{(t_c-t)^{\beta} -(t_c-t_0)^{\beta}\right\}\right. \\
\nonumber && \hspace*{1 cm} \left. - \frac{B'C'}{\omega^2+\beta^2 }
\left\{(t_c-t)^{\beta} \left(\omega\sin \psi(t) +\beta \cos \psi(t) \right)\right.\right.\\
\nonumber && \hspace*{3 cm} \left. \left.
- (t_c-t_0)^{\beta}\left(\omega\sin \psi(t_0) +\beta \cos \psi(t_0)\right)\right\}\right],\\
\nonumber \log p(t_c) &=& \kappa \left[\frac{B'(t_c-t_0)^{\beta} }{\beta}
+ \frac{B'C'(t_c-t_0)^{\beta}}{\omega^2+\beta^2 }
\left(\omega\sin \psi(t_0) + \beta \cos \psi(t_0) \right)\right] ,\\
\nonumber \log p(t) &=&
\log p(t_c) - \frac{\kappa B' (t_c-t)^{\beta}}{\beta}
-\frac{\kappa B'C'}{\omega^2+\beta^2 }
\left\{(t_c-t)^{\beta}\left(\omega\sin \psi(t) + \beta \cos \psi(t) \right)\right\} \\
\nonumber &=& A + B(t_c-t)^{\beta}[1+C''(\omega\sin \psi(t) + \beta \cos\psi(t) )] \\
\nonumber &\approx& A + B(t_c-t)^{\beta}[1+C\cos(\psi(t)+ \phi' )] \\
\label{p3} &\approx&
A + B(t_c-t)^{\beta}[1+C\cos(\omega \log(t_c-t) + \phi)],
\end{eqnarray}
\noindent where
$A = \log p(t_c),\
B= -\kappa B' /\beta,\ $ and
$C''= \beta C'/(\omega^2+\beta^2 )$,
which is the LPPL of Eq. \ref{lppl} with $y_t = \log(p_t)$.
\subsection{Index: raw versus log}
From Eq. (11), it is the log of the price index that needs to be fitted to the LPPL although the LPPL is often fitted to the raw index data.
Johansen and Sornette \cite{JS:2001} justify the use of the raw data by assuming that the price drop in the crash is proportional to the price over and above the fundamental value rather than being proportional simply to the price.
That is, they replace the condition \ref{dp1} by:
\begin{equation}
\label{dp2} dp \leftarrow \kappa (p(t)-p_1) h(t) dt,
\end{equation}
\noindent where $p_1$ is the fundamental value, which they do not further define.
Johansen and Sornette \cite{JS:2001} introduce the assumption that the rise in price since the beginning of the bubble is much less than the amount by which the price at the beginning of the bubble is above the fundamental value. Thus
\begin{equation}
\label{small price rise} p(t)-p(t_0) \ll p(t_0) -p_1,
\end{equation}
where $t_0$ is the time of the beginning of the bubble.
Even if the asset's fundamental value is not estimated in the model, the above assumption is weakly testable.
If the price rise during the bubble is greater than the price at the beginning of the bubble, i.e. $p(t) > 2 p(t_0)$, then the condition of Eq. \ref{small price rise} cannot be fulfilled unless the fundamental price is negative.
We assume that this is not what is intended.
So we can test whether or not this assumption is met.
Integrating Eq. \ref{dp2} from the moment when the bubble starts, $t_0$, and using Eq. \ref{small price rise} gives:
\begin{eqnarray}
\nonumber p(t) &=& p(t_0) + \int^t_{t_0} dp \\
\nonumber &=& p(t_0)+\kappa \int^t_{t_0} (p(t')-p_1) h(t') dt' \\
\label{p4} &\approx & p(t_0) +\kappa (p(t_0) - p_1) \int^t_{t_0} h(t') dt'.
\end{eqnarray}
Provided the assumption in Eq. \ref{small price rise} is met, Eq. \ref{p4} can be used to fit the LPPL to raw price (as done, e.g., in \cite{JS:2001}) rather than the log price data.
\subsection{Tests of the underlying mechanism}
\label{ss:underlying mechanism}
Chang and Feigenbaum \cite{CF:2006} tested the mechanism underlying the LPPL using S\&P index data for the bubble preceding the 1987 crash.
They tested whether the observed price changes could be fitted to a LPPL and whether price changes could be predicted by a random walk model.
To do so they first extended the LPPL as given in Eq. \ref{lppl} by adding:
\LI
\item a random term with zero mean and variance estimated from the data.
This noise term is necessary to compute a likelihood for the observed data deviations from the predicted LPPL.
\item a positive upward drift term estimated from the data.
This addition to the LPPL, while frequently made in financial time series, is unnecessary here as exponential growth is posited in the LPPL.
\end{list}
Then they estimated the likelihood of the observed change in price since the previous day $t-1$, and selected parameters that maximized the sum of these likelihoods over the entire bubble.
With a time series there is a choice of which next point to take as being the most likely:
either the predicted value or the predicted change since $t-1$.
Using the model's prediction of the value at $t$ ignores the value at $t-1$;
this is what Johansen et al \cite{JLS:2000} implicitly assume when they minimize the RMSE for the fitted LPPL against the data.
On the other hand, using the predicted change since $t-1$ ignores any deviation that the price at $t-1$ already has from the model's prediction for $t-1$.
This is what Chang and Feigenbaum \cite{CF:2006} explicitly do.
So they have established simply that the mechanism underlying their adaptation of the LPPL, when judged for each time point separately, is not to be preferred to the random walk model,
which is hardly surprising given what is known about the random nature of stock market prices.
While most of the assumptions underlying the mechanism from which the LPPL is derived are untestable (or even questionable), there is one that is testable: the hazard rate $h$ must be positive.
This implies that the price must always rise.
If the fitted LPPL does not have this property, then the assumption that $h(t)$ in Eq. \ref{dp1} is a probability, must be rejected.
Figure \ref{fig:HS1989} shows the LPPL fitted to the raw Hang Seng index data for the bubble preceding the 1989 crash. The fit is similar to Figure 8 of Sornette and Johansen \cite{SJ:2001}. The LPPL in Figure 1 has a negative slope some of the time.
The same is true in 16 of the 30 cases reported by Sornette and Johansen \cite{JS:2001, SJ:2001}.\footnote{These 16 pre-crash bubbles are: the Dow Jones (1929, '62), S \& P ('37, '87), Hang Seng ('89, '94), Argentina ('91, '92, '97) and various stock market crashes of 1994 (Indonesia, Korea, Malaysia, Philippines) and 1997 (Indonesia, Mexico, Peru).}
That is, the fitted LPPL predicts that the price decreases at some time points.
Unless we are mistaken, this empirical fact is sufficient to reject the martingale condition as being the mechanism underlying the LPPL fit to pre-crash bubbles.
However, this does not affect the ability or otherwise of the LPPL to fit the data.
\begin{figure}[t]
\centering
\resizebox{\textwidth}{!}{\includegraphics{best_fit_1989.pdf}}
\caption{LPPL fit to the bubble preceding the 1989 crash on Hang Seng.}
\label{fig:HS1989}
\end{figure}
\section{Fitting the LPPL Parameters}
\label{s:fitting}
The seven parameters of the LPPL in Eq. \ref{lppl} have to be estimated from the window of data points in the bubble.
The chosen values of these parameters should be the ones that minimize the root mean squared error (RMSE) between the data and the LPPL prediction for each day of the bubble.
The squared error between the prediction from the fitted curve from Eq. \ref{lppl} and the data is:
\begin{equation}
\label{rmse} SE = \displaystyle \sum^{t_n}_{t=t_1} (y_t -\hat{y}_t)^2
= \displaystyle \sum^{t_n}_{t=t_1} \left\{y_t - A -B(t_c-t)^\beta
\left(1 + C\cos(\omega \log(t_c-t) + \phi\right)\right\}^2,
\end{equation}
\noindent
\begin{tabular}{lll}
where: &$y_t $ &is the data point, either the price index or its log; \\
&$\hat{y_t} $ &is the data point as predicted by the model; \\
&$n$ &is the number of weekdays in the bubble; \\
&$t_i$ &is the calendar day date of the $i^{th}$ weekday \\
& &from the beginning of the bubble. \\
\end{tabular}
\noindent
Partially differentiating Eq. \ref{rmse} with respect to the parameters $A, B$ and $C$ gives us three linear equations from which the values of $A, B$ and $C$ that minimize the RMSE are derived, given the other four parameters: $\beta, \omega, t_c$ and $\phi$.
To find suitable values for these four parameters a search method is required.
The method used in \cite{JS:2001} and \cite{SJ:2001}, hereafter collectively called the {\em JS studies}, was:
\LI
\item First to make a grid of points for the parameters $\omega$ and $t_c$, from each of which a Taboo search was conducted to find the best value of $\beta$ and $\phi$, i.e. the ones for which, with $A, B$ and $C$ chosen to minimize the RMSE, gave the lowest RMSE.
\item To select from these points those for which $0<\beta<1$.
\item From these points to conduct a Nelder-Mead Simplex search \cite{NM:1965}, with all the four search parameters free (and $A, B$ and $C$ chosen to minimize the RMSE).
\end{list}
\noindent We presume that the reason that any fit with $\beta\ge 1$ was rejected is because then the increase in the index is exponentially {\em declining} whereas the underlying mechanism requires it to be increasing. An alternative technique would have been to place no restriction on the value of $\beta$, and if a value of $\beta \ge 1$ is found, to reject the model, as we have done for the requirement that the fitted LPPL never decreases (see Section \ref{ss:underlying mechanism}).
Similar to the JS studies, we use a preliminary search procedure based on a grid to provide seeds for the Nelder-Mead Simplex method, as implemented in Matlab \cite{LRWW:1998}.
It is based on choosing different values for the two parameters $\omega$ and $\beta$, as these are the critical parameters for determining whether the fitted LPPL is a crash precursor or not (see Eq. \ref{lppl}).
The algorithm and the parameter values used are shown in the Appendix.
Note that instead of the crash date, $t_c$, we use $t2c$, the number of days between the day on which the estimate is being made and the predicted critical date.
\section{Empirical Results}
\label{s:results}
\subsection{Data and descriptive statistics}
\begin{table}[ht]
\caption{Descriptive statistics for changes in the log of the Hang Seng stock index.}
\label{tab:Hang Seng stats}
~\\
{\footnotesize
\begin{tabular}{l|cccccc}
&N &Mean &Variance &Skewness &Kurtosis &J-B \\
\hline
HS Index returns
&10152 &0.00045$^b$ &0.00035 &-1.25934$^a$ &31.58011$^a$ &424542.78.9$^a$\\
\hline
\end{tabular}
\begin{description}
\item[~~~~] The mean and variance are multiplied by 100.
\item[J-B] is the Jarque-Bera statistic for testing the null hypothesis of normality.
\item[$^a {\rm and~} ^b$] denote statistical significance at a 1- and 5- percent level, respectively.
\end{description}
}
\end{table}
To test the LPPL, daily observations for the Hang Seng were downloaded from Data-stream. The data spans the period 1$^{\mbox{\footnotesize st}}$ January 1970 to 31$^{\mbox{\footnotesize st}}$ December 2008. We analyze the Hang Seng index since it is commonly believed that this market has had several crashes, thus giving us ample opportunity to test the LPPL.\footnote{
This suggests that stock market crashes can be common. Indeed, using a statistical method to identify outliers, Schluter and Trede \cite{ST:2008} show that the 1987 stock market crash of the Dow Jones Industrial index was not a structurally unusual event.}
Descriptive statistics, shown in Table~\ref{tab:Hang Seng stats}, reveal that the mean log changes of the Hang Seng index series are significantly different from zero. Both skewness and (excess) kurtosis are significant such that the Jarque-Bera test rejects the null of normality at a 1 percent level. Notice that skewness is highly significant and negative.
This finding suggests that the Hang Seng stock market can be very sensitive to stock market crashes. That is, volatility feedback can increase the probability of large negative returns and in turn, increase the potential for crashes \cite{CH:1992}.
\subsection{Identifying a crash}
\label{ss:what is a crash}
As indicated earlier, we use a definition of stock market crashes similar to that of Hong and Stein \cite{HS:2003}. To test whether or not the LPPL can predict crashes we first need to identify the crash itself.
Usually a stock market crash is taken to mean a very large and unusual price fall. In our application, a crash can span more than one day. This is consistent with the October 1987 stock market crash.
\begin{figure}[t]
\centering
\resizebox{\textwidth}{9 cm}{\includegraphics{HS_plot.pdf}}
\caption{The Hang Seng index 1970 to 2008, showing those peaks that are initiators of crashes.}
\label{fig:Hang Seng_index}
\end{figure}
There are two situations when we might falsely claim that a crash has occurred.
One is when the index is on the way up in a bubble and then there is a large drop, but it turns out that the drop is temporary and the bubble continues.
The other is when, on the way down during a crash, the index experiences a recovery and so we identify the beginning of a new bubble but the recovery is temporary and the anti bubble is still in effect.
To avoid those situations, we identify a peak as one initiating a crash as follows:
\LI
\item A period of 262 weekdays prior to the peak for which there is no value higher than the peak's.
\item A drop in price of 25\%, i.e. down to 0.75 of the peak price, which is in line with the 1987 crash.
\item A period of 60 weekdays within which the drop in price needs to occur.
\end{list}
\begin{figure}
\centering
\resizebox{\textwidth}{9 cm}{\includegraphics{HS_crashes.pdf}}
\caption{Drops from peaks on Hang Seng index 1970 to 2008.}
\label{fig:Hang Seng_drops}
\end{figure}
We first want to determine whether the application of these criteria enables us to capture the eight crashes on the Hang Seng index, as identified in the JS studies.
Indeed, we identify crashes at the same time points as in the JS studies, except for one additional crash in 1981 (see Figure~\ref{fig:Hang Seng_index}). To exclude the price fall in 1981 from being classified as a crash, we would have to increase the drop-to criterion or reduce the drop-by criterion.
Doing either would also exclude some of the other peaks as initiating crashes, viz. those peaks that immediately preceded the crashes of 1978, 1994, 1997, all of which are identified as crash initiators in the JS studies (see Figure~\ref{fig:Hang Seng_drops}). Thus the rule they apply seems somewhat imprecise.
It is true that the 1981 crash occurs shortly after the 1980 crash, so we might exclude the 1980 peak as initiating a crash, but rather being a part of the bubble preceding the 1981 crash, but this is not what was done in Sornette and Johansen \cite{SJ:2001}.
It would also be possible to exclude fitting an LPPL to the bubble preceding the 1981 crash on the grounds that this bubble is too short -- just 7 months long.
However, another bubble (the one preceding the crash 1971) was fitted even though it lasted only 6 months.
As such, the bubble preceding the 1981 crash should have been included in the JS studies, unless one insists on having more than say 7 months of data preceding a crash.
On balance, we believe that it is appropriate to include the 1981 crash we have identified, giving us nine crashes for the period of the JS studies. Overall, the criteria for identifying a crash does not appear to be consistently applied in the JS studies.
In the period after the JS studies, i.e. between 2000 and 2008, our criteria identify two additional peaks as initiating crashes; these are in 2000 and in 2007. The two bubbles preceding these crashes provide a post-hoc test of the hypothesis underlying the LPPL Eq. \ref{lppl}.
\subsection{Troughs and bubble beginnings}
Having decided that a peak is the initiator of a crash, the data window to be used for fitting the LPPL to the preceding bubble needs to be carefully selected.
In the JS studies the start of the data window is taken to be the day on which the index reaches its lowest value ``prior to the change in trend" \cite{JS:2001}.
In real time matters are not so simple, since one does not know if the index will drop still further in the future.
So for real time analysis we would need to take as the end of the previous crash the lowest point since the last crash, up until now.
\begin{figure}
\centering
\resizebox{\textwidth}{9 cm}{\includegraphics{HS_troughs.pdf}}
\caption{Troughs and other beginnings of bubbles on Hang Seng 1970 to 2008.}
\label{fig:Hang Seng_troughs}
\end{figure}
Moreover, Johansen and Sornette \cite{JS:2001} sometimes move the beginning of the bubble from the lowest point since the previous crash to a later time as in their Asian and Latin-American study. This was done if ``at the trough the next bubble had not yet begun" (Johansen, personal communication).
From the JS studies, we deduce that this was done for four of the eight crashes
they identified on the Hang Seng:
\LI
\item 1971 crash: forward 2 months, from 5/1/1971 to 10/3/1971;
\item 1978 crash: forward 3 years and 1 month, from 10/12/1974 to 13/1/1978;
\item 1987 crash: forward 1 year and 8 months, from 2/12/1982 to 23/7/1984;
\item 1994 crash: forward 2 years and 2 months, from 5/6/1989 to 19/8/1991.
\end{list}
\noindent
These are indicated by squares in Figure \ref{fig:Hang Seng_troughs}.
It is clear why Johansen and Sornette \cite{JS:2001} moved the beginning of the bubbles for the 1978 and 1987 crashes to times later than the trough proceeding the crash.
For 1978 there was a long period of stable prices which is clearly not part of a bubble.
For 1987 the year and 8 months following the trough are characterized by two mini bubbles and two peaks (which with other crash criteria would themselves be considered initiators of crashes).
It is not so clear why they moved the start points of the other two bubbles (preceding the 1971 and 1994 crashes) forward.
In the JS studies, a model fit is only made if there are at least 131 weekdays of data between the trough and the crash. Changing the number of days could lead to different bubbles being considered as crash precursors. To illustrate this for the Hang Seng data, there are only 155 weekdays between the end of the 1980 crash and the peak in 1981 when it appears that another crash occurred. To require (say) 262 weekdays would result in insufficient data, and thus exclude the bubbles before both the 1981 and the 1971 crashes, thus affecting the results. This means that one needs to be very careful in implementing the rule, given the data under consideration.
\subsection{Fitting to the raw index}
\label{ss:index}
\begin{table}[ht]
\caption{Ratio of raw Hang Seng index on the last day to index at the beginning of the bubble.}
\label{tab:Hang Seng_raw_assumption}
\begin{center}
$
\begin{array}{rr|rr|r}
\multicolumn{2}{c|}{\mbox{Bubble:}} &\multicolumn{2}{c|}{\mbox{Raw Hang Seng:}}
&\mbox{Ratio:} \\
\mbox{beginning at } t_0 &\mbox{~~~~ending on } t_e &~~~~p(t_0) &p(t_e) &p(t_e)/p(t_0)\\
\hline
\mbox{*10-Mar-1971} &\mbox{20-Sep-1971} & 201 & 406 &2.02! \\
\mbox{22-Nov-1971} &\mbox{09-Mar-1973} & 279 & 1775 &6.36! \\
\mbox{*13-Jan-1978} &\mbox{04-Sep-1978} & 383 & 707 &1.85 \\
\mbox{20-Nov-1978} &\mbox{13-Nov-1980} & 468 & 1655 &3.54! \\
\mbox{12-Dec-1980} &\mbox{17-Jul-1981} &1222 & 1810 &1.48 \\
\mbox{*23-Jul-1984} &\mbox{01-Oct-1987} & 747 & 3950 &5.29! \\
\mbox{07-Dec-1987} &\mbox{15-May-1989} &1895 & 3310 &1.75 \\
\mbox{*19-Aug-1991} &\mbox{04-Jan-1994} &3723 &12201 &3.28! \\
\mbox{23-Jan-1995} &\mbox{07-Aug-1997} &6968 &16673 &2.39! \\
\mbox{13-Aug-1998} &\mbox{28-Mar-2000} &6660 &18302 &2.75! \\
\mbox{23-Apr-2003} &\mbox{30-Oct-2007} &8520 &31638 &3.71! \\
\hline
\end{array}
$
\end{center}
{\small
\begin{description}
\item[$t_0$] the day the bubble began.
\item[$t_e$] the last day of the bubble.
\item[*] Bubble beginning moved to a time later than the trough between peaks.
\item[!] The ratio $p(t_e)/p(t_0)> 2$, so
the raw index should not be used (see Eq. \ref{small price rise}).
\end{description}
}
\end{table}
In the JS studies, for all but the 1973 crash, the LPPL has been fitted to the bubble in the raw index rather than to the log of the index.
For this to be justified, the inequality in Eq. \ref{small price rise} must hold. That is, the price rise during the bubble must be considerably less than the difference between the price at the beginning of the bubble and the fundamental price.
If we make the reasonable assumption that the fundamental price cannot be negative, then at any time during the bubble the price must at the very least not be more than double that at the beginning of the bubble.
If we compare the price at the time of fitting, prior to the crash, to that at the beginning of the bubble, then we see in Table \ref{tab:Hang Seng_raw_assumption} that this condition is met for only two of the eight bubbles found in the JS studies.
For the remaining six bubbles this condition does not hold, i.e. the price more than doubled during the bubble, so the inequality in Eq. \ref{small price rise}, which is the assumption upon which the raw rather than the log of the index can be chosen, was violated. Despite this, in the JS studies five of these six fits of the LPPL are made to the raw index rather than to its log; they should not have been.
\subsection{Sensitivity to search parameter values}
\label{ss:sensitivity}
Identifying an LPPL fit to a bubble as one that precedes a crash depends on the two critical parameters $\beta$ and $\omega$; so it is important to examine how sensitive the RMSE of the fit is to variations in these parameters.
We use the bubble preceding the 1989 crash on the Hang Seng to examine the sensitivity of the RMSE of the LPPL fit to variations in each of the four search parameters ($\beta, \omega, \phi$ and $t_c$); the other three parameters ($A, B$ and $C$) are always set using these four (see Section~\ref{s:fitting}). The results are shown in Figure~\ref{fig:sensitivity_1989_phi.jpg}. The circle indicates the chosen parameter value.
While the chosen values of the search parameters are at global minima, the RMSE is highly sensitive to small fluctuations in the value chosen for $\omega$.
The sensitivity diagrams for the other Hang Seng bubbles listed in Table~\ref{tab:Hang Seng_raw_assumption} are similar to those shown in Figure~\ref{fig:sensitivity_1989_phi.jpg}.
In general, the search procedure can easily get trapped in a local minimum for the $\omega$ parameter. Consequently the value found by the search procedure for $\omega$ may not be the one that leads to the minimum RMSE.
As the value of $\omega$ is used in predicting whether or not the bubble will be followed by a crash, this is a serious problem.
\begin{figure}[t]
\centering
\resizebox{\textwidth}{!}{\includegraphics{sensitivity_1989_phi.pdf}}
\caption{Sensitivity of the RMSE to the parameters of the LPPL for 1989 Hang Seng crash.}
\label{fig:sensitivity_1989_phi.jpg}
\end{figure}
\subsection{The `best' fits of the LPPL}
We now fit the LPPL to the raw data for each of the bubbles preceding the 11 crashes identified for the Heng Seng index over the period 1970 to 2008 (as selected by the criteria in Section~\ref{ss:what is a crash}), using the minimum RMSE as the criterion for best fit. For each crash:
\begin{table}[t]
\begin{center}
\caption{The bubbles and crashes of the Hang Seng index and LPPL fits to the raw bubble data.}
\label{tab:Hang Seng_fits}
~\\
{\small
$\begin{array}{rrrrr|r|r|rrr}
& &A &B &C &\beta &\omega &t2c &\phi &\mathrm{RMSE}\\
& &\mbox{HSI}&\mbox{HSI}& & &\mbox{rads} &\mbox{days} &\mbox{rads} &\mbox{HSI} \\
\mbox{Bubble:} &\mbox{Ref:} &\mbox{low:} & & &0.15 &4.80 &1 &0 &\\
\mbox{from/to} & &\mbox{high:} &0 & &0.51 &7.92 &? &\pi &\\
\hline
\mbox{*10-Mar-1971}&\mbox{SJ}
&594&-132 &-0.033 &0.20 &4.30 &7 &0.50 &7.58 \\
\mbox{20-Sep-{\bf 1971}} & &539 &-101 &-0.047 &0.22 &4.30 &3 &0.25 &6.11 \\
\hline
\mbox{22-Nov-1971}&\mbox{SJ}
&11 &-3 &0.003 &0.11 &8.70 &2 &0.05 &0.0722 \\
\mbox{09-Mar-{\bf 1973}} &log &65 &-56 &-0.001 &{\bf 0.01} &{\bf 11.1} &20 &1.32 &0.0538 \\
&log &8 &-0 &-0.177 &0.57 &{\bf 1.47} &2 &3.14 &0.0549 \\
&raw &2443 &-485 &-0.114 &0.26 &{\bf 1.45} &2 &3.14 &40.91 \\
\hline
\mbox{*13-Jan-1978}&\mbox{SJ}
&816 &-50 &-0.053 &0.40 &5.90 &6 &0.17 &10.09 \\
\mbox{04-Sep-{\bf 1978}} & &741 &-23 &0.072 &0.51 &5.30 &1 &0.00 &10.12 \\
\hline
\mbox{20-Nov-1978}&\mbox{SJ}
&1998 &-231 &-0.044 &0.29 &7.24 &3 &1.80 &46.72 \\
\mbox{13-Nov-{\bf 1980}} & &41164 &-38080 &0.001 &{\bf 0.01} &7.51 &52 &3.06 &35.02 \\
& &7929 &-5352 &0.008 &{\bf 0.05} &6.79 &26 &1.55 &35.55 \\
& &1998 &-231 &-0.044 &0.29 &7.24 &3 &2.63 &37.00 \\
\hline
\mbox{12-Dec-1980}& & & & & & & & & \\
\mbox{17-Jul-{\bf 1981}}& &1753 &-0 &-0.890 &{\bf 2.41} &{\bf 3.02} &1 &3.14 &40.46 \\
& &1817 &-3 &-0.567 &{\bf 1} &4.75 &12 &0.35 &49.24 \\
& &1946 &-11 &-0.399 &{\bf 0.76} &5.89 &36 &0.00 &54.95 \\
\hline
\mbox{*23-Jul-1984} &\mbox{JS}
&5262 &-542 &-0.007 &0.29 &5.60 &22 &1.60 &133.86 \\
\mbox{01-Oct-{\bf 1987}}& &5779 &-711 &0.048 &0.27 &5.68 &34 &2.63 &68.47 \\
\hline
\mbox{07-Dec-1987}&\mbox{SJ}
&3403 &-32 &-0.023 &0.57 &4.90 &34 &0.50 &133.21 \\
\mbox{15-May-{\bf 1989}} & &3575 &-53 &-0.195 &0.52 &4.95 &31 &1.74 &76.33 \\
\hline
\mbox{*19-Aug-1991}&\mbox{JS}
&21421&-7614 &0.024 &0.12 &6.30 &4 &0.60 &322.80 \\
\mbox{04-Jan-{\bf 1994}} & &212635 &-194575 &-0.002 &0.27 &5.95 &1 &3.13 &272.82 \\
& &14038 &-1717 &-0.028 &0.26 &6.43 &4 &3.14 &281.36 \\
\hline
\mbox{23-Jan-1995} &\mbox{JS}
&20359&-1149 &-0.019 &0.34 &7.50 &51 &0.80 &531.79\\
\mbox{07-Aug-{\bf 1997}} & &20255 &-1201 &-0.048 &0.33 &7.47 &51 &2.29 &438.79 \\
\hline
\mbox{13-Aug-1998} & & & & & & & & & \\
\mbox{28-Mar-{\bf 2000}} & &21918 &-16 &0.073 &{\bf 1.00} &{\bf 18.35}&290&0.00 &710.99 \\
& &24095 &-97 &-0.057 &{\bf 0.76} &{\bf 17.51}&264&3.14 &720.17 \\
& &19503 &-372 &0.111 &0.52 &5.7 &9 &2.07 &744.15 \\
\hline
\mbox{23-Apr-2003} & & & & & & & & & \\
\mbox{30-Oct-{\bf 2007}} & &38940 &-6408 &0.019 &0.20 &5.41 &1 &3.14 &693.61 \\
\hline
\end{array}
$
}
\end{center}
{\small\begin{description}
\item[$\rm HSI$] is the units of the Hang Seng Index.
\item[$\rm Ref:$] JS denotes Johansen and Sornette \cite{JS:2001}; SJ denotes Sornette and Johansen \cite{SJ:2001}.
\item[*] Bubble beginning moved to a time later than the trough between peaks.
\item[$\beta$] was constrained to be $\ge 0.01$, so a $\beta=0.01$ indicates that the optimal value of $\beta\le 0.01$.
\item[$t2c$] number of days from date of the fit until predicted crash date, i.e.
$t2c = t_c$ - today.
\item[{\bf Bold}] highlights those values of $\beta$ and $\omega$ that fall well outside the range specified in Eq.~\ref{lppl}, shown in the top two rows of the table.
\end{description}}
\end{table}
\clearpage
\LI
\item The first line of Table~\ref{tab:Hang Seng_fits} shows the parameters of the LPPL fit as given in the JS studies, but with the linear parameters $A, B$ and $C$ recalculated for time expressed in days rather than years.
As the RMSE was not reported for the JS studies (except for the LPPL fitted to the bubble preceding the 1997 crash) this too has been recalculated by us.
\item The second line shows the parameters for our best fit to the raw data.
The results are based on the raw data, despite our reservations about its appropriateness (Section~\ref{ss:index}), because we want to compare our results with those of the JS studies.\footnote{
For the crash of 1973 Johansen and Sornette \cite{JS:2001} used the log instead of the raw index, so we report both log and raw fits specifically for that year.}
\item If this is not within the bounds for a crash prediction, then subsequent lines show the next best fit that is (or might be).
\end{list}
Variation in the values of the critical parameters $\beta$ and $\omega$ sufficiently large to take them across their acceptable boundaries lead to only quite small fluctuations in the RMSE. This can bee seen, for example, for the crashes of 1973 and 1980 (see Table~\ref{tab:Hang Seng_fits}).
We were interested in comparing our LPPL fits to those found in the JS studies. However, given the high sensitivity of the RMSE to small changes in the value of $\omega$ (see Section~\ref{ss:sensitivity}) and as the values for $\beta$ and $\omega$ were reported to only one decimal place in the JS studies, our re-calculated RMSEs will be different from those that were obtained in these studies.
We can see this in the bubble ending in the crash of 1997, where we have not only our recalculated RMSE using the parameters rounded to one decimal place, but also the RMSE using the unrounded parameter values as found by Johansen et al. \cite{JLS:2000}; the latter fit is considerably better than our recalculation (RMSE=436 rather than 532 Hang Seng Index units). This improvement is almost certainly due to using the exact rather than the rounded value of $\omega$.
So caution needs to be taken when comparing the RMSEs for the fits reported in the JS studies and our fits.
Of the eight pre-crash bubbles fitted in the JS studies we find virtually the same parameters for the LPPL for six of them;
namely, those preceding the crashes of 1971, 1978, 1987, 1989, 1994 and 1997.
However, for their other two bubbles we found different parameters as follows:
\begin{description}
\item[1973:] For this bubble, Sornette and Johansen \cite{SJ:2001} report the fit to the log of the Hang Seng index, rather than to the raw index. We have used both the log and the raw index. When we fit the log of the index we find a better fit than that reported in \cite{SJ:2001} with values of both $\beta$ and $\omega$ outside their acceptable ranges.
For comparison with other bubbles we also fitted the raw index; we find that the best fitting LPPL has a value for $\beta=0.26$, which is within the acceptable range of 0.15 -- 0.51, but for $\omega=1.45$, which is well below the lower bound of its critical range of 4.8 -- 8.0 (see Equation~\ref{lppl}).
\item[1980:] We were able to reproduce the fit reported in \cite{SJ:2001}, with a crash predicted 3 days later, but it was not the best fit that we found.
Our best fit predicted a crash after 52 days, and had critical parameter values $\omega= 7.51$, which is acceptable, but $\beta=0.01$, which is outside the acceptable range.
\end{description}
There are three pre-crash bubbles that were not considered in the JS studies; one, in 1981, they did not consider a crash (but see Section~\ref{ss:what is a crash}), and two others were later than their period:
\begin{description}
\item[1981:] We find a best fit for which both $\beta(=2.41)$ and $\omega(=3.02)$ are well outside their acceptable ranges.
As $\beta > 1$, we surmise that this fit would have been rejected by the criteria used in the JS studies (see Section~\ref{s:fitting}).
The first fit that has a $\beta <= 1$ has $\omega=4.75$, which is just acceptable, but with a $\beta=1$, i.e. no power law, so well outside its acceptable range.
It might be argued that this peak was too soon (8 months) after the trough following the previous crash of 1980 for an LPPL to be fitted on the grounds of there being insufficient data. But, as we have argued in Section~\ref{ss:what is a crash}, we believe it should have been.
\item[2000:] Our best fit to the bubble has both critical parameters $\beta (=1.00)$ and $\omega (=18.35)$ well outside their respective acceptable ranges.
There is a fit that does have these parameters within their acceptable ranges, and predicts a crash after only 9 days; but it is not the best fit.
\item[2007:] Our best fit to this bubble has parameters well within the ranges required for a crash and the crash is predicted for the day it actually occurred.
\end{description}
\section{Conclusion}
\label{s:conclusion}
The LPPL for pre-crash bubbles on stock markets, as reported in Johansen et al. \cite{JLS:2000} and the JS studies, has important consequences. Our analysis has led us to the following conclusions.
The mechanism proposed to lead to the LPPL fluctuations as reported in Johansen et al. \cite{JLS:2000} must be incorrect as it requires the price to be increasing throughout the bubble (a constraint recognized later in \cite{SZ:2006}), but in half the studies reported the LPPL fitted to the index (or its log) decreases at some point during the bubble. Hence either another explanation is required or the fits have to be redone with a constraint on the parameters that leads to LPPL fits that never decrease.
Also, in the JS studies the fits were made to the raw rather than the log of the index for all but one (1973) of the eight bubbles, even though the assumption upon which the use of the raw rather than the log should be used was certainly {\bf not} met in six of these seven bubbles.
So, on both counts, these studies should no longer be used to support a conclusion that the proposed mechanism underlies the LPPL.
Identifying crashes and bubble beginnings was not well specified in the JS studies. In particular it is not clear why one peak, that of 1981, was not identified as a crash initiator. Moreover, moving the trough that marks the beginning of a bubble forward by `eye' in half the data sets is not really satisfactory. While we have taken more care in identifying those peaks that initiated crashes, we have still, for comparison, used the same bubble beginnings as used in the JS studies; in future a clear criterion needs to be established.
In the JS studies, the fits of the LPPL to the data were only accepted if the exponential parameter $\beta$ was $< 1$. That is, the fits showed an exponential {\em increase}. It would be stronger to reject the LPPL if a $\beta \ge 1$ is found.
In our study the two critical parameters of the fitted LPPLs, $\beta$ and $\omega$, do fall within acceptable ranges in 7 of the 11 bubbles.
Of the remaining four bubbles, an LPPL with critical parameters within their respective acceptable ranges could be found for all but one crash (1973).
However, these LPPLs did not have the best fits (minimum RMSE).
For one crash (1980) the best fit would be acceptable if the lower end of the acceptable range of $\beta$ was decreased, i.e. a range of 0.01 -- 0.51.
For another (1981), a fit with $\beta>1$ would also have to be ruled out to save the hypothesis.
For two crashes (1973 and 2000), there seems to be no saving strategy. That the bubbles leading to the 1981 and 2000 crashes do not satisfy the criteria is particularly negative as these are two of the three crashes for which the ranges on the critical parameters were not set {\em post hoc} in the JS studies.
Finally, while the objection that with seven parameters a curve can be fitted to any data \cite{LPC:1999} is not directly relevant as no goodness of fit is measured here, it is indirectly highly relevant. The RMSE of the fit of the LPPL model (Eq. \ref{lppl}) to the data is highly sensitive to small but not to large fluctuations in one of the critical parameters ($\omega$); this makes the search for the LPPL that minimizes the RMSE unreliable. Moreover, substantial fluctuations in both parameters together can result in quite small changes in the RMSE. This suggests that the permissible ranges for these parameters should not be independent of one another.
Despite these criticisms, and because of the partial success of the hypothesis, in particular for predicting the 2007 crash, we believe that it is worth investigating whether fitted LPPLs with critical parameters in acceptable non-independent ranges can be used to give a probabilistic, rather than an all-or-none prediction of an impending crash.
This will be the subject of on going work.
\newpage
|
\section*{Acknowledgments}
We would like to thank the anonymous referees for the helpful
suggestions. R. da Silva and G. I. Wirth are financially supported
by the following projects of CNPq: 490440/2007, 480258/2008-2,
311343/2006-6, and 577473/2008-5.
|
\section{Introduction}
The ultimate goal of nanotechnology is control over molecular-scale mechanical and electronic components. An oft-proposed route to this goal is the construction of components from controllable single molecules. An important class of such molecules with obvious applications is formed by those that have properties that are reversibly and bi-stably modifiable by external stimuli -- so-called \emph{molecular switches}. One example is the azobenzene molecule (H$_{5}$C$_{6}$-N$\!\!=\!\!$N-C$_{6}$H$_{5}$), which can be bi-stably photo-isomerized between its planar, $C_{2h}$ symmetric \emph{trans} and torsioned-twisted, $C_{2}$ \emph{cis} conformers in both solution and gas-phase. The high yield and stability of this reaction have rendered azobenzene an archetype of molecular switch research with proposed technical applications including e.g.~light-driven actuators\cite{yu03} and information storage media\cite{liu90,ikeda95}.
\begin{figure}
\begin{centering}
\includegraphics[width=1\columnwidth]{pics/fig1}
\end{centering}
\caption{\label{fig1}
Perspective views of {\em trans} and {\em cis} azobenzene and TBA, together with an illustration of the diazo-bridge bond length $d_{\rm NN}$ and the dihedral CNNC angle $\omega$. The latter is defined as the smallest angle between two planes spanned by the $-\!{\rm N}\!\!=\!\!{\rm N}\!-$ bridge and $-$C$-$N$-$ and $-$N$-$C$-$ bonds to the first and second phenyl-ring, respectively. The C atoms of the phenyl-rings have been darkened to allow easier distinction of azobenzene backbone and functional butyl groups in TBA.}
\end{figure}
For many such applications, switching of molecules at solid interfaces -- adsorbed at metal surfaces, for example -- is of particular interest. Unfortunately however, the switching properties of azobenzene have proven highly sensitive to the adsorbate-substrate interaction: Even at nearly chemically inert close-packed noble metal surfaces, switching of surface-adsorbed azobenzene by light has never been achieved, and switching by excitation with a scanning tunneling microscope (STM) tip has been successful only at Au(111)\cite{choi06}. A natural route to restoring the adsorbate switching ability is to further decouple the frontier $\pi,n$ and $\pi^{*}$ orbitals responsible\cite{ishikawa01,cembran04} for the gas- or liquid phase photo-isomerization from the substrate electronic structure. With these frontier orbitals largely located at the central diazo ($-$N=N$-$) bridge an intuitive idea to achieve such a decoupling is to functionalize the molecule with bulky spacer groups that prevent a closer encounter of the photochemically active unit with the substrate. This is precisely the notion behind the arguably to date most studied such adsorbate, $3,3',5,5'$-tetra-\emph{tert}-butyl-azobenzene (TBA)
\cite{jung97,moresco01,alemani06}. TBA consists of azobenzene functionalized with four {\em tert}-butyl (-C-(CH$_{3}$)$_{3}$) groups in the phenyl-ring meta positions as illustrated in Fig. \ref{fig1}. These 'table legs' were indeed found to enhance the switching efficiency of the adsorbed species, as e.g. indicated by the successful TBA switching by light at Au(111) \cite{hagen07,comstock07}. However, attempts at STM-tip induced switching at ostensibly comparable substrates such as Ag(111)\cite{tegeder07} and Au(100)\cite{alemani08} have been unsuccessful, indicating that TBA is not significantly more robust to specifics of the substrate interaction than pure azobenzene.
These circumstances beg the question, how and to what degree the TBA butyl groups really 'decouple' the photochromic moiety from the substrate. For a corresponding atomic-scale understanding the detailed characterization of the adsorbate geometry and binding constitutes a prerequisite and is the main objective of the present contribution. In contemporary surface science, corresponding analyses are increasingly performed by quantitative first-principles electronic structure calculations. Particularly density-functional theory (DFT) with present-day local or semi-local exchange-correlation (xc) functionals has developed into an unparalleled workhorse for this task, with often surprising accuracy particularly with respect to structural properties of the surface adsorption system. For TBA at Au(111) corresponding calculations are already challenged by the extension of the functionalized molecule and the simultaneous necessity to describe the metal band structure within a periodic supercell approach. On a more fundamental level, the real limitation comes nevertheless from sizable dispersive van der Waals (vdW) contributions to the surface chemical bond as characteristic for organic molecules containing highly polarizable aromatic ring systems \cite{jenkins09}. With local and semi-local xc functionals inherently unable to account for such contributions \cite{kristyan94} skewed, if not qualitatively wrong results must therefore be suspected. We have recently quantified this for azobenzene at the close-packed coinage metal surfaces \cite{mcnellis09,mcnellis09_2}. Comparison to detailed structural and energetic reference data from normal-incidence x-ray standing wave (NI-XSW) and temperature programmed desorption (TPD) measurements for azobenzene at Ag(111) demonstrates that the prevalent semi-local DFT xc treatment leads indeed to a significant underbinding with key structural parameters deviating by more than 0.5\,{\AA} \cite{mercurio10}.
For system sizes as those implied by the adsorption of azobenzene or TBA an appealing and computationally tractable possibility to improve on this situation is a semi-empirical account of dispersive interactions within the framework of so-called DFT-D schemes \cite{mcnellis09_2,mercurio10,wu02,grimme04,jurecka07,atodiresei08,tkatchenko09}. In this approach the vdW interactions not described by present-day xc functionals are approximately considered by adding a pairwise interatomic $C_6 R^{-6}$ term to the DFT energy. At distances below a cut-off, motivated by the vdW radii of the atom pair, this long-range dispersion contribution is heuristically reduced to zero by multiplication with a short-range damping function. While the applicability of this approach to adsorption at metal surfaces is uncertain ({\em vide infra}), our recent benchmark study for azobenzene at Ag(111) revealed that in particular the most recent DFT-D scheme due to Tkatchenko and Scheffler (TS) \cite{tkatchenko09} yields excellent structural properties, albeit at a notable overbinding \cite{mercurio10}.
In this contribution we further explore the generality of this finding by analyzing the adsorption geometry, vibrations and energetics of TBA at Au(111). Comparison against our recent near edge x-ray absorption fine structure (NEXAFS) \cite{schmidt10} and high-resolution electron energy loss spectroscopy (HREELS) \cite{ovari07} measurements, as well as a complete analysis of new TPD data confirms the accurate structural and vibrational predictions reached by the DFT-D TS scheme. The again obtained significant overbinding furthermore supports the interpretation \cite{mercurio10} that the neglect of metallic screening is the main limitation in the application of this scheme to the adsorption of organic molecules at metal surfaces. Comparing the TBA data to those for pure azobenzene at Au(111) we find a qualitatively different adsorption geometry for the functional backbone in the case of the {\em cis} isomer. For the {\em trans} isomer, on the other hand, we determine an intriguing structural and vibrational insensitivity of the photochemically active central diazo-bridge to the presence of the bulky spacer groups. The role of the latter for the switching efficiency is therefore more subtle than simple geometric decoupling and will be the topic of continuing work in our groups.
\section{Method}
\subsection{Theory}
The DFT-D methodology followed in this work corresponds exactly to that employed in our preceding work for the adsorption of azobenzene. We therefore restrict ourselves here to a brief account and refer to our previous publications \cite{mcnellis09,mcnellis09_2,mercurio10} for an in-depth description of the underlying concepts and technical details.
The DFT calculations were performed using a plane-wave basis set and library ultrasoft pseudopotentials\cite{vanderbilt90} as implemented in and distributed with the CASTEP\cite{clark05} code. The generalized gradient approximation (GGA) to the xc-functional with the parametrization suggested by Perdew, Burke and Ernzerhof\cite{perdew96} (PBE) was used throughout. In the spirit of the DFT-D approach, the lack of vdW interactions in this semi-local functional is approximately corrected with an additional analytical, two-body inter-atomic potential. At long range, this potential equals the leading $C_{6}R^{-6}$-term of the London series, where $R$ is the inter-atomic distance, and $C_{6}$ a so-called dispersion coefficient. At short range, this long range potential is matched to the DFT inter-atomic potential by a damping function $f(R,R^{0})$, typically modulated by the vdW radii $R^{0}$ of the atom pair. In this work we use the material-specific $C_{6}$ and $R^{0}$ parameters, as well as the damping function form suggested by Tkatchenko and Scheffler \cite{tkatchenko09} (henceforth denominating corresponding results as PBE+TS). This scheme accounts to some degree for the bonding environment through a Hirshfeld-analysis based adjustment of the $C_6$ parameters, which in our previous work on azobenzene at coinage metal surface gave a superior performance compared to other DFT-D schemes \cite{mcnellis09_2,mercurio10}.
The analytical form of the dispersion correction potential brings the advantage that dispersion-corrected total energies of geometries that are fully relaxed with respect to the dispersion-corrected forces can be obtained (employing an in-house extension to the CASTEP code) at essentially the same computational cost as a regular GGA calculation. On the other hand the assumption of a simple two-body form for the dispersion potential inherently neglects the effect of electronic screening of the vdW interactions\cite{rehr75}, which particularly for the here studied adsorption at metal surfaces is expected to lead to an overestimation of the binding energy \cite{mercurio10}. A second limitation of the semi-empirical DFT-D approach might arise for adsorbate molecules which also interact covalently with the substrate. This typically leads to molecule-substrate bond distances that are so short that the uncertainties in the heuristic damping function of the dispersion term may mingle in an uncontrolled way with deficiencies of the employed semi-local DFT xc functional.
The surface calculations were performed within supercells, using (111)-oriented metal slabs with $(6 \times 5)$ surface unit-cells and at least 18\,{\AA} vacuum. We verified that lateral interactions between adsorbed TBA and its periodic images are negligble at the GGA-PBE level. In the dispersion correction potential corresponding interactions were also switched off, so that our calculations should give a fair account of TBA adsorption in the low-coverage limit. As in our preceding work we neglected the subtle effects of the long-range Au(111) surface reconstruction. Full geometry optimizations (to within residual forces below 30 meV/{\AA}) of all molecular degrees of freedom were correspondingly performed with TBA adsorbed on one side of a four layer bulk-truncated slab. Test calculations with appropriately saturated {\em tert}-butyl groups indicated only a weakly expressed site specificity of these TBA functional groups at Au(111). This suggests that the photochemically active diazo-bridge plays a prominent role in anchoring the molecule at the metal substrate. The geometry optimizations were correspondingly initiated with the TBA diazo-bridge laterally placed as in the previously determined optimal adsorption geometry of azobenzene, corresponding to a 1:1 N-metal atom coordination \cite{mcnellis09_2}. Harmonic vibrational spectra of the adsorbed species in these relaxed geometries (and in the gas-phase) were subsequently obtained from numerical Hessians calculated by finite differences and neglecting any degrees of freedom of substrate atoms. To efficiently parallelize over the 432 displacements required for each spectrum, we have interfaced with the Atomistic Simulation Environment including the
'phonopy' extension to it \cite{ASE_phonopy}. Not aiming to reproduce the HREELS intensities \cite{ovari07}, simple Lorentzian broadening with a width of 1 meV was applied to the spectra for better visualization.
For the energetic and electronic structure calculations the thus determined relaxed adsorbate geometries were inverted in the bottom layer, forming inversion-symmetric seven layer slabs with adsorbates at both sides. This zeroes the internal dipole moment of the slab and results in a substantially improved substrate electronic structure. The two central energetic quantities obtained with the resulting structures are the adsorption energy
\begin{equation}
E_{{\rm ads}} \;=\; \frac{1}{2}\left[E_{{\rm azo@(111)}}-E_{(111)}\right]-E_{{\rm azo(gas)}} \quad,
\label{eq:E_ads}
\end{equation}
and the relative stability of adsorbed {\em cis} (C) and {\em trans} (T) conformer
\begin{equation}
\Delta E_{\mathrm{C-T}} \;=\; \frac{1}{2} \left[E_{{\rm azo@(111)}}(\mathrm{C}) - E_{{\rm azo@(111)}}(\mathrm{T})\right] \quad.
\label{eq:delta_E}
\end{equation}
Here $E_{\rm azo@(111)}$ is the total energy of the relaxed, double-sided azobenzene-surface system, $E_{(111)}$ the total energy of the clean slab, and $E_{\rm azo(gas)}$ the total energy of the correspondingly relaxed gas-phase isomer (computed within $\Gamma$-point sampled $(35\,{\rm \AA} \times 45\,{\rm \AA} \times 35\,{\rm \AA})$ supercells). Where applicable, TS DFT-D corrections calculated in optimized four layer slab and gas-phase geometries are added to $E_{{\rm azo@(111)}}$ and $E_{\rm azo(gas)}$, respectively. Both quantities were also consistently zero-point energy corrected with the previously obtained vibrational frequencies. In the convention of Eq. (\ref{eq:E_ads}) the adsorption energy of either {\em cis} or {\em trans} isomer at the surface is thus measured relative to its stability in the gas-phase at both pure PBE and dispersion corrected PBE+TS levels of theory, and a negative sign indicates that adsorption is exothermic. Consistently, a negative sign of $\Delta E_{\mathrm{C-T}}$ indicates a higher stability of the {\em cis} isomer. Convergence tests show that at the employed plane wave cutoff of 450\,eV and $(2 \times 3 \times 1)$ Monkhorst-Pack (MP) grid\cite{monkhorst76} both energetic quantities are numerically converged to within $\pm 30$\,meV.
\subsection{Experiment}
The TPD measurements were carried out under ultrahigh vacuum conditions. The Au(111) crystal was mounted on a liquid nitrogen cooled cryostat, which in conjunction with resistive heating enables temperature control from 90\,K to 750\,K. The crystal was cleaned by a standard procedure of Ar$^{+}$ sputtering and annealing. The TBA was dosed by means of a home-built effusion cell held at 380\,K at a crystal temperature of 260\,K. In the TPD experiments, the substrate was resistively heated with a linear heating rate of 1 K/s and desorbing TBA was detected with a quadrupole mass spectrometer at the TBA-fragment mass of 190
amu (3,5-di-{\em tert}-butyl-phenyl ion). This procedure was repeated for different dosing times corresponding to different initial TBA-coverages.
As further discussed below, three desorption features ($\alpha_{1}$--$\alpha_{3}$) are observed in the TPD. They are assigned to desorption from the multilayer ($\alpha_{1}$) and the first monolayer (ML) ($\alpha_{2}$ + $\alpha_{3}$), where $\alpha_{2}$ represents the desorption of $\approx$ 10\% of the monolayer coverage (for details see Fig.~\ref{fig5} and Ref.~\onlinecite{hagen07}). The NEXAFS and HREELS measurements of Refs. \onlinecite{schmidt10,ovari07} discussed in Section III were performed at a coverage of 1.0 and 0.9\,ML, respectively, which were prepared by heating the multilayer-covered surface to 340\,K or to 420\,K.
\section{Results and Discussion}
\subsection{Adsorption Geometry}
\begin{figure}
\begin{centering}
\includegraphics[width=1\columnwidth]{pics/fig2}
\end{centering}
\caption{\label{fig2}
Side view of adsorbed {\em cis} TBA at Au(111), illustrating key structural parameters defining the adsorption geometry (see text): The vertical height $z_i$ of the two diazo-bridge N atoms, as well as the out-of-horizontal phenyl plane bend angles $\tilde{\omega}_i$.}
\end{figure}
Our previous work on azobenzene at coinage metal surfaces points to an understanding of the surface chemical bond in terms of a balance of four major contributions: A covalent bond between the diazo-bridge and the metal, the vdW attraction between the metal and the phenyl-rings, the Pauli repulsion between the phenyl-rings and the metal, and the energetic penalty due to the distortion of the gas-phase molecular geometry \cite{mcnellis09,mcnellis09_2}. This understanding should largely carry over to TBA at Au(111), with the bulky tert-butyl groups particularly adding to the vdW attraction and the molecular deformation upon adsorption. Key structural parameters to characterize the adsorption geometry for both {\em trans} and {\em cis} isomer are therefore the location and orientation of the central diazo-bridge, as well as the orientation of the planes of the two phenyl-rings with respect to the surface. As further illustrated in Figs. \ref{fig1} and \ref{fig2} we concentrate on the $-$N=N$-$ bond length $d_{\mathrm{NN}}$ and the vertical N atom$-$surface plane distances $z_{i}$ for the prior. Here, the indices $i=1$ and $i=2$ represent the value for the N atom connected to the phenyl-ring closer to and further away from the surface, respectively. Obviously, $z_1 = z_2$ reflects an diazo-bridge that is oriented parallel to the surface, as we obtain for the more symmetric {\em trans} isomer throughout. To specify the position of the phenyl-rings we focus on the CNNC dihedral angle $\omega$ and the out-of-horizontal bend angles $\tilde{\omega}_{i}$ of the two ohenyl planes, with the same convention for the index $i = 1,2$ to distinguish the closer and more distant phenyl-ring in the asymmetric adsorption geometry of. Fig. \ref{fig2}. Adapting to the convention employed in Refs. \onlinecite{mercurio10} and \onlinecite{schmidt10}, a dihedral angle of $\omega = 180^{\circ}$ indicates a planar TBA molecule, while $\tilde{\omega}_{i} = 0^{\circ}$ denotes a phenyl-ring that lies parallel to the surface plane.
\begin{table}
\begin{centering}
\begin{tabular}{l|ccccc}
& \multicolumn{5}{c}{{\em Trans} @ Au(111)} \\
& \multicolumn{2}{c}{$z_{1}=z_{2}$ ({\AA})} & $\omega$ ($^{\circ}$) & \multicolumn{2}{c}{$\tilde{\omega}_{1}=\tilde{\omega}_{2}$ ($^{\circ}$)} \\ \hline
TBA (PBE) & \multicolumn{2}{c}{3.97} & 172 & \multicolumn{2}{c}{9} \\
TBA (PBE+TS) & \multicolumn{2}{c}{3.22} & 169 & \multicolumn{2}{c}{12} \\
TBA (Exp.)\cite{schmidt10} & \multicolumn{2}{c}{$-$} & $-$ & \multicolumn{2}{c}{$4\pm5$} \\[0.1cm]
Azo (PBE+TS)\cite{mcnellis09_2} & \multicolumn{2}{c}{3.28} & 180 & \multicolumn{2}{c}{3} \\ \hline
& \multicolumn{5}{c}{{\em Cis} @ Au(111)} \\
& $z_{1}$ ({\AA}) & $z_{2}$ ({\AA}) & $\omega$ ($^{\circ}$) & $\tilde{\omega}_{1}$ ($^{\circ}$) & $\tilde{\omega}_{2}$ ($^{\circ}$) \\ \hline
TBA (PBE) & 3.25 & 3.74 & 10 & 21 & 79 \\
TBA (PBE+TS) & 2.34 & 2.85 & 8 & 24 & 81 \\
TBA (Exp.)\cite{schmidt10} & $-$ & $-$ & $-$ & $30\pm5$ & $90\pm5$ \\[0.1cm]
Azo (PBE+TS)\cite{mcnellis09_2} & 2.23 & 2.23 & 18 & 32 & 68 \\ \hline
\end{tabular}
\end{centering}
\caption{\label{table1}
Structural parameters as defined in Figs. \ref{fig1} and \ref{fig2}. Values for adsorbed TBA at PBE+TS level of theory are compared to corresponding data obtained at PBE level of theory, as well as for adsorbed pure azobenzene \cite{mcnellis09_2}. Additionally shown are the out-of-horizontal phenyl plane bend angles $\tilde{\omega}_{i}$ as determined recently by NEXAFS measurements \cite{schmidt10}.}
\end{table}
\begin{figure}
\begin{centering}
\includegraphics[width=1\columnwidth]{pics/fig3}
\end{centering}
\caption{\label{fig3}
Perspective view of the {\em trans} TBA adsorption geometry at Au(111).}
\end{figure}
Table \ref{table1} summarizes the detailed geometric parameters obtained from our calculations with the exception of the diazo-bridge bond length. For the latter we consistently compute only insignificant changes away from the TBA gas-phase value of 1.29\,{\AA} ({\em trans}) and 1.28\,{\AA} ({\em cis}) at both PBE and PBE+TS level of theory. While this indicates an overall minor activation of the $-$N=N$-$ bond upon adsorption, it also demonstrates an insensivity to the degree of vdW interactions accounted for in the calculations. Starting the more detailed presentation with the {\em trans} isomer, Fig.~\ref{fig3} displays a perspective view of the overall TBA adsorption geometry. In line with the interpretation of STM \cite{alemani06,comstock07,alemani08}, surface vibrational \cite{ovari07} and NEXAFS \cite{schmidt10} data, the central azobenzene moiety essentially maintains its gas-phase planarity, with its long axis aligned to about $5^\circ$ along the direction of close-packed atom rows on the (111) surface. Also consistent with the preferential orientation deduced from NEXAFS the tert-butyl groups contact the surface with one C-H bond pointing towards the substrate. At the pure PBE level of theory, the central diazo-bridge correspondingly floats at a rather high height of about 4\,{\AA} parallel to the surface. This is further away than for pure azobenzene, where this height was 3.5\,{\AA} \cite{mcnellis09_2}. With the PBE functional not providing attractive vdW components to the molecule-surface interaction, the main effect of the bulky tert-butyl groups conforms therefore with the intuitive expectation to simply lift the functional azobenzene backbone further away from the surface.
This picture is obviously prone to change, once some account of vdW attraction is added to the theoretical description. At the PBE+TS level of theory, the height of the diazo-bridge is indeed significantly reduced to 3.22\,{\AA}, which is nevertheless still large on the scale of a typical N-transition metal bond length. In contrast, the differential interaction with the surface is not significantly altered by the additional bonding component. The slight bending of the azobenzene moiety away from the ideal gas-phase planarity is basically the same in both the PBE and PBE+TS adsorption geometry. Here, the parallel orientation of the diazo-bridge to the surface plane and the corresponding slight upward bending of both phenyl-rings by $\sim 10^\circ$ are in very good agreement to the recent NEXAFS experiments \cite{schmidt10}. Indirectly, this corroborates some of the assumptions made in the NEXAFS data analysis, without which twice as large tilt angles would have resulted \cite{schmidt10}. On the other hand, due the insensitivity of both structural parameters due to the degree of dispersive interaction in the calculations this agreement with experiment does unfortunately not allow any conclusion as to the accuracy of the description provided by the semi-empirical TS scheme. This is particularly unfortunate, as the latter predicts an intriguing similarity of the diazo-bridge height for TBA (3.22\,{\AA}) and pure azobenzene (3.28\,{\AA}), cf. Table \ref{table1}. If correct, this obviously largely contradicts the afore mentioned intuitive view of the bulky spacer groups in terms of 'table legs' that decouple the photochemically active unit in a geometric sense.
\begin{figure}
\begin{centering}
\includegraphics[width=1\columnwidth]{pics/fig4}
\end{centering}
\caption{\label{fig4}
Perspective view of the {\em cis} TBA adsorption geometry at Au(111).}
\end{figure}
The similarity between functionalized and pure molecule does not extend to the adsorption geometry of the {\em cis} isomer at Au(111). As shown in Fig. \ref{fig4} this geometry is skewed for TBA with both the position and orientation of the two phenyl-rings distinctly different. This feature is consistently obtained at both PBE and PBE+TS level of theory, while for pure azobenzene both theoretical descriptions agreed on an adsorption structure with only a small sideward tilt away from a $C_2$ rotational symmetry around the central diazo-bridge. In the asymmetric adsorption mode of TBA the lower phenyl-ring is tilted out of the surface plane, yet without significant torsion that would place its two butyl groups at different heights above the surface. This is much different for the upper phenyl-ring, which does not only stand essentially upright, but is so torsioned that one of its butyl groups points largely towards the diazo-bridge and the other one away from it. Overall, the internal structure of the azobenzene backbone in this adsorbed state is therefore very similar to that in gas-phase TBA with even the CNNC dihedral angle $\omega$ not much affected by the surface potential. Another noteworthy feature of the skewed adsorption mode is the prominent tilt of the central diazo-bridge, i.e. in contrast to adsorbed {\em trans} and {\em cis} azobenzene and adsorbed {\em trans} TBA the two N atoms are at a notably different vertical height, cf. Table \ref{table1}. Qualitatively, this overall structure is perfectly consistent with the understanding reached in the recent STM \cite{alemani06,comstock07,alemani08} and NEXAFS \cite{schmidt10} measurements. As to the NEXAFS, the agreement is even quantitative in all respects. The determined out-of-surface-plane bend angles of both phenyl-rings agree very well with the assignments made, cf. Table \ref{table1}. This also extends to the inclination of the central CNNC plane with respect to the surface plane\cite{footnote}. For this, NEXAFS determines about $60^\circ$, close to the $61^\circ$ and $68^\circ$ determined at the PBE and PBE+TS level, respectively.
To some extent unfortunate and similar to the situation for the {\em trans} conformer, none of these qualitative features, as well as tilt and inclination angles are very sensitive to the description of vdW interactions in the calculations. Both PBE and PBE+TS give essentially identical results. The major feature introduced again by the account of vdW attraction is an essentially rigid downward shift of the entire molecule by about 0.9\,{\AA}. This is slightly larger than is the case for {\em trans} TBA ($\sim 0.8$\,{\AA}) and leads at least for the lower N atom of the diazo-bridge to a vertical height above the surface (2.34\,{\AA}) that is close to the value determined for pure {\em cis}-azobenzene (2.23\,{\AA}). With such prominent differences between PBE and PBE+TS, measurements of the vertical heights as was done by NI-XSW for pure azobenzene \cite{mercurio10} would therefore again provide a critical benchmark to judge on the accuracy of the account of vdW interactions introduced by the semi-empirical TS scheme. In turn, however, precisely the lacking sensitivity of the qualitative features, as well as tilt and inclination angles of both {\em trans} and {\em cis} adsorption geometries suggests that they are not much affected by the shortcoming of the employed semi-local xc functional with respect to dispersive interactions. This increases the confidence that a quite reliable understanding of the adsorption geometry of both conformers has been reached by the present calculations.
\subsection{Energetics}
\begin{figure}
\includegraphics[width=0.8\columnwidth]{pics/fig5}
\caption{Thermal desorption spectra (raw data) of TBA adsorbed on Au(111) at different coverages as recorded with a linear heating rate of 1 K/s at the fragment-mass of 190 amu ((C$_{4}$H$_{9})_{2}$-C$_{6}$H$_{4}^{+}$). The inset shows the desorption rate (ln$(d r_{\rm des}/dt$) plotted against the reciprocal temperature at a coverage $\theta$ of 0.1 monolayer (ML). The slope of the line, -$E_{\rm des}/R$ with $R$ the gas constant, then determines the activation energy for desorption $E_{\rm des}$ at this coverage, in this case $E_{\rm des}$(0.1 ML) = 1.65$\pm$0.08 eV.}
\label{fig5}
\end{figure}
In view of the preceding study on azobenzene \cite{mercurio10} it is clear that the binding energetics provided by the semi-empirical DFT-D approach deserves particular scrutiny. We performed TPD measurements to obtain an experimental reference value for the binding energy of the thermodynamically favored {\em trans} TBA on Au(111). Figure \ref{fig5} shows the TPD spectra as a function of TBA surface coverage recorded at the fragment mass of 190 amu ((C$_{4}$H$_{9})_{2}$-C$_{6}$H$_{4}^{+}$), with a linear heating rate of 1\,K/s. In the low coverage regime a broad desorption peak ($\alpha_{3}$) is observed around 542\,K which extends to lower temperature with increasing coverage, as has also been observed for TBA on Ag(111) \cite{tegeder07} and other azobenzene (derivatives) on noble metal surfaces \cite{ovari08,mercurio10}. This behavior is attributed to repulsive interactions between the adsorbed molecules (for example due to dipole-dipole interactions).
After saturation of the desorption peak $\alpha_{3}$ a second feature $\alpha_{2}$ develops at 394\,K. Further increase in coverage leads to saturation of this peak and to the appearance of a new feature around 308\,K labeled as $\alpha_{1}$. The $\alpha_{1}$ peak increases in height and width with increasing coverage, showing a typical zero-order desorption behavior (data not shown). We therefore assign the $\alpha_{1}$ peak to desorption from the multilayer, while $\alpha_{2}$ and $\alpha_{3}$ are associated with desorption from the monolayer. The $\alpha_{2}$ component represents the desorption of about 10 \% of the monolayer coverage and is attributed to desorption out of a densely packed TBA structure \cite{hagen07}.
In order to derive the activation energy for desorption, $E_{\rm des}$, of TBA on Au(111) in the low-coverage regime from this TPD data we utilize the so-called complete analysis \cite{christmann91}, which has the advantage that knowledge about the absolute coverage is not required. For this analysis a family of TPD curves are measured as a function of initial coverage ($\theta_{i}$) as shown in Fig. \ref{fig5}. These curves are subsequently used to construct a family of $\theta(t)$ curves via $A_{p}\propto \int_{0}^{\infty} r_{\rm des}(t)dt= \theta$ with $A_{p}$ the peak area and $r_{\rm des}$ the desorption rate. Due to the known linear heating rate (in our case 1\,K/s), this corresponds to a knowledge of $\theta(T_{s}$) as a function of $\theta_{i}$, where $T_{s}$ is the surface temperature. Following this, an arbitrary coverage value $\theta_{1}$ is chosen that is contained in each of the desorption curves. The desorption rate at this coverage, $r_{\rm des}(\theta_{1})$, and the temperature at which this rate was obtained, $T_{1}$, are then read off from each of these desorption curves. Plotting ln$[r_{\rm des}(\theta_{1}$)] versus $1/T_{1}$ (Arrhenius-plot) finally allows to determine $E_{\rm des}(\theta_{1}$) as follows directly from the Polanyi-Wigner equation \cite{christmann91}. This is exemplarily shown for a coverage of 0.1\,ML in the inset of Fig. \ref{fig5}, which yields an activation energy of 1.65$\pm$0.08\,eV.
\begin{table}
\begin{centering}
\begin{tabular}{ll|ccc}
& & {\em Trans} & & {\em Cis} \\
& & $E_{\mathrm{ads}}$ & $\Delta E_{\mathrm{C-T}}$ & $E_{\mathrm{ads}}$ \\ \hline
Gas-phase & PBE & $-$ & 0.58 & $-$ \\
& PBE+TS & $-$ & 0.29 & $-$ \\ \hline
@ Au(111) & PBE & $-0.16$ & 0.52 & $-0.21$ \\
& PBE+TS & $-3.00$ & 0.99 & $-2.07$ \\ \hline
Experiment& & $-1.70\pm0.1$ & $-$ & $-$ \\ \hline
\end{tabular}
\par\end{centering}
\caption{\label{table2}
Energetics of TBA in gas-phase, and adsorbed at Au(111). All numbers are in eV.}
\end{table}
In the low-coverage regime $\leq$ 0.2 ML this analysis determines an essentially constant desorption energy of $E_{\rm des}=1.70\pm$0.1\,eV. Assuming no additional activation barrier for adsorption, this then provides the aspired benchmark number against which the calculated $E_{\rm ads}$ for {\em trans} TBA may be referenced. As apparent from Table \ref{table2} the values obtained at the PBE and PBE+TS level are very consistent with the findings of the previous studies of benzene and pure azobenzene at coinage metal surfaces \cite{mcnellis09_2,mercurio10}: The lack of vdW interactions in the semi-local PBE functional leads to an overall negligible binding of both TBA isomers, with the actual values for $E_{\rm ads}$ in fact essentially identical to those computed for pure azobenzene at Au(111) before \cite{mcnellis09_2}. Adding dispersive attraction within the DFT-D approach these binding energies are dramatically increased, such that in the end the intended dispersion 'correction' amounts to more than 90\,\% of the total binding energy. Compared to the experimental reference the PBE+TS approach overbinds the {\em trans} isomer by about as much as PBE underbinds. In absolute numbers this is a disconcerting deviation of more than 1\,eV. In our previous work we had assigned the corresponding overbinding determined for pure azobenzene to the neglect of metallic screening in the DFT-D approach\cite{mercurio10}. This argument was largely based on the intriguing accuracy of the PBE+TS adsorption geometry compared to the bond distances derived from NI-XSW measurements. As discussed above, such a conclusion on the determined TBA adsorption geometry is presently not possible, as all hitherto measured structural parameters are not very sensitive to the additional vdW attraction provided by the PBE+TS scheme. On the other hand, the additional butyl groups lead to a significantly increased overall binding energy of TBA compared to azobenzene, and in both cases PBE+TS overbinds by roughly 40\,\% (azo: 1.71\,eV vs. 1.08\,eV \cite{mercurio10}; TBA: 3.00\,eV vs. 1.70\,eV). Such a rather geometry-unspecific deviation between theoretical and experimental data quite well fits the anticipated effect of an overestimated uniform background potential. The latter is attributed to the neglect of screening of vdW contributions from more distant substrate atoms.
For gas-phase TBA PBE+TS yields a somewhat smaller {\em cis}-{\em trans} energy difference $\Delta E_{\rm C-T}$ than for pure azobenzene, 0.29\,eV compared to 0.49\,eV\cite{mcnellis09_2}, respectively. This arises predominantly from the additional vdW attractions due to the butyl groups in the bend {\em cis} conformer, cf. Fig. \ref{fig1}, and is not found at the PBE level of theory, where $\Delta E_{\mathrm{C-T}}=0.58$\,eV for both TBA and azobenzene. Compared to the gas-phase reference, the stability difference of the two TBA conformers is with 0.99\,eV substantially increased at Au(111) at PBE+TS level of theory. This is primarily due to the larger vdW attraction possible in the planar {\em trans} adsorption mode and was equivalently obtained for pure azobenzene adsorption \cite{mcnellis09_2}. This interpretation in terms of vdW is corroborated by the essentially unchanged $\Delta E_{\mathrm{C-T}}$ of gas-phase and adsorbed TBA at PBE level of theory, where these bonding contributions are absent, cf. Table \ref{table2}. Not withstanding, in light of the suspected overestimation of the vdW attraction within PBE+TS we also expect that this increase of $\Delta E_{\mathrm{C-T}}$ upon adsorption is overestimated. Extrapolating the roughly 40\,\% overshoot seen in the {\em trans} TBA binding energy, we would therefore cautiously conclude on only a moderate $\sim 0.3-0.4$\,eV increase of the {\em cis}-{\em trans} energy difference upon adsorption of TBA at Au(111). While the therewith implied higher stability of adsorbed {\em trans} TBA is consistent with the existing experimental data, there are to date unfortunately no dedicated measurements of $\Delta E_{\mathrm{C-T}}$ against which this estimate could be compared.
\subsection{Vibrations}
\begin{figure}
\includegraphics[width=1.0\columnwidth]{pics/fig6}
\caption{Comparison of calculated surface vibrational modes of {\em trans} (top, red lines) and {\em cis} (center, blue lines) TBA at Au(111) at PBE (dashed lines) and PBE+TS (solid lines) level of theory to the corresponding HREELS spectra from ref. \onlinecite{ovari07} (bottom, black solid lines. Upper: {\em trans}, Lower: {\em cis}). Center inset: The highest energy peaks of the spectrum, shown at larger scale. Note the C-H stretch peak on the left. The spectra have been vertically displaced for clarity. Not aiming to reproduce the HREELS intensities, the computed modes are convoluted with a Lorentzian broadening of 1 (meV).}
\label{fig6}
\end{figure}
The picture arising from the geometric and energetic characterization of TBA at Au(111) points to a surface chemical bond that is predominantly due to unspecific dispersive interactions. Particularly for the {\em trans} conformer, the structural similarity of the azobenzene and TBA adsorption complex with respect to the photochemically active diazo-bridge moiety (together with the in both cases negligible $-$N=N$-$ bond activation) questions the oft-quoted function of the tert-butyl 'spacers' in terms of a geometric decoupling from the surface. Complementary information confirming this view can come from an analysis of the surface vibrational modes. Figure \ref{fig6} compares corresponding results computed for {\em trans} TBA at Au(111) at PBE+TS level of theory with experimental data from HREELS \cite{ovari07}. At first glance, we find overall agreement with all major features nicely reproduced by the calculations. As had already been noticed in the experimental work, the vibrational spectrum exhibits only minor changes between adsorbed {\em cis} and {\em trans} TBA and even TBA in the condensed phase \cite{ovari07,kuebler60, kellerer71}. This is similarly obtained in the calculations, as illustrated by the spectrum for {\em cis} TBA at Au(111) also shown in Fig. \ref{fig6}.
The detailed inspection of the C-H vibrational modes in the energy range of 350 to 400~meV reveals nevertheless the subtle influence of the Au(111) substrate on the vibrational spectra of the adsorbed species. Based on the calculations the three frequency bands centered at 368, 378, and 390\,meV are assigned to the symmetric and asymmetric C-H vibrations of the tert-butyl groups, as well as the C-H bending modes at the phenyl-rings, respectively. There is a clear change of the former two vibrational bands when going from the PBE to the PBE+TS description, cf. the inset of Fig. \ref{fig6}. The overall reduced distance of the TBA molecule to the gold substrate upon inclusion of the attractive vdW forces leads to a clear splitting of the C-H vibrations of the {\em tert}-butyl moieties. In the adsorbed species the frequencies of the C-H bonds pointing towards the substrate are red shifted by 5-10\,meV and the corresponding vibrations develop low-energy sidebands. These subtle but important changes are confirmed by the experimental HREELS spectra. Due to the mainly in-plane character of the C-H phenyl-ring bending modes we observe only a small shoulder at 390\,meV. In contrast the symmetric and asymmetric C-H stretch vibrations of the {\em tert}-butyl legs at 368 and 380\,meV result in pronounced dipole-active bands. Upon {\em trans} to {\em cis} isomerization the number of C-H bonds pointing towards and noticeably contacting the substrate is roughly halved, cf. Fig. \ref{fig2}. This leads in both experiment and theory to a reduction of the low-energy side bands best seen for the symmetric C-H stretch vibrations which constitute the lowest band at 368\,meV. As the splitting of the C-H stretch vibrations happens only at a reduced distance to the substrate (PBE vs. PBE+TS adsorption structure) it confirms the importance of vdW interaction in the bonding of TBA. Unfortunately this is not a quantitative benchmark for the accuracy of our semi-empirical PBE+TS approach.
More insight into the specific bonding of the diazo-bridge moiety can specifically come from a detailed analysis of the $-$N=N$-$ stretch mode. However, identically obtained with PBE and PBE+TS, the computed value of 184\,meV in adsorbed {\em trans} and {\em cis} TBA again reflects a negligible activation of the NN bond at the surface. Compared to the respective gas-phase values, the mode red-shifts only by about 3\,meV for both isomers. Correspondingly, the obtained shift between this stretch mode in {\em trans} and {\em cis} TBA at Au(111) is very similar to the one computed for free TBA, 6 vs. 5 meV, respectively. This in turn compares very well to the 8\,meV measured for TBA in the condensed phase with IR and Raman spectroscopy \cite{kuebler60,kellerer71}. Such a negligible change of the {\em cis}-{\em trans} stretch frequency difference upon adsorption is not found for pure azobenzene at Au(111). Here, the closer encounter of the diazo-bridge in the {\em cis} adsorption geometry softens the stretch mode by more than 6 meV. With the mode unaffected in adsorbed {\em trans} azobenzene, this leads to a concomitant increase of the {\em cis}-{\em trans} stretch frequency shift compared to the gas-phase. In {\em cis} TBA at Au(111) a corresponding softening does not occur as the diazo-bridge does not come as close to the surface in the skewed adsorption mode. In this respect, one can really attest some geometric decoupling effect to the butyl groups, yet only for the {\em cis} isomer. For the {\em trans} isomer though, an equivalent analysis of the diazo-bridge against surface stretch or other torsional azo modes shows always the same similarity between TBA and azobenzene at Au(111) as for the just discussed $-$N=N$-$ stretch mode. Also the characterization of the surface vibrational properties supports therefore the understanding that the effect of the bulky {\em tert}-butyl groups for the switching properties of {\em trans} TBA at Au(111) must be more subtle than a mere geometric decoupling of the central photochemically active molecular moiety.
\section{Summary and Conclusions}
A methodological motivation of this work was to further explore the capabilities of the semi-empirical dispersion correction approach to semi-local DFT in describing the adsorption of complex organic molecules at metal surfaces. For this we have presented a detailed characterization of the geometric, energetic and vibrational properties of the {\em trans} and {\em cis} isomers of the azobenzene derivate TBA at Au(111). The findings are in all respects in line with the experience from preceding studies on benzene and pure azobenzene at coinage metal surfaces \cite{mcnellis09,mcnellis09_2,mercurio10}: The additional account of attractive vdW interaction introduced by the PBE+TS scheme leads to a significant modification of the adsorption geometry, primarily in terms of bringing the molecule closer to the surface. The concomitantly dramatically increased binding energy is notably overestimated compared to the reference value from TPD measurements. For azobenzene at Ag(111) a corresponding overbinding -- in fact in relative terms very much comparable to the one found here for TBA at Au(111) -- was attributed to the neglect of metallic screening of dispersive interactions in the semi-empirical DFT-D approach \cite{schmidt10}. This argument was largely based on the very accurate PBE+TS adsorption geometry as compared to detailed structural data from NI-XSW measurements. Such a conclusion is, unfortunately, not yet possible in full for the here studied TBA at Au(111). The qualitative features of the determined {\em trans} and {\em cis} adsorption geometries and even detailed tilt and inclination angles agree all very well with existing data from STM \cite{alemani06,comstock07,alemani08}, HREELS \cite{ovari07} and NEXAFS \cite{schmidt10} experiments. The energetic lowering of the C-H stretch vibrations of the {\em tert}-butyl groups pointing towards the substrate is only observed in the PBE+TS approach. This points towards the importance of the vdW component introduced by the DFT-D approach. However, no experimental reference data exist to date for the vertical height of the adsorbed molecule, that would allow to directly confirm the present working hypothesis, namely that the PBE+TS approach is a very useful and computationally tractable tool to provide accurate structural information for adsorbed complex organic molecules.
If this working hypothesis proves correct and the determined PBE+TS adsorption geometries are indeed accurate, the prevailing preconception of the role of the bulky functional groups for the experimentally observed improved isomerization ability of adsorbed TBA needs revision. Motivated by the gas-phase structure, this view discusses the butyl groups as spacers that help to decouple the molecular switch from the metal substrate. For the TBA {\em cis} isomer we indeed compute a skewed adsorption mode, in which one phenyl-moiety is much more tilted, up to the point of standing essentially perpendicular to the surface plane. This is quite different to the more symmetric adsorption mode of {\em cis} azobenzene at Au(111), such that here one can attest some geometric decoupling from the surface due to the butyl 'table legs'. Quite distinctly, the almost planar adsorption mode of {\em trans} TBA is very much comparable to the adsorption geometry of pure azobenzene. Particularly for the photochemically active diazo-bridge moiety this similarity goes even down to essentially unchanged structural and vibrational properties. In this respect, a core message arising from the present study is that especially for the photo-switching of {\em trans} TBA the effect of the bulky spacer groups must be more subtle than the anticipated mere geometric decoupling of the functional backbone from the metal surface.
\section{Acknowledgements}
Funding by the Deutsche Forschungsgemeinschaft through Sfb 658 - Elementary Processes in Molecular Switches at Surfaces - is gratefully acknowledged. Ample computing time has been provided at the HLRB-II architecture of the Leibniz Supercomputing Centre in Garching.
|
\section{Introduction
Interference channels model a network of simultaneously communicating node pairs.
In these channels, each transmitter has data to send to only one receiver, which also observes interference from the other transmitters in the network.
Analysis of interference channels has shown that interference is not a fundamental limitation.
In particular, with any sized interference channel with any number of users,
the capacity for any given user will scale at half the rate of its interference-free capacity
in the high transmit power regime~\cite{CadJaf:Interference-Alignment-and-Degrees:08}.
The key to achieving a linear capacity scaling is interference alignment (IA)~\cite{MadMotKha:Communication-Over-MIMO:08,MadMotKha:Communication-over-X-channel::06}.
With IA, interfering transmitters precode their signals to align in the unwanted users' receive space, allowing these receivers to completely
cancel more interferers than they otherwise could. The signals can be aligned in any dimension, including time, frequency, or space.
This can be viewed as a cooperative approach because the transmitters neglect the performance of their own link to allow other users
to perfectly cancel interference.
This is in contrast to a provably suboptimal ``selfish'' approach where a transmitter
ignores the interference it causes and aims simply to maximize its own data rate~\cite{YeBlu:Optimized-signaling-for-MIMO:03}.
Interference alignment has been shown to achieve the maximum capacity scaling, also known as degrees of freedom, of the $K$-user interference channel,
but at finite transmit power it offers suboptimal achievable sum rate. Consequently, there is interest in finding precoders for the interference channel
that relax the perfect alignment constraint with the objective of obtaining better nonasymptotic sum rate performance.
Alternative IA precoder designs have been proposed for the single-antenna interference channel with time or frequency
selectivity~\cite{SheHosVid:An-improved-interference-alignment:08,ChoJafChu:On-the-beamforming-design-for-efficient:09}.
Closed-form IA precoders and achievable degrees of freedom for the multiple-input multiple-output (MIMO) interference channel with
infinitely selective channels have also been found for
some asymmetric antenna arrangements~\cite{GouJaf:Degrees-of-Freedom-of-the-K-User:08}.
Interference channels where precoding can only be done over one transmission slot are said to have constant or static coefficients.
In this case, the degrees of freedom with linear precoding are unknown but have been hypothesized to be less than that with infinite
selectivity~\cite{YetJafKay:Feasibility-Conditions-for-Interference:09,TreGuiRie:On-the-achievability-of-interference-alignment:09},
while non-linear precoding might achieve $KM/2$ degrees of freedom with $M$ antennas at each transmitter and
receiver~\cite{MotGhaMad:Real-interference-alignment::09,GhaMotKha:Interference-Alignment-for-the:09}.
A challenge in constant coefficient MIMO interference channels is that closed form solutions have been found in only a few special
cases~\cite{CadJaf:Interference-Alignment-and-Degrees:08}. Algorithmic techniques, such as alternating
minimization~\cite{CsiTus:Information-geometry-and-alternating:84}, have been proposed to find
precoders and explore possible
degrees of freedom for the general case~\cite{GomCadJaf:Approaching-the-Capacity-of-Wireless:08,GomCadJaf:Approaching-the-capacity-of-wireless:09,
PetHea:Interference-Alignment-Via-Alternating:09}.
Such algorithms are promising both for their ability to provide precoder solutions in a practical setting and their flexibility in application
to arbitrary networks for which closed-form solutions are unknown. The subspace algorithms
of~\cite{GomCadJaf:Approaching-the-Capacity-of-Wireless:08,GomCadJaf:Approaching-the-capacity-of-wireless:09,
PetHea:Interference-Alignment-Via-Alternating:09}, however, still use alignment as the main objective, which is asymptotically optimal
for the interference channel but has suboptimal throughput at finite SNR and other regimes. They also neglect colored
noise, possibly caused by co-channel interference from outside the coordinating nodes.
A maximum-SINR algorithm was proposed in~[12], but this algorithm does not optimize a global objective, assumes white Gaussian noise, and is not shown to converge.
In this paper we propose several alternative linear precoding designs for MIMO interference channels. While maximizing the sum rate is the primary
objective, we do not directly maximize sum rate due to analytical intractability. Instead we approximate a sum rate maximization via algorithms
with varying performance and complexity tradeoffs.
First, we derive a generalization of subspace alignment that includes colored noise, which biases the preferred alignment subspaces.
The resulting objective, which minimizes the interference plus noise leakage (INL),
results in orthogonal precoders amenable to quantized CSI. This algorithm is shown to be a special type of minimum mean squared error (MMSE) design,
and at high SNR or white noise at all receivers is shown to reduce to the IA subspace
methods of~\cite{GomCadJaf:Approaching-the-Capacity-of-Wireless:08,PetHea:Interference-Alignment-Via-Alternating:09}.
From this, interference alignment is shown to be an MMSE-type solution at infinite SNR, where interference-suppression filters are optimal.
As with previous forms of interference alignment, the proposed minimum INL algorithm does not consider
the signal power at any given user and is thus suboptimal with finite transmit power. Further, this algorithm and previous designs derive precoders but neglect
receiver design, which could be optimized jointly with the precoders.
Inspired by the connection between mutual information and mean-square error~\cite{GuoShaVer:Mutual-information-and-minimum:05},
we derive an explicit joint MMSE precoder/receiver design for the interference channel.
Although it does not directly maximize the sum rate, the joint-MMSE design results in a higher sum rate than subspace methods.
It does not lead to orthonormal precoders, making quantized feedback design more difficult.
The MMSE design is shown to be a generalization of previous approaches for point-to-point and multiuser
settings~\cite{Sal:Digital-transmission-over:85,YanRoy:On-joint-transmitter-and-receiver:94,SamPau:Joint-transmit-and-receive:99,
TenAdv:Joint-multiuser-transmit-receive:04,ZhaWuZho:Joint-linear-transmitter:05}.
Further, the design is more computationally complex and requires more iterations at high SNR than subspace designs. MMSE-based designs have also recently been developed independently
in~\cite{SchShiBer:Minimum-mean-squared:09,SheLiTao:The-new-interference-alignment-scheme:10}.
To more directly optimize the sum rate we formulate a maximum SINR algorithm, which is proven to converge via alternating minimization of a global performance function.
The maximum SINR algorithm derived in~\cite{GomCadJaf:Approaching-the-Capacity-of-Wireless:08} is
shown to be an approximation to the one derived in this paper. On average, the two are shown to have the same performance, but for any given channel
realization may result in different sum rates.
This design often has increased throughput relative to MMSE and subspace approaches, but finds nonorthogonal precoders and requires more
channel state information if run in a distributed manner.
In summary, this paper proposes three algorithms that span the tradeoff between performance and complexity for the static MIMO interference channel.
The minimum INL algorithm has the same complexity as previous work but has improved performance when colored noise exists at any
receiver. The joint-MMSE design has further rate enhancements regardless of the noise covariance matrices but has a computationally more complex optimization
and non-orthogonal precoders. The maximum SINR design has the best overall performance of all proposed strategies (in most cases, as shown in the simulations),
but requires more channel state information than the previous designs and also results in non-orthogonal precoders which are difficult to quantize in a practical
setting~\cite{LovHeaStr:Grassmannian-beamforming-for-multiple-input:03}. The proposed algorithms are then simulated alongside existing methods in regimes
previously unconsidered in the literature.
For example, the algorithms are simulated in an environment with an uncoordinated interferer that is not participating in the alignment protocol.
This colors the noise at each receiver, and if its power is scaled with the rest of the transmitters, resulting in reduced capacity scaling. Each of the algorithms can
outperform the others in different regimes, and each of these regimes is simulated and enumerated.
The rest of this paper is organized as follows: Section~\ref{sec:model} presents the system model under consideration;
Section~\ref{sec:algorithms} presents the new MMSE and INL algorithms and derives a
maximum SINR algorithm with proven convergence and analyzes each of the methods;
Section~\ref{sec:sims} presents simulations under uncoordinated interference and colored noise; and Section~\ref{sec:conclusion}
concludes the paper and gives directions for future work.
Before proceeding, we introduce notation. The $\log$ refers to $\log_2$. Bold uppercase letters, such as ${\bf A}$, denote
matrices, bold lowercase letters, such as ${\bf a}$, denote column vectors, and normal letters $a$ denote scalars.
The letter $\mathbb{E}$ denotes expectation, $\mathbb{C}$ is the complex field, $\mathbb{R}\{a\}$ is the real component of
complex scalar $a$, $\min\{a,b\}$ denotes the minimum of $a$ and $b$,
$\nu_{\rm min}^R\left({\mathbf{A}}\right)$ is the matrix whose columns are the eigenvectors corresponding to the $R$ smallest eigenvalues of matrix
${\mathbf{A}}$, $\mathrm{tr}\left({\mathbf{A}}\right)$ is the trace of matrix ${\mathbf{A}}$, $|a|$ is the magnitude of the complex number $a$,
$\|\bf a\|$ is the Euclidean norm of vector ${\bf a}$, and $\left|{\mathbf{A}}\right|$ is the
determinant of square matrix ${\mathbf{A}}$. ${\mathbf{A}}^*$ is the Hermitian transpose of matrix ${\mathbf{A}}$ and ${\mathbf{A}}^{-1}$ is its inverse.
The matrices ${\mathbf{I}}$ and $\bf 0$ are the identity matrix and all zero matrix, respectively,
of appropriate dimension. Finally, we use $\{{\mathbf{F}}_\ell\}$ when referring to the set of precoders and ${\mathbf{F}}_\ell$ when referring to the precoder
at transmitter $\ell$, and similarly for receive spatial filters $\{{\mathbf{G}}_k\}$.
\section{System Model}\label{sec:model}
Consider the $K$-user MIMO interference channel illustrated in Figure~\ref{fig:channel_model},
with $K$ transmit-receive pairs. A wireless channel links each receiver to each transmitter, but a given transmitter
only intends to have its signal decoded by a single receiver. The $k$th transmitter possesses $M_k$ antennas with which to transmit
$S_k\le M_k$ spatial streams, and the $k$th receiver (which
is to decode the signal from the $k$th transmitter) possesses $N_k\ge S_k$ antennas. In some analysis and simulations, all
users will have the same antenna configurations so that
$M_k=M, N_k=N$, and $S_k = S, \forall k$; we denote this symmetric case as an $(M,N,K)$ interference channel with $S$ streams per user.
This paper considers the narrowband MIMO interference channel where each link is static for the duration of a transmission,
but may change between successive transmissions. This is the block fading model where all the links in the network
are constant for the period of transmission, creating a tractable approximation to more realistic continuous fading models.
Linear precoding is done independently over each channel realization, favoring simplicity over the possible degrees of freedom gained by jointly precoding
over realizations.
This is the same model as previous work on algorithms for the interference
channel~\cite{GomCadJaf:Approaching-the-Capacity-of-Wireless:08,GomCadJaf:Approaching-the-capacity-of-wireless:09,
PetHea:Interference-Alignment-Via-Alternating:09}.
The transmission of all $K$ users is synchronized such that each begins and ends each transmission simultaneously, and no frequency
offsets exist in the network.
We therefore take the standard approach~\cite{GomCadJaf:Approaching-the-Capacity-of-Wireless:08,PetHea:Interference-Alignment-Via-Alternating:09}
and focus on the transmission of a single vector symbol ${\mathbf{s}}_k$ from transmitter
$k\in\{1,\dots,K\}$, neglecting any time dependency.
\begin{figure}
\centering
\includegraphics[width=3.5in]{MIMO_Interference_Channel.pdf}
\caption{The MIMO interference channel. Each transmitter is paired with a single receiver, and all links are non-negligible.}
\label{fig:channel_model}
\end{figure}
Transmitter $k$ uses linear precoder ${\mathbf{F}}_k\in\mathbb{C}^{M_k\times S_k}$ to map $S_k$ symbols in ${\mathbf{s}}_k$ to its $M_k$ transmit antennas,
\begin{equation}
{\mathbf{x}}_k={\mathbf{F}}_k{\mathbf{s}}_k,
\end{equation}
where
the transmitted symbols are i.i.d.~such that $\mathbb{E}{\mathbf{s}}_k{\mathbf{s}}_k^*={\mathbf{I}}$,
the precoder is normalized such that $\|{\mathbf{F}}_k\|^2_F\le \rho_k$,
and $\rho_k$ is the transmit power at transmitter $k$.
Receiver $k$ observes the signal
\begin{equation}
{\mathbf{y}}_k=\sum_{\ell=1}^K{\mathbf{H}}_{k,\ell}{\mathbf{F}}_\ell{\mathbf{s}}_\ell + {\mathbf{v}}_k,
\label{eq:y_k}
\end{equation}
where
${\mathbf{H}}_{k,\ell}$ is the channel between transmitter $\ell$ and receiver $k$
and ${\mathbf{v}}_k$ is Gaussian noise at receiver $k$ with spatial covariance matrix ${\mathbf{R}}_k=\mathbb{E}{\mathbf{v}}_k{\mathbf{v}}_k^*$.
For the analysis in this paper we assume that the channels $\{{\mathbf{H}}_{k,\ell}\}$ are each full rank and mutually independent,
the transmitters send independent data ($\mathbb{E}{\mathbf{s}}_k{\mathbf{s}}_\ell^*={\bf 0}$ for $k\ne\ell$)
and all transmitted signals are statistically independent from the noise at any receiver ($\mathbb{E}{\mathbf{s}}_\ell{\mathbf{v}}_k^*={\bf 0}$ for all $(k,\ell)\in\{1,\dots,K\}^2$).
No assumptions are made on the noise power or covariance at any receiver.
Rewriting~(\ref{eq:y_k}), receiver $k$ sees
\begin{equation}
{\mathbf{y}}_k={\mathbf{H}}_{k,k}{\mathbf{F}}_k{\mathbf{s}}_k + \sum_{\substack{\ell=1\\\ell\ne k}}^K{\mathbf{H}}_{k,\ell}{\mathbf{F}}_\ell{\mathbf{s}}_\ell + {\mathbf{v}}_k.
\label{eq:y_k2}
\end{equation}
The vector ${\mathbf{s}}_k$ is the signal to be decoded by receiver $k$, and the summation term in~(\ref{eq:y_k2}) is called \emph{coordinated interference}, since it is
caused by transmitters that may coordinate to minimize its effect.
Once the precoders are designed, the instantaneous sum rate of the system is
\begin{equation}
R_{\rm sum} = \sum_{k=1}^K\log\left|{\mathbf{I}} +
\tilde{{\mathbf{R}}}_k^{-1}{\mathbf{H}}_{k,k}{\mathbf{F}}_k{\mathbf{F}}_k^*{\mathbf{H}}_{k,k}^*\right|,
\label{eq:sum_rate}
\end{equation}
where
\begin{equation}
\tilde{{\mathbf{R}}}_k = {\mathbf{R}}_k + \sum_{\substack{\ell=1\\\ell\ne k}}^K {\mathbf{H}}_{k,\ell}{\mathbf{F}}_\ell{\mathbf{F}}_\ell^*{\mathbf{H}}_{k,\ell}^*
\end{equation}
is the interference plus noise covariance matrix at receiver $k$.
The instantaneous sum rate is an important metric for multiuser systems because of its ability to capture the total network throughput in a single scalar.
Notice that $R_{\rm sum}$ assumes ideal non-linear decoding of the signal.
Although the proposed algorithms of Section~\ref{sec:algorithms} will design linear processing matrices that can form a part of a linear receiver design,
they serve mainly to simplify the optimization and design of the precoders.
Design of high performance linear receivers is left to future work, except for the MMSE design presented in Section~\ref{subsec:mmse_opt}.
Thus for fair comparison, the sum rate equations assume an ideal decoding for all precoder designs.
Previous authors have shown that $KM/2$ spatial degrees of freedom are achievable in an $(M,M,K)$ interference channel
Degrees of freedom $d$ is
\begin{equation}
d = \lim_{{\rm SNR}\rightarrow\infty}\frac{C_{\rm sum}\left({\rm SNR}\right)}{\log {\rm SNR}},
\end{equation}
where $C_{\rm sum}$ is the sum capacity of the network, rather than the sum rate for our linear precoding model presented in~(\ref{eq:sum_rate}).
The key idea of interference alignment is to make $\sum_{\ell\ne k} S_\ell$ interferers appear as $N_k-S_k$ interferers at receiver $k$ for
each $k$ by having them span a subspace of dimension $N_k-S_k$ of the $N_k$-dimensional receive space.
Mathematically,
\begin{equation}
\sum_{\ell\ne k}{\mathbf{H}}_{k,\ell}{\mathbf{F}}_\ell = \sum_{i=1}^{N_k-S_k}a_i{\mathbf{c}}_k^{(i)}, \forall k,
\end{equation}
where $\{{\mathbf{c}}_k^{(i)}\}$ are basis vectors for the subspace at receiver $k$ in which all interference must lie.
Then receiver $k$ can then resolve its $S_k$ streams with a linear receiver interference-free~\cite{CadJaf:Interference-Alignment-and-Degrees:08}.
For the three user interference channel it is possible to directly find closed-form solutions to $\{{\mathbf{F}}_\ell\}$.
for any $S\le M/2$. Such solutions for obtaining $KM/2$ degrees of freedom, however, in the $(M,M,K)$ interference channel
with $K>3$ users do not appear to be possible. Closed-form solutions, even for a reduced multiplexing gain, are
unknown~\cite{YetJafKay:Feasibility-Conditions-for-Interference:09} except in special cases~\cite{TreGuiRie:On-the-achievability-of-interference-alignment:09}.
A viable alternative for the general case are alternating minimizations.
The next section reviews the existing designs and proposes new algorithms for finding high-rate solutions at finite-SNR
in the MIMO interference channel.
\section{Iterative Algorithms Via Alternating Minimization}\label{sec:algorithms}
This section presents iterative solutions for precoders in the MIMO interference channel using an alternating minimization to solve various optimization objectives.
This section proposes three new metrics which aim to approximate a sum rate maximization with better finite-SNR rates than previous work.
The algorithms presented may be implemented in a distributed or centralized manner similar to~\cite{PetHea:Interference-Alignment-Via-Alternating:09}.
These algorithms share a common structure. Each algorithm is designed to optimize a global objective
$\mathcal{J}$ that incorporates the performance of each data link in the network. The objective is a function of the precoders $\{{\mathbf{F}}_\ell\}$,
the channels $\{{\mathbf{H}}_{k,\ell}\}$ between all nodes\footnote{Some of the algorithms will not make use of the data links,
instead focusing on minimizing post-processing interference.}, and a processing matrix at each receiver, the structure of which will
vary across designs\footnote{These matrices are not necessarily designed to function as spatial equalizers, instead serving mainly to
simplify the design and optimization of the precoders. With the exception of our joint-MMSE algorithm, design of high performance linear receivers
is left to future work.}. The free variables are the $K$ precoders and $K$ receive processing matrices.
A closed-form solution for a global optimization of any of the objectives in this section is unknown. We therefore turn to an
alternating minimization\footnote{Alternating maximizations can be converted into alternating minimizations,
so we focus on alternating minimization.} approach for the $2K$ variables~\cite{CsiTus:Information-geometry-and-alternating:84}.
In general, an alternating minimization arbitrarily initializes $2K-1$ variables
and, assuming these variables are fixed, solves for the remaining one. It stores this solution, and moves to another variable, finding a
new solution for it assuming the rest of the variables are fixed. Each variable in turn is solved for during each iteration.
Note that this procedure is convenient only if there is a simple or even closed-form solution for each of the variables assuming
the rest are fixed. Finally, for each of the designs, with the exception of the proposed maximum SINR design, the precoders
may be derived in parallel, since their solutions at any step of the algorithm do not depend on each other.
\subsection{Subspace Optimization}\label{sec:subspace_opt}
A direct algorithm for the interference channel inspired by interference alignment~\cite{CadJaf:Interference-Alignment-and-Degrees:08}
is to precode the signal at transmitter $\ell$ such that the coordinated interference caused by transmitter $\ell$ at receiver $k\ne\ell$ is nearly orthogonal
to a subspace (with orthonormal basis ${\bf\Phi}_k$) of its receive
space~\cite{GomCadJaf:Approaching-the-Capacity-of-Wireless:08,PetHea:Interference-Alignment-Via-Alternating:09}.
This subspace is then jointly designed along with the precoders to optimize an appropriate cost function.
One way of performing this optimization is to minimize the total ``leakage interference''~\cite{GomCadJaf:Approaching-the-Capacity-of-Wireless:08}
that remains at each receiver after attempting to cancel the coordinated interference by left-multiplication with ${\bf\Phi}_k^*$ for each $k$.
The global function to optimize is thus
\begin{equation}
\mathcal{J}_{\rm IA} = \sum_{k=1}^{K}\mathbb{E}\left\|{\bf\Phi}_k^*\sum_{\substack{\ell=1\\ \ell\ne k}}^{K}
{\mathbf{H}}_{k,\ell}{\mathbf{F}}_\ell{\mathbf{s}}_\ell\right\|^2_F.
\end{equation}
The expectation in $\mathcal{J}_{\rm IA}$ and all subsequent analysis is over ${\mathbf{s}}_k$ (and ${\mathbf{v}}_k$ where applicable), $k\in\{1,\dots,K\}$.
Evaluating the expectation and exploiting independence of the signals,
\begin{equation}
\mathcal{J}_{\rm IA} = \sum_{k=1}^K\sum_{\substack{\ell=1\\ \ell\ne k}}^K\|{\bf\Phi}_k^*{\mathbf{H}}_{k,\ell}{\mathbf{F}}_\ell\|_F^2,
\label{eq:j_ia}
\end{equation}
which is termed ``interference leakage'' in~\cite{GomCadJaf:Approaching-the-Capacity-of-Wireless:08}.
The precoders $\{{\mathbf{F}}_\ell\}$ are constrained to have mutually orthogonal columns with a per-stream power constraint so that
${\mathbf{F}}_\ell^*{\mathbf{F}}_\ell=\frac{\rho_\ell}{S_\ell}{\mathbf{I}},\forall\ell$. Although we could enforce a total power constraint on the precoders
(and, coincidentally in this case, get the same solution), orthogonality is desired in MIMO precoding designs to aid with
feedback of channel state~\cite{LovHeaSan:What-is-the-value-of-limited:04}. The receive subspace bases $\{{\bf\Phi}_k\}$ are orthonormal
by definition so that ${\bf\Phi}_k^*{\bf\Phi}_k = {\mathbf{I}}$.
The objective is thus
\begin{eqnarray}
\mathrm{minimize} & \mathcal{J}_{\rm IA}\left(\{{\mathbf{F}}_\ell\},\{{\bf\Phi}\}\right)\nonumber\\
\mathrm{subject~to} & {\mathbf{F}}_\ell^*{\mathbf{F}}_\ell=\frac{\rho_\ell}{S_\ell}{\mathbf{I}}, \ell\in\{1,\dots,K\}\label{eq:ia_obj}\\
{} & {\bf\Phi}_k^*{\bf\Phi}_k={\mathbf{I}}, k\in\{1,\dots,K\}\nonumber.
\end{eqnarray}
The optimization~(\ref{eq:ia_obj}) is intuitively pleasing since, with perfect interference alignment, $\mathcal{J}_{\rm IA}=0$, and without interference
alignment, $\mathcal{J}_{\rm IA}>0$. That is, interference alignment, if possible, achieves the global minimum for this function.
Deriving a closed-form solution to~(\ref{eq:ia_obj}) for $K>3$ users is difficult due to the inter-dependence of each precoder and
receive interference-free subspace.
A simple approach, which is guaranteed to converge, is to use an alternating minimization~\cite{CsiTus:Information-geometry-and-alternating:84}.
The derivation of this solution is in~\cite{GomCadJaf:Approaching-the-Capacity-of-Wireless:08} and our previous work in~\cite{PetHea:Interference-Alignment-Via-Alternating:09}
and is not included here for efficiency. At each step, the solution for each ${\mathbf{F}}_\ell$ is
\begin{equation}
{\mathbf{F}}_\ell = \nu_{\rm min}^{S_k}\left(\sum_{\substack{k=1\\ k\ne\ell}}^K{\mathbf{H}}_{k,\ell}^*{\bf\Phi}_k{\bf\Phi}_k^*{\mathbf{H}}_{k,\ell}\right),
\label{eq:subspace_precoder}
\end{equation}
and, with all precoders given, the solution for each ${\bf\Phi}_k$ is
\begin{equation}
{\bf\Phi}_k = \nu_{\rm min}^{S_k}\left(\sum_{\substack{\ell=1\\\ell\ne k}}^K{\mathbf{H}}_{k,\ell}{\mathbf{F}}_\ell{\mathbf{F}}_\ell^*{\mathbf{H}}_{k,\ell}^*\right).
\label{eq:subspace_decoder}
\end{equation}
To run the algorithm, arbitrary receive subspaces for each receiver are used for initialization and an arbitrary orthonormal basis
${\bf\Phi}_k$ for each subspace is found.
This subspace is ideally reserved for user $k$'s signal, thus coordinated interference at receiver $k$ is ideally orthogonal to this subspace.
Then, for each $\ell$, the algorithm finds the precoder matrix ${\mathbf{F}}_\ell$ such
that total coordinated interference caused at each node (other than at node $\ell$) has maximum squared Euclidean distance between it and the
subspace spanned by the columns of each ${\bf\Phi}_k$ using~(\ref{eq:subspace_precoder}). Given these new precoders, the algorithm can update the
receive subspaces to be those that span the columns of the matrices with minimum sum squared Euclidean distance to the interference caused
by the fixed
precoders using~(\ref{eq:subspace_decoder})~\cite{PetHea:Interference-Alignment-Via-Alternating:09}.
This can be carried out until $\mathcal{J}_{\rm IA}(t)<\epsilon$ if feasibility conditions are met, or
$\mathcal{J}_{\rm IA}(t-1)-\mathcal{J}_{\rm IA}(t)<\epsilon$ otherwise, for an arbitrary convergence threshold $\epsilon$.
Note that each receiver must still separate the desired spatial streams after the coordinated interference has been
canceled with left multiplication of ${\bf\Phi}^*$. Standard linear designs, such as zero forcing or MMSE, can be employed for this purpose.
Thus, the receiver can form a linear receive filter ${\mathbf{G}}_k$ by multiplying ${\bf\Phi}_k$ and
the linear spatial filter ${\mathbf{W}}_k$, which neglects coordinated inter-user interference and equalizes only the desired signal, so that ${\mathbf{G}}_k={\bf\Phi}_k{\mathbf{W}}_k$.
Then the vector $\hat{{\mathbf{s}}}_k={\mathbf{G}}_k^*{\mathbf{y}}_k$ is the interference-free estimate of the original transmitted vector ${\mathbf{s}}_k$.
\subsection{Minimum Interference Plus Noise Leakage (INL)}\label{sec:colored_noise}
\begin{figure}
\centering
\includegraphics[width=3.5in]{colored_source.pdf}
\caption{Receivers modeled by an interference channel may experience uncoordinated interference, modeled as colored noise, from part of the network not modeled as
being a part of the same interference channel.}
\label{fig:colored_source}
\end{figure}
The subspace approach of~\cite{GomCadJaf:Approaching-the-Capacity-of-Wireless:08,PetHea:Interference-Alignment-Via-Alternating:09},
outlined in Section~\ref{sec:subspace_opt}, aims at aligning interference, which is capacity-optimal as the ratio of signal power to receiver noise
power tends to infinity.
If colored noise exists in any receiver, however, the IA subspaces might be chosen to align with the noise to cancel it as well as the
interference. Such colored noise may be due to an interference source outside of the coordinated portion of the network modeled as an interference
channel, as shown in Figure~\ref{fig:colored_source}. This interference is referred to as \emph{uncoordinated interference}.
We therefore focus on algorithms that take noise into account in their optimization.
Note that these approaches have a ``global'' objective function limited to the users cooperating in interference channel, such as inside a single cluster
in Figure~\ref{fig:colored_source}, and thus assume the uncoordinated interferers of other clusters have fixed covariance over the optimization and transmission time.
The objective of the subspace algorithm of Section~\ref{sec:subspace_opt} is to minimize the total post-processing coordinated interference power,
also known as \emph{interference leakage} or \emph{interference power}
in~\cite{GomCadJaf:Approaching-the-Capacity-of-Wireless:08,GomCadJaf:Approaching-the-capacity-of-wireless:09}.
Thus, one intuitive solution is to minimize the total interference plus noise leakage, or INL.
Mathematically, this is represented with the global performance function
\begin{equation}
\mathcal{J}_{\rm INL}= \sum_{k=1}^{K}{\mathbb E}
\left\|{\bf\Phi}_k^*\left(\sum_{\substack{\ell=1\\ \ell\ne k}}^{K}{\mathbf{H}}_{k,\ell}{\mathbf{F}}_\ell{\mathbf{s}}_\ell+{\mathbf{v}}_k\right)\right\|^2_F,
\end{equation}
where ${\mathbf{v}}_k$ is the received noise vector observed at receiver $k$. Expanding the expectation and exploiting the independence of the signal and noise vectors,
the objective becomes
\begin{equation}
\mathcal{J}_{\rm INL}=\sum_{k=1}^K\sum_{\substack{\ell=1\\ \ell\ne k}}^K\left\|{\bf\Phi}_k^*{\mathbf{H}}_{k,\ell}{\mathbf{F}}_\ell\right\|^2_F +
{\rm tr}\left({\bf\Phi}_k^*{\mathbf{R}}_k{\bf\Phi}_k\right),
\label{eq:j_ian}
\end{equation}
where ${\mathbf{R}}_k=\mathbb{E}{\mathbf{v}}_k{\mathbf{v}}_k^*$ is the covariance matrix of the noise at receiver $k$.
The objective is then
\begin{eqnarray}
\mathrm{minimize} & \mathcal{J}_{\rm INL}\left(\{{\mathbf{F}}_\ell\},\{{\bf\Phi}_k\}\right)\nonumber\\
\mathrm{subject~to} & {\mathbf{F}}_\ell^*{\mathbf{F}}_\ell=\frac{\rho_\ell}{S_\ell}{\mathbf{I}}, \ell\in\{1,\dots,K\}\label{eq:noise_obj}\\
{} & {\bf\Phi}_k^*{\bf\Phi}_k={\mathbf{I}}, k\in\{1,\dots,K\}\nonumber.
\end{eqnarray}
The constraints on the precoders and receive subspaces are identical to those in Section~\ref{sec:subspace_opt}.
Further, since $\mathcal{J}_{\rm INL}$ is rotation-invariant to
each of the variables, the solutions lie on the Grassmann manifold and techniques derived for it can be used.
Since $\|{\mathbf{A}}\|_F^2={\rm tr}\left({\mathbf{A}}\bA^*\right)$, $\mathcal{J}_{\rm INL}$ can be rewritten as
\begin{equation}
\mathcal{J}_{\rm INL} = \sum_{k=1}^K\sum_{\substack{\ell=1\\\ell\ne k}}^K
{\rm tr}\left({\bf\Phi}_k^*\left({\mathbf{H}}_{k,\ell}{\mathbf{F}}_\ell{\mathbf{F}}_\ell^*{\mathbf{H}}_{k,\ell}^* + {\mathbf{R}}_k\right){\bf\Phi}_k\right),
\end{equation}
which for fixed $\{{\mathbf{F}}_\ell\}$ is minimized by~\cite{Lut:Handbook-of-Matrices:97}
\begin{equation}
{\bf\Phi}_k^{opt} = \nu_{\rm min}^{S_k}\left({\mathbf{R}}_k+\sum_{\substack{\ell=1\\\ell\ne k}}^K{\mathbf{H}}_{k,\ell}{\mathbf{F}}_\ell{\mathbf{F}}_\ell^*
{\mathbf{H}}_{k,\ell}^*\right).
\end{equation}
For the precoders $\{{\mathbf{F}}_\ell\}$, it is sufficient to note that, for fixed $\{{\bf\Phi}_k\}$, minimizing $\mathcal{J}_{\rm INL}$ with respect
to $\{{\mathbf{F}}_\ell\}$ is equivalent to minimizing $\mathcal{J}_{\rm IA}$ with respect to $\{{\mathbf{F}}_\ell\}$, as is seen by comparing
(\ref{eq:j_ian}) and (\ref{eq:j_ia}). Thus, the precoder solution is identical to~(\ref{eq:subspace_precoder}).
This solution effectively tries to align the coordinated interference with the dominant directions of the noise (or uncoordinated interference)
if the noise has significant energy. In particular, if the noise is highly correlated spatially with a rank-one covariance matrix, then
${\mathbf{R}}_k=\sigma^2_k{\mathbf{a}}_k{\mathbf{a}}_k^*$ and this algorithm will attempt to align the interference to ${\mathbf{a}}_k$ if possible. Such noise, which may
correspond to a single-stream uncoordinated interferer not part of the cooperating network, might then be mitigated, although
full removal is unlikely.
We can also prove the following quantitative conclusions.
\begin{proposition}
If ${\mathbf{R}}_k=\sigma_k^2{\mathbf{I}}$ $\forall k\in\{1,\dots,K\}$, then minimizing $\mathcal{J}_{\rm INL}$ is
equivalent to minimizing $\mathcal{J}_{\rm IA}$.
\label{prop:diag_noise}
\end{proposition}
\begin{IEEEproof}
By definition,
\begin{eqnarray}
\mathcal{J}_{\rm INL} & = & \sum_{k=1}^K\left\|{\bf\Phi}_k^*\sum_{\ell\ne k}{\mathbf{H}}_{k,\ell}{\mathbf{F}}_\ell\right\|_F^2 +
{\rm tr}\left({\bf\Phi}_k^*{\mathbf{R}}_k{\bf\Phi}_k\right)\nonumber\\
{} & = & \sum_{k=1}^K\left\|{\bf\Phi}_k^*\sum_{\ell\ne k}{\mathbf{H}}_{k,\ell}{\mathbf{F}}_\ell\right\|_F^2 +
\sigma_k^2{\rm tr}\left({\bf\Phi}_k^*{\bf\Phi}_k\right)\nonumber\\
{} & = & \mathcal{J}_{\rm IA} + \sum_{k=1}^K\sigma_k^2S_k.\label{eq:same_subspace}
\end{eqnarray}
Since the summation in~(\ref{eq:same_subspace}) is independent of any of the free variables, minimizing $\mathcal{J}_{\rm INL}$ is
equivalent to minimizing $\mathcal{J}_{\rm IA}$ when the noise is spatially white at each receiver.
\end{IEEEproof}
\begin{proposition}
As $\rho_k\rightarrow\infty$ or $\|{\mathbf{R}}_k\|_F\rightarrow 0$ for all $k$, $\mathcal{J}_{\rm INL}$ converges to $\mathcal{J}_{IA}$. Thus,
the subspace algorithm with noise consideration has the same SNR scaling as the pure interference alignment algorithm.
\label{prop:no_noise}
\end{proposition}
\begin{IEEEproof}
Define $\lambda_k$ as the largest eigenvalue of Hermitian matrix ${\mathbf{R}}_k$. Then,
\begin{eqnarray}
\mathcal{J}_{\rm INL} & = & \mathcal{J}_{\rm IA} + \sum_{k=1}^K{\rm tr}\left({\bf\Phi}_k^*{\mathbf{R}}_k{\bf\Phi}_k\right)\nonumber\\
{} & \le & \mathcal{J}_{\rm IA} + \sum_{k=1}^K\lambda_kS_k.
\end{eqnarray}
For any arbitrary $\{{\mathbf{R}}_k\}$, we define a sequence of functions
\begin{equation}
\mathcal{J}_{\rm INL}^{(n)}\doteq \mathcal{J}_{\rm IA} + \frac{1}{n}\sum_{k=1}^K\lambda_kS_k,
\end{equation}
corresponding to a sequence of noise covariance matrices ${\mathbf{R}}_k^{(n)} = {\mathbf{R}}_k/n$, so that
$\|{\mathbf{R}}_k\|_F\rightarrow 0$ as $n\rightarrow\infty$.
Then for any $\epsilon>0$,
\begin{equation}
\left|\mathcal{J}_{\rm INL}^{(n)} - \mathcal{J}_{\rm IA}\right| \le \epsilon
\end{equation}
for all $n>\sum_{k=1}^K\lambda_kS_k/\epsilon$.
\end{IEEEproof}
From the proof, we also note that $\mathcal{J}_{\rm INL}\ge\mathcal{J}_{\rm IA}$ and $\min \mathcal{J}_{\rm INL}=0$ iff
${\mathbf{R}}_k$ is singular for all $k$ and the columns of the interference-aligning receiver matrices $\{{\bf\Phi}_k\}$ lie in the null spaces of their
respective noise covariance matrices $\{{\mathbf{R}}_k\}$.
The metric $\mathcal{J}_{\rm INL}$ is, in fact, likely to have a positive global minimum unless the total number of streams is reduced below the
degrees of freedom of the network, even if the noise is correlated, because the noise subspaces at different receivers will be not perfectly
alignable almost surely. Adapting the number of streams in the network to improve finite-SNR performance is an interesting problem that is
beyond the scope of this paper.
In an idealized system with
Gaussian signaling, colored noise may correspond to uncoordinated interference from outside the network of interest. For instance,
consider the scenario of a cellular network across a metropolitan area. The strategy for this network may be to coordinate three adjacent sectors
to use interference alignment (via subspace optimization) to transmit to one mobile per sector in the downlink. For a regular IA solution,
the uncoordinated interference arriving at each receiver from sectors outside the coordination area would be ignored or modeled as spatially white.
The min-INL algorithm would be able to exploit the knowledge of this uncoordinated interference and account for it as necessary.
The algorithms of Sections~\ref{sec:subspace_opt} and~\ref{sec:colored_noise} aim to align the coordinated interference, which in turn maximizes capacity
in a fully connected high-SNR network. We have seen in Proposition~\ref{prop:diag_noise} that in finite-SNR environments, white Gaussian noise
does not change the solutions of subspace methods. Although this section has presented an approach for networks with colored noise,
algorithms with better throughput performance in finite-SNR regimes are desired, especially since most networks are not likely to be fully
connected and thus may operate with low interference-to-noise ratio (INR), where subspace methods are not likely optimal even with colored noise considerations.
To illustrate the problem of implementing subspace algorithms in a real network, consider the following argument.
Suppose all interfering links $\{{\mathbf{H}}_{k,\ell}\}, k\ne\ell$ have a path loss coefficient $\beta$ whereas direct links
have a path loss coefficient of 1. The subspace precoder design will then not depend at all on the value of $\beta$ since the scalar
multiplication does not change the direction of the signal. If the receivers use their interference suppression filters $\{{\mathbf{U}}_k\}$
to cancel the interference, then the throughput of the system will be independent of $\beta$. Thus, subspace algorithms treat weak and
strong interferers equally, without exploiting the possible capacity gains available when interference is weakened.
If no noise exists in the system, this is perfectly fine, since the receiver will still have an interference-free signal that it could decode
perfectly. Realistically, however, a dynamic network would benefit from adapting its behavior to the relative interference energy. As shown numerically in Section~\ref{sec:sims},
the algorithms proposed in Sections~\ref{subsec:mmse_opt} and~\ref{subsec:max_sinr} are more suited to such adaptation than the subspace method of Section~\ref{sec:subspace_opt}.
\subsection{Mean Squared Error Minimization}\label{subsec:mmse_opt}
A common metric for accounting for noise in linear receivers in wireless communication systems is the mean squared error. For example,
a zero-forcing linear MIMO receiver simply inverts the channel, and results in coloring and amplification of noise. An MMSE receiver
balances the effects of noise with that of inverting the channel depending on the relative energy of each.
This same concept can be applied to interference alignment, where the transmitter and receiver balance their wish to align the coordinated interference
with the need for keeping the signal level well above the noise.
Joint MMSE designs for MIMO channels have been studied for years and have been applied to the point-to-point
model~\cite{Sal:Digital-transmission-over:85,YanRoy:On-joint-transmitter-and-receiver:94,SamPau:Joint-transmit-and-receive:99}
and the broadcast channel~\cite{TenAdv:Joint-multiuser-transmit-receive:04,ZhaWuZho:Joint-linear-transmitter:05}.
The development for the interference channel is distinguished from previous work
in that precoders and receivers need to be designed for multiple transmitters and receivers, rather than just the multiple transmitters
\emph{or} receivers as in the multi-user case, or a single transmitter and receiver in the point-to-point case.
As opposed to objectives discussed in Sections~\ref{sec:subspace_opt} and \ref{sec:colored_noise}, the MMSE directly designs the receive
spatial filters $\{{\mathbf{G}}_k\}$. That is, the output of the product ${\mathbf{G}}^*_k{\mathbf{y}}_k$ is the estimate $\hat{{\mathbf{s}}}_k$ of ${\mathbf{s}}_k$, and the MMSE criterion
minimizes the expected sum of the norms between each $\hat{{\mathbf{s}}}_k$ and ${\mathbf{s}}_k$ for all $k$, yielding the objective
\begin{equation}
\mathcal{J}_{\rm MSE} = \sum_{k=1}^K{\mathbb E}\|{\mathbf{G}}_k^*{\mathbf{y}}_k - {\mathbf{s}}_k\|^2.
\end{equation}
Substituting~(\ref{eq:y_k2}) for ${\mathbf{y}}_k$ results in a global performance function of
\begin{equation}
\mathcal{J}_{\rm MSE} = \sum_{k=1}^{K}{\mathbb E}
\biggl\|{\mathbf{G}}_k^*\biggl({\mathbf{H}}_{k,k}{\mathbf{F}}_k{\mathbf{s}}_k+\sum_{\substack{\ell=1\\ \ell\ne k}}^{K}{\mathbf{H}}_{k,\ell}{\mathbf{F}}_\ell{\mathbf{s}}_\ell+{\mathbf{v}}_k\biggr)
-{\mathbf{s}}_k\biggr\|^2_F,
\end{equation}
and an optimization objective of
\begin{eqnarray}
\mathrm{minimize} & \mathcal{J}_{\rm MSE}\left(\{{\mathbf{F}}_\ell\},\{{\mathbf{G}}_k\}\right)\nonumber\\
\mathrm{subject~to} & \|{\mathbf{F}}_\ell\|_F^2\le \rho_\ell, \ell\in\{1,\dots,K\}.
\label{eq:mmse_obj}
\end{eqnarray}
Expanding the expectation and simplifying, the optimization is equivalent to
\begin{eqnarray}
\mathrm{minimize} & \sum_{k=1}^K
{\rm tr}\left({\mathbf{G}}_k^*\left(\tilde{{\mathbf{R}}}_k+{\mathbf{H}}_{k,k}{\mathbf{F}}_k{\mathbf{F}}_k^*{\mathbf{H}}_{k,k}^*\right){\mathbf{G}}_k\right)\nonumber\\
{} & -2\mathbb{R}\left\{{\rm tr}\left({\mathbf{G}}_k^*{\mathbf{H}}_{k,k}{\mathbf{F}}_k\right)\right\}\nonumber\\
\mathrm{subject~to} & \|{\mathbf{F}}_\ell\|_F^2\le \rho_\ell, \ell\in\{1,\dots,K\}.
\label{eq:j_mmse}
\end{eqnarray}
In general, MMSE solutions with an orthogonality constraint are more difficult to derive. Thus, we relax the orthogonality
constraint to a total power inequality constraint $\|{\mathbf{F}}_\ell\|_F^2\le \rho_\ell, \forall\ell$, and resort to a solution satisfying the
Karush-Kuhn-Tucker (KKT) conditions as in previous joint MMSE solutions for different channel models~\cite{UluYen:Iterative-transmitter-and-receiver:04}.
As shown in Appendix~\ref{app:mmse}, at each step the optimal precoders are
\begin{equation}
{\mathbf{F}}_\ell = \left(\mu_\ell{\mathbf{I}} + \sum_{k=1}^K{\mathbf{H}}_{k,\ell}^*{\mathbf{G}}_k{\mathbf{G}}_k^*{\mathbf{H}}_{k,\ell}\right)^{-1}{\mathbf{H}}_{\ell,\ell}^*{\mathbf{G}}_\ell,
\label{eq:mmse_f}
\end{equation}
where $\mu_\ell$ is the Lagrangian multiplier chosen to meet the power constraint. This may require a simple optimization
(detailed in Appendix~\ref{app:mmse}) and has no known closed form. The optimal receivers are
\begin{equation}
{\mathbf{G}}_k = \left(\sum_{\ell=1}^K{\mathbf{H}}_{k,\ell}{\mathbf{F}}_\ell{\mathbf{F}}_\ell^*{\mathbf{H}}_{k,\ell}^* + {\mathbf{R}}_k\right)^{-1}{\mathbf{H}}_{k,k}{\mathbf{F}}_k,
\label{eq:mmse_g}
\end{equation}
where no further optimization needs to be performed because there is no constraint on the receiver.
As the following proposition shows, this design can be viewed as a generalization of previous designs for the point-to-point case.
\begin{proposition}
With ${\mathbf{H}}_{k,\ell}={\bf 0}$ for all $\ell, k$ such that $k\ne\ell$, (\ref{eq:mmse_f}) and (\ref{eq:mmse_g}) are equivalent to an MMSE design for
a point-to-point scenario. Further, as $\rho_k\rightarrow\infty$, the precoders and receivers diagonalize their respective information
links.
\end{proposition}
\begin{IEEEproof}
This is proven by substituting ${\bf 0}$ for each ${\mathbf{H}}_{k,\ell}$, $k\ne\ell$, and referring to previous point-to-point
results~\cite{YanRoy:On-joint-transmitter-and-receiver:94,SamPau:Joint-transmit-and-receive:99}.
\end{IEEEproof}
We also note that at high SNR and no coordinated inter-user interference, the MMSE algorithm will converge with one step, since
any initialization precoder ${\mathbf{F}}_\ell$ is a fixed point of the algorithm and will minimize the MSE.
The MMSE design is unique among those discussed in this paper.
As discussed before, the MMSE receiver gives a direct estimate of ${\mathbf{s}}_k$, while the others require a conventional
MIMO receiver after ${\mathbf{G}}_k$ is applied. The MMSE receiver solution for fixed ${\mathbf{F}}_\ell, \ell\in\{1,\dots,K\},$ is simply the conventional
MMSE MIMO receiver with colored noise. Further, the solution at each step is not in closed form, as an optimization needs to be
done to meet the power constraint for the precoders. Lastly, the precoder solution is not orthogonal (or, conversely, a solution
with orthogonal constraints is difficult to find). This algorithm may be difficult to implement because of these properties.
Finally, the min-INL optimization is equivalent to an MMSE problem that compares the post-processing output
to ${\bf\Phi}_k^*{\mathbf{H}}_{k,k}{\mathbf{F}}_k{\mathbf{s}}_k$ instead of simply ${\mathbf{s}}_k$. That is,
\begin{equation}
\mathcal{J}_{\rm INL} = \sum_{k=1}^K\mathbb{E}\left\|{\bf\Phi}_k^*{\mathbf{y}}_k - {\bf\Phi}_k^*{\mathbf{H}}_{k,k}{\mathbf{F}}_k{\mathbf{s}}_k\right\|_F^2.
\end{equation}
Thus, the receiver ${\bf\Phi}_k$ is expected to remove only the effects of coordinated interference
and white noise instead of having to correct for distortion created by the channel as well. This output must then be sent to a
MIMO equalizer to remove inter-stream interference before symbol-by-symbol demodulation.
\subsection{Signal-to-Interference-Plus-Noise-Ratio Maximization}\label{subsec:max_sinr}
The original subspace algorithm presented in Section~\ref{sec:subspace_opt} minimizes post-processing coordinated interference energy.
The min-INL algorithm in Section~\ref{sec:colored_noise} adds consideration for noise leakage as well, which can improve performance under colored noise.
The MMSE solution in Section~\ref{subsec:mmse_opt} indirectly accounts for signal
power by attempting to force the received signal to look like the intended signal before precoding and transmission. It is clear, however, that
a more desirable metric for maximizing the sum throughput would directly account for the post-processing signal-to-interference-plus-noise ratio (SINR).
This section presents an algorithm for maximizing total SINR in the network.
The optimization we use is not the only one that could be considered ``maximum SINR'', however, since total SINR for multiple nodes is not
strictly defined in the literature. One may construct any number of global SINR metrics.
Previous authors have considered the inter-stream interference for each transmit/receive pair and solved for the
precoding and receiver matrices one column at a time~\cite{GomCadJaf:Approaching-the-Capacity-of-Wireless:08}, resulting in
non-orthogonal precoders and receive spatial filters, as in the MMSE case.
That approach, however, is not an alternating optimization of a global objective function, and its convergence is unproven.
We therefore reformulate the problem into a maximization of the sum signal power across the network divided by the sum interference power,
incorporating the inter-stream interference for each user.
The performance function becomes
\begin{equation}
\mathcal{J}_{\rm SINR} = \frac{\sum_{k=1}^{K}\sum_{n=1}^{S_k}\mathbb{E}\bigl|{\mathbf{g}}_k^{(n)*}{\mathbf{H}}_{k,k}{\bf f}_k^{(n)}s_k^n\bigr|^2}
{\sum_{k=1}^K\sum_{n=1}^{S_k}\mathbb{E}\bigl|{\mathbf{g}}_k^{(n)*}\bigl(\sum_{\substack{\ell=1\\ \ell\ne k}}^{K}
{\mathbf{r}}_{k,\ell}+{\mathbf{r}}_{k,k}^{(n)}
+{\mathbf{v}}_k\bigr)\bigr|^2},
\label{eq:j_sinr}
\end{equation}
where ${\mathbf{g}}_k^{(n)}$ is the $n$th column of matrix ${\mathbf{G}}_k$,
\begin{equation}
{\mathbf{r}}_{k,\ell}=\sum_{m=1}^{S_\ell}{\mathbf{H}}_{k,\ell}{\bf f}_\ell^{(m)}s_\ell^m
\end{equation}
is the pre-processing interference at receiver $k$ from transmitter $\ell$,
\begin{equation}
{\mathbf{r}}_{k,k}^{(n)}=\sum_{\substack{w=1\\ w\ne n}}^{S_k}{\mathbf{H}}_{k,k}{\bf f}_k^{(w)}s_k^w
\end{equation}
is the pre-processing self-interference from streams $w\ne n$ at receiver $k$,
and $s_k^n$ is the $n$th entry of vector ${\mathbf{s}}_k$.
Notice that
\begin{eqnarray}
R_{\rm sum} & = & \sum_{i=1}^K\sum_{n=1}^{S_i}\log\left(1+\frac{P_i^{(n)}}{I_i^{(n)} + N_i^{(n)}}\right)\nonumber\\
{} & \ge & \log\left(1+\sum_{i=1}^K\sum_{n=1}^{S_i}\frac{P_i^{(n)}}{I_i^{(n)} + N_i^{(n)}}\right)\nonumber\\
{} & \ge & \log\left(1+\frac{\sum_{i=1}^K\sum_{n=1}^{S_i}P_i^{(n)}}{\sum_{i=1}^K\sum_{n=1}^{S_i}I_i^{(n)} + N_i^{(n)}}\right)\nonumber\\
{} & = & \log\left(1+\mathcal{J}_{\rm SINR}\right)\nonumber,
\end{eqnarray}
where $P_i^{(n)}$ is the post-processing signal energy of the $n$th stream at the $i$th receiver, $I_i^{(n)}$ is the post-processing
interference energy, and $N_i^{(n)}$ is the post-processing noise energy seen by the stream.
The new objective~(\ref{eq:j_sinr}) is the sum of signal power in the network divided by the sum coordinated inter-user interference power
and inter-stream interference power after processing.
By maximizing this ratio the algorithm can design the precoders to either decrease
post-processing interference (the denominator) or increase signal power (the numerator) to improve total network performance.
The function $\mathcal{J}_{\rm SINR}$ is a generalized Rayleigh quotient and can be solved using generalized eigen decomposition.
and the optimization problem is
\begin{eqnarray}
\mathrm{maximize} & \mathcal{J}_{\rm SINR}\left(\{{\bf f}_\ell^{(n)}\},\{{\mathbf{g}}_k^{(n)}\}\right)\nonumber\\
\mathrm{subject~to} & \|{\bf f}_\ell^{(n)}\|^2 = \frac{\rho_\ell}{S_\ell},\forall n,\ell.
\label{eq:sinr_obj}
\end{eqnarray}
For tractability we constrain each stream's precoder to have an norm equality constraint so that $\|{\mathbf{F}}_\ell\|^2_F=\rho_\ell$. For a larger
objective function,
and increased complexity, we could also introduce an inequality constraint on each column and vary the transmit power over the streams.
As shown in Appendix~\ref{app:sinr}, the solutions to the columns of the precoders are
\begin{eqnarray}
{\bf f}_\ell^{(n)} = \sqrt{\frac{\rho_\ell}{S_\ell}}\nu_{\rm max}\biggl(
\biggl(~S_\ell q_\ell^{(n)}{\mathbf{I}} +
\sum_{\substack{w=1\\ w\ne n}}^{S_\ell}{\mathbf{H}}_{\ell,\ell}^*{\mathbf{g}}_\ell^{(w)}{\mathbf{g}}_\ell^{(w)*}{\mathbf{H}}_{\ell,\ell}\nonumber\\
+\sum_{\substack{k=1\\ k\ne\ell}}^K\sum_{m=1}^{S_k}{\mathbf{H}}_{k,\ell}^*{\mathbf{g}}_k^{(m)}{\mathbf{g}}_k^{(m)*}{\mathbf{H}}_{k,\ell}\biggr)^{-1}\nonumber\\
\left({\mathbf{H}}_{\ell,\ell}^*{\mathbf{g}}_\ell^{(n)}{\mathbf{g}}_\ell^{(n)*}{\mathbf{H}}_{\ell,\ell} + S_\ell r_\ell^{(n)}{\mathbf{I}}\biggr)
\right),
\label{eq:sinr_precoders}
\end{eqnarray}
where $q_\ell^{(n)}$ is the sum of the terms in the denominator
of~(\ref{eq:j_sinr}) that do not directly involve ${\bf f}_\ell^{(n)}$, and $r_\ell^{(n)}$ is the sum of the terms in the numerator
of~(\ref{eq:j_sinr}) that do not directly involve ${\bf f}_\ell^{(n)}$. The solutions to the columns of the receivers are
\begin{eqnarray}
{\mathbf{g}}_k^{(n)} = \nu_{\rm max}\Biggl(\Biggl(~\hat{q}_k^{(n)}{\mathbf{I}} +
\sum_{\substack{w=1\\ w\ne n}}^{S_k}{\mathbf{H}}_{k,k}{\bf f}_k^{(w)}{\bf f}_k^{(w)*}{\mathbf{H}}_{k,k}^* +\nonumber\\
\sum_{\substack{\ell=1\\ \ell\ne k}}^K\sum_{m=1}^{S_\ell}{\mathbf{H}}_{k,\ell}{\bf f}_\ell^{(m)}{\bf f}_\ell^{(m)*}{\mathbf{H}}_{k,\ell}^*\Biggr)^{-1}\nonumber\\
\left({\mathbf{H}}_{k,k}{\bf f}_k^{(n)}{\bf f}_k^{(n)*}{\mathbf{H}}_{k,k}^* +
\hat{r}_k^{(n)}{\mathbf{I}}\right)\Biggr),
\label{eq:sinr_receivers}
\end{eqnarray}
where $\tilde{\nu}_{\rm max}\left({\mathbf{A}},{\mathbf{B}}\right)$ is the generalized eigenvector corresponding to the
largest generalized eigenvalue of the matrix pair $\left({\mathbf{A}},{\mathbf{B}}\right)$,
$\hat{q}_k^{(n)}$ and $\hat{r}_k^{(n)}$ are defined similarly as~(\ref{eq:sinr_precoders}) but with respect to ${\mathbf{g}}_k^{(n)}$ instead
of ${\bf f}_\ell^{(n)}$.
With all other variables fixed, the solutions in~(\ref{eq:sinr_precoders}) and~(\ref{eq:sinr_receivers}) maximize the global SINR
function~(\ref{eq:j_sinr}), whereas the solutions in~\cite{GomCadJaf:Approaching-the-Capacity-of-Wireless:08} give a suboptimal approximation
to this solution. As shown in Section~\ref{sec:sims}, this does not imply that an iterative algorithm using the proposed solutions will converge to
a larger objective than that of~\cite{GomCadJaf:Approaching-the-Capacity-of-Wireless:08}. For any given channel realization and initialization,
the two algorithms may give an identical result, or either may outperform the other. The simulation results in Section~\ref{sec:sims}
suggest, however, that the two algorithms perform similarly on average. Since the proposed design requires more network knowledge than that
of~\cite{GomCadJaf:Approaching-the-Capacity-of-Wireless:08}, the latter is more attractive for implementation. If the extra network
state knowledge is available, however, an intelligent design would be to run both algorithms and choose the design that works best for
each channel realization, resulting in a sum throughput higher than either algorithm could produce individually.
Note that the IA algorithm will minimize the left-hand term of the denominator in $\mathcal{J}_{\rm SINR}$,
and the min-INL algorithm will minimize the entire denominator (minus inter-stream interference). Certainly, with no noise (or,
more rigorously, as total signal to noise ratio goes to infinity), the two solutions are equivalent since maximizing the SINR will
reduce to maximizing the SIR, which, as discussed before, IA does. This fact was proven in~\cite{GomCadJaf:Approaching-the-capacity-of-wireless:09}.
\subsection{Convergence and Initialization}\label{subsec:discuss}
This section analyzes some important details of the algorithms proposed in this paper. In particular, the focus is on variable initialization,
algorithm convergence, method of execution, obtainment of channel state, and precoder constraints.
\begin{figure}
\centering
\includegraphics[width=3.5in]{ia_test_initializations_11152009.pdf}
\caption{Sum rate vs. initialization for each algorithm discussed in Section~\ref{sec:algorithms} run on a single channel realization of the $(2,2,3)$ MIMO IC
at $10$ and $40$ dB. Although the MMSE algorithm varies the most for this channel realization, this is not a general trend.}
\label{fig:init}
\end{figure}
We have found heuristically that arriving at a globally optimum point for the minimization algorithms (global optimality cannot be identified with the
max SINR algorithm) is highly likely even when initializing the precoders to truncated identity matrices;
the throughput, however, is the real objective we wish to optimize, and these algorithms only approximate that optimization.
Thus, different initializations of an algorithm may result in drastically different throughputs, even if they result in the same
final objective (or cost) function. For example, consider Figure~\ref{fig:init}. Each of the algorithms discussed in Section~\ref{sec:algorithms}
was run on a fixed channel with 10 different random precoder initializations for the (2,2,3) MIMO IC at $\rho=40$ dB and $\rho=10$ dB.
For $\rho=40$ dB, the MMSE algorithm varied most between
different initializations, but this is not indicative of the algorithms on the whole, just the behavior for this particular channel.
It appears that finding ``good''
initializations is not difficult; experimentation has shown that random initializations give as good of rates in these
algorithms as any ``intelligent'' initialization tried. If possible, multiple runs of the
algorithm should be made with different initializations for the best performance in terms of \emph{throughput}, as shown in Figure~\ref{fig:init}.
Each of the algorithms from Section~\ref{sec:algorithms} are guaranteed to converge because the objectives are bounded
and at each step are moving
monotonically in the direction of that bound. Convergence to a global optimum is not guaranteed except when the
objective has certain convexity-like properties~\cite{CsiTus:Information-geometry-and-alternating:84} that these algorithms
are not proven to possess. Also, convergence of the cost function does not automatically imply convergence of the precoder designs, the analysis of
which is beyond our scope.
\section{Simulations}\label{sec:sims}
This section presents simulations of the algorithms presented in Section~\ref{sec:algorithms} to substantiate our claims and show that
each of the algorithms can outperform the others in different regimes since none explicitly maximizes throughput.
All of the simulations evaluate the expected sum rate with i.i.d. zero-mean unit-variance complex Gaussian coefficients for each channel,
with the precoders for
each realization calculated with perfect CSI and as if the realization was flat in time and frequency.
More realistic channel scenarios are considered in our related work~\cite{ElAPetHea:A-Study-of-the-Practicality-of-Interference:09}.
Transmitter $k$ is assigned a deterministic transmit power $\rho_k$ and the link from transmitter $\ell$ to receiver $k$
has a deterministic path loss coefficient $\alpha_{k,\ell}$. Whereas in preceding analysis $\alpha_{k,\ell}$ was absorbed into ${\mathbf{H}}_{k,\ell}$,
in this section we pull it out for exposition.
We also define $\gamma_{k,\ell}=\alpha_{k,\ell}\rho_{k,\ell}$ to be the expected SNR at receiver $k$ from transmitter $\ell$.
Thus, the sum rate is
\begin{equation}
R_{\rm sum} = \mathbb{E}_{\{{\mathbf{H}}_{k,\ell}\}}\left\{\sum_{k=1}^K\log\left|{\mathbf{I}} +
\hat{{\mathbf{R}}}_k^{-1}\alpha_{k,k}{\mathbf{H}}_{k,k}{\mathbf{F}}_k{\mathbf{F}}_k^*{\mathbf{H}}_{k,k}^*\right|\right\},
\label{eq:sims_rate}
\end{equation}
where
\begin{equation}
\hat{{\mathbf{R}}}_k={\mathbf{R}}_k + \sum_{\ell\ne k}\alpha_{k,\ell}{\mathbf{H}}_{k,\ell}{\mathbf{F}}_\ell{\mathbf{F}}_\ell^*{\mathbf{H}}_{k,\ell}^*
\end{equation}
is the interference plus noise covariance.
Precoders are initialized randomly with orthonormal columns, as discussed in Section~\ref{subsec:discuss}, and each algorithm is presented with identical
initializations. Five random initializations are used for each channel realization, as motivated in Figure~\ref{fig:init}, and the initialization
that maximizes~(\ref{eq:sims_rate}) is kept while the others are thrown away.
In each plot presented in this section, $R_{\rm sum}$ is computed via Monte Carlo simulations using 1000 independent channel realizations.
Each iterative algorithm is run with 100 iterations each.
Each algorithm from Section~\ref{sec:algorithms} is compared with a random precoding scenario where
each precoder ${\mathbf{F}}_\ell$ is chosen as the left singular vectors of a random Gaussian matrix, to enforce an orthogonality constraint.
That is, by \emph{Random Beamforming}, we mean that
\begin{equation}
{\mathbf{F}}_\ell = \sqrt{\frac{\rho_\ell}{S_\ell}}{\mathbf{U}}_\ell^{(S_\ell)},
\end{equation}
where ${\mathbf{U}}_\ell^{(S_\ell)}$ are the first $S_\ell$ columns of the left singular matrix of a random matrix with i.i.d. zero-mean unit-variance
complex Gaussian coefficients.
A greedy approach is also included to show the benefit of cooperation in the MIMO interference
channel~\cite{YeBlu:Optimized-signaling-for-MIMO:03,RosUluYat:Wireless-systems-and-interference:02}.
In this design, each precoder ${\mathbf{F}}_k$, $k\ne\ell$, is held fixed when designing ${\mathbf{F}}_\ell$. Then
\begin{equation}
{\mathbf{U}}_\ell{\bf \Sigma}_\ell{\mathbf{V}}_\ell^*=\biggl({\mathbf{R}}_\ell+\sum_{k\ne\ell}\alpha_{\ell,k}{\mathbf{H}}_{\ell,k}{\mathbf{F}}_k{\mathbf{F}}_k^*{\mathbf{H}}_{\ell,k}^*\biggr)^{-1/2}\sqrt{\alpha_{\ell,\ell}}{\mathbf{H}}_{\ell,\ell},
\end{equation}
and
\begin{equation}
{\mathbf{F}}_\ell = \sqrt{\frac{\rho_\ell}{S_\ell}}{\mathbf{V}}_\ell^{(S_\ell)}.
\end{equation}
The greedy algorithm is not guaranteed to converge since it is not optimizing a global function, but it requires less channel estimation.
Finally, when the $K=3$ user interference channel is considered, the closed-form solution from~\cite{CadJaf:Interference-Alignment-and-Degrees:08}
is also used for a baseline comparison.
We first introduce colored noise into the interference channel via an uncoordinated rank-one interferer in the network, as discussed in Section~\ref{sec:colored_noise}.
Defining ${\mathbf{H}}_{k,E}$ as the MIMO channel from the uncoordinated rank-one interferer to receiver $k$ in the
interference channel, then receiver $k$ observes
\begin{equation}
{\mathbf{y}}_k = \sum_{\ell=1}^K\sqrt{\alpha_{k,\ell}}{\mathbf{H}}_{k,\ell}{\mathbf{F}}_\ell{\mathbf{s}}_\ell + \sqrt{\alpha_{k,E}}{\mathbf{H}}_{k,E}{\bf f}_Es_E + {\mathbf{v}}_k,
\end{equation}
where since the uncoordinated interferer is rank-one, it is sending a single stream $s_E$ precoded with vector ${\bf f}_E$. Each receiver sees spatially white additive noise
on top of the signal and interference (coordinated and uncoordinated). Pure interference
alignment will ignore the uncoordinated interference, implicitly assuming it is spatially white. The rest of the algorithms will
take the uncoordinated interferer into account but will not be able to fully suppress it without reducing the number of streams in the network
since the uncoordinated interferer, which is scaling its power with the transmitters inside the network, is reducing the degrees of freedom
of the network, making it interference-limited.
Figure~\ref{fig:if_scaled} illustrates results for the rank-one interferer scenario for each algorithm
discussed in Section~\ref{sec:algorithms} with $K=3$ users, $M=N=2$ antennas at each node, and $S=1$ stream being transmitted
between each transmit/receive pair for $\rho_k=\rho_E=\rho$, $\forall k$ and $\alpha=1$. That is, the transmit power is equal at all transmitters, including
the uncoordinated interferer, and the path loss and fading statistics are identical on all links.
\begin{figure}
\centering
\includegraphics[width=3.5in]{ia_test_interferer_scaled_10142009.pdf}
\caption{Sum rate vs. $\rho_k=\rho_E=\rho$ for each algorithm discussed in Section~\ref{sec:algorithms} for the case where a rank-one uncoordinated interferer
is introduced into the (2,2,3) network with $S=1$ stream per user. The interferer's transmit power is scaled with the transmitters in the
network so the degrees of freedom are reduced, and the network is interference-limited at high values of $\rho$. The MMSE algorithm is an exception because it
has a power inequality constraint on its precoders and can thus allow two transmitters to turn off, giving the remaining transmitter one degree of freedom, so the
sum capacity scales linearly with $\rho$.}
\label{fig:if_scaled}
\end{figure}
The MMSE algorithm has higher degrees of freedom in this case because of its power inequality constraint on the precoders. This allows two transmitters to effectively shut
off while the third has a degree of freedom and can cancel the external interferer with its extra receive antenna. This shows the flexibility of the MMSE design.
Other than MMSE, the max-SINR algorithm outperforms the others in the power ranges considered. Note that, although on average the max SINR
algorithm and the approximate max SINR algorithm
have nearly identical performance, for any given channel realization they may have very different sum rates. IA performs the worst of all four iterative algorithms
since it is neglecting the uncoordinated interference. At high $\rho$, considering the colored noise in the algorithm objective results in a roughly 20\% increase
in sum rate for this scenario. Note that the two best-performing algorithms, MMSE and max-SINR, do not have orthogonal precoders and thus may be
more complex to implement in a real system with feedback requirements. With its orthogonal design and improved performance over IA,
the min-INL algorithm is a good tradeoff between complexity and performance in this scenario.
Next, we keep the same scenario but with fixed uncoordinated interference power, so that the degrees of freedom are not reduced.
Figure~\ref{fig:if_fixed} gives the results of this experiment. It shows that the uncoordinated interference, which is fixed at $\rho_E=0$ dB, has little
effect on the system, even at low $\rho_k=\rho$. The algorithms, except random beamforming, all scale at the same rate, and thus all exploit
the maximum degrees of freedom in the network. For a fixed number of iterations, however, the MMSE algorithm does not scale, as it appears to require
more iterations to converge than the others at high $\rho$. In particular, as shown in Figure~\ref{fig:if_fixed}, when the MMSE design is run with 500 iterations,
its performance approaches that of the rest of the designs, while the other algorithms benefit very little from the increase in iterations.
This is consistently seen in the rest of the simulations in this section. Analysis of this
longer convergence is left to future work. Finally, we note that iterative IA outperforms the closed-form solution because multiple IA solutions exist,
and iterative IA is better able to find the best one because of the multiple random initializations. If the closed-form algorithm is modified to explore
multiple possible solutions, it would perform equally well in this case.
\begin{figure}
\centering
\includegraphics[width=3.5in]{interferer_fixed.pdf}
\caption{Sum rate vs. $\rho_k=\rho$ for each algorithm discussed in Section~\ref{sec:algorithms} for the case where a rank-one uncoordinated
interferer with fixed transmit power of $\rho_E = 0$ dB and $\rho_E = 20$ dB is introduced into the (2,2,3) network with $S=1$ stream per user.
The degrees of freedom in this network are the same as if the interferer did not exist, and each algorithm, with the exception of random beamforming, performs
very similarly and exploits all the degrees of freedom in the network.}
\label{fig:if_fixed}
\end{figure}
Now we remove the uncoordinated interferer from all but one receiver in the network, and allow that uncoordinated interference
power to scale with internal network transmit power, so that $\rho_k=\rho_E=\rho$, but $\alpha_{k,E}=0$ for $k>1$ and $\alpha_{1,E}=1$.
Figure~\ref{fig:if_scaled_one} shows the results.
\begin{figure}
\centering
\includegraphics[width=3.5in]{ia_test_interferer_scaled_one_2antennas_07112009.pdf}
\caption{Sum rate vs. $\rho_k=\rho_E=\rho$, with $\alpha_{k,E}=0$ for $k>1$ and $\alpha_{1,E}=1$, for each algorithm discussed in
Section~\ref{sec:algorithms}. In this case, a rank-one uncoordinated
interferer is sensed at only receiver 1 in the (2,2,3) MIMO interference channel with $S=1$. The interferer's transmit power is
scaled with the transmitters in the network so the degrees of freedom are reduced. The network is not interference-limited, however,
since only one receiver sees the interference.}
\label{fig:if_scaled_one}
\end{figure}
The maximum SINR and minimum INL algorithms suffer at high $\rho$ relative to pure interference
alignment and MMSE. This is because these algorithms see the large interferer at receiver $k=1$ as something to be overcome; these algorithms
are effectively concerned with the average performance of the network. IA is equally concerned about the average performance
but only in terms of coordinated interference, whereas MMSE has flexibility to overcome the interference by reducing transmit power of receiver 1.
To maximize sum rate in this case, it appears one should either ignore the external interference or include a power inequality constraint in the precoder design.
We now turn to the case of no uncoordinated interference, considering only the conventional interference channel in isolation. In the first
experiment, the transmit power is kept fixed but the path loss coefficient $\alpha_{k,\ell}$ is varied on the interfering links ($k\ne\ell$) only.
Figure~\ref{fig:scale_if} illustrates the results. The
IA algorithm has constant throughput regardless of the interference path loss coefficient, but the other iterative algorithms
are able to exploit the decrease in interference, converging to IA when the interference power is high.
\begin{figure}
\centering
\includegraphics[width=3.5in]{test_optim1_scaled07112009.pdf}
\caption{Sum rate vs. $\alpha=\alpha_{k,\ell}$, $k\ne\ell$, for the $(2,2,3)$ MIMO interference channel with $S=1$ stream per user.
The iterative IA algorithm is not able to exploit reduced interference power, while all the other iterative algorithms can substantially
improve throughput. The SNR on the data links is fixed at $\gamma_{k,k} = 40$ dB.}
\label{fig:scale_if}
\end{figure}
\section{Conclusions and Future Work}\label{sec:conclusion}
This paper has discussed the application and performance of iterative algorithms in the MIMO $K$-user constant-coefficient interference channel
under various operating regimes. The convergence and optimality of the algorithms has been discussed, and similarities
between all of them have been derived. If an iterative solution for the interference channel is ever practical in a real system, it is unlikely that a direct
interference alignment approach is desirable because of its suboptimality in environments where one or more links have little
energy relative to the others. Instead, the max SINR or MMSE metrics are desirable in most environments because they
flexibly adapt the solution between interference alignment (high interference power)
and SVD precoding (no interference, fixed number of streams), and the MMSE solution in particular has a transmit power inequality constraint.
These algorithms, however, have relatively high implementation complexity because
of their nonorthogonality and lack of closed-form solutions at each step in general cases. In particular, the MMSE algorithm requires
some optimization for meeting the power constraint, and the max-SINR algorithm requires more channel state knowledge at each iteration than the others.
The min-INL algorithm is a good tradeoff between the three algorithms, since it has improved performance over
IA in scenarios where there is uncoordinated interference or colored noise, but still has relatively low implementation complexity because of
its simpler solutions and orthogonal precoders.
Future work will focus on analyzing and reducing the overhead associated with solutions such as the ones presented in this paper.
Although some studies have been carried out on the application of interference alignment to a cellular
network~\cite{SuhTse:Interference-Alignment-for-Cellular:08,CaiRamPap:Multiuser-MIMO-downlink:08,TreGui:Cellular-interference-alignment:09},
overhead and feedback analyses need to be performed to find out if the achievable gains are worth the effort.
\appendices
\section{Derivation of Mean Squared Error Minimization}\label{app:mmse}
\begin{IEEEproof}
For completeness, we restate the optimization from~(\ref{eq:mmse_obj}),
\begin{eqnarray}
\mathrm{minimize} & \mathcal{J}_{\rm MSE}\left(\{{\mathbf{F}}_\ell\},\{{\mathbf{G}}_k\}\right)\nonumber\\
\mathrm{subject~to} & \|{\mathbf{F}}_\ell\|_F^2\le \rho_\ell, \ell\in\{1,\dots,K\}.
\label{eq:app_mmse_obj}
\end{eqnarray}
where $\mathcal{J}_{\rm MSE}$ is defined in~(\ref{eq:j_mmse}).
We use the Karush-Kuhn-Tucker conditions to solve the optimization at each step with all but one variable fixed.
The Lagrangian of~(\ref{eq:mmse_obj}) is
\begin{eqnarray}
\mathcal{L} = \sum_{k=1}^K
{\rm tr}\left({\mathbf{G}}_k^*\left(\sum_{\ell=1}^K{\mathbf{H}}_{k,\ell}{\mathbf{F}}_\ell{\mathbf{F}}_\ell^*{\mathbf{H}}_{k,\ell}^* + {\mathbf{R}}_k\right){\mathbf{G}}_k\right) -\nonumber\\
2\mathbb{R}\left\{{\rm tr}\left({\mathbf{G}}_k^*{\mathbf{H}}_{k,k}{\mathbf{F}}_k\right)\right\}
+ \sum_{\ell=1}^K\mu_\ell\left({\rm tr}\left({\mathbf{F}}_\ell^*{\mathbf{F}}_\ell\right)-\rho_\ell\right),
\end{eqnarray}
where $\mu_\ell$ is the Lagrangian multiplier for the power constraint for precoder $\ell$.
The KKT conditions are
\begin{eqnarray}
\nabla\mathcal{L} & = & {\bf 0}\label{eq:kkt1}\\
\mu_\ell\left({\rm tr}\left({\mathbf{F}}_\ell^*{\mathbf{F}}_\ell\right)-1\right) & = & 0, \forall\ell\label{eq:kkt2}\\
{\rm tr}\left({\mathbf{F}}_\ell^*{\mathbf{F}}_\ell\right) & \le & \rho_\ell, \forall\ell\label{eq:kkt3}\\
\mu_\ell & \ge & 0, \forall\ell\label{eq:kkt4}.
\end{eqnarray}
For fixed $\{{\mathbf{F}}_\ell\}$ and $\{\mu_\ell\}$, $\{{\mathbf{G}}_k\}$ can be found by solving $\nabla_{{\mathbf{G}}_k}\mathcal{L}=\nabla_{{\mathbf{G}}_k}\mathcal{J}_{\rm MSE} =
{\bf 0}$ for
$k\in\{1,\dots,K\}$ and the KKT conditions will be automatically met since there are no constraints on $\{{\mathbf{G}}_k\}$. This yields
\begin{equation}
{\mathbf{G}}_k = \left(\sum_{\ell=1}^K{\mathbf{H}}_{k,\ell}{\mathbf{F}}_\ell{\mathbf{F}}_\ell^*{\mathbf{H}}_{k,\ell}^* + {\mathbf{R}}_k\right)^{-1}{\mathbf{H}}_{k,k}{\mathbf{F}}_k,
\end{equation}
In solving for $\{{\mathbf{F}}_\ell\}$, we must ensure all of~(\ref{eq:kkt1})--(\ref{eq:kkt4}) are satisfied. To satisfy
$\nabla_{{\mathbf{F}}_\ell}\mathcal{L} = {\bf 0}$, we must have
\begin{equation}
{\mathbf{F}}_\ell = \left(\mu_\ell{\mathbf{I}} + \sum_{k=1}^K{\mathbf{H}}_{k,\ell}^*{\mathbf{G}}_k{\mathbf{G}}_k^*{\mathbf{H}}_{k,\ell}\right)^{-1}{\mathbf{H}}_{\ell,\ell}^*{\mathbf{G}}_\ell.
\end{equation}
If $\mu_\ell=0$ satisfies~(\ref{eq:kkt3}), then all the KKT conditions are satisfied and the optimal ${\mathbf{F}}_\ell$ has been found for this step of the alternating
minimization. Otherwise, we must solve for $\mu_\ell>0$ such that $\left\|{\mathbf{F}}_\ell\right\|_F = \rho_\ell$ to satisfy the KKT conditions. Although there is no
known closed-form solution for $\mu_\ell$ in this case~\cite{UluYen:Iterative-transmitter-and-receiver:04}, $\left\|{\mathbf{F}}_\ell\right\|_F$ is a monotonically
decreasing function of $\mu_\ell$ for $\mu_\ell>0$, so simple one-dimensional searches such as the bisection method can be done to solve for $\{\mu_\ell\}$.
\end{IEEEproof}
\section{Derivation of SINR Maximization}\label{app:sinr}
\begin{IEEEproof}
For completeness, we restate the optimization from~(\ref{eq:sinr_obj}),
\begin{eqnarray}
\mathrm{maximize} & \mathcal{J}_{\rm SINR}\left(\{{\bf f}_\ell^{(n)}\},\{{\mathbf{g}}_k^{(n)}\}\right)\nonumber\\
\mathrm{subject~to} & \|{\bf f}_\ell^{(n)}\|^2 = \frac{\rho_\ell}{S_\ell}, \forall n,\ell.
\label{eq:app_sinr_obj}
\end{eqnarray}
where $\mathcal{J}_{\rm SINR}$ is defined in~(\ref{eq:j_sinr}).
The optimization is performed on the columns of each precoder ${\mathbf{F}}_\ell$ and spatial equalizer ${\mathbf{G}}_k$.
In solving for the $n$th column of precoder ${\mathbf{F}}_\ell$, we hold fixed every other precoder ${\mathbf{F}}_k$, $k\ne\ell$, and
other columns ${\bf f}_\ell^{(w)}$, $w\ne n$ of precoder ${\mathbf{F}}_\ell$, as well as every receive combining matrix ${\mathbf{G}}_k$, $\forall k$.
The objective of~(\ref{eq:j_sinr}) can be rewritten as
\begin{equation}
\mathcal{J}_{\rm SINR} = \frac{\sum_{\ell=1}^K\sum_{n=1}^{S_k}{\bf f}_\ell^{(n)*}\left({\mathbf{C}}_\ell^{(n)}+
S_\ell r_\ell^{(n)}{\mathbf{I}}\right){\bf f}_\ell^{(n)}}
{\sum_{\ell=1}^K\sum_{n=1}^{S_k}{\bf f}_\ell^{(n)*}\left(S_\ell q_\ell^{(n)}{\mathbf{I}} +
{\mathbf{A}}_\ell^{(n)}+{\mathbf{B}}_\ell^{(n)}\right){\bf f}_\ell^{(n)}},
\label{eq:app_sinr_expanded}
\end{equation}
where
\begin{eqnarray}
{\mathbf{A}}_\ell^{(n)} & = & \sum_{\substack{w=1\\ w\ne n}}^{S_\ell}{\mathbf{H}}_{\ell,\ell}^*{\mathbf{g}}_\ell^{(w)}{\mathbf{g}}_\ell^{(w)*}{\mathbf{H}}_{\ell,\ell}\\
{\mathbf{B}}_\ell^{(n)} & = & \sum_{\substack{k=1\\ k\ne\ell}}^K\sum_{m=1}^{S_k}{\mathbf{H}}_{k,\ell}^*{\mathbf{g}}_k^{(m)}{\mathbf{g}}_k^{(m)*}{\mathbf{H}}_{k,\ell}\\
{\mathbf{C}}_\ell^{(n)} & = & {\mathbf{H}}_{\ell,\ell}^*{\mathbf{g}}_\ell^{(n)}{\mathbf{g}}_\ell^{(n)*}{\mathbf{H}}_{\ell,\ell}
\end{eqnarray}
and $q_\ell^{(n)}$ is the remaining summation terms in the denominator of~(\ref{eq:j_sinr}) that are independent of ${\bf f}_\ell^{(n)}$,
contracted here for brevity. Similarly, $r_\ell^{(n)}$ is the remaining summation terms in the numerator of~(\ref{eq:j_sinr}) that are
independent of ${\bf f}_\ell^{(n)}$. The function in~(\ref{eq:app_sinr_expanded}) is the generalized Rayleigh quotient which is well
known to be solved by the generalized eigen-vector of the numerator and denominator matrices,
\begin{eqnarray}
{\bf f}_\ell^{(n)} = \sqrt{\frac{\rho_\ell}{S_\ell}}\nu_{\rm max}\biggl(\biggl(~S_\ell q_\ell^{(n)}{\mathbf{I}} +
{\mathbf{A}}_\ell^{(n)} + {\mathbf{B}}_\ell^{(n)}\biggr)^{-1}\nonumber\\
\biggl({\mathbf{C}}_\ell^{(n)} + S_\ell r_\ell^{(n)}{\mathbf{I}}\biggr) \biggr).
\end{eqnarray}
The derivation of the receive combiner columns follows the same structure as the precoders, but with a unit-norm constraint on the columns
for simplicity (this removes the $S_\ell$ multipliers in front of $q_\ell^{(n)}$ and $r_\ell^{(n)}$).
Here we define $\hat{r}_k^{(n)}$ to be the terms in the numerator of~(\ref{eq:j_sinr}) independent of ${\mathbf{g}}_k^{(n)}$ and similarly
for $\hat{q}_k^{(n)}$ in the denominator. The solution for the $n$th column of the $k$th receive combiner is then
\begin{eqnarray}
{\mathbf{g}}_k^{(n)} = \nu_{\rm max}\Biggl(\Biggl(~\hat{q}_k^{(n)}{\mathbf{I}} +
\sum_{\substack{w=1\\ w\ne n}}^{S_k}{\mathbf{H}}_{k,k}{\bf f}_k^{(w)}{\bf f}_k^{(w)*}{\mathbf{H}}_{k,k}^* +\nonumber\\
\sum_{\substack{\ell=1\\ \ell\ne k}}^K\sum_{m=1}^{S_\ell}{\mathbf{H}}_{k,\ell}{\bf f}_\ell^{(m)}{\bf f}_\ell^{(m)*}{\mathbf{H}}_{k,\ell}^*\Biggr)^{-1}\nonumber\\
\left({\mathbf{H}}_{k,k}{\bf f}_k^{(n)}{\bf f}_k^{(n)*}{\mathbf{H}}_{k,k}^* +
\hat{r}_k^{(n)}{\mathbf{I}}\right)\Biggr).
\end{eqnarray}
\end{IEEEproof}
\bibliographystyle{IEEEtran}
|
\section{Introduction}
The correlation between the masses of supermassive black holes (SMBH) with the global characteristics of their host galaxies \citep[e.g.][]{2000ApJ...539L...9F,2000ApJ...539L..13G} is crucial for our interpretation of galaxy formation and evolution \cite[e.g.][]{1998A&A...331L...1S,2006ApJS..163....1H}. M$_{\bullet}$ is usually determined using dynamical tracers such as a disk of gas clouds in Keplerian rotation around the SMBH or the line-of-sight velocity distribution (LOSVD) of stars. Both methods require observations of high spatial resolution since the dynamical models have to be constrained by resolved kinematics which probe the sphere of gravitational influence of the SMBH (R$_{sph}$), defined as the distance from SMBH at which the potential of the galaxy and SMBH are approximately equal. This spatial scale, usually defined as R$_{sph}= GM_{\bullet}/\sigma^2$, where $\sigma$ is the velocity dispersion of the galaxy, is typically significantly less than an arcsec in apparent size, even for nearby galaxies.
The influence of the SMBH, however, will be felt at larger radii than R$_{shp}$, albeit to a lessening degree. Here we show that it is possible to constrain the M$_{\bullet}$ when R$_{shp}$ is up to 3 times smaller than the spatial resolution of the observations if the observation consists of {\it (i)} low (arcsec) spatial resolution integral field data covering the galaxy out to about 1 effective radius required to determine the overall dynamical mass-to-light ratio of the system, {\it (ii)} high (sub-arcsec) resolution integral field data probing the stellar kinematics in the vicinity of the SMBH and have {\it (iii)} sufficient signal-to-noise ratio (S/N) to extract higher order moments of the LOSVD.
In this report we summarise the results of \cite{2009MNRAS.399.1839K} which present the usage of a new observational method with LGS AO system especially suitable for determination of M$_{\bullet}$ in the nuclei of nearby low mass early-type galaxies. For determination of M$_{\bullet}$ in massive early-type galaxies using similar method, but non-AO observations see the contribution of \cite{Cappellari2010} in this proceedings.
\section{Measuring small M$_{\bullet}$ from ground}
The main prohibitive issue with AO observations is that, even with LGS capabilities, a natural guide star is required for an optimal correction of atmospheric aberrations and there are only a handful of galaxies with a near-resolvable R$_{sph}$ that have a suitable guide star. This limitation has prompted innovative use of AO techniques, such as neglecting altogether the low-order corrections provided by the natural tip-tilt \citep{2008Msngr.131....7D}. In this study, we employ a novel method developed at Gemini Observatory for Altair LGS AO system.
\subsection{Open loop focus model of LGS}
To avoid the issue of suitable guide stars, it is, in principle, also possible to use the nucleus of the galaxy as a natural guide source for the system, provided it is sufficiently bright and compact (drop of $\geq 1$ magnitude within the central $\sim 1$ arcsec for Altair). The central light profiles of early-type galaxies show a general change from steeply rising `cusp' profiles at lower masses, towards a flatter central profile that can define a central `core' region \citep[e.g.][]{1997AJ....114.1771F}. In order to test what correction of the PSF could be expected for typical early-type galaxies, we chose NGC524 and NGC2549 which both satisfy the basic Altair requirement for nuclear guiding. Specifically, NGC524 has a core-like light profile, while NGC2549 a cusp-like profile at the HST resolution. In this proceeding we will consider the observations of NGC2549 only, the smaller of the two galaxies.
After initial observations were attempted, it became clear that the chosen nuclei, while suitable for tip-tilt correction, are too faint to be used for constraining the focus. Gemini staff implemented a procedure by which the focus correction during the science integration is controlled by a geometric function that takes into account the change in the distance to the sodium layer as the telescope position changes. In this, so-called `open-loop' focus model, the only time-dependent parameter is the altitude of the sodium layer, which is determined immediately before observing the galaxy by `tuning' the LGS AO system using a nearby bright star. When all the control loops of the system (tip-tilt, focus, and LGS) have converged with this reference source, the loops are opened and the science target is acquired. The tip-tilt and LGS control loops are then closed, using the galaxy nucleus and laser beacon respectively as reference sources. The focus loop is left open, being passively controlled by the open-loop model. After approximately one hour of science observations, the bright reference star is re-observed, so that the degradation of the PSF can be estimated (our tests show minimal degradation) and the LGS AO system can be re-optimised for further observations.
\begin{figure}
\includegraphics[height=.25\textheight]{001_dkrajnovic_fig1.ps}
\caption{\label{f:grid} Schwarzschild dynamical models and the determination of the best fitting parameters for NGC2549. Each symbol is a dynamical model. The agreement between the data and the models are described by overploted contours $\Delta \chi^2$ contours showing 1, 2 and 3$\sigma$ levels for two parameters. Further contours are spaced by a factor of 2. Large symbol on each panel marks the best fitting models. {\bf From left to right} panels show grids of models constrained by both SAURON and NIFS kinematics, models constrained by SAURON only kinematics and models constrained only using NIFS data.}
\end{figure}
We estimated the PSF of the science observations by convolving (with a double Gaussian) and rebining (to the pixel size of our observations) an HST/WFPC2 image. This image is compared with the reconstructed image of our observations, and the parameters of the PSF are varied until the best matching double-Gaussian is found. The final PSF of our LGS AO observations was: 0.17 and 0.80 arcsecs (FWHM) for the the narrow and broad Gaussian components and intensities of 0.53 and 0.47, respectively.
\begin{figure}
\includegraphics[height=.2\textheight]{001_dkrajnovic_fig2.ps}
\caption{\label{f:maps} Comparison between data and model velocity dispersion maps for NGC2549. From left to right: $\sigma$ (symmetrised) and model prediction maps for different M$_{\bullet}$ (at the best fitting M/L). The best fitting model is shown in the map adjacent to the observed $\sigma$ map, followed by a model with a too small M$_{\bullet}$ and a model with too large M$_{\bullet}$. The models were constrained using both SAURON and NIFS data. Note the colour change in the central few pixles which are a consequence of change in M$_{\bullet}$.}
\end{figure}
\subsection{Determining M$_{\bullet}$ when the resolution is smaller than R$_{sph}$}
We observed NGC2549 with NIFS IFU centered on CO absorption features starting at 2.29 microns. We spatially binned the data using Voronoi method \cite{2003MNRAS.342..345C} and achieved the SN of $\sim60$, while within the central arcsec the bins are no larger than $0.1 \times 0.1$ arcsecs. We extracted stellar kinematics using the pPXF method \cite{2004PASP..116..138C}. Large scale kinematic data were obtained from previous observations with SAURON IFU \cite{2001MNRAS.326...23B} and were previously presented in \cite{2004MNRAS.352..721E}. We constructed orbit-based Schwarzschild dynamical models following the method presented in \cite{2006MNRAS.366.1126C}.
Fig.~\ref{f:grid} shows grids of Schwarzschild dynamical models constrained using three different kinematic data sets: combined SAURON and NIFS data, SAURON data only and NIFS data only. The best fit model using SAURON and NIFS kinematics gives M$_{\bullet}$=($1.4^{+0.2}_{-1.3} )\times 10^7$ M$_{sun}$. It is evident that when using only SAURON data it is not possible to determine the lower limit of the M$_{\bullet}$, simply because the resolution of the SAURON data is 1.7 arcsec, while R$_{sph}$=0.05 arcsec, using the above best fit model. Using the NIFS dataset only it seems possible to determine even the lower limit, in spite of 0.17 arcsec resolution, but the uncertainty on M$_{\bullet}$ is very large because, due to the small field-of-view, NIFS data alone are not able to constrain the total orbital distribution and give the right mass-to-light ratio. These are, however, well constrained using SAURON large field-of-view (FoV) observations. Combining the two data sets, the high resolution and large FoV, one can constrain Schwarzschild models and determine the M$_{\bullet}$.
Fig.~\ref{f:maps} illustrates that our NIFS observations, although at about 3 times lower resolution than the estimated R$_{sph}$, are able to capture the change in the kinematics influenced by the SMBH. We illustrate this on the velocity dispersion: the $\sigma$ map of the best fit model is more similar to the observed data than $\sigma$ maps of models with a too small or a too big SMBH. To see this effect it was however, necessary to have high quality integral field data both at high resolution (NIFS) and with large FoV (SAURON).
\bibliographystyle{aipproc}
|
\section{Second law of thermodynamics}
The apparent horizon, $R_A$, in a non-flat universe is given by
\cite{Cai}
\begin{equation}
R_A=H^{-1}(1+\Omega_{k})^{-1/2},\label{ah}
\end{equation}
whereas Authors of ref. \cite{Setare1} considered $R_A=H^{-1}$ for a
non-flat universe. For the flat case, i.e. $\Omega_{k} = 0$, the
apparent horizon is same as the Hubble horizon. Therefore Eqs. (30)
and (31) in ref. \cite{Setare1} must be corrected, respectively, as
follows
\begin{equation}
{\rm d}S=\pi
(1+3\omega_{\Lambda}\Omega_{\Lambda}H^2R_{A}^2)R_{A}{\rm d}R_{A},
\end{equation}
\begin{eqnarray}
\frac{{\rm d}S}{{\rm
d}x}&=&-\pi(1+\Omega_{k}+3\omega_{\Lambda}\Omega_{\Lambda})H^{-2}\frac{\Big(H^{-1}\frac{{\rm
d}H}{{\rm
d}x}-\Omega_{k}\Big)}{(1+\Omega_{k})^3},\nonumber\\&=&-\pi
\Big(1+\Omega_{k}-\Omega_{\Lambda}-\frac{2}{c}\Omega_{\Lambda}^{3/2}\cos{y}\Big)H^{-2}\frac{\Big(H^{-1}\frac{{\rm
d}H}{{\rm
d}x}-\Omega_{k}\Big)}{(1+\Omega_{k})^3},\nonumber\\&=&-2\pi
qH^{-2}\frac{\Big(H^{-1}\frac{{\rm d}H}{{\rm
d}x}-\Omega_{k}\Big)}{(1+\Omega_{k})^3},
\end{eqnarray}
where $q$ is the deceleration parameter and from Eq. (32) in ref.
\cite{Setare1} it can be rewritten without approximation as
\begin{equation}
q=-1-H^{-1}\frac{{\rm d}H}{{\rm
d}x}=-\frac{\Omega_{\Lambda}^{3/2}\cos
y}{c}+\frac{1-\Omega_{\Lambda}}{2(1-a\gamma)}
=\frac{1}{2}\Big(1+\Omega_{k}-\Omega_{\Lambda}-\frac{2}{c}\Omega_{\Lambda}^{3/2}\cos{y}\Big).\label{qexact1}
\end{equation}
Using Eq. (\ref{ah}) one can obtain
\begin{equation}
R_A{\rm d}R_A=-H^{-2}\frac{\Big(H^{-1}\frac{{\rm d}H}{{\rm
d}x}-\Omega_{k}\Big)}{(1+\Omega_{k})^2}{\rm d}x,
\end{equation}
whereas in ref. \cite{Setare1}, $R_A{\rm d}R_A=-H^{-3}({\rm d}H/{\rm
d}x){\rm d}x$.
Using Eq. (\ref{ah}), the corrections of Eqs. (33) and (34) in ref.
\cite{Setare1} are obtained, respectively, as
\begin{equation}
\frac{{\rm d}S_{A}}{{\rm d}x}=-2\pi
H^{-2}\frac{\Big(H^{-1}\frac{{\rm d}H}{{\rm
d}x}-\Omega_{k}\Big)}{(1+\Omega_{k})^2},
\end{equation}
\begin{equation}
\frac{\rm d}{{\rm d}x}(S+S_{A})=\frac{2\pi
H^{-2}}{(1+\Omega_{k})^3}\Big(H^{-1}\frac{{\rm d}H}{{\rm
d}x}-\Omega_{k}\Big)^2.
\end{equation}
For the event horizon measured from the sphere of the horizon
named $L$, from Eq. (29) in ref. \cite{Setare1} and using
$\rho_{\Lambda}=\frac{3c^2}{8\pi}L^{-2}$ and $E=\frac{4}{3}\pi
L^3\rho=\frac{1}{2}c^2L$, Eq. (40) in ref. \cite{Setare1} is
corrected as
\begin{equation}
\frac{{\rm d}S}{{\rm d}x}=\frac{-\pi c^4
}{H^2\Omega_{\Lambda}^2}\Big(\frac{\Omega_{\Lambda}+3\omega_{\Lambda}\Omega_{\Lambda}}{H}\frac{{\rm
d}H}{{\rm
d}x}+\frac{1+3\omega_{\Lambda}}{2}\Omega_{\Lambda}^{'}\Big),\label{dSdx1}
\end{equation}
where $\Omega_{\Lambda}^{'}$ is given by Eq. (22) in ref.
\cite{Setare1} and can be rewritten as
\begin{equation}
\Omega_{\Lambda}^{'}=\frac{2}{c}\Omega_{\Lambda}^{3/2}\cos
y+2q\Omega_{\Lambda}.\label{OmegaLp}
\end{equation}
Due to have a correct dimension, the holographic dark energy (DE)
density, $\rho_{\Lambda}$, given by Eq. (5) in ref. \cite{Setare1}
should be corrected as $\rho_{\Lambda}=\frac{3c^2}{8\pi}L^{-2}$,
where we take $G=1$.
Using Eq. (\ref{dSdx1}) for the evolution of entropy of the DE
inside the universe enclosed by the horizon $L$ and Eq. (41) in
ref. \cite{Setare1} for the evolution of the geometric entropy of
the horizon, one can obtain
\begin{equation}
\frac{\rm d}{{\rm d}x}(S+S_{\rm L})=\frac{\pi c^4
}{H^2\Omega_{\Lambda}^2}\Big(\Omega_{\Lambda}+3\omega_{\Lambda}\Omega_{\Lambda}+\frac{2\Omega_{\Lambda}}{c^2}\Big)
\Big(1+q-\frac{\Omega_{\Lambda}^{'}}{2\Omega_{\Lambda}}\Big).\label{dStot1}
\end{equation}
Here we would like to correct Eq. (42) in ref. \cite{Setare1}
using the approximation given by Eq. (32) in ref. \cite{Setare1}
for the deceleration parameter $q$. This procedure is same as that
used by Authors of ref. \cite{Setare1}. To do this, from Eq. (32)
in ref. \cite{Setare1} we have
\begin{equation}
\frac{\Omega_{\Lambda}^{3/2}\cos
y}{c}\simeq-q+\frac{1-\Omega_{\Lambda}}{2},\label{qapprox}
\end{equation}
then substituting Eq. (\ref{qapprox}) in both Eq. (9) in ref.
\cite{Setare1} and Eq. (\ref{OmegaLp}) we obtain
\begin{equation}
3\omega_{\Lambda}\Omega_{\Lambda}\simeq 2q-1,\label{approx1}
\end{equation}
\begin{equation}
\Omega_{\Lambda}^{'}\simeq(2q-1)(\Omega_{\Lambda}-1).\label{approx2}
\end{equation}
Substituting Eqs. (\ref{approx1}) and (\ref{approx2}) in Eq.
(\ref{dStot1}), we get the corrected form of Eq. (42) in ref.
\cite{Setare1} with using the approximation given by Eq. (32) in
ref. \cite{Setare1} as
\begin{equation}
\frac{\rm d}{{\rm d}x}(S+S_{\rm L})=\frac{\pi c^4
}{H^2\Omega_{\Lambda}^2}\left\{\Big(2q+\frac{2\Omega_{\Lambda}}{c^2}+\Omega_{\Lambda}-1\Big)
\Big[\Big(\frac{1-\Omega_{\Lambda}}{2\Omega_{\Lambda}}\Big)(2q-1)+1+q\Big]\right\},\label{GSL11}
\end{equation}
or
\begin{eqnarray}
\frac{\rm d}{{\rm d}x}(S+S_{\rm L})=\frac{\pi c^4
}{H^2\Omega_{\Lambda}^2}\left\{(1+q)\Big(2q+\frac{2\Omega_{\Lambda}}{c^2}\Big)
+\Big(\frac{1-\Omega_{\Lambda}}{\Omega_{\Lambda}}\Big)(2q-1)\Big(1+q+\frac{\Omega_{\Lambda}}{c^2}\Big)
\right.\nonumber\\\left.-(1-\Omega_{\Lambda})
\Big[(2q-1)\Big(\frac{3-\Omega_{\Lambda}}{2\Omega_{\Lambda}}\Big)+1+q\Big]\right\}.\label{GSL12}
\end{eqnarray}
Taking $\Omega_{\Lambda}=0.73$ and $c=1$ given by ref.
\cite{Setare1} for the present time, Eq. (\ref{GSL11}) gives
\begin{eqnarray}
\frac{\rm d}{{\rm d}x}(S+S_{\rm L})&=&\frac{5.89528
}{H^2}(0.815068+1.36986q)(1.19+2q),\nonumber\\&=&\frac{16.1515
}{H^2}(q+0.5949)^2\geq 0,\label{GSL13}
\end{eqnarray}
which compared to Eq. (43) in ref. \cite{Setare1} shows that in
contrary to the conclusion of Authors of ref. \cite{Setare1}, the
generalized second law (GSL) of thermodynamics for the holographic
DE in a non-flat universe enveloped by the horizon $L$ is
satisfied for the present time independently of the deceleration
parameter $q$.
Here we would like to correct again Eq. (42) in ref. \cite{Setare1}
but this time without using the approximation given by Eq. (32) in
ref. \cite{Setare1} for the deceleration parameter $q$. To do this,
from the exact relation for $q$ given by Eq. (32) in ref.
\cite{Setare1} we have
\begin{equation}
\frac{\Omega_{\Lambda}^{3/2}\cos
y}{c}=-q+\frac{1-\Omega_{\Lambda}}{2(1-a\gamma)},\label{qexact2}
\end{equation}
then substituting Eq. (\ref{qexact2}) in both Eq. (9) in ref.
\cite{Setare1} and Eq. (\ref{OmegaLp}) we obtain
\begin{equation}
3\omega_{\Lambda}\Omega_{\Lambda}=2q-\Omega_{\Lambda}-\Big(\frac{1-\Omega_{\Lambda}}{1-a\gamma}\Big),\label{exact1}
\end{equation}
\begin{equation}
\Omega_{\Lambda}^{'}=(\Omega_{\Lambda}-1)\Big(2q-\frac{1}{1-a\gamma}\Big).\label{exact2}
\end{equation}
Substituting Eqs. (\ref{exact1}) and (\ref{exact2}) in Eq.
(\ref{dStot1}), we get the corrected form of Eq. (42) in ref.
\cite{Setare1} without using the approximation given by Eq. (32) in
ref. \cite{Setare1} as
\begin{equation}
\frac{\rm d}{{\rm d}x}(S+S_{\rm L})=\frac{\pi c^4
}{H^2\Omega_{\Lambda}^2}\left\{\Big[2q+\frac{2\Omega_{\Lambda}}{c^2}-\Big(\frac{1-\Omega_{\Lambda}}{1-a\gamma}\Big)\Big]
\Big[\Big(\frac{1-\Omega_{\Lambda}}{2\Omega_{\Lambda}}\Big)\Big(2q-\frac{1}{1-a\gamma}\Big)+1+q\Big]\right\},\label{GSL21}
\end{equation}
or
\begin{eqnarray}
\frac{\rm d}{{\rm d}x}(S+S_{\rm L})=\frac{\pi c^4
}{H^2\Omega_{\Lambda}^2}\left\{(1+q)\Big(2q+\frac{2\Omega_{\Lambda}}{c^2}\Big)
+\Big(\frac{1-\Omega_{\Lambda}}{\Omega_{\Lambda}}\Big)\Big(2q-\frac{1}{1-a\gamma}\Big)\Big(1+q+\frac{\Omega_{\Lambda}}{c^2}\Big)
\right.\nonumber\\\left.-(1+q)\Big(\frac{1-\Omega_{\Lambda}}{1-a\gamma}\Big)
-\Big(2q-\frac{1}{1-a\gamma}\Big)\Big(\frac{1-\Omega_{\Lambda}}{2\Omega_{\Lambda}}\Big)
\Big(2+\frac{1-\Omega_{\Lambda}}{1-a\gamma}\Big)\right\}.\label{GSL22}
\end{eqnarray}
Note that Eq. (\ref{GSL21}) reduces to Eq. (\ref{GSL11}) when
$\gamma:=\Omega_{k}^{0}/\Omega_{\rm m}^{0}$ goes to zero.
For $\Omega_{\Lambda}=0.73$, $c=1$ and $\gamma\sim0.04$ given by
ref. \cite{Setare1} for the present time $a=1$, Eq. (\ref{GSL21})
gives
\begin{eqnarray}
\frac{\rm d}{{\rm d}x}(S+S_{\rm L})&=&\frac{5.89528
}{H^2}(0.807363+1.36986q)(1.17875+2q),\nonumber\\&=&\frac{16.1515
}{H^2}(q+0.5894)^2\geq 0,\label{GSL23}
\end{eqnarray}
which shows that same as the result obtained by Eq. (\ref{GSL13})
and in contrary to the conclusion of Authors of ref.
\cite{Setare1}, the GSL for the horizon $L$ is satisfied again for
the present time independently of the deceleration parameter $q$.
Although Authors of ref. \cite{Setare1} have considered both of the
DE and dark matter (DM) in their model (see Eq. (18) in ref.
\cite{Setare1}), they have not taken into account the contribution
of the DM in the GSL (see again Eq. (42) in ref. \cite{Setare1}).
Therefore to complete the calculations, the contribution of the
entropy of the DM should be considered in the GSL. To do this, from
Eq. (29) in ref. \cite{Setare1} and using $P_{\rm m}=0$ and $E_{\rm
m}=\frac{4}{3}\pi L^3\rho_{\rm m}$, the evolution of entropy of the
DM inside the universe enclosed by the horizon $L$ is obtained as
\begin{equation}
\frac{{\rm d}S_{\rm m}}{{\rm d}x}=\frac{3\pi c^4
}{H^2\Omega_{\Lambda}^2}\Big(q-\frac{\Omega_{\Lambda}^{'}}{2\Omega_{\Lambda}}\Big)\Omega_{\rm
m}.\label{dSm}
\end{equation}
Finally, using Eqs. (\ref{dStot1}) and (\ref{dSm}), the GSL due to
different contributions of the DE, DM and horizon $L$ can be
obtained as
\begin{equation}
\frac{\rm d}{{\rm d}x}(S+S_{\rm L}+S_{\rm m})=\frac{\pi c^4
}{H^2\Omega_{\Lambda}^2}\left\{\Big(\Omega_{\Lambda}+3\omega_{\Lambda}\Omega_{\Lambda}+\frac{2\Omega_{\Lambda}}{c^2}+3\Omega_{\rm
m}\Big)
\Big(1+q-\frac{\Omega_{\Lambda}^{'}}{2\Omega_{\Lambda}}\Big)-3\Omega_{\rm
m}\right\}.\label{dStot2}
\end{equation}
Using Eq. (9) in ref. \cite{Setare1} and Eq. (\ref{OmegaLp}), we can
rewrite Eq. (\ref{dStot2}) as
\begin{eqnarray}
\frac{{\rm d}}{{\rm d}x}(S+S_{\rm L}+S_{\rm m})=\frac{\pi
c^4}{H^2\Omega_{\Lambda}^2}\left\{\frac{2\Omega_{\Lambda}}{c^2}\Big(1-\frac{\sqrt{\Omega_{\Lambda}}}{c}\cos{y}\Big)\Big(1-c\sqrt{\Omega_{\Lambda}}\cos{y}\Big)
\right.\nonumber\\\left.-3(1+\Omega_{k}-\Omega_{\Lambda})\frac{\sqrt{\Omega_{\Lambda}}}{c}\cos{y}\right\},\label{dStot3}
\end{eqnarray}
which is same as Eq. (1.6) in ref. \cite{Karami} with $b^2=0$.
Taking $\Omega_{\Lambda}=0.73$, $\Omega_{k}=0.01$, $c=1$
\cite{Setare1} and $\cos y=0.99$ \cite{Karami} for the present time,
Eq. (\ref{dStot3}) gives
\begin{equation}
\frac{\rm d}{{\rm d}x}(S+S_{\rm L}+S_{\rm
m})=-\frac{3.98421}{H^2}<0,
\end{equation}
which shows that same as the result obtained by \cite{Karami} and in
contrary to the conclusions of Eqs. (\ref{GSL13}) and (\ref{GSL23}),
the GSL is violated at the present time for a non-flat universe
containing the holographic DE and DM and enveloped by the event
horizon measured from the sphere of the horizon named $L$.
|
\section{Introduction}
The flares in the solar corona are believed to be due to sudden
restructuring of the stressed magnetic field. The energy is
released in the form of thermal as well as non-thermal
radiation and energetic charged particles. In powerful flares
the energetic particles can penetrate the dense chromosphere to
reach down to the photosphere where they heat-up the
photosphere leading to white-light flares. It has been observed
that photospheric changes are accompanied during these highly
energetic events in the form of: (i) change in morphology, (ii)
change in magnetic flux, (iii) change in magnetic shear angle
(the angle between observed field azimuth and potential field
azimuth) and (iv) proper motion.
Here we focus on the local changes, i.e., changes seen in
small-scale features like penumbral filaments during flares.
Such studies require seeing-free high-resolution observations
at a high-cadence. This is possible with the 50 cm Solar
Optical Telescope (SOT) onboard {\it Hinode} spacecraft
\citep{Kosugi2007,Tsuneta2008,Ichimoto2008,Suematsu2008}. Here,
we present the observations of a $\delta$-sunspot in active
region NOAA 10930 during a X-class flare on 13 December 2006
at 02:20 UT by {\it Hinode}. The two ribbons could be seen in
G-band and Fe I 630.2 nm Stokes-I and V images
\citep{Isobe2007}. Earlier, we had reported the lateral motion
of penumbral filaments during the flare interval
\citep{Gosain2009}. Here, we present the converging motion of
the two patches, one in either polarity, located on either side
of the PIL. Also, the pre-flare brightening in kernels located
along the PIL as seen in Ca II H line, is discussed.
\section{Observations and Data Analysis}
The X-class flare of 13 December 2006 occurred in an active
region numbered NOAA 10930 during 02:20 UT. The region
consisted of a $\delta$-sunspot with almost N-S orientation of
the bipole. The high-resolution filtergrams in G-band (430.5
nm), Fe I 630.2 nm and Ca II H (396.8 nm) wavelengths were
obtained by Filtergraph (FG) instrument onboard {\it Hinode}
Solar Optical Telescope (SOT). The images were sampled
spatially with 0.1 arc-sec per pixel and temporally with one
image every two minutes. The filtergrams were calibrated for
dark current, flat field and bad pixels using the standard
SolarSoft IDL libraries. A time sequence of filtergrams was
selected between 02:00 and 03:00 UT for analysis. The images
were aligned globally by choosing a large field-of-view for
registration. The registration procedure for the sunspot in
Hinode filtergrams has been described elsewhere in detail
\citep{Gosain2009}.
\begin{figure}[!ht]
\begin{center}
\includegraphics[scale=0.30,keepaspectratio=true,trim=300 55 0 35]{gosain_fig1.eps}\includegraphics[scale=0.7,keepaspectratio=true,trim=10 0 100 0]{gosain_fig2.eps}
\end{center}
\caption{The top and bottom panels on the left show the G-band intensity and Fe I Stokes-V images, respectively.
Boxes `1' and `2' mark the location of the two patches where the velocity $V_x$ and G-band intensity are tracked during the flare interval.
These are plotted in panels (a)-(d) on the right. }\label{fig1}
\end{figure}
\section{Results}
\subsection{Converging Motion of Opposite Polarity Pacthes}
The figure 1 shows the G-band filtergram of the
$\delta$-sunspot taken during 13 December 2006 at 02:00 UT. The
two boxes marked `1' and `2' correspond to patches located in
the two spots of opposite polarity. Within the two patches we
tracked the relative shifts between two subsequent frames
during the interval 02:06 to 03:00 UT. This shift $\Delta x$
within time interval $\Delta t$ gives velocity $V_x$. The
panels (a) and (b) of figure 2, show the time profile of $V_x$
corresponding to the patches within the boxes `1' and `2'.
It may be noticed that : (i) First signatures of motion begin at 02:18 UT which is
about two minutes earlier than the peak of the flare seen in microwave observations of \cite{Zhang2009}.
(ii) For both patches `1' and `2' there are two phases of motion, initially
in one direction and after about four minutes in the opposite
direction. The peak velocity of these two phases of motions are
represented by two vertical lines which are separated by about
four minutes. (iii) The motion in the two boxes are in opposite
direction in both phases. Initially, the motion in two boxes is
away from each other and later it is towards each other, like a
converging motion. (iv) The motion in box `2', which is closer
to the neutral line, starts earlier than motion in box `1' by
about four minutes. (v) The second phase of motion, i.e.,
converging motion, continues for long duration lasting more
than 40 minutes.
The panels (c) and (d) of figure 2, show the mean G-band
intensity within the two boxes. The two ribbons during the
flare are visible in G-band images. Therefore, the enhancement
in G-band intensity (marked by vertical lines) in panels (c)
and (d) of figure 2, correspond to the instant when the ribbons
move across these boxes. Here also, we notice that the
intensity in box `2', which is located closer to the neutral
line, peaks four minutes earlier than intensity in box `1',
which may explain why motion in box `2' starts earlier than
motion in box `1' by about four minutes.
\subsection{Pre-flare Brightening along PIL in Ca II H}
The figure 3 shows the Ca II H filtergrams of the sunspot
during 13 December 2006 at 02:04, 02:08, 02:14 and 02:20 UT. We
notice that although the flare onset time is 02:20 UT according
to microwave flux observations, initial brightenings can be
noticed in Ca II H filtergrams as early as 02:04 UT in panel
(a). The location of this brightening is marked by a
rectangular box. Further, it may be noticed that the
brightening is located along the neutral line, and in
subsequent panels (a)-(c) the length of the brightening
increases along the neutral line. Finally, the flare develops
into a two ribbon flare as seen in panel (d).
\begin{figure}[!ht]
\begin{center}
\includegraphics[scale=0.62,keepaspectratio=true,trim=0 45 0 15]{gosain_fig3.eps}
\end{center}
\caption{The panels (a)-(c) show the Ca II H filtergrams during pre-flare phase. The boxes mark the location of pre-flare
brightening along neutral line. The panel (d) shows the filtergram at 02:20 UT. }\label{fig3}
\end{figure}
\section{Discussion and Conclusions}
We have studied the evolution of small scale features in a
flaring sunspot using high resolution space based observations.
The fine structure of the sunspot penumbra as seen in high
resolution G-band images is believed to outline the magnetic
field lines. Using the penumbral filaments as a proxy for
magnetic field structure of the sunspot we study the changes in
the structure during X-class flare of 13 December 2006. A two
phased motion is seen in patches of either poalrity, i.e.,
boxes `1' and `2'. First phase of motion lasts for about four
minutes, directed away from the neutral line, and another phase
of motion lasts for more than 40 minutes, directed towards
neutral line, i.e., converging.
Further, we notice that in Ca II H images the pre-flare
brightening is clearly visible in elongated kernels, located
along the PIL. This brightening is seen as early as 16 minutes
prior to the flare onset.
To understand the changes in magnetic field configuration
during flares one requires high-cadence vector magnetograms
obtained with high-resolution which are expected from upcoming
space missions like Helioseismic and Magnetic Imager (HMI) and
Solar Orbiter.
\acknowledgements
{\it Hinode} is a Japanese mission developed and launched by
ISAS/JAXA, with NAOJ as a domestic partner and NASA and STFC
(UK) as international partners. It is operated by these
agencies in co-operation with ESA and NSC (Norway).
|
\section{Introduction}
Non-equilibrium phenomena in superconductors have been investigated intensively since the 1970s. Both experimental and theoretical investigations of charge imbalance have focused mostly on temperatures near the critical temperature $T_\mathrm{c}$ of the superconductor. In this regime, charge imbalance is easily accessible to experiments basically due to the divergence of the signal towards $T_\mathrm{c}$, and excellent theoretical approximations are available.\cite{clarke1972,tinkham1972,langenberg} Recently, the investigation of non-local transport properties of superconductors has gained new impetus from two separate but loosely related fields. One is the investigation of spin-dependent transport,\cite{johnson1994,*valenzuela2006,*poli2008,*luo2009} and in particular spin diffusion and accumulation in the context of spintronics. The second is the investigation of coherent non-local effects such as crossed Andreev reflection,\cite{byers1995,*deutscher2000,*beckmann2004,*beckmann2007,*russo2005,*cadden2006,*cadden2007,*cadden2009,*asulin2006,*kleine2009,*hofstetter2009,*herrmann2010,*huebler2010} which might be useful for quantum information processing. In the diffusive quasi-onedimensional structures typically used for such experiments, the magnitude of the signals due to these phenomena scale with the normal-state resistance of the superconductor over a characteristic length scale, which is given by the charge-imbalance relaxation length $\lambda_{Q^*}\sim 10~\mathrm{\mu m}$,\cite{langenberg} the spin-diffusion length $\lambda_\mathrm{sf}\sim 1~\mathrm{\mu m}$,\cite{johnson1985,valet1993,jedema2002,jedema2003} and the coherence length $\xi\sim 0.1~\mathrm{\mu m}$,\cite{golubev2009} respectively. Consequently, the signals due to charge imbalance are often the largest, and make an unambiguous identification of the other phenomena difficult. In particular for the investigation of crossed Andreev reflection, experiments far below $T_\mathrm{c}$ are necessary. Despite the vast amount of experimental and theoretical literature on charge imbalance, surprisingly little is known about the subject at very low temperatures.\cite{yagi2006} This is probably due to the fact that no simple theoretical models are available for this regime, and that the interpretation of the widely-used non-local resistance measurement scheme becomes increasingly difficult as temperature is lowered.
In this article, we report on a detailed investigation of non-local {\em conductance} rather than {\em resistance} in superconductor/normal-metal hybrid structures at temperatures $T\ll T_\mathrm{c}$. The samples consist of a quasi-onedimensional superconducting wire, with several normal-metal tunnel junctions attached to it. From these experiments, we deduce the charge-imbalance relaxation length $\lambda_{Q^*}$ as a function of bias voltage, temperature and magnetic field. The results allow a detailed comparison to theoretical predictions for different charge-imbalance relaxation mechanisms, and and an assessment of the possible impact on the interpretation of experiments on crossed Andreev reflection and spin diffusion.
\begin{figure}
\includegraphics[width=\columnwidth]{fig1.eps}
\caption{\label{fig_sem}(Color online) SEM image of sample A illustrating the experimental scheme.
Five copper (Cu) fingers are connected by tunnel contacts to an aluminum bar (Al). For one of the possible injector-detector pairs the bias and measurement scheme used for charge imbalance detection is shown.}
\end{figure}
\section{Theory}\label{sec_theory}
Charge imbalance (CI) can be described using two different theoretical frameworks, the quasiparticle, or two-fluid, approach,\cite{clarke1972,tinkham1972,tinkham1972b,pethick1979} and quasiclassical Green's functions.\cite{schmid1975} The relation of these two approaches has been discussed extensively in the literature.\cite{clarke1979,entin1979} We simply note here that the Green's function method is more general, and can be applied, e.g., to inhomogeneous superconducting states and situations with strong pair breaking. We will nevertheless use the quasiparticle approach, due to its conceptual (and computational) simplicity, and discuss its shortcomings where necessary.
We consider a quasi-onedimensional superconductor of length $L$ along the $x$ axis, with several normal-metal electrodes attached via tunnel junctions. These electrodes will serve both to inject nonequilibrium quasiparticles into the superconductor, and to detect them. The quasiparticle energies $E$ are given by
\begin{equation}
E=\sqrt{\epsilon^2+\Delta^2},
\end{equation}
where $\epsilon$ is the normal-state electron energy relative to the Fermi energy, and $\Delta$ is the pair potential. The normalized quasiparticle density of states is
\begin{equation}
n(E)=\Theta(E-\Delta)\frac{E}{|\epsilon|},
\end{equation}
where $\Theta$ is the Heaviside function. Charge imbalance is defined as
\begin{equation}
Q^* = 2N_0\int_{-\infty}^{\infty}q(\epsilon)f(\epsilon)d\epsilon,\label{equ_qstar}
\end{equation}
where $N_0$ is the density of states per spin at the Fermi energy in the normal state, $q(\epsilon)=\epsilon/E$ is the effective quasiparticle charge in units of the elementary charge $e$, and $f(\epsilon)$ is the quasiparticle distribution function. By using $\epsilon$ rather than $E$ as independent variable, we keep track of the electron-like ($\epsilon>0$) and hole-like ($\epsilon<0$) branch of the quasiparticle spectrum. In thermal equilibrium, $f(\epsilon)$ is given by the Fermi distribution $f_\mathrm{T}(E)$.
It is apparent from (\ref{equ_qstar}) that $Q^*$ is non-zero only if the populations of the electron- and hole-like branches are unequal, i.e., $f(\epsilon)-f(-\epsilon)\not\equiv 0$. Besides this charge-mode (transverse) non-equilibrium, there is also an energy-mode (longitudinal) non-equilibrium characterized by $f(\epsilon)+f(-\epsilon)-2f_\mathrm{T}(E)\not\equiv 0$. As will be shown below, the latter enters transport properties only indirectly via the self-consistency equation for the pair potential,
\begin{equation}
1+\mathcal{V}\int_{-\infty}^{\infty}\frac{1-2f(\epsilon)}{2\sqrt{\epsilon^2+\Delta^2}}d\epsilon=0,\label{equ_sc}
\end{equation}
where $\mathcal{V}$ is the pairing interaction.
The electric current through a tunnel junction between a normal metal held at bias voltage $V$ and a nonequilibrium superconductor is given by the sum of the usual "Giaever" tunnel current $I_\mathrm{T}(V)$,\cite{giaever1960} and a bias-independent extra current $I_{Q^*}$ due to CI,\cite{tinkham1972b}
\begin{equation}
I(V)=I_\mathrm{T}(V)+I_{Q^*}.\label{equ_current}
\end{equation}
Here,
\begin{equation}
I_\mathrm{T}(V) =\frac{G_\mathrm{N}}{e} \int\limits_{0}^\infty n(E) \left( f_\mathrm{T}(E-eV)-f_\mathrm{T}(E+eV) \right)dE \label{equ_itunnel}
\end{equation}
where $G_\mathrm{N}$ is the normal-state tunnel conductance, and the Fermi functions describe the electron occupation in the normal metal. The excess current is given by
\begin{equation}
I_{Q^*} =\frac{G_\mathrm{N}Q^*}{2eN_0}.\label{equ_iq}
\end{equation}
As mentioned above, the current $I_{Q^*}$ depends on the nonequilibrium distribution $f(\epsilon)$ only via the charge mode, whereas $I_\mathrm{T}(V)$ indirectly depends on the energy mode via the gap equation (\ref{equ_sc}).
In the vicinity of the critical temperature $T_\mathrm{c}$ of the superconductor, the deviation of the distribution function $f(\epsilon)$ from equilibrium is small and can be approximated by a Fermi function with a shift $\Delta\mu$ of the chemical potential of the quasiparticles relative to the chemical potential of the Cooper pairs.\cite{pethick1979} The most convenient measurement technique in this situation is the widely used non-local voltage detection scheme, in which the voltage between a normal-metal detector junction and the superconductor is measured. The voltage adjusts such that $I(V_\mathrm{det})=0$, i.e., the current $I_\mathrm{Q^*}$ is cancelled by the backflow $I_\mathrm{T}(V_\mathrm{det})$. In the regime where the chemical-potential model applies, this voltage is given by $eV_\mathrm{det}=\Delta\mu$, i.e., it directly measures the single parameter that characterizes CI. Voltage detection, however, is less useful in the low-temperature regime that we are interested in for two reasons. First, at low temperature the chemical-potential model breaks down, and the detector voltage has no longer a simple physical meaning. Second, the non-linearity and temperature dependence of the tunnel current $I_\mathrm{T}$ distort the measured signal. This has already been noted in the earliest experiment on charge imbalance,\cite{clarke1972} where the raw data at low temperature had to be corrected by the temperature dependence of the detector junction for comparison with theory. To avoid these shortcomings, we measure the detector current $I_\mathrm{det}(V_\mathrm{det}=0)=I_\mathrm{Q^*}$. Here, the only detector property that enters is the (constant) normal-state conductance, and therefore $I_\mathrm{det}$ is a direct measure of $Q^*$ even for arbitrary non-equilibrium distributions. Experimentally, we measure the differential non-local conductance, and we will now derive the expression used to evaluate our results.
The distribution function $f(\epsilon)$ is driven out of equilibrium by tunnel injection into a fixed volume $\Omega$ of the superconductor at a rate given by\cite{tinkham1972b,chi1980}
\begin{align}
\frac{\partial f}{\partial t}=\gamma_\mathrm{tun}=&\frac{1}{\tau_\mathrm{tun}}\bigg\{
\frac{1}{2}\left(1+\frac{\epsilon}{E}\right)\left(f_\mathrm{T}(E-eV)-f(\epsilon)\right)\nonumber\\*
& -\frac{1}{2}\left(1-\frac{\epsilon}{E}\right)\left(f(\epsilon)-f_\mathrm{T}(E+eV)\right)\bigg\},\label{equ_gamma_inj}
\end{align}
where $\tau_\mathrm{tun}^{-1}=G_\mathrm{N}/2N_0\Omega e^2$. Here, we use the full non-equilibrium distribution $f(\epsilon)$ in the superconductor, rather than the thermal distribution used in Ref. \onlinecite{chi1980}. However, we retain the assumption that the normal-metal electrode is at thermal equilibrium, described by the shifted Fermi functions $f_\mathrm{T}(E\pm eV)$. It is customary to define an injection efficiency $F^*=e\Omega\dot{Q}^*/I$. Since we will be interested in the differential conductance, we use the spectral quantity $f^*(E)=\epsilon^2/E^2$ rather than the usual integral definition. The expression for $f^*$ neglects the non-equilibrium contributions in (\ref{equ_current}) and (\ref{equ_gamma_inj}), which are unimportant for an injector junction biased at $|eV|\gtrsim \Delta$. On the other hand, a detector junction held at $V=0$ leads to charge imbalance relaxation at a rate $\tau_\mathrm{tun}^{-1}$, and care must be taken to ensure that this rate is negligible compared to bulk relaxation mechanisms for non-invasive detection.\cite{lemberger1984} Once injected, non-equilibrium quasiparticles diffuse along the wire with an energy-dependent diffusion constant $D(E)=v_\mathrm{g}D_\mathrm{N}$,\cite{entin1979,ullom1998} where $D_\mathrm{N}$ is the normal-state diffusion constant, and $v_\mathrm{g}=|\epsilon|/E$ is the normalized group velocity of quasiparticles. Concomitantly, the non-equlibrium distribution relaxes due to different mechanisms, including inelastic electron-phonon scattering,\cite{tinkham1972,tinkham1972b,kaplan1976} elastic impurity scattering in the presence of gap anisotropy,\cite{tinkham1972b,chi1979} and magnetic pair breaking.\cite{schmid1975,lemberger1981,lemberger1981b} Charge-imbalance relaxes over a characteristic time $\tau_{Q^*}=Q^*/\dot{Q}^*$, which we will also assume to be energy-dependent, leading to an exponential relaxation on the length scale $\lambda_{Q^*}=\sqrt{D\tau_{Q*}}$. The steady-state distribution is achieved when injection and relaxation rates are equal. Assuming that the effective injection volume is given by a wire section of length $2\lambda_{Q^*}$, we find that the non-local conductance due to charge imbalance is given by
\begin{equation}\label{equ_gnl}
g_\mathrm{nl} =\frac{dI_\mathrm{det}}{dV_\mathrm{inj}} = g^*G_\mathrm{inj}G_\mathrm{det}\frac{\rho_\mathrm{N}\lambda_{Q^*}}{2A}\exp\left(-\frac{d}{\lambda_{Q^*}}\right),
\end{equation}
where $G_\mathrm{inj}$ and $G_\mathrm{det}$ are the normal-state conductances of the injector and detector junctions, $\rho_\mathrm{N}$ is the normal-state resistivity of the superconductor, and $A$ is the cross-section of the wire. The factor $g^*$ accounts for thermal smearing, injection efficiency, etc., and is of order unity. At $T=0$, and neglecting energy relaxation, $g^*=n(E)f^*/v_\mathrm{g}=\Theta(eV_\mathrm{inj}-\Delta)$. Equation (\ref{equ_gnl}) will form the basis for our data analysis.
\section{Experiment}\label{sec_experiment}
Fig.~\ref{fig_sem} shows the relevant part of one of the three samples discussed, together with a scheme of the measurement setup. All samples are fabricated by e-beam lithography and shadow evaporation techniques. In the following, the processing details are explained with the help of the sample parameters listed in Table \ref{tab_prop}.
\begin{table
\caption{\label{tab_prop}Characteristic parameters of the three samples A-C.
Aluminum film thickness $t_\mathrm{Al}$, width $w_\mathrm{Al}$, and normal state resistivity $\rho_\mathrm{Al}$ at $T=4.2~\mathrm{K}$, range of contact distances $d$ and contact conductances $G$, critical temperature $T_\mathrm{c}$, critical field $B_\mathrm{c}$ and energy gap $\Delta_0$.}
\begin{ruledtabular}
\begin{tabular}{lcccccccc}
& $t_\mathrm{Al}$ & $w_\mathrm{Al}$ & $\rho_\mathrm{Al}$ & $d$ & $G$ & $T_\mathrm{c}$ & $B_\mathrm{c}$ & $\Delta_0$ \\
& (nm) & (nm) & ($\mathrm{\mu \Omega cm}$) & ($\mu$m) & ($\mu$S) & (K) & (T) & ($\mathrm{\mu eV}$) \\ \hline
A & 30 & 140 & 4.9 & 1-12 & 230-270 & 1.39 & 0.53 & 208 \\
B & 30 & 190 & 4.9 & 0.2-9.7 & 260-350 & 1.38 & 0.58 & 218 \\
C & 12.5 & 140 & 11.1 & 0.5-6.5 & 370-490 & 1.5 & 1.73 & 225 \\
\end{tabular}
\end{ruledtabular}
\end{table}
First, a copper film with thickness $t_\mathrm{Cu1}=25-30~\mathrm{nm}$ is evaporated onto a thermally oxidized silicon substrate. This first layer will form Ohmic interconnections to the subsequent layers. In a second evaporation step, the superconductor, an aluminum bar of thickness $t_\mathrm{Al}$ and width $w_\mathrm{Al}$, is deposited under a different angle, shifting the design to create intended overlaps only. In order to provide the formation of an insulating layer on the top, the aluminum is then oxidized {\em in situ} by applying the equivalent of $1~$Pa of oxygen for $10~$min. In the final evaporation step, a second layer of copper ($t_\mathrm{Cu2}$ = 30~nm) is deposited under a third angle forming five tunnel contacts with the aluminum. The contact distances between neighbouring copper fingers presented in sample A are about 1~$\mu$m, 2~$\mu$m, 4~$\mu$m and 5~$\mu$m from left to right, respectively.
For the transport experiment the samples are mounted into a shielded box thermally anchored to the mixing chamber of a dilution refrigerator. The measurement lines are fed through a series of filters to eliminate rf and microwave radiation from the shielded box.
A voltage $V_\mathrm{ex}$ consisting of a dc bias and a low-frequency ac excitation is applied to the injector contact, and the ac part of the resulting current $I_\mathrm{inj}$ is measured with a lock-in technique. Simultaneously, the ac current $I_\mathrm{det}$ is measured through the second contact, the detector. The local and non-local differential conductances $g_\mathrm{inj}=dI_\mathrm{inj}/dV_\mathrm{inj}$ and $g_\mathrm{nl}=dI_\mathrm{det}/dV_\mathrm{inj}$ are extracted from the ac signals.
Voltage and current polarities are indicated in Fig.~\ref{fig_sem} by plus signs and arrows, respectively.
All contacts are measured in a three-point configuration with a series resistance of about 90~$\Omega$ coming from the measurement line.
\section{Results}\label{sec_results}
\subsection{Contact and film characterization}
\begin{figure}
\includegraphics[width=\columnwidth]{fig2.eps}
\caption{\label{fig_local}(Color online)
Local differential conductance $g=dI/dV$ of one contact of sample B as a function of bias voltage $V$ for (a) different temperatures $T$ and (b) different applied magnetic fields $B$. Symbols represent measured data, lines are fits to the model described in the text.}
\end{figure}
To characterize the tunnel contacts and the properties of the superconducting film, we first discuss the tunnel spectra of the individual junctions. Fig.~\ref{fig_local} shows the local differential conductance $g_\mathrm{inj}$ as a function of bias $V_\mathrm{inj}$ for one of the tunnel junctions of sample B. Panel (a) represents the temperature dependence at zero magnetic field, whereas panel (b) depicts the variation with magnetic field applied in the plane of the substrate along the Cu wires for constant temperature $T = 50~\mathrm{mK}$. At lowest temperature and zero field, the differential conductance is completely suppressed at low bias, with sharp peaks at the energy gap, showing the high quality of the oxide tunnel barrier. Upon increasing the temperature or the magnetic field, the features are broadened and the gap is reduced. Since we are particularly interested in the dependence on magnetic field, we have used a slightly more elaborate model than (\ref{equ_itunnel}) to fit the data. In the presence of a magnetic field, the spin-resolved density of states in the superconductor can be described by\cite{maki1964b}
\begin{equation}
n_{\pm}(E)=\frac{1}{2}\mathrm{Re}\left(\frac{u_\pm}{\sqrt{u_\pm^2-1}}\right),
\end{equation}
where the complex quantities $u_\pm$ have to be determined from the implicit equation
\begin{equation}
\frac{E\mp\mu_\mathrm{B}B}{\Delta}=u_\pm\left(1-\frac{\Gamma}{\Delta}\frac{1}{\sqrt{1-u_\pm^2}}\right)
+b_\mathrm{so}\left(\frac{u_\pm-u_\mp}{\sqrt{1-u_\mp^2}}\right).\nonumber
\end{equation}
Here, $\mu_\mathrm{B}$ is the Bohr magneton, $\Gamma$ is the pair-breaking parameter, $b_\mathrm{so}=\hbar/3\tau_\mathrm{so}\Delta$ measures the spin-orbit scattering strength, and we have dropped a small higher-order term. The fits in Fig.~\ref{fig_local} were obtained by replacing the BCS density of states $n(E)$ by $n_+(E)+n_-(E)$ in (\ref{equ_itunnel}). During fitting, $T$ and $B$ were taken from the experiment, and the remaining parameters were varied. Including the Zeeman splitting was found to be necessary for the data at higher fields. The small spin-orbit term gave minor improvements of the fits, but could not be determined precisely. We simply chose a suitable value $b_\mathrm{so}\sim O(0.01)$ at high field, and kept it fixed for all other fits. Similar values can be found in the literature.\cite{meservey1975} The quality of the fits was excellent for samples A and B, as shown for the latter in Fig.~\ref{fig_local}. For the thin-film sample C, the quality of the fits was rather poor, and no reliable parameters could be obtained.
\begin{figure}
\includegraphics[width=\columnwidth]{fig3.eps}
\caption{\label{fig_gamma}(Color online)
(a) Pair potential $\Delta$ as a function of the applied magnetic field $B$. $\Delta$ is normalized to its zero-field value $\Delta_0$, and $B$ is normalized to the critical field $B_\mathrm{c}$.
(b) Normalized pair-breaking parameter $\Gamma/\Delta_0$ as a function of the $(B/B_\mathrm{c})^2$.
The lines are predictions from pair-breaking theory.}
\end{figure}
For samples A and B, where the fits were reliable, we show the pair-breaking parameter $\Gamma$ as well as the pair potential $\Delta$ as a function of the magnetic field in Fig.~\ref{fig_gamma}. In panel (a), the normalized pair potential $\Delta/\Delta_0$ is displayed as a function of $B/B_\mathrm{c}$, together with the expectation $\ln(\Delta/\Delta_0)=-(\pi/4)(\Gamma/\Delta)$.\cite{abrikosov1961} The critical pair-breaking strength for the suppression of superconductivity is given by $2\Gamma=\Delta_0$,\cite{abrikosov1961} and together with $\Gamma\propto B^2$ for a thin film in parallel magnetic field,\cite{maki1964a} we can rewrite $\Gamma$ as
\begin{equation}
\frac{\Gamma}{\Delta_0}=\frac{1}{2}\left(\frac{B}{B_\mathrm{c}}\right)^2.\label{equ_gammanrm}
\end{equation}
Fig.~\ref{fig_gamma}(b) shows $\Gamma/\Delta_0$ as a function of $(B/B_\mathrm{c})^2$. The solid line is the theoretical expectation (\ref{equ_gammanrm}). As can be seen, samples A and B can be described perfectly well by standard pair-breaking theory, and we will assume (\ref{equ_gammanrm}) to hold also for sample C later on.
\subsection{Energy-mode non-equilibrium}
\begin{figure}
\includegraphics[width=\columnwidth]{fig4.eps}
\caption{\label{fig_inj}(Color online)
a) Differential conductance $g_\mathrm{det}$ of the left-most contact of sample A as a function of bias voltage $V_\mathrm{det}$, with additional voltage bias $V_\mathrm{inj}$ applied to the neighbour contact at $1~\mathrm{\mu m}$ distance. b) Normalized pair potential $\Delta/\Delta_0$ as a function of injector bias $V_\mathrm{inj}$. The line is a guide to the eye.}
\end{figure}
While we are mostly interested in charge imbalance, we have also investigated the impact of energy-mode non-equilibrium in our samples. To this effect, we monitor the differential conductance $g_\mathrm{det}$ of a detector junction while non-equilibrium quasiparticles are injected through a second nearby injector contact. Fig.~\ref{fig_inj}(a) shows the differential conductance $g_\mathrm{det}$ of the left-most contact of sample A (see Fig.~\ref{fig_sem}) as a function of the local bias voltage $V_\mathrm{det}$ for different injector bias $V_\mathrm{inj}$ applied to the neighbouring contact at a distance of $1~\mathrm{\mu m}$. We focus here only on the bias region of the gap features. As the injector bias is increased, the density-of-states peak in the conductance shifts to lower bias, and broadens slightly. The increased broadening of the gap features might signify an increased temperature of the normal-metal side of the junction due to the injection of non-equlibrium quasiparticles tunneling out of the superconductor.\cite{takane2007} An alternative explanation would be an increased lifetime broadening of the density of states of the superconductor due to scattering of non-equilibrium quasiparticles. From our data, we can not make a clear decision between these scenarios. The evolution of the pair potential $\Delta$ as a function of injector bias is plotted in panel (b), normalized to its value $\Delta_0$ at $V_\mathrm{inj}=0$. No significant change is observed for $eV_\mathrm{inj}<\Delta_0$. As soon as $eV_\mathrm{inj}$ exceeds $\Delta_0\approx 200~\mathrm{\mu eV}$, the energy gap drops quickly by a few percent, and continues to decrease more slowly for higher bias. The gap reduction can be understood from the inspection of the self-consistency equation (\ref{equ_sc}). For injector voltages in the vicinity of $\Delta$, a large number of quasiparticles are injected due to the divergence of the density of states. In addition, these quasiparticles are very efficient in reducing the gap due to the energy denominator in the integral. Therefore, the initial decrease is steep, and then becomes more shallow as the density of states flattens, and the denominator increases.
\subsection{Charge-mode non-equilibrium}
\begin{figure}
\includegraphics[width=\columnwidth]{fig5.eps}
\caption{\label{fig_nonlocal}(Color online)
Non-local differential conductance $g_\mathrm{nl}=dI_\mathrm{det}/dV_\mathrm{inj}$ for an injector/detector pair of sample A as a function of bias voltage $V_\mathrm{inj}$ (a) for different temperatures $T$ and (b) for different applied magnetic fields $B$.}
\end{figure}
Figure \ref{fig_nonlocal} displays the non-local conductance $g_\mathrm{nl}$ as a function of the injector bias $V_\mathrm{inj}$ for one injector/detector-pair of sample A. Panel (a) shows data for different temperatures $T$ without an applied magnetic field. At $T=50~\mathrm{mK}$, the non-local conductance is zero within the experimental resolution for bias voltages below $\Delta/e \approx 200~\mathrm{\mu V}$. Above the energy gap, the signal increases continuously from zero with a finite initial slope up to a broad maximum at $V_\mathrm{inj} \approx 450~\mathrm{\mu V}$ before it decreases slowly again. With increasing temperature, the signal smears out around the gap, whereas the value at high bias remains unchanged. Panel (b) shows the impact of a magnetic field $B$ applied in the substrate plane along the direction of the copper wires. In contrast to temperature, the signal depends strongly on the magnetic field. The initial slope decreases, and the maximum decreases and shifts to higher bias until it is no longer observable within our bias range for $B\geq 100~\mathrm{mT}$.
\begin{figure}
\includegraphics[width=\columnwidth]{fig6.eps}
\caption{\label{fig_dist}(Color online) (a) Normalized nonlocal differential conductance $g_\mathrm{nl}/G_\mathrm{inj}G_\mathrm{det}$ as a function of injector bias voltage $V_\mathrm{inj}$ for different contact distances $d$. (b) Semi-logarithmic plot of $g_\mathrm{nl}/G_\mathrm{inj}G_\mathrm{det}$ as a function of contact distance $d$ for different injector bias $V_\mathrm{inj}$. The solid lines are fits to (\ref{equ_gnl}).
}
\end{figure}
Figure \ref{fig_dist}(a) shows the non-local differential conductance for several injector/detector contact pairs of sample A at $T=50~\mathrm{mK}$ and $B=0$. Since $g_\mathrm{nl}\propto G_\mathrm{inj}G_\mathrm{det}$, we have normalized the data accordingly to exclude the impact of small variations of the junction conductances. The overall signal magnitude decreases with increasing contact distance while the shape remains unchanged. Panel (b) shows the normalized non-local conductance as a function of contact distance $d$ for different injector bias on a semi-logarithmic scale. The solid lines are fits to the exponential decay predicted by (\ref{equ_gnl}). The quality of the fits is generally good, except for the very small signals at lowest bias. From the fits, the relaxation length $\lambda_{Q^*}$ and the amplitude
\begin{equation}\label{equ_amplitude}
a=g^*\frac{\rho_\mathrm{N}\lambda_{Q^*}}{2A}
\end{equation}
can be extracted.
\begin{figure}
\includegraphics[width=\columnwidth]{fig7.eps}
\caption{\label{fig_lambda}(Color online)
Charge imbalance relaxation length $\lambda_{Q^*}$ as a function of injector bias voltage $V_\mathrm{inj}$ for (a) different temperatures $T$ and (b) different applied magnetic fields $B$. The line is the result of a numerical simulation described in section \ref{sec_discussion}.}
\end{figure}
The charge imbalance relaxation length $\lambda_{Q^*}$ extracted from these fits is shown in Fig.~\ref{fig_lambda} as a function of injector bias for different temperatures $T$ (a) and magnetic fields $B$ (b). The data resemble those of the non-local conductance shown in Fig.~\ref{fig_nonlocal}. This is not surprising, since the signal amplitude $a$ is itself proportional to $\lambda_{Q^*}$. A noticeable difference is that $\lambda_{Q^*}$ is nearly independent of temperature. This indicates that the temperature dependence seen in Fig.~\ref{fig_nonlocal}(a) is mostly due to thermal broadening of the distribution in the injector contact rather than a change in relaxation rates. In contrast, the suppression of the non-local conductance upon increasing the magnetic field is reflected in the pronounced field dependence of $\lambda_{Q^*}$, indicating an increase of the relaxation rate.
\begin{table
\caption{\label{tab_results}Maximum values of the relaxation length $\lambda_{Q^*}$, the relaxation time $\tau_{Q^*}$,
and the ratio $\tau_{Q^*}/\tau_\mathrm{tun}$ for all three samples.}
\begin{ruledtabular}
\begin{tabular}{lccc}
sample & $\lambda_{Q^*}$ & $\tau_{Q^*}$ & $\tau_{Q^*}/\tau_\mathrm{tun}$ \\
& ($\mu$m) & (ns) & \\ \hline
A & 5.2 & 7.8 & 0.01 \\
B & 4.3 & 5.2 & 0.01 \\
C & 3.0 & 5.9 & 0.05 \\
\end{tabular}
\end{ruledtabular}
\end{table}
Similar results to those presented in Figs.~\ref{fig_nonlocal}-\ref{fig_lambda} were obtained for all three samples. From $\lambda_{Q^*}$, we can calculate $\tau_{Q^*}=\lambda_{Q^*}^2/D_\mathrm{N}v_\mathrm{g}$. The maximum values of $\lambda_{Q^*}$ and $\tau_{Q^*}$ obtained at lowest temperature and zero magnetic field are listed in table \ref{tab_results}, along with the maximum of the ratio $\tau_{Q^*}/\tau_\mathrm{tun}$. We find $\tau_{Q^*}/\tau_\mathrm{tun}\ll 1$ for all samples, confirming that our detector junctions are non-invasive.
\begin{figure}
\includegraphics[width=\columnwidth]{fig8.eps}
\caption{\label{fig_tau}(Color online)
(a) Normalized charge-imbalance relaxation rate $\hbar/\Delta_0\tau_{Q^*}$ as a function of the normalized pair-breaking parameter $\Gamma/\Delta_0$. The line is a guide to the eye.
(b) Normalized charge-imbalance relaxation rate $\hbar/\Gamma\tau_{Q^*}$ as a function of normalized injector bias $eV_\mathrm{inj}/\Delta_0$. The line is a joint fit to the data of all three samples.}
\end{figure}
We will now focus in more detail on the suppression of charge imbalance as a function of magnetic field. Figure \ref{fig_tau}(a) shows the normalized charge-imbalance relaxation rate $\hbar/\Delta_0\tau_{Q^*}$ as a function of the normalized pair-breaking parameter $\Gamma/\Delta_0$ for fixed injector bias. Here we have made use of equation (\ref{equ_gammanrm}) to calculate $\Gamma$ from $B$ for all three samples. The data from all samples fall onto a single line, and the relaxation rate at zero field is negligible on the scale of the plot. We note that in the magnetic-field range of the plot ($B\lesssim 0.5B_\mathrm{c}$) the spectral properties of the superconductor remain almost unchanged. The reduction of $\Delta$, for example, is less than $10\%$ in this range. The relaxation rate due to elastic pair-breaking perturbations such as magnetic impurities,\cite{schmid1975,artemenko1978,lemberger1981} supercurrent\cite{lemberger1981b} and applied magnetic field\cite{takane2008} has been calculated both within the quasiparticle description used by us, and from quasi-classical Green's functions. We note that by convention, rates from the Green's function formalism differ from those of the quasiparticle description by the factor $f^*$.\cite{clarke1979} When properly adjusted, the rate is predicted to be
\begin{equation}
\frac{1}{\tau_\mathrm{Q^*}}= \alpha\frac{\Gamma}{\hbar}\frac{\Delta^2}{E\epsilon},\label{equ_taub}
\end{equation}
where $\alpha$ is a numerical prefactor of order unity which we will use as a fit parameter. From linear fits of the data in panel (a) we can extract $\hbar/\Gamma\tau_{Q^*}$. The result extracted from such fits is plotted in panel (b) as a function of normalized injector bias $eV_\mathrm{inj}/\Delta_0$ for all three samples. The solid line is a joint fit of all data to (\ref{equ_taub}) (where we have set $E=eV_\mathrm{inj}$ and $\Delta=\Delta_0$), with $\alpha=0.73$.
\begin{figure}
\includegraphics[width=\columnwidth]{fig9.eps}
\caption{\label{fig_gstar}(Color online) $g^*$ as a function of normalized injector bias $eV_\mathrm{inj}/\Delta_0$ (a) for all samples at $B=0$ and (b) for sample A at different magnetic fields $B$.}
\end{figure}
From the signal amplitude $a$ extracted from the fits in Fig.~\ref{fig_dist}, we can calculate the prefactor $g^*$ using (\ref{equ_amplitude}) together with known sample parameters and $\lambda_{Q^*}$ extracted from the same fits. The results are plotted in Fig.~\ref{fig_gstar} as a function of normalized injector bias (a) for all samples at $B=0$ and (b) for sample A at different magnetic fields. At $B=0$, $g^*$ follows the expectation $g^*\approx \Theta(eV_\mathrm{inj}-\Delta)$ for samples A and B, whereas it deviates at low bias both for sample C, and in the presence of a magnetic field. Since $g^*$ depends on a combination of several spectral properties of the superconductor, we can not identify the precise cause of these deviations. For sample C, the enhanced energy-mode non-equilibrium due to the reduced film thickness may play a role. However, in all cases $g^*\approx 1$ at high bias, which justifies our choice of using a wire section of length $2\lambda_{Q^*}$ as injection volume.
\section{Discussion}\label{sec_discussion}
The bias dependence of the non-local conductance can be understood as follows: Charge-imbalance relaxation takes place mostly at energies close to $\Delta$, since the coherence factors for scattering between the electron- and hole-like branches, both elastic and inelastic, vanish at higher energies. In addition, in the low-temperature regime the inelastic contribution is expected to be negligible.\cite{takane2006} The elastic relaxation rate, both due to gap anisotropy and magnetic pair breaking, diverges for $E\rightarrow \Delta$, and quickly drops at higher energies. Consequently, charge imbalance rises continuously from zero as the bias is increased above $\Delta$, and in this regime relaxation is mainly due to direct scattering between the branches. As the bias is increased further, the direct relaxation rate decreases. On the other hand, energy relaxation due to inelastic scattering becomes important, and charge relaxation turns into a two-stage process.\cite{tinkham1972} Quasiparticles are first cooled to the vicinity of $\Delta$ by inelastic scattering, and are then scattered elastically between branches. If we assume that inelastic scattering is described by electron-phonon scattering in the Debye approximation, the inelastic rate quickly increases as more phonons become available at higher energy. Consequently, the non-local differential conductance begins to drop again after its initial increase. The bias at which the maximum appears scales roughly with the transition between the direct and two-stage relaxation regimes. The temperature dependence, or lack thereof, can be understood easily from this picture. Elastic relaxation depends only weakly on temperature via $\Delta$. For inelastic relaxation, at low temperatures the Bose factors for emission and absorption of phonons become $\approx 1$ and $\approx 0$, respectively. Relaxation is then dominated by phonon emission at a rate which only depends on the bias-dependent energy of the quasiparticles, but not on temperature. Consequently, the relaxation rate, and $\lambda_{Q^*}$, become practically independent of temperature over the entire bias range, as observed in Fig.~\ref{fig_lambda}(a). We have identified the quasiparticle energy with the bias voltage throughout our data analysis in section \ref{sec_results}. This has to be taken with some caution due to the presence of inelastic energy relaxation. Our approximation mainly affects the calculation of $\tau_{Q^*}=\lambda_{Q^*}^2/D_\mathrm{N}v_\mathrm{g}$, where $v_\mathrm{g}$ should actually be an average over energy. However, since inelastic relaxation is important mostly at higher energies, where $v_\mathrm{g}\approx 1$, we assume that the error is small. This is corroborated by the observation that the position of the maximum in $g_\mathrm{nl}$ does not depend much on contact distance, as seen in Fig.~\ref{fig_lambda}(a). We would nevertheless like to stress that $\lambda_{Q^*}$ and $\tau_{Q^*}$ are not really spectral quantities, but depend in detail on the non-equilibrium distribution and bias conditions.
To make a quantitative connection to microscopic theory, we have performed numerical simulations of the quasiparticle Boltzmann equation, basically following Ref.~\onlinecite{chi1979}. We have used the onedimensional form of the Boltzmann equation\cite{chi1979,entin1979}
\begin{equation}
\frac{\partial f}{\partial t}-v_\mathrm{g}D_\mathrm{N}\frac{\partial^2 f}{\partial x^2}=\gamma_\mathrm{tun}-\gamma_\mathrm{el}-\gamma_\mathrm{in}.\label{equ_boltzmann}
\end{equation}
The elastic and inelastic relaxation rates $\gamma_\mathrm{el}$ and $\gamma_\mathrm{in}$ are given by equations (2.16) and (2.17) of Ref.~\onlinecite{chi1979}, respectively. The injection rate is given by (\ref{equ_gamma_inj}). Steady-state solutions $f(\epsilon,x)$ of the Boltzmann equation (\ref{equ_boltzmann}) were obtained by numerical iteration on a discretized grid. Both injector and detector junction were included on an equal footing, with injection rates and currents given by (\ref{equ_gamma_inj}) and (\ref{equ_current}). Here, the injection volume $\Omega$ is given by the grid point size, and the spreading of charge imbalance over the length scale $\lambda_{Q^*}$ is included microscopically in the diffusion term. The granularity of the grid was chosen sufficiently small ($\Delta\epsilon=20~\mathrm{\mu eV}$, $\Delta x=500~\mathrm{nm}$) not to affect the results. Parameters such as wire geometry, diffusion constant, contact conductances $G_\mathrm{inj}$ and $G_\mathrm{det}$, etc., were taken directly from the experiment, leaving only the characteristic electron-phonon scattering time $\tau_0$ and the average gap anisotropy $\left<a^2\right>_0$ as free parameters. The results of the simulation are shown as a solid line in Fig.~\ref{fig_lambda}(b), where we have inserted the typical values $\tau_0=100~\mathrm{ns}$ and $\left<a^2\right>_0=0.03$ from Ref.~\onlinecite{chi1979}. The overall magnitude and shape is predicted correctly by the simulation, with $\lambda_{Q^*}\approx 5~\mathrm{\mu m}$, and a broad maximum at $V_\mathrm{inj}\approx 400~\mathrm{\mu V}$. However, in detail the agreement is rather poor. The slope is too steep, both below and above the maximum. The relaxation rate due to impurity scattering has been calculated both within the quasiparticle approach,\cite{tinkham1972b,chi1979} and using quasiclassical Green's functions.\cite{takane2006} In contrast to the quasiparticle approach, the Green's function approach predicts different energy dependences for the clean and dirty limits. This difference might explain the discrepancy at low bias. At high bias, electron-phonon scattering in the Debye approximation\cite{kaplan1976} apparently overestimates the energy dependence of the relaxation rate. A weaker energy dependence has been predicted for electron-phonon scattering in the presence of disorder.\cite{bergmann1971,schmid1973,belitz1987} Also, it has been argued that disorder-enhanced electron-electron interaction may dominate in particular in aluminum with its weak electron-phonon interaction.\cite{stuivinga1983,clarke1986,lee1989} A simulation based on microscopic theory for both electron-electron and electron-phonon interaction might provide detailed insight into the relaxation mechanisms here, but this is beyond the scope of this article.
We now focus on the dependence on magnetic field. Magnetic pair breaking adds an elastic contribution to the relaxation rate.\cite{schmid1975,artemenko1978,lemberger1981,lemberger1981b,takane2008} Consequently, as $B$ increases, the initial slope of the differential conductance decreases, and the maximum, i.e., the transition to the two-stage relaxation regime, shifts to higher bias. We find that the magnetic contribution to the relaxation rate is directly proportional to the pair-breaking parameter $\Gamma$, and that its energy dependence follows the theoretical prediction independent of sample details. Our observation $\tau_{Q^*}^{-1}\propto \Gamma$ is markedly different from the well-established approximation $\tau_{Q^*}^{-1}\propto \sqrt{1+2\tau_\mathrm{E}\Gamma/\hbar}$ valid for $T\rightarrow T_\mathrm{c}$,\cite{schmid1975} where $\tau_\mathrm{E}$ is the energy relaxation time (note that our $\Gamma/\hbar$ is the magnetic pair-breaking rate, i.e., the quantity $\tau_\mathrm{s}^{-1}$ of Ref.~\onlinecite{schmid1975}). Magnetic relaxation dominates all other contributions even at very low magnetic fields, where the spectral properties of the superconductor are almost unaffected by pair breaking.
We finally discuss the possible impact of charge imbalance on the observation of other phenomena. We first note that the non-local conductance at subgap energies remains negligible at lowest temperatures, even with an applied magnetic field. This is not surprising, since even in the presence of magnetic pair breaking the superconductor still has a well-defined energy gap for quasiparticle excitations, at least as long as pair breaking is not too strong. Also, the relaxation length is always larger than the coherence length ($\xi\approx 100~\mathrm{nm}$ for our samples). Thus, the dependence of non-local conductance on bias or contact distance remains a good criterion to distinguish coherent subgap transport from charge imbalance. On the other hand, the observation of spin-dependent quasiparticle transport necessarily involves injection at energies above the gap, and ferromagnetic electrodes must be used. At magnetic fields of $\sim 100~\mathrm{mT}$, which are easily reached by the fringing fields of electrodes made of elementary ferromagnets, $\lambda_{Q^*}$ can already be as small as $1~\mathrm{\mu m}$. This is similar to the spin-diffusion length $\lambda_\mathrm{sf}$ in aluminium,\cite{jedema2002,jedema2003} and consequently great care must be taken to distinguish charge imbalance and spin-dependent transport.
\section{Conclusion}
In conclusion, we have presented a detailed investigation of charge imbalance in superconductors in the low-temperature regime. From our measurements, we have extracted the charge-imbalance relaxation length as a function of bias voltage, temperature, and magnetic field. The bias-dependent results allow for a detailed comparison with different relaxation mechanisms. In particular, we have shown a transition from dominant elastic relaxation in the vicinity of the energy gap to an inelastic two-stage relaxation at high bias. The dependence on magnetic field follows theory with remarkable accuracy, and is clearly different from the known approximations for the high-temperature regime. The strong reduction of the relaxation length with magnetic field has possible implications for the interpretation of spin-diffusion experiments.
During the preparation of this manuscript we became aware of three related studies of charge imbalance at low temperatures.\cite{tsuboi2010,kleine2010,arutyunov2010} We thank K.~Yu.~Arutyunov, W.~Belzig and D.~S.~Golubev for useful discussions. This work was supported by the DFG Center for Functional Nanostructures, and by the DAAD within the RISE program.
|
\section{Introduction}
One commonly studied invariant of a finite, $(d-1)$-dimensional simplicial complex $\Delta$ is its $f$-vector $f(\Delta) = (f_{-1},f_0,\ldots,f_{d-1})$, where $f_i$ denotes the number of $i$-dimensional faces in $\Delta$. It is equivalent, and oftentimes more convenient, to study the $h$-vector $h(\Delta) = (h_0, \ldots, h_d)$ defined by the relation $\sum_{j=0}^dh_j\lambda^{d-j} = \sum_{i=0}^df_{i-1}(\lambda-1)^{d-i}$.
When studying the $h$-numbers of simplicial complexes, it is natural to study the class of simplicial complexes that are Cohen-Macaulay over a fixed field $\mathbf{k}$. This includes the classes of $\mathbf{k}$-homology balls and $\mathbf{k}$-homology spheres. A more specialized class of simplicial complexes is the class of complexes that are doubly Cohen-Macaulay (2-CM) over $\mathbf{k}$. This class was introduced and studied by Baclawski \cite{Bac82}. Any complex that is 2-CM over $\mathbf{k}$ is Cohen-Macaulay over $\mathbf{k}$ with non-vanishing top dimensional homology (computed with coefficients in $\mathbf{k}$). For example, $\mathbf{k}$-homology spheres are 2-CM over $\mathbf{k}$, but $\mathbf{k}$-homology balls are not 2-CM over $\mathbf{k}$.
The advantage of studying topological manifolds is that they are locally homeomorphic to topological balls. Using this local property as motivation, it is natural to define the class of Buchsbaum simplicial complexes. We say that a simplicial complex $\Delta$ is Buchsbuam over $\mathbf{k}$ if the link of each vertex of $\Delta$ is Cohen-Macaulay over $\mathbf{k}$. Hence $\mathbf{k}$-homology manifolds are Buchsbaum simplicial complexes over $\mathbf{k}$. If $\Delta$ is a Buchsbaum simplicial complex, it is convenient to study the $h^{\prime}$-numbers of $\Delta$, denoted by $h_j^{\prime}(\Delta)$, (defined in Section \ref{notation}) which encode both the $h$-numbers of $\Delta$ and the underlying geometric structure of $\Delta$.
Athanasiadis and Welker \cite{AthW} define the class of Buchsbaum* complexes that specialize Buchsbaum complexes in the same way that 2-CM complexes specialize Cohen-Macaulay complexes. They show that if $\Delta$ is Buchsbaum* over $\mathbf{k}$, then the link of each vertex of $\Delta$ is 2-CM over $\mathbf{k}$ and that a homology manifold is Buchsbaum* over $\mathbf{k}$ if and only if it is orientable over $\mathbf{k}$.
Stanley \cite{S79} introduces the class of \textit{balanced} simplicial complexes and shows that the rank selected subcomplexes of a balanced Cohen-Macaulay complex are Cohen-Macaulay. This result easily generalizes to the classes of 2-CM and Buchsbaum simplicial complexes. Our first goal in this paper is to show (Theorem \ref{rank-selected}) that the rank-selected subcomplexes of a Buchsbaum* simplicial complex are Buchsbaum*. This result answers a question posed in \cite{AthW} in the affirmative.
Barnette's Lower Bound Theorem \cite{B73} says that if $\Delta$ is a $(d-1)$-dimensional homology manifold without boundary and $d \geq 3$, then $h_2(\Delta) \geq h_1(\Delta)$. This gives a sharp lower bound on all the $f$-numbers of $\Delta$ in terms of $d$ and $n$, the number of vertices in $\Delta$. Equality in this lower bound is achieved by a \textit{stacked} simplicial $(d-1)$-sphere on $n$ vertices.
Kalai \cite{k87} used the theory of rigidity frameworks to prove that $h_2 \geq h_1$ for simplicial homology manifolds without boundary and $d \geq 3$. Nevo \cite{Ne07} extended this result to the class of 2-CM complexes, and Athanasiadis and Welker \cite{AthW} further extend this result to the class of connected Buchsbaum* complexes. Using these results, together with Theorem \ref{rank-selected}, we extend a result of Goff, Klee, and Novik \cite{GKN}, showing that $2h_2 \geq (d-1)h_1$ for all balanced, connected Buchsbaum* simplicial complexes with $d \geq 3$. As in the case of Barnette's LBT, this bound is sharp, with equality achieved by a so-called \textit{stacked cross-polytopal sphere}.
Athanasiadis and Welker \cite{AthW} show that $h^{\prime}_j(\Delta) \geq {d \choose j}$ when $\Delta$ is a flag, Buchsbaum* simplicial complex of dimension $d-1$. Equality is achieved in this bound by $\mathcal{P}^{\times}_d$, the boundary complex of a $d$-dimensional cross polytope. We generalize this result, showing that $h^{\prime}_j(\Delta) \geq {d \choose j}m^j$ when $\Delta$ is a flag, $m$-Buchsbaum* simplicial complex of dimension $d-1$ in Theorem \ref{m-buchs}. Moreover, we show that equality holds here only for the $d$-fold join of $m+1$ disjoint vertices, answering a question posed in \cite{AthW} in the case that $m=1$.
The paper is structured as follows. In Section 2, we review the definitions and background information that will be necessary for the remainder of the paper. In Section 3, we study balanced Buchsbaum* complexes, proving that the rank selected subcomplexes of Buchsbaum* complexes are Buchsbaum* (Theorem \ref{rank-selected}). In Section 4, we prove lower bounds on balanced Buchsbaum* complexes (Theorem \ref{BBS-LBT}) and flag Buchsbaum* complexes (Theorem \ref{m-buchs}).
\section{Notation and Definitions} \label{notation}
We begin by reviewing some basic definitions on simplicial complexes. A \textit{simplicial complex} $\Delta$ on vertex set $V=V(\Delta)$ is a collection of subsets $\tau \subseteq V,$ called \textit{faces}, that is closed under inclusion. We say that a simplicial complex $\Delta$ is \textit{pure} if all of its \textit{facets} (maximal faces under inclusion) have the same dimension.
The \textit{dimension} of a face $\tau \in \Delta$ is $\dim \tau = |\tau|-1$, and the dimension of $\Delta$ is $\dim \Delta = \max\{\dim \tau: \tau \in \Delta\}$. The $f$-vector of $\Delta$ is the vector $f(\Delta) = (f_{-1},f_0,\ldots,f_{d-1})$ where $\dim \Delta = d-1$ and the \textit{$f$-numbers} $f_i = f_i(\Delta)$ denote the number of $i$-dimensional faces of $\Delta$. It is oftentimes more convenient to study the $h$-numbers $h_j(\Delta)$ defined by the relation $\sum_{j=0}^dh_j\lambda^{d-j} = \sum_{i=0}^df_{i-1}(\lambda-1)^{d-i}$. It is easy to see that the $h$-numbers of $\Delta$ can be written as integer linear combinations of its $f$-numbers and that the $f$-numbers of $\Delta$ can be written as \textit{nonnegative} integer linear combinations of its $h$-numbers. In particular, bounds on the $h$-numbers of $\Delta$ immediately yield bounds on the $f$-numbers of $\Delta$.
If $\Delta$ is a simplicial complex and $\tau$ is a face of $\Delta$, the \textit{contrastar} of $\tau$ in $\Delta$ is $\cost_{\Delta}(\tau):= \{\sigma \in \Delta: \sigma \nsupseteq \tau\}$, and the \textit{link} of $\tau$ in $\Delta$ is $\mbox{\upshape lk}_{\Delta}(\tau):= \{\sigma \in \Delta: \sigma \cap \tau = \emptyset, \sigma \cup \tau \in \Delta\}$. If $A \subset V(\Delta)$ is a collection of vertices in $\Delta$, then $\Delta-A$ is the \textit{restriction} of $\Delta$ to $V(\Delta) \setminus A$. When $A$ consists of a single vertex $v \in V(\Delta)$, we simply write $\Delta-v$ instead of $\Delta - \{v\}$. If $\Gamma$ and $\Delta$ are simplicial complexes on disjoint vertex sets, their simplicial join is the $(\dim \Gamma + \dim \Delta +1)$-dimensional simplicial complex $$\Gamma * \Delta:= \{\sigma \cup \tau: \sigma \in \Gamma, \tau \in \Delta\}.$$
We are particularly interested in studying the class of balanced simplicial complexes, introduced by Stanley \cite{S79}.
\begin{definition}
A $(d-1)$-dimensional simplicial complex $\Delta$ is \textbf{balanced} if there is a coloring $\kappa: V(\Delta) \rightarrow [d]$ with the property that $\kappa(u) \neq \kappa(v)$ for all edges $\{u,v\} \in \Delta$. We assume that a balanced complex comes equipped with such a coloring $\kappa$.
\end{definition}
The order complex of a graded poset of rank $d$ is one example of a balanced simplicial complex. If $\Delta$ is a balanced simplicial complex and $S \subseteq [d]$, it is often important to study the \textit{$S$-rank selected subcomplex} of $\Delta$, which is defined as the collection of faces in $\Delta$ whose vertices are colored by $S$. Specifically,
\begin{displaymath}
\Delta_S = \{\tau \in \Delta: \kappa(\tau) \subseteq S\}.
\end{displaymath}
In \cite{S79} Stanley showed that
\begin{equation}\label{hnums}
h_i(\Delta) = \sum_{|S|=i}h_i(\Delta_S);
\end{equation}
and that if $\Delta$ is Cohen-Macaulay, then so are its rank-selected subcomplexes.
A more specialized class of Cohen-Macaulay complexes is the class of \textit{doubly Cohen-Macaulay} (2-CM) complexes. A $(d-1)$-dimensional simplicial complex $\Delta$ is 2-CM (over $\mathbf{k}$) if it is Cohen-Macaulay and $\Delta-v$ is Cohen-Macaulay of dimension $d-1$ for all vertices $v \in \Delta$. Walker \cite{W81} showed that double Cohen-Macaulayness is a topological property, meaning that it only depends on the homeomorphism type of the \textit{geometric realization} $|\Delta|$ of $\Delta$. In particular, if a $(d-1)$-dimensional simplicial complex $\Delta$ is 2-CM, then $\cost_{\Delta}\tau$ is Cohen-Macaulay of dimension $d-1$ for all nonempty faces $\tau \in \Delta$.
Athanasiadis and Welker \cite{AthW} define Buchsbaum* complexes as specializations of Buchsbaum complexes, much in the same sense that 2-CM complexes are specializations of Cohen-Macaulay complexes. For the purposes of this paper, we will use the following definition of a Buchsbaum* complex.
\begin{definition}
A $(d-1)$-dimensional simplicial complex $\Delta$ that is Buchsbaum over $\mathbf{k}$ is {\bf Buchsbaum*} over $\mathbf{k}$ if, for all $p \in |\Delta|$, the canonical map $\rho_*: \widetilde{H}_{d-1}(|\Delta|;\mathbf{k}) \rightarrow \widetilde{H}_{d-1}(|\Delta|,|\Delta|-p;\mathbf{k})$ is surjective.
\end{definition}
Henceforth we will fix a field $\mathbf{k}$. When we say that a simplicial complex $\Delta$ is Buchsbaum* without qualification, we implicitly mean that $\Delta$ is Buchsbaum* over $\mathbf{k}$. Moreover, we will implicitly compute all homology groups with coefficients in $\mathbf{k}$, and we will suppress this from our notation for convenience.
First we will give an equivalent definition of the Buchsbaum* property in a combinatorial language.
\begin{lemma}
Let $\Delta$ be a $(d-1)$-dimensional Buchsbaum simplicial complex. The following are equivalent.
\begin{enumerate}
\item[\rm{(a)}] $\Delta$ is Buchsbaum*;
\item[\rm{(b)}] For all faces $\sigma \subseteq \tau$ of $\Delta$, the map $$j_*: \widetilde{H}_{d-1}(\Delta,\cost_{\Delta}(\sigma)) \rightarrow \widetilde{H}_{d-1}(\Delta,\cost_{\Delta}(\tau)),$$ induced by inclusion, is surjective;
\item[\rm{(c)}] For all faces $\tau \in \Delta$, the map $$\rho_*:\widetilde{H}_{d-1}(\Delta) \rightarrow \widetilde{H}_{d-1}(\Delta,\cost_{\Delta}(\tau)),$$ induced by inclusion, is surjective.
\end{enumerate}
\end{lemma}
\begin{proof}
The equivalence of (a) and (b) is Proposition 2.8 in \cite{AthW}. Taking $\sigma = \emptyset$ shows that (b) implies (c). Next, consider any point $p \in |\Delta|$, and let $\tau$ be the unique minimal face of $|\Delta|$ whose relative interior contains $p$. Then $|\Delta|-p$ retracts onto $\cost_{\Delta}(\tau)$ and (c) implies (a).
\end{proof}
The $h^{\prime}$-vector of a Buchsbaum complex $\Delta$ encodes both the underlying geometry of $\Delta$ and the combinatorial data of $\Delta$. Let $\widetilde{\beta}_i(\Delta):= \dim_{\mathbf{k}}\widetilde{H}_i(\Delta;\mathbf{k})$ denote the (reduced) \textit{$\mathbf{k}$-Betti numbers} of $\Delta$. The $h^{\prime}$-numbers of $\Delta$ are defined by
\begin{displaymath}
h_j^{\prime}(\Delta) := h_j(\Delta) + {d \choose j} \sum_{i=0}^{j-1}(-1)^{j-i-1}\widetilde{\beta}_{i-1}(\Delta).
\end{displaymath}
We encode the $h^{\prime}$-numbers of $\Delta$ into the $h^{\prime}$-polynomial $h^{\prime}_{\Delta}(t):= \sum_{j=0}^dh^{\prime}_{j}(\Delta)t^j.$
Athanasiadis and Welker \cite{AthW} prove a number of very nice properties about Buchsbaum* complexes, which we summarize here.
\begin{theorem}
Let $\Delta$ be a $(d-1)$-dimensional Buchsbaum* complex with $d \geq 2$. Then
\begin{enumerate}
\item $\widetilde{H}_{d-1}(\Delta) \neq 0$;
\item $\mbox{\upshape lk}_{\Delta}v$ is 2-CM for all vertices $v \in \Delta$; and
\item $\Delta$ is doubly-Buchsbaum, meaning that $\Delta-v$ is a Buchsbaum complex of dimension $d-1$, for all vertices $v \in \Delta$.
\end{enumerate}
\end{theorem}
\section{Balanced Buchsbaum* Complexes}
It is well known (see, for example, \cite{S79}) that if $\Delta$ is a balanced, Cohen-Macaulay complex, then $\Delta_S$ is Cohen-Macaulay for any $S \subseteq [d]$. It is easy to see that this result generalizes to the classes of $2$-CM complexes and Buchsbaum complexes. The purpose of this section is to generalize this result to the class of Buchsbaum* complexes.
\begin{theorem} \label{rank-selected}
Let $\Delta$ be a $(d-1)$-dimensional balanced, doubly-Buchsbaum complex. For any $S \subset [d]$, the rank selected subcomplex $\Delta_S$ is Buchsbaum*.
\end{theorem}
In particular, since a Buchsbaum* complex is doubly-Buchsbaum, Theorem \ref{rank-selected} implies that the rank selected subcomplexes of a Buchsbaum* complex are Buchsbaum*.
We begin with a series of lemmas, the first of which is well-known (see, for example, \cite{Munkres84}).
\begin{lemma} \label{lem2}
Let $\Gamma$ be a simplicial complex, and let $\tau$ be a nonempty face in $\Gamma$. Then $\widetilde{H}_i(\Gamma, \cost_{\Gamma}(\tau)) \cong \widetilde{H}_{i-|\tau|}(\mbox{\upshape lk}_{\Gamma}\tau)$.
\end{lemma}
\begin{lemma} \label{lem3}
Let $\Delta$ be a balanced, doubly Buchsbaum simplicial complex of dimension $d-1$, and let $c \in [d]$. Choose vertices $v_1, \ldots, v_i$ of color $c$, and let $\Delta_{i-1} = \Delta \setminus\{v_1, \ldots, v_{i-1}\}$ and $\Delta_i = \Delta \setminus\{v_1, \ldots, v_i\}$. Then for $S = [d]-c$ and any nonempty face $\tau \in \Delta_S$,
\begin{displaymath}
\widetilde{H}_{d-2}(\cost_{\Delta_{i-1}}(\tau), \cost_{\Delta_i}(\tau)) = 0.
\end{displaymath}
\end{lemma}
\begin{proof}
By Lemma \ref{lem2},
\begin{eqnarray*}
\widetilde{H}_{d-2}(\cost_{\Delta_{i-1}}(\tau),\cost_{\Delta_i}(\tau)) \cong \widetilde{H}_{d-3}(\mbox{\upshape lk}_{\cost_{\Delta_{i-1}}(\tau)}v_i).
\end{eqnarray*}
Since $\Delta$ is balanced, $\mbox{\upshape lk}_{\cost_{\Delta_{i-1}}(\tau)}v_i = \mbox{\upshape lk}_{\cost_{\Delta}(\tau)}v_i$, and
\begin{displaymath}
\mbox{\upshape lk}_{\cost_{\Delta}(\tau)}v_i = \{\sigma \in \Delta: \sigma \nsupseteq \tau, v_i \notin \sigma, \sigma \cup v_i \in \Delta\} = \cost_{\mbox{\upshape lk} v_i}(\tau).
\end{displaymath}
\noindent Since $\mbox{\upshape lk}_{\Delta}v_i$ is 2-CM of dimension $d-2$, it follows that $\widetilde{H}_{d-3}(\cost_{\mbox{\upshape lk} v_i}(\tau)) = 0$.
\end{proof}
Now we proceed with the proof of Theorem \ref{rank-selected}.
\begin{proof}(Theorem \ref{rank-selected}) \\
Fix a coloring $\kappa: V(\Delta) \rightarrow [d]$. We need only consider those $S \subset [d]$ with $|S| = d-1$. Inductively, this is sufficient as any Buchsbaum* complex is doubly Buchsbaum. Suppose $S = [d]-\{c\}$ and consider the vertices $\{v_1, \ldots, v_k\} \in \Delta$ with $\kappa(v_i)=c$. For $1 \leq i \leq k$, let $\Delta_i: = \Delta - \{v_1, \ldots, v_{i}\}$, and when $i=0$, set $\Delta_0:= \Delta$.
Let $\tau$ be a nonempty face in $\Delta_S \subset \Delta_{k-1} \subset \cdots \subset \Delta_1 \subset \Delta$. We claim that for any $0 \leq i \leq k$, the canonical map $$\rho^i_*:\widetilde{H}_{d-2}(\Delta_i) \rightarrow \widetilde{H}_{d-2}(\Delta_i, \cost_{\Delta_i}(\tau))$$ is a surjection. We proceed by induction on $i$.
When $i=0$, $\Delta_0=\Delta$ is Buchsbaum so $\widetilde{H}_{d-2}(\Delta,\cost_{\Delta}(\tau))=0$, and the map $\rho^0_*$ is surely surjective.
Suppose now that $i>0$ and consider the long exact sequence for the pair $(\cost_{\Delta_{i-1}}(\tau),\cost_{\Delta_{i}}(\tau))$:
\begin{displaymath}
\rightarrow \widetilde{H}_{d-2}(\cost_{\Delta_{i-1}}(\tau),\cost_{\Delta_{i}}(\tau)) \rightarrow \widetilde{H}_{d-3}(\cost_{\Delta_{i}}(\tau)) \stackrel{\iota_*}{\rightarrow} \widetilde{H}_{d-3}(\cost_{\Delta_{i-1}}(\tau)) \rightarrow.
\end{displaymath}
\noindent By Lemma \ref{lem3}, $H_{d-2}(\cost_{\Delta_{i-1}}(\tau),\cost_{\Delta_{i}}(\tau)) \cong 0$, and hence the map $\iota_*$ is an injection.
Next, we consider the inclusion map $(\Delta_i,\cost_{\Delta_{i}}(\tau)) \hookrightarrow (\Delta_{i-1},\cost_{\Delta_{i-1}}(\tau))$, which induces the following commutative diagram of long exact sequences.
\[
\begin{diagram}
\node{\widetilde{H}_{d-2}(\Delta_i)} \arrow{r,t}{\rho^i_*} \arrow{s} \node{\widetilde{H}_{d-2}(\Delta_i,\cost_{\Delta_{i}}(\tau))} \arrow{s} \arrow{r,t}{\partial_i} \node{\widetilde{H}_{d-3}(\cost_{\Delta_{i}}(\tau))} \arrow{s,r}{\iota_*} \\
\node{\widetilde{H}_{d-2}(\Delta_{i-1})} \arrow{r,t}{\rho^{i-1}_*} \node{\widetilde{H}_{d-2}(\Delta_{i-1},\cost_{\Delta_{i-1}}(\tau))} \arrow{r,t}{\partial_{i-1}} \node{\widetilde{H}_{d-3}(\cost_{\Delta_{i-1}}(\tau))}
\end{diagram}
\]
By the inductive hypothesis, the map $\rho^{i-1}_*$ is surjective and so the map $\partial_{i-1}$ is the zero map. By commutativity of the above diagram, $\iota_* \circ \partial_i$ is the zero map, and since $\iota_*$ is an injection, the map $\partial_i$ is also the zero map. Thus by exactness, $\rho^i_*: \widetilde{H}_{d-2}(\Delta_i) \rightarrow \widetilde{H}_{d-2}(\Delta_i,\cost_{\Delta_{i}}(\tau))$ is a surjection.
This establishes the claim. In particular, when $i=k$, we have shown that the canonical map $\rho_*: \widetilde{H}_{d-2}(\Delta_S) \rightarrow \widetilde{H}_{d-2}(\Delta_S,\cost_{\Delta_S}(\tau))$ is surjective, and hence $\Delta_S$ is Buchsbaum*.
\end{proof}
\begin{definition} {\rm{(\cite{AthW}, Definition 5.5)}}
Let $\Delta$ be a $(d-1)$-dimensional simplicial complex, and let $m$ be a nonnegative integer. We say that $\Delta$ is $m$-Buchsbaum* if $\Delta$ is Buchsbaum and $\Delta-A$ is Buchsbaum* of dimension $d-1$ for any subset $A \subset V(\Delta)$ with $|A| < m$.
\end{definition}
\begin{corollary}
Let $\Delta$ be a $(d-1)$-dimensional balanced simplicial complex that is $m$-Buchsbaum* over $\mathbf{k}$. For any $S \subseteq [d]$, the rank selected subcomplex $\Delta_S$ is $m$-Buchsbaum* over $\mathbf{k}$.
\end{corollary}
\begin{proof}
By Lemma 5.6 in \cite{AthW}, $\Delta$ is $(m+1)$-Buchsbaum over $\mathbf{k}$. For any subset $A \subseteq V(\Delta_S)$ with $|A| < m$, the complex $\Delta-A$ is doubly Buchsbaum. Thus $(\Delta-A)_S = \Delta_S-A$ is Buchsbaum* by Theorem \ref{rank-selected}.
\end{proof}
\section{Lower Bounds}
Fix integers $n$ and $d$ such that $d$ divides $n$. Let $\mathcal{P}^{\times}_d$ denote the boundary complex of the $d$-dimensional cross polytope. Following \cite{GKN}, define a \textit{stacked cross-polytopal sphere} $\mathcal{ST}^{\times}(n,d-1)$ by taking the connected sum of $\frac{n}{d}-1$ copies of $\mathcal{P}^{\times}_d$. In each connected sum, we identify vertices of the same colors so that $\mathcal{ST}^{\times}(n,d-1)$ is a balanced $(d-1)$-sphere on $n$ vertices.
Athanasiadis and Welker (\cite{AthW}, Theorem 4.1) prove that if $\Delta$ is a connected, $(d-1)$-dimensional Buchsbaum* complex with $d \geq 3$, then the graph of $\Delta$ is generically $d$-rigid. This generalizes Nevo's result that $h_2(\Delta) \geq h_1(\Delta)$ when $\Delta$ is 2-CM (\cite{Ne07}, Theorem 1.3). Using Theorem 4.1 from \cite{AthW} in place of Nevo's result and the conclusion of Theorem \ref{rank-selected}, the techniques used to prove Theorem 5.3 in \cite{GKN} give the following Lower Bound Theorem for balanced Buchsbaum* complexes.
\begin{theorem} \label{BBS-LBT}
Let $\Delta$ be a connected, balanced, Buchsbaum* complex of dimension $d-1$ with $d \geq 3$. Then $d\cdot h_2(\Delta) \geq {d \choose 2}h_1(\Delta)$. In particular, if $d$ divides $n = f_0(\Delta)$, then $f_j(\Delta) \geq f_j(\mathcal{ST}^{\times}(n,d-1)$ for all $j$.
\end{theorem}
Hersh and Novik \cite{HN} define the \textit{short simplicial $h$-numbers} of a simplicial complex $\Delta$ as $\widetilde{h}_j(\Delta):= \sum_{v \in \Delta}h_j(\mbox{\upshape lk}_{\Delta}v)$ for $0 \leq j \leq d-1$. Swartz \cite{Sw04} proves that the short simplicial $h$-numbers satisfy
\begin{equation} \label{eq2}
\widetilde{h}_{j-1}(\Delta) = j\cdot h_j(\Delta) + (d-j+1)h_{j-1}(\Delta).
\end{equation}
We use this formula, together with Theorem \ref{BBS-LBT} to prove the following theorem.
\begin{theorem} \label{h3-h1}
Let $\Delta$ be a balanced Buchsbaum* complex of dimension $d-1$ with $d \geq 4$. Then $d\cdot h_3(\Delta) \geq {d \choose 3}h_1(\Delta)$.
\end{theorem}
\begin{proof}
The link of each vertex $v \in \Delta$ is 2-CM, and hence by Theorem 5.3 in \cite{GKN}, $2h_2(\mbox{\upshape lk}_{\Delta}v) \geq (d-2)h_1(\mbox{\upshape lk}_{\Delta}v)$ for all $v \in \Delta$. Thus
\begin{eqnarray*}
2\cdot(3h_3(\Delta)+(d-2)h_2(\Delta)) &=& 2 \cdot \widetilde{h}_2(\Delta) \geq (d-2)\widetilde{h_1}(\Delta)\\
&=& (d-2)(2h_2(\Delta)+(d-1)h_1(\Delta)),
\end{eqnarray*}
and the desired result follows.
\end{proof}
Bj\"orner and Swartz \cite{Sw06} have conjectured that the inequality $h_{d-1} \geq h_1$ holds for all 2-CM complexes with $d \geq 3$. When $d=4$, the conclusion of Theorem \ref{h3-h1} continues to hold without the assumption that $\Delta$ is balanced, establishing that $h_3 \geq h_1$ for all $3$-dimensional Buchsbaum* (and hence 2-CM) complexes.
Athanasiadis and Welker \cite{AthW} prove that if a $(d-1)$-dimensional Buchsbaum* simplicial complex $\Delta$ is flag, then $h^{\prime}_{\Delta}(t) \geq (1+t)^d$, where the inequality is interpreted coefficient-wise. Motivated by a question in \cite{Ath}, they pose the following question.
\begin{question} \label{lb}
{\rm{(\cite{AthW}, Question 6.5(i)) Let $\Delta$ be a $(d-1)$-dimensional flag, Buchsbaum* simplicial complex. If $h^{\prime}_j = {d \choose j}$ for some $1 \leq j \leq d-1$, is $\Delta$ necessarily isomorphic to $\mathcal{P}^{\times}_d$?}}
\end{question}
We will generalize the result of Athanasiadis and Welker to the class of $m$-Buchsbaum* simplicial complexes and answer Question \ref{lb} for this class. For fixed positive integers $m$ and $d$, we define the simplicial complex $\mathcal{P}(m+1,d)$ to be the $d$-fold join of $m+1$ disjoint vertices. We remark first that $\mathcal{P}(m+1,d)$ is $(m+1)$-CM and hence $m$-Buchsbaum* by Proposition 5.6 in \cite{AthW}, and second that $\mathcal{P}(2,d) = \mathcal{P}^{\times}_d$.
\begin{theorem} \label{m-buchs}
Let $\Delta$ be a $(d-1)$-dimensional flag simplicial complex that is $m$-Buchsbaum* over the field $\mathbf{k}$. Then $h^{\prime}_{\Delta}(t) \geq (1 + mt)^d$. Moreover, if $h^{\prime}_j(\Delta) = {d \choose j}m^j$ for some $1 \leq j \leq d-1$, then $\Delta$ is isomorphic to $\mathcal{P}(m+1,d)$.
\end{theorem}
\begin{proof}
We prove the claim by induction on $d$. When $d=1$, it is clear that $f_0(\Delta) \geq m+1$, so suppose that $d \geq 2$.
Let $F$ be a $(d-2)$-dimensional face of $\Delta$. Since $\mbox{\upshape lk}_{\Delta}F$ is $m$-Buchsbaum*, there are at least $m+1$ vertices $v_1, \ldots, v_{m+1}$ in $\mbox{\upshape lk}_{\Delta}F$. Since $\Delta$ is flag, no two of these vertices $v_i$ are connected by an edge in $\Delta$. In particular, this means that $\mbox{\upshape lk}_{\Delta-\{v_1, \ldots, v_i\}}(v_{i+1}) = \mbox{\upshape lk}_{\Delta}v_{i+1}$. Following the techniques of Theorem 1.3 in \cite{Ath} or Corollary 3.3 in \cite{AthW},
\begin{eqnarray*}
h^{\prime}_{\Delta}(t) &=& h^{\prime}_{\Delta-v_1}(t) + t\cdot h_{\mbox{\upshape lk} v_1}(t) \\
&=& h^{\prime}_{\Delta-\{v_1,v_2\}}(t) + t\cdot h_{\mbox{\upshape lk} v_2}(t) + t\cdot h_{\mbox{\upshape lk} v_1}(t) \\
&=& \cdots \\
&=& h^{\prime}_{\Delta-\{v_1,v_2,\ldots, v_m\}}(t) + t\cdot h_{\mbox{\upshape lk} v_m}(t) + \cdots + t\cdot h_{\mbox{\upshape lk} v_1}(t) \\
&\geq& h_{\mbox{\upshape lk} v_{m+1}}(t) + t\cdot h_{\mbox{\upshape lk} v_m}(t) + \cdots + t\cdot h_{\mbox{\upshape lk} v_1}(t) \qquad {\rm(\dagger)}\\
&\geq&(1+mt)^{d-1} + mt(1+mt)^{d-1} \\
&=& (1+mt)^d.
\end{eqnarray*}
To obtain line ${\rm(\dagger)}$, we use the fact that $\Delta$ is $m$-Buchsbaum* and hence $(m+1)$-Buchsbaum. Thus $\Delta - \{v_1,\ldots,v_m\}$ is a $(d-1)$-dimensional Buchsbaum complex and $h^{\prime}_i(\Delta - \{v_1, \ldots, v_m\}) \geq h_i(\mbox{\upshape lk}_{\Delta}v_{m+1})$ for all $0 \leq i \leq d$.
Suppose next that $h_j^{\prime}(\Delta) = {d \choose j}m^j$ for some $1 \leq j \leq d-1$. When $d=2$, the only case to consider is $j=1$. It is easy to see that the complete bipartite graph on two disjoint vertex sets of size $m+1$ is the only flag (i.e. triangle-free) $m$-Buchsbaum* graph with exactly $2(m+1)$ vertices.
Suppose now that $d \geq 3$. From the argument used above to show that $h_j^{\prime} \geq {d \choose j}m^j$, it follows that $h_j^{\prime}(\Delta) = {d \choose j}m^j$ if and only if $h_j(\mbox{\upshape lk}_{\Delta}u) = {d-1 \choose j}m^j$ and $h_{j-1}(\mbox{\upshape lk}_{\Delta}u) = {d-1 \choose j-1}m^{j-1}$ for all vertices $u \in \Delta$. In particular, one of the numbers $j$ and $j-1$ lies in the set $\{1,2,\ldots, d-2\}$, and so $\mbox{\upshape lk}_{\Delta}u$ is isomorphic to $\mathcal{P}(m+1,d-1)$ for all vertices $u \in \Delta$ by our inductive hypothesis.
Choose a vertex $u_1 \in \Delta$, and let $\Gamma:= \mbox{\upshape lk}_{\Delta}u_1$. Let $v$ be a vertex of $\Gamma$. Then $v$ has $(m+1)(d-1)$ neighbors in $\Delta$, and $(m+1)(d-2)$ of these neighbors lie in $\Gamma$. Let $u_1, u_2, \ldots, u_{m+1}$ be the neighbors of $v$ in $\Delta$ that do not lie in $\Gamma$. Now consider a vertex $v^{\prime} \in \Gamma$ that is adjacent to $v$. Then $\mbox{\upshape lk}_{\Delta}\{v,v^{\prime}\}$ has $(m+1)(d-2)$ vertices, and $(m+1)(d-3)$ of these vertices lie in $\Gamma$. The remaining $m+1$ vertices of $\mbox{\upshape lk}_{\Delta}\{v,v^{\prime}\}$ are adjacent to $v$, and the only such vertices are $\{u_1,\ldots,u_{m+1}\}$. Thus $u_iv^{\prime}$ is an edge for all $i$. Since $\Gamma$ is connected, it follows that $u_iw$ is an edge for all $w \in \Gamma$. We claim that $\Delta$ is connected, and since $\Delta$ is flag, it follows that $\Delta$ is isomorphic to $\mathcal{P}(m+1,d)$.
Finally, we show that $\Delta$ is connected. When $j=1$, this is obvious as each connected component of $\Delta$ requires $(m+1)\cdot d$ vertices. When $j \geq 2$, it is relatively easy to see that $h_j^{\prime}(\Delta) = \sum h_j^{\prime}(\Delta_t)$, where the sum is taken over all connected components $\Delta_t$ of $\Delta$. Each connected component of $\Delta$ is $m$-Buchsbaum*, and the claim follows.
\end{proof}
Taking $m=1$ answers Question \ref{lb}. We note that the assumption that $\Delta$ is flag in Theorem \ref{m-buchs} can easily be replaced by the assumption that $\Delta$ is balanced to yield the same conclusion. The following corollary is immediate and very interesting, especially when $m$ is large.
\begin{corollary}
Let $\Delta$ be a $(d-1)$-dimensional flag (or balanced) simplicial complex that is $m$-Buchsbaum* over $\mathbf{k}$. Then $(-1)^{d-1}\widetilde{\chi}(\Delta) \geq m^d$. In particular, if $\Delta$ is Cohen-Macaulay over $\mathbf{k}$, then $\widetilde{\beta}_{d-1}(\Delta) \geq m^d$.
\end{corollary}
\section{Concluding Remarks}
Recall that a set of vertices $\tau$ in a simplicial complex $\Delta$ is called a \textit{missing face} if $\tau \notin \Delta$ but $\sigma \in \Delta$ for all $\sigma \subsetneq \tau$. A flag simplicial complex, for example, only has missing faces of size two. For any simplicial complex $\Delta$, let $\Delta^{*q}$ denote the simplicial join of $q$ disjoint copies of $\Delta$. Let $\Sigma^j$ denote the $j$-dimensional simplex and $\partial\Sigma^j$ its boundary complex.
In \cite{Ne08}, Nevo studies the class of $(d-1)$-dimensional simplicial complexes with no missing faces of dimension larger than $i$. For fixed integers $d$ and $i$, write $d=qi+r$ where $q$ and $r$ are integers with $1 \leq r \leq i$. Nevo defines a certain $(d-1)$-dimensional sphere $S(i,d-1)$ with no missing faces of dimension larger than $i$ by $$S(i,d-1):= (\partial\Sigma^i)^{*q}*\partial\Sigma^r.$$
Goff, Klee, and Novik (\cite{GKN}, Theorem 3.1(2)) prove that $h_{\Delta}(t) \geq h_{S(i,d-1)}(t)$ for all $(d-1)$-dimensional simplicial complexes $\Delta$ that are 2-CM with no missing faces of dimension larger than $i$. This result and Theorem \ref{m-buchs} motivate the following question.
Fix integers $m, i,$ and $d$ and consider the class of $(d-1)$-dimensional simplicial complexes that are $m$-CM with no missing faces of dimension larger than $i$. As before, write $d = qi+r$ with $1 \leq r \leq i$, and consider the simplicial complex
\begin{displaymath}
S(m,i,d-1):= (\text{Skel}_{i-1}(\Sigma^{m+i-2}))^{*q}*\text{Skel}_{r-1}(\Sigma^{m+r-2}).
\end{displaymath}
Notice that $S(2,i,d-1)$ is Nevo's $S(i,d-1)$, and $S(m,1,d-1)$ is $\mathcal{P}(m,d)$. Each join-summand $\text{Skel}_{i-1}(\Sigma^{m+i-2})$ is $m$-CM, and hence $S(m,i,d-1)$ is a $(d-1)$-dimensional simplicial complex that is $m$-CM with no missing faces of dimension larger than $i$. It seems natural, therefore, to pose the following question.
\begin{question}
{\rm{Let $\Delta$ be a $(d-1)$-dimensional simplicial complex that is $m$-CM with no missing faces of dimension larger than $i$. Is it necessarily true that $$h_{\Delta}(t) \geq h_{S(m,i,d-1)}(t)?$$}}
\end{question}
\section*{Acknowledgements}
We are grateful to Christos Athanasiadis and Volkmar Welker for introducing us to the problems discussed in this paper. Our thanks also go to Christos Athanasiadis and Isabella Novik for a number of helpful conversations during the development of this paper. Jonathan Browder's research is partially supported by VIGRE NSF Grant DMS-0354131.
\bibliographystyle{plain}
|
\section{Introduction}
\label{sec1}
The combined symmetry CPT is supposed to be an exact symmetry of any local
axiomatic quantum field theory. This is indeed supported by the experiments:
all possible tests so far \cite{pdg} have yielded negative results, consistent
with no CPT violation. Why then should we be interested in the possibility of
CPT violation in the B system? There are three main reasons: first, any
symmetry which is supposed to be exact ought to be questioned and investigated,
and we may get a surprise, just like the discovery of CP violation; second,
it is not obvious that CPT will still be an exact symmetry in the bound
state of quarks and antiquarks, where the asymptotic states are not
uniquely defined \cite{quark}; third, there may be some nonlocal
and nonrenormalisable string-theoretic effects at the Planck scale which have a ramification at
the weak scale through the effective Hamiltonian \cite{cpt-hamil}.
Moreover, this effect can
very well be flavour-sensitive, and so the constraints obtained from the
K system \cite{nussinov} may not be applicable to the B systems. A comprehensive
study of CPT violation in the neutral K meson system, with a formulation
that is closely analogous to that in the B system, may be found in \cite{lavoura}.
There are already some investigations on CPT violation in B systems. Datta
{\em et al.} \cite{datta} have shown how CPT violation can lead to a significant
lifetime difference $\Delta\Gamma/\Gamma$ in the generic $M^0$-$\overline{M}
{}^0$ system, where $M^0=K^0, B^0$, or $B_s$. It was discussed in \cite{balaji}
how direct CP asymmetries and semileptonic decays can act as a
probe of CPT violation. Signatures of CPT violation in non-CP eigenstate channels
was discussed in \cite{xing1}. The role of dilepton asymmetry as a test of
CPT violation and possible discrimination from $\Delta B = - \Delta Q$ processes
were investigated in \cite{xing2}. The BaBar experiment at SLAC has tried to look for
CPT violation in the diurnal variations of CP-violating observables and set
some limits \cite{cpt-babar}.
Right now, there is no signature of CPT violation, or for that matter any
type of new physics, in the width difference of $B^0-\overline{B}{}^0$ and decay channels of
$B_d$ \footnote{We use $B^0$ and $\overline{B}{}^0$ to indicate the
flavour eigenstates, $B_d$ as a generic symbol for both of them, and similarly
for $B_s$. The symbol $B_q$ will mean either a $B_d$ or a $B_s$.}.
The width difference for the $B_d$ system, $\Delta\Gamma_d$, is too small
yet to be measured experimentally, and the bound is
compatible with the Standard Model (SM). On the other hand, it is expected
that the width difference $\Delta\Gamma_s$ would be significant for the $B_s$
system, but at
the same time we know that the theoretical uncertainties are quite
significant \cite{lenz}. If there is some new physics (NP) that does not
contribute to the absorptive part of the $B_s^0-\overline{B_s}{}^0$ box, the width difference
can only go down \cite{grossman}, while there are models where this conclusion
may not be true \cite{dighe}. To add to this murky situation, the CP-violating phase
$\beta_s$, which is expected to be very small from the CKM paradigm, has
been measured \cite{cdf-bs} to be large, compatible with the SM expectations
only at the $2.1\sigma$ level. The situation is interesting: there is hint of
some NP, but we are yet to be certain of its exact nature, not to mention
whether it exists at all.
In this situation, let us try to see what we can expect at the LHC, where
the $B_s$ meson, along with the $B_d$, will be copiously produced.
We are helped by the fact that the time resolution in ATLAS and
CMS are of the order of 40 fs, so one can track the time evolution of
even the rapidly oscillating
$B_s$. Thus, we expect excellent tagged and untagged measurements of both
$B_d$ and $B_s$ mesons. It is best to focus upon the single-amplitude
observables: $B_d\to J/\psi K_S$ and $B_s\to J/\psi\phi$ or $B_s\to D_s^+
D_s^-$ \footnote{They are not strictly single-channel as there is a penguin
process whose dominant part has the same phase as the leading Cabibbo-allowed
tree process, but on the other hand these channels are easy to measure, and
the penguin pollution is quite small and well under control.}.
For the $J/\psi\phi$ mode, one has to perform the angular
analysis and untangle the CP-even and CP-odd channels.
In this paper, we will discuss how one can detect the presence of a CPT
violating new physics from the tagged and untagged measurements of the
decay. We will confine our discussion to the case where CPT violation is
small compared to the SM amplitude, just to make the results more transparent.
The conclusions do not change qualitatively if the CPT violation is large,
which, we must say, is a far-off possibility based on the data from the
other experiments \cite{cpt-babar}.
We will also show how the nature of the CPT violating term
can be probed through these measurements.
In Section 2, we mention the relevant expressions, and introduce CPT violation,
with relevant expressions, in Section 3. The analysis for both $B_d$ and $B_s$
systems is performed in Section 4, while we summarise and conclude in Section
5.
\section{Basic Formalism}
Let us introduce CPT violation in the Hamiltonian matrix through the
parameter $\delta$ which can potentially be complex:
\begin{equation}
\delta = \frac{H_{22}-H_{11}}{\sqrt{H_{12}H_{21}}}\,,
\end{equation}
so that
\begin{equation}
\mathcal M = \left[\left(
\begin{array}{cc}
M_0-\delta' & M_{12}\\
M_{12}^* & M_0+\delta'
\end{array}
\right) - \frac{i}{2} \left(
\begin{array}{cc}
\Gamma_0 & \Gamma_{12}\\
\Gamma_{12}^* & \Gamma_0
\end{array}
\right)\right]\,,
\end{equation}
where $\delta'$ is defined by
\begin{equation}
\delta = \frac{2\delta'}{\sqrt{H_{12}H_{21}}}\,.
\end{equation}
Solving the eigenvalue equation of $\mathcal M$, we get,
\begin{eqnarray}
\lambda &=& \left(M_0- \frac{i}{2}\Gamma_0\right) \pm H_{12}\alpha y
\nonumber
\\
{\rm or},\ \ \lambda &=& \left[H_{11} + H_{12}\alpha \left(y +
\frac{\delta}{2}\right)\right]\,,\ \
\left[H_{22} - H_{12}\alpha \left(y + \frac{\delta}{2}\right)\right]\,,
\end{eqnarray}
where
$y = \sqrt{ 1 + \frac{\delta^2}{4} }$ and
$\alpha = \sqrt{ H_{21} / H_{12} }$.
Hence, corresponding eigenvectors or the mass eigenstates are defined as
\begin{eqnarray}
|B_H\rangle & =& p_1|B^0\rangle + q_1|\overline{B}{}^0\rangle\,,
\nonumber\\
|B_L\rangle &=& p_2|B^0\rangle - q_2|\overline{B}{}^0\rangle\,.
\end{eqnarray}
The normalisation satisfies
\begin{equation}
|p_1|^2 + |q_1|^2 = |p_2|^2 + |q_2|^2 = 1\,.
\end{equation}
Let us define,
\begin{equation}
\eta_1 = \frac{q_1}{p_1} = \left(y + \frac{\delta}{2}\right) \alpha\,;
\ \ \
\eta_2 = \frac{q_2}{p_2} = \left(y - \frac{\delta}{2}\right) \alpha\,;
\ \ \
\omega = \frac{\eta_1}{\eta_2}\,.
\label{fac1}
\end{equation}
The convention of \cite{cpt-babar} leads to $z_0=\delta/2$, where $z_0$ is a measure of CPT violation as
used in \cite{cpt-babar}. The limits imply that $|z_0|\ll 1$. Even if the origin of CPT violation is something
different, it is not unrealistic to assume $|\delta| \ll 1$.
One could even relax the assumption of $H_{21}=H_{12}^\ast$. However, there are two points that one
must note. First, the effect of expressing
$H_{12} = h_{12}+\bar\delta$, $H_{21} = h_{12}^\ast - \bar\delta$ appears as ${\bar\delta}^2$ in
$\sqrt{H_{12}H_{21}}$, the relevant expression in eq.\ (1), and can be neglected if we assume
$\bar\delta$ to be small. The second point, which is more important, is that CPT conservation constrains only
the diagonal elements and puts no constraint whatsoever on the off-diagonal elements.
It has been shown in \cite{lavoura,balaji} that $H_{12}\not= H_{21}^\ast$ leads to
T violation, and only $H_{11}\not= H_{22}$ leads to unambiguous CPT violation. Thus, we will
focus on the parametrization used in eqs.\ (1) and (2) to discuss the effects of CPT violation.
The time-dependent flavour eigenstates are given by
\begin{eqnarray}
|B_q(t)\rangle & = &f_+(t) |B_q\rangle + \eta_1 f_-(t) |\overline{B_q}\rangle \nonumber\\
|\overline{B_q}(t)\rangle & =& \frac{f_-(t)}{\eta_2} |B_q\rangle + \bar f_{+}(t)
|\overline{B_q}\rangle\,,
\end{eqnarray}
where
\begin{eqnarray}
\nonumber f_-(t) & = &
\frac{1}{(1 + \omega)} \left(e^{-i\lambda_1 t} - e^{-i\lambda_2 t}\right)\,, \\
\nonumber f_+(t) & = &\frac{1}{(1 + \omega)} \left(e^{-i\lambda_1 t} + \omega e^{-i\lambda_2 t}\right)\,, \\
\bar f_+(t) & = &\frac{1}{(1 + \omega)} \left(\omega e^{-i\lambda_1 t} + e^{-i\lambda_2 t}\right)\,.
\label{funct}
\end{eqnarray}
So, the decay rate of the meson $B_q$ at time $t$ to a CP eigenstate $f$
is given by
\begin{eqnarray}
\Gamma(B_q(t)\rightarrow f_{CP}) & = &
\left[|f_+(t)|^2 + |\xi_{f_1}|^2 |f_-(t)|^2 + 2{\rm Re}\left(\xi_{f_1} f_-(t) f_+^*(t)\right)\right] |A_f|^2\,, \nonumber\\
\Gamma(\overline{B_q}(t)\rightarrow f_{CP}) & = &
\left[|f_-(t)|^2 + |\xi_{f_2}|^2 |\bar f_+(t)|^2 + 2{\rm Re}\left(\xi_{f_2} \bar f_+(t) f_-^*(t)\right)\right] \left|\frac{A_f}{\eta_2}\right|^2\,,
\label{decay1}
\end{eqnarray}
where
\begin{equation}
A_f = \langle f|H|B_q\rangle\,,\ \ \
\bar A_f = \langle f|H|\overline{B_q}\rangle\,.
\end{equation}
Also,
\begin{equation}
\xi_{f_1} = \eta_1 \frac{\bar A_f}{A_f}\,,\ \ \
\xi_{f_2} = \eta_2 \frac{\bar A_f}{A_f}\,.
\end{equation}
In the SM, both are equal and $\xi_{f_1}=\xi_{f_2}=\xi_f$. For single-channel processes, $|\xi_f|=1$.
Now using eq.\ (\ref{fac1}) and eq.\ (\ref{funct}) one gets
\begin{eqnarray}
\left|f_-(t)\right|^2 & = &
\frac{2 e^{-\Gamma t}}{|1+\omega|^2} \left[\cosh\left(\frac{\Delta\Gamma t}{2}
\right) - \cos\left(\Delta{m} t\right)\right]\,,\nonumber\\
\left|f_+(t)\right|^2 & = &
\frac{e^{-\Gamma t}}{|1+\omega|^2} \Bigg[ \cosh\left(\frac{\Delta \Gamma t}{2}
\right)(1+|\omega|^2) + \sinh\left(\frac{\Delta \Gamma t}{2}
\right)(1-|\omega|^2) \nonumber \\
&& + 2 {\rm Re}(\omega) \cos\left(\Delta m t\right) - 2 {\rm Im}(\omega)
\sin\left(\Delta m t\right)\Bigg]\,, \nonumber \\
\left|\bar f_+(t)\right|^2 & =&
\frac{e^{-\Gamma t}}{|1+\omega|^2} \Bigg[ \cosh\left(\frac{\Delta \Gamma t}{2}
\right)(1+|\omega|^2) - \sinh\left(\frac{\Delta \Gamma t
}{2}\right)(1-|\omega|^2) \nonumber \\
&& + 2{\rm Re}(\omega) \cos\left(\Delta m t\right) + 2 {\rm Im}(\omega)
\sin\left(\Delta m t\right)\Bigg]\,,\nonumber \\
f_+^*(t) f_- (t) & =& \frac{e^{-\Gamma t}}{|1+\omega|^2}
\Bigg[ \cosh\left(\frac{\Delta \Gamma t}{2}\right)(1-\omega^{\ast}) -
\sinh\left(\frac{\Delta \Gamma t }{2}\right)(1+\omega^{\ast}) \nonumber \\
&& + \cos\left(\Delta m t\right) (-1 + \omega^{\ast}) - i \sin
\left(\Delta m t\right)(1+ \omega^{\ast}) \Bigg]\,, \nonumber \\
\bar{f_+}(t) f^{\ast}_{-} (t) & =&
\frac{e^{-\Gamma t}}{|1+\omega|^2} \Bigg[ \cosh\left(\frac{\Delta \Gamma t}{2}
\right)(\omega-1) -
\sinh\left(\frac{\Delta \Gamma t }{2}\right)(1+\omega) \nonumber \\
&& + \cos\left(\Delta m t\right) (1 - \omega) -
i \sin\left(\Delta m t\right)(1+ \omega) \Bigg]\,.
\label{fac2}
\end{eqnarray}
Here, $\Delta m$ and $\Delta \Gamma$ are defined through;
\begin{equation}
\lambda_1-\lambda_2 = \Delta m + \frac{i}{2}\Delta \Gamma\,,
\end{equation}
with
\begin{equation}
\lambda_{(1,2)} = m_{(1,2)} - \frac{i}{2} \Gamma_{(1,2)}\,, \ \ \
\Delta{m} = m_1-m_2\,, \ \ \
\Delta\Gamma = \Gamma_2 - \Gamma_1\,.
\end{equation}
\section{Introducing CPT Violation}
If we consider a time-independent CPT violation so that $\delta$ is a constant, there are only
two unknowns in the picture: Re($\delta$) and Im($\delta$), over those in the SM. We will try to see
how one can extract informations about them. For our analysis, let us take $\delta$ to be
complex; it will be straightforward to go to the simpler limiting cases where $\delta$ is purely
real or imaginary. For example, if the width difference $\Delta\Gamma$ is much smaller
than $\Delta m$, the model of \cite{cpt-babar} makes $\delta$ completely real.
When $B_q$ and $\bar{B}_q$ are produced in equal numbers, the untagged decay rate can be defined as
\begin{align}
\Gamma_U[f,t]=\Gamma(B_q(t)\to f) + \Gamma(\bar{B}_q(t)\to f)\,,
\label{untag}
\end{align}
using the above expression one could define the branching fraction as
\begin{align}
Br[f]= \frac{1}{2}\int^{\infty}_{0}{dt ~\Gamma[f,t]}\,.
\label{branch}
\end{align}
The above equation is useful to fix the overall normalization.
We assume, $\delta \ll 1$ and expand any function $f(\delta)$ using Taylor series expansion and drop all the terms ${\cal{O}}(\delta^n)$ for $n >2$.
From eq. (\ref{untag}), eq. (\ref{decay1}) and eq. (\ref{fac2}) we will get the untagged decay rate for the decay $B_q\to f$,
\begin{align}
\Gamma_U[f,t] &= |A_f|^2 e^{-\Gamma_q t} \Bigg[ \left\{(1+ |\xi_f|^2)(1+ \frac{({\rm Im}(\delta))^2}{4})- {\rm Im}(\delta) {\rm Im}(\xi_f)\right\} \cosh\left(\frac{\Delta\Gamma_q t}{2}\right) \nonumber \\
& \hskip 60pt - \left\{(1+ |\xi_f|^2)\frac{({\rm Im}(\delta))^2}{4}- {\rm Im}(\delta) {\rm Im}(\xi_f)\right\} \cos\left(\Delta m_q t\right) \nonumber \\
& \hskip 60pt + \left\{2 {\rm Re}(\xi_f)- \frac{1}{2}(1- |\xi_f|^2){\rm Re}(\delta)-\frac{1}{4} {\rm Re}(\xi_f)(({\rm Re}(\delta))^2- ({\rm Im}(\delta))^2)\right\} \times \nonumber \\
& \hskip 60pt \sinh\left(\frac{\Delta\Gamma_q t}{2}\right)- \frac{1}{2}{\rm Im}(\delta)\left\{(1- |\xi_f|^2)+ {\rm Re}(\delta){\rm Re}(\xi_f)\right\} \sin\left(\Delta m_q t\right)\Bigg]\,.
\end{align}
Thus, for the $B_s$ system, where the hyperbolic functions are not negligible, we get (keeping up to first order of
terms in $\Delta\Gamma_s$):
\begin{align}
Br[f] &= \frac{1}{2}\int^{\infty}_{0}{dt~ \Gamma[f,t]}\nonumber\\
&= \frac{|A_f|^2}{2} \Bigg[ \frac{1}{\Gamma_s}\left\{(1+ |\xi_f|^2)(1+ \frac{({\rm Im}(\delta))^2}{4})- {\rm Im}(\delta) {\rm Im}(\xi_f)\right\} \nonumber \\
& \hskip 40pt - \frac{\Gamma_s}{(\Delta m)^2+ (\Gamma_s)^2}\left\{(1+ |\xi_f|^2)\frac{({\rm Im}(\delta))^2}{4}- {\rm Im}(\delta) {\rm Im}(\xi_f)\right\} \nonumber \\
& \hskip 40pt + \frac{\Delta \Gamma_s}{2 (\Gamma_s)^2}\left\{2 {\rm Re}(\xi_f)- \frac{1}{2}(1- |\xi_f|^2){\rm Re}(\delta)-\frac{1}{4} {\rm Re}(\xi_f)(({\rm Re}(\delta))^2- ({\rm Im}(\delta))^2)\right\} \nonumber \\
& \hskip 40pt - \frac{1}{2}{\rm Im}(\delta)\left\{(1- |\xi_f|^2)+ {\rm Re}(\delta){\rm Re}(\xi_f)\right\} \frac{\Delta m}{(\Delta m)^2+ (\Gamma_s)^2}\Bigg]
\label{br-eq1}
\end{align}
Theoretically, one can obtain the coefficients of the trigonometric and the hyperbolic terms by fitting
the untagged decay rate. In actual cases this is a difficult task. However, there is one other observable which
may help us. Before we go to that, let us note that the
above expression is further simplified in the following four cases.
\begin{itemize}
\item For the $B_d$ system: We can neglect $\Delta\Gamma_d$
so that the cosh term is unity and the sinh term is zero. Thus, there are only two time-dependent
terms, $\cos(\Delta mt)$ and $\sin(\Delta mt)$, and the fitting is easier. Note that $\Delta\Gamma_d$ is
measured to be small, so we need not consider the case where it is enhanced to a significant value
because of the CPT violation. In fact, if $\delta$ is small, $\Delta\Gamma_d$ is bound to be that
coming from the SM, as the correction is proportional only to $\delta^2$ and higher.
\item For one-amplitude processes: We can put
$|\xi_f|=1$, and only one of Re($\xi_f$) and Im($\xi_f$) remains a free parameter
\footnote{$\xi_f$ is a SM quantity, so it is theoretically calculable, but it may also contain other new physics
which is CPT conserving, so it is better to obtain both real and imaginary parts of $\xi_f$ and
check whether $|\xi_f|=1$.}.
\item For $\delta$ being either purely real or purely imaginary: The expressions are straightforward.
For example, if $\delta$ is purely real, there is no trigonometric dependence on the untagged
rate.
\item Finally, for $|\delta| \ll 1$: We can neglect terms higher than linear in either Re($\delta$) or Im($\delta$)
in eq.\ (\ref{br-eq1}).
This is expected to be the case according to \cite{cpt-babar}. For example, the expression for the branching
fraction for a one-amplitude process simplifies to
\begin{equation}
Br[f]
= \frac{|A_f|^2}{2} \Bigg[ \frac{1}{\Gamma_s}\left\{2- {\rm Im}(\delta) {\rm Im}(\xi_f)\right\} +
\frac{\Gamma_s}{(\Delta m)^2+ (\Gamma_s)^2} {\rm Im}(\delta) {\rm Im}(\xi_f)+ \frac{\Delta \Gamma_s}{(\Gamma_s)^2} {\rm Re}(\xi_f)
\Bigg]\,.
\label{br-eq2}
\end{equation}
\end{itemize}
One can also tag the B mesons and define a tagged decay rate asymmetry
\begin{align}
\Gamma_T[f,t] &= \Gamma(B_q(t)\to f) - \Gamma(\bar{B}_q(t)\to f) \nonumber \\
& = |A_f|^2 e^{-\Gamma_q t} \Bigg[ \left\{(1- |\xi_f|^2)\frac{({\rm Re}(\delta))^2}{4}- {\rm Re}(\delta) {\rm Re}(\xi_f)\right\} \cosh\left(\frac{\Delta\Gamma_q t}{2}\right) \nonumber \\
& \hskip 60pt + \left\{(1 - |\xi_f|^2)(1- \frac{({\rm Re}(\delta))^2}{4}) + {\rm Re}(\delta){\rm Re}(\xi_f)\right\}\cos\left(\Delta m_q t\right) \nonumber \\
& \hskip 60pt - \frac{1}{2}{\rm Re}(\delta)\left\{(1+|\xi_f|^2)- {\rm Im}(\delta){\rm Im}(\xi_f)\right\}\sinh\left(\frac{\Delta\Gamma_q t}{2}\right)+ \Big\{2 {\rm Im}(\xi_f) \nonumber \\
& \hskip 60pt - \frac{1}{2}{\rm Im}(\delta)(1 +|\xi_f|^2)- \frac{1}{4} {\rm Im}(\xi_f)(({\rm Re}(\delta))^2- ({\rm Im}(\delta))^2)\Big\} \sin\left(\Delta m_q t\right)\Bigg]\,.
\label{untag5}
\end{align}
Note that (i) for Re($\delta$)=Im($\delta$)=0, this reverts back to the SM expression, as it should, and
(ii) If $|\delta|\ll 1$ and $\Delta\Gamma/\Gamma \ll 1$ as in the $B_d$ system, the tagged rate can measure
both Re($\delta$) and Im($\delta$).
For one-amplitude processes with $|\delta|\ll 1$, one may write a simplified expression:
\begin{align}
\Gamma_U[f,t] &= |A_f|^2 e^{-\Gamma_q t}\Bigg[ (2 - {\rm Im}(\delta) {\rm Im}(\xi_f)) \cosh\left(\frac{\Delta\Gamma_q t}{2}\right) \nonumber \\
& \hskip 60pt + {\rm Im}(\delta){\rm Im}(\xi_f)\cos\left(\Delta m_q t\right)
+ 2 {\rm Re}(\xi_f) \sinh\left(\frac{\Delta\Gamma_q t}{2}\right)
\Bigg]\,, \nonumber \\
\Gamma_T[f,t] & =|A_f|^2 e^{-\Gamma_q t}\Bigg[ - {\rm Re}(\delta) {\rm Re}(\xi_f) \cosh\left(\frac{\Delta\Gamma_q t}{2}\right) + {\rm Re}(\delta){\rm Re}(\xi_f)\cos\left(\Delta m_q t\right) \nonumber \\
& \hskip 60pt - {\rm Re}(\delta) \sinh\left(\frac{\Delta\Gamma_q t}{2}\right) + \left\{2 {\rm Im}(\xi_f) - {\rm Im}(\delta) \right\} \sin\left(\Delta m_q t\right)\Bigg]\,.
\label{untag6}
\end{align}
It is clear from eq.\ (\ref{untag6}) how one can extract ${\rm Re}(\delta)$ and
${\rm Im}(\delta)$ by comparing the untagged and tagged analyses. Suppose weconsider the $B_s$ system where $\Delta\Gamma_s$ is non-negligible. The coefficient
of the sinh term in $\Gamma_T$ gives ${\rm Re}(\delta)$. However, there is an
overall normalisation uncertainty given by $|A_f|^2$. To remove this, one can consider
a combined study of the coefficients of $\sinh\left(\frac{\Delta\Gamma_s t}{2}\right)$ and
$\cos\left(\Delta m_s t\right)$ from the untagged and tagged decay rates respectively; their
ratio allows for a clean extraction of ${\rm Re}(\delta)$. On the other
hand, the ratio of the coefficients of
$\cos(\Delta m_st)$ in $\Gamma_U$ and $\sin(\Delta m_st)$ in $\Gamma_T$
gives a clean measurement of ${\rm Im}(\delta)$, as ${\rm Im}(\xi_f)$ is known
from the SM dynamics. A further check about the one-amplitude nature is provided
from $|{\rm Re}(\xi_f)|^2 + |{\rm Im}(\xi_f)|^2 = 1$. In fact, as long as $\delta$ is small,
one can extract both ${\rm Re}(\delta)$ and ${\rm Im}(\delta)$ even if $|\xi_f|\not= 1$,
from the coefficients of the sine, cosine, and hyperbolic sine terms in $\Gamma_U$
and $\Gamma_T$.
One may also define the time-dependent CPT asymmetry as
\begin{equation}
A_{CPT}(f,t) = \frac{\Gamma_T[f,t]}{\Gamma_U[f,t]}
= \frac{ \Gamma(B_q(t)\to f) - \Gamma(\bar{B}_q(t)\to f) }
{ \Gamma(B_q(t)\to f) + \Gamma(\bar{B}_q(t)\to f) }\,,
\end{equation}
and the time-independent CPT asymmetry as
\begin{equation}
A_{CPT}(f) = \frac{\int_0^\infty dt~\Gamma_T[f,t]}{\int_0^\infty dt~\Gamma_U[f,t]}
= \frac{ \int_0^\infty dt~[\Gamma(B_q(t)\to f) - \Gamma(\bar{B}_q(t)\to f)] }
{ \int_0^\infty dt~[\Gamma(B_q(t)\to f) + \Gamma(\bar{B}_q(t)\to f)] }
\,.
\end{equation}
This goes to the usual CP asymmetry $A_{CP}$ if $\delta=0$; thus, any deviation from the expected CP
asymmetry calculated from the SM would signal new physics, but one must check all the boxes to
pinpoint the nature of the new physics. For example, there would not be any change in the semileptonic
CP asymmetry if the new physics is only CPT violating in nature.
\section{Analysis}
There are five {\em a priori} unknowns: ${\rm Re}(\delta)$, ${\rm Im}(\delta)$, ${\rm Re}(\xi_f)$, ${\rm Im}(\xi_f)$,
and $|A_f|^2$. For a one-amplitude process $|\xi_f|^2=1$ and the number of unknowns reduce to four. The
tagged and untagged decay rates, the branching fraction, and the time-independent CPT asymmetry would
provide informations on all of these unknowns. Assuming the CPT-conserving physics to be purely that
of the SM, one may calculate $\xi_f$ following the CKM picture. In the analysis that follows, we take $\xi_f$
to be known from the SM. We would like to point out the following features.
\begin{itemize}
\item
The overall amplitude $|A_f|^2$ cancels in the CPT asymmetry. This, therefore, is going to be the observable one
needs to measure most precisely.
\item
It is enough to measure the coefficients of the trigonometric terms only. For the $B_d$ system, $\Delta\Gamma_d$
is small anyway, and for the $B_s$ system, $\Delta\Gamma_s$ has a large theoretical uncertainty.
\item
The analysis holds even if the process under consideration is not a one-amplitude process. In fact, one may
check whether there is a second CPT conserving new physics amplitude just by looking at the extracted
values of ${\rm Re}(\xi_f)$ and ${\rm Im}(\xi_f)$.
\item
The coefficient of $\sin(\Delta m_qt)$ in the expression for the tagged decay rate $\Gamma_T$ gives the
mixing phase in the box diagram. Thus, ${\rm Im}(\delta)$ may be constrained by the CP asymmetry
measurements in the $B_d$ system. On the other hand, even those constrained values generate a large
mixing phase for the $B_s$ system compatible with the CDF data.
\end{itemize}
\subsection{The $B_s$ system}
For the $B_s$ system, we take
\begin{eqnarray}
&&\Delta m_s = 17.77\pm 0.12 {\rm ps}^{-1}\,,\
\Delta \Gamma_s =0.096\pm 0.039 {\rm ps}^{-1}\,, \
\frac{\Delta \Gamma_s}{\Gamma_s} = 0.147\pm 0.060\,, \nonumber\\
&&\frac{1}{\Gamma_s} = 1.530\pm 0.009 {\rm ps}\,, \
{\rm Re}(\xi_f)=0.99\,,\
{\rm Im}(\xi_f) = -0.04\,.
\end{eqnarray}
\begin{figure}[htbp]
\vspace{-10pt
\centerline{\hspace{-3.3mm}
\rotatebox{0}{\epsfxsize=12cm \epsfbox{acpt-red-bs.eps}}}
\caption{Variation of $A_{CPT}$ with ${\rm Re}(\delta)$ for the $B_s$ system.
The three lines, from top to bottom,
are for ${\rm Im}(\delta) = -0.1, 0$ and $0.1$ respectively.}
\label{acpt-bs}
\end{figure}
In figure \ref{acpt-bs}, we show the variation of $A_{CPT}$ with ${\rm Re}(\delta)$. For our analysis, we take
both $|{\rm Re}(\delta)|, |{\rm Im}(\delta)| < 0.1$, which is consistent with \cite{cpt-babar}. The variation of
$A_{CPT}$ with $\Delta m_s$ and $\Delta\Gamma_s$ is negligible, of the order of 0.2\%, so we fix
them to their respective central values. Effects of $\delta$ in both $\Delta m_s$ and $\Delta\Gamma_s$
are quadratic in $\delta$, and hence we can use the SM values for them. In fact, $A_{CPT}$ does not
depend significantly on the choice of ${\rm Im}(\delta)$ either; the variation is less than 1\%.
This is due to the fact that here, $|{\rm Im}(\xi_f)| \ll |{\rm Re}(\xi_f)|$ and hence the coefficient of
${\rm Re}(\delta)$ is much greater than the coefficient of ${\rm Im}(\delta)$ in the expression of
$A_{CPT}$. This feature does not hold for the $B_d$ system. Note that
$A_{CPT}$ clearly gives the sign of ${\rm Re}(\delta)$.
The small nonzero value of $A_{CPT}$
for $\delta=0$ indicates the small mixing phase in the $B_s^0-\overline{B_s}{}^0$ box diagram.
However, the apparent phase, i.e., the coefficient of $\sin(\Delta m_st)$, can increase with ${\rm Im}(\delta)$,
as can be seen from figure \ref{s2betas-imd}.
\begin{figure}[htbp]
\vspace{-10pt
\centerline{\hspace{-3.3mm}
\rotatebox{0}{\epsfxsize=12cm\epsfbox{s2betas-imd.eps}}}
\caption{Variation of $\sin(2\beta_s)$ with ${\rm Im}(\delta)$.}
\label{s2betas-imd}
\end{figure}
\subsection{The $B_d$ system}
The inputs that we use for the $B_d$ system are
\begin{equation}
\Delta m_d = 0.507 {\rm ps}^{-1}\,,\
\Delta \Gamma_d =0\,,\
{\rm Re}(\xi_f)=0.72\,,\
{\rm Im}(\xi_f) = 0.695\,.
\end{equation}
\begin{figure}[htbp]
\vspace{-10pt
\centerline{\hspace{-3.3mm}
\rotatebox{0}{\epsfxsize=12cm\epsfbox{acpt-red-bd.eps}}}
\hspace{3.3cm}
\caption{Variation of $A_{CPT}$ with ${\rm Re}(\delta)$ for the $B_d$ system.
The three lines, from top to bottom,
are for ${\rm Im}(\delta) = -0.1, 0$ and $0.1$ respectively.}
\label{acpt-bd1}
\end{figure}
\begin{figure}[htbp]
\vspace{-10pt
\centerline{\hspace{-3.3mm}
\rotatebox{0}{\epsfxsize=12cm\epsfbox{acpt-imd-bd.eps}}}
\hspace{3.3cm}
\caption{Variation of $A_{CPT}$ with ${\rm Im}(\delta)$ for the $B_d$ system.
The three lines, from top to bottom,
are for ${\rm Re}(\delta) = -0.1, 0$ and $0.1$ respectively.}
\label{acpt-bd2}
\end{figure}
This follows from the CKM expectation of $\sin(2\beta_d)=0.695\pm 0.020$. The constraint on $\delta$ comes
from the measurement of $\sin(2\beta_d)$ in the $b\to c\bar{c}s$ channel: $0.668 \pm 0.028$ \cite{utfit}
\footnote{We do not take the measurements coming from $b\to s$ penguin channels because of their
inherent uncertainties.}.
Again, we can fix $\Delta m_d$ at its central value. This time, due to the comparable values of
Re($\xi_f$) and Im($\xi_f$), $A_{CPT}$ is sensitive to both Re($\delta$) and Im($\delta$). The variations
are shown in figure \ref{acpt-bd1} for three values of Im($\delta$) and figure \ref{acpt-bd2} for three
values of Re($\delta$). It turns out that $A_{CPT}$ is always positive for ${\rm Re}(\delta), {\rm Im}(\delta)
< 1$; this is a consistency check for the CPT violation.
Note that the measured value of $\sin(2\beta_d)$ can go down from its CKM
expectation for ${\rm Im}(\delta)>0$, in fact, for ${\rm Im}(\delta)\approx 0.07$, $\sin(2\beta_d)\approx
0.66$, as can be seen from figure \ref{s2betad-imd}. While this value of ${\rm Im}(\delta)$ generates a
mixing phase for the $B_s$ system that is consistent with the CDF and D0 measurements at $1\sigma$,
one must remember that $\delta$ need not be a flavour-blind parameter.
\begin{figure}[htbp]
\vspace{-10pt
\centerline{\hspace{-3.3mm}
\rotatebox{0}{\epsfxsize=12cm\epsfbox{s2betad-imd.eps}}}
\hspace{3.3cm}
\caption{Variation of $\sin(2\beta_d)$ with ${\rm Im}(\delta)$.}
\label{s2betad-imd}
\end{figure}
\section{Summary and Conclusions}
We have investigated the possibility of CPT violation in neutral B systems. CPT is a symmetry that is
expected to be exact and the violation, even if it exists, should be quite small. However, it is possible
to measure even a small CPT violation from the tagged and untagged decay rates of the neutral B
mesons. In particular, for single-amplitude decay channels, the coefficients of the trigonometric terms
$\sin(\Delta mt)$ and $\cos(\Delta mt)$ can effectively pinpoint the nature of the CPT violating
parameter $\delta$. This is an interesting possibility for the decays $B_s\to D_s^+ D_s^-$ and $B_S
\to J/\psi \phi$ (with an angular analysis). Even a small CPT violation, allowed by the mixing
constraints for the $B_d$ system, can make the $B_s$ mixing phase more compatible with the
Tevatron measurements, at the level of about $1\sigma$.
On the other hand
CPT violation should not affect the semileptonic CP asymmetries, as the corrections are
quadratic in nature, and expected to be negligible for small $\delta$. Thus, a correlated study of
the CP asymmetries in $B_s\to J\psi\phi$ and $B_s\to D_s^+ D_s^-$ vis-a-vis
$B_s\to D_s \ell\nu$ might be useful to pinpoint the CPT violating effects.
This, we feel, is something that the
experimentalists should look for in the coming years at the LHC.
\centerline{\bf{Acknowledgements}}
SKP acknowledges CSIR, Govt.\ of India, for a research fellowship.
SN would like to thank Ulrich Nierste for useful discussions. His work is supported by a European Community's Marie-Curie Research Training Network under contract MRTN-CT-2006-035505 ``Tools and Precision Calculations for Physics Discoveries at Colliders". The work of AK was supported by BRNS, Govt.\ of
India; CSIR, Govt.\ of India; and the DRS programme of the University Grants Commission.
|
\section{\@startsection{section}{1
\z@{.7\linespacing\@plus\linespacing}{.5\linespacing
{\normalfont\bfseries \boldmath}}
\renewcommand\subsection{\@startsection{subsection}{2
\z@{.5\linespacing\@plus.7\linespacing}{-.5em
{\normalfont\bfseries \boldmath}}
\renewcommand\subsubsection{\@startsection{subsubsection}{3
\z@{.3\linespacing\@plus.5\linespacing}{-.5em
{\normalfont\bfseries \boldmath}}
\makeatother
\newtheorem{thm}{Theorem}[subsection]
\newtheorem{assump}[thm]{Assumption}
\newtheorem{lem}[thm]{Lemma}
\newtheorem{cor}[thm]{Corollary}
\newtheorem{hyp}[thm]{Hypothesis}
\newtheorem{prop}[thm]{Proposition}
\setcounter{section}{0}
\theoremstyle{definition}
\newtheorem{example}[thm]{Example}
\newtheorem{examples}[thm]{Examples}
\newtheorem{defn}[thm]{Definition}
\newtheorem{prob}[thm]{Problem}
\newtheorem{conj}[thm]{Conjecture}
\newtheorem{ques}[thm]{Question}
\newtheorem{rem}[thm]{Remark}
\newtheorem{rmk}[thm]{Remark}
\newtheorem{rems}[thm]{Remarks}
\newtheorem{subsec}[thm]{\!}
\numberwithin{equation}{subsection}
\newcounter{txtctr}[section] \setcounter{txtctr}{0}
\newenvironment{numtxt}
\refstepcounter{txtctr} \vspace{0.2cm} {\noindent
(\arabic{section}.\arabic{subsection}.\arabic{txtctr})
}\hspace{-36pt}\begin{minipage}[c]{\textwidth}\centering}
\end{minipage} \vspace{0.2cm}}
\renewcommand{\qedsymbol}{$_{\square}$}
\newcommand{\gl}{\mathfrak{g}}
\newcommand{\gll}{(\Lambda^2\gl)^*}
\newcommand{\glll}{(\Lambda^3\gl)^*}
\newcommand{\ul}{\mathfrak{u}}
\newcommand{\ull}{(\Lambda^2\ul)^*}
\newcommand{\ulll}{(\Lambda^3\ul)^*}
\newcommand{\uj}{\mathfrak{u}_J}
\newcommand{\gfq}{G(\mathbb{F}_q)}
\newcommand{\gfp}{G(\mathbb{F}_p)}
\newcommand{\pj}{\mathfrak{p}_J}
\newcommand{\lj}{\mathfrak{l}_J}
\newcommand{\Uz}{\mathcal{U}_{\zeta}}
\newcommand{\hh}{{\mathfrak h}}
\newcommand{\Fr}{{\text{\rm Fr}}}
\newcommand{\nc}{\mathcal{N}_1(\mathfrak{g})}
\newcommand{\rk}{\operatorname{rk}}
\newcommand{\Lie}{\operatorname{Lie}}
\newcommand{\mt}{\mapsto}
\newcommand{\ch}{\text{\rm ch}}
\newcommand{\ind}{\operatorname{ind}}
\newcommand{\Ext}{\operatorname{Ext}}
\newcommand{\into}{\hookrightarrow}
\newcommand{\onto}{\twoheadrightarrow}
\newcommand{\opH}{\operatorname{H}}
\renewcommand{\theenumi}{\roman{enumi}}
\renewcommand{\labelenumi}{(\theenumi)}
\newcommand{\Hom}{\operatorname{Hom}}
\newcommand{\End}{\operatorname{End}}
\newcommand{\sgn}{\operatorname{sgn}}
\newcommand{\soc}{\operatorname{soc}}
\newcommand{\head}{\operatorname{head}}
\renewcommand{\mod}{\operatorname{mod}}
\newcommand{\GH}{{\mathcal G}_{\operatorname{Hom}}}
\newcommand{\GT}{{\mathcal G}_{\otimes}}
\newcommand{\Tor}{\operatorname{Tor}}
\newcommand{\Lambdar}{\Lambda_{\operatorname{reg}}}
\newcommand{\Lambdas}{\Lambda_{\operatorname{sing}}}
\def\sgp#1#2{\left[\!\smallmatrix #1\cr #2\cr\endsmallmatrix\!\right]}
\newcommand{\ga}{\gamma}
\newcommand{\Ga}{\Gamma}
\newcommand{\la}{\lambda}
\newcommand{\La}{\Lambda}
\newcommand{\al}{\alpha}
\newcommand{\be}{\beta}
\newcommand{\ep}{\epsilon}
\newcommand{\Si}{\Sigma}
\newcommand{\si}{\sigma}
\newcommand{\om}{\omega}
\newcommand{\Om}{\Omega}
\newcommand{\vare}{\varepsilon}
\newcommand{\fee}{\varphi}
\newcommand{\de}{\delta}
\newcommand{\De}{\Delta}
\newcommand{\ka}{\kappa}
\newcommand{\tw}{\widetilde{\omega}}
\newcommand{\bZ}{{\mathbb Z}}
\newcommand{\BU}{{\mathbb U}}
\newcommand{\gr}{\text{\rm gr}\,}
\newcommand{\ad}{\operatorname{Ad}}
\newcommand{\adw}{\overline{\operatorname{Ad}}}
\newcommand{\Id}{\text{\rm Id}}
\newcommand{\cal}[1]{\mathcal{#1}}
\newcommand{\Ll}{\cal{L}(\lambda)}
\newcommand{\Hwl}{H^0(X(w),\Ll)}
\newcommand{\hwl}{H^0(w,\lambda)}
\newcommand{\Xw}{X(w)}
\newcommand{\XwP}{X(w)_P}
\newcommand{\XyP}{X(y)_P}
\newcommand{\V}{\mathcal{V}}
\newcommand{\N}{\mathcal{N}_1}
\newcommand{\NNu}{\mathcal{N}_1(\mathfrak{u})}
\newcommand{\Nu}{\boldsymbol{\mathfrak{u}}}
\newcommand{\supp}{\operatorname{supp}}
\begin{document}
\title[On the support varieties of Demazure modules]
{\bf On the support varieties of Demazure modules}
\begin{abstract} In \cite{NPV, UGA}, the support varieties for the induced modules/Weyl modules
for a reductive algebraic group $G$ were computed over the first Frobenius kernel $G_{1}$. A natural
generalization of this computation is the calculation of the support varieties of Demazure modules
over the first Frobenius kernel, $B_{1}$, of the Borel subgroup $B$. In the paper we initiate the study of
such computations. We complete the entire picture for reductive groups with underlying root
systems $A_{1}$ and $A_{2}$. Moreover, we give complete answers for Demazure modules
corresponding to a particular (standard) element in the Weyl group, and provide results relating support varieties
between different Demazure modules which depends on the Bruhat order.
\end{abstract}
\author{\sc Benjamin F Jones}
\address
{Department of Mathematics, Statistics, and Computer Science\\ University of Wisconsin-Stout \\
Menomonie\\ WI~54751, USA}
\thanks{Research of the first author was supported in part by NSF
VIGRE grant DMS-0738586.}
\email{<EMAIL>}
\author{\sc Daniel K. Nakano}
\address
{Department of Mathematics\\ University of Georgia \\
Athens\\ GA~30602, USA}
\thanks{Research of the second author was supported in part by NSF
grant DMS-1002135.} \email{<EMAIL>}
\subjclass[2000]{Primary 17B56, 17B10; Secondary 20G10}
\maketitle
\section{Introduction}
\subsection{} Let $G$ be a connected, simply connected, simple algebraic group
scheme defined over ${\mathbb F}_{p}$. Moreover, let $W$ be the associated
Weyl group, $B$ a Borel subgroup and $X(T)_{+}$ be the set of dominant
weights. Given $w\in W$ and $\lambda\in X(T)_{+}$, a natural set of
$B$-modules that arise are the Demazure modules labelled by $H^{0}(w,\lambda)$
which can be constructed using iterated inductions involving parabolics
corresponding to simple reflections occuring in a reduced decomposition of
$w$. When $w=w_{0}$ is the long element of $W$ one recovers the induced $G$-modules
$H^{0}(\lambda)=\text{ind}_{B}^{G} \lambda$ which can be realized as global
sections of the line bundle ${\mathcal L}(\lambda)$ over $G/B$.
Demazure modules arise naturally as the global sections on a line bundle
${\mathcal L}(\lambda)$ on the Schubert scheme $X(w)$ \cite[Ch. 14]{Jan}.
The structure of Demazure modules, and $B$-modules with excellent filtration
in general, is closely related to the geometry of the underlying Schubert
varieties (resolution of singularities, sheaf cohomology, normality, and
rational singularities). For example, Mehta and Ramanathan, using the
technique of Frobenius splittings, and later Andersen, using
representation-theoretic techniques, showed that the analog of Kempf's
vanishing theorem holds for sections of a dominant line bundle restricted to
a Schubert variety. This result was applied to complete Demazure's proof of
his character formula. As another example, Polo \cite{Polo} and van der
Kallen \cite{vanderKallen} use the normality of Schubert varieties in a
crucial way in their investigation of the category of $B$-modules with
excellent filtration.
\subsection{} In 2002, at a workshop in Seoul Korea, B. Parshall proposed the
problem of computing the support varieties of the Demazure modules
$H^{0}(w,\lambda)$ over the first Frobenius kernel $B_{1}$. This problem is
a natural and interesting extension of the ``Jantzen Conjecture" on support varieties
which predicted the support varieties of $H^{0}(\lambda)$ over $G_{1}$ when the
characteristic of the field is good. The conjecture was verified by Nakano,
Parshall and Vella \cite{NPV} and the support varieties of $H^{0}(\lambda)$
over $G_{1}$ were shown to be closures of Richardson orbits. This computation
was later extended to fields of bad characteristic by the University of Georgia VIGRE Algebra Group
\cite{UGA}. In the later case, the support variety of $H^{0}(\lambda)$ is
still irreducible and is the closure of an orbit, but the orbits need not be
Richardson.
Support varieties are natural with respect to the inclusion of $B_{1}$ in
$G_{1}$, so one can deduce from the aforementioned results that the $B_{1}$
support varieties of $H^{0}(\lambda)$ will be unions of closures of \emph{orbital varieties} (see \cite{Mel}). Indeed, orbital varieties should play an important role in the
general theory of support varieties of Demazure modules. This will be more
evident in the results in this paper.
The main obstacle in computing support varieties for general Demazure modules is that these modules
are rarely $G$-modules (i.e., their support varieties are not $G$-invariant, and not
closures of finitely many $G$-orbits). In general there are infinitely many $B$-orbits on
the nilpotent radical of $\Lie(B)$. At present it is not known how to classify these $B$-orbits.
The aim of the paper is to study the behavior of support varieties of Demazure modules.
In many instances we will be able to provide an explicit description of the supports.
The paper is organized as follows. In Section 2, we present various properties of
Schubert varieties that will be used throughout the paper. We then discuss properties
of support varieties over the Frobenius kernels $B_{r}$ and $P_{r}$. Several of the
main results in \cite{FP} and \cite{NPV} need to be modified and generalized for the purposes of this paper
(cf. Theorem 3.2.1 and Theorem 3.3.1). In Section 4, we prove a $G$-saturation result for the
$B_{r}$ support varieties of Demazure modules. In particular, we show that if $w_{1}<w_{2}$ (in
the Bruhat order) then $G\cdot {\mathcal V}_{B_{r}}(H^{0}(w_{2},\lambda))\subseteq
G\cdot {\mathcal V}_{B_{r}}(H^{0}(w_{1},\lambda))$. This result is subtle and we indicate by example that this inclusion
does not hold if one ignores the process of $G$-saturation (cf. Example 4.1.2). With these results, we describe the
supports of the Demazure modules in the $A_{1}$ case. Calculations of support varieties
${\mathcal V}_{B_{1}}(H^{0}(w,\lambda))$ are given for specific $w\in W$ in Section 5. Finally,
in Section 6, we provide a complete description of ${\mathcal V}_{B_{1}}(H^{0}(w,\lambda))$ for algebraic groups of type $A_{2}$.
An interesting facet of the $A_{2}$-computation is the need to analyze and use information about higher sheaf cohomology groups.
\section{Schubert Schemes}
\subsection{Notation}
Throughout this paper, let $k$ be an algebraically closed field of characteristic $p > 0$. For an algebraic group $H$,
the notation $\operatorname{Mod}(H)$ denotes the category of rational $H$-modules and
$\operatorname{mod}(H)$ denotes the category of finite dimensional, rational
$H$-modules.
Let $\Phi$ be a finite irreducible root system for a Euclidean space ${\mathbb E} $.
The inner product on ${\mathbb E} $ will be denoted by $(\ , \ )$.
For $\alpha\in\Phi$,
let
$\alpha^{\vee}=2\alpha/(\alpha,\alpha)$ be the corresponding coroot.
Fix a set $\Delta=\{\alpha_1,\cdots, \alpha_\ell\}$
of simple roots, and let $\Phi^{+}$ be the
corresponding set of positive roots.
The Weyl group $W\subset O(\mathbb E)$ is the group generated by the reflections
$s_\alpha:{\mathbb E}\to{\mathbb E}$, $\alpha\in\Phi$, given by
$s_\alpha(x)=x-2(x,\alpha^\vee)\alpha$.
Unless otherwise stated, $G$ will denote a reductive algebraic group over $k$.
We will always assume that the derived group $G'$ is simply connected.
Also, assume that $G$ has root system $\Phi$ with respect to a maximal split torus $T$.
Let $B\supset T$ be the Borel subgroup defined by $-\Phi^{+}$. The positive
Borel subgroup containing $T$ will be denoted $B^+$. Moreover,
let $X(T)=X(B)$
be the group of integral characters of $T$ or, equivalently, of $B$.
Given $\lambda\in X(T)$, we will let $\lambda$
also denote the one-dimensional $B$-module defined by regarding
$\lambda$ as a character on $B$. Then the set of dominant
integral weights is defined by
$$X_{+}:=X(T)_{+}=\{\lambda\in X(T)\,\,|\,\,\,
0\leq (\lambda,\alpha_i^{\vee}),\quad 1\leq i\leq\ell
\}.$$
Let $\rho$ be the half sum of the positive roots. We partially order
$X(T)$ by setting $\lambda\geq \mu$ if and only if
$\lambda-\mu\in \sum_{\alpha\in \Delta}\mathbb{N}\alpha$. Let $h$ be the
Coxeter number of $G$. Thus, if $G'$ is simple, $h=(\rho,\alpha_0^\vee)+1$
where $\alpha_0$ is the maximal short root in $\Phi$; otherwise,
$h$ is the maximal of the Coxeter numbers for the simple factors
of $G'$.
Each subset $J \subset \Delta$ gives rise to a standard parabolic
subgroup $P = P_J$ containing $B$ whose Lie algebra is generated by
$\mathfrak{t} = \Lie(T)$, the negative root spaces
$\mathfrak{g}_{-\alpha}$ ($\alpha \in \Phi_+$), and the positive root
spaces in the span of $J$: $\mathfrak{g}_{\alpha}$ for $\alpha \in
\Phi_{J}$. The subgroup $P_J$ has a Levi decomposition $P_J = L_J U_J$
where $\Lie(L_J)$ is generated by $\mathfrak{t}$ and the root spaces
$\mathfrak{g}_{\pm\alpha}$ for $\alpha \in J$ and $\Lie(U_J)$ is
generated by the root spaces $\mathfrak{g}_{-\alpha}$ for $\alpha \in
\Phi_+ \backslash \Phi_{J}$. We denote by $W_J$ the subgroup of $W$
generated by reflections $s_\alpha$ for $\alpha \in J$ and identify it
with the Weyl group of $L_J$. We denote the set of minimal length
right coset representatives for $W/W_J$ by $W^J$. When $P = P_J$ we also use
notations $W_P$ and $W^P$. We denote the opposite parabolic subgroup
that contains $B^+$ by $P_J^+$.
For $G$ as given above, the dominant weights $\lambda\in X(T)_+$ index the simple modules $L(\lambda)$
by their highest weight. If $\ind_B^G:\mod(B) \to
\mod(G)$ is the induction functor, let $H^0(\lambda)=\ind_B^G \lambda$ for
$\lambda\in X(T)$. If $\lambda\notin X(T)_+$, then
$H^{0}(\lambda)=0$, while if $\lambda\in X(T)_{+}$ then $H^0(\lambda)$
has socle $L(\lambda)$.
Let $F:G\to G$ be the Frobenius morphism
on $G$ induced by its ${\mathbb F}_p$-structure. For
$r\geq 1$, put $G_r={\text{\rm ker}}(F^r)$. If $H$ is an $F$-stable
subgroup of $G$, write similarly $H_r={\text{\rm
ker}}(F^r|_H)$---e.g., $B_r={\text{\rm ker}}(F^r|_B)$.
The group scheme $H_r$ is a finite $k$-group, i.e., an affine algebraic group
scheme over $k$ with finite dimensional coordinate algebra $k[H_r]$. Also, it has
height $\leq r$. In what follows, all affine $k$-groups $A$ will, by definition, be assumed to be algebraic, i.e.,
the coordinate algebra $k[A]$ is assumed to be finitely generated over $k$.
If $M\in \text{Mod}(H)$, let $M^{(r)}$ be the module in $\text{Mod}(H)$ obtained by composing the
representation corresponding to $M$ with $F^{r}$.
\subsection{Schubert Schemes}
\label{subsec:schubert}
In this section we follow the notation and conventions of \cite[II. Chapters 13-14]{Jan}. Fix a parabolic subgroup $P$. The group $G$ has a Bruhat decomposition:
\[ G = \bigcup_{w \in W^P} B \dot{w} P \] where $\dot{w}$ denotes a
chosen representative of $w$ in $N_G(T)$. This induces a decomposition
$G/P = \cup B\dot{w}P/P$ into $B$-stable affine subschemes (cells). We
denote by $X(w)_P$ the closure of the cell $B\dot{w}P/P$ in
$G/P$. These are the Schubert varieties of $G/P$. When $P=B$ is a
Borel subgroup, we simply use the notation $X(w)=X(w)_B$.
Let $M \in \mod(P)$. The variety $G \times_P M$ is naturally a vector bundle over $G/P$. We denote this vector bundle by $\cal{L}(M)$. The most important case is when $P=B$ and $M = k_\lambda$ for $\lambda \in X(T)$ in which case $\cal{L}(M)$ is a line bundle on $G/B$. If $J \subset \Delta$ and $\lambda$ satisfies $(\lambda, \alpha^\vee) = 0$ for all $\alpha \in J$, then there is a line bundle
$\cal{L}(\lambda)_P$ on $G/P$ where $P = P_J$. This bundle pulls back to $\cal{L}(\lambda)$ on $G/B$ under the quotient map $G/B \to G/P$ which is locally trivial. Therefore, by \cite[I 5.17]{Jan}, there is a canonical isomorphism $H^0(G/B, \Ll) \cong H^0(G/P, \Ll_P)$.
The cohomology groups $H^i(G/B, \cal{L}(M))$ are naturally $G$-modules.
For each $y \in W^P$ the inclusion $X(y)_P \into G/P$ induces the restriction map $H^i(G/P, \cal{L}(M)) \to
H^i(X(y)_P, \cal{L}(M))$.
The schemes $X(y)_P$ admit resolutions of singularities $\phi: X \to X(y)_P$ which are equivariant with respect to $B$ and depend on a reduced decomposition of $\dot{y}$, a minimal length coset representative of $y$ in $W$ (cf. \cite[13.6]{Jan}). The resolution $X$ is defined as a subset of a variety $Z$ which is a fiber bundle over $G/B$:
\begin{equation} \label{fig:resolution-diagram}
\xymatrix{
X \ar[r] \ar[d]_{\dot{\phi}} & Z \ar[d] \\
X(\dot{y}) \ar@{^{(}->}[r] \ar[d]_{\pi_P} & G/B \ar[d]^{\pi_P} \\
X(y)_P \ar@{^{(}->}[r]_i & G/P \quad .}
\end{equation}
In the diagram, $\pi_P$ is the natural projection $G/B \to G/P$ which is birational when restricted to $X(\dot{y})$ and the resolution $\phi$ is $\phi = \pi_P \circ \dot{\phi}$.
We need the following well known geometric results on Schubert varieties and sheaf cohomology.
\begin{prop} \label{prop:jan-cohomology} \cite[II. 14.15]{Jan} Let
$y \in W^P$, let $\dot{y}$ be a minimal length right coset representative
of $y$ in $W$, and let $w \in W$. Then the following hold:
\begin{enumerate}
\item $X(y)_P$ is normal, closed subscheme of $G/P$.
\item For every vector bundle $V$ on $G/P$ and $i\geq 0$, $H^i( X(y)_P, V ) \cong
H^i( X(\dot{y}), \pi_P^*V ) \cong H^i( X, \phi^*V ).$
\item For all $\lambda \in X(T)_+$, $H^i( X(w), \Ll ) = 0$ for $i>0$.
\item Given $\lambda \in X(T)_+$ such that $( \lambda, \alpha^\vee ) = 0$ for all $\alpha \in J$ where $J\subseteq \Delta$,
the restriction map $H^i( G/P, \Ll_P ) \to H^i( \XyP, \Ll_P )$ is surjective and moreover
$$H^i( \XyP, \Ll_P ) = 0$$ for all $i > 0$ where $P = P_J$ is the standard parabolic subgroup associated to $J$.
\end{enumerate}
\end{prop}
We also need the identification of the $G$-module $H^i(G/P,\cal{L}(M)_P)$ with induction from $P$ to $G$.
\begin{prop} \cite[I.5.12]{Jan}
\label{prop:ind-vs-cohomology}
\begin{enumerate}
\item For any $P$-module $M$ and $i\geq 0$ there is a canonical isomorphism
\[ R^i \ind^G_P M \cong H^i( G/P, \cal{L}(M) ) .\]
\item Let $H \subset K$ be $k$-group schemes such that $K/H$ is Noetherian (e.g., $K$ is reductive and $H$ is a parabolic, or $H \subset K \subset G$ are both parabolic in a reductive group) and let $M$ be a rational $H$-module. Then,
\[ R^i \ind^K_H M = 0 \]
for $i > \dim K/H$.
\end{enumerate}
\end{prop}
\section{Support Varieties over $P_{r}$}
\subsection{} In this section let $A$ be an arbitrary
finite $k$-group scheme and $\text{mod}(A)$ be the category of finite-dimensional
$A$-modules. We will consider maximal ideals in the commutative
part of the cohomology ring so set
$$R:=\operatorname{H}(A,k)=\begin{cases} \operatorname{H}^{2\bullet}(A,k) & \text{if $\text{char }k\neq 2$} \\
\operatorname{H}^{\bullet}(A,k) & \text{if $\text{char }k=2$}. \\
\end{cases}
$$
Friedlander and Suslin \cite{FS} proved that $R$ is a finitely generated $k$-algebra \cite{FS}.
Let $\V_{A}$ denote the variety associated to the maximum ideal spectrum of $R$.
Given $M,M'\in \text{mod}(A)$ we define the {\em relative support variety} $\V_A(M,M')
=\text{Maxspec}(R/J_{M,M^{\prime}})$ where $J_{M,M^{\prime}}$ is the annhilator
of the action of $R$ on $\text{Ext}^{\bullet}_{A}(M,M^{\prime})$. The action (Yoneda product) of
$R=\text{Ext}_{A}^{\bullet}(k,k)$ on $\text{Ext}^{\bullet}_{A}(M,M^{\prime})$ is given by taking an
extension in $R$ applying $- \otimes_{k} M^{\prime}$ then concatenating the new class with
an extension class in $\text{Ext}^{\bullet}_{A}(M,M^{\prime})$ (cf. \cite[Section 2.6]{Ben}).
The ordinary {\em support variety} of $M\in \text{mod}(A)$ is $\V_{A}(M):=\V_{A}(M,M)$.
In general for any $M,M'\in \text{mod}(A)$, $\V_A(M,M')$ is a homogeneous closed subvariety
contained in $\V_{A}=\V_{A}(k)$. For the basic properties of support varieties
for finite $k$-group schemes we refer the reader to \cite[Section 5]{FPe} and \cite[\S2.2]{NPV}.
Let $H$ be a closed subgroup of a finite $k$-group $A$ of height $\leq r$.
Suslin, Friedlander and Bendel \cite[(5.4)]{SFB2} proved that
the image of the restriction map $\text{res}:\text{H}(A,k)_{\text{red}}\to \text{H}(H,k)_{\text{red}}$
contains all $p^r$th powers $x^{p^r}$ of elements $x\in H(H,k)_{\text{red}}$, and the
induced morphism $\text{res}^{*}:\V_H\to \V_A$ maps $\V_H$ homeomorphically onto its image as
a closed subvariety of $\V_A$. In this paper we will identify the image of
$\V_{H}$ with $\text{res}^{*}(\V_{H})$ in $\V_{A}$. Under this map we have the
following naturality property.
\begin{prop}
\label{prop:naturality}
Let $H$ be a closed subgroup of $A$. Then $\V_{H}(M) =\V_{H}\cap \V_{A}(M)$.
\end{prop}
For infinitesimal group schemes of height 1, one can make the descriptions
of support varieties quite explicit. Let $H$ be an affine algebraic group scheme
defined over ${\mathbb F}_p$, $H_{1}=\text{ker }H_{1}$, and $\mathfrak{h}=\text{Lie }H$
(which is a restricted Lie algebra with $[p]$ operator). Let ${\cal N}_{1}({\mathfrak{h}})$
be the closed subvariety of nilpotent elements in $\mathfrak h$ of $H$ defined by
$${\cal N}_1(\mathfrak{h}):=\{x\in{\mathfrak h}\,|\,x^{[p]}=0\}.$$
We have following identification of varieties:
\begin{prop}
\label{prop:VH1} \cite[(1.6), (5.11)]{SFB1}
$\V_{H_1}$ is homeomorphic to ${\cal N}_1(\mathfrak{h})$.
\end{prop}
Finally, we can use the identification in (3.1.2) to identify $\V_{H_{1}}(M)$ as a closed subvariety of ${\cal N}_1(\mathfrak{h})$.
\begin{prop}
\label{prop:VH1M} \cite[(1.3) Theorem]{FP}
$\V_{H_1}(M)$ is homeomorphic to $$\{x\in {\mathcal N}_1({\mathfrak h}):\ M\ \text{is not }
x\text{-projective} \} \cup\{0\}.$$
\end{prop}
\subsection{} For the purposes of this paper we need to analyze the relationship of
support varieties over $B_{r}$ versus $P_{r}$ where $P$ is a parabolic subgroup of $G$.
The following result is a generalization of \cite[(1.2) Theorem]{FP} and \cite[Proposition 4.5.2]{Be}.
\begin{thm}
\label{thm:fp-bend-generalization}
Let $J\subseteq \Delta$, $P=P_{J}$ be the associated parabolic subgroup,
and $M\in \operatorname{mod}(P)$. Then
$$\V_{P_{r}}(M)=P\cdot \V_{B_{r}}(M).$$
\end{thm}
\begin{proof} The proof follows along the same line of reasoning as in \cite[(1.2) Theorem]{FP}.
We will indicate what modifications are necessary.
Let $\Psi=\text{res}^{*}:\V_{B_{r}}(M)\rightarrow
{\mathcal V}_{P_{r}}(M)$ be the map on varieties induced from the
restriction map $\text{res}:\text{H}^{\bullet}(P_{r},k)\rightarrow
\text{H}^{\bullet}(B_{r},k)$. According to
\cite[(1.6), (5.11)]{SFB1}, we can identify ${\mathcal V}_{B_{r}}(M)$ with
$\Psi({\mathcal V}_{B_{r}}(M))$ in ${\mathcal V}_{P_{r}}(M)$. Since ${\mathcal V}_{P_{r}}(M)$ is invariant
under $P$ we have
$$P\cdot {\mathcal V}_{B_{r}}(M)\subseteq {\mathcal V}_{P_{r}}(M).$$
We need to show that the reverse inclusion holds.
Following the proof of \cite[(1.2) Theorem]{FP}, set
\begin{eqnarray*}
I_{M}&=&\text{ker}\{\text{H}^{\bullet}(B_{r},k)\rightarrow \text{Ext}^{\bullet}_{B_{r}}(M,M)\}\\
J_{M}&=&\text{ker}\{\text{H}^{\bullet}(P_{r},k)\rightarrow \text{Ext}^{\bullet}_{P_{r}}(M,M)\}\\
K_{M}&=&\{x\in \text{H}^{\bullet}(P_{r},k):\ p\cdot \text{res}(x)\in I_{M} \ \forall p\in P\}\\
L_{M}&=&\{x\in \text{H}^{\bullet}(P_{r},k):\ p\cdot \text{res}(x)\in \sqrt{I_{M}} \ \forall p\in P\}
\end{eqnarray*}
Now replace ``$G$'' by ``$P$'', remove the ``symmetric algebras'', and use the fact
that $H^{m}(P/B,-)=0$ for $m> \dim P/B$. Then we can conclude that
$K_{M}\subseteq \sqrt{J_{M}}$, thus ${\mathcal V}_{P_{r}}(M)\subseteq
P\cdot {\mathcal V}_{B_{r}}(M)$.
\end{proof}
\subsection{}
\label{subsec:npv-541}
For $M$ a rational $B$-module, the relationship between the (relative) $B_r$ support variety of a module induced from $M$ and the $G_r$ support variety is described in \cite[Theorem 5.4.1]{NPV}. We generalize this result to the parabolic case as follows.
\begin{thm} \label{thm:npv-541}
Let $M$ be a rational $B$-module and $P$ be a parabolic subgroup of $G$ which contains $B$. Suppose that $R^m \ind^P_B M = 0$ for $m \ne t$, where $t$ is a fixed integer. Then,
\[ \V_{P_r}( R^t \ind^P_B M ) = P \cdot \V_{B_r}( R^t \ind^P_B M, M) .\]
\end{thm}
\begin{proof}
The proof of \cite[Theorem 5.4.1]{NPV} is formal and carries over after replacing $G$ by $P$.
The main issue involves the use of a spectral sequence which in our case is:
\[
E_2^{m,n}=R^m\ind^{P/P_r}_{B/B_r}\Ext^n_{B_r}(R^t
\ind_B^P M,M)\Rightarrow
\Ext^{m+n-t}_{P_r}(R^t\ind_B^P M,R^t\ind_B^P M),
\]
and an increasing filtration whose finiteness depends on a vanishing result,
\[ R^m \ind^{P/P_r}_{B/B_r} = 0 \quad \text{for} \quad m > \dim P/B. \]
This vanishing result holds by Proposition \ref{prop:ind-vs-cohomology}(ii).
\end{proof}
\section{$G$-Saturation}
\subsection{} We are interested in determining the support varieties $\V_{B_1}(H^0(X(w), \Ll))$ for all $w \in W$
and $\lambda \in X_+$. In particular, we want to understand the inclusion relations among support varieties for different $w$ and $\lambda$
of particular interest. In some instances we will use $H^{0}(w,\lambda):=H^{0}(X(w),{\mathcal L}(\lambda))$ as
a short hand notation. In the following theorem, we prove that for a fixed weight $\lambda$, the inclusion relation on
the $G$-saturation of support varieties for Demazure modules respects the Bruhat order on $W$.
\begin{thm}
\label{thm:saturation}
Let $\lambda \in X_+$ and $w_1 < w_2 $ in the Bruhat order on $W$. Then,
\[
G \cdot \V_{B_r}(H^0(w_2,\lambda)) \subseteq G \cdot \V_{B_r}(H^0(w_1,\lambda)) .
\]
\end{thm}
\begin{proof}
By induction on $\ell(w_2) - \ell(w_1)$, it suffices to prove the result when
$w_2 = s_\alpha w_1$ and $\ell(w_2) = \ell(w_1) + 1$. Let $P_\alpha$ be the minimal parabolic corresponding to
$\alpha$. By Theorem~\ref{thm:fp-bend-generalization},
\begin{equation}
\label{eq:sat1}
\V_{(P_\alpha)_r}( H^0(w_2,\lambda) ) = P_\alpha \cdot \V_{B_r}( H^0(w_2,\lambda) ) .
\end{equation}
Since $H^0(w_2,\lambda) \cong \ind^{P_\alpha}_B H^0(w_1,\lambda)$, Theorem \ref{thm:npv-541} implies:
\begin{equation}
\label{eq:sat2}
\begin{array}{ll}
P_\alpha \cdot \V_{B_r}( H^0(w_2, \lambda) ) & = P_\alpha \cdot \V_{B_r}( H^0(w_2, \lambda) , H^0(w_1, \lambda) ) \\
& \subseteq P_\alpha \cdot \V_{B_r}( H^0(w_1, \lambda) ).
\end{array}
\end{equation}
Combining \eqref{eq:sat1} and \eqref{eq:sat2} we have
\[ \V_{(P_\alpha)_r}( H^0(w_2, \lambda) ) \subseteq P_\alpha \cdot \V_{B_r}( H^0(w_1, \lambda) ) ,\]
so acting by $G$ on both sides we certainly have:
\[ G \cdot \V_{(P_\alpha)_r}( H^0(w_2, \lambda) ) \subseteq G \cdot \V_{B_r}( H^0(w_1, \lambda) ) .\]
Finally, by (3.1.1) $\V_{B_r}(M) \subseteq \V_{(P_\alpha)_r}(M)$ for all $M \in \mod(P_\alpha)$. Thus,
\[ G \cdot \V_{B_r}( H^0(w_2, \lambda) ) \subseteq G \cdot \V_{B_r}( H^0(w_1, \lambda) ) .\]
\end{proof}
\subsection{} We should remark that the result above is rather subtle in the sense that inclusion of the
$B_{1}$-support varieties of Demazure modules need not be preserved under the Bruhat order.
This can be seen in the following example.
\begin{example}
\label{ex:A2-steinberg}
Let $p\geq 3$, $\lambda = (p-1)\rho$ (the Steinberg weight), and $G = SL(3)$.
Let ${\mathfrak u}_{\alpha}$ (resp. ${\mathfrak u}_{\beta}$) be the
unipotent radical of the Lie algebra of $P_{\alpha}$ (resp. $P_{\beta}$).
The computation in Section \ref{subsec:A2} gives the support varieties
$\V_{B_1}(H^0(w, (p-1)\rho))$ for all $w \in W$, see Table 1 (below).
\begin{table}[H]
\label{tab:A2-steinberg}
\begin{tabular}{c|c}
$w$ & $\V_{B_1}(H^0(w, (p-1)\rho))$ \\
\hline
$e$ & $\mathfrak{u}$ \\
$s_\alpha$ & $\mathfrak{u}_\alpha$ \\
$s_\beta$ & $\mathfrak{u}_\beta$ \\
$s_\alpha s_\beta$ & $\mathfrak{u}_\alpha \cup \mathfrak{u}_\beta$ \\
$s_\beta s_\alpha$ & $\mathfrak{u}_\alpha \cup \mathfrak{u}_\beta$ \\
$w_0$ & $\{0\}$
\end{tabular}
\vspace{10pt}
\caption{Support varieties for Demazure modules in type $A_2$ with highest
weight $(p-1)\rho$.}
\end{table}
In particular, the pair $s_\beta$ and $s_\alpha s_\beta$ illustrate that $w_1 <
w_2$ does not neccesarily imply that $\V_{B_1}(H^0(w_2,\lambda)) \subseteq
\V_{B_1}(H^0(w_1,\lambda))$. Note, however, that the saturations in these two
cases agree:
\[ G \cdot \mathfrak{u}_\beta = G \cdot \mathfrak{u}_\alpha = G \cdot
(\mathfrak{u}_\alpha \cup \mathfrak{u}_\beta) .\]
\end{example}
\subsection{The Regular Case}
Fix a dominant weight $\lambda$. The subset
\[ \Phi_{\lambda, p} = \{ \alpha \in \Phi^+ \mid (\lambda + \rho, \alpha^\vee) \in p\mathbb{Z} \} \]
is a subroot system of $\Phi$ which, when the prime $p$ is good relative to $\Phi$, and conjugate under the Weyl group to a
root system $\Phi_I$ spanned by a subset $I \subseteq \Delta$ of simple roots, see \cite[Prop. 24, pg. 165]{Bo}. The weight $\lambda$ is called \emph{$p$-regular} if $\Phi_{\lambda,p} = \emptyset$.
\begin{prop} \label{prop:regular-case} Let $\lambda$ be a $p$-regular weight in $X_{+}$,
then $\V_{B_1}(H^0(w,\lambda)) =\V_{B_{1}}$.
\end{prop}
\begin{proof} If $w_0$ denotes the
longest element of the Weyl group, then $w \le w_0$ so Theorem
\ref{thm:saturation} gives us an inclusion of the saturated
supports:
\[ G \cdot \V_{B_1}(H^0(w_0, \lambda)) \subseteq G \cdot
\V_{B_1}(H^0(w, \lambda)).\] Since $X(w_0) = G/B$, $H^0(w_0, \lambda)$
is a $G$-module and we have
$$\V_{G_1}(H^0(w_0, \lambda)) =\V_{G_1}(H^0(G/B, \Ll)).$$
Moreover, by \cite[(1.2) Theorem]{FP},
$\V_{G_1}(H^0(G/B, \Ll)) = G \cdot \V_{B_1}(H^0(G/B, \Ll))$. Putting
these results together, we have
\begin{equation}
\label{eq:regular-case-eq-1}
\V_{G_1}(H^0(w_0, \lambda)) \subseteq G \cdot \V_{B_1}(H^0(w,
\lambda)) .
\end{equation}
Since $\lambda$ is $p$-regular, $\V_{G_1}(H^0(w_0, \lambda)) = \V_{G_{1}}$ by
\cite[Proposition (4.1.2)]{NPV} and thus
\[ \V_{G_{1}} \subseteq G \cdot \V_{B_1}(H^0(w, \lambda)) \subseteq \V_{G_{1}}.\]
Therefore, we must have
\[ G \cdot \V_{B_1}(H^0(w, \lambda))=\V_{G_{1}}. \]
Since $\lambda$ is $p$-regular we have $p\geq h$ (cf. \cite[6.2 (9)]{Jan}), and
$\V_{G_{1}}$ identifies with the nilpotent cone in ${\mathfrak g}$. Therefore,
the closed conical $B$-stable variety $\V_{B_1}(H^0(w, \lambda))$ must contain a regular nilpotent element.
It follows that
$$\V_{B_1}(H^0(w, \lambda))={\mathfrak u}={\mathcal N}_{1}({\mathfrak u})=\V_{B_{1}}.$$
\end{proof}
\subsection{The Root System $A_{1}$}
\label{subsec:A1}
We conclude this section by illustrating Proposition
\ref{prop:regular-case} in the situation when the root
system $\Phi$ is $A_1$ (i.e., for the group $G = \operatorname{SL}(2)$).
Let $G = \operatorname{SL}(2)$ and $\lambda$ be a dominant integral
weight (represented by a non-negative integer). In
this case $G/B \cong \mathbb{P}^1$ and $W = \{ e, s_\alpha \}$.
Let $w=e$, we have $X(e) = eB/B \cong \{ \text{pt}. \}$. It follows
that $\dim H^0(w, \lambda) = 1$ and so by the rank variety
description, $\V_{B_1}(H^0(w,\lambda)) = {\mathfrak u}$ which is
independent of $\lambda$.
The case $w = s_\alpha$ is the long element of $W$ so we have
$X(s_\alpha) = G/B$. Now the weight $\lambda$ is $p$-regular if
and only if $p \nmid \lambda + 1$. So by Proposition
\ref{prop:regular-case}, $\V_{B_1}(H^0(s_\alpha, \lambda)) = \V_{B_1}
= \mathfrak{u}$ unless $p \mid \lambda + 1$. When $p \mid \lambda +
1$, a simple application of \cite[Theorem 6.2.1]{NPV} gives
\[ \V_{B_1}(H^0(w, \lambda ) ) = \{ 0 \} .\]
We summarize the situation for type $A_1$ in Table \ref{tab:A1}.
\begin{table}[!ht]
\begin{tabular}{|c||c|c|c|}
\hline
& $w$ & & \\
\hline
\hline
$\mathbf{\lambda}$ & & $p \nmid \lambda+1$ & $p \mid \lambda+1$ \\
\hline
& $e$ & $\mathfrak{u}$ & $\mathfrak{u}$ \\
\hline
& $s_\alpha$ & $\mathfrak{u}$ & $\{0\}$ \\
\hline
\end{tabular}
\bigskip
\caption{Calculation of support varieties for all Demazure modules in
type $A_1$.}
\label{tab:A1}
\end{table}
\section{Calculation of Support Varieties}
\label{sec:calc}
In this section we determine the support varieties of Demazure modules for arbitrary
reductive groups $G$ when the underlying Schubert scheme corresponds either to the longest element in $W_I$ (for any $I\subseteq\Delta$) or to the longest element in $W^J$ (for certain subsets $J \subseteq \Delta$).
\subsection{Long elements in $W^J$} Let $\lambda \in X_+$ and define
\[ J_\lambda = \left\{ \alpha \in \Delta \mid \langle \lambda, \alpha^\vee \rangle = 0 \right\}. \]
For any subset $J \subseteq \Delta$, let $w_{0,J}$ denote
the Weyl group element of maximal length in $W^J$.
\begin{prop} \label{prop:w-is-w0J} There is an isomorphism of $B$-modules
\[ H^0(X(w_{0,J_\lambda})_{P_{J_\lambda}}, \Ll) \cong H^0(G/B, \Ll) .\]
\end{prop}
\begin{proof}
For simplicity, let $w = w_{0,J_\lambda}$ and $P = P_{J_\lambda}$ for the remainder of this section.
The resolution diagram \eqref{fig:resolution-diagram} induces a diagram involving cohomology groups:
\begin{equation} \label{fig:cohomology-diagram}
\xymatrix{
H^0(X(\dot{w}),\Ll) & H^0(G/B,\Ll) \ar[l]_{j^*} \\
H^0(X(w)_P,\Ll_P) \ar[u]_{(\pi_P\lvert_{X(\dot{w})})^*} & H^0(G/P,\Ll_P) \ar[l]_{i^*} \ar[u]_{\pi_P^*}} .
\end{equation}
By Proposition \ref{prop:jan-cohomology}, $(\pi_P\lvert_{X(\dot{w})})^*$ is an
isomorphism. Also, the choice of $w$ implies that $X(w)_P = G/P$, hence $i$
is the identity and $i^*$ is an isomorphism. By local triviality (cf. \cite[I
5.17]{Jan}), the map $\pi_P^*$ is an isomorphism. The diagram
commutes, thus $j^*$ is an isomorphism.
\end{proof}
The proposition and \cite[Theorem 6.2.1]{NPV} allows us to identify the support
variety of $H^0(X(w)_P, \Ll_P) $ in this special case.
Choose $x \in W$ such that $x(\Phi_{\lambda,p}) = \Phi_I$
for some subset $I \subseteq \Delta$.
\begin{thm} \label{thm:w-is-w0J} With $J=J_\lambda$, $w=w_{0,J}$, and $P = P_{J}$ as above,
\[ {\mathcal V}_{B_1}( H^0(X(w)_P, \Ll_P) ) = ( G\cdot \mathfrak{u}_I )
\cap \NNu .\]
\end{thm}
\begin{proof} By \cite[(6.2.1) Theorem]{NPV},
${\mathcal V}_{G_1}( H^0(G/B, \Ll) ) =G\cdot\mathfrak{u}_I$. The isomorphism of
Proposition \ref{prop:w-is-w0J} along with naturality of supports, see (3.1.1), implies that
\begin{align*} {\mathcal
V}_{B_1}( H^0(X(w)_P, \Ll_P) ) & = {\mathcal V}_{B_1}( H^0(G/B, \Ll) ) \\ & =
{\mathcal V}_{G_1}( H^0(G/B, \Ll) ) \cap \NNu \\ & = ( G\cdot \mathfrak{u}_I ) \cap
\NNu . \end{align*}
\end{proof}
Theorem~\ref{thm:w-is-w0J} implies that the $B_1$ support varieties of
certain Demazure modules are unions of the closures of orbital varieties. Recall from the introduction
that the $B_1$ support varieties of induced modules $H^0(G/B, \Ll)$
are also unions of orbital variety closures. It remains an interesting open problem
whether or not the support varieties of all Demazure modules are unions of orbital variety
closures and whether one can realize all such closures as support varieties of certain
modules.
\subsection{Longest element in $w_{I}$} In this section, let $I\subseteq \Delta$ be an arbitrary subset and let
$w=w_{I}\in W_{I}$ such that $w_{I}(\alpha)<0$ for all $\alpha\in I$. The
element $w_{I}$ is the long element in the group $W_{I}$. First, note that in
this case by \cite[II 13.3 (4)]{Jan}
\begin{equation*}
H^{0}(X(w),{\mathcal L}(\lambda))\cong \text{ind}_{B}^{P_{I}}\lambda .
\end{equation*}
Consequently, $H^{0}(X(w),{\mathcal L}(\lambda))$ is a
$P_{I}$-module with $U_{I}$ acting trivially. The following theorem describes
the support variety of $H^{0}(X(w),{\mathcal L}(\lambda))$ as a
$(P_{I})_{1}$-module by reducing down to case of \cite[Theorem 6.2.1]{NPV} for
the Levi subgroup $L_{I}$.
\begin{thm} Let $I\subseteq \Delta$ with ${\mathfrak u}_{I}=\operatorname{Lie
}U_{I}$, and $w=w_{I}$. Then
$$\V_{(P_{I})_{1}}(H^{0}(X(w),\Ll))=
[\V_{(L_{I})_{1}}(H_{I}^{0}(\lambda))+{\mathfrak u}_{I}] \cap {\mathcal N}_{1}({\mathfrak p}_{I}).$$
\end{thm}
\begin{proof}
Set ${\mathfrak l}_{I}=\text{Lie }L_{I}$ and ${\mathfrak
u}_{I}=\operatorname{Lie }U_{I}$. First observe by \cite[(4.2) Examples]{CPS}
that
\begin{equation*} \text{ind}_{B}^{P_{I}}\lambda|_{L_{I}}\cong
\text{ind}_{L_{I}\cap B}^{L_{I}}\lambda:=H^{0}_{I}(\lambda).
\end{equation*}
Let $z=x+y$ where $x\in {\mathfrak l}_{I}$, $y\in {\mathfrak u}_{I}$ and $z\in
{\mathcal N}_{1}({\mathfrak p}_{I})$. Then by \cite[Proposition 5.2(a)]{CLNP},
$x\in {\mathcal N}_{1}({\mathfrak l}_{I})$. Since ${\mathfrak u}_{I}$ acts trivially
on $H^{0}_{I}(\lambda)$ we have
\begin{equation*}\label{E:compareim}
z.H^{0}_{I}(\lambda)=x.H^{0}_{I}(\lambda).
\end{equation*}
In particular, $H^0_I(\lambda)$ is $z$-projective if and only if
it is $x$-projective. By
the realization of the support varieties in terms of rank varieties, we can
conclude that $z\in {\mathcal V}_{(P_{I})_{1}}(H^{0}(X(w),{\mathcal
L}(\lambda)))$ if and only if $x\in {\mathcal
V}_{(L_{I})_{1}}(H^{0}(X(w),{\mathcal L}(\lambda)))$.
Therefore,
\begin{equation*} {\mathcal V}_{(P_{I})_{1}}(H^{0}(X(w),{\mathcal
L}(\lambda))) = [{\mathcal
V}_{(L_{I})_{1}}(H_{I}^{0}(\lambda))+{\mathfrak u}_{I}] \cap
{\mathcal N}_{1}({\mathfrak p}_{I}).
\end{equation*}
\end{proof}
Using \cite[Theorem 6.2.1]{NPV} we obtain the following description of the
support variety. Recall that when the prime $p$ is good there exists $x \in W$ such that $x(\Phi_{\lambda,p}) =
\Phi_J$ for some subset $J \subseteq \Delta$.
\begin{cor} \label{cor:w-is-wI}
Let $w = w_I$ as above, let $\lambda \in X(T)_+$, and suppose $p$ is a good prime for $\Phi$. Let $x \in W_I$ be such that $x( (\Phi_I)_{\lambda,p} ) = (\Phi_I)_J$ for some
subset $J \subset I$. Let $\mathfrak{u}_{I,J}$ be the nilradical of the
parabolic in $\mathfrak{l}_I$ corresponding to $J$. Then,
\begin{eqnarray*}
\V_{(P_{I})_{1}}(H^{0}(X(w),\Ll)) & = & (L_I \cdot \mathfrak{u}_{I,J} + {\mathfrak u}_{I})\cap
{\mathcal N}_{1}({\mathfrak p}_{I}) \\
& = & \left( L_I \cdot (\mathfrak{u}_{I,J} + {\mathfrak u}_{I}) \right) \cap {\mathcal N}_{1}({\mathfrak p}_{I}) \\
& = & \left( L_I \cdot \mathfrak{u}_J \right) \cap {\mathcal N}_{1}({\mathfrak p}_{I}).
\end{eqnarray*}
\end{cor}
\subsection{Parabolic Upper and Lower Bounds}
The explicit calculation of Corollary \ref{cor:w-is-wI} and the inclusions among saturated support varieties in Theorem \ref{thm:saturation} give upper and lower bounds on the saturation $G \cdot \V_{B_1}(\hwl)$ for arbitrary $w \in W$ and $\lambda \in X(T)_+$. To state the bounds obtained we introduce some notation. For $v \in W$, let
$v = s_{\gamma_1} \cdots s_{\gamma_n}$ be a reduced expression. Define the support of $v$ by $S(v) = \{ \gamma_1, \ldots,
\gamma_n \}$. This definition is independent of the reduced expression chosen (cf. \cite[Theorem 3.3.1]{Bj}). As in the previous section, $w_I$ denotes the long element of $W_I$ for a subset $I \subseteq \Delta$.
\begin{lem}
\label{lem:w-bounds}
If $v \in W$ then $v \le w_{S(v)}$. Moreover, $v \le w_I$ implies that $S(v) \subseteq I$ and $w_{S(v)} \le w_I$.
\end{lem}
\begin{proof}
This is a consequence of \cite[Theorem 5.10]{Hum}.
The expression $v = s_{\gamma_1} \cdots s_{\gamma_n}$ implies that $v \in W_{S(v)}$ and hence $v \le w_{S(v)}$ since the latter is the unique longest element of
$W_{S(v)}$. Similarly, $v \le w_I$ implies that the generators of $W_{S(v)}$ are contained in $W_I$ hence $w_{S(v)} \le w_I$.
\end{proof}
The lemma gives us a precise characterization of the least upper bound by
elements of the form $w_I$ where $I \subseteq \Delta$.
In general there is no unique greatest lower bound as Example \ref{ex:no-unique-lower-bound} shows.
\begin{example}
\label{ex:no-unique-lower-bound}
Let $W$ be the Weyl group of type $A_3$ generated by simple reflections $s_{\alpha_1}, s_{\alpha_2}, s_{\alpha_3}$
such that $s_{\alpha_1}$ and $s_{\alpha_{3}}$ commute.
\begin{itemize}
\item The element $w = s_{\alpha_{1}}s_{\alpha_{2}}$ has support $S(s_{\alpha_{1}}s_{\alpha_{2}}) =
\{s_{\alpha_{1}},s_{\alpha_{2}}\}$ and its unique parabolic upper bound in the Bruhat order is
$w_{\{\alpha_1,\alpha_2\}} = s_{\alpha_{1}}s_{\alpha_{2}}s_{\alpha_{1}}$. On the other hand, $w$ has
maximal lower bounds
$w_{\{\alpha_1\}} = s_{\alpha_{1}}$ and $w_{\{\alpha_2\}} = s_{\alpha_{2}}$ which are incomparable.
\item The element $w = s_{\alpha_{1}}s_{\alpha_{2}}s_{\alpha_{3}}$ has support
$S(s_{\alpha_{1}}s_{\alpha_{2}}s_{\alpha_{3}}) = \Delta$ so its unique upper bound is
$w_0 = s_{\alpha_1}s_{\alpha_2}s_{\alpha_3}s_{\alpha_1}s_{\alpha_2}s_{\alpha_3}$.
Moreover, $w$ has a unique maximal lower bound given by $w_{\{\alpha_1,\alpha_3\}} =
s_{\alpha_1}s_{\alpha_3}$. The set of all
parabolic elements bounded above by $w$ is
$\{ e, s_{\alpha_{1}}, s_{\alpha_{2}}, s_{\alpha_{3}}, s_{\alpha_{1}}s_{\alpha_{3}} \}$.
\end{itemize}
\end{example}
As an application, the explicit description of supports given by Corollary \ref{cor:w-is-wI} implies the
following explicit upper and lower bounds on the $G$-saturated support variety of a Demazure module.
\begin{prop}
\label{prop:support-lower-bound}
Let $v \in W$ and $\lambda \in X(T)_+$ then,
\begin{equation*}
G \cdot \V_{B_1}(H^0(w_{S(v)}, \lambda ) ) \subseteq G \cdot \V_{B_1}(H^0(v, \lambda ) ) \subseteq \bigcap\limits_{w_I \le v} G \cdot \V_{B_1}(H^0(w_I,\lambda))
\end{equation*}
where the intersection may be taken over the set of $w_I \le v$ which are
maximal with respect to that property.
\end{prop}
Recall that the varieties of the form $\V_{B_1}(H^0(w_I,\lambda))$ are explicitly determined in Corollary \ref{cor:w-is-wI}.
\section{Support varieties of Demazure modules for the root system $A_2$}
\label{sec:A2}
In this section we present explicit calculations of the support
varieties for Demazure modules when the group $G$ has a root system
of type $A_2$. We proceed by applying our results from Section
\ref{sec:calc} in the case when the prime $p$ is good. For type $A_2$
this means that $p \ge 3$. We return to the case when $p=2$ in
Subsection \ref{subsec:A2-p2}.
\subsection{(Type $A_2$, $p \ge 3$)}
\label{subsec:A2}
Let $G = SL(3)$ with $p\geq 3$, and $\lambda = (\lambda_1, \lambda_2)$
be a dominant integral weight expressed in terms of the fundamental
weights. Let us identify $\Delta = \{ \alpha, \beta \}$ and
$W = \{ e, s_\alpha, s_\beta, s_\alpha s_\beta, s_\beta s_\alpha,
s_\alpha s_\beta s_\alpha \}$. The cases where $\ell(w) \ne 2$,
i.e., $w \in \{ e, s_\alpha, s_\beta, s_\alpha s_\beta s_\alpha \}$,
are covered by Corollary \ref{cor:w-is-wI}. For such a $w$, set $\V =
\V_{B_1}( H^0(w, \lambda ) )$. We summarize in Table \ref{tab:A2} below.
\begin{table}[ht]
\centering
\begin{tabular}{c|c|l}
$w$ & $\V_{B_1}(H^0(w,\lambda )$ & $\lambda$ \\[2pt]
\hline
$e$ & ${\mathfrak u}$ & \text{all} $\lambda$ \\[2pt]
$s_\alpha$ & ${\mathfrak u}_\alpha$ &
$p \mid \lambda_1 + 1$ \\[2pt]
& ${\mathfrak u}$ & $p \nmid \lambda_1 + 1$ \\[2pt]
$s_\beta$ & ${\mathfrak u}_\beta$ & $p \mid \lambda_2 + 1$ \\[2pt]
& ${\mathfrak u}$ & $p \nmid \lambda_2 + 1$ \\[2pt]
$s_\alpha s_\beta s_\alpha$ & $\V_{G_1}(H^0(\lambda)) \cap {\mathfrak u}$ &
all $\lambda$ \\[2pt]
\end{tabular}
\vspace{0.2cm}
\caption{$B_1$-support varieties for $A_2$ when $\ell(w) \ne 2$, $p\geq 3$}
\label{tab:A2}
\end{table}
In the $w = s_\alpha s_\beta s_\alpha$ case, $J \subset \Delta$
depends on $\lambda$ and $p$ as in the discussion before Corollary
\ref{cor:w-is-wI}.
For the cases where $\ell(w) = 2$, we analyze the regularity of $\lambda$ with respect to the prime
$p$ and $p$-divisibility of the dimension of $\hwl$. We treat the case $w = s_\alpha s_\beta$, the other case
being symmetric upon switching $\alpha$, $\beta$ and $\lambda_1, \lambda_2$. For
convenience, denote by $M(\lambda) = M(\lambda_1, \lambda_2)$ the $B$-module
$H^0(s_\alpha s_\beta, \lambda)$ which we also identify with
$\ind_B^{P_\alpha} \ind_B^{P_\beta} \lambda$.
In the root system of type $A_2$, a weight $\lambda$ is $p$-regular if
and only if
\begin{equation*}
\label{eq:A2-regularity}
\tag{A}
\left\{ \begin{array}{l}
p \nmid \lambda_1 + 1, \\
p \nmid \lambda_2 + 1, \\
p \nmid \lambda_1 + \lambda_2 + 2 .\end{array} \right.
\end{equation*}
We may apply the Demazure character formula (\cite{A}) in this situation to conclude that $\dim M(\lambda) = \frac{(\lambda_2+1)(2 \lambda_1 + \lambda_2 + 2)}{2}$. Thus $p$ does not divide $\dim M(\lambda)$ if and only if
\begin{equation*}
\label{eq:A2-dimension}
\tag{B}
\left\{
\begin{array}{l}
p \nmid \lambda_2+1, \\
p \nmid 2 \lambda_1 + \lambda_2 + 2 .
\end{array} \right.
\end{equation*}
\begin{thm}
\label{thm:A2}
Let $p\geq 3$. The $B_1$-support variety $\V = \V_{B_1}(M(\lambda))$ is
${\mathfrak u}$ if either \eqref{eq:A2-regularity} or \eqref{eq:A2-dimension} hold. Otherwise $\V$ is a proper subvariety of ${\mathfrak u}$ given by the conditions below:
\begin{equation}
\label{eq:A2-length2-calculation}
\V_{B_1}(M(\lambda)) = \left\{
\begin{array}{ll}
{\mathfrak u}_\alpha, & \text{if}\,\, \lambda = (np - 1, 0) \quad (n \ge 1), \\[3pt]
{\mathfrak u}_\alpha \cup {\mathfrak u}_\beta, & \text{if}\,\, \lambda_2 \ne 0 \,\text{and neither
\eqref{eq:A2-regularity} nor \eqref{eq:A2-dimension} hold.} \\[3pt]
\end{array} \right.
\end{equation}
\end{thm}
The rest of the section is devoted to proving Theorem \ref{thm:A2}. First, if either
\eqref{eq:A2-regularity} or \eqref{eq:A2-dimension} holds,
we conclude that $\V = {\mathfrak u}$ by Proposition \ref{prop:regular-case} or the rank variety
description of $\V$, respectively. For the rest of the section we assume that neither \eqref{eq:A2-regularity}
nor \eqref{eq:A2-dimension} holds and calculate $\V$ which will turn out to be a proper
subvariety of ${\mathfrak u}$.
Our analysis uses the $B$-stability of support varieties in a crucial way, in
particular the action of positive root subgroups and certain one-parameter
groups in the maximal torus. The nilradical $\mathfrak{u}$ is spanned by root
spaces $\mathfrak{u} = k X_\alpha \oplus k X_\beta \oplus k X_{\alpha+\beta}$.
There is a one-parameter subgroup $k^* \subset T \subset B$ such that \[
t.X_\gamma = t^{\operatorname{ht}(\gamma)} X_\gamma, \] for all $t \in k^*$
and $\gamma \in \Phi$. This group is generated by the element usually denoted
by $H_\rho$, where $\rho$ is the half sum of positive roots ($\rho = \alpha +
\beta$ in this case).
As a preliminary step we classify the $(B,k^*)$-stable subvarieties of
$\mathfrak{u}$. Let $v = a X_\alpha + b X_\beta + c X_{\alpha+\beta}$ be an
arbitrary point of $X \subset \mathfrak{u}$ an irreducible $B$-stable
subvariety of $\mathfrak{u}$. Here, $\rk v$ denotes the rank of a matrix
representative for $v$. The claim is that $X$ is equal to one of the following
($B$-stable) subspaces: $\mathfrak{u}$, $\overline{B \cdot X_\alpha} = k
X_\alpha \oplus k X_{\alpha+\beta}$, $\overline{B \cdot X_\beta} = k X_\beta
\oplus k X_{\alpha+\beta}$, and $\overline{B \cdot X_{\alpha+\beta}} = k
X_{\alpha+\beta}$, or $\{0\}$. There are five mutually exclusive cases:
\begin{enumerate}
\item Suppose that $a,b \ne 0$. Then $\rk v = 2$ and the
$B$-orbit through $v$ is dense in $\mathfrak{u}$. Thus $X = \mathfrak{u}$.
\item Suppose that $a \ne 0, b = 0$. Using the action of $k^*$, we see that
the element $v' = a X_\alpha$ is in the closure of $B \cdot v$. Hence,
$X_\alpha \in X$ and so $\overline{B \cdot X_\alpha} \subset X$.
\item Suppose
that $a = 0, b \ne 0$. Then as in the previous case we conclude that
$\overline{B \cdot X_\beta} \subset X$.
\item Suppose that $a, b = 0$ and $c
\ne 0$. In this case $\overline{B \cdot X_{\alpha + \beta}} \subset X$.
\item Suppose that $a,b,c = 0$. Then, $v = 0$ and $\{0\} \subset X$.
\end{enumerate}
Therefore, every $B$-stable, irreducible subvariety $X
\subset {\mathfrak u}$ is a union of the five subspaces above, thus it must
equal one of them.
Now we treat a number of cases depending on $\lambda$ and $p$ to determine
which root vectors are in the support variety. By the analysis of the previous
paragraph, this suffices to determine the variety as a union of $B$-stable
subvarieties.
First, suppose $\lambda_2 = 0$. In this case, $\ind_B^{P_\beta} \lambda \cong \lambda$ as a $B$-module and so $M(\lambda_1, 0) \cong \ind_B^{P_\alpha} (\lambda_1, 0)$. Thus by the $\ell(w) = 1$ calculation in Table \ref{tab:A2},
\[ \V_{B_1}(M(\lambda_1,0)) = \left\{ \begin{array}{ll}
{\mathfrak u}_\alpha, & \text{if}\,\, p \mid \lambda_1 + 1 \\[3pt]
{\mathfrak u}, & \text{if}\,\, p \nmid \lambda_1+1 .\\[3pt]
\end{array} \right.
\]
This proves the first part of \eqref{eq:A2-length2-calculation}.
Note that if $p$, $\lambda$ are such that $p \nmid \lambda_1+1$ and $\lambda_2 = 0$,
then they satisfy both \eqref{eq:A2-regularity} and \eqref{eq:A2-dimension}.
Now suppose $\lambda_2 \ne 0$. $M(\lambda)$ is induced from $H^0(s_\beta, \lambda)$ as a
$P_\alpha$-module and as an
$L_\alpha$-module we have: \[ M(\lambda)\lvert_{L_\alpha} \cong \bigoplus_{i=1}^{\lambda_2+1}
\ind^{L_\alpha}_{B \cap L_\alpha} (\lambda_1+i) ,\] where the right hand side
is a direct sum of irreducible $L_\alpha$-modules indexed by the integers
$\lambda_1+1, \cdots, \lambda_1 + \lambda_2 + 2$. The assumption that $\lambda_2
\ne 0$ implies that $M(\lambda)$ has at least two $L_\alpha$ summands whose dimensions
differ by $1$ and so it cannot be projective over $\langle X_\alpha \rangle$. Therefore,
$X_\alpha \in \V$ and by the analysis above of the $B$-stable, conical subvarieties of ${\mathfrak u}$
we get $\overline{B \cdot X_\alpha}
\subset \V$. Thus $X_{\alpha+\beta} \in \overline{B \cdot X_\alpha} \subset
\V$. Using the fact that $\V$ is $P_\alpha$-stable we can conclude that
$X_\beta \in \overline{P_\alpha \cdot X_{\alpha + \beta}} \subset \V$. Hence,
independent of $p$ we have:
\begin{equation*}
\overline{B \cdot X_\alpha} \cup \overline{B
\cdot X_\beta} = {\mathfrak u}_\alpha \cup {\mathfrak u}_\beta \subseteq \V.
\end{equation*}
Suppose that $p \mid(\lambda_2+1)$. In this case, $\lambda_2$ is a Steinberg weight for $L_\beta$
and we have $\V_{B_1}(H^0(s_\beta, \lambda )) = {\mathfrak u}_\beta$. By Theorem \ref{thm:fp-bend-generalization},
\[ \V_{(P_\alpha)_1}(M(\lambda)) \subseteq P_\alpha \cdot \V_{B_1}(H^0(s_\beta,
\lambda )) = P_\alpha \cdot {\mathfrak u}_\beta .\] Now,
observe that the right hand side is contained in the subvariety \[ R_1 :=
\left\{ v \in \mathfrak{g} \mid \rk v \le 1 \right\} .\] Since $\rk (X_\alpha
+ X_\beta) = 2$ we have that $(X_\alpha + X_\beta) \notin \V$ so $\V$ is a
proper subvariety of $\mathfrak{u}$. We conclude in this case that \[ \V =
{\mathfrak u}_\alpha \cup {\mathfrak u}_\beta .\]
\subsection{Properness of Supports} We continue the proof of Theorem \ref{thm:A2} with all assumptions from Subsection \ref{subsec:A2}; in particular, $w=s_\alpha s_\beta$.
Now assuming that $\lambda_2 \ne 0$ and $p \nmid \lambda_2+1$, we reduce to
two families of modules which also satisfy neither \eqref{eq:A2-regularity} nor \eqref{eq:A2-dimension}.
\begin{lem}
\label{lem:type-i-ii}
If $\lambda = (\lambda_1, \lambda_2)$ satisfy $\lambda_2 \ne 0$, $p \nmid \lambda_2+1$, and neither \eqref{eq:A2-regularity} nor \eqref{eq:A2-dimension}, then either
\begin{equation*}
\label{eq:type1}
\tag{i}
\lambda_1 \equiv -1 \pmod{p}, \quad \lambda_2 \equiv 0 \pmod{p}
\end{equation*}
or
\begin{equation*}
\label{eq:type2}
\tag{ii}
\lambda_1 \equiv 0 \pmod{p}, \quad \lambda_2 \equiv -2 \pmod{p}
\end{equation*}
\end{lem}
\begin{proof}
First, if $\lambda$, $p$ violate condition \eqref{eq:A2-dimension} and $p \nmid
\lambda_2+1$, then $p \mid 2\lambda_1+\lambda_2+2$. Since the pair also violates \eqref
{eq:A2-regularity}, there are two possibilities.
\begin{enumerate}
\item Suppose that $p \mid \lambda_1 + 1$. Then $p \mid (2\lambda_1+\lambda_2+2) = 2
(\lambda_1+1) + \lambda_2$ if and only if $p \mid \lambda_2$. This is case \eqref{eq:type1}
above.
\item Suppose that $p \mid \lambda_1 + \lambda_2 + 2$. Then $p \mid (2\lambda_1+
\lambda_2+2) = (\lambda_1+\lambda_2+2) + \lambda_1$ if and only if $p \mid \lambda_1$.
Hence, $p \mid \lambda_2+2$. These two conditions are equivalent to case \eqref{eq:type2}
above.
\end{enumerate}
\end{proof}
To complete the proof, we make two reductions. First, we prove that if the support variety
for all modules of type \eqref{eq:type1} in Lemma~\ref{lem:type-i-ii} are proper then the support varieties for all
modules of type \eqref{eq:type2} are proper, and vice versa. Next, we show by induction
that it suffices to prove the properness of the support variety for modules of the form
$M(np, p-2)$ for $n \ge 0$. These are modules of type \eqref{eq:type2}. Finally, we analyze
the support of $M(np, p-2)$ using filtrations on the tensor product $M(np, mp-2) \otimes L
(0,1)^{(1)}$ where $L(0,1)^{(1)}$ denotes the $G$-module $L(0,1)$ (with highest weight $\mu
= (0,1)$) twisted once by the Frobenius morphism.
\begin{lem}
\label{lem:typei-typeii}
The support $\V_{B_1}(M(\lambda))$ is a proper subvariety of ${\mathfrak u}$ (and hence equal to ${\mathfrak u}_\alpha
\cup {\mathfrak u}_\beta$) for all $\lambda$ of type (i) if and only if the same holds for all $\lambda$ of type (ii).
\end{lem}
\begin{proof}
Consider the $B$-module $M(np, mp-1)$ for some $n \ge 0$, $m > 0$. This module has proper support by the argument given for the case $p \mid \lambda_2+1$. Let $L(1,0)$ denote the irreducible $G$-module with highest weight $(1,0)$ and consider the tensor product $M(np, mp-1) \otimes L(1,0)$. The $G$-module structure on $L(1,0)$ allows to use the tensor identity (\cite[I.4.8]{Jan}) to identify
\begin{align*}
M(np, mp-1) \otimes L(1,0) & = \left[ \ind_B^{P_\alpha} \ind_B^{P_\beta} (np, mp-1) \right] \otimes L(1,0) \\
& \cong \ind_B^{P_\alpha} \ind_B^{P_\beta} \left[
(np, mp-1) \otimes L(1,0)_{\vphantom{B}}^{\vphantom{P_\alpha}} \right] .
\end{align*}
Now, $L(1,0)$ has a filtration as a $B$-module as follows
\[
\begin{array}{ccc}
L(1,0) = \left[
\vcenter{
\def\objectstyle{\scriptstyle}
\def\labelstyle{\scriptstyle}
\xymatrix{
(1,0) \ar@{-}[d] \\
(-1,1) \ar@{-}[d] \\
(0,-1)
}} \right. &
\text{which induces} &
(np,mp-1) \otimes L(1,0) = \left[
\vcenter{
\def\objectstyle{\scriptstyle}
\def\labelstyle{\scriptstyle}
\xymatrix{
(np+1,mp-1) \ar@{-}[d] \\
(np-1,mp) \ar@{-}[d] \\
(np,mp-2)
}} \right. .
\end{array}
\]
Let $F(\cdot)$ denote the functor $\ind_{B}^{P_\alpha} \ind_B^{P_\beta} (\cdot)$. Since
the weights in the filtration for \mbox{$(np,mp-1) \otimes L(1,0)$} are all dominant, Kempf's
vanishing theorem implies that $R^1F(\cdot)$ vanishes on each of the subquotients
(\cite[I.4.4]{Jan}). Thus there is an induced filtration:
\[ M(np,mp-1) \otimes L(1,0) = \left[
\vcenter{
\def\objectstyle{\scriptstyle}
\def\labelstyle{\scriptstyle}
\xymatrix{
M(np+1,mp-1) \ar@{-}[d] \\
M(np-1,mp) \ar@{-}[d] \\
M(np,mp-2)
}} \right. .\]
Let $N$ denote the quotient $( M(np,mp-1) \otimes L(1,0) ) / M(np,mp-2)$. We have an
exact sequence
\[ 0 \to M(np, mp-2) \to M(np,mp-1) \otimes L(1,0) \to N \to 0 .\]
The support variety of the middle term $M(np,mp-1) \otimes L(1,0)$ is proper. Now, $N$
sits in an exact sequence:
\[ 0 \to M(np-1, mp) \to N \to M(np+1,mp-1) \to 0 .\]
The support variety of the last term in this sequence, $M(np+1,mp-1)$, is proper by the $p \mid \lambda_2+1$ case.
Thus if the support of $M(np, mp-2)$ is proper, then the same hold for $N$
and hence by the second sequence, the same holds for $M(np-1, mp)$. On the other hand,
if the support of $M(np-1, mp)$ is proper the second sequence
implies that the same holds for $N$ and thus, by the first sequence, the same holds for $M(np, mp-2)$
(cf. \cite[(2.2.7)]{NPV} for properties of support varieties and exact sequences).
\end{proof}
\begin{lem}
\label{lem:reduce-to-base}
The support variety $\V_{B_1}(M(np, mp-2))$ is proper for all $n \ge 0$, $m > 0$ if $
\V_{B_1}(M(np, p-2))$ is proper for all $n \ge 0$.
\end{lem}
\begin{proof}
The results follow by induction on $m$. Suppose that $\V_{B_1}(M(np, kp-2))$ is proper
for all $n \ge 0$ and all $0 \le k \le m$. We prove that $\V_{B_1}(M(np,(m+1)p-2))$ is
proper. Consider the tensor product
$M(np, mp-2) \otimes L(0,1)^{(1)}$. As in Lemma
\ref{lem:typei-typeii}, we use the tensor identity and a filtration on $L(0,1)^{(1)}$.
We have
\[
\begin{array}{ccc}
L(0,1)^{(1)} = \left[
\vcenter{
\def\objectstyle{\scriptstyle}
\def\labelstyle{\scriptstyle}
\xymatrix{
(0,p) \ar@{-}[d] \\
(p,-p) \ar@{-}[d] \\
(-p,0)
}} \right. &
\Longrightarrow &
M(np,mp-2) \otimes L(0,1)^{(1)} = \left[
\vcenter{
\def\objectstyle{\scriptstyle}
\def\labelstyle{\scriptstyle}
\xymatrix{
M(np,(m+1)p-2) \ar@{-}[d] \\
M((n+1)p-1,(m-1)p-2) \ar@{-}[d] \\
M((n-1)p,mp-2)
}} \right.
\end{array} .
\]
Let $N$ be the submodule such that $(M(np,mp-2) \otimes L(0,1)^{(1)}) / N
\cong M(np,(m+1)p-2)$. The filtration on $N$ has subquotients whose
supports are proper by the induction hypothesis, hence $N$ has proper
support. It follows that $M(np,(m+1)p-2)$ has proper support.
\end{proof}
Finally, we prove that modules of the form $M(np, p-2)$ have proper
support. This will finish off the calculation for $l(w)=2$ when $\Phi=A_{2}$ with $p\geq 3$.
\begin{lem}
\label{lem:final-reduction}
The support variety $\V_{B_1}(M(np, p-2))$ is proper, hence
\[ \V_{B_1}(M(np, p-2)) = {\mathfrak u}_\alpha \cup {\mathfrak u}_\beta .\]
\end{lem}
\begin{proof}
We argue by induction on $n$. The base case is $M(0,p-2)$. This module has proper support by Corollary \ref{thm:w-is-w0J} since $\Phi_{\lambda, p} =
\{ \alpha + \beta \}$. Assume that $M(kp,p-2)$ has proper support for
all $0 \le k \le n$. We show that $M((n+1)p,p-2)$ has proper support.
As in the previous two lemmas, we consider a tensor product, in this case
$M(np,p-2) \otimes L(1,0)^{(1)}$. The filtration on $(np,p-2) \otimes L(1,0)^{(1)}$ now
has socle the 1-dimensional $B$-module $(np,-2)$ which is not a dominant weight so we
are forced to consider the higher derived functors $R^iF$, $i > 0$.
The $G$-module $L(1,0)^{(1)}$ has a $B$-filtration with sections of the form:
\[
L(1,0)^{(1)} = \left[
\vcenter{
\def\objectstyle{\scriptstyle}
\def\labelstyle{\scriptstyle}
\xymatrix{
(p,0) \ar@{-}[d] \\
(-p,p) \ar@{-}[d] \\
(0,-p)
}} \right. .
\]
Tensoring with $(np, p-2)$ gives an exact sequence of $B$-modules:
\[ 0 \to (np, -2) \to L(1,0)^{(1)} \otimes (np, p-2) \to
\left[ \vcenter{
\def\objectstyle{\scriptstyle}
\def\labelstyle{\scriptstyle}
\xymatrix{ (p,0) \ar@{-}[d] \\ (-p, p) }}\right. \otimes (np, p-2) \to 0 .\]
Applying the induction functor $F(\cdot)$ we have a long exact sequence in cohomology:
\begin{equation}
\label{eq:les1}
\xymatrix{
0 \ar[r] & F(np,-2) \ar[r] & L(1,0)^{(1)} \otimes F(np, p-2) \ar[r] &
F\left(
\left[
\def\objectstyle{\scriptstyle}
\def\labelstyle{\scriptstyle}
\xy
{\ar@{-} (0,5)*{(p,0)}; (0,-5)*{(-p,p)}};
\endxy
\right. \otimes (np, p-2)
\right) \\
\ar[r] & \ar[r] R^1F(np, -2) \ar[r] & 0 .& }
\end{equation}
The first term, $F((np,-2))$, vanishes since $(np,-2)$ is not $\beta$-dominant, so \eqref{eq:les1} is a
short exact sequence. Also note that
the second term has proper support by the induction hypothesis. Now, we claim that
the module $R^1F((np,-2))$ has proper support.
Recall that $F(\cdot) = \ind_B^{P_\alpha} \circ \ind_B^{P_\beta}( \cdot )$. Consider the
spectral sequence:
$$E_{2}^{i,j}=R^{i}\text{ind}_{B}^{P_{\alpha}}R^{j}\text{ind}_{B}^{P_{\beta}}(np,-2)\Rightarrow
R^{i+j}F(np,-2).$$
Set $E_1 = R^1F((np,-2))$. The spectral sequence yields a five term exact sequence of the form:
\[
\xymatrix{
0 \ar[r] & R^1\ind_B^{P_\alpha} \left( \ind_B^{P_\beta}( (np, -2) ) \right) \ar[r] &
E_1 & \\
\ar[r] & \ind_B^{P_\alpha}\left( R^1\ind_B^{P_\beta}( (np, -2) ) \right) \ar[r] &
R^2\ind_B^{P_\alpha}\left( \ind_B^{P_\beta}( (np,-2) ) \right) \ar[r] & \cdots .
}
\]
Since $\ind_B^{P_\beta}( (np,-2) ) = 0$, the first and last term vanish so we have
$$E_{1} \cong \ind_B^{P_\alpha}\left( R^1\ind_B^{P_\beta}( (np, -2) ) \right).$$
By Serre Duality (\cite[Prop. 5.2(c)]{Jan}), $R^1\ind_B^{P_\beta}( (np, -2) ) \cong (np-1,0)$. Consequently,
from the $\ell(w) = 1$ case we can conclude that $E_1 \cong \ind_B^{P_\alpha}( (np-1,0) )$ has proper support.
Now \eqref{eq:les1} implies that the module
\[
F\left(
\left[
\def\objectstyle{\scriptstyle}
\def\labelstyle{\scriptstyle}
\xy
{\ar@{-} (0,5)*{(p,0)}; (0,-5)*{(-p,p)}};
\endxy
\right. \otimes (np, p-2)
\right)
\]
has proper support. We have an exact sequence:
\[
0 \to F((n-1)p,2p-2) \to F\left( \left[
\def\objectstyle{\scriptstyle}
\def\labelstyle{\scriptstyle}
\xy
{\ar@{-} (0,5)*{(p,0)}; (0,-5)*{(-p,p)}};
\endxy
\right. \otimes (np, p-2) \right)
\to F((n+1)p, p-2) \to 0 .\]
The middle term has proper support and we to show that the last term has proper support. Thus it suffices to show that $F( ((n-1)p,2p-2) )$ has proper support. This is the other base case in our double induction.
We argue as in Lemma \ref{lem:reduce-to-base} with $n$ replaced by $n-1$ and $m=1$. Now, the $B$-filtration on $((n-1)p,p-2) \otimes L(0,1)^{(1)}$
has a non-dominant weight in the middle layer $((n+1)p,-2)$. Let $N$ denote the quotient $N := (((n-1)p,p-2) \otimes L(0,1)^{(1)})/((n-2)p,p-2)$ so that $N$ has socle consisting of $((n+1)p, -2)$. We have an exact sequence
\begin{equation}
\label{eq:FN}
0 \to F( (n-2)p, p-2 ) \to F\left[ ((n-1)p,p-2) \otimes L(0,1)^{(1)}
\right] \to F(N) \to 0
\end{equation}
with first and middle terms having proper support, thus $F(N)$ has proper support. Furthermore, $F(N)$ sits in a sequence
\[ 0 \to ((n+1)p, -2) \to N \to ((n-1)p, 2p-2) \to 0 \]
and applying $F(\cdot)$ we have
\[
\xymatrix{
0 \ar[r] & F((n+1)p, -2) \ar[r] & F(N) \ar[r] & F((n-1)p, 2p-2) \\
\ar[r] & R^1F((n+1)p, -2) \ar[r] & R^1F(N) \ar[r] & 0 .}
\]
The term $F((n+1)p, -2)$ vanishes and so does $R^1F(N)$ by extending
the sequence \eqref{eq:FN}. As before, we identify $R^1F((n+1)p,-2) \cong
\ind_B^{P_\alpha}((n+1)p-1,0)$ which has proper support. Thus the term
$F((n-1)p, 2p-2)$ has proper support and the proof is concluded.
\end{proof}
\vskip 1cm
\subsection{(Type $A_{2}$, $p=2$)}
\label{subsec:A2-p2}
When $p=2$, ${\mathcal N}_{1}({\mathfrak u})={\mathfrak u}_{\alpha}\cup {\mathfrak u}_{\beta}$. One can
apply results from Sections 4 and 5 to give explicit descriptions of the $B_{1}$-supports of
$H^{0}(w,\lambda)$ when $l(w)\neq 2$. We summarize the results in Table \ref{tab:A2p=2}.
\begin{table}[!h]
\centering
\begin{tabular}{c|c|l}
$w$ & $\V_{B_1}(H^0(w,\lambda ))$ & $\lambda$ \\[2pt]
\hline
$e$ & ${\mathfrak u}_{\alpha}\cup {\mathfrak u}_{\beta}$ & \text{all} $\lambda$ \\[2pt]
$s_\alpha$ & ${\mathfrak u}_\alpha$ &
$p \mid \lambda_1 + 1$ \\[2pt]
& ${\mathfrak u}_{\alpha}\cup {\mathfrak u}_{\beta}$ & $p \nmid \lambda_1 + 1$ \\[2pt]
$s_\beta$ & ${\mathfrak u}_\beta$ & $p \mid \lambda_2 + 1$ \\[2pt]
& ${\mathfrak u}_{\alpha}\cup {\mathfrak u}_{\beta}$ & $p \nmid \lambda_2 + 1$ \\[2pt]
$s_\alpha s_\beta s_\alpha$ & $\V_{G_1}(H^0(\lambda)) \cap ({\mathfrak u}_{\alpha}\cup {\mathfrak u}_{\beta})$ &
all $\lambda$ \\[2pt]
\end{tabular}
\vspace{0.2cm}
\caption{$B_1$-support varieties for $A_2$ when $\ell(w) \ne 2$, $p=2$.}
\label{tab:A2p=2}
\end{table}
In the $l(w)=2$ case it suffices to consider (by symmetry) $w=s_{\alpha}s_{\beta}$. We note that
there are no $p$-regular weights and $\dim M(\lambda)$ is always divisible by 2. Moreover,
we don't need to show properness because all $B_{1}$-support varieties are already contained in
${\mathfrak u}_{\alpha}\cup {\mathfrak u}_{\beta}$. The following result summarizes the $l(w)=2$ case.
\begin{thm} Let $w=s_{\alpha}s_{\beta}$. The $B_1$-support variety $\V = \V_{B_1}(M(\lambda))$ is given by:
\begin{equation}
\label{eq:A2-length2-calculationp=3}
\V_{B_1}(M(\lambda)) = \left\{
\begin{array}{ll}
{\mathfrak u}_\alpha, & \text{if}\,\, \lambda = (2n- 1, 0) \quad (n \ge 1), \\[3pt]
{\mathfrak u}_\alpha \cup {\mathfrak u}_\beta, & \text{if}\, \, \lambda=(2n,0) \quad (n \geq 0), \\[3pt]
{\mathfrak u}_\alpha \cup {\mathfrak u}_\beta, & \text{if}\,\, \lambda_2 \ne 0 .\\[3pt]
\end{array} \right.
\end{equation}
\end{thm}
|
\section{Introduction}
Schr\"{o}dinger had pointed out long time ago \cite{intro1} that quantum entanglement is a crucial element of quantum mechanics. In fact quantum entanglement is one of the most peculiar feature that distinguishes quantum physics from classical physics and lies at the heart of what is now quantum information theory. Quantum physics allows correlations between spatially separated systems that are fundamentally different from classical correlations, and this difference becomes evident when entangled states violate Bell-type inequalities that place an upper bound on the correlations compatible with local hidden variable (or local realistic) theories \cite{intro2}. In the last two decades research has been very focused on quantum entanglement because the field of quantum information theory (cf. \cite{intro3,intro4}) has developed rather quickly to be an important one.
It is very important to get a good understanding of entanglement properties of the quantum states, under effects of accelerations (Lorentz transformations) and magnetic fields (constant and homogeneous), e.g., in studying quantum entangled states of the particles that are produced and detected in Stern-Gerlach experiments \cite{key4}. The issues of quantum entanglement in (constant) external magnetic field were addressed in a previous paper \cite{key1}, through the approach of Wigner rotations of canonical spin. Considering frames of observers related through Lorentz transformations, it was shown that the entanglement is frame \textit{independent} but the violation of Bell's inequality is frame \textit{dependent}. Similar features and correlations (or degrees of entanglement) were noted for spins undergoing precession in a magnetic field in that study. However, the study of quantum entanglement in magnetic fields was limited to 2-particle states of total spin zero. Here we study entangled 3-particle states in constant external magnetic fields and display the emergence of remarkable \textit{periodic} correlations and density matrices. Note that consequent features of the 2-particle sub-systems can also be extracted from them. To our knowledge, such a study of periodicities in correlations and density matrices is being presented and analysed for the first time. We then introduce an electric field orthogonal to the magnetic field, and obtain the corresponding Wigner rotations of the spin states. In this paper, we also propose a new scheme of systematic classification of the entangled states corresponding to well-known ones in the study of 3-particle entanglements (viz. Greenberger-Horne-Zeilinger $\left| GHZ \right \rangle$, Werner $\left| W \right \rangle$, flipped Werner $\left| \widetilde{W} \right \rangle$ and their variant states), using a particular coupling of three angular momenta.
The plan of the paper is as follows: We introduce the notations and formalism in Section II. In Sections III and IV, the emerging patterns of periodicity in correlations and density matrices, respectively, are studied explicitly for different entangled states, and particularly for certain initial configuration of the velocities. In Section V, we study the case with an electric field orthogonal to the magnetic field, linking to the preceding case via Lorentz transformation, and obtain the corresponding Wigner rotations of the spin states. In Section VI, we demonstrate the proposed scheme of systematic classification of the entangled states corresponding to well-known ones in the study of 3-particle entanglements, using the particular coupling of three angular momenta. Finally, we make some concluding remarks and give an outlook of future directions of work in Section VII.
\section{Spin-1/2 particles in constant magnetic field: Formalism}
In this Section, we describe the formalism and notations:
\begin{enumerate}
\item The unitary transformation matrices acting on the spin states of three particles of spin-1/2 and positive rest mass will be constructed for arbitrary initial velocities $(\overrightarrow{v_{1}},\overrightarrow{v_{2}},\overrightarrow{v_{3}})$ of the particles $(1,2,3)$ respectively and a constant magnetic field $\overrightarrow{B}$, and the basic precession equations (see Ref. \cite{key1}, and review articles \cite{key2,key3}) will be used
and generalized.
\item A particularly simple configuration of the initial velocities is then selected for detailed study (later in Section III). This would permit us to display the contents of the generalized equations, for a few selected cases of particular interest without obscuring the basic features due to a profusion of parameters.
This will be achieved by restricting the initial velocities to a plane orthogonal to the magnetic field $\overrightarrow{B}$ and assuming them to be of equal magnitude, namely by imposing
\begin{equation}
\overrightarrow{B}.\overrightarrow{v_{i}}=0, \qquad \left| \overrightarrow{v_{i}} \right| = v \qquad (i=1,2,3).
\label{eq1}
\end{equation}
Since $\overrightarrow{B}.\overrightarrow{v}$ and $\left| \overrightarrow{v} \right|$ are constants of motion, the conditions \ref{eq1} will hold for all time $t$.
Eight initial entangled 3-particle states would be selected for detailed study in such a context. They are encoded using the following notations (for our three spin-1/2 particles)
\begin{equation}
\left| ijk \right \rangle \equiv \left| i \right \rangle \otimes \left| j \right \rangle \otimes \left| k \right \rangle ,
\label{eq2}
\end{equation}
where
\begin{equation}
\left| i \right \rangle \subset \left( \left| + \right \rangle, \left| - \right \rangle \right) \equiv \left( \left| 1 \right \rangle, \left| \overline{1} \right \rangle \right) ,
\label{eq3}
\end{equation}
and similarly for $\left( \left| j \right \rangle, \left| k \right \rangle \right) $.
In such notations the states at time $t=0$ are
\begin{equation}
\frac{1}{\sqrt{2}} \left( \left| 111 \right \rangle + \epsilon \left| \overline{1} \overline{1} \overline{1} \right \rangle \right) ,
\label{eq4}
\end{equation}
\begin{equation}
\frac{1}{\sqrt{3}} \left( \left| 11\overline{1} \right \rangle + \exp({-i \phi}) \left| 1\overline{1} 1 \right \rangle + \exp({i \phi}) \left| \overline{1} 11 \right \rangle \right) ,
\label{eq5}
\end{equation}
and
\begin{equation}
\frac{1}{\sqrt{3}} \left( \left| \overline{1}\overline{1}1 \right \rangle + \exp({i \phi}) \left| \overline{1} 1 \overline{1} \right \rangle + \exp({-i \phi}) \left| 1 \overline{1} \overline{1} \right \rangle \right),
\label{eq6}
\end{equation}
where $\left( \epsilon = \pm \right) $ and $\left( \phi = 0,\pm \frac{2 \pi}{3}\right)$.
Not only do these states have direct correspondence to the $\left| GHZ \right \rangle, \left| W \right \rangle, \left| \widetilde{W} \right \rangle$ states familiar in the study of 3-particle entanglements \cite{key4,key5} (Ref. \cite{key5} cites original sources), but they have also been constructed long ago \cite{key6,key7} in the study of coupling of three angular momenta involving eigenvalues of the operator
\begin{equation}
Z= \left( \overrightarrow{J_1} \times \overrightarrow{J_2} \right) . \overrightarrow{J_3},
\label{eq7}
\end{equation}
of three angular momenta $\left( \overrightarrow{J_1} , \overrightarrow{J_2} , \overrightarrow{J_3} \right)$, as is briefly explained in Section VI.
\item We will follow the time-evolution of the states \eq{eq4}-\eq{eq6} with special attention to {\it periodic oscillations} (in Section III), and also study the corresponding {\it periodic density matrices} (in Section IV), displaying many interesting features.
\item We will generalize the background field to include a constant electric field $\overrightarrow{E}$, orthogonal to the magnetic field $\overrightarrow{B}$, such that
$$\overrightarrow{E}. \overrightarrow{B}=0, \qquad \left| \overrightarrow{E} \right| < \left| \overrightarrow{B} \right|.$$
This constraint would permit us to obtain the results via an appropriate Lorentz transformation (in Section V), and adapt the results from Ref. \cite{key1} (citing original sources) for the present 3-particle case.
\end{enumerate}
First, following the lines of Ref. \cite{key1}, for a \textit{single} particle we denote $(\overrightarrow{B},\overrightarrow{v},\overrightarrow{\Sigma})$ to be respectively the constant, homogeneous magnetic field, the velocity and the polarization. With unit vectors $(\hat{B},\hat{v})$ and $c=1$, we have
\begin{equation}
\overrightarrow{B}= B. \hat{B} \qquad \overrightarrow{v}= v. \hat{v}, \qquad \gamma = \left( 1-v^{2}\right)^{1/2}.
\label{eq8}
\end{equation}
The anomalous magnetic moment is denoted by
\begin{equation}
\hat{\alpha}= (g-2)/2.
\label{eq9}
\end{equation}
We define, for mass $m$ and charge $e$,
\begin{eqnarray}
\overrightarrow{\omega} &=&\frac{eB}{m\gamma}\hat{B} \nonumber \\
\overrightarrow{\Omega} &=&\frac{\hat{\alpha} eB}{m\gamma}\left( \gamma \hat{B}-(\gamma-1) (\hat{B}.\hat{v}) \hat{v}\right).
\label{eq10}
\end{eqnarray}
The equations for $\hat{v}$ and $\overrightarrow{\Sigma}$ are
\begin{eqnarray}
\frac{\rm{d}\overrightarrow{v}}{\rm{dt}}&=& -\overrightarrow{\omega} \times \overrightarrow{v} \nonumber \\
\frac{\rm{d}\overrightarrow{\Sigma}}{\rm{dt}} &=& -(\overrightarrow{\omega}+ \overrightarrow{\Omega})\times \overrightarrow{\Sigma}.
\label{eq11}
\end{eqnarray}
The constants of motion are $(v,\hat{B}.\hat{v})$ and the moduli
\begin{eqnarray}
\omega &=& \left( \frac{eB}{m\gamma} \right) \nonumber \\
\Omega &=& \left( \frac{\hat{\alpha} eB}{m\gamma} \right) \left( \gamma^{2} -(\gamma^{2}-1) (\hat{B}.\hat{v})^{2} \right)^{1/2}
\label{eq12}
\end{eqnarray}
such that $\overrightarrow{\omega}= \omega \hat{\omega}, \overrightarrow{\Omega}= \Omega \hat{\Omega}$.
Introducing an intermediate set of rotating axes, one can finally obtain the time-dependent unitary transformation matrix acting on spin states $(\left| \frac{1}{2} \right \rangle, \left| -\frac{1}{2} \right \rangle )$ of a particle moving with velocity $\overrightarrow{v}$ as
\begin{equation}
M= \exp \left( -i \frac{\Omega t}{2} (\hat{\Omega}.\overrightarrow{\sigma})\right) \exp \left( -i \frac{\omega t}{2} (\hat{B}.\overrightarrow{\sigma})\right)
\label{eq13}
\end{equation}
where $\overrightarrow{\sigma}$ denote the Pauli matrices and the components of $\overrightarrow{\Omega}$ are
(assuming $(\hat{B}.\hat{v}) \neq \pm 1$)
\begin{eqnarray}
\Omega_1 &=& (\hat{B}.\overrightarrow{\Omega}) = \frac{\alpha eB}{m\gamma} \left( \gamma -(\gamma-1) (\hat{B}.\hat{v})^{2} \right) \nonumber \\
\Omega_2 &=& \frac{(\hat{B} \times \hat{v}). \overrightarrow{\Omega}}{\sqrt{1- (\hat{B}.\hat{v})^{2}}} = 0 \nonumber \\
\Omega_3 &=& \frac{ (\hat{B} \times (\hat{B} \times \hat{v})).\overrightarrow{\Omega}}{\sqrt{1- (\hat{B}.\hat{v})^{2}}} \nonumber \\
~ &=& \frac{\alpha eB}{m\gamma}(\gamma-1) (\hat{B}.\hat{v})\sqrt{1- (\hat{B}.\hat{v})^{2}} .
\label{eq14}
\end{eqnarray}
Below we leave aside the particularly simple case arising for $(\hat B . \hat v ) =\pm 1$.
We follow the same prescription as in Ref. \cite{key1}, and denote
\begin{eqnarray}
\hat{B}.\overrightarrow{\sigma} &=& (b_1 \sigma_1+ b_2 \sigma_2+ b_3 \sigma_3) \nonumber \\
\hat{\Omega}.\overrightarrow{\sigma} &=& (l_1 \sigma_1+ l_3 \sigma_3) ,
\label{eq15}
\end{eqnarray}
so that
\begin{eqnarray}
b_1^{2}+ b_2^{2} + b_3^{2} &=& 1 \nonumber \\
l_1^{2}+ l_3^{2} &=& 1 , \qquad l_2=0.
\label{eq15a}
\end{eqnarray}
We also introduce the notations
\begin{eqnarray}
(c, s) & \equiv & \left( \cos (\frac{\omega t}{2}), \sin (\frac{\omega t}{2})\right) \nonumber \\
(c', s') & \equiv & \left( \cos (\frac{\Omega t}{2}), \sin (\frac{\Omega t}{2})\right) ,
\label{eq16}
\end{eqnarray}
so that we can now write
\begin{eqnarray}
M \equiv \left| \begin{array}{cc}
\alpha & -i\beta \\
-i\beta^{*} & \alpha^{*}
\end{array} \right|,
\label{eq18}
\end{eqnarray}
where we define
\begin{eqnarray}
\alpha &=& (c'-il_3s')(c-ib_3s)-l_1s'(b_1+ib_2)s \nonumber \\
\beta &=& (c'-il_3s')(b_1-ib_2)s +l_1s'(c+ib_3s) ,
\label{eq19}
\end{eqnarray}
which will be the \textit{precession parameters}.
Using the notations defined above, we obtain the unitarity constraint
\begin{equation}
M^{\dagger} M \equiv (\alpha \alpha^{*}+ \beta \beta^{*} ) \left| \begin{array}{cc}
1 & 0 \\
0 & 1
\end{array} \right|
= \left| \begin{array}{cc}
1 & 0 \\
0 & 1
\end{array} \right|.
\label{eq20}
\end{equation}
The action of $M$ on the spin states
\begin{equation}
\left( \left| + \right \rangle, \left| - \right \rangle \right) \equiv \left( \left| 1 \right \rangle, \left| \overline{1} \right \rangle \right) = \left( \left| \begin{array}{c}
1 \\
0
\end{array} \right \rangle, \left| \begin{array}{c}
0 \\
1
\end{array} \right \rangle \right),
\label{eq21}
\end{equation}
is given by
\begin{equation}
M \left( \left| 1 \right \rangle, \left| \overline{1} \right \rangle \right) = \left( (\alpha \left| 1 \right \rangle - i \beta^{*} \left| \overline{1} \right \rangle ), (-i \beta \left| 1 \right \rangle + \alpha^{*} \left| \overline{1} \right \rangle )\right) .
\label{eq22}
\end{equation}
So far we have been considering a single particle with velocity $\overrightarrow{v}$. For three particles with velocities $(\overrightarrow{v_{1}},\overrightarrow{v_{2}},\overrightarrow{v_{3}})$ we denote the corresponding precession parameters as
$$(\alpha_1, \beta_1), (\alpha_2, \beta_2), (\alpha_3, \beta_3)$$
generalizing appropriately
$(\alpha, \beta)$ of \eq{eq19}.
Thus, for example, an inital state at $t=0$,
\begin{equation}
\left| A \right \rangle \equiv \frac{1}{\sqrt{2}} \left( \left| 111 \right \rangle + \left| \overline{1} \overline{1} \overline{1} \right \rangle \right) ,
\label{eq23}
\end{equation}
will evolve as
\begin{widetext}
\begin{eqnarray}
\left( M_{(1)} \otimes M_{(2)} \otimes M_{(3)} \right) \left| A \right \rangle = \frac{1}{\sqrt{2}} (\alpha_1 \left| 1 \right \rangle - i \beta_1^{*} \left| \overline{1} \right \rangle ) (\alpha_2 \left| 1 \right \rangle - i \beta_2^{*} \left| \overline{1} \right \rangle )(\alpha_3 \left| 1 \right \rangle - i \beta_3^{*} \left| \overline{1} \right \rangle ) \nonumber \\+ \frac{1}{\sqrt{2}} (-i \beta_1 \left| 1 \right \rangle + \alpha_1^{*} \left| \overline{1} \right \rangle )(-i \beta_2 \left| 1 \right \rangle + \alpha_2^{*} \left| \overline{1} \right \rangle )(-i \beta_3 \left| 1 \right \rangle + \alpha_3^{*} \left| \overline{1} \right \rangle ),
\label{eq24}
\end{eqnarray}
or
\begin{eqnarray}
M_{(1)} \otimes M_{(2)} \otimes M_{(3)} \left| A \right \rangle
&=& d_{111}\left| 111 \right \rangle + d_{1 \overline{1} \overline{1}}\left| 1 \overline{1} \overline{1} \right \rangle +d_{\overline{1} 1 \overline{1}}\left| \overline{1} 1 \overline{1} \right \rangle +d_{\overline{1} \overline{1} 1}\left| \overline{1} \overline{1} 1 \right \rangle \nonumber \\
&+&d_{\overline{1} \overline{1} \overline{1}} \left| \overline{1} \overline{1} \overline{1}\right \rangle +d_{\overline{1} 11}\left| \overline{1} 11 \right \rangle +d_{1 \overline{1} 1}\left| 1 \overline{1} 1 \right \rangle +d_{11\overline{1}}\left| 11\overline{1} \right \rangle,
\label{eq25}
\end{eqnarray}
\end{widetext}
where
\begin{eqnarray}
d_{111} = \frac{1}{\sqrt{2}} (\alpha_1 \alpha_2 \alpha_3 +i \beta_1 \beta_2 \beta_3) \nonumber \\
d_{1 \overline{1} \overline{1}} = -\frac{1}{\sqrt{2}} (\alpha_1\beta_2^{*} \beta_3^{*} +i\beta_1 \alpha_2^{*} \alpha_3^{*}) \nonumber \\
d_{\overline{1} 1 \overline{1}} = -\frac{1}{\sqrt{2}} ( \beta_1^{*}\alpha_2 \beta_3^{*}+i \alpha_1^{*} \beta_2 \alpha_3^{*}) \nonumber \\
d_{\overline{1} \overline{1} 1} = -\frac{1}{\sqrt{2}} ( \beta_1 \beta_2 \alpha_3+i\alpha_1 \alpha_2 \beta_3) \nonumber \\
d_{\overline{1} \overline{1} \overline{1}} = \frac{1}{\sqrt{2}} (\alpha_1^{*} \alpha_2^{*} \alpha_3^{*} +i \beta_1^{*} \beta_2^{*} \beta_3^{*}) \nonumber \\
d_{\overline{1} 11} = -\frac{1}{\sqrt{2}} (i\beta_1^{*} \alpha_2 \alpha_3+\alpha_1^{*} \beta_2 \beta_3) \nonumber \\
d_{1 \overline{1} 1} = -\frac{1}{\sqrt{2}} (i \alpha_1 \beta_2^{*}\alpha_3 + \beta_1 \alpha_2^{*} \beta_3) \nonumber \\
d_{11\overline{1}} = -\frac{1}{\sqrt{2}} (i\alpha_1 \alpha_2 \beta_3^{*}+ \beta_1 \beta_2 \alpha_3^{*}).
\label{eq26}
\end{eqnarray}
The coefficients above now involve \textit{three distinct periodic time dependences} through
\begin{eqnarray}
(c, s)_{i} & \equiv & \left( \cos (\frac{\omega_{i} t}{2}), \sin (\frac{\omega_{i} t}{2})\right) \nonumber \\
(c', s')_{i} & \equiv & \left( \cos (\frac{\Omega_{i} t}{2}), \sin (\frac{\Omega_{i} t}{2})\right),
\label{eq27}
\end{eqnarray}
where $(i=1,2,3)$, determined by the initial velocities $(\overrightarrow{v_{1}},\overrightarrow{v_{2}},\overrightarrow{v_{3}})$ and their orientations with respect to $\overrightarrow{B}$. Note that we ignore the mutual interactions of the particles assumed to be weak enough as compared to that with a strong magnetic field $\overrightarrow{B}$.
The probability associated to the state $\left| ijk \right \rangle $ is defined to be
\begin{equation}
P_{ijk}=d_{ijk}^*d_{ijk} .
\label{eq28}
\end{equation}
We note that in $\left| ijk \right \rangle $, the sum of the probablities of $\left| i \right \rangle $ being either $\left| 1 \right \rangle $ or $\left| \overline{1} \right \rangle $ must be $1$. Using
the relation
\begin{equation}
(\alpha_i \alpha_i^{*}+ \beta_i \beta_i^{*} )=1 \qquad (i=1,2,3)
\label{eq29}
\end{equation}
systematically, one may check that, for example,
\begin{eqnarray}
P_{11} &=& P_{111}+P_{11\overline{1}} \nonumber \\
&=& \frac{1}{2}(\alpha_1 \alpha_1^{*}\alpha_2 \alpha_2^{*}+ \beta_1 \beta_1^{*}\beta_2 \beta_2^{*} ) \\
P_{1\overline{1}} &=& \frac{1}{2}(\alpha_1 \alpha_1^{*}\beta_2 \beta_2^{*}+ \beta_1 \beta_1^{*}\alpha_2 \alpha_2^{*} ) \\
P_{1} &=& P_{11}+P_{1\overline{1}} \nonumber \\
&=& \frac{1}{2}(\alpha_1 \alpha_1^{*}+ \beta_1 \beta_1^{*})(\alpha_2 \alpha_2^{*}+\beta_2 \beta_2^{*} ) \nonumber \\
&=& \frac{1}{2}.
\label{eq30}
\end{eqnarray}
Similarly,
\begin{equation}
P_{\overline{1}} = P_{\overline{1}1}+P_{\overline{1}\overline{1}} = \frac{1}{2}.
\label{eq31}
\end{equation}
Hence
\begin{equation}
P_{1}+ P_{\overline{1}}=1 .
\label{eq32}
\end{equation}
Analogous results hold for $(j,k)$.
For the general configuration the interplay of the periods (see \eq{eq27})
\begin{equation}
\left( \frac{4\pi}{\omega_i}, \frac{4\pi}{\Omega_i} \right) \qquad (i=1,2,3)
\label{eq33}
\end{equation}
imply a \textit{rich} structure in the variation with $t$ of the coefficients $d_{ijk}$ in \eq{eq25} and the correlations \eq{eq28}. Such variations will of course depend on the velocities and masses involved.
\section{Periodic correlations for 3-particle states in a magnetic field}
In this section we will select, to start with, a simple case and try to follow closely the periodicities associated with the intial entangled 3-particle states given by \eq{eq4}-\eq{eq6}.
The constant magnetic field $\overrightarrow{B}$ is taken to be along the x-axis,
\begin{equation}
\overrightarrow{B}= (B,0,0).
\label{eq3_1}
\end{equation}
Three equal mass spin-1/2 particles (created, say, by the disintegration of a single one at rest) are assumed to have their intial velocities in the yz-plane and to be of equal magnitude, so that
\begin{equation}
\overrightarrow{B}.\overrightarrow{v_{(i)}}=0, \qquad \left| \overrightarrow{v_{(i)}} \right|=v \qquad (i=1,2,3).
\label{eq3_2}
\end{equation}
Since $\overrightarrow{v}^2$ and $\overrightarrow{B}.\overrightarrow{v}$ are constants of motion, the velocities wil stay in the yz-plane and remain of equal magnitude. The three velocities rotate uniformly in the yz-plane.
The precession equations \eq{eq11} now simplify (for each case) to
$$ \frac{\rm{d}\overrightarrow{\Sigma}}{\rm{dt}} = -(\overrightarrow{\omega}+ \overrightarrow{\Omega})\times \overrightarrow{\Sigma},$$
where
\begin{eqnarray}
\qquad \overrightarrow{\omega} &=&\frac{eB}{m\gamma}\hat{B} \equiv \omega \hat{B} \nonumber \\
\overrightarrow{\Omega} &=&\frac{\hat{\alpha} eB}{m} \hat{B}=\hat{\alpha} \gamma \omega \hat{B} \equiv \Omega \hat{B}.
\label{eq3_3}
\end{eqnarray}
and hence
\begin{eqnarray}
\frac{\rm{d}\overrightarrow{\Sigma}}{\rm{dt}} &=& -(\omega+ \Omega)\hat{B} \times \overrightarrow{\Sigma}.
\label{eq3_4}
\end{eqnarray}
Now in \eq{eq24} the unitary transformation acting on a state $\left| ijk \right \rangle$ is given by the matrix
$$M \otimes M \otimes M $$
where
\begin{equation}
M \equiv \left| \begin{array}{cc}
\alpha & -i\beta \\
-i\beta & \alpha
\end{array} \right|,
\label{eq3_5}
\end{equation}
and
\begin{eqnarray}
\alpha &=& \cos (\frac{\omega+ \Omega)}{2}t) \nonumber \\
\beta &=& \sin (\frac{\omega+ \Omega)}{2}t) .
\label{eq3_6}
\end{eqnarray}
\subsection{Case 1}
We start by studying the time-evolution of the state \eq{eq4}. One obtains at time $t$
\begin{align}
M \otimes M \otimes M \frac{1}{\sqrt{2}} \left( \left| 111 \right \rangle +\epsilon \left| \overline{1} \overline{1} \overline{1} \right \rangle \right)
= f_{111}\left| 111 \right \rangle \nonumber \\
+ f_{1 \overline{1} \overline{1}}\left| 1 \overline{1} \overline{1} \right \rangle +f_{\overline{1} 1 \overline{1}}\left| \overline{1} 1 \overline{1} \right \rangle
+f_{\overline{1} \overline{1} 1}\left| \overline{1} \overline{1} 1 \right \rangle
+f_{\overline{1} \overline{1} \overline{1}} \left| \overline{1} \overline{1} \overline{1}\right \rangle \nonumber \\ +f_{\overline{1} 11}\left| \overline{1} 11 \right \rangle
+f_{1 \overline{1} 1}\left| 1 \overline{1} 1 \right \rangle +f_{11\overline{1}}\left| 11\overline{1} \right \rangle,
\label{eq3_7}
\end{align}
where
\begin{eqnarray}
f_{111} = \frac{1}{\sqrt{2}} (\alpha^3 +i \epsilon \beta^3) = \epsilon f_{\overline{1} \overline{1} \overline{1}}
\label{eq3_8}
\end{eqnarray}
and
\begin{align}
f_{1 \overline{1} \overline{1}} = f_{\overline{1} 1 \overline{1}} = f_{\overline{1} \overline{1} 1} = f_{\overline{1} 11} = f_{1 \overline{1} 1} = f_{11\overline{1}} \nonumber \\
= -\frac{1}{\sqrt{2}} \alpha \beta( \alpha + \epsilon i \beta).
\label{eq3_9}
\end{align}
The corresponding probabilities given by $P_{ijk}=f_{ijk}^*f_{ijk}$ are
\begin{align}
P_{111} &=& P_{\overline{1} \overline{1} \overline{1}} = \frac{1}{2} (1-3\alpha^2 \beta^2) \nonumber \\
~&=& \frac{1}{2} (1- \frac{3}{4} (\sin^2(( \omega+ \Omega)t)))
\label{eq3_10}
\end{align}
\begin{align}
P_{1 \overline{1} \overline{1}} = P_{\overline{1} 1 \overline{1}} = P_{\overline{1} \overline{1} 1} = P_{\overline{1} 11} = P_{1 \overline{1} 1} = P_{11\overline{1}} \nonumber \\
= \frac{1}{8}(\sin^2(( \omega+ \Omega)t)) ,
\label{eq3_11}
\end{align}
consistent with the constraint of
\begin{equation}
\sum_{ijk} P_{ijk}=1.
\label{eq3_12}
\end{equation}
\begin{figure}[h]
\includegraphics[angle=270,width=0.45\textwidth]{figure1}
\caption{Plot of the variations of the probabilities $P_{ijk}$ defined in the text by \eq{eq3_10} (Curve 1) and \eq{eq3_11} (Curve 2) with $\theta \equiv (\frac{\omega+ \Omega)}{2}t)$.}
\end{figure}
Figure 1 shows the variations of the probabilities $P_{ijk}$ defined in \eq{eq3_10} and \eq{eq3_11} with $\theta \equiv (\frac{\omega+ \Omega)}{2}t)$. The periodic variations of the coefficients are easy to follow. Let us
emphasize some special points:
\begin{enumerate}
\item At $t=0$, one starts with $\alpha=1, \beta=0$
\begin{equation}
\left| \psi \right \rangle \equiv \sum_{ijk} f_{ijk}\left| ijk \right \rangle=\frac{1}{\sqrt{2}} \left( \left| 111 \right \rangle +\epsilon \left| \overline{1} \overline{1} \overline{1} \right \rangle \right).
\label{eq3_13}
\end{equation}
\item At $t=\frac{\pi}{2( \omega+ \Omega)}$, ($\alpha= \beta= \frac{1}{\sqrt{2}}$)
\begin{align}
\left| \psi \right \rangle \equiv \frac{\exp{(i \epsilon \pi /4)} }{2\sqrt{2}} \left\{ \left( \left| 111 \right \rangle +\epsilon \left| \overline{1} \overline{1} \overline{1} \right \rangle \right) - \left( \left| 1 \overline{1} \overline{1} \right \rangle +\left| \overline{1} 1 \overline{1} \right \rangle
+\left| \overline{1} \overline{1} 1 \right \rangle \right) - \left( \left| \overline{1} 11 \right \rangle
+ \left| 1 \overline{1} 1 \right \rangle +\left| 11\overline{1} \right \rangle \right) \right\}.
\label{eq3_14}
\end{align}
Note the negative signs before the two triplets, whose consequences will be discussed in Section IV.
Now all the 8 possible states are present with equal probability
\begin{equation}
P_{ijk}= \frac{1}{8} \qquad (i,j,k)=1 \quad {\rm or} \quad \overline{1}.
\label{eq3_15}
\end{equation}
\item At $t=\frac{\pi}{( \omega+ \Omega)}$, ($\alpha= 0 , \beta= 1$)
\begin{equation}
\left| \psi \right \rangle =\frac{\epsilon }{\sqrt{2}} \left( \left| 111 \right \rangle +\epsilon \left| \overline{1} \overline{1} \overline{1} \right \rangle \right).
\label{eq3_16}
\end{equation}
Note that apart from an overall sign ($\epsilon$) one is back at the starting point.
\end{enumerate}
\subsection{Case 2}
We will now study the time-evolution with the next initial state given by \eq{eq5}:
\begin{align}
\left| \psi \right \rangle_{(t=0)} &= \left| \psi \right \rangle_{(0)} \nonumber \\
&= \frac{1}{\sqrt{3}} \left( \left| 11\overline{1} \right \rangle + \exp({-i \phi}) \left| 1\overline{1} 1 \right \rangle + \exp({i \phi}) \left| \overline{1} 11 \right \rangle \right) ,
\label{eq3_17}
\end{align}
where $ \qquad \left( \phi = 0, \pm \frac{2 \pi}{3}\right). $
We define
\begin{eqnarray}
f &=& (1+\exp({-i \phi})+\exp({i \phi})) \nonumber \\
&=& 3 \qquad \left( \phi = 0\right) \nonumber \\
&=& 0 \qquad \left( \phi = \pm \frac{2 \pi}{3}\right) .
\label{eq3_18}
\end{eqnarray}
At time $t$, the periodic evolution gives (using \eq{eq3_5})
\begin{eqnarray}
\left| \psi \right \rangle_{(t)} &=& M \otimes M \otimes M \left| \psi \right \rangle_{(0)} \nonumber \\
&=& c_0 \left| 111 \right \rangle + \overline{c_0} \left| \overline{1} \overline{1} \overline{1} \right \rangle \nonumber \\
&+& c_1 \left| 11\overline{1} \right \rangle + c_2 \left| 1 \overline{1} 1 \right \rangle + c_3 \left| \overline{1} 11 \right \rangle \nonumber \\
&+& \overline{c_1} \left| \overline{1} \overline{1} 1 \right \rangle+ \overline{c_2} \left| \overline{1} 1 \overline{1} \right \rangle + \overline{c_3} \left| 1 \overline{1} \overline{1} \right \rangle.
\label{eq3_19}
\end{eqnarray}
where with
$$\alpha = \cos (\frac{\omega+ \Omega)}{2}t), \beta = \sin (\frac{\omega+ \Omega)}{2}t)$$
and implementing systematically $\alpha^2 +\beta^2 = 1$, we have
\begin{align}
\sqrt3 c_0 = -i \alpha^2 \beta f &,& \sqrt3 \overline{c_0} = - \alpha \beta^2 f\nonumber \\
\sqrt3 c_1 = \alpha(1 -\beta^2f) &,& \sqrt3 \overline{c_1} = i \beta(1- \alpha^2f)\nonumber \\
\sqrt3 c_2 = \alpha(\exp({-i \phi})-\beta^2f) &,& \sqrt3 \overline{c_2} = i \beta(\exp({-i \phi})- \alpha^2f) \nonumber \\
\sqrt3 c_3 = \alpha(\exp({i \phi})-\beta^2f) &,& \sqrt3 \overline{c_3} = i \beta(\exp({i \phi})- \alpha^2f) .
\label{eq3_20}
\end{align}
The correlations have periodicities determined by ($\alpha, \beta$) above and are
\begin{eqnarray}
P_{111} = \frac{1}{3} \alpha^4\beta^2f^2 , P_{\overline{1} \overline{1} \overline{1}} = \frac{1}{3} \alpha^2\beta^4f^2 &,& \nonumber \\
P_{1 \overline{1} \overline{1}} = \frac{1}{3} \beta^2(1-\alpha^2f\exp({-i \phi}))(1-\alpha^2f\exp({i \phi})) &,& \nonumber \\
P_{\overline{1} 11} = \frac{1}{3} \alpha^2(1-\beta^2f\exp({-i \phi}))(1-\beta^2f\exp({i \phi})) &,& \nonumber \\
P_{1 \overline{1} 1} = \frac{1}{3} \alpha^2(1-\beta^2f\exp({-i \phi}))(1-\beta^2f\exp({i \phi})) &,&
\nonumber \\
P_{\overline{1} 1 \overline{1}} = \frac{1}{3} \beta^2(1-\alpha^2f\exp({-i \phi}))(1-\alpha^2f\exp({i \phi})) &,& \nonumber \\
P_{\overline{1} \overline{1} 1} = \frac{1}{3} \beta^2(1-\alpha^2f)^2,
P_{11\overline{1}} = \frac{1}{3} \alpha^2(1-\beta^2f)^2 .
\label{eq3_21}
\end{eqnarray}
After this derivation, one notes that for:
\begin{enumerate}
\item ($\phi= 0 , f= 3$)
\begin{eqnarray}
P_{111} = 3 \alpha^4\beta^2 &,& P_{\overline{1} \overline{1} \overline{1}} = 3 \alpha^2\beta^4 \nonumber \\
P_{\overline{1} \overline{1} 1} = P_{1 \overline{1} \overline{1}} = P_{\overline{1} 1 \overline{1}}= \frac{1}{3} \beta^2(1-3\alpha^2)^2 \nonumber \\
P_{\overline{1} 11} =P_{1 \overline{1} 1} = P_{11\overline{1}} = \frac{1}{3} \alpha^2(1-3\beta^2)^2 .
\label{eq3_22}
\end{eqnarray}
consistent with the constraint \eq{eq3_12} of
$\sum_{ijk} P_{ijk}=1.$
\begin{figure}[h]
\includegraphics[angle=270,width=0.45\textwidth]{figure2}
\caption{Plot of the variations of the probabilities $P_{ijk}$ defined in the text by \eq{eq3_22} with $\theta \equiv (\frac{\omega+ \Omega)}{2}t)$, where $P_{111}$ is Curve 1, $P_{\overline{1} \overline{1} \overline{1}}$ is Curve 2, $P_{\overline{1} \overline{1} 1} = P_{1 \overline{1} \overline{1}} = P_{\overline{1} 1 \overline{1}}$ are Curve 3 and $P_{\overline{1} 11} =P_{1 \overline{1} 1} = P_{11\overline{1}}$ are Curve 4.}
\end{figure}
Figure 2 shows the variations of the probabilities $P_{ijk}$ defined in \eq{eq3_22} with $\theta \equiv (\frac{\omega+ \Omega)}{2}t)$.
\item ($\phi= \pm \frac{2\pi}{3} , f= 0$)
\begin{eqnarray}
P_{111} = P_{\overline{1} \overline{1} \overline{1}} = 0 \nonumber \\
P_{\overline{1} \overline{1} 1} = P_{1 \overline{1} \overline{1}} = P_{\overline{1} 1 \overline{1}}= \frac{1}{3} \beta^2 \nonumber \\
P_{\overline{1} 11} =P_{1 \overline{1} 1} = P_{11\overline{1}} = \frac{1}{3} \alpha^2 .
\label{eq3_23}
\end{eqnarray}
consistent with the constraint \eq{eq3_12} of
$\sum_{ijk} P_{ijk}=1.$
\end{enumerate}
Case (i) exhibits a richer pattern of periodicity which we analyse now, pinpoiting some special values of $t$.
At $t=0$, one starts with $\alpha=1, \beta=0$:
\begin{equation}
\left| \psi \right \rangle_{(0)} =\frac{1}{\sqrt{3}} \left( \left| 11 \overline{1} \right \rangle + \left| 1 \overline{1} 1 \right \rangle + \left| \overline{1}11 \right \rangle \right).
\label{eq3_24}
\end{equation}
At $t=\frac{\pi}{( \omega+ \Omega)}$, ($\alpha= 0 , \beta= 1$):
\begin{equation}
\left| \psi \right \rangle =\frac{i }{\sqrt{3}} \left( \left| \overline{1}\overline{1}1 \right \rangle + \left| \overline{1}1 \overline{1} \right \rangle + \left| 1 \overline{1} \overline{1} \right \rangle\right).
\label{eq3_25}
\end{equation}
we get the flipped over complementary triplet.
At $t=\frac{\pi}{2( \omega+ \Omega)}$, ($\alpha= \beta= \frac{1}{\sqrt{2}}$):
\begin{eqnarray}
\left| \psi \right \rangle = \frac{-i}{2\sqrt{6}} \left( 3 \left| 111 \right \rangle + \left| \overline{1} \overline{1} 1 \right \rangle + \left| \overline{1} 1 \overline{1} \right \rangle +\left| 1 \overline{1} \overline{1} \right \rangle \right) \nonumber \\
+ \frac{1}{2\sqrt{6}} \left( 3 \left| \overline{1} \overline{1} \overline{1} \right \rangle - \left| 11\overline{1} \right \rangle -\left| 1 \overline{1} 1 \right \rangle -\left| \overline{1}11 \right \rangle \right).
\label{eq3_26}
\end{eqnarray}
Now all the possible $\left| ijk \right \rangle$ (8 in number) appear, grouped as above according as the multiplicity of the index $\left| 1 \right \rangle$ is odd (3 or 1) or even (0 or 2).
Other interesting points are provided by the extrema of ($P_{111} , P_{\overline{1} \overline{1} \overline{1}}$).
For $\alpha^2=2/3, \beta^2=1/3$,
\begin{eqnarray}
P_{111} = \frac{4}{9}, P_{\overline{1} \overline{1} \overline{1}} = \frac{2}{9} \nonumber \\
P_{\overline{1} \overline{1} 1} = P_{1 \overline{1} \overline{1}} = P_{\overline{1} 1 \overline{1}}= \frac{1}{9} \nonumber \\
P_{\overline{1} 11} =P_{1 \overline{1} 1} = P_{11\overline{1}} = 0.
\label{eq3_27}
\end{eqnarray}
Considering the positive roots for example, $\alpha=\sqrt{2/3}, \beta=1/\sqrt{3}$,
\begin{equation}
\left| \psi \right \rangle = \frac{-i}{3} \left( 2 \left| 111 \right \rangle + \left| \overline{1} \overline{1} 1 \right \rangle + \left| \overline{1} 1 \overline{1} \right \rangle +\left| 1 \overline{1} \overline{1} \right \rangle \right) - \frac{\sqrt{2}}{3} \left| \overline{1} \overline{1} \overline{1} \right \rangle .
\label{eq3_28}
\end{equation}
For $\alpha=1/\sqrt{3}, \beta=\sqrt{2/3}$,
\begin{eqnarray}
P_{111} = \frac{2}{9}, P_{\overline{1} \overline{1} \overline{1}} = \frac{4}{9} \nonumber \\
P_{\overline{1} \overline{1} 1} = P_{1 \overline{1} \overline{1}} = P_{\overline{1} 1 \overline{1}}= 0 \nonumber \\
P_{\overline{1} 11} =P_{1 \overline{1} 1} = P_{11\overline{1}} = \frac{1}{9},
\label{eq3_29}
\end{eqnarray}
and
\begin{equation}
\left| \psi \right \rangle = \frac{-i\sqrt{2}}{3} \left| 111 \right \rangle - \frac{1}{3} \left( 2 \left| \overline{1} \overline{1} \overline{1} \right \rangle + \left| 11\overline{1} \right \rangle + \left| 1 \overline{1} 1 \right \rangle + \left| \overline{1}11 \right \rangle\right).
\label{eq3_30}
\end{equation}
Apart from the factor $i$ changing place, the passage from \eq{eq3_28} to \eq{eq3_30} corresponds to
\begin{equation}
(\left| 1 \right \rangle, \left| \overline{1} \right \rangle) \rightarrow (\left| \overline{1} \right \rangle, \left| 1 \right \rangle).
\label{eq3_31}
\end{equation}
The coefficients for the cases above will be further discussed in Section IV.
The time-evolution of the initial state
\begin{equation}
\left| \psi \right \rangle_{(0)} = \frac{1}{\sqrt{3}} \left( \left| \overline{1}\overline{1}1 \right \rangle + \exp({i \phi}) \left| \overline{1} 1 \overline{1}\right \rangle + \exp({-i \phi}) \left| 1\overline{1}\overline{1} \right \rangle \right)
\label{eq3_32}
\end{equation}
can be studied in closely analogous fashion, but we will not repeat the steps again.
The periods considered above, typically generated by ($\alpha, \beta$), of \eq{eq3_6} where ($\omega, \Omega$) are given by \eq{eq12} can be long but diminish as $B$ and the velocities increase in magnitude.
\section{Periodic density matrices}
We have seen how, starting with a restricted initial state, spin precessions in a magnetic field lead to periodic appearances and disappearances of all the possible eight states $\left| ijk \right \rangle$ of three spin-1/2 particles . Since such periodicities are induced through ``local'' unitary transformations, acting separately on each state $\left| \pm \right \rangle$, the sum of the varying correlations remains unity. Moreover, \textit{the basic invariant measures of entanglement (3-tangle, 2-tangles) must also be conserved}. The corresponding \textit{constrained} periodic variations of the elements of the density matrices is briefly studied below for special cases.
For
\begin{eqnarray}
\left| \psi \right \rangle = \left( f_0\left| 11 \right \rangle+ f_1\left| 1\overline{1} \right \rangle+ f_2\left|\overline{1}1 \right \rangle+ f_3\left| \overline{1}\overline{1} \right \rangle\right) \left| 1 \right \rangle \nonumber \\
+ \left( g_0\left| \overline{1}\overline{1} \right \rangle+ g_1\left| \overline{1}1 \right \rangle+ g_2\left| 1\overline{1} \right \rangle+ g_3\left| 11 \right \rangle \right) \left| \overline{1} \right \rangle
\label{eq4_1}
\end{eqnarray}
tracing out the index 3, the density matrix for the $(12)$ subsystem is
\begin{equation}
\rho_{12}= \left| \begin{array}{cccc}
a_{00} & a_{01} & a_{02} & a_{03}\\
a_{01}^{*} & a_{11} & a_{12} & a_{13}\\
a_{02}^{*} & a_{12}^{*} & a_{22} & a_{23}\\
a_{03}^{*} & a_{13}^{*} & a_{23}^{*} & a_{33}
\end{array} \right|
\label{eq4_2}
\end{equation}
where
\begin{eqnarray}
a_{00}=f_0f_0^{*}+g_3g_3^{*},\qquad a_{01}=f_0f_1^{*}+g_3g_2^{*} \nonumber\\
a_{02}=f_0f_2^{*}+g_3g_1^{*},\qquad a_{03}=f_0f_3^{*}+g_3g_0^{*} \nonumber\\
a_{11}=f_1f_1^{*}+g_2g_2^{*},\qquad a_{12}=f_1f_2^{*}+g_2g_1^{*} \nonumber\\
a_{13}=f_1f_3^{*}+g_2g_0^{*},\qquad a_{22}=f_2f_2^{*}+g_1g_1^{*} \nonumber\\
a_{23}=f_2f_3^{*}+g_1g_0^{*},\qquad a_{33}=f_3f_3^{*}+g_0g_0^{*}.
\label{eq4_3}
\end{eqnarray}
Similarly, one can obtain $\rho_{23},\rho_{31}$.
The 3-tangle (invariant under permutations of particles $1,2,3$) is obtained as follows \cite{key8}:
We define
\begin{equation}
\widetilde{\rho_{12}}= \left| \begin{array}{cc}
0 & -i\\
i & 0
\end{array} \right| \otimes
\left| \begin{array}{cc}
0 & -i\\
i & 0
\end{array} \right| \rho_{12}^{*}
\left| \begin{array}{cc}
0 & -i\\
i & 0
\end{array} \right| \otimes
\left| \begin{array}{cc}
0 & -i\\
i & 0
\end{array} \right| .
\label{eq4_4}
\end{equation}
For the subsystems considered ($\rho_{12}\widetilde{\rho_{12}}$) has at most two non-zero eigenvalues.
Let ($\lambda_{2},\lambda_{2}$) be the square roots (positive) of these two. Then the 3-tangle is given by
\begin{equation}
\tau_{123}=4\lambda_{1}\lambda_{2}.
\label{eq4_5}
\end{equation}
This can also be directly expressed in terms of the coefficients of \eq{eq4_1} above:
\begin{equation}
\tau_{123}=4\left|d_1-2d_2+4d_3 \right|,
\label{eq4_6}
\end{equation}
where (in our notations of \eq{eq4_1})
\begin{eqnarray}
d_1 &=& (f_0g_0)^{2}+(f_1g_1)^{2}+(f_2g_2)^{2}+(f_3g_3)^{2} \nonumber\\
d_2 &=& f_0g_0 (f_1g_1+f_2g_2+f_3g_3)\nonumber\\
&+&(f_1g_1f_2g_2+f_2g_2f_3g_3+f_3g_3f_1g_1)\nonumber\\
d_3 &=& f_0f_3g_1g_2+f_1f_2g_0g_3.
\label{eq4_7}
\end{eqnarray}
The conversion of \eq{eq4_5} to \eq{eq4_6} permits us to relate directly certain central features of the periodicities studied in Section III to constraints imposed by the invariance of \eq{eq4_5}. In particular, let us now evaluate the crucial role of the negative signs signalled below \eq{eq3_14}.
For the initial state \eq{eq3_13}, at $t=0$
\begin{eqnarray}
f_0= \epsilon g_0 = \frac{1}{\sqrt2} \nonumber\\
f_1=f_2=f_3=g_1=g_2=g_3=0
\label{eq4_8}
\end{eqnarray}
and thus
\begin{equation}
d_1 = \frac{1}{4}, d_2=d_3=0.
\label{eq4_9}
\end{equation}
Hence from \eq{eq4_6}
\begin{equation}
\tau_{123}=1
\label{eq4_10}
\end{equation}
a well-known result for GHZ states \eq{eq3_13}.
At $t=\frac{\pi}{2(\omega+\Omega)}$, from \eq{eq3_14} and \eq{eq4_7},
\begin{equation}
(d_1,d_2,d_3) = \frac{\exp{i\pi/4}}{{(2\sqrt2})^{4}}(4,6,-2).
\label{eq4_11}
\end{equation}
Hence again (as expected),
\begin{equation}
\tau_{123}= \frac{4}{64}\left| 4-2.6-4.2 \right|=1.
\label{eq4_12}
\end{equation}
If all the terms have the same sign, as in
\begin{eqnarray}
\left| \psi \right \rangle' &=& \frac{1}{2\sqrt2} \left( \left| 111 \right \rangle+ \left| \overline{1}\overline{1}\overline{1} \right \rangle+ \left|1\overline{1}\overline{1} \right \rangle+ \left| \overline{1}1\overline{1} \right \rangle \right) \nonumber \\
&+& \frac{1}{2\sqrt2} \left( \left| \overline{1}\overline{1}1 \right \rangle
+ \left| \overline{1}11 \right \rangle + \left| 1\overline{1}1 \right \rangle+ \left| 11\overline{1} \right \rangle \right)
\label{eq4_13}
\end{eqnarray}
then $d_3$ changes sign and
\begin{equation}
\tau_{123}= \frac{4}{64}\left| 4-2.6+4.2 \right|=0.
\label{eq4_14}
\end{equation}
Thus the two negative signs in \eq{eq3_14} alters $\tau_{123}$ from a minimum (0) to maximum (1).
For the initial state \eq{eq3_17}, using \eq{eq3_18}-\eq{eq3_20} the f's in \eq{eq4_1} have cubic periodic terms given by ($\alpha\beta^{2},\alpha^{2}\beta$), where
\begin{equation}
\alpha=\cos\left( \frac{\omega+\Omega)}{2}t\right) ,\qquad \beta= \sin \left( \frac{\omega+\Omega)}{2}t\right).
\label{eq4_15}
\end{equation}
Hence $P_{12}$ in \eq{eq4_2} and \eq{eq4_3} has sixth order periodic terms ($\alpha^{2}\beta^{4},\alpha^{4}\beta^{2}, \alpha^{3}\beta^{3}$). Such periodicities indeed turn out to be compatible with \cite{key9}:
\begin{equation}
\tau_{123}= 0, \tau_{12}=\tau_{13}=\tau_{23}=\frac{4}{9}.
\label{eq4_16}
\end{equation}
One realizes now more fully how elaborate and subtle patterns of periodicities are compatible with constraints of \textit{local} unitary transformations involved in spin-precessions.
\section{Constant orthogonal electric and magnetic fields}
We briefly indicate below how the periodicities studied so far, for a magnetic field alone are affected by the presence of an electric field $\overrightarrow{E}$ satisfying
\begin{equation}
\overrightarrow{E}. \overrightarrow{B}=0, \qquad \left| \overrightarrow{E} \right| < \left| \overrightarrow{B} \right| .
\label{eq5_1}
\end{equation}
Consider a Lorentz transformation corresponding to the 4-velocity
\begin{equation}
u''= (1- \frac{E^{2}}{B^{2}})^{-1/2}(1, \frac{E}{B} \hat{E} \times \hat{B})
\label{eq5_2}
\end{equation}
denoting $ \overrightarrow{E}= E. \hat{E}, \overrightarrow{B}= B. \hat{B}\qquad (\hat{E}^{2}=1= \hat{B}^{2})$.
In the transformed frame the tensor $(\overrightarrow{E}, \overrightarrow{B})$ reduces to $(\overrightarrow{E}', \overrightarrow{B}')$ where
\begin{equation}
\overrightarrow{E}'=0,\qquad \overrightarrow{B}' = (1- \frac{E^{2}}{B^{2}})^{-1/2} \overrightarrow{B}
\label{eq5_3}
\end{equation}
such that
\begin{eqnarray}
\overrightarrow{B}'^{2}= \overrightarrow{B}'^{2}-\overrightarrow{E}'^{2}= \overrightarrow{B}^{2}-\overrightarrow{E}^{2}\nonumber \\
\overrightarrow{B}'. \overrightarrow{E}'=0=\overrightarrow{B}. \overrightarrow{E}.
\label{eq5_4}
\end{eqnarray}
So in this frame, one finds back the situation studied in the previous Sections II-IV, with
\begin{equation}
B'=(B^{2}-E^{2})^{1/2}.
\label{eq5_5}
\end{equation}
The velocities and the spins of the particles are transformed according to standard rules \cite{key1,key2}. We now recapitulate some essential points.
A 4-velocity $u$ is transformed by a Lorentz transformation corresponding to $u''$ to $u'$ such that
\begin{equation}
u_0'= (u_0u''_0+\overrightarrow{u}.\overrightarrow{u''}).
\label{eq5_6}
\end{equation}
Define
\begin{eqnarray}
a &=& (1+u_0)(1+u''_0)(1+u'_0)\nonumber \\
b &=& (1+u_0+u''_0+u'_0),
\label{eq5_7}
\end{eqnarray}
and
\begin{equation}
\cos \frac{\delta}{2}= \frac{b}{\sqrt{2a}}.
\label{eq5_8}
\end{equation}
The spin states are conserved if $$\overrightarrow{u} \times \overrightarrow{u''} = 0.$$ Otherwise they undergo a Wigner rotation $\delta$ about the axis
\begin{equation}
\hat{k} = \frac{\overrightarrow{u} \times \overrightarrow{u''} }{\left| \overrightarrow{u} \times \overrightarrow{u''}\right| },
\label{eq5_9}
\end{equation}
such that
\begin{eqnarray}
\left| + \right \rangle \rightarrow \left| + \right \rangle' &=& \left( \cos \frac{\delta}{2} +i\hat{k_3} \sin \frac{\delta}{2} \right) \left| + \right \rangle \nonumber \\ &+& i\left( \hat{k_1}-i \hat{k_2}\right) \sin \frac{\delta}{2}\left| - \right \rangle \nonumber \\
\left| - \right \rangle \rightarrow \left| - \right \rangle' &=& i\left( \hat{k_1}+i \hat{k_2}\right) \sin \frac{\delta}{2}\left| + \right \rangle \nonumber \\ & +& \left( \cos \frac{\delta}{2} -i\hat{k_3} \sin \frac{\delta}{2} \right) \left| - \right \rangle.
\label{eq5_10}
\end{eqnarray}
The inverse rotation $(\delta \rightarrow -\delta)$ expresses $(\left| + \right \rangle , \left| - \right \rangle)$ in terms of $(\left| + \right \rangle' , \left| - \right \rangle')$.
The crucial point to note is that in \eq{eq5} apart from the $(\gamma ', \hat{v}')$ corresponding to the transformed velocities (depending on the intial velocity $\overrightarrow{v}$ and $u''$ given by \eq{eq5_2}), $B$ is replaced by $ B'=(B^{2}-E^{2})^{1/2}$.
The magnitudes $(\omega ', \Omega ')$ thus obtained determine the modified periodicities. This is the consequence of the presence of $\overrightarrow{E}$ in the initial frame.
In the transformed frame our preceeding results (for $\overrightarrow{E}=0$) can be implemented systematically along with the velocities transformed corresponding to \eq{eq5_2}. Then the inverting of \eq{eq5_10} gives the results for the initial frame ($\overrightarrow{E} \neq 0$).
\section{Classification scheme of 3-particle entangled states}
In this Section, we propose a new classification scheme of the 3-particle entangled states, by using the eigenstates of $Z=\left( \overrightarrow{J_1} \times \overrightarrow{J_2} \right) . \overrightarrow{J_3}$ of three angular momenta.
One can systematically construct eigenstates \cite{key6,key7} of three coupled angular momenta ($\overrightarrow{J_1},\overrightarrow{J_2},\overrightarrow{J_3}$) by diagonalizing the operators
\begin{eqnarray}
(\overrightarrow{J_1}+\overrightarrow{J_2}+\overrightarrow{J_3})^2\nonumber \\
({J_1^{(0)}}+{J_2^{(0)}}+{J_3^{(0)}})\nonumber \\
\rm{and} \qquad Z=\left( \overrightarrow{J_1} \times \overrightarrow{J_2} \right) . \overrightarrow{J_3} ,
\label{eqa_1}
\end{eqnarray}
where $J_i^{(0)}, (i=1,2,3)$ are the projections on the z-axis.
The states are denoted by the respective eigenvalues of the above operators ($j(j+1),j^{(0)}, \zeta$) as
$ \left| jm \zeta \right \rangle.$
Not only one obtains a complete mutually orthogonal set of eigenstates for each $j$ but along with reduction with respect to the rotation group one obtains {\it simultaneously a reduction with respect to $S_3$, the permutation group of three particles}. This is in sharp contrast with the usual 2-step reduction via 3-j coefficients where such permutations lead to 6-j coefficients.
For our purposes, we need here only the results for
$ j_1=j_2=j_3=\frac{1}{2}.$
In the table \ref{tab:coupling}, the states on the left correspond to eigenvalues ($j,m, \zeta$) respectively of the operators \eq{eqa_1} and those on the right to the values of $\pm\frac{1}{2}$ of ($m_1,m_2, m_3$) of $J_i^{(0)}$, ($i=1,2,3$) denoted by ($\left| 1 \right \rangle, \left| \overline{1} \right \rangle)$.
\begin{table}[ht]
\caption{The states on the left correspond to eigenvalues ($j,m, \zeta$) respectively of the operators \eq{eqa_1} and those on the right to the values of $\pm\frac{1}{2}$ of ($m_1,m_2, m_3$) of $J_i^{(0)}$, ($i=1,2,3$) denoted by ($\left| 1 \right \rangle, \left| \overline{1} \right \rangle)$.
}
\begin{tabular}{|l | c|}
\hline
$ \left| jm \zeta \right \rangle $ & $ \left| m_1 m_2 m_3 \right \rangle $ \\
\hline
\hline
$ \left| \frac{3}{2} \frac{3}{2} 0 \right \rangle $ & $ \left| 111 \right \rangle $ \\
$ \left| \frac{3}{2} \frac{-3}{2} 0 \right \rangle $ & $ \left| \overline{1} \overline{1} \overline{1} \right \rangle $ \\
$ \left| \frac{3}{2} \frac{1}{2} 0 \right \rangle $ & $ \frac{1}{\sqrt3}\left( \left| \overline{1}11 \right \rangle+\left| 1 \overline{1}1 \right \rangle+\left| 11\overline{1} \right \rangle\right) $ \\
$ \left| \frac{3}{2} \frac{-1}{2} 0\right \rangle $ & $ \frac{1}{\sqrt3}\left( \left| 1 \overline{1}\overline{1} \right \rangle+\left| \overline{1}1 \overline{1} \right \rangle+\left| \overline{1}\overline{1}1 \right \rangle \right) $ \\
$ \left| \frac{1}{2} \frac{1}{2} \frac{\pm \sqrt3}{4} \right \rangle $ & $ \frac{1}{\sqrt3}\left( \exp({\pm i\frac{2\pi}{3}})\left| \overline{1}11 \right \rangle+\exp({\mp i\frac{2\pi}{3}})\left| 1 \overline{1}1 \right \rangle+\left| 11\overline{1} \right \rangle \right) $ \\
$ \left| \frac{1}{2} -\frac{1}{2} \frac{\pm \sqrt3}{4} \right \rangle $ & $ \frac{1}{\sqrt3}\left( \exp({\mp i\frac{2\pi}{3}}) \left| 1 \overline{1}\overline{1} \right \rangle+\exp({\pm i\frac{2\pi}{3}}) \left| \overline{1}1 \overline{1} \right \rangle+\left| \overline{1}\overline{1}1 \right \rangle\right) $ \\
\hline
\end{tabular}
\label{tab:coupling}
\end{table}
This provides the complete set of 8 states spanning the space of possible values of ($m_1,m_2, m_3$).
Thus
\begin{equation}
\frac{1}{\sqrt2}\left( \left| \frac{3}{2} \frac{3}{2} 0 \right \rangle\pm \left| \frac{3}{2} -\frac{3}{2} 0 \right \rangle \right)=\frac{1}{\sqrt2}\left( \left| 111 \right \rangle \pm \left| \overline{1} \overline{1} \overline{1} \right \rangle\right),
\label{eqa_5}
\end{equation}
giving the $\left| GHZ \right \rangle$ states as a doublet.
The others correspond directly to the Werner ($\left| W \right \rangle$) and the flipped Werner ($\left| \widetilde{W} \right \rangle$) and their variants with relative phases $\exp({\pm i\frac{2\pi}{3}})$, which we had mentioned in the remarks below \eq{eq6} in Section II.
Of the three operators in \eq{eqa_1} the first two are invariant under all permutations of the particles ($1,2,3$). The remaining one ($Z$) is invariant under circular permutations of ($1,2,3$) and just changes sign under ($12$),($23$) and ($31$). This is in sharp contrast to the standard 2-step couplings where one has to pass from one 3-j coupling scheme to another under permutations. This is at the root of the simultaneous reduction under $S_3$ via the implementation of $Z$. This also helps to explain the direct relations of the $Z$-eigenstates with 3-tangles invariant under permutations of ($1,2,3$).
\section{Concluding remarks and outlook}
We summarize the essential features of the present work, and give an outlook of future directions of work:
\begin{enumerate}
\item An external (constant) magnetic field induces a precession of the spin of each particle of the entangled states considered, through local unitary transformations. Note that for two particle states of total spin zero, studied in earlier paper, the periodic precession of individual spins was not present. Here we analysed three particle states where such \textit{periodicities} (2 and 4 periods respectively for the particular cases illustrated) in correlations and density matrices were remarkably displayed and intertwined. The patterns of periodicity thus emerging were shown to be remarkably rich and subtle. Such a study was presented for the first time.
\item The individual precessions being implemented by local unitary matrices, the initial 3-tangle was conserved. But simply the verification of this fact was not the aim of the analyses of sections III and IV. We went beyond that and displayed in details how the conservation left scope for component periodic correlations to appear, increase and decrease. We pinpointed the crucial roles of the signs of the coefficients of different components, again for the first time.
\item The generalization to include orthogonal electric field was presented in Section V, using the Wigner rotation. For $E=B$, the Lorentz transformation via \eq{eq5_2} is not well-defined and for $E>B$ it becomes complex. The limiting case $E=B$ (e.g. for a plane wave field) will be studied elsewhere using exact solutions \cite{key10} of the Dirac-Pauli equations in such fields.
\item The remarkable and systematic correspondence of famous entangled states to a specific coupling scheme for 3-angular momenta was presented in Section VI, and a new classification scheme was proposed. We intend to study this aspect in details elsewhere.
\item Finally one may note that unitary transformations may be {\it non-local} when induced via unitary braid matrices \cite{key11}. Acting on pure product states, they can then {\it generate} entanglement. In the present work we started with states already entangled and then followed their periodic ramifications as they evolved in the magnetic fields, which is completely different.
\end{enumerate}
|
\section{Introduction}
Ten years after its discovery, cosmic acceleration remains one of the central puzzles in cosmology.
Although the data is thus far consistent with dark energy plus General Relativity (GR), a tantalizing alternative is that new
gravitational degrees of freedom on cosmological scales are responsible for late-time acceleration.
The most compelling idea along these lines is that the graviton has a small mass or width, of the order of today's
Hubble parameter, which could account for the apparent smallness of the cosmological constant via the degravitation mechanism~\cite{dgs,addg,degrav}.
The central difficulty in constructing any consistent infrared (IR) modified theory of gravity is the avoidance of ghosts. For instance, massive gravity theories
with a hard mass for the graviton~\cite{FP} are well-known to be unstable~\cite{BoulwareDeser,nima,Creminellipaper}. More promising constructions
rely on branes and extra dimensions~\cite{DGP,cascade1,cascade2,deRham:2008qx,intersecting,cascade3,aux1,aux2}, such as the
Dvali-Gabadadze-Porrati (DGP) model~\cite{DGP}. The normal branch of the DGP model is perturbatively ghost free, in contrast to
the self-accelerating branch~\cite{lpr,dgpghost}, and thus represents a perturbatively consistent infrared modification of gravity in which the gravity has
a soft mass, {\it i.e.}, it is a resonance.
The Cascading Gravity framework~\cite{cascade1,cascade2,deRham:2008qx,intersecting,cascade3} proposed recently generalizes the DGP model to
higher dimensions. There are two motivations for considering this broader set-up.
As theoretical motivation, it offers a new class of IR modified gravity theories whose properties are therefore worth exploring. In particular, the
lower momenta dependance of the mass could lead to important consequences for degravitation where the fundamental cosmological constant, here arising with tension of the 3-brane, could be large but give rise only to a small backreaction on the geometry on timescales of the order of the graviton Compton wavelength.
As observational motivation, the model carries distinguishable signatures that could be observable at late times~\cite{niayeshghazal}.
In Cascading Gravity, our 3-brane is embedded in a succession of higher-dimensional branes, each
with their own intrinsic Einstein-Hilbert term. In the simplest realization, with 6D bulk space-time, the action is
\begin{eqnarray}
\nonumber
S_{\rm 6D \; Cascading} &=& \frac{M_6^4}{2}\int_{\rm bulk} {\rm d}^6x \sqrt{-g_6} R_6 + \frac{M_5^3}{2}\int_{\rm 4-brane} {\rm d}^5x \sqrt{-g_5} R_5 \\
&+& \int_{\rm 3-brane} {\rm d}^4x \sqrt{-g_4}\left(\frac{M_4^2}{2}R_4 + {\cal L}^{(4)}_{\rm matter}\right)\,.
\label{6Dcascade}
\end{eqnarray}
There are two characteristic crossover scales, $m_5 = M_5^3/M_4^2$ and $m_6 = M_6^4/M_5^3$. Assuming $m_6\ll m_5$, the gravitational potential
on our brane cascades from a $1/r$ (4D gravity) regime at short distances, to a $1/r^2$ (5D gravity) regime at intermediate distances, and
finally to a $1/r^3$ (6D gravity) regime at large distances. Remarkably, the codimension-1 kinetic term makes the 4D propagator
finite~\cite{cascade1,cascade2}, thereby regulating the logarithmic divergence characteristic of pure codimension-2 branes~\cite{cod2eft,massimo}.
Some of the cosmological implications of this model have been explored in~\cite{niayeshghazal,markwyman,nishant}.
The ghost issue is trickier. Perturbing around the flat space solution with empty branes, one finds that
a ghost scalar mode propagates. This is seen most directly from the tensor structure of the one-graviton exchange amplitude between conserved sources
in the UV limit:
\begin{equation}
\mathcal A \; \sim \;
{\cal T}_4^{\mu\nu}\cdot \frac{1}{k^2-i\varepsilon}\cdot \left({\cal T}'_{4 \mu\nu}-\frac
13 \eta_{\mu\nu}{\cal T}'_4\right)
- \frac{1}{6}\;{\cal T}_4\cdot \frac{1}{k^2-i\varepsilon}\cdot {\cal T}_4'\,.
\label{ampghost}
\end{equation}
We have conveniently separated terms as the sum of a massive spin-2 contribution, with the well-known $1/3$ coefficient, plus a contribution from a conformally-coupled
scalar $\pi$. The problem is with this scalar mode. The last term in~(\ref{ampghost}) is negative, indicating
that $\pi$ is a ghost. This UV behavior is identical to other higher-dimensional scenarios~\cite{gigashif,sergei}.
However, it was immediately noticed~\cite{cascade1} that the ghost is removed if the codimension-2 is endowed with sufficiently large tension $\lambda$. (In analogy with a cosmic string in
ordinary 4D gravity, adding tension to a codimension-2 defect leaves the induced geometry flat but creates a deficit angle in the extra dimensions.) Instead of~(\ref{ampghost}) the
exchange amplitude now depends on the tension $\lambda$:
\begin{equation}
\mathcal A \; \sim \;
{\cal T}_4^{\mu\nu}\cdot \frac{1}{k^2-i\varepsilon}\cdot \left({\cal T}'_{4 \mu\nu}-\frac
13 \eta_{\mu\nu}{\cal T}'_4\right)
+\frac{1}{6\left(\frac{3\lambda}{2m_6^2M_4^2}-1\right)}\;{\cal T}_4\cdot \frac{1}{k^2-i\varepsilon}\cdot {\cal T}_4'\,.
\label{ampnoghost}
\end{equation}
The $\pi$ contribution is healthy, and the ghost is absent, provided that
\begin{equation}
\lambda \geq \frac{2}{3}m_6^2M_4^2\,.
\label{lambound}
\end{equation}
Note the essential role of the codimension-1 brane. In the limit $M_5\rightarrow 0$ where its induced gravity term disappears, the lower
bound in~(\ref{lambound}) diverges. In particular, adding tension cannot cure the ghost in the pure codimension-2 DGP case~\cite{gigashif,sergei}.
A limitation of the calculation presented in~\cite{cascade1} is that~(\ref{ampnoghost}) and~(\ref{lambound}) were obtained through a decoupling limit of~(\ref{6Dcascade}),
namely $M_5,M_6\rightarrow \infty$ keeping $\Lambda_6 = (m_6^4M_5^3)^{1/7}$ fixed.
(The derivation of these results using the decoupling limit is reviewed in Appendix A.)
Although the decoupling limit in DGP has been shown to faithfully reproduce much of the phenomenology of the full-fledged
higher-dimensional theory~\cite{lpr,nathan}, a complete 6D calculation is clearly needed to establish unequivocally the consistency of the cascading gravity framework.
In particular, the analysis of~\cite{cascade1} could not demonstrate the absence of ghost modes that disappear in the decoupling limit.
In this paper, we study the issue of stability rigorously by perturbing the full 6D action~(\ref{6Dcascade}) around a background including tension
on the 3-brane. The background geometry is flat everywhere, but the extra dimensions show a deficit angle due to the codimension-2 source, with
the 3-brane located at the conical singularity.
We are immediately faced with a puzzle: brane tension, by virtue of being a source for gravity, cannot by itself modify the graviton kinetic term. Indeed, the
3-brane part of the quadratic lagrangian density takes the initial form
\begin{equation}
S_{\rm 3-brane} = \int \mathrm{d} ^4x \(-\frac {3M_4^2{}} 4\, \pi \Box_4
\pi- 2\pi^2 \lambda\)\,,
\label{L3ini}
\end{equation}
where $\pi$ is the scalar perturbation of the 4D metric.
Thus $\pi$ acquires a localized mass term from $\lambda$, but as expected its kinetic term is unscathed and remains negative.
Incidentally, even the mass term is puzzling --- $\pi$ is a Goldstone mode and should enjoy a shift symmetry. How can~(\ref{L3ini})
be consistent with the conclusions of the decoupling analysis?
The resolution lies in the other scalar modes: integrating out the non-dynamical degrees of freedom in the bulk
results in contributions localized on the 3-brane, which generate $\lambda$-dependent additions to the $\pi$ kinetic term.
The final 3-brane lagrangian for $\pi$ is
\begin{equation}
S_{\rm 3-brane} = \int \mathrm{d} ^4x\, \frac{3M_4^2}{4}\left(\frac{3\lambda}{2m_6^2M_4^2}-1\right)\pi\Box_4\pi\,.
\end{equation}
Provided that $\lambda \ge 2m_6^2M_4^2/3$, the kinetic term for $\pi$ is positive, in precise agreement with the decoupling limit.
Meanwhile, the puzzling mass term gets canceled by a bulk contribution. Our analysis
shows unambiguously that the cascading framework is ghost-free, at least perturbatively,
provided that the codimension-2 brane is endowed with sufficiently large tension. Whether the theory is
stable non-linearly remains of course an open issue that deserves further study.
The full 6D calculation we perform here is necessarily delicate because of the fact that the background geometry contains a conical singularity at the location of the 3-brane on which we intend to localize matter. This is the usual conical singularity associated with the tension of a codimension-2 defect.
The orbifold symmetry imposed across the codimension-1 brane creates a $\mathds{Z}_2$ reflection axis in the background conical solution which is further felt at the level of perturbations. Although there are many different ways of representing the background geometry, none is particularly straightforward for performing the perturbative analysis. To get a sense of the difficulty, consider the usual way of writing a conical defect,
\begin{equation}
{\rm d}s^2={\rm d}r^2 + r^2(1-\delta)^2 {\rm d}\theta^2+\eta_{\mu\nu} {\rm d}x^{\mu}{\rm d}x^{\nu}\,.
\label{1stcoord}
\end{equation}
The angular variable $\theta$ runs from 0 to $2\pi$, and $\delta$ is the deficit angle proportional to the tension.
The 3-brane is understood to be located at $r=0$. The problem with this coordinate system is that the orbifold brane is located on two disjoint surfaces: $\theta=0$ and $\theta=\pi$. Even if the geometry is smoothed out at $r=0$, it is clear that this metric will provide a poor description of physics on the codimension-1 brane near the codimension-2 brane.
Alternatively, the geometry can be put in the form
\begin{equation}
{\rm d}s^2=e^{\phi(y,z)}\left({\rm d}y^2+{\rm d}z^2\right)+\eta_{\mu\nu} {\rm d}x^{\mu}{\rm d}x^{\nu}\,, \quad\,\,\, \text{with } \,\, \phi(y,z)=-\delta \ln (y^2+z^2)\,.
\label{2ndcoord}
\end{equation}
One advantage of this coordinate system is that the codimension-1 brane is now at $z=0$. The drawback is clear, however:
the conformal factor $\phi$ is singular at the location of the codimension-2 brane. One can deal with this by regulating the conformal factor, {\it e.g.}, $\phi \rightarrow -\delta \ln (y^2+z^2+\varepsilon^2)$, but this need to regulate the background geometry comes at the price of introducing extra terms in the perturbation equations which may or may not be negligible
in the limit $\varepsilon \rightarrow 0$. Although the final equations shall be independent of $\varepsilon$, it is necessary to introduce the artifice of regularization to obtain them.
Instead we will use a different form of the background metric, one that we have found is best suited for the analysis of perturbations:
\begin{equation}
\mathrm{d} s^2=\left(1+\beta^2 \left(1-\epsilon(y)^2\right)\right)\mathrm{d} z^2+\left(\mathrm{d} y+\beta \epsilon(y) \mathrm{d} z\right)^2 +\eta_{\mu\nu} {\rm d}x^{\mu}{\rm d}x^{\nu}\,.
\label{ourcoord}
\end{equation}
In this coordinate system, the codimension-1 brane is still at $z=0$, and we work in the half-picture $z> 0$. The above metric is understood to be orbifolded at $z=0$. The parameter
$\beta$ is related to the 3-brane tension. Meanwhile, $\epsilon(y)$ is a regulating function with the following properties: $\epsilon(\infty)=1$, $\epsilon(-y)=-\epsilon(y)$
and $\epsilon'(y)=2\delta_{\epsilon}(y)$, where $\delta_{\epsilon}(y)$ is an explicit regularization of the Dirac delta function. The upshot of this coordinate system is that
the induced metric on the codimension-1 brane is simply Minkowski space-time, and similarly for the induced geometry on the codimension-2 brane at $z=y=0$.
As we will see below, this greatly simplifies the analysis of perturbations.
As always the presence of ghosts in a theory can be determined by checking the sign of the kinetic term of all propagating degrees of freedom in the action. In the case of gravitational theories, however, the inevitable presence of gauge modes and constraints complicates matters. Only after the gauge has been completely fixed and the constraints fully solved will the action reduce to the true physically propagating degrees of freedom whose kinetic terms can be inspected.
As is well known, before the constraints have been solved the action may well contain wrong-sign kinetic terms, {\it e.g.} the famous conformal factor problem of Euclidean quantum gravity.
One way to bypass this problem is to compute the coupling to a conserved source. Since gauge invariance enforces conservation of stress energy, any gauge degrees of freedom will not couple to a conserved stress-energy source, and, in particular, will not contribute to the single-graviton exchange amplitude. In the following we shall do precisely this to verify the absence of ghost degrees of freedom.
The paper is organized as follows. In Sec.~\ref{background}, we describe the background geometry~(\ref{ourcoord}), starting from the general ADM form of
the 6D metric. In Sec.~\ref{pertn1}, we study metric perturbations, focusing first on 4D vector and tensor modes. We then turn to 4D scalar perturbations in
Sec.~\ref{pertn2}, and derive the lower bound on the tension. In Sec.~\ref{externalcoupling}, we discuss couplings to external sources.
We conclude with a summary of the results and discuss future avenues in Sec.~\ref{conclu}. The Appendices include a review of the decoupling
calculation (Sec.~\ref{decoup}), an argument for why certain naively ghostly terms do not contribute to the imaginary part of the exchange amplitude (Sec.~\ref{sigma1}), and an argument in (Sec.~\ref{sigma2}) for why one of the scalar modes,
which at first sight looks problematic and ghostly, is in fact pure gauge and does not contribute to the exchange amplitude. Our notation throughout is as follows: $A,B,\ldots$ denote 6D space-time indices, $a,b\ldots$ are 5D indices, and $\mu,\nu,\ldots$ are 4D indices. The coordinates along the worldvolume of the 3-brane are denoted
by $x^\mu$, whereas the extra dimensions have coordinates $y$ and $z$. The 4-brane is located at $z=0$, and the 3-brane at $y=z=0$.
\section{Cascading Setup}
\label{background}
Our approach is to start with the full 6D action for the cascading setup written in the radial ADM formalism, and then perturb this action to quadratic order.
We work in the ``half (or reduced)-picture", where we restrict ourselves to the $z>0$ side of the $\mathds{Z}_2$-symmetric 4-brane.
Crucially this implies that all quantities evaluated at $z=0$ are understood in the following sense: $\phi(z=0)=\lim_{z \rightarrow 0^+} \phi(z)$.
In the ADM form, with $z$ playing the role of a ``time" variable, the 6D metric in the region $z>0$ is
\begin{equation}
\label{6Dbackground}
\mathrm{d} s^2= \gamma_{AB}\mathrm{d} x^A \mathrm{d} x^B=N^2 \mathrm{d} z^2+g_{ab}(\mathrm{d} x^a+ N^a \mathrm{d} z)(\mathrm{d} x^b+N^b \mathrm{d} z)\,.
\end{equation}
In what follows, covariant objects are constructed using the 5D metric $g_{ab}$ induced on constant-$z$ surfaces.
The 6D part of the action, including the appropriate Gibbons-Hawking boundary term, is then
\begin{eqnarray}
S_6=\frac{M_6^4}{2}\int^{+}\hspace{-5pt}\mathrm{d} z \int \mathrm{d}^5x \, \mathcal{L}_6\,,
\end{eqnarray}
with $\int^+ \mathrm{d} z=\int_0^{+\infty} \mathrm{d} z$, and where
\begin{eqnarray}
\mathcal{L}_6\equiv \sqrt{-g_6}\, \( R^{(5)}+ {K^a_{\;a}}^2-K^a_{\;b} K^b_{\;a}\)\,.
\end{eqnarray}
The full action is given by
\begin{eqnarray}
\nonumber
S &=& \int^{+} \mathrm{d}^6 x \left( \frac{M_6^4}{2} \mathcal{L}_6 + {\mathcal L}_{\rm matter}\right)+\frac{M_5^3}{2} \int_{z=0} \mathrm{d}^5 x \sqrt{-g_5} R_5 \\
& & + \int_{z=y=0} \mathrm{d}^4 x \sqrt{-g_4}\(\frac{M_4^2}{2} R_4-\lambda \),
\end{eqnarray}
where $\mathcal{L}_{\rm matter}$ is the Lagrangian for the external sources which may be distributed in 6D, or localized on either brane. In what follows we will compute the coupling for an arbitrary conserved 6D, 5D or 4D source.
\subsection{Background Solution}
{\label{backgroundsolution}}
There are several ways to write the background solution. However, as discussed earlier,
great care must be taken because of the singular nature of the geometry. Consider the following background metric
\begin{equation}
\mathrm{d} s^2=\bar \gamma_{AB}\mathrm{d} x^A \mathrm{d} x^B=(1+\beta^2 (1-\epsilon(y)^2))\mathrm{d} z^2+(\mathrm{d} y+\beta \epsilon(y) \mathrm{d} z)^2 + \eta_{\mu\nu} {\rm d}x^{\mu}{\rm d}x^{\nu}\,.
\label{ourcoord2}
\end{equation}
By comparing with~(\ref{6Dbackground}), we can read off the background expressions for the ADM variables:
\begin{eqnarray}
\nonumber
\bar N^{2}&=&1+\beta^2(1-\epsilon(y)^2)\;; \\
\nonumber
\bar N^{y}&=& \beta \epsilon(y)\;; \qquad \bar N^{\;\mu}= 0\;; \\
\overline{g}_{ab}&=&\eta_{ab}\,.
\end{eqnarray}
The regulating function $\epsilon(y)$ is such that $\epsilon(\infty)=1$, $\epsilon(-y)=-\epsilon(y)$ and $\epsilon'(y)=2\delta_{\epsilon}(y)$, where $\delta_{\epsilon}(y)$ is an explicit regularization of the delta function. The induced metrics on the codimension-1 ($z=0$) and codimension-2 ($z=y=0$) branes are both flat Minkowski space.
The 6D Riemann tensor vanishes on the background for $z>0$, as it should. For instance, since $\bar N^{2}+\bar N_{y}^{2}=1+\beta^2$, the Ricci scalar
is given by
\begin{eqnarray}
\bar{R}_6=-\frac{1}{\bar N} \partial_y\(\frac{ \partial_y \left(\bar N^{2}+\bar N_{y}^{2}\right)}{\bar N}\)=0\,.
\end{eqnarray}
Because this metric is orbifolded at $z=0$, it follows from the Isra\"el junction conditions that~(\ref{ourcoord2}) correctly describes the metric of a codimension-2 object with tension localized on the orbifold plane. Explicitly, the Isra\"el junction conditions are
\begin{equation}
M_6^4 (K \delta^a_{\;b} - K^a_{\;b})\Big|_{z=0}=T^a_{\;b}-M_5^3 G^a_{\;b}\,.
\label{Israel1}
\end{equation}
In the background, we have $T^a_{\;b}=-\lambda\delta(y)$ and $G^a_{\;b}=0$. Meanwhile,
in the gauge \eqref{6Dbackground}, the extrinsic curvature on the surface $z=0$ is
\begin{eqnarray}
K_{ab}&=&\frac{1}{2N} \(\partial_z g_{ab}-\nabla_{a} N_b-\nabla_b N_a \) \nonumber \\
&=& \frac{1}{2N} \( \partial_z g_{a b}-g_{a c} \partial_{b}N^{c}-g_{b c} \partial_{a}N^{c}-N^{c}\partial_{c} g_{a b}\)\nonumber\\
&=&-\frac{\partial_y \bar N_{y}}{\bar N} \delta^y_{\; a} \delta^y_{\; b}\,.
\end{eqnarray}
At the background level, the Isra\"el matching conditions~(\ref{Israel1}) therefore read
\begin{eqnarray}
\label{Israel}
K_{yy}=-\frac{\partial_y \bar N_{y}}{\bar N} =-\frac{\beta \epsilon'(y)}{\sqrt{1+\beta^2(1-\epsilon^2(y))}}=-\frac{\lambda}{M_6^4{}}\delta(y)\,.
\end{eqnarray}
Since $\epsilon'(y) = 2\delta_\epsilon(y)\rightarrow \delta(y)$, the extrinsic curvature does indeed encode the codimension-2 source.
The relation between the brane tension $\lambda$ and the required value for $\beta$ requires some care, but can be determined
unambiguously by integrating the extrinsic curvature,
\begin{eqnarray}
\int _{-\infty}^{+\infty}\hspace{-6pt}\mathrm{d} y \frac{\partial_y \bar{N}_y}{\bar N} =\int_{-\infty}^{+\infty}\hspace{-6pt}\mathrm{d} y \frac{\beta \epsilon'(y)}{\sqrt{1+\beta^2(1-\epsilon^2(y))}}
=\int_{-1}^{1}\frac{\beta \, \mathrm{d} \epsilon(y)}{\sqrt{1+\beta^2(1-\epsilon^2(y))}}=2\arctan \beta \,.
\end{eqnarray}
Thus the relation is
\begin{eqnarray}
\lambda = 2 M_6^4{}\arctan \beta\,.
\end{eqnarray}
In particular, we recover the standard result that this static solution can only hold a maximal positive codimension-2 tension: $\lambda< \pi M_6^4{}$. (This is one-half of the usual condition because the $\mathds{Z}_2$ orbifolding projects out half of the space.)
\section{Tensor and Vector Perturbations}
\label{pertn1}
To study perturbations, we perform a scalar-vector-tensor decomposition of the full 6D metric $\gamma_{AB}=\bar \gamma_{AB}+\delta \gamma_{AB}$
with respect to the 4D Lorentz group. The tensor and vector modes are straightforward as none of them couple to the positions of the branes and are given by a simple generalization of the massless scalar field. In particular, the tensor and vector sectors are manifestly ghost free, regardless of the value of the tension on the codimension-2 brane.
\paragraph*{Tensors:}
The metric for the tensor perturbations takes the form
\begin{eqnarray}
\nonumber
\delta \gamma_{AB}\mathrm{d} x^A \mathrm{d} x^B &=& h^{\rm T}_{\mu \nu}\mathrm{d} x^\mu\mathrm{d} x^\nu \nonumber \,,
\end{eqnarray}
where $\eta^{\mu\nu} h^{\rm T}_{\mu\nu}=0$ and $\partial^{\mu} h^{\rm T}_{\mu\nu}=0$.
The action for the tensor perturbations is given by
\begin{eqnarray}
S&=&\frac{M_6^4}{8} \int^+ \mathrm{d} z \int \mathrm{d}^5 x \( -\bar N \partial_a h_{\rm T}^{\mu \nu} \partial^a h^{\rm T}_{\mu\nu}-\frac{1}{\bar N}\( \mathcal{L}_n h_{\rm T}^{\mu\nu} \)^2 \)\,, \\
& +&\frac{M_5^3}{8} \int_{z=0} \mathrm{d}^5 x \( h_{\rm T}^{\mu \nu} \Box_5 h^{\rm T}_{\mu\nu}\)+\frac{M_5^3}{8} \int_{z=y=0} \mathrm{d}^4 x \( h_{\rm T}^{\mu \nu} \Box_4 h^{\rm T}_{\mu\nu} \) \, ,\nonumber
\end{eqnarray}
where $\mathcal{L}_n$ is proportional to the Lie derivative: $\mathcal{L}_n X=\partial_z X-\bar{N}_y \partial_y X$. Notice that $\mathcal{L}_n$ and $\partial_y$ do not commute:
\begin{eqnarray}
\partial_y \mathcal{L}_n X=\mathcal{L}_n \partial_y X-\partial_y \bar{N}_y \, \partial_y X\,.
\label{Lniden}
\end{eqnarray}
The tensor modes constitute 5 degrees of freedom, each of which satisfies the same equation which is manifestly ghost free since the action is only composed of a sum of positive kinetic terms. These five spin-2-degrees of freedom can be further decomposed into two helicity-2, two helicity-1 and one helicity-0 modes. The helicity-0 degree of freedom is expected to exhibit the Vainshtein mechanism as in usual massive gravity~\cite{vainshtein,nima,babichev}, DGP model~\cite{lpr,ddgv,gruzinov,lue,greg} and general theories of resonance graviton~\cite{degrav,gd}.
\paragraph*{Vectors:}
For the vector perturbations the metric takes the form
\begin{eqnarray}
\delta \gamma_{AB}\mathrm{d} x^A \mathrm{d} x^B &=& 2 \( A_{\mu}+\bar N_y B_{\mu}\) \mathrm{d} x^{\mu}\mathrm{d} z+2 B_{\mu} \mathrm{d} x^{\mu}\mathrm{d} y +\( \partial_{\mu} C_{\nu}+ \partial_{\nu} C_{\mu}\)\mathrm{d} x^\mu\mathrm{d} x^\nu \,,
\end{eqnarray}
where the vectors $A_{\mu},B_{\mu}$ and $C_{\mu}$ are all transverse: $\partial_{\mu}A^{\mu}=\partial_{\mu}B^{\mu}=\partial_{\mu}C^{\mu}=0$. There is sufficient gauge freedom in the vector sector to set $B_{\mu}=0$. The action for the remaining perturbations is then given by
\begin{eqnarray}
S&=&\frac{M_6^4}{8} \int^+ \mathrm{d}^6x \( -\frac{2}{\bar N}\( \partial_y {A}_{\mu}\)^2+\frac{2}{\bar N} \(A_{\mu}-\mathcal{L}_n C_{\mu}\) \Box_4 \(A^{\mu}-\mathcal{L}_n C^{\mu}\)+2\bar N \partial_y C^{\mu}\Box_4 \partial_y C_{\mu} \) \nonumber \\
&&+\frac{M_5^3}{8} \int_{z=0} \mathrm{d}^5 x \(2 \partial_y C^{\mu}\Box_4 \partial_y C_{\mu}\) +\int^+ \mathrm{d}^6x \, \bar N \(A_\mu T^{\mu z}-C_\mu \partial_\nu T^{\mu\nu}\) \, ,
\end{eqnarray}
where $T^{AB}$ is an external 6D source, (which can include localized contributions on either of the branes).
Note that neither vector field has a 4D source and so we expect that the vectors will be appropriately smooth at $y=0$. With this in mind, we may neglect the mild $y$-dependence of $\bar N$, since its departure from $\bar N =1$ will be of zero measure in the integral. Then, on performing the following local field redefinition,
\begin{equation}
\hat{C}_{\mu}=\partial_y C_{\mu}\,; \, \quad \hat{A}_{\mu}=A_{\mu}-\mathcal{L}_n C_{\mu}\,,
\end{equation}
the kinetic terms are diagonal and manifestly positive definite and so is the coupling to the source
\begin{eqnarray}
S_{\rm kin}&=&\frac{M_6^4}{8} \int^+ \mathrm{d}^6x \( \frac{2}{\bar N} \hat A_\mu \Box_4 \hat A^\mu+2\bar N \hat C^{\mu}\Box_4 \hat C_{\mu} \)
+\frac{M_5^3}{8} \int_{z=0} \mathrm{d}^5 x \(2 \hat C^{\mu}\Box_4 \hat C_{\mu}\) \\
S_{\rm sources}&=&-\int^+ \mathrm{d}^6x\, \bar N \hat C_\mu \(T^{\mu y}+\bar{N}_y T^{\mu z}\) \, ,
\end{eqnarray}
where we used conservation of energy along the transverse direction $\nabla_A T^{A\mu}=0$.
Notice that $\hat{A}_{\mu}$ lives entirely in 6D, whereas $\hat{C}_{\mu}$ has a kinetic term both in 6D and 5D. Crucially all the kinetic terms are positive, as required, confirming that there are no ghosts present in the vector sector.
\section{Scalar Perturbations}
\label{pertn2}
The scalar perturbations are tricky because the positions of the brane transform as 4D scalars. We choose to utilize the bulk gauge freedom to work in a gauge where the codimension-1 and -2 branes remain at fixed position, respectively $z=0$ and $z=y=0$.
At the perturbed level, the metric has seven 4D scalar modes:
\begin{eqnarray}
\nonumber
\delta \gamma_{AB}\mathrm{d} x^A \mathrm{d} x^B &=& h_{zz}\mathrm{d} z^2+2 h_{zy}\mathrm{d} z \mathrm{d} y + h_{yy}\mathrm{d}
y^2+2 h_{\mu z}\mathrm{d} x^\mu\mathrm{d} z+ 2 h_{\mu y}\mathrm{d} x^\mu\mathrm{d} y \\
&+&\(\pi \eta_{\mu \nu} +\partial_\mu \partial_\nu \varpi\)\mathrm{d} x^\mu\mathrm{d} x^\nu \nonumber \,.
\end{eqnarray}
One can set one of these modes to zero by fixing the gauge while
keeping the brane positions unperturbed. We use this freedom to set
$\varpi=0$, such that $h_{\mu\nu} = \pi \eta_{\mu\nu}$. This is an analogue
of conformal Newtonian/longitudinal gauge used in cosmological perturbation theory.
We are therefore left with six scalar degrees of freedom, which will be
denoted by $\chi, \phi, V, \sigma, \tau$ and $\pi$:
\begin{eqnarray}
\nonumber
h_{ab}\mathrm{d} x^a \mathrm{d} x^b & =& \pi \eta_{ab}\mathrm{d} x^a \mathrm{d} x^b+V \mathrm{d} y^2+2
\partial_\mu \tau\ \mathrm{d} x^\mu \mathrm{d} y\\
\nonumber
h_{zz}&=& 2\bar N^{\;2} \phi+ \bar N^{a} \bar N^{b} h_{ab}+2\bar N^{a} A^b
\eta_{ab}\\
h_{az}&=&A^b\eta_{ab}+\bar N^{b} h_{ab}\,,
\end{eqnarray}
where $A_a\mathrm{d} x^a = \partial_a \sigma \mathrm{d} x^a+\chi \mathrm{d} y$.
Note that $A^a$ is the perturbation in the shift vector, $N^a = \bar N^a + A^a$, while
$\phi$ is the perturbation is the lapse function, $N = \bar N(1+\phi)$.
To quadratic order in perturbations, the full action is given by
\begin{eqnarray}
\label{Linitial}
\nonumber
S_{\rm quad} &=& \frac{M_6^4{}}{2} \int^+ \mathrm{d}^6x \mathcal{L}_6
- \frac {3 M_5^3{}} 4 \int_{z=0} \mathrm{d}^5x \, \pi \(\Box_4 V+2 \Box_5 \pi-2 \Box_4\partial_y
\tau\) \\
&-&\int_{z=y=0} \mathrm{d}^4x \(\frac {3M_4^2{}} 4\, \pi \Box_4
\pi+ \lambda \pi^2 \)\,.
\end{eqnarray}
The kinetic term for $\pi$ on the codimension-2 brane has manifestly the wrong sign. However, as we will see,
the presence of brane tension $\lambda$ can change the sign of this kinetic term, after integrating
out the non-dynamical degrees of freedom. Meanwhile, the mass term $\lambda\pi^2$ localized on the codimension-2 brane,
which seemingly breaks the shift symmetry, will cancel out after integrating by parts a bulk contribution.
To see how this works in detail, let us expand the 6D action to quadratic order, still focusing on the region $z>0$.
It is convenient to perform the following field redefinitions
\begin{eqnarray}
\tilde\chi&=&\chi-(\mathcal{L}_n +\partial_y \bar{N}_y)\tau-2\frac{\partial_y \bar N}{\bar N}\tilde \sigma + \frac{4}{\Box_4} \mathcal{L}_n \partial_y \pi\\
\Box_4 \tilde \sigma &=& \Box_4 \sigma-2 \mathcal{L}_n\pi=-\bar N \delta K^\mu_{\; \mu}\\
\Box_4 \tilde \tau &=& \Box_4 \tau-2 \partial_y \pi\,.
\end{eqnarray}
In terms of these new field variables, the result for ${\cal L}_6$ at quadratic order is then
\begin{eqnarray}
\label{6Dcurvature}
\mathcal{L}_6&=&\frac {1}{2\bar N} \tilde\chi\Box_4\tilde\chi
-\bar N \( V+4\pi -2 \partial_y \tilde \tau+2 \frac{\partial_y \bar{N}_y}{\bar N^2}\tilde \sigma\)\Box_4 \phi \nonumber \\
&&-\frac {1}{\bar N} \partial_y \bar{N}_y \mathcal{L}_n (\tilde \tau \Box_4 \tilde \tau)
+3 \bar N \pi (\partial_y^2-\Box_4) \pi-\frac{3}{\overline N} (\mathcal{L}_n \pi)^2
-2\frac{\partial_y \bar{N}_y}{\bar N}\partial_z \pi^2\nonumber\\
&&
-\frac{1}{\bar N}\Box_4 \tilde \sigma (\mathcal{L}_n - \partial_y \bar{N}_y) (V+4\pi)-\frac{3}{\bar N} \partial_y \bar{N}_y \pi \Box_4 \tilde \sigma
+\frac{2}{\bar N} \Box_4 \tilde \sigma \partial_y \mathcal{L}_n \tilde \tau
+3 \bar N \pi \Box_4 \partial_y \tilde \tau\nonumber\\
&&
-\frac 3 2 \bar N \pi \Box_4 V
+2 \frac{\partial_y \bar{N}_y}{\bar N}\frac{\partial_y \bar N}{\bar N} \tilde \sigma \Box_4 \tilde \tau-\frac{\partial_y^2 \bar N}{\bar N}\tilde \sigma \Box_4 \tilde \sigma
+\partial_y \bar N \tilde \tau \Box_4 (V+4\pi) \,.
\end{eqnarray}
The issue of the $\pi$ mass term in~(\ref{Linitial}) can be immediately resolved. The last term in
the second line of~\eqref{6Dcurvature}, when integrated by parts, generates a mass term for $\pi$
on the 3-brane that precisely cancels that in~\eqref{Linitial}:
\begin{eqnarray}
-2M_6^4{}\int^+\hspace{-3pt}\mathrm{d} z \int \mathrm{d} y\frac{\partial_y \bar{N}_y}{\bar N}\partial_z \pi^2&=&2M_6^4{}\int^+ \hspace{-3pt} \mathrm{d} z \int \mathrm{d} y \(\partial_z \frac{\partial_y \bar{N}_y}{\bar N}\) \pi^2
-2M_6^4{}\int \mathrm{d} y \Big[\frac{\partial_y \bar{N}_y}{\bar N} \pi^2 \Big]_{z=0}^{\infty}\nonumber\\
&=&2M_6^4{}\int_{z=0} \mathrm{d} y \frac{\partial_y \bar{N}_y}{\bar N} \pi^2=\lambda \pi^2\,.
\end{eqnarray}
This cancelation makes manifest the shift symmetry of $\pi$. We now proceed to decipher the $\pi$ kinetic term,
which requires a few field redefinitions and integrating out non-dynamical fields.
\subsection{Lagrange multipliers $\phi$ and $V$}
As it stands,~(\ref{6Dcurvature}) is linear in $\phi$ and $V$.
Varying the action with respect to either of these therefore yields a constraint
on the other degrees of freedom. We choose to vary with respect to $\phi$
and interpret the result as a constraint for $V$. Upon substituting $V$
by its constrained value back into the action, we will see that the action picks up a quadratic term in $\tilde \sigma$.
The expression for $V$ that follows from the $\phi$ variation is
\begin{eqnarray}
\label{EqforV}
V=-4 \pi+ 2 \partial_y \tilde \tau-\frac{2\partial_y \bar{N}_y}{\bar N^2} \tilde \sigma\,.
\end{eqnarray}
In the presence of an external source, there is an additional $T_6^{zz}$ that enters in this equation;
we shall take this into account in Sec.~\ref{externalcoupling} when we compute the relevant couplings --- see Eq.~(\ref{newV}).
Substituting this expression for $V$ in the 6D action, we can simplify the last two lines in~\eqref{6Dcurvature} as follows
\begin{eqnarray}
\nonumber
A&\equiv & -\frac{1}{\bar N}\Box_4 \tilde \sigma (\mathcal{L}_n - \partial_y \bar{N}_y) (V+4\pi)-\frac{3}{\bar N} \partial_y \bar{N}_y \pi \Box_4 \tilde \sigma
+\frac{2}{\bar N} \Box_4 \tilde \sigma \partial_y \mathcal{L}_n \tilde \tau
+3 \bar N \pi \Box_4 \partial_y \tilde \tau\nonumber\\
&&
-\frac 3 2 \bar N \pi \Box_4 V
+2 \frac{\partial_y \bar{N}_y}{\bar N}\frac{\partial_y \bar N}{\bar N} \tilde \sigma \Box_4 \tilde \tau-\frac{\partial_y^2 \bar N}{\bar N}\tilde \sigma \Box_4 \tilde \sigma
+\partial_y \bar N \tilde \tau \Box_4 (V+4\pi) \nonumber\\
&=& -\frac 1 \bar N\Box_4 \tilde \sigma (\mathcal{L}_n - \partial_y \bar{N}_y) \(2 \partial_y \tilde \tau-\frac{2\partial_y \bar{N}_y}{\bar N^2} \tilde \sigma\)-\frac 3 \bar N\partial_y \bar{N}_y \pi \Box_4 \tilde \sigma
+\frac 2 \bar N \Box_4 \tilde \sigma \partial_y \mathcal{L}_n \tilde \tau \nonumber
\\
&&+3 \bar N \pi \Box_4 \partial_y \tilde \tau- \frac 3 2 \bar N \pi \Box_4 \(-4 \pi+ 2 \partial_y \tilde \tau-\frac{2\partial_y \bar{N}_y}{\bar N^2} \tilde \sigma\)
+2\frac{\partial_y \bar{N}_y}{\bar N}\frac{\partial_y \bar N}{\bar N}\tilde \sigma \Box_4 \tilde \tau\nonumber\\
&&-\frac{\partial_y^2 \bar N}{\bar N^2}\tilde \sigma \Box_4 \tilde \sigma
+\partial_y \bar N \tilde \tau \Box_4 \(2 \partial_y \tilde \tau-2 \frac{\partial_y \bar{N}_y}{\bar N^2}\tilde \sigma\) \nonumber \\
&=& 6\bar N \pi \Box_4 \pi-\partial_y^2 \bar N \tilde \tau\Box_4 \tilde \tau+\frac 2 \bar N \Box_4 \tilde \sigma (\mathcal{L}_n-\partial_y \bar{N}_y) \(\frac{\partial_y \bar{N}_y}{\bar N^2} \tilde \sigma\)-\frac{\partial_y^2\bar N}{\bar N}\tilde \sigma \Box_4 \tilde \sigma\,,
\end{eqnarray}
where in the last step we have used $(\mathcal{L}_n-\partial_y \bar{N}_y)\partial_y \tilde \tau=\partial_y(\mathcal{L}_n \tilde \tau)$ from~(\ref{Lniden}).
We can further simplify this expression by integrating by parts:
\begin{eqnarray}
\int^+ \hspace{-5pt}\mathrm{d}^6x A &=&\int^+ \hspace{-5pt}\mathrm{d}^6x\(6\bar N \pi \Box_4 \pi-\partial_y^2 \bar N \tilde \tau\Box_4 \tilde \tau+ \frac 2 \bar N \Box_4 \tilde \sigma (\mathcal{L}_n-\partial_y \bar{N}_y) \(\frac{\partial_y \bar{N}_y}{\bar N^2} \tilde \sigma\)
-\frac{\partial_y^2\bar N}{\bar N}\tilde \sigma \Box_4 \tilde \sigma\)\nonumber\\
&=&\int^+ \hspace{-5pt}\mathrm{d}^6x \(6\bar N \pi \Box_4 \pi-\partial_y^2 \bar N \tilde \tau\Box_4 \tilde \tau - \frac {\tilde \sigma \Box_4 \tilde \sigma}{2\bar N^2}\, \partial_y\left[\frac{\partial_y (\bar N^2+\bar{N}_y^2)}{\bar N}\right]\) - \int_{z=0} \hspace{-5pt}\mathrm{d}^5x \frac{\partial_y \bar{N}_y}{ \bar N^3}\tilde \sigma \Box_4 \tilde \sigma\nonumber \\
&=&\int^+ \hspace{-5pt}\mathrm{d}^6x \(6\bar N \pi \Box_4 \pi-\partial_y^2 \bar N \tilde \tau\Box_4 \tilde \tau\)- \int_{z=0} \mathrm{d}^5x \frac{\partial_y \bar{N}_y}{ \bar N^3}\tilde \sigma \Box_4 \tilde \sigma \,,
\label{Asimp}
\end{eqnarray}
where the last follows because $\bar N^2+\bar{N}_y^2=1+\beta^2 = {\rm constant}$.
Substituting~(\ref{Asimp}) back into~(\ref{6Dcurvature}), we note that the quadratic terms in $\tilde \tau$ simplify through integration by parts:
\begin{eqnarray}
-\int^+\hspace{-5pt}\mathrm{d} z \, \(\partial_y^2 \bar N+\frac{\partial_y \bar{N}_y}{\bar N}\mathcal{L}_n \)(\tilde \tau \Box_4 \tilde \tau)
&=&-\frac 12 \int^+\hspace{-5pt}\mathrm{d} z \, \partial_y \(\frac{\partial_y(\bar N^2+\bar{N}_y^2)}{\bar N}\)\tilde \tau \Box_4 \tilde \tau
-\left[\frac{\partial_y \bar{N}_y}{\bar N}\tilde \tau \Box_4 \tilde \tau\right]_0^{\infty} \nonumber\\
&=&\int_{z=0} \mathrm{d}^5x \frac{\partial_y \bar{N}_y}{\bar N} \tilde \tau \Box_4 \tilde \tau \,.
\end{eqnarray}
Combining all of these results, the 6D part of the quadratic action reduces to
\begin{eqnarray}
\label{6Daction2}
\int^+ \mathrm{d}^6 x \mathcal{L}_6&=&\int^+ \mathrm{d} z \mathrm{d}^5 x \(\frac {1}{2\bar N}\tilde\chi\Box_4\tilde\chi
+3 \bar N \pi \Box_5 \pi-\frac 3\bar N (\mathcal{L}_n \pi)^2\) \nonumber \\
&+&\int_{z=0} \mathrm{d}^5x \frac{\partial_y \bar{N}_y}{\bar N}\(\tilde \tau \Box_4 \tilde \tau-\frac{1}{\bar N^2}\tilde \sigma \Box_4 \tilde \sigma\)\,.
\end{eqnarray}
Similarly, we must also substitute~(\ref{EqforV}) for $V$ into the 5D part of~(\ref{Linitial}):
\begin{eqnarray}
\label{action5}
\mathcal{L}_5= \left.\frac 12 \sqrt{-g_5}R_5\right\vert_{\rm quadratic}=\frac 32 \pi \Box_5 \pi +\frac 32 \frac{\partial_y \bar{N}_y}{\bar N^2} \pi \Box_4 \tilde \sigma \,.
\end{eqnarray}
As we will see, the above kinetic mixing term between $\pi$ and $\tilde \sigma$ is what provides an additional 4D kinetic term for $\pi$.
Putting everything together, and using the background Isra\"el Matching condition \eqref{Israel}, $M_6^4{} \partial_y \bar{N}_y/\bar N = \lambda\delta(y)$,
the complete action for the scalar modes is given by
\begin{eqnarray}
S_{\rm quad} &=&\frac{M_6^4{}}{2} \int^+ \mathrm{d}^6 x \left[ \frac {1}{2\bar N}\tilde\chi\Box_4\tilde\chi
+3\(\bar N \pi \Box_5 \pi-\frac 1\bar N (\mathcal{L}_n \pi)^2\) \right] \nonumber \\
&+&M_5^3{} \int_{z=0} \mathrm{d}^5x\, \frac 32 \pi \Box_5 \pi\nonumber \\
&+& M_4^2{} \int_{z=y=0} \mathrm{d}^4 x \left[-\frac{3}{4}\pi \Box_4 \pi +
\frac{\lambda}{2M_4^2{}}
\(\tilde \tau \Box_4 \tilde \tau-\frac{1}{\bar N^2}\tilde \sigma \Box_4 \tilde \sigma+\frac{3}{m_6 \bar N}\pi \Box_4 \tilde \sigma\)\right]\,.
\end{eqnarray}
\subsection{$\tilde \sigma$ degree of freedom}
The last step consists of diagonalizing the kinetic matrix for $\pi$ and $\tilde \sigma$. This is achieved by the following
change of variable
\begin{eqnarray}
\label{sigmabar}
\hat \sigma=\tilde \sigma - \frac{2 \bar N}{4 m_6}\, \pi\,,
\end{eqnarray}
in terms of which the action becomes
\begin{eqnarray}
\label{action3}
S_{\rm quad} &=&\frac{M_6^4{}}{2} \int^+ \mathrm{d}^6 x \left[ \frac {1}{2\bar N}\tilde\chi\Box_4\tilde\chi
+3\(\bar N \pi \Box_5 \pi-\frac 1\bar N (\mathcal{L}_n \pi)^2\) \right] \nonumber \\
&+&M_5^3{} \int_{z=0} \mathrm{d}^5x\, \frac 32 \pi \Box_5 \pi\nonumber \\
&+&M_4^2{} \int_{z=y=0} \mathrm{d}^4 x \left[ \frac 3 4 \(\frac{3\lambda}{2m_6^2M_4^2{}}-1\) \pi\Box_4 \pi +
\frac{\lambda}{2M_4^2{}}
\(\tilde \tau \Box_4 \tilde \tau-\frac{1}{\bar N^2}\hat\sigma \Box_4 \hat\sigma\) \right] \,.
\end{eqnarray}
This is our main result. Through a series of field redefinitions, and after integrating out auxiliary fields, the resulting
kinetic term for $\pi$ on the codimension-2 brane acquires a $\lambda$-dependent contribution:
\begin{eqnarray}
\mathcal L_2^{{\rm kin}}=\frac{3M_4^2{}}{4} \(\frac {3\lambda}{2m_6^2M_4^2{}}-1\)\pi \Box_4
\pi\,,
\end{eqnarray}
Thus the $\pi$ mode is healthy as long as the tension is larger than the lower bound
\begin{eqnarray}
\lambda_{\rm min}=\frac{2}{3}M_4^2{} m_6^2\,.
\end{eqnarray}
\section{Coupling to an external source}
\label{externalcoupling}
It is immediately apparent from~(\ref{action3}) that $\pi$, $\tilde\chi$ and $\tilde \tau$ are manifestly unitary. However, $\hat \sigma$ appears as a ghost and would seem to present a serious concern. (This issue never arose in~\cite{cascade1} since this mode disappears in the decoupling limit.) As already discussed, though, it is common in the analysis of gravitational systems to find apparently ghostly degrees of freedom that nevertheless decouple when a physical, gauge-invariant calculation is performed.
In this case we will see that neither $\hat \sigma$ nor $\tilde \tau$ couple to conserved matter, and consequently these modes are pure gauge, at least to this order in perturbation theory. To see this, let us introduce the matter content in the action
\begin{eqnarray}
\int^+ \mathrm{d} z \int \mathrm{d} ^5x \, \mathcal{L}_{\rm matter}=\int^+ \mathrm{d} z\int \mathrm{d} ^5 x \, \frac{\bar N}{2} h_{AB} T^{AB}_6\,,
\end{eqnarray}
where we consider a 6D stress-energy tensor in all generality. In particular the total stress-energy $T^{AB}_6$ can include terms localized on each of the branes:
\begin{eqnarray}
T^{AB}_6=\mathcal{T}^{AB}_6+\bar N^{-1}\delta_+(z)\, \mathcal{T}^{ab}_5 \delta_a^A\delta_b^B+
\bar N^{-1}\delta(y)\delta_+(z)\, \mathcal{T}^{\mu\nu}_4 \delta_\mu^A\delta_\nu^B\,,
\end{eqnarray}
where $\delta_+(z)$ is the delta function normalized on the half-line: $\int_0^\infty \mathrm{d} z \,\delta_+(z) =1$. However, it is not necessary for us to distinguish between these different contributions since, in any case, any source must be covariantly conserved in a 6-dimensional sense. In Appendix B we discuss a subtle issue having to do with conservation of 5-dimensional external sources of the above form that appears to arise because of the singular nature of the background geometry.
In the presence of an external source, the constraint equation obtained by varying with respect to $\phi$ gets modified from~\eqref{EqforV} to
\begin{eqnarray}
\label{newV}
V=-4 \pi+ 2 \partial_y \tilde \tau-\frac{2\partial_y \bar{N}_y}{\bar N^2}\tilde \sigma+\frac{\bar N^2}{\Box_4+i \varepsilon}T^{zz}_6 \,.
\end{eqnarray}
On substituting this back into the action, everything that is linear in $\phi$ drops out, including part of the matter action.
The remaining matter contribution which sources~(\ref{action3}) is
\begin{eqnarray}
\mathcal{L}'_{\rm matter}&=&-\bar N \hat \sigma \Big[\partial_A T^{Az}_6+\frac{\partial_y \bar{N}_y}{\bar N^2}(T^{yy}_6-\bar{N}_y T^{yz}_6)\Big] \nonumber \\
&&-\bar N \tilde \tau\Big[\partial_A T^{Ay}_6+\bar{N}_y \partial_A T^{Az}_6- \frac{\bar{N}_y \partial_y \bar{N}_y}{\bar N^2} (T^{yy}_6+\bar{N}_y T^{yz}_6)\Big]
+\bar N \tilde \chi \Big[T^{yz}_6+\bar{N}_y T^{zz}_6\Big]\nonumber\\
&&-\frac{3\bar N^2}{2m_6} \pi \Big[\partial_A T^{Az}_6+\frac{\partial_y \bar{N}_y}{\bar N^2}(T^{yy}_6-\bar{N}_y T^{yz}_6)\Big]
+\bar N \pi \(2T^{\;\mu}_{6\; \mu}-\frac{3}{2} \bar \gamma_{AB}T^{AB}_6\)\nonumber\\
&&+2\bar N \frac{\pi}{\Box_4}\Big[\partial_\mu \partial_z T^{\mu z}_6+\partial_\mu \partial_y T^{\mu y}_6-\frac{\bar{N}_y \partial_y \bar{N}_y}{\bar N^2}\partial_\mu T^{\mu y}_6\Big] \nonumber\\
&&+\frac{\bar N^3}{M_6^4} T^{zz}_6 \frac{1}{\Box_4+i \varepsilon} \( T^{yy}_6+2\bar N_y T^{yz}_6+\bar N_y^2 T^{zz}_6\) \nonumber \\
&=&-\bar N \hat \sigma \nabla^{(6)}_A T^{Az}_6-\bar N \tilde \tau \Big[\nabla^{(6)}_A T^{Ay}_6+\bar{N}_y \nabla^{(6)}_A T^{Az}_6\Big]
+\bar N \tilde \chi \Big[T^{yz}_6+\bar{N}_y T^{zz}_6\Big] \nonumber\\
&&-\frac{3\bar N^2}{2m_6}\pi \nabla^{(6)}_A T^{Az}_6
+2\bar N\frac{\pi}{\Box_4} \partial_\mu \nabla^{(6)}_A T^{A\mu}_6
+\bar N\pi\Big[2T^{\;\mu}_{6\; \mu}-\frac{3}{2} \bar \gamma_{AB}T^{AB}_6-2\frac{\partial_\mu\partial_\nu}{\Box_4}T^{\mu\nu}_6\Big] \nonumber \\
&&+\frac{\bar N^3}{M_6^4} T^{zz}_6 \frac{1}{\Box_4+i \varepsilon} \( T^{yy}_6+2\bar N_y T^{yz}_6+\bar N_y^2 T^{zz}_6\)\,,
\end{eqnarray}
where $T^{\;\mu}_{6\; \mu}=\eta_{\mu\nu} T^{\mu\nu}_6$, and $\nabla^{(6)}_A$ is the 6D covariant derivative.
In performing several integration by parts we have made the assumption that $\left.T^{zA}_6\right\vert_{z=0}=0$, which ensures that there is no external force on the orbifold plane. This condition is only necessary in the `half-picture' we are using here. In the doubled picture these terms are canceled by identical contributions of opposite sign on the other side of the orbifold plane.
We see that both modes $\hat \sigma$ and $\tilde \tau$ couple only to the transverse part of the stress-energy tensor. For conserved matter, $\nabla^{(6)}_A T^{AB}_6=0$, both $\hat \sigma$ and $\tilde \tau$ decouple, and the matter sector only sources $\tilde \chi$ and $\pi$
\begin{eqnarray}
\mathcal{L}'_{\rm matter}&=&\bar N \tilde\chi \Big[T^{yz}_6+\bar{N}_y T^{zz}_6\Big]
+\bar N \pi\Big[2T^{\;\mu}_{6\; \mu}-\frac{3}{2} \bar \gamma_{AB}T^{AB}_6-2\frac{\partial_\mu\partial_\nu}{\Box_4}T^{\mu\nu}_6\Big] \\
&&+\frac{\bar N^3}{M_6^4} T^{zz}_6 \frac{1}{\Box_4+i \varepsilon}
\( T^{yy}_6+2\bar N_y T^{yz}_6+\bar N_y^2 T^{zz}_6\)\,. \nonumber
\end{eqnarray}
The sign of the kinetic term of $\hat \sigma$ in the final action~\eqref{action3} is therefore irrelevant to this analysis, since this mode is clearly pure gauge and brings no instability.
Furthermore, we notice that $\tilde \chi$ only couples to 6D matter in the bulk, and thus $\pi$ is the only scalar mode to be excited when including an external source on the branes only. Since $\tilde \chi$ is not a true 6D degree of freedom, we may as well integrate it out, leaving only an action for $\pi$ with additional source contributions:
\begin{eqnarray}
\label{action4}
S_{\pi}&=& \frac{3 M_6^4{}}{2}\int^+ \mathrm{d}^6 x \(\bar N \pi \Box_5 \pi-\frac 1\bar N (\mathcal{L}_n \pi)^2\) +\frac 32M_5^3{} \int_{z=0} \mathrm{d}^5 x \; \pi \Box_5 \pi \nonumber \\
&&+ \frac 3 4 M_4^2{} \int_{z=y=0} \mathrm{d}^4 x \(\frac{3}{2m_6^2}\frac{\lambda}{M_4^2{}}-1\) \pi\Box_4 \pi \nonumber\\
&&+\int^+ \mathrm{d}^6x\; \bar N \pi\Big[2T^{\;\mu}_{6\; \mu}-\frac{3}{2} \bar \gamma_{AB}T^{AB}_6-2\frac{\partial_\mu\partial_\nu}{\Box_4}T^{\mu\nu}_6\Big] \nonumber \\
&&+\int^+ \mathrm{d}^6x\; \frac{\bar N^3}{M_6^4} \( T^{zz}_6 \frac{1}{\Box_4+i\varepsilon} T^{yy}_6-T^{yz}_6 \frac{1}{\Box_4+i\varepsilon} T^{yz}_6\)\,.
\end{eqnarray}
Now the additional source contributions may appear to have a 4-dimensional pole, but this is in fact not the case. Such terms arise whenever 6D perturbations are decomposed into 4D scalar-vector-tensors, and appear non-local because of the inherent non-locality of the split. In Appendix A we demonstrate that these do not contribute a pole to the single-graviton exchange amplitude.
When considering matter on the 3-brane only, $T^{AB}_6= \bar N^{-1}\delta(y)\delta_+(z)\, \mathcal{T}^{\mu\nu}_4 \delta_\mu^A\delta_\nu^B$, with $\partial_\mu \mathcal{T}^{\mu\nu}_4=0$, the matter Lagrangian is simply
\begin{eqnarray}
\mathcal{L}^{'(4)}_{\rm matter}=\frac 12 \pi \mathcal{T}_4\,.
\end{eqnarray}
This agrees with the flat space-time limit, as well as with the decoupling limit. Meanwhile, the Lagrangian for 5D matter is
\begin{eqnarray}
\mathcal{L}^{'(5)}_{\rm matter}=\frac 12 \pi \Big[\mathcal{T}_5^{ab}\eta_{ab}-4\frac{\Box_5}{\Box_4}\mathcal{T}_5^{yy}
\Big]\,,
\end{eqnarray}
where we used $\partial_\mu\partial_\nu \mathcal{T}_5^{\mu\nu}=\partial_y^2 \mathcal{T}_5^{yy}$, which follows from 5D stress-energy conservation.
\section{Discussion}
\label{conclu}
In this paper we have demonstrated the perturbative unitarity and stability of the 6D cascading gravity framework. This puts the claim that this model is a consistent infrared modification of gravity on a firm footing. While we cannot rule out from the previous analysis the existence of an non-perturbative instability, we at least expect a finite range of energies for which this model can be understood as consistent effective field theory. A potentially profound result of this analysis is the confirmation of the result, first obtained in the decoupling limit~\cite{cascade1}, that unitarity requires a minimum tension for the codimension-2 brane localized on an orbifold plane, as soon as we introduce a localized kinetic term for the graviton, {\it i.e.} a 4D induced gravity term. It will be interesting to explore whether a similar bound on the tension applies to higher-codimension branes, and whether there is any fundamental reason for this bound, or whether it is possible to relax this condition by other means (see \cite{gigashif,cascade2}). In any case, our results establish this model as a consistent framework
within which to explore the phenomenology of infrared theories of modified gravity and to confront its predictions against cosmological observations.
\acknowledgments
We would like to thank Gia Dvali, Gregory Gabadadze, Kurt Hinterbichler, Stefan Hofmann, Kazuya Koyama, Oriol Pujolas, Michele Redi, Mark Trodden and Daniel Wesley for useful discussions. A.~J.~T. would like to thank the University of Geneva and LMU, Munich for hospitality whilst this work was being completed. The work of A.~J.~T. at the Perimeter Institute is supported in part by the Government of Canada through NSERC and by the Province of Ontario through MRI. C.~dR is supported by the SNF. The work of J.~K. is supported in part by funds from the University of Pennsylvania and NSERC of Canada.
\section{Appendix A: The decoupling argument}
\label{decoup}
We briefly review the analysis of~\cite{cascade1} using the decoupling limit.
(Note the slight notational differences compared to~\cite{cascade1}. In particular, we have interchanged the roles of $\tilde{h}_{ab}$ and $h_{ab}$ below.)
In the decoupling limit $M_5,M_6\rightarrow \infty$ with $\Lambda_6 = (m_6^4M_5^3)^{1/7}$ fixed, the action~(\ref{6Dcascade}) reduces
to a local field theory on the orbifold brane, describing 5D weak-field gravity and a self-interacting scalar field $\pi$:
\begin{eqnarray}
S_{\rm decoup}^{\rm Jordan} & = & \frac{M_{5}^{3}}{2} \int {\rm d}^{5}x \left[ -\frac{1}{2} h^{ab}(\mathcal{E}h)_{ab} + \pi\eta^{ab}(\mathcal{E}h)_{ab} - \frac{27}{16m_6^2} (\partial_a\pi)^{2} \Box_{5}\pi \right] \nonumber \\
& & + \int_{y=0} {\rm d}^{4}x \left[ -\frac{M_{4}^{2}}{4} h^{\mu\nu}(\mathcal{E}h)_{\mu\nu} + \frac{1}{2} h^{\mu\nu}T_{\mu\nu} \right]\,,
\label{5Dcov1}
\end{eqnarray}
where $\left(\mathcal{E}h\right)_{ab} = -\frac{1}{2} \left(\Box_{5} h_{ab} - \eta_{ab}\Box_{5} h -\partial_a \partial^c h_{cb} - \partial_b\partial^c h_{ac} +\eta_{ab}\partial^c\partial^d h_{cd}+\partial_a\partial_b h\right)$ is the linearized Einstein tensor in 5D, and $(\mathcal{E}h)_{\mu\nu}$ that in 4D. All other interactions in~(\ref{6Dcascade}) are suppressed by powers of $1/M_5$, $1/M_6$ and therefore drop out in the decoupling limit.
With the action written in terms of the physical metric $h_{ab}$, as in~(\ref{5Dcov1}), we notice that $\pi$ is non-minimally coupled to the linearized Ricci scalar, {\it i.e.}, $h_{ab}$ is
the Jordan-frame metric. The kinetic matrix for the metric and $\pi$ can be diagonalized as usual by shifting to the Einstein frame variable, $\tilde{h}_{ab} = h_{ab} -\pi \eta_{ab}$.
Expressed in terms of $\tilde{h}_{ab}$, the decoupling action becomes
\begin{eqnarray}
S_{\rm decoup}^{\rm Einstein} & = & \int {\rm d}^{5}x \left[ -\frac{M_5^3}{4} \tilde{h}^{ab}(\mathcal{E}\tilde{h})_{ab} - \frac{3M_5^3}{2} (\partial_a\pi)^{2}\left(1+ \frac{9}{16m_6^2} \Box_{5}\pi\right) \right] \nonumber \\
& & + \int_{y=0} {\rm d}^{4}x \left[ -\frac{M_{4}^{2}}{4} h^{\mu\nu}(\mathcal{E}h)_{\mu\nu} + \frac{1}{2} \tilde{h}^{\mu\nu}T_{\mu\nu} + \frac{1}{2}\pi T\right]\,.
\label{5Dcov2}
\end{eqnarray}
As usual, the transformation to Einstein frame generates a conformal coupling of $\pi$ to matter.
In the presence of tension on the 3-brane, $T_{\mu\nu} = -\lambda \eta_{\mu\nu}$, the solution to the equations of motion is
\begin{equation}
\pi^{(0)} = \frac{\lambda}{3M_5^3}|y|\;; \qquad \tilde{h}_{\mu\nu}^{(0)} = - \frac{\lambda}{3M_5^3}|y|\eta_{\mu \nu} \; ; \qquad \tilde{h}_{y y}^{(0)} = 0\,.
\label{backlamb}
\end{equation}
These combine to give a flat Jordan-frame metric, $h_{\mu\nu}^{(0)} = \tilde{h}_{\mu\nu}^{(0)} + \pi^{(0)}\eta_{\mu\nu} = 0$, consistent with the fact
in 6D that the codimension-2 source generates a deficit angle but leaves the geometry Riemann flat.
To disentangle the different degrees of freedom, let us define
\begin{equation}
\tilde{h}_{ab}\mathrm{d} x^a \mathrm{d} x^b =\tilde{h}_{\mu\nu} \mathrm{d} x^{\mu} \mathrm{d} x^{\nu}+V\mathrm{d} y^2+2V_{\mu}\mathrm{d} x^{\mu}\mathrm{d} y\,.
\end{equation}
Substituting into the action (\ref{5Dcov2}) gives
\begin{eqnarray}
S_{\rm decoup}^{\rm Einstein} & = & \int {\rm d}^{5}x \left[ \frac{M_5^3}{8} \(2V(\partial^{\mu}\partial^{\nu}\tilde h_{\mu\nu}-\Box_4 \tilde h_{(4)}) +4V^{\mu}( \partial_y \partial_{\mu}\tilde h_{(4)}-\partial_y \partial^{\nu}\tilde h_{\mu \nu} ) \right. \right. \nonumber \\
&&\left. \left. +\tilde h^{\mu\nu}\Box_5 \tilde h_{\mu\nu}-\tilde h_{(4)}\Box_5 \tilde h_{(4)}+2 \tilde h_{(4)}\partial^{\mu}\partial^{\nu}\tilde h_{\mu\nu}+2(\partial^{\mu}\tilde h_{\mu\nu})^2 -F_{\mu \nu}^2\)\right.\nonumber\\
&&\left.- \frac{3M_5^3}{2} (\partial_a\pi)^{2}\left(1+ \frac{9}{16m_6^2} \Box_{5}\pi\right) \right] \nonumber \\
& & + \int_{y=0} {\rm d}^{4}x \left[ -\frac{M_{4}^{2}}{4} h^{\mu\nu}(\mathcal{E}h)_{\mu\nu} + \frac{1}{2} \tilde{h}^{\mu\nu}T_{\mu\nu} + \frac{1}{2}\pi T\right]\,,
\end{eqnarray}
with $F_{\mu \nu}=\partial_\mu V_\nu-\partial_\nu V_\mu$.
Concentrating on scalar perturbations only, we may choose a gauge in which the Einstein metric is pure trace: $\tilde h_{\mu\nu}=\frac{1}{4} \tilde h_{(4)} \eta_{\mu\nu}$. However, now it is clear that $V$ and the scalar part of $V_{\mu}$ are Lagrange multipliers which enforce the constraint $\Box_4 \tilde h_{(4)}=0$. Thus in the scalar sector we have $\Box_4 h_{\mu\nu}=\Box_4 \pi \eta_{\mu\nu}$ and $\Box_4 \tilde h_{\mu\nu}=0$. Note that these equations are consistent with the above background solution, at least for $|y|>0$.
The action therefore reduces to the simple form (neglecting the vector part of $V_{\mu}$ which does not couple to 4D matter)
\begin{equation}
\label{decouplingpi}
S_{\rm decoup}^{\pi} = \int {\rm d}^{5}x \left[ - \frac{3M_5^3}{2} (\partial_a\pi)^{2}\left(1+ \frac{9}{16m_6^2} \Box_{5}\pi \right) \right] + \int_{y=0} {\rm d}^{4}x \left[ -\frac{3M_{4}^{2}}{4} \pi \Box_4 \pi + \frac{1}{2}\pi T\right]\,.
\end{equation}
Expanding~(\ref{decouplingpi}) to quadratic order in perturbations around the background~(\ref{backlamb}), we find that the $\pi$ cubic term yields a contribution
to the kinetic term of the perturbations that is localized on the 3-brane:
\begin{equation}
\Delta S_5 = - \int {\rm d}^{5}x \, \frac{9\lambda}{8m_6^2}(\partial_{\mu}\hat{\pi})^2\delta(y)\,,
\end{equation}
where $\hat{\pi}$ now denotes a perturbation around the background profile $\pi^{(0)}$. Combining with the rest of~(\ref{5Dcov2}), the quadratic action is thus given by
\begin{equation}
\label{decouplingpi2}
S_{\rm decoup}^{\pi} = \int {\rm d}^{5}x \left[ - \frac{3M_5^3}{2} (\partial_a \hat \pi)^{2} \right] + \int_{y=0} {\rm d}^{4}x \left[ +\frac{3M_{4}^{2}}{4} \(\frac{3\lambda}{2m_6^2M_4^2}-1 \)\hat \pi \Box_4 \hat \pi + \frac{1}{2}\hat \pi \delta T\right]\,.
\end{equation}
This is now manifestly equivalent to the decoupling limit of our final result, reproducing the same minimum bound on the tension.
\section{Appendix B: No extra pole in $TT$ amplitude}
\label{sigma1}
In the final action for $\pi$, given in~(\ref{action4}), we obtained the following term
\begin{equation}
W\equiv \int^+ \mathrm{d} z \int \mathrm{d}^5 x \frac{\bar N^3}{M_6^4} \( T^{zz}_6 \frac{1}{\Box_4+i\varepsilon} T^{yy}_6-T^{yz}_6 \frac{1}{\Box_4+i\varepsilon} T^{yz}_6\) \,.
\end{equation}
Superficially this looks like a contribution to the $TT$ amplitude which has a 4 dimensional pole and will in general be ghostly. However, since this term already arises in Minkowski space-time it is clear that this cannot be the case. This non-local term arises because the scalar-vector-tensor decomposition is inherently nonlocal.
Absence of ghosts is equivalent to the condition that
\begin{equation}
{\rm Im} [W] \geq 0\,.
\end{equation}
Since the imaginary part comes entirely from the pole at $\Box_4=0$, it is sufficient to consider the form of the stress-energy on the pole surface, {\it i.e.}, in 4-momentum space with $k^2=0$. Since we are looking at scalar perturbations it then follows that on this surface $\partial_{\mu}T^{\mu y}=\partial_{\mu}T^{\mu z}=0$.
Now conservation of energy implies
\begin{eqnarray}
\partial_z (\bar{N} T_6^{zz})+\partial_y (\bar{N} T_6^{yz})+\frac{\partial_y \bar N_y}{\bar N}T_6^{yy}=0\,;\\
\partial_z (\bar N^2 T_6^{zy})+ \partial_y (\bar N^2 T_6^{yy})=0\,.
\end{eqnarray}
We can solve these equations as follows
\begin{equation}
T_6^{yy}=\bar N^{-2}\partial_z^2 U, \quad T_6^{yz}=-\bar N^{-2}\partial_z\partial_y U, \quad T_6^{zz}=\frac{1}{\bar N}\partial_y (\bar N^{-1} \partial_y U)-\frac{\partial_y \bar N_y}{\bar N^4} \partial_z U.
\end{equation}
Computing the relevant piece, and using the fact that the $\mathds{Z}_2$ symmetry implies $T^{zy}_{z=0}=0$, then $\partial_z U=0$ at $z=0$, and so we find when expressed in momentum space
\begin{eqnarray}
&& \int^+ \mathrm{d} z \mathrm{d} y\bar N^3 \( T^{zz}_6(-k) T^{yy}_6(k)-T^{yz}_6(-k) T^{yz}_6(k)\) \nonumber \\
&&=\int^+ \mathrm{d} z \mathrm{d} y\( -\partial_z \(\bar N^{-1} \partial_y U(-k)\partial_z \partial_y U(k) \) -\frac{\partial_y \bar N_y}{2\bar N^3}\partial_z |\partial_z U(k)|^2\) \nonumber \\
&&=\(\bar N^{-1} \partial_y U(-k)\partial_z \partial_y U(k) \) _{z=0}+\frac{\partial_y \bar N_y}{2\bar N^3}\left\vert\partial_z U(k)\right\vert^2_{z=0}=0\,,
\end{eqnarray}
where $k$ is the on-shell 4-momentum. In other words, conservation of energy and the orbifold condition ensures that there is no anomalous 4D pole in the propagator.
\section{Appendix C: No pole from $\tilde{\sigma}$ in $TT$ amplitude}
\label{sigma2}
Although the stress-energy on the 4-brane is automatically conserved in 5D and 6D, it is worth stressing that the presence of the singularity at $y=z=0$ imposes an extra constraint for the allowed matter on the orbifold brane. Indeed, the 6D conservation of energy condition applied on the 5D stress-energy implies
\begin{eqnarray}
\nabla^{(6)}_A\Big[\frac{1}{\bar N} \mathcal{T}_5^{Az} \delta_+(z) \Big] &=&\Gamma^{(6)\, z}_{\ \ AB}\, \mathcal{T}^{AB}_5 \bar N^{-1} \delta_+(z) \nonumber\\
&=&\frac{\partial_y \bar{N}_y}{\bar N^3}\mathcal{T}_5^{yy} \delta_+(z)=\frac{\lambda}{2M_6^4{} \bar N^2}\mathcal{T}_5^{yy} \delta(y)\delta_+(z)=0\,.
\end{eqnarray}
In the background gauge~\eqref{6Dbackground}, the 6D conservation of energy imposes the $yy$-component of the 5D stress-energy to vanish at the singularity $\mathcal{T}_5^{yy}\Big|_{y=z=0}=0$. If this is not so, then this must be compensated by a nonzero contribution to $T^{zA}$. This is a surprising condition, and only really becomes of importance because of the need to regulate the background geometry. Although at first sight it would seem to forbid putting a very small tension on the codimension-1 brane, it is very likely that in practice in the presence of a codimension-1 tension the background will be sufficiently modified that this condition is removed.
If we do not make the assumption that $T^{yy}_5=0$ at $y=z=0$, then on coupling to the metric induced on the orbifold brane, it would appear
that $\hat \sigma$, which decouples from 6D conserved matter, is nevertheless sourced by a term of the form
\begin{equation}
\int \mathrm{d}^5 x - \frac{\partial_y \bar N_y}{\bar{N}^2} \tilde{\sigma} {\cal T}_5^{yy}=-\int \mathrm{d}^4 x \frac{\lambda}{M_6^4 \bar{N}} \hat{\sigma} {\cal
T}^{yy}_5\,.
\end{equation}
Since $\hat{\sigma}$ enters with the wrong sign kinetic term, such a coupling would seem to indicate a violation of unitarity.
Integrating out $\hat{\sigma}$ this piece generates a term of the form
\begin{equation}
\int \mathrm{d}^5x\, \delta(y) \frac{\lambda}{2} {\cal T}^{yy}_{5}\frac{1}{\Box_4+i\varepsilon}{\cal T}^{yy}_{5}=\int \mathrm{d}^4 x\, \frac{\lambda}{2} {\cal T}^{yy}_{5}\frac{1}{\Box_4+i\varepsilon}{\cal T}^{yy}_{5}\,.
\label{finaleq}
\end{equation}
However, following the same reasoning as in Appendix B, since the unitarity violation comes entirely from the pole, we may look at the implications of the 5D conservation
equations on the pole surface $k^2=0$. Since we are looking at scalar perturbations it again follows that $\partial_{\mu} {\cal T}_5^{y \mu}=0$
on this surface, and so we have
\begin{equation}
\partial_y {\cal T}^{yy}_5=0\,.
\end{equation}
For a localized source we may integrate~(\ref{finaleq}) by parts to rewrite it as
\begin{equation}
{\rm Im}\( \int \mathrm{d}^5x \delta(y) \frac{\lambda}{2} {\cal T}^{yy}_{5}\frac{1}{\Box_4+i\varepsilon}{\cal T}^{yy}_{5}\)=-{\rm Im} \( \int d^5x \epsilon(y) {\cal T}^{yy}_{5}\frac{1}{\Box_4+i\varepsilon}\partial_y {\cal T}^{yy}_{5} \)=0\,.
\end{equation}
Thus there is no pole contribution to the $TT$ amplitude coming from $\hat{\sigma}$, further confirming that its apparent
wrong sign kinetic term does not indicate the presence of a ghost.
|
\section{Introduction}
A reinterpretation of the Weinberg--Salam model has recently appeared in the literature \cite{Chernodub:2008rz, Faddeev:2008qc, Masson:2010vx}, in which a change of variables is used to transform the action into one depending only on SU(2) {\it invariant} fields. In this way, the local SU(2) symmetry is factored out of the action. In addition, the approach of \cite{Chernodub:2008rz, Faddeev:2008qc} omits the usual Higgs symmetry breaking terms, and reinterprets the Higgs field as a dilaton. The action then describes a gravitational theory in which the electroweak fields interact in a locally conformally flat spacetime. In this picture, it is the condition of asymptotic flatness which is responsible for mass generation.
The transformation of \cite{Chernodub:2008rz, Faddeev:2008qc, Masson:2010vx} was actually considered some time ago in \cite{Vlasov:1987vt, Lunev:1994ty} and \cite{Lavelle:1994rh}, where it was indeed used to construct gauge invariant variables, and in \cite{Langguth:1985dr, Montvay:1985nk, Hasenfratz:1987uc} where it was put to various uses on the lattice. (See \cite{Goldstone:1978he, Frohlich:1981yi, Farhi:1995pt, Gromov:2007he, Gromov:2010gf} for related and other approaches to gauge invariant variables.) In this paper we will bring together the various interpretations of the transformation, clarify its relation to gauge fixing and point out that the transformation potentially suffers from an ambiguity at the quantum level, which can be interpreted as a Gribov problem. From this we will obtain an important physical consequence of the asymptotic flatness condition: it is precisely this which allows us to sidestep the Gribov ambiguity in this situation, and so obtain nonperturbatively locally gauge invariant definitions of the electroweak fields, including the $Z$ and $W$ gauge bosons. This is in contrast to unbroken SU(3), where there is no gauge invariant description of the gluon, which consequently is not observable.
We begin in Sect.~\ref{review} with a brief review of the classical transformation to gauge invariant variables, the quantum contributions arising from the functional determinant for the transformation, and the interpretation of the Higgs as a conformal metric factor. In Sect.~\ref{gauge-sect} we explain the relations between this approach, unitary gauge fixing in the matter sector, and the definition of physical charges in gauge theories. In Sect.~\ref{check-sect} we show that a careful treatment of boundary conditions is necessary in order for the given transformation to make sense nonperturbatively. Here we also explore the connections between the scalar field's v.e.v.\ and the Gribov ambiguity. We also discuss the (non) existence of similar transformations in other gauge theories. The important details of the new interpretation also holds for the SU(2) Higgs model, and we work for the most part with this theory in order to simplify the presentation and clarify what is really going on. The extension to SU(2)$\times$U(1) is direct and will be described in Sect.~\ref{check-sect}. We conclude in Sect. \ref{Concs}.
\section{To gauge invariant variables}\label{review}
\subsection{The classical transformation}
We begin with the transformation to gauge invariant variables in the SU(2) Higgs model, which has the Lagrangian
\begin{equation}\label{lag1}
\mathcal{L} = |D_\mu\Phi |^2 - \frac{1}{4}F^a_{\mu\nu}F^{\mu\nu}_a + \mathcal{L}_{ssb}(|\Phi|)\;.
\end{equation}
Here $\Phi=(\Phi_1,\Phi_2)$ is a doublet of complex scalars, with covariant derivative $D_\mu = \partial_\mu + g A_\mu$, and our $A_{\mu}$ are anti--hermitian. We consider two forms of the potential terms,
\begin{eqnarray}
\label{ssb1} \mathcal{L}_{ssb}(|\Phi|) &=& 0\;,\\
\label{ssb2} \mathcal{L}_{ssb}(|\Phi|) &=& -m^2 |\Phi|^2 -\lambda |\Phi|^4\;,
\end{eqnarray}
the second being the usual symmetry breaking potential. We address this aspect of the action in a later section. The essential idea here is to separate the scalar into gauge invariant and noninvariant pieces, then make a change of variables which removes the latter from the action. This transformation will also remove the gauge noninvariant components of the gauge field, seemingly for free. To do so, we follow \cite{Faddeev:2008qc}\ and \cite{Masson:2010vx}, where the field $\Phi$ is decomposed as follows (up to conventions), introducing the scalar $\rho\equiv |\Phi|$ and a matrix we will call $h$,
\begin{equation}\label{h-def}
\Phi = \rho \left(\begin{array}{c} \Phi_1/\rho \\ \Phi_2/\rho \end{array}\right) = \rho \left(\begin{array}{cc} \Phi_1/\rho &-\bar{\Phi}_2/\rho \\ \Phi_2/\rho & \bar\Phi_1/\rho \end{array}\right) \left(\begin{array}{c} 1 \\ 0\end{array}\right) \equiv \rho\, h[\Phi] \left(\begin{array}{c} 1 \\ 0\end{array}\right)\;.
\end{equation}
The question of whether this decomposition is allowed at $|\Phi|=0$ is actually crucial, and we return to it shortly. (The appearance of a fixed vector in (\ref{h-def}) suggests a relation to unitary gauge fixing, which we explain in the next section.) For now we observe that $\rho$ is gauge invariant, while the matrix $h$ is SU(2) valued, and under gauge transformations
\begin{equation}
\Phi \to \Phi^U \equiv U^{-1}\Phi\;, \qquad A_\mu \to A_\mu^U\equiv U^{-1}A_\mu U + \frac{1}{g}U^{-1}\partial_\mu U\;,
\end{equation}
it can be checked that $h$ transforms in the same way as the scalar, i.e.
\begin{equation}\label{h-trans}
h[\Phi] \rightarrow U^{-1}\, h[\Phi]\;.
\end{equation}
This, as we will see, is the crucial property of $h$. Using this field-dependent matrix, we define from the gauge potential $A_\mu$ a new field $B_\mu$,
\begin{equation}\label{B-def}
B_\mu := h^{-1}A_\mu h + \frac{1}{g}h^{-1}\partial_\mu h\;.
\end{equation}
Now, because of the transformation property of $h$, it follows that $B_\mu$ is {\it invariant} under local SU(2) transformations. If we rewrite the Lagrangian ({\ref{lag1}}) in terms of the new fields $\rho$, $h$ and $B_\mu$, we find
\begin{equation}\label{lag2}
\mathcal{L} = \partial_\mu\rho \partial^\mu\rho -\frac{1}{4}B^a_{\mu\nu}B_a^{\mu\nu}+\frac{g^2}{4}\rho^2 B^a_\mu B_a^\mu + \mathcal{L}_{ssb}(\rho)\;,
\end{equation}
where $B_{\mu\nu}$ is the nonabelian field strength calculated from $B_\mu$ in the usual way, and all the dependence on $h$ has vanished. The remaining fields are all SU(2) invariant. In this way, the gauge symmetry has been factored out of the action, but no gauge has been fixed. With the usual potential terms (\ref{ssb2}), the classical vacuum solutions following from (\ref{lag2}) are
\begin{equation}\label{vacsolnew}
B_\mu^a(x)=0\;,\qquad \rho(x) = \sqrt{\frac{-m^2}{\lambda}}\equiv v\;,
\end{equation}
with $v$ being the scalar's v.e.v. Note, though, that these solutions are unique, in that there is a single value for the scalar in the minimum of the potential -- there is no U(1) of equivalent vacua. Expanding the scalar about this minimum,
\begin{equation}\label{expan1}
\rho(x) = v + H(x)\;,
\end{equation}
(\ref{lag2}) acquires mass terms and describes a massive vector and a single scalar, which are the expected degrees of freedom in the broken sector of the SU(2) Higgs theory.
It is worth recalling here Elitzur's theorem, which states that it is impossible to spontaneously break a local symmetry \cite{Elitzur:1975im}. Essentially, the reason for this is that the functional integral averages over gauge orbits, so that even when the scalar's potential has a nontrivial minimum, the average of the scalar field will still be zero. Thus, the scalar acquires no v.e.v. and the gauge symmetry is preserved. (See \cite{Fradkin:1978dv, Kajantie:1995kf, Gurtler:1997hr} for related investigations and \cite{Caudy:2007sf} for a recent review.) Starting from (\ref{lag1}), though, and proceeding to (\ref{lag2}), no gauge fixing is performed, nor are any properties of the scalar's potential employed. Performing only a change of variables, the result is a theory which is completely independent of the SU(2) symmetry. As there is no gauge symmetry, Elitzur's theorem has no consequence for the theory (\ref{lag2}). Taking the usual potential (\ref{ssb2}) we find the nontrivial vacuum solution (\ref{vacsolnew}), and expanding about this vacuum generates a mass for the field $B_\mu^a$. We will return to Elitzur's theorem later in the paper.
\subsection{The measure and conformal geometry}
In the interpretation of \cite{Chernodub:2008rz, Faddeev:2008qc}, we perform the above transformation but take $\mathcal{L}_{ssb}=0$. The classical vacuum solutions of (\ref{lag2}) are then degenerate, being,
\begin{equation}
B_\mu^a(x) =0\;, \qquad \rho^2(x) = \Lambda^2\;,
\end{equation}
for some $\Lambda$, the choice of which corresponds to a choice of vacuum. A reason for taking $\Lambda\not=0$, as required for mass generation, may be seen by quantising the theory. The functional measure for the new variables is
\begin{equation}\label{measure1}
\prod\limits_x\ \rho^2 \mathrm{d} \rho^2 \mathrm{d} B^a_\mu\;.
\end{equation}
This is multiplied by an integral over the $h$ degrees of freedom, but since the Lagrangian is independent of $h$, this gives simply the volume of SU(2) and can be dropped. More interesting is the local factor of $\rho^2$ appearing in the measure. In \cite{Chernodub:2008rz, Faddeev:2008qc} this was interpreted as the conformal factor of a spacetime metric, $G_{\mu\nu}=\rho^2 \eta_{\mu\nu}$, since the Lagrangian (\ref{lag2}) is in just the right form to be rewritten
\begin{equation}\label{lag3}
\mathcal{L} = \sqrt{-G}\bigg( \frac{\mathcal{R}}{6} -\frac{1}{4}B^a_{\mu\nu}B_a^{\mu\nu}+ \frac{g^2}{4}B^a_\mu B_a^\mu\bigg)\;,
\end{equation}
where all indices are now raised and lowered with respect to the metric $G$, and $\mathcal{R}$ is the Ricci scalar\footnote{See \cite{Chernodub:2008rz} for a discussion of the differences between the Euclidean and Minkowskian theories.} for $G$. This gives us a theory of a gauge boson living in a locally conformally flat spacetime. For phenomenological investigations of this idea see \cite{Foot:2008tz, Ryskin:2009kw}. Now, in order to maintain asymptotic flatness of the metric, we require $\rho^2(x) \to \Lambda^2$ asymptotically. In this way we can expand $\rho(x)=\Lambda + H(x)$, with $H(x)$ vanishing asymptotically, in analogy to (\ref{expan1}). The presence of $\Lambda$ generates the required mass terms in the action, in analogy to $v$ appearing in (\ref{vacsolnew}). Below, we will provide further physical motivation for the appearance of $\Lambda$.
\section{Relation to unitary gauge}\label{gauge-sect}
\subsection{The transformation}
Although we have only performed a change of variables, it is clear from (\ref{h-def}) and (\ref{B-def}) that there is a connection here to unitary gauge fixing. In this section we make this connection explicit. The purpose is to derive and understand a criteria, connected to scalar field's potential, for when the above transformation is valid.
We begin by noting that an implicit choice was made in the definition of $h$ in (\ref{h-def}) -- there are many other choices for the vector piece and ways to fill up the matrix. The only property of $h$ on which everything rests is that it should transform as in (\ref{h-trans}). The matrix $h$ is in fact just one example of a much more general object, called a `dressing', used to construct gauge invariant, physical fields in gauge theories \cite{Lavelle:1995ty}. The idea is that in order to describe physical particles, the gauge noninvariant Lagrangian fields must be combined into gauge invariant composites, or dressed, fields. The canonical example is Dirac's static electron $\Psi_s$ which is constructed from the Lagrangian fermion $\psi$ and the U(1) gauge field as \cite{Dirac:1955uv}
\begin{equation}\label{dirac}
\Psi_s = h^{-1}_s [A] \psi\equiv \exp\bigg[ie \frac{\nabla.A}{\nabla^2}\bigg]\, \psi\;.
\end{equation}
The exponential prefactor, or `dressing' $h^{-1}_s$, compensates for local gauge transformations of the fermion (but not for global transformations, so this electron does indeed carry a charge). The properties of dressed fields in both QED and QCD are discussed thoroughly in \cite{Lavelle:1995ty, Bagan:1999jf, Bagan:1999jk}. In particular, they allow for the construction of infrared finite asymptotic states and hence offer a route to solving the infrared problems which plague gauge theories \cite{Lavelle:2005bt}.
Returning to our non--abelian theory, one could choose, in analogy to (\ref{dirac}), to construct a dressing using the gauge fields $A^a_\mu$. This would give a transformation of (\ref{lag1}) to another set of variables which would, in complete analogy to $B^a_\mu$ and $\rho$, be gauge invariant. This is illustrated for the abelian Higgs model in \cite{Lavelle:1994rh}, but as it amounts to {\it directly} identifying the gauge invariant part of the gauge field, the same approach is much more difficult in nonabelian theories. In fact, while it can be approached in perturbation theory, there is in general a nonperturbative obstruction, which we will encounter shortly. (In the present case we will also find a way around the problem.) However, \cite{Lavelle:1994rh} also shows that the interpretation of the invariant fields is more natural if one dresses using the scalars, i.e. with the $h$ given in (\ref{h-def}). Indeed, this choice makes the expected degrees of freedom of the broken sector manifest, just as we saw in (\ref{lag2}).
As shown in \cite{Lavelle:1995ty, Ilderton:2007qy}, dressings can be constructed using gauge fixing conditions. If we have a gauge condition $\chi[f]=0$, for some field $f$, then solving $\chi[f^h]=0$ gives the field dependent transformation $h\equiv h[f]$ which takes $f$ into the chosen gauge slice. From this one can show that $h^{-1}$ is a dressing. So, our matrix $h[\Phi]$ is indeed a gauge transformation, albeit a {\it field dependent} one: it is the transformation which takes an arbitrary field into unitary gauge. This is easily seen for the scalar field:
\begin{equation}
\Phi\to \Phi^h \equiv h^{-1}[\Phi] \Phi =\rho \left(\begin{array}{c} 1 \\ 0\end{array}\right)\;,
\end{equation}
by definition, which is the single expected physical scalar of unitary gauge, while a short calculation gives the vector potential components, e.g.,
\begin{eqnarray*}
B^3_\mu = &\frac{1}{|\Phi|^2}\bigg[ A^3_\mu(|\Phi_1|^2- |\Phi_2|^2) +\bar\Phi_1\Phi_2(A^1_\mu-i A^2_\mu)+ \bar\Phi_2\Phi_1(A^1_\mu+i A^2_\mu) \bigg] \\
&+ \frac{i}{g|\Phi|^2}\bigg[ \bar\Phi_1 \partial_\mu \Phi_1 - \Phi_1 \partial_\mu \bar\Phi_1 + \bar\Phi_2 \partial_\mu \Phi_2 - \Phi_2 \partial_\mu \bar\Phi_2\bigg]\;,
\end{eqnarray*}
and the others similarly.
\subsection{The measure as a Faddeev--Popov determinant}
We now offer a more standard interpretation of the functional measure in (\ref{measure1}). This will allow us to use familiar techniques to establish when the transformation discussed above holds nonperturbatively. Given the interpretation of the classical transformation, it is not too hard to see that the measure factor is just the Faddeev--Popov determinant for unitary gauge fixing. Let us briefly confirm this. Expanding the complex scalars as $\Phi_1= \varphi_1 + i\varphi_2$, $\Phi_2 = \varphi_3+ i\varphi_4$, unitary gauge eliminates the components $\varphi_2$, $\varphi_3$ and $\varphi_4$, so that $\rho \equiv |\Phi| \to |\varphi_1|$. Using the explicit form of the Pauli sigma matrices, one can calculate the variation of the gauge fixing conditions from their Poisson brackets with Gauss' law, e.g.,
\begin{equation}\label{gauss}
\{G^a\epsilon^a(x), \varphi_2(y)\} = (\epsilon^1\varphi_3 +\epsilon^2\varphi_4 + \epsilon^3\varphi_1) \delta(x-y)\;,
\end{equation}
and check that the Faddeev--Popov determinant for this gauge condition is
\begin{equation}\label{thedet}
\left| \begin{array}{ccc} \varphi_3 & \varphi_4 & \varphi_1 \\
-\varphi_2 & -\varphi_1 & \varphi_4 \\
\varphi_1 & -\varphi_2 & -\varphi_3 \end{array} \right|_{\varphi_2=\varphi_3=\varphi_4=0} = |\varphi_1|^3\,.
\end{equation}
The resulting measure on the remaining scalar is\footnote{We may trade $\varphi_1$ for $|\varphi_1|$, since the transformed action sees only $|\Phi|\to |\varphi_1|$.},
\begin{equation}
\prod\limits_x \varphi_1^3 \mathrm{d} \varphi_1 \sim \prod\limits_x \varphi_1^2 \mathrm{d} \varphi_1^2 \sim \prod\limits_x \rho^2 \mathrm{d} \rho^2\,,
\end{equation}
which is just the measure over the gauge invariant field $\rho$ given in (\ref{measure1}).
We are now ready to bring everything together. With some insight gained from the relations to unitary gauge, we can address when the above transformation is allowed, relate this to symmetry breaking and the condition of asymptotic flatness in the conformal interpretation.
\section{Physical degrees of freedom}\label{check-sect}
Recall that a good gauge fixing is identified by its Faddeev--Popov determinant being {\it non vanishing} for all field configurations. If it does vanish, we have a Gribov problem \cite{Singer:1978dk}. See \cite{Dell'Antonio:1991xt, Canfora:2008vt, Holdom:2009ws} for investigations of the distribution of Gribov copies, \cite{Heinzl:2007cp, vonSmekal:2007ns, Sorella:2009vt} for the connections to BRST invariance, and \cite{Dudal:2008sp, Bornyakov:2008yx, Greensite:2010hn} plus references therein for the effect of copies on the infra--red behaviour of ghost and gluon propagators, and the implications for the Gribov--Zwaniger confinement conditions \cite{Zwanziger:1989mf}.
Now, since our functional measure is a Faddeev--Popov determinant, we may use the above condition to determine when our transformation is sensible quantum mechanically. It is important to stress that if there is a Gribov problem, the new variables will {\it fail} to be gauge invariant nonperturbatively. Let us explain why this happens, and give a simple example.
Consider a U(1) gauge theory, with the {\it local} gauge transformation $U=\exp(ie\, k\cdot x)$ which sends $A_\mu \to A_\mu + k_\mu$. If a gauge field starts in Coulomb gauge, it returns to Coulomb gauge after this transformation, and so we {\it appear} to have Gribov copies. We will remove them in a moment, but first let us see how the copies affect dressed fields. Take Dirac's static electron (\ref{dirac}), and perform the above transformation. The dressed field behaves as
\begin{equation}
\Psi_s \to \exp(-ie k\cdot x)\, \Psi_s\;,
\end{equation}
because $h^{-1}_s$ is {\it insensitive} to the transformation between copies. Our supposed electron then acquires a gauge dependence due to the Gribov ambiguity, and therefore does not describe a physical field. The resolution of this strange result is well known -- the fields and transformations in this example are not an allowed part of configuration space. The fields must fall off sufficiently fast at spatial infinity, and the gauge transformations must tend to the identity, or more generally the centre \cite{Lavelle:1995ty}, neither of which holds here, and so our copies are removed.
Contrastingly, one can construct in SU(2) explicit examples of allowed, finite energy configurations, their Gribov copies and the transformations between them \cite{Henyey:1978qd, vanBaal:1991zw, Grotowski:1999ay, Ilderton:2007qy}. These non--trivial copies are explicitly nonperturbative, and, under transformations between them, the dressed fields indeed fail to be gauge invariant.
This examples shows both the importance of boundary conditions when discussing copies, and the effect copies have on the gauge invariance of our transformed variables, which we have learnt are dressed fields. So, the Gribov ambiguity, whose presence is signalled by the vanishing of the Faddeev--Popov determinant, could introduce a nonperturbative gauge dependence into the supposedly invariant fields of (\ref{lag2}). In the case of SU(2)$\times$U(1), this would include the $Z$ and $W$ bosons, which are observable. There must, therefore, be a way to circumvent the Gribov problem in this case. We now want to show that this is the case, and how it is related to the scalar's potential. (For other approaches to treating the Gribov ambiguity see, e.g. \cite{Jackiw:1977ng, Reinhardt:2008pr}).
We note that a gauge--covariant treatment of the Goldstone theorem shows that a spontaneously broken potential for the scalar (i.e. the potential (\ref{ssb2}) with $m^2<0$) will produce {\it gauge invariant} masses for the gauge fields in the Goldstone directions \cite{O'Raifeartaigh:1990ht}. Thus, we can generate mass terms which are insensitive to Elitzur's averaging. (It is of course common to work in a particular gauge, and other aspects of the theory will reflect this choice.) To analyse when our Faddeev--Popov determinant vanishes, we therefore follow here the usual approach to the Higgs mechanism, beginning with (\ref{lag1}) and expanding the scalar around the minimum of its potential as so,
\begin{equation}
\varphi_1(x)\equiv v + H(x)\;,
\end{equation}
in analogy to (\ref{expan1}). Now, finite energy considerations imply that the scalar fields of the theory vanish at infinity, see e.g.\ \cite{Ryder:1985wq}. Hence, we have $H(x)\to 0$. From the expressions (\ref{gauss}) we then have
\begin{equation}
\{G^a\epsilon^a(x), \varphi_2(y)\} = (\epsilon^1\varphi_3 +\epsilon^2\varphi_4 + \epsilon^3\left(v + H(x)\right) \delta(x-y)\;,
\end{equation}
etc., and so the Faddeev--Popov determinant (\ref{thedet}) is nonvanishing for any allowed $H(x)$. This is because the only potentially difficult configuration, $H(x)=-v$, is forbidden by the boundary conditions. If, however, $v=0$ then the field
\begin{equation}
\varphi_1(x) \equiv H(x)=0
\end{equation}
remains an allowed configuration and the determinant vanishes. (Going back to the original transformation, we noted that $|\Phi|=0$ was potentially problematic.) Therefore, when the scalar's potential has a trivial minimum, the Faddeev--Popov determinant, equivalently our functional measure, can vanish, which is a Gribov problem.
In the same fashion, suppose we take $\mathcal{L}_{ssb}=0$ and turn to the conformal interpretation, the functional determinant vanishes when there is a trivial asymptotic boundary condition $\rho(x)\to 0$. If, however, we impose asymptotic flatness via $\rho(x)\equiv \Lambda+H(x)$, with $\Lambda\not=0$ and $H(x)$ vanishing asymptotically, then the functional measure is nonvanishing and we can circumvent the Gribov problem. As we saw in Sect.~\ref{review}, $v$ and $\Lambda$ play the same role in mass generation, in the action. We have seen here that they also play the same role in avoiding the Gribov problem, and ensuring that the new variables are genuinely gauge invariant.
\subsection{Extension to the Weinberg--Salam model}
The classical transformation to SU(2) invariant variables proceeds very similarly in the Weinberg--Salam model \cite{Faddeev:2008qc}. The matrix $h[\Phi]$ has precisely the same form as in (\ref{h-def}), and the same transformation property (\ref{h-trans}), which is the essential ingredient. The initial fields -- the doublet $\Phi$, SU(2) gauge field $A_\mu^a$ and U(1) gauge field $Y_\mu$ -- are transformed into
\begin{equation}
\rho\,,\quad Z_\mu\,,\quad W^+_\mu\,,\quad W^-_\mu\,,\quad A_\mu\;.
\end{equation}
The U(1) of the original Lagrangian is not fixed by the transformation, and the resulting Lagrangian therefore describes a U(1) gauge theory, the photon $A_\mu$ being the remaining gauge non--invariant field. This abelian theory can be quantised by gauge fixing in the photon sector, as usual. We have the correct number $1 + 9 + 2=12$ degrees of freedom belonging to, respectively, the real scalar, three massive vector bosons and the photon $A_\mu$, with three SU(2) degrees of freedom factored out from the original Lagrangian. (The inclusion of fermions is straightforward \cite{Masson:2010vx}.)
As we have learnt, though, such a transformation only makes sense when conditions on the scalar field are such that we can circumvent the Gribov problem. Otherwise, Gribov copies introduce a nonperturbative gauge dependence into the new variables of the theory. Thus, we see that in the approach of \cite{Faddeev:2008qc}, the infra--red behaviour of the gravity sector has important consequences for the nonperturbative properties of the electroweak theory: the condition of asymptotic flatness is essential in order for the electroweak fields to be gauge invariant.
\subsection{Discussion}
In \cite{Masson:2010vx}\ it is suggested that the gauge invariant theory (\ref{lag2}), with the usual symmetry breaking potential (\ref{ssb2}), possesses two phases. If $m^2<0$ then the scalar can be expanded around the minimum of the potential, as above, generating the required mass terms. The other phase has $m^2>0$, and therefore no mass is generated. In this case, though, $\Phi\equiv 0$ is still part of configuration space, and while the classical Lagrangian (\ref{lag2}) appears to make sense, the transformation to it is not well defined because of the Gribov problem.
Indeed, the approach considered here will not work in all gauge theories, nor should it be expected to, as a straightforward counting of degrees of freedom can show \cite{Lunev:1994ty, Vlasov:1987vt}. Consider for example the $O(3)$ model which has three real scalar fields $\phi^a$ in the adjoint representation. After eliminating the $A^a_0$ multipliers, the theory describes a system with nine degrees of freedom -- there are twelve fields and three constraints coming from Gauss' law. The unitary gauge eliminates only two scalar degrees of freedom, and hence is not a complete gauge fixing. It is, however, a good partial gauge fixing, provided the scalar has a v.e.v./asymptotic condition, i.e.\ we can expand $\phi^3(x) = v + H(x)$, with $\phi^1$, $\phi^2$ unchanged. For then we can check that the Faddeev--Popov determinant of these gauge conditions with the $G^1$ and $G^2$ components of Gauss' law does not vanish (hence, the set $\phi^1,\phi^2,G^1,G^2$ form a second class set of constraints). Direct calculation yields the nonvanishing Poisson brackets
\begin{eqnarray}
\{G^1(x),\phi^2(y)\} &= (v+H(x))\delta(x-y)\;,\\
\{G^2(x),\phi^1(y)\} &= -(v+H(x))\delta(x-y)\;.
\end{eqnarray}
Thus, just as before, unitary gauge is a `good' gauge, but does not completely fix the gauge completely. In terms of the interpretations studied here, we could attempt to construct, from the scalar fields, a transformation to a gauge invariant set of variables. We know that this would be the field dependent transformation to $O(3)$ unitary gauge. While this exists, since unitary is a good partial gauge fixing here, the resulting variables would still transform under the unbroken part of the gauge group. They would not be gauge invariant. Hence, there will be no dressing analogous to (\ref{h-def}) which can be used to define gauge invariant fields; equivalently, there does not exist a transformation to gauge invariant variables of the type considered above.
\section{Conclusions}\label{Concs}
The electroweak sector of the standard model may be expressed entirely in terms of fields which are explicitly SU(2) gauge invariant. We have shown that the transformation to these variables is the field--dependent gauge transformation to unitary gauge and, as such, can suffer a Gribov problem at the quantum level. This reintroduces, nonperturbatively, an unphysical gauge dependence into the new fields. We have shown, though, that this problem is circumvented precisely when the scalar field has a spontaneously broken potential. It is not surprising that the ability to identify physical variables in Higgs models has a subtle, but in this case explorable, dependence on the topology of the configuration space, as this is which really distinguishes non--abelian gauge theories from their abelian counterparts.
Note that it is not the breaking mechanism {\it per se} which is required to give a good transformation, but rather the existence of the nontrivial field minimum. Typically, of course, this comes about because of the potential terms. When the invariant scalar is interpreted as a conformal metric factor, though, the required `v.e.v.' of the field is generated by the condition of asymptotic flatness. Again, this makes manifest the importance of topology on resolving the Gribov ambiguity: here it is the boundary conditions living on the `boundary' of spacetime which play a key role.
The claim in \cite{Masson:2010vx} that similar transformations to those given here can be applied to the whole standard model Lagrangian is a little suspect -- we know there are theories in which no such transformation exists. Additionally, when it does, the boundary conditions on the fields need to be carefully handled, see also \cite{Vlasov:1987vt}, as they have direct implication for the observability of the fields. Consider SU(3), where there is no symmetry breaking. In this case there are no physical descriptions of an observable asymptotic gauge boson and hence gluons are confined.
\subsubsection*{Acknowledgements}
A. I. is supported by the European Research Council under Contract No. 204059-QPQV.
\section*{References}
|
\section{Introduction}
CGH (Computer-generated-hologram) has the ability to correctly record and reconstruct a light wave for a 3D object.
Electroholography\cite{benton} using the CGH technique is attractive as a 3D display, because the CGH technique has remarkable features ; however, due to two significant problems, it is difficult to develop a practical 3D display system using electroholography.
One problem is the need for an SLM (spatial light modulator) that can display a CGH with large area and high resolution, because the resolution of a CGH is that of wavelength-order\cite{maeno, active,lee,takaki}.
The other problem is the enormous computational time required for generating a CGH.
This paper focuses on this problem.
Assuming that a 3D object is composed of $N$ point light sources, the formula for computing a CGH is expressed as:
\begin{eqnarray}
I(x_h,~y_h)&=&\sum_j^N {A_j}{\rm cos}(\frac{2 \pi}{\lambda} (\frac{(p x_h-p x_j)^2+(p y_h-p y_j)^2}{2z_j}))\nonumber \\
&=&\sum_j^N {A_j}{\rm cos}(P_j ((x_h-x_j)^2+(y_h-y_j)^2)),
\label{eqn:cgh_basic}
\end{eqnarray}
where, $I(x_h, y_h)$ is the light intensity on a CGH, $(x_h,y_h)$ and $(x_j,y_j,z_j)$ are the coordinates for the CGH and a 3D object, $A_j$ is the light intensity of the 3D object, $\lambda$ is the wavelength of the reference light, and $P_j=\pi p^2 / (\lambda z_j)$, where $p$ is the sampling interval on the CGH plane.
Note that the coordinates $(x_j, y_j)$ and $(x_h, y_h)$ are normalized by $p$.
The computational complexity of the above formula is O($N N_x N_y$), where $N_x$ and $N_y$ are the horizontal and vertical sampling numbers of the CGH.
This creates very enormous computational complexity.
To solve this problem, several software approaches have been proposed: for example, recurrence approaches\cite{yoshi, matsu, recurrence}, and the look-up table methods\cite{lut, image_hol, lutgpu}.
Another approach to dramatically reduce the calculation time is the hardware approach, such as FPGA (field-programmable gate array) and GPU (graphics processing unit).
We have designed and built special-purpose computers for holography using FPGA technology, called HORN (HOlographic ReconstructioN) \cite{ shimo_unit, horn5, horn6}.
The FPGA-based approaches showed excellent computational speed, however, the approach has the following restrictions: the high cost for developing the FPGA board, long development time and technical know-how required for the FPGA technology.
GPU-based approaches have already been applied to the optics field.
Especially, CGH calculations \cite{masuda3, Ahrenberg,lutgpu, Onural} and reconstruction calculations in digital holography \cite{ gwo, shimo_dhm} are used to accelerate the calculation.
In 2007, NVIDIA released a new architecture of GPU and its software development environment, CUDA (Compute Unified Device Architecture).
Using CUDA allows us to program GPU easier than prior software developments, such as HLSL, Cg language and so forth.
Since the release, many papers using NVIDIA GPU and CUDA have been published in optics.
On the other hand, more recently in December 2009, a new GPU of the HD5000 series (RV870) made by AMD was released.
The RV870 GPU has new architecture and its software environment, OpenCL (Open Computing Language).
The architecture of the RV870 GPU is different from that of the NVIDIA GPU.
The RV870 GPU has huge potential for fast calculation because one GPU chip has over 1,000 floating-point number processors, while one NVIDIA GPU chip has about 200 floating-point number processors.
However, fast CGH calculation using the RV870 GPU has not been reported so far.
In this paper, we report fast CGH calculation using RV870 GPU and OpenCL.
Using these, we can calculate $1,920 \times 1,024$ resolution of a CGH from a 3D object consisting of $1,024$ points in $30$ milli-seconds.
To the best of our knowledge, this article is the first report of using the RV870 GPU and OpenCL in optics.
In addition, we compare the calculation performance between the RV870 GPU and the GPU made by NVIDIA.
In Section 2, we describe a fast CGH calculation on AMD RV870 and OpenCL.
In Section 3, we show and compare the performance between the RV870 GPU and the GPU made by NVIDIA.
In Section 4, we conclude this work.
\section{Fast calculation of computer-generated-hologram on AMD RV870 and OpenCL}
\begin{figure}[htb]
\centerline{
\includegraphics[width=12cm]{fig_amd.eps}}
\caption{Architecture of RV870 GPU. (a) Outline of RV870 GPU chip (b) Thread processor.}
\label{fig:amd}
\end{figure}
The architecture of RV870 GPU is shown in Fig.\ref{fig:amd}.
The top level of the GPU consists of many SIMD (Single Instruction Multiple Data) engines.
The SIMD engine has 16 thread processors (TP) and a shared memory, which is small and high-speed.
In addition, the thread processor has four stream cores and one T-stream core.
The stream core is a simple floating-point-number operation unit.
And, the T-stream core also has a floating-point-number operation unit and special function unit.
The special function unit can calculate special functions at high speed, such as trigonometric function, logarithm function and so on.
The stream cores in the same SIMD engine operate by the same instructions; therefore, the SIMD engine is similar to a SIMD processor.
Calculation on the GPU using OpenCL is executed using the following steps:
(1) We initialize a GPU using OpenCL API (Application Program Interface) functions.
(2) We allocate the required amount of memory on a device memory in Fig.\ref{fig:amd}. The device memory is large amount, but large latency access of memory.
(3) We send an input data to the device memory.
(4) We send a kernel function from the host computer to the GPU. The kernel function is compiled to native code of GPU using the OpenCL compiler. The GPU executes the kernel function.
(5) We receive a calculated result from the device memory.
(6) We release the device memory and GPU resources.
Figure \ref{fig:block_list}(a) shows the outline of the CGH calculation on the RV870 GPU with OpenCL.
When calculating a CGH with the resolution of $N_x \times N_y$, we need to divide the CGH area into $group$s with the size of $T_x \times T_y$.
Therefore, the number of $group$s is $N_x/T_x \times N_y/T_y$.
In addition, each $group$ has $T_x \times T_y$ $item$s
(In the CUDA, $group$ and $item$ are equivalent to $block$ and $thread$, respectively).
Each $group$ is allocated to SIMD engines and each $item$ simultaneously calculate Eq.(\ref{eqn:cgh_basic}) by each stream core on an SIMD engine.
In Fig.\ref{fig:block_list}(b), we show the kernel source code of the CGH calculation on the RV870 GPU with OpenCL.
The source code is not optimized because we understand it easily.
The optimization is shown in the next subsection.
Each $group$ and $item$ have the indices, group\_id and local\_id.
The OpenCL functions, get\_group\_id(0) and get\_group\_id(1), give us the horizontal and vertical indices of group\_ids respectively.
The OpenCL functions, get\_local\_id(0) and get\_local\_id(1), also give us the horizontal and vertical indices of local\_ids respectively.
The arguments of the kernel function are a CGH data ($d\_hol$), an object data ($d\_obj$), the number of object points ($N$) and the CGH size ($N_x, N_y$).
An object data ($d\_obj$) consists of the coordinates and the intensity as four float data ($float4$).
In lines 5, 6 and 7 of the Fig.\ref{fig:block_list}(b), the variables $x$ and $y$ calculate the coordinates ($x_h, y_h)$ on the CGH plane and $adr$ calculates the address of the device memory for storing the calculation result $I(x_h,y_h)$.
In lines 11 to 16, a CGH point $I(x_h,y_h)$ can be calculated by iterating for $N$.
Although seeming to execute only one kernel, in fact, each stream core corresponding to $local\_id$ and $global\_id$ can perform the kernel in parallel.
When calculating a CGH with $1,920 \times 1,024$ from a 3D object composed of $1,024$ points, the kernel with $T_x \times T_y=16 \times 16$ took about $\rm 215 ms$.
The calculation speed of the kernel is slow.
\begin{figure}[htb]
\centerline{
\includegraphics[width=15cm]{fig_block_list.eps}}
\caption{(a) Outline of CGH calculation on RV870 GPU and OpenCL. (b) Kernel for the CGH calculation using OpenCL without optimization.}
\label{fig:block_list}
\end{figure}
\subsection{Optimization}
\label{sec:opt}
The previous source code is not optimized.
In this section, we optimize the previous source code to obtain more acceleration speed.
Figure \ref{fig:list2} shows the optimized kernel function from Fig.\ref{fig:block_list}(b).
For more acceleration, we applied the following optimization techniques to the previous kernel: (1) Recurrence algorithm (2) Shared memory (3) Loop unrolling (4) Vectorization (5) Native instruction.
We proposed a fast CGH computation method using two recurrence formulas \cite{recurrence, shimo_unit, horn5,horn6}.
Our recurrence algorithm can compute the phase component of the cosine function in Eq.(\ref{eqn:cgh_basic}) by two recurrence formulas.
The recurrence algorithm is as follows:
\begin{equation}
I(x_h+n, y_h)=\sum_j^N {A_j}{\rm cos}(\Gamma_n) ~~~ \Gamma_n=\Gamma_{n-1}+\delta_{n-1} ~~~ \delta_{n}=\delta_{n-1}+\Delta
\label{eqn:cgh_rec}
\end{equation}
Here, we define $\Gamma_0=P_j((x_h-x_j)^2+(y_h-y_j)^2)$, $\delta_0 = P_j (2 (x_h-x_j) + 1)$, $\Delta=2P_j$.
Eventually, we can compute the phase $\Gamma_n$ at the next coordinate by the two recurrence formulas.
For more details, see Ref.\cite{recurrence}
In lines 15 to 18, we copy the object data from the device memory ($d\_obj$) to a shared memory ($s\_obj$).
The shared memory can store 256 object points at a time because the shared memory is small and high-speed.
Therefore, in the 13, we must iterate $N/256$ times.
Note that $barrier(CLK\_LOCAL\_MEM\_FENCE)$ means a barrier synchronization in line 18.
It is equivalent to the $syncthreads$ function in the CUDA.
Loop unrolling is a well-known technique for optimizing a kernel function.
It can be realized by reducing the number of iterations and replicating the body of the loop.
Benefits of the loop unrolling are the capable to decrease the loop frequency, branch instructions and conditional instructions.
In the optimized kernel, we applied the loop unrolling to the loop of object points.
In lines 20 to 51 in Fig.\ref{fig:list2}, we can perform four object points per one iteration of the loop.
In addition, we vectorize the operations in the loop using the $float4$ type, in order to handle four object points at a time.
For example, in line 22, we can calculate the four subtractions simultaneously.
In the same way, the kernel can handle eight CGH points using the $float8$ type at same time in lines 40 to 50.
In lines 42 to 45, we used native cosine functions, instead of the normal cosine function shown in Fig.\ref{fig:block_list}(b).
The native cosine function can compute the fast cosine function using the hardware.
\begin{figure}[htb]
\centerline{
\includegraphics[width=13cm]{fig_list2.eps}}
\caption{Kernel for the CGH calculation using OpenCL with optimization.}
\label{fig:list2}
\end{figure}
\subsection{Results}
Table \ref{tbl:time} shows a comparison of the calculation times for a CPU alone, NVIDIA GPU and an AMD RV870 GPU.
The size of the CGH is $1,920 \times 1,024$.
The specifications of the personal computer are as follows: Intel Core 2 Quad Q6600 (We used one core for the calculation), 2 GB of memory, Microsoft Windows XP SP3.
We used a GeForce GTX260 as the NVIDIA GPU board and its software development environment of CUDA version 2.3, and a RADEON HD5850 as the AMD GPU board and its software development environment of StreamSDK version2.0.
The RADEON HD5850 GPU has 1,440 stream cores (namely, 18 SIMD engines) with the clock frequency of 725MHz.
We can see that the optimization method for the AMD GPU described in Section \ref{sec:opt} can perform more than ten times faster than that without the optimization.
In the calculation times for the NVIDIA GPU in the table, we optimized the kernel for the NVIDIA GPU using the same method as described in Section \ref{sec:opt}: namely, recurrence algorithm, shared memory, loop unrolling, vectorization, native instruction.
And, in the calculation times for the CPU alone in the table, we used Eq.(\ref{eqn:cgh_rec}) for the CGH calculation.
All calculation times using AMD and NVIDIA are superior to those using the CPU alone.
In addition, the AMD GPU can calculate a CGH approximately two times faster than the NVIDIA GPU.
\begin{table}[htb]
\caption{Comparison of calculation times on CPU alone, NVIDIA GPU and AMD GPU.}
\begin{center}
\scalebox{0.7}{
\begin{tabular}{|c|c|c|c|c|}
\hline
\multicolumn{ 1}{|c|}{Number of object points} & \multicolumn{ 4}{c|}{Time(ms)} \\ \cline{ 2- 5}
\multicolumn{ 1}{|c|}{} & \multicolumn{ 1}{c|}{CPU} & \multicolumn{ 1}{c|}{NVIDIA GPU} & \multicolumn{ 2}{c|}{AMD GPU} \\ \cline{ 4- 5}
\multicolumn{ 1}{|c|}{} & \multicolumn{ 1}{c|}{} & \multicolumn{ 1}{c|}{} & \multicolumn{1}{c|}{Without optimization} & \multicolumn{1}{c|}{With optimization} \\ \hline
512 & $30 \times 10^3$ & 33 & 215 & 19 \\ \hline
1024 & $59 \times 10^3$ & 59 & 422 & 31 \\ \hline
1536 & $88 \times 10^3$ & 85 & 630 & 44 \\ \hline
2048 & $115 \times 10^3$ & 112 & 838 & 56 \\ \hline
2560 & $146 \times 10^3$ & 139 & 1045 & 68 \\ \hline
3072 & $174 \times 10^3$ & 165 & 1252 & 80 \\ \hline
3584 & $204 \times 10^3$ & 192 & 1464 & 93 \\ \hline
\end{tabular}
}
\end{center}
\label{tbl:time}
\end{table}
\section{Conclusion}
In this paper, we described a fast CGH calculation using an AMD RV870 GPU with new architecture and its new software development environment, OpenCL.
Many fast CGH calculation methods using a NVIDIA GPU and the CUDA have already been reported in optics field; however, a study using the RV870 GPU has not been reported so far.
To the best of our knowledge, this article is the first report of using the RV870 GPU and OpenCL in optics.
Using the RV870 GPU and OpenCL, we can calculate $1,920 \times 1,024$ resolution of a CGH from a 3D object consisting of $1,024$ points in about 30 ms.
The calculation speed can realize approximately two times faster than the NVIDIA GPU.
This research was partially supported by the Ministry of Internal Affairs and Communications, Strategic Information and Communications R\&D Promotion Programme (SCOPE), 2009, and Japan Society for the Promotion of Science (JSPS), Grant-in-Aid for Scientific Research
(C) (21500094).
|
\section*{Acknowledgement}
We are grateful to K. Motegi, M. Shiroishi, and M. Takahashi for discussions and helpful comments. This work was supported by
World Premier International Research Center Initiative on Materials Nanoarchitectonics,
the University of Tsukuba Research Initiative, and KAKENHI 18340112, 20029004, 20046002, 20046015, 20654034, 20740206, 21740281, and 21740285.
|
\section{Introduction}
Let $\mathfrak{S}_n$ denote the set of permutations of $\{ 1,2,\ldots, n\}$, and let $\mathfrak{S}_n(\tau)$ denote the set of permutations of $\{ 1,2,\ldots, n\}$ which avoid the pattern $\tau$, i.e., which do not contain a subsequence order-isomorphic to $\tau$. More generally, let $\mathfrak{S}_n(\tau_1,\tau_2,\ldots, \tau_k)$ denote the set of permutations of $\{ 1,2,\ldots, n\}$ which avoid all patterns $\tau_i$ for $i=1,2,\ldots, k$.
For any permutation $\pi \in \mathfrak{S}_n$, the \emph{reverse-complement} of $\pi$ is $\mathrm{rc}(\pi)=\pi'$ where $\pi'(i) = (n+1) - \pi(n+1-i)$ for each $i=1,2,\ldots, n$. Also, a permutation $\pi \in \mathfrak{S}_n$ is said to be \emph{centrosymmetric} if and only if $\mathrm{rc}(\pi) = \pi$. Denote the set of centrosymmetric permutations by $\mathcal{C}_{n}$, and the set of centrosymmetric involutions by $\mathcal{CI}_{n}$. Also $\mathcal{C}_{n}(\tau)$, $\mathcal{CI}_{n}(\tau_1,\tau_2,\ldots, \tau_k)$ etc. are defined in the usual way.
It is well known that a permutation is centrosymmetric if and only if both of the standard Young tableaux yielded by the Robinson-Schensted algorithm are invariant under the Sch\"{u}tzenberger involution (see \cite{Schutzenberger} and \cite{Knuth} for more details). In work by Egge \cite{egge:restricted_symmetric_perms_L3}, permutations and involutions were enumerated which are invariant under the natural action of a subgroup of the symmetry group of a square. This included enumeration and Wilf-equivalence classification of centrosymmetric permutations and involutions avoiding all patterns of length $3$. In a recent contribution also by the same author, $\left| \mathcal{C}_{2n}(k \; k-1 \cdots 2 1) \right|$ and $\left| \mathcal{CI}_{2n}(k \; k-1 \cdots 2 1) \right|$ were evaluated by counting self-evacuating standard Young tableaux and using the Robinson-Schensted correspondence \cite{egge:restricted_symmetric_perms_L4}. Other results along this line which have been achieved are the enumeration of the \emph{vexillary involutions} (i.e., the set $\mathcal{CI}_{2n}(2143)$) by Guibert and Pergola \cite{Guibert}, and the set $\mathcal{C}_{2n}(123,2413)$ by Ostroff and Lonoff \cite{Lonoff_ostroff}. In \cite{Barnabei_bonetti_silimbani}, Barnabei \emph{et al.} enumerated many classes of pattern-avoiding centrosymmetric involutions by using a bijection with labeled Motzkin paths.
In this paper, centrosymmetric permutations and involutions avoiding the patterns $1243$ and $2143$ are enumerated. We begin by characterizing the set of permutations avoiding $1243$ and $2143$ whose images under the reverse-complement operation also avoid these patterns; to this end we make use of a result by Egge and Mansour \cite{egge_mansour} which puts permutations avoiding $1243$ and $2143$ in bijective correspondence with the set of Schr\"{o}der paths of an appropriate length. The characterization we require is particularly simple in the Schr\"{o}der path domain. We then enumerate centrosymmetric permutations avoiding $1243$ and $2143$ by enumerating the corresponding Schr\"{o}der paths. The corresponding enumeration for involutions is subsequently achieved by first proving that a Schr\"{o}der path $p$ corresponds to an involution under the bijection of Egge and Mansour if and only if $p$ is symmetric with respect to path reversal. In particular, the centrosymmetric involutions which avoid $1243$ and $2143$ are shown to be enumerated by the Pell numbers.
It was proved in \cite{Albert} that the cardinality of the full class of permutations avoiding $1243$ and $2143$ whose reverse complements also avoid these patterns (i.e., the sequence $\left| \mathfrak{S}_n(1243, 2143, 2134) \right|$) has a \emph{rational} generating function; it would be interesting to investigate whether similar methods can prove the rationality or otherwise of the generating functions of $\left| \mathcal{C}_{n}(1243,2143) \right|$ and $\left| \mathcal{CI}_{n}(1243,2143) \right|$.
\section{Permutations avoiding $1243$ and $2143$, and Schr\"{o}der paths}
The large Schr\"oder numbers $r_{n}$ are defined by $r_0=1$ and for all $n \ge 1$,
\begin{equation*}
r_{n} = r_{n-1} + \sum_{k=1}^{n} r_{k-1} r_{n-k}.
\end{equation*}
A \emph{Schr\"oder prefix} is a lattice path beginning at the point $(0,0)$ which may take only a finite number of steps from the set $\{\mathsf{e} =(1,0), \mathsf{n} =(0,1), \mathsf{d} =(1,1) \}$ and which does not pass below the line $y=x$. Denote by $\mathcal{S}$ the set of all Schr\"oder prefixes. For $n \ge 1$, a \emph{Schr\"oder path} of length $n$ is a Schr\"oder prefix which terminates at the point $(n,n)$. Let $\mathcal{S}_n$ denote the set of Schr\"{o}der paths of length $n$. In this paper, such Schr\"{o}der paths will sometimes be denoted by the corresponding sequence of letters from $\{\mathsf{e}, \mathsf{n}, \mathsf{d}\}$. Also denote by $\mathcal{S}_0$ the set containing the null path $\emptyset$ having length $0$. For any pair of Schr\"{o}der paths $p \in \mathcal{S}_m$ and $q \in \mathcal{S}_n$, denote by $p \, q \in \mathcal{S}_{m+n}$ the \emph{concatenation} of these Schr\"{o}der paths. The set $\mathcal{S}_n$ is enumerated by $r_n$ for $n \ge 0$. Also, the permutations $\mathfrak{S}_{n+1}(1243,2143)$ for $n \ge 0$ are called \emph{Schr\"oder permutations} since they are enumerated by the large Schr\"oder numbers $r_n$.
\begin{lemma}
A permutation $\pi$ lies in $\mathfrak{S}_n(1243,2143)$ if and only if $\pi^{-1}$ lies in $\mathfrak{S}_n(1243,2143)$.
\label{lemma:pi_inv}
\end{lemma}
\begin{proof}
This follows immediately from the observation that any occurrence of a pattern $\tau \in \{1243,2143\}$ in the permutation $\pi$ corresponds directly to an occurrence of the same pattern $\tau$ in $\pi^{-1}$.
\end{proof}
\begin{lemma}
Let $\pi_1, \pi_2 \in \mathfrak{S}_n$. Then $\pi_1 = \mathrm{rc}(\pi_2)$ if and only if $\pi_1^{-1} = \mathrm{rc}(\pi_2^{-1})$.
\label{lemma:rc_inv}
\end{lemma}
\begin{proof}
This follows immediately from the observation that the substitution $j = \pi_1(i)$ allows the condition $\pi_1(i) + \pi_2(n+1-i) = n+1$ for all $i=1,2,\ldots,n$ to be rearranged as $\pi_1^{-1}(j) + \pi_2^{-1}(n+1-j)=n+1$ for all $j=1,2,\ldots,n$.
\end{proof}
\begin{definition}
For $t\in \{ 1,2,\ldots, n-1 \}$, a Schr\"oder path $p \in \mathcal{S}_n$ which contains an occurrence of $\mathsf{d}$ joining $(t-1,t)$ to $(t,t+1)$ is said to have a \emph{level feature} at $t$, and a Schr\"oder path $p\in \mathcal{S}_n$ which contains an occurrence of $\mathsf{e}\mathsf{n}$ joining points $(t-1,t)$ to $(t,t)$ to $(t,t+1)$ is said to have a \emph{notch feature} at $t$. A Schr\"oder path $p\in \mathcal{S}_n$ is said to have a \emph{feature} at $t\in \{ 1,2,\ldots, n-1 \}$ if it has either a level feature or a notch feature at $t$.
\label{def:features}
\end{definition}
\begin{definition}
For any Schr\"oder path $p\in \mathcal{S}_n$, the \emph{latest} (resp. level or notch) \emph{feature} of $p$ is the largest $t\in \{ 1,2,\ldots, n-1 \}$ at which $p$ contains a (resp. level or notch) feature, or is equal to $0$ if there is no such $t$. Also, for any Schr\"oder path $p\in \mathcal{S}_n$, the \emph{earliest} (resp. level or notch) \emph{feature} of $p$ is the smallest $t\in \{ 1,2,\ldots, n-1 \}$ at which $p$ contains a (resp. level or notch) feature, or is equal to $n$ if there is no such $t$.
\label{def:latest_earliest_features}
\end{definition}
\begin{definition}
For $p \in \mathcal{S}_n$, the \emph{reversed path} $\mathrm{rev}(p)$ is the path obtained by applying the $\{ \mathsf{e},\mathsf{n},\mathsf{d} \}$ steps of the path in reverse order, then replacing all occurrences of $\mathsf{e}$ by $\mathsf{n}$ and vice versa. Also for $p \in \mathcal{S}_n$, $\psi(p)$ is the path obtained by replacing all level features at $t\in \{ 1,2,\ldots, n-1 \}$ by notch features at $t$ and vice versa.
\label{def:bijective_involutions}
\end{definition}
Egge and Mansour~\cite [\S 4]{egge_mansour} define a bijection $\varphi:\mathcal{S}_n \mapsto \mathfrak{S}_{n+1}(1243,2143)$ from the set of Schr\"oder paths of length $n$ to the set of permutations {\bf of length} $\boldsymbol{n+1}$ avoiding $1243$ and $2143$. In the following we briefly describe the bijection $\varphi$; more details may be found in~\cite{egge_mansour}. An illustrative example of the bijection $\varphi$ will follow this formal description.
Let $p\in \mathcal{S}_n$ and let $s_i$ denote the transposition $(i,i+1)$ for each $i=1,2,\ldots, n$. We also use the convention $s_i s_j \pi = s_i(s_j(\pi))$.
\begin{description}
\item[Step 1] Let $\Gamma_{r,s}$ denote the unit square which has diagonally opposite corners at $(r-1,s-1)$ and $(r,s)$. For each such square $\Gamma_{r,s}$ whose top-left corner lies below the path $p$ and above the line $y=x$, place the label $r$ in the top-left corner. We will next construct a sequence of permutations $\sigma_i$, $i=1,2,\ldots,k$; initially set $i=1$.
\item[Step 2] Locate the labeled square $\Gamma_{r,s}$ with largest label $r$. The permutation $\sigma_i$ is equal to the sequence of transpositions $s_t$ where the subscript $t$ sequentially takes on all label values starting with this $r$ and continuing diagonally to the lower left until an $\mathsf{n}$ step is encountered. Remove the labels which were used as subscripts. If there are now no more labeled squares, we are finished; otherwise set $i$ to $i+1$ and repeat Step 2.
\item[Step 3]
The image $\varphi(p)$ of the Schr\"oder path $p$ under the bijection $\varphi$ is defined as
\begin{equation}
\varphi(p) = \sigma_k \sigma_{k-1} \cdots \sigma_1 (n+1,n,n-1,\ldots, 2,1) \; .
\label{eq:Egge_Mansour_bijection}
\end{equation}
\end{description}
\begin{figure}
\begin{center}\includegraphics[%
width=0.5\columnwidth,
keepaspectratio]{Schroder1.eps}
\end{center}
\caption{\label{cap:Schroder1}The Schr\"oder path $p$ of Example \ref{ex:running_example}.}
\end{figure}
The following example illustrates Definitions \ref{def:features} and \ref{def:latest_earliest_features} as well as the bijection $\varphi$, and will also serve as a useful example when we consider the reverse-complements of permutations in $\mathfrak{S}_{n+1}(1243,2143)$.
\begin{example}\label{ex:running_example}
Consider the path $p = \mathsf{n}\mathsf{e}\mathsf{n}\nn\mathsf{n}\nn\mathsf{e}\ee\mathsf{e}\mathsf{d}\mathsf{e}\mathsf{n}\nn\mathsf{e}\mathsf{d}\mathsf{e} \in \mathcal{S}_9$ shown in Figure \ref{cap:Schroder1}.
The path $p$ has an (earliest) level feature at $t=5$ and another (latest) level feature at $t=8$. Also, the path $p$ has an (earliest) notch feature at $t=1$ and another (latest) notch feature at $t=6$. To compute $\varphi(p)$, note that here we have
$\sigma_1=s_9s_8s_7$, $\sigma_2=s_7$, $\sigma_3= s_6s_5s_4s_3s_2$, $\sigma_4=s_4s_3s_2$, $\sigma_5=s_3s_2$, $\sigma_6=s_2$ and $\sigma_7=s_1$. So
\begin{eqnarray*}
\varphi(p) &=& \sigma_7 \sigma_6 \sigma_5 \sigma_4 \sigma_3 \sigma_2 \sigma_1 (10,9,8,7,6,5,4,3,2,1) \\
&=& s_1 \; s_2 \; s_3s_2 \; s_4s_3s_2 \; s_6s_5s_4s_3s_2 \; s_7 \; s_9s_8s_7 (10,9,8,7,6,5,4,3,2,1) \\
&=& (5,10,6,7,8,2,9,3,1,4) \in \mathfrak{S}_{10}(1243,2143).
\end{eqnarray*}
A detailed evolution of $(10,9,8,7,6,5,4,3,2,1)$ to $\varphi(p)$ under the application of the permutations $\sigma_i$ for $i=1,2,\ldots,k$ is provided in Table \ref{ttable_1}.
\end{example}
\begin{lemma}
A Schr\"oder path $p \in \mathcal{S}_n$ contains an occurrence of $\mathsf{d}$ joining $(t-1,t-1)$ to $(t,t)$ for $t \in \{ 1,2,\ldots,n \}$ if and only if the largest $t$ numbers in $\{ 1,2,\ldots, n+1\}$ occupy the first $t$ positions of $\pi = \varphi(p)$.
\label{lemma:no_mixing}
\end{lemma}
\begin{proof}
With reference to (\ref{eq:Egge_Mansour_bijection}), note that $\pi = \varphi(p)$ is obtained by applying the permutations $\sigma_1, \sigma_2, \ldots, \sigma_k$ to a permutation in which the largest $t$ numbers in $\{ 1,2,\ldots, n+1\}$ occupy the first $t$ positions. Next observe that the path $p \in \mathcal{S}_n$ contains an occurrence of $\mathsf{d}$ joining $(t-1,t-1)$ to $(t,t)$ if and only there is no occurrence of $s_t$ among the permutations $\sigma_1, \sigma_2, \ldots, \sigma_k$. This is exactly the condition under which there will be no ``mixing'' of the two parts of the permutation, i.e., the largest $t$ numbers in $\{ 1,2,\ldots, n+1\}$ will remain in the first $t$ positions of $\pi = \varphi(p)$.
\end{proof}
\begin{theorem}
If $p$ is a Schr\"{o}der path of length $n \ge 1$, and $\pi = \varphi(p)$, then $\varphi(\mathrm{rev}(p)) = \pi^{-1}$.
\label{lemma:inverse_through_EM_bijection}
\end{theorem}
\begin{proof}
We prove this by induction on the length $n$ of the path $p$. The base case $n=1$ holds trivially. Next, assume the result holds for all paths of length at most $n-1$ (where $n > 1$), and consider a path $p$ of length $n$. Denote $\pi = \varphi(p)$ and $\tilde{\pi} = \varphi(\mathrm{rev}(p))$. Our goal is to show that $\tilde{\pi} = \pi^{-1}$.
If no point $(t,t)$ lies on the path $p$ for $t = 1,2,\ldots,n-1$, then we may write $p = \mathsf{n} \, p_1 \, \mathsf{e}$ for some Schr\"{o}der path $p_1$ of length $n-1$. Let $\pi_1 = \varphi(p_1)$. Then, with reference to obtaining $\pi = \varphi(p)$ from $(n+1,n,n-1,\ldots,2,1)$ via (\ref{eq:Egge_Mansour_bijection}), first $\sigma_1 = s_n s_{n-1} \cdots s_2 s_1$ is applied; this is a cyclic shift which yields $\pi(n+1) = n+1$. The subsequent sequence of permutations $\{\sigma_i \; : \; i \ge 2 \}$ is the same as that involved in obtaining $\varphi(p_1)$ from $(n,n-1,\ldots,2,1)$ via (\ref{eq:Egge_Mansour_bijection}); thus $\pi(i) = \pi_1(i)$ for all $i=1,2,\ldots,n$. Since $\mathrm{rev}(p) = \mathsf{n} \, \mathrm{rev}(p_1) \, \mathsf{e}$, using the same argument together with the induction hypothesis yields $\tilde{\pi}(i) = \pi^{-1}_1(i)$ for all $i=1,2,\ldots,n$, and $\tilde{\pi}(n+1) = n+1$. Thus $\tilde{\pi} = \pi^{-1}$.
\begin{table}
\begin{center}
\caption{\label{ttable_1}
Illustration of the bijection $\varphi(p)$ for the Schr\"oder path $p \in \mathcal{S}_9$ given in Example \ref{ex:running_example}. The table shows the evolution of the permutation from $(10,9,8,7,6,5,4,3,2,1)$ towards $\varphi(p)$ as each permutation $\sigma_i$ is applied for $i=1,2,\ldots,k$.}
{\small
$\begin{array}{|c|c|}
\hline
\begin{array}[t]{c}
\mathrm{Permutation} \\
\hline
\mathrm{Start} \\
\sigma_1 = s_9 s_8 s_7 \\
\sigma_2 = s_7 \\
\sigma_3 = s_6 s_5 s_4 s_3 s_2 \\
\sigma_4 = s_4 s_3 s_2 \\
\sigma_5 = s_3 s_2 \\
\sigma_6 = s_2 \\
\sigma_7= s_1 \\
\end{array} &
\begin{array}[t]{c}
\mathrm{Result} \\
\hline
10 \; 9 \; 8 \; 7 \; 6 \; 5 \; 4 \; 3 \; 2 \; 1 \\
10 \; 9 \; 8 \; 7 \; 6 \; 5 \; {\mathbf 3} \; {\mathbf 2} \; {\mathbf 1} \; {\mathbf 4} \\
10 \; 9 \; 8 \; 7 \; 6 \; 5 \; {\mathbf 2} \; {\mathbf 3} \; 1 \; 4 \\
10 \; {\mathbf 8} \; {\mathbf 7} \; {\mathbf 6} \; {\mathbf 5} \; {\mathbf 2} \; {\mathbf 9} \; 3 \; 1 \; 4 \\
10 \; {\mathbf 7} \; {\mathbf 6} \; {\mathbf 5} \; {\mathbf 8} \; 2 \; 9 \; 3 \; 1 \; 4 \\
10 \; {\mathbf 6} \; {\mathbf 5} \; {\mathbf 7} \; 8 \; 2 \; 9 \; 3 \; 1 \; 4 \\
10 \; {\mathbf 5} \; {\mathbf 6} \; 7 \; 8 \; 2 \; 9 \; 3 \; 1 \; 4 \\
{\mathbf 5} \; \mathbf{10} \; 6 \; 7 \; 8 \; 2 \; 9 \; 3 \; 1 \; 4 \\
\end{array} \\
\hline
\end{array}$
}
\end{center}
\end{table}
If $p$ terminates in a $\mathsf{d}$ step, then we may write $p = p_1 \, \mathsf{d}$ for some Schr\"{o}der path $p_1$ of length $n-1$. Let $\pi_1 = \varphi(p_1)$. Then, the sequence of permutations $\{\sigma_i\}$ involved in obtaining $\varphi(p)$ from $(n+1,n,n-1,\ldots,2,1)$ via (\ref{eq:Egge_Mansour_bijection}) is the same as that involved in obtaining $\varphi(p_1)$ from $(n,n-1,\ldots,2,1)$ via (\ref{eq:Egge_Mansour_bijection}); thus $\pi(i) = \pi_1(i) + 1$ for all $i=1,2,\ldots,n$, and $\pi(n+1) = 1$. Also, since $\mathrm{rev}(p) = \mathsf{d} \, \mathrm{rev}(p_1)$, the sequence of permutations used to obtain $\varphi(\mathrm{rev}(p))$ from $(n+1,n,n-1,\ldots,2,1)$ via (\ref{eq:Egge_Mansour_bijection}) is the same as that used to obtain $\varphi(\mathrm{rev}(p_1))$ from $(n,n-1,\ldots,2,1)$ via (\ref{eq:Egge_Mansour_bijection}), but with each $s_t$ replaced by $s_{t+1}$. Together with the induction hypothesis, this yields $\tilde{\pi}(i) = \pi^{-1}_1(i-1)$ for all $i=2,3,\ldots,n+1$, and $\tilde{\pi}(1) = n+1$. Thus $\tilde{\pi} = \pi^{-1}$.
Therefore, hereafter we need only prove the inductive step for Schr\"{o}der paths $p$ of length $n$ which terminate in an $\mathsf{e}$ step and such that the point $(t,t)$ lies on $p$ for some $t \in \{ 1,2,\ldots,n-1 \}$. For such paths, there exists a largest value of $t \in \{1,2,\ldots,n-1\}$ for which $(t,t)$ lies on the path $p$; denote by $n-i$ this largest value, where $i \in \{ 1,2,\ldots,n-1 \}$. Note that an $\mathsf{n}$ step must connect $(n-i,n-i)$ to $(n-i,n-i+1)$.
First consider the case where the point $(n-i,n-i)$ on $p$ is reached via an $\mathsf{e}$ step. Note that we may write $p = q \, r$ where $q$ and $r$ are Schr\"oder paths of length $n-i$ and $i$ respectively.
With reference to (\ref{eq:Egge_Mansour_bijection}), the permutation $\pi_2=\varphi(q)$ is obtained by applying some sequence of permutations $\{ \tau_j \}$ to the permutation $(n+1-i,n-i,n-1-i,\ldots, 2,1)$. Similarly, the permutation $\pi_1=\varphi(r)$ is obtained by applying some sequence of permutations $\{ \sigma_j \}$ to the permutation $(i+1,i,i-1,\ldots, 2,1)$. The image of $p$ under the bijection $\varphi$ is then obtained by applying the sequence of permutations $\{ \nu_j \}$, followed by the sequence of permutations $\{ \tau_j \}$, to the permutation $(n+1,n,n-1,\ldots, 2,1)$, where the sequence $\{ \nu_j \}$ is simply the sequence $\{ \sigma_j \}$ with each transposition $s_t$ replaced by $s_{t+n-i}$.
The application of the permutations $\{ \nu_j\}$ (corresponding to the path $r$) results in the final $i+1$ values of $(n+1,n,n-1,\ldots, 2,1)$ being replaced with the permutation $\pi_1$. Then, the application of the permutations $\{ \tau_j \}$ (corresponding to the path $q$) results in the initial $n+1-i$ values of the resulting permutation being replaced with $\pi_2 + i$, except for the unique position $t\in \{1,2,\ldots,n+1-i\}$ with $\pi_2(t) = 1$, which takes the value $\pi_1(1)$.
We may summarize the results of the previous paragraph as follows: for all $t\in\{1,2,\ldots,n+1\}$,
\begin{equation}
\pi(t) = \left\{ \begin{array}{ll}
\pi_1(1) & \textrm{ if } \pi_2(t) = 1 \\
\pi_1(t-n+i) & \textrm{ if } t > n+1-i \\
\pi_2(t) + i & \textrm{ otherwise. }\end{array}\right. \;
\label{eq:case_1_pi}
\end{equation}
Since $\mathrm{rev}(p) = \mathrm{rev}(r) \, \mathrm{rev}(q)$, applying the same argument and invoking the induction hypothesis yields that, for all $s\in\{1,2,\ldots,n+1\}$,
\begin{equation}
\tilde{\pi}(s) = \left\{ \begin{array}{ll}
\pi_2^{-1}(1) & \textrm{ if } \pi_1^{-1}(s) = 1 \\
\pi_2^{-1}(s-i) & \textrm{ if } s > i+1 \\
\pi_1^{-1}(s) + n-i & \textrm{ otherwise. }\end{array}\right. \;
\label{eq:case_1_pi_tilde}
\end{equation}
The reader may check that \eqref{eq:case_1_pi} and \eqref{eq:case_1_pi_tilde} together imply (through the identification $\pi(t) = s$) that $\tilde{\pi} = \pi^{-1}$, as required.
Finally, consider the case where the point $(n-i,n-i)$ on $p$ is reached via a $\mathsf{d}$ step. In this case, we may write $p = q \, \mathsf{d} \, r$, where $q$ and $r$ are Schr\"oder paths of length $n-i-1$ and $i$ respectively. Again denote $\pi_1=\varphi(r)$ and $\pi_2=\varphi(q)$. Arguing in a similar manner to the previous case, we find that in obtaining $\pi=\varphi(p)$ via (\ref{eq:Egge_Mansour_bijection}), the sequence of permutations corresponding to the path $r$ results in the final $i+1$ values of $(n+1,n,n-1,\ldots, 2,1)$ being replaced with the permutation $\pi_1$. Then, the sequence of permutations corresponding to the path $q$ results in the initial $n-i$ values of the resulting permutation being replaced with $\pi_2 + i + 1$. Summarizing,
\begin{equation}
\pi(t) = \left\{ \begin{array}{ll}
\pi_1(t-n+i) & \textrm{ if } t > n-i \\
\pi_2(t) + i + 1 & \textrm{ otherwise. }\end{array}\right. \;
\label{eq:case_2_pi}
\end{equation}
Since $\mathrm{rev}(p) = \mathrm{rev}(r) \, \mathsf{d} \, \mathrm{rev}(q)$, applying the same argument and invoking the induction hypothesis yields that, for all $s\in\{1,2,\ldots,n+1\}$,
\begin{equation}
\tilde{\pi}(s) = \left\{ \begin{array}{ll}
\pi_2^{-1}(s-i-1) & \textrm{ if } s > i + 1 \\
\pi_1^{-1}(s) + n-i & \textrm{ otherwise. }\end{array}\right. \;
\label{eq:case_2_pi_tilde}
\end{equation}
As before, \eqref{eq:case_2_pi} and \eqref{eq:case_2_pi_tilde} together imply (through the identification $\pi(t) = s$) that $\tilde{\pi} = \pi^{-1}$, as required. The result then follows by the principle of induction.
\end{proof}
\begin{corollary}\label{cor:involutions}
Let $p \in \mathcal{S}_n$ and $\pi = \varphi(p)$. Then $\pi$ is an involution if and only if $\mathrm{rev}(p) = p$.
\end{corollary}
It follows that the number of involutions on $\{1,2,3,\ldots,n+1\}$ which avoid $1243$ and $2143$ is equal to the number of Schr\"{o}der paths $p \in \mathcal{S}_n$ which are symmetric with respect to path reversal. This latter fact may also be deduced from the bijection given in \cite[\S 2]{deng_dukes:Schroder_involutions}.
The following example illustrates the inductive step in the proof of Theorem \ref{lemma:inverse_through_EM_bijection}.
\begin{example}\label{ex:inverse_through_EM_bijection}
Consider the Schr\"{o}der path $p = \mathsf{n}\mathsf{d}\mathsf{n}\mathsf{d}\mathsf{n}\mathsf{e}\ee\mathsf{e}\mathsf{n}\nn\mathsf{n}\mathsf{e}\mathsf{n}\mathsf{e}\ee\mathsf{e} \in \mathcal{S}_9$ illustrated in Figure \ref{cap:Schroder_general}. The path $p$ terminates in an $\mathsf{e}$ step, and the largest $t \in \{ 1,2,\ldots,n-1 \} = \{ 1,2,\ldots,8 \}$ for which the point $(t,t)$ lies on $p$ is given by $t = n-i = 5$, i.e., $i=4$. The point $(5,5)$ on $p$ is reached via an $\mathsf{e}$ step. Here we have $p = q \, r$ where $q = \mathsf{n}\mathsf{d}\mathsf{n}\mathsf{d}\mathsf{n}\mathsf{e}\ee\mathsf{e} \in \mathcal{S}_5$ and $r = \mathsf{n}\nn\mathsf{n}\mathsf{e}\mathsf{n}\mathsf{e}\ee\mathsf{e} \in \mathcal{S}_4$. The reader may verify that $\pi_2 = \varphi(q) = (5,3,1,2,4,6)$ and that $\varphi(\mathrm{rev}(q)) = \varphi(\mathsf{n}\nn\mathsf{n}\mathsf{e}\mathsf{d}\mathsf{e}\mathsf{d}\mathsf{e}) = (3,4,2,5,1,6) = \pi_2^{-1}$. Also $\pi_1 = \varphi(r) = (1,3,2,4,5)$ and since $\mathrm{rev}(r) = r$ we have $\varphi(\mathrm{rev}(r)) = \pi_1$; the reader may verify that $\pi_1 = \pi_1^{-1}$.
\begin{figure}
\begin{center}\includegraphics[%
width=0.5\columnwidth,
keepaspectratio]{Schroder_general.eps}
\end{center}
\caption{\label{cap:Schroder_general}The Schr\"oder path $p$ of Example \ref{ex:inverse_through_EM_bijection}.}
\end{figure}
The application of the first set of permutations $\{ \nu_j\}$ (corresponding to the path $r$) results in the final $i+1=5$ values of $(10,9,8,7,6,5,4,3,2,1)$ being replaced with $\pi_1 = (1,3,2,4,5)$ (c.f. row 2 of Table \ref{ttable_general}). Then, the application of the second set of permutations $\{ \tau_j \}$ (corresponding to the path $q$) results in the initial $n+1-i=6$ values of the resulting permutation being replaced with $\pi_2 + i = \pi_2 + 4 = (9,7,5,6,8,10)$, except for position $\pi_2^{-1}(1) = 3$ which takes the value $\pi_1(1) = 1$ (c.f. row 3 of Table \ref{ttable_general}). Thus \eqref{eq:case_1_pi} holds in this case.
\end{example}
\begin{table}[h]
\begin{center}
\caption{\label{ttable_general}
Illustration of the steps involved in obtaining $\pi=\varphi(p)$ for the Schr\"oder path $p \in \mathcal{S}_9$ given in Example \ref{ex:inverse_through_EM_bijection}. The table shows the evolution of the permutation from the starting point of $(10,9,8,7,6,5,4,3,2,1)$.}
{\small
$\begin{array}{|c|c|}
\hline
\begin{array}[t]{c}
\mathrm{Permutation} \\
\hline
\mathrm{Start} \\
\{ \nu_j \} \\
\{ \tau_j \} \\
\end{array} &
\begin{array}[t]{c}
\mathrm{Result} \\
\hline
10 \; 9 \; 8 \; 7 \; 6 \; 5 \; 4 \; 3 \; 2 \; 1 \\
10 \; 9 \; 8 \; 7 \; 6 \; \underline{{\mathbf 1}} \; {\mathbf 3} \; {\mathbf 2} \; {\mathbf 4} \; {\mathbf 5} \\
{\mathbf 9} \; {\mathbf 7} \; \underline{{\mathbf 1}} \; {\mathbf 6} \; {\mathbf 8} \; \mathbf{10} \; 3 \; 2 \; 4 \; 5 \\
\end{array} \\
\hline
\end{array}$
}
\end{center}
\end{table}
\section{Centrosymmetric permutations and involutions avoiding $1243$ and $2143$}
\begin{definition}
The set $\mathcal{P}$ is the set of Schr\"oder paths formed by concatenating a finite sequence of elements from $\mathsf{d} \cup \{ \mathsf{n}^k \mathsf{e}^k \mid k > 0 \}$. The set $\mathcal{D}_n \subseteq \mathcal{S}_n$ is the set of Schr\"oder paths of length $n$ formed by concatenating a finite sequence of elements from $\mathsf{d} \cup \mathsf{n} \mathcal{P} \mathsf{e}$.
\label{def:peaks_etc}
\end{definition}
Less formally, the set $\mathcal{D}_n$ is the set of Schr\"{o}der paths of length $n$ which, whenever they rise above the ``main superdiagonal'' $y = x + 1$, do so by a sequence of $\mathsf{n}$ steps immediately followed by an equal number of $\mathsf{e}$ steps.
\begin{lemma}
Let $p \in \mathcal{D}_n$ be a Schr\"{o}der path with no features, and which does not contain a $\mathsf{d}$ step connecting $(t-1,t-1)$ to $(t,t)$ for any $t \in \{1,2,\ldots,n\}$. Furthermore, let $\varphi(p) = \pi$, $p' = \mathrm{rev} (\psi(p))$, and $\varphi(p') = \pi'$. Then $\pi' = \mathrm{rc}(\pi)$, and $\pi(1) = 1$.
\label{lemma:induction_base_case}
\end{lemma}
\begin{proof}
The Schr\"{o}der path $p$ is of the form $\mathsf{n} \alpha_1 \alpha_2 \cdots \alpha_{r-2} \alpha_{r-1} \mathsf{e}$ where $\alpha_j = \mathsf{n}^{m_{j+1}-m_{j}} \mathsf{e}^{m_{j+1}-m_{j}}$ for some $r$ numbers $1 = m_1 < m_2 < \cdots < m_{r-1} < m_r = n$. In terms of the permutations $\{\sigma_i\}$ involved in obtaining $\varphi(p)$ via (\ref{eq:Egge_Mansour_bijection}), note that $\alpha_j$ ($1 \le j < r$) corresponds to a \emph{reversal} of the numbers in positions $\{ m_j, m_j+1, \ldots, m_{j+1} \}$; denote this reversal by $\delta(m_j, m_{j+1})$. Thus
\begin{equation}
\varphi(p) = \pi = \delta(m_1, m_2) \delta(m_2, m_3) \cdots \delta(m_{r-1}, m_r) \sigma_1 (n+1,n,n-1,\ldots, 2,1) \; ,
\label{eq:base_case_step_1}
\end{equation}
where $\sigma_1 = s_n s_{n-1} \cdots s_2 s_1$.
Since $p$ has no features we have
\[
p' = \mathrm{rev} (\psi(p)) = \mathrm{rev} (p) = \mathsf{n} \alpha_{r-1} \alpha_{r-2} \cdots \alpha_2 \alpha_1 \mathsf{e} \; ,
\]
and thus
\begin{align}
&\varphi(p') = \pi' = \delta(n+1-m_r, n+1-m_{r-1}) \delta(n+1-m_{r-1}, n+1-m_{r-2}) \nonumber \\
&\cdots \delta(n+1-m_{2}, n+1-m_{1}) \sigma_1 (n+1,n,n-1,\ldots, 2,1) \; .
\label{eq:base_case_step_2}
\end{align}
From comparison of (\ref{eq:base_case_step_1}) and (\ref{eq:base_case_step_2}), it may be observed that $\pi' = \mathrm{rc}(\pi)$ due to the mirror symmetry of the applied reversal operations, together with the fact that $\pi(n+1) = \pi'(n+1) = n+1$ and $\pi(1) = \pi'(1) = 1$.
\end{proof}
\begin{figure}
\begin{center}
\includegraphics[%
width=0.48\columnwidth,
keepaspectratio]{Schroder_no_features1.eps}
\includegraphics[%
width=0.48\columnwidth,
keepaspectratio]{Schroder_no_features2.eps}
\vspace{2mm} \\
(a) \hspace{5cm} (b)
\end{center}
\caption{\label{cap:Schroder_no_features1} (a) The Schr\"oder path $p \in \mathcal{D}_9$ and (b) the Schr\"oder path $p' = \mathrm{rev} (\psi(p))$ of Example \ref{ex:base_case}.}
\end{figure}
\begin{example}
Consider the Schr\"{o}der path $p \in \mathcal{D}_9$ illustrated in Figure \ref{cap:Schroder_no_features1} (a). The path $p$ has no features, and does not contain a $\mathsf{d}$ step connecting $(t-1,t-1)$ to $(t,t)$ for any $t \in \{1,2,\ldots,9\}$. The path $p' = \mathrm{rev} (\psi(p)) = \mathrm{rev}(p)$ is shown in Figure \ref{cap:Schroder_no_features1} (b). Table \ref{ttable_3} shows the sequence of reversals applied to the permutation $(10,9,8,7,6,5,4,3,2,1)$ to obtain $\pi = \varphi(p)$ (upper Table \ref{ttable_3}) and $\pi' = \varphi(p')$ (lower Table \ref{ttable_3}) respectively. It may be observed that $\pi' = \mathrm{rc}(\pi)$ holds due to the mirror symmetry of the applied reversal operations as well as the fact that both permutations $\pi$ and $\pi'$ interchange the values $1$ and $n+1=10$.
\label{ex:base_case}
\end{example}
\begin{lemma}\label{lemma:concatenation}
Let $p \in \mathcal{D}_n$ be a Schr\"{o}der path which does not contain a $\mathsf{d}$ step connecting $(t-1,t-1)$ to $(t,t)$ for any $t \in \{1,2,\ldots,n\}$, and which contains no level features. If $\pi = \varphi(p)$, then $\pi(1) = 1$.
\end{lemma}
\begin{proof}
Write $p = q_1 q_2 \cdots q_l$ where each Schr\"{o}der path $q_i$ is of the form described in Lemma \ref{lemma:induction_base_case}; therefore, each permutation $\pi_i = \varphi(q_i)$ satisfies $\pi_i(1) = 1$. Letting $m_i>0$ denote the length of each path $q_i$, we have $\sum_{j=1}^{l} m_j = n$. Then, with reference to obtaining $\varphi(p)$ from $(n+1,n,n-1,\ldots,2,1)$ via \eqref{eq:Egge_Mansour_bijection}, permutations are applied corresponding to each path $q_i$ for $i=l,l-1,\ldots,2,1$ respectively, in each case moving the entry $1$ from position $\sum_{j=1}^i m_j + 1$ to position $\sum_{j=1}^{i-1} m_j + 1$. Therefore, after all permutations are applied we must have $\pi(1) = 1$.
\end{proof}
\begin{table}
\begin{center}
\caption{\label{ttable_3}
Illustration of the subsequence reversal operations involved in obtaining $\pi=\varphi(p)$ (upper) and $\pi' = \varphi(p')$ (lower) for the Schr\"oder paths $p,p' \in \mathcal{D}_9$ given in Example \ref{ex:base_case}. The table shows the evolution of the permutation from $(10,9,8,7,6,5,4,3,2,1)$ with the initial cyclic shift $\sigma_1$ followed by the reversals $\delta(\cdot,\cdot)$ in each case.}
{\small
$\begin{array}{|c|c|}
\hline
\begin{array}[t]{c}
\mathrm{Permutation} \\
\hline
\mathrm{Start} \\
\sigma_1 = s_9 s_8 s_7 s_6 s_5 s_4 s_3 s_2 s_1 \\
\delta(8,9) \\
\delta(3,8) \\
\delta(1,3) \\
\end{array} &
\begin{array}[t]{c}
\mathrm{Result} \\
\hline
10 \; 9 \; 8 \; 7 \; 6 \; 5 \; 4 \; 3 \; 2 \; 1 \\
{\mathbf 9} \; {\mathbf 8} \; {\mathbf 7} \; {\mathbf 6} \; {\mathbf 5} \; {\mathbf 4} \; {\mathbf 3} \; {\mathbf 2} \; {\mathbf 1} \; \mathbf{10} \\
9 \; 8 \; 7 \; 6 \; 5 \; 4 \; 3 \; {\mathbf 1} \; {\mathbf 2} \; 10 \\
9 \; 8 \; {\mathbf 1} \; {\mathbf 3} \; {\mathbf 4} \; {\mathbf 5} \; {\mathbf 6} \; {\mathbf 7} \; 2 \; 10 \\
{\mathbf 1} \; {\mathbf 8} \; {\mathbf 9} \; 3 \; 4 \; 5 \; 6 \; 7 \; 2 \; 10 \\
\end{array} \\
\hline
\end{array}$
}
{\small
$\begin{array}{|c|c|}
\hline
\begin{array}[t]{c}
\mathrm{Permutation} \\
\hline
\mathrm{Start} \\
\sigma_1 = s_9 s_8 s_7 s_6 s_5 s_4 s_3 s_2 s_1 \\
\delta(7,9) \\
\delta(2,7) \\
\delta(1,2) \\
\end{array} &
\begin{array}[t]{c}
\mathrm{Result} \\
\hline
10 \; 9 \; 8 \; 7 \; 6 \; 5 \; 4 \; 3 \; 2 \; 1 \\
{\mathbf 9} \; {\mathbf 8} \; {\mathbf 7} \; {\mathbf 6} \; {\mathbf 5} \; {\mathbf 4} \; {\mathbf 3} \; {\mathbf 2} \; {\mathbf 1} \; \mathbf{10} \\
9 \; 8 \; 7 \; 6 \; 5 \; 4 \; {\mathbf 1} \; {\mathbf 2} \; {\mathbf 3} \; 10 \\
9 \; {\mathbf 1} \; {\mathbf 4} \; {\mathbf 5} \; {\mathbf 6} \; {\mathbf 7} \; {\mathbf 8} \; 2 \; 3 \; 10 \\
{\mathbf 1} \; {\mathbf 9} \; 4 \; 5 \; 6 \; 7 \; 8 \; 2 \; 3 \; 10 \\
\end{array} \\
\hline
\end{array}$
}
\end{center}
\end{table}
\begin{lemma}\label{thm:earliest_level_feature}
Let $p \in \mathcal{D}_n$ be a Schr\"{o}der path which does not contain a $\mathsf{d}$ step connecting $(t-1,t-1)$ to $(t,t)$ for any $t \in \{1,2,\ldots,n\}$. If the earliest level feature of $p$ is at $t=k$, then $\pi(1) = n-k+1$.
\end{lemma}
\begin{proof}
Let $q \in \mathcal{D}_k$ denote the Schr\"{o}der path of length $k$ obtained by following $p$ up the point $(k-1,k)$ and terminating via an $\mathsf{e}$ step at the point $(k,k)$. Note that $q$ is of the form described in Lemma \ref{lemma:concatenation}, and thus $\pi_1 = \varphi(q)$ satisfies $\pi_1(1) = 1$. Next, with reference to obtaining $\varphi(p)$ from $(n+1,n,n-1,\ldots,2,1)$ via \eqref{eq:Egge_Mansour_bijection}, initially the entry $n-k+1$ lies in position $k+1$. First a sequence of permutations is applied which uses only transpositions $s_t$ with $t > k+1$; after this, the entry $n-k+1$ remains in position $k+1$. Then, a sequence of transpositions is applied which contains $s_k$; this moves the entry $n-k+1$ into position $k$. Immediately after this, a sequence of permutations is applied containing only transpositions $s_t$ with $t>k$; during this process, entry $n-k+1$ remains in position $k$. For the remaining sequence of permutations, observe that since entry $n-k+1$ starts in position $k$, its final position is the same as the final position of entry $1$ when obtaining $\pi_1 = \varphi(q)$ via \eqref{eq:Egge_Mansour_bijection}, i.e., $\pi(n-k+1) = 1$.
\end{proof}
\begin{example}
Consider the Schr\"{o}der path $p = \mathsf{n}\nn\mathsf{e}\mathsf{n}\mathsf{e}\ee\mathsf{n}\nn\mathsf{e}\mathsf{d}\mathsf{n}\mathsf{e}\ee\mathsf{n}\nn\mathsf{e}\ee \in \mathcal{D}_9$ illustrated in Figure \ref{cap:Schroder_concat}. The earliest level feature of $p$ is at $t=k=5$. Here we have $q = \mathsf{n}\nn\mathsf{e}\mathsf{n}\mathsf{e}\ee\mathsf{n}\nn\mathsf{e}\ee$ and $\pi_1 = \varphi(q) = (1,5,4,6,2,3)$. The Schr\"{o}der path $q$ is of the form described in Lemma \ref{lemma:concatenation}, and accordingly $\pi_1(1) = 1$. Table \ref{ttable_concat} summarizes the steps involved in obtaining $\pi = \varphi(p)$ from $(10,9,8,7,6,5,4,3,2,1)$ via \eqref{eq:Egge_Mansour_bijection}. The permutations $\{ \sigma_i \}$ prior to that containing transposition $s_k = s_5$ involve only transpositions $s_8$ and $s_9$, and thus do not affect the position of entry $n-k+1 = 5$. Next, the permutation $s_7 s_6 s_5 s_4$ is applied which moves entry $n-k+1 = 5$ to position $k=5$. The next permutation, $s_6$, does not alter the position of the entry $5$. The final sequence of permutations is simply that used to obtain $\pi_1 = \varphi(q)$ via \eqref{eq:Egge_Mansour_bijection}, but with $\sigma_1 = s_5 s_4$ omitted; therefore we finally obtain $\pi(1) = 5$.
\begin{figure}
\begin{center}\includegraphics[%
width=0.5\columnwidth,
keepaspectratio]{Schroder_concat.eps}
\end{center}
\caption{\label{cap:Schroder_concat}The Schr\"oder path $p$ of Example \ref{ex:earliest_level_feature}.}
\end{figure}
\label{ex:earliest_level_feature}
\end{example}
\begin{lemma}\label{thm:latest_level_feature}
Let $p \in \mathcal{D}_n$ be a Schr\"{o}der path which does not contain a $\mathsf{d}$ step connecting $(t-1,t-1)$ to $(t,t)$ for any $t \in \{1,2,\ldots,n\}$. If the latest level feature of $p$ is at $t=k$, then $\pi(k+1) = 1$.
\end{lemma}
\begin{proof}
Note that $\mathrm{rev}(p) \in \mathcal{D}_n$ does not contain a $\mathsf{d}$ step connecting $(t-1,t-1)$ to $(t,t)$ for any $t \in \{1,2,\ldots,n\}$, and the earliest level feature of $\mathrm{rev}(p)$ is at $t=n-k$. Applying Lemma \ref{thm:earliest_level_feature}, and noting that $\varphi(\mathrm{rev}(p)) = \pi^{-1}$ by Theorem \ref{lemma:inverse_through_EM_bijection}, we obtain $\pi^{-1}(1) = k+1$, i.e., $\pi(k+1) = 1$.
\end{proof}
\begin{table}
\begin{center}
\caption{\label{ttable_concat}
Illustration of the steps involved in obtaining $\pi=\varphi(p)$ for the Schr\"oder path $p \in \mathcal{S}_9$ given in Example \ref{ex:earliest_level_feature}. The table shows the evolution of the permutation from the starting point of $(10,9,8,7,6,5,4,3,2,1)$.}
{\small
$\begin{array}{|c|c|}
\hline
\begin{array}[t]{c}
\mathrm{Permutation} \\
\hline
\mathrm{Start} \\
\mathrm{permutations \; prior \; to \; that \; including} \; s_5 \\
\mathrm{after \; permutation \; including} \; s_5 \\
\mathrm{permutations \; containing} \; s_t, \; t>5 \\
\mathrm{remainder \; of \; permutations} \\
\end{array} &
\begin{array}[t]{c}
\mathrm{Result} \\
\hline
10 \; 9 \; 8 \; 7 \; 6 \; \underline{5} \; 4 \; 3 \; 2 \; 1 \\
10 \; 9 \; 8 \; 7 \; 6 \; \underline{5} \; 4 \; {\mathbf 1} \; {\mathbf 2} \; {\mathbf 3} \\
10 \; 9 \; 8 \; {\mathbf 6} \; \underline{{\mathbf 5}} \; {\mathbf 4} \; {\mathbf 1} \; {\mathbf 7} \; 2 \; 3 \\
10 \; 9 \; 8 \; 6 \; \underline{5} \; {\mathbf 1} \; {\mathbf 4} \; 7 \; 2 \; 3 \\
\underline{{\mathbf 5}} \; {\mathbf 9} \; {\mathbf 8} \; \mathbf{10} \; {\mathbf 6} \; 1 \; 4 \; 7 \; 2 \; 3 \\
\end{array} \\
\hline
\end{array}$
}
\end{center}
\end{table}
The following example illustrates Lemmas \ref{thm:earliest_level_feature} and \ref{thm:latest_level_feature} in the context of the Schr\"oder path $p \in \mathcal{S}_9$ of Example \ref{ex:running_example}.
\begin{example}
Consider the path $p \in \mathcal{S}_9$ defined in Example \ref{ex:running_example}. Note that $p \in \mathcal{D}_9$, and that $p$ does not contain a $\mathsf{d}$ step connecting $(t-1,t-1)$ to $(t,t)$ for any $t \in \{1,2,\ldots,n\}$. The earliest and latest level features of $p$ are at $t=5$ and $t=8$ respectively. We also have $\pi(1)=5$ and $\pi(9)=1$ in accordance with Lemmas \ref{thm:earliest_level_feature} and \ref{thm:latest_level_feature} respectively.
\end{example}
\begin{theorem}\label{thm:main_transformation}
For $n \ge 0$, let $\pi \in \mathfrak{S}_{n+1}(1243,2143)$. Then $\pi' = \mathrm{rc}(\pi)$ lies in $\mathfrak{S}_{n+1}(1243,2143)$ if and only if $p = \varphi^{-1}(\pi) \in \mathcal{D}_n$. Furthermore, if $p \in \mathcal{D}_n$, then $\varphi^{-1}(\pi') = p'$ where $p' = \mathrm{rev} (\psi(p))$.
\end{theorem}
Before giving a proof, we provide an example in order to illustrate this Theorem.
\begin{example}\label{ex:example_contd}
Consider the path $p \in \mathcal{S}_9$ defined in Example \ref{ex:running_example}. Note that $p \in \mathcal{D}_9$. The path $p' = \mathrm{rev}(\psi(p)) = \mathsf{n}\mathsf{e}\mathsf{n}\nn\mathsf{e}\mathsf{d}\mathsf{e}\mathsf{n}\nn\mathsf{n}\nn\mathsf{e}\ee\mathsf{e}\mathsf{d}\mathsf{e}$ is illustrated in Figure \ref{cap:Schroder2}.
Here we have
$\sigma_1=s_9s_8s_7s_6s_5$, $\sigma_2=s_7s_6s_5$, $\sigma_3= s_6s_5$, $\sigma_4=s_5$, $\sigma_5=s_4s_3s_2$, $\sigma_6=s_2$ and $\sigma_7=s_1$. So
\begin{eqnarray*}
\varphi(p') &=& \sigma_7 \sigma_6 \sigma_5 \sigma_4 \sigma_3 \sigma_2 \sigma_1 (10,9,8,7,6,5,4,3,2,1) \\
&=& s_1 \; s_2 \; s_4s_3s_2 \; s_5 \; s_6s_5 \; s_7s_6s_5 \; s_9s_8s_7s_6s_5 (10,9,8,7,6,5,4,3,2,1) \\
&=& (7,10,8,2,9,3,4,5,1,6) \in \mathfrak{S}_{10}(1243,2143)
\end{eqnarray*}
The evolution of $(10,9,8,7,6,5,4,3,2,1)$ towards $\varphi(p')$ under the application of the permutations $\sigma_i$ for $i=1,2,\ldots,k$ is shown in Table \ref{ttable_2}. The reader may check that $\varphi(p') = \mathrm{rc}(\pi)$, as expected.
\end{example}
\begin{figure}
\begin{center}\includegraphics[%
width=0.5\columnwidth,
keepaspectratio]{Schroder2.eps}
\end{center}
\caption{\label{cap:Schroder2}The Schr\"oder path $p'$ of Example \ref{ex:example_contd}.}
\end{figure}
\begin{lemma}
If $p$ is any Schr\"oder path of length $n \ge 0$ and $\pi = \varphi(p)$, then $\mathrm{rc}(\pi) \in \mathfrak{S}_{n+1}(1243,2143)$ implies that $p \in \mathcal{D}_n$.
\label{lemma:one_direction}
\end{lemma}
\begin{table}
\begin{center}
\caption{\label{ttable_2}
Illustration of the bijection $\varphi(p')$ for the Schr\"oder path $p' \in \mathcal{S}_9$ given in Example \ref{ex:example_contd}. The table shows the evolution of the permutation from $(10,9,8,7,6,5,4,3,2,1)$ as each permutation $\sigma_i$ is applied for $i=1,2,\ldots,k$.}
{\small
$\begin{array}{|c|c|}
\hline
\begin{array}[t]{c}
\mathrm{Permutation} \\
\hline
\mathrm{Start} \\
\sigma_1 = s_9 s_8 s_7 s_6 s_5 \\
\sigma_2 = s_7 s_6 s_5 \\
\sigma_3 = s_6 s_5 \\
\sigma_4 = s_5 \\
\sigma_5 = s_4 s_3 s_2 \\
\sigma_6 = s_2 \\
\sigma_7= s_1 \\
\end{array} &
\begin{array}[t]{c}
\mathrm{Result} \\
\hline
10 \; 9 \; 8 \; 7 \; 6 \; 5 \; 4 \; 3 \; 2 \; 1 \\
10 \; 9 \; 8 \; 7 \; {\mathbf 5} \; {\mathbf 4} \; {\mathbf 3} \; {\mathbf 2} \; {\mathbf 1} \; {\mathbf 6} \\
10 \; 9 \; 8 \; 7 \; {\mathbf 4} \; {\mathbf 3} \; {\mathbf 2} \; {\mathbf 5} \; 1 \; 6 \\
10 \; 9 \; 8 \; 7 \; {\mathbf 3} \; {\mathbf 2} \; {\mathbf 4} \; 5 \; 1 \; 6 \\
10 \; 9 \; 8 \; 7 \; {\mathbf 2} \; {\mathbf 3} \; 4 \; 5 \; 1 \; 6 \\
10 \; {\mathbf 8} \; {\mathbf 7} \; {\mathbf 2} \; {\mathbf 9} \; 3 \; 4 \; 5 \; 1 \; 6 \\
10 \; {\mathbf 7} \; {\mathbf 8} \; 2 \; 9 \; 3 \; 4 \; 5 \; 1 \; 6 \\
{\mathbf 7} \; \mathbf{10} \; 8 \; 2 \; 9 \; 3 \; 4 \; 5 \; 1 \; 6 \\
\end{array} \\
\hline
\end{array}$
}
\end{center}
\end{table}
\begin{proof}
Note that since $\pi \in \mathfrak{S}_{n+1}(1243,2143)$ and $\mathrm{rc}(\pi) \in \mathfrak{S}_{n+1}(1243,2143)$, this implies that $\pi$ avoids the pattern $\mathrm{rc}(1243) = 2134$ also. Suppose that points $(\alpha,\alpha+1)$ and $(\beta,\beta+1)$ lie on $p$ for some $0 \le \alpha < \beta \le n-1$, and that no part of $p$ lies on or below the line segment joining these two points. [In particular, this means that no point $(t,t+1)$ lies on $p$ for $\alpha < t < \beta$; note that this implies an occurrence of $\mathsf{n}$ joining $(\alpha,\alpha+1)$ to $(\alpha,\alpha+2)$, and an occurrence of $\mathsf{e}$ joining $(\beta-1,\beta+1)$ to $(\beta,\beta+1)$.] Then $\varphi(p)$ is obtained by applying a sequence of permutations $\sigma_i$ to $(n+1,n,n-1,\ldots,2,1)$ as per (\ref{eq:Egge_Mansour_bijection}). A subsequence of these permutations is as follows: first a permutation $P_1$ is applied which contains $s_{\beta+1} s_{\beta} \cdots s_{\alpha+1}$ as a subsequence, later the permutation $P_2 = s_{\beta} s_{\beta-1} \cdots s_{\alpha+1}$ is applied, and immediately after application of $P_2$, a permutation $P_3$ is applied which consists of a sequence of transpositions from $\{ s_{\alpha+1}, s_{\alpha+2}, \ldots, s_{\beta-1} \}$. The permutation $P_1$ moves a number $q > n-\alpha$ to the right of position $\beta+1$. Permutation $P_2$ moves the number $n-\alpha$ to position $\beta+1$. At this point, the numbers in positions $\alpha+1, \alpha+2, \ldots, \beta$ are in \emph{decreasing} order. If after application of $P_3$ any two of these numbers (say $f > g$) end up in decreasing order, the subsequence $(f, g, n-\alpha, q)$ would be an occurrence of the pattern $2134$, which is a contradiction. Therefore the numbers in positions $\alpha+1, \alpha+2, \ldots, \beta$ must end up in \emph{increasing order}, i.e., the points $(\alpha,\alpha+1)$ and $(\beta,\beta+1)$ on $p$ must be joined\footnote{Note that reversal of ordering of the numbers in positions $\{m, m+1, \ldots, n\}$ is effected through application of the sequence of permutations $\sigma_{m} \sigma_{m+1} \cdots \sigma_{n-1}$ where $\sigma_i = s_i s_{i-1} \cdots s_m$ for each $i=m,m+1,\ldots,n-1$; this corresponds to the path $\mathsf{n}^{n-m} \mathsf{e}^{n-m}$.} by an occurrence of $\mathsf{n}^{\beta-\alpha} \mathsf{e}^{\beta-\alpha}$. We conclude that for every pair of points $(\alpha,\alpha+1)$ and $(\beta,\beta+1)$ on $p$ such that no part of $p$ lies on or below the line segment joining these two points, the two points must be joined by an occurrence of $\mathsf{n}^{\beta-\alpha} \mathsf{e}^{\beta-\alpha}$. This is equivalent to the condition $p \in \mathcal{D}_n$.
\end{proof}
The following example illustrates Lemma \ref{lemma:one_direction} in the context of the Schr\"oder path $p \in \mathcal{S}_9$ of Example \ref{ex:running_example}.
\begin{example}
Consider the Schr\"oder path $p \in \mathcal{S}_9$ of Example \ref{ex:running_example}. In particular, note that points $(1,2)$ and $(4,5)$ lie on $p$, while points $(2,3)$ and $(3,4)$ do not; therefore we may apply the reasoning in the proof of Lemma \ref{lemma:one_direction} with $\alpha=1$ and $\beta=4$. After application of permutations $\sigma_1$ and $\sigma_2$, the permutation $(10,9,8,7,6,5,4,3,2,1)$ has changed to $(10,9,8,7,6,5,2,3,1,4)$ (c.f. row 3 of Table \ref{ttable_1}). Here $P_1=\sigma_3=s_6 s_5 s_4 s_3 s_2$ contains $s_{\beta+1} s_{\beta} \cdots s_{\alpha+1} = s_5 s_4 s_3 s_2$ as a subsequence. This permutation moves the number $q = 9 > 8 = n-\alpha$ (via a sequence of adjacent transpositions) to position $7$ which lies to the right of position $\beta+1=5$. Then, permutation $P_2 = \sigma_4 = s_{\beta} s_{\beta-1} \cdots s_{\alpha+1} = s_4 s_3 s_2$ is applied, which moves the number $n-\alpha=8$ to position $\beta+1 = 5$. To avoid the subsequence $x y 8 9$ being an occurrence of the pattern $2134$ for some $x$ and $y$, the subsequence $7 6 5$ must next be rearranged as $5 6 7$; this requires $\sigma_5= s_3 s_2$ and $\sigma_6 = s_2$, i.e., the points $(1,2)$ and $(4,5)$ on $p$ must be joined by an occurrence of $\mathsf{n}^{3} \mathsf{e}^{3}$.
\label{ex:must_have_spike}
\end{example}
\begin{proof}[\bf{Proof of Theorem \ref{thm:main_transformation}}]
First suppose that the Schr\"oder path $p$ contains an occurrence of $\mathsf{d}$ joining $(t-1,t-1)$ to $(t,t)$ for $t \in \{ 1,2,\ldots,n \}$, i.e., the path $p$ may be viewed as the concatenation of a Schr\"oder path $q$ (of length $t-1$), a $\mathsf{d}$ step, and a Schr\"oder path $r$ (of length $n-t$). From Lemma \ref{lemma:no_mixing}, this occurs if and only if the largest $t$ numbers in $\{ 1,2,\ldots, n+1\}$ occupy the first $t$ positions of $\pi = \varphi(p)$. From the definition of the reverse-complement, this latter condition is obtained if and only if the largest $n+1-t$ numbers in $\{ 1,2,\ldots, n+1\}$ occupy the first $n+1-t$ positions of $\pi' = \mathrm{rc}(\pi)$. Again invoking Lemma \ref{lemma:no_mixing}, this occurs if and only if the Schr\"oder path $p' = \varphi^{-1}(\pi')$ contains an occurrence of $\mathsf{d}$ joining $(n-t,n-t)$ to $(n-t+1,n-t+1)$. The problem is then seen to reduce to proving the proposition for $q$ and $r$ separately (note that $p\in \mathcal{D}_n$ if and only if both $q \in \mathcal{D}_{t-1}$ and $r \in \mathcal{D}_{n-t}$). For this reason, in the following we need consider only those Schr\"oder paths $p$ which do not contain any occurrence of $\mathsf{d}$ joining $(t-1,t-1)$ to $(t,t)$ for $t \in \{ 1,2,\ldots,n \}$. By Lemma \ref{thm:latest_level_feature}, if the latest level feature of such a path $p$ is at $t=k$, then $\pi(k+1) = 1$.
From Lemma \ref{lemma:one_direction}, if $p$ is any Schr\"oder path of length $n \ge 0$ and $\pi = \varphi(p)$, then $\mathrm{rc}(\pi) \in \mathfrak{S}_{n+1}(1243,2143)$ implies that $p \in \mathcal{D}_n$. To prove the other direction, let $p \in \mathcal{D}_n$, and let $\varphi(p) = \pi$. Let $p' = \mathrm{rev} (\psi(p))$ and $\varphi(p') = \pi'$. We wish to show that $\pi' = \mathrm{rc}(\pi)$. We proceed by induction on the number $\gamma$ of features of $p$. The base case $\gamma = 0$ of the induction has already been proved in Lemma \ref{lemma:induction_base_case}.
Next assume the result holds for all Schr\"oder paths with $\gamma \ge 0$ features, and consider a Schr\"oder path $p$ with $\gamma+1$ features. The latest feature of $p$ occurs at $t=n-i$ for some integer $i\in\{ 1,2,\ldots, n-1 \}$. We assume also that this is a notch feature (this assumption will be justified later). It follows that we may write $p$ as the concatenation of two Schr\"oder paths: $q$ (of length $n-i$, and containing $\gamma$ features) and $r$ (of length $i$, and containing no features).
The permutation $\pi_2=\varphi(q)$ is obtained by applying some sequence of permutations $\{ \tau_j \}$ to the permutation $(n+1-i,n-i,n-1-i,\ldots, 2,1)$. Similarly, the permutation $\pi_1=\varphi(r)$ is obtained by applying some sequence of permutations $\{ \sigma_j \}$ to the permutation $(i+1,i,i-1,\ldots, 2,1)$. The image of $p$ under the bijection $\varphi$ is then obtained by applying the sequence of permutations $\{ \nu_j \}$, followed by the sequence of permutations $\{ \tau_j \}$, to the permutation $(n+1,n,n-1,\ldots, 2,1)$, where the sequence $\{ \nu_j \}$ is simply the sequence $\{ \sigma_j \}$ with each transposition $s_t$ replaced by $s_{t+n-i}$.
The application of the permutations $\{ \nu_j\}$ (corresponding to the path $r$) results in the final $i+1$ values of $(n+1,n,n-1,\ldots, 2,1)$ being replaced with the permutation $\pi_1$. Suppose that the latest level feature of $q$ is at $t=k$; by Lemma \ref{thm:latest_level_feature}, this implies that $\pi_2(k+1) = 1$. Then, since $\pi_1(1)=1$ (by Lemma \ref{lemma:induction_base_case}), the application of the permutations $\{ \tau_j \}$ (corresponding to the path $q$) results in the initial $n+1-i$ values of the resulting permutation being replaced with $\pi_2+i$, except for $\pi(k+1)$ which takes the value $1$.
Formally, we may summarize these results as
\begin{equation}
\varphi(p) = (f(1), f(2), \ldots, f(n+1-i), \pi_1(2), \pi_1(3), \ldots, \pi_1(i+1))
\label{eq:main_result_proof_1}
\end{equation}
where, for each $t \in \{ 1,2, \ldots, n+1-i\}$,
\begin{equation}
f(t) = \left\{ \begin{array}{cc}
1 & \textrm{ if } t = k+1 \\
\pi_2(t) + i & \textrm{ otherwise. }\end{array}\right. \;
\label{eq:ft_definition}
\end{equation}
Next consider $p' = \mathrm{rev} (\psi(p))$; this Schr\"oder path may be considered as the concatenation of Schr\"oder paths $r' = \mathrm{rev} (\psi(r))$ and $q' = \mathrm{rev} (\psi(q))$, followed by the replacement of the resulting notch feature at $t=i$ by a level feature. Let $\pi_2' = \varphi(q')$ and $\pi_1' = \varphi(r')$.
The permutation $\pi_2'=\varphi(q')$ is obtained by applying some sequence of permutations $\{ \tau_j' \}$ to the permutation $(n+1-i,n-i,n-1-i,\ldots, 2,1)$, and the permutation $\pi_1'=\varphi(r')$ is obtained by applying some sequence of permutations $\{ \sigma_j' \}$ to the permutation $(i+1,i,i-1,\ldots, 2,1)$. The permutation $\pi=\varphi(p')$ is then obtained by applying a sequence of permutations $\{ \mu_j' \}$ to the permutation $(n+1,n,n-1,\ldots, 2,1)$, followed by the sequence of permutations $\{ \sigma_j' \}$ with $\sigma'_1 = s_i s_{i-1} \cdots s_2 s_1$ omitted. The application of the permutations in $\{ \mu_j' \}$ results in the final $n+1-i$ values of the resulting permutation being replaced with $\pi_2'$, except for position $n+1-k$ which holds the value $n+1$. The reason for the latter condition is that the level feature at $t=i$ causes the number $n+1$ to move into position $n+1-k$. The reason it moves into this particular position is that the latest level feature of $q$ lying at $t=k$ implies that the earliest notch feature of
$p'$ is at $t = n-k$.
Note that at this point, the first $i$ values are in decreasing order starting with $n$. Therefore, application of the permutations $\{ \sigma_j'\}$ with $\sigma'_1 = s_i s_{i-1} \cdots s_2 s_1$ omitted (corresponding to the path $r'$) results in the first $i$ values of the resulting permutation being replaced with the first $i$ numbers in $\pi_1'+n-i$.
Formally, we may summarize these results as
\begin{align}
\varphi(p') &= (\pi'(1) + n-i,\pi'(2) + n-i, \ldots, \pi'(i) + n-i, \nonumber \\
& \quad \quad \quad \quad \quad \quad \quad \quad g(1), g(2), \ldots, g(n+1-i)) \; ,
\label{eq:main_result_proof_2}
\end{align}
where, for each $t \in \{ 1,2, \ldots, n+1-i\}$,
\begin{equation}
g(t) = \left\{ \begin{array}{cc}
n+1 & \textrm{ if } t = n+1-i-k \\
\pi_2'(t) & \textrm{ otherwise. }\end{array}\right. \;
\label{eq:gt_definition}
\end{equation}
Comparing (\ref{eq:main_result_proof_1}) and (\ref{eq:ft_definition}) with (\ref{eq:main_result_proof_2}) and (\ref{eq:gt_definition}) while applying the induction hypothesis, which guarantees $\pi_1' = \mathrm{rc}(\pi_1)$ and $\pi_2' = \mathrm{rc}(\pi_2)$, yields $\varphi(p') = \mathrm{rc}(\pi)$, as required.
Finally, note that the preceding proof assumed that the latest feature of $p$ is a notch feature. If this is not the case, we may replace $p$ by $\psi(p)$ (whose latest feature \emph{is} a notch feature) and repeat the argument from the beginning, thus establishing (due to Theorem \ref{lemma:inverse_through_EM_bijection}) that $(\pi')^{-1} = \mathrm{rc}(\pi^{-1})$. By Lemma \ref{lemma:rc_inv}, this implies that $\pi' = \mathrm{rc}(\pi)$, as required. The result then follows by the principle of induction.
\end{proof}
The following example illustrates the inductive proof of Theorem \ref{thm:main_transformation} for the case of $\gamma=2$.
\begin{example}
Consider the Schr\"{o}der path $p = \mathsf{n}\nn\mathsf{e}\mathsf{d}\mathsf{e}\mathsf{n}\nn\mathsf{n}\mathsf{e}\ee\mathsf{e}\mathsf{n}\nn\mathsf{e}\mathsf{n}\mathsf{e}\ee \in \mathcal{D}_9$ with $\gamma+1=3$ features shown in Figure \ref{cap:main_result} (a). This path is a concatenation of a path $q \in \mathcal{D}_6$ and a featureless path $r \in \mathcal{D}_3$, i.e., $i=3$. The reader may verify that $\pi_1 = \varphi(r) = (1,3,2,4)$ and $\pi_2 = \varphi(q) = (5,6,1,7,2,3,4)$. Here the latest level feature of $q$ is at $k=2$, so Lemma \ref{thm:latest_level_feature} guarantees $\pi_2(k+1) = \pi_2(3) = 1$. The application of the first set of permutations $\{ \nu_j \}$ (corresponding to the path $r$) results in the final $i+1=4$ values of $(10,9,8,7,6,5,4,3,2,1)$ being replaced with $\pi_1$ (c.f. row 2 of upper Table \ref{ttable_4}). Then, since $\pi_1(1)=1$, the application of the second set of permutations $\{ \tau_j \}$ (corresponding to the path $q$) results in the initial $n+1-i=7$ values of the resulting permutation being replaced with $\pi_2+i = \pi_2+3 = (8,9,4,10,5,6,7)$, except for $\pi(k+1) = \pi(3)$ which takes the value $1$ (c.f. row 3 of upper Table \ref{ttable_4}).
\begin{figure}
\begin{center}
\includegraphics[%
width=0.48\columnwidth,
keepaspectratio]{Schroder_p.eps}
\includegraphics[%
width=0.48\columnwidth,
keepaspectratio]{Schroder_pprime.eps}
\vspace{2mm} \\
(a) \hspace{5cm} (b)
\end{center}
\caption{\label{cap:main_result} (a) The Schr\"oder path $p$ and (b) the Schr\"oder path $p' = \mathrm{rev} (\psi(p))$ of Example \ref{ex:main_result}. The corresponding evolution of permutations via the bijection $\varphi$ is shown in Table \ref{ttable_4}.}
\end{figure}
The path $p' = \mathrm{rev}(\psi(p))$ is shown in Figure \ref{cap:main_result} (b); this may be seen as the concatenation of $r'=\mathrm{rev}(\psi(r))$ and $q'=\mathrm{rev}(\psi(q))$, followed by the replacement of the resulting notch feature at $i=3$ by a level feature. The reader may verify that $\pi_1' = \psi(r') = (1,3,2,4)$ and $\pi_2' = \psi(q') = (4,5,6,1,7,2,3)$. The application of the first set of permutations $\{ \mu_j' \}$ results in (i) the number $n+1 = 10$ moving into position $n+1-k = 8$ (c.f. row 2 of lower Table \ref{ttable_4}) -- the reason it moves into this position is that the latest level feature of $q$ lying at $k=2$ implies that the earliest notch feature of $p'$ lies at $n-k=7$; and (ii) the final $n+1-i=7$ values of the resulting permutation being replaced with $\pi_2'$, except for position $n+1-k = 8$ which holds the value $n+1=10$. At this point, the first $i=3$ values are in decreasing order starting with $n=9$ (again c.f. row 2 of lower Table \ref{ttable_4}). Application of the permutations $\{ \sigma_j'\}$ with $\sigma_1 = s_3 s_2 s_1$ omitted (corresponding to the path $r'$) results in the first $i=3$ values of the resulting permutation being replaced with the first $i=3$ numbers in $\pi_1'+n-i = \pi_1'+6 = (7,9,8,10)$ (c.f. row 3 of lower Table \ref{ttable_4}). Application of the induction hypothesis, which guarantees $\pi_1' = \mathrm{rc}(\pi_1)$ and $\pi_2' = \mathrm{rc}(\pi_2)$, then yields the result.
\label{ex:main_result}
\end{example}
\begin{theorem}\label{Schroder_perms}
\[
\left| \mathcal{C}_{2n}(1243,2143) \right| = \left| \mathcal{C}_{2n+1}(1243,2143) \right| = q_n
\]
for all $n \ge 1$, where the sequence $q_n$ is defined by $q_1=2$, $q_2=7$ and for every $n \ge 3$, $q_n = 4q_{n-1} - q_{n-2}$.
\end{theorem}
\begin{table}
\begin{center}
\caption{\label{ttable_4}
Illustration of the steps involved in obtaining $\pi=\varphi(p)$ (upper) and $\pi' = \varphi(p')$ (lower) for the Schr\"oder paths $p,p' \in \mathcal{S}_9$ given in Example \ref{ex:main_result}. The table shows the evolution of the permutation from $(10,9,8,7,6,5,4,3,2,1)$ in each case.}
{\small
$\begin{array}{|c|c|}
\hline
\begin{array}[t]{c}
\mathrm{Permutation} \\
\hline
\mathrm{Start} \\
\{ \nu_j \} \\
\{ \tau_j \} \\
\end{array} &
\begin{array}[t]{c}
\mathrm{Result} \\
\hline
10 \; 9 \; 8 \; 7 \; 6 \; 5 \; 4 \; 3 \; 2 \; 1 \\
10 \; 9 \; 8 \; 7 \; 6 \; 5 \; \underline{{\mathbf 1}} \; {\mathbf 3} \; {\mathbf 2} \; {\mathbf 4} \\
{\mathbf 8} \; {\mathbf 9} \; \underline{{\mathbf 1}} \; \mathbf{10} \; {\mathbf 5} \; {\mathbf 6} \; {\mathbf 7} \; 3 \; 2 \; 4 \\
\end{array} \\
\hline
\end{array}$
}
{\small
$\begin{array}{|c|c|}
\hline
\begin{array}[t]{c}
\mathrm{Permutation} \\
\hline
\mathrm{Start} \\
\{ \mu_j' \} \\
\{ \sigma'_j \} \\
\end{array} &
\begin{array}[t]{c}
\mathrm{Result} \\
\hline
10 \; 9 \; 8 \; 7 \; 6 \; 5 \; 4 \; 3 \; 2 \; 1 \\
{\mathbf 9} \; {\mathbf 8} \; {\mathbf 7} \; {\mathbf 4} \; {\mathbf 5} \; {\mathbf 6} \; {\mathbf 1} \; \underline{\mathbf{10}} \; {\mathbf 2} \; {\mathbf 3} \\
{\mathbf 7} \; {\mathbf 9} \; {\mathbf 8} \; 4 \; 5 \; 6 \; 1 \; \underline{10} \; 2 \; 3 \\
\end{array} \\
\hline
\end{array}$
}
\end{center}
\end{table}
\begin{proof}
Denote by $\mathcal{D}$ the set of all Schr\"{o}der prefixes which may be extended to form Schr\"{o}der paths in $\mathcal{D}_n$ for any finite $n$. For $i \ge 0$, let $a_i$ denote the number of Schr\"oder prefixes in $\mathcal{D}$ terminating at the point $(i,i)$ (this is simply $|D_i|$). For $i \ge 1$, let $b_i$ denote the number of Schr\"oder prefixes in $\mathcal{D}$ terminating at the point $(i-1,i)$. We have $a_0 = 1$; also for completeness we define $b_0 = 0$. For $i \ge 1$ the point $(i,i)$ may be reached either by a $\mathsf{d}$ step or by an $\mathsf{e}$ step, and so we have
\begin{equation}
a_i = a_{i-1} + b_i \quad \mbox{for} \: i \ge 1 \; .
\label{eq:a_recursion}
\end{equation}
The justification of this recursion is illustrated in Figure \ref{cap:paths_and_recursion2} for the case of $i=7$. Summing (\ref{eq:a_recursion}) over $i=1,2,\ldots, k$ we obtain
\begin{equation}
a_k = 1 + \sum_{i=0}^{k} b_i \quad \mbox{for} \: k \ge 0 \; .
\label{eq:a_recursion2}
\end{equation}
We have $b_1 = a_0$ since the point $(0,1)$ may only be reached by an $\mathsf{n}$ step. For $i \ge 1$ consider the point $(i-1,i)$; there are $a_{i-1}$ paths which reach this point via an $\mathsf{n}$ step, $b_{i-1}$ paths which reach this point via a $\mathsf{d}$ step, and $b_j$ paths which reach this point via the steps ${\mathsf{n}}^{i-j}{\mathsf{e}}^{i-j}$ for each $j=1,2,\ldots, i-1$. Therefore, for each $i \ge 1$,
\begin{equation}
b_{i} = a_{i-1} + b_{i-1} + \sum_{j=1}^{i-1} b_j = 1 + b_{i-1} + 2\sum_{j=0}^{i-1} b_j
\label{eq:recursion_bis}
\end{equation}
where the second equality is obtained using (\ref{eq:a_recursion2}). The justification of this recursion is illustrated in Figure \ref{cap:paths_and_recursion2} for the case of $i=5$. Subtracting (\ref{eq:recursion_bis}) for $i=k$ from (\ref{eq:recursion_bis}) for $i=k+1$ we obtain
\begin{equation}
b_{k+1} = 4b_k - b_{k-1}
\label{eq:b_recursion}
\end{equation}
for all $k \ge 1$, with $b_0=0$ and $b_1 = 1$.
\begin{figure}
\begin{center}
\includegraphics[%
width=0.52\columnwidth,
keepaspectratio]{paths_and_recursion2.eps}
\end{center}
\caption{\label{cap:paths_and_recursion2} The figure shows graphically the justification of (\ref{eq:a_recursion}) and (\ref{eq:recursion_bis}). In the figure, each lattice point $(i,i)$ (resp. $(i,i-1)$) is labeled with the number $a_i$ (resp. $b_i$) of Schr\"{o}der prefixes in $\mathcal{D}$ which terminate at that lattice point. The figure shows all possible terminations of such prefixes at the points $(7,7)$ and $(4,5)$, thus illustrating that (respectively) $a_7 = a_6 + b_7$ and $b_5 = a_4 + 2 b_4 + b_3 + b_2 + b_1$.}
\end{figure}
\begin{figure}
\begin{center}
\includegraphics[%
width=0.75\columnwidth,
keepaspectratio]{paths_and_recursion1.eps}
\end{center}
\caption{\label{cap:paths_and_recursion1} The figure shows graphically the justification of (\ref{eq:qn_recursion}) for the example of $n=4$. Any Schr\"{o}der prefix in $\mathcal{D}$ which terminates on the line $x+y=2n$ must have one of the possible terminations shown in the figure. Each such prefix then has a unique completion to form a Schr\"{o}der path $p \in \mathcal{D}_{2n}$. Therefore, in the case illustrated, $q_4 = a_3 + b_4 + b_3 + b_2 + b_1$. A similar illustration may be made for the case of odd Schr\"{o}der path length.}
\end{figure}
Next, by Theorem \ref{thm:main_transformation}, the number of permutations $\pi \in \mathfrak{S}_{n+1}(1243,2143)$ which satisfy $\mathrm{rc}(\pi) = \pi$ is equal to the number of Schr\"oder paths $p \in \mathcal{D}_{n}$ which satisfy $\mathrm{rev}(\psi(p)) = p$. First let $q_n$ denote the number of Schr\"oder paths $p \in \mathcal{D}_{2n}$ which satisfy $\mathrm{rev}(\psi(p)) = p$. The initial steps of any such path must form a Schr\"oder prefix in $\mathcal{D}$ terminating on the line $x+y=2n$. This termination occurs either at the point $(n,n)$ (there are $a_{n-1}$ of these -- note that the point $(n,n)$ may not be reached by an $\mathsf{e}$ step) or at the point $(i,2n-i)$ for some $i \in \{0,1,\ldots, n-1 \}$ (there are $b_{i+1}$ of these, as they must join $(i,i+1)$ to $(i,2n-i)$ via the steps $\mathsf{n}^{2(n-i)-1}$). Note that the point $(n,n+1)$ may not be reached via a $\mathsf{d}$ step. It is easy to see that each of these Schr\"oder prefixes has a unique completion to form a Schr\"oder path with $\mathrm{rev}(\psi(p)) = p$. Therefore
\begin{equation}
q_n = a_{n-1} + \sum_{i=0}^{n} b_i = 2b_n - b_{n-1}
\label{eq:qn_recursion}
\end{equation}
for $n \ge 1$ (using (\ref{eq:a_recursion2}) and (\ref{eq:recursion_bis})). The justification of this recursion is illustrated in Figure \ref{cap:paths_and_recursion1}. From (\ref{eq:b_recursion}) we then have $q_{n+1} = 4q_n - q_{n-1}$ with $q_1=2$ and $q_2=7$.
Similarly, let $u_n$ denote the number of Schr\"oder paths $p \in \mathcal{S}_{2n-1}$ which satisfy $\mathrm{rev}(\psi(p)) = p$. The initial steps of any such path must form a Schr\"oder prefix terminating at the point $(n-1,n-1)$ (with the next step joining $(n-1,n-1)$ to $(n,n)$ via a $\mathsf{d}$ step -- there are $a_{n-1}$ of these), or at the point $(i,2n-1-i)$ for some $i \in \{0,1,\ldots, n-1 \}$ (there are $b_{i+1}$ of these, as they must join $(i,i+1)$ to $(i,2n-1-i)$ via the steps $\mathsf{n}^{2(n-i-1)}$). Again, each of these Schr\"oder prefixes has a unique completion to form a Schr\"oder path with $\mathrm{rev}(\psi(p)) = p$; thus $u_n = a_{n-1} + \sum_{i=0}^{n} b_i = 2b_n - b_{n-1}$ for $n \ge 1$ and so $u_n=q_n$ for $n \ge 1$.
\end{proof}
\begin{theorem}\label{Schroder_invs}
\[
\left| \mathcal{CI}_{2n}(1243,2143) \right| = \left| \mathcal{CI}_{2n+1}(1243,2143) \right| = p_n
\]
for all $n \ge 1$, where $p_n$ denotes the $n$-th Pell number, i.e., $p_1=2$, $p_2=5$ and for every $n \ge 3$, $p_n = 2p_{n-1} + p_{n-2}$.
\end{theorem}
\begin{proof}
This proceeds similarly to the proof of Theorem \ref{Schroder_perms}. From Corollary \ref{cor:involutions}, a permutation $\pi \in \mathfrak{S}_n(1243,2143)$ is an involution if and only if $p = \varphi^{-1}(\pi)$ is \emph{symmetric}, i.e., if and only if it satisfies $\mathrm{rev}(p) = p$. Therefore, the number of involutions in $\mathfrak{S}_{n+1}(1243,2143)$ which satisfy $\mathrm{rc}(\pi) = \pi$ is equal to the number of Schr\"oder paths in $p \in \mathcal{S}_n$ which satisfy $\psi(p) = p$; our task is to count these Schr\"oder paths. To this end, let $\tilde{\mathcal{D}}$ denote the set of Schr\"oder prefixes in $\mathcal{D}$ with no features.
For $i \ge 0$, let $c_i$ denote the number of Schr\"{o}der prefixes in $\tilde{\mathcal{D}}$ terminating at the point $(i,i)$. For $i \ge 1$, let $d_i$ denote the number of prefixes in $\tilde{\mathcal{D}}$ terminating at the point $(i-1,i)$. We have $c_0 = 1$; also for completeness define $d_0 = 0$. For $i \ge 1$ the point $(i,i)$ may be reached either by a $\mathsf{d}$ step or by an $\mathsf{e}$ step, and so we obtain
\begin{equation}
c_k = 1 + \sum_{i=0}^{k} d_i \quad \mbox{for} \: k \ge 0
\label{eq:c_recursion}
\end{equation}
by the same method as that which obtained (\ref{eq:a_recursion2}). We have $d_1 = c_0$ since the point $(0,1)$ may only be reached by an $\mathsf{n}$ step. For $i \ge 2$ consider the point $(i-1,i)$; there are $c_{i-2}$ paths which reach this point via an $\mathsf{n}$ step (since such a step must be preceded by a $\mathsf{d}$ step), and $d_j$ paths which reach this point via the steps ${\mathsf{n}}^{i-j}{\mathsf{e}}^{i-j}$ for each $j=1,2,\ldots, i-1$. Therefore, for each $i \ge 2$,
\begin{equation}
d_{i} = c_{i-2} + d_{i-1} + \sum_{j=1}^{i-1} d_j = 1 + d_{i-1} + 2\sum_{j=0}^{i-2} d_j
\label{eq:recursion_dis}
\end{equation}
where we have used (\ref{eq:c_recursion}). Subtracting (\ref{eq:recursion_dis}) for $i=k$ from (\ref{eq:recursion_dis}) for $i=k+1$ we obtain
\begin{equation}
d_{k+1} = 2d_k + d_{k-1}
\label{eq:d_recursion}
\end{equation}
for all $k \ge 1$, with $d_0=0$ and $d_1 = 1$.
Let $p_n$ denote the number of Schr\"oder paths $p \in \mathcal{S}_{2n}$ which satisfy $\psi(p) = p$. The initial steps of any such path must form a Schr\"oder prefix in $\tilde{\mathcal{D}}$ terminating at the point $(n,n)$ (there are $c_{n-1}$ of these -- note that the point $(n,n)$ may not be reached by an $\mathsf{e}$ step) or at the point $(i,2n-i)$ for some $i \in \{0,1,\ldots, n-1 \}$ (there are $d_{n+1}$ of these, as they must join $(i,i+1)$ to $(i,2n-i)$ via the steps $\mathsf{n}^{2(n-i)-1}$). It is easy to see that each of these Schr\"oder prefixes has a unique completion to form a Schr\"oder path with $\psi(p) = p$. Therefore $p_n = c_{n-1} + \sum_{i=0}^{n} d_i = d_{n+1}$ for $n \ge 1$ (again using (\ref{eq:c_recursion})). From (\ref{eq:d_recursion}) we then have $p_{n+1} = 4p_n - p_{n-1}$ with $p_1=2$ and $p_2=5$.
Similarly, let $v_n$ denote the number of Schr\"oder paths $p \in \mathcal{S}_{2n-1}$ which satisfy $\psi(p) = p$. The initial steps of any such path must form a Schr\"oder prefix terminating at the point $(n-1,n-1)$ (with the next step joining $(n-1,n-1)$ to $(n,n)$ via a $\mathsf{d}$ step -- there are $c_{n-1}$ of these), or at the point $(i,2n-1-i)$ for some $i \in \{0,1,\ldots, n-1 \}$ (there are $d_{n+1}$ of these, as they must join $(i,i+1)$ to $(i,2n-1-i)$ via the steps $\mathsf{n}^{2(n-i-1)}$). Again, each of these Schr\"oder prefixes has a unique completion to form a Schr\"oder path with $\psi(p) = p$; thus $v_n = c_{n-1} + \sum_{i=0}^{n} d_i = d_{n+1}$ for $n \ge 1$ and so $v_n=p_n$ for $n \ge 1$.
\end{proof}
\section*{Acknowledgments}
The authors would like to thank the anonymous referees for their helpful suggestions which have greatly improved the accuracy and presentation of this work. They would also like to thank Marilena Barnabei, W. M. B. Dukes and T. Mansour for helpful discussions.
|
\section{Motivations}
For a long time, lattice perturbation theory was the only available
tool for the computation of Lattice QCD Renormalization Constants
(RC's). Since the introduction of methods that allow the non
perturbative computation of the RC's for generic composite
operators~\cite{Martinelli:1994ty}\cite{Luscher:1992an}, these
techniques are preferred. Nevertheless there is no theoretical
obstacle to the perturbative computation of either finite or
logarithmically divergent RC's. In principle, Lattice Perturbation
Theory (LPT) provides the connection between lattice simulations and
continuum perturbative QCD, that works only at high energy. The main
difficulties of LPT are actually practical. First of all, LPT
requires much more effort than in the continuum. Because of this, a
lot of perturbative results are known only at one loop. Second, LPT
series present bad convergence proprieties, so that we should
emphasize that \emph{a fortiori} one loop computations cannot give the
correct results. A good example is provided by the computation of the
quark mass renormalization constant: there appear to be discrepancies
in between determinations coming from perturbative and
non-perturbative techniques~\cite{Blossier:2007vv}. The very point is
that as long as a comparison is made taking into account one loop, it
is virtually impossible to assess the
systematics of both results.\\
In order to obtain higher orders in PT expansions, one can make use of
Numerical Stochastic Perturbation Theory (NSPT)~\cite{Di
Renzo:2004ge}, a numerical implementation of Stochastic
PT~\cite{Parisi:1980ys}. Renormalization constants can be computed in
NSPT at 3 (or even 4) loops, as it has been done in the case of Wilson
fermions with non-improved gauge actions~\cite{Di Renzo:2006wd}. A
nice aspect is that one can work in the massless limit, where RC's are
often defined, and no chiral extrapolation is needed. We will discuss
how one can assess to finite $a$ effects by means of \emph{hypercubic
Taylor expansion}. Moreover, finite volume effects can be corrected
by repeating the computations at different lattice sizes. Measuring
RC's for different values of $n_f$ is also possible to know the
dependence on the number of flavors.
\section{RI'-MOM scheme}
The scheme we will adhere to is the so called RI'-MOM scheme, which
became much popular since the development of non-perturbative
renormalization. RI emphasizes that the scheme is \emph{regulator
independent}, which makes the lattice a viable regulator, while the
prime signals a choice of renormalization conditions which is slightly
different from the original one. An important feature of this scheme
is the fact that the relevant
anomalous dimensions are known up to three loops in the literature~\cite{Gracey:2003yr}.\\
The main observables in our computations are the quark bilinears
between states at fixed off shell momentum $p$:
\begin{equation}\label{Propagator}
\int dx\langle p|\bar\psi(x)\Gamma\psi(x)|p\rangle = G_\Gamma(pa)
\end{equation}
where $\Gamma$ stands for any of the 16 Dirac matrices, returning
the S,~V,~P,~A,~T currents. Our notation points out the $pa$ lattice space
dependence.\\
Since these quantities are gauge dependent we need to fix the
gauge. We will work in Landau gauge, which is easy to fix on the
lattice~\cite{Davies:1987vs}. This gauge condition also gives an
advantage: the anomalous dimension for the quark field is zero at one
loop.\\
Given the quark propagator $S(pa)$, one can obtain the amputated
functions
\begin{equation}\label{Amputated}
G_\Gamma(pa)\to\Gamma_\Gamma(pa) = S^{-1}(pa)G_\Gamma(pa)S^{-1}(pa).
\end{equation}
The $\Gamma_\Gamma(pa)$ are then projected on the tree-level structure
by means of a suitable projector $\hat P_{O_\Gamma}$
\[
O_\Gamma(pa) = \mbox{Tr}\left( \hat P_{O_\Gamma}
\Gamma_\Gamma(pa)\right).
\]
We can finally express the renormalization conditions in terms of the
$O_\Gamma(pa)$ operator:
\[
\left.Z_{O_\Gamma}(\mu a,g(a))Z^{-1}_q(\mu
a,g(a))O_\Gamma(pa)\right|_{p^2=\mu^2} = 1.
\]
The $Z$'s depend on the scale $\mu$ via the dimensionless parameter
$\mu a$, while the dependence on the coupling $g(a)$ is given by the
perturbative expansion. The quark field renormalization constant
$Z_q$ in the formula above is defined by
\begin{equation}\label{Zq:def}
\left.Z_q =
-i\frac{1}{12}\frac{\mbox{Tr}(\slash\!\!\!\!\!pS^{-1}(pa))}{p^2}\right|_{p^2=\mu^2}.
\end{equation}
To obtain a mass independent scheme, we impose these conditions at the
massless point. In the case of Wilson fermions this requires the knowledge of
the critical mass. While one and two loop results are known
from the literature (and can be used as a consistency test) the third loop
is a byproduct of these computations.
\section{Finite lattice size effects}
At the generic n$th$-loop the RC's take the form
\[
z_n = c_n + \sum_{j=1}^n d_j(\gamma)\log^j(pa) + F(pa),
\]
where logarithmic contributions are known from the literature and one is
mainly interested in the finite number $c_n$. The first thing
to do is then to subtract the divergent $\log$s (again,
we take them from the literature). After such a
subtraction we still have the irrelevant contribution $F(pa)$, that
can be fitted by means of a hypercubic Taylor expansion. We show how
this technique works by an example.\\
Consider the two points function in
the continuum limit:
\[
\Gamma_2(p^2) = S^{-1}(p^2).
\]
On the lattice it depends on the dimensionless quantity
$\hat{p} = pa$. Furthermore, we explicit write the dependence on the coupling
\begin{eqnarray*}
\hat{\Gamma}_2(\hat{p}, \hat{m}_{cr},\beta^{-1}) & = & \hat{S}^{-1}(\hat p, \hat{m}_{cr},\beta^{-1}) \\
& = & i\hat{\slash\!\!\!\!\!p} + \hat{m_W}(\hat p) - \hat{\Sigma}(\hat{p}, \hat{m}_{cr},\beta^{-1}),
\end{eqnarray*}
where the $hat$ stands for a dimensionless quantity. Here $\hat m_W$
is the mass term generated by the Wilson prescription (it is $\mathcal
O(\hat{p}^2)$), $\hat{\Sigma}(\hat{p}, \hat{m}_{cr},\beta^{-1})$ is the self
energy and $\hat m_{cr}$ is the quark critical mass. The fermion mass
counterterm arises because Wilson regularization breaks chiral
symmetry. Both these last two terms are $\mathcal O(\beta^{-1})$.\\
The self energy can be written as
\[ \hat{\Sigma}(\hat{p}, \hat{m}_{cr},\beta^{-1}) =
\hat{\Sigma}_c(\hat{p}, \hat{m}_{cr},\beta^{-1}) +
\hat{\Sigma}_V(\hat{p}, \hat{m}_{cr},\beta^{-1}) +
\hat{\Sigma}_{other}(\hat{p}, \hat{m}_{cr},\beta^{-1}).
\] These are the different contribution along the different elements
of the Dirac base. In particular, $\hat{\Sigma}_c$ is the contribution
along the identity, $\hat{\Sigma}_V$ the contribution along the gamma
matrices.\\ The self energy $\hat{\Sigma}_c$ contains the contribution of
the mass:
\[ \hat{\Sigma}(0, \hat{m}_{cr},\beta^{-1}) = \hat{\Sigma}_c(0,
\hat{m}_{cr},\beta^{-1}) = \hat m_{cr} = am_{cr}.
\] To show how hypercubic Taylor expansions work, we can concentrate
on the $\hat\Sigma_V$ term, from which we can in this way extract the
quark field RC. We can consider a Taylor expansion of $\hat\Sigma_V$
in the form
\begin{equation}\label{SigmaV:expansion}
\hat\Sigma_V(\hat{p}, \hat{m}_{cr},\beta^{-1}) =
i\sum_\mu\gamma_\mu\hat p_\mu\left( \hat\Sigma_V^{(0)}(\hat{p},
\hat{m}_{cr},\beta^{-1})+ \hat p^2_\mu\hat\Sigma_V^{(1)}(\hat{p},
\hat{m}_{cr},\beta^{-1})+ \hat p^4_\mu\hat\Sigma_V^{(2)}(\hat{p},
\hat{m}_{cr},\beta^{-1})+ \ldots \right)
\end{equation}
where the expansion entails an expansion in powers of $a$. The
functions $\hat\Sigma_V^{(i)}(\hat{p}, \hat{m}_{cr},\beta^{-1})$ are
combinations of hypercubic invariants. As an example, the first term
in the expansion~(\ref{SigmaV:expansion}) can be written as
\begin{equation}\label{SigmaV0:expansion}
\hat\Sigma_V^{(0)}(\hat{p}, \hat{m}_{cr},\beta^{-1}) =
\alpha_1^{(0)} 1 +
\alpha_2^{(0)} \sum_\nu p_\nu^2 +
\alpha_3^{(0)} \sum_\nu p_\nu^4 + \alpha_4^{(0)} \left(\sum_{\nu} p_\nu^2\right)^2 +
\mathcal{O}(a^6)
\end{equation}
As a general recipe, all the possible covariant polynomials can be
found via a character's projection of the polynomial representation of
the hypercubic group onto the defining representation of the group.
One can see that plugging this expansion in the definition~(\ref{Zq:def}) of
$Z_q$ the only term that doesn't vanish in the continuum limit is
$\alpha_1^{(0)}$, the first coefficient of the expansion of
$\hat\Sigma_V^{(0)}$.
\section{Finite volume effects}
In the previous section we have shown how one can overcome the effects
due to the discretization. These are not the only systematic effects in a
lattice simulation. Though large lattices are available, one
can't neglect the finite volume effects. This means that on a finite
lattice we have to consider also a $pL$ dependence in our
quantities.\\
We will now show how the analysis of $\Sigma = \hat{\Sigma}(pa, pL,
am_{cr},\beta^{-1})$ can be modified to take care of the finite volume
effects. Let us take $\hat{\Sigma}^{(0)}_V(pa, pL,
am_{cr},\beta^{-1})$ as an example. In the spirit of
\cite{Kawai:1980ja}, consider the ansatz
\begin{eqnarray*}
\hat{\Sigma}^{(0)}_V(pa, pL) & = & \hat{\Sigma}^{(0)}_V(pa) + \left(\hat{\Sigma}^{(0)}_V(pa, pL) - \hat{\Sigma}^{(0)}_V(pa) \right)\\
& = & \hat{\Sigma}^{(0)}_V(pa) + \Delta\hat{\Sigma}^{(0)}_V(pa, pL)
\end{eqnarray*}
where $\hat{\Sigma}^{(0)}_V(pa) = \hat{\Sigma}^{(0)}_V(pa, \infty)$
and to keep the formula simpler we omit the dependence on
$\hat m_{cr}$ and $\beta^{-1}$ .\\
We can now expand as in~(\ref{SigmaV0:expansion}):
\begin{equation}\label{SigmaVpL}
\hat\Sigma_V^{(0)}(\hat{p}, pL) =
\alpha_1^{(0)} 1 +
\alpha_2^{(0)} \sum_\nu p_\nu^2 +
\alpha_3^{(0)} \sum_\nu p_\nu^4 + \alpha_4^{(0)} \left(\sum_{\nu} p_\nu^2\right)^2 +
\Delta\hat{\Sigma}^{(0)}_V(pL) +
\mathcal{O}(a^6)
\end{equation}
where the continuum limit $\Delta\hat{\Sigma}^{(0)}_V(pL) =
\Delta\hat{\Sigma}^{(0)}_V(pa = 0, pL)$ has been taken into
account. The rationale for this is the following: $pL$ effects are
present also in the continuum limit, and $pa$ corrections on top
of those are regarded as \emph{corrections on corrections}. In principle, also
these corrections can be taken into account, but this requires a
larger number of parameters.\\
The latter assumption has a strong consequence: measurements on
lattices of different size are affected by the same $pL$ effect once
one consider the same tuples $\left(n_1,n_2,n_3,n_4\right)$
\[
p_\mu L = \frac{2 \pi n_\mu}{L} L = 2\pi n_\mu.
\]
The practical implementation is the following. First of all one has to
select a collection of lattice sizes and an interval
$[(pa)^2_{min},(pa)^2_{max}]$. Then one considers all the tuples $\vec
n = \left(n_1, n_2, n_3, n_4\right)$ which, for all the lattice sizes,
fall in $\left[(pa)^2_{min},(pa)^2_{max}\right]$ and chooses one
representative for all the tuples connected by an H4
transformation. We further add to these data the measure taken at the
highest value of $(pa)^2$ which falls in the interval
$\left[(pa)^2_{min},(pa)^2_{max}\right]$ on the lattice with the
biggest value of $N=L/a$. Assuming that this tuple
(which we will call $n^*$) is a good approximation to $pL = \infty$, the measure will be considered as a normalization point. Is then possible to fit the parameters in the expansion (in our case the $\alpha_i^{(0)}$), and the finite volume corrections, which came in the same number of the considered tuples.\\
To be explicit,~(\ref{SigmaVpL}) can then be formulated in terms of
the $(\vec n, N)$ dependence:
\begin{equation}\label{SigmaVn}
\hat\Sigma_V^{(0)}(\vec n\neq\vec{n^*}, N) =
\alpha_1^{(0)} 1 +
\alpha_2^{(0)} p^2(\vec n, N) +
\alpha_3^{(0)} p^4(\vec n, N) + \alpha_4^{(0)} \left(p^2(\vec n, N)\right)^2 +
\Delta^{(0)}_{\vec n} +
\mathcal{O}(a^6)
\end{equation}
where $\vec n^*$ means the tuple taken as normalization point. For the measure taken at the tuple $\vec n^*$ the formula is slightly different:
\begin{equation}\label{SigmaVnstar}
\hat\Sigma_V^{(0)}(\vec{n^*}, N) =
\alpha_1^{(0)} 1 +
\alpha_2^{(0)} p^2(\vec{n^*}, N) +
\alpha_3^{(0)} p^4(\vec{n^*}, N) + \alpha_4^{(0)} \left(p^2(\vec{n^*}, N)\right)^2 +
\mathcal{O}(a^6)
\end{equation}
In~(\ref{SigmaVn}) and~(\ref{SigmaVnstar}) $p^2 = \sum_\nu p_\nu^2$ and $p^4 = \sum_\nu p_\nu^4$.
Two issues of the procedure can only be assessed a posteriori: the
assumption that the measure at the normalization point is free from
finite volume effects and the stability of the fit.\\
\section{Results}
In the previous sections, we discussed the procedure to obtain results and keep finite $a$ and $pL$ effects under control. Since the work is still in progress, we only display the first technique at work. We computed at every order relevant expectation values given by~(\ref{Propagator}) and amputated to obtain~(\ref{Amputated}). Finally we projected on the tree level structure and performed the $pa$ analysis by means of the hypercubic Taylor expansion.\\
We are not presenting any new result (apart a preliminar value for
critical mass at three loop), but we can compare some numerical results at leading order with analytical results~\cite{Aoki:1998ar}.\\
The first quantity we consider is the simpler vector one: the measure
of the RC for the quark propagator. This is a log free quantity, so we
only need to extract the correct $a\to 0$ limit. In figure~\ref{Zq} we
show the computation of $Z_q$. It's easy to recognize the effect of
different lengths of $\hat p_\mu$ in the relevant direction.\\
\begin{figure}[!htbc]
\begin{centering}
\includegraphics[width=0.7\textwidth]{f2.eps}
\end{centering}
\caption{Computation at one loop of $Z_q$. Different ``families'' of
point correspond to the different components of momentum $p$ along
the relevant direction. In red measured values at given
$(p^2,p_\mu)$, in red the fit results and in blue the extrapolation
at $a = 0$.}
\label{Zq}
\end{figure}
As an example of a logarithmically divergent quantity, in
figure~\ref{Zs} we show the computation of $Z_s$. In the figure, red
point are the measurements before the $\log$ subtraction, blue the
result after the $\gamma_s^{(1)}\log{\hat{p}^2}$ subtraction. Since
this is a scalar quantity, we haven't the effect of the length of the
components in a given direction.\\
\begin{figure}[!htbc]
\begin{centering}
\includegraphics[width=0.7\textwidth]{scalare.eps}
\end{centering}
\caption{Computation at one loop of $Z_s$. The anomalous dimension at
one loop is different from zero, and the measure shows a logarithmic
divergence (red points). In blue the results after the
$\gamma_s\log{\hat p}^2$ contribution subtraction.}
\label{Zs}
\end{figure}
In figure~\ref{mcr} we present the preliminar result for the critical
mass at three loop, $m_{cr}^{(3)}$. This quantity doesn't present
either vector structure or $\log$ divergences. Critical mass is a
byproduct of all the previous computations. The introduction of this
counterterm is needed in the computation of the next loop quantities.
\begin{figure}[!htbc]
\begin{centering}
\includegraphics[width=0.7\textwidth]{m6a.eps}
\end{centering}
\caption{Preliminar result for the critical mass at three loop. In red the result of measurements, in black the result of the fitting procedure.}
\label{mcr}
\end{figure}
\section{Work in progress}
The aim of this work is to compute high order renormalization
coefficients for masses, fields and bilinears for different
actions. Analytic results are known at most up to 2 loops, while there
are NSPT results for Wilson gauge-Wilson fermions up to 3 loops at
various $n_f$. We are in the process of taking into account $pL$
effects for the latter action. We are now applying the method also to
Tree Level Symanzik (current work) and Iwasaki gauge actions. For
these actions we're also interested in computing the $n_f$
dependence. Current data are obtained from $32^4$ lattices measured on
an APE machine, while a C++ code is working on smaller lattices on
standard workstations.
\newpage
|
\section{Introduction}
The existence of dark matter (DM) has been established by many
astrophysical observations, but the nature of DM paticle is still
unclear. Among the large amount of candidates proposed in many
theories of new physics, the weakly interacting massive particle
(WIMP) is the most popular and attractive one
\cite{1996PhR...267..195J,2005PhR...405..279B}. The mass of WIMP
is generally from a few GeV to TeV, and the interaction strength
is of the weak scale, which can give the right relic density of
DM. In this scenario, the weak interaction of DM particles would
produce observable standard model particles, such as charged
anti-matter particles, photons and neutrinos. Investigating
such particles from the cosmic rays (CRs) is the task of DM
indirect detection.
The recently reported new signatures of CR positrons, antiprotons
and electrons by PAMELA
\cite{2009Natur.458..607A,2009PhRvL.102e1101A}, ATIC
\cite{2008Natur.456..362C}, HESS \cite{2008PhRvL.101z1104A,
2009A&A...508..561A} and Fermi-LAT \cite{2009PhRvL.102r1101A} have
stimulated great interests and extensive studies of the DM indirect
searches. The DM scenario with mass $O$(TeV), leptonic
annihilation/decay final states and a high annihilation/decay rate
can well explain the observational data (e.g.,
\cite{2009NuPhB.813....1C,2009PhRvD..79b3512Y}). Furthermore, more
quantitative constraints on the DM model parameters can be derived
through a global fitting method
\cite{2010PhRvD..81b3516L,2009arXiv0911.1002L}.
Regardless of detailed models of DM to explain the data, it is
essential to find observable signals to test the models. Since the
charged particles will gyrate in the magnetic field and lose most of
the source information, it is difficult to test the DM models using
only the data of charged CRs. Gamma-rays and neutrinos seem to be
very good probes. There are several advantages of using $\gamma$-ray
photons and neutrinos to investigate the DM models. Firstly, photons
and neutrinos propagate along straight line and can trace back to
the source sites where the DM annihilation/decay takes place.
Secondly there is little interaction during the propagation and most
of the primary source information hold. Thirdly the effective volume
of which photons and neutrinos can reach is much larger than that of
charged particles, e.g., from the Milky Way to extragalactic space,
and even the early Universe. It has been shown in some works that
$\gamma$-rays and neutrinos can be powerful tools to test the DM
scenarios explaining the CR lepton data (e.g.,
\cite{2009JCAP...03..009B,2009PhRvD..80b3007Z,2009PhRvD..79h1303B,
2009arXiv0908.1236Z,2009arXiv0912.0663C,2010JCAP...03..014P,
2009arXiv0912.4504Z,2009arXiv0908.4317C},
\cite{2009PhRvD..79d3516H,2009PhRvD..79f3522L,2009arXiv0905.4764S,
2010PhRvD..81a6006B,2010PhRvD..81d3508M,2010PhRvD..81h3506S,
2010JCAP...04..017C}).
There are many sites proposed to be good candidates for the search
of $\gamma$-rays and neutrinos from DM, such as the Galactic center
\cite{2009JCAP...03..009B,2009PhRvD..80b3007Z,2009PhRvD..79h1303B},
Galactic halo \cite{2009arXiv0908.1236Z,2009ApJ...699L..59B,
2010JCAP...03..014P}, satellite galaxies or substructures
\cite{2009PhRvD..80b3506E,2009MNRAS.399.2033P,
2009Sci...325..970K,2009arXiv0908.0195P}, the extragalactic space
\cite{2009PhRvD..80b3517K,2009JCAP...07..020P,2010MNRAS.405..593Z}
and the emissions at the early Universe
\cite{2009A&A...505..999H,2009JCAP...10..009C,2009arXiv0912.2504Y}.
As the largest gravitational bounding system in the Universe, galaxy
clusters may also be useful for DM indirect searches
\cite{2009PhRvD..80b3005J}. Pinzke et al. investigated the
$\gamma$-ray emission from nearby clusters and used EGRET upper
limits to set constraints on the DM model parameters
\citep{2009PhRvL.103r1302P}. They found that if the DM annihilation
was responsible for the electron/positron excesses and the
luminosity-mass distribution of DM substructures in clusters
followed the extrapolation of numerical simulation results, the
minimum mass of DM subhalos should be larger than $10^{-2}$
M$_{\odot}$ in order not to exceed the EGRET limits. This is a
useful way to study the particle nature of DM through structures.
After more than one year's operation, Fermi-LAT reported some
results about the $\gamma$-ray emission from galaxy clusters
\cite{Fermi-LAT:cluster}. Non detection of significant $\gamma$-ray
emission from galaxy clusters was reported except for Perseus
cluster, in which the emission from the central galaxy NGC 1257 was
discovered \cite{2009ApJ...699...31A}. The upper limits given by
Fermi-LAT are lower by more than one order of magnitude than that
given by EGRET. It can be expected that the new results from
Fermi-LAT will set much stronger constraints on the DM models.
In Ref. \cite{Fermi-LAT:cluster} the constraints on DM mass and
annihilation cross section were presented assuming $\mu^+\mu^-$ and
$b\bar{b}$ channels. In this work we will also use the Fermi-LAT upper
limits to constrain DM model parameters. Different from Ref.
\cite{Fermi-LAT:cluster}, we will pay more attention on the implication
of DM structure properties such as the minimal mass of subhalo
$M_{\rm min}$, which would be important for understanding the nature
of DM particle. This is one of the motivations of this study.
Another motivation of this work is the neutrino emission. Neutrinos
can be served as an independent diagnostic of DM indirect searches
besides photons. It has been shown that the measured atmospheric
neutrino background can set effective constraints on the DM
annihilation cross section
\cite{2007PhRvL..99w1301B,2007PhRvD..76l3506Y}. There are no high
energy astrophysical neutrinos being detected currently, so it is
valuable to explore the sensitivity of the forthcoming neutrino
detectors to the neutrino signals from the DM annihilation. Due to
the very weak interaction cross section between neutrinos and
matter, we generally need large detector volume. The ongoing
experiment IceCube has an effective volume $\sim$km$^3$, which would
give unprecedented sensitivity for the neutrino detection up to very
high energies. The detectability of neutrino signals from DM
annihilation in galaxy clusters by e.g. IceCube, will be discussed
in this work.
This paper is organized as follows. In Sec. II, we discuss the
$\gamma$-ray emission from several galaxy clusters, and employ the
recent Fermi-LAT limits of these clusers to constrain the DM model
parameters. In Sec. III, we discuss the detectability of neutrino
emission from the galaxy clusters by the neutrino detectors. The
last section is our conclusions and discussions.
\section{Gamma rays from galaxy clusters}
\subsection{Cluster sample}
It is known that the objects with high masses and small distances will
be very efficient for the DM searches. Therefore nearby massive clusters
are the first choice of study. Here we adopt a sample of $6$ clusters
with redshift from $0.0031$ to $0.0231$ (corresponding to distance from
$13$ to $100$ Mpc for a standard $\Lambda$CDM cosmology), which are
reported with flux upper limits by Fermi-LAT. The basic parameters
of these clusters are compiled in Table \ref{table:sample}.
The flux of DM annihilation from cluster is generally scaled with
$M_{200}^{\alpha}/d^2$, where $\alpha$ depends on the concentration-mass
relation and DM profile of the halo. In this work we will assume NFW
profile for the halo of cluster. The final results are expected not
sensitively dependent on the profile since most of the cluster lies in
the angular window of Fermi-LAT ($3.5^{\circ}$ for 100 MeV,
\cite{2009ApJ...697.1071A}). As a benchmark configuration, we adopt the
concentration-mass relation fitted from X-ray observations
\cite{2007ApJ...664..123B}
\begin{equation}
c_{\rm vir}=\frac{9.0}{1+z}\times\left(\frac{M_{\rm vir}}{10^{14}h^{-1}
{\rm M}_{\odot}}\right)^{-0.172}.
\end{equation}
After correcting the definition of virial overdensity in Ref.
\cite{2007ApJ...664..123B} ($\Delta\approx 100$) to $\Delta=200$
we have \cite{2003ApJ...584..702H}
\begin{equation}
c_{200}=\frac{6.9}{1+z}\times\left(\frac{M_{200}}{10^{14}
{\rm M}_{\odot}}\right)^{-0.178}.
\end{equation}
For this concentration-mass relation we find $\alpha\approx 0.65$.
Comparing the quantity $M_{200}^{\alpha}/d^2$ among these clusters,
we find that DM signals from clusters NGC 4636, M49 and Fornax are
of the same level, and are several times larger than the rest three
clusters. In the following we will see that these three clusters
will indeed give stronger constraints on the DM models.
\begin{table}[htb]
\centering
\caption{Cluster sample}
\begin{tabular}{cccccc}
\hline \hline
Name & $z$\footnotemark[1] & R.A.\footnotemark[1] & Dec.\footnotemark[1] & $M_{200}$($10^{14}$M$_{\odot}$)\footnotemark[2] & $r_{200}$(Mpc)\footnotemark[2] \\
\hline
NGC 4636 & $0.0031$ & $12^h43^m$ & $2^{\circ}41'$ & $0.25$ & $0.60$\\
M49 & $0.0033$ & $12^h30^m$ & $8^{\circ}00'$ & $0.46$ & $0.73$\\
Fornax & $0.0046$ & $03^h39^m$ & $-35^{\circ}27'$ & $1.00$ & $0.95$\\
Centaurus & $0.0114$ & $12^h49^m$ & $-41^{\circ}18'$ & $2.66$ & $1.32$\\
AWM 7 & $0.0172$ & $02^h55^m$ & $41^{\circ}35'$ & $4.28$ & $1.54$\\
Coma & $0.0231$ & $13^h00^m$ & $27^{\circ}59'$ & $13.65$ & $2.27$\\
\hline
\hline
\end{tabular}
\footnotetext[1]{Redshift and coordinates are adopted from NASA/IPAC
Extragalactic Database, http://nedwww.ipac.caltech.edu/}
\footnotetext[2]{Virial mass and radius parameters are taken from
Ref. \cite{2002ApJ...567..716R}.}
\label{table:sample}
\end{table}
\subsection{Gamma-ray emission from DM distribution in clusters}
There are generally two kinds of $\gamma$-ray emission from DM
annihilation: one is produced directly from the annihilation final
state particles which is called {\it primary} emission (such as
the $\gamma$ rays by $\pi^0$ decay after hadronization or emission
directly from final charged leptons), and the other is produced
through interactions of final state particles with external medium
or radiation field such as the inverse Compton (IC) radiation
which is called {\it secondary} emission hereafter. The primary
$\gamma$-ray flux observed on the Earth from DM annihilation in a
galaxy cluster can be expressed as
\begin{equation}
\phi^{pri}=\frac{\langle\sigma v\rangle}{2m_{\chi}^2}\frac{{\rm d}N}
{{\rm d}E}\times \frac{\int \rho^2(r) {\rm d}V}{4\pi d_L^2},
\label{phi}
\end{equation}
where $m_\chi$ is the mass of DM particle, $\langle\sigma
v\rangle$ is the annihilation cross section of DM, $\frac{{\rm
d}N}{{\rm d}E}$ is the yield spectrum of $\gamma$-rays per
annihilation which is simulated using PYTHIA
\cite{2006JHEP...05..026S}, $d_L$ is the luminosity distance of
the cluster, $\rho(r)$ is the density distribution of DM inside
the cluster with $r$ the distance from the cluster center. All of the
cluster is taken into account in the integral since the analysis of
Fermi-LAT was done in a $10$ degree radius of each cluster
\cite{Fermi-LAT:cluster}, which is large enough to contain the whole
cluster halo. For the smooth halo we assume the density distribution
to be NFW profile \cite{1997ApJ...490..493N}
\begin{equation}
\rho_{\rm sm}(r)=\frac{\rho_s}{(r/r_s)(1+r/r_s)^2},
\label{nfw}
\end{equation}
where parameters $r_s$ and $\rho_s$ can be determined by the
concentration-mass relation and normalization of total mass.
Since there are substructures in the clusters, such as galaxy
groups and galaxies, we have to take these into account. We will
see later that the existence of substructures enhances the
annihilation luminosity of DM and is the main reason that affects
the $\gamma$ flux. To take into account the effect of
substructures, we replace $\rho^2$ in Eq. (\ref{phi}) with
$\rho^2_{\rm tot}\equiv\rho_{\rm sm}^2 +\langle\rho_{\rm
sub}^2\rangle$, where the average density square of substructures
reads
\begin{equation}
\langle\rho_{\rm sub}^2(r)\rangle=\int{\rm d}M\frac{{\rm d}N}
{{\rm d}V{\rm d}M}\times L(M),
\label{sub}
\end{equation}
in which $\frac{{\rm d}N}{{\rm d}V{\rm d}M}$ is the number density of
subhalos in mass bin ${\rm d}M$, $L(M)=\int_{V_{\rm sub}}\rho_{\rm sub}^2
{\rm d}V'$ is the intrinsic annihilation luminosity of a subhalo with mass
$M$. In this work we will employ the results from recent high resolution
simulation, {\it Aquarius} \cite{2008Natur.456...73S,2008MNRAS.391.1685S}
to treat the subhalos. Because the concentration and density profile of
subhalos are very complicated inside the host halo, the detailed
computation using Eq. (\ref{sub}) is difficult. Thus we directly
adopt the counted results of luminosity distribution from the
simulation\footnote{This relation is different from that given in Ref.
\cite{2008Natur.456...73S}, where $L(>M)\approx M^{-0.226}$ was found.
According to this fit we have $L(>M)\approx \left(M^{-0.16}-
M_{\rm max}^{-0.16}\right)$, which gives similar behavior as that in
Ref. \cite{2008Natur.456...73S} in the large (resolved) mass range of
the simulation, but is different when extrapolating to low (unresolved)
mass range.}
\cite{Yuan:neutrino}
\begin{equation}
\frac{{\rm d}{\mathcal L}}{{\rm d}M}(r,M)\equiv\frac{{\rm d}N}{{\rm d}V
{\rm d}M}\times L(M)\propto \left(\frac{r}{0.2r_{200}}\right)^{-0.1}
\left(1+\frac{r}{0.2r_{200}}\right)^{-2.9}\times M^{-1.16}.
\label{lumin_sub}
\end{equation}
Similar with Ref. \cite{2009PhRvL.103r1302P} we adopt a scale between
the Milky Way like halo given in {\it Aquarius} simulation and the case
of clusters, i.e., the ratio of ${\mathcal L}_{\rm sub}/{\mathcal L}_
{\rm sm}$ keeps unchanged whatever the mass is. The maximum mass of
subhalos found in simulation is about $0.01M_{\rm host}$. But the
minimum mass is not well known due to the limit of resolution of the
numerical simulation. From the observational point of view, we have
observed DM halos with mass $\sim 10^7$ M$_{\odot}$, e.g. dwarf
galaxies. While the study of free streaming of cold DM (CDM) particles
indicates a minimum halo mass down to $\sim 10^{-7}$ M$_{\odot}$
\cite{2001PhRvD..64h3507H}. In this work we leave $M_{\rm min}$ to
be a free parameter and investigate the effects of $M_{\rm min}$ on
the DM signals.
Besides the primary $\gamma$-ray emission, there is also secondary
production of $\gamma$-ray photons through the IC scatterings between
the DM induced electrons/positrons and the cosmic microwave background
(CMB) field. For the calculation of the fluxes of secondary IC emission
please see the Ref. \cite{2009arXiv0908.1236Z}. Note that when
calculating the energy loss
rate of electrons/positrons, both the IC loss induced by scattering with CMB
photons and the synchrotron loss in the magnetic field in clusters are
considered. The average value of magnetic field strength is assumed to
be $\sim 1$ $\mu$G \cite{1998APh.....9..227C}.
\begin{figure}[!htb]
\begin{center}
\includegraphics[width=0.47\columnwidth]{spec_mu.eps}
\includegraphics[width=0.47\columnwidth]{spec_bb.eps}
\caption{Integral spectra of the IC and FSR components from DM
annihilation in Fornax cluster. The left panel is for $\mu^+\mu^-$
channel, and the right panel is for $b\bar{b}$ channel respectively.
Also shown are the $95\%$ upper limits (arrows) of Fermi-LAT 1 yr
observations. See the text for details.
\label{fig:spec}}
\end{center}
\end{figure}
For illustration we show the integral spectra of $\gamma$-rays
from DM annihilation in Fornax cluster in Fig. \ref{fig:spec}.
Here we adopt a sample DM model with $m_\chi=1$ TeV,
$\langle\sigma v\rangle=10^{-23}$ cm$^3$ s$^{-1}$, and the
annihilation channels with $\mu^+\mu^-$ (left) and $b\bar{b}$
(right) respectively. The {\it primary} and {\it secondary}
components are shown separately. In each group we show three
curves which represent the smooth halo contribution, the total
emission with subhalos down to two different $M_{\rm min}$. The
$95\%$ confidence level upper limits from Fermi-LAT are shown by
arrows. It is shown that in the energy range interested here,
i.e. $0.1-10$ GeV, the {\it secondary} radiation from IC is dominant
for $\mu^+\mu^-$ channel, while for $b\bar{b}$ channel the
{\it primary} contribution is dominant. This is because for $\mu^+\mu^-$
final state the spectra of photons and electrons are relatively hard,
and the {\it secondary} produced photons through IC can just lie in
the interested energy range. For $b\bar{b}$ final state the energies
of photons and electrons from the hadronic cascade are generally much
lower than the mass of DM, so the IC component dominates at even lower
energies ($<0.1$ GeV). This conclusion will always hold for $m_\chi$
ranging from $100$ GeV to $10$ TeV for $\mu^+\mu^-$ channel ($10$ GeV
to $1$ TeV for $b\bar{b}$), which is the sensitive region explored by
Fermi-LAT.
\begin{figure}[!htb]
\begin{center}
\includegraphics[width=0.47\columnwidth]{NGC4636_con_mu.eps}
\includegraphics[width=0.47\columnwidth]{M49_con_mu.eps}
\includegraphics[width=0.47\columnwidth]{Fornax_con_mu.eps}
\includegraphics[width=0.47\columnwidth]{Centaurus_con_mu.eps}
\includegraphics[width=0.47\columnwidth]{AWM7_con_mu.eps}
\includegraphics[width=0.47\columnwidth]{Coma_con_mu.eps}
\caption{Fermi-LAT constraints on the DM model parameters $m_{\chi}$ and
$\langle\sigma v\rangle$ for different minimum halo mass $M_{\rm min}$.
The DM annihilation channel is $\mu^+\mu^-$. Dashed circles are the
$3\sigma$ and $5\sigma$ parameters regions which can fit the
PAMELA/Fermi-LAT/HESS data of the CR positrons/electrons
\cite{2010NuPhB.831..178M}.
\label{fig:mu}}
\end{center}
\end{figure}
\begin{figure}[!htb]
\begin{center}
\includegraphics[width=0.47\columnwidth]{Fornax_con_mu_M08.eps}
\includegraphics[width=0.47\columnwidth]{Fornax_con_mu_VL.eps}
\caption{Comparison of the constraints for Maccio et al. (2008)
concentration model (left, \cite{2008MNRAS.391.1940M}) and {\it Via
Lactea} substructure model (right, \cite{2008ApJ...686..262K}).
\label{fig:comp}}
\end{center}
\end{figure}
\begin{figure}[!htb]
\begin{center}
\includegraphics[width=0.47\columnwidth]{NGC4636_con_bb.eps}
\includegraphics[width=0.47\columnwidth]{M49_con_bb.eps}
\includegraphics[width=0.47\columnwidth]{Fornax_con_bb.eps}
\includegraphics[width=0.47\columnwidth]{Centaurus_con_bb.eps}
\includegraphics[width=0.47\columnwidth]{AWM7_con_bb.eps}
\includegraphics[width=0.47\columnwidth]{Coma_con_bb.eps}
\caption{The same as Fig. \ref{fig:mu} but for $b\bar{b}$ final state
of DM annihilation.
\label{fig:bb}}
\end{center}
\end{figure}
In \cite{Fermi-LAT:cluster} the Fermi-LAT has derived preliminary
constraints on the $\gamma$-ray fluxes from DM annihilation from
these clusters by assuming the $\mu^+\mu^-$ and $b\bar{b}$ final
states. According to these results, we derive the constraints on
the $m_{\chi}-\langle\sigma v\rangle$ plane. In Figs. \ref{fig:mu}
we show the constraints for the $\mu^+\mu^-$ channel. It is shown
that the constrains of the smooth component (that
$M_{\rm min}=10^{12}M_\odot$) is still very weak. If the substructures
are taken into account the constraints can be stronger by more
than one order of magnitude, depending on the free streaming mass
of DM subhalo. If the minimum mass of DM subhalos can be as low as
$10^{-7}-10^{-6}$ M$_\odot$ as predicted in the supersymmetric DM
scenario, the constraint on the cross section of TeV DM can reach
$\sim 10^{-24}$ cm$^3$ s$^{-1}$. This constraint is actually
powerful enough to explore the DM models which are proposed to
explain the CR lepton excesses. In Fig. \ref{fig:mu} we plot the
favored parameter regions of the DM model to explain the
PAMELA/Fermi-LAT/HESS data of the CR positrons/electrons with
$\mu^+\mu^-$ final state \cite{2010NuPhB.831..178M}. We can see
that for almost all of the 6 clusters the PAMELA/Fermi-LAT/HESS
favored parameter regions can be excluded if $M_{\rm min}$ is down
to $\sim 10^{-6}$ M$_{\odot}$. If DM annihilation is responsible
for the observational positron/electron excesses, it may indicate
that the cutoff of $M_{\rm min}$ should be much larger
\cite{2009PhRvL.103r1302P}. For clusters Fornax and NGC 4636, the
constraint of $M_{\rm min}$ is about $10^{2}-10^{3}$ M$_\odot$ for
the best fit mass and cross section\footnote{Note that here we assume
a constant boost factor of the DM models responsible to the
electron/positron excesses. However, as proposed in the literature
the large boost factor might be due to velocity-dependent annihilation
cross section such as the Sommerfeld effect \cite{2004PhRvL..92c1303H,
2009PhRvD..79a5014A}. In that case the boost factor might be different
in clusters from in the Galaxy. We do not discuss this effect in detail
in this study.}. This constraint is much
stronger than that derived using EGRET upper limit about Virgo
cluster \cite{2009PhRvL.103r1302P}.
This cutoff of the mass of substructures may have important
implication of the particle nature of DM. It gives an estimate of
the free streaming scale of the matter power spectrum as $k<750$
Mpc$^{-1}$, which is not very far from the lower limits given by
Lyman-$\alpha$ power spectrum measurements
\cite{1999ApJ...520....1C,2006ApJS..163...80M}. Compared with the
canonical value expected in CDM picture, it favors a warm massive
DM scenario which may be produced non-thermally in the early
Universe \cite{2001PhRvL..86..954L}. Such nonthermally produced DM
particles have large initial velocity and large free streaming.
Thus the matter power spectrum is suppressed at small scales and
leads to less low mass subhalos \cite{2004PhRvD..69l3521B,
2009PhRvD..80j3502B}.
As a check of the model uncertainties of the halo structure
configuration, we compare the results using Maccio (2008)
concentration-mass relation given in Ref. \cite{2008MNRAS.391.1940M}
\begin{equation}
c_{200}=\frac{3.56}{1+z}\times\left(\frac{M_{200}}{10^{15}
{\rm M}_{\odot}}\right)^{-0.098}.
\end{equation}
For substructures we also use the results from another high-resolution
simulation {\it Via Lactea}. According to Fig. 3 of Ref.
\cite{2008ApJ...686..262K}, the substructure enhancement is simply
extracted to be $B=10\times\left(M_{\rm min}^{-0.048}-M_{\rm max}^{-0.048}
\right)$ with $M_{\rm max}=0.01M_{\rm host}$. Note that there is a host
halo mass dependence of $B$ as given in Ref. \cite{2008ApJ...686..262K}.
However, since the results with $M_{\rm host}$ larger than the galaxy scale
halo were not calibrated in the simulation, we adopt $B(M_{\rm host})$ at
$M_{\rm host}=10^{12}$ M$_{\odot}$ and apply it to the cluster scale halos.
This treatment is also consistent with the above assumption that we
adopt the scaling relation of Eq. (\ref{lumin_sub}) to be the same for
Milky Way halo and cluster halos. The extrapolation to the cluster masses
following the $M_{\rm host}$ dependence of $B$ in Ref.
\cite{2008ApJ...686..262K} would lead to a two times larger boost factor.
The constraints for $\mu^+\mu^-$ channel from Fornax cluster for these
two models are shown in Fig. \ref{fig:comp}. It is shown that for Maccio
(2008) concentration the constraint is about $2$ times weaker than the
benchmark model. And for {\it Via Lactea} substructures the total
substructure enhancement is some weaker than that of {\it Aquarius}
simulation. However, in both of these cases we see that the
PAMELA/Fermi-LAT/HESS favored parameter regions can be constrained,
given $M_{\rm min}\sim 10^{-6}$ M$_{\odot}$. For the best fit mass and
cross section the constraint on $M_{\rm min}$ is about $10^0-10^2$
M$_{\odot}$.
The results for $b\bar{b}$ channel are shown in Fig. \ref{fig:bb}.
We can see that for $m_{\chi}\approx 100$ GeV the most stringent
constraint of $\langle\sigma v\rangle$ can reach the thermal scale
of $3\times 10^{-26}$ cm$^3$ s$^{-1}$ for $M_{\rm min}=10^{-6}$
M$_{\odot}$. The future observation of Fermi-LAT can put a strong
constraint on the DM model even for the thermal scenario.
Note that for $b\bar{b}$ channel only the {\it primary}
emission is assumed when deriving the flux limits. If the IC component
is included the constraints will be stronger for relatively heavy DM
mass ($\sim$TeV). Finally we should point out that $b\bar{b}$ channel
is typically not suitable to explain the lepton excesses observed by
PAMELA/Fermi-LAT/HESS, due to both the constraint from PAMELA
$\bar{p}/p$ data \cite{2009PhRvL.102e1101A,2009NuPhB.813....1C}
and the spectral shape required by lepton excesses
\cite{2009PhRvD..79b3512Y}. Here we include the study of $b\bar{b}$
channel is just to show the power of Fermi-LAT to the supersymmetric-like
DM particles.
\section{Neutrinos}
In this section, we discuss the sensitivity of detecting neutrino
signals from clusters. The cluster could be treated as high energy
neutrino point source, and it is possible to be observed at the
on-going large volume neutrino telescopes, such as IceCube. To
suppress the large atmospheric muon background, the neutrino
telescopes usually detect the upward muons induced by muon neutrinos
through interacting with the matter surrounding the detectors.
Therefore, the south pole based detector IceCube is more suitable to
probe the neutrino sources in the northern hemisphere. For the
sample of clusters in Table \ref{table:sample}, Fornax and Centaurus
locating in the southern hemisphere are not good candidates of
Icecube \footnote{The IceCube + DeepCore has the capability to
search the downward neutrinos, but the angular resolution is fairly
bad \cite{2009NIMPA.602....7F}. It is very difficult to distinguish
the high energy neutrino sources from the high atmospheric neutrino
background without powerful angular resolution.}. AWM 7 with
declination of $41^{\circ}35'$ and Coma with declination of
$27^{\circ}59'$ are suitable for IceCube, but such two clusters are
more distant away from us than other clusters. Taking into account
the location, mass and distance, we find M49 with declination of
$8^{\circ}00'$ is better to be detected than others.
Similar to the {\it primary} $\gamma$-ray flux, the neutrino flux
from DM annihilation in the cluster can also be calculated
according to Eq. (\ref{phi}), by replacing $\left.\frac{{\rm
d}N}{{\rm d}E} \right|_{\gamma}$ with $\left.\frac{{\rm d}N}{{\rm
d}E}\right|_{\nu}$. In the following we will mainly discuss two DM
annihilation channels, $\mu^+\mu^-$ and
$\mu^+\mu^-+\nu_{\mu}\bar{\nu}_{\mu}$. The $b\bar{b}$ channel as
discussed in the previous section will also be mentioned. However,
as we will see below, it gives negligible neutrino signals. We use
the PYTHIA \cite{2006JHEP...05..026S} to simulate the initial
neutrino spectra from decay of annihilation final states. We
further assume the neutrino flavor distribution is $1:1:1$ at the
Earth due to vacuum oscillation during the propagation.
The through-going upward muon rate at the detector can be
calculated as
\begin{equation}
\frac{dN_{\mu}}{dE_{\mu}}=\int d\Omega\int_{E_{\mu}}^{m_\chi}
dE_{\nu_{\mu}} \frac{dN_{\nu_{\mu}}}{dE_{\nu_{\mu}}}\left[
\frac{d\sigma_{\text{CC}}^{\nu p}(E_{\nu_{\mu}},E_{\mu}^0)}{dE_{\mu}^0}
\,n_p+(p\rightarrow n)\right]R(E_\mu)+(\nu\rightarrow\bar{\nu}),
\label{throughmuon}
\end{equation}
where $n_p$ ($n_n$) is the number density of protons (neutrons) in
the matter around the detector, $R(E_{\mu})$ named muon range is the
distance that a muon can travel in matter before its energy drops
below the threshold energy of detector $E_{\rm th}$, which is given by
\begin{equation}
R(E_\mu) = \frac{1}{\rho \beta} \ln {\frac{ \alpha + \beta E_{\mu}}{
\alpha + \beta E_{\rm th}} } ,
\end{equation}
with $\alpha , \beta$ the parameters describing the energy loss of
muons as $dE_{\mu}/dx=-\alpha-\beta E_{\mu}$.
The main background for high energy neutrino detection is the
atmospheric neutrinos. The atmospheric neutrino flux decreases
rapidly as energy increasing. We use a parametrization of atmospheric
neutrino flux \cite{2009PhRvD..80d3514E} which describes the results
of Ref. \cite{2007PhRvD..75d3006H} as
\begin{equation}
\frac{{\rm d} N_{\nu}}{{\rm d} E_\nu {\rm d} \Omega}=N_0
E_\nu^{-2.74}\left(\frac{0.018}{1+0.024 E_\nu |\cos
\theta|}+\frac{0.0069}{1+0.00139 E_\nu |\cos \theta|}\right),
\end{equation}
where $N_0=1.95\times 10^{17}$ ($1.35\times 10^{17}$)
GeV$^{-1}$km$^{-2}$yr$^{-1}$sr$^{-1}$ for $\nu_\mu$
($\bar{\nu}_\mu$) respectively, $\theta$ is the zenith angle.
IceCube could have an angular resolution $\sim 1^\circ$, which is
effective to suppress the diffuse atmospheric neutrino background.
Notice the cluster is not an exact point source, we utilize an
angular resolution of 3.0$^\circ$ (1.5$^\circ$) for M49
(AWM47/Coma) cluster. The number of atmospheric neutrinos for
3$^\circ$ resolution angle is $\sim 4$ times larger than it for
1.5$^\circ$ cone.
\begin{figure}[!htb]
\begin{center}
\includegraphics[width=0.47\columnwidth]{M49_1000_DM_mu.eps}
\includegraphics[width=0.47\columnwidth]{M49_1000_DM_bb.eps}
\caption{Muon flux induced by neutrinos from DM annihilation in M49
cluster, for muon (left) and $b\bar{b}$ (right) final states
respectively. The mass of DM is $1$ TeV and cross section is $10^{-23}$
cm$^{3}$ s$^{-1}$. The results of smooth halo (green) and smooth halo
together with subhalos with two values of $M_{\rm min}$, $10^{-1}$ M$_{\odot}$
(blue) and $10^{-6}$ M$_{\odot}$ (red) are shown. In the left panel, the
solid lines denote $\mu^+ \mu^-$ channel, and dashed lines denote
$\mu^+ \mu^-$ plus $\nu_\mu\bar{\nu}_\mu$ with equal branching ratios.
Also shown is muon flux induced by atmospheric muon neutrino background
in 3$^\circ$ cone.
\label{fig:M49spec}}
\end{center}
\end{figure}
In Fig. \ref{fig:M49spec}, we show the through-going muon flux
induced by DM annihilation in M49 cluster. Similar as in Fig.
\ref{fig:spec} we adopt $m_\chi=1$ TeV, $\langle\sigma
v\rangle=10^{-23}$ cm$^3$ s$^{-1}$, and annihilation channels are
$\mu^+\mu^-$ (left solid), $\mu^+\mu^-+\nu_\mu \bar{\nu}_\mu$
($B_{\mu}=B_{\nu}=0.5$, left dashed) and $b \bar{b}$ (right). For
$b\bar{b}$ channel, the neutrinos are produced through decay of
hadrons induced by $b \bar{b}$ hadronization. It is shown that the
muon spectrum of such channel is very soft and difficult to detect
given the high level of background. The case for $\mu^+\mu^-$
channel seems better but the signal is still very weak. Even for
$M_{\rm min}=10^{-6}$ M$_{\odot}$, the muon flux from DM
annihilation is $\sim 100$ times smaller than the background in
energy range $(200,\,800)$ GeV. Only for $\mu^+\mu^-+\nu_\mu
\bar{\nu}_\mu$ the situation is better. The reason is that
monochromatic neutrino spectrum is harder than other channels, and
is easier to be detected. Compared with Fig. \ref{fig:spec}, it is
not strange to see that the neutrino detection sensitivity would
be much weaker than the $\gamma$-ray detection.
\begin{figure}[!htb]
\begin{center}
\includegraphics[width=0.47\columnwidth]{M49_mu.eps}
\includegraphics[width=0.47\columnwidth]{M49_nu.eps}
\caption{$\langle\sigma v\rangle$ required to discover neutrinos from DM
annihilation in M49 cluster as a function of DM mass. The left panel
is for $\mu^+ \mu^-$ channel, and the right panel is for $\mu^+\mu^- +
\nu_\mu \bar{\nu}_\mu$ with branching ratios $B_{\mu}=B_{\nu}=0.5$.
Dashed circles are the $3\sigma$ and $5\sigma$ parameters regions which
can fit the PAMELA/Fermi-LAT/HESS data of the CR positrons/electrons
\cite{2010NuPhB.831..178M}. The dot-dashed curves represent the
$\gamma$-ray constraints for $M_{\rm min}=10^{-1}$ and $10^{-6}$
M$_{\odot}$ respectively (see Fig. \ref{fig:mu}). The circles and
$\gamma$-ray constraints in the right panel are scaled
upwards by a factor $2$ due to the branching ratio $B_{\mu}=0.5$.
\label{fig:M49nu}}
\end{center}
\end{figure}
\begin{figure}[!htb]
\begin{center}
\includegraphics[width=0.47\columnwidth]{AWM7_mu.eps}
\includegraphics[width=0.47\columnwidth]{AWM7_nu.eps}
\caption{The same as Fig. \ref{fig:M49nu} but for AWM7 cluster.
\label{fig:AWM7nu}}
\end{center}
\end{figure}
\begin{figure}[!htb]
\begin{center}
\includegraphics[width=0.47\columnwidth]{Coma_mu.eps}
\includegraphics[width=0.47\columnwidth]{Coma_nu.eps}
\caption{The same as Fig. \ref{fig:M49nu} but for Coma cluster.
\label{fig:Comanu}}
\end{center}
\end{figure}
The total muon event rate in a specific energy bin at the detector is
\begin{equation}
N=\int {\rm d}E_\mu \frac{{\rm d}N_\mu}{{\rm d}E_\mu} A_{\rm eff}(E_{\mu},
\theta) \Delta T,
\end{equation}
where $A_{\rm eff}$ is the effective muon detecting area taken
from Ref. \cite{2009APh....31..437G}, $\Delta T$ is the operation time
which is set as $10$ years here. We take the threshold energy to be
$E_{\rm th}\sim$40 GeV, and assume the energy resolution is $\Delta
\log_{10}E=0.35$ \cite{2009NIMPA.602....7F}.
In Fig. \ref{fig:M49nu} we give the IceCube sensitivity of detecting
neutrinos from DM annihilations in M49 for $10$-yr exposure. The left
and right panels show the results for annihilation channels $\mu^+\mu^-$
and $\mu^+\mu^- + \nu_\mu \bar{\nu}_\mu$ respectively. For $\mu^+\mu^-$
channel and $M_{\rm min}=10^{-6}$ M$_{\odot}$, a $5\sigma$ detection
requires $\langle\sigma v\rangle$ as large as $O(10^{-22})$ cm$^3$
s$^{-1}$. We can see such a cross section is much larger than that
required to explain PAMELA/Fermi-LAT/HESS data. For the channel to
equal $\mu^+\mu^- + \nu_\mu \bar{\nu}_\mu$, the sensitivity is $\sim 10$
times better than $\mu^+\mu^-$ channel. Because the neutrino-nucleon
cross section and muon range are approximately proportional to $E_\nu$,
while the atmospheric neutrino background decreases as $\sim E_\nu^{-3}$,
the neutrino telescope is more powerful to detect the high energy neutrinos.
For the same reason, the detector is more sensitive to explore heavy DM.
It is shown that if DM annihilation products have a large branching ratio
to $\nu \bar{\nu}$, IceCube could reach the parameter space to explain
PAMELA/Fermi-LAT/HESS results.
The sensitivities for other two clusters, AWM 7 and Coma, are shown in
Figs. \ref{fig:AWM7nu} and \ref{fig:Comanu} respectively. These two
clusters are much more distant from us than M49. Note for these two
clusters we take $1.5^\circ$ cone, which is enough to include all of
the cluster, to count the atmospheric background. Due to a much lower
level of background and larger masses of AWM 7 and Coma, the sensitivities
are only several tens percent weaker than M49. Because the declinations
of AWM 7 and Coma are larger than M49 which is close to the horizon,
it would be more effective to reject the CR muon background and could
provide clearer detection of signals.
Compared with the $\gamma$-ray sensitivity discussed in Sec. II, the
sensitivity of neutrino detection is relatively poor. If we employ the
$\gamma$-ray constraint of DM annihilation in clusters (e.g., $M_{\rm min}
\approx 10^3$ M$_{\odot}$ for PAMELA/Fermi-LAT/HESS best fit parameters),
there would be little chance to detect the neutrino signals. Only for very
heavy DM ($m_\chi\sim 10$ TeV), the sensitivities of the two ways are
comparable.
\section{Summary and discussion}
In this work we investigate the $\gamma$-ray and neutrino emission
from DM annihilation in clusters of galaxies. A sample of several
nearby clusters is considered. Both the annihilations in the host
halo and substructures are taken into account. For the
annihilation luminosity of substructures we adopt the result of
{\it Aquarius} simulation, and scale it from Milky Way like halo
to the cluster like halo. There is also a component of unresolved
substructures which is not seen due to the limit of resolution of
numerical simulation. For the contribution from the unresolved
substructures we adopt an extrapolation of the luminosity-mass
relation fitted from the resolved subhalos in the simulation. The
minimum mass of the subhalo $M_{\rm min}$ is left to be a free
parameter. The value of $M_{\rm min}$ may catch the information
about free streaming length, and may reflect the generation
history of the DM particle.
For the $\gamma$-ray emission we consider two typical annihilation
channels, quark final states $b\bar{b}$ and lepton final states
$\mu^+\mu^-$. Both the {\it primary} component of photons from
hadron decay and final state radiation, and the {\it secondary} component which
is produced by the IC scattering of DM-induced
electrons/positrons and CMB field are taken into account. The
Fermi-LAT upper limits of the $\gamma$-ray emission from these
clusters are employed to constrain the $m_{\chi}-\langle\sigma
v\rangle$ parameter plane. The constraints depend on the value of
$M_{\rm min}$. For $M_{\rm min}$ down to $\sim 10^{-6}$
M$_{\odot}$, which is expected for the structure formation of
neutralino-like CDM picture, the constraints are $\sim 50$ times
stronger than the case of smooth halo only. Typically for
$m_{\chi}=1$ TeV and $M_{\rm min}=10^{-6}$ M$_{\odot}$ the
strongest constraint on cross section from the cluster sample is
$10^{-24}$ ($10^{-24}$) cm$^3$ s$^{-1}$ for $\mu^+\mu^-$
($b\bar{b}$).
It is of great interest for the $\mu^+\mu^-$ channel which is
proposed to explain the positron and electron excesses reported by
PAMELA, Fermi-LAT and HESS
\cite{2009PhRvL.103c1103B,2010NuPhB.831..178M}. A very large
annihilation cross section (or boost factor) is needed to explain
the data. It is shown that the Fermi-LAT observations about
$\gamma$-rays from galaxy clusters can strongly constrain the
model parameters. If we fix the mass and cross section to the
values explaining the $e^{\pm}$ excesses, the minimum mass of
substructures $M_{\rm min}$ is constrained to be larger than
$10^2-10^3$ M$_{\odot}$. Such a large value of halo mass means a
very large free streaming length of DM particle. It may indicate
the nature of DM particles is warm instead of cold
\cite{2001PhRvL..86..954L,2009PhRvL.103r1302P}.
Finally we calculate the sensitivity of detecting neutrinos from
these clusters by the IceCube detector. It is shown to detect
neutrinos would be much more difficult than $\gamma$-rays. For
$b\bar{b}$ final state the sensitivity is extremely poor due to
the neutrino spectrum from $b\bar{b}$ hadronization is soft and
suffers from a very high atmospheric background. The case becomes
better for $\mu^+\mu^-$ final state. However, the signal is still
very weak. For example, even for $M_{\rm min}=10^{-6}$
M$_{\odot}$, a $5\sigma$ detection with $\sim 10$-yr exposure
requires $\langle\sigma v\rangle$ as large as $O(10^{-22})$ cm$^3$
s$^{-1}$. This result does not have enough capability to explore
the PAMELA/Fermi-LAT/HESS favored parameter region. If we consider
the model with a large fraction of line neutrino emission the
detectability would be much better (e.g. improved by an order of
magnitude). However, compared with $\gamma$-rays the constraint
from neutrinos is still some weaker. Only for very heavy DM
($m_\chi\sim 10$ TeV), the sensitivity of neutrino detection can
be comparable with $\gamma$-rays.
\acknowledgments
This work was supported in part by the Natural Sciences Foundation
of China under the grant Nos. 10773011, 10775001, 10635030, the
973 project under grant 2010CB833000, and the trans-century fund
of Chinese Ministry of Education.
|
\section{Introduction}
\label{sec:intro}
How is the stellar content of galaxies assembled? Two extreme,
contrasting models for the assembly of stars in galaxies
have been considered over the years with no conclusive evidence as
yet as to which mode dominates in which systems at which times. At one
extreme, we can treat the process as totally dissipational with
regard to energy; gas flows in from the virial radius, radiating away
the kinetic and thermal energy it acquires while descending into the
deep potential well of the dark matter halo. Once in place, the gas is
transformed into stars `in situ'--in approximately the regions in
which we see stars in fully-formed galaxies today. The other extreme
model postulates that stars are formed in smaller stellar systems far
outside the
effective radius of the ultimate galaxy. From there, they lose orbital
energy via dynamical friction and lose their potential energy only
by heating or expelling other matter. These stars could be considered
`accreted,' whether by minor or major mergers. The balance between
these two processes of galaxy formation (both of which surely occur) is unknown;
determining that balance should help us to unravel some
apparent paradoxes in the standard \ensuremath{\Lambda\mathrm{CDM}}{} model of cosmology.
Although the \ensuremath{\Lambda\mathrm{CDM}}{} model of cosmology has been very successful explaining
large scale observations, such as the cosmic microwave background and
the large scale structure of galaxies, it has not enjoyed as much
success on smaller, galactic scales. One notable discrepancy between
simply-modeled \ensuremath{\Lambda\mathrm{CDM}}{} predictions and observations is the difference
between the predicted and observed dark matter content in the centers
of galaxies. Many N-body simulations of dark matter particles have
been performed, and they show that cold dark matter will
collapse into self-similar halos with a central density cusp
\citebut{Fukushige97, NFW97, Moore99, Diemand05}{Springel08}. The
steepness of the
cusp may vary from $r^{-1}$ in the case of the NFW profile \citep{NFW97}
to as steep as $r^{-1.5}$ \citep{Moore99}. Other studies
\citep{Subramanian00, Ricotti03} have shown a
range of central slope indices, $1<\alpha<1.5$ for $r^{-\alpha}$,
depending on time, mass,
and environment.
Cold dark matter simulations uniformly predict that dark matter halos
should be universally cuspy, but observations have yet unambiguously to
find these cusps. Indeed, observations of gravitational lensing, stellar
velocity dispersions, and gas dynamics suggest inner dark matter
density profiles are cored ($r^\alpha$, $\alpha > -1$) instead of
cuspy \citebut{FloresPrimack94, Romanowsky03, Swaters03, Gentile04,
SandTreu04, Simon05, Cappellari06, Gilmore07, Gentile07, deBlok08,
Oh08, Napolitano09}{Rhee04, Spekkens05, Valenzuela07}. For the Milky
Way, it has been shown that within the errors of the microlensing
observations, stars alone can more than account for the total mass
density of the galaxy, leaving little room
for a dark matter component in the center \citep{Binney01}.
Acceptance of both the dark matter simulations and the observations of
dark matter
density profiles leaves us with an apparent discrepancy. One way to solve this
discrepancy is to abandon \ensuremath{\Lambda\mathrm{CDM}}{}. For example, the self-interactions
of warm dark matter will smooth central density cusps
\citep{SpergelSteinhardt00, Bode01}. Another solution to
this discrepancy may lie in a misunderstanding of galaxy formation and
the assembly of stellar material at the center of dark matter
halos. The addition of baryons to dark matter halos can have a profound
effect on the dark matter profile of a galaxy. One such effect is the
adiabatic contraction of dark matter by the slow addition of baryons to
the center of the potential well \citep{Blumenthal86}. Adiabatic
contraction has been observed in simulations with both dark matter and
baryons \citep{Navarro91, Navarro94, Jesseit02, Abadi03, Gnedin04}. This process
conserves the adiabatic invariants of the dark matter orbits. With
regard to energy, however, it is a
dissipational process, as the orbital energy of the infalling baryonic
material is radiated away and lost from the system. Because the dark
matter density increases in the center of the galaxy under
adiabatic contraction, the discrepancy between simulations and
observations is only made worse, in the sense that the inner slope
would be steeper than the NFW slope. For example, if the ultimate mass
profile is ``isothermal,'' ($\rho_{\mathrm{tot}}(r)\propto r^{-2}$), then a dark
matter profile with an initial slope index $1<\alpha<1.5$,
($r^{-\alpha}$), would have a post-adiabatic-contraction index of
$1.67<\alpha<1.8$.
Processes which lower the dark matter density in the central regions
of galaxies have also been explored. These processes are dissipationless; energy
from stellar baryons is transferred to the dark matter, heating it and lowering
the central dark matter density. Such processes include interactions
of the dark matter with a
stellar bar \citep{Weinberg02, HolleyBocklemann05, McMillanDehnen05},
baryon energy feedback from AGN\citep{Peirani08}, decay of binary
black hole orbits after galaxies merge \citep{Milosavljevic01},
scattering of dark matter particles by gravitational heating from
infalling subhalos \citep{MaBoylanKolchin04}, and dynamical friction
of stellar/dark matter clumps against the smooth background dark
matter halo \citep{El-Zant01, El-Zant04, Tonini06, Romano-Diaz08,
RomanoDiaz09, Jardel09}. In order to be successful, these methods
must retain the strong central stellar concentration observed in large
galaxies.
\citet{El-Zant01} propose erasing the dark matter cusp via
dynamical friction of incoming stellar/dark matter clumps against the
dark matter background. \citet{Romano-Diaz08} tested this
hypothesis with dark matter/baryon N-body/SPH simulations and found
that the introduction of baryons can flatten the dark matter cores in
the inner 3 kpc. \citet{Romano-Diaz08} claim that this cusp-flattening
is not seen in other baryon/DM simulations because of a low numerical
resolution and a focus on early, dissipational galaxy
formation. Recent simulations \citep{Naab07, Abadi09, Johansson09}
including both dark matter
and baryons have shown similar departures from the standard adiabatic
contraction model. In their cosmological simulations for building up
elliptical galaxies,
\citet{Johansson09} note that their results show reasonably low dark
matter fractions in the inner $10\ \mathrm{kpc}$ and that the assembly
of their elliptical galaxies at late times is dominated by accretion of stellar
lumps, not gas, which as noted above, tend to reduce the dark matter
concentration. Similarly,
accretion at late times has been invoked to
erase the dark matter cusp formed by early adiabatic contraction, bringing
dark matter halos to a universal (e.g. NFW) profile
\citep{LoebPeebles03, Gao04}.
Dynamical friction of stellar lumps
against the dark
matter halo represents a dissipationless method of stellar build-up in
which the orbital energy of the infalling stars is transferred to the
dark matter, thereby heating it and driving it out of the central
parts of the galaxies. This is in direct contrast to the case
of adiabatic contraction, which arises from a dissipational build-up of
stellar material. Although both processes undoubtedly take place, the
fully dissipational (adiabatic contraction) and fully
dissipationless (dynamical friction) scenarios for galaxy
formation represent the two extrema. Since these two
processes have opposite effects on the central dark matter content of
galaxies, the balance between them will determine the present-day dark
matter density profiles in galaxies.
In this paper we explore the physics of galaxy assembly, presenting
two toy models for the two extremes of galaxy
formation constrained to produce the same observed final stellar
distributions: one model for the fully dissipational build-up of
stellar matter with star
formation occurring in-situ and one model for the fully
dissipationless build-up of stellar matter with stars added by
accretion. The models are described in \S\ref{sec:models}. We focus
on the structure of giant elliptical galaxies whose light profiles are
well-described by Sersic models. Taking the observed stellar
profile of the galaxy as given, but leaving the stellar mass-to-light
ratio, \ensuremath{M_*/L}{}, as a free
parameter, we model the assembly of the
stellar component via the two extreme formation methods. Then, by comparing
the properties of the galaxies formed by these two methods to
observations, we can begin to determine which method of assembling stars
dominates and help to resolve the discrepancies between the simulated
and observed dark
matter density profiles.
Throughout this paper, we adopt the \ensuremath{\Lambda\mathrm{CDM}}{} cosmology with $H_0 = 70\
\ensuremath{\mathrm{h}_{70}}\ \ensuremath{\mathrm{km\ s}^{-1}}$, $\Omega_{\mathrm{baryon}}/\Omega_{\mathrm{matter}} =
0.17$, $\Omega_{\mathrm{matter}} = 0.26$, and $\Omega_{\Lambda}=0.74$.
\section{Models}
\label{sec:models}
In this section, we describe the two toy models of galaxy
formation used in this paper. The first assumes that the orbital
energy of the in-falling baryons is deposited entirely in the dark matter halo,
while the second assumes energy is radiated away, leaving the galaxy-halo
system.
In both models, we assume that both the dark matter and the baryonic
matter follow circular orbits. For simplicity, we also assume the initial
conditions (before the formation of the galaxy) for both the dark
matter and the baryons are NFW density profiles, $\rho \propto
[(r/r_{\mathrm{NFW}})(1+r/r_{\mathrm{NFW}})^2]^{-1}$, with the same
concentrations. The ratio of baryon to dark matter density is equal to
the universal fraction, which we assume to be
$\Omega_{\mathrm{baryon}}/\Omega_{\mathrm{dark matter}} = 0.20$. In
reality, the baryons will have a broader distribution than the dark
matter, because the baryons are coupled to the radiation
background. However, the additional thermal energy of the baryons
corresponds to a velocity less than $10\ \ensuremath{\mathrm{km\ s}^{-1}}$, which is negligible
for halos more massive than $\sim10^8\ \ensuremath{M_\odot}$. In this study, we
restrict ourselves to large halos, and can safely assume the baryons
are initially distributed identically to the dark matter. We have
verified this is a valid assumption by placing all the baryons
initially outside the virial radius where they have negligible binding
energy. This does not affect any of the results in this paper,
altering the dark matter mass with $5\ \ensuremath{{R_{\mathrm{e}}}}$ by a few parts in $10^3$.
The final condition to which the models evolve is an elliptical galaxy
with a predetermined stellar luminosity profile in the center of a
dark matter halo. In both the dissipational and dissipationless
models, we assume that the final galaxies contain no gas; all the
baryons are accounted for in stars. This assumption is justified
because the gas mass in elliptical
galaxies today is usually only a few percent of the total baryon
mass in these galaxies \citep{Georgakakis01}.
For both models, the dark matter and stellar matter profiles are
discretized into spherical shells such that the radius of a shell is
at least two orders of magnitude larger than its width, and each shell
is taken to be of homogeneous density. In order to
form the galaxy, shells of stellar
matter are moved inwards. Each time a shell of stellar matter passes through a
shell of dark matter, the stars' orbital energy is either deposited in the
dark matter shell (dissipationless model) causing the dark matter shell to
move outwards, or the stars' orbital and thermal energy is
radiated away causing the dark matter shell's average radius to decrease as it
undergoes adiabatic contraction. Since truly spherical shells would
not suffer from dynamical
friction, we are in fact considering a shell composed of stellar
clumps, all of which (at any given time) have the same energy and
angular momentum per unit mass, but whose angular momentum vectors are
oriented randomly, giving the shell a total angular momentum of
zero. Additionally, we assume that the stellar clumps do not contain
any dark matter, as they formed far from the center of the dark matter
halo, where the density is low. In reality, these clumps will contain
some dark matter, much of which will be stripped before the stellar
clump merges with the galaxy. Any remaining dark matter will undergo
dynamical friction against the background dark matter.
Because the system is spherically symmetric, a given dark matter shell
is only affected by matter that crosses through it, but not by the
rearrangement of matter interior or exterior to it. Therefore a
procedure that describes a single shell crossing can be repeated many
times for many shells, until the desired galaxy has been assembled.
\subsection{Dissipationless Model: Dynamical Friction}
\label{ssec:modelDF}
In the dissipationless model, all the orbital energy lost by the
stars in forming the galaxy is deposited in the dark matter halo. In
this case, the galaxy is built up by `dry,' dissipationless
mergers.
We imagine constructing the Sersic profile of the stars in a shell by
shell fashion, sequentially moving each shell of stars from a large
radius to its final position, in such a way that two stellar shells
never cross. Therefore, the problem can first be idealized as moving
one infinitely thin stellar shell from a large radius $R_{\mathrm{i}}$
to a smaller, final radius $\ensuremath{R_{\mathrm{fin}}}$. As the stellar shell moves, it
passes through a dark matter distribution with a mass distribution
$M_{\mathrm{dark}}(r)$, to which the stellar shell gives its orbital
energy. The dark matter can be subdivided into a series of mass shells
of thickness $\Delta R_k$, such that $\sum_k \Delta R_k = \ensuremath{R_{\mathrm{int}}} -
\ensuremath{R_{\mathrm{fin}}}$. Therefore, we only need to compute the effects of one stellar
shell passing through one dark matter shell uniformly distributed
between $R_{k}$ and $R_{k+1}$, with thickness $\Delta R_k$, and then
execute a double sum over the dark matter shells and all the stellar
shells. A depiction of this process is shown in Figure
\ref{fig:cartoon}. In the figure, the sum over dark matter shells is
indexed by $k$, while the sum over the stellar shells is indexed by
$i$.
\middfig{\figcartoon}
The total energy lost by a single stellar shell as it moves from $R_k$ to
$R_{k+1}$ can be calculated by taking the difference between the
energy change of the stellar shell as it moves from a large initial
radius, $\ensuremath{R_{\mathrm{int}}}$, with zero potential and zero kinetic energy, to $R_{k+1}$
and the energy change as it moves from $\ensuremath{R_{\mathrm{int}}}$ to $R_{k}$. The changes in
kinetic and potential
energy per unit mass of the stellar shell as it moves from $\ensuremath{R_{\mathrm{int}}}$ to $R_k$ are
\begin{eqnarray}
\label{eq:deltaEstars}
\Delta T_k &=& \frac{1}{2}\frac{\ensuremath{\mathrm{G}} M(R_k)}{R_k}\,,\qquad\mathrm{and} \\
\Delta W_k &=& -\frac{\ensuremath{\mathrm{G}} M(R_k)}{R_k} - \int_{R_k}^{\ensuremath{R_{\mathrm{int}}}}\frac{\ensuremath{\mathrm{G}}
\, \ensuremath{\mathrm{d}} M_{\mathrm{dark}}(r)}{r} \, ,
\end{eqnarray}
where $M(r)$ includes both the dark matter and any stellar matter
interior to $r$. By assumption, there is no stellar matter between
$\ensuremath{R_{\mathrm{int}}}$ and $R_k$, so the integral in the potential energy depends only on
$M_{\mathrm{dark}}$. The change
in total energy of the stellar shell as it moves from $R_k$ to
$R_{k+1}$ is thus
\begin{eqnarray}
\label{eq:deltaEdeltaRk}
\Delta E_k &=& \frac{\ensuremath{\mathrm{G}} M_{\mathrm{dark}}(R_k)}{2 R_k} - \frac{\ensuremath{\mathrm{G}}
M_{\mathrm{dark}}(R_{k+1})}{2 R_{k+1}} + \\\nonumber
&&\, \frac{\ensuremath{\mathrm{G}}
M_{\mathrm{star}}(R_k)}{2}\left(\frac{1}{R_{k+1}}-\frac{1}{R_k}\right) +
\\\nonumber
&&\, \int_{R_k}^{R_{k+1}} \frac{\ensuremath{\mathrm{G}} \,\ensuremath{\mathrm{d}} M_{\mathrm{dark}}(r)}{r} \, .
\end{eqnarray}
The stellar shell is made of up of small lumps of stars
that will undergo dynamical friction against the dark matter
background as they move; the energy lost by the stars as
they move from $R_k$ to $R_{k+1}$ must be deposited in the dark
matter. We assume that the exchange of energy between stars and dark
matter is a local process. Therefore, all the energy of the stars is
deposited in the dark matter initially orbiting between $R_k$ and
$R_{k+1}$. As the dark matter gains energy, it's orbit will expand,
and the uniform density dark matter shell between $R_k$ and $R_{k+1}$
will become wider. Since we are assuming the energy exchange is
local, the dark matter orbiting at the inner radius ($R_{k+1}$) will
not move. In
reality, this dark matter is affected by both adiabatic contraction,
since the stars add mass interior to $R_{k+1}$, and dynamical
friction, since the
energy exchange is not purely local; this layer will move, but the
movement will be second order in $\Delta R_k = R_{k} - R_{k+1}$. Therefore,
by conserving energy the only quantity which changes is thickness of
the dark matter shell the stars have moved through and consequently
its mean radius.
In order to calculate the new dark matter layer thickness, we repeat
the procedure above, but this time also account for the energy of the
dark matter shell both before and after the stars move through it. As
above, we assume a spherical dark matter shell of mass $\ensuremath{M_{\mathrm{DM}}}$ and
uniform density, with an inner radius of $R$ (corresponding to
$R_{k+1}$ above), and a thickness $\Delta R << R$. Directly exterior
to the dark matter shell is an infinitesimally thin stellar shell of
mass $\Delta M_*$, at the radius $R+\Delta R$. The kinetic+potential energy
of the two-shell system can be broken into the self-interaction energy
of the dark matter shell, $E_{\mathrm{d}}$, the self-interaction
energy of the stellar shell, $E_{\mathrm{s}}$, and the interaction
energy of the two shells, $E_{\mathrm{ds}}$. To first order in $\Delta
R$, these are given by:
\begin{eqnarray}
\label{eq:EdB}
E_{\mathrm{d}} &=& -\frac{\ensuremath{\mathrm{G}} \ensuremath{M_{\mathrm{DM}}}^2}{4 R}\left(1-\frac{2}{3}\frac{\Delta
R}{R}\right)\, ;\\
\label{eq:EsB}
E_{\mathrm{s}} &=& -\frac{\ensuremath{\mathrm{G}} \Delta M_*^2}{4 R}\left(1-\frac{\Delta
R}{R}\right)\, ;\\
\label{eq:EdsB}
E_{\mathrm{ds}} &=& -\frac{\ensuremath{\mathrm{G}} \Delta M_* \ensuremath{M_{\mathrm{DM}}}}{2 R}\left(1-\frac{\Delta
R}{R}\right)\, .
\end{eqnarray}
The shells are embedded in a spherically symmetric galaxy which also
contributes to the energy. We assume that the mass distribution of the
galaxy remains fixed as the shells interact. This assumption is
exactly true for the mass interior to the shells and the mass far
outside the shells. We define $\ensuremath{M_{\mathrm{int}}}$ to be the total mass (dark
matter $+$ baryons ) interior to $R$. This mass contributes kinetic+potential energy $\ensuremath{E_{\mathrm{int}}}$ to the shells while the
mass external to $R+\Delta R$ contributes a potential energy
$\ensuremath{E_{\mathrm{ext}}}$. These energies are given by
\begin{eqnarray}
\label{eq:Eint}
\ensuremath{E_{\mathrm{int}}} &=& -\frac{\ensuremath{\mathrm{G}} \ensuremath{M_{\mathrm{int}}}}{2R}\bigg[\ensuremath{M_{\mathrm{DM}}}\left(1-\frac{\Delta
R}{2R}\right) + \\\nonumber
&&\qquad\qquad\left.\Delta M_*\left(1-\frac{\Delta
R}{R}\right)\right]\,,\ \quad\mathrm{and}\\
\label{eq:Ext}
\ensuremath{E_{\mathrm{ext}}} &=& -\ensuremath{\mathrm{G}}\left(\Delta M_*+\ensuremath{M_{\mathrm{DM}}}\right) \int_{R+\Delta
R}^{\infty}\frac{\ensuremath{\mathrm{d}}\,M(r)}{r} \,.
\end{eqnarray}
Therefore, the total initial energy of the two shells is given by the sum of
equations (\ref{eq:EdB})-(\ref{eq:Ext}).
We now move the stellar shell through the dark matter shell until
the stellar shell is orbiting at the radius $R$, as in the above example.
This time, however, we will include the changes in the dark matter
distribution in the energy difference. After the move, the dark
matter shell will widen from $\Delta R$ to $\Delta R^{\prime}$.
Because $\Delta R^{\prime}$ is still small compared to $R$, the
density of the dark matter layer is still uniform after the move. The final
total energy of the stellar and dark matter shells after moving the stars
is
\begin{eqnarray}
\label{E:after}
E &=&
-\frac{G}{2R}\bigg[\frac{\ensuremath{M_{\mathrm{DM}}}^2}{2}\left(1-\frac{2}{3}\frac{\Delta
R^{\prime}}{R}\right) + \frac{\Delta M_*^2}{2} + \\\nonumber
&&\qquad \Delta M_*\ensuremath{M_{\mathrm{DM}}}\left(1-\frac{\Delta R^{\prime}}{2R}\right) + \\\nonumber
&&\qquad \ensuremath{M_{\mathrm{int}}}\left(\ensuremath{M_{\mathrm{DM}}}\left(1-\frac{\Delta R^{\prime}}{2R}\right) +
\Delta M_*\right)\bigg] \\\nonumber
&& -\ensuremath{\mathrm{G}} \left(\Delta M_*+\ensuremath{M_{\mathrm{DM}}}\right) \int_{R+\Delta
R}^{\infty}\frac{\ensuremath{\mathrm{d}}\,M(r)}{r} + \Delta \ensuremath{E_{\mathrm{ext}}}\, .
\end{eqnarray}
If we assume that the mass initially exterior to $R+\Delta R$ is
unaffected by the dark matter expanding to $R+\Delta R^{\prime}$, the
change in external energy, $\Delta \ensuremath{E_{\mathrm{ext}}}$, depends only on the mass
initially between $R+\Delta R$ and $R+\Delta R^{\prime}$, and the dark matter that moves beyond $R + \Delta R$. This
assumption is valid to first order in $\Delta R$. Since both $\Delta
R$ and $\Delta R^{\prime}$ are small compared
to $R$, we assume that mass between $R+\Delta R$ and $R+\Delta
R^{\prime}$ is of homogeneous density and a total mass
$\ensuremath{M_{\mathrm{ext}}}$. Therefore,
\begin{equation}
\label{eq:Eextdiff}
\Delta \ensuremath{E_{\mathrm{ext}}} = \frac{\ensuremath{\mathrm{G}} \ensuremath{M_{\mathrm{DM}}} \ensuremath{M_{\mathrm{ext}}}}{4 R}\frac{\Delta R^{\prime}-\Delta
R}{\Delta R^{\prime}}\,.
\end{equation}
Substituting this into equation \ref{E:after} and taking the
difference between the total energy of the two shells before and after
the move yields
\begin{eqnarray}
\label{eq:Ediff}
\Delta E &=& \frac{\ensuremath{\mathrm{G}}}{12 R^2}\Bigg\{\left(\Delta R^{\prime}-\Delta
R\right)\ensuremath{M_{\mathrm{DM}}}\bigg[2\ensuremath{M_{\mathrm{DM}}} + \\\nonumber
&&\qquad\qquad\qquad 3\left(\ensuremath{M_{\mathrm{int}}}+\Delta M_* +\frac{R}{\Delta R^{\prime}}
\ensuremath{M_{\mathrm{ext}}}\right)\bigg]- \\\nonumber
&& \qquad\qquad 3\Delta R \Delta M_*\left(\ensuremath{M_{\mathrm{DM}}} + 2 \ensuremath{M_{\mathrm{int}}}+\Delta M_*\right)\Bigg\}\,.
\end{eqnarray}
This can be solved numerically for $\Delta R^{\prime}$, keeping in
mind that $\ensuremath{M_{\mathrm{ext}}}$ is a
function of $\Delta R^{\prime}$. In the limit that $\ensuremath{M_{\mathrm{ext}}}\to 0$,
equation \ref{eq:Ediff} can be solved analytically for $\Delta R^{\prime}$:
\begin{eqnarray}
\frac{\Delta R^{\prime}}{\Delta R} &=& \frac{1/2 \ensuremath{M_{\mathrm{int}}} + 1/3 \ensuremath{M_{\mathrm{DM}}} +
3\Delta M_*
}{1/2\ensuremath{M_{\mathrm{int}}}+1/3\ensuremath{M_{\mathrm{DM}}}+3/2\Delta M_*}+\\\nonumber
&&\quad\frac{(\Delta M_*/\ensuremath{M_{\mathrm{DM}}})(\ensuremath{M_{\mathrm{int}}}+1/2\Delta
M_*)}{1/2\ensuremath{M_{\mathrm{int}}}+1/3\ensuremath{M_{\mathrm{DM}}}+3/2\Delta M_*} \, .
\end{eqnarray}
In the limit that $\Delta M_* << \ensuremath{M_{\mathrm{DM}}}$ the ratio of widths goes to $1$
as expected. This procedure can be
easily scaled up to a series of interleaved dark matter and stellar
shells. As the stellar shells move inwards to form a galaxy, they
expand each dark matter shell they cross, thereby slowly moving
dark matter outward.
In the example above, all the energy from the stars is deposited in
the dark matter layer which the stars cross. In our numerical
calculations, the stars deposit their energy in a set of layers
surrounding the layer they cross. The width of this set is
proportional to the current radius and the amount of energy deposited
in each layer is proportional to the mass of that layer. This approximates
the wake created by infalling stellar material, which is responsible
for the forces causing dynamical friction \citep{Weinberg86}. The size
of the wake scales
as $\ensuremath{\mathrm{G}} \Delta M_*/\sigma_{\mathrm{dm}}^2 \approx (\Delta M_*/\ensuremath{M_{\mathrm{int}}})
R$. We assume that
the stars are added by a series of minor mergers, in which each added
stellar shell of mass $\Delta M_*$ constitutes a minor merger. The
mass ratio of such
mergers is approximately $(\Delta M_*/\ensuremath{M_{\mathrm{int}}})\approx 1/10$, thus setting the
size of the wake.
In a more accurate treatment, there would also be
a diffusive term; each dark matter shell would spread out in radius as
it was on average moved outwards when passed by a stellar
shell.
In the calculations above, we assume no stellar
shells cross each other. This is equivalent to assuming all the
infalling material remains on spherical orbits. This is not true in
reality, especially since galaxies are not spherical but often
triaxial systems, which do not allow purely circular orbits as assumed
above. If
we allowed for triaxial systems in our models and allowed for the
accretion of material on radial orbits, the infalling stars would
deposit energy interior and
exterior to their final mean orbital radius. This would have an effect
on the final dark matter density profile, but the effect would depend
on the fraction of energy deposited interior and exterior to the final
orbital radius. If all the energy is deposited interior to the final
orbital radius, then more dark matter would be displaced from the
center, leading to a lower central dark matter density. The opposite
is true if the energy is deposited outside the final orbital
radius.
Additionally, if we relax the requirement that no stellar
shells cross, we must take into account energy deposited in the stars,
not just the dark matter. This will expand the stellar orbits, in the
same way the dark matter orbits are expanded, and lower the stellar
density in the center of the galaxy in the same way the dark matter
density is lowered. Therefore, in order to make a galaxy with a given
stellar density, we would first have to make a more
concentrated stellar system, and then add stellar clumps which would
undergo dynamical friction against the highly concentrated stellar
system, thereby lowering the central stellar density to the
desired value. Indeed, observations have
been made of highly concentrated stellar systems at higher redshift
\citep{vanDokkum08, Cappellari09}. In order to make highly
concentrated stellar systems, the galaxies would have to undergo an
early period of dissipational formation. This would also increase the
central dark matter density before the onset of dissipationless
formation, making
the final dark matter density dependent on the relative importance of
dissipational and dissipationless formation mechanisms. Simulations
show that early type
galaxy formation can be divided into two phases, an initial
dissipational formation of a centrally concentrated system, followed
by accretion via `dry' mergers of additional stellar material
\citep{Naab07, Naab09, Cook09}. If we allowed stellar shell crossing, we would have
to take into account the two phase growth of galaxies and combine the
dissipational and dissipationless models into a single model. This
work attempts to determine the relative importance of the
dissipational and dissipationless formation mechanisms; we are only
concerned with the two extreme formation models, which can be easily
modeled assuming spherical galaxies and infalling material on circular
orbits. The effects of triaxiality and radial orbits would require
combining the dissipational and dissipationless formation mechanisms,
which is left for future work.
Of course, dynamical friction on incoming stellar clumps is
intrinsically a three-dimensional process
\citep{TremaineWeinberg84, Aubert06, Aubert07}, and the treatment in
spherical shells
above is
not intended to accurately mimic the actual assembly of a galaxy via dynamical
friction. Rather, the adopted model is designed to be correct with
respect to the total energy deposited in the dark matter.
In the numerical calculations presented below, the total binding
energy is conserved. As the width of the stellar shells decreases, the
calculations conserve energy to approximately one part in $10^5$ of
the binding energy
of the stars in the final galaxy.
\subsection{Dissipational Model: Adiabatic Contraction}
\label{ssec:modelAC}
For the dissipational build-up of galaxies we present the following
picture: baryons in the form of gas slowly fall into the centers of
dark matter potential wells, where the gas condenses to form stars.
In this case, the gas radiates away its orbital and thermal energy as it
falls inwards, so the total energy of the system is not
conserved. However, as long as the gas is
accreted into the center of the galaxy on a timescale that is long compared
to the local dynamical time, the adiabatic invariants of the dark matter orbits
will be conserved.
In our model for adiabatic contraction, we follow the
prescription of \citet{Blumenthal86} and again assume circular orbits
for the dark matter and a spherically symmetric mass
distribution. Instead of energy being conserved, the
adiabatic invariants of the dark matter orbits are conserved. For
periodic orbits, $\oint\,p\mathrm{d}q$ is an adiabatic invariant,
where $q$ is a coordinate and $p$ is its conjugate momentum. For a
particle in a circular orbit at a radius $r$ around a spherical mass
distribution $M(r)$, we take the conjugate momentum to be the angular
momentum and its corresponding coordinate, the angular position. The
adiabatic invariant is then:
\begin{equation}
\label{eq:adiabatInvariant}
J^2 = \Big(\oint\,\sqrt{M(r)r}\mathrm{d}\theta\Big) \propto rM(r)\,.
\end{equation}
Therefore, if the
mass interior to the orbit, $M(r)$, increases, the orbital radius must
decrease. Following the same set-up as in the previous model, we start
with a dark matter shell of constant density and mass $\ensuremath{M_{\mathrm{DM}}}$ directly
interior to an infinitesimally thin shell of stars/gas of mass
$\Delta M_*$. After the baryons move interior to the dark matter layer,
the inner radius of the dark matter shell becomes $R^{\prime}= R \, \ensuremath{M_{\mathrm{int}}} / (
\ensuremath{M_{\mathrm{int}}} + \Delta M_* )$ and the width of the layer becomes
\begin{equation}
\label{eq:adContractWidth}
\Delta R^{\prime} = \frac{(\ensuremath{M_{\mathrm{int}}}+\ensuremath{M_{\mathrm{DM}}})(R+\Delta R)}
{\ensuremath{M_{\mathrm{int}}}+\ensuremath{M_{\mathrm{DM}}}+\Delta M_*}
- R^{\prime}\, .
\end{equation}
Overall, the dark matter shell moves inwards and becomes thicker or
thinner depending on the mass ratios of the dark matter shell, the
baryon shell, and the mass interior to both. As in the previous model,
two shells only interact when their orbits cross, so this prescription
for interchanging two
shells can be scaled up to many shells.
If the constraint of circular orbits is removed, the adiabatic
invariant is no longer $r M(r)$. Using N-body simulations,
\citet{Gnedin04} show that $M(\bar{r}) r$, where $\bar{r}$ is the
orbit-averaged position, is a good proxy for the adiabatic
invariant. If this quantity is conserved for isotropic orbits, the
prescription for adiabatic contraction remains the same. However, the
prescription above will overestimate the amount of adiabatic
contraction if the orbits are radially biased \citep{Gnedin04}.
\subsection{Models Fit to Example of Massive Galaxy}
\label{ssec:example}
The difference in the dark matter density in the central regions of
a galaxy for each model is clearly shown by the comparison of the velocity
profiles of model galaxies. In the top panel of Figure \ref{fig:vcirc},
the circular
velocity curves for both models are compared to the velocity curve
produced by the stars alone. The models shown in Figure
\ref{fig:vcirc} are taken to fit very massive galaxies and we have
adjusted \ensuremath{M_*/L}{} such that the central velocity dispersions of both
models are the same. The central velocity dispersions are computed
assuming the stars are on isotropic orbits. In both cases, the stars
dominate in the
very central regions, but the differences in dark matter significantly
affect both curves. The lower panel of Figure \ref{fig:vcirc} shows the ratio
of the dark matter to stellar matter projected densities as a function
of radius for both models. In reality, since galaxies certainly form by a
combination of dissipational and dissipationless methods, the velocity
curves and the dark matter to stellar matter density ratios will fall
somewhere in between the two extreme toy models examined here.
\middfig{\figvcirc}
In order to compute the two models, we require a set of input
parameters describing the initial matter distribution and the final
stellar distribution. If we assume the initial distribution of both
components fit a single NFW profile, then the two required input
parameters are the total mass of the halo and the initial
concentration. Assuming that the galaxy is
spherically symmetric, the final stellar mass
distribution is completely described by the surface brightness profile
of the galaxy
and a stellar mass-light ratio. The luminosity profile can be
directly observed and \ensuremath{M_*/L}{} can be determined from observations of the
central velocity dispersions of
galaxies. Since stars dominate the mass in the central regions, they
also are the dominant contribution to the central velocity
dispersion. Thus, the
observational input parameters for these models are the total
luminosity of the galaxy, the luminosity profile (in the case of
Sersic profiles, the necessary terms are the Sersic index, $n$, and the
half-light (effective) radius, $\ensuremath{{R_{\mathrm{e}}}}$), and the central velocity
dispersion. The free
parameters are the total halo mass, the initial NFW concentration, and
the stellar mass-to-light ratio (which we assume to be constant
throughout a given system). Therefore, for a given galaxy, the
two models will by construction have the same luminosity and central
velocity dispersion, but
the total halo masses and the mass-to-light ratios will be different,
and in some cases, outside the bounds set by other
observational and
model constraints. Table \ref{table:models} gives the parameters for
the dissipational and dissipationless models which best fit the strong
lensing cluster MS-2137-23 discussed in \S\ref{ssec:strongLens}.
\subsection{Minimum Radius for Galaxies Formed by Dissipationless
Mergers}
\label{ssec:minRad}
In order for a galaxy to form by purely dissipationless processes, the
incoming stars must deposit their orbital energy in the dark matter
halo. Therefore, the dark matter halo must initially have sufficient
\emph{binding} energy to give to the infalling stars. The stars
in the final galaxy can have no more binding energy than the initial
dark matter halo. This sets a lower limit on the size of a galaxy
formed by purely dissipationless processes. A galaxy formed by
dissipationless accretion would approach the minimum size and would
have very little dark matter in the center, which is in agreement with
observations mentioned in \S\ref{sec:intro}. However, if
dissipational processes play a dominant role in
galaxy formation, there is no minimum size for galaxies, as the
infalling baryons can dissipate all of their orbital and thermal energy.
Given a cluster mass, concentration, and final galaxy stellar mass,
and Sersic index,
a lower limit on the galaxy's effective radius, $\ensuremath{{R_{\mathrm{e}}}}$, can be obtained by
setting the change in binding energy of the dark matter halo equal to change
in binding energy of the stars and baryons in the cluster. For
galaxies with $L\approx 10^{11}\ \ensuremath{L_\odot}$ in the $K$-band, the minimum
effective radius for a dissipationlessly formed galaxy yields a
reasonable minimum for the effective radius of observed early type
galaxies. \citep{Bernardi03I}.
We should note that there is another limiting radius for the formation
of accreted halos. A satellite may be destroyed by
tidal shocks before it reaches the energetically allowed minimum
radius, thereby depositing stars in the outer region of
the growing galaxy \citep{Kormendy77, Gnedin99, Wetzel09}. This
process is difficult to compute but gives a limiting
radius comparable to the dynamical friction limits computed above. For
the most massive systems (BCGs), the tidal shock limiting radius is
more severe than the dynamical friction one, so we can expect that the
dark matter to stellar matter ratio in such systems will be higher
than in more moderate mass galaxies.
\section{Comparisons to Observations of BCGs}
\label{sec:ComparetoObs}
One test of the dissipational and dissipationless models of galaxy
formation is the build-up of
brightest cluster galaxies (BCGs). BCGs offer a good comparison sample
for these
spherically symmetric test models for several
reasons. First, BCGs almost always sit at or nearly at the center of
their host cluster. Therefore, they are also centered in a dark matter
halo, so there are no contributions to the potential from an
off-center halo, not included in our models. Furthermore,
BCGs represent a uniform sample, so much so that they have been suggested
as standard candles \citep{Postman95}, allowing comparisons to be made
to the entire population instead of individual galaxies. Finally,
BCGs are thought to have been formed by a series of galaxy mergers
during the build-up of clusters \citep{OstrikerHausman77, Nipoti04,
Cooray05}, making them
good candidate systems in which to observe dissipationless galaxy formation.
\subsection{Scaling Relations for Input Parameters}
\label{ssec:scaling}
In order to compare the models to observations, we normalize the
models to $K$-band and $r$-band data of BCGs. Given the luminosity of a
BCG and choosing a constant stellar mass-to-light ratio for the models,
we use empirical relations to derive the cluster and the BCG
properties. From \citet{LinMohr04}, the cluster mass is related to the
observed
$K$-band BCG luminosity by
\begin{equation}
\label{eq:MclLumLin}
\frac{\ensuremath{L_{\mathrm{BCG}}}}{10^{11}\ensuremath{\mathrm{h}_{70}}^{-2}\ \ensuremath{L_\odot}} = \left(4.9 \pm
0.2\right) \left(\frac{\ensuremath{M_{\mathrm{cluster}}}}{10^{14}\ \ensuremath{M_\odot}}\right)^{0.26\pm0.04} \, .
\end{equation}
The luminosities of BCGs are 6-10 times brighter than $\mathrm{L}_*$
in the $K$-band. For galaxies typically in clusters, $M_{K*} =
-24.34$ \citep{LinMohr04}, including the BCG, or about
$1.16\times10^{11}\ \ensuremath{L_\odot}$ in
the $K$-band.
From the cluster mass, the cluster virial radius, $r_{200}$, can be
calculated assuming the critical density $\rho_{\mathrm{cr}} = 1.36\ \ensuremath{\mathrm{h}_{70}}^2
\times10^{11}\ \ensuremath{M_\odot}\ \mathrm{Mpc}^{-3}$. Simulations have shown that
the
dark matter halo concentration, $c = r_{\mathrm{200}}/r_{\mathrm{NFW}}$, scales
approximately with the
mass as \citep{Neto07}
\begin{equation}
\label{eq:concentration}
c = 4.67 \left(\frac{\ensuremath{M_{\mathrm{cluster}}}}{10^{14}\ \ensuremath{M_\odot}}\right)^{-0.11}\, .
\end{equation}
Equations \ref{eq:MclLumLin} and \ref{eq:concentration} set the
initial conditions for the models.
The properties of the stellar component of the BCG can also be derived
from \ensuremath{L_{\mathrm{BCG}}}. We assume that the BCGs are well-modeled by a single
Sersic profile \citep[$\mathrm{I}(R)\sim
\exp(R^{1/n}),$][]{Sersic69}, ignoring the ICL
and outer components of the BCG \citep[see][]{Gonzalez05}. The
two-dimensional surface
brightness profile defined by Sersic can be deprojected numerically into a
three-dimensional luminosity density profile, assuming the galaxy is
spherically symmetric. This numerical deprojection is well-approximated by the
analytic formula \citep{LimaNeto99}:
\begin{eqnarray}
\label{eq:Sersicdensity}
\rho_*(r) &\propto& (r/\ensuremath{{R_{\mathrm{e}}}})^ {1-1.188/(2n)+0.22/(4n^2)} \nonumber\\
&&\exp\left((0.327-2n)(r/\ensuremath{{R_{\mathrm{e}}}})^{1/n}\right)\,,
\end{eqnarray}
where $n$ is the Sersic index ($n=4$ for a de Vaucouleurs profile),
and \ensuremath{{R_{\mathrm{e}}}}{} is the half-light radius of the
surface brightness profile. Observations show that the Sersic
properties of BCGs are correlated with the galaxy's luminosity. Using data from
\citet{LinMohr04} and \citet{Graham96}, the luminosity can be related
to the half-light radius and the Sersic index by
\begin{eqnarray}
\label{eq:LReffn}
\log\ensuremath{{R_{\mathrm{e}}}} &=& -10.30 +
1.01\log\left(\frac{\ensuremath{L_{\mathrm{BCG}}}}{\ensuremath{L_\odot}}\right) \quad \mathrm{and}\\
n &=& 2.9\log\ensuremath{{R_{\mathrm{e}}}}+1.98\, ,
\end{eqnarray}
which are in good agreement with scaling relations from
\citet{ValeOstriker08}, \citet{Bernardi07} and \citet{Desroches07}.
Also modeled in each galaxy is a central supermassive black hole,
which adds a minor correction to the velocity dispersion of the
galaxy. The black hole mass is determined from the galaxy luminosity
by the relation \citep{Graham07}:
\begin{eqnarray}
\label{MbhLumGraham07}
&&\log\left(\frac{\ensuremath{M_{\mathrm{BH}}}}{\ensuremath{M_\odot}}\right) = \\\nonumber
&&\quad (0.95\pm 0.15) \log\left( \frac{\ensuremath{L_{\mathrm{BCG}}}}{10^{10.91}\
\ensuremath{L_\odot}{}_{\mathrm{,}K}} \right) + (8.26\pm0.11) \,.
\end{eqnarray}
Thus, by
supplying a galaxy luminosity and a stellar mass-to-light ratio,
we can obtain all the other input parameters needed to compare the
dissipational and dissipationless models to observations of BCGs.
\subsection{The $L$--$\sigma$ Relation}
\label{ssec: FaberJacksonBCGs }
The innermost probe of the mass profile is the central velocity
dispersion of a galaxy. Elliptical galaxies fall on the
fundamental plane \citep{DjorgovskiDavis87, Dressler87} and one
projection of the plane is the Faber-Jackson
relation: the
relation between a galaxy's luminosity, $L$, and velocity
dispersion, $\sigma$ \citep{FaberJackson76}. In the $K$-band, the
Faber-Jackson relation observed by \citet{Pahre98} is
\begin{equation}
\label{eq:FJ}
M_K = -10.35\pm0.55\log\sigma_0\, .
\end{equation}
For BCGs, \citet{Lauer07} show that the
velocity dispersion saturates at
high luminosities, leading to the relation
\begin{equation}
\label{eq:lauerFJ}
M_V = -2.5 (6.5\pm1.3) \log\left(\frac{\sigma}{250\ \ensuremath{\mathrm{km\ s}^{-1}}}\right) -
22.45 \pm 0.18 \,.
\end{equation}
\citet{Desroches07} find a similar relation. In order to
compare to the $L$--$\sigma$ relation for BCGs, we normalize our
models to observed BCGs using the scaling relations described in
\S\ref{ssec:scaling}. The line-of-sight central velocity dispersions
averaged over an aperture of $1.64\ \mathrm{kpc}$ are then calculated
for both the dissipationless and dissipational models. The comparison
to the $L$--$\sigma$ relation from \citet{Lauer07} is shown
in Figure \ref{fig:Lsigma}. Although both the dissipational and
dissipationless models are slightly steeper than the
$L$--$\sigma$ relation found by Lauer, the dissipationless
model has a slope that more closely matches the observed
$L$--$\sigma$ for BCGs. To match each model
to the observations, the stellar mass-to-light ratios can be
adjusted. For the dissipational models, the best fit \ensuremath{M_*/L}{} in the
$K$-band is 1.43, while for the dissipationless models the best fit to the
$L-\sigma$ relation is for $\ensuremath{M_*/L}{}=2.40$. These are equivalent to
stellar mass-to-light ratios of 7.57 and 12.71 in the $V$-band, assuming
$V-K = 3.31$ for the BCG population. In the $K$-band,
the stellar mass to
light ratio measured by \citet{Cole01} is
$0.73$ for a Kennicut IMF and $1.32$ for a Salpeter IMF.
The stellar mass-to-light ratios derived for these models are simply
the mass in stars needed to reproduce the dynamics (in this case, the
central velocity dispersion) divided by the total observed luminosity
of the galaxy. Although the dissipationally formed galaxies were
brighter at high redshift due to star formation, we are only concerned
with the $z\approx0$ luminosity and dynamical state of the
galaxy. This corresponds to the luminosity of the evolved population;
therefore, we have implicitly included passive evolution in the
dissipational model and do not need to passively evolve the \ensuremath{M_*/L}{}
values derived above.
However, the mass-to-light ratios are sensitive to the empirical
scaling relations. For example, if the relation for $\ensuremath{{R_{\mathrm{e}}}}(\ensuremath{L_{\mathrm{BCG}}})$ is
replaced with that derived by \citet{Bernardi07} for galaxies fit by
de Vaucouleurs profiles ($n=4$), the best fit stellar mass-to-light
ratios become $0.98$ and $1.76$ for the dissipational and
dissipationless models respectively.
Additionally, the above calculations rely on the scaling relation
between the total halo mass and the BCG luminosity. Instead, we can
assume that the cluster is built
hierarchically out of galaxies formed at $z\approx2$. At $z=2$, the
concentration of a dark matter halo is a weak function of halo mass;
\citet{GaoNavarro08} find that $c \propto \ensuremath{M_{\mathrm{halo}}}^{-0.031}$. The
fraction of mass in a halo which will form stars is given by
$M_*/\ensuremath{M_{\mathrm{halo}}} \propto \ensuremath{M_{\mathrm{halo}}}^{-0.26}$ \citep{Lin03,Bode09}. Using these
relations for the concentration and stellar mass, the best-fit \ensuremath{M_*/L}{}
values become $0.69$ and
$1.55$ for the dissipational and dissipationless models respectively.
These values are in better agreement with the measured values given above.
It is not surprising that neither value for \ensuremath{M_*/L}{}
can be discarded based on the Faber-Jackson
relation. In the central regions of the galaxies, both models are
stellar-dominated, as shown in the lower panel of Figure
\ref{fig:vcirc}. Therefore, even though the dark matter density
can differ by more than a factor of 10, it only makes up $\sim10\%$ of
the total mass in the central regions, and thus does not determine the
central dynamics.
\middfig{\figLsigma}
\subsection{Microlensing Optical Depth}
\label{ssec:microlens}
Microlensing of quasars, which has been observed in
multiply imaged systems \citep[and references
therein]{Wozniak00, Wambsganss06},
in principle provides a probe of the mass function
of microlenses (MACHOS, stars, or dark matter substructure), and the
density of these microlenses relative to a
smooth background density \citep{SchechterWambsganss02,Dobler07,Pooley09}. Searches in the Sloan
Digital Sky Survey have found $\sim220$
strongly lensed quasars \citep{Inada08}, which are lensed by
individual galaxies or entire clusters. In the following, we use the
best-fit models for BCGs as an
example to show the expected differences in microlensing results
between the dissipational and dissipationless formation models.
The difference in stellar mass-to-light ratios between the dissipational and dissipationless models leads to
differences in the microlensing optical depth. The optical depth,
$\tau$, is proportional to the number density of lenses, stars in
this case, times the Einstein radius, $\theta_{\mathrm{E}}$, of
each lens. Assuming that the distance between lens and source is
large compared to the size of the
galaxy, the microlensing optical
depth is \citep{Paczynski86}:
\begin{equation}
\label{optDepthML}
\tau = \Sigma_* \frac{4\pi \mathrm{G}}{c^2}
\frac{D_{\mathrm{ls}}D_{\mathrm{l}}}{D_{\mathrm{s}}}\,,
\end{equation}
where $\Sigma_*$ is the projected stellar density and $D_i$ are the
angular diameter distances to the lens, to the source, and between the lens
and the source. If $(D_{\mathrm{ls}}D_{\mathrm{l}})/D_{\mathrm{s}}= D$
and the Sersic index of the lens galaxy is
assumed to be 4.0, then the microlensing optical depth at the
half-light radius is
\begin{eqnarray}
\label{tau1gpc}
\tau &=& 1.33\times 10^{-2} \left(\ensuremath{M_*/L}\right)_K \left(
\frac{\ensuremath{L_{\mathrm{BCG}}}}{10^{11} \ \ensuremath{L_\odot}}\right) \\\nonumber
&&\qquad\qquad \left( \frac{\ensuremath{{R_{\mathrm{e}}}}}{10 \ \mathrm{kpc}} \right)^{-2} \left(
\frac{D}{0.5 \ \mathrm{Gpc}}\right) \,.
\end{eqnarray}
For the dissipational and dissipationless models that best fit BCGs,
the ratio of the microlensing optical depths equals the ratio of
\ensuremath{M_*/L}{}, or $1.43/2.40 = 0.60$. In most cases, the microlensing
optical depth at the position of the
image is of order unity. Occasionally, individual microlensing events can be
observed.
Additionally, the relative
density of smoothly distributed matter (dark matter) to microlenses
(stars) can be probed
\citep{SchechterWambsganss02, Dobler07, Pooley09}. Using
the dissipational and dissipationless models shown in Figure
\ref{fig:vcirc} as an example
($L=6.0\ L_*$), the fraction of dark
matter to total matter along
a line-of-sight is 0.81 at 0.1 \ensuremath{{R_{\mathrm{e}}}}{} and 0.98 at 1.0 \ensuremath{{R_{\mathrm{e}}}}{} for the
dissipational model. For the
dissipationless model, the ratios at 0.1 \ensuremath{{R_{\mathrm{e}}}}{} and 1.0 \ensuremath{{R_{\mathrm{e}}}}{} are 0.48
and 0.97, respectively. These large differences should be measurable in
microlensing studies of multiply-imaged quasars.
\subsection{Strong Lensing}
\label{ssec:strongLens}
\middfig{\figlenses}
Strong lensing measurements provide a clean method of probing the
total mass
distribution of BCGs and their host clusters. \citet{SandTreu04}
present observations of radial and tangential arcs for six
clusters acting as lenses. The positions of the radial and tangential
arcs are given by the solutions to
\begin{eqnarray}
\label{eq:lenseqn}
0 &=& 1 -
\frac{\mathrm{d}}{\mathrm{d}R}\frac{M_{\mathrm{proj}}(R_{\mathrm{rad}})}{\pi
R_{\mathrm{rad}}}
\quad\mathrm{and}\\\nonumber
0 &=& 1 - \frac{M_{\mathrm{proj}}(R_{\mathrm{tan}})}{\pi R_{\mathrm{tan}}^2} \,,
\end{eqnarray}
where $M_{\mathrm{proj}}(R)$ is the projected mass interior to $R$ scaled by the
critical surface density,
\begin{equation}
\label{eq:sigmaCR}
\Sigma_{\mathrm{cr}} = \frac{c^2}{4\pi
G}\frac{D_{\mathrm{s}}}{D_{\mathrm{ls}}D_{\mathrm{l}}}\, .
\end{equation}
Together, the radial and tangential lenses constrain the slope of
the density profile and its normalization. \citet{SandTreu04} use the
lensing information as well as the velocity dispersion profile of the BCG to
create density models for the stars and the dark matter in each
lens. They find that the mean inner dark matter density profile
for six lensing clusters is
$r^{-0.52\pm0.3}$, significantly shallower than the NFW profile.
We repeat the analysis of \citet{SandTreu04},
fitting our dissipational and dissipationless models for the
dark matter profile to the three clusters with both radial and
tangential arcs. As in the previous section, we assume that each BCG
in the center of the cluster is built hierarchically, either a series of purely
dissipationless mergers of smaller stellar systems or by the
dissipational accretion of gaseous streams which lead to in situ star
formation. The dissipationless model for formation will yield a lower dark
matter density in the center of the cluster, while the dissipational
model for formation will lead to adiabatic contraction of the dark
matter. At $z\sim2$,
both dissipational and dissipationless formation mechanisms will be
important, but the ratio between the two mechanisms is unknown, and by
comparing data to the purely dissipationally and dissipationlessly
formed BCGs, we hope to constrain how much each mechanism contributes
to galaxy formation.
For the dissipational and dissipationless models, the fixed input
parameters for the models are the BCG
luminosity, half-light radius, and Sersic index ($n=4$). The free
parameters are the total cluster mass, \ensuremath{M_{\mathrm{cluster}}}, the dark halo concentration,
$c$, and \ensuremath{M_*/L}{}. By randomly selecting these input parameters from a
reasonable range, we can find both dissipational and
dissipationless models that are within $1$
and $2\sigma$ of the measured central
velocity dispersion and radial and tangential arc
locations. Projections of these points in \ensuremath{M_{\mathrm{cluster}}}--$c$
space and \ensuremath{M_{\mathrm{cluster}}}--\ensuremath{M_*/L}{} space are shown in Figure \ref{fig:lenses}. The
cluster Abell 383 does not have any models which lie within $1\sigma$
of the observations, so the $2\sigma$ models are plotted instead. These
projections show that the best-fit models lie on a tight
relation between \ensuremath{M_{\mathrm{cluster}}}{} and $c$. From equation \ref{eq:concentration},
halos in this mass range should have a concentration between 3.2 and
5.0, eliminating most of the best-fit models for MS 2137-23 and
RXJ-1133. In the case of MS 2137-23, this eliminates
almost all of the dissipationless models (squares). However, weak
lensing measurements of MS-2137-23 predict a concentration about twice
as large as simulations, thereby only eliminating a few models
\citep{Gavazzi03}.
Also illustrated in
Figure \ref{fig:lenses} is the difference in \ensuremath{M_*/L}{} between the two
models. The heavy lines indicate the median and
$25-75^{\mathrm{th}}$ percentile (SIQR) for the best-fits for each toy
model. As with the comparison to the $L$--$\sigma$ relation, the
dissipational models
have lower \ensuremath{M_*/L}{} values than the dissipationless models. For example,
the median $(\ensuremath{M_*/L})_V$ ratios for MS 2137-23 are
$4.3\pm2.4$ and $1.2\pm0.6$ for the dissipationless and dissipational
models, respectively. For RX-J1133, $(\ensuremath{M_*/L})_B=2.1\pm0.8$ for the
dissipational models and $(\ensuremath{M_*/L})_B=5.2\pm1.7$ for the dissipationless
models. The best-fit dissipational models
all have an \ensuremath{M_*/L}{} below 2.5(3.5) in the $V(B)$-band. Assuming passive
evolution, the expected value
of \ensuremath{M_*/L}{} is $(\ensuremath{M_*/L})_B\approx4.1\pm0.95$ at
$z\sim0.35$ \citep{TreuKoopmans04,vanderWel04, Treu06}. The purely
dissipational model for RX-J1133 is therefore marginally
inconsistent with expected \ensuremath{M_*/L}{} values. The shaded regions in Figure
\ref{fig:lenses} show the expected ranges for the stellar
mass-to-light ratios for the three clusters; MS 2137-23 is consistent
with both the dissipational and dissipationless models. Neither the dissipational nor the dissipationless models represent an adequate fit to Abell 383; however, the \ensuremath{M_*/L}{} values for the dissipational models within $2\sigma$ of the observations are in better agreement with the expected \ensuremath{M_*/L}{} value.
Taking MS 2137-23 as a specific example, Figure \ref{fig:msbestfit} plots
the stellar and dark matter density of the two models. Both models
selected have tangential and radial arcs and velocity dispersions
within the error bars of the observations. The best-fit model
parameters and the observed model parameters are given in Table
\ref{table:models}. For the dissipational model, the
dark matter density dominates over the stellar density at all radii.
From \citet{SandTreu04}, the best-fit inner slope of the dark matter
profile is 0.57. This is shown along with the $2\sigma$
error bars. The inner slope derived by \citet{SandTreu04} depends on
the concentration remaining fixed at 400 kpc. If the concentration is
allowed to vary, the best-fit inner slope will generally increase by
0.15, bringing it closer to the dissipationless model. However, the
dissipationless model also depends strongly on concentration and there
exist choices for \ensuremath{M_{\mathrm{cluster}}}, \ensuremath{M_*/L}{}, and $c$, which fit the observations
equally well, such that the dissipationless model almost matches an
NFW profile.
\middfig{\tablemodels}
\middfig{\figmsbestfit}
As illustrated by the strong lensing, the differences between the two
models for BCGs are small. This is
due to the fact that the ratio between the stellar mass of the galaxy
and the dark matter halo mass is very small, on the order of
$0.002$. Figure \ref{fig:Mratio} shows the ratio of the dark matter
to stellar mass inside $0.25\ensuremath{{R_{\mathrm{e}}}}$ as a function of galaxy luminosity
using the scaling relations from \S\ref{ssec:scaling}. As the
luminosity and cluster mass increase, the differences
between the two models and the initial NFW profile become
smaller. Therefore, the differences between the dissipational and
dissipationless methods of
galaxy formation will be most pronounced in smaller dark matter halos,
such as fossil groups and isolated ellipticals. In these cases the stellar
component of the central galaxy is much larger relative to the halo
component, making the differences between the dissipational and
dissipationless models more pronounced.
\middfig{\figMratio}
\section{Comparison to SAURON Data}
\label{sec:sauron}
Figure \ref{fig:Mratio} illustrates that the total mass-to-light
ratio is an increasing function of galaxy luminosity. This is in
agreement with the trend found by the SAURON project
\citep{Cappellari06}, which uses integrated-field spectroscopic
observations of 25 E/S0 galaxies. Using these observations and stellar
population models to determine the stellar mass-to-light
ratios for their sample of galaxies, \citet{Cappellari06} find that the total(dynamical) \ensuremath{M/L}{} is
consistently larger than \ensuremath{M_*/L}{} and that this difference increases with
increasing stellar mass. This trend is shown in Figure
\ref{fig:sauron}. The `$\times$'-symbols denote the SAURON data and the
shaded region is the best fit. The stellar mass-to-light ratio
($I$-band) is never larger than 3.4 for the brightest galaxies. The
best fitting line is given by \citep{Cappellari06}
\begin{equation}
\label{eq:SAURON}
\left(\ensuremath{M/L}\right)_I = (2.35\pm0.19)\left(\frac{L_I}{10^{10}L_\odot}\right)^{0.32\pm0.06}\,.
\end{equation}
This fit ignores the galaxy (M32) at $M_I \sim -17.5$.
\middfig{\figSAURON}
We can compare our dissipational and dissipationless models to the
SAURON data to determine whether the models recover the trend given by
equation \ref{eq:SAURON}. As in the previous section, we assume that
each of the 25 galaxies in the \citet{Cappellari06} study is formed
either by purely dissipational or purely dissipationless processes. We
fix the stellar mass-to-light ratio, the total $I$-band luminosity,
and the
effective radius (\ensuremath{{R_{\mathrm{e}}}}{}) for
each galaxy to the values given in \citet{Cappellari06}. We then vary
the dark matter halo mass for the dissipational and
dissipationless models of each galaxy until the velocity dispersion
within \ensuremath{{R_{\mathrm{e}}}}{} for both models matches the value reported in
\citet{Cappellari06}. Since dissipational formation increases the dark matter content in the center of a galaxy
compared to the dissipationless model, a smaller total halo mass is required to
recover the same central velocity dispersion in the dissipationally
formed galaxies than in the dissipationlessly formed galaxies. The
differences in halo mass lead to differences in dynamical \ensuremath{M/L}{}, which
are shown in Figure \ref{fig:sauron}. Both the dissipational and
dissipationless models reproduce the same trend in \ensuremath{M/L}{} with luminosity
as is shown in the SAURON data. However, the dissipationless models
yield slightly lower \ensuremath{M/L}{} values than the dissipational models, leading
to better agreement with the observed values. The standard deviation
of the SAURON points around the best fit line is 0.11. The
standard deviation of the dissipationless model points (squares)
around the best-fit line to the SAURON data (dotted line) is
0.14. The same value for the dissipational models is 0.25. However,
both the dissipational and dissipationless models have \ensuremath{M/L}{} values
higher than those from the SAURON data. Therefore, no combination of
these models will yield the measured SAURON galaxies. However, both
the dissipational and dissipationless models used here assume the
stellar orbits are
isotropic and that the galaxies are spherically symmetric. Both of
these assumptions will affect the model-calculated \ensuremath{M/L}{} values; in
the case of rotating galaxies, the calculated \ensuremath{M/L}{} values will be
lowered and possibly brought into better agreement with the SAURON
observations.
\section{Example Galaxy: NGC 4494}
\label{sec:ngc4494}
NGC 4494 is an ordinary elliptical galaxy with a $B$-band luminosity of
$2.37\times10^{10}\ \ensuremath{L_\odot}$. Because it is an isolated
galaxy instead of a BCG, the mass ratio between the dark matter halo
and the stars is smaller and, therefore, the difference between the
dissipationless and dissipational models of formation will be larger
than those found for BCGs in massive clusters. As with the BCGs, the
input model parameters for NGC 4494
can be constrained by observations.
\subsection{Velocity Dispersion using Planetary Nebulae}
\label{ssec:PNe}
Planetary nebulae have been established as a good mass tracer in the
outer regions of galaxies. These observations provide a good
comparison case for our extreme models of galaxy
formation. \citet{Napolitano09} measure positions and velocities of
planetary nebulae out to $\sim7\ \ensuremath{{R_{\mathrm{e}}}}$ in the elliptical NGC
4494, probing the velocity dispersion for the galaxy much farther out
than the central velocity dispersion. At these
large radii, the dark matter will be comparable to, and dominate over,
the stellar matter (see Figure \ref{fig:PNeDMdensity}), thus
emphasizing the differences between the dissipational and
dissipationless models. By fixing
the luminosity ($L_B = 2.37\times10^{10}\ \ensuremath{L_\odot}$), the effective
radius ($\ensuremath{{R_{\mathrm{e}}}}=3.68\ \mathrm{kpc}$),
and the Sersic index ($n=3.30$), of
the model galaxies to observations
from \citet{Napolitano09}, we can fit our
dissipational and dissipationless models to the
planetary nebulae velocity dispersion curves by varying the total halo
mass
and the stellar mass-to-light ratios. These fits also include a point
for the central
velocity dispersion, $\sigma=150.2\pm3.7\ \ensuremath{\mathrm{km\ s}^{-1}}$, as reported in the
Hyperleda\footnotemark\footnotetext{http://leda.univ-lyon1.fr}
database \citep{Paturel03}. The results are shown in Figure
\ref{fig:PNe}. For the dissipational
model, $(\ensuremath{M_*/L})_B=2.97$, while the
dissipationless model has a mass-to-light ratio of $3.96$. The total
halo masses are $6.0\times10^{11}\ \ensuremath{M_\odot}$ for the dissipational model
and $1.0\times10^{13}\ \ensuremath{M_\odot}$ for the dissipationless model. The
slightly
poorer quality fit for the dissipationless model is due to the fact
that the galaxy's effective radius is close to the minimum allowed for
the galaxy to form via dissipationless mergers ($\sim
2.7\ \mathrm{kpc}$), as discussed above in \S\ref{ssec:minRad}. As the
galaxy approaches this minimum size, there is insufficient binding energy
in some of the central dark matter layers to allow the stellar layers
to cross. The fit of the
dissipationless model could be improved if we relaxed our
model requirement that all the energy from the stars is deposited
locally, instead allowing more energy to be deposited in the outer
regions of the halo. This could be the case if the orbits of the
in-falling material were radial orbits instead of perfectly
circular orbits as assumed in this work.
Although the dissipational model shown in Figure \ref{fig:PNe} provides a better fit to the data, the $B$-band mass-to-light ratio required for the dissipational model is significantly lower than the value derived from stellar
population models, $4.3\pm0.7$ \citep{Napolitano09}. Thus, a purely
dissipational formation of NGC 4494 appears to be ruled out at the
$\sim1.9\sigma$ level.
\middfig{\figPNe}
Although adjusting the stellar mass-to-light ratio eliminates the
differences in the velocity dispersion profile for these two models,
the difference in dark matter
density between them remains large (see Figure
\ref{fig:PNeDMdensity}). The inner dark matter density profile
for the dissipationless model follows $\sim r^{-0.2}$, while that of the
dissipational model follows $\sim r^{-1.7}$, making the central dark
matter densities in the two models very different. The slope
index of the dark matter in the dissipational model is in good
agreement with that predicted for a final isothermal mass distribution
in \S\ref{sec:intro}.
\subsection{Dark Matter Annihilation}
\label{ssec:darkAnnihilation}
Although only available in the Milky Way, one direct method of
probing WIMP dark matter currently being explored
is the observation of gamma rays from the self-annihilation of WIMP dark
matter particles \citep{Stoehr03,Colafranceso06,Diemand07}. The
signal strength from such annihilations will be
proportional to $\rho_{\mathrm{dark}}^2$. Assuming a smooth distribution of
dark matter, the ratio of the annihilation signal strength within an
aperture of $\ensuremath{{R_{\mathrm{e}}}}$ for the dissipational and dissipationless models of
NGC 4494 is around $1890$, similar to what would be expected for a
Milky Way sized halo. Although not observable today, the Fermi
gamma-ray space
telescope hopes to measure the dark matter annihilation signal from
our own galaxy. The large difference in signal strength between the
dissipational and the dissipationless models calculated here dominates
over the boost in signal strength expected from unresolved
substructure, which is of order $10$ \citep{Strigari07,Kuhlen08},
providing another possible test of the formation history of the
stellar component of galaxies.
\middfig{\figPNeDMdensity}
\section{Conclusions}
\label{sec:conclusions}
We have shown that the two extreme cases for the assembly of the
stellar content of galaxies lead to large differences in the dark
matter density profiles of galaxies, assuming that the initial halo
conditions are well-described by N-body simulations. The stellar mass
density dominates over the dark matter
density in the central regions of both models; the dark matter
density in the dissipational models can be as much as two orders of
magnitude lower at $r\approx1\,\mathrm{kpc}$ than the dark matter
density in the dissipational models. However, because
galaxies are undoubtedly built up by both dissipational and
dissipationless accretion, most observations will not easily distinguish
between these two models. For example, although the best-fit models for
BCGs have different stellar mass-to-light ratios, neither is outside the
acceptable range of values from stellar population models and
observations. Strong gravitational lensing observations of
BCGs and their host clusters show that the dissipational formation
models have \ensuremath{M_*/L}{} values that are marginally too low compared to
those expected for passively evolving ellipticals. For RX-J1133, the
median $(\ensuremath{M_*/L})_B = 2.1\pm0.8$ and $5.2\pm1.7$ for the
dissipational and
dissipationless models, respectively, while the expected
$(\ensuremath{M_*/L})_B$ from
passive evolution is $4.1\pm1.0$ \citep{TreuKoopmans04}. This
discrepancy between \ensuremath{M_*/L}{} values
marginally rules out a purely dissipational formation history for
BCGs, in agreement with both theoretical expectations and
other observational evidence. Observations of strong
lensing by BCGs have been used to
effectively rule out dark matter density profiles as steep and steeper
than an NFW \citep{SandTreu04}, further strengthening arguments
against
a purely dissipational formation for the stellar component of BCGs.
Although extreme values for the
concentration ($\sim10$)
and $(\ensuremath{M_*/L})_B$ ($\sim1.0$) are allowed by the lensing and dynamics data, the
dissipational model can be ruled out for a more constrained and
plausible set of model parameters.
Both models adequately reproduce the trend of increasing total \ensuremath{M/L}{}
with galaxy luminosity for E/S0 galaxies, observed using
integrated-field spectroscopy by the SAURON project
\citep{Cappellari06}. However,
the lower \ensuremath{M/L}{} values found for the dissipationless models are in
better agreement with the data.
Constraints on the stellar mass-to-light ratios can also be used to
exclude the purely dissipational model of galaxy formation in the case of
the isolated elliptical, NGC 4494. Fitting the dissipational and
dissipationless models to observations of planetary nebulae yields
$(\ensuremath{M_*/L})_B$ values of $2.97$ and $3.96$ for the dissipational and
dissipationless models, respectively. Compared to $4.3\pm0.7$, the
$(\ensuremath{M_*/L})_B$
inferred from stellar synthesis models, the purely
dissipational model can be ruled out at the $1.9\sigma$ level.
Since the change in the dark matter density for both models is
directly related to the change in the central mass of the halo, the
larger the stellar component is relative to the dark matter halo, the
larger the
differences between the dissipational and dissipationless extremes
will be. Therefore, instead of examining the properties of BCGs, we
propose looking for the differences between dissipational and
dissipationless formation mechanisms using the brightest galaxies of
fossil groups
and isolated elliptical galaxies. The large differences attainable in
this mass range of galaxies is clearly illustrated by the study of NGC
4494. The dark matter density profiles
shown in Figure \ref{fig:PNeDMdensity} have inner slope indices of
$\alpha\approx0.2$ and $1.7$ for the dissipationless and dissipational
models, respectively. The differences in the dark matter density
profiles for galaxies in this mass range are significant enough
that they could be probed by galaxy-galaxy weak
lensing studies, provided the difference in dark matter slopes is not
removed by averaging over many galaxies with different formation
histories. Finally, the difference in dark matter content between the
dissipational and dissipationless models yields differences in the
signal strength from dark matter annihilation of order $\sim1890$, far
larger than the boost factor expected from the unresolved dark matter
substructure in the Milky Way halo.
The focus of this paper has been the energetics of the
dissipational and dissipationless galaxy formation mechanisms, not
the mechanisms themselves. For dissipational galaxy formation, we have
assumed that baryons cool and condense in the center of halos, leading
to adiabatic contraction of the surrounding dark matter. This behavior
has been confirmed in cosmological simulations. Although simplified,
the model presented here is correct, on average, for more complicated
galaxy formation scenarios, including major as well as minor mergers,
and accretion from filaments instead of spherical shells \citep{Gnedin04}.
The physical mechanism we propose for dissipationless galaxy formation is the
dynamical friction of small stellar clumps against a smooth dark
matter background. In the models used here, we assume circular orbits
for the incoming stellar material. The inclusion of radial orbits and
non-spherical galaxies is left for future work, as it requires
modeling a combination of dissipational and dissipationless
formation mechanisms. We assume that the build-up of
the galaxy occurs via a series of small, minor mergers
\citep{Bezanson09, Cook09, Naab09}, not allowing
for equal-mass mergers, which more violently disrupt the
system. Indeed, it has been shown in dissipationless N-body
simulations that equal-mass merger remnants will retain the profile of
the steepest progenitor \citep{BoylanKolchinMa04, Dehnen05, Kazantzidis06,
Vass09}. Therefore, cuspy
dark matter profiles are robust under major mergers. However, if
baryons are added to dark matter halos, they will presumably condense
more than the dark matter, making up the bulk of the central,
high-density matter in merging halos. As two halos merge, the outer,
less tightly bound and dark-matter-dominated components will be
tidally stripped, but the central high density, predominantly stellar
components will settle into the center of the merger remnant, undergoing
dynamical friction along the way. Therefore, the baryons are an
important ingredient to dry merger scenarios because they ensure the
merging clump is sufficiently tightly bound to reach the central regions of
the nascent galaxy.
The extreme differences in the inner dark matter halo densities for
the dissipational and dissipationless models emphasize the importance
of the addition of baryons to dark matter halos. Without introducing
modifications to the \ensuremath{\Lambda\mathrm{CDM}}{} paradigm, dark matter halo cusps can be
reduced to cores via the dissipationless formation of the central
stellar regions
of galaxies. The balance between dissipational and dissipationless
formation mechanisms can be probed by observations. Current
observations of BCGs and ellipticals galaxies are sufficient to
exclude a purely dissipational formation mechanism for these
galaxies. Future measurements of stellar mass-to-light ratios from
microlensing observations, and direct detection of dark matter in the
Milky Way will help to constrain the balance between dissipational and
dissipationless formation mechanisms and the dependence of this
balance on time and environment.
\acknowledgments We thank S. Tremaine and J. Binney for their valuable
comments and
corrections. We would also like to thank the referee for his/her
comments and suggestions. CNL acknowledges support from the NDSEG
fellowship.
\bibliographystyle{apj}
|
\section{The main conjecture}
In this paper we propose a conjectural way for exact computing of
the $S$-matrix and the Green functions of quantum field theory.
Recall that the Schrodinger functional differential equation reads
\begin{equation}
ih\frac{\partial\Psi}{\partial t}=\hat H(t)\Psi,
\end{equation}
where $\Psi=\Psi(t,\varphi(\cdot))$ is the unknown ``half-form''
on the space of functions $\varphi(\x)$, $\x=(x_1,\ldots,x_d)$,
$\hat H(t)=H(t,\hat\varphi(\cdot),\hat\pi(\cdot))$ is the quantum
Hamiltonian of the theory, and the operators
$\hat\varphi(\x)=\varphi(\x)$ and
$\hat\pi(\x)=-ih\frac{\delta}{\delta\varphi(\x)}$ satisfy the
canonical commutation relations
\begin{equation}
[\hat\varphi(\x),\hat\pi(\x')]=ih\delta(\x-\x'),\ \
[\hat\varphi(\x),\hat\varphi(\x')]=[\hat\pi(\x),\hat\pi(\x')]=0.
\end{equation}
For the relativistically invariant generalization of the
functional differential Schrodinger equation, see [1]. For
example, for the scalar field with self-action in $\R^{d+1}$ the
Hamiltonian reads
\begin{equation}
\begin{aligned}{}
H(t,\varphi(\cdot),\pi(\cdot))&=H_0(\varphi(\cdot),\pi(\cdot))\\
&+\int\left(\frac1{k!}g(t,\x)\varphi(\x)^k+j(t,\x)\varphi(\x)\right)d\x,\\
H_0(\varphi(\cdot),\pi(\cdot))&=\int\frac12\left(\pi(\x)^2+\sum_{j=1}^d\varphi_{x_j}(\x)^2+m^2\varphi(\x)^2\right)d\x.
\end{aligned}
\end{equation}
Here $g(t,\x)$ (the interaction cutoff function) and $j(t,\x)$
(the source) are smooth functions with compact support. For
simplicity of exposition, below we restrict ourselves by this
model. (One can see that equation (1,3) has no nonzero solutions
if $\Psi$ is a usual functional of $\varphi(\x)$, see [1].)
Let us regularize the operators $\hat\pi(\x)$ and
$\hat\varphi(\x)$, as in [2], as follows: consider the delta-like
family of smooth functions with compact support
$f_\Lambda(\x)\to\delta(\x)$, where $\Lambda\to\infty$ is the
regularization parameter (the ultraviolet regularization at small
distances), and a family of smooth functions with increasing
compact support $g_L(\x)\to1$ as $L\to\infty$ (the infrared regularization
at big distances), and put
\begin{equation}
\begin{aligned}{}
\hat\varphi_{\Lambda,L}(\x)&=g_L(\x)\int f_\Lambda(\x-\x_1)\hat\varphi(\x_1)d\x_1,\\
\hat\pi_{\Lambda,L}(\x)&=g_L(\x)\int
f_\Lambda(\x-\x_1)\hat\pi(\x_1)d\x_1.
\end{aligned}
\end{equation}
Consider the regularized Schrodinger functional differential
equation
\begin{equation}
ih\frac{\partial\Psi}{\partial t}=\hat H^{\Lambda,L}(t)\Psi,
\end{equation}
where
\begin{equation}
\hat
H^{\Lambda,L}(t)=H(t,\hat\varphi_{\Lambda,L}(\cdot),\hat\pi_{\Lambda,L}(\cdot)).
\end{equation}
The regularized quantum Hamiltonian $\hat H^{\Lambda,L}(t)$ and the
regularized free quantum Hamiltonian
\begin{equation}
\hat
H_0^{\Lambda,L}=H_0(\hat\varphi_{\Lambda,L}(\cdot),\hat\pi_{\Lambda,L}(\cdot))
\end{equation}
are well-defined and regular operators in the Fock Hilbert space
of functionals
$$
\Psi(\varphi(\cdot))=\Psi_0(\varphi(\cdot))\exp\left(-\frac1{2h}\int\tilde\varphi(\p)\tilde\varphi(-\p)\omega_\p
d\p\right),
$$
where $\p=(p_1,\ldots,p_d)$,
$\tilde\varphi(\p)=\frac1{(2\pi)^{n/2}}\int
e^{-i\p\x}\varphi(\x)d\x$, $\omega_\p=\sqrt{\p^2+m^2}$.
Denote by $U_{\Lambda,L}(T_1,T_2)$ the evolution unitary operator of
equation (5) in the Fock space from $t=T_1$ to $t=T_2$, and choose
the numbers $-T_1,T_2$ so large that the supports of the functions
$g(t,\x)$ and $j(t,\x)$ be contained in the interval $(T_1,T_2)$.
Denote
\begin{equation}
S_{\Lambda,L}(g(\cdot),j(\cdot))=e^{iT_2\hat
H_0^{\Lambda,L}/h}U_{\Lambda,L}(T_1,T_2)e^{-iT_1\hat H_0^{\Lambda,L}/h}.
\end{equation}
Clearly, this unitary operator in the Fock space does not depend
on $T_1,T_2$.
{\bf The Main Conjecture.} {\it The strong limit
\begin{equation}
S(g(\cdot),j(\cdot))=\lim\limits_{\Lambda,L\to\infty}S_{\Lambda,L}(g(\cdot),j(\cdot))
\end{equation}
is correctly defined modulo multiplication by a phase factor
$e^{ic}$, for $c$ a real number, and does not depend on the way of
regularization \emph(i.~e., on the choice of the functions
$f_\Lambda(\x)$, $g_L(\x)$\emph). The strong limit
\begin{equation}
S(g,j(\cdot))=\lim\limits_{g(t,\x)\to g}S(g(\cdot),j(\cdot))
\end{equation}
exists and coincides with the generating functional for the
operator Green functions, and the unitary operator
\begin{equation}
S(g)=S(g,j\equiv 0)
\end{equation}
coincides with the physical $S$-matrix. }
This Conjecture is partly a mathematical conjecture, and partly a
conjectural physical law.
\section{Discussion}
In this Section we present heuristic arguments in favor of the
Main Conjecture from \S1, and discuss the mathematical and
physical contents of this Conjecture.
Regarding the mathematical contents of the Conjecture, one can
imagine that there exists a space of distribution ``half-forms''
(or ``half-densities'') $\Psi$ on the Schwartz space of functions
$\varphi(\x)$, and that there exists a mathematical theory of
functional differential equations (for example, like the
Schrodinger functional differential equation above) with solutions
in this space of half-forms. (It was the main aim of Dirac in his
book [3] to construct a similar space for fermions.) The Fock
spaces are parts of this space of half-forms. Then a surprising
and mysterious fact which follows from the physical picture and
which I do not understand, is that the result of evolution of the
Schrodinger functional differential equation with the initial
conditions in the Fock space at $t=T_1$, returns to the Fock space
at $t=T_2$. It is clear that between $t=T_1$ and $t=T_2$ the
vector $\Psi$ leaves the Fock space. In [4,5] it is conjectured
that under the evolution of the Schrodinger functional
differential equation and its relativistically invariant
generalization from the surfaces $t=const$ to curved space-like
surfaces in space-time, the Fock space evolves into a family of
Hilbert spaces parameterized by space-like surfaces, and the
generalized Schrodinger equation yields an integrable flat
connection in this family. Even for the free scalar field, it is
proved in [6] that the result of evolution of the generalized
Schrodinger equation from the surface $t=const$ to a curved
space-like surface, for $d>1$ leaves the Fock space. The fact that
the solution of the functional differential Schrodinger equation
returns to the Fock space, is mathematically confirmed by results
of the theory of complex germ of Maslov and Shvedov ([2], cf.
[7]), which state that the result of quasiclassical evolution of
the functional differential Schrodinger equation along any
classical trajectory in the phase space returns to the Fock space.
Regarding the physical contents of the Conjecture, one should
prove that for the renormalizable theories, the Taylor series of
the $S$-matrix $S(g,j(\cdot))$ at $g=0$, $j\equiv 0$ coincides
with the renormalized perturbation series for the generating
functional of operator Green functions of the theory, since these
renormalized perturbation series are well checked by experiment.
Let us sketch a plan of such a proof.
In the book [8] by Bogolubov and Shirkov, the renormalized
perturbation series for the $S$-matrix and the Green functions are
constructed as the limit as $g(t,\x)\to g=const$ of a more general
object, the renormalized perturbation series $\tilde
S(g(\cdot),j(\cdot))$ with non-constant interaction cutoff
function $g(t,\x)$. This object is almost uniquely (up to the
change of parameters $m$, $g(x)$) characterized by the properties
of unitarity, causality, Lorentz invariance, and the
correspondence principle stating that the coefficient before
$g(x)$ in $\tilde S(g(\cdot),0)$ coincides with the normally
ordered interaction Lagrangian. After taking the limit $g(x)\to
g$, the parameters are fixed uniquely by conditions on the Green
functions of the theory (e.~g., for the $\varphi^4$ theory in
$\R^{3+1}$, the condition that the two-point Green function
$G^{(2)}(p_1,p_2)$ has poles at $p_i^2=m^2$, and the four-point
one-particle irreducible Green function
$\Gamma^{(4)}(p_1,p_2,p_3,p_4)$ equals $g$ at the point
$p_1=p_2=p_3=p_4=0$).
Note that the conditions of unitarity and causality are fulfilled
for any evolution operator (or limit of evolution operators) of
unitary evolution differential equations with $g(t,\x)$, $j(t,\x)$
as coefficients. Hence our operator $S(g(\cdot),j(\cdot))$ and its
Taylor series at $0$ satisfy these conditions. Regarding Lorentz
invariance for $S(g(\cdot),j(\cdot))$, it follows from the fact
that the Schrodinger functional differential equation (and its
regularizations) admit a relativistically invariant
generalization, the generalized Schrodinger equation [1] which
forms an integrable flat connection over the family of space-like
surfaces. Finally, the correspondence principle, say, for the
$\varphi^4$ model in $\R^{3+1}$ is an easy direct computation.
Therefore, the Taylor series of our $S(g(\cdot),j(\cdot))$
coincides with one of Bogolubov $S$-matrices $\tilde
S(g(\cdot),j(\cdot))$. The remaining check of parameters as
$g(x)\to g$ should not be a difficult task. We are so sure that we
obtain the right result due to our final argument which is the
inner conceptual simplicity of the theory.
Finally, note that regarding computational part of our approach,
it yields an algorithm of computation different from the
renormalization in the Feynman diagram technique. This is seen,
for example, already on the $\varphi^4$ model (see below). However, this part
of our investigation is not finished yet, so we leave it as a
challenging problem, especially for physically interesting
theories such as Yang--Mills theory or quantum gravity.
For further problems closely related to this paper, see [9].
\section{No-Counterterm perturbation series for the $\varphi^4$ model: the setup}
Traditional perturbation series for the $\varphi^4$ model is obtained by renormalization of the expression
\begin{equation}
T\exp\int g\,:\varphi(x)^4:/4!\,dx,
\end{equation}
where dots denote the normal ordering, and $\varphi(x)$ is the free scalar field, $x\in\R^{d+1}$.
It is easy to see that the perturbation series for the $S$-matrix in the
No-Counterterm approach is obtained by developing the expression
\begin{equation}
T\exp\int g\varphi(x)^4/4! dx
\end{equation}
(without normal orderings) into power series with respect to $g$.
This means that we first regularize the operator
$$
\varphi(x)^4\to\reg\varphi(x)^4=\varphi_{\Lambda,L}(x)^4,
$$
where $\Lambda\to\infty$ is the parameter of the ultraviolet regularization at small distances (and large momenta),
and $L\to\infty$ is the parameter of the infrared regularization at large distances (and small momenta). Next, we
develop the regularized integral (13) into series over powers of $g$, and finally we omit the regularization.
To perform this procedure, note first that we have
\begin{equation}
\varphi_{\Lambda,L}(x)^4/4!=:\varphi_{\Lambda,L}(x)^4:/4!+C_{\Lambda,L}:\varphi_{\Lambda,L}(x)^2:/2+\const,
\end{equation}
where $C_{\Lambda,L}$ and $\const$ are certain divergent constants. The latter constant can be neglected, since we are
interested in the expression only modulo an overall phase factor (see \S1).
Now the regularized integral (13) can be developed into
series by usual Feynman diagram techniques, using (14) (see, for example, [8]).
The quadratic term in (14) means that we change the propagator as
follows:
\begin{equation}
\begin{aligned}{}
&\frac1{p^2-m^2+i\varepsilon}\to\frac1{p^2-m^2+i\varepsilon}
+\frac1{p^2-m^2+i\varepsilon}gC_{\Lambda,L}\frac1{p^2-m^2+i\varepsilon}\\
&+\frac1{p^2-m^2+i\varepsilon}gC_{\Lambda,L}\frac1{p^2-m^2+i\varepsilon}gC_{\Lambda,L}\frac1{p^2-m^2+i\varepsilon}+\ldots.
\end{aligned}
\end{equation}
This sum of a geometric progression converges, for $g$ small enough, to the new propagator
\begin{equation}
\frac1{p^2-m^2+i\varepsilon-gC_{\Lambda,L}}.
\end{equation}
In the next Section the regularized integrals corresponding to Feynman diagrams with this new propagator are tested to
converge as $\Lambda,L\to\infty$.
\section{Rough estimates}
We consider the ultraviolet cutoff regularization ($d=3$)
\begin{equation}
\begin{aligned}{}
&\reg f(p)=\reg f(p_0,\p)=\reg f(p_0,p_1,p_2,p_3)\\
&=0\text{ if }|p_i|\ge\Lambda\text{ for some }i, 0\le i\le3.
\end{aligned}
\end{equation}
If the mass $m>0$, then we shall not need the infrared regularization at all.
The computation shows that (for any time $t$)
\begin{equation}
\begin{aligned}{}
&C_{\Lambda,L}=6\reg\int[\varphi_-(t,\p),\varphi_+(t,\p')]d\p d\p'\\
&=\reg\int\frac{6h\delta(\p+\p')}{2\sqrt{\p^2+m^2}}d\p d\p'\sim 3 h\Lambda^2.
\end{aligned}
\end{equation}
Substituting this into the propagator, one obtains for the one-loop ``fish'' diagram the following expression:
\begin{equation}{}
\reg\int\frac1{(p^2-m^2+i\varepsilon-3 gh\Lambda^2)((k-p)^2-m^2+i\varepsilon-3 gh\Lambda^2)}dp.
\end{equation}
(Here $k$ is the sum of ingoing $4$-momenta of the diagram.)
Let us divide each of the two brackets in the denominator by $\Lambda^2$.
Then the integrand becomes $\sim1$, and the integration domain
is a $4$-cube of size $\Lambda$. Hence the whole integral is $\sim\Lambda^{-4}\Lambda^4\sim1$, and it is finite.
However, for the simplest two-loop diagram with two outgoing edges we have the integral
\begin{equation}
\begin{aligned}{}
\reg\int&\frac1{(p^2-m^2+i\varepsilon-3 gh\Lambda^2)(q^2-m^2+i\varepsilon-3 gh\Lambda^2)}\\
&\times\frac1{(k-p-q)^2-m^2+i\varepsilon-3 gh\Lambda^2}dpdq,
\end{aligned}
\end{equation}
and the same argument shows that the integral diverges as $\Lambda^{-6}\Lambda^8\sim\Lambda^2$.
\section{Conclusion}
Thus, if we believe into the No-Counterterm Conjecture, we should conclude that the estimate above is
too rough for the two-loop diagram. Otherwise, if all the estimates above are correct, we see that for $d=3$
the ``No-Count\-er\-term approach'' is not valid and requires
counterterms, as well as the traditional approach.
It seems that our estimate is correct if considered as an upper bound for the integral. As a lower bound it can be
incorrect.
For a general $\varphi^4$ diagram
in $(d+1)$-dimensional space-time, the same argument as above gives the following estimate of the diagram integral.
Denote by $E_i$ ($E_e$) the number of internal (respectively external) edges of the diagram, by $L$ the number of
independent loops, by $V$ the number of vertices. Assume $d>2$. Then the following Theorem holds:
{\bf Theorem.} {\it The integral is no greater than
$O(\Lambda^m)$, where
\begin{equation}
\begin{aligned}{}
m&=-(d-1)E_i+(d+1)L\\
&=(d+1)(L-E_i)+2E_i\\
&=(d+1)(1-V)+2E_i\ \ (\text{since }V-E_i+L=1)\\
&=(d+1)(1-V)+4V-E_e\ \ (\text{since }4V=2E_i+E_e)\\
&=d+1-(d-3)V-E_e.
\end{aligned}
\end{equation}
Therefore, if the sign of $m$ is negative, then the limit of the integral is zero.}
At least, for $d=3$, $L=1$ and for $d\ge 6$ all the diagrams seemingly converge.
|
\part{\wt{\partial}}
\def\parta{\part_a}
\def\partb{\part_b}
\def\partc{\part_c}
\def\partv{\part_v}
\def\partw{\part_w}
\def\sa{{_\alpha}}
\def\sb{{_\beta}}
\def\sab{_{\alpha\beta}}
\def\sba{_{\beta\alpha}}
\def\parx{\partial_{k_x}}
\def\pary{\partial_{k_y}}
\def\para{\partial_a}
\def\parb{\partial_b}
\def\parc{\partial_c}
\def\park{\partial_{\rm k}}
\def\partx{\wt{\partial}_{k_x}}
\def\party{\wt{\partial}_{k_y}}
\def\Mt{{\wt{M}}}
\def\plam{{^{(\lambda)}}}
\begin{document}
\title{Theory of orbital magnetoelectric response}
\author{Andrei Malashevich$^1$, Ivo Souza$^1$, Sinisa Coh$^2$
and David Vanderbilt$^2$}
\address{$^1$ Department of Physics,
University of California,
Berkeley, CA 94720-7300, USA}
\address{$^2$ Department of Physics \& Astronomy,
Rutgers University,
Piscataway, NJ 08854-8019, USA}
\ead{<EMAIL>}
\begin{abstract}
We extend the recently-developed theory of bulk orbital
magnetization to finite electric fields, and use it to calculate the
orbital magnetoelectric response of periodic insulators. Working in
the independent-particle framework, we find that the finite-field
orbital magnetization can be written as a sum of three
gauge-invariant contributions, one of which has no counterpart at
zero field. The extra contribution is collinear with and explicitly
dependent on the electric field. The expression for the orbital
magnetization is suitable for first-principles implementations,
allowing to calculate the magnetoelectric response coefficients by
numerical differentiation. Alternatively, perturbation-theory
techniques may be used, and for that purpose we derive an expression
directly for the linear magnetoelectric tensor by taking the first
field-derivative analytically. Two types of terms are obtained.
One, the `Chern-Simons' term, depends only on the unperturbed
occupied orbitals and is purely isotropic. The other, `Kubo' terms,
involve the first-order change in the orbitals and give isotropic as
well as anisotropic contributions to the response. In ordinary
magnetoelectric insulators all terms are generally present,
while in strong $Z_2$ topological insulators only the
Chern-Simons term is allowed, and is quantized. In order to
validate the theory we have calculated under periodic boundary
conditions the linear magnetoelectric susceptibility for a 3-D
tight-binding model of an ordinary magnetoelectric insulator, using
both the finite-field and perturbation-theory expressions. The
results are in excellent agreement with calculations on bounded
samples.
\end{abstract}
\pacs{75.85.+t,03.65.Vf,71.15.Rf}
\submitto{\NJP}
\maketitle
\section{Introduction}
\label{sec:intro}
In insulating materials in which both spatial inversion and
time-reversal symmetries are broken, a magnetic field ${\bi B}$ can
induce a first-order electric polarization ${\bi P}$, and conversely
an electric field $\eb$ can induce a first-order magnetization
$\mb$~\cite{odell70,fiebig05}. This linear magnetoelectric (ME)
effect is described by the susceptibility tensor
\beq
\label{eq:alpha}
\alpha_{da}=\left.\frac{\partial P_d}{\partial
B_a}\right|_{{\bi B}=0} =\left.\frac{\partial M_a}{\partial
\e_d}\right|_{\eb=0}
\eeq
where indices label spatial directions. This tensor can be divided
into a ``frozen-ion'' contribution that occurs even when the ionic
coordinates are fixed, and a ``lattice-mediated'' contribution
corresponding to the remainder. Each of these two contributions can
be decomposed further according to whether the magnetic interaction is
associated with spins or orbital currents, giving four
contributions to $\alpha$ in total.
All of those contributions, except the frozen-ion orbital one, are
relatively straightforward to evaluate, at least in principle, and
{\it ab initio} calculations have started to appear. For example, the
lattice-mediated spin-magnetization response was calculated in
\cite{iniguez08} for Cr$_2$O$_3$ and in \cite{wojdel09} for BiFeO$_3$
(including the strain deformation effects that are present in the
latter), and calculations based on the converse approach (polarization
response to a Zeeman field) were recently
reported~\cite{delaney09}. One generally expects the lattice-mediated
couplings to be larger than the frozen-ion ones, and insofar as the
spin-orbit interaction can be treated perturbatively, interactions
involving spin magnetization are typically larger than the orbital
ones. However, we shall see that there are situations in which the
spin-orbit interaction cannot be treated perturbatively, and in which
the frozen-ion orbital contribution is expected to be dominant.
Therefore, it is desirable to have a complete description which
accounts for all four contributions.
The frozen-ion orbital contribution is, in fact, the one part of the
ME susceptibility for which there is at present no satisfactory
theoretical or computational framework, although some progress towards
that goal was made in two recent works\cite{qi08,essin09}. Following
Essin {\it et al.}~\cite{essin09} we refer to it as the ``orbital
magnetoelectric polarizability'' (OMP). For the remainder of this
paper, we will focus exclusively on this contribution to
\eref{eq:alpha}, and shall denote it simply by $\alpha$. Accordingly,
the symbol $\mb$ will be used henceforth for the orbital component of
the magnetization.
The question we pose to ourselves is the following: what is the
quantum-mechanical expression for the tensor $\alpha$ of a generic
three-dimensional band insulator? We note that the conventional
perturbation-theory expression for $\alpha$~\cite{raab2005,barron04}
does not apply to Bloch electrons,
as it involves matrix elements of unbounded operators. The proper
expressions for ${\bi P}$~\cite{King-Smith} and ${\bi
M}$~\cite{xiao05,timo05,ceresoli06,shi07} in periodic crystals
have been derived, but so far only at ${\bi B}=0$ and $\eb=0$
respectively. The evaluation of equation \eref{eq:alpha} remains
therefore an open problem.
Phenomenologically, the most general form of $\alpha$ is a $3\times 3$
matrix where all nine components are independent. Dividing it into
traceless and isotropic parts, the latter is conveniently expressed
in terms of a single dimensionless parameter $\theta$ as
\beq
\label{eq:alpha-iso}
\alpha^\theta_{da}=\frac{\theta e^2}{2\pi hc}\delta_{da}.
\eeq
The presence of an isotropic ME coupling is equivalent to the addition
of a term proportional to $\theta\eb\cdot{\bi B}$ to the
electromagnetic Lagrangian. Such a term describes ``axion
electrodynamics'' \cite{wilczek87} and \eref{eq:alpha-iso} may
therefore also be referred to as the ``axion OMP.'' The
electrodynamic effects of the axion field are elusive (in fact, the
very existence of $\alpha^\theta$ was debated until recently: see
\cite{raab97,hehl08} and references therein). For example, in a
finite, static sample cut from a uniform ME medium those effects are
only felt at the surface\cite{wilczek87,obukhov05}. In
particular, $\alpha^\theta$ gives rise to a surface Hall
effect~\cite{widom86}.
An essential feature of the axion theory is that a change of $\theta$
by $2\pi$ leaves the electrodynamics invariant~\cite{wilczek87}. The
profound implications for the ME response of materials were recognized
by Qi {\it et al.}~\cite{qi08}, and discussed further by Essin
{\it et al.}~\cite{essin09}. These authors showed that there is a
part of the isotropic OMP which remains ambiguous up to integer
multiples of $2\pi$ in the corresponding $\theta$ until the surface
termination of the sample is specified. For example, a change by
$2\pi n$ occurs if the surface is modified by adsorbing a quantum
anomalous Hall layer. Hence this particular contribution to
$\theta$ can be formulated as a bulk quantity only modulo a quantum of
indeterminacy, in much the same way as the electric polarization ${\bi
P}$~\cite{King-Smith,resta-review07}. A microscopic expression for
it was derived in the framework of single-particle band theory by
the above authors. It is given by the Brillouin-zone integral of
the Chern-Simons form~\cite{csforms} in $k$-space, which is a
multivalued global geometric invariant reminiscent of the Berry-phase
expression for ${\bi P}$~\cite{King-Smith}. We denote henceforth this
``geometric'' contribution to the OMP as the Chern-Simons OMP (CSOMP).
A remarkable outcome of this analysis is the prediction~\cite{qi08} of
a purely isotropic ``topological ME effect,'' associated with the
CSOMP, in a newly-discovered class of time-reversal invariant
insulators known as $Z_2$ topological
insulators~\cite{physicstoday,hasan10,moore10}. As a result of the
multivaluedness of $\theta$, the presence of time-reversal symmetry in
the bulk, which takes $\theta$ into $-\theta$, is consistent with two
solutions: $\theta=0$, corresponding to ordinary insulators, and
$\theta=\pi$, corresponding to strong $Z_2$ topological
insulators.\footnote{An analogous situation occurs in the theory of
polarization: inversion symmetry, which takes ${\bf P}$ into $-{\bf
P}$, allows for a nontrivial solution which does not include ${\bf
P}=0$ in the ``lattice'' of values~\cite{resta-review07}. An
important difference is that while $\theta$ is a directly measurable
response, only {\it changes} in ${\bf P}$ are detectable, so that
the experimental implications of the nontrivial solution are less
clear in this case.} The latter case is non-perturbative in the
spin-orbit interaction, and $\theta=\pi$ amounts to a rather large ME
susceptibility (in Gaussian units it is $1/4\pi$ times the fine
structure constant, or $\sim$6$\times$10$^{-4}$, to be compared with
$\sim$1$\times$10$^{-4}$ for the total ME response of Cr$_2$O$_3$ at
low temperature~\cite{wiegelmann94}).
It is not clear from these recent works, however, whether the
isotropic CSOMP constitutes the full OMP response of a generic
insulator. It does appear to do so for the tight-binding model
studied in \cite{essin09}, whose ME response was correctly reproduced
by the Chern-Simons expression even when the parameters were tuned to
break time-reversal and inversion symmetries (i.e., for generic
$\theta$ not equal to 0 or $\pi$). On the other hand, other
considerations seem to demand additional contributions. For example,
it is not difficult to construct tight-binding models of molecular
crystals in which it is clear that the OMP cannot be purely isotropic.
In this work we derive, using rigorous quantum-mechanical arguments,
an expression for the OMP tensor $\alpha$ of band insulators, written
solely in terms of bulk quantities (the periodic Hamiltonian and
ground state Bloch wavefunctions, and their first-order change in an
electric field). We restrict our derivation to non-interacting
Hamiltonians, as the essential physics we wish to describe occurs
already at the single-particle level. We find that in crystals with
broken time-reversal and inversion symmetries there are, in addition
to the CSOMP term discussed in \cite{qi08,essin09}, extra terms which
generally contribute to both the trace and the traceless parts of
$\alpha$.
Our theoretical approach closely mimics one type of ME response
experiment: a finite electric field $\eb$ is applied to a bounded
sample, and the (orbital) magnetization is calculated in the presence
of the field. Then the thermodynamic limit is taken at fixed field.
This key step in the derivation must be done carefully, so that
crucial surface contributions are not lost in the process, and
here we follow the Wannier-based approach of
references~\cite{timo05,ceresoli06}, adapted to $\eb\not=0$. Finally the linear response
coefficient $\alpha_{da}=\partial M_a/\partial \e_d$ is extracted in
the limit that $\eb$ goes to zero.
In a concurrent work by Essin, Turner, Moore, and one of
us~\cite{essin10} an alternative approach was taken, which is closer
in spirit to the calculation in \cite{King-Smith} of the change in
polarization as an integrated current: the adiabatic current induced
in an infinite crystal by a change in its Hamiltonian in the
presence of a magnetic field is computed, and then expressed as a
total time derivative. The two approaches are complementary and
lead to the same expression for $\alpha$, illuminating it from
different angles.
The paper is organized as follows. In
section~\ref{sec:finite-field} we derive the bulk expression for
${\bf M}(\eb)$, and reorganize it into three gauge-invariant
contributions, one of which yields directly the CSOMP response.
The gauge-invariant decomposition of $\mb(\eb)$ is done at first in
$k$-space for periodic crystals, and then also for bounded samples
working in real space. In section~\ref{sec:linear-response} we derive
a $k$-space formula for the OMP tensor $\alpha$ by taking analytically
the field-derivative of ${\bi M}(\eb)$. Numerical tests on a
tight-binding model of a ME insulator are presented at appropriate
places throughout the paper in order to validate the bulk expressions
for $\mb(\eb)$ and $\alpha$. In \ref{app:model} we describe the
tight-binding model, as well as technical details on how the various
formulas are implemented on a $k$-point grid. \ref{app:derivation} and
\ref{app:band-sum-consistency} contain derivations of certain results
given in the main text.
\section{Orbital magnetization in finite electric field}
\label{sec:finite-field}
\subsection{Preliminaries}
The orbital magnetization $\mb$ is defined as the orbital moment per
unit volume,
\beq
\label{eq:M-finite}
\mb=-\frac{e}{2cV}\sum_i\,\bra{\psi_i}\r\times\v\ket{\psi_i}.
\eeq
Here $e>0$ is the magnitude of the electron charge, $V$ is the sample
volume, and $\ket{\psi_i}$ are the occupied eigenstates. While this
expression can be directly implemented when using open boundary
conditions, the electronic structure of crystals is more conveniently
calculated and interpreted using periodic boundary conditions,
in order to take advantage of Bloch's theorem. This poses however
serious difficulties in dealing with the circulation operator
$\r\times\v$, because of the unbounded and nonperiodic nature of the
position operator $\r$. These subtle issues were fully resolved only
recently, with the derivation of a bulk expression for $\mb$ directly
in terms of the extended Bloch
states~\cite{xiao05,timo05,ceresoli06,shi07}.
In previous derivations the crystal was taken to be under shorted
electrical boundary conditions. We shall extend the derivation given
in \cite{timo05,ceresoli06} to the case where a static homogeneous
electric field $\eb$ is present, so that the full Hamiltonian reads
\beq
\label{eq:ham}
\ch=\ch^0+e\eb\cdot\r.
\eeq
The derivation, carried out for an insulator with $N$ valence bands
within the independent-particle approximation, involves transforming
the set of occupied eigenstates $\ket{\psi_i}$ of $\ch$ into a set of
Wannier-type (i.e., localized and orthonormal) orbitals $\ket{w_i}$
and expressing $\mb(\eb)$ in the Wannier representation. This is done
at first for a finite sample cut from a periodic crystal, and
eventually the thermodynamic limit is taken at fixed field.
Before continuing, two remarks are in order. First, the assumption
that it is possible to construct well-localized Wannier functions
(WFs) spanning the valence bands is only valid if the Chern invariants
of the valence bandstructure vanish identically~\cite{timo06}. This
requirement is satisfied by normal band insulators as well as by $Z_2$
topological insulators, but not by quantum anomalous Hall
insulators~\cite{haldane88}, which thus far remain
hypothetical. Second, because of Zener tunnelling, an insulating
crystal does not have a well-defined ground state in a finite electric
field. Nonetheless, upon slowly ramping up the field to the desired
value, the electron system remains in a quasistationary state which
is, for all practical purposes, indistinguishable from a truly
stationary state. This is the state we shall consider in the ensuing
derivation. As discussed in \cite{souza02,souza04}, it is Wannier-
and Bloch-representable, even though the Hamiltonian \eref{eq:ham} is
not lattice-periodic.
\subsection{$k$-space expression}
\label{sec:finite-field-k}
Our derivation of a $k$-space (bulk) expression for $\mb(\eb)$ is carried out
mostly in real space, using a Wannier representation. It is only in
the last step that we switch to reciprocal space, by expressing the
crystalline WFs $\ket{\R n}$ in terms of the cell-periodic Bloch
functions $\ket{u_{n\k}}$ via \cite{mv97}
\beq
\ket{\R n}=V_{\rm c}\int [\rmd k]\rme^{\rmi\k\cdot(\r-\R)}\ket{u_{n\k}},
\eeq
where $\R$ is a lattice vector, $V_{\rm c}$ is the unit-cell volume,
$[\rmd k]\equiv \rmd^3k/(2\pi)^3$, and the integral is over the first
Brillouin zone.
We begin with a finite sample immersed in a field $\eb$, divide it up
into an interior region and a surface region, and assign each WF to
either one. The boundary between the two regions is chosen in such a
way that the fractional volume of the surface region goes to zero as
$V\rightarrow\infty$, but deep enough that WFs near the boundary are
bulk-like. Following \cite{timo05,ceresoli06}, equation
\eref{eq:M-finite} for the orbital magnetization can then be rewritten
as an interior contribution plus a surface contribution, denoted
respectively as the ``local circulation'' (LC) and the ``itinerant
circulation'' (IC). Remarkably, in the thermodynamic limit {\it both}
can be expressed solely in terms of the interior-region crystalline
WFs, or equivalently, in terms of the bulk Bloch functions, as shown
in the above references at $\eb=0$ and below for $\eb\not=
0$. Specifically, we shall show that \numparts
\beq
\label{eq:M-tot-a}
\mb=\lcb+\icOb+\iceb,
\eeq
where
\beq
\label{eq:M-lc-bulk}
\lc_a=-\gamma\epsilon_{abc}\im
\sum_{n}^N\,\int [\rmd k]\bra{\partial_b u_{n\k}}
\hk\ket{\partial_c u_{n\k}}
\eeq
is the contribution from the interior WFs,
\beq
\label{eq:M-ic0-bulk}
M^{{\rm IC},0}_a=-\gamma\epsilon_{abc}\im
\sum_{nm}^N\,\int [\rmd k]\ip{\partial_b u_{n\k}}
{\partial_c u_{m\k}}H^0_{mn\k}
\eeq
is the part of the surface contribution
coming from the zero-field Hamiltonian, and
\beq
\label{eq:M-ice-bulk}
M^{{\rm IC},\eb}_a=-\gamma\epsilon_{abc}\im
\sum_{nm}^N\,\int [\rmd k]\ip{\partial_b u_{n\k}}
{\partial_c u_{m\k}}e\eb\cdot{\bi A}_{mn\k}
\eeq
\endnumparts
is the part of the surface contribution coming from the electric field
term in the Hamiltonian \eref{eq:ham}. In the above
expressions
$\gamma=-e/(2\hbar c)$,
\beq
\label{eq:hk}
\hk=\rme^{-\rmi\k\cdot\r}\ch^0 \rme^{\rmi\k\cdot\r},
\eeq
\beq
\label{eq:h0}
H^0_{mn\k}=\bra{u_{m\k}}\hk\ket{u_{n\k}},
\eeq
and ${\bf A}_{mn\k}$ is the Berry connection matrix defined in
equation \eref{eq:Amnb} below.
Having stated the result we now present the derivation, starting with
the interior contribution $\lcb$. Using $[r_i,r_j]=0$, the velocity
operator $\v=(\rmi/\hbar)[\ch,\r]$ becomes $(\rmi/\hbar)[\ch^0,\r]$, so that the
circulation operator $\r\times\v$ is unaffected by the electric
field. It immediately follows that the local circulation part $\lcb$
is given in terms of the field-polarized states $\ket{u_{n\k}}$ by
the same expression, equation~\eref{eq:M-lc-bulk}, as was derived in
\cite{ceresoli06} for the zero-field case.
Consider now the contribution
$\icb=\icOb+\iceb$ from the surface WFs $\ket{w_s}$.
For large samples it takes the form~\cite{ceresoli06}
\beq
\icb=-\frac{e}{2cN_{\rm c}V_{\rm c}}\sum_s^{\rm surf}\,\r_s\times\v_s,
\eeq
where $N_{\rm c}$ is the number of crystal cells of volume $V_{\rm c}$,
$\r_s=\bra{w_s}\r\ket{w_s}$, and
\beq
\label{eq:v_s}
\v_s=\bra{w_s}\v\ket{w_s}=\frac{2}{\hbar}\im\bra{w_s}\r \ch\ket{w_s}.
\eeq
Note that $\ch\ket{w_s}$ already belongs to the occupied manifold
spanned by $P=\sum_j^{\rm occ}\,\ket{w_j}\bra{w_j}$, since we
assume a (quasi)stationary state. Thus we can insert a $P$
between $\r$ and $\ch$ above, and using \eref{eq:ham}
we obtain
\beq
\v_s=
\sum_j^N\,\left(
\v_{\langle js\rangle}^0+\v_{\langle js\rangle}^\eb
\right),
\eeq
where $\v_{\langle js\rangle}^0=(2/\hbar)\im[\r_{sj}\ch^0_{js}]$ is the
same as in \cite{timo05,ceresoli06}
and $\v_{\langle js\rangle}^\eb=(2e/\hbar)\im[\r_{sj}(\r_{js}\cdot\eb)]$
is a new term.
The reasoning~\cite{timo05,ceresoli06} by which $\icb$ can be recast in
terms of the bulk WFs $\ket{\R n}$ relies on the exponential
localization of the WFs and on certain properties of $\v_{\langle js\rangle}^0$
(antisymmetry under $j\leftrightarrow s$ and invariance under lattice
translations deep inside the crystallite) which are shared by
$\v_{\langle js\rangle}^\eb$. Hence we can follow similar steps
as in those works, arriving at
\beq
\label{eq:ice_a}
\ice_a=\frac{e}{4cV_{\rm c}}\epsilon_{abc}
\sum_\R\sum_{mn}^N\, v^{\eb}_{\langle
\O m,\R n\rangle,b}R_c,
\eeq
and similarly for $\icO_a$ with $v^0$ substituting for $v^{\eb}$. The
latter is identical to the expression for $\ic_a$ valid at
$\eb=0$~\cite{timo05,ceresoli06}, and upon converting to $k$-space
becomes \eref{eq:M-ic0-bulk}.
Let us now turn to $\ice_a$ and write \eref{eq:ice_a} as
$(e^2/2c\hbar V_{\rm c})\epsilon_{abc}\e_d\im W_{bd,c}$ where
\beq
\label{eq:W_bdc}
W_{bd,c}=\sum_\R\sum_{mn}^N\,\bra{\R n}r_b\ket{\O m}
\bra{\O m}r_d\ket{\R n}R_c.
\eeq
In order to recast this expression as a $k$-space integral it is
useful to introduce the $N\times N$ Berry connection matrix
\beq
\label{eq:Amnb}
A_{mn\k,b}=\rmi\bra{u_{m\k}}\partial_b u_{n\k}\rangle=A_{nm\k,b}^*,
\eeq
where $\partial_b\equiv\partial/\partial k_b$.
It satisfies the
relation~\cite{mv97,blount1962}
\beq
\label{eq:A-w1}
\bra{\R n}r_b\ket{\O m}=V_c\int[\rmd k]A_{nm\k,b}\rme^{\rmi\k\cdot\R}.
\eeq
We also need
\beq
\label{eq:A-w2}
R_c\bra{\R n}r_d\ket{\O m}=\rmi V_c\int [\rmd k](\partial_c A_{nm\k,d})\rme^{\rmi\k\cdot\R},
\eeq
which follows from \eref{eq:A-w1}.
Using these two relations, \eref{eq:W_bdc} becomes
\beq
W_{bd,c}=\rmi V_c\sum_{mn}^N\int[\rmd k]A_{mn\k,d}\partial_c A_{nm\k,b},
\eeq
and we arrive at \eref{eq:M-ice-bulk}.
The sum of equations~\eref{eq:M-lc-bulk}--\eref{eq:M-ice-bulk} gives the
desired $k$-space expression for $\mb(\eb)$.
In the limit $\eb\rightarrow 0$ the term $\iceb$ vanishes, and
equation (31) of \cite{ceresoli06} is recovered.
We have implemented \eref{eq:M-lc-bulk}--\eref{eq:M-ice-bulk} for the
tight-binding model of \ref{app:model}. Since for small electric
fields $\mb(\eb)$ differs only slightly from $\mb(0)$, in order to
observe the effect of the electric field we consider differences in
magnetization rather than the absolute magnetization. Therefore, in
all our numerical tests we evaluated the OMP tensor
$\alpha_{da}$. With the help of
\eref{eq:M-lc-bulk}--\eref{eq:M-ice-bulk} we calculated it as $\Delta
M_a/\Delta \e_d$, using small fields $\e_d=\pm 0.01$. We then
repeated the calculation on finite samples cut from the bulk crystal,
using \eref{eq:M-finite} in place of
\eref{eq:M-lc-bulk}--\eref{eq:M-ice-bulk}. Figure~\ref{fig:al_zz}
shows the value of the $zz$ and $zy$ components of $\alpha$ plotted as
a function of the parameter $\varphi$, the phase of one of the complex
hopping amplitudes (see \ref{app:model} for details). The very precise
agreement between the solid and dashed lines confirms the correctness
of the $k$-space formula. The same level of agreement was found for
the other components of $\alpha$.
\begin{figure}
\centering\includegraphics{figure1.eps}
\caption{The $zz$ and $zy$ components of the OMP tensor
$\alpha$ of the tight-binding model described in
\ref{app:model}, as a function of the parameter $\varphi$.
The two lower bands are treated as occupied. Solid line:
extrapolation from finite-size samples using numerical
differentiation of the finite-field magnetization calculated from
\eref{eq:M-finite}. Dashed line: numerical differentiation of
the finite-field magnetization calculated using
\eref{eq:M-lc-bulk}--\eref{eq:M-ice-bulk} discretized on a
$k$-space grid. Open circles: linear-response calculation in
$k$-space using discretized versions of
\eref{eq:theta-cs}--\eref{eq:alpha-ic}. }
\label{fig:al_zz}
\end{figure}
\subsection{Gauge-invariant decomposition}
\subsubsection{Periodic crystals}
\label{sec:finite-field-k-inv}
\Eref{eq:M-tot-a} for $\mb(\eb)$ is valid in an arbitrary
gauge, that is, the sum of its three terms given by
\eref{eq:M-lc-bulk}--\eref{eq:M-ice-bulk} --~but not each term
individually~-- remains invariant under a unitary transformation
\beq
\label{eq:gauge}
\ket{u_{n\k}}\rightarrow \sum_m^N\,\ket{u_{m\k}}U_{mn\k}
\eeq
among the valence-band states at each $\k$. In order to make the
gauge invariance of \eref{eq:M-tot-a} manifest, it is convenient to
first manipulate it into a different form, given in terms of certain
canonical objects which we now define. We begin by introducing the
covariant $k$-derivative of a valence state~\cite{souza04},
\beq
\label{eq:cov-der}
\ket{\partb u_{n\k}} =Q_\k\,\ket{\parb u_{n\k}},
\eeq
where $Q_\k=1-P_\k$ and
\beq
\label{eq:proj}
P_\k=\sum_{j=1}^N\ket{u_{j\k}}\bra{u_{j\k}}.
\eeq
The covariant and ordinary derivatives are related by
\beq
\label{eq:cov_deriv}
\ket{\parb u_{n\k}} = \ket{\partb u_{n\k}}
-\rmi \sum_m^N A_{mn\k,b}\ket{u_{m\k}}.
\eeq
The generalized metric-curvature tensor is~\cite{mv97}
\beq
\label{eq:metric-curv}
F_{nm\k,bc}=\ip{\partb u_{n\k}}{\partc u_{m\k}} =
F_{mn\k,cb}^*.
\eeq
Viewed as an $N\times N$ matrix over the band indices, $F$ is
gauge-covariant, changing as
\beq
\label{eq:gauge-covariant}
F_{nm\k,bc}\rightarrow \left(U_\k^\dagger F_{\k,bc} U_\k\right)_{nm}
\eeq
under the transformation \eref{eq:gauge}. We also note the relation
\beq
\label{eq:F-ident}
\ip{\parb u_{n\k}}{\parc u_{m\k}} = F_{nm\k,bc}+(A_{\k,b}A_{\k,c})_{nm}.
\eeq
We shall make use of two more gauge-covariant objects,
\beq
H^0_{nm\k,b}=\rmi\me{u_{n\k}}{\hk}{\partb u_{m\k}}
\label{eq:Xa}
\eeq
and
\beq
\label{eq:Xab}
H^0_{nm\k,bc}=\me{\partb u_{n\k}}{\hk}{\partc u_{m\k}},
\eeq
which enter the relation
\beq
\label{eq:Xnmab-ident}
\fl\me{\parb u_{n\k}}\hk{\parc u_{m\k}}=
H^0_{nm\k,bc
+\left[A_{\k,b} H^0_{\k,c} +
\left( H^0_{\k,b}\right)^\dagger A_{\k,c}+
A_{\k,b}\hk A_{\k,c}\right]_{nm}.
\eeq
Coming back to equations \eref{eq:M-tot-a}--\eref{eq:M-ice-bulk}, for
$\lc_a$ we use \eref{eq:Xnmab-ident} and for $\ic_a$ we use
\eref{eq:F-ident}, leading to
\beq
\fl M_a=-\gamma\epsilon_{abc}\int [\rmd k]\,
\tri\Big[
H^0_{bc} + 2A_bH^0_c
+H^0F_{bc
+ e\e_dA_dF_{bc} +
e\e_d A_d A_b A_c
\Big],
\eeq
where ``tr'' denotes the electronic trace over the occupied valence
bands and we have dropped the subscript $\k$. The second term can be
rewritten using
\beq
\label{eq:H0nmc}
H^0_{nm,c}=-e\e_dF_{nm,dc}.
\eeq
(To obtain this relation start from the generalized Schr\"odinger equation
satisfied by the valence states at $\eb\not=0$~\cite{nunes01},
\beq
H^0\ket{u_n}=\sum_m^N\,
(H^0_{mn}+e\eb\cdot{\bi A}_{mn})
\ket{u_m}-
\rmi e\e_d\ket{\partial_d u_n},
\eeq
and multiply through by $\bra{\partc u_m}$.)
Let us define the quantities
\beq
\label{eq:M-lc-a}
\lct_a=-\gamma\epsilon_{abc}\int [\rmd k]\tri \left[ H^0_{bc}\right],
\eeq
\beq
\label{eq:M-ic-a}
\ict_a=-\gamma\epsilon_{abc}\int [\rmd k]\tri \left[ H^0F_{bc}\right],
\eeq
and
\beq
\label{eq:M-cs-a}
\mcs_a=-e\gamma\epsilon_{abc}\e_d\int [\rmd k]\tri
\bigl[
2A_bF_{cd}+F_{bc}A_d
+A_bA_cA_d
\bigr].
\eeq
The total magnetization is given by their sum
\numparts
\beq
\label{eq:M-tot}
M_a=\lct_a+\ict_a+\mcs_a.
\eeq
Referring to \eref{eq:metric-curv} and \eref{eq:Xab} the first
two terms read, in a more conventional notation,
\beq
\label{eq:M-lc}
\lct_a=-\gamma\epsilon_{abc}\int [\rmd k]\sum_n^N\,\im
\bra{\partb u_{n\k}}\hk\ket{\partc u_{n\k}}
\eeq
and
\beq
\label{eq:M-ic}
\ict_a=-\gamma\epsilon_{abc}\int [\rmd k]\sum_{nm}^N\,\im
\left(
H^0_{nm\k}\bra{\partb u_{m\k}}\partc u_{n\k}\rangle
\right).
\eeq
These are the only terms that remain in the limit $\eb\rightarrow 0$,
in agreement with equation~(43) of \cite{ceresoli06}.
At finite field they depend on $\eb$ implicitly via the wavefunctions.
We now show that the
term $\mcsb$, which gathers all the contributions with an
explicit dependence on $\eb$, can be recast as
\beq
\label{eq:M-cs}
\mcs_a=e\gamma\e_a\int [\rmd k]\epsilon_{ijk}\tr\left[A_i\partial_jA_k-
\frac{2\rmi}{3}A_iA_jA_k\right].
\eeq
\endnumparts
To do so it is convenient to introduce the Berry curvature tensor
\beq
\label{eq:omega}
\Omega_{nm,ab}=\rmi F_{nm,ab}-\rmi F_{nm,ba}=-\Omega_{nm,ba},
\eeq
where $F_{nm,ab}$ was defined in \eref{eq:metric-curv}.
A few lines of algebra show that
\beq
\label{eq:omega-b}
\Omega_{nm,ab} = \para A_{nm,b} - \parb A_{nm,a} -\rmi[A_a,A_b]_{nm}.
\eeq
In order to go from \eref{eq:M-cs-a} to \eref{eq:M-cs}, use
\eref{eq:omega} to write $\tri[F_{bc}A_d]$ as
$-\frac12\tr[A_d\Omega_{bc}]$ and $-2\tri[A_bF_{dc}]$ as
$\tr[A_b\Omega_{dc}]$, and then replace $\Omega_{nm,bc}$ in these
expressions with $\epsilon_{abc}\Omega_{nm,a}$,
where $\Omega_{nm,a}=\frac12\epsilon_{abc}\Omega_{nm,bc}$ is the Berry
curvature tensor written in axial-vector form. This leads to
\beq
\label{eq:M-cs-alt}
\mcs_a=
e\gamma\int [\rmd k]
\big(
\e_a\tr[\bOmega\cdot{\bi A}]-
\e_d\epsilon_{abc}\tri[A_bA_cA_d]
\big).
\eeq
The first term is parallel to the field, and can be rewritten with the
help of \eref{eq:omega-b}:
\beq
\label{eq:omega-A}
\tr[\bOmega\cdot{\bi A}]=\epsilon_{ijk}
\tr[A_i\partial_j A_k-\rmi A_iA_jA_k].
\eeq
While not immediately apparent, the second term in
\eref{eq:M-cs-alt} also points along the field. To see this,
write
\beq
\fl\sum_{bcd}\,
\e_d\epsilon_{abc}\tri[A_bA_cA_d]=
\e_a\sum_{bc}\,\epsilon_{abc}\tri[A_aA_bA_c
+
\sum_{d\not= a}\sum_{bc}\,\epsilon_{abc}\tri[A_bA_cA_d],
\eeq
where we suspended momentarily the implied summation convention. The
last term vanishes because the factor $\epsilon_{abc}$ forces $d\not=
a$ to equal either $b$ or $c$, producing terms such as
$\tri[A_bA_bA_c]$ which vanish identically as $A_b$ is
Hermitian. Rewriting $\e_a\sum_{bc}\,\epsilon_{abc}\tri[A_aA_bA_c]$ as
$(\e_a/3)\sum_{ijk}\epsilon_{ijk}\tri[A_iA_jA_k]$ and restoring the
summation convention, we arrive at \eref{eq:M-cs}.
Equations~\eref{eq:M-lc}--\eref{eq:M-cs}, which constitute the main
result of this section, are separately gauge-invariant. For
$\wt{\mb}^{\rm LC}$ and $\wt{\mb}^{\rm IC}$ this is apparent already
from \eref{eq:M-lc-a} and \eref{eq:M-ic-a}, whose {\it integrands} are
gauge-invariant, being traces over gauge-covariant matrices. In
contrast, equation~\eref{eq:M-cs} for $\mcsb$ only becomes invariant
after taking the integral on the right-hand-side over the entire
Brillouin zone (the integrand being familiar from differential
geometry as the Chern-Simons 3-form~\cite{nakahara,csforms}).
The Chern-Simons contribution \eref{eq:M-cs} has several remarkable
features: (i) as already noted, it is perfectly isotropic, remaining
parallel to $\eb$ for arbitrary orientations of $\eb$ relative to the
crystal axes; (ii) being isotropic, it vanishes in less than three
dimensions, which intuitively can be understood because already in two
dimensions polarization must be in the plane of the system and
magnetization must be out of the plane; (iii) for $N>1$ valence bands
it is a multivalued bulk quantity with a quantum of arbitrariness
$(e^2/hc)\e_a$, a fact that is connected with the possibility of a
cyclic adiabatic evolution that would change \eref{eq:theta-cs} below
for $\theta$ by $2\pi$~\cite{qi08}.
We have repeated the calculation of the OMP presented in
figure~\ref{fig:al_zz} using \eref{eq:M-lc}--\eref{eq:M-cs} instead of
\eref{eq:M-lc-bulk}--\eref{eq:M-ice-bulk}, finding excellent agreement
between them. The electric field derivative of the
decomposition~\eref{eq:M-tot} gives the corresponding decomposition of
the OMP tensor \eref{eq:alpha},
\beq
\label{eq:omp-tot}
\alpha=\alc+\aic+\acs,
\eeq
where each term is also gauge-invariant. The $zz$ components of these
terms are plotted separately in figure~\ref{fig:al_zz_c}.
\begin{figure}
\centering\includegraphics{figure2.eps}
\caption{Decomposition of the $\alpha_{zz}$ curve in
figure~\ref{fig:al_zz} into the gauge-invariant contributions
$\alc_{zz}$ (solid lines), $\aic_{zz}$ (dashed line), and
$\acs_{zz}$ (dotted line), calculated in $k$-space using finite
differences in $\eb$. Symbols denote the same contributions
evaluated for bounded samples, also using finite differences.
}
\label{fig:al_zz_c}
\end{figure}
\subsubsection{Finite samples}
It is natural to ask whether the gauge-invariant decomposition of the
orbital magnetization given in equation \eref{eq:M-tot} can be made
already for finite samples, before taking the thermodynamic limit and
switching to periodic boundary conditions. This has previously been
done in the case $\eb=0$, where $\mcsb=0$ and $\lctb$ and $\ictb$ take
the form~\cite{souza08}
\numparts
\beq
\label{eq:M-lcfin}
\lct_a=\frac{e}{2\hbar cV}\epsilon_{abc}\mathrm{Im\,}\Tr\,[Pr_bQ\ch^0Qr_c]
\eeq
and
\beq
\label{eq:M-icfin}
\ict_a=\frac{e}{2\hbar cV}\epsilon_{abc}\mathrm{Im\,Tr}\,[P\ch^0Pr_bQr_c].
\eeq
Here $P$ and $Q=1-P$ are the projection operators onto the occupied and
empty subspaces, respectively, and
``Tr'' denotes the electronic trace over the entire Hilbert space.
These two expressions, which are manifestly gauge-invariant, remain
valid at finite field, reducing to \eref{eq:M-lc} and
\eref{eq:M-ic} in the thermodynamic limit.
We now complete this picture for $\eb\not= 0$ by showing that the
remaining contribution $\mcsb=\mb-\lctb-\ictb$ can also be written in
trace form, as
\beq
\label{eq:cs-finite}
\mcs_a=-\frac{e^2}{3\hbar cV}\e_a\epsilon_{ijk}\Tri\,[Pr_iPr_jPr_k].
\eeq
\endnumparts
We first recast the orbital magnetization \eref{eq:M-finite} as
\beq
M_a=-\frac{e}{2cV}\epsilon_{abc}\mathrm{Tr}\,[Pr_bv_c]
=\frac{e}{2\hbar cV}\epsilon_{abc}\Tri\,\bigl[Pr_b\ch^0r_c\bigr]
\eeq
and then subtract \eref{eq:M-lcfin} and \eref{eq:M-icfin}
from it to find, after some manipulations,
\beq
M^{\mathrm{CS}}_a=
-\frac{e}{\hbar cV}\epsilon_{abc}\mathrm{Im\,Tr}\,[Q\ch^0Pr_bPr_c].
\eeq
Replacing $\ch^0$ with $\ch-e\e_dr_d$ and using $Q\ch P=0$,
\beq \mcs_a
=-\frac{e^2}{\hbar cV}\epsilon_{abc}\e_d\mathrm{Im\,Tr}\,[Pr_dPr_bPr_c].
\eeq
The imaginary part of the trace vanishes if any two of the indices
$b$, $c$, or $d$ are the same, and therefore $d$ must be equal to $a$.
Using the cyclic property
we conclude that all non-vanishing terms in the sum over $b$ and $c$
are identical, leading to \eref{eq:cs-finite}. This part of the
field-induced magnetization is clearly isotropic, with a coupling
strength (see equation~\eref{eq:alpha-iso}) given by
\beq
\label{eq:theta-cs-finite}
\tcs=-\frac{4\pi^2}{3V}\epsilon_{ijk}\Tri[Pr_iPr_jPr_k].
\eeq
This expression can assume nonzero
values because the Cartesian components of the projected position
operator $P\r P$ do not commute~\cite{mv97}.
We have used \eref{eq:M-lcfin}--\eref{eq:cs-finite} to evaluate
the OMP contributions $\alc$, $\aic$, and $\acs$ for finite samples,
finding excellent agreement with the $k$-space calculations using
\eref{eq:M-lc}--\eref{eq:M-cs}. As an example, the finite-sample results
for the $zz$ component are plotted as the symbols in figure~\ref{fig:al_zz_c}.
\section{Linear-response expression for the OMP tensor}
\label{sec:linear-response}
In sections~\ref{sec:finite-field-k} and \ref{sec:finite-field-k-inv}
expressions were given for evaluating $\mb(\eb)$ under periodic
boundary conditions. Used in conjunction with finite-field
{\it ab-initio} methods for periodic insulators~\cite{umari02,souza02},
they allow to calculate the OMP tensor by finite differences.
Alternatively, the electric field may be treated
perturbatively~\cite{nunes01}. With this approach in mind, we shall now
take the $\eb$-field derivative in \eref{eq:alpha} analytically
and obtain an expression for the OMP tensor which is amenable to
density-functional perturbation-theory implementation~\cite{baroni01}.
It should be kept in mind that in the context of self-consistent-field
(SCF) calculations the ``zero-field'' part of the Hamiltonian
(\ref{eq:ham}),
\beq
\label{eq:h-scf}
\ch^0=-\frac{\hbar^2}{2m}\nabla^2+V_{\rm SCF}(\r),
\eeq
does depend on $\eb$ implicitly, through the charge density. As we
will see, this dependence gives rise to additional local-field
screening terms in the expression for the OMP.
\begin{figure}
\centering\includegraphics{figure3.eps}
\caption{Contributions to the isotropic OMP from the
Chern-Simons term $\acs$ and from the Kubo-like terms $\alc$ and
$\aic$, expressed in terms of the dimensionless coupling strength
$\theta$ in \eref{eq:alpha-iso}. Model parameters are the same
as for figure~\ref{fig:al_zz}.}
\label{fig:theta}
\end{figure}
We shall only consider the case where the OMP is calculated for a
reference state at zero field, which we indicate by a superscript
``0.'' Upon inserting \eref{eq:M-tot} into \eref{eq:alpha} we obtain
the three gauge-invariant OMP terms in \eref{eq:omp-tot}. The term
$\acs$ is clearly of the isotropic form \eref{eq:alpha-iso}, with
\numparts
\beq
\label{eq:theta-cs}
\tcs=-\frac{1}{4\pi}\int \rmd^3k\,\epsilon_{ijk}\tr\left[A_i^0\partial_jA_k^0-
\frac{2\rmi}{3}A_i^0A_j^0A_k^0\right].
\eeq
This is the same expression as obtained previously by
heuristic methods~\cite{qi08,essin09}\footnote{An inconsistency
in the published literature regarding the numerical prefactor in
\eref{eq:theta-cs} has been resolved: see~\cite{essin09E}}.
The other two terms were not considered in the previous works. They are
\begin{eqnarray}
\label{eq:alpha-lc}
\alc_{da}=\gamma\epsilon_{abc}\int [\rmd k]\,
\sum_n^N\,\mathrm{Im}
\Big(
2\bra{\partb u^0_{n\k}}(\partial_c\hk)\ket{\wt{\partial}_Du_{n\k}^0}\nonumber\\
\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;
\;\;
-\bra{\partb u^0_{n\k}}(\partial_D \hk)\ket{\wt{\partial}_cu_{n\k}^0}
\end{eqnarray}
and
\begin{eqnarray}
\label{eq:alpha-ic}
\aic_{da}=\gamma\epsilon_{abc}\int [\rmd k]\,
\sum_{mn}^N\,\mathrm{Im}
\Big(
2\bra{\partb u_{n\k}^{0}}\wt{\partial}_D u_{m\k}^0\rangle
\bra{u_{m\k}^0}(\partial_c\hk)\ket{u_{n\k}^0}\nonumber\\
\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\,\,
-\bra{\widetilde{\partial}_b u_n^0}\widetilde{\partial}_c u_m^0\rangle
\bra{u_m^0}(\partial_D \hk)\ket{u_n^0}
\Big),
\end{eqnarray}
\endnumparts
where $\partial_D$ denotes the field-derivative $\partial/\partial
\e_d$ and
\beq
\label{eq:proj-field-der}
\ket{\wt{\partial}_D u_{n\k}^0}\equiv
Q_\k\left. \ket{\partial_D u_{n\k}}\right|_{\eb=0}
\eeq
are the first-order field-polarized states projected onto the
unoccupied manifold. The terms containing $\partial_D \hk$ describe
the screening by local fields. They vanish for tight-binding models
such as the one in this work, but should be included in
self-consistent calculations, in the way described in \cite{baroni01}.
We shall sometimes refer to $\alc$ and $\aic$ as `Kubo' contributions
because, unlike the Chern-Simons term, they involve first-order
changes in the occupied orbitals and Hamiltonian, in a manner
reminiscent of conventional linear-response theory.\footnote{The
terminology `Kubo terms' for $\alc$ and $\aic$ is only meant to be
suggestive. A Kubo-type linear-response calculation of the OMP
should produce all three terms, including $\acs$.}
Equations~\eref{eq:theta-cs}--\eref{eq:alpha-ic} are the main result
of this section. The derivation of \eref{eq:alpha-lc} and
\eref{eq:alpha-ic} is somewhat laborious and is sketched in
\ref{app:derivation}. We emphasize that the Kubo-like terms, besides
endowing the tensor $\alpha$ with off-diagonal elements, also
generally contribute to its trace, which therefore is not purely
geometric. Writing the isotropic part of the OMP response in the form
\eref{eq:alpha-iso}, we then have
\beq
\theta=\tcs+\tkubo.
\eeq
The two contributions are plotted for our model in
figure~\ref{fig:theta}. Moreover, the open circles in
figure~\ref{fig:al_zz} show the $zz$ and $zy$ components of the OMP
tensor computed from \eref{eq:theta-cs}--\eref{eq:alpha-ic},
confirming that the analytic field derivative of the magnetization was
taken correctly.
In the case of an insulator with a single valence band, the partition
\eref{eq:omp-tot} of the OMP tensor acquires some interesting
features. The terms $\aic$ and $\acs$ become purely itinerant, i.e.,
they vanish in the limit of a crystal composed of non-overlapping
molecular units, with one electron per molecule. Also, the first term
in the expression \eref{eq:alpha-ic} for $\aic$ --~the only term for
tight-binding models~-- becomes traceless, as can be readily verified
in a Hamiltonian gauge (where
$\hk\ket{u^0_{n\k}}=E^0_{n\k}\ket{u^0_{n\k}}$) with the help of the
perturbation theory formula~\cite{nunes01}
\beq
\label{eq:sternheimer-efield}
\ket{\wt{\partial}_Du_{n\k}^0}=\rmi e\sum_{m>N}
\frac{\ket{u_{m\k}^0}\bra{u_{m\k}^0}}{E_n^0-E_m^0}\ket{\partial_d u_{n\k}^0}.
\eeq
In order to verify these features numerically, we calculated the
various contributions treating only the lowest band of our
tight-binding model as occupied. The molecular limit was taken by
setting to zero the hoppings between neighbouring eight-site cubic
``molecules.''
\begin{figure}
\centering\includegraphics{figure4.eps}
\caption{Comparison between $\alpha_{zz}$ calculated treating the two
lowest bands as occupied (crosses) and the sum
$\alpha_{zz}^{(1)}+\alpha_{zz}^{(2)}$ (thick solid line), where
$\alpha_{zz}^{(n)}$ (thin solid lines) correspond to treating only
the lowest band ($n=1$, upper line) or the second-lowest band
($n=2$, lower line) as occupied. Model parameters are the same as
for figure~\ref{fig:al_zz} except that the second lowest on-site energy
in table~\ref{tab:tb_model} is raised from $-$6.0 to $-$5.0
in order to keep the two lowest bands well-separated.}
\label{fig:bsc}
\end{figure}
It could have been anticipated from the outset that
the Chern-Simons term \eref{eq:theta-cs} could not be the
entire expression for the OMP, based on the following
argument~\cite{essin10}.
Consider an insulator with $N>1$ valence bands, all of which are
isolated from one another. By looking at $\alpha_{da}$ as
$\partial P_d/\partial B_a$ one can argue that, since
each band carries a certain amount of polarization ${\bi P}^{(n)}$,
the total OMP should satisfy the relation
\beq
\label{eq:bsc}
\alpha=\sum_n^N\,\alpha^{(n)},
\eeq
where $\alpha^{(n)}$ is the OMP one would obtain by filling band $n$
while keeping all other bands empty. We shall refer to this property
as the ``band-sum-consistency'' of the OMP. It only holds exactly for
models without charge self-consistency (see the analytic proof in
\ref{app:band-sum-consistency}), but that suffices for the purpose of
the argument. We note that the Chern-Simons contribution
\eref{eq:theta-cs} alone is {\it not} band-sum-consistent, because the
second term therein vanishes for a single occupied band. Hence an
additional contribution, also band-sum-inconsistent, must necessarily
exist. Indeed, both $\alc$ and $\aic$ are band-sum-inconsistent, in
such a way that the total OMP satisfies \eref{eq:bsc}. This is
illustrated in figure~\ref{fig:bsc} for our tight-binding model.
\section{Summary and outlook}
In summary, we have developed a theoretical framework for calculating
the frozen-ion orbital-magnetization response (OMP) to a static
electric field. This development fills an important gap in the
microscopic theory of the magnetoelectric effect, paving the way to
first-principles calculations of the full response. While the OMP is
often assumed to be small compared to the lattice-mediated and
spin-magnetization parts of the ME response, there is no {\it a
priori} reason why it should always be so. In fact, in strong $Z_2$
topological insulators it is the only contribution that survives, and
the predicted value is large compared to that of prototypical
magnetoelectrics. Although the measurement of the $\theta=\pi$ ME
effect in topological insulators is challenging, as time-reversal
symmetry must be broken to gap the
surfaces~\cite{qi08,essin09,hasan10}, there may be other related
materials where those symmetries are broken already in the bulk. The
present formalism should be helpful in the ongoing computational
search for such materials with a large and robust OMP.
A key result of this work is a $k$-space expression for the orbital
magnetization of a periodic insulator under a finite electric field
$\eb$ (equations \eref{eq:M-tot-a}--\eref{eq:M-ice-bulk}, or
equivalently, \eref{eq:M-tot}--\eref{eq:M-cs}). In addition to the
terms \eref{eq:M-lc}--\eref{eq:M-ic} already present at zero
field~\cite{ceresoli06}, in three dimensions the field-dependent
magnetization comprises an additional purely isotropic `Chern-Simons'
term, given by \eref{eq:M-cs}. This new term depends explicitly on
$\eb$ and only implicitly on $\hk$, while the converse is true for the
other terms. Moreover, it is a multivalued quantity, with a quantum
of arbitrariness $\mb_0=\eb e^2/hc$ along $\eb$. Thus, the analogy
with the Berry-phase theory of electric
polarization~\cite{King-Smith,resta-review07}, where a similar
quantum arises, becomes even more profound at finite electric
field.
The Chern-Simons term $\mcsb$ is responsible for the geometric part of
the OMP response discussed in \cite{qi08,essin09} in connection with
topological insulators. We have clarified that in materials with
broken time-reversal and inversion symmetries in the bulk the CSOMP
does not generally constitute the full response, as the remaining
orbital magnetization terms, $\lctb$ and $\ictb$, can also depend
linearly on $\eb$. Their contribution to the OMP, given by
\eref{eq:alpha-lc}--\eref{eq:alpha-ic}, is twofold: (i) to modify the
isotropic coupling strength $\theta$; and (ii) to introduce an
anisotropic component of the response.
Another noteworthy result is equation~\eref{eq:theta-cs-finite} for
the Chern-Simons OMP of finite systems. One appealing feature of this
expression is that it allows one to calculate the CSOMP without the
need to choose a particular gauge. Instead, its $k$-space counterpart,
equation~\eref{eq:theta-cs}, requires for its numerical evaluation a
smoothly varying gauge for the Bloch states across the Brillouin zone.
\Eref{eq:theta-cs-finite} is also the more general of the two, as it
can be applied to noncrystalline or otherwise disordered systems.
We conclude by enumerating a few questions that are raised by the
present work.
Do the individual gauge-invariant OMP terms identified here in a
one-electron picture remain meaningful for interacting systems, and
can they be separated experimentally? (This appears to be the case
for $\lctb$ and $\ictb$ at $\eb=0$~\cite{souza08}.) How does the
expression \eref{eq:theta-cs}--\eref{eq:alpha-ic} for the linear OMP
response change when the reference state is under a finite electric
field $\eb$? Finally, we note that equation~\eref{eq:cs-finite} for
the CSOMP of finite systems has a striking resemblance to a formula
given by Kitaev~\cite{kitaev06} for the 2-D Chern invariant
characterizing the integer quantum Hall effect. Can this connection
be made more precise, in view of the fact that the quantum of
indeterminacy in $\tcs$ is associated with the possibility of changing
the Chern invariant of the surface layers? These questions are left
for future studies.
\ack
This work was supported by NSF Grant Nos. DMR 0706493 and 0549198.
Computational resources have been provided by NERSC. The authors
gratefully acknowledge illuminating discussions with Andrew Essin,
Joel Moore, and Ari Turner.
|
\section{Introduction}
Open string tachyon condensation has been studied from many viewpoints, see \cite{Sen:2004nf} for a review.
Here we consider a holographic (AdS/CFT) setup where the bulk theory contains an open string tachyon, and ask
what is the counterpart of tachyon condensation in the boundary field theory. We will identify
a sector of the boundary theory as a ``holographic open string field theory'' capturing the tachyon dynamics.
Since the bulk is weakly coupled when the boundary is strongly coupled, and viceversa,
we are bound to learn something new from their comparison.
We introduce the open string tachyon by adding to the $AdS_5 \times S^5$ background two probe $D7$ branes intersecting at general angles.
Probe branes are the familiar way to include a small number of fundamental flavors in the AdS/CFT correspondence \cite{Karch:2002sh}.
If the closed string background is supersymmetric,
it is possible, and often desirable, to consider configurations of probe branes that preserve some supersymmetry, as {\it e.g.} in \cite{Karch:2002sh, Sakai:2003wu, Kuperstein:2004hy,
Arean:2004mm, Ouyang:2003df, Canoura:2006es, Apreda:2006bu, Sieg:2007by, Penati:2007vj, Wang:2003yc, Nunez:2003cf, Karch:2000gx, DeWolfe:2001pq, Erdmenger:2002ex,
Skenderis:2002vf, Constable:2002xt}.
Instead, we are after supersymmetry breaking and the ensuing tachyonic instability. Another way to motivate
our work is then as a natural susy-breaking generalization of the standard supersymmetric setup of \cite{Karch:2002sh}.
This generalization is technically challenging, and the technical aspects have some interest of their own.
Intersecting brane systems have many other applications in string theory, from string phenomenology to string cosmology, and
the technical lessons learnt in our problem may be useful in those contexts as well.
The system that we study is as an open string analogue of the AdS/CFT
pairs involving closed string tachyons considered in \cite{Adams:2001jb, Dymarsky:2005nc, Dymarsky:2005uh, Pomoni:2008de}. Let us briefly review that analysis.
In all non-supersymmetric orbifolds of ${\cal N}= 4$ SYM, there is an instability for large $N$
and small 't Hooft coupling $\lambda$. The instability is
triggered by the renormalization of double-trace couplings, of the form $f \int d^4 x \;{\cal O}^2$,
where ${\cal O} \sim {\rm Tr} X^2$ is a scalar bilinear. At leading order for large $N$,
the 't Hooft coupling $\lambda$ is exactly marginal, but the double-trace coupling $f$ runs.
The one-loop beta function $\beta_f (f, \lambda) \equiv \mu \frac{\partial f}{\partial \mu}$
does not admit zeros for real values of $f$ \cite{Dymarsky:2005nc, Dymarsky:2005uh}, so conformal invariance is inevitably broken for arbitrarily small (but non-zero) $\lambda$.
On the field theory side, the instability can be seen in two equivalent ways. The most
direct is as the Coleman-Weinberg instability of the double-trace part
of the scalar potential, implying that the scalars $X$ must acquire a non-zero vev. Alternatively \cite{Pomoni:2008de},
we can insist
in formally preserving conformal invariance by tuning $f$ to the zero of its beta function, which is a complex number;
it then turns out the anomalous dimension $\Delta$ of ${\cal O}$ takes a complex value of the form $\Delta=2 + i b \lambda + O(\lambda^2)$.
By AdS/CFT, the bulk scalar field dual to ${\cal O}$ has $m^2 = \Delta (\Delta-4) = -4-b^2 \lambda^2 + O(\lambda^3)$
(in AdS units), and is thus a ``true'' tachyon, since its squared mass is below the Breitenlohner-Freedman \cite{Breitenlohner:1982bm}
stability bound $m^2_{BF} = -4$.
The field theory analysis holds for small 't Hooft coupling $\lambda$, when
the bulk string background is strongly curved and the direct evaluation of its spectrum difficult. By contrast
for $\lambda \to \infty $ calculations are easy in the bulk. The bulk analysis reveals that in some cases (orbifolds with fixed points on $S^5$)
the tachyonic instability persists at large $\lambda$, but it disappears
in others (freely acting orbifolds). The upshot is that AdS/CFT
makes interesting predictions both at weak and a strong coupling.
The $AdS_5 \times S^5$ background with two probe intersecting $D7$s can be viewed as
an open string version of this story. For general angles the bulk theory is unstable via condensation of
an open string tachyon, or at least this is the picture for large $\lambda$ where
we can calculate the string spectrum. In this paper we focus on the the field theory analysis at small $\lambda$,
with the goal of detecting the expected instability.
The first challenge is to write down the Lagrangian
of the dual field theory. As is well-known, adding $N_f$ parallel $D7$ branes to $AdS_5 \times S^5$ corresponds
to adding to ${\cal N} =4$ SYM action $N_f$ extra ${\cal N}=2$ hyper multiplet in the fundamental representation of the $SU(N)$ gauge group.
The resulting action preserves an ${\cal N}=2$ subalgebra of the original ${\cal N}=4$ supersymmetry algebra -- which
particular ${\cal N}=2$ being a matter of convention so long as it is the same for all the hyper multiplets.
Introducing relative angles between the $D7$ branes corresponds to choosing {\it different} embeddings for the ${\cal N}=2$ subalgebras
of each different hyper multiplet. In general supersymmetry will be completely broken, while for special angles ${\cal N}=1$
susy is preserved. When ${\cal N} =1$ is preserved we can use ${\cal N} =1$ superspace
to write the Lagrangian. When supersymmetry is completely broken the determination
of the Lagrangian turns out to be a difficult technical
problem that we are unable to solve completely. We cannot fix the quartic terms $\sim Q^4$
where $Q$ are the hyper multiplet scalars. The difficulty is related to the lack of an off-shell superspace formulation of ${\cal N}=4$ SYM.
Nevertheless, by making what we believe is a mild technical assumption, we can fix the {\it sign} of the classical quartic potential.
This is sufficient to argue that the theory is indeed unstable from the renormalization
of ``double-trace'' terms $f \int d^4 x \; {\cal O}^2$, where now ${\cal O} \sim \bar Q^\gaugeind{a} Q_\gaugeind{a}$ with $\gaugeind{a}=1, \dots N$ a color index.
We are now using ``double-trace'' in quotes since of course
the fields $Q$ are not matrices but vectors, but the logic is much the same. The renormalization of $f$ has the same
twofold interpretation as above. We identify the mesonic operator ${\cal O}$ as the dual of the open string tachyon
between the two $D7$ branes. The Coleman-Weinberg potential for $Q$ plays the role a holographic effective action for the tachyon.
\section{AdS/CFT with Flavor Branes Intersecting at General Angles}
We begin with a review the Karch-Katz setup \cite{Karch:2002sh},
where parallel probe $D7$ branes are used to engineer an ${\cal N} = 2$ supersymmetric field theory with flavor.
We then break supersymmetry by introducing a relative angle between the $D7$ branes. We derive the dual Lagrangian,
up to an ambiguity in the quartic potential for the fundamental scalars. We end the section with a review of the
basic bulk-to-boundary dictionary.
\begin{figure}[h]
\begin{center}
\includegraphics[height=4cm,angle=0]{rotationtable.eps}\\
\caption{\it The brane configuration.}
\end{center}
\end{figure}
\subsection{Parallel flavor branes}
We start with the familiar $D3/D7$ supersymmetric brane configuration
with $N$ ``color'' $D3$s and $N_f$ ``flavor'' $D7$s, arranged as shown in Figure 1. For
now $\theta_1 = \theta_2 = 0$, that is, all $D7$ branes are parallel to one another.
Taking the decoupling limit on the $D3$s wordvolume, the $D3$ branes are replaced by their near-horizon geometry.
If $N_f \ll N$, we can treat the $D7$ branes as probes in the $AdS_5 \times S^5$ background, neglecting their backreaction \cite{Karch:2002sh}.
This background preserves ${\cal N} = 2$ supersymmetry in four dimensions.
The dual field theory is ${\cal N}=4$ $SU(N)$ SYM coupled to $N_f$ ${\cal N} = 2$ hyper multiplets in the fundamental representation of the $SU(N)$ color group,
arising from the $D3$-$D7$ open strings. We are interested in the case of massless hyper multiplets, corresponding to the brane setup where the $D7$s
coincide with the $D3$s (at the origin of the 89 plane). After decoupling,
the probe $D7$s fill the whole $AdS_5$ and wrap an $S^3 \subset S^5$.
Let us briefly recall the field content of the boundary theory. A more detailed treatment and the full Lagrangian can be found in Appendix A.
The ${\cal N}=4$ vector multiplet consists of the gauge field $A_\mu$, four Weyl spinors $\lambda^A_\alpha$, $A = 1, \dots, 4$
and six real scalars $X_{m}$, $m=4, \dots, 9$ corresponding to the six transverse directions to the $D3$ branes. It
is convenient to represent the scalars as a self-dual antisymmetric tensor $X^{AB}$ of the $R$-symmetry group $SU(4)_R \cong Spin(6)$,
\begin{equation} \label{selfduality}
(X^{AB})^\dagger = \bar{X}_{AB} \equiv \frac{1}{2}\epsilon_{ABCD}\,X^{CD} \,.
\end{equation}
The explicit change of variables is
\begin{equation}
\label{Xmatrix}
X^{AB} = \frac{1}{\sqrt{2}} \left(\begin{array}{cc|cc}
0 & X_8 + iX_9 & X_6 + iX_7 & X_4 + iX_5
\\
-X_8 - iX_9 & 0 & X_4 - iX_5 & -X_6 + iX_7
\\
\hline
-X_6 - iX_7 & -X_4 + iX_5 & 0 & X_8 - iX_9
\\
-X_4 - iX_5 & X_6 - iX_7 &-X_8 + iX_9 & 0
\end{array}\right) \,.
\end{equation}
Each ${\cal N} =2$ flavor hyper multiplet consists of two Weyl spinors and two complex scalars,
\begin{equation}
\begin{array}{ccc} & \psi^i_\alpha &
\\ q^i & & \left( \tilde{q}_i \right)^\dagger
\\ & \left( \tilde{\psi}_{i\,\alpha} \right)^\dagger &
\end{array}
\end{equation}
Here $i = 1, \dots, N_f$ is the flavor index. The scalars form an $SU(2)_R$ doublet,
\begin{equation}
Q^{{\cal I}} \equiv \left(\begin{array}{c} q
\\
\tilde{q}^\dagger
\end{array}\right) \, , \quad {{\cal I}} = 1,2 \,.
\end{equation} The flavor hyper multiplets
are minimally coupled to the ${\cal N}=2$ vector multiplet that sits
inside the ${\cal N} = 4$ vector multiplet. This coupling breaks the $R$-symmetry
$SU(4)_R$ to $SU(2)_L \times SU(2)_R \times U(1)_R$, where $SU(2)_R \times U(1)_R$
is the $R$-symmetry of the resulting ${\cal N} = 2$ theory.
There is a certain arbitrariness
in the choice of embedding $SU(2)_L \times SU(2)_R \times U(1)_R \subset SU(4)_R \cong Spin(6)$.
This corresponds to the choice of
orientation of the whole stack of $D7$ branes in the 456789 directions
(we need to pick an ${\mathbb R}^4 \subset {\mathbb R}^6$).
For example if we choose the configuration of Figure 1, we identify
$SU(2)_L \times SU(2)_R \cong SO(4)$ with rotations in the $4567$ directions and $U(1)_R \cong SO(2)$ with a rotation on the $89$ plane.
A short calculation using
our parametrization of the scalars (\ref{Xmatrix}) shows that this
corresponds to
the following natural embedding of $SU(2)_L \times SU(2)_R \times U(1)_R \subset SU(4)_R$:
\begin{equation} \label{choice}
\begin{array}{c} 1
\\ 2
\\ 3
\\ 4;p[
\end{array} \,
\left(\begin{array}{cc|cc} SU_R(2) \times U(1)_R & & &
\\ & & & \\
\hline & & &
\\ & & & SU_L(2) \times U(1)_R^*
\end{array}\right) \,.
\end{equation}
Of course, any other choice would be equivalent, so long as it is performed simultaneously for all $D7$ branes.
With the choice (\ref{choice}), the ${\cal N} = 4$ vector multiplet splits into
the ${\cal N} = 2$ {vector multiplet}
\begin{equation}
\begin{array}{ccc} & A_\mu &
\\ \lambda^1_\alpha& & \lambda^2_\alpha
\\ & \frac{X_8 + i X_9}{\sqrt{2}} &
\end{array} \, ,
\end{equation}
and the
${\cal N} = 2$ {hyper multiplet}
\begin{equation}
\begin{array}{ccc} & \lambda^3_\alpha &
\\ \frac{X_4 + i X_5}{\sqrt{2}} & & \frac{X_6 + i X_7}{\sqrt{2}}
\\ & \lambda^4_\alpha &
\end{array} \,.
\end{equation}
The two Weyl spinors in the vector multiplet form an $SU(2)_R$ doublet
\begin{equation}
\Lambda_{{\cal I}} \equiv \left(\begin{array}{c} \lambda_1
\\
\lambda_2
\end{array}\right) \, , \quad {\cal I} = 1,2 \,,
\end{equation}
while the two spinors in the hyper multiplet form an $SU(2)_L$ doublet,
\begin{equation}
\hat{\Lambda}_{\hat {\cal I}} \equiv \left(\begin{array}{c} \lambda_3
\\
\lambda_4
\end{array}\right) \, , \quad \hat{{\cal I}} = 1,2 \,.
\end{equation}
We use ${\cal I}\,, {\cal J}\, \dots =1,2$ for $SU(2)_R$ indices and $\hat {\cal I} \,, \hat {\cal J}\, \dots= 1,2$ for $SU(2)_L$ indices.
To make the $SU(2)_L \times SU(2)_R$ quantum numbers of the scalars more transparent we also introduce
the $ 2 \times 2 $ complex matrix ${\cal X}_{{\cal I} \hat {\cal I}}$, defined as the off-diagonal block of $X^{A B}$,
\begin{equation} \label{calX}
{\cal X}^{\hat {\cal I} {\cal I} } = \left(\begin{array}{cc}
X_6 + iX_7 & X_4 + iX_5
\\
X_4 - iX_5 & -X_6 + iX_7
\end{array}\right) \,.
\end{equation}
Note that $ {\cal X}^{\hat {\cal I} {\cal I} } $ obeys the reality condition
\begin{equation}
\left( {\cal X}^{\hat {\cal I} {\cal I} } \right)^{*}= - {\cal X}_{\hat {\cal I} {\cal I} } = - \epsilon_{\hat {\cal I} \hat{\cal J}} \epsilon_{{\cal I} {\cal J}} {\cal X}^{\hat {\cal J} {\cal J}} \, .
\end{equation}
We summarize in the following table the transformation properties of the fields:
\begin{table}[h]
\begin{center}
\label{table}
\begin{tabular}{c|c|c|c|c|c}
& $SU(N)$ & $SU(N_f)$ & $SU(2)_L$ & $SU(2)_R$ & $U(1)_R$\\
\hline $A_{\mu}$ & Adj & $\bf{1}$ & $\bf{1}$ & $\bf{1}$ & 0
$\vphantom{\raisebox{3pt}{asymm}}$\\
$X^{12}$ & Adj & $\bf{1}$ & $\bf{1}$ & $\bf{1}$ & +2
$\vphantom{\raisebox{3pt}{asymm}}$\\
${\cal X}^{{\cal I} \hat {\cal I}} $ & Adj & $\bf{1}$ & {\bf 2} & $\bf{2}$ & 0
$\vphantom{\raisebox{3pt}{asymm}}$\\
$\Lambda_{{\cal I}}$ &Adj & $\bf{1}$ & {\bf 1} & $\bf{2}$ & +1
$\vphantom{\raisebox{3pt}{asymm}}$\\
$\hat{\Lambda}_{\hat{\cal I}}$ &Adj & $\bf{1}$ & {\bf 2} & $\bf{1}$ & --1
$\vphantom{\raisebox{3pt}{asymm}}$\\
$Q^{\cal I}$ & $\Box$ & $\Box$ & {\bf 1} & $\bf{2}$ & 0
$\vphantom{\raisebox{3pt}{asymm}}$\\
$\psi$ & $\Box$ & $\Box$ & {\bf 1} & $\bf{1}$ & --1
$\vphantom{\raisebox{3pt}{asymm}}$\\
$\tilde{\psi}$ & $\overline{\Box}$ & $\overline{\Box}$ & {\bf 1} & $\bf{1}$ & +1
$\vphantom{\raisebox{3pt}{}}$\\
\end{tabular}
\end{center}
\caption{\it Quantum numbers of the fields. }
\end{table}
\subsection{Rotating the flavor branes \label{rotate}}
We now describe a non-supersymmetric open string deformation of this background.
For simplicity we consider the case $N_f=2$. While keeping the two $D7$ branes coincident with the $D3$s in the 0123 directions,
we rotate them with respect to each other in the transverse six directions, see Figure 1.
There are two independent angles, so without loss of generality we may
perform a rotation of angle $\theta_1 = \theta_{49}$ in the 49 plane and a rotation of angle $\theta_2 = \theta_{85}$ in the 58 plane. For generic angles
supersymmetry is completely broken; for $\theta_1= \theta_2 $ it is broken to ${\cal N} = 1$.
As we rotate the branes, some $D7$-$D7'$ open string modes become tachyonic.
The main goal of this paper is to study this tachyonic instability from the viewpoint of the dual field theory.
On the field theory side, rotating the second brane $D7'$ amounts to choosing a different embedding of $SU(2)_R \subset SU(4)$
for the second hyper multiplet, while keeping the standard embedding (\ref{choice}) for the first.
In the Lagrangian, we must perform an $SU(4)$ rotation of the ${\cal N}=4$ fields
that couple to the second hyper multiplet, leaving the ones that couple to the first unchanged. The rotation is
of the form
\begin{equation} \label{rotation}
X'_m = {\cal R}^{ ({\bf 6} ) \;n}_{\;m} (\theta_{1}, \theta_{2}) \, X_n \, , \qquad \lambda'_A = {\cal R}^{ ({\bf 4}) \; B}_{\;A} (\theta_{1}, \theta_{2}) \, \lambda_B\, .
\end{equation}
The explicit form of the rotation matrices $ {\cal R}^{ ({\bf 6} )}$ and $ {\cal R}^{ ({\bf 4} )}$ is given in Appendix B.
Naively, the $Q^4$ terms are not affected by the rotation, but this is incorrect. This is seen clearly in ${\cal N}=1$ superspace.
The ${\cal N}=4$ multiplet is built out of three chiral multiplets $\Phi^a$, $a=1,2,3$ and one vector multiplet $V$.
The $Q^4$ terms arise from integrating out the auxiliary fields $F^a$ ($a=1,2,3$) of the chiral multiplets and $D$ of the vector multiplet,
which transform under the $SU(4)_R$ rotation. For example, a rotation that preserves
${\cal N} =1$ supersymmetry ($\theta_1 = \theta_2$) corresponds to a matrix ${\cal R}^{({\bf 4})} \subset SU(3)$,
which acts on $F^a$ leaving $D$ invariant. The correct Lagrangian is obtained by performing
the rotation on the $X_m$, $\lambda_A$ {\it and} $F^a$ fields that couple to the primed hyper multiplet,
and only then can the auxiliary fields be integrated out. The $Q^4$ terms get modified accordingly.
Under a more general $SU(4)_R$ rotation, the $F^a$ and $D$ auxiliary fields are expected to mix in a non-trivial fashion.
There exists a formalism developed in \cite{Marcus:1983wb, Gates:1983nr}
that provides the generic R-symmetry
transformations action in ${\cal N} =1$ superspace.
Unfortunately, for ${\cal N}=4$ supersymmetry we cannot rely on this formalism because the transformations do not close off-shell.
This technically involved point is explained in detail in Appendix C. There we also provide an ${\cal N}=2$ supersymmetry toy example where the formalism works perfectly since the ${\cal N}=2$ R-symmetry algebra closes off-shell.
To proceed, we parametrize our ignorance of the $Q^4$ terms. The exact form of the full Lagrangian, including the parametrized $Q^4$ potential, is spelled out in Appendix B.
Schematically, we write the $Q^4$ potential as
\begin{equation}
V_{Q^4} = Q^4_1 + Q^4_2 + \left(Q_1Q_2 \right)^2_F \,f\left( \theta_1 , \theta_2 \right) + \left(Q_1Q_2 \right)^2_D \,d\left( \theta_1 , \theta_2 \right) \, ,
\end{equation}
for some unknown functions $f(\theta_1, \theta_2)$ and $d(\theta_1, \theta_2)$. Here $Q_1$ and $Q_2$ are shorthands for the scalars in the first and second hyper multiplets
and the subscripts $F$ and $D$ refer to different ways to contract the indices, see (\ref{the potential}) for the exact expressions. The letters $F$ and $D$ are chosen as reminders
of the (naive) origin of the two structures from integrating out the ``rotated'' $F$ and $D$ ${\cal N}=1$ auxiliary fields,
but this form of the potential follows from rather general symmetry considerations, as we explain in Appendix B.
When $\theta_1 = \theta_2$, ${\cal N}=1$ supersymmetry is preserved and ${\cal N}= 1$ superspace allows to fix the two functions,
\begin{equation}
f(\theta, \theta) = \cos \theta \, ,\qquad d(\theta, \theta) =1 \,.
\end{equation}
For general angles, we can constrain $f$ and $d$ somewhat, using bosonic symmetries (see Appendix B),
but unfortunately we are unable to fix them uniquely.
The most important assumption we will make in the following is {\it positivity} of the classical potential, $V_{Q^4} \geq 0$,
implying $f(\theta_1, \theta_2) \leq 1$ and $d(\theta_1, \theta_2) \leq 1$ for all $\theta_1$, $\theta_2$.
Positivity would follow from the mere existence of any reasonable off-shell superspace formulation,
as the scalar potential would always be proportional to the square of the auxiliary fields,
even when supersymmetry
is broken by the relative R-charge rotation between the two hyper multiplets.\footnote{To illustrate how this would work we consider
in section C.2 ${\cal N}=2$ SYM theory coupled to two fundamental ${\cal N}=1$
chiral multiplets, with different choices of the two ${\cal N}=1$ subalgebras.}
Note also that the classical potential $V_{Q^4}$ is a homogeneous function of the $Q$s, so it is everywhere positive
if and only if it is bounded from below, which is another plausible requirement.
\subsection{Bulk-boundary dictionary}
The basic bulk-to-boundary dictionary for the parallel brane case has been worked
out in \cite{Aharony:1998xz, Kruczenski:2003be}. A brief review is in order.
In the closed string sector, Type IIB closed string fields map to
single-trace operators of ${\cal N} = 4$ SYM, as usual. In the open string sector, open string fields on the $D7$ worldvolume
map to gauge-singlet mesonic operators, of the schematic form $\bar Q X^n Q$, where $Q$ stands for a generic fundamental field and $X$ for a generic adjoint field.
The massless bosonic fields on the $D7$ worldvolume are a scalar $\Phi$
and a gauge field $(A_{\hat{\mu}} , A_{\hat{\alpha}})$, where $\hat{\mu}$ are $AdS_5$ indices and $\hat{\alpha}$ are $S^3$ indices.
Kaluza Klein reduction on the $S^3$ generates the following tower of states, labeled in terms of $\left(j_1\,,\,j_2 \right)_s$ representations of $SU(2)_L \times SU(2)_R \times U(1)_R$:
\begin{equation}
\Phi \rightarrow \Phi^\ell = \left(\frac{\ell}{2}\,,\,\frac{\ell}{2} \right)_2 \, ,\quad A_{\hat{\mu}} \rightarrow A_{\hat{\mu}}^\ell = \left(\frac{\ell}{2}\,,\,\frac{\ell}{2} \right)_0 \, ,\quad A_{\hat{\alpha}} \rightarrow A_{\pm}^\ell = \left(\frac{\ell \pm 1}{2},\,\frac{\ell \mp 1}{2} \right)_0 \,.
\end{equation}
(The longitudinal component of $A_{\hat{\alpha}} $ is not included because it can be gauged away). These states (and their fermionic partners, which we omit) can be organized
into short multiplets of the ${\cal N} = 2$ superconformal algebra,
\begin{equation}
\left(A_-^{\ell+1},A_{\hat \mu}^\ell,\Phi^\ell,A_+^{\ell-1} \right) \,, \quad \ell = 0,1,2, \dots
\end{equation}
of conformal dimensions
\begin{equation}
(\ell +2, \ell+3, \ell+3, \ell +4) \,.
\end{equation}
For $\ell = 0$ the $A_+$ state is absent. Note that all states in a given multiplet have the same $SU(2)_L$ spin, indeed the ${\cal N} = 2$ supercharges are neutral under $SU(2)_L$.
The lowest member of each multiplet, namely $A_-^{\ell + 1}$, is dual to the chiral primary operator
\begin{equation} \label{chiralmeson}
\bar Q_{\{ {\cal I} } {\cal X}_{ {\cal I}_1 \, \hat {\cal I}_1 } \dots {\cal X}_{ {\cal I}_\ell \, \hat {\cal I}_\ell } Q_{ {\cal J} \} } \, ,
\end{equation}
where
$Q^{\cal I}$
is the $SU(2)_{R}$ doublet of complex fundamental scalars. In (\ref{chiralmeson}) the $SU(2)_L$ and $SU(2)_R$ indices are separately symmetrized.
In particular for $\ell =0$, we have the triplet of mesonic operators
\begin{equation}
\label{triplet}
{\cal O}_{\bf{3}} \equiv \bar{Q}_{ \{ {\cal I} } Q_{{\cal J} \} } = \bar{Q}_{\cal I} Q^{\cal J} - \frac{1}{2}\bar{Q}_{\cal K} Q^{\cal K} \delta^{\cal J}_{\cal I} \,.
\end{equation}
The singlet operator
\begin{equation}
{\cal O}_{\bf{1}} \equiv \bar{Q}_{\cal I} Q^{\cal I} \, ,\quad
\end{equation}
is not a chiral primary and maps to a massive open string state.
\begin{figure}[t]
\begin{center}
\includegraphics[height=4cm,angle=0]{renoper.eps}\\
\caption{\it Diagrams contributing to the one-loop renormalization of the mesonic operators.}
\label{renormoper}
\end{center}
\end{figure}
In Appendix D we compute the one-loop dilatation operator acting on the basis of states ${\cal O} \equiv \bar{Q}_{\cal I} Q^{\cal J}$, evaluating
the diagrams schematically drawn in Figure 2. We find
\begin{equation}
\Gamma^{(1)} = \frac{\lambda}{4\pi^2} \,\mathbb{K} \, , \quad \mathbb{K}\equiv \delta^{\cal I}_{\cal J} \delta^{\cal K}_{\cal L} \,.
\end{equation}
The eigenstates are the triplet and the singlet, with eigenvalues
\begin{equation}
\gamma_{\bf{3}} = 0
\, ,\quad
\gamma_{\bf{1}} = \frac{\lambda}{2\pi^2} \,.
\end{equation}
As expected, the chiral triplet operator has protected dimension. At one-loop,
this result does not change as we turn on non-zero angles $\theta_1$ and $\theta_2$.
So far we have considered the case of a single $D7$ brane, or a single flavor. For multiple $D7$ branes
(multiple flavors) the Chan-Paton labels of the open strings are interpreted as the bifundamental flavor indices
of the mesonic operators, ${\cal O}^{ij}$, $i,j=1,\dots N_f$. In our setup, with $N_f=2$, the lowest mode of the open string with off-diagonal Chan-Paton labels,
which is the massless gauge field for parallel branes, becomes tachyonic as we turn on a relative angle between the $D7$s. In the dual field theory
we expect to find an instability associated with the operator ${\cal O}_{\bf 3}^{12}$, the lowest dimensional operator dual to the off-diagonal open string mode.
\section{``Double-trace'' Renormalization and the Open String Tachyon}
Our setup is an open string version of the phenomena studied in \cite{Adams:2001jb, Dymarsky:2005nc, Dymarsky:2005uh, Pomoni:2008de}.
Motivated by the work of \cite{Dymarsky:2005nc,Dymarsky:2005uh}, we considered in \cite{Pomoni:2008de} a generic large $N$, non-supersymmetric field theory with all matter in the adjoint (or bifundamental) representation. We further assumed the theory to be ``conformal in its single-trace sector", by which we mean that all single-trace couplings have vanishing beta function for large $N$. In such a theory, quantum effects induce double-trace couplings of the schematic form
\begin{equation}
\delta S = f \int d^4x \, {\cal O} \bar{{\cal O}} \, , \quad {\cal O} \sim \mbox{Tr}\, \phi^2\,,
\end{equation}
where $\phi$ is a scalar field. The beta function for $f$ may or may not admit a real fixed point. If $\beta_f$ has no real zeros, conformal invariance is broken.
A closely related phenomenon is the generation of a Coleman-Weinberg potential $\mathcal{V}\left( \langle \phi \rangle \right)$ \cite{Adams:2001jb}. It is not difficult to show \cite{Pomoni:2008de} that the symmetric vacuum $\langle \phi \rangle=0$ is stable if and only if $\beta_f$ has a fixed point; conversely, if $\beta_f$ has no real zeros, dynamical symmetry breaking occurs.
Field theories of the kind just described arise in several examples of the AdS/CFT correspondence. The best known cases are non-supersymmetric orbifolds of ${\cal N}=4$ SYM, dual to Type IIB string theory on $AdS_5\times S^5/\Gamma$ with $\Gamma $ a subgroup of $SU(4)_R$ (but not a subgroup of $SU(3)$).
The field theory instability associated with double-trace renormalization is the boundary counterpart of the instability associated with a closed string tachyon in the AdS bulk. By a tachyon we mean a bulk scalar that {\it violates} the Breinlohner-Freedman bound. As usual, the weakly coupled boundary theory (small 't Hooft coupling $\lambda$) gives information about the high curvature regime of the bulk theory and vice versa. In some examples (non-freely acting orbifolds of ${\cal N}=4$ SYM), the instability is visible both in the weakly curved bulk theory and in the one-loop analysis of the boundary theory; presumably the theory is unstable for all couplings. In other examples (freely acting orbifolds of ${\cal N}=4$) an instability shows up in the one-loop analysis of the boundary theory, but the spectrum of the weakly curved bulk theory has no tachyon; the tachyon must become massive for $\lambda$ greater than some critical value.
We showed in \cite{Pomoni:2008de} that at large $N$ the conformal dimension $\Delta_{{\cal O}}$ and the beta function $\beta_f$ take the general forms
\begin{equation} \label{DeltaOform}
\Delta_{{\cal O} } = 2 + \gamma(\lambda) + \frac{ v(\lambda) }{1+ \gamma(\lambda)}f \, ,
\end{equation}
\begin{equation} \label{betafform}
\beta_f =\frac{ v(\lambda) }{1+ \gamma(\lambda)}f^2 + 2 \gamma(\lambda) f + a(\lambda) \, .
\end{equation}
Here $ v(\lambda)$ is defined as the normalization coefficient of $ {\cal O}$,
\begin{equation}
\langle {\cal O}(x) \bar{\cal O}(0) \rangle = \frac{ v(\lambda) }{2\pi^2x^{2 \Delta (\lambda)}} \, \, ;
\end{equation}
$\gamma(\lambda)$ is the contribution to the anomalous dimension of ${\cal O}$ from single-trace interactions; finally $a(\lambda)$ is the coefficient of the induced
double-trace terms, coming from the single trace interactions, in the quantum effective potential.
The expressions (\ref{DeltaOform}, \ref{betafform}) are valid to all orders in planar perturbation theory: large $N$ factorization implies that $\beta_f$ depends at most quadratically on $f$, and $\Delta_{{\cal O}}$ at most linearly \cite{Pomoni:2008de} .
The coefficients $v(\lambda)$, $\gamma(\lambda)$ and $a(\lambda)$ have planar perturbative expansions
\begin{equation}
v(\lambda) = \sum_{L=1}^\infty v^{(L)} \lambda^{L-1} \,,\quad \gamma(\lambda) = \sum_{L=1}^\infty \gamma^{(L)} \lambda^L \,,\quad a(\lambda) = \sum_{L=1}^\infty a^{(L)} \lambda^{L+1} \,,
\end{equation}
where the $L$ denotes the loop order. Consider the discriminant of the quadratic equation $\beta_f=0$,
\begin{equation}
D (\lambda) \equiv \gamma(\lambda)^2 - \frac{a(\lambda)\, v(\lambda) }{1+ \gamma(\lambda)} \, .
\end{equation}
If $D(\lambda) <0$, there are no physical (real) values of $f$ for which the theory is conformal. But if we insist on formally preserving conformal invariance by
tuning $f$ to one of its two complex fixed points,
then the operator dimension also becomes complex,
\begin{equation}
\Delta_{{\cal O}} = 2 \pm \, i \, b(\lambda)\,, \qquad b(\lambda) \equiv \sqrt{|D|}\,.
\end{equation}
Using the usual AdS/CFT dictionary
\begin{equation}
\Delta_{{\cal O}} = \frac{d}{2} \pm \sqrt{\frac{d^2}{4}+m^2R^2} = 2 \pm \sqrt{4+m^2R^2} \, ,
\end{equation}
we find that the $AdS_5$ scalar dual to ${\cal O}$ has mass \cite{Pomoni:2008de}
\begin{equation} \label{tachyonmass}
m^2(\lambda) R^2 = m^2_{BF} R^2 +D(\lambda)= - 4 +D(\lambda) \,.
\end{equation}
For negative discriminant $m^2(\lambda) < m^2_{BF} $: the scalar field dual to ${\cal O}$ is a true tachyon and the bulk theory is unstable.
We now generalize this story
to AdS/CFT dual pairs containing an open string sector. In the presence of flavor branes in the AdS bulk, the dual field theory contains extra fundamental matter. An {\it open} string tachyon corresponds to an instability in the {\it mesonic} sector of the boundary theory. In this paper we illustrate this phenomenon in the example of the intersecting $D7$ brane system. The classical Lagrangian of the boundary theory takes the schematic form
\begin{equation}
{\cal L} = {\cal L}_{adjoint}+ {\cal L}_{fund} = -\mbox{Tr} \left[ F^2 + \left(D X \right)^2 +\dots \right] -\left(D Q \right)^2 - \frac{\lambda}{N} {\cal O}^{ij} {\bar{\cal O}}^{ij}+ \dots
\end{equation}
where $ {\cal O}^{ij} = Q^{i \,\gaugeind{a}} \bar Q_{j \, \gaugeind{a}}$ are the gauge-invariant mesonic operators made from the fundamental scalars and for simplicity we have ignored
the $SU(2)_R$ structure, which will be restored shortly.\footnote{To avoid cluttering in some expressions below we always write the flavor indices as upper indices in ${\cal O}^{ij}$.}
The whole $ {\cal L}_{fund} $ is $1/N$ suppressed with respect to ${\cal L}_{adjoint}$, in harmony with the fact that the classical D-brane
effective action arises from worldsheets with disk topology, and is thus suppressed by a power of $g_s\sim 1/N$ with respect to the classical closed string effective action,
arising from worldsheets with sphere topology. Nevertheless, as always
in the tachyon condensation problem, it makes perfect sense to focus on classical open string field theory. The classical open string dynamics
is dual to the quantum planar dynamics of the mesonic sector of the field theory.
The 't Hooft coupling $\lambda$ does not run at leading order in $N$, indeed the hyper multiplet contribute to $\beta_\lambda$ at order $O(1/N)$.
For generic angles, the term in the $Q^4$ potential that mix the two flavors run, so perturbative renormalizability forces the introduction of a new coupling constant
$f$,
\begin{equation}\label{extradt}
\delta {\cal L}_{fund} = -\frac{f}{N} {\cal O}^{12} {\bar {\cal O}}^{12}\,.
\end{equation}
Note on the other hand that no extra terms diagonal in flavor (namely ${\cal O}^{11} {\bar {\cal O}}^{11}$ and ${\cal O}^{22} {\bar {\cal O}}^{22}$
are induced at one-loop. For the first flavor this is immediate to see: the diagrams contributing to the term ${\cal O}^{11} {\bar {\cal O}}^{11}$ of the effective
potential are independent of $\theta_1$, $\theta_2$ (they do not involved any coupling mixing the two flavors)
and thus their sum must vanish, as it does in the ${\cal N}=2$ supersymmetric theory with $\theta_1 = \theta_2 =0$.
For the second flavor this follows by symmetry, since the two flavors are of course interchangeable.\footnote{In more detail, the diagrams
contributing to ${\cal O}^{22} {\bar {\cal O}}^{22}$ do not involve the first flavor, which could then be set to zero as the calculation of this terms
of the effective potential
is concerned. The Lagrangian with the first flavor set to zero is ${\cal N}=2$ supersymmetric, only with an unconventional choice of $SU(2)_R$ embedding into $SU(4)_R$ --
it can be turned into the standard Lagrangian by an R-symmetry rotation of the ${\cal N}=4$ fields.}
This extra ``double-trace'' term (\ref{extradt}) arise at
the same order in $N$ order as the classical ${\cal L}_{fund}$, indeed inspection of the Feynman diagrams shows that the one-loop bare coupling $f_0$ behaves as
\begin{equation}
f_0 \sim \lambda^2 \log \Lambda \,.
\end{equation}
The analysis \cite{Pomoni:2008de} can be applied in its entirety to this ``open string'' case. The ``double-trace'' beta function $\beta_f$ takes again the form (\ref{betafform}),
and its discriminant $D(\lambda)$ computes now (through (\ref{tachyonmass}) the mass of the {\it open} string tachyon dual to mesonic operator ${\cal O}^{12}$.
Let us turn to explicit calculations.
\subsection{The one-loop ``double-trace'' beta function}
To proceed, we need to be more precise about the structure of the``double-trace'' terms induced at one-loop, restoring their $SU(2)_R$ structure.
For general angles $\theta_1$ and $\theta_2$, there are three independent structures,
\begin{equation}
\label{deformation}
\delta {\cal L}_{fund} = -\frac{1}{N} \left[ f_{ {\bf 3}^\pm } \left( {\cal O}_{ {\bf 3}^+}^{12} {\cal O}^{21}_{{\bf 3}^-} + {\cal O}_{{\bf 3}^-}^{12} { {\cal O}}^{21}_{{\bf 3}^+} \right) +
f_{ {\bf 3}^0 } {\cal O}_{{\bf 3}^0}^{12} { {\cal O}}^{21}_{{\bf 3}^0} + f_{ {\bf 1}}{\cal O}_{{\bf 1} }^{12} { {\cal O}}^{21}_{{\bf 1}} \right] \,.
\end{equation}
We have imposed neutrality under the Cartan of $SU(2)_R$, since this is an exact symmetry for generic angles, corresponding geometrically
to rotations in the 67 plane (more precisely a 67 rotation is a linear combination of the Cartan of $SU(2)_L$ and $SU(2)_R$, but the hyper multiplets
are neutral under $SU(2)_L$). When one of the angles is zero, say $\theta_2=0$, rotations in the 567 directions a symmetry (again a diagonal combination of $SU(2)_L$ and $SU(2)_R$),
implying $f_{{\bf 3}^\pm} = f_{{\bf 3}^0} \equiv f_{\bf 3}$. We focus on the triplet mesons, which are dual to the open string tachyon. For a single non-zero angle
there is one beta function $\beta_{f_{{\bf 3}}}$ to compute, since the three components of the triplet are related by symmetry. For generic angles there are in principle two distinct
beta functions $\beta_{f_{{\bf 3}^\pm}}$ and $\beta_{f_{{\bf 3}^0}}$; we will illustrate our method computing the first, which is a slightly simpler calculation.
At one-loop, the ``double-trace'' beta function takes the form
\begin{equation}
\beta_f = v^{(1)} f^2 + 2 \gamma^{(1)} \lambda f + a^{(1)}\lambda^2 \,.
\end{equation}
We have seen that $\gamma^{(1)} = 0$ at one loop for the triplet mesons. The normalization coefficient $v^{(1)}$ is easily evaluated
by free Wick contractions,
\begin{equation}
\langle {\cal O}^{12}_{\bf{3}^0}\,(x) {\cal O}^{21}_{\bf{3}^0}\,(y) \rangle = \langle {\cal O}^{12}_{\bf{3}^+}\,(x) {\cal O}^{21}_{\bf{3}^-}\,(y) \rangle = \langle {\cal O}^{12}_{\bf{3}^-}\,(x) {\cal O}^{21}_{\bf{3}^+}\,(y) \rangle = \frac{1}{16 \pi^4 |x-y|^4} \, ,
\end{equation}
implying
\begin{equation}
v_{\bf{3}^+}^{(1)}= v_{\bf{3}^-}^{(1)}= v_{\bf{3}^0}^{(1)}=\frac{1}{8\pi^2}\,.
\end{equation}
It remains to evaluate the coefficient $a^{(1)}$. We are going to extract $a^{(1)}$ from the one-loop
Coleman-Weinberg potential along the ``Higgs branch'' of the gauge theory, $\langle X_{A B} \rangle = 0$, ${\cal Q}^{\cal I} \neq 0$. We put ``Higgs branch'' in quotes because
for general angles it is in fact lifted already at the classical level. Let us first recall the analysis
for the ${\cal N} =2$ supersymmetric theory corresponding to two parallel flavor branes are parallel ($\theta_1 = \theta_2 =0$).
As always in a supersymmetric theory, flat directions are parametrized by holomorphic gauge-invariant composite operators. In our case the relevant operators are the mesons
\begin{equation}
\label{mesons}
{\cal O}^{ij} =q^i\cdot \tilde{q}_j\, ,\quad i,j=1,2
\end{equation}
The dot stands for color contraction $q \cdot {q}^* \equiv q^\gaugeind{a}\,q^*_\gaugeind{a} $ and $i,j$ are the flavor indices.
The holomorphic, gauge invariant mesons that parameterize the Higgs flat directions are subject to F-flatness conditions
\begin{equation}
q^{\gaugeind{a}\,i} \tilde{q}_{\gaugeind{b}\,i} =0 \Leftrightarrow \, {\rm tr}\, {{\cal O}} = {\rm det}\, {{\cal O}} =0\, ,
\end{equation}
thus there are $4-2=2$ complex parameters for the moduli space of the supersymmetric theory ($\theta_1 = \theta_2=0$). We may parameterize the flat directions by
\begin{equation}
\label{unrotated flat}
Q_1 = U \left(\begin{array}{c} q
\\
0
\end{array}\right)
\, ,\quad
Q_2 = U \left(\begin{array}{c} 0
\\
-q
\end{array}\right)
\, ,\quad
U \in SU(2) \quad \mbox{and} \quad q\in \mathbb{R} \, .
\end{equation}
Color indices are kept implicit. In color space we may take $q^{\gaugeind{a}=1}=q$ and $q^{\gaugeind{a} \neq 1}=0$.
For generic $\theta_1$, $\theta_2$ supersymmetry is explicitly broken in the classical Lagrangian and the Higgs branch is completely lifted.
To select $\beta_{f_{{\bf 3}^+}}$ (which is of course equal to $\beta_{f_{{\bf 3}^-}}$), we calculate the effective
potential around a classical background such that $\langle {\cal O}_{\bf 1}^{12} \rangle =\langle {\cal O}_{{\bf 3}^0}^{12} \rangle = \langle {\cal O}_{{\bf 3}^-}^{12} \rangle= 0$, but $\langle {\cal O}_{{\bf 3}^+}^{12} \rangle \neq 0$,
namely
\begin{equation}
\label{back}
Q_1 = \left(\begin{array}{c} q
\\
0
\end{array}\right)
\, ,\quad
Q_2 = \left(\begin{array}{c} 0
\\
-q
\end{array}\right)
\, ,\quad
\quad q\in \mathbb{C}
\end{equation}
This choice corresponds to the
flat direction for the ${\cal N}=1$ susy case $\theta_1 =\theta_2$.
The F-terms of the classical potential vanish for general angles, but for $\theta_1 \neq \theta_2 $ the D-terms do not, ${\cal V}^D_{Q^4} =g^2 |q|^4 ( 1- d\left(\theta_1,\theta_2 \right) )$.
In Appendix E we evaluate the one-loop contribution to the effective potential along this background (at large $N$).
With the help of the Callan-Symanzik equation we find
\begin{equation}
a^{(1)}_{{\bf 3}^{\pm}}= \frac{1}{16\pi^2} \left[ \Big(1-d(\theta_1 , \theta_2) \Big)
+ \frac{1}{2} \Big(1-d(\theta_1 , \theta_2) \Big)^2
+ 4\,\sin^2{\left(\frac{\theta_1 + \theta_2}{2} \right)} \sin^2{\left(\frac{\theta_1 - \theta_2}{2} \right)}\, \right] \,.
\end{equation}
From our (mild) assumption that the classical potential be positive we have $d(\theta_1, \theta_2) \leq 1$, has the crucial implication
\begin{equation}
a^{(1)}_{{\bf 3}^{\pm}} \geq 0\,.
\end{equation}
In the supersymmetric case ($\theta_1=\theta_2$), $a^{(1)}_{{\bf 3}^{\pm}} = 0$, as it must. For $\theta_2 =0$, the $SU(2)$ symmetry is restored, so
\begin{equation}
a^{(1)}_{{\bf 3}^{\pm}}= a^{(1)}_{{\bf 3}^0}=
\frac{1}{16\pi^2} \left[ \Big(1-d(\theta , 0) \Big)
+ \frac{1}{2} \Big(1-d(\theta , 0) \Big)^2
+ 4\,\sin^4{\left(\frac{\theta }{2} \right)} \, \right] \geq 0\,.
\end{equation}
The one-loop triplet beta function (let us focus on the single-angle case)
\begin{equation}
\label{beta3}
\beta_{f_{\bf 3}} = v^{(1)}_{\bf{3}} \,f^2_{\bf 3} + a^{(1)}_{\bf 3} \lambda^2
\end{equation}
does not admit real fixed points for $f_{\bf 3}$, so conformal
invariance is inevitably broken in the quantum theory.\footnote{Conformal invariance is already broken
in the adjoint (``closed string'') sector by the hyper multiplet contribution
to $\beta_\lambda$, but this is subleading effect (of order $O(1/N)$) with respect to the classical Lagrangian.
In the fundamental (``open string'') sector the breaking of conformal
invariance is at leading order in $N$ (quantum effects arise as the same order as the classical Lagrangian).
Of course the {\it whole} fundamental sector is $O(1/N)$ with respect to the adjoint sector, but we can meaningfully separate the effect
we are interested in. This is the field theory counterpart of focussing on the classical open string dynamics of the D-branes, while
ignoring the backreaction of the branes on the bulk background.}
The running coupling
\begin{equation}
\bar f(\mu) = \frac{ a^{(1)}} {\sqrt{v^{(1)}_{\bf{3}}}} \, \lambda^2 \, \tan \left[ \frac{a^{(1)}\lambda^2} {\sqrt{v^{(1)}_{\bf{3}}}} \ln (\mu/\mu_0) \right]
\end{equation}
is a monotonically increasing function interpolating between IR and UV Landau poles, at
energies
\begin{equation}
\mu_{IR} = \mu_0 \, \exp\left(-\frac{\pi \sqrt{v^{(1)}_{\bf3} } }{ \lambda \sqrt{a^{(1)}_{\bf{3} } } } \right) \, , \quad \mu_{UV} = \mu_0 \,\exp \left(\frac{\pi \sqrt{v^{(1)}_{\bf{3}}}}{ \lambda\sqrt{a^{(1)}_{\bf{3}}}} \right) \,. \end{equation}
For small coupling $\lambda \to 0$, the Landau poles are pushed respectively to zero and infinity.
\subsection{The tachyon mass}
As reviewed above, the mass of the field dual to ${\cal O}^{12}_{{\bf 3}^\pm}$ is
directly related to the discriminant of $\beta_{{\bf 3}^\pm}$,
\begin{equation} \label{tachyonmass2}
m^2_{{\bf 3}^\pm} R^2 = m^2_{BF}R^2 + D_{{\bf 3}^\pm}(\lambda; \theta_1, \theta_2)=
-\,4\, -\frac{ \lambda^2}{16\pi^4} \, \mathcal{D}^{(1)}_{\bf{3}}\left(\theta_1, \theta_2 \right)+ {\cal O}(\lambda^3) \, ,
\end{equation}
where
\begin{equation}
\mathcal{D}_{\bf{3}}^{(1)} = \Big(1-d(\theta_1 , \theta_2) \Big)
+ \frac{1}{2} \Big(1-d(\theta_1 , \theta_2) \Big)^2
+ 4\,\sin^2{\left(\frac{\theta_1 + \theta_2}{2} \right)} \sin^2{\left(\frac{\theta_1 - \theta_2}{2} \right)} \, .
\end{equation}
For $\theta_1 \neq \theta_2$ the discriminant is negative, implying that the bulk
field violates the BF stability bound.
Whenever supersymmetry is broken, the bulk field dual to ${\cal O}^{12}_{{\bf 3}^\pm}$ is a true tachyon.
For $\theta_2 =0$ the $\pm$ and $0$ components of the triplet are related by the $SU(2)$ symmetry and are all tachyonic.
We expect the field dual to ${\cal O}^{12}_{{\bf 3}^0}$ to be tachyonic for general angles.
For small angles, the $O(\lambda^2)$ tachyon mass depends on a single unknown parameter $\alpha$ (which enters the parametrization of the classical $Q^4$ potential, see Appendix B),
\begin{equation} \label{weak}
R^2 m_{\bf 3}^2(\lambda) =-4 - \alpha \, \frac{ \lambda^2}{16\pi^4}\left( \theta_1- \theta_2\right)^2 + {\cal O}(\lambda^3)\, , \qquad \theta_1\, ,\theta_2 \ll 1\,.
\end{equation}
This expression applies to all three components of the triplet.
For the $\pm$ components it is just the expansion of (\ref{tachyonmass2}) for small angles. For the $0$ component it follows by imposing the symmetry constraints
$m_{{\bf 3}^0}^2(\theta, \theta)=0$ and $m_{{\bf 3}^0}^2(\theta, 0)=m_{{\bf 3}^\pm}^2(\theta, 0)$. By AdS/CFT, we get an interesting prediction
for the mass of the open string tachyon for large AdS curvature (small $\lambda$).
Conversely, for large $\lambda$ (small AdS curvature) we can compute the mass of the open string tachyon using the dual string picture. The open string spectrum
of branes intersecting at small angles in flat space is well-known. The lowest tachyon mode has mass
(see {\it e.g.} \cite{Epple:2003xt} for a review),
\begin{equation}
m^2 = -\frac{| \theta_1- \theta_2|}{\pi^2 \alpha'} \, , , \qquad \theta_1\, ,\theta_2 \ll 1\,.
\end{equation}
This becomes a good approximation to the mass in the exact AdS sigma model in the
limit $\alpha'/R^2 \sim \lambda^{-1/2} \to 0$. Thus
\begin{equation} \label{strong}
\lim_{\lambda \to \infty} R^2 m_{\bf 3}^2(\lambda) = -\frac{| \theta_1- \theta_2|}{\pi^2} \frac{R^2}{\alpha'} = -\frac{| \theta_1- \theta_2|}{\pi^2} \lambda^{1/2}\, ,
, \qquad \theta_1\, ,\theta_2 \ll 1\,.
\end{equation}
This can be regarded as a prediction for the large $\lambda$ behavior of the discriminant $D_{\bf 3}(\lambda)$, which is a purely field-theoretic quantity.
Note that apart from the $\lambda$ dependence, which could have been anticipated on general grounds, the weak coupling result (\ref{weak}) and the strong coupling
result (\ref{strong}) differ in their angular dependence.
\section{Discussion}
The main technical question that we leave answered is the precise form of the classical $Q^4$ potential for generic angles.
As we have emphasized, a superspace formulation of ${\cal N}=4$ SYM with manifest $SU(4)_R$ symmetry
would offer a solution. It would be interesting to see whether the new off-shell formalism for ${\cal N}=1$ SYM in ten dimensions
introduced in \cite{Berkovits:1993zz, Baulieu:2007ew} could be applied to our problem.
In principle, another way to obtain the $Q^4$ potential is by taking the decoupling limit
of the intersecting brane effective action. This would first require the calculation of a four-point function of twist fields, two twist fields corresponding to D3-D7 open strings
and two twist fields corresponding to D3-D7' open strings. This problem has been solved for branes intersecting at right angles
(see {\it e.g.} \cite{Dixon:1986qv, Cvetic:2003ch, Abel:2003vv, Antoniadis:2000jv}). The generalization to arbitrary angles is an interesting
and difficult problem in boundary conformal field theory. Taking the decoupling limit may also be challenging in the presence
of tachyons -- it is not clear to us whether the result would be unambiguous or it would require some renormalization prescription.
Even without a complete knowledge of the classical $Q^4$ potential, by making a plausible positivity assumption
we argued that the field theory is unstable at the quantum level. By AdS/CFT, we obtained a non-trivial
prediction for the tachyon squared mass $m_{\bf 3}^2(\lambda)$ at small $\lambda$. Its behavior at large $\lambda$ is known
from flat-space string theory. There must exist an
interpolating function $m_{\bf 3}^2(\lambda)$ valid for all $\lambda$. It would be extremely
interesting to apply integrability techniques to find the whole function. There is a large
literature on open spin chains arising in the calculation of anomalous
dimensions of mesonic operators, see in particular \cite{Erler:2005nr, Mann:2006rh, Correa:2008av, Correa:2009dm} for our system in the ${\cal N} =2$ supersymmetric case $\theta_1 = \theta_2 =0$.
It remains to be seen whether the susy-breaking rotation preserves integrability.
Another direction for future work is to study the actual tachyon condensation process
on the field theory side. In the bulk, after tachyon condensation the intersecting $D7$ branes recombine (see {\it e.g.} \cite{Epple:2003xt}).
For small $\lambda$, the tachyon vacuum corresponds on the field theory side to the local minimum of the one-loop effective potential.
It would be interesting to expand the Lagrangian around the minimum and relate this field theory calculation
to the bulk phenomenon of brane recombination.
\section*{Acknowledgements}
It is pleasure to thank Igor Klebanov, William Linch III, Andrei Parnachev, Martin Rocek and Warren Siegel for useful discussions.
This work is supported in part by the DOE grant DEFG-0292-ER40697 and by the NSF grant PHY-0653351-001.
Any opinions, findings, and conclusions or recommendations expressed in this material are those of the authors
and do not necessarily reflect the views of the National Science Foundation.
\label{global}
|
\section{Introduction}
Electron transfer in biological systems is rather different from electron transport in an electronic device. In addition to the intrinsic materials
differences between the respective conducting media, namely soft molecules in biology against inorganic semiconductors in electronics, the
electron transport in these two situations differs for the effects and the relevance of the environment. In fact, while one wants to keep
electronic devices in dry conditions to avoid electrostatic disorder, biological electron transfer is characterized by a time-evolving
local dielectric environment.
The boundary between the two fields, electronics and biology, becomes however blurred when one looks at the nanoscale. On the one
hand, ``conventional'' nanotechnology has expanded in the biology domain with a growing number of electronic applications requiring
operation in wet conditions. In addition to scanning tunnel microscopy in water \cite{Bruce}, which has been available for the last two
decades, several biological sensors have been recently proposed. These for instance include cancer markers detectors \cite{Cancer}
and protocols for DNA sequencing \cite{DNA}. On the other hand, when a solution is confined at the nanoscale, highly ordered structures
can form at room temperature \cite{h2o}. This essentially means that under such stringent confinement a biological system
almost behaves like a solid.
Interestingly, numerous experiments to date on electronic transport through molecules are carried out in solution \cite{bdtexpxiao,laka}.
Still, with a few exceptions \cite{bdtwater,colin,Cao}, most of the theoretical calculations are performed in the dry, i.e. without explicitly including
water molecules in the simulations. Therefore the question on how water affects the current-voltage characteristic in a
molecular junction remains. In general one may expect that solutions made with polar molecules, such as $\mathrm{H}_2\mathrm{O}$, may affect
significantly the transport because of the generation of local dipole fields. However, since the time scale of a transport measurement
is far longer than the typical molecular rearrangement of a solution, one should ask what is the time-averaged dipolar field near
the molecule of interest. This, together with the degree of localization of the relevant molecular orbitals, will determine whether
or not the wetting conditions influence the junction electronic transport.
In order to address those fundamental questions we have performed a number of combined molecular dynamics (MD) and
quantum transport simulations for molecular junctions in water. Our computational strategy is to investigate first the
dynamics of $\mathrm{H}_2\mathrm{O}$ and then to evaluate the electronic transport of a representative number
of configurations, i.e. for a number of MD snapshots. In particular we have considered two rather different junctions,
both using gold electrodes but differing for the local charge arrangement of the molecule of interest. The first is made
from a polar molecule, namely benzene-dithiol (BDT), while the second includes locally charge neutral carbon nanotubes (CNTs),
in particular a (3,3) metallic and a (8,0) insulating one.
We find that the effect of water on the transport is that of effectively gating the molecule, therefore shifting almost rigidly
its transmission spectrum. This is rather pronounced for BDT, but only tiny in the case of CNTs, and reflects the different charge distribution
of the two classes of molecules. We associate those differences to the time-averaged dipole field of the water.
Notably all the calculations are performed with density functional theory including appropriately corrections for the electronic structure
of water. This is an essential condition for quantitative predictions.
The paper is organized as follows. In the next section we describe the device set-up for the calculations and briefly the computational
tools used. Then we analyze the transport. First we look at the electronic structure of a gold capacitor with water wetting the two
electrodes and then we consider finite bias conductance across BDT and the CNTs. Finally we present our main conclusions.
\section{Methods}
\label{sec:methods}
Classical MD calculations are performed with the {\sc namd2} package\cite{NAMD2005}. BDT is attached to the two
gold electrodes at the Au(111) hollow site, which has been previously calculated to be the low energy bonding position for
BDT on Au(111) \cite{cormacbdtprb}. We define our coordinate system with the $z$ axis along the (111) direction and the
$x$-$y$ plane orthogonal to it. All the
calculations are carried out with periodic boundary conditions in the $x$-$y$ plane. TIP water molecules are added to a $(25 \times 17.3 \times 7.3)$~\AA$^{3}$ box intercalated
between the Au electrodes. Note that the dimension along the $z$ axis is chosen in such a way that the
Van der Waals distance between the Au(111) surface and the water molecules is taken into account. In total there are 83
$\mathrm{H}_2\mathrm{O}$ molecules in the simulation box. Since our main objective is that of examining the effects of water over the
conductance of a Au/molecule/Au device, we always fix the atomic positions of both the molecule and the electrodes.
The interaction between the H$_2$O molecules and gold is treated at the level of a 12-6 Lennard-Jones potential
\begin{equation}
U^\mathrm{LJ}=4\epsilon\left[\left(\frac{\sigma}{r}\right)^{12}-\left(\frac{\sigma}{r}\right)^{6}\right]\:,
\end{equation}
with parameters for Au ($\epsilon =0.039$~kcal/mol and $\sigma = 2.934$~\AA) taken from the literature [\onlinecite{GoldLJ}].
Periodic boundary conditions are applied with a cutoff of 12 \AA~for long-range interactions. In order to maintain the size and shape
of the cell constant, we perform simulations in the micro-canonical ensemble, with re-initialized velocities to 300~K for every
1000 time-steps and with a time-step of 2~fs. The trajectory is recorded every 4~ps from the initial equilibration of 1~ns to a total
simulation time of 20~ns. For the systems comprising the CNTs, Au(111)/CNT(3,3)-H$_2$O/Au(111) and Au(111)/CNT(8,0)-H$_2$O/Au(111),
and for the reference Au(111)/H$_2$O/Au(111) capacitor, similar conditions and procedures are followed. The parameters for the
aromatic carbon atoms are used for CNTs.
For each system we calculate the transport properties for a set of representative MD configurations taken after equilibration. From these
we can then estimate the fluctuations in the transmission and current over time, as well as their time-averaged values. We use the {\sc smeagol}
\textit{ab initio} electronic transport code to calculate the zero bias transmission coefficients and the current-voltage ($I$-$V$) characteristics.
{\sc smeagol} combines the non-equilibrium Green's function method with density functional theory (DFT)\cite{smeagol1,smeagol2,senepaper}
and has the pseudopotential code {\sc siesta}\cite{siesta} as its electronic structure platform.
The local density approximation (LDA) to the exchange and correlation functional is adopted throughout. The atomic self-interaction
correction (ASIC)\cite{dasasic} scheme however is used for the water and the BDT molecule, in order to bring their ionization potentials (IPs)
in closer agreement to experiments. This has been already proved to be a successful strategy for aligning correctly the highest
occupied molecular orbital (HOMO) energy, $\epsilon_\mathrm{HOMO}$, of the molecule under consideration to the Fermi level
($E_\mathrm{F}$) of the electrodes\cite{cormacbdtprb,cormacbdtprl}. Note that the same corrections also reproduce well the
band-gap of many insulating oxides, after the ASIC potential is rescaled appropriately\cite{dasasic}. Such a rescaling
in bulk crystals is attributed to charge screening, which in solids is usually stronger than in molecules. In general we use a scaling
parameter, $\alpha$, ranging between 0 and 1, to adjust the amount of self-interaction correction included
($\alpha=0$ corresponds to the LDA, $\alpha=1$ is the full ASIC). Usually $\alpha$ is 1 for small molecules, it is around 0.5 for
insulating oxides and it vanishes for metals. For this reason ASIC is never applied to Au.
For the transport calculations we use a real space mesh cutoff of 200~Ry and an electronic temperature of $300$~K. The unit cell
includes 5 Au atomic layers on each side of the Au(111) surface, which are enough to screen charging at the Au-molecule interface.
In order to reduce the system size the Au $5d$ shell is kept in the core, so that we consider only $6s$ orbitals in the valence.
We use a single-$\zeta$ basis for Au $6s$, specifically optimized to give the correct work function for the Au(111) surface.\cite{magmol}
The rest of the basis set is double-$\zeta$ for the C $s$ and $p$ orbitals, and double-$\zeta$ plus polarization for S ($s$ and $p$). For
the $\mathrm{H}_2\mathrm{O}$ molecules we use a double-$\zeta$ basis for both H and O. However when calculating transport through CNTs we reduce it
to single-$\zeta$, in order to keep the size of the density matrix tractable.
The charge density is obtained by splitting the integral of the Green's function into a contribution calculated over the
complex energy plane and one along the real axis \cite{smeagol1,smeagol2}. The complex part of the integral is computed
by using 16 energy points on the complex semi-circle, 16 points along the line parallel to the real axis and 16 poles. The integral
over real energies necessary at finite bias is evaluated over at least 1000 points\cite{smeagol1,smeagol2} per eV.
\section{Results and Discussion}
\subsection{Au capacitor in water}
\label{sec:watergold}
\begin{figure}
\center
\includegraphics[width=8.0cm,clip=true]{fig1}
\caption{(Color online) Au capacitor with water as dielectric medium. The top panel shows the unit cell used for the
MD simulations, which includes $\mathrm{H}_2\mathrm{O}$ confined between the two Au electrodes. In panel (a) we show the probability to find
a O (solid curve) or a H (dashed curve) atom at a given $z$ position. Panels (b-d) are the planar averages
of the Hartree electrostatic potential $\bar{V}_{\mathrm H}$ as function of position along the transport direction, $z$, for different setups:
(b) entire junction, (c) junction without the electrodes, and (d) junction without water. The grey curves, merging in a shadow, are the
results for 21 snapshots taken at different times after equilibration, the black solid curves are their time averages. The dashed curve
in (c) shows the difference between the time-averages of panels (b) and (d).}
\label{Fig1a}
\end{figure}
In a transport calculation it is crucial to describe accurately both the electrodes' work function, $W$, and the IP of the molecule,
so that the correct alignment of the molecular levels to the electrodes' $E_\mathrm{F}$ is reproduced. With this in mind we first
analyze the electronic structure of water sandwiched between Au electrodes. Such a setup corresponds to calculating the electronic
structure of a Au parallel plate capacitor, where the two plates are separated by water. The unit cell, containing 408 $\mathrm{H}_2\mathrm{O}$
molecules and 480 Au atoms, is shown in Fig. \ref{Fig1a}. The statistical distribution of $\mathrm{H}_2\mathrm{O}$ about the Au plates
is described in Fig. \ref{Fig1a}(a), where we plot the normalized probability, $p(z)$, to find O (solid curve) or H (dashed curve)
at a given position $z$ in the cell ($\int p(z)~dz=1$). Such a distribution is obtained by using the MD data for all the time-steps
included in the 16~ns simulation after equilibration. We note that $p(z)$ is constant, in the middle of the gap between the plates, indicating an
average random arrangement of the $\mathrm{H}_2\mathrm{O}$ molecules. In contrast close to the Au interface there are marked oscillations in
$p(z)$, signaling a correlation of the water position with respect to the Au surface. Note that the peaks in $p(z)$ are found at the same positions for O and H atoms. This indicates that on average there is no net dipole at the interface.
Next the Au work function is calculated by using the same Au parallel plate capacitor of Fig.~\ref{Fig1a} after we have removed
the water, i.e. with vacuum as spacer between the plates. $W$ is then the energy difference between $E_\mathrm{F}$ and the
vacuum potential, $V_\mathrm{vacuum}$. Such an exercise is reported in Fig.~\ref{Fig1a}(d) where we show the planar average
$\bar{V}_H$ of the electrostatic Hartree potential along the $x$-$y$ plane. Note that the absolute value of $\bar{V}_H$ in Au is
arbitrary, since it depends on the pseudopotentials. However, in the middle of the capacitor $\bar{V}_H=V_\mathrm{vacuum}$.
Therefore, setting $V_\mathrm{vacuum}=0$ we have $W=-E_\mathrm{F}$. We find
$W=5.3$~eV, in good agreement with experiments.
We now turn our attention to the electronic structure of water. This must be extracted for its liquid phase, i.e. from the MD
simulations. We consider data averaged over 21 structural configurations corresponding to
21 equally spaced MD snapshots. The so-calculated planar average of the electrostatic potential is shown in Fig. \ref{Fig1a}(b).
The grey lines, merging in a shadow, are the superimposed curves for all the snapshots, while the black line is their time-average.
We note oscillations of $\bar{V}_\mathrm{H}$ close to Au. These are due to the arrangement of the water molecules
with respect to the Au surface. In contrast in the middle of the junction $\bar{V}_H$ is rather flat due to the average random
orientation of the molecules [see also Fig.~\ref{Fig1a}(a)].
\begin{figure}
\center
\includegraphics[width=6.0cm,clip=true]{fig2}
\caption{(Color online) DOS projected onto $\mathrm{H}_2\mathrm{O}$ for the capacitor of Fig. \ref{Fig1a} and obtained for different
values of the ASIC scaling parameter $\alpha$. The grey lines are the superimposed curves for 21 MD snapshots and
the solid black one is their average. The dashed curve is the average DOS for the same MD snapshots for a system where
Au is replaced by vacuum.}
\label{Fig2a}
\end{figure}
From the same simulations we can extract the IP of liquid water. In transport one wants $\epsilon_\mathrm{HOMO}$ of the molecule of
interest to correspond to the actual negative of its IP \cite{TSFB,CormacSTM}. Although this should be the case
for exact DFT, it happens only rather rarely in practice for the standard local approximations of the exchange and correlation
functional\cite{DD}. In the case of liquid H$_2$O the experimental value of the IP
is in the range between 9.3 and 10~eV, in rather good agreement with recent \textit{ab initio} calculations of the electron
removal energy (see [\onlinecite{wateriedft,wateriedft2}] and references therein).
We evaluate $\epsilon_\mathrm{HOMO}$ for liquid H$_2$O by calculating the density of states (DOS) projected onto the water molecules. This is
shown in Fig.~\ref{Fig2a} ($V_\mathrm{vacuum}=0$), where again the grey curves are the superimposed data for
all the MD snapshots and the black one is their average. We note that there are large
fluctuations in the DOS over time with energy shifts of the order of 1~eV. The average value however is smooth and can be
reliably taken as the liquid water DOS. In this way the top of the water ``valence band'' is calculated to be only -6~eV in
LDA, in agreement with previous DFT calculations\cite{waterdft}, but still far from the experimental value.
We then apply ASIC to $\mathrm{H}_2\mathrm{O}$ and find that $\epsilon_\mathrm{HOMO}$ moves to -8.5~eV for $\alpha=0.5$ and to -11~eV for $\alpha=1.0$, so that
$\alpha=0.7$ fits the average experimental value (-9.5~eV). Such an optimal value, as usually with ASIC\cite{dasasic},
in general improves the entire electronic structure and returns an HOMO-LUMO gap of about 6.4~eV, in good agreement with
the experimental value of 6.9 eV.\cite{waterexpgap} Note that $\alpha=0.7$ is typical for moderately ionic insulating oxides\cite{dasasic}.
In order to analyze the effects of the Au/H$_2$O interface over the water IP and DOS we perform a second set of calculations, where
we remove the Au plates and we replace them by vacuum. This corresponds to an hypothetical H$_2$O slab. In this case [see
Fig.~\ref{Fig1a}(c)] $\bar{V}_\mathrm{H}$ at a given MD time-step has a finite slope in the vacuum region, which is caused by
the non-compensated dipoles at the H$_2$O external surface. These dipoles produce a long-range electric field outside slab,
so that $V_\mathrm{vacuum}$ of a single snapshot is not defined. However the time-averaged $\bar{V}_\mathrm{H}$ (black curve)
is approximately flat in the vacuum, demonstrating that, although at each time-step surface charge may lead to long-ranged
electric fields, its time-average is actually zero.
If we now take the average $\bar{V}_\mathrm{H}$ away from the water molecules as $V_\mathrm{vacuum}$, we can plot the time-averaged
DOS for the water slab and superimpose it to that calculated for the Au/H$_2$O/Au capacitor [see Fig.~\ref{Fig2a}]. We find that the
two DOSs overlap on each other, confirming our results for the water IP and the fact that on average there is little electronic interaction
between Au and H$_2$O. Our results also suggest that one should ideally use periodic boundary conditions to simulate the electronic
structure of molecules in solution. These eliminate the possible spurious electric fields in the vacuum, which can
lead to an unphysical rearrangement of the energy levels. Furthermore for simulations of $\mathrm{H}_2\mathrm{O}$ surfaces it is essential to consider
time-averages, so that the water electric field vanishes in vacuum.
Fig.\ref{Fig1a}(c) also shows the difference between the time-averaged $\bar{V}_\mathrm{H}$ of the Au capacitors with
and without water (dashed line). This difference is almost identical to the time-averaged $\bar{V}_\mathrm{H}$ for the water slab
and is consistent with the fact that the Au electrodes do not induce any noticeable change in the average DOS of H$_2$O.
\section{Benzene-dithiol}
\label{sec:BDT}
\begin{figure}
\center
\includegraphics[width=5.5cm,clip=true]{fig3}
\caption{(Color online) Unit cell used for the BDT molecule in water attached to Au electrodes.}
\label{Fig3a}
\end{figure}
Electron transport through BDT attached to Au electrodes has been extensively studied both
experimentally\cite{bdtexpreed,bdtexpxiao,bdtexp2,bdtexp3} and theoretically.\cite{cormacbdtprl,cormacbdtprb} In fact, because of its simple
structure, BDT is an ideal system for comparing theory with experiments. However, also for BDT the LDA description is not
adequate and it is necessary to employ ASIC\cite{cormacbdtprl,cormacbdtprb}. In order to limit the number of adjustable
parameters we set the same $\alpha$ for both BDT and H$_2$O, and check that such a value reproduces well previous transport
calculations for the Au/BDT/Au junction in dry conditions\cite{cormacbdtprl,cormacbdtprb}.
We first investigate the $V=0$ transport (the cell used is shown in Fig. \ref{Fig3a}). In Fig. \ref{Fig4a} the transmission
coefficient $T(E;V=0)$ for one MD snapshot is shown as function of energy, $E$, for different values of the ASIC scaling parameter
$\alpha$ (solid curves). The same quantity is compared to that calculated for the same cell, this time without including water (dashed curves).
In general the effects of water are two-fold: firstly there is a shift of the BDT transmission peaks to lower energies, and secondly there appear
additional sharp transmission peaks, which are attributed to resonant transport through the electronic states of H$_2$O. Interestingly the
height, the width and the relative position of the transport peaks with respect to each other is unchanged when water is present.
Therefore the main effect of adding water is to shift the energy levels of BDT, so that water acts as an external gate. Since the BDT molecular
orbitals extend over the entire molecule and are strongly coupled to Au\cite{cormacbdtprb}, all the levels shift by approximately the
same amount. If the levels were more localized, we might have expected a change in their relative position, sensitively dependent
on the local configuration of $\mathrm{H}_2\mathrm{O}$ \cite{LocField}.
\begin{figure}
\center
\includegraphics[width=6.0cm,clip=true]{fig4}
\caption{(Color online) Transmission coefficient for transport through BDT in water as function of energy for one MD snapshot
and for different values of the ASIC scaling parameter $\alpha$. The dashed curves are for BDT in dry conditions
(no water), while the solid curves are for BDT in solution.}
\label{Fig4a}
\end{figure}
We now analyze in more detail the electronic properties of H$_2$O for the same MD snapshot. In Fig. \ref{Fig5a} the DOS projected
onto the water molecules is shown over the same energy range as that of $T$ in Fig. \ref{Fig4a}. It is clear that the additional
peaks in the transmission (Fig.~\ref{Fig4a}) are at energies where H$_2$O has a finite DOS, confirming that these are due to transport
through the electronic states of water. The additional transmission peaks below $E_\mathrm{F}$ are rather close to the Fermi level
when $\alpha=0$, whereas they move down in energy as $\alpha$ is increased. In contrast the position of the peaks above
$E_\mathrm{F}$ is almost constant for different $\alpha$, since the ASIC mainly affects occupied states. We also find that
for $\alpha=0$ the water HOMO is about -6.3~eV from $V_\mathrm{vacuum}$, while it is at -8.7~eV for $\alpha=0.5$.
These values agree with the values obtained for the water slab.
\begin{figure}
\center
\includegraphics[width=6.0cm,clip=true]{fig5}
\caption{Density of states projected onto $\mathrm{H}_2\mathrm{O}$ for the junction of Fig. \ref{Fig3a} as function of energy for one MD snapshot
and for different values of the ASIC scaling parameter $\alpha$.}
\label{Fig5a}
\end{figure}
The self-consistent current-voltage, $I$-$V$, curve is shown next in Fig. \ref{Fig6a} for the same MD snapshot and for different values of $\alpha$.
Generally speaking the presence of water leads to a reduction of the current, which is more pronounced for small $\alpha$. The reason for
such a reduction is that $\epsilon_\mathrm{HOMO}$ is very close to $E_\mathrm{F}$ for small $\alpha$, so that a tiny shift of the BDT levels to lower energies
considerably reduces $T(E_\mathrm{F};V\approx0)$. For $\alpha=1$ there is almost no change in the current, since $E_\mathrm{F}$
is approximately at mid-gap already in dry conditions.
\begin{figure}
\center
\includegraphics[width=6.0cm,clip=true]{fig6}
\caption{(Color online) Current, $I$, as function of the bias voltage, $V$, through the BDT molecule calculated at the first MD
time-step and for different values of the SIC scaling parameter $\alpha$. The dashed curves are for the molecule in the dry
while the solid curves are for BDT in solution.}
\label{Fig6a}
\end{figure}
We now briefly discuss the most appropriate value of $\alpha$ in this context. For BDT in the gas phase $\alpha=1$ gives
$\epsilon_\mathrm{HOMO}$ close to the experimental -IP.\cite{dasasic,cormacbdtprl} However, when BDT is immersed in water additional
screening lowers down the value of $\alpha$. We then take $\alpha=0.7$ (the optimal value for $\mathrm{H}_2\mathrm{O}$), which provides
a good IP for water and also accounts for the additional screening in BDT due to the solution. Importantly our results
depend little on the exact choice of $\alpha$, as long as it is of the order of 0.5, i.e. such that $\epsilon_\mathrm{HOMO}$ for H$_2$O is well
below the Au $E_\mathrm{F}$.
Note that using $\alpha=0$ would lead to the erroneous prediction that water becomes conducting at about 1 V of bias.
We now move to calculate the time-averaged transmission coefficient and the $I$-$V$ curves. In this case $T(E; V=0)$ and the $\mathrm{H}_2\mathrm{O}$
DOS are evaluated over 201 snapshots taken in the last 16~ns of our MD simulations, while the $I$-$V$ curves are evaluated over
only 21. Transmission and DOS are presented in Fig.~\ref{Fig7a}, where again the curves for the single snapshot calculations are
plotted in grey to form a shadow, while their average is a solid black line. In general, when $\mathrm{H}_2\mathrm{O}$ is introduced in the simulation,
there is a rigid shift of the entire spectrum towards lower energies with respect to the dry situation (dashed line).
This is because BDT and $\mathrm{H}_2\mathrm{O}$ are both polar molecules and in time the water molecules arrange around BDT so to screen the
local dipole field. Such screening moves the average BDT molecular levels to lower energies.
\begin{figure}
\center
\includegraphics[width=7.0cm,clip=true]{fig7}
\caption{(Color online) Time averaged (a) transmission coefficient and (b) DOS projected onto the $\mathrm{H}_2\mathrm{O}$ molecules for BDT attached to Au.
The calculations are obtained with ASIC and $\alpha=0.7$; the solid black curves are the time-averages over 201
time-steps while the grey ones forming a shadow are for each of the 201 time-steps. The dashed curve is for the dry situation.}
\label{Fig7a}
\end{figure}
This analysis is confirmed in Fig.~\ref{Fig8a}, where we present the O and H position distributions along the $y$ direction, for those H and O atoms lying either above or below the plane of the BDT (shadowed region in Fig. \ref{Fig8a}). Note that we define the $y$ axis as the direction perpendicular to the plane of the BDT. In contrast to the
case of the Au capacitor, now the $p(y)$'s of O and H ions differ near BDT. In particular we find that H approaches
BDT approximately 1~\AA\ closer than O. This means that on average the first solvation layer is oriented with the H atoms
of $\mathrm{H}_2\mathrm{O}$ molecules pointing towards the BDT, as suggested by elementary electrostatics, since the C and S atoms are fractionally negatively charged, while the H atoms have a positive charge. The second peak of the H atoms overlaps with the first peak of O, indicating that while one of the H atoms points towards the BDT, the second one aligns with the negative O atoms. It is also interesting to note that
the O distribution has a second pronounced peak in addition to that close to the BDT, signaling a relative large degree of order
also in the second solvation layer.
\begin{figure}
\center
\includegraphics[width=7.0cm,clip=true]{fig8}
\caption{(Color online) Probability to find an O (solid curve) or a H (dashed curve) atom at a given position $y$ in the Au/BDT/Au junction, for atoms whose $x$ coordinate lies within the shadowed region. In the top panel we show a representative snapshot of the atomic configuration of the water.
Note that in the first solvation layer the $\mathrm{H}_2\mathrm{O}$ molecules are oriented with the H atoms pointing towards the BDT.}
\label{Fig8a}
\end{figure}
We finally turn our attention to the fluctuations. Generally, time-fluctuations in the position of the BDT single particle levels result in
zero-bias conductance fluctuations\cite{bdtwater}. These cause both a reduction in the average height of the various transmission peaks and at the
same time an increase of their widths. For BDT attached to Au this second effect is rather small, since the transmission peaks at each time-step
are already rather broad, due to the strong electronic coupling with the electrodes. However we expect that for molecules weakly coupled
to the electrodes and thus presenting sharp peaks in $T(E)$ this effect will be more pronounced, probably dominating the energy level
broadening.
As already mentioned before the electrostatic interactions of water with the BDT mimics a gate potential.
At each MD time-step such an effective gate voltage changes, depending on the relative position of $\mathrm{H}_2\mathrm{O}$.
In order to quantify the fluctuations of the BDT molecular levels we track the position of $\epsilon_\mathrm{HOMO}$ [from the peak in $T(E)$]
over time, and display the result in the form of a histogram in Fig.~\ref{Fig9a}. In the plot $N$ is the number of counts $\epsilon_\mathrm{HOMO}$
is found in a particular energy window, the dashed red line indicates the position of $\epsilon_\mathrm{HOMO}$ in the dry, while the solid
black line indicates the time averaged $\epsilon_\mathrm{HOMO}$ in $\mathrm{H}_2\mathrm{O}$ solutions. Clearly $\epsilon_\mathrm{HOMO}$ fluctuates between -1.8~eV and -2.4~eV,
i.e. in an energy range of 0.6~eV. The time-averaged $\epsilon_\mathrm{HOMO}$ is about 0.6 eV below $\epsilon_\mathrm{HOMO}$ for the dry molecule, which means that the effective water-induced gating potential is about 0.6 eV.
\begin{figure}
\center
\includegraphics[width=6.0cm,clip=true]{fig9}
\caption{(Color online) Histogram of $\epsilon_\mathrm{HOMO}$ extracted from the maximum in $T(E)$ at around -2~eV (see Fig.~\ref{Fig7a}) for $\alpha=0.7$
and for 201 MD time-steps. $N$ is the number of times $\epsilon_\mathrm{HOMO}$ is found in the given energy window specified by the bin width. The solid
line indicates the time-averaged $\epsilon_\mathrm{HOMO}$, while the dashed red one marks the result in dry conditions.}
\label{Fig9a}
\end{figure}
\begin{figure}
\center
\includegraphics[width=6.0cm,clip=true]{fig10}
\caption{(Color online) $I$-$V$ curve for a BDT molecule attached to gold in presence of water obtained with ASIC ($\alpha=0.7$); the solid black
curve corresponds to the average current over 21 time-steps, the grey curves merging in a shadow are the $I$-$V$ curves of each individual configuration and the red dashed curve is for BDT in dry conditions.}
\label{Fig10a}
\end{figure}
Finally in Fig.~\ref{Fig10a} we present the self-consistent $I$-$V$ characteristics, where we conclude that the current fluctuates about its time average
by approximately $\pm20\%$, while it is reduced from that in the dry by about 35\%. This discussion is based on ASIC calculations with the optimal
value of $\alpha=0.7$. As discussed previously, the most appropriate correction for BDT in the dry is $\alpha\approx1$. If the same correction is
exported into the wet situation the reduction of the current due to water becomes almost negligible.
\section{Carbon nanotubes}
\label{sec:cnt}
We now move to the analysis of the effects of $\mathrm{H}_2\mathrm{O}$-wetting on non-polar molecules, i.e. molecules presenting local charge neutrality. In particular we choose
two different CNTs: i) (3,3) metallic armchair and ii) (8,0) insulating zigzag.
In the case of the metallic (3,3) CNT we again perform 20~ns long MD simulations and calculate the observables over 201 equally spaced snapshots in the
last 16~ns. The unit cell used is shown in Fig.~\ref{Fig11a} for one particular MD snapshot. This has a (20.0$\times$17.3)~\AA$^2$ cross section and
contains 480 Au atoms, 192 C atoms and 360 $\mathrm{H}_2\mathrm{O}$ molecules. The Au-CNT distance is simply obtained by adding the Au and C atomic radii
(respectively 1.44~\AA\ and 0.7~\AA) and it is close to that obtained by total energy minimization \cite{cnt33fp}. We note that the exact conformation of
the Au-CNT bonding is not known and that changes in bond structure lead to quantitative changes in the transmission spectra.\cite{cntautightbinding}
Here however we are mainly interested in investigating how the transmission is affected by the water so that the precise bonding
geometry is less important.
As already mentioned before, because of the large system size here we use a single-$\zeta$ basis for $\mathrm{H}_2\mathrm{O}$. We verified that this gives a similar IP
to that obtained with the double-$\zeta$ basis. In what follows we will use $\alpha=0.7$ for $\mathrm{H}_2\mathrm{O}$, but no ASIC for the CNTs, since their
electron screening is good. We have verified that the band-structure of the (3,3) CNT agrees well with previous calculations.\cite{cnt33bands}
In particular we obtain a CNT work function of 4.4~eV in good agreement with previous calculations\cite{calcswc33}. The IP for (3,3) CNT is not
available experimentally, but that of similar CNTs ranges between 4.8~eV and 5.0~eV,\cite{expwc33,expwc33p2,expwc33p3} thus is not
far from what calculated here.
Since $W$ of Au is about 1~eV larger than that of the CNT, electrons transfer from the CNT into Au, leading to a substantial band-bending.
This is demonstrated in Fig.~\ref{Fig11a}(a), where the planar average of the Hartree potential is plotted for the Au/CNT/Au junction in dry conditions.
Close to the Au/CNT interface the oscillating $\bar{V}_\mathrm{H}$ is higher than in the middle of the junction, whereas for an infinite CNT $\bar{V}_\mathrm{H}$ oscillates around a constant average.
We note that such charging effects have been neglected in previous tight-binding calculations.\cite{cntautightbinding}
Importantly however charging leads to a shift in the transmission spectrum, so that it is important to include such an effect in a self-consistent way.
\begin{figure}
\center
\includegraphics[width=8.0cm,clip=true]{fig11}
\caption{(Color online) Au/CNT/Au junction. The top panel shows the unit cell used for the MD simulations, which includes the CNT, Au electrodes and
water molecules. In panels (a) and (b) we present the planar averages of the Hartree electrostatic potential, $\bar{V}_\mathrm{H}$, as function of position
along the transport direction $z$: (a) junction without water, and (b) junction with water. The grey curves, merging in a shadow, are the results for all 201
MD snapshots, the black solid curves are their time-averages.}
\label{Fig11a}
\end{figure}
Next we look at the wet situation of Fig.~\ref{Fig11a}(b). In this case $\bar{V}_\mathrm{H}$ for a single MD snapshot oscillates dramatically along the CNT.
However, when the time average is considered [solid black curve in Fig. \ref{Fig11a}(b)] a regular pattern emerges, where $\bar{V}_\mathrm{H}$
resembles closely that obtained in the dry. This confirms that on average the position of the $\mathrm{H}_2\mathrm{O}$ molecules away from the interface is random.
Since CNTs are hydro-phobic, we expect the interaction between the water and the CNT to be weak. This is confirmed by taking the difference between
$\bar{V}_\mathrm{H}$ calculated with and without $\mathrm{H}_2\mathrm{O}$ molecule and observing that the resulting curve matches closely that of the water slab
calculated previously [see Fig.~\ref{Fig1a}(c)].
\begin{figure}
\center
\includegraphics[width=6.0cm,clip=true]{fig12}
\caption{(Color online) Transmission coefficient as function of energy for the Au(111)/CNT(3,3)-H$_2$O/Au(111) junction. The solid black curve corresponds
to the average $T$ over 201 MD snapshots, the dashed red curve is the transmission for the same junction in the dry and the dot-dashed green
line indicates the number of channels (spin-degenerate) for the infinite CNT. The grey curves, merging in a shadow, are $T(E;V=0)$ for each of the MD snapshot.}
\label{Fig12a}
\end{figure}
The transmission coefficients for all the 201 MD snapshots are shown in Fig. \ref{Fig12a} as super-imposed grey curves together with their average (solid black line),
the same quantity calculated in dry conditions (dashed red line) and the total number of open channel in the CNT (dot-dashed green line). In the figure we shift
$E_\mathrm{F}$ in such a way that $\bar{V}_\mathrm{H}$ for the infinite CNT matches $\bar{V}_\mathrm{H}$ of the CNT attached to Au without water in the middle of the
junction. The necessary shift is about 0.8~eV, which correctly corresponds to the difference in the work functions between the CNT and Au.
The main result is that the average transmission in wet conditions and that of the dry junction overlap almost exactly, demonstrating that in this case $\mathrm{H}_2\mathrm{O}$ has
no gating effect. This can be easily understood by recalling that, since the CNT has no polar edges, the average $\mathrm{H}_2\mathrm{O}$ conformation presents no net electrical
dipole, so that on average there is no shift of the CNT energy levels. Of course, each individual MD snapshot displays a dipole and the CNT energy levels get shifted.
This leads to fluctuations in the transmission. As a result of the dipole fluctuations, we find that that some of the sharp transmission peaks visible in the dry are
broadened up and have an average reduced height in solution. In some extreme cases (see for instance the sharp peak at about -1~eV) they are
completely washed out by the fluctuations.
Again in order to quantify the fluctuations of $T(E)$, we choose a particular molecular level (transmission peak) and follow its
time fluctuations. Here we select the well-defined peak at -1.7~eV below $E_\mathrm{F}$ and present its energy distribution histogram
in Fig.~\ref{Fig13a}. This time the peak position fluctuates over the tiny energy range of 0.06~eV, which is one order of magnitude smaller than
that of the HOMO of BDT (see Fig. \ref{Fig9a}). The origin of such small fluctuations is twofold: firstly the interaction between $\mathrm{H}_2\mathrm{O}$ and the CNT
is very weak due to the hydro-phobic nature of the nanotube and secondly the CNT $\pi$-like molecular states are delocalized, so that local fluctuations
in the electrostatic potential largely cancel out over the entire molecule. We also find that the difference between the average peak position (solid
black line in Fig. \ref{Fig13a}) and that in the dry (dashed red line) is only 0.01 eV. This is also much smaller then the same quantity calculated
for BDT (0.6~eV).
\begin{figure}
\center
\includegraphics[width=6.0cm,clip=true]{fig13}
\caption {(Color online) Histogram of the energy position of the transmission peak located at -1.7~eV below $E_\mathrm{F}$ (see Fig.~\ref{Fig12a}).
The histogram has been constructed from 201 MD snapshots. $N$ is the number of times the peak is found in the given energy window specified
by the bin width. The solid line indicates the time-averaged position, while the dashed red one marks the result in dry conditions.}
\label{Fig13a}
\end{figure}
Finally we discuss results for the insulating (8,0) CNT. The simulation cell is identical to that of the (3,3) case, but this time we have 288 C atoms
and 322 $\mathrm{H}_2\mathrm{O}$ molecules. The MD simulations are run for 20~ns and only 41 snapshots are taken within the last 16~ns. We have reduced the number of
snapshots from 201 to 41 because this time we do not perform a detailed statistical analysis of the peak position. Our main results are shown in Fig.~\ref{Fig14a},
where we present $T(E;V=0)$ for all the snapshots (grey lines), their average (solid black line), that in the dry conditions (dashed red line) together
with the total number of scattering channels. The transmission coefficient is plotted in logarithmic scale in order to emphasize
the tunneling behavior in the gap. In general the quantitative features of the Au/CNT(8,0)-H$_2$O/Au junction are similar to those of the
Au/CNT(3,3)-H$_2$O/Au one. In particular also here the average transmission almost overlaps with that of the dry junction meaning that there
is a negligible average gating.
\begin{figure}
\center
\includegraphics[width=6.0cm,clip=true]{fig14}
\caption{(Color online) Transmission coefficient as function of energy for the Au(111)/CNT(8,0)-H$_2$O/Au(111) junction. The solid (black) curve
corresponds to the average $T$ over 41 MD snapshots, the dashed (red) curve is the transmission for the same junction in the dry and the
dot-dashed (green) line indicates the number of channels (spin-degenerate) for the infinite CNT. The grey curves, merging in a shadow, are
$T(E;V=0)$ for each of the MD snapshot.}
\label{Fig14a}
\end{figure}
However for energies corresponding to the CNT gap, the transmission fluctuations are rather large due to the tunneling nature of the transport.
For instance $T(E_\mathrm{F};V=0)$ fluctuates between $6.0~10^{-5}$ and $2.0~10^{-3}$ within the 41 MD snapshots considered. This means that
in tunneling conditions the presence of $\mathrm{H}_2\mathrm{O}$ produces substantial variations in the instantaneous conductance amplitude. However and most
importantly the time-averaged transmission at $E_\mathrm{F}$ is only about 30\% larger than that of the dry junction. This gives us the important result that
in general the transport in CNTs is little affected by $\mathrm{H}_2\mathrm{O}$ solution regardless of the metallic state of the CNT.
\section{Conclusions}
In conclusion, we have investigated the effects of water on the transport properties of two types of molecules. This is done by combining classical
molecular dynamics with \textit{ab initio} electron transport calculations. Firstly, as an important technical result, we find that self-interaction
corrections are fundamental for describing the $\mathrm{H}_2\mathrm{O}$ ionization energy and its band-gap. This is a pre-requisite for quantitative transport calculations.
Then our main result is the finding that the $\mathrm{H}_2\mathrm{O}$-wetting conditions effectively produce electrostatic gating to the molecular junction, with a gating potential
determined by the time-averaged water dipole field. Such a field is rather large for the polar BDT molecule, resulting in an average transmission
spectrum shifted by about 0.6~eV with respect to that of the dry junction. In contrast, the hydro-phobic nature of the CNTs leads to almost negligible
gating, so that the average transmission spectrum for Au/CNT/Au is essentially the same as that in dry conditions, regardless of the CNT metallic state.
This suggests that CNTs can be used as molecular interconnects also in water wet situations, for instance as tips for scanning tunnel microscopy in
solutions or in biological sensors.
\section{Acknowledgments}
This work is sponsored by Science Foundation of Ireland (grants 07/RFP/PHYF235 and 07/IN.1/I945) and by the EU FP7 (NANODNA). Computational resources have been provided by KAUST. We thank C.D.~Pemmaraju for useful discussions.
\small
|
\section{Introduction}
\label{sec:intro}
Magnetic fields in the hot plasma associated
with groups and clusters of galaxies are poorly understood, but are thought to
play a vital role in regulating thermal conduction (e.g.\ Balbus 2000; Bogdanovic et
al. 2009) and influencing the dynamics of
cavities formed by radio jets (e.g.\ Dursi \& Pfrommer 2008; O'Neill et al.\ 2009).
The existence of magnetic fields can be demonstrated in several different ways
(e.g.\ Carilli \& Taylor 2002; Govoni \& Feretti 2004
and references therein).
One of the ways of studying these fields is
via the Faraday effect: rotation of the plane of linearly polarized radiation by
a magnetized plasma. Synchrotron emission from radio sources (either behind or
embedded within the group/cluster medium) can be used to probe the distribution
of foreground Faraday rotation. These can be combined with X-ray
observations (which provide the thermal gas density profile) to infer the strength
and fluctuation properties of the magnetic field.
Faraday rotation studies of clusters have been carried out using both
statistical samples of background radio sources (e.g. Lawler \& Dennison 1982, Vall\'ee et al. 1986,
Kim et al. 1990, Kim et al. 1991, Clarke et al. 2001) or individual radio
sources within the clusters
(e.g. Taylor \& Perley 1993; Feretti et al. 1995; Feretti et al. 1999a,b; Govoni et
al. 2001; Eilek \& Owen 2002; Pollack et al. 2005; Govoni et al. 2006; Guidetti
et al. 2008).
The central magnetic field strengths deduced from these data are usually a few
$\mu$G, but can exceed 10\,$\mu$G in the inner regions
of relaxed cool-core clusters (e.g.\ Taylor et al. 2002).
The RM distributions of radio galaxies in both interacting and relaxed clusters
are generally patchy, indicating that cluster magnetic
fields show structure on scales $\la 10$\,kpc.
Several studies of Abell clusters (Murgia et al. 2004; Govoni et al. 2006;
Guidetti et al. 2008) have shown that detailed RM images of radio galaxies can
be used to infer not only the strength of the cluster magnetic field, but also
its power spectrum. The analysis of Vogt \& En{\ss}lin (2003, 2005) suggests
that the power spectrum has a power law form with the slope appropriate for
Kolmogorov turbulence and that the auto-correlation length of the magnetic field
fluctuations is a few kpc. The deduction of a Kolmogorov slope could be
premature, however: there is a degeneracy between the slope and the outer
scale which is difficult to resolve with current Faraday-rotation data (Murgia
et al. 2004; Guidetti et al. 2008; Laing et al. 2008). Indeed, Murgia et
al. (2004) pointed out that shallower magnetic field power spectra are possible if
the magnetic field fluctuations have structure on scales of several tens of kpc. Recently,
Guidetti et al. (2008) showed that a power-law power spectrum with a Kolmogorov
slope and an abrupt long-wavelength cut-off at 35\,kpc gave a very good fit to
their Faraday rotation and depolarization data for the radio galaxies in A2382,
although a shallower slope extending to longer wavelengths was not ruled out.
While most work until recently has been devoted to rich clusters of galaxies,
little attention has been given in the literature to sparser environments,
although similar physical processes are likely to be at work. Faraday-rotation
fluctuations have previously been detected in galaxy groups (e.g. Perley et al.,
1984; Feretti et al. 1999a), but without deriving in
detail the geometry and structure of the magnetic field. The first detailed
work on galaxy groups was done by Laing et al. (2008), who analysed the radio
emission of 3C\,31.
They found that the three-dimensional magnetic-field power spectrum
$\widehat{w}(f)$,defined in Sect.~\ref{2d-general}, might be
described in terms of spatial frequency $f$ by a broken power law
$\widehat{w}(f)\propto D_0 f^{~-q}$ with $q=11/3$\footnote{$q=11/3$
is the slope of the three-dimensional power spectrum for Kolmogorov turbulence.} for
$f>$0.062\,arcsec $^{-1}$ (corresponding to a spatial scale of about 17\,kpc) and q=2.32 at lower frequencies, although a
power spectrum with a slope of 2.39 and an abrupt cut-off at high frequencies could
not be ruled out. Their results are qualitatively similar to those for sources
in Abell clusters.
Magnetic fields associated with galaxy groups deserve to be
investigated in more detail, since their environments are more representative than
those of rich clusters. Moreover, observations, analytical models and MHD simulations
of galaxy clusters all suggest
that the magnetic-field intensity should scale with the thermal
gas density
(e.g. Brunetti et al. 2001, Dolag 2006, Guidetti et al. 2008). A key
question is whether the relation between magnetic field strength and density in
galaxy groups is a continuation of this trend.\\
This paper presents a detailed analysis of Faraday rotation
in 3C\,449, a bright, extended radio source hosted by the central galaxy
of a nearby group. With the aim of shedding new light on the environment
around this source, we derive the statistical properties of the magnetic field
from observations of Faraday rotation, following the method developed by Murgia et al. (2004).
We use numerical simulations to predict the Faraday rotation for different
strengths and power spectra of the magnetic field.
The paper is organized as follows.
In Sect.\,\ref{general} the general properties of the radio source under investigation are presented.
Sect.\,\ref{vla} presents the radio images
on which our analysis is based. In Sect.\,\ref{sec:rm_obs}, we discuss the
observed Faraday rotation distribution of 3C\,449 and assess the contribution
from our Galaxy. The observed depolarization and its relation to
the RM properties are investigated in Sect.\,\ref{sec:dp}. Our two-dimensional
analysis of the structure of the RM fluctuations and a three-dimensional model of the magnetic
field consistent with these results are presented in
Sect.\,\ref{2d} and Sect.\,\ref{sec:model}, respectively. Sect.\,\ref{sec:sum} summarizes our conclusions and briefly compares the Faraday-rotation properties
of 3C\,449 with those of other sources.
Throughout this paper we assume a cosmology with $H_0$ = 71 km
s$^{-1}$Mpc$^{-1}$, $\Omega_m$ = 0.3, and $\Omega_{\Lambda}$ = 0.7, which
implies that 1\,arcsec corresponds to 0.342\,kpc at the distance of 3C\,449.
\section{The radio source 3C\,449: general properties}
\label{general}
In this paper, we image and model the Faraday rotation distribution across the
giant Fanaroff-Riley Class I (FR\,I; Fanaroff \& Riley 1974) radio source
3C\,449, whose environment is very similar to that of 3C\,31.
The optical counterpart of 3C\,449, UGC\,12064, is a dumb-bell galaxy and is the most prominent member of the group of galaxies 2231.2+3732 (Zwicky \& Kowal 1968).
The source is relatively nearby (z=0.017085, RC3.9, De Vaucouleurs et al. 1991)
and quite extended, both in angular (30\arcmin) and linear size, so it is an
ideal target for an analysis of the Faraday rotation distribution: detailed
images can be constructed that can serve as the basis of an accurate study of
magnetic field power spectra.
3C\,449 was one of the first radio galaxies studied in detail with the VLA (Perley et. al 1979).
High- and low-resolution radio data already exist
and the source has been mapped at many frequencies.
The radio emission of 3C\,449 (Fig.~\ref{XMM}) is elongated in the N--S direction and is characterized by long,
two-sided \textit{jets} with striking mirror symmetry close to the nucleus. The jets
terminate in well-defined inner \textit{lobes}, which fade
into well polarized \textit{spurs}, of which the southern one is more
collimated. The spurs in turn expand to form diffuse outer lobes.
The brightness ratio of the radio jets is very nearly 1, implying that they
are close to the plane of the sky if they are intrinsically symmetrical and have
relativistic flow velocities similar to those derived for other FR\,I jets (Perley et
al. 1979, Feretti et al. 1999a, Laing \& Bridle 2002).
In this paper, we therefore assume that the jets lie exactly in the plane of the sky,
simplifying the geometry of the Faraday-rotating medium.
Hot gas associated with the galaxy has been detected on both the group and galactic scales by X-ray imaging (Hardcastle et al. 1998; Croston et al. 2003).
These observations revealed deficits in the X-ray surface brightness
at the positions of the outer radio lobes, suggesting interactions with the surrounding
material. Fig.\,\ref{XMM} shows radio contours at 1.365\,GHz overlaid on the X-ray
emission as observed by the XMM-Newton satellite (Croston et al. 2003).
The X-ray radial surface brightness profile
of 3C\,449 derived from these data can be fitted with the sum of a point-source convolved with
the instrumental response and a
$\beta$ model (Cavaliere \& Fusco-Femiano, 1976),
\begin{equation}
\label{beta}
n_e(r) = n_0 (1+r^2/r_c^2)^{-\frac{3}{2}\beta}\,,
\end{equation}
where $r$, $r_c$ and $n_0$ are the
distance from the group X-ray centre, the group core radius and the central electron density, respectively.
Croston et al.\ (2008) found a best fitting model with $\beta = 0.42 \pm 0.05$,
$r_c$ = 57.1\,arcsec and $n_0 =3.7 \times 10^{-3}$cm$^{-3}$.
In what follows, we assume that the group gas density is described
by the model of Croston et al. (2008). The X-ray depressions noted by Croston et
al.\ (2003) are at distances larger than those at which we can measure linear polarization, and
there is no direct evidence of smaller cavities close to the nucleus. We
therefore neglect any departures from the spherically-symmetrical density model,
noting that this approximation may become increasingly inaccurate where the
source widens (i.e.\ in the inner lobes and spurs).
3C\,449 resembles 3C\,31 in environment and in radio morphology: both sources
are associated with the central members of groups of galaxies and their
redshifts are very similar.
The nearest neighbours are at a projected distances of about 30\,kpc
in both cases. Both radio sources have large angular extents, bending jets and
long, narrow tails with low surface brightnesses and steep spectra, although
3C\,31 appears much more distorted on large scales. There is one significant
difference: the inner jets of 3C\,31 are thought to be inclined by
$\approx$50$^\circ$ to the line of sight (Laing \& Bridle 2002), whereas those
in 3C\,449 are likely to be close to the plane of the sky (Feretti et al.\
1999a). We therefore expect that the magnetized foreground medium will be very
similar in the two sources, but that the geometry will be significantly
different, leading to a much more symmetrical distribution of Faraday rotation
in 3C\,449 compared with that observed in 3C\,31 by Laing et al.\ (2008).
\begin{figure*}
\centering
\includegraphics[width=13cm]{fig1_low.ps}
\caption[]{Radio contours of 3C\,449 at 1.365\,GHz superposed on the XMM-Newton X-ray image (courtesy of J. Croston,
Croston et al. 2003).
The radio contours start at 3$\sigma_{I}$ and increase by factors of 2. The restoring beam is 5.5\,arcsec FWHM.
The main regions of 3C\,449 discussed in the text are labelled.\label{XMM}}
\end{figure*}
\section{Total intensity and polarization properties}
\label{vla}
The Very Large Array (VLA\footnote{The Very Large Array is a facility of the National Science Foundation, operated under
cooperative agreement by Associated Universities, Inc.}) observations and their reduction have been presented by Feretti et al. (1999a).
The high quality of these data make this source suited for a very detailed analysis of the statistics of the Faraday rotation.
We produced total intensity ($I$) and polarization ($Q$ and $U$) images at
frequencies in the range 1.365 -- 8.385\,GHz from the combined, self-calibrated
u-v datasets described by Feretti et al.\ (1999a). The centre frequencies and
bandwidths are listed in Table~\ref{pol}. Each frequency channel was imaged
separately, except for those at 8.245 and 8.445\,GHz, which were averaged. The
analysis below confirms that these frequency-bandwidth combinations lead to
negligible Faraday rotation across the channels, as already noted by Feretti et
al. (1999a). All of the datasets were imaged with Gaussian tapering in the u-v
plane to give resolutions of 1.25\,arcsec and 5.5\,arcsec FWHM, {\cal CLEAN}ed
and restored with circular Gaussian beams. The first angular resolution is the
highest possible at all frequencies and provides good signal-to-noise for the
radio emission within 150\,arcsec ($\simeq$50\,kpc) of the radio core (the well
defined radio jets and the inner lobes), while minimizing beam
depolarization. The lower resolution of 5.5\,arcsec allows imaging of the
extended emission as far as 300\,arcsec ($\simeq$100\,kpc) from the core at
frequencies from 1.365 -- 4.985\,GHz (the 8.385-GHz dataset does not have
adequate sensitivity to image the outer parts of the source). We can therefore
study the structure of the magnetic field in the spur regions, which lie well
outside the bulk of the X-ray emitting gas. Noise levels for both sets of
images are given in Table~\ref{pol}. Note that the maximum scales of structure
which can be imaged reliably with the VLA at 8.4 and 5\,GHz are $\approx$180 and
$\approx$300\,arcsec, respectively (Ulvestad, Perley \& Chandler 2009). For
this reason, we only use the Stokes $I$ images for quantitative analysis
within half these distances of the core. The $Q$ and $U$ images have much less
structure on such large scales and are reliable to distances of $\pm$150\,arcsec
at 8.4\,GHz and $\pm$300\,arcsec at 5\,GHz, limited by sensitivity rather than
systematic errors due to missing flux as in the case of $I$ image.
Images of polarized intensity $P = (Q^2+U^2)^{1/2}$ (corrected for Ricean bias, following Wardle \&
Kronberg 1974),
fractional polarization $p=P/I$ and polarization angle $\Psi=(1/2)\arctan(U/Q)$ were derived from the $I$, $Q$, and $U$ images.
\begin{table*}
\caption{Parameters of the total intensity and polarization images. Col. 1:
Observation frequency. Col. 2: Bandwidth (note that the images at 8.385\,GHz are
derived from the average of two frequency channels, both with bandwidths of
50\,MHz, centred on 8.285 and 8.485\,GHz). Cols.3, 4 : rms noise levels in total
intensity ($\sigma_{I}$) and linear polarization ($\sigma_{QU}$, the average of
$\sigma_{Q}$ and $\sigma_{U}$) at 1.25\,arcsec FWHM resolution; Col. 5: mean
degree of polarization at 1.25\,arcsec; Col. 6, 7: rms noise levels for the
5.5\,arcsec images; Col. 8: mean degree of polarization at 5.5\,arcsec. We
estimate that the uncertainty in the degree of polarization, which is dominated
by systematic deconvolution errors on the $I$ images, is $\approx$0.02 at each
frequency.\label{pol}} \centering
\begin{tabular} {c c c c c c c c}
\hline\hline
$\nu$ & Bandwidth & \multicolumn{3}{c} {1.25 arcsec} & \multicolumn{3}{c} {5.5 arcsec} \\
& & $\sigma_{I}$ & $\sigma_{QU}$ & $\langle p \rangle$ & $\sigma_{I}$ & $\sigma_{QU}$ & $\langle p \rangle$ \\
(GHz) & (MHz) & (mJy/beam) & (mJy/beam) & & (mJy/beam) & (mJy/beam) & \\
\hline
&&&&&&&\\
1.365 & 12.5 & 0.037 & 0.030 & 0.24 & 0.018 & 0.014 & 0.26 \\
1.445 & 12.5 & 0.021 & 0.020 & 0.25 & 0.020 & 0.011 & 0.27 \\
1.465 & 12.5 & 0.048 & 0.049 & 0.25 & 0.019 & 0.013 & 0.25 \\
1.485 & 12.5 & 0.035 & 0.027 & 0.21 & 0.014 & 0.010 & 0.26 \\
4.685 & 50.0 & 0.017 & 0.017 & 0.32 & 0.017 & 0.013 & 0.37 \\
4.985 & 50.0 & 0.018 & 0.017 & 0.33 & 0.017 & 0.016 & 0.39 \\
8.385 & 100.0 & 0.014 & 0.011 & 0.31 & 0.015 & 0.013 & $-$ \\
&&&&&&&\\
\hline
\end{tabular}
\end{table*}
All of the polarization images (P, $p$, $\Psi$) at a given frequency were blanked
where the rms error in $\Psi >$ 10$^\circ$ at any frequency. We then calculated the scalar mean
degree of polarization $\langle p \rangle$ for each frequency and resolution;
the results are listed in Table~\ref{pol}.
The values of $\langle p \rangle$ are higher at 5.5\,arcsec resolution than at 1.25\,arcsec
because of the contribution of the extended and highly polarized emission which is not seen at the higher resolution.
At 1.25\,arcsec, where the beam depolarization is minimized, the mean fractional polarization shows a steady increase from 1.365 to 4.685\,GHz, where it reaches an average value of 0.32 and
then remains roughly constant at higher frequencies, suggesting that the
depolarization between 4.685 and 8.385\,GHz is insignificant.
\section{The Faraday rotation in 3C\,449}
\label{sec:rm_obs}
\subsection{Rotation measure images}
\label{sec:rm_images}
A magnetized, ionized medium rotates the plane of polarization of linearly
polarized radiation passing through it as follows:
\begin{equation}
\label{eq:rm}
\Delta\Psi = \Psi(\lambda) - \Psi_0 = \rm RM~\lambda^2\,,
\end{equation}
where $\Psi(\lambda)$ is the position angle observed at a wavelength $\lambda$ and
$\Psi_0$ is the intrinsic position angle. The rotation measure (RM) is related to the electron
density ($n_e$), the magnetic field along the line-of-sight ($B_{\parallel}$),
and the path-length ($L$)
through the Faraday-rotating medium according to:
\begin{equation}
\label{equaz}
{\rm RM_{~[rad/m^2]}}=812\int_{0}^{L_{[kpc]}}n_{e~[cm^{-3}]}B_{\parallel~[\mu G]}dl\,.
\end{equation}
Images of rotation measure can be obtained for radio sources by fitting to the polarization
angle as a function of $\lambda^2$, taking into account the well-known problem
of n$\pi$ ambiguities in the observed $\Psi$, as is done for example by the
{\cal AIPS} task {\cal RM}.
Removal of these ambiguities requires observations at many wavelengths well-spaced
in $\lambda^{2}$.
\begin{figure*}
\includegraphics[width=18cm]{fig2_low.ps}
\caption[]{(a): Image of the rotation measure of 3C\,449 at a resolution of
1.25\,arcsec FWHM,
computed at the seven
frequencies between 1.365 and 8.385\,GHz.
(b): Image of the rotation measure of 3C\,449 at a resolution of 5.5\,arcsec FWHM,
computed at the six frequencies between 1.365 and 4.985\,GHz.
In both of the RM images, the sub-regions used for the two-dimensional analysis of Sect.\,\ref{2d} are labelled.
(c) and (d): profiles of $\sigma_{\rm RM}~$ as a function of the projected distance from the radio source centre.
The points represent the values of $\sigma_{\rm RM}~$ evaluated in boxes as described in the text.
The horizontal and vertical bars represent the bin widths and the rms on the
mean expected from fitting errors,
respectively. Positive distances are in the direction of the north jet and the vertical dashed lines show the position of the nucleus.\label{highlowrm}
}
\end{figure*}
We produced images of RM and its associated rms error with resolutions of
1.25\,arcsec and 5.5\,arcsec (Fig.\,\ref{highlowrm}a and b) using a version of the {\cal AIPS} task {\cal RM}
modified by G. B. Taylor.
The 1.25\,arcsec-RM map was made
by combining the maps of polarization ${\bf E}$-vector ($\Psi$) at all seven
frequencies available to us, so our sampling of $\lambda^2$ is very good.
The RM map was calculated using a weighted least-squares fit
at pixels with polarization angle uncertainties $<$10\,$^{\circ}$ at all
frequencies. It is essentially the same as the RM image of Feretti et
al. (1999a), but with more stringent blanking.
The average fitting error is $\simeq$1.4\,rad\,m$^{-2}$ and is
almost constant over the whole RM image.
The image of RM at 5.5\,arcsec resolution
was produced using the polarization position angles at the 6 frequencies between
1.365 and 4.985\,GHz (see Table\,\ref{pol}), using the same blanking criterion
as at higher resolution.
Patches with different size are apparent in the 1.25\,arcsec resolution map, with
fluctuations down to scales of a few kpc.
The bulk of the RM values range from about $-$220\,rad\,m$^{-2}$ up to
$-$90\,rad\,m$^{-2}$, dominated by the Galactic contribution (see Sect.~\ref{sec:gal}).
The RM distribution peaks at $-$161.7\,rad\,m$^{-2}$, with rms dispersion
$\sigma_{\rm RM}~$=19.7\,rad\,m$^{-2}$. Note that we have not corrected the values of $\sigma_{\rm RM}~$
for the fitting error $\sigma_{\rm RM_{fit}}$. A first order correction would be
$\sigma_{\rm RM_{true}}=(\sigma_{\rm RM}^2-\sigma^2_{\rm RM_{fit}})^{1/2}$.
Given the low value for $\sigma_{\rm RM_{fit}}$, the effect of this correction
would be very small.
As was noted by Feretti et al. (1999a), the RM distribution in the inner jets is highly
symmetric about the core with RM $\simeq -197$ rad\,m$^{-2}$ at distances $\la$15\,arcsec.
The symmetry of the RM distribution in the jets is broken at larger distances from the core: while
the RM structure in the southern jet is homogeneous, with values around
~$-$130\,rad\,m$^{-2}$, fluctuations on scales of $\simeq$10\,arcsec ($\simeq$
3\,kpc) around a $\langle{\rm RM}\rangle~$ of ~$-$160\,rad\,m$^{-2}$ are present in the northern jet.
In both lobes, we observe similar patchy RM structures with
mean values $\langle{\rm RM}\rangle~$ $\simeq -164$\,rad\,m$^{-2}$ and $\sigma_{\rm RM}~$
$\simeq$16\,rad\,m$^{-2}$.
At 5.5\,arcsec resolution, more extended polarized regions of 3C\,449 can be
mapped with good sampling in $\lambda^2$. The average fitting error is $\simeq$1.0\,rad\,m$^{-2}$.
Both the spurs are characterized by $\langle{\rm RM}\rangle~$ $\simeq -$160\,rad\,m$^{-2}$, with $\sigma_{\rm RM}~$=15 and 10\,rad\,m$^{-2}$
in the north and south, respectively.
The overall mean and rms for the 5.5\,arcsec image, $\langle{\rm RM}\rangle~$ $= -$160.7\,rad\,m$^{-2}$ and $\sigma_{\rm RM}~$=18.9\,rad\,m$^{-2}$,
are very close to those determined at higher resolution and consistent with the
integrated value of $-162 \pm 1$\,rad\,m$^{-2}$ derived by Simard-Normandin et
al. (1981).
It was demonstrated by Feretti et al. (1999a) that the polarization position
angles at 1.25\,arcsec resolution accurately follow the relation $\Delta \Psi \propto \lambda^2$ over a
large range of rotation. We find the same
effect at lower resolution: plots of ${\bf E}$-vector position angle
$\Psi$ against $\lambda^2$ at representative points of the 5.5\,arcsec-RM image are shown in Fig\,\ref{fittini}.
As at the higher resolution, there are no significant deviations from
the relation $\Delta\Psi\propto\lambda^2$ over a range of rotation
$\Delta\Psi$ of 600\,$^{\circ}$, confirming that a foreground magnetized medium is responsible for
the majority of the Faraday rotation and extending this result to regions of
lower surface brightness.
\begin{figure}
\includegraphics[height=12.3cm]{fig3.ps}
\caption[]{Plots of ${\bf E}$-vector position angle $\Psi$ against $\lambda^2$ at representative points of the 5.5-arcsec RM map.
Fits to the relation $\Psi(\lambda) = \Psi_0 + {\rm RM}\lambda^2$ are shown. The values of
RM are given in the individual panels.\label{fittini}
}
\end{figure}
In Fig.\,\ref{highlowrm}(c) and (d), we show profiles of $\sigma_{\rm RM}~$ for both low and
high resolution RM images. The 1.25\,arcsec profile was obtained by averaging
over boxes with lengths ranging from 9 to 13\,kpc along the radio axis; for the
5.5\,arcsec profile we used boxes with a fixed length of 9\,kpc (these sizes were
chosen to give an adequate number of independent points per box).
The boxes extend far enough perpendicular to the source axis to include all unblanked pixels.
In both plots, there is clear evidence for a decrease in the observed $\sigma_{\rm RM}~$ towards the
periphery of the source, the value dropping from $\simeq$30\,rad\,m$^{-2}$ close
to the nucleus to $\simeq$10\,rad\,m$^{-2}$ at 50\,kpc. This is qualitatively as
expected for foreground Faraday rotation by a medium whose density (and presumably
also magnetic field strength) decreases with radius. The symmetry of the $\sigma_{\rm RM}~$
profiles is consistent with our assumption
that the radio source lies in the plane of the sky.
\subsection{The Galactic Faraday rotation}
\label{sec:gal}
For the purpose of this work,
3C\,449 has an unfortunate line of sight within our Galaxy.
Firstly, the source is located at $l=95.4$\,$^{\circ}$~, $b=-15.9$\,$^{\circ}$\ in Galactic coordinates,
where the Galactic magnetic field
is known to be aligned almost along the line of sight.
Secondly, there is evidence from radio and optical imaging for a diffuse,
ionized Galactic feature in front of 3C\,449, perhaps associated with the nearby
HII region S126 (Andernach et al.\ 1992).
Estimates of the Galactic foreground RM at the position of 3C\,449 from observations of other radio
sources are uncertain: Andernach et al.\ (1992) found a mean value of
$-$212\,rad\,m$^{-2}$ for six nearby sources, but the spherical harmonic
models of Dineen \& Coles (2005), which are derived by fitting to
the RM values of large numbers of extragalactic sources, predict
$-$135\,rad\,m$^{-2}$. Nevertheless, it is clear that the bulk of the mean RM
of 3C\,449 must be Galactic.
In order to investigate the magnetized plasma local to 3C\,449, we need to
constrain the value and possible spatial variation of this Galactic
contribution. The profiles of $\sigma_{\rm RM}~$ (Fig.~\ref{highlowrm}) show that the
small-scale fluctuations of RM drop rapidly with distance from the nucleus. We
might therefore expect the Galactic contribution to dominate on the largest scales.
At low resolution, we can determine the RM accurately out to $\approx$100\,kpc
from the core. This is roughly 5 core radii for the X-ray emission and
therefore well outside the bulk of the intra-group gas.
In order to estimate the Galactic RM contribution, we averaged the 5.5-arcsec RM
image in boxes of length 20\,kpc along the radio axis (the box size has been
increased from that of Fig.~\ref{highlowrm} to improve the display of
large-scale variations). The profile of $\langle{\rm RM}\rangle~$ against the distance from the
radio core is shown in Fig.\,\ref{low}. The large deviations from the mean in
the innermost two bins are associated with the maximum in $\sigma_{\rm RM}~$ and are almost
certainly due to the intra-group medium. The dispersion in $\langle{\rm RM}\rangle~$ is quite small
in the south and the value of $\langle{\rm RM}\rangle~$ = $-$160.7\,rad\,m$^{-2}$ for the whole
source is very close to that of the outer south jet. There are significant
fluctuations in the north, however. Given their rather small scale
($\sim$300\,arcsec), it is most likely that these arise in the local environment
of 3C\,449 and we include them in the statistical analysis given below.
There is some evidence for linear gradients in Galactic RM on arcminute scales:
Laing et al.\ (2006) found a gradient of magnitude
0.025\,rad\,m$^{-2}$\,arcsec$^{-1}$ along the jets of the radio galaxy NGC\,315
($l=124.6$\,$^{\circ}$~, $b=-32.5$\,$^{\circ}$). They argued that this gradient is
almost certainly Galactic in origin, since the amplitude of the linear variation
exceeds that of the small-scale fluctuations associated with NGC\,315. In order
to check the effect of a large-scale Galactic RM gradient on our results, we
computed an unweighted least-squares fit of a function $\langle{\rm RM}\rangle~$ $= {\rm RM}_{0} + ax$,
where $a$ and $\rm RM_{0}$ are constant and $x$ is measured along the radio
axis. The two innermost bins in Fig.\,\ref{low} were excluded from the fit. Our
best estimate for the gradient is very small:
$a$=0.0054\,rad\,m$^{-2}$\,arcsec$^{-1}$. We have
verified that subtraction of this gradient has a negligible effect on the
structure-function analysis given in Sect.~\ref{sec:sfuncov}.
We therefore adopt a constant value of $-$160.7\,rad\,m$^{-2}$ as the Galactic
contribution.
\begin{figure}
\includegraphics[height=8cm]{fig4.ps}
\caption[]{Profile of RM averaged over boxes of length 20\,kpc along the radio axis for the 5.5\,arcsec image. The horizontal bars represent the bin width. The vertical bars are the errors on the mean calculated from the dispersion in the boxes, the contribution from the fitting error is negligible and is not taken into account.
Positive distances are in the direction of the north jet. The black vertical dashed line indicates the position of the nucleus; the green dashed line shows our adopted mean value for the Galactic RM.\label{low}}
\end{figure}
\section{Depolarization}
\label{sec:dp}
Faraday rotation generally leads to a decrease of the degree of polarization
with increasing wavelength, or \textit{depolarization}.
We define DP$^{\lambda_{1}}_{\lambda_{2}}=p(\lambda_{1})/p(\lambda_{2})$,
where $p(\lambda)$ is the degree of polarization at a given wavelength
$\lambda$. We adopt the conventional usage in which \textit{higher} depolarization
corresponds to a \textit{lower} value of DP.
Laing (1984) has summarized the interpretation of polarization data. Faraday
depolarization of radio emission from radio sources can occur in three principal
ways:
\begin{enumerate}
\item thermal plasma is mixed with the synchrotron
emitting material (\textit{internal depolarization});
\item there are fluctuations of the foreground
Faraday rotation
across the beam (\textit{beam depolarization}) and
\item the polarization angle varies across the finite band of the
receiving system (\textit{bandwidth depolarization}).
\end{enumerate}
We first estimated the bandwidth effects on the polarized emission of 3C\,449
using the RM measurements from Sect.~\ref{sec:rm_images}. In
the worst case (the highest absolute RM value of $-$240\,rad\,m$^{-2}$) at the
lowest frequency of 1.365\,GHz) the
rotation across the band is $\approx$10\,$^{\circ}$. This results
in a depolarization of 1.7\%, negligible compared with errors due to noise.
If $\lambda^2$ rotation is
observed over a position-angle range $\gg$90$^{\circ}$,
then a foreground screen must be responsible for
the bulk of the observed RM.
In that case, depolarization can still
result from unresolved inhomogeneities of thermal density or magnetic field
in the surrounding medium.
\begin{figure*}
\includegraphics[width=18cm]{fig5_low.ps}
\caption[]{(a): image of the Burn law $k$ in rad$^2$\,m$^{-4}$
computed from a fit to the relation $p(\lambda)=p(0)\exp(-k\lambda^4)$ for seven frequencies between 1.365 and 8.385\,GHz.
(b): as (a) but the angular resolution is 5.5\,arcsec FWHM, and the $k$ image has been
computed from the fit to the six frequencies between 1.365 and 4.985\,GHz.
(c) and (d): profiles for $k$ as functions of the projected distance from the radio source centre (boxes as in Fig.\,\ref{highlowrm}).
The horizontal and vertical bars represent the bin widths and the error on
the mean, respectively. Positive distances are in the direction of the north jet and the vertical dashed lines show the position of the nucleus.\label{kb}}
\end{figure*}
Our analysis of the depolarization of 3C\,449 is based on the approach of Laing
et al.\ (2008). In the presence of a foreground Faraday screen with a small
gradient of RM across the beam, it is still possible to observe $\lambda^2$
rotation over a wide range of polarization angle and the wavelength dependence
of the depolarization is expected to follow the Burn law (Burn 1966):
\begin{equation}
\label{equadp}
p(\lambda)=p(0)\exp(-k\lambda^4),
\end{equation}
where $p(0)$ is the intrinsic value of the degree of polarization and
$k$=2$\arrowvert\nabla{RM}\arrowvert^2 \sigma^2$, with ${\rm FWHM} = 2\sigma
(2\ln 2)^{1/2}$. Since $k \propto\arrowvert\nabla{RM}\arrowvert^2 $, Eq.\,\ref{equadp}
clearly illustrates that higher RM gradients across the beam generate higher $k$
values and hence higher depolarization. The variation of $p$ with wavelength can
potentially be used to estimate fluctuations of RM across the beam which are
below the resolution limit. We can determine the intrinsic polarization $p(0)$
and the proportionality constant $k$ by a linear fit to the logarithm of the
observed fractional polarization as a function of $\lambda^4$.
We made images of $k$ at both standard resolutions by weighted least-squares
fitting to the fractional polarization maps, using
the FARADAY code by M. Murgia. The same frequencies were used as
for the RM images: 8.385 -- 1.365\,GHz and 4.985 --
1.365\,GHz at 1.25 and 5.5\,arcsec resolution, respectively.
By simulating the error distributions for $p$, we established
that the mean values of $k$ were biased significantly
at low signal-to-noise (cf.\ Laing et al. 2008), so only
data with $p>4\sigma_p$ at each frequency are included in the fits.
We estimate that any bias is negligible compared with the fitting error.
We also derived profiles of $k$ using the same sets of boxes as for
the $\sigma_{\rm RM}~$ profiles in Fig.~\ref{highlowrm}.
The 1.25\,arcsec resolution $k$ map is shown in
Fig.\,\ref{kb}(a), together with the profile of the $k$ values (Fig.\,\ref{kb}c).
The fit to a $\lambda^4$ law is very good everywhere: examples of
fits at selected pixels in the jets and lobes are shown in in
Fig.\,\ref{fittini_dph}. The symmetry observed in the $\sigma_{\rm RM}~$ profiles is also seen in
the 1.25\,arcsec $k$ image (Fig.\,\ref{kb}):
the mean values of $k$
are $\simeq$50\,rad$^2$\,m$^{-4}$ for both lobes, 107 and 82\,rad$^2$\,m$^{-4}$ for the
northern and southern jet, respectively. The region with the highest depolarization
is in the northern jet,
very close to the core
and along the west side.
The integrated value of $k$ at this resolution is $\simeq$56\,rad$^2$\,m$^{-4}$,
corresponding to a mean depolarization $DP^{\rm 20 cm}_{\rm 3 cm} \simeq 0.87$.
\begin{figure}
\includegraphics[height=8.5cm]{fig6.ps}
\caption[]{Plots of degree of polarization, $p$ (log scale) against $\lambda^4$ for representative points at 1.25-arcsec resolution.
Burn law fits (Eq.\,\ref{equadp}) are also plotted. The values of
$k$ are quoted in the individual panels.\label{fittini_dph}}
\end{figure}
\begin{figure}
\includegraphics[height=8.5cm]{fig7.ps}
\caption[]{Plots of degree of polarization, $p$ (log scale) against $\lambda^4$ for representative points at 5.5-arcsec resolution.
Burn law fits (equation\,\ref{equadp}) are also plotted. The values of
$k$ are quoted in the individual panels.\label{fittini_dpl}}
\end{figure}
The image and profile of $k$ at 5.5-arcsec resolution are shown in Fig.\,\ref{kb}(b) and (d).
The fit to a $\lambda^4$ law is in general
good and examples are shown in in Figs.\,\ref{fittini_dph}.
As mentioned earlier,
the maximum scale of structure imaged accurately in
total intensity at 5\,GHz is $\sim$300\,arcsec (100\,kpc) and there are
likely to be significant systematic errors in the degree of polarization on
larger scales. We therefore show the profile only for the inner $\pm$50\,kpc.
Over this range, the $k$ profiles are quite symmetrical, as at higher resolution.
Note also that the
small regions of very high $k$ at the edge of the northern and southern spurs in
the map shown in Fig.\,\ref{kb} are likely to be spurious.
The mean values of $k$
are $\simeq$184\,rad$^2$\,m$^{-4}$ and 178\,rad$^2$\,m$^{-4}$ for the northern and southern lobes, respectively;
and $\simeq$238 and 174\,rad$^2$\,m$^{-4}$ in the northern and in the southern spurs.
The integrated value of $k$ is $\simeq$194\,rad$^2$\,m$^{-4}$,
corresponding to a depolarization $DP^{\rm 20 cm}_{\rm 6 cm} \simeq 0.64$.
To summarize, we observe depolarization between 20\,cm and 3\,cm. Since we
measure lower values of $k$ at 1.25\,arcsec than 5.5\,arcsec, there is less
depolarization at high resolution, as expected in the case of beam
depolarization. The highest depolarization is observed in a region of the
northern jet, close to the radio core and associated with a large RM
gradient. Depolarization is significantly higher close to the nucleus,
consistent with the higher path length through the group gas observed in X-rays.
Aside from this global variation, we have found no evidence for detailed correlation of depolarization with
source structure. Depolarization and RM data are therefore both consistent with a
foreground Faraday screen. We show in Sect.\,\ref{sec:sfunc} that the residual depolarization at 1.25-arcsec resolution
can be produced by RM fluctuations on scales smaller than the beamwidth, but
higher-resolution observations are needed to establish this conclusively.
\section {Two Dimensional Analysis}
\label{2d}
\subsection{General considerations}
\label{2d-general}
In order to interpret the fluctuations of the magnetic field responsible for the
observed RM and depolarization of 3C\,449, we first discuss the statistics of the
RM fluctuations in two dimensions. We use the notation of Laing et al.\ (2008),
in which ${\bf f}=(f_{x},f_{y},f_{z})$ is a vector in the spatial frequency
domain, corresponding to the position vector ${\bf r} = (x, y, z)$. We take the
$z$-axis to be along the line of sight, so that the vector ${\bf r}_\perp = (x,
y)$ is in the plane of the sky and ${\bf f}_\perp=(f_{x},f_{y})$ is the
corresponding spatial frequency vector.
Our goal is to estimate the RM power spectrum $\widehat{C}({\bf
f}_{\perp})$, where $\widehat{C}({\bf f}_{\perp})df_{x}df_{y}$ is the power in
the area $df_{x}df_{y}$ and in turn to derive the three dimensional magnetic
field power spectrum $\widehat{w}({\bf f})$, defined so that $\widehat{w}({\bf
f})df_{x}df_{y}df_{z}$ is the power in a volume $df_{x}df_{y}df_{z}$ of
frequency space.
The relation between the magnetic field
statistics and the observed RM distribution is in general quite complicated,
depending on fluctuations in the thermal gas density, the geometry of the source
and the surrounding medium and the effects of incomplete sampling. In order to
derive the magnetic-field power spectrum, we make the following simplifying
assumptions, as in Guidetti et al. (2008) and Laing et al.\ (2008).
\begin{enumerate}
\item The observed Faraday rotation is due entirely to a foreground ionized medium (in
agreement with our results in Sects\,\ref{sec:rm_obs} and \ref{sec:dp}).
\item The magnetic field is an isotropic, Gaussian random variable, and can
therefore be characterized by a power spectrum $\widehat{w}(f)$ which is a
function only of scalar frequency $f$.
\item The form of the magnetic field power spectrum is independent of position.
\item The magnetic field is distributed throughout the Faraday-rotating medium, whose density
is a smooth, spherically symmetric function.
\item The amplitude of $\widehat{w}(f)$ is spatially variable, but is a function
only of the thermal electron density.
\end{enumerate}
These assumptions guarantee that the spatial distribution of the magnetic field
can be described entirely by its power spectrum $\widehat{w}(f)$, and that for a
medium of constant depth and density, the power spectra of magnetic field and RM
are proportional (En\ss lin and Vogt 2003).
\begin{figure*}
\includegraphics[height=14cm]{fig8.ps}
\caption[]{(a)-(f): Plots of the RM structure functions for the sub-regions showed in Fig.\ref{highlowrm}. The horizontal
bars represent the bin widths and the crosses the centroids for data included in the bins. The red lines are the predictions
for the CPL power spectra described in the text, including the effects of the convolving beam. The vertical error bars
are the rms variations for the structure functions derived using a CPL power
spectrum with the quoted value of $q$ on the observed grid of points for each sub-region.
(g)-(l): as (a)-(f) but using a BPL power spectra with fixed
slopes and break frequency, but variable normalization.
\label{sfunc}}
\end{figure*}
If the fluctuations are isotropic, the RM power spectrum $\widehat{C}(f_{\perp})$ is the Hankel transform of the autocorrelation function ${C}(r_{\perp})$, defined as
\begin{equation}
C(r_{\perp})= \langle \rm{RM}({\bf r_{\perp}}+{\bf r'_{\perp}}){\rm RM}({\bf r'_{\perp}}) \rangle,
\label{auto}
\end{equation}
where ${\bf r_{\perp}}$ and ${\bf r'_{\perp}}$ are vectors in the plane of the sky and $\langle\rangle$ is an average over ${\bf r'_{\perp}}$.
In an ideal case, it would be possible to derive the RM power spectrum and,
consequently, that of the magnetic field, directly from $C(r_{\perp})$. In
reality, the observations are affected first by the effects of convolution with
the beam, which modify the spatial statistics of RM, and secondly, by the
limited size and
irregular shape of the sampling region for 3C\,449, which results in a
complicated window function (En{\ss}lin \& Vogt, 2003) and limits the accuracy
with which the zero-level can be determined. In Sect.\,\ref{sec:gal}, we showed
that the
Galactic contribution to the 3C\,449 RM is large, and we argued that
a constant value of $-$160.7\,rad\,m$^{-2}$ is the best estimate for its value.
Fluctuations in the Galactic magnetic field on scales comparable with the size of
the radio sources could be present; conversely, the local environment of the
source might make a significant contribution to the mean RM. Both of these
possibilities lead to difficulties in the use of the autocorrelation function.
Laing et al.\,(2008) demonstrated a procedure which takes into account the convolution
effects and minimises the effects of uncertainties in the zero-level. In
particular, they showed that:
\begin{enumerate}
\item In the
short-wavelength limit (meaning that changes in Faraday rotation across the beam
are adequately represented as a linear gradient), the measured RM distribution is closely approximated by
the convolution of the true RM distribution with the observing beam.
\item The \textit{structure
function} is a powerful and reliable statistical tool to
quantify the two dimensional fluctuations of RM, given that it is independent
of the zero level and structure on scales larger than the area under
investigation.
\end{enumerate}
The structure function is defined by
\begin{equation}
S(r_\perp)=\rm{<[RM({\bf r}_\perp + {\bf r}_\perp^\prime)-RM({\bf r}_\perp^\prime)]^2>}
\label{sfunction}
\end{equation}
(Simonetti, Cordes \& Spangler, 1984; Minter and Spangler 1996).
It is related to the autocorrelation function $C(r_{\perp})$ for a
sufficiently large averaging region by $S(r_\perp) = 2[C(r_\perp)-C(0)]$.
Laing et al. (2008) also derived the effects of convolution with the observing
beam on the observed structure function. For the special case of a power-law
power spectrum (their Eq.\ B2), they showed that the observed structure function
after convolution can be heavily modified even at separations up to a few times
the FWHM of the observing beam. This effect must be taken into account when
comparing observed and predicted structure functions. Laing et al.\ (2008) and
Guidetti et al.\ (2008) also showed that numerical simulation of depolarization
provides complementary information on RM fluctuations on scales smaller than the
beam.
Following the approach of Laing et al. 2008, we initially used the RM structure
function to determine the form of $\widehat{w}(f)$ (Sect.~\ref{sec:sfunc}),
while for its normalization (determined by global variations of density and
magnetic field strength), we made use of three-dimensional simulations
(Sect.\,\ref{sec:model}).
\subsection{Structure functions}
\label{sec:sfunc}
We calculated the structure function for discrete regions of 3C\,449 over which
we expect the spatial variations of thermal gas density, rms magnetic field
strength and path length to be reasonably small. For each of these regions, we
first made unweighted fits of model structure functions derived from power
spectra with simple, parameterized functional forms, accounting for convolution
with the observing beam. We then generated multiple realizations of a Gaussian,
isotropic, random RM field, with the best-fitting power spectrum on the
observed grids, again taking into account the effects of the convolving
beam. Finally, we made a weighted fit using the dispersion of the synthetic
structure functions as estimates of the statistical errors for the \textit{observed}
structure functions, which are impossible to quantify analytically (Laing et
al. 2008). These errors, which result from incomplete
sampling, are much larger than those due to noise, but depend only weakly on the
precise form of the underlying power spectrum. Our measure of goodness of fit
is $\chi^2$, summed over a range of separations from $r_\perp =$ FWHM to roughly
half of the size of the region: there is no information in the structure
function for scales smaller than the beam, and the upper limit is set by
sampling. The errors are, of course, much higher at the large spatial scales,
which are less well sampled. Note, however, that estimates of the structure
function from neighbouring bins are not statistically independent, so it is not
straightforward to define the effective number of degrees of freedom.
We selected six regions for the structure-function analysis, as shown in
Fig.\,\ref{highlowrm}. These are symmetrically placed about the nucleus,
consistent with the orientation of the radio jets close to the plane of the sky.
For the north and south jets, we derived the structure functions only at
1.25-arcsec resolution, as the low-resolution RM image shows no additional
structure and has poorer sampling. For the north and south lobes, we computed
the structure functions at both resolutions over identical areas and compared
them. The agreement is very good, and the low-resolution RM images do not sample
significantly larger spatial scales, so we show only the 1.25-arcsec
results. Finally, we used the 5.5-arcsec RM images to compute the structure
functions for the north and south spurs, which are not detected at the higher
resolution.
The structure function has a positive bias given by 2$\sigma_{\rm noise}^2$
where $\sigma_{\rm noise}$ is the uncorrelated random noise in the RM image
(Simonetti, Cordes\& Spangler, 1984). The mean noise of the 1.25 and 5.5-arcsec RM
maps is $<$1\,rad\,m$^2$ and essentially uncorrelated on scales larger than the
beam. For each region we therefore subtracted 2$\sigma_{\rm noise}^2$ from the
structure functions, although this correction is always small. The
noise-corrected structure functions are shown in Fig.\,\ref{sfunc}.
The individual observed structure functions have approximately power-law
forms. Given that the structure function for a power-law power spectrum with no
frequency limits is itself a power law (Minter \& Spangler 1996; Laing et al.\
2008), we first tried to fit the observed data with a RM power spectrum of the
form
\begin{equation}
\label{pure}
\widehat{C}(f_{\perp})\propto~{f_{\perp}^{~-q}}
\end{equation}
over an infinite frequency range. This last assumption allows us to use the
analytical solution of the structure function, including convolution (Laing et
al. 2008) and therefore to avoid numerical integration.
The fits were quite good, but systematically gave slightly too much power on
small spatial scales and over-predicted the depolarization. We therefore
fit a \textit{cut-off power law} (CPL) power spectrum
\begin{eqnarray}
\widehat{C}(f_{\perp}) &= & 0 ~~~~~~~~~~~~~~~~
f_{\perp}<f_{\rm min}\nonumber \\
&= & C_{0}f_{\perp}^{~-q} ~~~~~~~~
f_{\perp}\leq{f_{\rm max}} \nonumber \\
&= & 0 ~~~~~~~~~~~~~~~~
f_{\perp}>f_{\rm max}\,.
\label{cpl}
\end{eqnarray}
Initially, we consider values of $f_{\rm min}$ sufficiently small
that their effects on the structure functions over the observed range of separations
are negligible.
The free parameters of the fit in this case are the slope, $q$, the cut-off
spatial frequency $f_{\rm max}$ and the normalization of the power spectrum,
$C_0$.
In Table\,\ref{fittingsf}, we give the best-fitting parameters for CPL fits to
all of the individual regions.
The fitted model structure functions are plotted in
Fig.\,\ref{sfunc}(a)--(f), together
with error bars derived from multiple realizations of the power spectrum as in
Laing et al.\,(2008).
\begin{table*}
\caption{CPL power spectrum parameters for the six individual sub-regions of 3C\,449 (lower and upper limits are quoted at $\sim$90\% confidence).
\label{fittingsf}}
\centering
\begin{tabular}{c c c c c c c c c c c}
\hline\hline
Region & FWHM & \multicolumn{9}{c} {CPL} \\
& (arcsec ) & \multicolumn{3}{c} {Best Fit} & \multicolumn{3}{c} {Min Slope}
& \multicolumn{3}{c} {Max Slope} \\
& & $q$ &$f_{\rm max}$ & & $q^{\rm -}$ & $f_{\rm max}$ & &
$q^{\rm +}$ & $f_{\rm max}$ & \\
\hline
N SPUR & 5.50 & 2.53 & 1.96 & & 1.58 & 0.23 & & 3.44 & $\infty$ \\
N LOBE & 1.25 & 2.87 & 1.60 & & 2.31 & 0.55 & & 3.35 & $\infty$ \\
N JET & 1.25 & 3.15 & 1.21 & & 2.29 & 0.30 & & 4.27 & $\infty$ \\
S JET & 1.25 & 2.76 & 1.95 & & 2.02 & 0.3 & & 3.69 & $\infty$ \\
S LOBE & 1.25 & 2.71 & 1.68 & & 2.36 & 0.65 & & 3.05 & $\infty$ \\
S SPUR & 5.50 & 2.17 & 1.53 & & 0.20 & 0.12 & & 3.95 & 0.12 \\
\hline\hline
\end{tabular}
\end{table*}
In order to constrain RM structure on spatial scales below the beamwidth, we
estimated the depolarization expected from the best power spectrum for each of
the regions with 1.25-arcsec RM images, following the approach of Laing et al.\
(2008). To do this, we made multiple realizations of RM images on an 8192$^2$
grid with fine spatial sampling. We then derived the $Q$ and $U$ images at our
observing frequencies, convolved to the appropriate resolution and compared the
predicted and observed mean degrees of polarization.
These values are given in
Table~\ref{fit_burnt}.
The uncertainties in the expected $<k>$ in Table~\ref{fit_burnt} represent statistical errors determined from multiple realizations of RM images with the same set of power spectrum parameters.
The predicted and observed values are in excellent agreement.
A constant value
of $f_{\rm max} = 1.67$\,arcsec$^{-1}$ predicts very similar values, also listed
in Table~\ref{fit_burnt}. We have not compared the depolarization data at
5.5-arcsec resolution in the spurs because of limited coverage of large spatial
scales in the $I$ images (Sect.~\ref{vla}), which is likely to introduce
systematic errors at 4.6 and 5.0\,GHz.
We performed a joint fit of the CPL power spectra, minimizing the $\chi^2$ summed over all six
sub-regions, giving equal weight to each and allowing the normalizations to vary
independently. In this case the free parameter of the fit are:
the six normalizations (one for each sub-region) the slope and the maximum spatial frequency.
The joint best-fitting single power-law power spectrum has $q = 2.68$.
A single power law slope does not give a good fit to all of the regions
simultaneously, however.
It is clear
from Fig.\,\ref{sfunc} and Table\,\ref{fittingsf} that there is a flattening in
the slope of the observed structure functions on the largest scales (which are
sampled primarily by the spurs).
In order to fit all of the data accurately with a single
functional form for the power spectrum, we adopt a
\textit{broken power law} form (BPL) for the RM power spectrum:
\begin{eqnarray}
\widehat{C}(f_{\perp}) &= & 0 ~~~~~~~~~~~~~~~~~~~~~~~~~~~~~
f_{\perp}<f_{\rm min}\nonumber \\
&= & D_{0}f_b^{(q_{\rm l}-q_{\rm h})}f_{\perp}^{~-q_{\rm{l}}} ~~~~~~~~~~~
f_{b}\geq{f_{\perp}} \nonumber \\
&= & D_{0}f_{\perp}^{~-q_{\rm{h}}} ~~~~~~~
f_{\rm max} \geq f_{\perp} > f_{b} \nonumber \\
&= & 0 ~~~~~~~~~~~~~~~~~~~~~~~~~~~~~
f_{\perp} > f_{\rm max}\,.
\label{bpl}
\end{eqnarray}
We performed a BPL joint fit, in the same way as for the CPL power spectra.
In this case the free parameters of the fit are:
the six normalizations, $D_0$, one for each sub-region, the high and low-frequency slopes, $q_{\rm
h}$ and $q_{\rm l}$, the break and maximum spatial frequency $f_{\rm b}$ and $f_{\rm
max}$.
We found best fitting parameters of $q_{\rm{l}}$= 2.07,
$q_{\rm{h}}$=2.98, $f_{\rm b}$=0.031\,$\rm{arcsec^{-1}}$.
As noted earlier, we
also fixed $f_{\rm max} = 1.67$\,arcsec$^{-1}$ to ensure consistency with the
observed depolarizations at 1.25-arcsec resolution.
The corresponding structure functions are plotted in Fig.\,\ref{sfunc}(g)--(l)
and the normalizations for the individual regions are given in
Table~\ref{fit_burnt}.
As for the CPL fits, the errors bars are derived from the rms scatter of the
structure functions of multiple convolved RM realizations.
It is evident from Fig.\,\ref{sfunc}
that the structure functions corresponding to the BPL power spectrum, which gives less power
on large spatial scales, are in much better agreement with the data.
The joint BPL fit has a $\chi^2$ of 17.7, compared with 33.5 for the joint CPL fit (the former has only
two extra parameters) confirming this result.
We have so far ignored the effects of any outer scale of magnetic-field
fluctuations. This is justified by the fact that the structure functions for the
spurs continue to rise at the largest observed separations, indicating that the
outer scale must be $\ga$10\,arcsec ($\simeq$30\,kpc). The model structure functions fit to the
observations assume that the outer scale is infinite and the realizations are
generated on sufficiently large grids in Fourier space that the effects of the
implicit outer scale are negligible over the range of scales we sample. We use
structure-function data for the entire source to determine an approximate value
for the outer scale in Sect.~\ref{sec:sfuncov}.
We now adopt the BPL power spectrum with these parameters and investigate the
spatial variations of the RM fluctuation amplitude using three-dimensional
simulations.
\begin{table*}
\caption{Best-fitting parameters for the joint CPL and BPL fits to all six sub-regions of
3C449 (lower and upper limits are quoted at $\sim$90\% confidence).
The values of $q$ and $f_{\rm max}$ for the joint CPL fit and $q_h$, $q_l$ and $f_{\rm b}$ for the joint BPL fit are the same for all sub-regions, while the normalizations are varied to minimize the overall $\chi^2$. In the joint BPL fit, the maximum frequency is fixed at $f_{\rm max}=
1.67$\,arcsec$^{-1}$.}
\centering
\begin{tabular}{c c c c c c c c c c c c c}
\hline\hline
& \multicolumn{5}{c} {Best Fit} & & \multicolumn{3}{c} {Min Slope}&
\multicolumn{3}{c} {Max Slope} \\
& $q$ &$f_{\rm max}$ & & $\chi^2$ & & $q^{\rm -}$ & $f_{\rm max}$ &
& $q^{\rm +}$ & $f_{\rm max}$ & \\
&&&&&&&&&&\\
joint CPL & 2.68 & 1.67 & & 33.5 & & 2.55 & 1.30 & & 2.81 & 2.00 & \\
\hline
&\multicolumn{5}{c} {Best Fit} & & \multicolumn{3}{c} {Min Slope} & \multicolumn{3}{c} {Max Slope} \\
& $q_l$ &$f_{\rm b}$ & $q_h$ & $\chi^2$ & & $q_l^{\rm -}$ & $f_{\rm b}$ & $q_h^{\rm -}$ &
$q_l^{\rm +}$ & $f_{\rm b}$ & $q_l^{\rm +}$ \\
&&&&&&&&&&\\
joint BPL & 2.07 & 0.031 & 2.97 & 17.7 & & 1.99 & 0.044 & 2.91 & 2.17 & 0.021 & 3.09 \\
&&&&&&&&&&\\
\hline\hline
\label{fittingcomb}
\end{tabular}
\end{table*}
\begin{table*}
\centering
\caption{Normalizations $C_0$ and $D_0$ for the individual fit parameters corresponding to the CPL, joint CPL and joint BPL fits at a resolution of 1.25\,arcsec.
Observed and expected depolarizations are also given.
Col.1: region; Col. 2: observed Burn law $<k>$. Col. 3, 4 and 5: normalization constant $C_0$, fitted maximum spatial frequency $f_{\rm max}$ for the best CPL power spectrum of each region and the predicted $<k>$ for each power spectrum.
Col. 6, 7 as Col. 3 and 5 but for the joint fit to each CPL power spectrum; Col. 8 and 9 as Col. 3, and 5 but for
the joint BPL power spectrum. For both the joint CPL and BPL fits, the maximum frequency is fixed at $f_{\rm max}=
1.67$\,arcsec$^{-1}$. In calculating each value of $<k>$ only data with $p>4\sigma_p$ are included.}
\label{fit_burnt}
\begin{tabular}{c c|c c c|c c|c c c}
\hline\hline
Region & Observed $<k>$ & \multicolumn{3}{c} {CPL} & \multicolumn{2}{c} {JOINT CPL} & & \multicolumn{2}{c} {JOINT BPL} \\
& & $C_0$ & $f_{\rm max}$ & $<k>$ & $C_0$ & $<k>$ & & $D_0$ & $<k>$ \\
& (rad$^2$\,m$^{-4}$) & & ($\rm arcsec^{-1}$) & (rad$^2$\,m$^{-4}$) & & (rad$^2$\,m$^{-4}$) & & & (rad$^2$\,m$^{-4}$) \\
\hline
&&&&&&\\
N LOBE & 61$\pm$6 & 0.96 & 1.60 & 63$\pm$3 & 1.52 & 66$\pm$3 & & 1.91 & 52$\pm$4\\
N JET & 106$\pm$12 & 1.34 & 1.21 & 106$\pm$5 & 4.76 & 110$\pm$5 && 0.5 & 109$\pm$5 \\
S JET & 91$\pm$11 & 1.50 & 1.95 & 70$\pm$5 & 1.94 & 73$\pm$5 & & 1.52 & 65$\pm$4 \\
S LOBE & 50$\pm$5 & 1.18 & 1.68 & 53$\pm$2 & 1.28 & 50$\pm$2 & & 2.20 & 45$\pm$3 \\
&&&&&&\\
\hline\hline
\end{tabular}
\end{table*}
\section{Three-dimensional analysis}
\label{sec:model}
\subsection{Models}
\label{3Dcode}
We used the software package {\cal FARADAY} (Murgia et al. 2004) to compare the
observed RM with simulated images derived from three-dimensional multi-scale
magnetic-field models. Given a field model and the density distribution of the
thermal gas, {\cal FARADAY} calculates an RM image by integrating
Eq.\,\ref{equaz} numerically. As in Sect.~\ref{2d}, we model the fluctuations of RM on the
assumption that the magnetic field responsible for the foreground rotation is an
isotropic, Gaussian random variable and therefore characterized entirely by its
power spectrum. Each point in a cube in Fourier space is first
assigned components of the magnetic vector potential. The amplitudes are
selected from a Rayleigh distribution of unit variance and the phases are random
in $[0, 2\pi]$. The amplitudes are then multiplied by the square root of the
power spectrum of the vector potential, which is simply related to that of the
magnetic field. The corresponding components of the magnetic field along the
line of sight are then calculated and transformed to real space. This procedure
ensures that the magnetic field is divergence-free. The field components in real
space are then multiplied by the model density distribution and integrated along
the line of sight to give a synthetic RM image at the full resolution of the
simulation, which is then convolved to the observing resolution.
For 3C\,449, we assumed that the source is in a plane perpendicular to the line
of sight which passes through the group centre and simulated the field and
density structure using a 2048$^3$ cube with a real-space pixel size of
0.1\,kpc. We used the best-fitting BPL power spectrum found in
Sect.\,\ref{sec:sfunc}, but with a spatially-variable normalization, as
described below (Sect.~\ref{sec:radial}), and a low-frequency cut-off $f_{\rm
min}$, corresponding to a maximum scale of the magnetic field
fluctuations,\footnote{Here we refer to the scale length $\Lambda$ as a complete
wavelength, i.e. $\Lambda = 1/f$. This differs by a factor of 2 from the
definition in Guidetti et al. 2008, where $\Lambda$ is the reversal scale of the
magnetic field, so $\Lambda = 1/2f$.} $\Lambda_{\rm max}$ ($ = f_{\rm
min}^{-1}$). The power spectrum of Eq.~\ref{bpl} is then set to 0 for $f <
f_{\rm min}$. We fixed the minimum scale of the fluctuations $\Lambda_{\rm
min} = 0.2$\,kpc. This is equivalent to the value $f_{\rm max} =
1.67$\,arcsec$^{-1}$ found in Sect.\,\ref{sec:sfunc} and also consistent with
the requirement that the minimum scale can be no larger than twice the pixel
size for adequate sampling.
We made multiple synthetic RM images at resolutions of 1.25 and 5.5\,arcsec over
the fields of view of the observations for each combination of parameters. In
order to estimate the spatial variation of the magnetic-field strength, we first
made a set of simulations with a large, fixed value of $\Lambda_{\rm max}$ and
compared the predicted and observed profiles of $\sigma_{\rm RM}~$
(Sect.~\ref{sec:radial}). We then fixed the radial variation of the field at its
best-fitting form and estimated the value of $\Lambda_{\rm max}$ using a
structure-function analysis for the whole source (Sect.~\ref{sec:sfuncov}).
\begin{table*}
\caption{Summary of magnetic field power spectrum and density scaling parameters.\label{simul}}
\centering
\begin{tabular}{c c}
\hline
\hline
\multicolumn{2}{c}{BPL power spectrum}\\
\hline
$q_{\rm l}=$2.07 & low-frequency slope \\
$q_{\rm h}$=2.98 & high-frequency slope \\
$f_{b}$=$\rm 0.031\,arcsec^{-1}$ & break frequency ($\Lambda_b=1/{f_{b}}$=11\,kpc)\\
$f_{\rm max}$=$\rm 1.67\,arcsec^{-1}$ & maximum frequency ($\Lambda_{\rm min}=1/{f_{\rm max}}$=0.2\,kpc)\\
$f_{\rm min}$ fitted & minimum frequency ($\Lambda_{\rm max}=1/{f_{\rm min}}$) \\
\hline
\multicolumn{2}{c}{Scaling of the magnetic field}\\
\hline
$B_{0}$ fitted & Average magnetic field at group centre\\
$\eta$ fitted & Magnetic field exponent of the radial profile: $\langle B\rangle(r)=B_1186
{0}\left[\frac{n_e(r)}{n_0}\right]^{\eta}$ \\
&\\
\hline
\hline
\label{param}
\end{tabular}
\end{table*}
\subsection{Magnetic field strength and radial profile}
\label{sec:radial}
In order to estimate the radial variation of field strength, we first fixed the
value of the outer scale to be $\Lambda_{\rm max} = 205$\,kpc, the largest
allowed by our simulation grid. Our approach was to make a large number of
simulations for each combination of field strength and radial profile and to
compare the predicted and observed values of $\sigma_{\rm RM}~$ evaluated over the boxes used
in Sect.~\ref{sec:rm_images} (Fig.~\ref{highlowrm}). We used $\chi^2$ summed
over the boxes as a measure of the goodness of fit. This
procedure is independent of the precise value of the outer scale provided that
it is much larger than the averaging boxes. We express our results in
terms of $\chi^2_{\rm
red}$, which is the value of $\chi^2$ divided by the number of degrees of
freedom.
We initially tried a radial field-strength
variation of the form:
\begin{equation}\label{br}
\langle B^2(r)\rangle^{1/2} = B_{0} \left[\frac{n_e(r)}{n_0}\right]^{~\eta}
\end{equation}
as used by Guidetti et al. (2008) and Laing et al. (2008). Here, $B_{0}$ is the
rms magnetic field strength at the group centre and $n_{e}(r)$ is the thermal
electron gas density, assumed to follow the $\beta$-model profile derived by
Croston et al. (2008; see Sect.\,\ref{general}).
This functional form is consistent with other observations, analytical
models and numerical simulations. In particular, $\eta = 2/3$
corresponds to flux-freezing and $\eta = 1/2$ to equipartition between thermal
and magnetic energy. Dolag et al. (2001) and Dolag (2006) found $\eta \approx 1$
from the correlation between the observed rms RM and X-ray surface brightness in
galaxy groups and clusters and showed that this is consistent with the results of
MHD simulations.
We produced simulated RM images for each combination of $B_0$ and $\eta$ in the ranges
0.5 -- 10\,$\mu$G in steps of 0.1\,$\mu$G and 0 -- 2 in steps of 0.01, respectively.
We then derived the synthetic $\sigma_{\rm RM}~$ profiles and, by comparing them with
the observed one, calculated the unweighted $\chi^2$.
We repeated this procedure 35 times at each angular resolution,
noting the ($B_0$, $\eta$) pair which gave the lowest $\chi^2$ in each case.
These values are plotted in Fig.\,\ref{first_cd}.
As in earlier work (Murgia et al. 2004; Guidetti et
al. 2008; Laing et al. 2008), we found a degeneracy between values of $B_{0}$ and
$\eta$, in the sense that the fitted values are positively correlated, but there
are clear minima in $\chi^2$ at both resolutions. We therefore adopted the
mean values of $B_0$ and $\eta$, weighted by $1/\chi^2$, as the best overall
estimates. These are also plotted in Fig.\,\ref{first_cd}
as blue crosses.
Although the central magnetic field strengths derived for the two RM images are
consistent at the 1$\sigma$ level ($B_{0}$=2.8$\pm$0.5\,$\mu$G and $B_{0}$=
4.1\,$\pm$1.2$\mu$G at 5.5 and 1.25-arcsec resolution, respectively), the
values of $\eta$ are not. The best-fitting values are $\eta$=0.0$\pm$0.1 at
5.5\,arcsec FWHM and $\eta$=0.8$\pm$0.4 at
1.25\,arcsec FWHM.
We next produced 35 RM simulations at each angular resolution by fixing $B_0$ and $\eta$
at their best values for that resolution. This allowed us to calculate
weighted $\chi^2$'s for the $\sigma_{\rm RM}~$ profiles, evaluating the errors for
each box by summing in quadrature the rms due to sampling (determined from the dispersion in
the realizations) and the fitting error of the observations.
These values are listed in Table\,\ref{3dfit}.
The observed and best-fitting model profiles at both the angular resolutions
are shown in Fig.\,\ref{first}.
A model with $\eta \simeq 0$ at all radii is unlikely a priori:
previous work has found values of $0.5 \la \eta \la 1$ in other sources
(Dolag 2006; Guidetti et al. 2008; Laing et al. 2008). The most likely
explanation for the low value of $\eta$ inferred from the low-resolution image
is that the electron density distribution is not well represented by a
$\beta$-model (which describes a spherical and smooth distribution) at large
radii. In support of this idea, Fig.\,\ref{XMM} shows that the morphology of the
X-ray emission is not spherical at large radii, but quite irregular. Croston et
al. (2003) and Croston et al. (2008) pointed out that the quality of the fit of a single $\beta$-model to
the X-ray surface brightness profile was poor in the
outer regions, suggesting small-scale deviations in the gas distribution. The
single $\beta$-model gave a better fit to the inner region of the X-ray surface
brightness profile, where the polarized emission of 3C\,449 can be observed at
1.25-arcsec resolution.
Our aim is to fit the $\sigma_{\rm RM}~$ profiles at both resolutions with the same
distribution of
$n_{e}(r) \langle B(r)^2\rangle^{1/2}$.
In the rest of this subsection we assume that the density profile
$n_{e}(r)$ is still represented by the single $\beta$-model,
even though we have argued that it might not be appropriate
in the outer regions of the hot gas distribution.
Although the resulting estimates of field strength at large radii
may be unreliable, the fit is still necessary for the calculation of
outer scale described in Sec.\,\ref{sec:sfuncov}, which depends only on
the combined spatial variation of density and field strength.
The best-fitting model at 1.25-arcsec resolution,
which is characterized by a more physically reliable $\eta$, gives a very bad
fit to the low resolution profile at almost all distances from the
core (Fig.\,\ref{first}a). Conversely, the model determined at 5.5-arcsec resolution
gives a very poor fit to the sharp peak in $\sigma_{\rm RM}~$ observed
within 20\,kpc of the nucleus at 1.25-arcsec resolution, where the radio and
X-ray data give the strongest constraints (Fig.\,\ref{first}b). We
have also verified that no single intermediate value of $\eta$ gives an adequate
fit to the $\sigma_{\rm RM}~$ profiles at all distances from the nucleus.
A better description of the observed $\sigma_{\rm RM}~$ profile is provided by the empirical
function:
\begin{eqnarray}
\langle B^2(r)\rangle^{1/2} &= & B_{0}\left[\frac{n_e(r)}{n_0}\right]^{~\eta_{int}} ~~~~~~~
r\leq{r_{\rm m}} \nonumber \\
&= & B_{0} \left[\frac{n_e(r)}{n_0}\right]^{~\eta_{out}} ~~~~~~
r>r_{\rm m}\,,~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~\nonumber \\
\label{brb}
\end{eqnarray}
where $\eta_{\rm int}$, $\eta_{\rm out}$ are the inner and outer scaling index of the
magnetic field and $r_{\rm m}$ is the break radius.
We fixed $\eta_{\rm int}$=1.0 and $\eta_{\rm out}$=0.0, consistent with
our initial results, in order to reproduce both the inner sharp peak and the
outer flat decline of the $\sigma_{\rm RM}~$ observed at the two resolutions, keeping $r_{\rm
m}$ as a free parameter. We made three sets of three-dimensional simulations
for values of the outer scale $\Lambda_{\rm max}$= 205, 65 and
20\,kpc. Anticipating the result of Sect.~\ref{sec:sfuncov}, we plot the results
only for $\Lambda_{\rm max}$= 65\,kpc, but the derived $\sigma_{\rm RM}~$ profiles are in any
case almost independent of the value of the outer scale in this range. The new simulations
were made only at a resolution of 5.5\,arcsec, since the larger field of view at
this resolution is essential to define the change in slope of the profile.
In order to determine the best-fitting break radius, ${r_{\rm m}}$ in
Eq.\,\ref{brb}, we produced 35 sets of synthetic RM images for a grid of values
of $B_0$ and ${r_{\rm m}}$ for each outer scale, noting the pair of values which gave the minimum
unweighted $\chi^2$ for the $\sigma_{\rm RM}~$ profile for each set of simulations. These
values are plotted in Fig.~\ref{200_e}, which shows that there is a degeneracy
between the break radius $r_{\rm m}$ and $B_0$. As with the similar degeneracy
between $B_0$ and $\eta$ noted earlier, there is a clear minimum in $\chi^2$,
and we therefore adopted the mean values of $B_0$ and $r_{\rm m}$ weighted by
$1/\chi^2$ as our best estimates of the magnetic-field parameters.
\begin{figure*}
\centering
\includegraphics[width=11cm]{fig9.eps}
\caption[]{(a) and (b): Distributions of
the best-fitting values of $B_{0}$ and $\eta$ from 35 sets of simulations,
each covering ranges of 0.5 -- 10\,$\mu$G in $B_0$ and 0 -- 2 in $\eta$, at
5.5 and 1.25\,arcsec respectively. The sizes of the circles are proportional to
$\chi^2$ for the fit
and the blue crosses represent the means of the distributions
weighted by $1/\chi^2$.
The plot shows the expected degeneracy between $B_{0}$ and
$\eta$. \label{first_cd}}
\end{figure*}
\begin{figure*}
\centering
\includegraphics[width=11cm]{fig10.eps}
\caption[]{(a): Observed and synthetic radial profiles for rms Faraday $\sigma_{\rm RM}~$ at
5.5\,arcsec as functions of the projected distance from the radio source
centre. The outer scale is $\Lambda_{\rm max}$=205\,kpc.
The black points represent the data with vertical bars
corresponding to the rms error of the RM fit. The magenta triangles represent the
mean values from 35 simulated profiles and the vertical bars are the rms
scatter in these profiles due to sampling. (b) as (a) but
at 1.25\,arcsec.\label{first}}
\end{figure*}
We then made 35 simulations with the best-fitting values of $B_0$ and
$\eta$ for each outer scale and evaluated the weighted $\chi^2$'s for the
resulting $\sigma_{\rm RM}~$ profiles. All three values of $\Lambda_{\rm max}$ we
investigated give reasonable fits to the observed $\sigma_{\rm RM}~$ profile along the whole
radio source. The fit for $\Lambda_{\rm max} = 65$\,kpc is marginally better
than for the other two values ($\chi^2_{\rm red} = 1.8$), consistent with the results of
Sect.~\ref{sec:sfuncov}, below. In
this case the central magnetic field strength is 3.5$\pm1.2$\,$\mu$G and the
break radius is 16$\pm$11\,kpc. For the power spectrum with $\Lambda_{\rm max}
= 65$\,kpc and these best-fitting parameters, we also produced three-dimensional
simulations at a resolution of 1.25\,arcsec. Even though the fitting procedure
is based only on the low-resolution data, this model also reproduces the
1.25-arcsec profile very well ($\chi^2_{\rm red}$=0.7). Combining the values of
$\chi^2$ for the two resolutions, using the 1.25-arcsec profile close to the
core and the 5.5-profile at larger distances, we find $\chi^2_{\rm red}$=1.8.
Fig.\,\ref{200} shows a comparison of the observed radial profiles
for rms Faraday $\sigma_{\rm RM}~$ and $\langle{\rm RM}\rangle~$ with the synthetic ones
derived for this model.
The synthetic $\sigma_{\rm RM}~$ profile plotted in Fig.~\ref{200} is the mean over 35 simulations, and may
be compared directly with the observations.
In contrast, the $\langle{\rm RM}\rangle~$ profile is derived from a single example realization.
It is important to emphasize that the latter is {\em one example of a random process}, and is
not expected to fit the observations; rather, we aim to compare the fluctuation
amplitude as a function of position.
The values of $B_0$ and $\chi^2_{\rm red}$ for all
of the three-dimensional simulations, together with $\eta$ and $r_{\rm m}$ for
the single and double power-law profiles, respectively, are summarized in
Table\,\ref{3dfit}.
\begin{table*}
\centering
\caption{Results from the three-dimensional fits at both the angular resolutions of 1.25 and 5.5\,arcsec.\label{3dfit}}
\begin{tabular}{c c c c c c c c}
\hline\hline
FWHM & $\Lambda_{\rm max}$ & \multicolumn{3}{c} {single $\eta$} & \multicolumn{3}{c} {broken $\eta$} \\
(arcsec) & (kpc) & $B_{0}$ ($\mu$G) & $\eta$ & $\chi^2_{\rm red}$ & $B_{0}$ ($\mu$G) & $r_{\rm m}$(kpc) & $\chi^2_{\rm red}$\\
\hline
1.25 & 205 & 4.1$\pm$1.1 & 0.8$\pm$0.4 & 0.48 & - & - & - \\
5.50 & 205 & 2.8$\pm$0.5 & 0.1$\pm$0.1 & 2.2 & 3.5$\pm$0.7 & 17$\pm$9 & 1.9 \\
5.50 & 65 & - & - & - & 3.5$\pm$1.2 & 16$\pm$11 & 1.8 \\
5.50 & 20 & - & - & - & 3.5$\pm$0.8 & 11$\pm$8 & 2.1 \\
\hline
\end{tabular}
\end{table*}
\begin{figure*}
\centering
\includegraphics[height=6cm]{fig11.ps}
\caption[]{Distribution of the best-fitting values of $B_{0}$ and $r_{\rm m}$ from
35 sets of simulations, each covering ranges of 0.5 -- 10\,$\mu$G in $B_0$ and 0
-- 50\,kpc in $r_{\rm m}$. The sizes of the circles are proportional to $\chi^2$ and
the blue cross represents the means of the distribution
weighted by $1/\chi^2$.
The plot shows the degeneracy between $B_{0}$ and $r_{\rm m}$ described in the
text.}
\label{200_e}
\end{figure*}
\begin{figure*}
\centering
\includegraphics[height=10.5cm]{fig12.eps}
\caption[]{Comparison between observed and synthetic profiles of rms and mean
Faraday rotation at resolutions of 5.5\,arcsec (a and b) and 1.25\,arcsec (c and
d). The synthetic profiles are derived from the best-fitting model with
$\Lambda_{\rm max} = 65$\,kpc. The black points represent the data with
vertical bars corresponding to the rms fitting error.
(a) and (c): Profiles of $\sigma_{\rm RM}~$. The magenta
triangles represent the mean values from 35 simulations at 5.5-arcsec resolution
and the vertical bars are the rms scatter in these profiles due to sampling.
(b) and (d): Profiles of $\langle{\rm RM}\rangle~$ derived from single example realizations at 5.5 and 1.25-arcsec resolution.
}
\label{200}
\end{figure*}
\begin{figure*}
\centering
\includegraphics[width=9cm]{fig13_low.ps}
\caption[]{Comparison of observed and representative synthetic distribution of Faraday RM at 1.25\,arcsec and 5.5\,arcsec.
The synthetic images have been produced for the best-fitting model with $\Lambda_{\rm max}$=65\,kpc.
The colour scale is the same for all displays.
\label{comp}}
\end{figure*}
\subsection{The outer scale of magnetic-field fluctuations}
\label{sec:sfuncov}
The theoretical RM structure function for uniform field strength, density and
path length and a power spectrum with a low-frequency cut-off should
asymptotically approach a constant value ($2\sigma^2_{\rm RM}$ for a large enough averaging
region) at separations $\ga \Lambda_{\rm max}$.
The observed RM structure function
of the whole source is heavily modified from this theoretical one by the scaling
of the electron gas density and magnetic field at large separations, which acts
to suppress power on large spatial scales.
In Sect.\,\ref{2d} we therefore
limited the study of the structure function to sub-regions of 3C\,449 in which
uniformity of field strength, density and path length (and therefore of the
power-spectrum amplitude) is a reasonable assumption, inevitably limiting our
ability to constrain the power spectrum on the largest scales.
Now that we have an adequate model for the variation of $n_e(r)\langle B^2(r)
\rangle^{1/2}$ with radius (Eq.~\ref{brb}), we can correct for it to derive
what we call the pseudo-structure function -- that is the structure function for
a power-spectrum amplitude which is constant over the source. This can be
compared directly with the structure functions derived from the Hankel transform
of the power spectrum. To evaluate the pseudo-structure function, we
divided the observed 5.5-arcsec RM image by the function:
\begin{equation}
\label{detrend}
\left [\int_{0}^{L}n_{e}(r)^2B(r)^2dl\right ]^{1/2}
\end{equation}
(En{\ss}lin \& Vogt 2003), where the radial variations of $n_e$ and B are those of the best-fitting model
(Sect.\,\ref{sec:model}) and the upper integration limit $L$ has a length of 10 times the core radius.
The integral was normalized to unity at
the position of the radio core: this is equivalent to fixing the field and gas
density at their maximum values and holding them constant everywhere in the
group. The normalization of the pseudo-structure function should then be quite
close to that for the two central jet regions.
The pseudo-structure function is shown in Fig.\,\ref{sfuncov}(a) together
with the predictions for the BPL power spectra with $\Lambda_{\rm max}$=205, 65 and 20\,kpc.
As expected, the normalization of the pseudo-structure function is consistent
with that of the jets (Fig.\,\ref{sfunc}a and b).
The comparison between the synthetic and observed pseudo-structure functions
indicates firstly that they agree very well at small separations, independent of
the value of $\Lambda_{\rm max}$. This confirms that the best BPL power
spectrum found from a combined fit to all six sub-regions is a very good fit
over
the entire source. Secondly, despite the poor sampling on very large scales,
the asymptotic values of the predicted structure functions for the three values
of $\Lambda_{\rm max}$ are sufficiently different from each other that we can
determine an approximate outer scale. The model with
$\Lambda_{\rm max}$=65\,kpc gives the best representation of the data. The fit
is within the estimated errors except for a marginal discrepancy at the very
largest (and therefore poorly sampled) separations. That
with $\Lambda_{\rm max}$=20\,kpc is inconsistent with the observed
pseudo-structure function for any separation $\ga$20\,arcsec ($\simeq$6\,kpc) where
the sampling is still very good, and is firmly excluded. The model with
$\Lambda_{\rm max}$=205\,kpc has slightly, but significantly too much power on
large scales.
We emphasize that our estimate of the outer scale of the RM
fluctuations is essentially independent of the functional form assumed for the
variation of the field strength with radius in the central region, which
affects the structure function only for small separations. Our results are
almost identical if we fit the field-strength variation with
either the profile of Eq.\,\ref{br} (with $\eta \approx 0$) or that of Eq.\,\ref{brb}.
As for the structure functions of individual regions, the pseudo-structure
function at large separations is clearly affected by poor sampling: this
increases the errors but does not produce any bias in the derived
values. At large radii, however, the integral in equation~\ref{detrend} becomes
small, so the noise on the RM image is amplified. This is a potential source of
error, and we have therefore checked our
results using numerical simulations.
We calculated the mean and rms structure functions for sets of realizations of
RM images generated using the FARADAY code, as
in Sect.~\ref{3Dcode}, for different values of $\Lambda_{\rm max}$.
These structure functions are plotted in Figs.\,\ref{sfuncov}(b) and (c).
The main difference between the model structure functions derived from
simulations and the pseudo-structure functions described earlier is that the
former show a steep decline in power on large scales in place of a plateau. This
occurs because the smooth fall-off in density and magnetic field strength with
distance from the nucleus suppresses the fluctuations in RM on large scales.
The results of the simulations confirm our analysis using the pseudo-structure
function. The mean model structure function with $\Lambda_{\rm max} = 65$\,kpc
again fits the data very well, except for a marginal discrepancy at the largest
scales. In view of the deviations from spherical symmetry on large
scales evident in the X-ray emission surrounding 3C\,449 (Fig.~\ref{XMM}), we do
not regard this as a significant effect.
In order to check that the best-fitting density and field model also
reproduces the data at small separations, we repeated the analysis at
1.25-arcsec resolution. The observed pseudo-structure function is shown in
Fig. 14(d), together with the the predictions for the BPL power spectrum
with $\Lambda_{\rm max}$= 205, 65 and 20 kpc. The observed structure function is
compared with the mean from 35 simulations with $\Lambda_{\rm max}$= 65 kpc in
Fig. 14(e). In both cases, the agreement for $\Lambda_{\rm max}$= 65 kpc is
excellent.
In Fig.\,\ref{comp}, example realizations of this model with the best-fitting
field variation are shown for resolutions of 1.25 and 5.5\,arcsec alongside the
observed RM images.
\begin{figure*}
\centering
\includegraphics[width=17cm]{fig14.ps}
\caption[]{(a): Comparison of the observed pseudo-structure function of the
5.5-arcsec RM image as described in the text (points) with the predictions of the BPL model (curves, derived using a Hankel transform). The predicted curves are for
$\Lambda_{\rm max}$=205 (blue dotted), 65 (continuous magenta) and 20\,kpc (green dashed).
(b) and (c): structure functions of the
observed (points) and synthetic 5.5-arcsec RM images produced with the $\Lambda_{\rm max}$=205,20\,kpc
and $\Lambda_{\rm max}$=65\,kpc, respectively.
(d) and (e) as (a) and (c) respectively, but at 1.25-arcsec resolution.
The error bars are the rms from 35 structure functions at a given $\Lambda_{\rm max}$,
in (a) and (d) are representative error bars from the model with $\Lambda_{\rm max}$=65\,kpc.
}
\label{sfuncov}
\end{figure*}
\section{Summary and comparison with other sources}
\label{sec:sum}
\subsection{Summary}
In this work we have studied the structure of the magnetic field associated with
the ionized medium around the radio galaxy 3C\,449. We have analysed images of
linearly polarized emission with resolutions of 1.25\,arcsec and 5.5\,arcsec
FHWM at seven frequencies between 1.365 and 8.385 GHz, and produced images of
degree of polarization and rotation measure. The RM images at both the angular
resolutions show patchy and random structures. In order to study the spatial
statistics of the magnetic field, we used a structure-function analysis and
performed two- and three-dimensional RM simulations. We can summarize the
results as follows.
\begin{enumerate}
\item The absence of deviations from $\lambda^2$ rotation over a wide range of
polarization position angle implies that a pure foreground Faraday screen with no mixing of radio-emitting and
thermal electrons is a good approximation for 3C\,449
(Sect.\,\ref{sec:rm_obs}).
\item The dependence of the degree of polarization on wavelength
is well fitted by a Burn law. This is also consistent with pure foreground
rotation, with the residual depolarization observed at the higher resolution
being due to unresolved RM fluctuations across the beam
(Sect.~\ref{sec:dp}). There is no evidence for detailed correlation of
radio-source structure with either RM or depolarization.
\item There is no obvious anisotropy in the RM distribution, consistent with our
assumption that the magnetic field is an isotropic, Gaussian random variable.
\item Our best estimate for the Galactic contribution to the RM of 3C\,449 is a
constant value of $-$160.7\,rad\,m$^{-2}$ (Sect.\,\ref{sec:gal}).
\item The RM structure functions for six different regions of the source are
consistent with the hypothesis that only the amplitude of the RM power
spectrum varies across the source.
\item A broken power-law spectrum of
the form given in Eq.\,\ref{bpl} with $q_{\rm{l}}$= 2.07,
$q_{\rm{h}}$=2.98, $f_{\rm b}$=0.031\,$\rm{arcsec^{-1}}$ and $f_{\rm max} =
1.67$\,arcsec$^{-1}$ (corresponding to a spatial scale $\Lambda_{\rm
min}$=0.2\,kpc)
is consistent with the observed structure functions and
depolarizations for all six regions. No single power law provides a good fit
to all of the structure functions.
\item The high-frequency cut-off in the power spectrum is required to model the
depolarization data.
\item The profiles of $\sigma_{\rm RM}~$ strongly suggest that most of the fluctuating component of RM
is associated with the intra-group gas, whose core radius is comparable with
the characteristic scale of the profile (Sect.~\ref{sec:rm_images}). The
symmetry of the profile is consistent with the idea that the radio source axis
is close to the plane of the sky.
\item We therefore simulated the RM distributions expected for an isotropic,
random magnetic field in the hot plasma surrounding 3C\,449, assuming the
density model derived by Croston et al.\ (2008).
\item These three-dimensional simulations show that the dependence of magnetic
field on density is best modelled by a broken power-law function with $B(r)
\propto n_e(r)$ close to the nucleus and $B(r) \approx$ constant at larger
distances.
\item With this density model, our best estimate of the central magnetic field
strength is $B_0 = 3.5 \pm 1.2 \mu$G.
\item Assuming these variations of density and field strength with radius, a
structure-function analysis can be used to estimate the outer scale $\Lambda_{\rm max}$ of the
magnetic-field fluctuations. We find excellent agreement for $\Lambda_{\rm max}
\approx 65$\,kpc ($f_{\rm min} =
0.0053$\,arcsec$^{-1}$).
\end{enumerate}
\subsection{Comparison with other sources}
Our results are qualitatively similar to those of Laing et al.\ (2008) on
3C\,31. The maximum RM fluctuation amplitudes are similar in the two sources, as
are their environments. For spherically-symmetric gas density models, the
central magnetic fields are almost the same: $B_0 \approx 2.8\mu$G for 3C\,31
and $3.5\mu$G for 3C\,449. Both results are consistent with the idea that the
RM fluctuation amplitude in galaxy groups and clusters scales roughly linearly
with density, ranging from a few\,rad\,m$^{-2}$ in the much sparsest
environments (e.g.\ NGC\,315; Laing et al. 2006), through intermediate values
$\approx$~30 -- 100\,rad\,m$^{-2}$ in rich groups such as 3C\,31 and 3C\,449 to $\sim
10^4$\,rad\,m$^{-2}$ in the centres of clusters with cool cores.
The RM distribution of 3C\,31 is asymmetrical, the
northern (approaching) side of the source showing a much lower fluctuation
amplitude, consistent with the inclination of $\approx 50^\circ$ estimated by
Laing \& Bridle (2002). Detailed modelling of the RM profile led Laing et al.\
(2008) to suggest that there is a cavity in the X-ray gas, but this would have
to be significantly larger than the observed extent of the radio lobes and is,
as yet, undetected in X-ray observations. A broken power-law scaling of magnetic
field with density, similar to that found for 3C\,449 in the present paper,
would also improve the fit to the $\sigma_{\rm RM}~$ profile for 3C\,31; alternatively, the
effects of cavities around the inner lobes and spurs of 3C\,449 might be
significant. Deeper X-ray observations of both sources are needed to resolve
this issue. In neither case is the magnetic field dynamically important: for
3C\,449 we find that the ratio of the thermal and magnetic-field pressures is
$\approx$30 at the nucleus
and $\approx$400 at the core radius of the group gas, $r_c = 19$\,kpc. The
magnetic field is therefore not dynamically important, as in 3C\,31.
The magnetic-field power spectrum in both sources can be fit by a broken
power-law form. The low-frequency slopes are 2.1 and 2.3 for 3C\,449 and 3C\,31
respectively. In both cases, the power spectrum steepens at higher spatial
frequencies, but for 3C\,31 a Kolmogorov index (11/3) provides a good fit,
whereas we find that the depolarization data for 3C\,449 require a cut-off below
a scale of 0.2\,kpc and a high-frequency slope of 3.0. The break scales are
$\approx$5\,kpc for 3C\,31 and $\approx$11\,kpc for 3C\,449. It is important to
note that the simple parametrized form of the power spectrum is not unique, and
that a smoothly curved function would fit the data equally well.
The gas-density structure on large scales in the 3C\,31 group is uncertain, so
Laing et al.\ (2008) could only give a rough lower limit to the outer scale of
magnetic-field fluctuations, $\Lambda_{\rm max} \ga 70$\,kpc. For 3C\,449, we
find $\Lambda_{\rm max} \approx 65$\,kpc. The projected distance between 3C\,449 and
its nearest neighbour is $\approx$33\,kpc (Birkinshaw et al. 1981), similar to
the scale on which the jets first bend through large angles (Fig.~\ref{XMM}). As
in 3C\,31, it is plausible that the outer scale of magnetic-field fluctuations
is set by interactions with companion galaxies in the group.
\begin{acknowledgements}
This work is part of the ``Cybersar'' Project, which is managed by the
COSMOLAB Regional Consortium with the financial support of the Italian
Ministry of University and Research (MUR), in the context of the ``Piano
Operativo Nazionale Ricerca Scientifica, Sviluppo Tecnologico, Alta
Formazione (PON 2000-2006)''.
We thank Luigina Feretti for providing the
VLA data of 3C449, Greg Taylor for the use of his rotation measure code
and Marco Bondi for many
helpful comments.
We also acknowledge the use of HEALPIX package (http://healpix.jpl.nasa.gov)
and the provision of the models of Dineen \& Coles (2005) in HEALPIX format.
\end{acknowledgements}
|
\section{Quadratic Witt group of graded fields}
\label{sec:gradedWitt}
Let $\Gamma$ be a divisible torsion-free abelian group, which will
contain the degrees of all the graded structures we shall
consider. Graded commutative rings in which every nonzero homogeneous
element is invertible are called \emph{graded fields}, and graded
modules over graded fields are called \emph{graded vector
spaces}. Since $\Gamma$ is torsion-free, graded fields are domains
and graded vector spaces are free modules, see \cite[\S1]{HW}. The
rank of a graded vector space is called its \emph{dimension}. Let
\[
\mathsf{F}=\bigoplus_{\gamma\in\Gamma}\mathsf{F}_\gamma
\]
be a graded field and
\[
\mathsf{V}=\bigoplus_{\gamma\in\Gamma}\mathsf{V}_\gamma
\]
be a graded $\mathsf{F}$-vector space. We let $\Gamma_\mathsf{F}$, $\Gamma_\mathsf{V}$ denote
the sets of degrees of $\mathsf{F}$ and $\mathsf{V}$, i.e.,
\[
\Gamma_\mathsf{F}=\bigl\{\gamma\in\Gamma\mid \mathsf{F}_\gamma\neq\{0\}\bigr\},
\qquad
\Gamma_\mathsf{V}=\bigl\{\gamma\in\Gamma\mid \mathsf{V}_\gamma\neq\{0\}\bigr\}.
\]
The set $\Gamma_\mathsf{F}$ is a subgroup of $\Gamma$, and $\Gamma_\mathsf{V}$ is
a union of cosets of $\Gamma_\mathsf{F}$. For each coset
$\Lambda\in\Gamma/\Gamma_F$, let
\[
\mathsf{V}_\Lambda=\bigoplus_{\lambda\in\Lambda}\mathsf{V}_\lambda
\]
(so $\mathsf{V}_\Lambda=\{0\}$ if $\Lambda\not\subset\Gamma_V$). If
$\Gamma_\mathsf{V}$ is the disjoint union of cosets $\Lambda_1$, \ldots,
$\Lambda_n\in\Gamma/\Gamma_F$, i.e.,
$\Gamma_\mathsf{V}=\Lambda_1\sqcup\ldots\sqcup \Lambda_n$, we have a canonical
decomposition of $\mathsf{V}$ into graded sub-vector spaces
\begin{equation}
\label{eq:candec}
\mathsf{V}=\bigoplus_{i=1}^n\mathsf{V}_{\Lambda_i}.
\end{equation}
Note that the homogeneous component $\mathsf{F}_0$ of $\mathsf{F}$ is a field, and
each $\mathsf{V}_\gamma$ for $\gamma\in\Gamma$ is an $\mathsf{F}_0$-vector
space. Picking an element $\lambda_i\in\Lambda_i$ for each $i=1$,
\ldots, $n$, we have
\[
\dim_\mathsf{F}\mathsf{V}=\sum_{i=1}^n\dim_\mathsf{F}\mathsf{V}_{\Lambda_i} =
\sum_{i=1}^n\dim_{\mathsf{F}_0} \mathsf{V}_{\lambda_i}.
\]
Let $\mathsf{V}$ be a finite-dimensional graded vector space over a graded
field $\mathsf{F}$. A \emph{graded quadratic form} on $\mathsf{V}$ is a map
\[
q\colon \mathsf{V}\to\mathsf{F}
\]
satisfying the following conditions involving $q$ and its polar form
$b\colon \mathsf{V}\times\mathsf{V}\to\mathsf{F}$ defined by
\[
b(x,y)=q(x+y)-q(x)-q(y):
\]
\begin{enumerate}
\item[(i)]
$q(xa)=q(x)a^2$ for all $x\in\mathsf{V}$, $a\in\mathsf{F}$;
\item[(ii)]
$b$ is an $\mathsf{F}$-bilinear form on $\mathsf{V}$;
\item[(iii)]
$q(\mathsf{V}_\gamma)\subseteq \mathsf{F}_{2\gamma}$ for all $\gamma\in\Gamma$;
\item[(iv)]
$b(\mathsf{V}_\gamma,\mathsf{V}_\delta)\subseteq \mathsf{F}_{\gamma+\delta}$ for all
$\gamma$, $\delta\in\Gamma$.
\end{enumerate}
These conditions imply in particular that
\begin{equation}
\label{eq:conseq}
q(\mathsf{V}_\gamma)=\{0\}\text{ if
$\gamma\notin\textstyle\frac12\Gamma_\mathsf{F}$} \quad\text{and}\quad
b(\mathsf{V}_\gamma,\mathsf{V}_\delta)=\{0\}\text{ if $\gamma+\delta\notin\Gamma_\mathsf{F}$}.
\end{equation}
The graded quadratic form $q$ is called \emph{nonsingular} if its
polar form $b$ is nondegenerate, i.e., $x=0$ is the only vector such
that $b(x,y)=0$ for all $y\in V$. It is called \emph{hyperbolic} if it
is nonsingular and $\mathsf{V}=\mathsf{U}\oplus\mathsf{W}$ for some graded subspaces
$\mathsf{U}$, $\mathsf{W}$ such that $q(\mathsf{U})=q(\mathsf{W})=\{0\}$.
\begin{prop}
Let $q$ be a nonsingular graded quadratic form on $\mathsf{V}$. The
canonical decomposition \eqref{eq:candec} yields an orthogonal
decomposition
\[
\mathsf{V}=\Bigl(\bigoplus_{\Lambda\in\frac12\Gamma_\mathsf{F}/\Gamma_\mathsf{F}}^\perp \mathsf{V}_\Lambda\Bigr)
\stackrel{\perp}{\oplus} \mathsf{W}
\qquad\text{where}\quad
\mathsf{W}=\bigoplus_{\Lambda\notin\frac12\Gamma_\mathsf{F}/\Gamma_\mathsf{F}} \mathsf{V}_\Lambda.
\]
For each $\Lambda\in \frac12\Gamma_\mathsf{F}/\Gamma_\mathsf{F}$, the restriction
of $q$ to $\mathsf{V}_\Lambda$ is
nonsingular. Moreover, the restriction of $q$ to $\mathsf{W}$ is hyperbolic.
\end{prop}
\begin{proof}
All assertions except the last one really concern the polar form
$b$; their easy proof (based on the observation \eqref{eq:conseq})
can be found in \cite[Proposition~1.1]{RTW}. It is also proven there
that the restriction of $b$ to $\mathsf{W}$ is
hyperbolic. Since moreover $q(\mathsf{V}_\Lambda)=\{0\}$ for all cosets
$\Lambda\not\subset\frac12\Gamma_\mathsf{F}$ by
\eqref{eq:conseq}, the proposition follows.
\end{proof}
For each coset $\Delta\in\Gamma_\mathsf{F}/2\Gamma_\mathsf{F}$, fix an element
$\delta\in\Delta$ and a nonzero homogeneous element $\pi_\delta\in
\mathsf{F}_{\delta}$ ($\delta=0$ and $\pi_0=1$ for $\Delta=2\Gamma_\mathsf{F}$) and let
\begin{equation}
\label{eq:resa}
q_\Delta\colon\mathsf{V}_{\frac12\delta}\to\mathsf{F}_0,\qquad x\mapsto
\pi_\delta^{-1}q(x).
\end{equation}
Since the restriction of $q$ to $\mathsf{V}_{\frac12\Delta}$ is nonsingular,
it follows that $q_\Delta$ is a nonsingular quadratic form on the
$\mathsf{F}_0$-vector space $\mathsf{V}_{\frac12\delta}$. Of course, this
quadratic form depends on the choice of $\pi_\delta$, except for
$\Delta=2\Gamma_\mathsf{F}$.
\medbreak
\par
Mimicking the usual construction, we may define a Witt equivalence of
nonsingular graded quadratic forms over a given graded field $\mathsf{F}$
and endow the set of Witt-equivalence classes with a group structure
using the orthogonal sum. We let $I_q(\mathsf{F})$ denote the quadratic Witt
group of $\mathsf{F}$, consisting of Witt-equivalence classes of
even-dimensional nonsingular graded quadratic forms over $\mathsf{F}$.
\begin{thm}
\label{thm:gradedWitt}
The map that carries each nonsingular graded quadratic form $q$ to
the collection of nonsingular $\mathsf{F}_0$-quadratic forms
$(q_\Delta)_{\Delta\in
\Gamma_\mathsf{F}/2\Gamma_\mathsf{F}}$ defines a group isomorphism:
\[
I_q(\mathsf{F})\stackrel{\sim}{\to} \bigoplus_{\Gamma_\mathsf{F}/2\Gamma_\mathsf{F}}
I_q(\mathsf{F}_0).
\]
This isomorphism depends on the choice of the homogeneous elements
$\pi_\delta\in \mathsf{F}_{\delta}$.
\end{thm}
The proof is routine. See \cite[Proposition~1.5(iv)]{RTW} for the
(more complicated) case of even hermitian forms over graded division
algebras with involution.
\section{Norms and residues}
\label{sec:norms}
Let $(F,v)$ be an arbitrary valued field and $V$ a finite-dimensional
$F$-vector space. We recall from \cite{RTW} and \cite{TWgr} that a
\emph{$v$-value function} on $V$ is a map
\[
\alpha\colon V\to\Gamma\cup\{\infty\}
\]
satisfying the following properties, for $x$, $y\in V$ and $\lambda\in
F$:
\begin{enumerate}
\item[(i)] $\alpha(x)=\infty$ if and only if $x=0$;
\item[(ii)] $\alpha(x\lambda)=\alpha(x)+v(\lambda)$;
\item[(iii)] $\alpha(x+y)\geq \min\bigl(\alpha(x),\alpha(y)\bigr)$.
\end{enumerate}
The $v$-value function $\alpha$ is called a \emph{$v$-norm} if there
is a base $(e_i)_{i=1}^n$ of $V$ that splits $\alpha$ in the following
sense:
\[
\alpha\Bigl(\sum_{i=1}^ne_i\lambda_i\Bigr) =
\min\bigl(\alpha(e_i\lambda_i) \mid i=1,\ldots,n\bigr) \qquad\text{for
$\lambda_1$, \ldots, $\lambda_n\in F$.}
\]
Any $v$-value function $\alpha$ on $V$ defines a filtration of $V$ by
modules over the valuation ring of $F$: for any $\gamma\in\Gamma$ we
let
\[
V^{\geq\gamma}=\{x\in V\mid \alpha(x)\geq\gamma\},\quad
V^{>\gamma} = \{x\in V\mid \alpha(x)>\gamma\},\quad
V_\gamma=V^{\geq\gamma}/V^{>\gamma},
\]
and we define
\[
\gr_\alpha(V)=\bigoplus_{\gamma\in\Gamma} V_\gamma.
\]
Similarly, let $\gr(F)$ be the graded ring associated to the
filtration of $F$ defined by the valuation. Since every nonzero
homogeneous element of $\gr(F)$ is invertible, this ring is a
graded field. The $F$-vector space structure on $V$ induces on
$\gr_\alpha(V)$ a structure of graded $\gr(F)$-module, so
$\gr_\alpha(V)$ is a graded $\gr(F)$-vector space. For any nonzero
$x\in V$ we let
\[
\tilde x = x+ V^{>\alpha(x)} \in V_{\alpha(x)}\subseteq \gr_\alpha(V).
\]
We also let $\tilde 0 = 0$ and use a similar notation for elements in
$\gr(F)$. It is shown in \cite[Corollary~2.3]{RTW} that a base
$(e_i)_{i=1}^n$ of $V$ splits $\alpha$ if and only if $(\tilde
e_i)_{i=1}^n$ is a $\gr(F)$-base of $\gr_\alpha(V)$, and that $\alpha$
is a norm if and only if
\[
\dim_{\gr(F)}\gr_\alpha(V) = \dim_FV.
\]
Now, let $q\colon V\to F$ be an arbitrary quadratic form, with polar
form $b\colon V\times V\to F$. We say that a $v$-value function
$\alpha$ is \emph{bounded by $q$}, and write $\alpha\prec q$, if the
following two conditions hold:
\begin{enumerate}
\item[(a)]
$v\bigl(b(x,y)\bigr) \geq \alpha(x)+\alpha(y)$ for all $x$, $y\in V$,
and
\item[(b)]
$v\bigl(q(x)\bigr) \geq 2\alpha(x)$ for all $x\in V$.
\end{enumerate}
Of course, letting $x=y$ in (a) yields (b) if $\charac F\neq2$. On the
other hand, (b) does not imply (a): see Example~\ref{ex:notbounded}
below. Note that Springer in \cite{S2} requires only~(b), whereas
Goldman--Iwahori in \cite{GI} require only~(a). Both conditions are
required by Bruhat--Tits in \cite[D\'efinition~2.1]{BT}.
For each $\gamma\in\Gamma$, let $p_\gamma\colon F^{\geq\gamma}\to
F_\gamma$ be the canonical map. When $\alpha\prec q$ we define maps
\[
\tilde q_\alpha\colon V_\gamma\to F_{2\gamma}\quad\text{and}\quad
\tilde b_\alpha\colon V_\gamma\times V_\delta\to F_{\gamma+\delta}
\quad\text{for $\gamma$, $\delta\in\Gamma$}
\]
by
\[
\tilde q_\alpha(\tilde x) = p_{2\gamma}\bigl(q(x)\bigr)
\quad\text{and}\quad \tilde b_\alpha(\tilde x, \tilde y) =
p_{\gamma+\delta}\bigl(b(x,y)\bigr)
\]
for $x$, $y\in V$ with $\alpha(x)=\gamma$ and $\alpha(y)=\delta$.
We extend $\tilde b_\alpha$ to $\gr_\alpha(V)$ by bilinearity and
define
\[
\tilde q_\alpha\colon\gr_\alpha(V)\to\gr(F)
\]
as follows: for $\tilde x_\gamma\in V_\gamma$,
\[
\tilde q_\alpha\Bigl(\sum_{\gamma\in\Gamma} \tilde x_\gamma\Bigr) =
\sum_\gamma \tilde q_\alpha(\tilde x_\gamma) + \sum_{\gamma<\delta}
\tilde b_\alpha(\tilde x_\gamma,\tilde x_\delta).
\]
The map $\tilde q_\alpha$ is a quadratic form on $\gr_\alpha(V)$ with
polar form $\tilde b_\alpha$. It is a graded quadratic
form since $\tilde q_\alpha(V_\gamma)\subseteq F_{2\gamma}$ and
$\tilde b_\alpha(V_\gamma,V_\delta)\subseteq F_{\gamma+\delta}$ for
$\gamma$, $\delta\in \Gamma$. The straightforward verifications are
omitted.
\begin{example}
\label{ex:notbounded}
Suppose $v$ is a discrete valuation on $F$ with
$\Gamma_F=\mathbb{Z}$ and let $V=F^{\oplus2}$ with the hyperbolic
quadratic form $q(x_1,x_2)=x_1x_2$. Define a norm $\alpha\colon
V\to\frac12 \mathbb{Z}\cup\{\infty\}$ by
\[
\alpha(x_1,x_2) = \min(v(x_1), v(x_2)+{\textstyle\frac12})\qquad
\text{for $x_1$, $x_2\in F$.}
\]
Clearly, we have $v\bigl(q(x)\bigr)\geq2\alpha(x)$ for all $x\in V$;
but $b\bigl((1,0),(0,1)\bigr) = 1$, so
\[
v\bigl(b\bigl((1,0),(0,1)\bigr)\bigr) = 0 < \alpha(1,0) +
\alpha(0,1) = \textstyle\frac12.
\]
Thus, condition~(b) holds but not (a). Also, note that for $x=(1,0)$
and $x'=(1,1)$ we have $\tilde x = \tilde x'$ in $\gr_\alpha(V)$,
but $q(x)=0$ and $q(x')=1$, so $\tilde q_\alpha(\tilde x)$ is not
well-defined.
By contrast, the map $\beta\colon V\to\mathbb{Z}\cup\{\infty\}$
defined by
\[
\beta(x_1,x_2)=\min\bigl(v(x_1), v(x_2)\bigr)\qquad \text{for $x_1$,
$x_2\in F$}
\]
is bounded by $q$. The induced quadratic form $\tilde q_\beta$ is
hyperbolic.
\end{example}
\begin{example}
\label{ex:quadext}
Let $(F,v)$ be an arbitrary valued field and let $K/F$ be an
arbitrary quadratic extension. Let $N\colon K\to F$ be the norm
form, which is a quadratic form. We consider two cases:
\noindent\emph{Case 1: The valuation $v$ extends to two different
valuations $v_1$, $v_2$ on $K$.}
Then $K/F$ is not purely inseparable, hence it is a Galois
extension. Let $\sigma\colon K\to K$ be the nontrivial automorphism
of $K/F$, so $v_2=v_1\circ\sigma$. We have $v\bigl(N(x)\bigr) =
v_1(x)+v_2(x)$ for all $x\in V$. The map $\alpha\colon
K\to\Gamma\cup\{\infty\}$ defined by
\[
\alpha(x)=\min\bigl(v_1(x),v_2(x)\bigr) \qquad\text{for $x\in K$}
\]
is a $v$-norm on $K$ by \cite[Corollary~1.7]{TWgr}, and it is
readily checked that
$\alpha\prec N$. Let $u\in K$ be such that $v_1(u)\neq v_2(u)$. Then
$\sigma(u)\neq -u$, hence after scaling by a suitable factor in
$F^\times$ we may assume $u+\sigma(u)=1$. Since
$v_1\bigl(\sigma(u)\bigr) = v_2(u)\neq v_1(u)$, this equation yields
\[
0 = \min\bigl(v_1(u), v_1\bigl(\sigma(u)\bigr)\bigr) = \alpha(u).
\]
It follows that $\tilde b_\alpha(\tilde 1,\tilde u)=\tilde 1$, so
the form $\tilde N_\alpha$ is nonsingular. Moreover, $\tilde
N_\alpha(\tilde u) =0$, so $\tilde N_\alpha$ is hyperbolic.
\noindent\emph{Case 2: The valuation $v$ has a unique extension to
$K$.}
We again write $v$ for the valuation on $K$ extending $v$. This
valuation is a $v$-value function on $K$ and it is readily checked
that $v\prec N$; in fact $v\bigl(N(x)\bigr) = 2v(x)$ for all $x\in
K$. If $K/F$ is immediate (i.e., $\overline{K}=\overline{F}$ and
$\Gamma_K=\Gamma_F$), then $\gr_v(K)=\gr(F)$, so $v$ is not a
norm. Otherwise, $\gr_v(K)$ is a quadratic extension of $\gr(F)$,
and $\tilde N_\alpha$ is the norm form of that extension. It is
nonsingular if and only if $K/F$ is tame, i.e., either $K/F$ is
totally ramified and $\charac\overline{F}\neq2$, or
$\overline{K}/\overline{F}$ is a separable quadratic extension
(i.e., $K/F$ is inertial).
\end{example}
This last example suggests the following definition:
\begin{definition}
Let $(F,v)$ be an arbitrary valued field and $(V,q)$ be a quadratic
space over $F$. A $v$-norm $\alpha$ on $V$ is called a \emph{tame norm
compatible with $q$} if $\alpha\prec q$ and the induced quadratic
form $\tilde q_\alpha$ on $\gr_\alpha(V)$ is nonsingular.
\end{definition}
For any quadratic space $(V,q)$ and any $v$-value function $\alpha$ on
$V$ such that $\alpha\prec q$, it is clear that $\tilde
b_\alpha(\tilde x, \eta)=0$ for all $\eta\in \gr_\alpha(V)$ if
$b(x,y)=0$ for all $y\in V$. Therefore, tame compatible norms exist
only for nonsingular forms. If $\charac\overline{F}\neq2$, the tame
norms compatible with a quadratic form are exactly the norms that are
compatible with its polar form, in the terminology of \cite{RTW}; see
\cite[Remark~3.2]{RTW}. In that case, for every nonsingular quadratic
form there is a tame compatible norm, see
\cite[Corollary~3.6]{RTW}. If $\charac\overline{F}=2$, the tame norms
compatible with a quadratic form $q$ are those that are compatible
with its polar form and that moreover satisfy condition~(b):
$v\bigl(q(x)\bigr)\geq2\alpha(x)$ for all $x\in V$. There are
nonsingular forms for which there is no tame compatible norm, for
instance the norm forms of totally ramified quadratic extensions of
Henselian dyadic fields: see Theorem~\ref{thm:main}.
\begin{lem}
\label{lem:sum}
Let $(V_1,q_1)$ and $(V_2,q_2)$ be quadratic spaces over $F$, and
let $\alpha_1$, $\alpha_2$ be tame $v$-norms on $V_1$, $V_2$ that
are compatible with $q_1$ and $q_2$ respectively. Define
$\alpha_1\oplus\alpha_2\colon V_1\oplus V_2\to \Gamma\cup\{\infty\}$
by
\[
(\alpha_1\oplus\alpha_2)(x_1,x_2) = \min\bigl(\alpha_1(x_1),
\alpha_2(x_2)\bigr)\qquad \text{for $x_1\in V_1$ and $x_2\in V_2$.}
\]
Then $\alpha_1\oplus\alpha_2$ is a tame $v$-norm on $V_1\oplus V_2$
compatible with $q_1\perp q_2$, and there is a canonical
identification of graded quadratic spaces
\[
(\gr_{\alpha_1\oplus\alpha_2}(V_1\oplus V_2), \tilde{(q_1\perp
q_2)}_{\alpha_1\oplus\alpha_2}) =
(\gr_{\alpha_1}(V_1),\tilde{q_1}_{\alpha_1}) \perp
(\gr_{\alpha_2}(V_2), \tilde{q_2}_{\alpha_2}).
\]
\end{lem}
The easy proof is omitted; see \cite[Example~3.7(iii)]{RTW}.
\begin{lem}
\label{lem:hyperb}
Let $(V,q)$ be a quadratic space over $F$ and let $\alpha$ be a tame
$v$-norm on $V$ compatible with $q$. The Witt indices of $q$ and
$\tilde q_\alpha$ are related by
\[
\mathfrak{i}_0(q)\leq \mathfrak{i}_0(\tilde q_\alpha).
\]
In particular, if $q$ is hyperbolic, then
$\tilde q_\alpha$ is hyperbolic.
\end{lem}
\begin{proof}
If $U\subseteq V$ is a totally $q$-isotropic subspace, then
$\gr_\alpha(U)\subseteq \gr_\alpha(V)$ is a
totally $\tilde q_\alpha$-isotropic subspace with $\dim\gr_\alpha(U)
= \dim U$.
\end{proof}
Let $I_q(F)$ denote the quadratic Witt group of $F$, consisting of
Witt-equivalence classes of even-dimensional nonsingular quadratic forms;
see \cite[\S8.B]{EKM}. From Lemma~\ref{lem:sum}, it follows that the
Witt-equivalence classes of even-dimensional nonsingular quadratic forms
that admit a tame compatible norm form a subgroup of $I_q(F)$. We let
$I_{qt}(F)$ denote this subgroup, which we call the \emph{tame
quadratic Witt group of $F$}.
\begin{prop}
\label{prop:residue}
There is a well-defined group epimorphism
\[
\partial\colon I_{qt}(F) \to I_q\bigl(\gr(F)\bigr)
\]
that carries the Witt class of any even-dimensional nonsingular
quadratic form
$q$ with a tame compatible norm $\alpha$ to the
Witt class of $\tilde q_\alpha$.
\end{prop}
\begin{proof}
If $\charac\overline{F}\neq2$, this is shown in
\cite[Theorem~3.11]{RTW} (and holds for odd-dimensional nondegenerate
quadratic forms as well). The arguments also hold if
$\charac\overline{F}=2$: to see that the Witt equivalence class of
$\tilde q_\alpha$ does not depend on the choice of the tame
compatible norm $\alpha$, suppose $\beta$ is another tame compatible
norm. Then $\alpha\oplus\beta$ is a tame norm compatible with the
hyperbolic form $q\perp-q$, hence Lemmas~\ref{lem:sum} and
\ref{lem:hyperb} show that $\tilde q_\alpha\perp-\tilde q_\beta$ is
hyperbolic.
To prove surjectivity of $\partial$ when $\charac\overline{F}=2$, it
suffices to show that the
Witt class of every nonsingular binary form over $\gr(F)$ is in the
image. Let $\mathsf{V}$ be a $2$-dimensional graded vector space over
$\gr(F)$, and let $\varphi\colon\mathsf{V}\to\gr(F)$ be a nonsingular
quadratic form. Since the corresponding polar form $b_\varphi$ is
nonsingular, we may find homogeneous vectors $\varepsilon_1$,
$\varepsilon_2\in\mathsf{V}$ such that
$b_\varphi(\varepsilon_1,\varepsilon_2)=1$. These vectors form a
base of $\mathsf{V}$. If $\varphi(\varepsilon_1)=0$ or
$\varphi(\varepsilon_2)=0$, then $\varphi$ is hyperbolic hence its
Witt class is zero, which lies in the image of $\partial$. If
$\varphi(\varepsilon_1)$, $\varphi(\varepsilon_2)$ are nonzero, then
we may find $a_1$, $a_2\in F^\times$ such that
$\varphi(\varepsilon_1)=\tilde a_1$ and
$\varphi(\varepsilon_2)=\tilde a_2$. Note that the condition
$b_\varphi(\varepsilon_1,\varepsilon_2)=1$ implies that
$\deg\varepsilon_1+\deg\varepsilon_2=0$, hence
$v(a_1)+v(a_2)=0$. Consider a $2$-dimensional $F$-vector space $U$
with base $e_1$, $e_2$ and quadratic form
\[
q(e_1x_1+e_2x_2)=a_1x_1^2+x_1x_2+a_2x_2^2\qquad\text{for $x_1$,
$x_2\in F$.}
\]
Straightforward computations show that $q$ is a nonsingular
quadratic form and that the map $\alpha\colon
U\to\Gamma\cup\{\infty\}$ defined by
\[
\alpha(e_1x_1+e_2x_2)=\min\bigl(\textstyle{\frac12} v(a_1)+v(x_1),
\textstyle{\frac12}v(a_2)+v(x_2)\bigr) \qquad\text{for $x_1$,
$x_2\in F$}
\]
is a tame norm compatible with $q$, such that $(\gr_\alpha(U),\tilde
q_\alpha) \simeq (\mathsf{V},\varphi)$ under the map $e_1x_1+e_2x_2\mapsto
\varepsilon_1\tilde x_1+\varepsilon_2\tilde x_2$. Thus, the Witt
class of $\varphi$ is in the image of $\partial$.
\end{proof}
Our next goal is to show that $\partial$ is an isomorphism when the
valuation $v$ on $F$ is Henselian.
\begin{lem}
\label{lem:Hensel}
Let $(V,q)$ be a quadratic space over $F$. Suppose $v$ is Henselian
and $x$, $y\in V$ are such that
\[
2 v\bigl(b(x,y)\bigr) < v\bigl(q(x)\bigr) + v\bigl(q(y)\bigr).
\]
Then there is a $q$-isotropic vector in the span of $x$ and $y$.
\end{lem}
\begin{proof}
This was observed by several authors, in particular Springer
\cite[Proposition~1]{S} (see also \cite[Lemma~(2.2)]{Tietze}). We
give a proof for completeness. First,
note that if $z_1$, $z_2\in V$ are
multiple of each other, then
\begin{equation}
\label{eq:mult}
2v\bigl(b(z_1,z_2)\bigr) = v\bigl(q(z_1)\bigr) + v\bigl(q(z_2)\bigr) +
2v(2) \geq v\bigl(q(z_1)\bigr) + v\bigl(q(z_2)\bigr).
\end{equation}
In particular, under the hypotheses of the lemma, $x$ and $y$ are
not multiple of each other. Set
$z=yq(x)b(x,y)^{-1}$, so $b(x,z)=q(x)$. For all $\lambda\in F$ we have
\begin{equation}
\label{eq:q}
q(x\lambda +z) = q(x) (\lambda^2+\lambda + q(x)q(y)b(x,y)^{-2}).
\end{equation}
The second factor on the right side is a polynomial $P(\lambda)$
with coefficients in the valuation ring of $F$. Its image in
$\overline{F}[\lambda]$ is $\lambda(\lambda+1)$. By Hensel's Lemma,
$P(\lambda)$ has a root $\lambda_0\in F$. The vector $x\lambda_0+z$
is nonzero since $x$ and $y$ are not multiple of each other, and it
is an isotropic vector of $q$.
\end{proof}
\begin{thm}
\label{thm:Wittindex}
Let $(V,q)$ be a quadratic space over $F$. Assume the valuation $v$
on $F$ is Henselian. Suppose $\alpha\colon
V\to\Gamma\cup\{\infty\}$ is a tame norm compatible with $q$, and
consider the induced quadratic form $\tilde q_\alpha$ on
$\gr_\alpha(V)$. The forms $q$ and $\tilde q_\alpha$ have the same
Witt index:
\[
\mathfrak{i}_0(q) = \mathfrak{i}_0(\tilde q_\alpha).
\]
\end{thm}
\begin{proof}
Lemma~\ref{lem:hyperb} already yields $\mathfrak{i}_0(q) \leq
\mathfrak{i}_0(\tilde q_\alpha)$. To prove the reverse inequality,
we argue by induction on $\dim V$,
and copy the proof of \cite[Proposition~4.3]{RTW}. If $\tilde
q_\alpha$ is isotropic, we may find a homogeneous isotropic vector
$\tilde x$ by taking the homogeneous component of smallest degree of
an arbitrary isotropic vector. Since $\alpha$ is a tame norm
compatible with $q$, the polar form $\tilde b_\alpha$ is
nondegenerate. Therefore, we may find a homogeneous vector $\tilde y$
such that $\tilde b_\alpha(\tilde x,\tilde y)\neq0$. Let $W\subseteq
V$ be the subspace spanned by $x$ and $y$, so
$\gr_\alpha(W)\subseteq \gr_\alpha(V)$ is the graded subspace
spanned by $\tilde x$ and $\tilde y$. We claim that $W$ is a
hyperbolic plane. To see this, observe that the equation $\tilde
q_\alpha(\tilde x)=0$ yields $v\bigl(q(x)\bigr)>2\alpha(x)$, while
$\tilde b_\alpha(\tilde x,\tilde y)\neq0$ shows that
$v\bigl(b(x,y)\bigr) =\alpha(x)+\alpha(y)$. Since
$v\bigl(q(y)\bigr)\geq2\alpha(y)$, it follows that
$2v\bigl(b(x,y)\bigr) < v\bigl(q(x)\bigr) +
v\bigl(q(y)\bigr)$. Therefore, Lemma~\ref{lem:Hensel} shows that $W$
contains a
$q$-isotropic vector. The restriction of $q$ to $W$ is nonsingular
since $b(x,y)\neq0$, hence $W$ is a hyperbolic plane.
Let $q'$ denote the restriction of $q$ to $W^\perp$, so
$\mathfrak{i}_0(q) = 1+\mathfrak{i}_0(q')$. By the choice of
$x$ and $y$, the bilinear form $\tilde b_\alpha$ is nondegenerate on
$\gr_\alpha(W)$. Therefore, the norm $\alpha\rvert_W$ is compatible
with the restriction of the polar form $b\rvert_W$, and
\cite[Proposition~3.8]{RTW} shows that $W^\perp$ is a splitting
complement of $W$ with respect to $\alpha$, i.e.,
$\alpha=\alpha\rvert_W\oplus \alpha\rvert_{W^\perp}$. Thus,
\[
\gr_\alpha(V) = \gr_\alpha(W) \perp \gr_\alpha(W^\perp),
\]
hence $\mathfrak{i}_0(\tilde q_\alpha) = 1+ \mathfrak{i}_0(\tilde
q'_\alpha)$ since $\gr_\alpha(W)$ is a hyperbolic plane. The
induction hypothesis yields $\mathfrak{i}_0(q') \geq
\mathfrak{i}_0(\tilde q'_\alpha)$, hence $\mathfrak{i}_0(q) \geq
\mathfrak{i}_0(\tilde q_\alpha)$ and the proof is complete.
\end{proof}
\begin{cor}
\label{cor:deliso}
The homomorphism $\partial$ of Proposition~\ref{prop:residue} is an
isomorphism if $F$ is Henselian.
\end{cor}
\begin{proof}
Surjectivity of $\partial$ was shown in
Proposition~\ref{prop:residue}, and injectivity follows from
Theorem~\ref{thm:Wittindex}.
\end{proof}
\section{Tame quadratic forms over Henselian fields}
\label{sec:tame}
In this section, we show that the tame quadratic Witt group of a
Henselian field is the Witt kernel of the scalar extension map to the
maximal tame extension.
Let $F$ be a field with a valuation $v\colon
F\to\Gamma\cup\{\infty\}$. Throughout this section, we assume $v$ is
Henselian. Let $(V,q)$ be a quadratic space over $F$ with polar form
$b$. If $q$ is anisotropic, we define a map $\widehat{\alpha}\colon
V\to\Gamma\cup\{\infty\}$ by
\begin{equation}
\label{eq:alpha}
\widehat{\alpha}(x)={\textstyle\frac12}v\bigl(q(x)\bigr)\qquad\text{for $x\in
V$.}
\end{equation}
\begin{prop}
\label{prop:properties}
The map $\widehat{\alpha}$ is a $v$-value function and $\widehat{\alpha}\prec q$.
\end{prop}
\begin{proof}
This was proved by Springer
\cite[Proposition~1]{S} (and attributed to M.~Eichler) in the case
where the valuation is discrete and $F$ is complete. Springer's
arguments hold without change under the more general hypotheses of this
section. We include the proof for the reader's convenience.
By definition of $\widehat{\alpha}$, we clearly have $\widehat{\alpha}(x\lambda) =
\widehat{\alpha}(x)+v(\lambda)$ for $x\in V$ and $\lambda\in F$, and
$\widehat{\alpha}(x)=\infty$ if and only if $x=0$ because $q$ is
anisotropic. It is also clear from the definition that
$v\bigl(q(x)\bigr)\geq2\widehat{\alpha}(x)$ for all $x\in V$. Thus, it
only remains to show
\begin{equation}
\label{eq:iv}
v\bigl(b(x,y)\bigr) \geq \widehat{\alpha}(x)+\widehat{\alpha}(y)\qquad\text{for
$x$, $y\in V$}
\end{equation}
and
\begin{equation}
\label{eq:iii}
\widehat{\alpha}(x+y)\geq\min\bigl(\widehat{\alpha}(x),\widehat{\alpha}(y)\bigr)
\qquad\text{for $x$, $y\in V$.}
\end{equation}
The inequality~\eqref{eq:iv} readily follows from
Lemma~\ref{lem:Hensel} since $q$ is anisotropic.
Property~\eqref{eq:iii} follows, since \eqref{eq:iv} implies
$v\bigl(b(x,y)\bigr)\geq \min\bigl(v\bigl(q(x)\bigr),
v\bigl(q(y)\bigr)\bigr)$ for $x$, $y\in V$, hence
\[
v\bigl(q(x+y)\bigr) = v\bigl(q(x)+b(x,y) + q(y)\bigr) \geq
\min\bigl(v\bigl(q(x)\bigr), v\bigl(q(y)\bigr)\bigr).
\]
\end{proof}
Now, consider the following property for an anisotropic quadratic form
$q$ over $F$:
\begin{equation}
\tag{S}
v\bigl(b(x,y)\bigr) > \widehat{\alpha}(x) + \widehat{\alpha}(y) \qquad\text{for all
nonzero $x$, $y\in V$ such that $\widehat{\alpha}(x)=\widehat{\alpha}(y)$.}
\end{equation}
Note that this property cannot hold if $\charac\overline{F}\neq2$
since for $x=y$ it implies $v(2)>0$ (see
\eqref{eq:mult}). By~\eqref{eq:iv}, we have
$v\bigl(b(x,y)\bigr)>2\min\bigl(\widehat{\alpha}(x),\widehat{\alpha}(y)\bigr)$ if
$\widehat{\alpha}(x)\neq\widehat{\alpha}(y)$, hence Property~(S) has the following
equivalent formulation: for all nonzero $x$, $y\in V$,
\begin{equation}
\tag{S'}
v\bigl(b(x,y)\bigr)>2\min\bigl(\widehat{\alpha}(x),\widehat{\alpha}(y)\bigr) =
\min\bigl(v\bigl(q(x)\bigr), v\bigl(q(y)\bigr)\bigr).
\end{equation}
Our goal is to prove that Property~(S) characterizes the quadratic
forms that remain anisotropic over any inertial extension. Thus,
\emph{until the end of this section, except for the last theorem, we
assume $\charac\overline{F}=2$.}
We first
consider inertial \emph{quadratic} extensions $K/F$. The residue
extension $\overline{K}/\overline{F}$ is obtained by adjoining to
$\overline{F}$ a root of an irreducible polynomial of the form
$X^2+X+\overline{u}$ for some $\overline{u}\in\overline{F}$, hence
$K=F(\mu)$ where $\mu^2+\mu+u=0$ for some $u\in F$ with $v(u)=0$. The
Henselian valuation on $F$ has a unique extension to $K$, for which we
also use the notation $v$. Since $1$, $\overline{\mu}$ are linearly
independent over $\overline{F}$, we have
\begin{equation}
\label{eq:vmin}
v(a+b\mu) = \min\bigl(v(a), v(b)\bigr)\qquad\text{for $a$, $b\in F$.}
\end{equation}
\begin{lem}
\label{lem:quadext}
An anisotropic quadratic form satisfies~(S) if and only if it
remains anisotropic over every inertial quadratic extension.
\end{lem}
\begin{proof}
If (S) does not hold, then we can find nonzero vectors $x$, $y\in V$
such that $\widehat{\alpha}(x)=\widehat{\alpha}(y)$ and $v\bigl(b(x,y)\bigr) =
\widehat{\alpha}(x)+\widehat{\alpha}(y)$. In view of \eqref{eq:mult}, the vectors $x$
and $y$ are not multiple of each other. Letting $z=yq(x)b(x,y)^{-1}$
as in the proof of Lemma~\ref{lem:Hensel}, we see
from~\eqref{eq:q} that $q$ becomes isotropic over $F(\lambda_0)$,
where $\lambda_0$ is a root of the polynomial
$\lambda^2+\lambda+q(x)q(y)b(x,y)^{-2}$. Since
$v(q(x)q(y)b(x,y)^{-2})=0$, the field $F(\lambda_0)$ is an inertial
extension of $F$.
Conversely, suppose $q$ becomes isotropic over an inertial quadratic
extension $K$ of $F$. We have $K=F(\mu)$ where $\mu^2+\mu+u=0$ for
some $u\in F$ with $v(u)=0$. Let $x_0\otimes1+ x_1\otimes\mu\in
V\otimes_FK$ be an isotropic vector of $q$. We have
\begin{equation}
\label{eq:qK}
q(x_0\otimes 1+x_1\otimes\mu) = \bigl(q(x_0)-uq(x_1)\bigr) +
\bigl(b(x_0,x_1)-q(x_1)\bigr)\mu,
\end{equation}
hence
\[
q(x_0)=uq(x_1)\qquad\text{and}\qquad b(x_0,x_1)=q(x_1).
\]
Since $v(u)=0$, it follows that $v\bigl(q(x_0)\bigr) =
v\bigl(q(x_1)\bigr) = v\bigl(b(x_0,x_1)\bigr)$, hence (S) does not hold.
\end{proof}
\begin{lem}
\label{lem:quadextS}
If an anisotropic quadratic form satisfies~(S), then its scalar
extension to any inertial quadratic extension also satisfies~(S).
\end{lem}
\begin{proof}
Suppose $q$ is an anisotropic quadratic form over $F$
satisfying~(S), and $K$ is an inertial quadratic extension of
$F$. By Lemma~\ref{lem:quadext}, we know $q$ remains anisotropic
over $K$. We may therefore extend to $V\otimes_FK$ the definition of
the $v$-value function $\widehat{\alpha}$ of \eqref{eq:alpha}. Let $K=F(\mu)$
where $\mu^2+\mu+u=0$ for some $u\in F$ with $v(u)=0$. For
$x=x_0\otimes1 +x_1\otimes\mu\in V\otimes_FK$ we have by
\eqref{eq:qK} and \eqref{eq:vmin}
\[
\widehat{\alpha}(x)={\textstyle\frac12}\min\bigl(v\bigl(q(x_0)-uq(x_1)\bigr),
v\bigl(b(x_0,x_1)-q(x_1)\bigr)\bigr).
\]
We claim that
\[
\widehat{\alpha}(x) = \min\bigl(\widehat{\alpha}(x_0), \widehat{\alpha}(x_1)\bigr).
\]
We check this formula case-by-case:
if $v\bigl(q(x_0)\bigr) = v\bigl(q(x_1)\bigr)$, then
$v\bigl(q(x_0)-uq(x_1)\bigr)\geq v\bigl(q(x_1)\bigr)$ and, by~(S),
$v\bigl(b(x_0,x_1)-q(x_1)\bigr)=v\bigl(q(x_1)\bigr)$; thus,
$\widehat{\alpha}(x)=\widehat{\alpha}(x_0)=\widehat{\alpha}(x_1)$.
If $v\bigl(q(x_0)\bigr)> v\bigl(q(x_1)\bigr)$, then
$v\bigl(q(x_0)-uq(x_1)\bigr) = v\bigl(q(x_1)\bigr)$ and
$v\bigl(b(x_0,x_1)\bigr)> v\bigl(q(x_1)\bigr)$ by~(S'), hence
$\widehat{\alpha}(x)=\widehat{\alpha}(x_1)$.
Likewise, if $v\bigl(q(x_0)\bigr)< v\bigl(q(x_1)\bigr)$, then
$v\bigl(b(x_0,x_1)\bigr)> v\bigl(q(x_0)\bigr)$, hence
$\widehat{\alpha}(x)=\widehat{\alpha}(x_0)$. Thus, the claim is proved.
Now, suppose $x=x_0\otimes 1 + x_1\otimes\mu$ and $y=y_0\otimes1 +
y_1\otimes \mu$ are nonzero vectors in $V\otimes_FK$ such that
$\widehat{\alpha}(x)=\widehat{\alpha}(y)$, and let $\gamma=\widehat{\alpha}(x)$, hence
\[
\gamma=\min\bigl(\widehat{\alpha}(x_0),\widehat{\alpha}(x_1)\bigr) =
\min\bigl(\widehat{\alpha}(y_0),\widehat{\alpha}(y_1)\bigr).
\]
We have
\[
b(x,y) = \bigl(b(x_0,y_0) -ub(x_1,y_1)\bigr) + \bigl(b(x_0,y_1) +
b(x_1,y_0) - b(x_1,y_1)\bigr) \mu,
\]
hence by \eqref{eq:vmin}
\[
v\bigl(b(x,y)\bigr) = \min\bigl(v\bigl(b(x_0,y_0) -
ub(x_1,y_1)\bigr), v\bigl(b(x_0,y_1) + b(x_1,y_0) - b(x_1,y_1)\bigr)
\bigr).
\]
By Property~(S') we have
\[
v\bigl(b(x_i,y_j)\bigr) > \min\bigl(\widehat{\alpha}(x_i), \widehat{\alpha}(y_j)\bigr)
\geq2\gamma \qquad \text{for $i$, $j=0$, $1$,}
\]
hence
\[
v\bigl(b(x_0,y_0) -
ub(x_1,y_1)\bigr)>2\gamma \quad\text{and}\quad
v\bigl(b(x_0,y_1) + b(x_1,y_0) - b(x_1,y_1)\bigr)>2\gamma.
\]
Therefore, $v\bigl(b(x,y)\bigr)>2\gamma$, and it follows that
Property~(S) holds for the extension $q_K$ of $q$ to $K$.
\end{proof}
We now turn to odd-degree extensions. Let $L/F$ be an inertial
extension of odd degree~$d$. The Henselian valuation $v$ has a unique
extension to $L$, for which we also use the notation $v$, and we have
$L=F(\lambda)$ for some $\lambda$ with $v(\lambda)=0$ such that
$\overline{L}=\overline{F}(\overline{\lambda})$ and the minimal
polynomial of $\overline{\lambda}$ over $\overline{F}$ has
degree~$d$. Every anisotropic quadratic form $q\colon V\to F$ remains
anisotropic over $L$ by a theorem of Springer (see
\cite[Corollary~18.5]{EKM}), hence we
may extend the value function $\widehat{\alpha}$ of \eqref{eq:alpha} to
$V\otimes_FL$.
\begin{lem}
\label{lem:oddext}
For $x=x_0\otimes1+x_1\otimes\lambda+\cdots+ x_{d-1}\otimes
\lambda^{d-1} \in V\otimes_FL$, we have
\[
\widehat{\alpha}(x)=\min\bigl(\widehat{\alpha}(x_0),\ldots, \widehat{\alpha}(x_{d-1})\bigr).
\]
If $q$ satisfies~(S), then its extension $q_L$ to $L$ also
satisfies~(S).
\end{lem}
\begin{proof}
If $x=0$, the formula trivially holds. We may therefore assume
$x_0$, \ldots, $x_{d-1}$ are not all zero and set
\[
\gamma=\min\bigl(\widehat{\alpha}(x_0),\ldots, \widehat{\alpha}(x_{d-1})\bigr).
\]
We have
\[
q_L(x) = \sum_{0\leq i\leq d-1} q(x_i)\lambda^{2i} + \sum_{0\leq i<j
\leq d-1} b(x_i,x_j) \lambda^{i+j}.
\]
Since $v\bigl(q(x_i)\bigr)\geq 2\gamma$ by definition of $\gamma$
and
\[
v\bigl(b(x_i,x_j)\bigr) \geq \widehat{\alpha}(x_i)+\widehat{\alpha}(x_j) \geq 2\gamma
\]
by \eqref{eq:iv}, it follows that
$v\bigl(q_L(x)\bigr)\geq2\gamma$, hence $\widehat{\alpha}(x)\geq\gamma$.
To prove that this inequality is an equality, consider the
homogeneous component $V_\gamma$ of the graded vector space
$\gr_{\widehat{\alpha}}(V)$ and the quadratic map $\tilde q_{\widehat{\alpha}}\colon
V_\gamma\to F_{2\gamma}$ induced by $q$, as in
\S\ref{sec:norms}. This map is anisotropic since $v\bigl(q(z)\bigr)
= 2\widehat{\alpha}(z)$ for all $z\in V$, by definition of
$\widehat{\alpha}$. Since $\overline{L}/\overline{F}$ is an odd-degree
extension, this map remains anisotropic after scalar extension to
$\overline{L}$, by a theorem of Springer (see
\cite[Corollary~18.5]{EKM}). Therefore, letting
$x'_i=x_i+V^{>\gamma}\in V_\gamma$ for $i=0$, \ldots, $d-1$, so that
$x'_i=\tilde x_i\neq 0$ if $\widehat{\alpha}(x_i)=\gamma$ and $x'_i=0$ if
$\widehat{\alpha}(x_i)>\gamma$, we have
\[
(\tilde q_{\widehat{\alpha}})_{\overline{L}}(x'_0\otimes 1 + x'_1\otimes
\overline{\lambda}+ \cdots+x'_{d-1}\otimes\overline{\lambda}^{d-1})
\neq 0.
\]
The left side is the image of $q_L(x)$ under the canonical map
$L^{\geq2\gamma}\to L_{2\gamma}$, hence
$v\bigl(q_L(x)\bigr)=2\gamma$, and $\widehat{\alpha}(x)=\gamma$. The
formula for $\widehat{\alpha}(x)$ is thus established.
Now, let $x$, $y\in V\otimes_FL$ be nonzero vectors with
$\widehat{\alpha}(x)=\widehat{\alpha}(y)$. Let $\gamma=\widehat{\alpha}(x)=\widehat{\alpha}(y)$ and
\[
x=x_0\otimes 1 +\cdots + x_{d-1}\otimes\lambda^{d-1}, \qquad
y=y_0\otimes 1+\cdots+ y_{d-1}\otimes \lambda^{d-1}
\]
with $x_0$, \ldots, $y_{d-1}\in V$ and
\[
\gamma=\min\bigl(\widehat{\alpha}(x_0), \ldots, \widehat{\alpha}(x_{d-1})\bigr) =
\min\bigl(\widehat{\alpha}(y_0), \ldots, \widehat{\alpha}(y_{d-1})\bigr).
\]
We have
\[
b(x,y) = \sum_{i,j=0}^{d-1} b(x_i, y_j)\lambda^{i+j}.
\]
If Property~(S) (hence also (S')) holds for $q$, we have
\[
v\bigl(b(x_i,y_j)\bigr) > 2\min\bigl(\widehat{\alpha}(x_i),\widehat{\alpha}(y_j)\bigr)
\geq 2\gamma \qquad\text{for all $i$, $j$},
\]
hence also $v\bigl(b(x,y)\bigr)>2\gamma$. Thus, Property~(S) holds
for $q_L$.
\end{proof}
\begin{thm}
\label{thm:inertnonisot}
Suppose $F$ is a field with a Henselian dyadic valuation.
An aniso\-tropic quadratic form over $F$ satisfies~(S) if and only if it
remains anisotropic over every inertial extension of $F$.
\end{thm}
\begin{proof}
The ``if'' part follows from Lemma~\ref{lem:quadext}. For the
converse, suppose $q$ is an anisotropic form over $F$
satisfying~(S), and let $M$ be an inertial extension of $F$. We have
to show that $q$ remains anisotropic over $M$. Substituting for $M$
its Galois closure, we may assume $M$ is Galois over $F$. Let
$L\subseteq M$ be the subfield fixed under some $2$-Sylow subgroup
of the Galois group. Then $L/F$ is an odd-degree extension and $M/L$
is a Galois $2$-extension, hence there is a sequence of field
extensions
\[
L=L_0\subseteq L_1\subseteq \cdots \subseteq L_r = M
\]
with $[L_i:L_{i-1}] = 2$ for $i=1$, \ldots, $r$. By a theorem of
Springer (see \cite[Corollary~18.5]{EKM}), $q$ remains anisotropic
over $L$, and its
extension $q_L$ satisfies~(S) by Lemma~\ref{lem:oddext}. Therefore,
$q_{L_1}$ is anisotropic by Lemma~\ref{lem:quadext}, and it
satisfies~(S) by Lemma~\ref{lem:quadextS}. Arguing iteratively, we
see that $q_M$ is anisotropic (and satisfies~(S)).
\end{proof}
Theorem~\ref{thm:inertnonisot} also yields precise information on the
quadratic forms that are split by an inertial extension of the base
field.
\begin{cor}
\label{cor:inertsplit}
Let $q$ be an anisotropic quadratic form over the Henselian dyadic field
$F$. If $q$ becomes hyperbolic over some inertial extension of $F$,
then there exist inertial quadratic extensions $K_1$, \ldots, $K_n$
of $F$ and $a_1$, \ldots, $a_n\in F^\times$ such that
\[
q\simeq \langle a_1\rangle N_1 \perp \ldots \perp \langle a_n\rangle
N_n
\]
where $N_i$ is the (quadratic) norm form of $K_i/F$.
\end{cor}
\begin{proof}
Since $q$ becomes hyperbolic over some inertial extension of $F$,
Theorem~\ref{thm:inertnonisot} shows that it does not satisfy
Property~(S). Therefore, by Lemma~\ref{lem:quadext}, $q$ becomes
isotropic over some inertial quadratic extension $K_1/F$. By
\cite[Proposition~34.8]{EKM}, we
may then find $a_1\in F^\times$ and a quadratic form $q_1$ such that
\[
q\simeq \langle a_1\rangle N_1\perp q_1.
\]
Since $q$ becomes hyperbolic over
some inertial extension of $F$ and $N_1$ becomes hyperbolic over the
inertial extension $K_1$, the form $q_1$ also becomes hyperbolic
over some inertial extension of $F$. The corollary follows by
induction on the dimension.
\end{proof}
If we include the split quadratic $F$-algebra $F\times F$, with
hyperbolic norm form, among the inertial quadratic extensions of $F$,
Witt's decomposition theorem shows that Corollary~\ref{cor:inertsplit}
also holds for isotropic quadratic forms that become
hyperbolic over some inertial extension. It also holds in the
nondyadic case, because then Springer's theorem can be used to show
that every anisotropic quadratic form split by an inertial extension
becomes isotropic over some inertial quadratic extension. Thus, the
same argument as in the proof of Corollary~\ref{cor:inertsplit} applies.
Our last theorem holds without the hypothesis that
$\charac\overline{F}=2$.
\begin{thm}
\label{thm:main}
Let $F$ be a field with a Henselian valuation $v$ (with arbitrary
residue characteristic), and let $(V,q)$
be an even-dimensional nondegenerate quadratic space over $F$. The
following conditions are equivalent:
\begin{enumerate}
\item[(a)]
$q$ becomes hyperbolic over some tame extension of $F$;
\item[(b)]
there exists a tame norm $\alpha$ on $V$ compatible with $q$.
\end{enumerate}
\end{thm}
\begin{proof}
If $\charac\overline{F}\neq2$, both conditions hold for all
even-dimensional nondegenerate
quadratic spaces: (a)~because the quadratic
closure of $F$ is tame and (b)~by \cite[Corollary~3.6]{RTW}. For
the rest of the proof, we may thus assume
$\charac\overline{F}=2$. We may also assume $q$ is anisotropic
because if (b)~holds for an anisotropic form it also holds for every
Witt-equivalent form by Lemma~\ref{lem:sum} since hyperbolic forms
admit tame compatible norms (see Example~\ref{ex:notbounded}).
If (a) holds, then by Corollary~\ref{cor:inertsplit} we may find a
decomposition
\[
q\simeq\langle a_1\rangle N_1\perp\ldots\perp\langle a_n\rangle N_n
\]
for some $a_1$, \ldots, $a_n\in F^\times$ and for $N_1$, \ldots,
$N_n$ the norm forms of some inertial quadratic extensions $K_1$,
\ldots, $K_n$ of $F$. Example~\ref{ex:quadext} shows that for each
$i=1$, \ldots, $n$ the unique valuation on $K_i$ extending $v$ is a
tame norm compatible with $N_i$, hence also with $\langle a_i\rangle
N_i$. The direct sum of these norms is a tame norm compatible with
$q$ by Lemma~\ref{lem:sum}, hence condition~(b) holds.
Conversely, suppose there is a tame norm $\alpha$ on $V$ compatible
with $q$. Then we may find a separable extension $L'$ of
$\overline{F}$ such that $\tilde q_\alpha$ splits after scalar
extension to $\gr(F)\otimes_{\overline{F}}L'$. Let $L$ be an
inertial lift of $L'$, i.e., $L/F$ is an inertial extension such
that $\overline{L}=L'$. Write again $v$ for the unique extension of
$v$ to $L$. Then $\alpha\otimes v$ is a tame norm on $V\otimes_FL$
compatible with $q_L$, and $(\tilde{q_L})_{\alpha\otimes v} =
(\tilde q_\alpha)_{\gr(L)}$ is split since
$\gr(L)=\gr(F)\otimes_{\overline{F}} L'$. Therefore,
Theorem~\ref{thm:Wittindex} shows that $q_L$ is hyperbolic.
\end{proof}
As an example, consider the field $F=\mathbb{Q}_2((t))$ of Laurent series in
one indeterminate over the field of $2$-adic numbers. The composite of
the $2$-adic valuation on $\mathbb{Q}_2$ and the $t$-adic valuation on $F$ is
a Henselian valuation $v$ on $F$ with value group $\mathbb{Z}^2$ and residue
field $\mathbb{F}_2$. We have $I_q(\mathbb{F}_2)\simeq\mathbb{Z}/2\mathbb{Z}$, and the unique nontrivial
Witt class is represented by the norm form of $\mathbb{F}_4$. The unique
inertial quadratic extension of $F$ is $F(\sqrt{5})$, and it follows
from Theorem~\ref{thm:gradedWitt}, Corollary~\ref{cor:deliso},
Corollary~\ref{cor:inertsplit} and Theorem~\ref{thm:main} that
\[
I_{qt}(F)\simeq(\mathbb{Z}/2\mathbb{Z})^4,
\]
with generators the Witt classes of the following forms:
$\langle1,-5\rangle$,
$\langle2\rangle\langle1,-5\rangle$, $\langle
t\rangle\langle1,-5\rangle$, $\langle2t\rangle\langle1,-5\rangle$.
By contrast, the full Witt groups $W(F)$, $I_q(F)$ may also be
determined by using Springer's theorem for the $t$-adic valuation,
since the Witt group of $\mathbb{Q}_2$ is known (see for instance
\cite[Ch.~VI, Remark~2.31]{Lam}); we thus get
\[
W(F)\simeq (\mathbb{Z}/8\mathbb{Z})^2\oplus(\mathbb{Z}/2\mathbb{Z})^4,
\]
with generators the Witt classes of $\langle1\rangle$, $\langle
t\rangle$, $\langle1,-2\rangle$, $\langle t\rangle\langle1,-2\rangle$,
$\langle1,-5\rangle$, and $\langle t\rangle\langle1,-5\rangle$. Note
that $4\langle1\rangle\simeq\langle1,-2,-5,10\rangle$ over $\mathbb{Q}_2$ (see
\cite[\emph{loc. cit.}]{Lam}), hence
$\langle2\rangle\langle1,-5\rangle$ is Witt-equivalent to
$\langle1,-5\rangle - 4\langle1\rangle$. Therefore,
\[
I_q(F)/I_{qt}(F)\simeq(\mathbb{Z}/4\mathbb{Z})\oplus(\mathbb{Z}/2\mathbb{Z})^3
\]
with generators represented by $\langle1,t\rangle$,
$\langle1,1\rangle$, $\langle1,-2\rangle$, and $\langle
t\rangle\langle1,-2\rangle$.
|
\section{Introduction}
In modern theoretical physics one often tries to make statements about ``naturalness" or ``fine-tuning" of the observed values of fundamental parameters, where fine-tuning of a parameter is interpreted as an indication for incompleteness of the theory. Popular examples of fine-tuning problems include the quark mass hierarchy and the cosmological constant problem in particle physics. Since statements about naturalness are fundamentally of a statistical character, to make them mathematically precise one has to assume a well-motivated probability distribution on the parameter space relevant for the theory. A fine-tuning problem then indicates that the probability distribution one has used should be modified by introducing new physical considerations. As an example, as long as observation was consistent with a vanishing cosmological constant, it seemed reasonable to assume that a postulated symmetry would constrain it to vanish. With more recent observations indicating that it must be taken to be very small and positive, there seems to be an issue of fine-tuning\footnote{For an alternative interesting but presumably non-mainstream viewpoint, see \cite{rovellilambda}}. We note that much of the motivation to extend the standard model of particle physics is driven by such considerations, that it is by no means necessary to contemplate a ``Multiverse'' where all possible values of a given parameter are actually realised, and that we need not consider anthropic arguments. Of course we can only ever observe and make measurements in a single Universe, and if a parameter takes a value that appears unlikely maybe this just means that an unlikely possibility is realised in our Universe. One is merely doing statistics. But one should recall Bayes' theorem \cite{bayes}
\begin{equation}
P(A|B)=\frac{P(B|A)P(A)}{\sum_{A_i}P(B|A_i)P(A_i)},
\end{equation}
where one can take $A$ as a hypothesis and $B$ as an observation. Once we specify {\sc a priori} probabilities for a full set of possible hypotheses $A_i$, and compute the probability of observing $B$ given $A_i$, this allows us to make a statement about the probability of the hypothesis following from the observation. If the priors $P(A_i)$ are modified, one might get a very different answer for $P(A|B)$.
The particular example we will investigate in this paper is the apparently weak magnitude of $CP$ violation in the electroweak sector of the standard model, measured by the Jarlskog invariant $J$. The observation by Kobayashi and Maskawa \cite{kobayashimaskawa} that $CP$ violation is only possible for at least three quark families has led to a Nobel prize, but the issue of possible fine-tuning in the magnitude of $CP$ violation is much less understood.
It is true that, if one is talking about parameters in quantum field theory, they should really not be regarded as constants, but have an evolution with energy scale given by renormalization group equations. However, since fine-tuning means a discrepancy of several orders of magnitude, it may well be that a fine-tuning problem is present at all energy scales. This is true for the quark mass hierarchy \cite{massref}, and also for the case of $CP$ violation: Recent numerical studies \cite{cpviolref} indicate that $J^2$ does not run strongly with energy scale, but that the value at extremely high energies $(\sim 10^{15} {\rm GeV})$ is merely about twice the value at low energies. It is then meaningful to talk about ``naturalness" of the value of such a parameter.
Since any statements one tries to make depend very directly on the choice of measure, it is helpful if geometric considerations allow for a natural choice of probability distribution. For instance, if the parameter space is a homogeneous space $G/H$ for $G$ a compact Lie group and $H$ a closed subgroup, the natural requirement on the measure determining the probability distribution is invariance under the left action of $G$, which leads to a unique measure (up to normalization). Probabilities for a given function on $G/H$ to take certain values are then well-defined.
In the case of $CP$ violation in electroweak theory, the parameter space is the space of Cabibbo-Kobayashi-Maskawa (CKM) matrices, a double quotient $H\backslash G/H$, which makes the problem of determining a natural probability distribution more involved than for a homogeneous space $G/H$. This problem was discussed in \cite{paper}, where several possible choices were investigated. It was found that while there is no clearly preferred choice of measure, there always seems to be fine-tuning in the observed value for $J$, unless additional input is used.
In the second part of \cite{paper} (summarised in \cite{prl}), the observed values for the quark masses were taken into account by considering not the space of CKM matrices, but the space of mass matrices, as the fundamental parameter space. This was motivated by the observation that the mass matrices are directly linked to the Yukawa couplings and the Higgs vacuum expectation value, whereas the CKM matrix is only a derived quantity. The observed values for the quark masses were then taken as given, and a probability distribution was constructed that could reproduce these values. The choice made for this distribution was as simple as possible in the following sense: The natural group action on the space of Hermitian mass matrices is the action of the unitary group $U(3)$ by conjugation, $M\rightarrow UMU^{\dagger}$. There is essentially a unique measure invariant under this action. This measure was then modified by introducing the simplest possible function that would allow for a modification of the expectation values for quark masses fitting observation. No further assumptions were needed. It was then found that this simple choice for the measure gave an expectation value for $J$ that was remarkably close to the observed value. Hence the conclusion of \cite{prl} was that once one assumes the quark masses as given, one does not face an additional fine-tuning problem with $J$. This statement, while not new, had been made precise using a geometrically motivated measure on the parameter space.
The calculations done in \cite{paper} heavily relied on the fact that the mass matrices in the standard model can be taken to be Hermitian without loss of generality. Indeed, the starting point was the most natural measure on the space of $3\times 3$ Hermitian matrices. However, in left-right symmetric extensions of the standard model, such as Pati-Salam, such an assumption can no longer be made and the mass matrices have to be regarded as a priori arbitrary complex matrices\footnote{We thank Ben Allanach for pointing this out.}. In this paper, we investigate the consequences for statements about naturalness of $J$ by redoing the analysis of \cite{paper} for general complex matrices. We again use the most symmetric measure on the space of mass matrices, here the space of general complex matrices, and modify it in the simplest possible way to incorporate the observed quark mass hierarchy. While this is a choice that could of course be made very differently, it is a simple choice that uses as few assumptions as possible, and that has worked very well for the case of Hermitian mass matrices. What we find is that the resulting probability distribution is different, and the result is different too: The observed value for $J$ now appears to be unnaturally large, since $CP$ violation should be more heavily suppressed by the quark mass hierarchy. One faces a fine-tuning problem, and needs additional assumptions to modify the measure appropriately.
We should point out that the only real input we use from the physical theory (standard model or a left-right symmetric extension of it) is, apart from the very definition of $CP$-violating parameters, how the theory restricts the type of mass matrices that appear. In particular, for any extension of the standard model (such as left-right symmetric models where parity is the left-right symmetry\footnote{Thanks to the referee for clarification on this point.}) that also has Hermitian mass matrices we would not see any modification in the results.
The structure of the paper is as follows: In section 2, we review $CP$ violation in the electroweak sector and detail why mass matrices may be assumed to be Hermitian in the standard model, but not if one has an extended left-right symmetry. In section 3, we construct a measure on the space of $3\times 3$ complex mass matrices, taking into account both the invariance of a natural measure under left or right multiplication of a complex matrix by a unitary matrix as well as a decaying function that is necessary for the calculation, includes the observed values for the quark masses, and partially breaks this invariance. In section 4, we compute the expectation value for the square of the Jarlskog invariant $J$ in the given probability distribution, showing that it is much smaller that the observed value. This main part is very similar to the calculations in \cite{paper} and the reader may benefit from comparing with this paper. We summarise in section 5.
\section{$CP$ Violation in the Standard Model and Beyond}
We summarise how $CP$ violation arises, first in the standard model and then in the more general case of left-right symmetric extensions, essentially following \cite{jarlskogbook}. In the standard model, the quark fields appear as left-handed $SU(2)$ doublets and right-handed $SU(2)$ singlets:
\begin{equation}
\left(\begin{matrix}q_{jL}\\ q'_{jL}\end{matrix}\right),\quad q_{jR}\,,\quad q'_{jR}\,,\qquad j=1,2,\ldots,N\,.
\end{equation}
Here $N$ is the number of quark families, which is arbitrary in the standard model, and normally taken to be three. The fields are written in a flavour basis which can be considered unphysical, since flavour eigenstates do not correspond to mass eigenstates.
The coupling of the Higgs doublet $H$ to quarks, through spontaneous symmetry breaking by the Higgs potential, gives masses to the quark fields:
\begin{equation}
\mathcal{L}_{{\rm Higgs}}\stackrel{{\rm SSB}}{\longrightarrow}-\sum_{j,k=1}^N \left(m_{jk} \overline{q_{jL}} q_{kR} + m_{jk}' \overline{q'_{jL}} q'_{kR}\right) +{\rm h.c.}
\label{ssb}
\end{equation}
The mass matrices $m$ and $m'$ are determined by the original Yukawa couplings and the Higgs vacuum expectation value. Thus, it seems appropriate to regard either the set of Yukawa couplings together with the Higgs vacuum expectation value or the collection of elements of $m$ and $m'$ as fundamental parameters of the theory. This viewpoint was supported by the results of \cite{paper}.
This part of the Lagrangian is formally $CP$ invariant if and only if $m$ and $m'$, which are so far arbitrary complex matrices, are real. Since this condition is not satisfied in Nature, one has formal $CP$ violation. However, as remarked before, the Lagrangian has been written in the unphysical flavour basis. One can, for general $m$ and $m'$, pass to a different basis, namely the basis of mass eigenstates, by diagonalising the mass matrices with unitary matrices:
\begin{equation}
m=U_L^{\dagger}\Delta U_R\,,\quad m'=(U'_L)^{\dagger}\Delta'U_R'\,,
\label{massdiag}
\end{equation}
thus the basis of mass eigenstates is related to the previously considered basis by
\begin{equation}
q_{L {\rm phys}}=U_L q_L\,, \quad q_{R {\rm phys}}=U_R q_R\,,\quad {\rm etc.}
\end{equation}
It is always possible to choose the unitary matrices so that $\Delta$ and $\Delta'$ are real, and in this new basis this part of the Lagrangian is invariant under $C$ and $P$ separately, and hence also under $CP$. However, the electroweak Lagrangian also contains charged current terms mixing up- and down-type quarks, coupled to the $W$ boson fields via (we are now using the basis of mass eigenstates)
\begin{equation}
X_C:=(W_{\mu}^1-iW_{\mu}^2)J_c^{\mu}+{\rm h.c.}\,,\quad J_c^{\mu}:= \overline{(u,c,t)_L}\gamma^{\mu}V\left(\begin{matrix}d_L\\ s_L \\ b_L \end{matrix}\right)\,,
\label{chargedc}
\end{equation}
where $V:=U_L (U_L')^{\dagger}$ is the {\it Cabibbo-Kobayashi-Maskawa} (CKM) matrix. In the basis of mass eigenstates, this term $X_C$ is not invariant under $CP$ unless $V$ is real. Since we consider the mass eigenstates as physical, $CP$ is violated through these charged current terms.
An important observation made in \cite{jarlskogframpton} is that one can redefine the right-handed quark bases by arbitrary unitary transformations,
\begin{equation}
U_R\rightarrow OU_R\,,\quad U'_R\rightarrow O'U_R'\,,
\label{redefine}
\end{equation}
obtaining a new basis which is to be regarded as equally physical. This is due to the absence of charged current terms involving right-handed quarks, since they are singlets under $SU(2)$. It is therefore no loss of generality to set $U_R=U_L$ and $U'_R=U'_L$ in (\ref{massdiag}), and to assume that $m$ and $m'$ are Hermitian.
A natural way to extend the standard model is to assume the existence of a second $SU(2)$ symmetry which acts on the right-handed quarks, as in the Pati-Salam model \cite{patisalam}. In such extensions, one adds a term (\ref{chargedc}) for right-handed quarks to the Lagrangian. This has the important consequence that a general transformation (\ref{redefine}) for arbitrary unitary transformations $O$ and $O'$ can no longer be regarded as giving an equivalent quark basis, since it modifies this new charged current term. The mass matrices cannot in general be taken to be Hermitian, but are arbitrary complex matrices.
There are now also two possibly $CP$ violating terms, and two CKM matrices. We focus on $V=U_L (U_L')^{\dagger}$ which involves the processes that are actually observed and disregard $V_R=U_R(U_R')^{\dagger}$ in the following.
Mathematically, $V$ is an element of $SU(N)$, but since the phases of the quark fields are arbitrary (even for non-Hermitian mass matrices), $V$ is only defined up to left or right multiplication by a diagonal element of $SU(N)$, {\it i.e. } an element of the maximal torus $U(1)^{N-1}$. The space of CKM matrices is therefore the double quotient $U(1)^{N-1}\backslash SU(N)/U(1)^{N-1}$, characterised by $(N-1)^2$ parameters, out of which $\frac{1}{2}N(N-1)$ may be taken to be real (Euler) angles and the remaining $\frac{1}{2}(N-1)(N-2)$ appear as complex phases. It follows that the matrix $V$ can be taken to be real for $N=2$, and so in the formalism explained here one needs at least three quark families to have a possibility of $CP$ violation. Kobayashi and Maskawa were awarded the 2008 Nobel Prize for using this observation to predict the existence of a third quark family \cite{kobayashimaskawa}. We set $N=3$ in what follows.
The mathematical theory of observable measures of $CP$ violation in the standard model was developed by Jarlskog \cite{jarlskogbook}. She showed \cite{jarlsprl} that all necessary and sufficient conditions for $CP$ violation can be summarised as the following condition on the commutator of the Hermitian matrices $m,m'$:
\begin{equation}
\det C := \det \left(-{\rm i}\left[m,m'\right]\right)\neq 0\,.
\end{equation}
One finds that explicitly
\begin{equation}
\det C=-2J(m_t-m_c)(m_c-m_u)(m_u-m_t)(m_b-m_s)(m_s-m_d)(m_d-m_b)\,,
\end{equation}
where $J:=\frak{Im}(V_{11}V_{22}V_{12}^*V_{21}^*)$ is the Jarlskog invariant which is invariant under left or right multiplication of $V$ by a diagonal matrix, {\it i.e. } an element of $U(1)^2$. The geometrical interpretation of the quantity $J$ is given by the so-called unitarity triangles. These express the requirement on $V$ to be unitary, so that for instance
\begin{equation}
(VV^{\dagger})_{12}=V_{11}V^*_{21}+V_{12}V^*_{22}+V_{13}V^*_{23}=0\,.
\end{equation}
In the complex plane, the three complex numbers that sum to zero form the sides of a triangle. The absolute value $|J|$ is twice the area of this triangle. Since there are different unitarity triangles corresponding to different elements of $VV^{\dagger}$, all with the same area, there are several ways of expressing $J$ in terms of the elements of $V$. A general formula is given by \cite{jarlsprl}
\begin{equation}
J\sum_{\gamma, l}\epsilon_{\alpha\beta\gamma}\epsilon_{jkl}=\frak{Im}(V_{\alpha j}V_{\beta k}V_{\alpha k}^*V_{\beta j}^*)\,.
\end{equation}
The quantities describing the CKM matrix which are invariant under rephasing of the quark fields are $J$ and the absolute values $|V_{\alpha j}|$.
In the general case where $m$ and $m'$ are not assumed to be Hermitian, the corresponding quantity would be
\begin{equation}
\det {\bf C} := \det \left(-{\rm i}\left[m m^{\dagger},m' m'^{\dagger}\right]\right)=-2J(m_t^2-m_c^2)(m_c^2-m_u^2)(m_u^2-m_t^2)(m_b^2-m_s^2)(m_s^2-m_d^2)(m_d^2-m_b^2)\,,
\end{equation}
which of course leads to the same conditions on the mass matrices as before (given that we took all masses to be positive before). In the literature, the use of ${\bf C}$ is perhaps more common than the use of $C$, and one may well argue that this second measure of $CP$ violation should be considered more fundamental as its value does not depend on the arbitrary signs of the mass terms in the Lagrangian.
One common parametrization of the CKM matrix, which we use in the following, is given by
\begin{equation}
V = \left( \begin{matrix} \cos y \cos z & \cos y \sin z & e^{-{\rm i}w} \sin y \\ -\cos x \sin z - e^{{\rm i}w}\sin x \sin y \cos z & \cos x \cos z - e^{{\rm i}w} \sin x \sin y \sin z & \sin x \cos y \\ \sin x \sin z - e^{{\rm i}w} \cos x \sin y \cos z & - \sin x \cos z - e^{{\rm i} w} \cos x \sin y \sin z & \cos x \cos y\end{matrix} \right)\,,
\end{equation}
where the ranges of the Euler angles $x, y, z$ and the complex phase $w$ are
\begin{equation}
0 \le x, y, z \le \frac{\pi}{2}\,, \quad 0 \le w < 2\pi\,.
\end{equation}
An arbitrary $SU(3)$ matrix can then be written as
\begin{equation}
U = T_L\,V\,T_R\,,
\end{equation}
where
\begin{equation}
T_L = {\rm diag}(e^{2{\rm i}p},e^{{\rm i}(q-p)},e^{-{\rm i}(p+q)})\,,\quad T_R = {\rm diag}(e^{{\rm i}(r+t)},e^{{\rm i}(r-t)},e^{-2{\rm i}r})\,,
\end{equation}
and $p,q,r$, and $t$ are phases which can take all values between $0$ and $2\pi$. This shows that the coordinates $x,y,z$, and $w$ indeed parametrise representatives of the double quotient $U(1)^2 \backslash SU(3) / U(1)^2$.
In this parametrization, the Jarlskog invariant $J$ is given by
\begin{equation}
J = \frac{1}{4} \sin 2x \sin 2z \sin y \cos^2 y \sin w\,.
\end{equation}
It appears that the observed value for $J$ is very small, since the maximal value would be $\frac{1}{6\sqrt{3}}\approx 0.1$, whereas in Nature \cite{databook}
\begin{equation}
J=3.05_{-0.20}^{+0.19}\times 10^{-5}\,.
\label{jobserved}
\end{equation}
In a general discussion where the values of the quark masses are not fixed, $J$ is not an appropriate measure of $CP$ violation, since even with nonvanishing $J$ one could have $CP$ conservation if, for example, the up and charm quark masses were coinciding. It was suggested in \cite{jarlskog2} to use an appropriately normalised form of $\det C$, namely
\begin{equation}
a_{CP}= 3\sqrt{6}\frac{\det C}{({{\rm Tr }}\,C^2)^{3/2}}
\end{equation}
for three quark families as the unique basis independent measure of $CP$ violation. This is a dimensionless number which takes values between $-1$ and $+1$, and is again observed to be very close to zero. When written out in terms of the CKM matrix parameters and quark masses, it is a rather complicated expression which is therefore not extremely useful in practical computations. In the present analysis, we assume the quark masses as known and regard $J$ as the measure of $CP$ violation.
\section{Measure on the Space of $3\times 3$ Complex Matrices}
In this section we determine a natural measure on the space of $3\times 3$ complex matrices in order to make statements about likely or natural values for the magnitude of the Jarlskog invariant $J$. Following \cite{paper}, this measure is a product of the geometrically most natural measure with maximal symmetry and a factor that involves the observed values for the quark masses. This second factor is introduced for two reasons; firstly, to make the total volume of the parameter space finite, secondly, to allow for a modification of the distribution on the space of CKM matrices through the quark masses. The analysis is analogous to the case of $3\times 3$ Hermitian mass matrices discussed in \cite{paper}, and we will see shortly where differences arise that eventually lead to different results.
The space of $3\times 3$ complex matrices has a natural left and right action by $U(3)$ corresponding to changes of basis. We first determine a metric invariant under these group actions. We start with Jarlskog's representation of an arbitrary complex matrix as
\begin{equation}
M=U_L^{\dagger}DU_R\,,
\end{equation}
where $U_L,U_R$ are unitary and $D={\rm diag }(D_1,D_2,D_3)$ is real diagonal. Here, as suggested in \cite{jarlsprl}, $M$ is the dimensionless mass matrix $M=m/\Lambda$, where $\Lambda$ is a scale which may be chosen for convenience. A natural choice would be $\Lambda=m_t$ for the up-type or $\Lambda'=m_b$ for the down-type quarks, but since we in principle allow arbitrary values for the quark masses we leave $\Lambda$ arbitrary. Clearly, $U_L$ and $U_R$ are only defined up to simultaneous left multiplication by a diagonal unitary matrix
\begin{equation}
U_L\rightarrow AU_L\,,\quad U_R\rightarrow AU_R\,,\quad A\in U(1)^3
\end{equation}
which reduces the number of (real) parameters from 21 to 18. There are additional discrete ambiguities, corresponding to a permutation or change of sign of the elements of $D$, given by elements of the group $\frak{S}_3\times\mathbb{Z}_2^3$, where $\frak{S}_3$ is the symmetric group of three elements. (We will see shortly that the measure vanishes whenever elements of $D$ coincide up to sign, so we can restrict to matrices with $D_1^2\neq D_2^2\neq D_3^2\neq D_1^2$.) Hence we can identify the relevant subspace of the space of $3\times 3$ complex matrices with $\mathbb{R}^3\times (U(1)^3\times\frak{S}_3\times \mathbb{Z}_2^3)\backslash(U(3)\times U(3))\simeq\mathbb{R}_+^3\times (U(1)^3\times\frak{S}_3)\backslash(U(3)\times U(3))$. However, since all expressions will only involve absolute values of the elements of $D$ and $D'$ (precisely due to the $\mathbb{Z}_2^3$ symmetry), we will integrate over all of $\mathbb{R}^3$ for simplicity. This only leads to an extra factor 8 which drops out in expectation values. The metric we will use to determine a measure is
\begin{equation}
ds^2 = {{\rm Tr }} (dM\cdot dM^{\dagger})\,,
\end{equation}
which clearly is invariant under $M\rightarrow OMO'$ for $O,O'\in U(3)$. To evaluate this, use
\begin{eqnarray}
dM & = & -U_L^{\dagger} dU_L U_L^{\dagger} D U_R + U_L^{\dagger} dD U_R + U_L^{\dagger} D dU_R\,,\nonumber
\\ dM^{\dagger} & = & -U_R^{\dagger} dU_R U_R^{\dagger} D U_L + U_R^{\dagger} dD U_L + U_R^{\dagger} D dU_L
\end{eqnarray}
and $[D,dD]=0$ to obtain
\begin{equation}
{{\rm Tr }} (dM\cdot dM^{\dagger}) = {{\rm Tr }}(dD\cdot dD) - {{\rm Tr }}(D^2(dU_L U_L^{\dagger})^2) - {{\rm Tr }}(D^2(dU_R U_R^{\dagger})^2) + 2 {{\rm Tr }} (D dU_L U_L^{\dagger} D dU_R U_R^{\dagger})\,.
\end{equation}
We introduce right-invariant one-forms
\begin{equation}
dU_L U_L^{\dagger}={{\rm i}} \lambda_a \tau^a_L\,,\quad dU_R U_R^{\dagger}={{\rm i}} \lambda_b \tau^b_R\,,
\end{equation}
where $\lambda_1,\ldots,\lambda_8$ are the Gell-Mann matrices, and
\begin{equation}
\lambda_9=\sqrt{\frac{2}{3}}\left(\begin{matrix} 1 & 0 & 0 \cr 0 & 1 & 0 \cr 0 & 0 & 1 \end{matrix}\right)\,,
\end{equation}
so that ${\rm i}\lambda_a$ are a basis for the Lie algebra $\frak{u}(3)$. Then
\begin{equation}
ds^2 = {{\rm Tr }} (dM\cdot dM^{\dagger}) = {{\rm Tr }}(dD\cdot dD) +{{\rm Tr }}(D^2\lambda_{(a}\lambda_{b)})(\tau^a_L\tau^b_L+\tau^a_R\tau^b_R) - {{\rm Tr }} (D \lambda_a D \lambda_b)(\tau^a_L\tau^b_R+\tau^a_R\tau^b_L)\,.
\end{equation}
The only nonvanishing traces are
\begin{equation}
{{\rm Tr }}(D^2\lambda_1\lambda_1)={{\rm Tr }}(D^2\lambda_2\lambda_2)={{\rm Tr }}(D^2\lambda_3\lambda_3)={{\rm Tr }}(D\lambda_3 D\lambda_3)=D_1^2+D_2^2\,,\nonumber
\end{equation}
\[{{\rm Tr }}(D^2\lambda_{(3}\lambda_{8)})={{\rm Tr }}(D\lambda_{(3}D\lambda_{8)})=\frac{1}{\sqrt{3}}(D_1^2-D_2^2)\,,\quad {{\rm Tr }}(D^2\lambda_{(3}\lambda_{9)})={{\rm Tr }}(D\lambda_{(3}D\lambda_{9)})=\sqrt{\frac{2}{3}}(D_1^2-D_2^2)\,,\]
\[{{\rm Tr }}(D^2\lambda_{(8}\lambda_{9)})={{\rm Tr }}(D\lambda_{(8}D\lambda_{9)})=\frac{\sqrt{2}}{3}(D_1^2+D_2^2-2D_3^2)\,,\]
\[{{\rm Tr }}(D^2\lambda_4\lambda_4)={{\rm Tr }}(D^2\lambda_5\lambda_5)=D_1^2+D_3^2\,,\quad {{\rm Tr }}(D^2\lambda_6\lambda_6)={{\rm Tr }}(D^2\lambda_7\lambda_7)=D_2^2+D_3^2\,,\]
\[{{\rm Tr }}(D^2\lambda_8\lambda_8)={{\rm Tr }}(D\lambda_8 D\lambda_8)=\frac{1}{3}(D_1^2+D_2^2+4D_3^2)\,,\quad {{\rm Tr }}(D^2\lambda_9\lambda_9)={{\rm Tr }}(D\lambda_9 D\lambda_9)=\frac{2}{3}(D_1^2+D_2^2+D_3^2)\,,\]
\[{{\rm Tr }}(D\lambda_1 D\lambda_1)={{\rm Tr }}(D\lambda_2 D\lambda_2)=2D_1 D_2\,,\]
\begin{equation}
{{\rm Tr }}(D^2\lambda_4\lambda_4)={{\rm Tr }}(D^2\lambda_5\lambda_5)=2 D_1 D_3\,,\quad {{\rm Tr }}(D^2\lambda_6\lambda_6)={{\rm Tr }}(D^2\lambda_7\lambda_7)=2 D_2 D_3\,.
\end{equation}
The metric can be written in the form
\begin{eqnarray}
ds^2 & = & dD_1^2+dD_2^2+dD_3^2 + \frac{1}{2}(D_1-D_2)^2(\tau_L^1+\tau_R^1)^2 + \frac{1}{2}(D_1+D_2)^2(\tau_L^1-\tau_R^1)^2\nonumber
\\ & & + \frac{1}{2}\left[(D_1-D_2)^2(\tau_L^2+\tau_R^2)^2 + (D_1+D_2)^2(\tau_L^2-\tau_R^2)^2\right.\nonumber
\\ & & \left.+ (D_1-D_3)^2(\tau_L^4+\tau_R^4)^2 + (D_1+D_3)^2(\tau_L^4-\tau_R^4)^2\right]\nonumber
\\ & & + \frac{1}{2}\left[(D_1-D_3)^2(\tau_L^5+\tau_R^5)^2 + (D_1+D_3)^2(\tau_L^5-\tau_R^5)^2\right.\nonumber
\\ & & \left.+ (D_2-D_3)^2(\tau_L^6+\tau_R^6)^2 + (D_2+D_3)^2(\tau_L^6-\tau_R^6)^2\right]\nonumber
\\ & & + \frac{1}{2}(D_2-D_3)^2(\tau_L^7+\tau_R^7)^2 + \frac{1}{2}(D_2+D_3)^2(\tau_L^7-\tau_R^7)^2\nonumber
\\ & & + 2 D_1^2 \left(\frac{1}{\sqrt{2}}(\tau_L^3 - \tau_R^3) + \frac{1}{\sqrt{6}}(\tau_L^8 - \tau_R^8)+\frac{1}{\sqrt{3}}(\tau_L^9 - \tau_R^9)\right)^2 \nonumber
\\ & & + 2 D_2^2 \left(-\frac{1}{\sqrt{2}}(\tau_L^3 - \tau_R^3) + \frac{1}{\sqrt{6}}(\tau_L^8 - \tau_R^8)+\frac{1}{\sqrt{3}}(\tau_L^9 - \tau_R^9)\right)^2\nonumber
\\ & & + 2 D_3^2 \left(\sqrt{\frac{2}{3}}(\tau_L^8 - \tau_R^8)+\frac{1}{\sqrt{3}}(\tau_L^9 - \tau_R^9)\right)^2.
\end{eqnarray}
The fact that the metric only depends on $\tau_L^3-\tau_R^3$ etc., and not on $\tau_L^3+\tau_R^3$ etc., again reflects the $U(1)^3$ that has to be factored out. As is easy to show, the volume form is proportional to
\[(D_1^2-D_2^2)^2(D_1^2-D_3^2)^2(D_2^2-D_3^2)^2 |D_1 D_2 D_3|\,dD_1\wedge dD_2\wedge dD_3\wedge \tau_L^1\wedge \tau_R^1\wedge\ldots\wedge(\tau_L^8-\tau_R^8)\wedge(\tau_L^9-\tau_R^9)\,.\]
Note that this expression only depends on the absolute values of $D_i$, as expected. Since the range of the $D_i$ is infinite, integration over these coordinates will give an infinity, so that a function decaying sufficiently fast for large $|D_i|$ is introduced. A natural and simple choice is a Gaussian.
Since there are actually two integrations over the space of mass matrices, corresponding to two mass matrices for the up-type and down-type quarks, the general expression for the measure considered in \cite{paper,prl}, where $M$ was assumed to be Hermitian, was
\[DM\,DM'\,\exp(-{{\rm Tr }}(M^2 A))\exp(-{{\rm Tr }}((M')^2 A))\,,\]
where $DM$ was the natural measure on the space of $3\times 3$ Hermitian matrices, with $A$ and $A'$ Hermitian with non-negative eigenvalues and commuting. $A$ and $A'$ could then be diagonalized by the transformation $U\rightarrow UW$, $U'\rightarrow U'W$ without changing any measurable quantities. The Gaussian broke the symmetry of the measure from invariance under $(U(3)\times U(3))$ to invariance under $(U(1)^3\times U(1)^3)$.
The analogous procedure in the case of general complex mass matrices is to use the measure
\[DM\,DM'\,\exp(-{{\rm Tr }}(MM^{\dagger}A+M'(M')^{\dagger}A'))=DM\,DM'\,\exp(-{{\rm Tr }}(U_L^{\dagger}D^2 U_L A+(U'_L)^{\dagger}(D')^2 U'_L A'))\,,\]
so that the measure is still invariant under the right $U(3)$ action on complex matrices, but the invariance under the left $U(3)$ action is again broken to $U(1)^3$. This symmetry breaking is necessary to obtain a distribution that can reproduce the observed quark masses. As before, we assume $A$ and $A'$ to be Hermitian with non-negative eigenvalues and commuting and use the redefinitions $U_L\rightarrow U_LW_L$, $U_L'\rightarrow U_L'W_L$ to diagonalise $A$ and $A'$.
Since the important measurable quantity involving the unitary matrices is the CKM matrix $V=U_L U_L'^{\dagger}$, everything is independent of the $U_R$ parameters and we integrate over these, obtaining a constant which is irrelevant in the averaging process.
We are left with integrating over the space of possible $U_L$, the coset $U(1)^3 \backslash U(3) = U(1)^2 \backslash SU(3)$ (both the Gaussian and the CKM matrix are invariant under left multiplication of $U_L$ by an element of $U(1)^3$), and the volume form is proportional to
\[(D_1^2-D_2^2)^2(D_1^2-D_3^2)^2(D_2^2-D_3^2)^2 |D_1 D_2 D_3|\,dD_1\wedge dD_2\wedge dD_3\wedge \tau_L^1\wedge \tau_L^2\wedge \tau_L^4 \wedge \tau_L^5\wedge\tau_L^6\wedge\tau_L^7\,.\]
In terms of the coordinates on $U(1)^2\backslash SU(3)$ introduced in section 2, the wedge product of right-invariant forms gives the usual bi-invariant measure on $SU(3)$, so that we finally get
\begin{equation}
DM = \left(\prod_{i<j}(D_i^2-D_j^2)^2\right) |D_1 D_2 D_3|\,\sin 2x \cos^3 y \sin y \sin 2z\,dD_1\, dD_2\, dD_3\,dx\,dy\,dz\,dw\,dr\,dt.
\label{su3meas}
\end{equation}
We also have to take the discrete $\frak{S}_3$ symmetry into account, integrating only over one sixth of the homogeneous space $U(1)^2\backslash SU(3)$, corresponding to
\begin{equation}
0\le y\le \arctan(\sin x)\,,\quad 0\le x\le\frac{\pi}{4}\,.
\end{equation}
This restriction amounts to removing unitary matrices that permute the elements of $D$ and hence to fixing an ordering.
Comparing to the case of Hermitian matrices \cite{prl}, the measure involves higher powers of elements of $D$. Therefore, we would expect a stronger influence of the quark mass hierarchy on expectation values of (powers of) $J$, namely a stronger suppression of large values of $J$. This naive expectation will be confirmed in the next section.
\section{Results for $J$}
The calculations go through exactly as in the case of the measure used in \cite{paper}, and our analysis will be completely analogous. We will obtain analytical approximations of the relevant integrations whose validity is then verified by numerical integration (using {\sc Mathematica}).
We want to compute the expectation value for $J^2$ in the given probability distribution (all odd powers of $J$ average to zero),
\begin{equation}
\langle J^2\rangle = \frac{\int DM\,DM'\,e^{-{{\rm Tr }}(MM^{\dagger}A)-{{\rm Tr }}(M'(M')^{\dagger}A')} J^2(M,M') }{\int DM\,DM'\,e^{-{{\rm Tr }}(MM^{\dagger}A)-{{\rm Tr }}(M'(M')^{\dagger}A')}}\,,
\label{jint}
\end{equation}
where $DM$ is given in (\ref{su3meas}) and $DM'$ is the same expression in terms of primed variables. It should be clear that multiplying $A$ or $A'$ by a (non-negative) constant is the same as rescaling the arbitrary scales $\Lambda$ and $\Lambda'$ used in defining $D_i$ or $D_i'$. Hence we may assume, without loss of generality, the form
\begin{equation}
A = \left(\begin{matrix} 1 & 0 & 0 \\ 0 & \mu_c^{-2} & 0 \\ 0 & 0 & \mu_u^{-2}\end{matrix}\right)\,,\quad A' = \left(\begin{matrix} 1 & 0 & 0 \\ 0 & \mu_s^{-2} & 0 \\ 0 & 0 & \mu_d^{-2}\end{matrix}\right)\,.
\label{aform}
\end{equation}
We have introduced four dimensionless parameters $\mu_c,\mu_u,\mu_s$, and $\mu_d$ that we are free to choose so as to reproduce the observed values of the quark masses as expectation values in our distribution.\footnote{We emphasise that we do not make any predictions about quark masses, but use them as an input to modify the probability distribution on the space of all mass matrices.} On dimensional grounds alone, we expect that we will have to set $\mu_c\approx m_c/m_t$ etc. In particular, we will have $\mu_u\ll 1$ and $\mu_d\ll 1$. The precise relation between the $\mu$ parameters and expectation values for (squared) quark masses will be determined shortly.
Now $J^2$, written in terms of coordinates on $M$ and $M'$, is a very complicated expression. However, since
\begin{equation}
{{\rm Tr }}(MM^{\dagger}A) = \sum_{ij}D_i^2 A_{j} |U_L|_{ij}^2 = \frac{D_1^2}{\mu_u^2}\sin^2 y + \ldots \,,\quad {{\rm Tr }}(M'(M')^{\dagger}A') = \frac{(D_1')^2}{\mu_d^2}\sin^2 y' + \ldots \,,
\end{equation}
where we assume $\mu_u\ll 1$ and $\mu_d\ll 1$ and $A_j$ denote the diagonal elements of $A$, the integrand in the numerator and denominator of (\ref{jint}) is negligibly small unless $y$ and $y'$ are close to zero. We therefore use the approximation $\sin y\approx y$ in the measure $DM$, setting $y=y'=0$ in the remaining part of the integrand. The integration over $y$ and $y'$ can be easily performed, and everything is independent of $w$ and $w'$ which may hence be dropped as integration variables. One is left with the integral
\begin{equation}
\langle J^2 \rangle \approx \frac{\int dD \,dD' \int d^4 x\, d^4 x'\sin 2x \sin 2z \sin 2x' \sin 2z' \left(e^{-{{\rm Tr }}(D^2 UAU^{\dagger}+(D')^2 U'A'{U'}^{\dagger})}\,J^2(U,U')\right)\big|_{y=y'=0}}{\int dD \, dD' \int d^4 x\, d^4 x'\sin 2x \sin 2z \sin 2x' \sin 2z' \left(e^{-{{\rm Tr }}(D^2 UAU^{\dagger}+(D')^2 U'A'{U'}^{\dagger})}\right)\big|_{y=y'=0}}\,,
\label{newjint}
\end{equation}
where
\begin{equation}
\int d^4 x\equiv \int\limits_0^{\pi/4}dx\int\limits_0^{\pi/2}dz \int\limits_0^{2\pi}dr\int\limits_0^{2\pi}dt
\end{equation}
and similarly for $\int d^4 x'$. The shorthand $dD$ denotes the measure over the $D_i$, namely
\begin{equation}
dD = \left(\prod_{i<j}(D_i^2-D_j^2)^2\right) |D_1 D_2 D_3|\,dD_1\, dD_2\, dD_3\,.
\end{equation}
It is possible to analytically integrate over both copies of ${{\mathbb R}}^3$ in (\ref{newjint}), using
\begin{eqnarray}
f_{\xi_1\xi_2\xi_3} & := &\int\limits_{-\infty}^{\infty}dD_1\int\limits_{-\infty}^{\infty}dD_2\int\limits_{-\infty}^{\infty}dD_3\,(D_1^2-D_2^2)^2(D_1^2-D_3^2)^2(D_2^2-D_3^2)^2 |D_1 D_2 D_3| e^{-\xi_1 D_1^2-\xi_2D_2^2-\xi_3D_3^2}\nonumber
\\& = &\frac{24}{\xi_1^{5}\xi_2^{5}\xi_3^{5}}\left(2(\xi_1^2\xi_2^2+\xi_1^2\xi_3^2+\xi_2^2\xi_3^2)(\xi_1^2+\xi_2^2+\xi_3^2-\xi_1\xi_2-\xi_1\xi_3-\xi_2\xi_3)\right.\nonumber
\\& & \left.-3\xi_1\xi_2\xi_3(\xi_1^3+\xi_2^3+\xi_3^3-\xi_2^2\xi_3-\xi_3^2\xi_1-\xi_1^2\xi_2-\xi_3^2\xi_2-\xi_1^2\xi_3-\xi_2^2\xi_1)-8\xi_1^2\xi_2^2\xi_3^2\right)\,.
\end{eqnarray}
The explicit expression for $J$ at $y=y'=0$ is
\begin{eqnarray}
J(U,U')\big|_{y=y'=0} & = &\frac{1}{4}s_{2x} s_{2x'} \left\{c^2_{z'} s^3_{z} s_{z'}\sin(3 \hat{r} + \hat{t}) + c^3_{z} c_{z'} s^2_{z'}\sin(3 \hat{r} - \hat{t}) \right.\nonumber
\\& & - c^2_{z} s_{z} s_{z'} (c^2_{z'} \left[\sin(3 \hat{r} + \hat{t}) + \sin(3\hat{r} -3\hat{t})\right] - s^2_{z'}\sin(3\hat{r} + \hat{t}) )\nonumber
\\& & \left. + c_{z} c_{z'} s^2_{z} (c^2_{z'} \sin(3\hat{r} - \hat{t}) - s^2_{z'}\left[\sin(3 \hat{r} + 3\hat{t}) + \sin(3 \hat{r} - \hat{t})\right])\right\}
\label{jexp}
\end{eqnarray}
where $s_x=\sin x, c_{z'}=\cos z'$, etc., $\hat r= r-r'$, and $\hat t=t-t'$. Integrating (\ref{jexp}) over $r,r',t$, and $t'$ indeed gives zero, which is why we choose to use $J^2$.
To fix the parameters appearing in the matrices $A$ and $A'$, we first observe that expectation values for squared mass matrices, again in the approximation $y\ll 1$, take the relatively simple form
\begin{equation}
\langle D_1^2 \rangle \approx \frac{\int_{{{\mathbb R}}^3} dD\, D_1^2 \int\limits_0^{\pi/4} dx \int\limits_0^{\pi/2} dz\,\sin 2x \sin 2z \left(e^{-{{\rm Tr }}(D^2 UAU^{\dagger})}\right)\big|_{y=0}}{\int_{{{\mathbb R}}^3} dD\, \int\limits_0^{\pi/4} dx \int\limits_0^{\pi/2} dz\,\sin 2x \sin 2z \left(e^{-{{\rm Tr }}(D^2 UAU^{\dagger})}\right)\big|_{y=0}}\,.
\end{equation}
The denominator is explicitly
\begin{equation}
I_D:=\int\limits_0^{\pi/4} dx \int\limits_0^{\pi/2} dz\,\sin 2x \sin 2z \,f_{\xi_1\xi_2\xi_3}\,,
\label{denom}
\end{equation}
where
\begin{eqnarray}
& \xi_1=A_1 \cos^2 z + A_2 \sin^2 z\,,\quad \xi_2=A_1 \cos^2 x \sin^2 z + A_2 \cos^2 x \cos^2 z + A_3 \sin^2 x\,,\nonumber
\\& \xi_3=A_1 \sin^2 x \sin^2 z + A_2 \sin^2 x \cos^2 z + A_3 \cos^2 x\,,
\end{eqnarray}
with $A_3 \gg A_2 \gg A_1$. The integral is dominated by very small $x$ and $z$ (we cannot have $x=\frac{\pi}{2}$), and we can approximate $I_D$ well by only keeping the terms of leading order in $x$ and $z$ in the trigonometric functions, and
\begin{equation}
f_{\xi_1\xi_2\xi_3} \approx \frac{24}{\xi_1^5\xi_2^5\xi_3^5}\times 2\xi_3^4\xi_2^2\approx \frac{48}{(A_1+A_2 z^2)^5 (A_2 + A_3 x^2)^3 A_3}\,,
\end{equation}
which are the leading terms (as we shall see, the combination $A_3 x^2$ is effectively of order $A_2$ etc.):
\begin{eqnarray}
I_D & \approx & \frac{48}{A_3} \int\limits_0^{\pi/4} dx \int\limits_0^{\pi/2} dz\,4 x z \,(A_1+A_2 z^2)^{-5} (A_2 + A_3 x^2)^{-3}\nonumber
\\ & \approx &\frac{48}{A_3} \int\limits_0^{\infty} dX \int\limits_0^{\infty} dZ\,(A_1+A_2 Z)^{-5} (A_2 + A_3 X)^{-3}\nonumber
\\ & = &\frac{48}{A_3}\cdot\frac{1}{4A_2 A_1^4}\cdot\frac{1}{2A_3 A_2^2} = \frac{6}{A_1^4 A_2^3 A_3^2}\,.
\end{eqnarray}
This is very well reproduced by numerical calculations. Similarly, we find
\begin{eqnarray}
I_D\langle D_1^2 \rangle & \approx & \frac{240}{A_3} \int\limits_0^{\infty} dX \int\limits_0^{\infty} dZ\,(A_1+A_2 Z)^{-6} (A_2 + A_3 X)^{-3}\nonumber
\\ & = & \frac{240}{A_3}\cdot\frac{1}{5A_2 A_1^5}\cdot\frac{1}{2A_3 A_2^2} = \frac{24}{A_1^5 A_2^3 A_3^2}\,,
\end{eqnarray}
hence
\begin{equation}
\langle D_1^2 \rangle \approx \frac{4}{A_1}\,.
\end{equation}
Redoing the same calculation for $D_2$ and $D_3$ gives
\begin{equation}
\langle D_2^2 \rangle \approx \frac{2}{A_2}\,,\quad \langle D_3^2 \rangle \approx \frac{1}{A_3}\,.
\end{equation}
Using the form (\ref{aform}) for $A$, this is
\begin{equation}
\langle D_1^2\rangle\approx 4\,,\quad \langle D_2^2 \rangle \approx 2\mu_c^2\,,\quad \langle D_3^2 \rangle \approx \mu_u^2\,.
\end{equation}
If $\langle D_1^2\rangle$ is to reproduce the squared top quark mass in units of $\Lambda$, our reference scale for the up-type quarks must be $\Lambda=\frac{1}{2}m_t$. Then setting $m_c^2/\Lambda^2=2\mu_c^2$ determines $\mu_c$, etc. Apart from numerical prefactors of order one, the $\mu$ parameters indeed correspond to the quark masses one wants to reproduce in the probability distribution. Note that factoring out the $\frak{S}_3$ above corresponds to fixing an ordering of the quark masses, so that it is possible to obtain unequal expectation values for $D_1,D_2$ and $D_3$. Integrating over the whole of $U(1)^2\backslash SU(3)$ would mean to also include permutations, so that necessarily $\langle D_1^2\rangle=\langle D_2^2\rangle=\langle D_3^2\rangle$.
Again, because we have to consider the dependence of masses on the energy scale in quantum field theory, described by the renormalization group, there is some ambiguity in what is meant by the ``quark masses" we want to reproduce. Following \cite{rosner}, for example, we take all the quark masses evolved to the scale of the $Z$ boson mass. These are given in \cite{massref}:
\begin{eqnarray}
&&(m_u,m_c,m_t)=(1.27_{-0.42}^{+0.50}\;{\rm MeV},\; 0.619 \pm 0.084\;{\rm GeV},\;171.7 \pm 3.0\;{\rm GeV})\,; \nonumber
\\&& (m_d,m_s,m_b)=(2.90_{-1.19}^{+1.24}\;{\rm MeV},\; 55_{-15}^{+16}\;{\rm MeV},\;2.89 \pm 0.09\;{\rm GeV})\,.
\label{qmass}
\end{eqnarray}
We use the central values
\begin{eqnarray}
&&(m_u,m_c,m_t):=(1.27\;{\rm MeV},\; 0.619\;{\rm GeV},\;171.7\;{\rm GeV})\,; \nonumber
\\&& (m_d,m_s,m_b):=(2.9\;{\rm MeV},\; 55\;{\rm MeV},\;2.89\;{\rm GeV})\,.
\label{cvalues}
\end{eqnarray}
The mass scales $\Lambda$ and $\Lambda'$ are now fixed by setting $\langle D_1^2 \rangle=(m_t/\Lambda)^2$ and $\langle (D'_1)^2 \rangle=(m_b/\Lambda')^2$. By comparing the results obtained by numerical integration with the values we want to reproduce, we can then fix the parameters $\mu_c,\mu_u,\mu_s$ and $\mu_d$.
In the case of the positively charged top, charm and up quarks, which exhibit a more extreme quark mass hierarchy, we find that numerical calculations (using {\sc Mathematica}) reproduce the results we have obtained analytically very well. For the negatively charged quarks, we find numerically that we have to use different relative factors to reproduce the observed masses. Comparing the numerical results with (\ref{cvalues}), we fix the parameters appearing in $A$ and $A'$ to
\begin{eqnarray}
\mu_c^2 = 2\left(\frac{m_c}{m_t}\right)^2\approx 2.60\times 10^{-5}\,,\quad \mu_u^2 = 4\left(\frac{m_u}{m_t}\right)^2\approx 2.19\times 10^{-10}\,,\nonumber
\\ \mu_s^2 = \left(\frac{m_s}{m_b}\right)^2\approx 3.62\times 10^{-4}\,,\quad \mu_d^2 = 4\left(\frac{m_d}{m_b}\right)^2\approx 4.03\times 10^{-6}\,.
\label{muvalues}
\end{eqnarray}
In order to obtain an analytical expression for expectation value of $J^2$, the next approximation is that the main contribution to the integral (\ref{newjint}) will come from small $z$. This again is seen by writing out ${{\rm Tr }}(MM^{\dagger} A)$ and using the mass hierarchy. We only take the term in (\ref{jexp}) that is non-zero at $z=0$ into account, setting $\sin 2z\approx 2z$ in the measure.
Averaging over $r,t,r'$,and $t'$ gives a factor of 1/2, as one might have expected, and therefore we use
\begin{equation}
J^2_{{\rm small}\;z}:=\frac{1}{2}\sin^2 x\cos^2 x\sin^2 x'\cos^2 x'\cos^2 z'\sin^4 z'
\end{equation}
for our calculations. Within this approximation for $J$, still taking $f_{\xi_1\xi_2\xi_3} \approx 48\xi_1^{-5}\xi_2^{-3}\xi_3^{-1}$, the numerator of (\ref{newjint}) is the product (using again that only small $z$ contributes)
\begin{eqnarray}
I_N & \approx &1152\times \int\limits_0^{\pi/2} dz \, \frac{2z}{(A_1 + A_2 z^2)^{5/2}} \times \int\limits_0^{\pi/2} dz' \, \frac{\sin 2z' \cos^2 z'\sin^4 z'}{(A_1 \cos^2 z'+ A_2 \sin^2 z')^5} \nonumber
\\ & & \times \int\limits_0^{\pi/4} dx \,\frac{\sin 2x\,\sin^2 x\,\cos^2 x}{(A_2 \cos^2 x + A_3 \sin^2 x)^3 (A_3 \cos^2 x + A_2 \sin^2 x)}\nonumber
\\ & & \times \int\limits_0^{\pi/4} dx' \,\frac{\sin 2x'\,\sin^2 x'\,\cos^2 x'}{(A'_2 \cos^2 x' + A'_3 \sin^2 x')^3 (A'_3 \cos^2 x' + A'_2 \sin^2 x')}
\end{eqnarray}
The first two factors are $1/(4A_1^4 A_2)$ and $1/(12(A_1')^2(A_2')^3)$, respectively; for the other two (which have the same form) we change variables to $X=\cos^2 x$ to obtain
\begin{equation}
\int\limits_{1/2}^{1} dX \,\frac{X(1-X)}{(A_2 X + A_3 (1-X))^3 (A_3 X + A_2 (1-X))}= \frac{1}{2A_3^3 A_2}\left(1-\frac{2A_2}{A_3}+O\left(\left(A_2/A_3\right)^2\right)\right)\,;
\end{equation}
the expressions for the denominator are similar but simpler. Putting everything together, we obtain the approximation to lowest order in quark mass ratios
\begin{equation}
\langle J^2_{{\rm small}\;z}\rangle \approx \frac{1}{6}\frac{A_2 (A_1')^2}{A_3 A_2' A_3'} = \frac{4}{3}\frac{m_s^2 m_u^2 m_d^2}{m_b^4 m_c^2}\,,
\label{japprox}
\end{equation}
where the numerical factors appearing in the last line come from the different factors chosen in (\ref{muvalues}). Note that the top quark mass does not appear in this approximate result. This compares with the scaling behaviour obtained in \cite{paper} (for Hermitian mass matrices),
\begin{equation}
\langle J^2_{{\rm small}\;z}\rangle \sim \frac{m_s^2 m_u m_d}{m_b^3 m_c}.
\end{equation}
For numerical calculations we use both the simplified expression $J^2_{{\rm small}\;z}$ and the expression for $J$ given in (\ref{jexp}). We find that for the first quantity, the numerically evaluated expectation value, $\langle J^2_{{\rm small}\;z}\rangle\approx 1.89\times 10^{-15}$, is about $94\%$ of (\ref{japprox}), and the numerical result for $\langle J^2 \rangle$ (taken at $y=y'=0$) is
\begin{equation}
\langle J^2 \rangle \approx 2.07\times 10^{-15}\,,
\end{equation}
which gives
\begin{equation}
\Delta J = \sqrt{\langle J^2 \rangle} \approx 4.55 \times 10^{-8}
\end{equation}
which is now almost three orders of magnitude {\it smaller} than the observed value (\ref{jobserved}). Assuming a Gaussian distribution peaked at zero, we get
\begin{equation}
P(|J|\le 10^{-7})\approx 97\%\,,
\end{equation}
When the measure presented here is used, there seems to be extreme fine-tuning in $J$, but now we would say that one observes unnaturally {\it large} CP violation! This result may look surprising, given that the maximal value for $J$ is around 0.1 and the observed value just $3\times 10^{-5}$, but it shows how strongly the quark mass hierarchy suppresses large values of $J$ in our distribution.
\section{Summary and Discussion}
We have tried to estimate the naturalness of the observed magnitude of $CP$ violation in the electroweak theory under the assumption that there is a left-right symmetry which implies that the quark mass matrices can not in general be taken to be Hermitian. We have constructed a probability distribution on the space of $3\times 3$ complex matrices which takes into account the geometrical structure of this space, but also includes a Gaussian factor which makes the total volume finite and leads to expectation values for quark masses that can reproduce the observed values if four free parameters are fitted accordingly. While this is a choice we make, and all results depend on this choice, our measure is the combination of a maximally symmetric measure, invariant under a redefinition of a complex matrix by left or right multiplication by unitary matrices, and a Gaussian incorporating the observed values of the quark masses. We would have to make additional rather strong assumptions to motivate a different choice of measure that would differ appreciably from this simple construction. It may well be that such assumptions are justified by the underlying mechanism determining the mass matrices, but we do not know of such a mechanism yet. Furthermore, such a simple choice for the measure was shown in \cite{paper}, where only Hermitian mass matrices were considered, to lead to expectation values for the Jarlskog invariant $J$ that make the observed value appear typical.
The conclusion for general complex mass matrices, as shown here, is rather different. Using the given probability distribution, one would now expect $J$ to be about three orders of magnitude smaller than the observed value. Hence, there is a fine-tuning problem: Without further assumptions, a fundamental theory leading to a left-right symmetric electroweak sector at low energy should generically be expected to reproduce very weak $CP$ violation. Invoking the principle of Occam's razor, ``{\sc entia non sunt multiplicanda praeter necessitatem}," we would like to conclude that, only looking at possible explanations for the magnitude of $CP$ violation in the electroweak sector, the standard model should be preferred to left-right symmetric extensions such as Pati-Salam: In the latter one needs additional assumptions on the fundamental parameters that resolve the issue of observing ``unnaturally large" $CP$ violation, that are not necessary in the standard model, or any extension of it that allows a restriction to Hermitian mass matrices only.
Although we have tried to argue that our results are independent of renormalization group flow since the relevant quantities do not run strongly with energy scale, there is another subtle issue: The low-energy limit of a left-right symmetric extension with non-Hermitian mass matrices would still be the standard model, where mass matrices can be assumed to be Hermitian, leading us back to the measure considered in \cite{paper, prl}. It would be desirable to incorporate this dependency of the assumptions one has used to construct the measure on energy scale into the analysis, namely to use a measure which depends on energy scale also. A starting point would be a quantification of ``non-Hermiticity" that could then flow from zero at low energies to some non-zero value at high energies. At present these ideas are however somewhat vague, so that we will have to leave them to exploration in future work.
\section*{Acknowledgments}
The groundwork for the simple calculations presented here was laid in the previous papers \cite{paper,prl}, and I thank my collaborators in this previous work for many fruitful discussions and suggestions. I am supported by EPSRC and Trinity College, Cambridge. I should also thank the referee for suggestions that hopefully led to an improvement of presentation.
|
\section{Introduction}
\label{s1}
Granular matter in general \cite{JNyB96,Du00}, and granular gases in particular \cite{Go03,ByP04,AyT06},
have been recently the object of intensive theoretical and experimental research. Not in the least this
is because of the opportunity they offer to investigate fundamental questions of non-equilibrium physics.
One of these issues is the nature and properties of non-equilibrium fluctuations and their relevance for
the description of the macroscopic behavior of the system. The simplest fluctuations that can be considered
are those of global properties of a system. The total energy fluctuations of an isolated
granular gas, modeled as an ensemble of smooth inelastic hard spheres, have already been investigated
elsewhere \cite{BGMyR04,VPBWyT06}. They are specific of granular systems, and a consequence of the localized
character of the energy dissipation in collisions.
In this paper, the volume fluctuations of a vibrated granular gas are
studied by means of event driven simulations \cite{AyT87,PyS05}. The volume of the system changes because the wall on
top of the grains is a piston that can move in the vertical direction. The equilibrium position of the piston
is determined by equating its weight per unit of area with the pressure of the granular gas just below it.
On the other hand, the properties of the fluctuations of the piston around that position, and therefore
the volume fluctuations of the system, are not known. If the system were a molecular gas at equilibrium,
therefore without being energized by vibration, the total volume fluctuations are known to be Gaussian, and with
a second moment that is proportional to the temperature and the isothermal compressibility of the
gas \cite{LyL79}. There is no
reason to expect the above properties to hold in granular gases, that are inherently in non-equilibrium states. Even
more, to give a meaning to them, a first point to be addressed is to specify the notion of temperature to be used.
Actually, the equilibrium
expression for the volume fluctuations can be used to define a temperature-like quantity, that can be experimentally
measured. This is somehow an extension of the usual way of defining ``effective temperatures'' from the extension to
non-equilibrium states of the Fluctuation-Dissipation Theorem. A revision of these ideas in the context of
granular media is given in \cite{BBALMyP05}. See also \cite{PByV07} for a discussion of the validity of the Einstein
relation in externally driven granular gases.
The emerging natural question is whether the effective temperature defined
from the relationship between volume fluctuations and compressibility has some other conceptual interpretation,
and if it is related to other possible sensible definitions of temperature in the system. The first obvious
candidate to be considered is the granular temperature of the gas, defined from the second moment of the
velocity distribution of the grains. Actually, this parameter is known to play in the hydrodynamic description of
granular gases a role similar to the usual temperature in molecular hydrodynamics \cite{Ca90}. But there is another
temperature parameter that is relevant for the description of the system, the one defined from the second moment of
the velocity distribution of the movable piston. Both temperatures, defined from the velocity of the gas and of the piston, respectively,
are the same in equilibrium systems, but they can differ strongly in granular systems, as a manifestation of
the violation of energy equipartition \cite{ByR08a}. Clarifying the relationship between the effective temperature
defined from the volume fluctuations, the granular temperature, and the temperature parameter of the piston, is
one of the aims of this work.
The structure of the remaining of this paper is as follows. In Sec. \ref{s2}, the system is described
and the macroscopic steady state to be considered is characterized. Also, the parameter region to be
investigated is specified. It is shown than when the mass of the piston is much larger than the mass of the
grains, the volume fluctuations of the system exhibit a Gaussian distribution. The second moment of this distribution
depends much stronger on the inelasticity of collisions between particles than on the elastic or inelastic
character of the collisions of the particles with the piston. In Sec. \ref{s3}, a compressibility factor is defined for the
granular gas. As usual, it is a measurement of the change in the volume of the system as a consequence of a change in
the external pressure under well defined constrains. From the values of the second moment of the volume fluctuations and of the compressibility factor,
an effective temperature is defined, as indicated above. This temperature turns out to be quite simply related with
the temperature parameter of the piston, but its relationship with the granular temperature of the gas
appears to be very intricate. Finally, Sec. \ref{s4} contains a short summary of the main results in the paper and some
general comments.
\section{Steady fluctuations of the position of the piston}
\label{s2}
Consider a system composed by $N$ smooth inelastic hard disks of mass $m$ and
diameter $d$, in presence of gravity, and confined by a movable piston of mass $M$
located on the top. By definition, the piston can only move in the direction of the
gravity field. Moreover, there is no friction between the piston and the lateral
walls of the vessel containing the gas. The system of particles is kept fluidized and at low density
by injecting energy through the
bottom wall, which is vibrating. Inelasticity in collisions between particles is modeled by means of a
constant coefficient of normal restitution,
$\alpha$, defined in the interval $0 < \alpha < 1$. Thus when two particles $i$ and $j$ collide
their velocities change instantaneously from the initial values ${\bm v}_{i}$, ${\bm v}_{j}$ to the
post-collisional ones given by
\begin{equation}
\label{2.1}
{\bm v}_{i}^{\prime} = {\bm v}_{i}- \frac{1+ \alpha}{2} \left( \widehat{\bm \sigma} \cdot {\bm v}_{ij}
\right) \widehat{\bm \sigma},
\end{equation}
\begin{equation}
\label{2.2}
{\bm v}_{j}^{\prime} = {\bm v}_{j} + \frac{1+ \alpha}{2} \left( \widehat{\bm \sigma} \cdot {\bm v}_{ij}
\right) \widehat{\bm \sigma},
\end{equation}
where ${\bm v}_{ij} \equiv {\bm v}_{i}-{\bm v}_{j}$ is the relative velocity and $\widehat{\bm \sigma}$ is
the unit vector joining the center of the two particles at contact. The $z$ axis will be taken in the direction
of the gravitational field, so that the particles are submitted to an external force of the form ${\bm f}=- m g_{0}
\widehat{\bm e}_{z}$, $g_{0}$ being a positive constant and $\widehat{\bm e}_{z}$ the positive unit vector
along the $z$ axis. Collisions of particles with the movable piston on the top are also considered as smooth and
inelastic, $\alpha_{P}$ being the coefficient of normal restitution for them.
Therefore, in a collision between particle $i$ and the piston,
the component $v_{i,x}$ of the velocity of the particle perpendicular to the $z$ axis remains unchanged,
\begin{equation}
\label{2.3} v_{i,x}^{\prime}= v_{i,x},
\end{equation}
while the component $v_{i,z}$ of the velocity of the particle and the velocity
$V_{z}$ of the piston are instantaneously modified accordingly with
\begin{equation}
\label{2.4} v^{\prime}_{i,z} =v_{i,z} -\frac{M}{m+M}\, (1+\alpha_{P})
(v_{i,z}-V_{z})
\end{equation}
and
\begin{equation}
\label{2.5} V^{\prime}_{z} =V_{z} +\frac{m}{m+M}\, (1+\alpha_{P}) (v_{i,z}-V_{z}),
\end{equation}
respectively. When the transversal section of the system, i.e. the size of the piston $W$, is smaller than
a critical value \cite{BRMyG02,LMyS02}, the gas reaches, after a transient time interval, a stationary state with gradients only in the
direction of the gravitational field. If, in addition, the inelasticity of the system is small, the inelastic hydrodynamic
Navier-Stokes equations with the appropriate boundary conditions \cite{BDKyS98,ByC01} provide an accurate description of
the stationary state \cite{ByR09a}. As the coefficient of normal
restitution of the gas $\alpha$ decreases, significant deviations from the predictions following from
the Navier-Stokes equations show up. They are due to the coupling between inelasticity and gradients that
exists in the stationary state, in such a way that strong inelasticity implies large gradients of the hydrodynamic fields.
This coupling is a peculiarity of the steady states of inelastic fluids, following from the balance between the energy
dissipated because of inelastic cooling and the hydrodynamic energy fluxes.
A previous analysis, carried out in \cite{ByR09a}, focussed on the macroscopic description of the granular gas
in terms of the density and granular
temperature fields, the velocity field being zero. At this level of description, the role of the movable piston
on top of the gas is to partially determine the boundary conditions needed to solve the hydrodynamic equations for
the steady state under consideration. Here, the interest will be on the fluctuations of the movable piston, namely on its
position fluctuations. Some results for the velocity fluctuations have been reported elsewhere \cite{ByR08a}.
There, it was shown that the steady state velocity fluctuations of the piston are gaussian with zero mean for
$\alpha_{P} \geq 0.6$ and $\alpha \geq 0.8$. Nevertheless, no simple relationship between the second moments
of the velocity distributions of the piston and the gas next to it was found. It is worth to remark that
there is no reason to expect such a relation to exist at a macroscopic level of description,
i.e. involving only the hydrodynamic fields and the parameters of the system. Actually, the simulation results
reported in \cite{ByR08a} indicate that the details of the velocity distribution of the gas, beyond its first
few moments, are relevant to determine the second moment of the velocity distribution of the piston.
It is clear that the position fluctuations of the movable piston are related with the volume
fluctuations of the inelastic gas. Actually, this relationship can be made direct and exact by properly choosing
the nature of the vibrating wall located at the bottom of the system. The mission of the latter is to energize the system,
keeping the particles fluidized. The expectation is that the behavior in the bulk of the system is independent of the details of the
way in which this wall is being vibrated. Consequently, the
simplest possible choice has been used in all the results to be reported in the following.
The bottom wall is vibrated with a sawtooth velocity profile, having a velocity $v_{W}$. This means that all the
particles colliding with the wall find it moving upwards with that velocity \cite{McyB97,McyL98}. Besides, the amplitude of
the wall motion is considered much smaller than the mean free path of the particles in its vicinity, so that
the position of the wall can be taken as fixed at $z=0$. Therefore, the dynamics of the
vibrating wall at the bottom does not induce directly any change in the volume (or area) occupied by the granular gas.
Also, and again for the sake of simplicity, collisions of the particles with this wall will be considered as elastic.
In the event-driven simulations carried out, periodic boundary conditions were
used in the $x$ direction. The width of the system and the number of particles in it were fixed to $W= 70 d$
and $N_{z} \equiv N / W =6 d^{-1}$, respectively. Moreover, the velocity of the
vibrating wall $v_{W}$ was chosen in each case large enough, not only to fluidize the system, but also
to guarantee that the density remains small throughout the granular gas and, consequently, the dilute limit can be expected to
be accurate. In this case, the dependence of the hydrodynamic profiles on $v_{W}$ is very simple
and follows by dimensional analysis \cite{ByR09a}. The dependence of the position fluctuations of the piston on this velocity
will be discussed later on. The value of the coefficient of normal restitution for the collisions
between particles has been varied within the interval $0.85 \leq \alpha <1$. This includes a range of values for
which the Navier-Stokes hydrodynamic description is not accurate, due to the coupling between inelasticity and
gradients already pointed out \cite{ByR09a}. The coefficient of restitution for the collisions of the particles with the
movable piston has been set to $\alpha_{P}=0.99$ in most of the simulations being reported, but it has been verified that
the results depend very
weakly on the value of this coefficient, remaining practically the same when $\alpha_{P}$ is decreased, at least
up to $\alpha_{P}=0.8$. Some examples of this behavior will be given below.
In all the simulations being presented, it was observed that the height $Z$ of the movable piston oscillates about
an average value $<Z> =L$, once the steady state is reached. As an example, in Fig.\ \ref{fig1} the time evolution
of the scaled position of the piston, $Z^{*} \equiv Z g_{0} /v_{W}^{2}$, is shown for $\alpha = 0.95$, $\alpha_{P}=0.99$,
and three
choices of the mass of the piston: $M=30 m$, $M=75 m$, and $M=150m$. Time $\tau$ is measured in accumulated
number of collisions per particle. It is observed that both, the value of $L$ and the amplitude of the fluctuations,
decrease as $M$ increases. Of course, this is the expected behavior. Also notice that the trajectory
of the piston does not exhibit systematic oscillations, but it is apparently random. This indicates that the motion
of the piston does not have any hydrodynamic component induced, for instance, by the vibrating wall at the bottom.
\begin{figure}
\includegraphics[scale=0.7,angle=0]{byr09bf1.eps}
\caption{Time evolution of the dimensionless scaled position $Z^{*} \equiv Z g_{0}/ v_{W}^{2}$ of the movable piston
located on top of the system, once the steady state has been reached. There are $420$ disks in a box of width
$W=70 d$. In all cases the coefficient of normal restitution of the particle collisions is $\alpha= 0.95$ and that
for the particle-piston collisions is $\alpha_{P}=0.99$. Results for three
values of the mass of the piston M are shown, as indicated. Time $\tau$ is measured in accumulated number of collisions
per particle. } \label{fig1}
\end{figure}
From the steady trajectory of the piston, the probability distribution for its position can be built. To increase
the statistics, several trajectories have been generated for each set of values of the parameters. As already mentioned,
the fluctuations of the piston depend on the parameters defining the system. To see whether this dependence occurs
only through the first two moments of the probability distribution, a normalized length $\ell$ has been defined as
\begin{equation}
\label{2.6}
\ell \equiv \frac{Z-L}{\sigma_{Z}}=\frac{Z^{*}-L^{*}}{\sigma_{Z}^{*}},
\end{equation}
where $\sigma_{Z}$ is the square root of the second central moment or standard deviation of $Z$, $\sigma_{Z}^{2} \equiv
<Z^{2}>-L^{2}$, and the star indicates that lengths are being measured in
the dimensionless scale defined above. In Fig. \ref{fig2}, the obtained probability distribution of $\ell$,
$P(\ell)$, is plotted for a system with $\alpha=0.94$. Again, three values of the mass ratio have been considered, namely
$M=36 m$, $M = 60 m$, and $M= 120 m$. It is seen that the probability distributions are accurately fitted by a Gaussian
(solid lines), at least up to values of the probability density of the order of $10^{-4}$. A similar behavior has been
found in all the simulated systems with parameters within the ranges mentioned above, although it seems that a
small but systematic deviation shows up as the mass of the piston $M$ becomes smaller, approaching the mass of the
particles. A possible explanation for this behavior is that, as the mass of the piston decreases, the amplitude of its position
fluctuations increases, and the effect of the external gravitational field breaks the symmetry of the fluctuations around the
average position. To check this idea and to quantify the deviations from the Gaussian of
the position fluctuations, the third and fourth moments
of $\ell$ have been computed from the simulation data. The results for some of the simulations are given in Table
\ref{table1}. For a Gaussian
distribution it is $<\ell^{3}> = 0$ and $<\ell^{4}> =3$. The deviations of the third moment from the Gaussian value are much stronger than those
of the fourth one, supporting the idea that the main cause of the deviation from the Gaussian is due to the symmetry breaking
produced by the external field, when the mass of the piston is not much larger than the mass of the particles. In any case,
the deviations are rather weak and it can be concluded that, in the explored parameter region, the position fluctuations
of the piston can be considered as Gaussian with a very good accuracy.
\begin{figure}
\includegraphics[scale=0.7,angle=0]{byr09bf2a.eps}
\includegraphics[scale=0.7,angle=0]{byr09bf2b.eps}
\caption{Steady position distribution of the piston $P(l)$ in both normal and logarithmic scales. The symbols are
from the simulations while the solid lines are Gaussian with unity dispersion. The data correspond to three systems differing
in the mass of the piston, as indicated. The coefficient of normal restitution for the particle collisions is $\alpha=0.94$.
The dimensionless length $\ell$ is defined in Eq.\ (\protect{\ref{2.6}}).} \label{fig2}
\end{figure}
\begin{table}
\caption{Third and fourth moments of the position distribution of the piston, obtained from the simulations.
The position $\ell$ is given in the dimensionless scale defined in Eq.\ (\protect{\ref{2.6}}).} \label{table1}
\begin{ruledtabular}
\begin{tabular}{lllll}
$ \alpha $ & $ \alpha_{P} $ & $ M/m $ & $ <\ell^{3}> $ & $ <\ell^{4}> $ \\
0.98 & 0.99 & 24 & 0.250 & 3.083 \\
& & 60 & 0.172 & 3.190 \\
& & 120 & 0.0385 & 2.898 \\
& 0.8 & 24 & 0.254 & 3.230 \\
& & 60 & 0.136 & 3.047 \\
& & 120 & 0.057 & 2.922 \\
0.94 & 0.99 & 24 & 0.242& 3.138 \\
& & 60 & 0.125 & 2.996 \\
& & 120 & 0.047 & 3.006 \\
& 0.8 & 24 & 0.145 & 2.964 \\
& & 60 & 0.213 & 3.088 \\
& & 120 & 0.101 & 2.985 \\
\end{tabular}
\end{ruledtabular}
\end{table}
In ref. \cite{ByR09a} it was shown that the values of the average position of the piston $L$ scale with $v_{W}^{2}$ in the
low density limit and, consequently, $L^{*}$ does not depend on the velocity of the vibrating wall in this limit.
The extension of this result to volume fluctuations requires to go beyond hydrodynamics.
To investigate whether $\sigma^{*}_{Z}$ also has the above scaling property, several series of simulations have been performed varying
the value of $v_{W}$, while keeping constant all the other parameters. It is important to stress that
$v_{W}$ was always chosen large enough as to fluidize the system and to avoid the presence of regions with density above
what is considered the low density range. In all the cases investigated, there was no dependence of $\sigma^{*}_{Z}$ on
$v_{W}$, within the
statistical uncertainties. Therefore, in the low density limit, the standard deviation of the piston position seems to
scale with the square of the velocity of the vibrating wall, i.e. in the same way as the average position $L$.
In Fig.\ \ref{fig3}, the relative standard deviation $\sigma_{Z} / L = \sigma_{Z}^{*} / L^{*}$ is plotted as a function of
the mass ratio $M/m$, for several values of the coefficient of normal restitution of the gas $\alpha$ in the interval
$0.85 \leq \alpha \leq 0.98$. The coefficient of normal restitution for particle-movable piston collisions is in all the
cases $\alpha_{P} =0.99$. It is observed that, for given $\alpha$, there is a region in which $\sigma_{Z} / L$ decreases
as the mass ratio increases. This effect is less pronounced the smaller the coefficient of restitution $\alpha$, i.e.
the more inelastic the collisions. For large values of $M/m$, the results in the figure indicate that $\sigma_{Z}/L$ tends to a plateau with
a constant value. The values of the mass ratio needed to reach the plateau monotonically decrease as $\alpha$
decreases. The data in Fig. \ref{fig3} also indicate that for constant mass ratio, the relative fluctuations increase as $\alpha$
decreases.
To show that the influence of the inelasticity of the particle-movable piston collisions is much weaker than that
of the particle-particle collisions, in Fig. \ref{fig4} the relative standard deviation is plotted as a function
of the mass ratio for two pairs of data. Each pair corresponds to the same value of $\alpha$ (namely, $0.98$ and $0.94$),
but different values of $\alpha_{P}$ (namely, $0.8$ and $0.99$). Although the variation of $\alpha_{P}$ is almost five times
the variation of $\alpha$, it is seen that the data corresponding to the same $\alpha$ are much closer than those
corresponding to the same $\alpha_{P}$. On the other hand, there is a relevant qualitative feature to be stressed.
While decreasing $\alpha$ produces an increase of the relative fluctuations, decreasing $\alpha_{P}$ has the
opposite effect: the relative fluctuations also decrease.
\begin{figure}
\includegraphics[scale=0.7,angle=0]{byr09bf3.eps}
\caption{(Color online) Dimensionless relative standard deviation $\sigma_{Z}/L$ of the position of the
movable piston as a function of the mass ratio $M/m$, for several values of the coefficient of normal restitution of
the gas $\alpha$, as indicated in the insert. The curves are guides for the eye. In all the cases, the coefficient of restitution
for the gas-movable piston collisions is $\alpha_{P}=0.99$.}
\label{fig3}
\end{figure}
\begin{figure}
\includegraphics[scale=0.7,angle=0]{byr09bf4.eps}
\caption{(Color online) Dimensionless relative standard deviation $\sigma_{Z}/L$ of the position of the
movable piston as a function of the mass ratio $M/m$. The symbols are simulation data and the lines guides for the eye.
The two upper curves correspond to $\alpha=0.94$, and the two lower ones to $\alpha =0.98$. In each case,
two values of $\alpha_{P}$ have been employed: $0.99$ (circles) and $0.8$ (triangles).}
\label{fig4}
\end{figure}
\section{Compressibility and effective temperature}
\label{s3}
To measure the facility of the system to be compressed, define a coefficient of compressibility
$k$ by
\begin{equation}
\label{3.1}
k \equiv - \frac{1}{<V>} \left( \frac{\partial <V>}{\partial p_{L}} \right)_{v_{W}},
\end{equation}
where $p_{L}$ is the pressure of the granular gas in the vicinity of the movable piston and $V$ is
the volume (area) of the system. The derivative in the above equation is computed at constant value
of all the parameters defining the system, $\alpha$, $\alpha_{P}$, $N_{z}$ and $v_{W}$, except
$p_{L}=Mg_{0}/W$, as it follows from the definition of the pressure. In the following, $p_{L}$ will be modified by changing the mass of the piston $M$,
keeping $g_{0}$ and $W$ unchanged. The reason for this choice is twofold. First, changing $g_{0}$ is equivalent
to modifying $v_{W}$ and, second, increasing $W$ can lead to the set up of transversal instabilities, as already
mentioned. The notation stresses the constancy of the velocity of the vibrating wall. Although other
compressibility coefficients could be defined, the one in (\ref{3.1}) has the
advantage of being easy to implement in experiments, at least at a conceptual level. On the other hand, it is worth
to remark that by keeping $v_{W}$ constant and changing $M$, both the hydrodynamic profiles inside the
fluid and the power injected into it through the vibrating wall are modified. In particular, the latter is given
by \cite{ByR09a}
\begin{equation}
\label{3.2}
Q_{0}= W \left( N_{z}+ \frac{M}{mW} \right) m g_{0}v_{W}.
\end{equation}
Therefore, increasing $M$ while keeping $v_{W}$ constant produces an increase of the power $Q_{0}$. For the two
dimensional systems being
considered here, the definition (\ref{3.1}) is equivalent to
\begin{equation}
\label{3.3}
k= -\frac{W}{Lg_{0} } \left( \frac{\partial L}{\partial M} \right)_{v_{W}}.
\end{equation}
The idea of introducing a compressibility for vibrated granular fluid was already used in ref. \cite{ACyS02} in the context
of hydrodynamical stability analysis.
In a molecular system at equilibrium, the isothermal compressibility
\begin{equation}
\label{3.4}
k_{T} = -\frac{1}{<V>} \left( \frac{\partial <V>}{\partial p} \right)_{T}
\end{equation}
is related with the volume fluctuations by
\begin{equation}
\label{3.5}
\sigma_{V}^{2} \equiv <(V-<V>)^2> = k_{B}T <V> k_{T},
\end{equation}
where $k_{B}$ is the Boltzmann constant. In non-equilibrium states, there is
no reason to expect the above relationship to hold, but it is tempting to
employ it to define an effective temperature parameter of the system, $T_{eff}$,
expecting it to have some intrinsic physical meaning. Therefore, taking into account
that the volume fluctuations in the case being considered are associated to fluctuations of the
position $Z$ of the movable piston, we define $T_{eff}$ through
\begin{equation}
\label{3.6}
\sigma_{Z}^{2}= - \frac{T_{eff}}{g_{0}} \left( \frac{\partial L}{\partial M} \right)_{{v}_{W}}.
\end{equation}
The Boltzmann constant has been set equal to unity as it is usually done when defining
the granular temperature from the average kinetic energy of the grains. From Eq.\ (\ref{3.6}) it follows that
$T_{eff}$ relates the volume response to a pressure perturbations with the steady volume fluctuations of the
system. Of course, Eq. (\ref{3.6}) by itself is just a mathematical definition and does not add anything to
the physical understanding of the system. On the other hand, the definition would become relevant if
this effective temperature were related to other temperature-like parameters of the system. Two main candidates
clearly stand out: the granular temperature of the gas in the vicinity of the piston, $T_{L}$, and
the temperature parameter of the piston, $T_{P}$, defined through $T_{P} = M <V_{z}^{2}>$ \cite{ByR09a}.
Both parameters, $T_{L}$ and $T_{P}$, are not at all the same, as it should be the case if the system
under consideration were at equilibrium and energy equipartition would apply. Violation of equipartition is
a general feature of granular systems known since long ago \cite{JyM87} and that has attracted a lot of attention
in the last years. A more detailed discussion of this issue for the set up being considered here is given in
\cite{ByR08a}. In Fig. \ref{fig5}, event driven simulation results for the ratio $T_{P}/T_{L}$ are plotted
as a function of $M/m$. Several values of the coefficient of normal restitution of the gas in the interval
$0.85 \leq \alpha \leq 0.98$ have been considered, while again $\alpha_{P}=0.99$ for all the data shown. It is observed that
the behavior of this temperature ratio is quite intricate. For instance, $T_{P}/T_{L}$ decreases as $M/m$ increases
for $\alpha>0.95$, but it happens the other way round for $\alpha < 0.95$. Actually, as $\alpha$ decreases below
this value, the increase of the temperature ratio is rather fast, and $T_{P}$ reaches to be up to four times larger
than $T_{L}$. If the parameters $T_{P}$ and $T_{L}$ were interpreted as real temperature parameters, this latter
behavior would be fully counterintuitive. The temperature of the heated body (the piston) is larger than the
temperature of the heating one (the gas next to the piston).
\begin{figure}
\includegraphics[scale=0.7,angle=0]{byr09bf5.eps}
\caption{(Color online) Ratio of the temperature parameter of the piston, $T_{P}$, to the granular temperature
of the gas next to it, $T_{L}$, versus the mass ratio $M/m$, for several values of the coefficient of restitution of the
gas $\alpha$, as indicated in the insert. The curves are guides for the eye. In all the cases, the coefficient of restitution
for the gas-movable piston collisions is $\alpha_{P}=0.99$.}
\label{fig5}
\end{figure}
From the values of $\sigma_{Z}$ and $(\partial L / \partial M)_{v_{W}}$ obtained from the event driven simulation data, the effective
temperature $T_{eff}$ has been computed by means of its definition, Eq.\ (\ref{3.6}). In the Appendix some details are
given of the way in which the above derivative was actually evaluated. Then, in Figs. \ref{fig6} and \ref{fig7} the
temperature ratios $T_{eff}/T_{L}$ and $T_{eff}/T_{P}$ are plotted, respectively, as a function of the mass ratio for the same
systems as in Fig. \ref{fig5}. A clear difference is observed in the behavior of the two temperature ratios for a given
constant coefficient of restitution $\alpha$. While $T_{eff}/T_{L}$ exhibits a strong dependence on $M/m$ and does not
seem to tend to a well defined limit as it increases, the dependence of $T_{eff}/T_{P}$ on $M/m$ is very weak, being only
appreciable for the least inelastic cases and when the mass ratio is small. Thus it is concluded that for
large enough mass ratio $M/m$, the effective temperature is proportional to the temperature
parameter of the piston with a coefficient of proportionality that is independent of the mass of the piston, i.e.
\begin{equation}
\label{3.7}
T_{eff} = b(\alpha,\alpha_{P}) T_{P},
\end{equation}
where a possible dependence of the coefficient on $\alpha_{P}$ has been included.
\begin{figure}
\includegraphics[scale=0.7,angle=0]{byr09bf6.eps}
\caption{(Color online) Ratio of the effective temperature, $T_{eff}$, defined in Eq. (\protect{\ref{3.6}}) to the
temperature of the gas in the vicinity of the piston, $T_{L}$, as a function of the mass ratio for the same systems as in
Fig. \protect{\ref{fig5}}.}
\label{fig6}
\end{figure}
\begin{figure}
\includegraphics[scale=0.7,angle=0]{byr09bf7.eps}
\caption{(Color online) Ratio of the effective temperature, $T_{eff}$, defined in Eq. (\protect{\ref{3.6}}) to the
temperature parameter of the piston, $T_{P}$, as a function of the mass ratio for the same systems as in
Fig. \protect{\ref{fig5}}. }
\label{fig7}
\end{figure}
To identify the dependence of the coefficient $b$ on $\alpha$, in Fig. \ref{fig8} the plateau values
of $T_{eff}/T_{P}$, reached upon increasing the value of the mass ratio $M/m$, are plotted versus $1- \alpha^2$.
The points are very well fitted by a straight line, indicating that $T_{eff}/T_{P}$ grows linearly with
$1-\alpha^{2}$, at least in the considered interval, $0.85 \leq \alpha \leq 0.98$. It is worth to note that
some care is needed when extrapolating to the elastic limit $\alpha \rightarrow 1$ these results. In this limit,
a stationary state is only possible if, in addition, the vibrating wall is arrested, i.e. also the limit
$v_{W} \rightarrow 0$ is taken.
But all the previous discussion has been carried out at constant velocity of the vibrating wall. This explains
why extrapolation of the linear fitting in Fig. \ref{fig8} does not lead to $b=1$, and this does not mean any
kind of contradiction.
\begin{figure}
\includegraphics[scale=0.7,angle=0]{byr09bf8.eps}
\caption{Ratio $T_{eff}/T_{P}$ for large values of $M/m$ as a function of $1-\alpha^{2}$ for the same
systems as in Fig. \protect{\ref{fig5}}. The symbols are simulation results and the dashed line a linear fit of them.}
\label{fig8}
\end{figure}
In Fig. \ref{fig9}, the dimensionless compressibility,
$-m L^{-1} \left( \partial L / \partial M \right)_{v_{W}} $ and the scaled second moment of the position fluctuations
$\sigma_{Z}^{2} m g_{0} / L b(\alpha,\alpha_{P}) T_{P}$ have been plotted as functions of $M/m$. A logarithmic
representation is employed. Data for several values of the coefficient $\alpha$ have been included. For all the
data $\alpha_{P} = 0.99$. If
$T_{eff}$ had been used instead of $ b(\alpha,\alpha_{P}) T_{P}$, the two plotted quantities would be the same
by definition, i.e. the filled and empty symbols would agree in all the cases. It is seen that the dependence on $M/m$
of the dimensionless compressibility is accurately described by a power law of the form $(M/m)^{-3/4}$, indicated in the
figure by the solid straight line. Nevertheless, it is important to realize that the interval of values of $M/m$
for which the above dependence is identified is rather narrow, just one order of magnitude, so that its range of
validity can be limited.
\begin{figure}
\includegraphics[scale=0.7,angle=0]{byr09bf9.eps}
\caption{(Color online) Dimensionless compressibility $-m L^{-1} \left( \partial L / \partial M \right)_{v_{W}} $ (filled
symbols) and scaled position fluctuations of the piston $\sigma_{Z}^{2} m g_{0} / L b(\alpha,\alpha_{P}) T_{P}$ (
empty symbols) as a function of the mass ratio, $M/m$. The different symbols correspond to different values of the
coefficient of restitution for the collision between particles, $\alpha$, as indicated in the insert. The straight
line has a slope $-3/4$, and it is a guide for the eye.}
\label{fig9}
\end{figure}
\section{Discussion and summary}
\label{s4}
The aim here has been to investigate the volume fluctuations of a vibrated low density
gas of inelastic hard disks in presence of gravity, and confined by a movable piston on the top. The study has
been restricted to the parameter region in which the system reaches a steady state with gradients only in
the direction of the external gravitational field, i.e. perpendicular to the movable piston.
In practice, this has limited the values of the coefficient of normal restitution of the gas particles to
the interval $ 0.85 \leq \alpha < 1$. Due to the coupling between inelasticity and hydrodynamic gradients,
which is peculiar of steady states of granular systems, the above limitation also implies restriction to small gradients.
Nevertheless, the analysis carried out in ref.\ \cite{ByR09a} indicates that the range of hydrodynamic gradients
considered here exceeds the limit of validity of the Navier-Stokes approximation.
Some of the main results can be summarized as follows: i) for large mass of the movable piston compared with the mass of the gas
particles, the volume fluctuations are Gaussian with very good accuracy, ii) the square
root of the second moment of their distribution scales with the square of the velocity of the vibrating wall at the
bottom, i.e. in the same way as the amplitudes of the hydrodynamic fields in the gas, iii) by requiring the same relation between volume
fluctuations and compressibility as in equilibrium systems to be verified, an ``effective temperature'' can be defined,
iv) the effective temperature turns out to be proportional to the second moment of the velocity fluctuations of the
piston, with a proportionality parameter that depends on the inelasticity of both the particle-particle and particle-piston
collisions, but it seems to be independent of the mass of the piston, and v) the effective temperature can not be
related in a simple way to the temperature of the granular gas; even more, the relationship between both parameters is
not monotonic.
A relevant open question is the relationship between the granular temperature of the gas in the vicinity
of the piston and the temperature parameter of the piston, the latter defined from the second moment of its velocity
distribution. An explanation of the simulation results seems to require a detailed knowledge of the velocity
distribution function of the gas next to the piston \cite{ByR08a}. If this is the case, approximated solutions of the
Boltzmann equation,
as provided by instance by the Chapman-Enskog procedure in the first Sonine approximation, would not
be of enough accuracy as to describe the deviation from equipartition between the gas and the movable piston.
The present study complements the one in ref. \cite{ByR08a}, in which the velocity fluctuations of the piston were
investigated in detail. A natural issue now is whether the velocity fluctuations and the position
fluctuations of the piston are correlated. We have computed from the simulation data the joint probability distribution
for the position and velocity of the piston and compared it with the product of the marginal
distributions for the position and the velocity. Both results agree within the statistical uncertainties, indicating
the absence of correlations.
\section{Acknowledgements}
This research was supported by the Ministerio de Educaci\'{o}n y
Ciencia (Spain) through Grant No. FIS2008-01339 (partially financed
by FEDER funds).
|
\section{Appendix}
\label{app}
In this section we prove the following lemmas.
\begin{lemma}
\label{l4}
The order ideal of $\mbox{Abs}\,(S_n)$ generated by all cycels $u\in S_n$ for which $\pi_n(u)=(1\,2\,\cdots\,n-1)$ is homotopy Cohen-Macaulay of rank $n-1$.
\end{lemma}
\begin{lemma}
\label{l4'}
\begin{enumerate}[leftmargin=*, itemsep=5pt]
\item[\emph{(i)}] The order ideal of $\mbox{Abs}\,(B_n)$ generated by all cycles $u\in B_n$ for which $\pi_n(u)=\lleft1,2,\dots,n-1) \! )$ is homotopy Cohen-Macaulay of rank $n-1$.
\item[\emph{(ii)}] The order ideal of $\mbox{Abs}\,(B_n)$ generated by all cycles $u$ of $B_n$ for which $\pi_n(u)=[1,2,\dots,n-1]$ is homotopy Cohen-Macaulay of rank $n$.
\end{enumerate}
\end{lemma}
\smallskip
\subsection{Proof of Lemma \ref{l4}}
\noindent We will show that the order ideal considered in Lemma \ref{l4} is in fact strongly constructible.
The following remark will be used in the proof.
\begin{remark}
\label{L0}
Let $u_1,u_2,\dots, u_m\in S_n$ be elements of absolute length $k$ and let $v\in S_n$ be a cycle of absolute length $r$
which is disjoint from $u_i$ for each $i\in\{1,2,\dots, m\}$.
Suppose that the union $\bigcup_{i=1}^m[e,u_i]$ is strongly constructible of rank $k$.
Then
\[\bigcup\limits_{i=1}\limits^m\,[e,vu_i]\,\cong\,\bigcup\limits_{i=1}\limits^m\left([e,v]\times[e,u_i]\right)\,=\,[e,v]\times \bigcup\limits_{i=1}\limits^m\,[e,u_i],\]
is strongly constructible of rank $k+r$, by Lemma \ref{strcon} (i).
\end{remark}
\begin{lemma}
\label{l1}
For $i\in\{1,2,\dots,n-1\}$, consider the element \[u_i=(1\,i+1\,\cdots\,n-1)(2\,3\,\cdots\,i\, n)\in S_n.\]
The union
$\bigcup_{i=1}^m[e,u_i]$ is strongly constructible of rank $n-2$ for all $1\leq m\leq n-1$.
\end{lemma}
\begin{proof}
We denote by $I(n,m)$ the union in the statement of the lemma and proceed by induction on $n$ and $m$, in this order.
We may assume that $n\geq 3$ and $m\geq 2$, since otherwise the result is trivial.
Suppose that the result holds for positive integers smaller than $n$.
We will show that it holds for $n$ as well.
By induction on $m$, it suffices to show that $[e,u_m]\cap I(n,m-1)$ is strongly constructible of rank $n-3$.
Indeed, we have $[e,u_m]\cap I(n,m-1)=\bigcup_{i=1}^{m-1}[e,u_m]\cap[e,u_i]$
and \[[e,u_m]\cap[e,u_i]=[e,(1\,m+1\,m+2\cdots\,n-1)(2\,3\,\cdots\,i\,n)(i+1\,i+2\,\cdots\,m)].\]
Since the cycle $(1\,m+1\,m+2\cdots\,n-1)$ is present in the disjoint cycle decomposition of each maximal element of
$[e,u_m]\cap I(n,m-1)$, the desired statements follows easily from Remark \ref{L0} by induction on $n$.
\end{proof}
\begin{example}
If $n=6$ and $m=3$, then $I(n,m)$ is the order ideal of $\mbox{Abs}\,(S_n)$ generated by the elements $u_1=(1\,2\,3\,4\,5)(6),\,u_2=(1\,3\,4\,5)(2\,6)$
and $u_3=(1\,4\,5)(2\,3\,6)$.
The intersection \[[e,u_3]\cap\left([e,u_1]\cup[e,u_2]\right)\,=\,[e,(1\,4\,5)(2\,3)(6)]\cup[e,(1\,4\,5)(3)(2\,6)]\]
is strongly constructible of rank $3$ and $I(n,m)$ is strongly constructible of rank $4$.
\end{example}
\begin{lemma}
\label{l2}
For $i\in\{1,2,\dots,n-2\}$, consider the element
\[v_i=(1\,n\,i+2\,\cdots\,n-1)(2\,3\,\cdots\,i+1)\in S_n.\]
The union
$\bigcup_{i=1}^m[e,v_i]$ is strongly constructible of rank $n-2$ for all $1\leq m\leq n-2$.
\end{lemma}
\begin{proof}
The proof is similar to that of Lemma \ref{l1} and is omitted.
\end{proof}
\begin{lemma}
\label{l3}
Let $u_1,u_2,\dots,u_{n-1}\in S_n$ and $v_1,v_2,\dots,v_{n-2}\in S_n$ be defined as in Lemmas \ref{l1} and \ref{l2}, respectively.
If $I_n=\bigcup_{i=1}^{n-1}[e,u_i]$ and $I'_n=\bigcup_{i=1}^{n-2}[e,v_i]$,
then $I_n\cap I'_n$ is strongly constructible of rank $n-3$.
\end{lemma}
\begin{proof}
We proceed by induction on $n$. For $n=3$ the result is trivial, so assume that $n\geq 4$.
For $i,j\in\{2,3,\dots, n-1\}$ we set
\[z_{ij}=(1\,j+1\,\cdots\, n-1)(2\,3\,\cdots\, i)(n)(i+1\,\cdots\, j)\]
and
\[w_{ij}=(1\,i+1\,\cdots\, n-1)(2\,3\,\cdots\, j)(j+1\,\cdots\, i\,n).\]
We observe that
\[[e,u_i]\cap[e,v_j]
=\left\{\begin{array}{ll}
z_{ij}, & \mbox{if $i<j$}, \\
w_{ij}, & \mbox{if $i\geq j$},
\end{array}
\right.\]
Let $M_i$ be the order ideal of $\mbox{Abs}\,(S_n)$ generated by the elements
$w_{ij}$ for $2\leq j\leq i-1$.
Since $z_{ij}\preceq w_{ij}$ for all $i,j\in\{2,3,\dots, n-1\}$ with $i\neq j$,
we have $I_n\cap I'_n=\bigcup_{i=2}^{n-1} M_i$.
Each of the ideals $M_i$ is strongly constructible of rank $n-3$, by Remark \ref{L0} and Lemma \ref{l2}.
We prove by induction on $k$ that $\bigcup_{i=2}^k M_i$ is strongly constructible of rank $n-3$ for every $k\leq n-1$.
Suppose that this holds for positive integers smaller than $k$.
We need to show that $M_k\cap \left(\bigcup_{i=2}^{k-1} M_i\right)$ is strongly constructible of rank $n-4$.
For $i\leq k-1$ we have
\[M_k\cap M_i=\langle v\,(2\,3\,\cdots\,j)(j+1\,\cdots\,i\,n)(i+1\,\cdots\, k):\,j=2,3,\dots,i-1\rangle,\]
where $v= (1\,k+1\,\cdots\, n-1)$. Remark \ref{L0} and Lemma \ref{l2} imply that $M_k\cap M_i$ is a strongly constructible poset of rank $n-3$.
Since $v$ is present in the disjoint cycle decomposition of each maximal element of $M_k\cap \left(\bigcup_{i=2}^{k-1} M_i\right)$,
it follows by Remark \ref{L0} and induction on $n$ that $M_k\cap \left(\bigcup_{i=2}^{k-1} M_i\right)$
is strongly constructible of rank $n-3$ as well.
This concludes the proof of the lemma.
\end{proof}
\noindent\emph{Proof of Lemma \ref{l4}.}
We denote by $C_n$ the order ideal in the statement of the lemma. We will show that $C_n$ is strongly constructible of rank $n-1$
by induction on $n$. The result is easy to check for $n\leq 3$, so suppose that $n\geq 4$.
We have $C_n=\bigcup_{i=1}^{n-1}[e,w_i]$, where
$w_1=(1\,2\,\cdots\,n-1\,n),\, w_2=(1\,2\,\cdots\,n\,n-1),\dots,w_{n-1}=(1\,n\,2\,\cdots\,n-1)$.
By induction and Remark \ref{L0}, it suffices to show that $[e,w_{n-1}]\cap\left(\bigcup_{i=1}^{n-2}[e,w_i]\right)$ is strongly constructible of rank $n-2$.
We observe that for $1\leq i\leq n-2$ the intersection $[e,w_{n-1}]\cap[e,w_i]$ is equal to the ideal generated by $(1\,2\,\cdots\,n-1)$ and
the elements
\[u_{n-i}=(1\,n-i+1\,\cdots\,n-1)(2\,\cdots\,n-i\,n),\]
\[v_{n-i-1}=(1\,n\,n-i+1\,\cdots\,n-1)(2\,\cdots\,n-i),\]
considered in Lemmas \ref{l1} and \ref{l2}, respectively.
Hence $[e,w_{n-1}]\cap\left(\bigcup_{i=1}^{n-2}[e,w_i]\right)=I_n\cup I'_n$ and
the result follows from Lemmas \ref{l1},\,\ref{l2} and \ref{l3}. \qed
\smallskip
\subsection{Proof of Lemma \ref{l4'}.}
Part (i) of Lemma \ref{l4'} is equivalent to Lemma \ref{l4}.
The proof of part (ii) is analogous to that of Lemma \ref{l4}, with the following minor modifications in the statements of the
various lemmas involved and the proofs.
\begin{remark}
\label{L0'}
Let $u_1,u_2,\dots, u_m\in B_n$ be elements of absolute length $k$ which are products of disjoint paired cycles
and let $v\in B_n$ be a cycle of absolute length $r$
which is disjoint from $u_i$ for each $i\in\{1,2,\dots, m\}$.
Suppose that the union $\bigcup_{i=1}^m[e,u_i]$ is strongly constructible of rank $k$.
Then
\[\bigcup\limits_{i=1}\limits^m\,[e,vu_i]\cong\bigcup\limits_{i=1}\limits^m\left([e,v]\times[e,u_i]\right)=[e,v]\times \bigcup\limits_{i=1}\limits^m\,[e,u_i],\]
is strongly constructible of rank $k+r$, by Lemma \ref{tomes} (i).
\end{remark}
\begin{lemma}
\label{l1'}
For $i\in\{1,2,\dots, n-1\}$ consider the element \[u_i=[1,i+1,\dots,n-1]\lleft2,3,\dots,i, n) \! )\in B_n.\]
The union
$\bigcup_{i=1}^m[e,u_i]$ is strongly constructible of rank $n-1$ for all $1\leq m\leq n-1$.
\end{lemma}
\begin{proof}
The proof is similar to that of Lemma \ref{l1}.
\end{proof}
\begin{example}
Let $I(n,m)$ be the union in the statement of the Lemma \ref{l1'}. If $n=6$ and $m=3$, then $I(n,m)$ is the order ideal of $\mbox{Abs}\,(B_n)$ generated by the elements
$u_1=[1,2,3,4,5]\lleft6) \! ),\,u_2=[1,3,4,5]\lleft2,6) \! )$ and $u_3=[1,4,5]\lleft2,3,6) \! )$.
We have \[[e,u_3]\cap\left([e,u_1]\cup[e,u_2]\right)=[e,[1,4,5]\lleft2,3) \! )\lleft6) \! )]\cup[e,[1,4,5]\lleft3) \! )\lleft2,6) \! )].\]
This intersection is strongly constructible of rank $4$ and $I(n,m)$ is strongly constructible of rank $5$.
\end{example}
\begin{lemma}
\label{l2'}
For $i\in\{1,2,\dots, n-2\}$ consider the element
\[v_i=[1,n,i+2,\dots,n-1]\lleft2,3,\dots,i+1) \! )\in B_n.\]
The union
$\bigcup_{i=1}^m[e,v_i]$ is strongly constructible of rank $n-1$ for all $1\leq m\leq n-2$.
\end{lemma}
\begin{proof}
The proof is similar to that of Lemma \ref{l2}.
\end{proof}
\begin{lemma}
\label{l3'}
Let $u_1,u_2,\dots,u_{n-1}\in B_n$ and $v_1,v_2,\dots,v_{n-1}\in B_n$ be defined as in Lemmas \ref{l1'} and \ref{l2'}, respectively.
If $I_n=\bigcup_{i=1}^{n-1}[e,u_i]$ and $I'_n=\bigcup_{i=1}^{n-2}[e,v_i]$,
then $I_n\cap I'_n$ is strongly constructible of rank $n-2$.
\end{lemma}
\begin{proof}
We proceed by induction on $n$. For $n=3$ the result is trivial, so assume that $n\geq 4$.
Let $M_i$ be the order ideal of $\mbox{Abs}\,(B_n)$ generated by the elements $w_{ij}$ for $j\in\{2,3,\dots,i-1\}$,
where \[w_{ij}=[1,i+1,\dots, n-1]\lleft2,3,\dots, j) \! )( \! ( j+1,\dots, i,n) \! ).\]
We observe that $I_n\cap I'_n=\bigcup_{i=2}^{n-1} M_i$.
Each of the ideals $M_i$ is strongly constructible of rank $n-2$, by Remark \ref{L0'} and Lemma \ref{l2'}.
As in the proof of Lemma \ref{l3}, it can be shown by induction on $k$ that $\bigcup_{i=2}^k M_i$ is strongly constructible for every $k\leq n-1$.
\end{proof}
\section{Introduction and results}
\label{intro}
Coxeter groups are fundamental combinatorial structures which appear in several areas of mathematics.
Partial orders on Coxeter groups often provide an important tool for understanding the questions of interest.
Examples of such partial orders are the Bruhat order and the weak order.
We refer the reader to \cite{bjo0, bb, Hu} for background on Coxeter groups and their orderings.
In this work we study the absolute order. Let $W$ be a finite Coxeter group and let $\mathcal{T}$
be the set of \emph{all} reflections in $W$. The absolute order on $W$ is denoted by $\mbox{Abs}\,(W)$ and
defined as the partial order on $W$ whose Hasse diagram is obtained from the Cayley graph of $W$
with respect to $\mathcal{T}$ by directing its edges away from the identity (see Section \ref{abs}
for a precise definition). The poset $\mbox{Abs}\,(W)$ is locally self-dual and graded. It has a minimum
element, the identity $e\in W$, but will typically not have a maximum, since every Coxeter element
of $W$ is a maximal element of $\mbox{Abs}\,(W)$. Its rank function is called the absolute length and is
denoted by $\ell_\mathcal{T}$. The absolute length and order arise naturally in combinatorics
\cite{arm}, group theory \cite{Be, brwtt}, statistics \cite{Di} and invariant theory \cite{Hu}. For
instance, $\ell_{\mathcal{T}}(w)$ can also be defined as the codimension of the fixed space of $w$,
when $W$ acts faithfully as a group generated by orthogonal reflections on a vector space $V$ by
its standard geometric representation. Moreover, the rank generating polynomial of $\mbox{Abs}\,(W)$
satisfies
\[\sum_{w \in W}\ t^{\ell_{\mathcal{T}} (w)} \ = \ \prod_{i=1}^\ell\ (1 + e_i t), \]
where $e_1, e_2,\dots,e_\ell$ are the exponents \cite[Section 3.20]{Hu}
of $W$ and $\ell$ is its rank.
We refer to \cite[Section 2.4]{arm} and \cite[Section 1]{ca} for further discussion of
the importance of the absolute order and related historical remarks.
We will be interested in the combinatorics and topology of $\mbox{Abs}\,(W)$. These have been studied
extensively for the interval $[e,c]:=NC(W,c)$ of $\mbox{Abs}\,(W)$, known as the poset of noncrossing
partitions associated to $W$, where $c\in W$ denotes a Coxeter element. For instance, it was shown
in \cite{cbw} that $NC(W,c)$ is shellable for every finite Coxeter group $W$. In particular,
$NC(W,c)$ is Cohen-Macaulay over $\mathbb{Z}$ and the order complex of $NC(W,c)\smallsetminus\{e,c\}$ has the
homotopy type of a wedge of spheres.
The problem to study the topology of the poset
$\mbox{Abs}\,(W)\smallsetminus\{e\}$ and to decide whether $\mbox{Abs}\,(W)$ is Cohen-Macaulay, or even shellable, was posed by
Reiner \cite[Problem 3.1] {arm2} and Athanasiadis (unpublished); see also \cite[Problem
3.3.7]{mwa}. Computer calculations carried out by Reiner showed that the absolute order is not
Cohen-Macaulay for the group $D_4$. This led Reiner to ask \cite[Problem 3.1] {arm2} whether the order ideal of $\mbox{Abs}\,(W)$
generated by the set of Coxeter elements is Cohen-Macaulay (or shellable) for every finite Coxeter group $W$.
In the case of the symmetric group $S_n$ this ideal coincides with $\mbox{Abs}\,(S_n)$, since every maximal element of $S_n$ is a Coxeter element.
Although it is not known whether $\mbox{Abs}\,(S_n)$ is shellable, the following results were obtained in \cite{ca}.
\begin{theorem}\emph{(\cite[Theorem 1.1]{ca}).}
\label{thca1} The poset $\mbox{Abs}\,(S_n)$ is homotopy Cohen-Macaulay for every $n \ge 1$. In particular, the order
complex of $\mbox{Abs}\,(S_n)\smallsetminus\{e\}$ is homotopy equivalent to a wedge of $(n-2)$-dimensional spheres and
Cohen-Macaulay over $\mathbb{Z}$.
\end{theorem}
\begin{theorem}\emph{(\cite[Theorem 1.2]{ca}).}
\label{thca2}
Let $\bar{P}_n=\mbox{Abs}\,(S_n)\smallsetminus\{e\}$. The reduced Euler characteristic of the order complex $\Delta
(\bar{P}_n)$ satisfies
\[\sum_{n \ge 1} \ (-1)^n \, \tilde{\chi} (\Delta (\bar{P}_n)) \,
\frac{t^n}{n!} \ = \ 1 - C(t) \exp \left\{ -2t \, C(t) \right\},\]
where $C(t) = \frac{1}{2t} \, (1 - \sqrt{1-4t})$ is the ordinary generating function for the
Catalan numbers.
\end{theorem}
In the present paper we focus on the hyperoctahedral group $B_n$.
We denote by $\mathcal{J}_n$ the order ideal of $\mbox{Abs}\,(B_n)$ generated by the Coxeter elements of $B_n$
and by $\bar{\mathcal{J}}_n$ its proper part $\mathcal{J}_n\smallsetminus \{e\}$. Contrary to the case of the symmetric
group, not every maximal element of $\mbox{Abs}\,(B_n)$ is a Coxeter element.
Our main results are as follows.
\begin{theorem}
\label{th3}
The poset $\mathcal{J}_n$ is homotopy Cohen-Macaulay for every $n\geq 2$.
\end{theorem}
\begin{theorem}
\label{th3b}
The reduced Euler
characteristic of the order complex $\Delta(\bar{\mathcal{J}}_n)$ satisfies
\[\sum_{n \geq 2} (-1)^n \tilde{\chi} (\Delta (\bar{\mathcal{J}}_n))
\frac{t^n}{n!} = 1 - \sqrt{C(2t)} \exp \left\{ -2t C(2t) \right\}\left(1+\sum_{n \geq 1} 2^{n-1}{2n-1\choose n} \frac{t^n}{n}\right),\]
where $C(t) = \frac{1}{2t} \, (1 - \sqrt{1-4t})$ is the ordinary generating function for the
Catalan numbers.
\end{theorem}
The maximal (with respect to inclusion) intervals in $\mbox{Abs}\,(B_n)$ include the posets $NC^B(n)$ of
noncrossing partitions of type $B$, introduced and studied by Reiner \cite{R}, and
$NC^B(p,q)$ of annular noncrossing partitions, studied recently by Krattenthaler \cite{krat} and by Nica and Oancea \cite{N0}.
We have the following result concerning the
intervals of $\mbox{Abs}\,(B_n)$.
\begin{theorem}
\label{th2}
Every interval of $\mbox{Abs}\,(B_n)$ is shellable.
\end{theorem}
Furthermore, we consider the absolute order on the group $D_n$ and give an example of a maximal
element $x$ of $\mbox{Abs}\,(D_4)$ for which the interval $[e,x]$ is not Cohen-Macaulay over any field
(Remark \ref{exd4}). This is in accordance with Reiner's computations, showing that $\mbox{Abs}\,(D_4)$ is
not Cohen-Macaulay and answers in the negative a question raised by Athanasiadis (personal
communication), asking whether all intervals in the absolute order on Coxeter groups are shellable.
Moreover, it shows that $\mbox{Abs}\,(D_n)$ is not Cohen-Macaulay over any field for every $n\geq 4$.
It is an open problem to decide whether $\mbox{Abs}\,(B_n)$ is Cohen-Macaulay for every $n\geq 2$ and whether
the order ideal of $\mbox{Abs}\,(W)$ generated by the set of
Coxeter elements is Cohen-Macaulay for every Coxeter group $W$ \cite[Problem 3.1]{arm2}.
This paper is organized as follows.
In Section \ref{prepre} we fix notation and terminology related
to partially ordered sets and simplicial complexes and discuss the absolute order on
the classical finite reflection groups.
In Section \ref{elel} we prove Theorem \ref{th2} by showing that every closed interval of
$\mbox{Abs}\,(B_n)$ admits an EL-labeling. Theorems \ref{th3} and \ref{th3b} are proved in Section \ref{cmcm}. Our method to establish
homotopy Cohen-Macaulayness is different from that of \cite{ca}. It is based on a poset fiber
theorem due to Quillen \cite[Corollary 9.7]{Q}.
The same method gives an alternative proof of Theorem \ref{thca1},
which is also included in Section \ref{cmcm}. In Section \ref{latlat} we characterize the closed intervals in $\mbox{Abs}\,(B_n)$ and $\mbox{Abs}\,(D_n)$
which are lattices. In Section \ref{spespe} we study a special case of
such an interval, namely the maximal interval $[e,x]$ of $\mbox{Abs}\,(B_n)$, where $x=t_1t_2\cdots t_n$
and each $t_i$ is a balanced reflection. Finally, in Section \ref{telostelos}
we compute the zeta polynomial, cardinality and M\"obius function of the intervals of $\mbox{Abs}\,(B_n)$
which are lattices. These computations are based on results of Goulden, Nica and Oancea \cite{N}
concerning enumerative properties of the poset $NC^B(n-1,1)$.
\section{Preliminaries}
\label{prepre}
\subsection{Partial orders and simplicial complexes}
Let $(P,\leq)$ be a finite partially ordered set (poset for short) and $x,y\in P$. We say that $y$
\emph{covers} $x$, and write $x\to y$, if $x<y$ and there is
no $z\in P$ such that $x< z< y$. The poset $P$ is called \emph{bounded} if there exist elements
$\hat{0}$ and $\hat{1}$ such that $\hat{0}\leq x\leq \hat{1}$ for every $x\in P$.
The elements of $P$ which cover $\hat{0}$ are called \emph{atoms}.
A subset $C$ of a poset $P$
is called a \emph{chain} if any two elements of $C$ are comparable in $P$. The length of a (finite)
chain $C$ is equal to
$|C|-1$. We say that $P$ is \emph{graded} if all maximal chains of $P$ have the same length.
In that case, the common length of all maximal chains of $P$ is called \emph{rank}. Moreover, assuming $P$ has a $\hat{0}$ element,
there exists a unique function $\rho:P\to \mathbb{N}$, called the \emph{rank function} of $P$, such that
\[\rho(y)=\left\{
\begin{array}{ll}
0 & \mbox{if $y=\hat{0}$}, \\
\rho(x)+1 & \mbox{if $x\to y$}.
\end{array}
\right.\]
We say that $x$ has \emph{rank} $i$ if
$\rho(x)=i$.
For $x\leq y$ in $P$ we denote by $[x,y]$ the closed interval $\{z \in P : x \leq z \leq y\}$ of $P$, endowed with the partial order induced from $P$.
If $S$ is a subset of $P$, then the \emph{order ideal} of $P$ generated by
$S$ is the subposet $\langle S\rangle$ of $P$ consisting of all $x\in P$ for which $x\preceq y$ holds for some $y\in S$.
We will write $\langle y_1,y_2,\dots,y_m \rangle$ for the order ideal of $P$ generated by the set
$\{y_1,y_2,\dots,y_m\}$.
Given two posets $(P,\leq_P)$ and $(Q,\leq_Q)$, a map $f:P\to Q$ is called a
\emph{poset map} if it is order preserving, i.e. $x\leq_Py$ implies $f(x)\leq_Qf(y)$ for all
$x,y\in P$. If, in addition, $f$ is a bijection with order preserving inverse, then $f$ is said to be a
\emph{poset isomorphism}.
The posets $P$
and $Q$ are said to be \emph{isomorphic}, and we write $P\cong Q$, if there
exists a poset isomorphism $f: P \to Q$. Assuming that $P$ and $Q$ are graded,
the map $f:P\to Q$ is called \emph{rank-preserving} if for every $x\in P$,
the rank of $f(x)$ in $Q$ is equal to the rank of $x$ in $P$.
The \emph{direct product} of
$P$ and $Q$ is the poset $P\times Q$ on the set $\{(x,y):x\in P,\, \ y\in Q\}$
for which $(x,y)\leq (x',y')$ holds in $P\times Q$ if $x\leq_P x'$ and $y\leq_Q y'$.
The \emph{dual} of $P$ is the poset $P^*$ defined on the same ground
set as $P$ by letting $x\leq y$ in $P^*$ if and only if $y\leq x$ in $P$. The poset $P$ is called \emph{self-dual} if $P$ and $P^*$ are isomorphic
and \emph{locally self-dual} if every closed interval of $P$ is self-dual.
For more information on partially ordered sets we refer the reader to \cite[Chapter 3]{St}.
We recall the notion of EL-shellability, defined by Bj\"orner \cite{bjo1}.
Assume that $P$ is bounded and graded and let
$C(P)=\{(a,b)\in P\times P:\ a\to b\}$ be the set of covering relations of $P$.
An \emph{edge-labeling} of $P$ is a map $\lambda:C(P)\to \Lambda$, where $\Lambda$ is some poset.
Let $[x,y]$ be a closed interval of $P$ of rank $n$.
To each maximal chain $c:\,x\to x_1\to\cdots\to x_{n-1}\to y$ of $[x, y]$ we associate the sequence
$\lambda(c)=(\lambda(x,x_1),\lambda(x_1,x_2),\dots, \lambda(x_{n-1},y)\,)$.
We say that $c$ is \emph{strictly increasing}
if the sequence $\lambda(c)$ is strictly increasing in the order of $\Lambda$.
The maximal chains of $[x, y]$ can be totally ordered by using the lexicographic order on the
corresponding sequences. An \emph{edge-lexicographic labeling (EL-
labeling)} of $P$ is an edge labeling such that in each closed interval $[x, y]$
of $P$ there is a unique strictly increasing maximal chain and this chain lexicographically precedes
all other maximal chains of $[x, y]$. The poset $P$ is called \emph{EL-shellable} if it admits an EL-labeling.
A finite poset $P$ of rank $d$ with a minimum element is called \emph{strongly constructible} \cite{ca}
if it is bounded and pure shellable, or it can be
written as a union $P = I_1 \cup I_2$ of two strongly
constructible proper ideals $I_1, I_2$ of rank $n$, such that $I_1
\cap I_2$ is strongly constructible of rank at least $n-1$.
Let $V$ be a nonempty finite set.
An \emph{abstract simplicial complex} $\Delta$ on the vertex set $V$ is a collection
of subsets of $V$ such that $\{v\} \in\Delta$ for every $v\in V$ and such that $G \in \Delta$ and $F\subseteq G$ imply $F\in\Delta$.
The elements of $V$ and $\Delta$ are called \emph{vertices} and \emph{faces} of $\Delta$, respectively.
The maximal faces are called \emph{facets}.
The dimension of a face $F\in \Delta$ is equal to $|F|-1$ and is denoted by $\dim F$.
The \emph{dimension} of $\Delta$ is defined as the maximum dimension of a face of $\Delta$ and is denoted by $\dim\Delta$.
If all facets of $\Delta$ have the same dimension, then $\Delta$ is said to be \emph{pure}.
The \emph{link} of a face $F$ of a simplicial complex
$\Delta$ is defined as $\mbox{link}_{\Delta}(F)=\{G\smallsetminus F:\,G\in\Delta,F\subseteq G\}$.
All topological properties of an abstract simplicial complex $\Delta$ we mention will refer to those of its geometric
realization $\|\Delta\|$. The complex $\Delta$ is said to be
\emph{homotopy Cohen-Macaulay} if for all $F\in\Delta$ the link of $F$ is topologically $(\dim
\mbox{link}_\Delta$$(F)-1)$-connected.
For a facet $G$ of a simplicial complex $\Delta$, we denote by $\bar{G}$ the Boolean interval $[\varnothing,G]$.
A pure $d$-dimensional simplicial complex $\Delta$ is
\emph{shellable} if there exists a total ordering $G_1, G_2,\dots,G_m$ of
the set of facets of $\Delta$ such that for all $1 < i \le m$, the
intersection of $\bar{G}_1\cup \bar{G}_2 \cup \, \cdots \, \cup \bar{G}_{i-1}$ with $\bar{G}_i$ is pure
of dimension $d-1$.
For a $d$-dimensional simplicial complex we have the following implications:
pure shellable $\Rightarrow$ homotopy Cohen-Macaulay $\Rightarrow$
homotopy equivalent to a wedge of $d$-dimensional spheres. For background concerning the topology of simplicial complexes we refer to
\cite{bjo2} and \cite{mwa}.
To every poset $P$ we associate an abstract simplicial complex $\Delta(P)$, called the \emph{order
complex} of $P$. The vertices of $\Delta(P)$ are the elements of $P$ and its faces are the chains
of $P$. If $P$ is graded of rank $n$, then $\Delta(P)$ is pure of dimension $n$. All topological
properties of a poset $P$ we mention will refer to those of the geometric realization of
$\Delta(P)$. We say that a poset $P$ is \emph{shellable} if its order complex $\Delta(P)$ is
shellable and recall that every EL-shellable poset is shellable \cite[Theorem 2.3]{bjo1}.
We also recall the following lemmas.
\begin{lemma}
\label{strcon}
Let $P$ and $Q$ be finite posets, each with a minimum element.
\begin{enumerate}[leftmargin=*, itemsep=5pt]
\item[\emph{(i)}] \emph{\cite[Lemma 3.7]{ca}}
If $P$ and $Q$ are strongly constructible, then so is
the direct product $P\times Q$.
\item [\emph{(ii)}] \emph{\cite[Lemma 3.8]{ca}} If $P$ is the union
of strongly constructible ideals $I_1, I_2,\dots,I_k$ of $P$ of rank $n$ and the
intersection of any two or more of these ideals is strongly constructible of rank $n$ or
$n-1$, then $P$ is also strongly constructible.
\end{enumerate}
\end{lemma}
\begin{lemma}
Every strongly constructible poset is homotopy Cohen-Macaulay.
\end{lemma}
\begin{proof}
It follows from \cite[Proposition 3.6]{ca} and \cite[Corollary 3.3]{ca}.
\end{proof}
\begin{lemma}
\label{tomes}
Let $P$ and $Q$ be finite posets, each with a minimum element.
\begin{enumerate}[leftmargin=*, itemsep=5pt]
\item[\emph{(i)}] If $P$ and $Q$ are homotopy Cohen-Macaulay, then so is
the direct product $P\times Q$.
\item [\emph{(ii)}] If $P$ is the union
of homotopy Cohen-Macaulay ideals $I_1, I_2,\dots,I_k$ of $P$ of rank $n$ and the
intersection of any two or more of these ideals is homotopy Cohen-Macaulay of rank $n$ or
$n-1$, then $P$ is also homotopy Cohen-Macaulay.
\end{enumerate}
\end{lemma}
\begin{proof}
The first part follows from \cite[Corollary 3.8]{bww2}. The proof of
the second part is similar to that of \cite[Lemma 3.4]{ca}.
\end{proof}
\subsection{The absolute length and absolute order}
\label{abs}
Let $W$ be a finite Coxeter group and let $\mathcal{T}$ denote the set of all reflections in $W$.
Given $w \in W$, the \emph{absolute length} of $w$ is defined as the smallest integer $k$ such that $w$ can be
written as a product of $k$ elements of $\mathcal{T}$; it is denoted by $\ell_{\mathcal{T}} (w)$.
The \emph{absolute order} $\mbox{Abs}\,(W)$ is the partial order $\preceq$ on $W$ defined by
\[ u \preceq v \ \ \ \mbox{if and only if} \ \ \ \ell_{\mathcal{T}}(u) \, + \,
\ell_{\mathcal{T}}(u^{-1} v) \, = \, \ell_{\mathcal{T}} (v) \]
for $u, v \in W$. Equivalently, $\preceq$ is the partial order on $W$ with covering relations $w
\to wt$, where $w \in W$ and $t \in\mathcal{T}$ are such that $\ell_{\mathcal{T}} (w) <
\ell_{\mathcal{T}} (wt)$. In that case we write $w\stackrel{t}{\to}wt$. The poset $\mbox{Abs}\,(W)$ is
graded with rank function $\ell_{\mathcal{T}}$.
Every closed interval in $W$ is isomorphic to one which contains
the identity. Specifically, we have the following lemma (see also \cite[Lemma 3.7]{cbw}).
\begin{lemma}
\label{lemarm}
Let $u,v\in W$ with $u\preceq v$.
The map $\phi: [u,v] \to [e, u^{-1}v]$ defined by $\phi(w)= u^{-1}w$ is a poset isomorphism.
\end{lemma}
\begin{proof}
It follows from \cite[Lemma 2.5.4]{arm} by an argument similar to that in the proof of \cite[Proposition 2.6.11]{arm}.
\end{proof}
For more information on the absolute order on $W$ we refer the reader to \cite[Section 2.4]{arm}.
\subsection*{The absolute order on $S_n$}
We view the group $S_n$ as the group of permutations of the set $\{1,2,\dots, n\}$. The set
$\mathcal{T}$ of reflections of $S_n$ is equal to the set of all transpositions $(i\,j)$, where
$1\leq i<j\leq n$. The length $\ell_{\mathcal{T}}(w)$ of $w\in S_n$ is equal to $n-\gamma(w)$,
where $\gamma(w)$ denotes the number of cycles in the cycle decomposition of $w$. Given a cycle $c =
(i_1\, i_2\, \cdots\, i_r)$ in $S_n$ and indices $1\leq j_1<j_2<\cdots<j_s\leq r$, we say that the
cycle $(i_{j_1}\,i_{j_2}\,\cdots\,i_{j_s})$ can be obtained from $c$ by deleting elements.
Given two disjoint cycles $a, b$ in $S_n$ each of which can be obtained from $c$ by deleting
elements, we say that $a$ and $b$ are noncrossing with respect to $c$ if there does not exist a
cycle $(i\, j\, k\, l)$ of length four which can be obtained from $c$ by deleting elements, such
that $i, k$ are elements of $a$ and $j, l$ are elements of $b$. For instance, if $n = 9$ and $c =
(3\, 5\, 1\, 9\, 2\, 6\, 4)$ then the cycles $(3\, 6\, 4)$ and $(5\, 9\, 2)$ are noncrossing with
respect to $c$ but $(3\, 2\, 4)$ and $(5\, 9\, 6)$ are not. It can be verified \cite[Section 2]{Br}
that for $u, v\in S_n$ we have $u \preceq v$ if and only if
\begin{itemize}
\item every cycle in the cycle decomposition for $u$ can be obtained from some cycle in the cycle
decomposition for $v$ by deleting elements and
\item any two cycles of $u$ which can be obtained
from the same cycle $c$ of $v$ by deleting elements are noncrossing with respect to $c$.
\end{itemize}
Clearly, the maximal elements of $\mbox{Abs}\,(S_n)$ are precisely the $n$-cycles, which are the Coxeter elements of $S_n$.
Figure \ref{s3} illustrates the Hasse diagram of the poset $\mbox{Abs}\,(S_3)$.
\begin{figure}[h]
\begin{center}
\includegraphics[width=2.7in]{s3}
\end{center}
\caption{}
\label{s3}
\end{figure}
\subsection*{The absolute order on $B_n$}
\label{relationbn}
We view the hyperoctahedral group $B_n$ as the group of permutations
$w$ of the set $\{\pm1,\pm2,\dots,\pm n\}$ satisfying $w(-i)=-w(i)$
for $1\leq i\leq n$.
Following \cite{brwtt}, the permutation which has cycle form
$(a_1\,a_2\,\cdots\,a_k)(-a_1\,-a_2\,\cdots\,-a_k)$ is denoted by
$( \! ( a_1,a_2,\dots, a_k) \! )$ and is called a \emph{paired} $k$-cycle,
while the cycle $(a_1\,a_2\,\cdots\, a_k\,-a_1\,-a_2\,\cdots\,-a_k)$
is denoted by $[a_1,a_2,\dots,a_k]$ and is called a \emph{balanced} $k$-cycle.
Every element $w \in B_n$ can be written as a product of disjoint paired or balanced cycles, called cycles of $w$.
With this notation, the set $\mathcal{T}$ of reflections of $B_n$ is equal to the union
\begin{equation}
\label{T_B}
\{[i]:\,1\leq i\leq n\}\,\cup\,\{( \! ( i,j) \! ),( \! ( i,-j) \! ):\,1\leq i<j\leq n\}.
\end{equation}
The length $\ell_{\mathcal{T}}(w)$ of $w\in B_n$ is equal to $n-\gamma(w)$, where $\gamma(w)$
denotes the number of paired cycles in the cycle decomposition of $w$. An element $w\in B_n$ is
maximal in $\mbox{Abs}\,(B_n)$ if and only if it can be written as a product of disjoint balanced cycles
whose lengths sum to $n$. The Coxeter elements of $B_n$ are precisely the balanced $n$-cycles.
The covering relations $w\stackrel{t}{\to}wt$ of $\mbox{Abs}\,(B_n)$, when $w$ and
$t$ are non-disjoint cycles, can be described as follows: For $1\leq i<j\leq m\leq n$, we have:
\
\begin{enumerate
\item[(a)]\label{a1}$( \! ( a_1,\dots, a_{i-1},a_{i+1},\dots,a_m) \! )\stackrel{( \! ( a_{i-1},a_i) \! )}{\longrightarrow}( \! ( a_1,\dots,a_m) \! )$
\item[(b)]\label{a2}$( \! ( a_1,\dots,a_m) \! )\stackrel{[a_i]}{\longrightarrow}[a_1,\dots,a_{i-1},a_i,-a_{i+1},\dots,-a_m]$
\item[(c)]\label{a3}$( \! ( a_1,\dots,a_m) \! )\stackrel{( \! ( a_i,-a_j) \! )}{\longrightarrow}[a_1,\dots,a_i,-a_{j+1},\dots,-a_m][a_{i+1},\dots,a_j]$
\item[(d)]\label{a4}$[a_1,\dots, a_{i-1},a_{i+1},\dots,a_m]\stackrel{( \! ( a_{i-1},a_i) \! )}{\longrightarrow}[a_1,\dots,a_m]$
\item[(e)]\label{a5}$[a_1,\dots, a_j]( \! ( a_{j+1},\dots,a_m) \! )\stackrel{( \! ( a_j,a_m) \! )}{\longrightarrow}[a_1,\dots,a_m]$
\item[(f)]\label{a6}$( \! ( a_1,\dots, a_j) \! )( \! ( a_{j+1},\dots,a_m) \! )\stackrel{( \! ( a_j,a_m) \! )}{\longrightarrow}( \! ( a_1,\dots,a_m) \! )$
\end{enumerate}
\
\noindent where $a_1,\dots,a_m$ are elements of $\{\pm1,\dots,\pm n\}$ with pairwise distinct
absolute values. Figure \ref{b2} illustrates the Hasse diagram of the poset $\mbox{Abs}\,(B_2)$.
\begin{remark}
\label{anal}
Let $w=bp$ be an element in $B_n$, where $b$ (respectively, $p$) is the product of all balanced (respectively, paired) cycles of $w$.
The covering relations of $\mbox{Abs}\,(B_n)$ imply the poset isomorphism $[e,w]\cong [e,b]\times[e,p]$.
Moreover, if $p=p_1\cdots p_k$ is written as a product of disjoint paired cycles, then
\[[e,w]\cong [e,b]\times[e,p_1]\times\cdots\times[e,p_k].\]
\end{remark}
\begin{figure}[h]
\begin{center}
\includegraphics[width=2.7in]{b2}
\end{center}
\caption{}
\label{b2}
\end{figure}
\subsection*{The absolute order on $D_n$}
\label{dn}
The Coxeter group $D_n$ is the subgroup of index two of the group $B_n$,
generated by the set of reflections
\begin{equation}
\label{T_D}
\{( \! ( i,j) \! ),( \! ( i,-j) \! ):1\leq i<j\leq n\}
\end{equation}
(these are all reflections in $D_n$). An element $w\in B_n$ belongs to $D_n$ if and only if $w$ has
an even number of balanced cycles in its cycle decomposition. The absolute length on $D_n$
is the restriction of the absolute length of $B_n$ on the set $D_n$ and hence $\mbox{Abs}\,(D_n)$ is a
subposet of $\mbox{Abs}\,(B_n)$. Every
Coxeter element of $D_n$ has the form $[a_1,a_2,\dots,a_{n-1}][a_n]$, where
$a_1,\dots,a_n$ are elements of $\{\pm1,\dots,\pm n\}$ with pairwise distinct absolute values.
\subsection*{Projections}
We recall that $\mathcal{J}_n$ denotes the order ideal of $\mbox{Abs}\,(B_n)$ generated by the Coxeter elements of $B_n$.
Let $P_n$ be $\mbox{Abs}\,(S_n)$ or $\mathcal{J}_n$ for some $n\geq 2$.
For $i\in\{1,2,\dots,n\}$ we define a map $\pi_i:P_n\to P_n$ by letting $\pi_i(w)$ be the permutation obtained when
$\pm i$ is deleted from the cycle decomposition of $w$. For example, if $n=i=5$ and
$w=[1,-5,2]( \! ( 3,-4) \! )\in \mathcal{J}_5$, then $\pi_i(w)=[1,2]( \! ( 3,-4) \! )$.
\begin{lemma}
\label{proj1}The following hold for the map $\pi_i:P_n\to P_n$.
\begin{enumerate}[leftmargin=*, itemsep=5pt]
\item [\emph{(i)}] $\pi_i(w)\preceq w$ for every $w\in P_n$.
\item [\emph{(ii)}] $\pi_i$ is a poset map.
\end{enumerate}
\end{lemma}
\begin{proof}
Let $w\in P_n$. If $w(i)=i$, then clearly $\pi_i(w)=w$. Suppose that $w(i)\neq i$. Then it follows from our description of $\mbox{Abs}\,(S_n)$ and from
the covering relations of types (a) and (d) of $\mbox{Abs}\,(B_n)$, that $\pi_i(w)$ is covered by $w$. Hence $\pi_i(w)\preceq w$.
This proves (i). To prove (ii), it suffices to show that for every covering relation $u\to v$ in $P_n$ we have either $\pi_i(u)=\pi_i(v)$ or $\pi_i(u)\to\pi_i(v)$.
Again, this follows from our discussion of $\mbox{Abs}\,(S_n)$ and from our list of covering relations of $\mbox{Abs}\,(B_n)$.
\end{proof}
\begin{lemma}
\label{antenatel}
Let $P_n$ stand for either $\mbox{Abs}\,(S_n)$ for every $n\geq 1$, or $\mathcal{J}_n$ for every $n\geq 2$.
Let also $w\in P_n$ and $u\in P_{n-1}$ be such that $\pi_n(w)\preceq u$.
Then there exists an element $v\in P_n$ which covers $u$ and satisfies $\pi_n(v)=u$ and $w\preceq v$.
\end{lemma}
\begin{proof}We may assume that $w$ does not fix $n$, since otherwise the result is trivial.
Suppose that $\pi_n(w)=w_1\cdots w_l$ and $u=u_1\cdots u_r$ are written as products of disjoint cycles in $P_{n-1}$.
\smallskip
\noindent{\bf~Case 1:} $P_n=\mbox{Abs}\,(S_n)$ for $n\geq 1$.
Then there is an index $i\in\{1,2,\dots, l\}$ such that
$w$ is obtained from $\pi_n(w)$ by inserting $n$ in the cycle $w_i$.
Let $y$ be the cycle of $w$ containing $n$, so that $\pi_n(y)=w_i$. From the description of the absolute order on $S_n$ given in this section, it follows that $w_i\preceq u_j$ for some
$j\in\{1,2,\dots,r\}$.
We may insert $n$ in the cycle $u_j$ so that the resulting cycle $v_j$ satisfies $y\preceq v_j$.
Let $v$ be the element of $S_n$ obtained by
replacing $u_j$ in the cycle decomposition of $u$ by $v_j$. Then $u$ is covered by $v,\,\pi_n(v)=u$ and $w\preceq v$.
\smallskip
\noindent{\bf~Case 2:} $P_n=\mathcal{J}_n$ for $n\geq 2$.
The result follows by a simple modification of the argument in the previous case, if $[n]$ is not a cycle of $w$.
Assume the contrary, so that $w=\pi_n(w)[n]$ and all cycles of $\pi_n(w)$ are paired.
If $u$ has no balanced cycle, then $w\preceq u[n]\in \mathcal{J}_n$ and hence $v=u[n]$ has the desired properties.
Suppose that $u$ has a balanced cycle in its cycle decomposition, say $b=[a_1,\dots,a_k]$.
We denote by $p$ the product of all paired cycles of $u$, so that $u=bp$.
If $\pi_n(w)\preceq p$, then the choice $v=[a_1,\dots,a_k,n]p$ works.
Otherwise, we may assume that there is an index $m\in \{1,2,\dots,l\}$ such that $w_1\cdots w_m\preceq b$ and $w_i$ and $b$ are disjoint for every $i>m$.
From the covering relations of $\mbox{Abs}\,(B_n)$ of types (a), (b) and (f) it follows that there is a paired cycle $c$ which is covered by $b$ and satisfies $w_1\cdots w_m\preceq c$. Thus $\pi_n(w)\preceq cp\preceq u$.
More specifically, $c$ has the form $( \! ( a_1,\dots, a_i,-a_{i+1},\dots,-a_k) \! )$ for some $i\in \{2,\dots, k\}$.
We set $v=[a_1,\dots,a_i,n,a_{i+1},\dots,a_l]p$. Then $v$ covers $u$ and $w\preceq cp[n]\preceq v$. This concludes the proof of the lemma.
\end{proof}
\section{Shellability}
\label{elel}
In this section we prove Theorem \ref{th2} by showing that every closed interval of $\mbox{Abs}\,(B_n)$
admits an EL-labeling.
Let $C(B_n)$
be the set of covering relations of $\mbox{Abs}\,(B_n)$
and $(a,b)\in C(B_n)$. Then $a^{-1}b$ is a reflection of $B_n$, thus
either $a^{-1}b=[i]$ for some $i\in\{1,2,\dots,n\}$, or there exist $i,j\in\{1,2,\dots,n\}$, with $i<j$, such that
$a^{-1}b=( \! ( i,j) \! )$ or $a^{-1}b=( \! ( i,-j) \! )$.
We define a map $\lambda:C(B_n)\to \{1,2,\dots,n\}$ as follows:
\[\lambda(a,b)=\left\{
\begin{array}{ll}
i & \mbox{if $a^{-1}b=[i]$}, \\
j & \mbox{if $a^{-1}b=( \! ( i,j) \! )$ or $( \! ( i,-j) \! )$}.
\end{array}
\right.\]
A similar labeling was used by Biane \cite{Bi} in order to study the maximal chains
of the poset $NC^B(n)$ of noncrossing $B_n$-partitions.
Figure \ref{figpxel} illustrates the Hasse diagram of the interval
$\left[e,[3,-4]( \! ( 1,2) \! )\right]$, together with the corresponding labels.
\
\begin{figure}[h]
\begin{center}
\includegraphics[width=4.5in]{pxel}
\end{center}
\caption{}
\label{figpxel}
\end{figure}
\begin{propo}
\label{el}
Let $u,v\in B_n$ with $u\preceq v$. Then, the restriction of the map $\lambda$ to the interval $[u,v]$ is an EL-labeling.
\end{propo}
\begin{proof}
Let $u,v\in B_n$ with $u\preceq v$. We consider the poset isomorphism $\phi:[u,v]\to[e,u^{-1}v]$ from Lemma \ref{lemarm}.
Let $(a,b)\in C([u,v])$.
Then we have $\phi(a)^{-1}\phi(b)=(u^{-1}a)^{-1}u^{-1}b=a^{-1}uu^{-1}b=a^{-1}b,$
which implies that $\lambda(a,b)=\lambda(\phi(a),\phi(b))$.
Thus, it suffices to show that $\lambda|_{[e,w]}$ is an EL-labeling for the interval $[e,w]$, where $w=u^{-1}v$.
Let $b_1b_2\cdots b_k\, p_1p_2\cdots p_l$ be the cycle decomposition of $w$, where
$b_i=[b_i^1,\dots,b_i^{k_i}]$ for $i\leq k$ and $p_j=( \! ( p_j^1,\dots,p_j^{l_j}) \! )$, with
$p_j^1=\min\{|p_j^m|:1\leq m\leq l_j\}$ for $j\leq l$.
We consider the sequence of positive integers obtained by placing the numbers $|b_i^h|$ and
$|p_j^m|$, for $i,j,h\geq 1$ and $m>1$, in increasing order. There are $r=\ell_{\mathcal{T}}(w)$ such integers.
To simplify the notation, we denote by $c(w)=(c_1,c_2,\dots,c_r)$ this sequence and say that
$c_{\mu}$ ($\mu=1,2,\dots,r$) belongs to a balanced (respectively, paired) cycle if it is equal to
some $|b_i^h|$ (respectively, $|p_j^m|$). Clearly, we have
\begin{equation}
\label{al}
c_1<c_2<\dots<c_r
\end{equation}
and $\lambda(a,b)\in\{c_1,c_2,\dots,c_r\}$ for all $(a,b)\in C([e,w])$.
To the sequence (\ref{al})
corresponds a unique maximal chain
\[\mathcal{C}_w:\ w_0=e\stackrel{c_1}{\to} w_1\stackrel{c_2}{\to} w_2\stackrel{c_3}{\to}\dots\stackrel{c_r}{\to} w_r=w,\]
which can be constructed inductively as follows (here, the integer $\kappa$ in
$a\stackrel{\kappa}{\to}b$ denotes the label $\lambda(a,b)$). If $c_1$ belongs to a balanced cycle,
then $w_1=[c_1]$. Otherwise, $c_1$ belongs to some $p_i$, say $p_1$, and we set $w_1$ to be either
$( \! ( p_1^1, c_1) \! )$ or $( \! ( p_1^1,-c_1) \! )$, so that $w_1\preceq p_1$ holds. Note that in
both cases we have $\lambda(e,w_1)=c_1$ and $\lambda(e,w_1)<\lambda(e,w)$ for any other atom
$t\in[e,w]$. Indeed, suppose that there is an atom $t\neq w_1$ such that $\lambda(e,t)=c_1$. We
assume first that $c_1$ belongs to a balanced cycle, so $w_1=[c_1]$. Then $t$ is a reflection of
the form $( \! ( c_0,\pm c_1) \! )$, where $c_0<c_1$ and, therefore, $c_0$ belongs to some paired
cycle of $w$ (if not then $c_1$ would not be minimum). However from the covering relations of $\mbox{Abs}\,(B_n)$ written
at the end of Section \ref{relationbn} it follows that $( \! ( c_0,\pm c_1) \! )\not\preceq w$, thus
$( \! ( c_0,\pm c_1) \! )\not\in[e,w]$, a contradiction.
Therefore $c_1$ belongs to a paired cycle of $w$, say $p_1$, and $w_1,t$ are both paired reflections.
Without loss of generality, let $w_1=( \! ( p_1^1,c_1) \! )$ and $t=( \! ( c_0,c_1) \! )$, for
some $c_0<c_1$. By the first covering relation written at the end of Section \ref{relationbn} and
the definition of $\lambda$, it follows that $c_0=p_1^1$, thus $w_1=t$, again a contradiction.
Suppose now that we have uniquely defined the elements $w_1,w_2,\dots, w_j$, so that for every
$i=1,2,\dots,j$ we have $w_{i-1}\to w_i$ with $\lambda(w_{i-1},w_i)=c_i$ and
$\lambda(w_{i-1},w_i)<\lambda(w_{i-1},z)$ for every $z\in[e,w]$ such that $z\neq w_i$ and
$w_{i-1}\to z$. We consider the number $c_{j+1}$ and distinguish two cases.
\smallskip
\noindent{\bf~Case 1:} $c_{j+1}$ belongs to a cycle whose elements have not been used.
In this case, if $c_{j+1}$ belongs to a balanced cycle, then we set $w_{j+1}=w_j[c_{j+1}]$,
while if $c_{j+1}$ belongs to $p_s$ for some $s\in\{1,2,\dots, l\}$, then we set
$w_{j+1}$ to be either $w_j\,( \! ( p_s^1, c_{j+1}) \! )$ or $w_j\,( \! ( p_s^1,-c_{j+1}) \! )$,
so that $w_j^{-1}w_{j+1}\preceq p_s$ holds.
\smallskip
\noindent{\bf~Case 2:} $c_{j+1}$ belongs to a cycle some element of which has been used. Then there
exists an index $i<j+1$ such that $c_i$ belongs to the same cycle as $c_{j+1}$. If $c_i,c_{j+1}$
belong to some $b_s$, then there is a balanced cycle of $w_j$, say $a$, that contains $c_i$. In
this case we set $w_{j+1}$ to be the permutation that we obtain from $w_j$ if we add the number
$c_{j+1}$ in the cycle $a$ in the same order and with the same sign that it appears in $b_s$. We
proceed similarly if $c_i,c_{j+1}$ belong to the same paired cycle.
\smallskip
In both cases we have $\lambda(w_j,w_{j+1})=c_{j+1}$. This follows from the covering relations of $\mbox{Abs}\,(B_n)$ given in
the end of Section \ref{relationbn}. Furthermore, we claim that if $z\in[e,w]$ with $z\neq w_{j+1}$
is such that $w_j\to z$, then $\lambda(w_j,w_{j+1})<\lambda(w_j,z)$.
Indeed, in view of the poset isomorphism $\phi: [u,v] \to [e, u^{-1}v]$ for $u=w_j$ and $v=w$,
this follows from the special case $j=0$ treated earlier. By definition of $\lambda$ and the
construction of $\mathcal{C}_u$, the sequence
\[\left(\lambda(e,w_1),\lambda(w_1, w_2),\dots,\lambda(w_{r-1}, w)\right)\] coincides with $c(w)$.
Moreover, $\mathcal{C}_w$ is the unique maximal chain having this sequence of labels.
This and the fact that the labels of any chain in $[e,w]$ are elements of the set $\{c_1, c_2,\dots,c_r\}$
imply that $\mathcal{C}_w$ is the unique strictly increasing maximal chain.
By what we have already shown, $\mathcal{C}_w$ lexicographically precedes all other maximal chains of $[e,w]$.
Thus $\mathcal{C}_w$ is lexicographically first and the unique strictly increasing chain in $[e,w]$.
Hence $\lambda$ is an EL-labeling for the interval $[e,w]$ and Proposition \ref{el} is proved.
\end{proof}
\begin{example}
\begin{enumerate}[leftmargin=*, itemsep=5pt]
\item [(i)] Let $n= 7$ and $w=[1,-7][3]\lleft2,\,-6,\,-5) \! )\lleft4) \! )\in B_7$.
Then $c(w)=(1,3,5,6,7)$ and
\[\mathcal{C}_w: e\stackrel{1}{\to}[1]\stackrel{3}{\to}[1][3]\stackrel{5}{\to}[1][3]\lleft2,-5) \! )\stackrel{6}{\to}[1][3]\lleft2,-6,-5) \! )\stackrel{7}{\to} w.\]
\item [(ii)] Let $n=4$ and $w=[3,-4]\lleft1,2) \! )$.
Then $c(w)=(2,3,4)$ and \[\mathcal{C}_w:e\stackrel{2}{\to} \lleft1,2) \! )\stackrel{3}{\to} \lleft1,2) \! )[3]\stackrel{4}{\to} w.\]
\end{enumerate}
\end{example}
\noindent{\emph{Proof of Theorem \ref{th2}.}\,
It follows from Proposition \ref{el}, since EL-shellability implies shellability.
\qed
\begin{figure}[h]
\begin{center}
\includegraphics[width=4.32in]{d4b}
\end{center}
\caption{}
\label{d4b}
\end{figure}
\begin{remark}
\label{exd4}
Figure \ref{d4b} illustrates the Hasse diagram of the interval $I=\left[e, u\right]$ of
$\mbox{Abs}\,(D_4)$, where $u= [1][2][3][4]$. Note that the Hasse diagram of the open interval $(e,u)$ is
disconnected and, therefore, $I$ is not Cohen-Macaulay over any field.
Since $\mbox{Abs}\,(D_n)$ contains an interval which is isomorphic to $I$ for any $n\geq 4$, it follows that $\mbox{Abs}\,(D_n)$
is not Cohen-Macaulay over any field for $n\geq 4$ either (see \cite[Corollary 3.1.9]{mwa}).
\end{remark}
\section{Cohen-Macaulayness}
\label{cmcm}
In this section we prove Theorems \ref{th3} and \ref{th3b}. Our method to show that $\mathcal{J}_n$ is homotopy Cohen-Macaulay is based on
the following theorem, due to Quillen
\cite[Corollary 9.7]{Q}; see also \cite[Theorem 5.1]{bww}.
The same method yields a new proof of Theorem \ref{thca1}, which we also include in this section.
\begin{theorem
\label{zup}
Let $P$ and $Q$ be graded posets and let $f:P\to Q$ be a surjective rank-preserving poset map. Assume
that for all $q\in Q$ the fiber $f^{-1}(\langle q\rangle)$ is homotopy Cohen-Macaulay. If
$Q$ is homotopy Cohen-Macaulay, then so is $P$.
\end{theorem}
For other poset fiber theorems of this type, see \cite{bww}.
\smallskip
To prove Theorems \ref{thca1} and \ref{th3}, we need the following.
Let $\{\hat{0},\hat{1}\}$ be a two element chain, with
$\hat{0}<\hat{1}$ and $i\in\{1,2,\dots,n\}$.
We consider the map $\pi_i:P_n\to P_n$ of Section \ref{abs}, where $P_n$ is either $\mbox{Abs}\,(S_n)$ or $\mathcal{J}_n$.
We define the map
\[f_i:P_n\to \pi_i(P_n)\times\{\hat{0},\hat{1}\}\]
by letting
\[f_i(w)=\left\{\begin{array}{ll}
(\pi_i(w),\,\hat{0}), & \mbox{if $w(i)=i$}, \\
(\pi_i(w),\,\hat{1}), & \mbox{if $w(i)\neq i$}
\end{array}
\right.\]
for $w\in P_n$. We first check that $f_i$ is a surjective rank-preserving poset map. Indeed, by definition $f_i$
is rank-preserving. Let $u,v\in P_n$ with $u\preceq v$. Lemma \ref{proj1} (ii) implies that
$\pi_i(u)\preceq \pi_i(v)$. If $u(i)\neq i$, then $v(i)\neq i$ as
well and hence $f_i(u)=(\pi_i(u),\hat{1})\leq(\pi_i(v),\hat{1})=f_i(v)$.
If $u(i)=i$, then $f_i(u)=(\pi_i(u),\hat{0})$
and hence $f_i(u)\leq f_i(v)$. Thus $f_i$ is a poset map.
Moreover, if $w\in \pi_i(P_n)$, then $f_i^{-1}\left(\{(w,\hat{0})\}\right)=\{w\}$
and any permutation obtained
from $w$ by inserting the element $i$ in a cycle of $w$ lies in
$f_i^{-1}\left(\{(w,\hat{1}\right)\})$. Thus $f_i^{-1}\left(\{q\}\right)\neq\varnothing$ for every $q\in \pi_i(P_n)\times\{\hat{0},\hat{1}\}$,
which means that $f_i$ is surjective.
Given a map $f: P \to Q$, we abbreviate by $f^{-1} (q)$
the inverse image $f^{-1} (\{q\})$ of a singleton subset $\{q\}$ of $Q$.
For subsets $U$ and $V$ of $S_n$ (respectively, of $B_n$), we write $U\cdot V=\{uv: u\in U, v\in V\}$.
\
The following lemmas will be used in the proof of Theorem \ref{thca1}.
\smallskip
\begin{lemma}
\label{f-1}
For every $q\in S_{n-1}\times \{\hat{0},\hat{1}\}$ we have $f_n^{-1}\left(\langle q\rangle\right)=\langle f_n^{-1}(q)\rangle$.
\end{lemma}
\begin{proof}
The result is trivial for $q=(u,\hat{0})\in S_{n-1}\times\{\hat{0},\hat{1}\}$, so suppose that $q=(u,\hat{1})$.
Since $f_n$ is a poset map, we have $\langle f_n^{-1}(q)\rangle\subseteq f_n^{-1}\left(\langle q\rangle\right)$.
For the reverse inclusion consider any element $w\in f_n^{-1}\left(\langle q\rangle\right)$. Then $f_n(w)\leq q$ and hence $\pi_n(w)\preceq u$.
Lemma \ref{antenatel} implies that there exists an element $v\in S_n$ which covers $u$ and satisfies $\pi_n(v)=u$ and $w\preceq v$.
We then have $v\in f_n^{-1}(q)$ and hence $w\in \langle f_n^{-1}(q)\rangle$. This proves that $f_n^{-1}\left(\langle q\rangle\right)\subseteq\langle f_n^{-1}(q)\rangle$.
\end{proof}
\smallskip
\begin{lemma}
\label{lsn}
For every $u\in S_{n-1}$, the order ideal
\[M(u)=\langle v\in S_n:\pi_n(v)=u\rangle\] of $\mbox{Abs}\,(S_n)$ is homotopy Cohen-Macaulay of rank $\ell_{\mathcal{T}}(u)+1$.
\end{lemma}
\begin{proof}
Let $u=u_1u_2\cdots u_l$ be written as a product of disjoint cycles in $S_{n-1}$.
Then \[M(u)=\bigcup\limits_{i=1}\limits^lC(u_i)\cdot\langle u_1\cdots\hat{u}_i\cdots u_l\rangle,\]
where $u_1\cdots\hat{u}_i\cdots u_l$ denotes the permutation obtained from $u$ by deleting the
cycle $u_i$ and $C(u_i)$ denotes the order ideal of $\mbox{Abs}\,(S_n)$ generated by the cycles $v$ of $S_n$ which cover $u_i$ and satisfy $\pi_n(v)=u_i$.
Lemma \ref{l4}, proved in the Appendix, implies that $C(u_i)$ is homotopy Cohen-Macaulay of rank $\ell_{\mathcal{T}}(u_i)+1$ for every $i$.
Each of the ideals $C(u_i)\cdot\langle u_1\cdots\hat{u}_i\cdots u_l\rangle$ is isomorphic to a direct product of
homotopy Cohen-Macaulay posets and
hence it is homotopy Cohen-Macaulay, by Lemma \ref{tomes} (i); their rank is equal to $\ell_{\mathcal{T}}(u)+1$.
Moreover, the intersection of any two or more of the ideals $C(u_i)\cdot\langle u_1\cdots\hat{u}_i\cdots u_l\rangle$ is equal to
$\langle u\rangle$, which is homotopy Cohen-Macaulay of rank $\ell_{\mathcal{T}}(u)$.
Thus the result follows from Lemma \ref{tomes} (ii).
\end{proof}
\noindent \emph{Proof of Theorem \ref{thca1}.}\ We proceed by induction on $n$. The result is trivial for
$n\leq 2$. Suppose that the poset $\mbox{Abs}\,(S_{n-1})$ is homotopy Cohen-Macaulay.
Then so is the direct product
$\mbox{Abs}\,(S_{n-1})\times\{\hat{0},\hat{1}\}$ by Lemma \ref{tomes} (i). We consider the map
\[f_n:\mbox{Abs}\,(S_n)\to \mbox{Abs}\,(S_{n-1})\times\{\hat{0},\hat{1}\}.\]
In view of Theorem \ref{zup} and Lemma \ref{f-1}, it suffices to show that for every
$q\in S_{n-1}\times\{\hat{0},\hat{1}\}$ the order ideal
$\langle f_n^{-1}(q)\rangle$ of $\mbox{Abs}\,(S_n)$ is homotopy
Cohen-Macaulay.
This is true in case $q=(u,\hat{0})$ for some $u\in S_{n-1}$, since then
$\langle f_n^{-1}(q)\rangle=\langle u\rangle$ and every interval in $\mbox{Abs}\,(S_n)$ is shellable.
Suppose that
$q=(u,\hat{1})$. Then $\langle f_n^{-1}(q)\rangle= M(u)$, which is homotopy Cohen-Macaulay by Lemma \ref{lsn}.
This completes the induction and the proof of the theorem.\qed
\
We now focus on the hyperoctahedral group.
The proof of Theorem \ref{th3} is based on the following lemmas.
\begin{lemma}
\label{g^{-1}}
For every $q\in \mathcal{J}_{n-1}\times \{\hat{0},\hat{1}\}$ we have $f_n^{-1}\left(\langle q\rangle\right)=\langle f_n^{-1}(q)\rangle$.
\end{lemma}
\begin{proof}
The proof of Lemma \ref{f-1} applies word by word, if one replaces $S_{n-1}$ by the ideal $\mathcal{J}_{n-1}$. We thus omit the details.
\end{proof}
\begin{lemma}
\label{yohoho}
For every $u\in \mathcal{J}_{n-1}$ the order ideal
\[M(u)=\langle v\in \mathcal{J}_n:\,\pi_n(v)=u\rangle\]
of $\mbox{Abs}\,(B_n)$ is homotopy Cohen-Macaulay of rank $\ell_{\mathcal{T}}(u)+1$.
\end{lemma}
\begin{proof}
Let $u=u_1u_2\cdots u_l\in\mathcal{J}_{n-1}$ be written as a product of disjoint cycles.
For $i\in\{1,\dots,l\}$, we denote by $C(u_i)$ the order ideal of $\mathcal{J}_n$ generated by
all cycles $v\in\mathcal{J}_n$ which can be obtained by inserting either $n$ or $-n$ at any place in the cycle $u_i$.
The ideal $C(u_i)$ is graded of rank $\ell_{\mathcal{T}}(u_i)+1$
and homotopy Cohen-Macaulay, by Lemma \ref{l4'} proved in the Appendix.
Let $u_1\cdots \hat{u}_i\cdots u_l$ denote the permutation obtained from $u$ by removing the cycle $u_i$.
Suppose first that $u$ has a balanced cycle in its cycle decomposition. Using Remark \ref{anal}, we find that
\[M(u)=\bigcup\limits_{i=1}\limits^lC(u_i)\cdot\langle u_1\cdots \hat{u}_i\cdots u_l\rangle.\]
Clearly, $M(u)$ is graded of rank $\ell_{\mathcal{T}}(u)+1$.
Each of the ideals $C(u_i)\cdot\langle u_1\cdots\hat{u}_i\cdots u_l\rangle$ is isomorphic to a direct product of homotopy Cohen-Macaulay posets and
hence it is homotopy Cohen-Macaulay, by Lemma \ref{tomes} (i).
Moreover, the intersection of any two or more of these ideals
is equal to $\langle u\rangle$, which is homotopy Cohen-Macaulay of rank $\ell_{\mathcal{T}}(u)$, by Theorem \ref{th2}.
Suppose now that $u$ has no balanced cycle in its cycle decomposition.
Then
\[M(u)=\bigcup\limits_{i=1}\limits^l C(u_i)\cdot\langle u_1\cdots \hat{u}_i\cdots u_l\rangle\cup\langle u[n]\rangle.\]
Again, $M(u)$ is graded of rank $\ell_{\mathcal{T}}(u)+1$,
each of the ideals $C(u_i)\langle u_1\cdots \hat{u}_i\cdots u_l\rangle$ and $\langle u[n]\rangle$ is homotopy Cohen-Macaulay and the intersection of any two or more of
these ideals is equal to $\langle u\rangle$. In either case, the result follows from Lemma \ref{tomes} (ii).
\end{proof}
\noindent\emph{Proof of Theorem \ref{th3}.}\
We proceed induction on $n$. The result is trivial for
$n\leq 2$. Suppose that the poset $\mathcal{J}_{n-1}$ is homotopy Cohen-Macaulay.
Then so is the direct product
$\mathcal{J}_{n-1}\times\{\hat{0},\hat{1}\}$ by Lemma \ref{tomes} (i). We consider the map
\[f_n:\mathcal{J}_n\to\mathcal{J}_{n-1}\times\{\hat{0},\hat{1}\}.\]
In view of Theorem \ref{zup} and Lemma \ref{g^{-1}}, it suffices to show that for every
$q\in\mathcal{J}_{n-1}\times\{\hat{0},\hat{1}\}$ the order ideal
$\langle f_n^{-1}(q)\rangle$ of $\mbox{Abs}\,(B_n)$ is homotopy
Cohen-Macaulay.
This is true in case $q=(u,\hat{0})$ for some $u\in \mathcal{J}_{n-1}$, since then
$\langle f_n^{-1}(q)\rangle=\langle u\rangle$ and every interval in $\mbox{Abs}\,(B_n)$ is shellable by Theorem \ref{th2}.
Suppose that
$q=(u,\hat{1})$. Then $\langle f_n^{-1}(q)\rangle= M(u)$, which is homotopy Cohen-Macaulay by Lemma \ref{lsn}.
This completes the induction and the proof of the theorem.\qed
\
\noindent\emph{Proof of Theorem \ref{th3b}.}\
Let us denote by $\hat{0}$ the minimum element of $\mbox{Abs}\,(B_n)$.
Let $\hat{\mathcal{J}}_n$ be the poset obtained from $\mathcal{J}_n$ by adding
a maximum element $\hat{1}$ and let $\mu_n$ be the
M\"obius function of $\hat{\mathcal{J}}_n$.
From Proposition 3.8.6 of \cite{St} we have that $\tilde{\chi} (\Delta (\bar{\mathcal{J}}_n)) = \mu_n (\hat{0},
\hat{1})$. Since $\mu_n(\hat{0},\hat{1})=-\sum\limits_{x \in \mathcal{J}_n} \ \mu_n
(\hat{0}, x)$, we have
\begin{equation}
\label{eq1}
\tilde{\chi} (\Delta (\bar{\mathcal{J}}_n)) = -\sum\limits_{x \in \mathcal{J}_n} \ \mu_n
(\hat{0}, x).
\end{equation}
\noindent Suppose that $x\in B_n$ is a cycle. It is known \cite{R} that
\[\mu(\hat{0},x)=\left\{\begin{array}{ll}
(-1)^m{2m-1\choose k}, & \mbox{if $x$ is a balanced $m$-cycle}, \\
(-1)^{m-1}C_{m-1}, & \mbox{if $x$ is a paired $m$-cycle},
\end{array}
\right.\]
where $C_m=\frac{1}{m+1}{2m\choose m}$ is the $m$th Catalan number.
We recall (Remark \ref{anal}) that if $x\in\mathcal{J}_n$
has exactly $k+1$ paired cycles, say $p_1,\dots,p_{k+1}$, and one balanced cycle, say $b$, then
$[\hat{0},x]\cong [\hat{0},b]\times[\hat{0},p_1]\times\cdots\times[\hat{0},p_k]$
and hence
\[\mu_n(\hat{0},x)=\mu_n(\hat{0},b)\,\prod\limits_{i=1}\limits^k\, \mu_n(\hat{0},p_i).\]
It follows that
\begin{equation}
\label{eq2}
\mu_n(\hat{0},x)=(-1)^{\ell_{\mathcal{T}}(b)}{2\ell_{\mathcal{T}}(b)-1\choose \ell_{\mathcal{T}}(b)} \prod\limits_{i=1}\limits^k (-1)^{\ell_{\mathcal{T}}(p_i)}C_{\ell_{\mathcal{T}}(p_i)}.
\end{equation}
From (\ref{eq1}), (\ref{eq2}), \cite[Proposition 5.1.1]{St2} and the exponential formula \cite[Corollary 5.1.9]{St2}, we conclude that
\begin{equation}
\label{eqa}
1 - \sum_{n \ge 2} \tilde{\chi} (\Delta (\bar{\mathcal{J}}_n)) \frac{t^n}{n!}
= \left( 1+\sum_{n \geq 1}2^{n-1}\alpha_n\frac{t^n}{n}\right) \exp\left(\sum_{n \geq 1}2^{n-1}\beta_n \frac{t^n}{n}\right),
\end{equation}
where $\alpha_n=(-1)^n{2n-1\choose n}$ is the M\"obius function of a balanced $n$-cycle and $\beta_n=(-1)^{n-1} C_{n-1}$
is the M\"obius function of a paired $n$-cycle.
Thus it suffices to compute $\exp\left(\sum_{n \geq 1}2^{n-1}\beta_n \frac{t^n}{n}\right)$.
From \cite[ Section 5]{ca} we have that
\[\exp\,\sum_{n \ge 1} \beta_n\frac{t^n}{n}=\frac{\sqrt{1+4t}-1}{2t}\,\exp\left(\sqrt{1+4t}-1\right)\]
and hence, replacing $t$ by $2t$,
\[\exp\left(\sum_{n \geq 1} 2^{n-1} \beta_n \frac{t^n}{n}\right)=
\left(\frac{\sqrt{1+8t}-1}{4t}\right)^{1/2}\exp\left(\frac{\sqrt{1+8t}-1}{2}\right).\]
The right-hand side of (\ref{eqa}) can now be written as
\
1-\left(\frac{\sqrt{1+8t}-1}{4t}\right)^{1/2}\exp\left(\frac{\sqrt{1+8t}-1}{2}\right)\left(1+\sum_{n \geq 1}2^{n-1}\alpha_n\frac{t^n}{n}\right).\]
\noindent The result follows by switching $t$ to $-t$.
\qed
\
\begin{remark}
Theorem \ref{th3} can also be proved using the notion of strong constructibility,
introduced in \cite{ca}. The details will appear in \cite{myr}.
\end{remark}
\section{Intervals with the lattice property}
\label{latlat}
Let $W$ be a finite Coxeter group and $c\in W$ be a Coxeter element. It is known \cite{Be,brwtt,brwtt0} that the interval $[e,c]$
in $\mbox{Abs}\,(W)$ is a lattice.
In this section we characterize the intervals in $\mbox{Abs}\,(B_n)$ and $\mbox{Abs}\,(D_n)$ which are lattices (Theorems \ref{th5} and \ref{th6}, respectively).
As we explain in the sequel, some partial results in this direction were obtained in \cite{Be, brwtt, brwtt0, N, R}.
To each $w\in B_n$ we associate the integer partition $\mu(w)$ whose parts are the absolute lengths of all balanced cycles of $w$,
arranged in decreasing order.
For example, if $n=8$ and $w=[1,-5][2,7][6]( \! ( 3,4) \! )$, then $\mu(w)=(2,2,1)$.
It follows from the results of \cite[Section 6]{N} that the interval $[e,w]$ in $\mbox{Abs}\,(B_n)$ is a lattice if $\mu(w)=(n-1,1)$
and that $[e, w]$ is not a lattice if $\mu(w)=(2,2)$.
Recall that a \emph{hook partition} is an integer partition of the form $\mu=(k,1,\dots,1)$, also written as $\mu=(k,1^r)$, where $r$ is one less than the total number of parts of $\mu$.
Our main results in this section are the following.
\begin{theorem}
\label{th5}
For $w\in B_n$, the interval $[e,w]$ in $\mbox{Abs}\,(B_n)$ is a lattice if and only if $\mu(w)$ is a hook partition.
\end{theorem}
\begin{theorem}
\label{th6}
For $w\in D_n$, the interval $[e,w]$ in $\mbox{Abs}\,(D_n)$ is a lattice if and only if $\mu(w)=(k,1)$ for some $k\leq n-1$, or $\mu(w)=(1,1,1,1)$.
\end{theorem}
We note that in view of Lemma \ref{lemarm}, Theorems \ref{th5} and \ref{th6} characterize all closed intervals in $\mbox{Abs}\,(B_n)$ and $\mbox{Abs}\,(D_n)$ which are lattices.
The following proposition provides one half of the first characterization.
\begin{propo}
\label{charlatt}
Let $w\in B_n$. If $\mu(w)$ is a hook partition,
then the interval $[e,w]$ in $\mbox{Abs}\,(B_n)$ is a lattice.
\end{propo}
\begin{proof}
Let us write $w=bp$, where $b$ (respectively, $p$) is the product of all balanced (respectively, paired) cycles of $w$.
We recall then that $[e,w]\cong[e,b]\times[e,p]$ (see Remark \ref{anal}). Since $[e,p]$ is isomorphic to a direct product of noncrossing partition lattices,
the interval $[e,w]$ is a lattice if and only if $[e,b]$ is a lattice.
Thus we may assume that $w$ is a product of disjoint balanced cycles.
Since $\mu(w)$ is a hook partition, we may further assume that $w=[1,2,\dots,k][k+1]\cdots[k+r]$ with $k+r\leq n$.
We will show that $L(k,r):=[e,w]$ is a lattice by
induction on $k+r$. The result is trivial for $k+r=2$.
Suppose that $k+r\geq 3$ and that the poset $L(k,r)$ is a lattice whenever $k+r< \kappa+\rho\leq n$.
We will show that $L(\kappa, \rho)$ is a lattice as well.
For $\rho\leq 1$, this follows from \cite[Proposition 2]{R} and the result of \cite{N} mentioned earlier.
Thus we may assume that $\rho\geq 2$. Let $u,v\in L(\kappa,\rho)$.
By \cite[Proposition 3.3.1]{St}, it suffices to show that $[e, u]\cap [e, v]=[e, z]$ for some $z\in L(\kappa, \rho)$.
Suppose first that $u(i)=i$ for some $i\in\{1,2,\dots, \kappa + \rho\}$
and let $v'$ be the signed permutation obtained by deleting the element $i$ from
the cycle decomposition of $v$.
We may assume that $u,v'\in L(\kappa_1,\rho_1)$,
where either $\kappa_1= \kappa-1$ and $\rho_1= \rho$, or $\kappa_1= \kappa $ and $\rho_1= \rho-1$.
We observe that $[e, u]\cap[e, v]=[e, u]\cap[e,v']$.
Since $L(\kappa_1,\rho_1)$ is a lattice by induction, there exists an element
$z\in L(\kappa_1,\rho_1)$ such that $[e, u]\cap[e, v']=[e, z]$.
We argue in a similar way if $v(i)=i$ for some $i\in\{1,2,\dots, \kappa + \rho\}$.
Suppose that $u(i)\neq i$ and $v(i)\neq i$ for every $i\in\{1,2,\dots, \kappa + \rho\}$.
Since $\rho\geq 2$, each of $u,v$ has at least one reflection in its cycle decomposition.
Without loss of generality, we may assume that
no cycle of $u$ is comparable to a cycle of $v$ in $\mbox{Abs}\,(B_n)$ (otherwise the result follows by induction).
Then at least one of the following holds:
\begin{itemize}[leftmargin=*, itemsep=5pt]
\item The reflection $[i]$ is a cycle of $u$ or $v$ for some $i\in\{\kappa +1, \kappa +2,\dots, \kappa + \rho\}$.
\item There exist $i,j\in\{\kappa +1, \kappa +2,\dots, \kappa + \rho\}$ with $i< j$, such that either
$( \! ( i,j) \! )$ or $( \! ( i,-j) \! )$ is a cycle of $u$ and $i$ and $j$ belong to distinct cycles of $v$, or conversely.
\item There exist $i,j\in\{\kappa +1, \kappa +2,\dots, \kappa + \rho\}$ with $i< j$, such that $( \! ( i,j) \! )$ is a cycle of $u$
and $( \! ( i,-j) \! )$ is a cycle of $v$, or conversely.
\end{itemize}
In any of the previous cases, let $u'$ and $v'$ be the permutations obtained from
$u$ and $v$, respectively, by deleting the element $i$ from their cycle decomposition.
We may assume once again that $u',v'\in L(\kappa_1,\rho_1)$,
where either $\kappa_1= \kappa-1$ and $\rho_1=\rho$, or $\kappa_1= \kappa $ and $\rho_1=\rho-1$.
As before, $[e, u]\cap[e, v]=[e,u']\cap[e, v']$.
By the induction hypothesis, $L(\kappa_1,\rho_1)$ is a lattice and hence
$[e,u']\cap[e,v']=[e,z]$
for some $z\in L(\kappa_1,\rho_1)$.
This implies that $L(\kappa,\rho)$ is a lattice and completes the induction.
\end{proof}
\noindent\emph{Proof of Theorem \ref{th5}}.
If $\mu(w)$ is a hook partition, then the result follows from Proposition \ref{charlatt}.
To prove the converse, assume that $w$ has at least two balanced cycles, say $w_1$ and $w_2$, with
$\ell_{\mathcal{T}}(w_1),\ell_{\mathcal{T}}(w_2)\geq 2$.
Then there exist $i,j,l,m\in \{\pm1,\pm2,\dots,\pm n\}$ with $|i|,|j|,|l|,|m|$ pairwise distinct, such that
$[i,j]\preceq w_1$ and $[l,m]\preceq w_2$.
However, in \cite[Section 5]{N0} it was shown that the poset $\left[e,[i,j][l,m]\right]$ is not a lattice.
It follows that neither $[e,w]$ is a lattice. This completes the proof.
\qed
\
In the sequel we denote by $L(k,r)$ the lattice $[e,w]\subset\mbox{Abs}\,(B_n)$, where
$w=[1,2,\dots,k][k+1]\cdots [k+r]\in B_n$.
Clearly, $L(k,r)$ is isomorphic to any interval of the form $[e,u]$, where $u\in B_n$ has no paired cycles and satisfies $\mu(u)=(k,1^r)$.
\
\noindent\emph{Proof of Theorem \ref{th6}}.
The argument in the proof of Theorem \ref{th5} shows that
the interval $[e,w]$ is not a lattice unless $\mu(w)$ is a hook partition.
Moreover, it is known \cite{Be,brwtt} that $[e, w]$
is a lattice if $\mu(w)=(k,1)$ for some $k \ge 1$.
Suppose that $\mu(w)=(k,1^r)$, where $r>1$ and $r+k\leq n$.
If $k \ge 2$, then there exist distinct elements
of $[e,w]$ of the form $u=[a_1,a_2][a_3]$ and $v=[a_1,a_2][a_4]$.
The intersection $[e, u]\cap [e, v]\subset \mbox{Abs}\,(D_n)$ has two maximal
elements, namely the paired reflections $( \! ( a_1,a_2) \! )$ and $( \! ( a_1,-a_2) \! )$.
This implies that $u$ and $v$ do not have a meet and therefore
the interval $[e, w]$ is not a lattice.
Suppose that $k=1$.
Without
loss of generality, we may assume that $[1][2]\cdots[r+1]\preceq w$. Suppose that $r+1\geq 5$.
We consider the elements $u=[1][2][3][4]$ and $v=[1][2][3][5]$ of $[e,w]$
and note that the intersection $[e, u]\cap[e, v]$ has three maximal
elements, namely $[1][2],[1][3]$ and $[2][3]$. This implies that the interval $[e,w]$ is not a lattice.
Finally, if $r+1=4$, then $\mu(w)=(1,1,1,1)$ and $[e,w]=[e,[1][2][3][4]]\times[e,p]$, where $p$ is a product
of disjoint paired cycles which fixes each $i\in\{1,2,3,4\}$.
Figure \ref{d4b} shows that the interval
$[e,[1][2][3][4]]$ is a lattice and hence, so is $[e, w]$. This
completes the proof.
\qed
\section{The lattice $\mathcal{L}_n$}
\label{spespe}
The poset $L(n,0)$ is the interval $[e,c]$ of $\mbox{Abs}\,(B_n)$, where $c$ is the Coxeter element $[1,2,\dots,n]$ of $B_n$.
This poset is isomorphic to the lattice $NC^B(n)$ of noncrossing partitions of type $B$.
Reiner \cite{R} computed its basic enumerative invariants listed below:
\begin{itemize}[leftmargin=*, itemsep=5pt]
\item The cardinality of $NC^B(n)$ is equal to ${2n\choose n}$.
\item The number of elements of rank $k$ is equal to ${n\choose k}^2$.
\item The zeta polynomial satisfies $Z(NC^B(n),m)={mn\choose n}$.
\item The number of maximal chains is equal to $n^n$.
\item The M\"obius function satisfies $\mu_n(\hat{0},\hat{1})=(-1)^n {2n-1\choose n}$.
\end{itemize}
\
In this section we focus on the enumerative properties of another interesting special case of
$L(k,r)$, namely the lattice $\mathcal{L}_n:=L(0,n)$.
First we describe this poset explicitly.
Each element of $\mathcal{L}_n$ can be obtained from $[1][2]\cdots[n]$ by applying repeatedly the following steps:
\begin{itemize}
\item delete some $[i]$,
\item replace a product $[i][j]$ with $( \! ( i,j) \! )$ or $( \! ( i,-j) \! )$.
\end{itemize}
Thus $w\in \mathcal{L}_n$ if and only if every nontrivial cycle of $w$ is a reflection.
In that case there is a poset isomorphism $[e,w]\cong\mathcal{L}_k\times \mathcal{B}_{l}$,
where $k$ and $l$ are the numbers of balanced and paired cycles of $w$, respectively and $\mathcal{B}_l$ denotes the lattice of subsets of the set
$\{1,2,\dots,l\}$, ordered by inclusion.
It is worth pointing out that $\mathcal{L}_n$ coincides with the subposet
of $\mbox{Abs}\,(B_n)$ induced on the set of involutions.
Figure \ref{el3} illustrates the Hasse diagram of
$\mathcal{L}_3$.
\begin{figure}[h]
\begin{center}
\includegraphics[width=5.2in]{el3}
\end{center}
\caption{}
\label{el3}
\end{figure}
In Proposition \ref{k=0} we give the analogue of the previous list for the lattice $\mathcal{L}_n$.
We recall that the zeta polynomial $Z(P,m)$ of a finite poset $P$ counts the number of multichains $x_1\leq x_2\leq\cdots\leq x_{m-1}$ of $P$.
It is known (see \cite{Ed}, \cite[Proposition 3.11.1]{St}) that $Z(P,m)$ is a polynomial function of $m$ of degree $n$, where $n$ is the length of $P$ and that $Z(P,2)=\#P$.
Moreover, the leading coefficient of $Z(P,m)$ is equal to the number of maximal chains divided by $n!$ and if $P$ is bounded, then $Z(P,-1)=\mu(\hat{0},\hat{1})$.
\begin{propo}
\label{k=0}
\emph{For the lattice $\mathcal{L}_n$ the following hold:}
\begin{enumerate}[leftmargin=*, itemsep=5pt]
\item [\emph{(i)}]
The number of elements of $\mathcal{L}_n$ is equal to
\[\sum\limits_{k=0}\limits^{\left\lfloor \nicefrac{n}{2}\right\rfloor} {n\choose 2k}\,2^{n-k}(2k-1)!!,\]
where $(2m-1)!!=1\cdot3\,\cdots\, (2m-1)$ for positive integers $m$.
\item [\emph{(ii)}] The number of elements of $\mathcal{L}_n$ of rank $r$ is equal to
\end{enumerate}
\[\displaystyle\sum\limits_{k=0}\limits^{\min\{r,n-r\}}\frac{n!}{k!(r-k)!(n-r-k)!}.\]
\begin{enumerate}[leftmargin=*, itemsep=5pt]
\item [\emph{(iii)}]
The zeta polynomial $Z_n$ of $\mathcal{L}_n$ is given by the formula
\[Z_n(m)=\displaystyle\sum\limits_{k=0}\limits^{\left\lfloor \nicefrac{n}{2}\right\rfloor} {n\choose 2k}\,m^{n-k}(m-1)^k(2k-1)!!.\]
\item [\emph{(iv)}]
The number of maximal chains of $\mathcal{L}_n$ is equal to\
\[n!\sum\limits_{k=0}\limits^{\left\lfloor \nicefrac{n}{2}\right\rfloor} {n\choose 2k}(2k-1)!!.\]
\item [\emph{(v)}]
For the M\"obius function $\mu_n$ of $\mathcal{L}_n$ we have
\[\displaystyle\mu_n(\hat{0},\hat{1})=(-1)^n\sum\limits_{k=0}\limits^{\left\lfloor \nicefrac{n}{2}\right\rfloor} {n\choose 2k}\,2^k(2k-1)!!,\]
where $\hat{0}$ and $\hat{1}$ denotes the minimum and the maximum element of $\mathcal{L}_n$, respectively.
\end{enumerate}
\end{propo}
\begin{proof}
Suppose that $x$ has $k$ paired reflections.
These can be chosen in $2^k{n\choose 2k}(2k-1)!!$ ways.
On the other hand, the balanced reflections of $w$ can be chosen in $2^{n-2k}$ ways. Therefore the cardinality of $\mathcal{L}_n$ is equal to
\[\sum\limits_{k=0}\limits^{\left\lfloor \nicefrac{n}{2}\right\rfloor} {n\choose 2k}\,2^{n-k}(2k-1)!!.\]
The same argument shows that the number of elements of $\mathcal{L}_n$ of rank $r$, where $r\leq \left\lfloor \nicefrac{n}{2}\right\rfloor$ is equal to
\begin{eqnarray*}
\sum\limits_{k=0}\limits^r2^k{n\choose 2k}(2k-1)!!\,{n-2k\choose r-k}&=&\sum\limits_{k=0}\limits^r2^k{n\choose 2k}\frac{(2k)!}{2^k\,k!}\,{n-2k\choose r-k}\\
&=&\sum\limits_{k=0}\limits^r\frac{n!}{k!(r-k)!(n-r-k)!}.
\end{eqnarray*}
Since $\mathcal{L}_n$ is self dual, the number of elements in $\mathcal{L}_n$ of rank
$r$ is equal to the number of those that have rank $n-r$.
The number of multichains in $\mathcal{L}_n$ in which $k$ distinct paired reflections appear, is equal to ${n\choose 2k}(2k-1)!!(m(m-1))^k m^{n-2k}$. Therefore, the zeta polynomial of $\mathcal{L}_n$ is given by
\[Z_n(m)=\sum\limits_{k=0}\limits^{\left\lfloor \nicefrac{n}{2}\right\rfloor} {n\choose 2k}(2k-1)!!\,m^{n-k}(m-1)^k.\]
Finally, computing the coefficient of $m^n$ in this expression for $Z_n(m)$ and multiplying by $n!$ we conclude that
the number of maximal chains of $\mathcal{L}_n$ is equal to
\[n!\sum\limits_{k=0}\limits^{\left\lfloor \nicefrac{n}{2}\right\rfloor} {n\choose 2k}(2k-1)!!\]
and setting $m=-1$ we get
\[\mu_n(\hat{0},\hat{1})=Z_n(-1)=(-1)^n\sum\limits_{k=0}\limits^{\left\lfloor \nicefrac{n}{2}\right\rfloor} {n\choose 2k}(2k-1)!!\,2^k.\]
\end{proof}
\begin{remark
\noindent By Proposition \ref{el}, the lattice $\mathcal{L}_n$ is EL-shellable.
We describe two more EL-labelings for $\mathcal{L}_n$.
\begin{enumerate}[leftmargin=*, itemsep=5pt]
\item [(i)] Let $\Lambda=\{[i]: i=1,2,\dots,n\}\cup\{( \! ( i,j) \! ):i,j=1,2,\dots,n,\,i<j\}$.
We linearly order the elements of $\Lambda$ in the following way. We first order the balanced reflections
so that $[i]<_{\Lambda}[j]$ if and only if $i<j$. Then we order the paired reflections lexicographically.
Finally, we define $[n]<_{\Lambda}\lleft1,2) \! )$.
The map $\lambda_1:C(B_n)\to \Lambda$ defined as:
\[\lambda_1(a,b)=\left\{\begin{array}{ll}
[i] & \mbox{if $a^{-1}b=[i]$}, \\
( \! ( i,j) \! ) & \mbox{if $a^{-1}b=( \! ( i,j) \! )$ or $( \! ( i,-j) \! )$}
\end{array}
\right.\]
is an EL-labeling for $\mathcal{L}_n$.
\end{enumerate}
\begin{enumerate}[leftmargin=*, itemsep=5pt]
\item [(ii)] Let $\mathcal{T}$ be the set of reflections of $B_n$.
We define a total order $<_{\mathcal{T}}$ on $\mathcal{T}$
which extends the order $<_{\Lambda}$, by ordering the reflections
$( \! ( i,-j) \! )$, for $1\leq i<j\leq n$, lexicographically and letting
$( \! ( n-1,n) \! )<_{\mathcal{T}}\lleft1,-2) \! )$.
For example, if $n=3$ we have the order
$[1]_{\mathcal{T}}<_{\mathcal{T}}[2]<_{\mathcal{T}}[3]<_{\mathcal{T}}\lleft1,2) \! )<_{\mathcal{T}}\lleft1,3) \! )<_{\mathcal{T}}\lleft2,3) \! )<_{\mathcal{T}}\lleft1,-2) \! )<_{\mathcal{T}}\lleft1,-3) \! )<_{\mathcal{T}}\lleft2,-3) \! )$.
Let $t_i$ be the $i$-th reflection in the order above. We define a map $\lambda_2:C(B_n)\to \{1,2,\dots,n^2\}$ as:
\end{enumerate}
\[\lambda_2(a,b)=\min\limits_{1\leq i\leq n^2}\{i:\,t_i\vee a=b\}.\]
\begin{enumerate}[leftmargin=*, itemsep=5pt]
\item []The map $\lambda_2$ is an EL-labeling for $\mathcal{L}_n$.
\end{enumerate}
See Figure \ref{elp(0,2)n} for an example of these two EL-labelings when $n=2$.
\end{remark}
\begin{figure}[h]
\begin{center}
\includegraphics[width=4.5in]{elp0,2}
\end{center}
\caption{}
\label{elp(0,2)n}
\end{figure}
\section{Enumerative combinatorics of $L(k,r)$}
\label{telostelos}
In this section we compute the cardinality, zeta polynomial and M\"obius function of the lattice $L(k,r)$,
where $k,r$ are nonnegative integers with $k+r=n$.
The case $k=n-1$ was treated by Goulden, Nica and Oancea in their work \cite{N} on the posets of annular noncrossing partitions;
see also \cite{KM,N0} for related work. We will use their results, as well as the formulas for cardinality and zeta polynomial
for $NC^B(n)$ and Proposition \ref{k=0},
to find the corresponding formulas for $L(k,r)$.
\begin{propo}
\label{k,r}
Let $\alpha_r=|\mathcal{L}_r|,\,\beta_r(m)=Z(\mathcal{L}_r,m)$
and $\mu_r=\mu_r(\mathcal{L}_r)$, where $\alpha_r=\beta_r(m)=\mu_r=1$ for $r=0,1$.
For fixed nonnegative integers $k,r$ such that $k+r=n$, the cardinality, zeta polynomial and M\"obius function of $L(k,r)$
are given by:
\begin{itemize}[leftmargin=*, itemsep=5pt]
\item
$\#L(k,r)=\displaystyle {2k\choose k}\left(\frac{2\,r\,k}{k+1}\,\alpha_{r-1}+a_r\right)$.
\item
$Z(L(k,r),m)=\displaystyle{mk\choose k}\left( \frac{2\,r\,k}{k+1}(m-1)\,\beta_{r-1}(m)+\beta_r(m)\right)$.
\item $\mu(L(k,r))=\displaystyle (-1)^n{2k-1\choose k}\left(\frac{4\,r\,k}{k+1}\,|\mu_{r-1}| + |\mu_r|\right)$.
\end{itemize}
\end{propo}
\begin{proof}
We denote by $A$ the subset of $L(k,r)$ which consists of the elements $x$ with the following property:
every cycle of $x$ that contains at least one of $\pm1,\pm2,\dots,\pm k$ is less than or equal to
the element $[1,2,\dots,k]$ in $\mbox{Abs}\,(B_n)$.
Let $x=x_1x_2\cdots x_{\nu}\in A$, written as a product of disjoint cycles.
Without loss of generality, we may assume that there is a
$\,t\in\{0,1,\dots,\nu\}$ such that $x_1x_2\cdots x_t\preceq [1,2,\dots,k]$ and
$x_{t+1}x_{t+2}\cdots x_{\nu}\preceq[k+1][k+2]\cdots[k+r]$.
Observe that if $t=0$ then $x\preceq[k+1][k+2]\cdots[k+r]$ in $\mbox{Abs}\,(B_n)$,
while if $t=\nu$ then $x\preceq[1,2,\dots, k]$.
Clearly, there exists a poset isomorphism
\begin{alignat*}{3}
f: A & \to NC^B(k) \ \ && \times && \ \ \langle[k+1]\cdots[k+r]\rangle \\
x & \mapsto (x_1\cdots x_t \ \ && \ , && \ \ \,x_{t+1}\cdots x_{\nu}),
\end{alignat*}
\noindent so that
\begin{equation}
\label{A}
A\cong NC^B(k)\times \mathcal{L}_r.
\end{equation}
Let $C=L(k,r)\smallsetminus A$ and $x=x_1x_2\cdots x_{\nu}\in C$, written as a product of disjoint cycles.
Then there is a exactly one paired cycle $x_1$ of $x$ and one reflection $( \! ( i, l) \! )$ with
$i\in\{\pm 1,\pm2,\dots, \pm k\},\ l\in\{k+1,k+2,\dots,k+r\}$, such that $( \! ( i, l) \! )\preceq x_1$.
For every $l\in \{k+1,k+2,\dots,k+r\}$ denote by $C_l$ the set of permutations
$x\in L(k,r)$ which have a cycle, say $x_1$, such that $( \! ( i, l) \! )\preceq x_1$ for some
$i\in \{\pm 1,\pm 2,\dots,\pm k\}$.
It follows that $C_l\cap C_{l'}=\varnothing$ for $l\neq l'$.
Clearly, $C_l\cong C_{l'}$ for $l\neq l'$ and $C=\bigcup_{l=k+1}^{k+r}C_l$.
Summarizing, for every $x\in C$ there exists an ordering
$x_1,x_2,\dots,x_{\nu}$ of the cycles of $x$ and a unique index $t\in\{1,2,\dots,\nu\}$
such that $x_1x_2\cdots x_t\preceq [1,2,\dots,k][l]$ and
$x_{t+1}x_{t+2}\cdots x_{\nu}\preceq[k+1][k+2]\cdots[l-1][l+1]\cdots[k+r]$.
Let \[E_l=\{x\in C: x\preceq[1,2,\dots,k][l]\}.\] We remark that no permutation of $E_l$ has a balanced cycle in its cycle decomposition.
Clearly, there exists a poset isomorphism
\begin{alignat*}{3}
g_l: C_l & \to\ \ \ \ \, E_l \ \ && \times && \ \ \langle[k+1]\cdots[l-1][l+1]\cdots[k+r]\rangle \\
x & \mapsto (x_1\cdots x_t \ \ && \ , && \ \ x_{t+1}\cdots x_{\nu})
\end{alignat*}
\noindent so that
\begin{equation}
\label{C_l}
C_l\cong E_l\times \mathcal{L}_{r-1}
\end{equation}
for every $l\in\{k+1,k+2,\dots,k+r\}$. Using (\ref{A}) and (\ref{C_l}), we proceed to the proof of Proposition \ref{k,r}
as follows.
From our previous discussion we have $L(k,r)=\#A+r\,(\#C_{k+1})$.
From (\ref{A}) we have \[\#A={2k\choose k}\alpha_r\]
and (\ref{C_l}) implies that $\#C_{k+1}=(\#E_{k+1})\,(\#\mathcal{L}_{r-1})=(\#E_{k+1})\, \alpha_{r-1}$.
Since $E_{k+1}$ consists of the permutations in $\langle[1,2,\dots,k][k+1]\rangle\cap C$,
it follows from \cite[Section 5]{N} that $\#E_{k+1}=2{2k\choose k-1}$.
Therefore,
\[\displaystyle \#L(k,r)= 2\,r\,{2k\choose k-1}\alpha_{r-1}+{2k\choose k}\alpha_r= {2k\choose k}\left(\frac{2r k}{k+1} \alpha_{r-1}+a_r\right).\]
\
Recall that the zeta polynomial $Z(L(k,r),m)$ counts the number of multichains
$\pi_1\preceq \pi_2\preceq\cdots\preceq\pi_{m-1}$ in $L(k,r)$.
We distinguish two cases. If $\pi_{m-1}\in C$, then $\pi_{m-1}\in C_l$ for some $l\in \{k+1,\dots,k+r\}$.
Isomorphism (\ref{C_l}) then implies that there are $Z(E_l,m)\, Z(\mathcal{L}_{r-1},m)$ such multichains.
From \cite[Section 5]{N} we have $Z(E_l,m)=2{mk\choose k+1}$, therefore
$Z(E_l,m)\, Z(\mathcal{L}_{r-1},m)=2{mk\choose k+1} \beta_{r-1}$.
Since there are $r$ choices for the set $C_l$, we conclude that the number of multichains
$\pi_1\preceq \pi_2\preceq\cdots\preceq\pi_{m-1}$ in $L(k,r)$ for which $\pi_{m-1}\in C$ is equal to
\begin{equation}
\label{z1}
2\, r\,{mk\choose k+1} \beta_{r-1}(m).
\end{equation}
\noindent If $\pi_{m-1}\in A$, then $\pi_{m-1}\in NC^B(k)\times \mathcal{L}_r$ and therefore
number of such multichains is equal to
\begin{equation}
\label{z2}
{mk\choose k} \beta_r(m).
\end{equation}
\noindent The proposed expression for the zeta polynomial of $L(k,r)$ follows by summing the expressions (\ref{z1}) and (\ref{z2})
and straightforward calculation.
The expression for the M\"obius function follows once again from that of the zeta polynomial by
setting $m=-1$.
\end{proof}
\input{appendix}
\
\subsection*{Acknowledgments}
I am grateful to Christos Athanasiadis for valuable conversations, for his encouragement and for
his careful reading and comments on preliminary versions of this paper. I would also like to thank
Christian Krattenthaler and Victor Reiner for helpful discussions and Volkmar Welker for bringing reference \cite{bww}
to my attention. A summary of the results of this paper has appeared in \cite{myrr}.
|
Subsets and Splits